Chapter Three - The Envelope Proteins of the Bunyavirales

https://doi.org/10.1016/bs.aivir.2017.02.002Get rights and content

Abstract

The Bunyavirales Order encompasses nine families of enveloped viruses containing a single-stranded negative-sense RNA genome divided into three segments. The small (S) and large (L) segments encode proteins participating in genome replication in the infected cell cytoplasm. The middle (M) segment encodes the viral glycoproteins Gn and Gc, which are derived from a precursor polyprotein by host cell proteases. Entry studies are available only for a few viruses in the Order, and in each case they were shown to enter cells via receptor-mediated endocytosis. The acidic endosomal pH triggers the fusion of the viral envelope with the membrane of an endosome. Structural studies on two members of this Order, the phleboviruses and the hantaviruses, have shown that the membrane fusion protein Gc displays a class II fusion protein fold and is homologous to its counterparts in flaviviruses and alphaviruses, which are positive-sense, single-stranded RNA viruses. We analyze here recent data on the structure and function of the structure of the phlebovirus Gc and hantavirus Gn and Gc glycoproteins, and extrapolate common features identified in the amino acid sequences to understand also the structure and function of their counterparts in other families of the Bunyavirales Order. Our analysis also identified clear structural homology between the hantavirus Gn and alphavirus E2 glycoproteins, which make a heterodimer with the corresponding fusion proteins Gc and E1, respectively, revealing that not only the fusion protein has been conserved across viral families.

Introduction

The Bunyavirales Order is one of the largest group of RNA viruses, including more than 350 members distributed among nine families: Feraviridae, Fimoviridae, Hantaviridae, Jonviridae, Nairoviridae, Peribunyaviridae, Phasmaviridae, Phenuiviridae and Tospoviridae (https://talk.ictvonline.org/taxonomy/). Four of the nine families contain viruses that cause disease in humans, including, among others: the highly pathogenic Crimean-Congo hemorrhagic fever virus (CCHFV) in the Nairoviridae family; the Rift Valley fever virus (RVFV), the severe fever with thrombocytopenia syndrome virus (SFTSV) and Toscana virus in the Phenuiviridae family (genus Phlebovirus); Lacrosse virus and Oropouche virus in the Peribunyaviridae family (genus orthobunyavirus); and several hantaviruses such as Hantaan virus, Sin Nombre virus, Andes virus (Hantaviridae family). Hantaviruses, which cause persistent infections in rodents or in insectivores and often spill over to humans where they can cause severe disease, are the only bunyaviruses that are not arboviruses (short for arthropod-borne viruses). All the others, including the plant viruses in the family Tospoviridae, are transmitted by insect vectors (mosquitoes, sand flies, thrips) or by tick vectors. Bunyaviruses thus infect a broad range of hosts: vertebrates (in particular, mammals), invertebrates, and plants. Among the newly discovered bunyaviruses are: Gouleako virus (family Phenuiviridae, genus Goukovirus), which is similar to the phleboviruses but infects only mosquitoes (Marklewitz et al., 2011) and is not transmitted to vertebrates or plants; the Jonchet (family Jonviridae), Ferak (family Feraviridae), and Herbert virus (family Peribunyaviridae, genus Herbevirus) groups, which have been isolated from mosquitoes in Côte d’Ivoire, Ghana, and Uganda and replicate efficiently in insect cells but not in mammalian cells (Marklewitz et al., 2013, Marklewitz et al., 2015); and phasmaviruses (Phasmaviridae family), which were discovered in arctic phantom midges (Ballinger et al., 2014). Many other bunyaviruses have been detected and sequenced using metagenomics approaches on arthropods (Li et al., 2015) and other invertebrate species (Shi et al., 2016a), but have not yet been isolated and characterized.

Bunyavirus particles are in general spherical, with a diameter of about 80–120 nm, and their morphology has been studied by a number of biochemical, biophysical, and electron microscopy studies (Battisti et al., 2011, Bowden et al., 2013, Freiberg et al., 2008, Hepojoki et al., 2010, Huiskonen et al., 2009, Huiskonen et al., 2010, Martin et al., 1985, Overby et al., 2008, Rusu et al., 2012, Salanueva et al., 2003, Sherman et al., 2009). Viruses in some of the families (like the hantaviruses) also exhibit elongated particles. The virion contains a negative- or ambisense single-stranded RNA genome divided into three segments, called S (small), M (medium), and L (large) (Fig. 1) (Elliott and Schmaljohn, 2013). The S and the L segments code for the proteins required for genome replication within a cell: the nucleocapsid protein (N), which encapsidates the genomic RNA into ribonucleoprotein complexes is encoded in the S segment, and the large RNA-dependent RNA-polymerase (or “L” protein), responsible for transcription and replication of the genomic RNA, in the L segment. Some genera code for a nonstructural protein (NSs) in the S segment. The L and S segments are very similar to the two segments of the Arenaviridae family of bisegmented ambisense RNA viruses, and very likely have the same origin, except that in arenaviruses the S gene also codes for the envelope protein (Buchmeier et al., 2013).

