Principles of polyoma-and papillomavirus uncoating

1 downloads 0 Views 340KB Size Report
Keywords Papillomavirus 4 Polyomavirus 4 Virus entry 4. Virus uncoating 4 Virus ... covered Merkel cell polyomavirus (MCV) is connected to an aggressive ..... ferentially before and after cell attachment, supports the notion of changes in the ...
Med Microbiol Immunol DOI 10.1007/s00430-012-0262-1

REVIEW

Principles of polyoma- and papillomavirus uncoating Carla Cerqueira • Mario Schelhaas

Received: 23 August 2012 / Accepted: 23 August 2012 Ó Springer-Verlag 2012

Abstract Virus particles are vehicles for transmission of the viral genetic information between infected and uninfected cells and organisms. They have evolved to selfassemble, to serve as a protective shell for the viral genome during transfer, and to disassemble when entering a target cell. Disassembly during entry is a complex, multi-step process typically termed uncoating. Uncoating is triggered by multiple host-cell interactions. During cell entry, these interactions occur sequentially in different cellular compartments that the viruses pass through on their way to the site of replication. Here, we highlight the general principles of uncoating for two structurally related virus families, the polyoma- and papillomaviruses. Recent research indicates the use of different compartments and cellular interactions for uncoating despite their structural similarity.

to transmit the viral genomes from infected to uninfected cells or organisms. These particles are simple in content and structure and are composed of nucleic acids (RNA or DNA), proteins, and—for enveloped viruses—membrane lipids. Viral structural proteins assemble into capsids that protect the genomic information. These capsids are mostly spherical, often icosahedral structures, which are composed of many subunits of one or more proteins. Enveloped viruses have an additional lipid membrane displaying viral membrane glycoproteins. Despite their simplicity, virus particles are fascinating nanomachineries capable of selfassembling from limited structural components, of protecting the genome during transmission, and of mediating complex host-cell interactions during entry.

Keywords Papillomavirus  Polyomavirus  Virus entry  Virus uncoating  Virus structure

The assembly-uncoating paradox

Introduction Viruses are intracellular parasites that depend on the assistance of the host to replicate and to ‘live.’ In infected cells, new virus particles are built which serve as vehicles

C. Cerqueira  M. Schelhaas Emmy-Noether Group ‘Virus Endocytosis’, Institutes of Molecular Virology and Medical Biochemistry, University of Mu¨nster, Mu¨nster, Germany M. Schelhaas (&) ZMBE, Institutes for Medical Biochemistry and Molecular Virology, University of Mu¨nster, Von-Esmarch-Str. 56, 48149 Mu¨nster, Germany e-mail: [email protected]

A virus particle serves seemingly opposing functions [1]. For assembly and transmission, the different structural components associate within the nucleus or cytosol of an infected cell into a stable unit, which protect the viral genome from a hostile environment that features, for example, adverse chemical conditions and hydrolases. For successful entry into uninfected cells, the stable interactions must be reversed to release the viral genome into the cytosol or the nucleus for transcription and replication, a stepwise process termed uncoating. In this review, we discuss for two structurally related virus families, the polyoma- and papillomaviruses (PY and PV, respectively), how the dramatic change in virion stability occurs during transmission between two homologous environments. Despite their structural similarities, these viruses have evolved to employ different strategies for their uncoating program.

123

Med Microbiol Immunol

PY and PV Both, PY and PV, are animal viruses with transforming potential (for an overview see [2, 3]). Most infections with human PY tend to cause mild diseases or to be asymptomatic in healthy individuals. However, the recently discovered Merkel cell polyomavirus (MCV) is connected to an aggressive malignancy of the skin, Merkel cell carcinoma [4]. More than 100 PV types are associated with a variety of infections that range from asymptomatic infections, over papilloma/wart formation to anogenital cancers [5]. Data on the structure and uncoating of PY have been mainly derived from Simian Virus 40 (SV40) and murine polyomavirus (mPY), but findings on the human viruses parallel the structural aspects and uncoating strategies seen for these model viruses (for recent reviews see [6–8]). The productive life cycle of PV requires differentiating human epidermal tissue allowing only limited virus production in vitro [9]. Thus, most structural information and insights into the uncoating program has been derived from viruslike particles (VLPs) and the so-called pseudoviruses. Pseudoviruses consist basically of VLPs that incorporate a reporter plasmid [10].

Structural similarities of polyomaand papillomaviruses PY and PV are small, non-enveloped double-stranded DNA viruses. Due to their overall similar structure, they were originally classified as a common virus family, the papovaviridae. The high divergence in genomic organization, in encoded genes and in the biology of their life cycles, has led to their classification as two separate virus families thereafter [11]. The major capsid proteins VP1 and L1 of PY and PV, respectively, form an icosahedral capsid (T = 7) built by 72 pentameric capsomers, from which 12 and 60 pentamers are pentavalently and hexavalently coordinated, respectively (Fig. 1; Table 1; [12–14]). Despite a lack of sequence similarity, VP1 and L1 both build the pentamer by forming a ring of five b-jellyroll domains. These are linked by interacting loops between the core unit of VP1 or L1. The core itself is composed of eight b-strands each [13, 15, 16]. For both, PY and PV, variable loops in the major capsid protein account for most of the neutralizing antibodies being induced [16–18]. Interpentameric links are provided by the C-terminal arm of VP1 or L1 that invades the neighboring pentamers [13, 15, 16, 19]. For PY, this interaction is stabilized by coordinating calcium ions [13, 15, 20–22]. For PV, stabilization by calcium ions seems to occur for bovine papillomavirus (BPV-1), whereas for HPV-11 and -16, calcium ions are most likely not required [23–25].

123

Fig. 1 PV and PY structure. Electron micrograph of negatively stained virus particles, a HPV-16 pseudovirion and b SV40. Scale bars correspond to 50 nm. c Interpentamer connections according to the invading arm model [13, 19]. Triangles represent individual major capsid proteins that form a pentameric capsomer. Depicted is the hexavalent arrangement of capsomers

Interpentamer disulfide bonds further stabilize the capsids of PY and PV by either covalently linking the invading C-terminal arm to a neighboring pentamer (PV and SV40) [13, 15, 19, 22, 26–30] or by clamping it into place (mPY, [31]). Besides the major capsid protein, PY and PV encode for minor capsid proteins that are important for assembly and entry. In PY capsids, two minor capsid proteins are found: VP2 and the splice variant VP3 that lacks an N-terminal extension of about 120 amino acid residues including a myristylation site [32–34]. For PV, the single minor capsid protein is called L2. For the most part, VP2/VP3 and L2 are thought to locate within a central cavity of the capsomers [13, 15, 35–38]. Interaction of the minor capsid proteins with the capsomer occurs primarily by hydrophobic interactions [38–41]. Different to PY, where VP2/VP3 seem to be completely hidden inside the capsid [42], a stretch of

Med Microbiol Immunol Table 1 Overview of the main structural characteristics of PY and PV

Listed are the icosahedral isometry, structural proteins, capsomer arrangement, and intercapsomeric contacts stabilizing the capsomer

Polyomaviruses

Papillomaviruses

Reference [12–16]

