Molecular characterization of new archaeal viruses from ...

64
INSTITUTE OF BIOTECHNOLOGY AND DIVISION OF GENERAL MICROBIOLOGY DEPARTMENT OF BIOSCIENCES FACULTY OF BIOLOGICAL AND ENVIRONMENTAL SCIENCES DOCTORAL PROGRAMME IN MICROBIOLOGY AND BIOTECHNOLOGY UNIVERSITY OF HELSINKI Molecular Characterization of New Archaeal Viruses from High Salinity Environments TATIANA DEMINA DISSERTATIONES SCHOLA DOCTORALIS SCIENTIAE CIRCUMIECTALIS, ALIMENTARIAE, BIOLOGICAE. UNIVERSITATIS HELSINKIENSIS 26/2016

Transcript of Molecular characterization of new archaeal viruses from ...

INSTITUTE OF BIOTECHNOLOGY AND DIVISION OF GENERAL MICROBIOLOGYDEPARTMENT OF BIOSCIENCESFACULTY OF BIOLOGICAL AND ENVIRONMENTAL SCIENCESDOCTORAL PROGRAMME IN MICROBIOLOGY AND BIOTECHNOLOGY UNIVERSITY OF HELSINKI

Molecular Characterization of New Archaeal Viruses from High Salinity Environments

TATIANA DEMINA

dissertationes schola doctoralis scientiae circumiectalis, alimentariae, biologicae. universitatis helsinkiensis 26/2016

26/2016

Helsinki 2016 ISSN 2342-5423 ISBN 978-951-51-2706-8

Recent Publications in this Series

9/2016 Satu OlkkolaAntimicrobial Resistance and Its Mechanisms among Campylobacter coli and Campylobacter upsaliensis with a Special Focus on Streptomycin10/2016 Windi Indra MuziasariImpact of Fish Farming on Antibiotic Resistome and Mobile Elements in Baltic Sea Sediment11/2016 Kari Kylä-NikkiläGenetic Engineering of Lactic Acid Bacteria to Produce Optically Pure Lactic Acid and to Develop a Novel Cell Immobilization Method Suitable for Industrial Fermentations12/2016 Jane Etegeneng Besong epse NdikaMolecular Insights into a Putative Potyvirus RNA Encapsidation Pathway and Potyvirus Particles as Enzyme Nano-Carriers13/2016 Lijuan YanBacterial Community Dynamics and Perennial Crop Growth in Motor Oil-Contaminated Soil in a Boreal Climate14/2016 Pia RasinkangasNew Insights into the Biogenesis of Lactobacillus rhamnosus GG Pili and the in vivo Effects of Pili15/2016 Johanna RytiojaEnzymatic Plant Cell Wall Degradation by the White Rot Fungus Dichomitus squalens16/2016 Elli KoskelaGenetic and Environmental Control of Flowering in Wild and Cultivated Strawberries 17/2016 Riikka KylväjäStaphylococcus aureus and Lactobacillus crispatus: Adhesive Characteristics of Two Gram-Positive Bacterial Species18/2016 Hanna AarnosPhotochemical Transformation of Dissolved Organic Matter in Aquatic Environment – From a Boreal Lake and Baltic Sea to Global Coastal Ocean19/2016 Riikka PuntilaTrophic Interactions and Impacts of Non-Indigenous Species in the Baltic Sea Coastal Ecosystems20/2016 Jaana KuuskeriGenomics and Systematics of the White-Rot Fungus Phlebia radiata: Special Emphasis on Wood-Promoted Transcriptome and Proteome21/2016 Qiao ShiSynthesis and Structural Characterization of Glucooligosaccharides and Dextran from Weissella confusa Dextransucrases22/2016 Tapani JokiniemiEnergy Efficiency in Grain Preservation23/2016 Outi NyholmVirulence Variety and Hybrid Strains of Diarrheagenic Escherichia coli in Finland and Burkina Faso24/2016 Anu RiikonenSome Ecosystem Service Aspects of Young Street Tree Plantings25/2016 Jing ChengThe Development of Intestinal Microbiota in Childhood and Host-Microbe Interactions in Pediatric Celiac Diseases

YE

BT

AT

IAN

A D

EM

INA

Molecu

lar Ch

aracterization of N

ew A

rchaeal V

iruses from

High

Salin

ity En

vironm

ents

Molecular characterization of new archaealviruses from high salinity environments

Tatiana A. Demina

Institute of Biotechnology andDivision of General Microbiology

Department of BiosciencesFaculty of Biological and Environmental Sciences and

Viikki Doctoral Programme in Molecular Biosciences andDoctoral Programme in Microbiology and Biotechnology

University of Helsinki

ACADEMIC DISSERTATION

To be presented, with the permission of the Faculty of Biological andEnvironmental Sciences of the University of Helsinki, for public examination

in lecture room 511 of Info Center Korona, Viikinkaari 11, Helsinki,on 16th of December 2016, at 12 noon.

Helsinki 2016

Supervisors

Docent Hanna M. OksanenInstitute of Biotechnology andDepartment of BiosciencesUniversity of HelsinkiFinland

Docent Maija K. PietiläDepartment of Food and EnvironmentalSciencesUniversity of HelsinkiFinland

PhD Nina S. AtanasovaInstitute of Biotechnology andDepartment of BiosciencesUniversity of HelsinkiFinland

Thesis advisory committee

Academy ProfessorDennis H. BamfordInstitute of Biotechnology andDepartment of BiosciencesUniversity of HelsinkiFinland

Docent Sari TimonenDepartment of Food and EnvironmentalSciencesUniversity of HelsinkiFinland

©Tatiana Demina

Cover image: Tatiana Demina

ISBN 978-951-51-2706-8 (paperback)ISBN 978-951-51-2707-5 (PDF)ISSN 2342-5423 (print)ISSN 2342-5431 (online)

Hansaprint

Helsinki 2016

Reviewers

Professor Kenneth StedmanDepartment of BiologyPortland State UniversityU.S.A.

