Skip to main content
Erschienen in: Wiener Medizinische Wochenschrift 7-8/2016

Open Access 01.05.2016 | main topic

Viral entry pathways: the example of common cold viruses

verfasst von: Dieter Blaas

Erschienen in: Wiener Medizinische Wochenschrift | Ausgabe 7-8/2016

Summary

For infection, viruses deliver their genomes into the host cell. These nucleic acids are usually tightly packed within the viral capsid, which, in turn, is often further enveloped within a lipid membrane. Both protect them against the hostile environment. Proteins and/or lipids on the viral particle promote attachment to the cell surface and internalization. They are likewise often involved in release of the genome inside the cell for its use as a blueprint for production of new viruses. In the following, I shall cursorily discuss the early more general steps of viral infection that include receptor recognition, uptake into the cell, and uncoating of the viral genome. The later sections will concentrate on human rhinoviruses, the main cause of the common cold, with respect to the above processes. Much of what is known on the underlying mechanisms has been worked out by Renate Fuchs at the Medical University of Vienna.

Introduction

The first encounter between a virus and its host cell usually takes place via a protein, a proteoglycan, an oligosaccharide, or a glycolipid exposed on the cell surface that is recognized by cognate viral surface components. Only few viruses can subsequently penetrate directly from the plasma membrane into the cytosol of the host cell; the large majority rather exploits cellular entry pathways by travelling inside a membrane vesicle. Under physiologic conditions, these cell surface molecules serve as nutrient transporters or signal transduction receptors, or in cell–cell interactions and attachment, amongst other things. In the context of viral infection they have been termed “viral receptors” although the cell obviously does not make them for the purpose of becoming infected.
Viruses not only abuse membrane receptors but also soluble molecules, as a bridge to a plasma membrane protein or to bind to virus-specific antibodies that are present in the serum as a consequence of a previous infection. The latter, in turn, can then bind Fcγ-receptors present at the surface of some specialized cells and thus link the virus to them. For instance, virus-specific immunoglobulin G (IgG) can enhance infection by various flaviviruses, including, e.g., Dengue virus, rather than protecting the host against infection [1]. Other examples are the binding of apolipoprotein-E to both hepatitis C virus and heparan sulfate proteoglycans, which also leads to a connection between the virus and a cell surface component [2], and growth arrest-specific 6 (Gas6) bridging TAM receptors (various receptor tyrosine kinases) with phosphatidylserine that is present in the viral lipid membrane of Dengue and Vaccinia virus [3]. So, viruses have evolved to make use of whatever is available for attaching to and entering the host cell. Clearly, the presence or absence of a cognate viral receptor is a major but not the only factor in species and tissue tropism of a given virus [4]. It is of note that targeting viruses with oncolytic potential to tumors for their infection and destruction has been achieved by copying and exploiting this strategy [5].
Following the mutual recognition, the viral particle is taken up into the cell, either constitutively or triggered through its interaction with the receptor. The many identical subunits of a virion (i.e., a viral particle) render it multivalent; this can lead to clustering of, e. g., receptor tyrosine kinases or of integrins. For example, Vaccinia virus triggers activation of phosphatidylinositol 3‑kinase (PI3K)/Akt in an integrin β1-dependent manner, suggesting that this particular signaling pathway is essential for virus endocytosis [6]. Entry into the cell then engenders lipid vesicle trafficking processes. Reminiscent of the natural ligand, most intruding viruses become trapped where shallow membrane pits form; upon further membrane bending, a deep invagination is produced and finally severed from the plasma membrane in the form of a closed vesicle. The vesicle pinches off the inner side of the plasma membrane and carries its viral cargo, together with some extracellular fluid, through the crowded cytoplasm towards its intracellular destination. The transport vesicles undergo maturation by fusing with other vesicles, severing, and content sorting, all taking place during their voyage inside the cell. The vesicular membrane composition and intravesicular milieu continuously change along this route. Molecular machines, such as dyneins and kinesins, are involved in transporting virus-containing vesicles—or free virus particles—along actin fibers or microtubules [79].
Material inside the vesicle lumen is topologically extracellular and thus requires that either the entire virus or at least its genome penetrate the delimiting membrane to attain the cytosol. This occurs as soon as membrane composition, pH, ionic environment, etc. have become optimal for this process and a given virus; it can happen in early endosomes, late endosomes, recycling endosomes, macropinosomes, the endoplasmic reticulum, in various subcompartments of the Golgi, and even in lysosomes with their interior full of hostile hydrolytic enzymes.
Membrane penetration is usually preceded by or occurs in concert with conformational changes of viral surface proteins and/or the whole virion shell itself. In many cases, these structural changes require that an envelope protein, such as the hemagglutinin of influenza viruses, has been proteolytically cleaved during maturation [10]. A hydrophobic or amphiphilic fusion peptide then becomes exposed and inserts into the plasma membrane; in non-enveloped viruses, membrane destabilization or even disruption is brought about by the release of “membranolytic peptides”, small amphipathic viral proteins [11]. Another means of membrane destabilization was shown for canine parvovirus, which has a phospholipase A(2)-like domain in the N‑terminus of its capsid protein VP1, whose enzymatic activity makes the membrane permeable for dextran of 3 kD but not of 10 kD [12]. Recently, the perfidiousness of viral entry has been demonstrated for adenovirus; this virus induces small pores via its membrane lytic protein-VI, which triggers calcium-mediated lysosomal exocytosis repair pathways and lipid signaling that finally facilitates uptake of the pathogen [13]. In all these cases, once the nucleic acid(s) has arrived in the cytosol, depending on the type of virus and the nature of its genome, replication is either initiated right there, or after it has been shuttled to the nucleus.
The present article is aimed at providing a brief overview on these early events of viral infection. Its extent is far from covering all aspects of these complex and in part poorly understood processes. I apologize for citing only few, mostly recent and in part arbitrarily chosen publications referring to examples out of the about 17,500 papers found in a PubMed literature search for “virus AND entry” at the time of writing. For recent excellent and more in-depth reviews on viral entry and uncoating, see, for example, [1423].

Naked and enveloped viruses

Viruses come essentially in two distinct flavors; either carrying a lipid membrane envelope decorated with viral proteins, or naked, i. e., without any lipid. The former have built-in machinery for penetration of their capsid into the cytosol via fusion of viral with cellular membranes. Most often, this process is triggered by the acid environment (pH between about 5 and 6) established inside the endocytic vesicles during their maturation into late endosomes [2426]; there are only few viruses, including human immunodeficiency virus (HIV) and herpes simplex virus-1, which do not require an acidic environment and can thus even fuse with the plasma membrane at neutral pH [27, 28]. In any case, a structural change of a conformationally metastable viral surface protein needs to occur; this results in exposure of a previously hidden fusion peptide, usually a stretch of amphipathic and/or hydrophobic amino acid residues. The metastable state is often prepared during maturation via cleavage of the protein, setting the trigger for this sequence to insert into the cellular membrane upon arrival in a low-pH environment. The ensuing conformational change forces the viral membrane and the cellular membrane into close apposition. The result is hemifusion (i. e., an intermediate state without content mixing), and finally the complete melding of the two membranes, including lipid mixing and delivery of the nucleocapsid into the cytosol (Fig. 1).
Whereas membrane fusion is quite well understood, penetration of naked viruses still holds a number of open questions; in principle, access of the viral genome to the cytoplasm might either occur via disruption of the virus-containing vesicle—with the entire virion, together with other endosomal content, being released into the cytosol—or via pores of limited size in the endosomal membrane. Such pores are presumably lined by domains of viral proteins forming a channel contiguous with a pore in the viral shell. The holes in the virus capsid open upon a conformational switch, again most often triggered by the acidic pH and sometimes assisted by the viral receptor [29]. As described in more detail at the end of this article, by using common cold viruses as examples of cellular entry by non-enveloped viruses, Renate Fuchs has been working for many years on unravelling viral uptake and trafficking of viruses inside the cell, and the transfer of the viral RNA genome into the cytosol (see reviews [30, 31]). Because at least three different receptors are used by the three rhinovirus species, it is no wonder that they enter host cells via different endocytic pathways [3234]. As a result, uncoating, i. e., release of their positive-sense single stranded RNA genome, might occur in different cellular compartments via different mechanisms, which is the subject of the last sections of this article.

Viral attachment—role and function of the receptors

Parameters governing interactions between viruses and their cognate receptors have been investigated in vitro mostly by surface plasmon resonance methodology, for an example, see [35]. Virus is immobilized on a dextran-modified metal surface, a solution containing recombinant soluble receptor (i. e., lacking the membrane anchor) is constantly flown over the chip, and binding is monitored online as a mass-dependent change of the surface plasmon resonance angle. On washing with plain buffer, previously bound receptor dissociates, returning the signal to baseline. From these binding/unbinding profiles, the on and off rates and the affinity constants can be derived.
Capillary electrophoresis was demonstrated to yield information on the number of soluble receptors bound per virion by resolving virus with zero and up to 12 attached receptors. In addition, a rough estimate of the affinity constant could be derived [36]. Here, both components are free in solution and their concentrations and those of their complexes at different stoichiometry are measured.
Atomic force microscopy can deliver a detailed energy landscape of the unbinding forces when the virus is pulled away from the receptors and the (noncovalent) bonds between them are ruptured sequentially [3739].
Finally, information on the spatial arrangement of the five ligand-binding modules of a soluble very-low density lipoprotein receptor (VLDLR) concatemer construct on the surface of a rhinovirus could be inferred from fluorescence resonance energy transfer between its N‑ and C‑termini. Quenching was recorded upon attachment. This demonstrated that the ends indeed come close to each other when the receptor molecules wrap around each of the vertices at the 5‑fold axes of icosahedral symmetry [40].
Although providing valuable information, the experimental conditions of at least the first two methods are not representative of the in vivo situation where the receptor is anchored in the cellular membrane. In this latter context, binding of a virus to its membrane receptor is not always as straightforward; it might be preceded by relatively unspecific electrostatic interactions with charged molecules at the cell surface, such as heparan sulfate or sialic acid, either tethering the virions to the plasma membrane and thus increasing their local concentration, or rather hampering their access to the receptor(s) and thus decreasing their local concentration [41]. Such effects are difficult to investigate; live-cell single-particle tracking methods [42] have suggested that some viruses remain on the cell surface for extended periods of time wandering. According to the “seek and stick” paradigm, a virus would thus diffuse in the plane of the membrane, bound weakly and rather unspecifically to main components of the glycocalyx, such as proteoglycans, glycolipids, etc., until it encounters its specific, higher-affinity receptor(s), usually present at lower density. Single receptor molecules might cluster during this diffusion in two dimensions, as a consequence of the multivalence of the virion. This can increase the avidity of the interaction, and, at the same time, slow down the speed of diffusion, i. e., the virus would collect several receptors on this excursion. In some specific cases, such avidity effects might turn on the endocytosis machinery at a site of multivalently attached virus by relaying signals into the cell [43]. There are also other ways one can envisage the sequence of events leading to formation of a virus–receptor complex ready to be engulfed; to give just one example, the high-affinity receptors might be confined to special microdomains (e. g., lipid rafts enriched in cholesterol), requiring the virus to first arrive in such microdomains by diffusion in order for tight binding to become possible [14]. High-speed particle tracking experiments have suggested that the plasma membrane is not a two-dimensional continuum fluid, but rather contains submicron compartments with different fluidity [44]. Reminiscent of signal transduction via bi- or oligomeric natural ligand molecules, the cytoskeleton might also be involved in limiting virus movements or in relaying viral entry via signals triggered through virus-induced clustering of the membrane receptors. It must be taken into account that the number of receptor molecules simultaneously attached to a single virion might impact on efficiency and speed of the uptake, as well as on uncoating later in the entry pathway, e. g., by exerting disruptive strain. For example, uncoating of HIV-1 was shown to depend on dynein and kinesin 1. It was suggested that these motors exert mechanical forces “tearing apart the virion” [45]. Another example is the pull of kinesin on adenovirus capsids when bound to the nuclear pore complex via Nup214 [46]. In addition, transport of receptor-bound virus to sites of active endocytosis (e. g., from filopodia to the cell body) has been observed.

Uptake of the virus into the cell

Uptake of cargo, be it for nutrition, signaling, or for downregulating signals by ferrying receptor–ligand complexes to lysosomes for degradation, can occur mainly by i) clathrin-mediated endocytosis (CME), ii) caveolin-mediated uptake (CavME), iii) macropinocytosis, and iv), poorly characterized uptake mechanisms involving neither clathrin nor caveolin. These processes are usually summarized under the name of endocytosis. It is of note that often none of these pathways is exclusively exploited by a given virus; even more so when one entry route is blocked, e. g., by a specific inhibitor. In this situation, the cell might compensate by upregulating another pathway that will then also be used by the virus [47]. By the same token, a given virus might prefer different pathways in different cells [48].
Various components of the internalization machinery can be targeted by chemical inhibitors or ablated through genetic approaches such as RNA knockdown [49]. Recently, high-throughput screening of gene knockout libraries generated via CRISPR/Cas9 allowed identification of factors involved in uptake of bacterial toxins [50]; obviously, the same techniques can be used to find genes involved in viral endocytosis. Nevertheless, such results are sometimes ambiguous, either because of low specificity, redundancy of the targeted factors, or because the inhibited components take part in more than one pathway. Systematic and bioinformatics-guided RNAi screens have recently been adopted to identify proteins of the host cell involved in uptake of viruses [5153]. This method involves cell transfection with short ‘interfering RNAs’ with complementarity to sequences of a target mRNA to be silenced. Their hybridization results in their degradation and consequently in downregulation of the encoded protein.
Finally, haploid mutant cell libraries—with each clone having a different gene inactivated—have been prepared. Screening such libraries for reduced viral replication has led to the identification of host components essential in virus uptake and to drugs specifically inhibiting the function of these factors [54, 55]. Such cellular components are clearly less prone to escape by viral mutation and are increasingly considered as possible targets for antiviral drugs.