Isometry

T = 7 icosahedral

T = 7 icosahedral

Capsomer

Pentamer of VP1

Pentamer of L1

[13, 15, 35, 132, 133]

Viral structural proteins

Major: VP1

Major: L1

[14, 35, 36]

Minor: VP2/VP3

Minor: L2

Stabilization of capsomer contacts

VP1 C-terminal arms invade neighboring capsomers

L1 C-terminal arms invade neighboring capsomers

Calcium binding

Interchain disulfide bonds

[13, 15, 16, 19–22, 26–30]

Interchain disulfide bonds Interaction between capsid proteins

Hydrophobic interactions

about 60 amino acids (aa) close to the N-terminus of L2 is exposed on the capsid surface of PV (e.g., aa 60–120 in the case of BPV-1) [43–45]. However, the extreme N-terminus is thought to fold back into the capsid lumen [46, 47]. Both viruses assemble in the nucleus of infected cells, although—at least for PY—pentamers may be formed already in the cytoplasm [48]. The minor capsid proteins help to incorporate the viral DNA [49, 50]. The formation of disulfide bonds is unlikely to occur in the nucleus due to its reducing environment. Rather, particles mature after cell lysis when the virions are exposed to the oxidizing cell exterior. For PV, it has been shown that maturation of PsV after cell lysis increases the resistance of the virus to proteases [51], whereas in vivo it may be helped by a redox gradient within epidermal tissue [52]. As summarized in Table 1, PY and PV resemble each other in most of the structural elements that stabilize the viral capsids despite a lack in sequence similarity. Virion stability is in principle conferred by strong hydrophobic interaction of the major capsid proteins within a pentamer, by interpentameric links through invading C-terminal arms of neighboring pentamers, and by interchain disulfide bond formation that occurs by maturation. At least some of these stabilizing features need to be reversed for uncoating.

Endoplasmic reticulum-associated machineries facilitate the uncoating of polyomaviruses For binding and internalization, most PY can use gangliosides, that is, glycosphingolipids containing one or more sialic acids [53–57], see also [6, 7, 58]. Although viral interactions with sialic acids on proteins are possible and may be functional, at least for mPY this interaction is futile by restriction of intracellular sorting events that prevent trafficking to infectious compartments [59]. In contrast, JC virus uses a sialic acid-based glycan, LSTc, as receptor, but requires in addition the serotonin receptor 5-HT2A [60–62]. MCV also seems to use two receptor

Hydrophobic interactions

[38–41]

molecules, it binds to heparan sulfate proteoglycans (HSPGs) but may engage sialic acid containing molecules in a second step [63]. Binding affinities between viruses and sialic acids are typically low, but highly specific for a given virus (discussed in detail by [58]). The available x-ray crystallographic data suggest that binding occurs within small contact areas and that they do not induce any conformational changes in the virion [31, 58, 64, 65]. Preincubation of mPY with oligosaccharide receptor fragments increases binding and infectivity, and confers protease resistance of the virion on cells. This has been previously interpreted as induction of a conformational change [66]. Given the structural data, alternative interpretations may be that protease consensus sites are covered by the receptor fragments, or that less specific binding to decoy receptors is blocked, whereas binding to gangliosides may be possible. In any case, the structure of polyomaviruses seems to remain largely unchanged prior to internalization and is thus most likely not activated for uncoating. Internalization of PY into cells occurs mostly by caveolar/lipid raft endocytosis [42, 67–70]. The exception is JC virus that is internalized by clathrin-mediated endocytosis [71, 72]. PY are routed through endosomal compartments and are eventually delivered to the endoplasmic reticulum (ER), which makes them unusual among other viruses [71, 73–76]. For mPY, the ganglioside receptor GD1a mediates sorting to the ER [75], and it is likely that sorting to the ER generally depends on the receptor usage. During endosomal passage, PY encounter a low pH and proteases. For many viruses, conformational changes and proteolytic shedding of virion proteins are induced in endosomes for partial uncoating [77]. Neutralization of the low endosomal pH by lysosomotropic agents typically blocks infection, which has also been observed for PY [73, 75, 78]. However, it is unclear, whether a low pH induces conformational changes in the capsids of most PY. For mPY, a conformational change in the viral particle has been suggested [75]. In contrast, for SV40, neutralization interferes

123

Med Microbiol Immunol

with the endocytic internalization of particles [73], which suggest that the effects observed by lysosomotropic agents may be related to trafficking defects rather than changes in virion structure. Likewise, proteolysis of viral components in endosomal compartments has not been observed [76, 78]. Taken together, it is generally assumed that PY particles arrive in the ER largely unmodified in their structure and that uncoating is in fact initiated in the ER. Most detailed information on the interactions that facilitate uncoating of PY within the ER is derived from mPY and SV40. The interactions largely resemble the behavior of certain bacterial toxins in that they pose as misfolded proteins to engage the ER protein folding and quality control machinery and hijack the ER-associated degradation (ERAD) machinery (see also [79]). Upon arrival in the ER, SV40 is released into the lumen [80]. For SV40, the protein disulfide isomerase (PDI) family member ERp57 isomerases the C9–C9 disulfide bond, which links the vertex to the adjacent pentamers, to a C9–C104 intrachain disulfide bond [76]. In effect, the vertex pentamer interactions with the rest of the capsid are destabilized. Similar observations have been made for mPY and BK virus [78, 81, 82]. In addition, multiple ER chaperones are required for mPY or SV40 infection, that is, ERp29 or PDI, BiP, BAP31, and DNAJ family members, respectively [76, 83–86]. They seem to help exposing the minor capsid proteins VP2/VP3 [42, 83–88]. In addition, they may mediate recognition of the viral particles by the quality control or ERAD machinery. Exposure of VP2 seems to lead to binding of and integration into the ER membrane, an interaction, which has the ability to perforate the membranes [85, 88]. These events are consistent with the initiation of viral penetration into the cytosol. Whether membrane insertion of VP2 alone allows membrane penetration, is unclear. SV40 is translocated into the cytosol as a largely intact structure [80], which may explain the involvement of ERAD factors [76, 78, 83, 89]. Transfer of a large structure across membranes requires an enormous amount of energy. Derlin, a retrotranslocation channel protein, has been implicated in PY infection [80, 89–91]. Interaction with derlin may facilitate a potential recruitment of cytosolic ERAD factors that could help to extract membrane bound PY particles. Alternatively, derlins might form channels large enough for a transfer. In the cytosol, the destabilized PY capsids disassemble, further rendering the DNA accessible for detection [80, 87]. How this occurs mechanistically is not clear. It has been proposed that cytosolic disassembly may be facilitated by a lower calcium concentration [76]. In addition, interactions with Hsp70 chaperones may contribute to this process [92, 93]. To date it is unknown, which remaining viral structure would be imported into the nucleus to initiate infection.