Associate Professor Josefa AntónDepartment of Physiology, Genetics, andMicrobiologyUniversity of AlicanteSpain

Opponent

Professor Jarkko HantulaNatural Resources Institute (LUKE)Finland

Custos

Academy ProfessorDennis H. BamfordInstitute of Biotechnology andDepartment of BiosciencesUniversity of HelsinkiFinland

“Science is the poetry of reality”

Richard Dawkins

i

List of original publicationsThis thesis is based on the following publications referred to by their Roman numerals:

I. Atanasova, N.S., Demina, T.A., Buivydas, A., Bamford, D.H., Oksanen, H.M.,2015. Archaeal viruses multiply: temporal screening in a solar saltern. Viruses 7,4, 1902-1926.

II. Demina, T.A., Pietilä, M.K., Svirskaite, J., Ravantti, J.J., Atanasova, N.S.,Bamford, D.H., Oksanen, H.M., 2016. Archaeal Haloarcula californiaeicosahedral virus 1 highlights conserved elements in icosahedral membrane-containing DNA viruses from extreme environments. mBio 7, 4.

III. Demina, T.A.*, Atanasova, N.S.*, Pietilä, M.K., Oksanen, H.M., Bamford, D.H.Vesicle-like virion of Haloarcula hispanica pleomorphic virus 3 preserves highinfectivity in saturated salt. Virology 499, 40-51. *These authors contributedequally.

Doctoral candidate’s contribution:

In I, TD participated in performing experiments, data analysis and handling, and writing themanuscript.

In II, TD contributed to the study design, acquisition of data, its analysis and interpretation,as well as drafring and editing the manuscript.

In III, TD was involved in project design, performing experiments, processing of obtaineddata, and writing the manuscript.

ii

AbbreviationsATP adenosine triphosphateATPase adenosine triphosphatase3-D three-dimensionalacc. no. accession numberBLAST Basic Local Alignment Search Toolbp base pairBREX bacteriophage exclusioncas clustered regularly interspaced short palindromic repeats-associated

proteinsCRISPR clustered regularly interspaced short palindromic repeatscryo-EM cryo-electron microscopyDNA deoxyribonucleic acidds double-strandedICTV International Committee on Taxonomy of VirusesITR inverted terminal repeatMCP major capsid proteinMEGA Molecular Evolutionary Genetics AnalysisMUSCLE Multiple Sequence Comparison by Log-Expectationn.r. non-redundantnt nucleotideORF open reading framep.i. post infectionPCR polymerase chain reactionPFU plaque forming unitRNA ribonucleic acidrRNA ribosomal RNASDS-PAGE sodium dodecyl sulfate polyacrylamide gel electrophoresisss single-strandedTEM transmission electron microscopyTLC thin layer chromatographytRNA transfer RNAVLP virus-like particleVP virion protein

iii

Abbreviated virus namesBTV Bluetongue virusEHP-1 environmental halophage 1EHP-12 environmental halophage 12EHP-6 environmental halophage 6HATV-1 Haloarcula head-tailed virus 1HATV-2 Haloarcula head-tailed virus 2HCIV-1 Haloarcula californiae icosahedral virus 1HCTV-1 Haloarcula californiae head-tailed virus 1HCTV-2 Haloarcula californiae head-tailed virus 2HCTV-5 Haloarcula californiae head-tailed virus 5HGPV-1 Halogeometricum pleomorphic virus 1HGTV-1 Halogeometricum head-tailed virus 1HHIV-2 Haloarcula hispanica icosahedral virus 2HHPV-1 Haloarcula hispanica pleomorphic virus 1HHPV-2 Haloarcula hispanica pleomorphic virus 2HHPV3 Haloarcula hispanica pleomorphic virus 3HHTV-1 Haloarcula hispanica head-tailed virus 1HHTV-2 Haloarcula hispanica head-tailed virus 2HJTV-1 Haloarcula japonica head-tailed virus 1HJTV-2 Haloarcula japonica head-tailed virus 2HRPV-1 Halorubrum pleomorphic virus 1HRPV-2 Halorubrum pleomorphic virus 2HRPV-3 Halorubrum pleomorphic virus 3HRPV-6 Halorubrum pleomorphic virus 6HRTV-1 Halorubrum head-tailed virus 1HRTV-10 Halorubrum head-tailed virus 10HRTV-11 Halorubrum head-tailed virus 11HRTV-12 Halorubrum head-tailed virus 12HRTV-2 Halorubrum head-tailed virus 2HRTV-3 Halorubrum head-tailed virus 3HRTV-4 Halorubrum head-tailed virus 4HRTV-5 Halorubrum head-tailed virus 5HRTV-6 Halorubrum head-tailed virus 6HRTV-7 Halorubrum head-tailed virus 7HRTV-8 Halorubrum head-tailed virus 8HRTV-9 Halorubrum head-tailed virus 9HSTV-1 Haloarcula sinaiiensis head-tailed virus 1HSTV-2 Halorubrum sodomense head-tailed virus 2HSTV-3 Halorubrum sodomense head-tailed virus 3HVTV-1 Haloarcula vallismortis head-tailed virus 1HVTV-2 Haloarcula vallismortis head-tailed virus 2NHV-1 nanohaloarchaeal virus 1OLPV Organic Lake phycodnavirus

iv

OLV Organic Lake virophagePBCV-1 Paramecium bursaria chlorella virus 1SCTP-1 Salicola head-tailed phage 1SCTP-2 Salicola head-tailed phage 2SCTP-3 Salicola head-tailed phage 3SIRV2 Sulfolobus islandicus rod-shaped virus 2SSIP-1 Salisaeta icosahedral phage 1STIV Sulfolobus turreted icosahedral virus

v

SummaryViruses are important ecological and biological actors and one of the major evolutionaryforces that cells encounter. In contrast to their overwhelming genetic diversity, all knownvirus isolates display a limited number of morphotypes. It has been suggested that due tostructural constraints and limitations in protein fold space, virus morphological diversity isrestricted to certain possible architectures. Thus, the huge and complex virosphere can beordered by grouping viruses into a reasonably small number of structure-based lineages. Totest this hypothesis, more virus-host systems have to be isolated and characterized.Hypersaline environments, where archaea tend to dominate, are especially rich in viruses.Still very little is known about archaeal viruses in general, and halophilic archaeal viruses inparticular. This thesis is devoted to the isolation and characterization of new archaeal virusesfrom high salinity environments.