Clathrin-mediated endocytosis

CME (Fig. 2 pathway 1) is the by far best understood mechanism of cellular uptake (for an instructive movie see ref [56]). Many ligands (e. g., low-density lipoprotein, chylomicrons, Fe3+-saturated transferrin, growth factors, etc.) are channeled into this system via receptors whose cytoplasmic tails feature amino acid (AA) sequence motives like Asn-Pro-X-Tyr (X is any AA) or two consecutive leucines (the di-leucine motive) [57]. These signatures are recognized by the adapter complex AP-2 that tethers soluble clathrin monomers to the inner side of the plasma membrane for polymerization into a coat easily seen under the electron microscope. More than fifty other components are involved in CME; the membrane is curbed to build a shallow basket or pit that subsequently grows to form an invagination that becomes constricted and pinches off the plasma membrane. The main players in this process are the GTPase dynamin, the membrane-bending protein amphiphysin, and the AP-2 binding protein eps15 [58]. Finally, a clathrin-coated vesicle is formed [59]; once free in the cytosol, this coat is again removed under ATP hydrolysis, making the membrane accessible for fusion and fission with other vesicles [60]. During these maturation/fusion steps new proteins are acquired, amongst them the vesicular ATPase complex that acidifies the vesicular lumen by pumping H+ into the vesicle, creating an increasingly acidic intravesicular environment and a membrane potential [26]. The endocytic vesicles move on and finally fuse with lysosomes, the disassembly and degradation factories of the cell. Thus, viruses need mechanisms to release their precious genome undamaged into the cytosol to avoid destruction at the end of this itinerary by the aggressive lysosome.
CME can be blocked by various more or less specific chemical inhibitors, amongst them chlorpromazine and dynasore, by siRNAs (small interfering RNAs, see also RNAi above) knocking down the clathrin heavy chain and/or dynamin-2 (dyn2), and by expression of a dominant-negative form of this latter protein (dynK44A) or of eps15 lacking the modules that interact with AP2 [61]. Chlorpromazine relocates clathrin and the adaptor complex AP-2 from coated pits to vesicles [62]; dynasore, the dyngos, dynols, and the iminodyns are all inhibitors of dynamin’s GTPase activity and, as such, target all processes involving dynamins [63, 64]. Pitstop inhibits interaction of amphiphysin with the amino terminal domain of clathrin and was thus expected to be specific for CME. However, it appears to also affect various forms of clathrin-independent endocytosis (CIE) [65]. Other measures, like potassium depletion or incubation of the cells with hypertonic sucrose, have been shown to dissociate polymerized clathrin and thereby inhibit entry of viruses by CME [66]. So, the best way to identify the main entry pathway exploited by a given virus is via employing several of the methods mentioned above in parallel.

Caveolae-mediated endocytosis (CavME)

There are substantially fewer reports on viral entry involving caveolae than clathrin-coated pits, and only few viruses have been found to preferentially exploit this route. In addition, early reports on caveosomes, vesicles presumably derived from caveolae [67], in viral entry have been partially superseded by later results, e. g., see the case of echovirus 1 [68, 69]. Morphologically, caveolae are easily distinguishable from clathrin-coated pits; in the transmission electron microscope they show a typical elongated form reminiscent of round-bottom flasks extending into the cytosol, with a cavin coat that is less electron-dense than clathrin [70]. Their necks are open at the cell surface ([71]; Fig. 2 pathway 2). Apart from caveolin-1, a main component, caveolae possess a particular lipid composition with high cholesterol and glycosphingolipid content, and accumulate membrane proteins with glycosylphosphatidyl-anchors; as caveolae originate from cholesterol-rich lipid rafts, their lipid composition is similar. Viruses that might enter via this route include simian virus 40 (SV40), polyomavirus, echovirus 1 (but see above), certain coxsackieviruses, and some others [67]. Compared to CME, entry via caveolae is slow, i. e., the viruses might remain for hours unchanged in caveosomes, a distinct class of caveolin-1-containing endosomes identified in some cell lines. Similar to the reports on echovirus 1, SV40 entry in the absence of caveosome formation was also demonstrated. This questions the role of caveolin in the uptake of these particular viruses [72].
There are inhibitors that are, to some low extent only, specific for the caveolar pathway, such as the protein kinase inhibitor genistein. A combination of the cholesterol-sequestering drugs nystatin or filipin with the cholesterol-synthesis inhibitor progesterone blocks the formation of caveolae as summarized in [73]. Since severing of caveolae from the plasma membrane also involves dynamin, drugs blocking its function also block the caveolar pathway (see above).

Macropinocytosis

During macropinocytosis cellular protrusions are actively formed and fold back onto the plasma membrane under the direction of actin assembly (Fig. 2 pathway 3; [74]); the process primarily takes up liquid but also material bound to the region where the vesicles form [75]. Few viruses have been shown to definitely enter via this pathway [76]; amongst them a murine amphotropic retrovirus [77], which this was inferred from its uptake into fibroblasts lacking caveolin or dynamin. Cytochalasin and latrunculin block both macropinocytosis and actin-dependent phagocytosis that is responsible for engulfment of larger particles, such as bacteria, and were shown to also reduce entry of this virus. Amiloride acts on Na+/H+ exchangers at the plasma membrane and has been used as a moderately specific inhibitor of macropinocytosis. At least in A431 cells it was shown to increase the submembranous pH that is normally lowered by metabolically generated acid; this, in turn, inhibits activation of GTPases involved in actin remodeling [78]. However, amiloride and it derivatives can also block later steps in viral synthesis, like replication of the nucleic acid; therefore, results with this drug need to be controlled with much care. The inhibitor profiles are often not exactly the same for the uptake of different viruses, making it difficult to term the process “typical” macropinocytosis. This might result from the virus exploiting other pathways in parallel to different degrees.

“Virus-made invaginations”

The unique structure of the GM1 ganglioside, which can act as an SV40 receptor, together with the multitude of binding sites on the virion can promote membrane curvature and internalization by itself, in the absence of any invagination-promoting coats. Remarkably, even long tubules containing a multitude of viruses arranged as pearls on a string were observed [79]. Similarly, virus-like particles derived from a norovirus were shown to induce negative membrane curvature on binding to glycosphinoglipids present in giant unilamellar vesicles [80]. Although different from “pure” lipid-mediated endocytosis studied in liposomes [81], a similar effect might be responsible for deep invaginations seen around a rhinovirus (RV-A2) when bound to concatemers of five copies of repeat 3 of VLDLR, recombinant high-affinity protein receptors that were attached to liposomes via a his6-tag, which, in turn, was bound to Ni-nitrilotriacetate (Ni-NTA) lipids incorporated in the lipid bilayer [82]. Clearly, all cellular components involved in membrane-bending and fission were lacking in these latter highly artificial systems and no vesicles were observed inside the liposomes; formation of these pits and tubules might be strongly dependent on concentration, affinity, and size of the receptors. Similar tubules filled with intercellular adhesion molecule 1 (ICAM-1)-binding rhinoviruses had been observed in rhabdomyosarcoma cells overexpressing this receptor [33]. It is unknown whether a similar entry mechanism operates in vivo in cells expressing the respective receptors—protein but not lipid—at high concentration.

Entry of common cold viruses—a paradigm for naked virus entry

Rhinoviruses (RVs) are the main cause of common colds. They are of enormous economic impact, estimated at about 25 billion USD per year in the USA due to absences from work, spending for medicines, and doctors’ visits [83]. Infections are recurrent and often involve different RV serotypes. Nevertheless, immunization with recombinant viral proteins seems to result in antibodies more broadly cross-reacting; however, again only with low cross-neutralization efficiency [8487]. A major challenge for production of a vaccine is the large number of different viral serotypes, which would require the formulation to contain at least one representative of each cluster of the (weakly) cross-reacting antigens, which is difficult to realize.

Rhinovirus receptors

RVs are of comparatively simple architecture; an icosahedral shell with T = 1, P = 3 symmetry is built from sixty copies of four different capsid proteins, VP1 through VP4, that enwrap a single-stranded positive-sense RNA genome of roughly 7100 bases in length. The particle is about 30 nm in diameter. Within the family picornaviridae, the Enterovirus genus includes more than 150 RVs divided into three species, RV-A, RV-B, and RV-C. It also includes the three poliovirus serotypes, the coxsackieviruses, and the well-known enterovirus EV71—causing epidemics of hand-foot-and-mouth disease with severe complications [88]—amongst many others. Despite high similarity of the nucleotide sequence of their RNA genomes [89] and three-dimensional structures, RVs recognize three different classes of receptors; twelve RV-A use members of the low-density lipoprotein receptor (LDLR) family for cell entry and constitute the minor receptor group [9093]. Ninety species, comprising representatives of both species A and B, the major receptor group, bind ICAM-1 [94]. And finally, about fifty RV-C might bind the cadherin-related family member 3 (CDHR3), as so far explicitly demonstrated for three representatives [89, 95, 96].
The different receptors are unrelated and exhibit different AA sequences, 3D-structures, and functions in the context of cell metabolism. Whereas LDLRs are ubiquitous and highly conserved throughout animal species, ICAM-1 is not expressed in all cell types and is less conserved; major-group RVs only bind the human and primate version of ICAM-1. Systematic humanization of the mouse protein allowed pinpointing AA residues important for recognition [97]. The natural human ICAM-1 Kilifi variant substantially differs in affinity for RV14 and RV16 [98], allowing for estimation of the impact of a single AA mutation on virus recognition. Whereas numerous 3D-structures of virions belonging to the major and minor receptor groups in complex with their cognate receptors are available, most of what is known on CDHR3 interaction with HRV-Cs is derived from homology modelling [99] and no 3D structure is yet available. Probably, this is due to the low yield of virus in cells transfected to express the receptor; no established cell line that would become infected by RV-Cs has been found so far.
Members of the LDLR family possess various numbers of highly conserved ligand binding modules, seven in LDLR, eight in very-LDLR (VLDLR), and 31 in LDLR-related protein (LRP) [100]. Single modules exhibit very low binding affinity; however, simultaneous binding of several modules results in a considerable increase in avidity [101, 102]. Although not anticipated in initial cryo-EM work [103], a recombinant concatemer of five copies of module 3 of VLDLR was later demonstrated to arrange around a five-fold symmetry axis of the virus in a ring-like structure [40, 104, 105]. Interestingly, binding of a recombinant concatemer (i.e., five identical repeats fused head to tail) of repeat 3 of VLDL neutralized infectivity not only via competition with the natural receptor, but also via inhibiting the structural changes necessary for the release of the RNA [102, 106, 107]. Mutations of single VLDLR modules and their display on phage [108], as well as systematic exchange of human LDLR repeats for their mouse homologues [109] allowed the requirements of particular AA residues for viral recognition to be defined. The different modules in LDLR and LRP exhibit various degrees of affinity; therefore, they most likely contribute to the overall binding avidity to a different extent. The footprint of the receptor modules indicates that the interactions are mostly governed by charge and shape complementarity [90, 105, 110]. Since VLDLR is probably not expressed in the nasal mucosa [111, 112], presumably only LDLR and LRP are used by minor-group RVs for cell entry and infection in vivo [92]. In contrast to ICAM-1 (see below), LDLRs merely function as vehicles for virus entry. However, they might have a role in release of the virus from its receptors in the late endosomal compartment through the pH-dependence of the binding affinity [113].
LDLR molecules wrap around each of the 12 star-like mesas at the five-fold axes by engaging several, presumably up to five, ligand-binding repeats [104, 105]. In LRP, the 31 repeats are arranged in groups with spacers in between. Thus, it is possible that LRP can bind simultaneously to more than one pentamer because of its length. At maximal occupancy the stoichiometry (virus: soluble receptor) is 1:12 for recombinant VLDLR and, presumably, LDLR [113116], but maybe less for LRP.
ICAM-1, the receptor of major-group RVs [117], belongs to the immunoglobulin (Ig) superfamily and has five Ig-domains [118], with only the first one (D1) being engaged in RV binding [119]; the second domain appears to be necessary for correct folding of D1. Cryo-EM reconstructions of complexes between representative RVs and soluble ICAM-1 show the latter sticking out from the canyon, upright with a slight outward inclination [120]. The theoretical stoichiometry at maximal occupation is 1:60 for ICAM-1 [121].
According to the canyon hypothesis, receptor-recognizing AA residues would be inaccessible for antibodies due to the narrowness of the canyon [122, 123]. Nevertheless, at least one monoclonal antibody (MAb) reaching far into this valley has been isolated. The paratopes of IgG molecules and of ICAM-1 are quite different and AA residues recognized by the MAb differ from those recognized by ICAM-1. This allows for escape from antibody neutralization via mutation without substantially changing receptor binding [124]. This also explains why ICAM-1 binds many different RVs, whereas MAbs targeting residues at the canyon floor and cross-reacting with several RV serotypes have not been described. On the other hand, for minor-group viruses, the receptor-binding site is entirely exposed and available for antibody recognition. Again, no cross-reacting antibodies targeting residues within the receptor footprint have been reported; in this case, this might be due to the necessity to bind more than one symmetrically-related site for appreciable avidity, which is impossible for IgGs because of their lower valence, geometry, and steric constraints. It has not been addressed whether IgMs, which would at least allow pentavalent attachment to such sites, might be better cross-reactors, assuming that the geometry and flexibility would allow such interactions.
A particular feature of ICAM-1 is its “catalytic” activity; in vitro, interaction with cognate RVs at a temperature above 30 °C leads in some, but not all, major-group RVs to uncoating at neutral pH, a process remotely similar to catalysis [121, 125127]. Since initial low-affinity binding is followed by an increase in affinity it is likely that the receptor stabilizes an “open” conformation temporarily adopted by the virus via “breathing” [127]. This might allow ICAM-1 to penetrate more deeply into the canyon [128]. Breathing is also responsible for the transient exposure of capsid-internal protein sequences at the N‑terminus of VP1 and the myristoylated VP4 that are not accessible at lower temperature [129, 130]. Thus, ICAM-1 can be seen as a wedge driven into the canyon just when it is expanded. This would shift the equilibrium between the two conformations towards the more “open” one. The “priming” for uncoating strongly depends on the temperature and the concentration of soluble ICAM-1, which agrees with the dynamic structural rearrangements [131, 132]. However, receptor-driven uncoating is not very efficient at neutral pH but is aided by acidification [133]. RV-A89 adapted to grow in cells devoid of ICAM-1 was more readily neutralized by soluble ICAM-1. This might indicate that a decrease in stability is necessary for becoming less dependent, and finally independent—like minor-group RVs—from the destabilizing activity of ICAM-1 and allowing the use of a receptor that “just binds but does not uncoat” [134]. The concentration range of soluble ICAM-1 required for 50 % neutralization of different serotypes at physiologic pH extends over about 1.7 logs [135]; whether competition with the cellular receptor or ICAM-1-catalyzed uncoating is primarily responsible for neutralization cannot be inferred from these simple measurements. It is noteworthy that dimerization of the ICAM-1 molecule increases its neutralization potency [136], which might be a mechanical phenomenon resulting from generation of strain upon bivalent binding with suboptimal geometry. The “hit-and-run” uncoating of poliovirus [137] and EV71 by monoclonal antibodies [138] might be related phenomena.
The ratio between the number of physical rhinovirus particles and the number of infectious particles is in the range of 24–240 under mild purification conditions using metrizamide density gradients, and can be much higher when harsh conditions such as CsCl-gradient centrifugation are used [94]. It is not known whether this ratio is lower for unpurified virus. It is thus necessary to not only follow the physical particles on their way into the cell but also to demonstrate that a given pathway leads to initiation of infection. Therefore, it is essential to assess “productive” infection as well. Preferentially, such assays should target an early stage of the viral replication cycle to avoid secondary effects of inhibitors and/or particular experimental incubation conditions on later stages such as RNA replication, assembly, maturation, and release of progeny. Considering that i) uncoating results in release of the viral RNA into the cytosol, that ii) this RNA is translated, and that iii) the resulting polyprotein is cleaved to release the 2A protease, which, in turn, cleaves  the eukaryotic initiation factor 4G (eIF4G), the latter can be correlated with uncoating. The more RNA arrives in the cytosol the more eIF4G is cleaved. The tiny amounts of the viral proteinase P2A that becomes translated from the incoming viral RNA have thus been used to demonstrate viral uncoating without having to wait until replication has set in [33].