123

Stepwise disassembly of PV in different cell compartments Many cell biological characteristics of PV entry have been identified during the last decade. A number of cellular co-factors facilitating endocytic uptake and initial infection has been identified [94]. Here, we will focus on the interactions that may mediate uncoating of PV. PV can bind directly to cells or to extracellular matrix (ECM) components [95–98]. Binding to tissue culture cells requires heparan sulfate proteoglycans (HPSG) for various types [99–104]. HPV-11 can bind to glypicans and syndecans for infection suggesting that binding may be independent of a specific proteoglycan [102]. However, since syndecan-1 is very abundant in keratinocytes or wounded epidermal tissue, it may be the primary interaction partner for cell binding during infection as it has been proposed recently for HPV-16 [105]. HPV-11 binds ECM secreted by keratinocytes through interactions with laminin-332, whereas HPV-16 and -45 bind ECM partially via heparan sulfates and laminin-332 [95–98]. However, data from a mouse vaginal challenge model suggest that binding of HSPGs may be the exclusive mode of interaction with basement membranes in vivo [103, 106]. Early studies showed that HPV-33 infection becomes resistant to competition by heparin after binding, which may be due to binding of a secondary receptor and/or a conformational change in the virion [101]. The existence of a putative secondary receptor for PV has been supported by experiments using heparin binding drugs that prevent transfer from HSPGs to the elusive receptor and that block infection [98]. Although most PV seem to require HSPGs for binding, this interaction may not be facultative for all subtypes as shown for HPV-31 and HPV-5 in tissue culture [107, 108]. Interestingly, in the mouse model HPV-5 and 31 bind HSPGs on the basement membrane [103]. That antibodies recognize and neutralize infection differentially before and after cell attachment, supports the notion of changes in the conformation of HPV-33 on the cell surface [109]. For HPV-16, these structural changes involve the action of a cell surface-associated peptidylprolyl isomerase, cyclophilin B, which appears to expose the N-terminus of L2 that is originally buried within the capsid structure [110, 111]. Furin, a cellular proprotein convertase, cleaves a conserved consensus site within the exposed C-terminus of L2 [112]. Furin-precleaved virus can infect cells independently of attachment to HSPGs, which further supports the use of a secondary receptor [113]. Both interactions, with furin and with cyclophilin B, are required for infection. Besides allowing transfer or binding of a secondary receptor, they constitute most likely the first changes toward uncoating. For a more detailed discussion

Med Microbiol Immunol

of cell surface-associated conformational changes, see ([6] and [94]). Infectious internalization occurs slowly and asynchronously, which suggests that one of the conformational changes on the cell surface/ECM and/or the subsequent transfer to the secondary receptor is rate limiting [114]. Endocytosis of PV has been attributed to many different mechanisms [115–120]. For HPV-16 entry, a recent study analysed comprehensively the contributions of most key factors involved in the endocytic pathways known to date [114]. Together with a previous study, the cumulative evidence indicates that HPV-16 entry occurs by a novel clathrin-, caveolin-, lipid raft-independent, but actindependent endocytic pathway that may involve tetraspanin microdomains [114, 121]. This pathway has not been described in detail in the existing literature before, which may explain the attribution of PV entry to various different endocytic mechanisms previously. After endocytosis, all PV types that have been tested require a low endosomal pH for infection which suggests viral passage through the endosomal pathway [114, 117, 119, 122–124]. PV appear to spend a protracted time in endosomal compartments, since HPV-16 is sensitive to pH neutralization for two to three hours after infectious internalization [114]. Viral particles accumulate in late endosomes/lysosomes [112, 114, 119, 125–127], which are able to support infection [114]. Unlike PY, PV localization to the ER has not been observed. Like for many other viruses, it has been suggested that low pH or pH-dependent proteases facilitate uncoating and/or membrane penetration by structurally altering the viral capsid. In fact, structural changes of PV virions in endosomal compartments have been observed though not causally linked to a defined compartment: (1) exposure of an epitope that is buried in the viral capsid (HPV-16, -33, [121, 128]); (2) exposure of the L2 C-terminus [46, 112, 125, 126, 129] and the viral pseudogenome [46, 112, 125, 126, 130]. It is interesting to note that while L2 becomes progressively more accessible, epitopes against L1 seem to disappear consistent with major structural alterations and/or degradation of L1 [112, 126]. A number of factors has been recently implicated to assist in the later stages of PV entry (for more information, see [94]). How several of these proteins contribute mechanistically to uncoating, however, remains to be largely speculative. The role of endosomal proteases has been addressed by two studies using, for example, inhibitors of cathepsins [46, 124]. Since the inhibitors do not affect infection or only to a minor extend, it seems unlikely that cathepsins play a major role in uncoating. Addressing the potential reversal of stabilizing disulfides in interpentameric contacts, Campos et al. [130] suggested an involvement of PDI family members analogous to PY.

Although PDIs reside mostly in the ER, the interaction was hypothesized to occur in endosomal compartments. Unfortunately, clear conclusions are hampered by the fact that PDIs are not specifically inhibited by bacitracin [131], the main inhibitor used in the PV entry study [130]. So it remains unclear, how the stabilizing interpentameric contacts are reversed. Interestingly, cyclophilins are able to release L1 pentamers from L2 in a low pH-dependent manner in vitro [122]. Further, inhibition or knockdown of cyclophilins increases the amount of L2 signals colocalizing with L1 signal, which indicates a failed separation of L1 and L2 [122]. Thus, cyclophilins appear to be chaperones that assist in shedding of pentamers from the viral structure. Perturbation of the intramembrane protease c-secretase affects nuclear localization of pseudogenomes but not intracellular trafficking of the virus, or L2 and pseudogenome exposure [125]. Hence, a role of c-secretase in membrane penetration has been suggested. How c-secretase actually facilitates this process remains to be investigated in more detail. A membrane destabilizing peptide in L2 is hypothesized to facilitate membrane penetration by insertion into membranes [132]. Translocation of the viral histone-genome complex through a membrane or more so organelle lysis would require a large amount of energy. Thus, it seems unlikely that insertion into the membrane of the limited number of L2 molecules present in a virion would be sufficient to provide this energy. This may suggest the involvement of further host factors to assist in membrane penetration. An interaction of sorting nexin 17 with L2 after the putative membrane insertion may retain the remnant virion in a compartment that favors penetration [126]. In the nucleus, only L2 and the viral genome has been detected [112]. Whether L2 and the histone complexes are sufficient to protect the viral genome through their passage in the cytsosol, is an open question. In summary, PV appear to initiate the uncoating process by a series of extracellular conformational changes in the virion. This is followed by further uncoating events in endosomal compartments that may be linked to membrane penetration. In which form the virus remnant structure is translocated, and whether further changes in the cytosol occur has not been investigated so far.

Summary and perspectives PY and PV have evolved to form capsids that exhibit a striking structural similarity despite a lack of significant sequence similarity between the respective structural proteins. Both viruses commonly use structural modifications of their capsid which occur through interactions with host-cell factors during entry into host cells for uncoating. Although many details in the uncoating process of PY and PV or more

123

Med Microbiol Immunol Table 2 Overview of the main structural changes during PY uncoating Structural change

PY

Cellular co-factor

Location

Reference

Conformational change

mPY

Low pH

Endolysosomes

[75]

Isomerization of disulfide bonds of VP1 pentamer

SV40, mPY

PDI family members

ER

[76, 81, 82]

Conformational change: exposure of C-terminal arm of VP1 Conformational change: VP2/VP3 exposure

mPY

ERp29

ER

[85]

SV40, mPY

ErP29, PDI, BiP, BAP31, DNA J family

ER

[42, 76, 83–88]

Conformational change: genome exposure

SV40

Calcium?