Thirty seven haloarchaeal viruses were obtained from a solar saltern in Samut Sakhon(Thailand), significantly increasing the total number of currently available haloarchaealvirus isolates. The majority of the isolated viruses had tailed icosahedral morphology, butpleomorphic and tailless icosahedral morphotypes were also found. When the infectivity ofthe isolated viruses was tested against haloarchaeal strains isolated from the same locationor distant hypersaline environments, numerous interactions were observed across samplesand host genera. Myoviruses and Halorubrum strains were the two components thatcontributed to ~75% of the observed interactions. From the isolated viruses, two werecharacterized in molecular detail: Haloarcula californiae icosahedral virus 1 (HCIV-1) andHaloarcula hispanica pleomorphic virus 3 (HHPV3),

HCIV-1 was isolated on Haloarcula californiae, but it could also infect three otherhaloarchaeal strains. The virus maintains its infectivity under high and low saltconcentrations, showing adaptability to environmental changes. HCIV-1 infection results incell lysis, which was demonstrated by traditional metrics, as well as by on-line monitoringof physiological changes in infected cells. The HCIV-1 virion is a tailless icosahedron of ~70nm in diameter, and it has an inner lipid membrane. The major virion components areproteins, lipids selectively acquired from the host, and a linear double-stranded (ds) DNAmolecule of 31,314 bp in length. The virus genome is predicted to contain 47 open readingframes (ORFs), the majority of which display similarities to the genes/ORFs ofalphasphaerolipoviruses SH1, PH1, and Haloarcula hispanica icosahedral virus 2 (HHIV-2). The genomes of these four viruses share gene synteny and demonstrate high conservationin the genes encoding core virion elements: major capsid proteins (MCPs) and putativeadenosine triphosphatases (ATPases). On the contrary, genes coding for (putative) vertexcomplex proteins share the lowest similarity among genes encoding structural proteins.HCIV-1 virion assembly model is predicted to mimic the one described for HHIV-2. TheInternational Committee on Taxonomy of Viruses (ICTV) has approved the proposal toassign HCIV-1 to the Alphasphaerolipovirus genus in Sphaerolipoviridae family, whichcomprises tailless icosahedral inner membrane-containing dsDNA viruses isolated fromextreme environments. In terms of structure-based virus classification, HCIV-1 conservedelements responsible for virion organization determine its place in the PRD1-adenovirus

vi

structural viral lineage, which unites archaeal, bacterial, and eukaryotic viruses that aredistributed world-wide.

HHPV3 was isolated on Har. hispanica, and its host range is limited to the isolationstrain. The HHPV3 life cycle is non-lytic, but virus infection results in significant host growthretardation. The virus withstands extremely high NaCl concentrations (3-5.5 M NaCl), butdemonstrates a peculiar, irreversible, pH-dependent loss of infectivity at 1.5 M NaCl. TheHHPV3 virion is a pleomorphic lipid vesicle of ~50 nm in diameter that contains only twomajor structural protein species: a membrane associated protein and a spike protein.HHPV3 acquires lipids non-selectively from the host membranes. The virus genome is acircular dsDNA molecule of 11,648 bp in length, which contains 17 predicted ORFs, includinga set of ORFs/genes characteristic to pleolipoviruses. HHPV3-related proviruses were foundin the chromosomes of many haloarchaeal strains, highlighting the natural abundance ofpleolipoviruses. We propose that HHPV3 is a new member in Betapleolipovirus genus ofthe Pleolipoviridae family. Whether pleolipoviruses form a distinct viral structural lineageremains to be seen.

The viruses isolated here provide a rich source for more detailed future studies, whichmay reveal valuable information on global virus diversity and evolutionary relationsbetween different viral groups. Accumulation of such information helps in resolving virusclassification, which is currently an extremely challenging task. Markedly, the 37 virusisolates obtained in this study display the previously known morphotypes. Moreover, HCIV-1 clearly resembles the other haloarchaeal tailless icosahedral viruses, while HHPV3 belongsto the group of pleolipoviruses. Notably, both viral groups are distributed world-wide. Thiswork supports the observation that highly similar virion architectures can be sampled allover the biosphere, highlighting limited morphological virus diversity. Further studies areneeded to portray true viral diversity and to verify whether all viruses can be efficientlyclassified into structure-based lineages.

vii

Table of contentsList of original publications ...................................................................................................... iAbbreviations ........................................................................................................................... iiAbbreviated virus names ....................................................................................................... iiiSummary .................................................................................................................................. vTable of contents .................................................................................................................... vii1. Introduction .......................................................................................................................... 1

1.1 Hypersaline environments and halophiles ..................................................................... 11.2 Viruses ............................................................................................................................. 31.3 Current virus classification ............................................................................................. 41.4 Viral structural lineages .................................................................................................. 51.5 Haloviruses ...................................................................................................................... 7

1.5.1 Eukaryotic and bacterial haloviruses ........................................................................ 71.5.2 Archaeal haloviruses ................................................................................................ 11

1.5.2.1 Tailed icosahedral viruses ................................................................................ 131.5.2.2 Spindle-shaped virus His1 ................................................................................ 141.5.2.3 Tailless icosahedral viruses .............................................................................. 151.5.2.4 Pleomorphic viruses ......................................................................................... 18

1.6 Virus-host interactions in hypersaline environments .................................................. 212. Aims of the Study ............................................................................................................... 243. Materials and Methods ...................................................................................................... 254. Results and Discussion ...................................................................................................... 27

4.1 New archaeal viruses isolated from high salinity ......................................................... 274.2 Virus-host interactions ................................................................................................ 284.3 Molecular characterization of HCIV-1 ......................................................................... 29

4.3.1 HCIV-1 infection cycle ............................................................................................ 294.3.2 HCIV-1 virion organization ................................................................................... 30

4.4 Molecular characterization of HHPV-3 ....................................................................... 324.4.1 HHPV3 infection cycle ........................................................................................... 324.4.2 HHPV3 infectivity at different salinities ............................................................... 334.4.3 HHPV3 virion organization ................................................................................... 33

5. Conclusions and Future Prospects .................................................................................... 376. Acknowledgements ............................................................................................................ 397. References .......................................................................................................................... 41