Uptake and uncoating of the minor receptor-group prototype RV-A2

LDLR and LDLR-related protein 1 (LRP1) possess internalization sequence motives in their N‑terminal cytoplasmic tails that typically associate with the AP-2 complex, which, in turn, assembles the clathrin coat at the cytoplasmic side of the plasma membrane giving rise to clathrin-coated pits. Indeed, uptake of the minor receptor-group prototype RV-A2 into HeLa-H1 cells was demonstrated to be impeded by most inhibitors and interventions generally accepted to block CME, as summarized in Tab. 1 and Fig. 3. However, as mentioned above, endocytosis of RV-2A was also seen to be diverted to a non-clathrin-dependent route when the former pathway was curtailed by overexpression of the dominant negative dynamin 1 mutant K44A [139]. So, at least in HeLa-H1 (and in Rhabdomyosarcoma cells), RV-A2 primarily, but not exclusively, enters via clathrin-dependent endocytosis [33, 140].
Tab. 1
Compilation of data referenced in the text. Upper panel, impact of pharmacological inhibitors, overexpression of dominant negative mutant proteins, or other manipulations on viral entry either tested via immunofluorescence microscopy (IFM) or cleavage of eIF4G and/or replication by wildtype RVs and variants binding heparan sulfate, as exemplified by RV-A8v. Note that the inhibitors of proton carriers and weak bases increasing the pH in endosomes—such as niclosamide, monensin, methylamine, etc.—inhibit uncoating of minor-group RVs strongly and of some major-group RVs weakly. Lower panel, colocalization of viruses with various markers observed by fluorescence microscopy. For a list on the mode of action of these inhibitors see [49]
Serotype/receptor group
RV-A2/minor-group
RV-B14 (major-group), entering via ICAM-1
RV-A8v & other major-group entering via HS
Process inhibited/tested via
Entry (obs. via IFM)
eIF4G clv./replication
Entry (obs. via IFM)
eIF4G clv./replication
Entry (obs. via IFM)
eIF4G clv./replication
Inhibitor
Chlorpromazine
Not done
Strong
Not done
Marginal
Not done
No
Filipin or nystatin
Not done
No
Not done
No
Not done
No
Methyl-ß-cyclodextrin
Strong
Weak
Not done
Weak
Not done
Weak
Amiloride or EIPA
No
Strong
Not done
Strong
Not done
Strong
Dynasore
Strong
Strong
No
Weak
Not done
Strong
Cytochalasin D
No
Weak
Not done
Weak
Not done
Strong
Bafilomycin
No
Strong
Not done
Weak
Not done
Not done
Amphi-SH3
Strong
Not done
No
Not done
No
Not done
AP180-C
Strong
Not done
No
Not done
No
Not done
DynK44A
Strong
Not done
No
Not done
Strong
Not done
K-depletion
Strong
Not done
Not done
Not done
Not done
Not done
Rab5 S34N
Strong
Not done
Not done
Not done
Not done
Not done
Colocalization with
Transferrin
Weak
Not done
No
Not done
No
Not done
Flotillin-1
Not done
Not done
No
Not done
No
Not done
CtxB
Not done
Not done
No
Not done
No
Not done
FITC-dextran
Not done
Not done
Strong
Not done
Strong
Not done
GFP-clathrin
Not done
Not done
Not done
Not done
No
Not done
Caveolin-1
No
Not done
Not done
Not done
No
Not done
List of abbreviations: CtxB choleratoxin B, IFM immunofluorescence microscopy, Rab5 S34N Rab5 with S34 mutated to N, cvl cleavage, HS heparan sulphate, obs observed
Once inside endosomal carrier vesicles (ECV) or late endosomes, RV-A2 is believed to dissociate from its receptor, a process partly promoted by the acid-dependent intramolecular competition of the beta-propeller domain of LDLR with the bound virus for the ligand binding domains [114, 115] and partly by the conversion of the native virus into the expanded A‑particle that lacks affinity for the receptor. Both effects occur at about pH 5.7, usually attained in ECVs and late endosomes, depending on the cell type [113, 141, 142], and are inhibited when the pH is increased by incubation with weak bases, proton carriers, and inhibitors of H+-ATPases [133, 143145]. The affinity for the receptor is lost because of the altered conformation of the subviral particle within the receptor footprint [146, 147]. Presumably the particle is handed over to the membrane during its formation, a process assisted by the myristoylated VP4 and N‑terminal sequences of VP1 both exiting from the virion and inserting into the lipid bilayer creating pores lined by amphiphilic segments of these proteins [148, 149]. Flickering pores were detected by electrophysiological measurements in experiments with the related poliovirus [150152] and with RV-A2 [30], and via the release of co-internalized FITC-dextran [153]. The amount of the dextran leaking from virus-containing endosomes into the cytosol was found to be inversely proportional to its size; whereas 10 kD dextran exited, 70 kD dextran did not. The membrane-disrupting adenovirus used as a control released both dextrans equally well. This led the authors to conclude that RV-A2 indeed formed pores in the endosomal membrane and that the RNA genome was most likely passed through such a pore into the cytosol. These results were supported by in vitro data demonstrating that the viral RNA could be transferred from membrane-bound virions into liposomes on acidification, and that the process was accompanied by membrane permeabilization but not disruption; the maximum RNA transfer occurred at a virus concentration that only marginally impacted on membrane integrity, again speaking for RNA transfer through pores of limited size [82]. Finally, in vivo, RV-A2 capsid proteins are readily degraded in lysosomes, indicating that the virus shell does not attain the cytosol. Proteolysis starts quite early; at about 30 min post infection, all of VP1 and a large proportion of VP2 had been cleaved [144]. Degradation of VP1 is inhibited by the microtubule-disrupting drug nocodazole, which blocks transport from ECV to lysosomes but leaves infection unaltered [143]. Nevertheless, it is possible that, similar to calicivirus infection [154], a marginal and thus difficult-to-detect cleavage of the subviral shell already occurs in (late) endosomes and facilitates RNA release in vivo.
Recently, it was demonstrated that RNA exit starts from the 3’-end on heating to 56 °C in vitro [155], as well as on physiologic endosomal acidification in vivo [156]. Interestingly, upon acidification in vitro, exit halted after about 700 bases had egressed. A similar phenomenon was seen in vivo; after conversion of the native virion into the subviral A‑particle, which was observed at about 5 min post infection at 34 °C, it took about 10 more minutes until the 700 3’-terminal bases became accessible to added nuclease; at lower temperature these times were prolonged. However, all the remaining genome sequences were lost from the viral shell at 34 °C within the next 2 min. It was also found that the partial egress in vitro quickly completed when a microsomal fraction prepared from HeLa cells was added. This suggests that cellular facilitators of the process might exist. Nevertheless, it remains enigmatic how an RNA molecule with a highly complex secondary structure with many stems, loops, and hairpins, should unfold to pass through a hole of about 0.1 × 0.2 nm in the absence of energy consumption in this short time.

Uptake and uncoating of the major receptor-group prototype RV-B14

ICAM-1 lacks a clathrin localization signal but its cytoplasmic tail associates with alpha-actinin, thus connecting it to the cytoskeleton [157]. However, this association is not required for viral uptake since the tail and the transmembrane region can be replaced with a glycosylphosphatidyl anchor without impact on viral infection [158]. Therefore, it is not unexpected that chlorpromazine, which is moderately specific for the clathrin-dependent pathway, and some dominant negative mutant proteins involved in clathrin-dependent endocytosis failed to block RV-B14 uptake (Tab. 1 and [33]). A particularly striking observation was the accumulation of virions in long tubules connected to the cell surface in RD and BHK-cells expressing human ICAM-1 [33, 159]. Such aggregates had not been reported for viruses internalized by clathrin-dependent endocytosis, but were observed during entry of SV40 when bound to the lipid receptor GM1 ([79] and see above); nevertheless, in some instances, RV-B14 in vesicles with a coat looking very similar to a clathrin coat were seen in HeLa cells [159]. The significance of this observation for viral uptake is unclear. Entry was strongly reduced by the Na(+)/H(+) ion exchange inhibitor amiloride, and moderately reduced by the actin polymerization inhibitor cytochalasin. Thus, it is highly likely that RV-B14 enters by a non-clathrin- non-caveolin-dependent pathway resembling macropinocytosis. Entry is slower than that of RV-A2 and a large fraction of the virus appears to be retained at the plasma membrane despite of “normal” infection efficiency. As mentioned above, RV-B14 probably arrives in the cytoplasm as a whole (subviral) particle by disruption of the endosome [153, 160]; this contention was also supported by the finding that RV-B14 promotes endosomal release of transfection complexes [161]. In summary, RV-B14 enters via a pathway similar but not identical to typical macropinocytosis (Fig. 3).

Uptake and uncoating of major receptor-group viruses adapted to heparan sulfate proteoglycan

Identical to RV-B14, the serotypes RV-A8, RV-A54, and RV-A89 all depend on ICAM-1 for entry and are thus major-group RV [32]. The latter three were adapted to use heparan sulfate (HS) as a receptor via blind passages alternating between HeLa and (ICAM-1-negative) Hep-2 cells. RV-A89 needed 34 such passages [134, 162], RV-A8 three [32], and RV-A54 was found to already be naturally adapted [34]. In all cases, binding to ICAM-1 was not lost but the viruses had become less resistant to moderately acidic pH. This suggests that in the absence of the “catalytic” activity of the receptor, the low pH alone must suffice for uncoating. HS-binding variants were shown to strictly depend on dynamin but not on clathrin, caveolin, and flotillin, as summarized in Tab. 1 for the HS-adapted variant RV-A8v, whose pathway was more extensively characterized; the same profile is probably seen with the other HS-adapted variants. The tabulated data show that entry via HS is similarly, but not identically sensitive to inhibitors and manipulations as entry via ICAM-1. Accumulation of the virions in tubules connected to the plasma membrane was also observed [32]. Recent work on uptake of wildtype RV-A89 (Conzemius et al. manuscript submitted) indicates that this major-group RV might involve clathrin for entry and uncoat in the endosomal recycling compartment (Fig. 3).

Conclusion

Similar to the natural ligands of LDLR, minor-group RVs follow the clathrin-dependent pathway for entry. Once inside endosomal carrier vesicles or late endosomes, the low-pH environment triggers conversion of the native virion into the A‑particle, from which, in a poorly-understood process, the RNA is extruded through the endosomal membrane into the cytosol. Major-group ICAM-1 binding viruses appear to be taken up via a somewhat untypical form of macropinocytosis. Both pathways converge in early endosomes and the viruses are further shuttled to late endosomes. Aided by the destabilizing activity of ICAM-1 they are converted into A‑particles with concomitant disruption of the endosomal membrane, releasing the subviral particle or the empty particle together with already uncoated RNA into the cytosol. Finally, HS-adapted RVs follow a very similar pathway but behave similar to minor-group RVs in being stringently dependent on low pH for uncoating of the viral genome. Repeated attempts to adapt a major-group virus to use LDLR as a receptor failed. This indicates that recognition of LDLR is not limited to mere attraction of opposite charges. On the other hand, the ubiquitous HS can be exploited for attachment and entry but, at least in vivo, there must be other more stringent requirements for not using it in nature. The future might hold surprises regarding how RV-C viruses are taken up by CDHR3 and routed inside the host cell. Knowledge of how viruses enter the cell and points of possible intervention is of paramount importance for finding new ways of blocking infection with drugs. Cellular targets are increasingly considered for this purpose as this minimizes resistance via viral mutation.

Acknowledgements

Open access funding provided by Medical University of Vienna.Much of the research cited by the author is the product of a long, intense, fruitful, and pleasant collaboration with Renate Fuchs. It has been funded over the years by a number of individual grants from the Austrian Science Fund (FWF) to Renate Fuchs and to the author. I express my gratitude to Heinrich Kowalski and Sofiya Fedosyuk for critically reading the manuscript, and to Alexander Otahal for drawing the figures.