Cytosol

[76, 87]

Capsomer shedding

SV40, mPY

Hsp70?

Cytosol

[80, 92, 93]

Listed are the major examples for changes in viral structure during cell entry which have been shown or hypothesized in the references provided. Given are also the detected or presumed locations of the change within the cell. The changes are ordered such that they reflect particles on their way toward the site of replication; ? proposed

Table 3 Overview of the main structural changes during PV uncoating Structural change

HPV type

Cellular co-factor

Location

Reference

Conformational change: heparin resistance

33

?

Cell surface

[101]

Conformational change: H33.J3 epitope exposure (aa 51–58)

33

?

Cell surface

[109, 135]

Conformational change: exposure of L2 N-terminus

16

Cyclophilin B

Cell surface/basement membrane

[110, 111]

L2 N-terminus cleavage

16, BPV

Furin

Cell surface/basement membrane

[46, 104, 113]

Conformational change: L1-7 epitope exposure (aa 329–339)

16

Low pH

Endosomes

[121, 128]

Genome exposure

16, BPV

Low pH

Endosomes

[46, 112, 125, 126, 130]

Capsomer shedding L1–L2 separation

16

Cyclophilins

Endosomes

[112, 122]

Membrane penetration

16

c-secretase

Endosomes

[125]

Listed are the major examples for changes in viral structure during cell entry which have been shown or hypothesized in the references provided. Given are also the detected or presumed locations of the change within the cell. The changes are ordered such that they reflect particles on their way toward the site of replication; ? unknown

so of virus types are still lacking, it has become apparent over the past decade that PY and PV use different sets of cellular compartments, interaction partners, and triggers to facilitate the stepwise process of uncoating (for overview, see Table 2, 3). PY use mostly gangliosides for transport to the ER. In contrast, PV are transported to late endosomal/ lysosomal compartments. The major structural changes in PY capsid seem to be initiated in the ER, whereas for PV uncoating most likely starts with changes initiated extracellularly. PY appear to shed capsomers in the cytosol, whereas for PV this may already occur within endosomes. The theme of divergent uncoating principles between PY and PV has become clear, but there are several key questions still unanswered. For example, how are the stabilizing interpentameric contacts in PV weakened? How are the viral genomes structurally protected until they enter

123

the nucleus? Are there cellular factors associated with them, and if so, which ones? How are they shed for transcriptional activation? We can also expect that different viruses from the same family utilize different but analogous host-cell factors such as ER chaperones for PY to achieve a similar outcome. Hence, more efforts are required to fully understand the uncoating process for the different viruses. Acknowledgments The authors would like to acknowledge the help of Dr. R. Mancini in acquisition of electron micrographs. We apologize to all those great individuals who have advanced the field of polyoma- and papillomavirus structure and entry, and who were not mentioned in the manuscript due to space limitations. MS and CC were supported by the German Science Foundation (DFG, Emmy-Noether grant SCHE 1552/2-1) and by the Portuguese Foundation for Science and Technology (PhD grant SFRH/BD/45921/2008), respectively.

Med Microbiol Immunol

References 1. Greber UF, Singh I, Helenius A (1994) Mechanisms of virus uncoating. Trends Microbiol 2(2):52–56 2. Howley PM, Lowy DR (2007) Papillomaviruses. In: Fields BN, Howley PM, Griffin DE, Lamb RA, Martin MA, Roizman B, Straus SE, Knipe DM (eds) Fields Virology, 5th edn. Lippincott Williams and Wilkins, Philadelphia, PA, USA, p 2300–2354 3. Imperiale MJ, Major EO (2007) Polyomavirus. In: Fields BN, Howley PM, Griffin DE, Lamb RA, Martin MA, Roizman B, Straus SE, Knipe DM (eds) Fields Virology, 5th edn. Lippincott Williams and Wilkins, Philadelphia, PA, USA, p 2268–2282 4. Feng H, Shuda M, Chang Y, Moore PS (2008) Clonal integration of a polyomavirus in human Merkel cell carcinoma. Science 319(5866):1096–1100. doi:10.1126/science.1152586 5. de Villiers EM, Fauquet C, Broker TR, Bernard HU, zur Hausen H (2004) Classification of papillomaviruses. Virology 324(1): 17–27. doi:10.1016/j.virol.2004.03.033 6. Sapp M, Day PM (2009) Structure, attachment and entry of polyoma- and papillomaviruses. Virology 384(2):400–409. doi: 10.1016/j.virol.2008.12.022 7. Tsai B, Qian M (2010) Cellular entry of polyomaviruses. Curr Top Microbiol Immunol 343:177–194. doi:10.1007/82_2010_38 8. Neu U, Stehle T, Atwood WJ (2009) The Polyomaviridae: contributions of virus structure to our understanding of virus receptors and infectious entry. Virology 384(2):389–399. doi: 10.1016/j.virol.2008.12.021 9. Longworth MS, Laimins LA (2004) Pathogenesis of human papillomaviruses in differentiating epithelia. Microbiol Mol Biol Rev 68(2):362–372 10. Buck CB, Pastrana DV, Lowy DR, Schiller JT (2004) Efficient intracellular assembly of papillomaviral vectors. J Virol 78(2): 751–757 11. Van Regenmortel M, Fauquet C, Bishop D, Carstens E, Estes M, Lemon S, Maniloff J, Mayo M, McGeoch D, Pringle C, Wickner R (2000) Virus taxonomy. Seventh Report of the International Committee on Taxonomy of Viruses Academic Press, San Diego, New York 12. Klug A, Finch JT (1965) Structure of Viruses of the PapillomaPolyoma Type. I. Human Wart Virus. J Mol Biol 11:403–423 13. Liddington RC, Yan Y, Moulai J, Sahli R, Benjamin TL, Harrison SC (1991) Structure of simian virus 40 at 3.8-A resolution. Nature 354(6351):278–284 14. Baker TS, Newcomb WW, Olson NH, Cowsert LM, Olson C, Brown JC (1991) Structures of bovine and human papillomaviruses. Analysis by cryoelectron microscopy and threedimensional image reconstruction. Biophys J 60(6):1445–1456 15. Stehle T, Gamblin SJ, Yan Y, Harrison SC (1996) The structure of simian virus 40 refined at 3.1 A resolution. Structure 4(2):165–182 16. Chen XS, Garcea RL, Goldberg I, Casini G, Harrison SC (2000) Structure of small virus-like particles assembled from the L1 protein of human papillomavirus 16. Mol Cell 5(3):557–567 17. Bishop B, Dasgupta J, Klein M, Garcea RL, Christensen ND, Zhao R, Chen XS (2007) Crystal structures of four types of human papillomavirus L1 capsid proteins: understanding the specificity of neutralizing monoclonal antibodies. J Biol Chem 282(43):31803–31811. doi:10.1074/jbc.M706380200 18. Murata H, Teferedegne B, Sheng L, Lewis AM Jr, Peden K (2008) Identification of a neutralization epitope in the VP1 capsid protein of SV40. Virology 381(1):116–122. doi:10.1016/j. virol.2008.07.032 19. Modis Y, Trus BL, Harrison SC (2002) Atomic model of the papillomavirus capsid. EMBO J 21(18):4754–4762 20. Brady JN, Kendall JD, Consigli RA (1979) In vitro reassembly of infectious polyoma virions. J Virol 32(2):640–647