1

1. Introduction1.1 Hypersaline environments and halophilesHypersaline environments, artificial and natural, are found all over our planet. Theexamples of such environments include saline lakes and soils, solar salterns, deep-seabrines, salt mines, halite deposits, and salty food products (Boetius and Joye, 2009).Hypersaline environments may be categorized into thalassohaline (also called thalassic) orathalassohaline (athalassic). Thalassohaline waters, such as solar salterns, haveapproximately the same ratio of ions as in seawater. On the contrary, athalassohalineenvironments have salts in proportions different from seawater, e.g. the Dead Sea is a low-Na+, high-Mg2+, and high-Ca2+ chloride brine (DasSarma and DasSarma, 2012, Oren, 2015).Some hypersaline environments are actually polyextreme, i.e. characterized by otherextreme conditions in addition to the extreme salinity. For example, halophiles thriving inthe Dead Sea are exposed not only to the saturated salt concentrations, but also to acidic pHand intensive UV radiation (Oren, 2010). Despite harsh conditions, the representatives ofall three domains of cellular life, Eukarya, Bacteria, and Archaea, as well as their virusesinhabit hypersaline environments (Boetius and Joye, 2009, Oren, 2015). The followingchapters will be devoted to halophilic viruses, while here, I will focus on halophilic cellularorganisms.

Halophilic, or halophile, literally means “salt-loving”. Halophiles may be classified intoslight, moderate, and extreme, which require 1-3 % (0.2-0.5 M) NaCl, 3-15% (0.5-2.5 M)NaCl, or 15-30% (2.5-5.2 M) NaCl, respectively, for the optimal growth (Kushner andKamekura, 1988). This categorization is, however, nominal, and more widely, halophilerefers to an organism that inhabits an environment with salinity higher than in seawater(>3.5 % total salts). To survive in high salinity, halophiles use two main strategies, “high-salt-in” and “organic-solutes-in” (DasSarma and DasSarma, 2012, Edbeib et al., 2016). The“high-salt-in” strategy implies accumulation of ions inside the cells in concentration equalto that in the environment, while “organic-solutes-in” refers to the accumulation (de novosynthesis or uptake from the environment) of compatible solutes (osmolytes), such as aminoacids and sugars (Edbeib et al., 2016).

Hypersaline environments are dominated by prokaryotes, but eukaryotes also thrive atmoderate and even extreme salinities (Boetius and Joye, 2009, DasSarma and DasSarma,2012, Oren, 2014a, Ventosa et al., 2014). Dense populations of green algae, e.g. Dunaliellasalina, are main food sources for brine shrimp Artemia and larvae of brine flies. Otherinvertebrates are also known to survive in hypersaline environments: rotifers, flat worms,copepods, ostracods. In addition, halophytic plants, e.g. Laguncularia mangrove, can growin moderately saline soils. Even some vertebrates, e.g. Tilapia species, are observed tosurvive in moderately high salinity. It is also worth mentioning that some hypersaline pondsare important feeding sources for various birds, including flamingo. Among othereukaryotes inhabiting salty ponds are diatoms, protozoa (ciliate), and fungi.

Moderately or extremely halophilic bacteria are classified to at least eight differentphyla, among which Actinobacteria, Firmicutes, and Proteobacteria include most genera(Ventosa et al., 2012). The group is very heterogeneous, and contains Gram-positive/Gram-negative, aerobic/anaerobic, heterotrophic/phototrophic bacteria of different shapes (cocci,

2

spirals, etc.) (Ventosa et al., 2012). Microbial mats formed by halophilic bacteria areobserved in hypersaline waters. Unicellular cyanobacteria constitute the upper level of themat, while filamentous cyanobacteria form the second layer. Anaerobic phototrophicbacteria (green and purple bacteria) grow beneath cyanobacterial layers, and strictlyanaerobic bacteria thrive in the bottom layers of the mat. In addition, methanogenic archaeaare also found in these bottom layers (Bolhuis et al., 2014, DasSarma and DasSarma, 2012).

In the saltiest ponds, halophilic archaea dominate, although some bacteria, e.g.Salinibacter ruber, may also be abundant (Ventosa et al., 2014, Zhaxybayeva et al., 2013).Halophilic archaea are taxonomically classified to the phylum Euryarchaeota, classesHalobacteria (extreme halophiles) and Methanomicrobia (halophilic or halotolerantmethanogens) (Oren, 2014b, Ventosa et al., 2012). In addition, a yet uncultivable lineage ofsmall-sized (~0.6 μm) archaea, called “Nanohaloarchaea”, seems to be globally distributedin hypersaline environments (Ghai et al., 2011, Grant et al., 1999, Narasingarao et al., 2012,Oh et al., 2010).

The majority of halophilic archaea belong to the class Halobacteria, which includes asingle order Halobacteriales with a single family Halobacteriaceae. By the year 2014, thefamily contained 47 genera and 165 species, with Halorubrum, Haloferax, and Haloarculabeing the largest genera (Oren, 2014b). Recently, it has been suggested to devide the classHalobacteria into three orders, Halobacteriales, Haloferacales, and Natrialbales (Gupta etal., 2015). The order Haloferacales has been proposed to include two families,Haloferacaceae and Halorubraceae, while the order Halobacteriales has been proposed toinclude three families, Halobacteriaceae, Haloarculaceae, and Halococcaceae (Gupta et al.,2016). The members of the Halobacteria are “the halophiles par excellence” (Oren, 2014b),meaning that many species may grow even in almost saturated brines. Haloarchaea displayvarious cell shapes: cocci, rods, disks, triangles (Haloarcula japonica), square(Haloquadratum walsbyi), and pleomorphic (DasSarma and DasSarma, 2012). Theproteins of haloarchaea are negatively charged, having an excess of acidic to basic aminoacids, which retains functionality at low water activity (DasSarma and DasSarma, 2015).Haloarchaeal lipids are ether-linked, as in other Archaea, and polar lipids profiles are usedfor chemotaxonomic differentiation (Ventosa et al., 2012). Some haloarchaea, e.g.Haloquadratum walsbyi and Halobacterium salinarum, may produce gas-filled rigidproteinaceous vesicles, which enables them to migrate vertically in water to the moreoxygenated surface layers (Pfeifer, 2015). Typically, the members of the Halobacteria areaerobic heterotrophs, but some species may have specific nutritional needs and have diversemetabolic pathways (Andrei et al., 2012). The species of the genera Haloferax andHaloarcula are usually easily grown in the laboratory conditions (Oren, 2002), while others,e.g. Halosimplex carlsbadense (Vreeland et al., 2002), have complex nutritional demands.Some haloarchaea may degrade aromic compounds, indicating a potential use inbioremediation of oil-polluted hypersaline environments (Al-Mailem et al., 2010). Somemembers of Halobacteria, e.g. Hlb. salinarum, are photoheterotrophs and use retinal-basedpigments to adsorb light energy (Andrei et al., 2012). These pink-pigmented archaea,together with Salinibacter ruber and Dunaliella salina, are responsible for unusual pink orred colors of highly saline ponds. Noticeably, a few halophilic archaea are polyextremophilic,as they grow at alkaline pH and high temperatures (Bowers and Wiegel, 2011). Interestingly,