Compliance with ethical guidelines

Conflict of interest

D. Blaas states that there are no conflicts of interest.

Ethical standards

This article does not include studies on humans or animals.
Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Unsere Produktempfehlungen

Abo für kostenpflichtige Inhalte

Literatur
1.
Zurück zum Zitat Goncalvez AP, Engle RE, Claire StM, Purcell RH, Lai CJ. Monoclonal antibody-mediated enhancement of dengue virus infection in vitro and in vivo and strategies for prevention. Proc Natl Acad Sci USA. 2007;104(22):9422–27.PubMedPubMedCentralCrossRef Goncalvez AP, Engle RE, Claire StM, Purcell RH, Lai CJ. Monoclonal antibody-mediated enhancement of dengue virus infection in vitro and in vivo and strategies for prevention. Proc Natl Acad Sci USA. 2007;104(22):9422–27.PubMedPubMedCentralCrossRef
2.
Zurück zum Zitat Jiang J, Wu X, Tang H, Luo G. Apolipoprotein E mediates attachment of clinical hepatitis C virus to hepatocytes by binding to cell surface heparan sulfate proteoglycan receptors. PloS One. 2013;8(7):e67982.PubMedPubMedCentralCrossRef Jiang J, Wu X, Tang H, Luo G. Apolipoprotein E mediates attachment of clinical hepatitis C virus to hepatocytes by binding to cell surface heparan sulfate proteoglycan receptors. PloS One. 2013;8(7):e67982.PubMedPubMedCentralCrossRef
3.
Zurück zum Zitat Meertens L, Carnec X, Lecoin MP, Ramdasi R, Guivel-Benhassine F, Lew E, et al. The TIM and TAM families of phosphatidylserine receptors mediate dengue virus entry. Cell Host Microbe. 2012;12(4):544–57.PubMedPubMedCentralCrossRef Meertens L, Carnec X, Lecoin MP, Ramdasi R, Guivel-Benhassine F, Lew E, et al. The TIM and TAM families of phosphatidylserine receptors mediate dengue virus entry. Cell Host Microbe. 2012;12(4):544–57.PubMedPubMedCentralCrossRef
4.
Zurück zum Zitat Schneider-Schaulies J. Cellular receptors for viruses: links to tropism and pathogenesis. J Gen Virol. 2000;81(6):1413–29.CrossRef Schneider-Schaulies J. Cellular receptors for viruses: links to tropism and pathogenesis. J Gen Virol. 2000;81(6):1413–29.CrossRef
5.
Zurück zum Zitat Vasiljevic S, Beale EV, Bonomelli C, Easthope IS, Pritchard LK, Seabright GE, et al. Redirecting adenoviruses to tumour cells using therapeutic antibodies: Generation of a versatile human bispecific adaptor. Mol Immunol. 2015;68(2A):1413–29. Vasiljevic S, Beale EV, Bonomelli C, Easthope IS, Pritchard LK, Seabright GE, et al. Redirecting adenoviruses to tumour cells using therapeutic antibodies: Generation of a versatile human bispecific adaptor. Mol Immunol. 2015;68(2A):1413–29.
6.
Zurück zum Zitat Izmailyan R, Hsao JC, Chung CS, Chen CH, Hsu PW, Liao CL, et al. Integrin beta1 mediates vaccinia virus entry through activation of PI3K/Akt signaling. J Virol. 2012;86(12):6677–87.PubMedPubMedCentralCrossRef Izmailyan R, Hsao JC, Chung CS, Chen CH, Hsu PW, Liao CL, et al. Integrin beta1 mediates vaccinia virus entry through activation of PI3K/Akt signaling. J Virol. 2012;86(12):6677–87.PubMedPubMedCentralCrossRef
9.
Zurück zum Zitat Tan SC, Scherer J, Vallee RB. Recruitment of dynein to late endosomes and lysosomes through light intermediate chains. Mol Biol Cell. 2011;22(4):467–77.PubMedPubMedCentralCrossRef Tan SC, Scherer J, Vallee RB. Recruitment of dynein to late endosomes and lysosomes through light intermediate chains. Mol Biol Cell. 2011;22(4):467–77.PubMedPubMedCentralCrossRef
10.
Zurück zum Zitat Garten W, Braden C, Arendt A, Peitsch C, Baron J, Lu Y, et al. Influenza virus activating host proteases: Identification, localization and inhibitors as potential therapeutics. Eur J Cell Biol. 2015;94(7–9):375–83.PubMedCrossRef Garten W, Braden C, Arendt A, Peitsch C, Baron J, Lu Y, et al. Influenza virus activating host proteases: Identification, localization and inhibitors as potential therapeutics. Eur J Cell Biol. 2015;94(7–9):375–83.PubMedCrossRef
11.
Zurück zum Zitat Banerjee M, Johnson JE. Activation, exposure and penetration of virally encoded, membrane-active polypeptides during non-enveloped virus entry. Curr Protein Pept Sci. 2008;9(1):16–27.PubMedCrossRef Banerjee M, Johnson JE. Activation, exposure and penetration of virally encoded, membrane-active polypeptides during non-enveloped virus entry. Curr Protein Pept Sci. 2008;9(1):16–27.PubMedCrossRef
12.
Zurück zum Zitat Suikkanen S, Antila M, Jaatinen A, Vihinen-Ranta M, Vuento M. Release of canine parvovirus from endocytic vesicles. Virology. 2003;316(2):267–80.PubMedCrossRef Suikkanen S, Antila M, Jaatinen A, Vihinen-Ranta M, Vuento M. Release of canine parvovirus from endocytic vesicles. Virology. 2003;316(2):267–80.PubMedCrossRef
13.
Zurück zum Zitat Luisoni S, Suomalainen M, Boucke K, Tanner LB, Wenk MR, Guan XL, et al. Co-option of Membrane Wounding Enables Virus Penetration into Cells. Cell Host Microbe. 2015;18(1):75–85.PubMedCrossRef Luisoni S, Suomalainen M, Boucke K, Tanner LB, Wenk MR, Guan XL, et al. Co-option of Membrane Wounding Enables Virus Penetration into Cells. Cell Host Microbe. 2015;18(1):75–85.PubMedCrossRef
14.
Zurück zum Zitat Boulant S, Stanifer M, Lozach PY. Dynamics of virus-receptor interactions in virus binding, signaling, and endocytosis. Viruses. 2015;7(6):2794–815.PubMedPubMedCentralCrossRef Boulant S, Stanifer M, Lozach PY. Dynamics of virus-receptor interactions in virus binding, signaling, and endocytosis. Viruses. 2015;7(6):2794–815.PubMedPubMedCentralCrossRef
16.
Zurück zum Zitat Sun Y, Tien P. From endocytosis to membrane fusion: emerging roles of dynamin in virus entry. Crit Rev Microbiol. 2013;39(2):166–79.PubMedCrossRef Sun Y, Tien P. From endocytosis to membrane fusion: emerging roles of dynamin in virus entry. Crit Rev Microbiol. 2013;39(2):166–79.PubMedCrossRef
17.
Zurück zum Zitat Schelhaas M. Come in and take your coat off – how host cells provide endocytosis for virus entry. Cell Microbiol. 2010;12(10):1378–88.PubMedCrossRef Schelhaas M. Come in and take your coat off – how host cells provide endocytosis for virus entry. Cell Microbiol. 2010;12(10):1378–88.PubMedCrossRef
18.
Zurück zum Zitat Mercer J, Schelhaas M, Helenius A. Virus entry by endocytosis. Annu Rev Biochem. 2010;79:803–33.PubMedCrossRef Mercer J, Schelhaas M, Helenius A. Virus entry by endocytosis. Annu Rev Biochem. 2010;79:803–33.PubMedCrossRef
19.
Zurück zum Zitat Yamauchi Y, Helenius A. Virus entry at a glance. J Cell Sci. 2013;126(6):1289–95.CrossRef Yamauchi Y, Helenius A. Virus entry at a glance. J Cell Sci. 2013;126(6):1289–95.CrossRef
21.
Zurück zum Zitat Suomalainen M, Greber UF. Uncoating of non-enveloped viruses. Curr Opin Virol. 2013;3(1):27–33.PubMedCrossRef Suomalainen M, Greber UF. Uncoating of non-enveloped viruses. Curr Opin Virol. 2013;3(1):27–33.PubMedCrossRef
22.
Zurück zum Zitat Endocytosis. In: Schmid SL, Sorkin A, Zerial M, editors. Cold Spring Harbor Perspectives in Biology: Cold Spring Harbor Laboratory Press; 2015. Endocytosis. In: Schmid SL, Sorkin A, Zerial M, editors. Cold Spring Harbor Perspectives in Biology: Cold Spring Harbor Laboratory Press; 2015.
23.
Zurück zum Zitat Yamauchi Y, Greber UF. Principles of Virus Uncoating: Cues and the Snooker Ball. Traffic. 2016; doi:10.1111/tra.12387.PubMed Yamauchi Y, Greber UF. Principles of Virus Uncoating: Cues and the Snooker Ball. Traffic. 2016; doi:10.1111/tra.12387.PubMed
24.
Zurück zum Zitat Schmid S, Fuchs R, Kielian M, Helenius A, Mellman I. Acidification of endosome subpopulations in wild-type Chinese hamster ovary cells and temperature-sensitive acidification-defective mutants. J Cell Biol. 1989;108(4):1291–300.PubMedCrossRef Schmid S, Fuchs R, Kielian M, Helenius A, Mellman I. Acidification of endosome subpopulations in wild-type Chinese hamster ovary cells and temperature-sensitive acidification-defective mutants. J Cell Biol. 1989;108(4):1291–300.PubMedCrossRef
25.
Zurück zum Zitat Fuchs R, Schmid S, Mellman I. A possible role for Na+, K+-ATPase in regulating ATP-dependent endosome acidification. Proc Natl Acad Sci USA. 1989;86(2):539–43.PubMedPubMedCentralCrossRef Fuchs R, Schmid S, Mellman I. A possible role for Na+, K+-ATPase in regulating ATP-dependent endosome acidification. Proc Natl Acad Sci USA. 1989;86(2):539–43.PubMedPubMedCentralCrossRef
26.
Zurück zum Zitat Mellman I, Fuchs R, Helenius A. Acidification of the endocytic and exocytic pathways. Annu Rev Biochem. 1986;55:663–700.PubMedCrossRef Mellman I, Fuchs R, Helenius A. Acidification of the endocytic and exocytic pathways. Annu Rev Biochem. 1986;55:663–700.PubMedCrossRef
28.
29.
Zurück zum Zitat Strauss M, Filman DJ, Belnap DM, Cheng N, Noel RT, Hogle JM. Nectin-like interactions between poliovirus and its receptor trigger conformational changes associated with cell entry. J Virol. 2015;89(8):4143–57.PubMedPubMedCentralCrossRef Strauss M, Filman DJ, Belnap DM, Cheng N, Noel RT, Hogle JM. Nectin-like interactions between poliovirus and its receptor trigger conformational changes associated with cell entry. J Virol. 2015;89(8):4143–57.PubMedPubMedCentralCrossRef
31.
32.
Zurück zum Zitat Khan AG, Pickl-Herk A, Gajdzik L, Marlovits TC, Fuchs R, Blaas D. Entry of a heparan sulphate-binding HRV8 variant strictly depends on dynamin but not on clathrin, caveolin, and flotillin. Virology. 2011;412(1):55–67.PubMedCrossRef Khan AG, Pickl-Herk A, Gajdzik L, Marlovits TC, Fuchs R, Blaas D. Entry of a heparan sulphate-binding HRV8 variant strictly depends on dynamin but not on clathrin, caveolin, and flotillin. Virology. 2011;412(1):55–67.PubMedCrossRef
33.
Zurück zum Zitat Khan AG, Pickl-Herk A, Gajdzik L, Marlovits TC, Fuchs R, Blaas D. Human rhinovirus 14 enters rhabdomyosarcoma cells expressing icam-1 by a clathrin-, caveolin-, and flotillin-independent pathway. J Virol. 2010;84(8):3984–92.PubMedPubMedCentralCrossRef Khan AG, Pickl-Herk A, Gajdzik L, Marlovits TC, Fuchs R, Blaas D. Human rhinovirus 14 enters rhabdomyosarcoma cells expressing icam-1 by a clathrin-, caveolin-, and flotillin-independent pathway. J Virol. 2010;84(8):3984–92.PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Khan AG, Pichler J, Rosemann A, Blaas D. Human rhinovirus type 54 infection via heparan sulfate is less efficient and strictly dependent on low endosomal pH. J Virol. 2007;81(9):4625–32.PubMedPubMedCentralCrossRef Khan AG, Pichler J, Rosemann A, Blaas D. Human rhinovirus type 54 infection via heparan sulfate is less efficient and strictly dependent on low endosomal pH. J Virol. 2007;81(9):4625–32.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Xing L, Tjarnlund K, Lindqvist B, Kaplan GG, Feigelstock D, Cheng RH, et al. Distinct cellular receptor interactions in poliovirus and rhinoviruses. EMBO J. 2000;19(6):1207–16.PubMedPubMedCentralCrossRef Xing L, Tjarnlund K, Lindqvist B, Kaplan GG, Feigelstock D, Cheng RH, et al. Distinct cellular receptor interactions in poliovirus and rhinoviruses. EMBO J. 2000;19(6):1207–16.PubMedPubMedCentralCrossRef
36.
Zurück zum Zitat Konecsni T, Kremser L, Snyers L, Rankl C, Kilar F, Kenndler E, et al. Twelve receptor molecules attach per viral particle of human rhinovirus serotype 2 via multiple modules. FEBS Lett. 2004;568(1–3):99–104.PubMedCrossRef Konecsni T, Kremser L, Snyers L, Rankl C, Kilar F, Kenndler E, et al. Twelve receptor molecules attach per viral particle of human rhinovirus serotype 2 via multiple modules. FEBS Lett. 2004;568(1–3):99–104.PubMedCrossRef
37.
Zurück zum Zitat Sieben C, Kappel C, Zhu R, Wozniak A, Rankl C, Hinterdorfer P, et al. Influenza virus binds its host cell using multiple dynamic interactions. Proc Natl Acad Sci USA. 2012;109(34):13626–31.PubMedPubMedCentralCrossRef Sieben C, Kappel C, Zhu R, Wozniak A, Rankl C, Hinterdorfer P, et al. Influenza virus binds its host cell using multiple dynamic interactions. Proc Natl Acad Sci USA. 2012;109(34):13626–31.PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat Rankl C, Wildling L, Neundlinger I, Kienberger F, Gruber H, Blaas D, et al. Determination of the kinetic on- and off-rate of single virus-cell interactions. Methods Mol Biol. 2011;736:197–210.PubMedCrossRef Rankl C, Wildling L, Neundlinger I, Kienberger F, Gruber H, Blaas D, et al. Determination of the kinetic on- and off-rate of single virus-cell interactions. Methods Mol Biol. 2011;736:197–210.PubMedCrossRef
39.