21. Brady JN, Winston VD, Consigli RA (1977) Dissociation of polyoma virus by the chelation of calcium ions found associated with purified virions. J Virol 23(3):717–724 22. Ishizu KI, Watanabe H, Han SI, Kanesashi SN, Hoque M, Yajima H, Kataoka K, Handa H (2001) Roles of disulfide linkage and calcium ion-mediated interactions in assembly and disassembly of virus-like particles composed of simian virus 40 VP1 capsid protein. J Virol 75(1):61–72. doi:10.1128/JVI.75. 1.61-72.2001 23. Hanslip SJ, Zaccai NR, Middelberg AP, Falconer RJ (2006) Assembly of human papillomavirus type-16 virus-like particles: multifactorial study of assembly and competing aggregation. Biotechnol Prog 22(2):554–560. doi:10.1021/bp0502781 24. Paintsil J, Muller M, Picken M, Gissmann L, Zhou J (1998) Calcium is required in reassembly of bovine papillomavirus in vitro. J Gen Virol 79(Pt 5):1133–1141 25. McCarthy MP, White WI, Palmer-Hill F, Koenig S, Suzich JA (1998) Quantitative disassembly and reassembly of human papillomavirus type 11 viruslike particles in vitro. J Virol 72(1): 32–41 26. Sapp M, Fligge C, Petzak I, Harris JR, Streeck RE (1998) Papillomavirus assembly requires trimerization of the major capsid protein by disulfides between two highly conserved cysteines. J Virol 72(7):6186–6189 27. Sapp M, Volpers C, Muller M, Streeck RE (1995) Organization of the major and minor capsid proteins in human papillomavirus type 33 virus-like particles. J Gen Virol 76(Pt 9):2407–2412 28. Li M, Beard P, Estes PA, Lyon MK, Garcea RL (1998) Intercapsomeric disulfide bonds in papillomavirus assembly and disassembly. J Virol 72(3):2160–2167 29. Ishii Y, Tanaka K, Kanda T (2003) Mutational analysis of human papillomavirus type 16 major capsid protein L1: the cysteines affecting the intermolecular bonding and structure of L1-capsids. Virology 308(1):128–136 30. Conway MJ, Cruz L, Alam S, Christensen ND, Meyers C (2011) Differentiation-dependent interpentameric disulfide bond stabilizes native human papillomavirus type 16. PLoS ONE 6(7):e22427. doi:10.1371/journal.pone.0022427 31. Stehle T, Harrison SC (1996) Crystal structures of murine polyomavirus in complex with straight-chain and branched-chain sialyloligosaccharide receptor fragments. Structure 4(2):183–194 32. Krauzewicz N, Streuli CH, Stuart-Smith N, Jones MD, Wallace S, Griffin BE (1990) Myristylated polyomavirus VP2: role in the life cycle of the virus. J Virol 64(9):4414–4420 33. Sahli R, Freund R, Dubensky T, Garcea R, Bronson R, Benjamin T (1993) Defect in entry and altered pathogenicity of a polyoma virus mutant blocked in VP2 myristylation. Virology 192(1):142–153. doi:10.1006/viro.1993.1016 34. Gasparovic ML, Gee GV, Atwood WJ (2006) JC virus minor capsid proteins Vp2 and Vp3 are essential for virus propagation. J Virol 80(21):10858–10861. doi:10.1128/JVI.01298-06 35. Baker TS, Drak J, Bina M (1989) The capsid of small papova viruses contains 72 pentameric capsomeres: direct evidence from cryo-electron-microscopy of simian virus 40. Biophys J 55(2):243–253. doi:10.1016/S0006-3495(89)82799-7 36. Griffith JP, Griffith DL, Rayment I, Murakami WT, Caspar DL (1992) Inside polyomavirus at 25-A resolution. Nature 355(6361):652–654. doi:10.1038/355652a0 37. Buck CB, Cheng N, Thompson CD, Lowy DR, Steven AC, Schiller JT, Trus BL (2008) Arrangement of L2 within the Papillomavirus capsid. J Virol 82(11):5190–5197 38. Chen XS, Stehle T, Harrison SC (1998) Interaction of polyomavirus internal protein VP2 with the major capsid protein VP1 and implications for participation of VP2 in viral entry. EMBO J 17(12):3233–3240. doi:10.1093/emboj/17.12.3233

123

Med Microbiol Immunol 39. Finnen RL, Erickson KD, Chen XS, Garcea RL (2003) Interactions between papillomavirus L1 and L2 capsid proteins. J Virol 77(8):4818–4826 40. Okun MM, Day PM, Greenstone HL, Booy FP, Lowy DR, Schiller JT, Roden RB (2001) L1 interaction domains of papillomavirus l2 necessary for viral genome encapsidation. J Virol 75(9):4332–4342. doi:10.1128/JVI.75.9.4332-4342.2001 41. Barouch DH, Harrison SC (1994) Interactions among the major and minor coat proteins of polyomavirus. J Virol 68(6):3982– 3989 42. Norkin LC, Anderson HA, Wolfrom SA, Oppenheim A (2002) Caveolar endocytosis of simian virus 40 is followed by brefeldin A-sensitive transport to the endoplasmic reticulum, where the virus disassembles. J Virol 76(10):5156–5166 43. Kawana Y, Kawana K, Yoshikawa H, Taketani Y, Yoshiike K, Kanda T (2001) Human papillomavirus type 16 minor capsid protein l2 N-terminal region containing a common neutralization epitope binds to the cell surface and enters the cytoplasm. J Virol 75(5):2331–2336 44. Kondo K, Ishii Y, Ochi H, Matsumoto T, Yoshikawa H, Kanda T (2007) Neutralization of HPV16, 18, 31, and 58 pseudovirions with antisera induced by immunizing rabbits with synthetic peptides representing segments of the HPV16 minor capsid protein L2 surface region. Virology 358(2):266–272. doi: 10.1016/j.virol.2006.08.037 45. Liu WJ, Gissmann L, Sun XY, Kanjanahaluethai A, Muller M, Doorbar J, Zhou J (1997) Sequence close to the N-terminus of L2 protein is displayed on the surface of bovine papillomavirus type 1 virions. Virology 227(2):474–483. doi:10.1006/viro. 1996.8348 46. Richards RM, Lowy DR, Schiller JT, Day PM (2006) Cleavage of the papillomavirus minor capsid protein, L2, at a furin consensus site is necessary for infection. Proc Nat Acad Sci USA 103(5):1522–1527 47. Yang R, Day PM, Yutzy WH, Lin KY, Hung CF, Roden RB (2003) Cell surface-binding motifs of L2 that facilitate papillomavirus infection. J Virol 77(6):3531–3541 48. Li PP, Nakanishi A, Clark SW, Kasamatsu H (2002) Formation of transitory intrachain and interchain disulfide bonds accompanies the folding and oligomerization of simian virus 40 Vp1 in the cytoplasm. Proc Nat Acad Sci USA 99(3):1353–1358. doi: 10.1073/pnas.032668699 49. Dean DA, Li PP, Lee LM, Kasamatsu H (1995) Essential role of the Vp2 and Vp3 DNA-binding domain in simian virus 40 morphogenesis. J Virol 69(2):1115–1121 50. Bordeaux J, Forte S, Harding E, Darshan MS, Klucevsek K, Moroianu J (2006) The l2 minor capsid protein of low-risk human papillomavirus type 11 interacts with host nuclear import receptors and viral DNA. J Virol 80(16):8259–8262. doi:10. 1128/JVI.00776-06 51. Buck CB, Thompson CD, Pang YY, Lowy DR, Schiller JT (2005) Maturation of papillomavirus capsids. J Virol 79(5):2839–2846. doi:10.1128/JVI.79.5.2839-2846.2005 52. Conway MJ, Alam S, Ryndock EJ, Cruz L, Christensen ND, Roden RB, Meyers C (2009) Tissue-spanning redox gradientdependent assembly of native human papillomavirus type 16 virions. J Virol 83(20):10515–10526. doi:10.1128/JVI.00731-09 53. Low JA, Magnuson B, Tsai B, Imperiale MJ (2006) Identification of gangliosides GD1b and GT1b as receptors for BK virus. J Virol 80(3):1361–1366. doi:10.1128/JVI.80.3.1361-1366.2006 54. Smith AE, Lilie H, Helenius A (2003) Ganglioside-dependent cell attachment and endocytosis of murine polyomavirus-like particles. FEBS Lett 555(2):199–203 55. Tsai B, Gilbert JM, Stehle T, Lencer W, Benjamin TL, Rapoport TA (2003) Gangliosides are receptors for murine polyoma virus and SV40. EMBO J 22(17):4346–4355