3

due to their remarkable ability to survive extreme desiccation, radiation, and starvation,halophilic archaea serve as model organisms in ground-based simulations, as well as in outerspace experiments, to explore whether microbial life could survive in space (Leuko et al.,2014, Stan-Lotter and Fendrihan, 2015).

1.2 VirusesSince the discovery of viruses in the 19th century, the concept of virus has changeddramatically from an infectious particle to the most abundant biological entity, importantecological actor, and a major factor in the evolution of cellular life (Dolja and Krupovič, 2013,Forterre and Prangishvili, 2013, Koonin and Dolja, 2013, Pennazio, 2003, Pradeu et al.,2016). Moreover, virus research has challenged our perception of life, i.e. its definition,boundaries, and origin (Kostyrka, 2016).

Many specific characteristics that were traditionally attributed to viruses now seem notto be universal. Historically, viruses were first determined as “filterable agents”, i.e. agentsthat bypass bacterial filters (Pennazio, 2003). However, this view is no longer applicablesince the discovery of giant amoebal viruses that are visible by light microscopy and possesslarge and complex genomes: e.g. Pandoravirus salinus genome is ~2.5 Mbp in length andcontains ~2,500 genes (Aherfi et al., 2016). Moreover, the perception of a virus particle, i.e.virion, as an inert infectious carrier of the virus genome has also been challenged with theobservation that the virions of hyperthermophilic spindle-shaped viruses, Acidianus two-tailed virus and Sulfolobus monocaudavirus 1, develop their tails outside (independently) ofthe host cells (Häring et al., 2005a, Prangishvili et al., 2006, Uldahl et al., 2016). The otherconceptual change lies in the definition of viruses as harmful obligate intracellular parasites:giant viruses may be called “quasi-autonomous” (Claverie and Abergel, 2010) and certainviruses may have relatively mutualistic relations with their hosts. For example,polydnaviruses are obligatory symbionts of parasitoid wasps and promote wasp larvaedevelopment in a body of another insect (Strand and Burke, 2012), and some cyanophagespromote photosynthesis in cyanobacteria during infection (Mann, 2003). In addition, giantviruses themselves may be preyed on by virophages (Aherfi et al., 2016).

Thus, recent discoveries suggest that “viruses are more about exceptions than therules” (Dolja and Krupovič, 2013). All this together with the accumulating data on capsid-less selfish elements has posed a reasonable need to revise the concept of virus. However,the way to define virus still raises hot debates. It has been suggested that virion may beconsidered as viral self, distinguishing viruses from all other replicative elements, such asplasmids, viroids, virusoids, and transposons (Bamford et al., 2002, Bamford, 2003,Krupovič and Bamford, 2010). Alternatively, virus can be defined as a type of replicatorwhich, together with other replicators, forms a continuum from the most selfish (virulentviruses) to the most cooperative (chromosomes, organelle DNA) replicators (Koonin andStarokadomskyy, 2016). Virus can also be viewed as a process, i.e. cyclic change of intra-and extracellular stages (Dupre and Guttinger, 2016). Raoult and Forterre (2008) haveproposed to define viruses and cells as capsid-encoding and ribosome-encoding organismsrespectively, while other replicating entities are grouped into “orphan replicons”.

4

Viruses are extremely abundant in various environments (Rosario and Breitbart, 2011),imposing significant selective pressure on their hosts. It has been estimated that in marineenvironments virus particles are 5-100 times more abundant than prokaryotic cells (Suttle,2007). Moreover, in sea-surface waters, viruses appear to kill ~20-40% of microbial biomassdaily, thus substantially affecting composition of microbial communities and, indirectly,biogeochemical cycles (Suttle, 2007). Viruses are not only the most dominant biologicalentities, but also reservoirs of diverse genetic information. Many currently known viralsequences have no homologs in cellular or other viral sequences (Yin and Fischer, 2008). Ithas been suggested that gene flux from viruses to cells is common, and cells might beregarded as “giant pickpockets” of viral genes (Forterre and Prangishvili, 2013). Virus-derived sequences may constitute large portions of cellular genomes (Gogvadze and Buzdin,2009), and there are several examples of how originally, most probably, viral genes were“domesticated” by cells and recruited for crucial functions. For instance, expression ofendogenous retroviral (former) coat proteins, syncytins, is essential for the formation ofplacenta (Lokossou et al., 2014). Moreover, it has been hypothesized that viruses may havehad direct roles in such important evolutionary events as the origin of DNA and itsreplication mechanisms (Forterre, 2002), as well as development of cellular antiviraldefence systems, including “invention” of nucleus in eukaryotes and formation of complexcell wall in bacteria (Forterre and Prangishvili, 2009). All this, together with many moreexamples of possible virus impact on the evolution of life (Forterre and Prangishvili, 2013,Koonin and Dolja, 2013), suggests that viruses are one of the major forces in shaping of ourworld as we know it.

1.3 Current virus classificationApparently, an intrinsic feature of the human mind is a need for systematization of thesurrounding world in order to understand its ultimate complexity. The discovery of newviruses and accumulation of data regarding their shapes and properties resulted in thedevelopment of a concept of virus species. However, there is still no universal definition ofvirus species that would be satisfactory for all virologists, as different criteria have beenproposed to be used for assigning viruses into taxonomical categories.