Zurück zum Zitat Rankl C, Kienberger F, Wildling L, Wruss J, Gruber HJ, Blaas D, et al. Multiple receptors involved in human rhinovirus attachment to live cells. Proc Natl Acad Sci USA. 2008;105(46):17778–83.PubMedPubMedCentralCrossRef Rankl C, Kienberger F, Wildling L, Wruss J, Gruber HJ, Blaas D, et al. Multiple receptors involved in human rhinovirus attachment to live cells. Proc Natl Acad Sci USA. 2008;105(46):17778–83.PubMedPubMedCentralCrossRef
40.
Zurück zum Zitat Wruss J, Pollheimer, Meindl I, Reichel A, Schulze K, Schofberger W, et al. Conformation of receptor adopted upon interaction with virus revealed by site-specific fluorescence quenchers and FRET analysis. J Am Chem Soc. 2009;131(15):5478–82.PubMedCrossRef Wruss J, Pollheimer, Meindl I, Reichel A, Schulze K, Schofberger W, et al. Conformation of receptor adopted upon interaction with virus revealed by site-specific fluorescence quenchers and FRET analysis. J Am Chem Soc. 2009;131(15):5478–82.PubMedCrossRef
41.
Zurück zum Zitat Mateo M, Generous A, Sinn PL, Cattaneo R. Connections matter–how viruses use cell-cell adhesion components. J Cell Sci. 2015;128(3):431–9.PubMedPubMedCentralCrossRef Mateo M, Generous A, Sinn PL, Cattaneo R. Connections matter–how viruses use cell-cell adhesion components. J Cell Sci. 2015;128(3):431–9.PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Brandenburg B, Lee LY, Lakadamyali M, Rust MJ, Zhuang X, Hogle JM. Imaging poliovirus entry in live cells. PLoS Biol. 2007;5(7):e183.CrossRef Brandenburg B, Lee LY, Lakadamyali M, Rust MJ, Zhuang X, Hogle JM. Imaging poliovirus entry in live cells. PLoS Biol. 2007;5(7):e183.CrossRef
44.
Zurück zum Zitat Kusumi A, Nakada C, Ritchie K, Murase K, Suzuki K, Murakoshi H, et al. Paradigm shift of the plasma membrane concept from the two-dimensional continuum fluid to the partitioned fluid: high-speed single-molecule tracking of membrane molecules. Annu Rev Biophys Biomol Struct. 2005;34:351–78.PubMedCrossRef Kusumi A, Nakada C, Ritchie K, Murase K, Suzuki K, Murakoshi H, et al. Paradigm shift of the plasma membrane concept from the two-dimensional continuum fluid to the partitioned fluid: high-speed single-molecule tracking of membrane molecules. Annu Rev Biophys Biomol Struct. 2005;34:351–78.PubMedCrossRef
45.
Zurück zum Zitat Lukic Z, Dharan A, Fricke T, Diaz-Griffero F, Campbell EM. HIV-1 uncoating is facilitated by dynein and kinesin 1. J Virol. 2014;88(23):13613–25.PubMedPubMedCentralCrossRef Lukic Z, Dharan A, Fricke T, Diaz-Griffero F, Campbell EM. HIV-1 uncoating is facilitated by dynein and kinesin 1. J Virol. 2014;88(23):13613–25.PubMedPubMedCentralCrossRef
46.
Zurück zum Zitat Strunze S, Engelke MF, Wang IH, Puntener D, Boucke K, Schleich S, et al. Kinesin-1-mediated capsid disassembly and disruption of the nuclear pore complex promote virus infection. Cell Host Microbe. 2011;10(3):210–23.PubMedCrossRef Strunze S, Engelke MF, Wang IH, Puntener D, Boucke K, Schleich S, et al. Kinesin-1-mediated capsid disassembly and disruption of the nuclear pore complex promote virus infection. Cell Host Microbe. 2011;10(3):210–23.PubMedCrossRef
47.
Zurück zum Zitat Bayer N, Schober D, Huttinger M, Blaas D, Fuchs R. Inhibition of clathrin-dependent endocytosis has multiple effects on human rhinovirus serotype 2 cell entry. J Biol Chem. 2001;276(6):3952–62.PubMedCrossRef Bayer N, Schober D, Huttinger M, Blaas D, Fuchs R. Inhibition of clathrin-dependent endocytosis has multiple effects on human rhinovirus serotype 2 cell entry. J Biol Chem. 2001;276(6):3952–62.PubMedCrossRef
48.
Zurück zum Zitat Garcin PO, Pante N. The minute virus of mice exploits different endocytic pathways for cellular uptake. Virology. 2015;482:157–66.PubMedCrossRef Garcin PO, Pante N. The minute virus of mice exploits different endocytic pathways for cellular uptake. Virology. 2015;482:157–66.PubMedCrossRef
49.
50.
Zurück zum Zitat Zhou Y, Zhu S, Cai C, Yuan P, Li C, Huang Y, et al. High-throughput screening of a CRISPR/Cas9 library for functional genomics in human cells. Nature. 2014;509(7501):487–91.PubMedCrossRef Zhou Y, Zhu S, Cai C, Yuan P, Li C, Huang Y, et al. High-throughput screening of a CRISPR/Cas9 library for functional genomics in human cells. Nature. 2014;509(7501):487–91.PubMedCrossRef
51.
Zurück zum Zitat Perreira JM, Aker AM, Savidis G, Chin CR, McDougall WM, Portmann JM, et al. RNASEK Is a V‑ATPase-Associated Factor Required for Endocytosis and the Replication of Rhinovirus, Influenza A Virus, and Dengue Virus. Cell Rep. 2015;12(5):850–63.PubMedCrossRef Perreira JM, Aker AM, Savidis G, Chin CR, McDougall WM, Portmann JM, et al. RNASEK Is a V‑ATPase-Associated Factor Required for Endocytosis and the Replication of Rhinovirus, Influenza A Virus, and Dengue Virus. Cell Rep. 2015;12(5):850–63.PubMedCrossRef
52.
Zurück zum Zitat Hackett BA, Yasunaga A, Panda D, Tartell MA, Hopkins KC, Hensley SE, et al. RNASEK is required for internalization of diverse acid-dependent viruses. Proc Natl Acad Sci USA. 2015;112(25):7797–802.PubMedPubMedCentralCrossRef Hackett BA, Yasunaga A, Panda D, Tartell MA, Hopkins KC, Hensley SE, et al. RNASEK is required for internalization of diverse acid-dependent viruses. Proc Natl Acad Sci USA. 2015;112(25):7797–802.PubMedPubMedCentralCrossRef
53.
Zurück zum Zitat Ramo P, Drewek A, Arrieumerlou C, Beerenwinkel N, Ben-Tekaya H, Cardel B, et al. Simultaneous analysis of large-scale RNAi screens for pathogen entry. BMC Genomics. 2014;15:1162.PubMedPubMedCentralCrossRef Ramo P, Drewek A, Arrieumerlou C, Beerenwinkel N, Ben-Tekaya H, Cardel B, et al. Simultaneous analysis of large-scale RNAi screens for pathogen entry. BMC Genomics. 2014;15:1162.PubMedPubMedCentralCrossRef
54.
Zurück zum Zitat Essletzbichler P, Konopka T, Santoro F, Chen D, Gapp BV, Kralovics R, et al. Megabase-scale deletion using CRISPR/Cas9 to generate a fully haploid human cell line. Genome Res. 2014;24(12):2059–65.PubMedPubMedCentralCrossRef Essletzbichler P, Konopka T, Santoro F, Chen D, Gapp BV, Kralovics R, et al. Megabase-scale deletion using CRISPR/Cas9 to generate a fully haploid human cell line. Genome Res. 2014;24(12):2059–65.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Burckstummer T, Banning C, Hainzl P, Schobesberger R, Kerzendorfer C, Pauler FM, et al. A reversible gene trap collection empowers haploid genetics in human cells. Nat Methods. 2013;10(10):965–71.PubMedCrossRef Burckstummer T, Banning C, Hainzl P, Schobesberger R, Kerzendorfer C, Pauler FM, et al. A reversible gene trap collection empowers haploid genetics in human cells. Nat Methods. 2013;10(10):965–71.PubMedCrossRef
57.
Zurück zum Zitat Bonifacino JS, Traub LM. Signals for sorting of transmembrane proteins to endosomes and lysosomes. Annu Rev Biochem. 2003;72:395–447.PubMedCrossRef Bonifacino JS, Traub LM. Signals for sorting of transmembrane proteins to endosomes and lysosomes. Annu Rev Biochem. 2003;72:395–447.PubMedCrossRef
59.
Zurück zum Zitat Xing Y, Bocking T, Wolf M, Grigorieff N, Kirchhausen T, Harrison SC. Structure of clathrin coat with bound Hsc70 and auxilin: mechanism of Hsc70-facilitated disassembly. EMBO J. 2010;29(3):655–65.PubMedPubMedCentralCrossRef Xing Y, Bocking T, Wolf M, Grigorieff N, Kirchhausen T, Harrison SC. Structure of clathrin coat with bound Hsc70 and auxilin: mechanism of Hsc70-facilitated disassembly. EMBO J. 2010;29(3):655–65.PubMedPubMedCentralCrossRef
61.
Zurück zum Zitat Benmerah A, Poupon V, Cerf-Bensussan N, Dautry-Varsat A. Mapping of Eps15 domains involved in its targeting to clathrin-coated pits. J Biol Chem. 2000;275(5):3288–95.PubMedCrossRef Benmerah A, Poupon V, Cerf-Bensussan N, Dautry-Varsat A. Mapping of Eps15 domains involved in its targeting to clathrin-coated pits. J Biol Chem. 2000;275(5):3288–95.PubMedCrossRef
62.
Zurück zum Zitat Wang LH, Rothberg KG, Anderson RG. Mis-assembly of clathrin lattices on endosomes reveals a regulatory switch for coated pit formation. J Cell Biol. 1993;123(5):1107–17.PubMedCrossRef Wang LH, Rothberg KG, Anderson RG. Mis-assembly of clathrin lattices on endosomes reveals a regulatory switch for coated pit formation. J Cell Biol. 1993;123(5):1107–17.PubMedCrossRef
63.
Zurück zum Zitat Robertson MJ, Deane FM, Robinson PJ, McCluskey A. Synthesis of Dynole 34-2, Dynole 2‑24 and Dyngo 4a for investigating dynamin GTPase. Nat Protoc. 2014;9(4):851–70.PubMedCrossRef Robertson MJ, Deane FM, Robinson PJ, McCluskey A. Synthesis of Dynole 34-2, Dynole 2‑24 and Dyngo 4a for investigating dynamin GTPase. Nat Protoc. 2014;9(4):851–70.PubMedCrossRef
64.
Zurück zum Zitat Hill TA, Mariana A, Gordon CP, Odell LR, Robertson MJ, McGeachie AB, et al. Iminochromene inhibitors of dynamins I and II GTPase activity and endocytosis. J Med Chem. 2010;53(10):4094–102.PubMedCrossRef Hill TA, Mariana A, Gordon CP, Odell LR, Robertson MJ, McGeachie AB, et al. Iminochromene inhibitors of dynamins I and II GTPase activity and endocytosis. J Med Chem. 2010;53(10):4094–102.PubMedCrossRef
65.
Zurück zum Zitat Dutta D, Williamson CD, Cole NB, Donaldson JG. Pitstop 2 is a potent inhibitor of clathrin-independent endocytosis. PloS One. 2012;7(9):e45799.CrossRef Dutta D, Williamson CD, Cole NB, Donaldson JG. Pitstop 2 is a potent inhibitor of clathrin-independent endocytosis. PloS One. 2012;7(9):e45799.CrossRef
66.
Zurück zum Zitat Vercauteren D, Vandenbroucke RE, Jones AT, Rejman J, Demeester J, Smedt SC De, et al. The use of inhibitors to study endocytic pathways of gene carriers: optimization and pitfalls. Mol Ther. 2010;18(3):561–9.PubMedPubMedCentralCrossRef Vercauteren D, Vandenbroucke RE, Jones AT, Rejman J, Demeester J, Smedt SC De, et al. The use of inhibitors to study endocytic pathways of gene carriers: optimization and pitfalls. Mol Ther. 2010;18(3):561–9.PubMedPubMedCentralCrossRef
67.
Zurück zum Zitat Pietiainen VM, Marjomaki V, Heino J, Hyypia T. Viral entry, lipid rafts and caveosomes. Ann Med. 2005;37(6):394–403.PubMedCrossRef Pietiainen VM, Marjomaki V, Heino J, Hyypia T. Viral entry, lipid rafts and caveosomes. Ann Med. 2005;37(6):394–403.PubMedCrossRef
68.
Zurück zum Zitat Krieger SE, Kim C, Zhang L, Marjomaki V, Bergelson JM. Echovirus 1 entry into polarized Caco-2 cells depends on dynamin, cholesterol, and cellular factors associated with macropinocytosis. J Virol. 2013;87(16):8884–95.PubMedPubMedCentralCrossRef Krieger SE, Kim C, Zhang L, Marjomaki V, Bergelson JM. Echovirus 1 entry into polarized Caco-2 cells depends on dynamin, cholesterol, and cellular factors associated with macropinocytosis. J Virol. 2013;87(16):8884–95.PubMedPubMedCentralCrossRef
69.
Zurück zum Zitat Pietiainen V, Marjomaki V, Upla P, Pelkmans L, Helenius A, Hyypia T. Echovirus 1 endocytosis into caveosomes requires lipid rafts, dynamin II, and signaling events. Mol Biol Cell. 2004;15(11):4911–25.PubMedPubMedCentralCrossRef Pietiainen V, Marjomaki V, Upla P, Pelkmans L, Helenius A, Hyypia T. Echovirus 1 endocytosis into caveosomes requires lipid rafts, dynamin II, and signaling events. Mol Biol Cell. 2004;15(11):4911–25.PubMedPubMedCentralCrossRef
70.
Zurück zum Zitat Kovtun O, Tillu VA, Jung W, Leneva N, Ariotti N, Chaudhary N, et al. Structural insights into the organization of the cavin membrane coat complex. Dev Cell. 2014;31(4):405–19.PubMedCrossRef Kovtun O, Tillu VA, Jung W, Leneva N, Ariotti N, Chaudhary N, et al. Structural insights into the organization of the cavin membrane coat complex. Dev Cell. 2014;31(4):405–19.PubMedCrossRef
71.
Zurück zum Zitat Shvets E, Bitsikas V, Howard G, Hansen CG, Nichols BJ. Dynamic caveolae exclude bulk membrane proteins and are required for sorting of excess glycosphingolipids. Nat Commun. 2015;6:6867.PubMedPubMedCentralCrossRef Shvets E, Bitsikas V, Howard G, Hansen CG, Nichols BJ. Dynamic caveolae exclude bulk membrane proteins and are required for sorting of excess glycosphingolipids. Nat Commun. 2015;6:6867.PubMedPubMedCentralCrossRef
72.
Zurück zum Zitat Damm EM, Pelkmans L, Kartenbeck J, Mezzacasa A, Kurzchalia T, Helenius A. Clathrin- and caveolin-1-independent endocytosis: entry of simian virus 40 into cells devoid of caveolae. J Cell Biol. 2005;168(3):477–88.PubMedPubMedCentralCrossRef Damm EM, Pelkmans L, Kartenbeck J, Mezzacasa A, Kurzchalia T, Helenius A. Clathrin- and caveolin-1-independent endocytosis: entry of simian virus 40 into cells devoid of caveolae. J Cell Biol. 2005;168(3):477–88.PubMedPubMedCentralCrossRef