123

56. Neu U, Woellner K, Gauglitz G, Stehle T (2008) Structural basis of GM1 ganglioside recognition by simian virus 40. Proc Nat Acad Sci USA 105(13):5219–5224. doi:10.1073/pnas.0710301 105 57. Campanero-Rhodes MA, Smith A, Chai W, Sonnino S, Mauri L, Childs RA, Zhang Y, Ewers H, Helenius A, Imberty A, Feizi T (2007) N-glycolyl GM1 ganglioside as a receptor for simian virus 40. J Virol 81(23):12846–12858. doi:10.1128/JVI.0131107 58. Neu U, Bauer J, Stehle T (2011) Viruses and sialic acids: rules of engagement. Curr Opin Struct Biol 21(5):610–618. doi: 10.1016/j.sbi.2011.08.009 59. Qian M, Tsai B (2010) Lipids and proteins act in opposing manners to regulate polyomavirus infection. J Virol 84(19): 9840–9852. doi:10.1128/JVI.01093-10 60. Liu CK, Wei G, Atwood WJ (1998) Infection of glial cells by the human polyomavirus JC is mediated by an N-linked glycoprotein containing terminal alpha(2–6)-linked sialic acids. J Virol 72(6):4643–4649 61. Neu U, Maginnis MS, Palma AS, Stroh LJ, Nelson CD, Feizi T, Atwood WJ, Stehle T (2010) Structure-function analysis of the human JC polyomavirus establishes the LSTc pentasaccharide as a functional receptor motif. Cell Host Microbe 8(4):309–319. doi:10.1016/j.chom.2010.09.004 62. Maginnis MS, Haley SA, Gee GV, Atwood WJ (2010) Role of N-linked glycosylation of the 5-HT2A receptor in JC virus infection. J Virol 84(19):9677–9684. doi:10.1128/JVI.00978-10 63. Schowalter RM, Pastrana DV, Buck CB (2011) Glycosaminoglycans and sialylated glycans sequentially facilitate Merkel cell polyomavirus infectious entry. PLoS Pathog 7(7):e1002161. doi: 10.1371/journal.ppat.1002161 64. Stehle T, Harrison SC (1997) High-resolution structure of a polyomavirus VP1-oligosaccharide complex: implications for assembly and receptor binding. EMBO J 16(16):5139–5148 65. Stehle T, Yan Y, Benjamin TL, Harrison SC (1994) Structure of murine polyomavirus complexed with an oligosaccharide receptor fragment. Nature 369(6476):160–163 66. Cavaldesi M, Caruso M, Sthandier O, Amati P, Garcia MI (2004) Conformational changes of murine polyomavirus capsid proteins induced by sialic acid binding. J Biol Chem 279(40):41573–41579. doi:10.1074/jbc.M405995200 67. Anderson HA, Chen Y, Norkin LC (1996) Bound simian virus 40 translocates to caveolin-enriched membrane domains, and its entry is inhibited by drugs that selectively disrupt caveolae. Mol Biol Cell 7(11):1825–1834 68. Damm EM, Pelkmans L, Kartenbeck J, Mezzacasa A, Kurzchalia T, Helenius A (2005) Clathrin- and caveolin-1-independent endocytosis: entry of simian virus 40 into cells devoid of caveolae. J Cell Biol 168(3):477–488 69. Pelkmans L, Kartenbeck J, Helenius A (2001) Caveolar endocytosis of simian virus 40 reveals a new two-step vesiculartransport pathway to the ER. Nat Cell Biol 3(5):473–483 70. Pelkmans L, Puntener D, Helenius A (2002) Local actin polymerization and dynamin recruitment in SV40-induced internalization of caveolae. Science 296(5567):535–539 71. Querbes W, Benmerah A, Tosoni D, Di Fiore PP, Atwood WJ (2004) A JC virus-induced signal is required for infection of glial cells by a clathrin- and eps15-dependent pathway. J Virol 78(1):250–256 72. Pho MT, Ashok A, Atwood WJ (2000) JC virus enters human glial cells by clathrin-dependent receptor-mediated endocytosis. J Virol 74(5):2288–2292 73. Engel S, Heger T, Mancini R, Herzog F, Kartenbeck J, Hayer A, Helenius A (2011) Role of endosomes in simian virus 40 entry and infection. J Virol 85(9):4198–4211. doi:10.1128/JVI.0 2179-10