Viruses display a variety of virion shapes and genomic structures. The majority of viruscapsids can be characterized by helical or icosahedral symmetry, although some morecomplex structures are also known. Notably, some viruses that infect thermophilic archaeadisplay quite unusual morphotypes, e.g. Acidianus bottle-shaped virus (Häring et al.,2005b, Pietilä et al., 2014, Prangishvili et al., 2016). In addition, a virion may harbour aninner or outer lipid envelope, derived from the host membranes. Virus genome may be aDNA or an RNA molecule, single-stranded (ss) or double-stranded (ds), circular or linear,continuous, nicked or segmented. The size of a genome and consequently, the size of a virionalso vary greatly. For example, porcine circoviruses have a circular ssDNA genome of ~2 kblong and an icosahedral capsid of only ~17 nm in diameter (Finsterbusch and Mankertz,2009), while Pithovirus sibericum has a circular dsDNA genome of ~610 kbp long and avirion of ~1500 nm in length and ~500 nm in diameter (Aherfi et al., 2016).

In the 1960s, Lwoff, Horne, and Tournier proposed a system of virus classificationbased on four essential virus features: a type of nucleic acid (DNA/RNA), a type of capsid

5

symmetry, presence/absence of envelope, and quantitative characteristics of capsid (e.g.number of capsomers) (Lwoff and Tournier, 1966). This system excluded some previouslyproposed criteria, such as symptoms that virus infection causes (Lwoff and Tournier, 1966).Another approach has been proposed by Baltimore, who suggested that all viruses can begrouped into seven classes based on how virus genetic information is expressed andreplicated (Baltimore, 1971).

In 1966, the International Union of Microbiological Societies delegated a task ofdeveloping and maintaining a universal, internationally agreed system of virus classificationto the newly established International Committee on Taxonomy of Viruses (ICTV). The maingoal of ICTV is “to categorize the multitude of known viruses into a single classificationscheme that reflects their evolutionary relationships” (http://www.ictvonline.org/). ICTVcurrently defines a virus species as “a monophyletic group of viruses whose properties canbe distinguished from those of other species by multiple criteria. […] These criteria mayinclude, but are not limited to, natural and experimental host range, cell and tissue tropism,pathogenicity, vector specificity, antigenicity, and the degree of relatedness of their genomesor genes” (http://www.ictvonline.org/codeOfVirusClassification.asp). The recognized taxa areorder, family, sub-family, genus, and species. The Virus Taxonomy Release 2015 reports 7orders, 111 families, 27 sub-families, 609 genera, and 3704 species. Of 111 families, 82 arenot assigned to any order.

Although some genes are widely present in virus genomes, e.g. genes encoding forcapsid proteins in icosahedral viruses, RNA and DNA polymerases, and other enzymesmediating genome replication or integration to the host genomes (Koonin and Dolja, 2013),there is, however, no single universal gene shared by all viruses. This makes it impossible toresolve virus phylogeny in a way similar to the usage of small subunit ribosomal RNAsequences in reconstructing the tree of cellular life. In addition, extensive horizontal genetransfer between different viruses and hosts (Hambly and Suttle, 2005) makes phylogenyresolution even more challenging. It has been proposed that instead of tree-like hierarchy, areticulated network topology might be more suitable for reflecting evolutionary pathways ofviruses (Morgan, 2016).

1.4 Viral structural lineagesWith the increasing number of resolved virion protein structures, striking similarities havebeen found in viruses previously thought to be unrelated (Benson et al., 1999). Inspired bythe observation that same structures are found all across the virosphere, Bamford and co-workers proposed to use similarities in virion architectural principles to group viruses intohigher-level taxa called viral structural lineages (Bamford et al., 2002, Bamford, 2003,Bamford et al., 2005a). The key point of this approach is that virion is what defines virusand distinguishes it from other replicators (Krupovič and Bamford, 2010). Virus geneticmodules responsible for forming a virion are considered to be vertically-inherited virus“self”, opposite to (often) horizontally exchanged modules involved in genome replicationand interactions with the host cell (Abrescia et al., 2010, Krupovič and Bamford, 2010). Thevirus “self” elements are conserved over evolutionary time, as only a limited number ofprotein folds are capable to constitute a functional virion (Bamford, 2003, Benson et al.,2004, Oksanen et al., 2012). Structural similarities may reflect deep evolutionary links, i.e.

6

common ancestry, even when there are no sequence similarities (Bamford, 2003).Moreover, one lineage may include viruses infecting hosts from all three domains of life,which suggests that viruses are ancient and that contemporary viruses are polyphyletic,originating from several types of viruses that infected the last universal common ancestor ofcells (Bamford, 2003, Krupovič and Bamford, 2010).

So far, four major lineages of icosahedral viruses have been established: HK97-like,Picorna-like, Bluetongue virus (BTV)-like, and PRD1-adenovirus-like (Abrescia et al., 2012,Bamford et al., 2005a, Oksanen et al., 2012). Each lineage unites viruses with a distinct foldof the major capsid protein (MCP) and its specific arrangement to form a capsid lattice.

(i) HK97-like viruses are dsDNA viruses with MCPs having a so-called canonical HK97fold, originally determined in bacteriophage HK97 (Wikoff et al., 2000). This lineage alsoincludes several tailed bacteriophages, e.g. T4 (Fokine et al., 2005), P22 (Jiang et al., 2003),and φ29 (Morais et al., 2005), eukaryotic herpesviruses (Baker et al., 2005), as well asarchaeal Haloarcula sinaiiensis tailed virus 1 (HSTV-1) (Pietilä et al., 2013c). Interestingly,HK97 fold was also identified in archaeal and bacterial encapsulins, proteins formingicosahedral nanocompartments inside cells, suggesting a common evolutionary origin ofencapsulins and capsid proteins of viruses belonging to the HK97-like structural lineage(Akita et al., 2007, McHugh et al., 2014).

(ii) Picorna-like viruses have MCPs of a single β-barrel fold, which are usuallypositioned tangentially to the capsid shell surface. Eukaryotic RNA viruses, such as foot-and-mouth disease virus (Acharya et al., 1989), rhinovirus (Rossmann et al., 1985), andpoliovirus (Hogle et al., 1985), as well as ssDNA tailless icosahedral bacteriophage ΦX174(Dokland et al., 1997) represent this lineage.