The M segment of the bunyaviruses codes for proteins required in particle budding and for entry into target cells, in the form of a polyprotein precursor of the envelope glycoproteins Gn and Gc. As discussed later, the M gene appears to be related to the structural genes of positive strand viruses of the Togaviridae (Kuhn, 2013) and Flaviviridae (Lindenbach et al., 2013) families, which appear to have evolved from a distant ancestor in common with the bunyavirus M gene. This finding is a further illustration of the mosaic nature of viral genomes, which are a patchwork of genes having different origins. Some bunyavirus families also encode one or several nonstructural proteins (NSm) in the M segment (Fig. 1). The individual glycoproteins Gn and Gc—which are the focus of this review—are generated by signalase cleavage in the ER, followed by additional processing depending on the family. Gn/Gc heterodimers associate in the ER and travel to the site of particle morphogenesis (Petterson, 1996). Budding was shown to occur in the Golgi apparatus, although some hantaviruses were reported to bud at the plasma membrane (Goldsmith et al., 1995). Gn/Gc heterodimers associate into larger complexes that interact laterally to drive budding of newly formed virions by inducing membrane curvature, resulting in closed particles with projections 5–10 nm long at the viral surface. There is no proteolytic maturation processing of the envelope proteins upon assembly in the Golgi, and the “priming” process necessary for activation by the endosomal acidic pH upon entering a new cell is not currently understood. This is an important difference with flaviviruses (Flaviviridae family) and alphaviruses (Togaviridae family), which have a similar envelope protein organization, but require cleavage of a companion, chaperone protein occupying the place of Gn in the precursor polyprotein of those viruses. In this respect, the bunyaviruses are more similar to Rubella virus (Togaviridae family), which also has a similar envelope protein organization with no maturation cleavage (Kuhn, 2013).

Section snippets

Bunyavirus Entry Into Cells

Cell entry has been studied for some bunyaviruses, showing in each case entry via receptor-mediated endocytosis, with release of the genomic RNA into the cytoplasm after fusion of the viral envelope with endosomal membranes. The nature of the cell receptors has not been identified for most viruses in the Order, with the best characterized being the hantaviruses. Pathogenic New York 1 and Sin Nombre hantaviruses were shown to use β3 integrins (αIIbβ3 or αVβ3) to infect endothelial cells (

Bunyavirus Gc Is a Class II Fusion Protein

Glycoprotein Gc from the RVFV phlebovirus was the first bunyavirus for which an X-ray structure was determined (Dessau and Modis, 2013), confirming predictions that it would exhibit a class II fusion fold (Garry and Garry, 2004). Later on, the crystal structures of Gc from the hantaviruses Hantaan virus (Guardado-Calvo et al., 2016) and Puumala virus (Willensky et al., 2016) were also determined and confirmed that they are also class II fusion proteins (see Box 1 for a description of this

The Target Membrane-Interacting Region

The fusion loop is a central piece in the mechanism of action of all viral fusion proteins because it is responsible for the initial insertion into target membranes. In class I fusion proteins it is generally at the N-terminus, and it is termed fusion “peptide” because it is not rigidly connected to the rest of the ectodomain and adopts a particular structure upon insertion into lipid bilayers (Apellaniz et al., 2014). In classes II and III, there are internal fusion loops connecting two β

pH-Sensing Mechanisms

The class II fusion machinery is activated by the acidic environment of the endosomes encountered during cell entry. It has been proposed that the different protonation states of strategically located histidine residues of the protein can be at the core of the triggering mechanism leading to fusion (Kampmann et al., 2006). The pK of a histidine side chain exposed to solvent is around 6.2, but an unexposed histidine side chain will have a pK value that depends on its immediate environment within

Lipid Sensing

Upon acid pH activation, the TMIR becomes exposed and inserts into the target membrane. To ensure that viral fusion takes place in the proper compartment, the viral fusion proteins have evolved mechanisms to sense the lipid composition of the membranes and to only fuse efficiently with membranes having a defined lipid composition. The lipid dependence in class II proteins has been studied for alphaviruses and flaviviruses. In alphaviruses, the fusion reaction depends on the presence of

The Stem Region

The stem is a region located between the end of domain III and the transmembrane domain (Box 1). It plays a crucial role in fusion by allowing completion of the postfusion hairpin conformation. In spite of the many structures of class II fusion proteins in their postfusion conformation, structural information for the stem region is only available for Rubella virus (DuBois et al., 2013) and the cellular fusion protein EFF-1 (Perez-Vargas et al., 2014) but was absent or not resolved in all the