Med Microbiol Immunol 74. Kartenbeck J, Stukenbrok H, Helenius A (1989) Endocytosis of simian virus 40 into the endoplasmic reticulum. J Cell Biol 109(6 Pt 1):2721–2729 75. Qian M, Cai D, Verhey KJ, Tsai B (2009) A lipid receptor sorts polyomavirus from the endolysosome to the endoplasmic reticulum to cause infection. PLoS Pathog 5(6):e1000465. doi: 10.1371/journal.ppat.1000465 76. Schelhaas M, Malmstrom J, Pelkmans L, Haugstetter J, Ellgaard L, Grunewald K, Helenius A (2007) Simian virus 40 depends on ER protein folding and quality control factors for entry into host cells. Cell 131(3):516–529 77. Schelhaas M (2010) Come in and take your coat off—how host cells provide endocytosis for virus entry. Cell Microbiol 12(10):1378–1388. doi:10.1111/j.1462-5822.2010.01510.x 78. Jiang M, Abend JR, Tsai B, Imperiale MJ (2009) Early events during BK virus entry and disassembly. J Virol 83(3):1350– 1358. doi:10.1128/JVI.02169-08 79. Inoue T, Moore P, Tsai B (2011) How viruses and toxins disassemble to enter host cells. Annu Rev Microbiol 65:287–305. doi:10.1146/annurev-micro-090110-102855 80. Inoue T, Tsai B (2011) A large and intact viral particle penetrates the endoplasmic reticulum membrane to reach the cytosol. PLoS Pathog 7(5):e1002037. doi:10.1371/journal.ppat.1002037 81. Gilbert J, Ou W, Silver J, Benjamin T (2006) Downregulation of protein disulfide isomerase inhibits infection by the mouse polyomavirus. J Virol 80(21):10868–10870. doi:10.1128/JVI.0 1117-06 82. Walczak CP, Tsai B (2011) A PDI family network acts distinctly and coordinately with ERp29 to facilitate polyomavirus infection. J Virol 85(5):2386–2396. doi:10.1128/JVI.01855-10 83. Geiger R, Andritschke D, Friebe S, Herzog F, Luisoni S, Heger T, Helenius A (2011) BAP31 and BiP are essential for dislocation of SV40 from the endoplasmic reticulum to the cytosol. Nat Cell Biol 13(11):1305–1314. doi:10.1038/ncb2339 84. Goodwin EC, Lipovsky A, Inoue T, Magaldi TG, Edwards AP, Van Goor KE, Paton AW, Paton JC, Atwood WJ, Tsai B, Di Maio D (2011) BiP and multiple DNAJ molecular chaperones in the endoplasmic reticulum are required for efficient simian virus infection. mBio 2(3):e00101–e00111. doi:10.1128/mBio.00101-11 85. Magnuson B, Rainey EK, Benjamin T, Baryshev M, Mkrtchian S, Tsai B (2005) ERp29 triggers a conformational change in polyomavirus to stimulate membrane binding. Mol Cell 20(2): 289–300 86. Rainey-Barger EK, Magnuson B, Tsai B (2007) A chaperoneactivated nonenveloped virus perforates the physiologically relevant endoplasmic reticulum membrane. J Virol 81(23): 12996–13004. doi:10.1128/JVI.01037-07 87. Kuksin D, Norkin LC (2012) Disassembly of simian virus 40 during passage through the endoplasmic reticulum and in the cytoplasm. J Virol 86(3):1555–1562. doi:10.1128/JVI.05753-11 88. Daniels R, Rusan NM, Wadsworth P, Hebert DN (2006) SV40 VP2 and VP3 insertion into ER membranes is controlled by the capsid protein VP1: implications for DNA translocation out of the ER. Mol Cell 24(6):955–966. doi:10.1016/j.molcel.2006.11. 001 89. Lilley BN, Gilbert JM, Ploegh HL, Benjamin TL (2006) Murine polyomavirus requires the endoplasmic reticulum protein Derlin-2 to initiate infection. J Virol 80(17):8739–8744. doi:10.1128/ JVI.00791-06 90. Lilley BN, Ploegh HL (2004) A membrane protein required for dislocation of misfolded proteins from the ER. Nature 429(6994):834–840 91. Ye Y, Shibata Y, Yun C, Ron D, Rapoport TA (2004) A membrane protein complex mediates retro-translocation from the ER lumen into the cytosol. Nature 429(6994):841–847

92. Chromy LR, Oltman A, Estes PA, Garcea RL (2006) Chaperone-mediated in vitro disassembly of polyoma- and papillomaviruses. J Virol 80(10):5086–5091. doi:10.1128/JVI.80.10. 5086-5091.2006 93. Li PP, Itoh N, Watanabe M, Shi Y, Liu P, Yang HJ, Kasamatsu H (2009) Association of simian virus 40 vp1 with 70-kilodalton heat shock proteins and viral tumor antigens. J Virol 83(1):37–46. doi:10.1128/JVI.00844-08 94. Florin L, Sapp M, Spoden G (2012) Host cell factors in papillomavirus entry. Med Microbiol Immunol. doi:10.1007/s00430012-0270-1 95. Broutian TR, Brendle SA, Christensen ND (2010) Differential binding patterns to host cells associated with particles of several human alphapapillomavirus types. J Gen Virol 91(Pt 2):531– 540. doi:10.1099/vir.0.012732-0 96. Culp TD, Budgeon LR, Christensen ND (2006) Human papillomaviruses bind a basal extracellular matrix component secreted by keratinocytes which is distinct from a membraneassociated receptor. Virology 347(1):147–159 97. Culp TD, Budgeon LR, Marinkovich MP, Meneguzzi G, Christensen ND (2006) Keratinocyte-secreted laminin 5 can function as a transient receptor for human papillomaviruses by binding virions and transferring them to adjacent cells. J Virol 80(18):8940–8950 98. Selinka HC, Florin L, Patel HD, Freitag K, Schmidtke M, Makarov VA, Sapp M (2007) Inhibition of transfer to secondary receptors by heparan sulfate-binding drug or antibody induces noninfectious uptake of human papillomavirus. J Virol 81(20): 10970–10980 99. Joyce JG, Tung JS, Przysiecki CT, Cook JC, Lehman ED, Sands JA, Jansen KU, Keller PM (1999) The L1 major capsid protein of human papillomavirus type 11 recombinant virus-like particles interacts with heparin and cell-surface glycosaminoglycans on human keratinocytes. J Biol Chem 274(9):5810–5822 100. Combita AL, Touze A, Bousarghin L, Sizaret PY, Munoz N, Coursaget P (2001) Gene transfer using human papillomavirus pseudovirions varies according to virus genotype and requires cell surface heparan sulfate. FEMS Microbiol Lett 204(1): 183–188 101. Giroglou T, Florin L, Schafer F, Streeck RE, Sapp M (2001) Human papillomavirus infection requires cell surface heparan sulfate. J Virol 75(3):1565–1570 102. Shafti-Keramat S, Handisurya A, Kriehuber E, Meneguzzi G, Slupetzky K, Kirnbauer R (2003) Different heparan sulfate proteoglycans serve as cellular receptors for human papillomaviruses. J Virol 77(24):13125–13135 103. Johnson KM, Kines RC, Roberts JN, Lowy DR, Schiller JT, Day PM (2009) Role of heparan sulfate in attachment to and infection of the murine female genital tract by human papillomavirus. J Virol 83(5):2067–2074. doi:10.1128/JVI.02190-08 104. Kines RC, Thompson CD, Lowy DR, Schiller JT, Day PM (2009) The initial steps leading to papillomavirus infection occur on the basement membrane prior to cell surface binding. Proc Nat Acad Sci USA 106(48):20458–20463. doi:10.1073/ pnas.0908502106 105. Surviladze Z, Dziduszko A, Ozbun MA (2012) Essential roles for soluble virion-associated heparan sulfonated proteoglycans and growth factors in human papillomavirus infections. PLoS Pathog 8(2):e1002519. doi:10.1371/journal.ppat.1002519 106. Roberts JN, Buck CB, Thompson CD, Kines R, Bernardo M, Choyke PL, Lowy DR, Schiller JT (2007) Genital transmission of HPV in a mouse model is potentiated by nonoxynol-9 and inhibited by carrageenan. Nat Med 13(7):857–861 107. Buck CB, Thompson CD, Roberts JN, Muller M, Lowy DR, Schiller JT (2006) Carrageenan is a potent inhibitor of

123

Med Microbiol Immunol

108.