(iii) BTV-like viral lineage comprises dsRNA viruses with an α-helical MCP (Grimes etal., 1998). In addition to BTV, the members of this lineage are e.g. eukaryotic reovirus(Reinisch et al., 2000), rice dwarf virus (Nakagawa et al., 2003), yeast L-A virus (Naitow etal., 2002), and bacteriophage Φ6 (Huiskonen et al., 2006).

(iv) PRD1-adenovirus-like viruses are dsDNA tailless icosahedral viruses with aninternal lipid membrane (except adenovirus) and MCPs of a vertical double β-barrel fold,arranged into trimers forming pseudo-hexameric capsomers (Benson et al., 1999). Thislineage comprises e.g. bacteriophages PRD1 (Benson et al., 1999) and PM2 (Abrescia et al.,2008), Paramecium bursaria chlorella virus 1 (PBCV-1) (Nandhagopal et al., 2002), humanadenovirus (Athappilly et al., 1994, Rux et al., 2003), and archaeal virus Sulfolobus turretedicosahedral virus (STIV) (Rice et al., 2004). The same MCP fold was also found in virophageSputnik (Zhang et al., 2012a). In addition, vaccinia scaffolding protein D13, which forms atransitory lattice during virus morphogenesis, has also a double β-barrel fold (Bahar et al.,2011).

Recent findings suggest that PRD1-adenovirus lineage may be divided into “double β-barrel viruses” (described above) and “single β-barrel viruses”. Both groups of viruses shareoverall PRD1-type virion architecture, but the single β-barrel viruses have two MCPs of avertical single β-barrel fold, both forming the lattice. The suggested members of this groupare prokaryotic extremophilic viruses: thermophilic phage P23-77 (Rissanen et al., 2013)

7

and haloviruses Salisaeta icosahedral phage 1 (SSIP-1) (Aalto et al., 2012), SH1 (Jäälinojaet al., 2008) and Haloarcula hispanica icosahedral virus 2 (HHIV-2) (Gil-Carton et al.,2015).

In PRD1, genome packaging into a pre-formed procapsid is aided by the packagingATPase P9, which is a virion protein located at the unique vertex (Gowen et al., 2003, Honget al., 2014, Strömsten et al., 2005). Comparison of P9 and other putative packagingadenosine triphosphatases (ATPases) of PRD1-type viruses revealed that in addition to theclassical Walker A and Walker B motifs (Walker et al., 1982), all these ATPases (exceptadenovirus ATPase) contain a region of conserved amino acid residues called P9/A32-specific motif (A32 for vaccinia virus ATPase A32) (Strömsten et al., 2005). A lack ofP9/A32-specific motif in adenovirus ATPase IVa2 may be related to the lack of an inner lipidmembrane in adenovirus virion (Strömsten et al., 2005).

Some other distinct viral structure-based lineages have been suggested in addition tothe four major established ones (Oksanen et al., 2012). Obviously, more information isneeded to identify all possible lineages, the number of which is, however, predicted to berelatively low (Bamford et al., 2005a). Current structure-based virus classification is limitedto the available high-resolution structures, but high-throughput methods for resolvingthree-dimensional (3-D) protein structures, as well as robust computational tools forstructure comparison, are under constant improvement (Abrescia et al., 2012, Ravantti etal., 2013). Recent advances in cryo-electron microscopy (cryo-EM) have already yieldedcomplex molecular structures of near-atomic resolution (Kuhlbrandt, 2014). Improvementsin the methodology within time might enable more comprehensive structure-basedreconstruction of viral phylogeny.

1.5 Haloviruses

1.5.1 Eukaryotic and bacterial halovirusesMany marine organisms may be called slight halophiles, but here, halophilic refers toorganisms inhabiting environments where salinity is higher than in seawater. To the best ofmy knowledge, and excluding this work, 74 haloviruses, i.e. viruses that infect halophilicorganisms, have been isolated and described to some extent (Atanasova et al., 2015b, Luk etal., 2014). The vast majority of these are archaeal viruses (54 viruses, see chapter 1.5.2).Very little is known about viruses that infect halophilic eukaryotes (5 viruses) or bacteria (15viruses) (Figures 1 A and B, Table 1). So far, eukaryotic haloviruses are represented by onlyfive amoebal icosahedral viruses, which were isolated from saline environments in Tunis(Boughalmi et al., 2013). Based on virion morphological features and comparisons of DNApolymerase or capsid protein sequences, four of these viruses were assigned to the familyMarseilleviridae and one to Mimiviridae family (Boughalmi et al., 2013) (Figure 1 A).

The known bacterial viruses isolated from hypersaline environments infect hosts fromgenera Halomonas, Pseudomonas, Tetragenococcus, Salinivibrio, Salicola, and Salisaeta,having either a lytic or lysogenic life cycle (Table 1). These viruses are tailed, morphologicallyresembling the members of the families Myoviridae (icosahedral head and contractile tail,further called myoviruses), Siphoviridae (icosahedral head and long non-contractile tail,further called siphoviruses), or Podoviridae (icosahedral head and short noncontractile tail,

8

further called podoviruses), except SSIP-1, which is a tailless icosahedral virus with an innerlipid membrane (Figure 1 B). Myovirus Salicola tailed phage 2 (SCTP-2) is a large virus witha head of ~125 nm in diameter and a tail of ~145 nm in length (Kukkaro and Bamford, 2009),while the other described bacterial haloviruses display smaller particles. The genomes ofbacterial haloviruses are either linear or circular DNA molecules of variable size up to 340kb, as in the temperate Halomonas phage ΦgspC (Seaman and Day, 2007) (Table 1). Cryo-EM-based reconstructions of virus particles are available for Salinivibrio phage CW02 (Shenet al., 2012) and for Salisaeta phage SSIP-1 (Aalto et al., 2012). CW02 is a member of theT7-like phage supergroup, and its virion reconstruction at 16-Å resolution permitted fittingof the HK97-fold (Shen et al., 2012), a conserved fold commonly found in dsDNAbacteriophages (see above, chapter 1.4). Salisaeta phage SSIP-1 shares common featureswith PRD1-like viruses, and the canonical double β-barrel MCP of bacteriophage PM2 wasfitted to SSIP-1 virion reconstruction at 12.5-Å resolution (Aalto et al., 2012). SSIP-1 has twoMCPs, thus it has been suggested to belong to the subgroup of single β-barrel viruses in thePRD1-adenovirus structural lineage (Aalto et al., 2012).