73.
74.
Zurück zum Zitat Mooren OL, Galletta BJ, Cooper JA. Roles for actin assembly in endocytosis. Annu Rev Biochem. 2012;81:661–86.PubMedCrossRef Mooren OL, Galletta BJ, Cooper JA. Roles for actin assembly in endocytosis. Annu Rev Biochem. 2012;81:661–86.PubMedCrossRef
75.
Zurück zum Zitat Mercer J, Helenius A. Gulping rather than sipping: macropinocytosis as a way of virus entry. Curr Opin Microbiol. 2012;15(4):490–9.PubMedCrossRef Mercer J, Helenius A. Gulping rather than sipping: macropinocytosis as a way of virus entry. Curr Opin Microbiol. 2012;15(4):490–9.PubMedCrossRef
76.
77.
78.
Zurück zum Zitat Koivusalo M, Welch C, Hayashi H, Scott CC, Kim M, Alexander T, et al. Amiloride inhibits macropinocytosis by lowering submembranous pH and preventing Rac1 and Cdc42 signaling. J Cell Biol. 2010;188(4):547–63.PubMedPubMedCentralCrossRef Koivusalo M, Welch C, Hayashi H, Scott CC, Kim M, Alexander T, et al. Amiloride inhibits macropinocytosis by lowering submembranous pH and preventing Rac1 and Cdc42 signaling. J Cell Biol. 2010;188(4):547–63.PubMedPubMedCentralCrossRef
79.
Zurück zum Zitat Ewers H, Romer W, Smith AE, Bacia K, Dmitrieff S, Chai W, et al. GM1 structure determines SV40-induced membrane invagination and infection. Nat Cell Biol. 2010;12(1):11–8.PubMedCrossRef Ewers H, Romer W, Smith AE, Bacia K, Dmitrieff S, Chai W, et al. GM1 structure determines SV40-induced membrane invagination and infection. Nat Cell Biol. 2010;12(1):11–8.PubMedCrossRef
80.
Zurück zum Zitat Rydell GE, Svensson L, Larson G, Johannes L, Human RWGII. 4 norovirus VLP induces membrane invaginations on giant unilamellar vesicles containing secretor gene dependent alpha1,2-fucosylated glycosphingolipids. Biochim Biophys Acta. 2013;1828(8):1840–5.PubMedCrossRef Rydell GE, Svensson L, Larson G, Johannes L, Human RWGII. 4 norovirus VLP induces membrane invaginations on giant unilamellar vesicles containing secretor gene dependent alpha1,2-fucosylated glycosphingolipids. Biochim Biophys Acta. 2013;1828(8):1840–5.PubMedCrossRef
82.
Zurück zum Zitat Bilek G, Matscheko NM, Pickl-Herk A, Weiss VU, Subirats X, Kenndler E, et al. Liposomal nanocontainers as models for viral infection: monitoring viral genomic RNA transfer through lipid membranes. J Virol. 2011;85(16):8368–75.PubMedPubMedCentralCrossRef Bilek G, Matscheko NM, Pickl-Herk A, Weiss VU, Subirats X, Kenndler E, et al. Liposomal nanocontainers as models for viral infection: monitoring viral genomic RNA transfer through lipid membranes. J Virol. 2011;85(16):8368–75.PubMedPubMedCentralCrossRef
83.
Zurück zum Zitat Bramley TJ, Lerner D, Sames M. Productivity losses related to the common cold. J Occup Environ Med. 2002;44(9):822–9.PubMedCrossRef Bramley TJ, Lerner D, Sames M. Productivity losses related to the common cold. J Occup Environ Med. 2002;44(9):822–9.PubMedCrossRef
84.
Zurück zum Zitat Katpally U, Fu TM, Freed DC, Casimiro DR, Smith TJ. Antibodies to the buried N terminus of rhinovirus VP4 exhibit cross-serotypic neutralization. J Virol. 2009;83(14):7040–8.PubMedPubMedCentralCrossRef Katpally U, Fu TM, Freed DC, Casimiro DR, Smith TJ. Antibodies to the buried N terminus of rhinovirus VP4 exhibit cross-serotypic neutralization. J Virol. 2009;83(14):7040–8.PubMedPubMedCentralCrossRef
85.
Zurück zum Zitat McLean GR, Walton RP, Shetty S, Peel TJ, Paktiawal N, Kebadze T, et al. Rhinovirus infections and immunisation induce cross-serotype reactive antibodies to VP1. Antiviral Res. 2012;95(3):193–201.PubMedCrossRef McLean GR, Walton RP, Shetty S, Peel TJ, Paktiawal N, Kebadze T, et al. Rhinovirus infections and immunisation induce cross-serotype reactive antibodies to VP1. Antiviral Res. 2012;95(3):193–201.PubMedCrossRef
86.
Zurück zum Zitat Glanville N, McLean GR, Guy B, Lecouturier V, Berry C, Girerd Y, et al. Cross-serotype immunity induced by immunization with a conserved rhinovirus capsid protein. Plos Pathog. 2013;9(9):e1003669.PubMedPubMedCentralCrossRef Glanville N, McLean GR, Guy B, Lecouturier V, Berry C, Girerd Y, et al. Cross-serotype immunity induced by immunization with a conserved rhinovirus capsid protein. Plos Pathog. 2013;9(9):e1003669.PubMedPubMedCentralCrossRef
87.
Zurück zum Zitat Edlmayr J, Niespodziana K, Popow-Kraupp T, Krzyzanek V, Focke-Tejkl M, Blaas D, et al. Antibodies induced with recombinant VP1 from human rhinovirus exhibit cross-neutralisation. Eur Respir J. 2011;37(1):44–52.PubMedCrossRef Edlmayr J, Niespodziana K, Popow-Kraupp T, Krzyzanek V, Focke-Tejkl M, Blaas D, et al. Antibodies induced with recombinant VP1 from human rhinovirus exhibit cross-neutralisation. Eur Respir J. 2011;37(1):44–52.PubMedCrossRef
88.
Zurück zum Zitat Yip CC, Lau SK, Woo PC, Yuen KY. Human enterovirus 71 epidemics: what’s next? Emerg Health Threats J. 2013;6:19780.PubMed Yip CC, Lau SK, Woo PC, Yuen KY. Human enterovirus 71 epidemics: what’s next? Emerg Health Threats J. 2013;6:19780.PubMed
89.
Zurück zum Zitat Palmenberg AC, Spiro D, Kuzmickas R, Wang S, Djikeng A, Rathe JA, et al. Sequencing and analyses of all known human rhinovirus genomes reveal structure and evolution. Science. 2009;324(5923):55–9.PubMedPubMedCentralCrossRef Palmenberg AC, Spiro D, Kuzmickas R, Wang S, Djikeng A, Rathe JA, et al. Sequencing and analyses of all known human rhinovirus genomes reveal structure and evolution. Science. 2009;324(5923):55–9.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Vlasak M, Roivainen M, Reithmayer M, Goesler I, Laine P, Snyers L, et al. The minor receptor group of human rhinovirus (HRV) includes HRV23 and HRV25, but the presence of a lysine in the VP1 HI loop is not sufficient for receptor binding. J Virol. 2005;79(12):7389–95.PubMedPubMedCentralCrossRef Vlasak M, Roivainen M, Reithmayer M, Goesler I, Laine P, Snyers L, et al. The minor receptor group of human rhinovirus (HRV) includes HRV23 and HRV25, but the presence of a lysine in the VP1 HI loop is not sufficient for receptor binding. J Virol. 2005;79(12):7389–95.PubMedPubMedCentralCrossRef
91.
Zurück zum Zitat Vlasak M, Blomqvist S, Hovi T, Hewat E, Blaas D. Sequence and structure of human rhinoviruses reveal the basis of receptor discrimination. J Virol. 2003;77(12):6923–30.PubMedPubMedCentralCrossRef Vlasak M, Blomqvist S, Hovi T, Hewat E, Blaas D. Sequence and structure of human rhinoviruses reveal the basis of receptor discrimination. J Virol. 2003;77(12):6923–30.PubMedPubMedCentralCrossRef
92.
Zurück zum Zitat Hofer F, Gruenberger M, Kowalski H, Machat H, Huettinger M, Kuechler E, et al. Members of the low density lipoprotein receptor family mediate cell entry of a minor-group common cold virus. Proc Natl Acad Sci USA. 1994;91(5):1839–42.PubMedPubMedCentralCrossRef Hofer F, Gruenberger M, Kowalski H, Machat H, Huettinger M, Kuechler E, et al. Members of the low density lipoprotein receptor family mediate cell entry of a minor-group common cold virus. Proc Natl Acad Sci USA. 1994;91(5):1839–42.PubMedPubMedCentralCrossRef
93.
Zurück zum Zitat Duechler M, Ketter S, Skern T, Kuechler E, Blaas D. Rhinoviral receptor discrimination: mutational changes in the canyon regions of human rhinovirus types 2 and 14 indicate a different site of interaction. J Gen Virol. 1993;74(10):2287–91.PubMedCrossRef Duechler M, Ketter S, Skern T, Kuechler E, Blaas D. Rhinoviral receptor discrimination: mutational changes in the canyon regions of human rhinovirus types 2 and 14 indicate a different site of interaction. J Gen Virol. 1993;74(10):2287–91.PubMedCrossRef
94.
Zurück zum Zitat Abraham G, Colonno RJ. Many rhinovirus serotypes share the same cellular receptor. J Virol. 1984;51(2):340–5.PubMedPubMedCentral Abraham G, Colonno RJ. Many rhinovirus serotypes share the same cellular receptor. J Virol. 1984;51(2):340–5.PubMedPubMedCentral
96.
Zurück zum Zitat Bochkov YA, Watters K, Ashraf S, Griggs TF, Devries MK, Jackson DJ, et al. Cadherin-related family member 3, a childhood asthma susceptibility gene product, mediates rhinovirus C binding and replication. Proc Natl Acad Sci USA. 2015;112(17):5485–90.PubMedPubMedCentralCrossRef Bochkov YA, Watters K, Ashraf S, Griggs TF, Devries MK, Jackson DJ, et al. Cadherin-related family member 3, a childhood asthma susceptibility gene product, mediates rhinovirus C binding and replication. Proc Natl Acad Sci USA. 2015;112(17):5485–90.PubMedPubMedCentralCrossRef
97.
Zurück zum Zitat Register RB, Uncapher CR, Naylor AM, Lineberger DW, Colonno RJ. Human-murine chimeras of ICAM-1 identify amino acid residues critical for rhinovirus and antibody binding. J Virol. 1991;65(12):6589–96.PubMedPubMedCentral Register RB, Uncapher CR, Naylor AM, Lineberger DW, Colonno RJ. Human-murine chimeras of ICAM-1 identify amino acid residues critical for rhinovirus and antibody binding. J Virol. 1991;65(12):6589–96.PubMedPubMedCentral
98.
Zurück zum Zitat Xiao C, Tuthill TJ, Bator KCM, Challinor LJ, Chipman PR, Killington RA, et al. Discrimination among rhinovirus serotypes for a variant ICAM-1 receptor molecule. J Virol. 2004;78(18):10034–44.PubMedPubMedCentralCrossRef Xiao C, Tuthill TJ, Bator KCM, Challinor LJ, Chipman PR, Killington RA, et al. Discrimination among rhinovirus serotypes for a variant ICAM-1 receptor molecule. J Virol. 2004;78(18):10034–44.PubMedPubMedCentralCrossRef
99.
Zurück zum Zitat Basta HA, Sgro JY, Palmenberg AC. Modeling of the human rhinovirus C capsid suggests a novel topography with insights on receptor preference and immunogenicity. Virology. 2014;448:176–84.PubMedCrossRef Basta HA, Sgro JY, Palmenberg AC. Modeling of the human rhinovirus C capsid suggests a novel topography with insights on receptor preference and immunogenicity. Virology. 2014;448:176–84.PubMedCrossRef
100.
Zurück zum Zitat Blacklow SC. Versatility in ligand recognition by LDL receptor family proteins: advances and frontiers. Curr Opin Struct Biol. 2007;17(4):419–26.PubMedPubMedCentralCrossRef Blacklow SC. Versatility in ligand recognition by LDL receptor family proteins: advances and frontiers. Curr Opin Struct Biol. 2007;17(4):419–26.PubMedPubMedCentralCrossRef
101.
Zurück zum Zitat Wruss J, Runzler D, Steiger C, Chiba P, Kohler G, Blaas D. Attachment of VLDL receptors to an icosahedral virus along the 5‑fold symmetry axis: multiple binding modes evidenced by fluorescence correlation spectroscopy. Biochemistry. 2007;46(21):6331–9.PubMedCrossRef Wruss J, Runzler D, Steiger C, Chiba P, Kohler G, Blaas D. Attachment of VLDL receptors to an icosahedral virus along the 5‑fold symmetry axis: multiple binding modes evidenced by fluorescence correlation spectroscopy. Biochemistry. 2007;46(21):6331–9.PubMedCrossRef
102.
Zurück zum Zitat Moser R, Snyers L, Wruss J, Angulo J, Peters H, Peters T, et al. Neutralization of a common cold virus by concatemers of the third ligand binding module of the VLDL-receptor strongly depends on the number of modules. Virology. 2005;338(2):259–69.PubMedCrossRef Moser R, Snyers L, Wruss J, Angulo J, Peters H, Peters T, et al. Neutralization of a common cold virus by concatemers of the third ligand binding module of the VLDL-receptor strongly depends on the number of modules. Virology. 2005;338(2):259–69.PubMedCrossRef
103.
Zurück zum Zitat Hewat EA, Neumann E, Conway JF, Moser R, Ronacher B, Marlovits TC, et al. The cellular receptor to human rhinovirus 2 binds around the 5‑fold axis and not in the canyon: a structural view. EMBO J. 2000;19(23):6317–25.PubMedPubMedCentralCrossRef Hewat EA, Neumann E, Conway JF, Moser R, Ronacher B, Marlovits TC, et al. The cellular receptor to human rhinovirus 2 binds around the 5‑fold axis and not in the canyon: a structural view. EMBO J. 2000;19(23):6317–25.PubMedPubMedCentralCrossRef
104.
Zurück zum Zitat Querol-Audi J, Konecsni T, Pous J, Carugo O, Fita I, Verdaguer N, et al. Minor group human rhinovirus-receptor interactions: geometry of multimodular attachment and basis of recognition. FEBS Lett. 2009;583(1):235–40.PubMedCrossRef Querol-Audi J, Konecsni T, Pous J, Carugo O, Fita I, Verdaguer N, et al. Minor group human rhinovirus-receptor interactions: geometry of multimodular attachment and basis of recognition. FEBS Lett. 2009;583(1):235–40.PubMedCrossRef
105.