109.

110.

111.

112.

113.

114.

115.

116.

117.

118.

119. 120.

121.

papillomavirus infection. PLoS Pathog 2(7):e69. doi:10.1371/ journal.ppat.0020069 Patterson NA, Smith JL, Ozbun MA (2005) Human papillomavirus type 31b infection of human keratinocytes does not require heparan sulfate. J Virol 79(11):6838–6847. doi:10. 1128/JVI.79.11.6838-6847.2005 Selinka HC, Giroglou T, Nowak T, Christensen ND, Sapp M (2003) Further evidence that papillomavirus capsids exist in two distinct conformations. J Virol 77(24):12961–12967 Day PM, Gambhira R, Roden RB, Lowy DR, Schiller JT (2008) Mechanisms of human papillomavirus type 16 neutralization by l2 cross-neutralizing and l1 type-specific antibodies. J Virol 82(9):4638–4646 Bienkowska-Haba M, Patel HD, Sapp M (2009) Target cell cyclophilins facilitate human papillomavirus type 16 infection. PLoS Pathog 5(7):e1000524. doi:10.1371/journal.ppat.1000524 Day PM, Baker CC, Lowy DR, Schiller JT (2004) Establishment of papillomavirus infection is enhanced by promyelocytic leukemia protein (PML) expression. Proc Nat Acad Sci USA 101(39):14252–14257. doi:10.1073/pnas.0404229101 Day PM, Lowy DR, Schiller JT (2008) Heparan sulfate-independent cell binding and infection with furin-precleaved papillomavirus capsids. J Virol 82(24):12565–12568. doi:10.1128/ JVI.01631-08 Schelhaas M, Shah B, Holzer M, Blattmann P, Kuhling L, Day PM, Schiller JT, Helenius A (2012) Entry of human papillomavirus type 16 by actin-dependent, clathrin- and lipid raftindependent endocytosis. PLoS Pathog 8(4):e1002657. doi: 10.1371/journal.ppat.1002657 Hindmarsh PL, Laimins LA (2007) Mechanisms regulating expression of the HPV 31 L1 and L2 capsid proteins and pseudovirion entry. Virol J 4:19. doi:10.1186/1743-422X-4-19 Bousarghin L, Touze A, Sizaret PY, Coursaget P (2003) Human papillomavirus types 16, 31, and 58 use different endocytosis pathways to enter cells. J Virol 77(6):3846–3850 Smith JL, Campos SK, Wandinger-Ness A, Ozbun MA (2008) Caveolin-1-dependent infectious entry of human papillomavirus type 31 in human keratinocytes proceeds to the endosomal pathway for pH-dependent uncoating. J Virol 82(19):9505– 9512. doi:10.1128/JVI.01014-08 Smith JL, Campos SK, Ozbun MA (2007) Human papillomavirus type 31 uses a caveolin 1- and dynamin 2-mediated entry pathway for infection of human keratinocytes. J Virol 81(18): 9922–9931 Day PM, Lowy DR, Schiller JT (2003) Papillomaviruses infect cells via a clathrin-dependent pathway. Virology 307(1):1–11 Laniosz V, Holthusen KA, Meneses PI (2008) Bovine papillomavirus type 1: from clathrin to caveolin. J Virol 82(13): 6288–6298. doi:10.1128/JVI.00569-08 Spoden G, Freitag K, Husmann M, Boller K, Sapp M, Lambert C, Florin L (2008) Clathrin- and caveolin-independent entry of human papillomavirus type 16–involvement of tetraspaninenriched microdomains (TEMs). PLoS ONE 3(10):e3313. doi: 10.1371/journal.pone.0003313

123

122. Bienkowska-Haba M, Williams C, Kim SM, Garcea RL, Sapp M (2012) Cyclophilins facilitate dissociation of the HPV16 capsid protein L1 from the L2/DNA complex following virus entry. J Virol. doi:10.1128/JVI.00980-12 123. Selinka HC, Giroglou T, Sapp M (2002) Analysis of the infectious entry pathway of human papillomavirus type 33 pseudovirions. Virology 299(2):279–287 124. Dabydeen SA, Meneses PI (2009) The role of NH4Cl and cysteine proteases in Human Papillomavirus type 16 infection. Virol J 6:109. doi:10.1186/1743-422X-6-109 125. Karanam B, Peng S, Li T, Buck C, Day PM, Roden RB (2010) Papillomavirus infection requires gamma secretase. J Virol 84(20):10661–10670. doi:10.1128/JVI.01081-10 126. Bergant Marusic M, Ozbun MA, Campos SK, Myers MP, Banks L (2012) Human papillomavirus L2 facilitates viral escape from late endosomes via sorting nexin 17. Traffic 13(3):455–467. doi: 10.1111/j.1600-0854.2011.01320.x 127. Campos SK, Ozbun MA (2009) Two highly conserved cysteine residues in HPV16 L2 form an intramolecular disulfide bond and are critical for infectivity in human keratinocytes. PLoS ONE 4(2):e4463. doi:10.1371/journal.pone.0004463 128. Sapp M, Kraus U, Volpers C, Snijders PJ, Walboomers JM, Streeck RE (1994) Analysis of type-restricted and cross-reactive epitopes on virus-like particles of human papillomavirus type 33 and in infected tissues using monoclonal antibodies to the major capsid protein. J Gen Virol 75(Pt 12):3375–3383 129. Schneider MA, Spoden GA, Florin L, Lambert C (2011) Identification of the dynein light chains required for human papillomavirus infection. Cell Microbiol 13(1):32–46. doi:10.1111/ j.1462-5822.2010.01515.x 130. Campos SK, Chapman JA, Deymier MJ, Bronnimann MP, Ozbun MA (2012) Opposing effects of bacitracin on human papillomavirus type 16 infection: enhancement of binding and entry and inhibition of endosomal penetration. J Virol 86(8): 4169–4181. doi:10.1128/JVI.05493-11 131. Karala AR, Ruddock LW (2010) Bacitracin is not a specific inhibitor of protein disulfide isomerase. FEBS J 277(11): 2454–2462. doi:10.1111/j.1742-4658.2010.07660.x 132. Kamper N, Day PM, Nowak T, Selinka HC, Florin L, Bolscher J, Hilbig L, Schiller JT, Sapp M (2006) A membrane-destabilizing peptide in capsid protein L2 is required for egress of papillomavirus genomes from endosomes. J Virol 80(2): 759–768. doi:10.1128/JVI.80.2.759-768.2006 133. Baker TS, Drak J, Bina M (1988) Reconstruction of the threedimensional structure of simian virus 40 and visualization of the chromatin core. Proc Nat Acad Sci USA 85(2):422–426 134. Rayment I, Baker TS, Caspar DL, Murakami WT (1982) Polyoma virus capsid structure at 22.5 A resolution. Nature 295(5845):110–115 135. Roth SD, Sapp M, Streeck RE, Selinka HC (2006) Characterization of neutralizing epitopes within the major capsid protein of human papillomavirus type 33. Virol J 3:83