Newly Identified pGc-Like Envelope Proteins

Taking advantage of the documented sequence similarity between phleboviruses, phasmavirus and goukovirus (Shi et al., 2016a), we identified a sequence motif conserved in the three groups (C-x(17)-[ST]-[RKH]-[RK]-C-x(3)-[GAS]-x-C-x(3,4)-C-x(10,11)-[ED]-x(12,14)-C). We found that this motif is present not only in Gc from the bunyaviruses used to generate it but also in multiple sequences of cellular genes from nematodes in the order Rhabditia, particularly from the genus Caenorhabditis and

Projections for Gc From Other Members of the Bunyavirales Order

Amino acid sequence alignments of hGc with its counterparts in the other families of the Bunyavirales Order allowed the identification of a motif present in the families Nairoviridae, Peribunyaviridae, and Tospoviridae but not in the family Phenuiviridae (Guardado-Calvo et al., 2016). This motif (C-x-[GE]-x(1,2)-C-x(20,24)-W-x-C-x(4)-C-x(5)-G-x(0,2)-C-x(2,4)-C-x(23,27)-C-x(8,9)-C-x(73,113)-C-x(7,12)-C), is based on the relative position of 10 conserved cysteine residues forming five disulfide

Hantavirus Gn Is Homologous to Alphavirus E2

Class II viral fusion proteins function together with an accompanying protein, which is released first from the same polyprotein precursor and can interact cotranslationally and in cis with the fusion protein, helping avoid premature activation in the secretory pathway. The evolution of both proteins is closely related because of structural and functional constraints. The glycoprotein complex in Alphaviruses is produced as an immature p62-E1 heterodimer, which trimerises to form the viral

Discussion

We have focused this review to a large extent on the analysis of the implications of recently reported envelope protein structural and functional data for understanding the biology of the bunyaviruses. These data include the structures of hGc from two Hantaviruses, Hantaan virus (Guardado-Calvo et al., 2016), and Puumala virus (Willensky et al., 2016), hGn from Puumala virus (Li et al., 2016), and pGc from SFTSV (Halldorsson et al., 2016). In addition, we have identified specific amino acid

References (105)

  • T. Kampmann et al.

    The role of histidine residues in low-pH-mediated viral membrane fusion

    Structure

    (2006)
  • J.F. Koellhoffer et al.

    Structural characterization of the glycoprotein GP2 core domain from the CAS virus, a novel arenavirus-like species

    J. Mol. Biol.

    (2014)
  • S. Li et al.

    A molecular-level account of the antigenic hantaviral surface

    Cell Rep.

    (2016)
  • P.Y. Lozach et al.

    Entry of bunyaviruses into mammalian cells

    Cell Host Microbe

    (2010)
  • P.Y. Lozach et al.

    DC-SIGN as a receptor for phleboviruses

    Cell Host Microbe

    (2011)
  • R.B. Medeiros et al.

    Immunoprecipitation of a 50-kDa protein: a candidate receptor component for tomato spotted wilt tospovirus (Bunyaviridae) in its main vector, Frankliniella occidentalis

    Virus Res.

    (2000)
  • Y. Modis

    Relating structure to evolution in class II viral membrane fusion proteins

    Curr. Opin. Virol.

    (2014)
  • W.A. Mohler et al.

    The type I membrane protein EFF-1 is essential for developmental cell fusion

    Dev. Cell

    (2002)
  • J. Perez-Vargas et al.

    Structural basis of eukaryotic cell-cell fusion

    Cell

    (2014)
  • M.L. Plassmeyer et al.

    Mutagenesis of the La Crosse Virus glycoprotein supports a role for Gc (1066-1087) as the fusion peptide

    Virology

    (2007)
  • H.N. Ramanathan et al.

    New and old world hantaviruses differentially utilize host cytoskeletal components during their life cycles

    Virology

    (2008)
  • J. Ravantti et al.

    Automatic comparison and classification of protein structures

    J. Struct. Biol.

    (2013)
  • F.A. Rey

    Virus/host interactions: a strong force driving the diversification of cellular organisms

    Curr. Opin. Microbiol.

    (2011)
  • R.I. Santos et al.

    Oropouche virus entry into HeLa cells involves clathrin and requires endosomal acidification

    Virus Res.

    (2008)
  • M.B. Sherman et al.

    Single-particle cryo-electron microscopy of Rift Valley fever virus

    Virology

    (2009)
  • M. Backovic et al.

    Class III viral membrane fusion proteins

    Adv. Exp. Med. Biol.

    (2011)
  • M.J. Ballinger et al.

    Discovery and evolution of bunyavirids in arctic phantom midges and ancient bunyavirid-like sequences in insect genomes

    J. Virol.

    (2014)
  • G.P. Barriga et al.