Zurück zum Zitat Verdaguer N, Fita I, Reithmayer M, Moser R, Blaas D. X‑ray structure of a minor group human rhinovirus bound to a fragment of its cellular receptor protein. Nat Struct Mol Biol. 2004;11(5):429–34.PubMedCrossRef Verdaguer N, Fita I, Reithmayer M, Moser R, Blaas D. X‑ray structure of a minor group human rhinovirus bound to a fragment of its cellular receptor protein. Nat Struct Mol Biol. 2004;11(5):429–34.PubMedCrossRef
106.
Zurück zum Zitat Nicodemou A, Petsch M, Konecsni T, Kremser L, Kenndler E, Casasnovas JM, et al. Rhinovirus-stabilizing activity of artificial VLDL-receptor variants defines a new mechanism for virus neutralization by soluble receptors. FEBS Lett. 2005;579(25):5507–11.PubMedCrossRef Nicodemou A, Petsch M, Konecsni T, Kremser L, Kenndler E, Casasnovas JM, et al. Rhinovirus-stabilizing activity of artificial VLDL-receptor variants defines a new mechanism for virus neutralization by soluble receptors. FEBS Lett. 2005;579(25):5507–11.PubMedCrossRef
107.
Zurück zum Zitat Ronacher B, Marlovits TC, Moser R, Blaas D. Expression and folding of human very-low-density lipoprotein receptor fragments: neutralization capacity toward human rhinovirus HRV2. Virology. 2000;278(2):541–50.PubMedCrossRef Ronacher B, Marlovits TC, Moser R, Blaas D. Expression and folding of human very-low-density lipoprotein receptor fragments: neutralization capacity toward human rhinovirus HRV2. Virology. 2000;278(2):541–50.PubMedCrossRef
108.
Zurück zum Zitat Nizet S, Wruss J, Landstetter N, Snyers L, Blaas D. A mutation in the first ligand-binding repeat of the human very-low-density lipoprotein receptor results in high-affinity binding of the single V1 module to human rhinovirus 2. J Virol. 2005;79(23):14730–6.PubMedPubMedCentralCrossRef Nizet S, Wruss J, Landstetter N, Snyers L, Blaas D. A mutation in the first ligand-binding repeat of the human very-low-density lipoprotein receptor results in high-affinity binding of the single V1 module to human rhinovirus 2. J Virol. 2005;79(23):14730–6.PubMedPubMedCentralCrossRef
109.
Zurück zum Zitat Reithmayer M, Reischl A, Snyers L, Blaas D. Species-specific receptor recognition by a minor-group human rhinovirus (HRV): HRV serotype 1A distinguishes between the murine and the human low-density lipoprotein receptor. J Virol. 2002;76(14):6957–65.PubMedPubMedCentralCrossRef Reithmayer M, Reischl A, Snyers L, Blaas D. Species-specific receptor recognition by a minor-group human rhinovirus (HRV): HRV serotype 1A distinguishes between the murine and the human low-density lipoprotein receptor. J Virol. 2002;76(14):6957–65.PubMedPubMedCentralCrossRef
110.
Zurück zum Zitat Weber C, Pickl-Herk A, Khan AG, Strauss S, Carugo O, Blaas D. Predictive bioinformatic identification of minor receptor group human rhinoviruses. FEBS Lett. 2009;583(15):2547–51.PubMedCrossRef Weber C, Pickl-Herk A, Khan AG, Strauss S, Carugo O, Blaas D. Predictive bioinformatic identification of minor receptor group human rhinoviruses. FEBS Lett. 2009;583(15):2547–51.PubMedCrossRef
111.
Zurück zum Zitat Iijima H, Miyazawa M, Sakai J, Magoori K, Ito MR, Suzuki H, et al. Expression and characterization of a very low density lipoprotein receptor variant lacking the O‑linked sugar region generated by alternative splicing. J Biochem. 1998;124(4):747–55.PubMedCrossRef Iijima H, Miyazawa M, Sakai J, Magoori K, Ito MR, Suzuki H, et al. Expression and characterization of a very low density lipoprotein receptor variant lacking the O‑linked sugar region generated by alternative splicing. J Biochem. 1998;124(4):747–55.PubMedCrossRef
112.
Zurück zum Zitat Nakamura Y, Yamamoto M, Kumamaru E. Very low-density lipoprotein receptor in fetal intestine and gastric adenocarcinoma cells. Arch Pathol Lab Med. 2000;124(1):119–22.PubMed Nakamura Y, Yamamoto M, Kumamaru E. Very low-density lipoprotein receptor in fetal intestine and gastric adenocarcinoma cells. Arch Pathol Lab Med. 2000;124(1):119–22.PubMed
113.
Zurück zum Zitat Konecsni T, Berka U, Pickl-Herk A, Bilek G, Khan AG, Gajdzig L, et al. Low pH-triggered beta-propeller switch of the low-density lipoprotein receptor assists rhinovirus infection. J Virol. 2009;83(21):10922–30.PubMedPubMedCentralCrossRef Konecsni T, Berka U, Pickl-Herk A, Bilek G, Khan AG, Gajdzig L, et al. Low pH-triggered beta-propeller switch of the low-density lipoprotein receptor assists rhinovirus infection. J Virol. 2009;83(21):10922–30.PubMedPubMedCentralCrossRef
114.
115.
Zurück zum Zitat Beglova N, Blacklow SC. The LDL receptor: how acid pulls the trigger. Trends Biochem Sci. 2005;30(6):309–17.PubMedCrossRef Beglova N, Blacklow SC. The LDL receptor: how acid pulls the trigger. Trends Biochem Sci. 2005;30(6):309–17.PubMedCrossRef
116.
Zurück zum Zitat Brabec M, Baravalle G, Blaas D, Fuchs R. Conformational changes, plasma membrane penetration, and infection by human rhinovirus type 2: role of receptors and low pH. J Virol. 2003;77(9):5370–7.PubMedPubMedCentralCrossRef Brabec M, Baravalle G, Blaas D, Fuchs R. Conformational changes, plasma membrane penetration, and infection by human rhinovirus type 2: role of receptors and low pH. J Virol. 2003;77(9):5370–7.PubMedPubMedCentralCrossRef
117.
Zurück zum Zitat Greve JM, Davis G, Meyer AM, Forte CP, Yost SC, Marlor CW, et al. The major human rhinovirus receptor is ICAM-1. Cell. 1989;56(5):839–47.PubMedCrossRef Greve JM, Davis G, Meyer AM, Forte CP, Yost SC, Marlor CW, et al. The major human rhinovirus receptor is ICAM-1. Cell. 1989;56(5):839–47.PubMedCrossRef
118.
Zurück zum Zitat Staunton DE, Marlin SD, Stratowa C, Dustin ML, Springer TA. Primary structure of ICAM-1 demonstrates interaction between members of the immunoglobulin and integrin supergene families. Cell. 1988;52(6):925–33.PubMedCrossRef Staunton DE, Marlin SD, Stratowa C, Dustin ML, Springer TA. Primary structure of ICAM-1 demonstrates interaction between members of the immunoglobulin and integrin supergene families. Cell. 1988;52(6):925–33.PubMedCrossRef
119.
Zurück zum Zitat Olson NH, Kolatkar PR, Oliveira MA, Cheng RH, Greve JM, McClelland A, et al. Structure of a human rhinovirus complexed with its receptor molecule. Proc Natl Acad Sci USA. 1993;90(2):507–11.PubMedPubMedCentralCrossRef Olson NH, Kolatkar PR, Oliveira MA, Cheng RH, Greve JM, McClelland A, et al. Structure of a human rhinovirus complexed with its receptor molecule. Proc Natl Acad Sci USA. 1993;90(2):507–11.PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Xing L, Casasnovas JM, Cheng RH. Structural analysis of human rhinovirus complexed with ICAM-1 reveals the dynamics of receptor-mediated virus uncoating. J Virol. 2003;77(11):6101–7.PubMedPubMedCentralCrossRef Xing L, Casasnovas JM, Cheng RH. Structural analysis of human rhinovirus complexed with ICAM-1 reveals the dynamics of receptor-mediated virus uncoating. J Virol. 2003;77(11):6101–7.PubMedPubMedCentralCrossRef
121.
Zurück zum Zitat Hoover-Litty H, Greve JM. Formation of rhinovirus-soluble ICAM-1 complexes and conformational changes in the virion. J Virol. 1993;67(1):390–7.PubMedPubMedCentral Hoover-Litty H, Greve JM. Formation of rhinovirus-soluble ICAM-1 complexes and conformational changes in the virion. J Virol. 1993;67(1):390–7.PubMedPubMedCentral
122.
Zurück zum Zitat Rossmann MG. The canyon hypothesis. Hiding the host cell receptor attachment site on a viral surface from immune surveillance. J Biol Chem. 1989;264(25):14587–90.PubMed Rossmann MG. The canyon hypothesis. Hiding the host cell receptor attachment site on a viral surface from immune surveillance. J Biol Chem. 1989;264(25):14587–90.PubMed
123.
Zurück zum Zitat Colonno RJ, Condra JH, Mizutani S, Callahan PL, Davies ME, Murcko MA. Evidence for the direct involvement of the rhinovirus canyon in receptor binding. Proc Natl Acad Sci USA. 1988;85(15):5449–53.PubMedPubMedCentralCrossRef Colonno RJ, Condra JH, Mizutani S, Callahan PL, Davies ME, Murcko MA. Evidence for the direct involvement of the rhinovirus canyon in receptor binding. Proc Natl Acad Sci USA. 1988;85(15):5449–53.PubMedPubMedCentralCrossRef
124.
Zurück zum Zitat Smith TJ, Chase ES, Schmidt TJ, Olson NH, Baker TS. Neutralizing antibody to human rhinovirus 14 penetrates the receptor-binding canyon. Nature. 1996;383(6598):350–4.PubMedPubMedCentralCrossRef Smith TJ, Chase ES, Schmidt TJ, Olson NH, Baker TS. Neutralizing antibody to human rhinovirus 14 penetrates the receptor-binding canyon. Nature. 1996;383(6598):350–4.PubMedPubMedCentralCrossRef
125.
Zurück zum Zitat Greve JM, Forte CP, Marlor CW, Meyer AM, Hoover-Litty H, Wunderlich D, et al. Mechanisms of receptor-mediated rhinovirus neutralization defined by two soluble forms of ICAM-1. J Virol. 1991;65(11):6015–23.PubMedPubMedCentral Greve JM, Forte CP, Marlor CW, Meyer AM, Hoover-Litty H, Wunderlich D, et al. Mechanisms of receptor-mediated rhinovirus neutralization defined by two soluble forms of ICAM-1. J Virol. 1991;65(11):6015–23.PubMedPubMedCentral
126.
Zurück zum Zitat Martin S, Casasnovas JM, Staunton DE, Springer TA. Efficient neutralization and disruption of rhinovirus by chimeric ICAM-1/immunoglobulin molecules. J Virol. 1993;67(6):3561–8.PubMedPubMedCentral Martin S, Casasnovas JM, Staunton DE, Springer TA. Efficient neutralization and disruption of rhinovirus by chimeric ICAM-1/immunoglobulin molecules. J Virol. 1993;67(6):3561–8.PubMedPubMedCentral
127.
Zurück zum Zitat Casasnovas JM, Springer TA. Pathway of rhinovirus disruption by soluble intercellular adhesion molecule 1 (ICAM-1): an intermediate in which ICAM-1 is bound and RNA is released. J Virol. 1994;68(9):5882–9.PubMedPubMedCentral Casasnovas JM, Springer TA. Pathway of rhinovirus disruption by soluble intercellular adhesion molecule 1 (ICAM-1): an intermediate in which ICAM-1 is bound and RNA is released. J Virol. 1994;68(9):5882–9.PubMedPubMedCentral
128.
Zurück zum Zitat Casasnovas JM, Springer TA. Kinetics and thermodynamics of virus binding to receptor. Studies with rhinovirus, intercellular adhesion molecule-1 (ICAM-1), and surface plasmon resonance. J Biol Chem. 1995;270(22):13216–24.PubMedCrossRef Casasnovas JM, Springer TA. Kinetics and thermodynamics of virus binding to receptor. Studies with rhinovirus, intercellular adhesion molecule-1 (ICAM-1), and surface plasmon resonance. J Biol Chem. 1995;270(22):13216–24.PubMedCrossRef
129.
Zurück zum Zitat Reisdorph N, Thomas JJ, Katpally U, Chase E, Harris K, Siuzdak G, et al. Human rhinovirus capsid dynamics is controlled by canyon flexibility. Virology. 2003;314(1):34–44.PubMedCrossRef Reisdorph N, Thomas JJ, Katpally U, Chase E, Harris K, Siuzdak G, et al. Human rhinovirus capsid dynamics is controlled by canyon flexibility. Virology. 2003;314(1):34–44.PubMedCrossRef
130.
Zurück zum Zitat Lewis JK, Bothner B, Smith TJ, Siuzdak G. Antiviral agent blocks breathing of the common cold virus. Proc Natl Acad Sci U SA. 1998;95(12):6774–8.CrossRef Lewis JK, Bothner B, Smith TJ, Siuzdak G. Antiviral agent blocks breathing of the common cold virus. Proc Natl Acad Sci U SA. 1998;95(12):6774–8.CrossRef
131.
Zurück zum Zitat Arruda E, Crump CE, Marlin SD, Merluzzi VJ, Hayden FG. In vitro studies of the antirhinovirus activity of soluble intercellular adhesion molecule-1. Antimicrob Agents Chemother. 1992;36(6):1186–91.PubMedPubMedCentralCrossRef Arruda E, Crump CE, Marlin SD, Merluzzi VJ, Hayden FG. In vitro studies of the antirhinovirus activity of soluble intercellular adhesion molecule-1. Antimicrob Agents Chemother. 1992;36(6):1186–91.PubMedPubMedCentralCrossRef
132.
Zurück zum Zitat Nurani G, Lindqvist B, Casasnovas JM. Receptor priming of major group human rhinoviruses for uncoating and entry at mild low-pH environments. J Virol. 2003;77(22):11985–91.PubMedPubMedCentralCrossRef Nurani G, Lindqvist B, Casasnovas JM. Receptor priming of major group human rhinoviruses for uncoating and entry at mild low-pH environments. J Virol. 2003;77(22):11985–91.PubMedPubMedCentralCrossRef
133.
Zurück zum Zitat Bayer N, Prchla E, Schwab M, Blaas D, Fuchs R. Human rhinovirus HRV14 uncoats from early endosomes in the presence of bafilomycin. FEBS Lett. 1999;463(1–2):175–8.PubMedCrossRef Bayer N, Prchla E, Schwab M, Blaas D, Fuchs R. Human rhinovirus HRV14 uncoats from early endosomes in the presence of bafilomycin. FEBS Lett. 1999;463(1–2):175–8.PubMedCrossRef
134.
Zurück zum Zitat Reischl A, Reithmayer M, Winsauer G, Moser R, Gosler I, Blaas D. Viral evolution toward change in receptor usage: adaptation of a major group human rhinovirus to grow in ICAM-1-negative cells. J Virol. 2001;75(19):9312–9.PubMedPubMedCentralCrossRef Reischl A, Reithmayer M, Winsauer G, Moser R, Gosler I, Blaas D. Viral evolution toward change in receptor usage: adaptation of a major group human rhinovirus to grow in ICAM-1-negative cells. J Virol. 2001;75(19):9312–9.PubMedPubMedCentralCrossRef
135.
Zurück zum Zitat Ohlin A, Hoover-Litty H, Sanderson G, Paessens A, Johnston SL, Holgate ST, et al. Spectrum of activity of soluble intercellular adhesion molecule-1 against rhinovirus reference strains and field isolates. Antimicrob Agents Chemother. 1994;38(6):1413–5.PubMedPubMedCentralCrossRef Ohlin A, Hoover-Litty H, Sanderson G, Paessens A, Johnston SL, Holgate ST, et al. Spectrum of activity of soluble intercellular adhesion molecule-1 against rhinovirus reference strains and field isolates. Antimicrob Agents Chemother. 1994;38(6):1413–5.PubMedPubMedCentralCrossRef
136.
Zurück zum Zitat Crump CE, Arruda E, Hayden FG. Comparative antirhinoviral activities of soluble intercellular adhesion molecule-1 (sICAM-1) and chimeric ICAM-1/immunoglobulin A molecule. Antimicrob Agents Chemother. 1994;38(6):1425–7.PubMedPubMedCentralCrossRef Crump CE, Arruda E, Hayden FG. Comparative antirhinoviral activities of soluble intercellular adhesion molecule-1 (sICAM-1) and chimeric ICAM-1/immunoglobulin A molecule. Antimicrob Agents Chemother. 1994;38(6):1425–7.PubMedPubMedCentralCrossRef
137.
Zurück zum Zitat Brioen P, Rombaut B, Boeye A. Hit-and-run neutralization of poliovirus. J Gen Virol. 1985;66(11):2495–9.PubMedCrossRef Brioen P, Rombaut B, Boeye A. Hit-and-run neutralization of poliovirus. J Gen Virol. 1985;66(11):2495–9.PubMedCrossRef
138.
Zurück zum Zitat Plevka P, Lim PY, Perera R, Cardosa J, Suksatu A, Kuhn RJ, et al. Neutralizing antibodies can initiate genome release from human enterovirus 71. Proc Natl Acad Sci U S A. 2014;111(6):2134–9.PubMedPubMedCentralCrossRef Plevka P, Lim PY, Perera R, Cardosa J, Suksatu A, Kuhn RJ, et al. Neutralizing antibodies can initiate genome release from human enterovirus 71. Proc Natl Acad Sci U S A. 2014;111(6):2134–9.PubMedPubMedCentralCrossRef
139.
Zurück zum Zitat Huber M, Brabec M, Bayer N, Blaas D, Fuchs R. Elevated endosomal pH in HeLa cells overexpressing mutant dynamin can affect infection by pH-sensitive viruses. Traffic. 2001;2(10):727–36.PubMedCrossRef Huber M, Brabec M, Bayer N, Blaas D, Fuchs R. Elevated endosomal pH in HeLa cells overexpressing mutant dynamin can affect infection by pH-sensitive viruses. Traffic. 2001;2(10):727–36.PubMedCrossRef
140.
141.
Zurück zum Zitat Gruenberger M, Pevear D, Diana GD, Kuechler E, Blaas D. Stabilization of human rhinovirus serotype 2 against pH-induced conformational change by antiviral compounds. J Gen Virol. 1991;72(2):431–3.PubMedCrossRef Gruenberger M, Pevear D, Diana GD, Kuechler E, Blaas D. Stabilization of human rhinovirus serotype 2 against pH-induced conformational change by antiviral compounds. J Gen Virol. 1991;72(2):431–3.PubMedCrossRef
142.
Zurück zum Zitat Prchla E, Kuechler E, Blaas D, Fuchs R. Uncoating of human rhinovirus serotype 2 from late endosomes. J Virol. 1994;68(6):3713–23.PubMedPubMedCentral Prchla E, Kuechler E, Blaas D, Fuchs R. Uncoating of human rhinovirus serotype 2 from late endosomes. J Virol. 1994;68(6):3713–23.PubMedPubMedCentral
143.
Zurück zum Zitat Bayer N, Schober D, Prchla E, Murphy RF, Blaas D, Fuchs R. Effect of bafilomycin A1 and nocodazole on endocytic transport in HeLa cells: implications for viral uncoating and infection. J Virol. 1998;72(12):9645–55.PubMedPubMedCentral Bayer N, Schober D, Prchla E, Murphy RF, Blaas D, Fuchs R. Effect of bafilomycin A1 and nocodazole on endocytic transport in HeLa cells: implications for viral uncoating and infection. J Virol. 1998;72(12):9645–55.PubMedPubMedCentral
144.
Zurück zum Zitat Neubauer C, Frasel L, Kuechler E, Blaas D. Mechanism of entry of human rhinovirus 2 into HeLa cells. Virology. 1987;158(1):255–8.PubMedCrossRef Neubauer C, Frasel L, Kuechler E, Blaas D. Mechanism of entry of human rhinovirus 2 into HeLa cells. Virology. 1987;158(1):255–8.PubMedCrossRef
145.
Zurück zum Zitat Jurgeit A, McDowell R, Moese S, Meldrum E, Schwendener R, Greber UF. Niclosamide is a proton carrier and targets acidic endosomes with broad antiviral effects. Plos Pathog. 2012;8(10):e1002976.PubMedPubMedCentralCrossRef Jurgeit A, McDowell R, Moese S, Meldrum E, Schwendener R, Greber UF. Niclosamide is a proton carrier and targets acidic endosomes with broad antiviral effects. Plos Pathog. 2012;8(10):e1002976.PubMedPubMedCentralCrossRef
146.
Zurück zum Zitat Hewat EA, Blaas D. Nonneutralizing human rhinovirus serotype 2‑specific monoclonal antibody 2G2 attaches to the region that undergoes the most dramatic changes upon release of the viral RNA. J Virol. 2006;80(24):12398–401.PubMedPubMedCentralCrossRef Hewat EA, Blaas D. Nonneutralizing human rhinovirus serotype 2‑specific monoclonal antibody 2G2 attaches to the region that undergoes the most dramatic changes upon release of the viral RNA. J Virol. 2006;80(24):12398–401.PubMedPubMedCentralCrossRef
147.
Zurück zum Zitat Neumann E, Moser R, Snyers L, Blaas D, Hewat EA. A cellular receptor of human rhinovirus type 2, the very-low-density lipoprotein receptor, binds to two neighboring proteins of the viral capsid. J Virol. 2003;77(15):8504–11.PubMedPubMedCentralCrossRef Neumann E, Moser R, Snyers L, Blaas D, Hewat EA. A cellular receptor of human rhinovirus type 2, the very-low-density lipoprotein receptor, binds to two neighboring proteins of the viral capsid. J Virol. 2003;77(15):8504–11.PubMedPubMedCentralCrossRef
148.
Zurück zum Zitat Panjwani A, Strauss M, Gold S, Wenham H, Jackson T, Chou JJ, et al. Capsid protein VP4 of human rhinovirus induces membrane permeability by the formation of a size-selective multimeric pore. Plos Pathog. 2014;10(8):e1004294.PubMedPubMedCentralCrossRef Panjwani A, Strauss M, Gold S, Wenham H, Jackson T, Chou JJ, et al. Capsid protein VP4 of human rhinovirus induces membrane permeability by the formation of a size-selective multimeric pore. Plos Pathog. 2014;10(8):e1004294.PubMedPubMedCentralCrossRef
149.
Zurück zum Zitat Davis MP, Bottley G, Beales LP, Killington RA, Rowlands DJ, Tuthill TJ. Recombinant VP4 of human rhinovirus induces permeability in model membranes. J Virol. 2008;82(8):4169–74.PubMedPubMedCentralCrossRef Davis MP, Bottley G, Beales LP, Killington RA, Rowlands DJ, Tuthill TJ. Recombinant VP4 of human rhinovirus induces permeability in model membranes. J Virol. 2008;82(8):4169–74.PubMedPubMedCentralCrossRef
150.
Zurück zum Zitat Tosteson MT, Wang H, Naumov A, Chow M. Poliovirus binding to its receptor in lipid bilayers results in particle-specific, temperature-sensitive channels. J Gen Virol. 2004;85(6):1581–9.PubMedCrossRef Tosteson MT, Wang H, Naumov A, Chow M. Poliovirus binding to its receptor in lipid bilayers results in particle-specific, temperature-sensitive channels. J Gen Virol. 2004;85(6):1581–9.PubMedCrossRef
151.
Zurück zum Zitat Danthi P, Tosteson M, Li QH, Chow M. Genome delivery and ion channel properties are altered in VP4 mutants of poliovirus. J Virol. 2003;77(9):5266–74.PubMedPubMedCentralCrossRef Danthi P, Tosteson M, Li QH, Chow M. Genome delivery and ion channel properties are altered in VP4 mutants of poliovirus. J Virol. 2003;77(9):5266–74.PubMedPubMedCentralCrossRef
152.
Zurück zum Zitat Tosteson MT, Chow M. Characterization of the ion channels formed by poliovirus in planar lipid membranes. J Virol. 1997;71(1):507–11.PubMedPubMedCentral Tosteson MT, Chow M. Characterization of the ion channels formed by poliovirus in planar lipid membranes. J Virol. 1997;71(1):507–11.PubMedPubMedCentral
153.
Zurück zum Zitat Brabec M, Schober D, Wagner E, Bayer N, Murphy RF, Blaas D, et al. Opening of size-selective pores in endosomes during human rhinovirus serotype 2 in vivo uncoating monitored by single-organelle flow analysis. J Virol. 2005;79(2):1008–16.PubMedPubMedCentralCrossRef Brabec M, Schober D, Wagner E, Bayer N, Murphy RF, Blaas D, et al. Opening of size-selective pores in endosomes during human rhinovirus serotype 2 in vivo uncoating monitored by single-organelle flow analysis. J Virol. 2005;79(2):1008–16.PubMedPubMedCentralCrossRef
154.
Zurück zum Zitat Shivanna V, Kim Y, Chang KO. Endosomal acidification and cathepsin L activity is required for calicivirus replication. Virology. 2014;464–465:287–95.PubMedCrossRef Shivanna V, Kim Y, Chang KO. Endosomal acidification and cathepsin L activity is required for calicivirus replication. Virology. 2014;464–465:287–95.PubMedCrossRef
155.
Zurück zum Zitat Harutyunyan S, Kumar M, Sedivy A, Subirats X, Kowalski H, Kohler G, et al. Viral uncoating is directional: exit of the genomic RNA in a common cold virus starts with the poly-(A) tail at the 3’-end. Plos Pathog. 2013;9(4):e1003270.PubMedPubMedCentralCrossRef Harutyunyan S, Kumar M, Sedivy A, Subirats X, Kowalski H, Kohler G, et al. Viral uncoating is directional: exit of the genomic RNA in a common cold virus starts with the poly-(A) tail at the 3’-end. Plos Pathog. 2013;9(4):e1003270.PubMedPubMedCentralCrossRef
156.
Zurück zum Zitat Harutyunyan S, Kowalski H, Blaas D. The Rhinovirus subviral a‑particle exposes 3’-terminal sequences of its genomic RNA. J Virol. 2014;88(11):6307–17.PubMedPubMedCentralCrossRef Harutyunyan S, Kowalski H, Blaas D. The Rhinovirus subviral a‑particle exposes 3’-terminal sequences of its genomic RNA. J Virol. 2014;88(11):6307–17.PubMedPubMedCentralCrossRef
157.
Zurück zum Zitat Carpen O, Pallai P, Staunton DE, Springer TA. Association of intercellular adhesion molecule-1 (ICAM-1) with actin-containing cytoskeleton and alpha-actinin. J Cell Biol. 1992;118(5):1223–34.PubMedCrossRef Carpen O, Pallai P, Staunton DE, Springer TA. Association of intercellular adhesion molecule-1 (ICAM-1) with actin-containing cytoskeleton and alpha-actinin. J Cell Biol. 1992;118(5):1223–34.PubMedCrossRef
158.
Zurück zum Zitat Staunton DE, Gaur A, Chan PY, Springer TA. Internalization of a major group human rhinovirus does not require cytoplasmic or transmembrane domains of ICAM-1. J Immunol. 1992;148(10):3271–4.PubMed Staunton DE, Gaur A, Chan PY, Springer TA. Internalization of a major group human rhinovirus does not require cytoplasmic or transmembrane domains of ICAM-1. J Immunol. 1992;148(10):3271–4.PubMed
159.
Zurück zum Zitat Grunert HP, Wolf KU, Langner KD, Sawitzky D, Habermehl KO, Zeichhardt H. Internalization of human rhinovirus 14 into HeLa and ICAM-1-transfected BHK cells. Med Microbiol Immunol. 1997;186(1):1–9.PubMedCrossRef Grunert HP, Wolf KU, Langner KD, Sawitzky D, Habermehl KO, Zeichhardt H. Internalization of human rhinovirus 14 into HeLa and ICAM-1-transfected BHK cells. Med Microbiol Immunol. 1997;186(1):1–9.PubMedCrossRef
160.
Zurück zum Zitat Prchla E, Plank C, Wagner E, Blaas D, Virus- FR. mediated release of endosomal content in vitro: different behavior of adenovirus and rhinovirus serotype 2. J Cell Biol. 1995;131(1):111–23.PubMedCrossRef Prchla E, Plank C, Wagner E, Blaas D, Virus- FR. mediated release of endosomal content in vitro: different behavior of adenovirus and rhinovirus serotype 2. J Cell Biol. 1995;131(1):111–23.PubMedCrossRef
161.
Zurück zum Zitat Zauner W, Blaas D, Kuechler E, Wagner E. Rhinovirus-mediated endosomal release of transfection complexes. J Virol. 1995;69(2):1085–92.PubMedPubMedCentral Zauner W, Blaas D, Kuechler E, Wagner E. Rhinovirus-mediated endosomal release of transfection complexes. J Virol. 1995;69(2):1085–92.PubMedPubMedCentral
162.
Zurück zum Zitat Vlasak M, Goesler I, Blaas D. Human rhinovirus type 89 variants use heparan sulfate proteoglycan for cell attachment. J Virol. 2005;79(10):5963–70.PubMedPubMedCentralCrossRef Vlasak M, Goesler I, Blaas D. Human rhinovirus type 89 variants use heparan sulfate proteoglycan for cell attachment. J Virol. 2005;79(10):5963–70.PubMedPubMedCentralCrossRef
Metadaten
Titel
Viral entry pathways: the example of common cold viruses
verfasst von
Dieter Blaas
Publikationsdatum
01.05.2016
Verlag
Springer Vienna
Erschienen in
Wiener Medizinische Wochenschrift / Ausgabe 7-8/2016
Print ISSN: 0043-5341
Elektronische ISSN: 1563-258X
DOI
https://doi.org/10.1007/s10354-016-0461-2

Weitere Artikel der Ausgabe 7-8/2016

Wiener Medizinische Wochenschrift 7-8/2016 Zur Ausgabe