    Inhibition of the hantavirus fusion process by predicted domain III and stem peptides from glycoprotein Gc

    PLoS Negl. Trop. Dis.

    (2016)
  • A.J. Battisti et al.

    Structural studies of Hantaan virus

    J. Virol.

    (2011)
  • T.A. Bowden et al.

    Orthobunyavirus ultrastructure and the curious tripodal glycoprotein spike

    PLoS Pathog.

    (2013)
  • S. Bressanelli et al.

    Structure of a flavivirus envelope glycoprotein in its low-pH-induced membrane fusion conformation

    EMBO J.

    (2004)
  • M.J. Buchmeier et al.

    Fields virology. Chapter 43—arenaviridae

  • L.H. Chao et al.

    Sequential conformational rearrangements in flavivirus membrane fusion

    Elife

    (2014)
  • M. Crispin et al.

    Uukuniemi Phlebovirus assembly and secretion leave a functional imprint on the virion glycome

    J. Virol.

    (2014)
  • S.M. De Boer et al.

    Heparan sulfate facilitates Rift Valley fever virus entry into the cell

    J. Virol.

    (2012)
  • M. Dessau et al.

    Crystal structure of glycoprotein C from Rift Valley fever virus

    Proc. Natl. Acad. Sci. U.S.A.

    (2013)
  • R.M. Dubois et al.

    Functional and evolutionary insight from the crystal structure of rubella virus protein E1

    Nature

    (2013)
  • K. El Omari et al.

    Unexpected structure for the N-terminal domain of hepatitis C virus envelope glycoprotein E1

    Nat. Commun.

    (2014)
  • R.M. Elliott et al.

    Fields Virology. Chapter 42—Bunyaviridae

    (2013)
  • J. Fedry et al.

    The ancient gamete fusogen HAP2 is a eukaryotic class II fusion protein

    Cell

    (2017)
  • C.M. Filone et al.

    Rift valley fever virus infection of human cells and insect hosts is promoted by protein kinase C epsilon

    PLoS One

    (2010)
  • A.N. Freiberg et al.

    Three-dimensional organization of Rift Valley fever virus revealed by cryoelectron tomography

    J. Virol.

    (2008)
  • R. Fritz et al.

    Identification of specific histidines as pH sensors in flavivirus membrane fusion

    J. Cell Biol.

    (2008)
  • C.E. Garry et al.

    Proteomics computational analyses suggest that the carboxyl terminal glycoproteins of Bunyaviruses are class II viral fusion protein (beta-penetrenes)

    Theor. Biol. Med. Model.

    (2004)
  • I.N. Gavrilovskaya et al.

    beta3 Integrins mediate the cellular entry of hantaviruses that cause respiratory failure

    Proc. Natl. Acad. Sci. U.S.A.

    (1998)
  • C.S. Goldsmith et al.

    Ultrastructural characteristics of Sin Nombre virus, causative agent of hantavirus pulmonary syndrome

    Arch. Virol.

    (1995)
  • P. Guardado-Calvo et al.

    Mechanistic insight into Bunyavirus-induced membrane fusion from structure-function analyses of the hantavirus envelope glycoprotein Gc

    PLoS Pathog.

    (2016)
  • S. Halldorsson et al.

    Structure of a phleboviral envelope glycoprotein reveals a consolidated model of membrane fusion

    Proc. Natl. Acad. Sci. U.S.A.

    (2016)
  • B. Harmon et al.

    Rift Valley fever virus strain MP-12 enters mammalian host cells via caveola-mediated endocytosis

    J. Virol.

    (2012)
  • E.E. Heldwein et al.

    Crystal structure of glycoprotein B from herpes simplex virus 1

    Science

    (2006)
  • Cited by (73)

    • Cryo-EM structure of glycoprotein C from Crimean-Congo hemorrhagic fever virus

      2022, Virologica Sinica
      Citation Excerpt :

      CCHFV contains a tripartite genome composed of small (S), medium (M) and large (L) negative-stranded RNA segments (Bente et al., 2013; Zivcec et al., 2016). The M segment encodes the glycoprotein precursor that is processed by cellular proteases into several proteins, including the N-terminal mucin-like domain (MLD), two non-structural proteins (GP38 and NSm) and two structural glycoproteins (Gn and Gc) (Fig. 1A) (Bergeron et al., 2007; Guardado-Calvo and Rey, 2017). Glycoproteins Gn and Gc form a lattice of putative tetrameric projections on the viral membrane surface according to recent cryo-electron tomography studies of a nairovirus (Hazara virus) (Punch et al., 2018).

    • Bunyavirus

      2022, Encyclopedia of Infection and Immunity
    • The surface glycoproteins of hantaviruses

      2021, Current Opinion in Virology
    View all citing articles on Scopus
    View full text