In addition, the analysis of a metavirome derived from Santa Pola solar saltern (Spain)allowed putative assignment of the identified halophages to the bacterium Salinibacterruber, one of the dominant players in the saltern microbiota (Garcia-Heredia et al., 2012,Santos et al., 2010). However, no viruses infecting S. ruber have yet been isolated. Similarly,metagenomic analysis of water surface samples from hypersaline Organic Lake in Antarcticaallowed reconstruction of a complete genome of virophage Organic Lake virophage (OLV),which is related to Sputnik, and a near-complete genome of its putative helper virus OrganicLake phycodnavirus (OLPV), which infects algae (Yau et al., 2011). Interestingly, thevirophage OLV MCP sequences were also found in metagenomes from other geographicallydistant hypersaline and freshwater reservoirs, suggesting its possibly global occurrence (Yauet al., 2011).

9

Figure 1. Morphotypes of isolated (A) eukaryotic, (B) bacterial, and (C) archaeal haloviruses.Representative viruses (virions) are drawn schematically. Scale bar, 100 nm. Tailed icosahedralviruses are coloured yellow, tailless icosahedral are green, and lipids are indicated by blue. Modifiedfrom (Atanasova et al., 2015b).

10

Tab

le 1

. Euk

aryo

tic

and

bact

eria

l hal

ovir

uses

.

Vir

us

Isol

atio

n h

ost

Mor

ph

otyp

e, s

izea

Lif

e cy

cle

Gen

ome,

size

Isol

atio

nso

urc

eR

efer

ence

Seb1

sol

Aca

ntha

moe

ba p

olyp

haga

Taill

ess

icos

ahed

ral (

Mrs

-lik

eb ), ~

220

nm

Ndc

DN

ASa

line

soil,

Tuni

sB

ough

alm

iet a

l.,20

13Se

b2so

lTa

illes

s ic

osah

edra

l (M

rs-l

ike)

, ~ 2

01 n

mN

dD

NA

Seb6

sol

Taill

ess

icos

ahed

ral (

Mrs

-lik

e), ~

206

nm

Nd

DN

ASe

b1ea

uTa

illes

s ic

osah

edra

l (M

rs-l

ike)

, ~ 2

26 n

mN

dD

NA

Salt

lake

wat

er, T

unis

Bou

g1Ta

illes

s ic

osah

edra

l (M

imi-

liked )

, ~ 5

19 n

mN

dD

NA

F9-1

1H

alom

onas

hal

ophi

laSi

phov

irus

,~ 6

0/20

0 nm

Lyso

geni

cN

dH

. hal

ophi

laF9

-11e

Cal

voet

al.,

198

8

G3

Pseu

dom

onas

sp. G

3M

yovi

rus

Lyti

cN

dSa

lt p

ond,

Can

ada

Kau

riet

al.,

199

1

φ71

16Te

trag

enoc

occu

s ha

loph

ilus

Myo

viru

s, 8

7-96

/200

nm

Lyti

cN

dFe

rmen

ting

soy

sauc

eU

chid

a an

d K

anbe

,19

93φ

D-8

6Te

trag

enoc

occu

s ha

loph

ilus

Siph

ovir

us 6

7/30

0 nm

Lyti

cN

dF5

-4H

alom

onas

hal

ophi

laSi

phov

irus

,~ 3

3/14

2 nm

Lyso

geni

cN

dH

. hal

ophi

lae

Cal

voet

al.,

199

4F1

2-9

Hal

omon

as h

alop

hila

Siph

ovir

us,~

42/

133

nmLy

soge

nic

Nd

UTA

KV

ibri

osp

. B1

Myo

viru

s, ~

86/

100

nmLy

tic

dsD

NA

,~

80 k

bpSo

lar

salt

ern,

Alic

ante

, Spa

inG

oele

t al.,

199

6

Φgs

pBH

alom

onas

sal

ina

Myo

viru

s,46

-51/

112

nmN

dds

DN

A,

~49

kb

Salin

e so

il,G

reat

Sal

tPl

ains

, USA

Seam

an a

nd D

ay,

2007

Φgs

pCH

alom

onas

sal

ina

Myo

viru

s, 5

3-86

/121

nm

Lyso

geni

cds

DN

A,

~34

0 kb

SCTP

-1Sa

licol

asp

. PV

3Si

phov

irus

, hea

d ~

95nm

Nd

Nd

Sola

r sa

lter

n,M

argh

erit

a di

Savo

ia, I

taly

Kuk

karo

and

Bam

ford

, 200

9SC

TP-2

Salic

ola

sp. P

V4

Myo

viru

s, ~

125

/145

nm

Nd

Nd

SCTP

-3Sa

licol

asp

. s3-

1M

yovi

rus

Nd

Nd

Ata

naso

vaet

al.,

2012

CW

02Sa

liniv

ibri

oSA

50Po

dovi

rus,

~60

nm

,T=

7la

evo

Nd

Line

ards

DN

A,

49,3

90 b

p

Gre

at S

alt

Lake

, USA

Shen

et a

l., 2

012

SSIP

-1Sa

lisae

tasp

. SP9

-1Ta

illes

s ic

osah

edra

l wit

h in

ner

lipid

mem

bran

e, ~

100

nm,T

= 4

9Ly

tic

Cir

cula

rds

DN

A,

43,7

88 b

p

Salt

wat

er,

Sedo

m P

onds

f ,Is

rael

Aal

toet

al.,

201

2

QH

HSV

-1H

alom

onas

ven

tosa

eSi

phov

irus

,~ 4

7/75

nm

Lyti

cds

DN

ASa

lt m

ine,

Qia

ohou

,C

hina

Fuet

al.,

201

6

a.C

apsi

d di

amet

er a

nd tr

iang

ulat

ion

num

ber

(T) f

or ta

illes

s ph

ages

and

pod

ovir

uses

; cap

sid

diam

eter

/tai

l len

gth

for

taile

d ph

ages

.b.

Mrs

-lik

e, M

arse

illev

irus

-lik

e.c.

Nd,

not

det

erm

ined

.d.

Mim

i-lik

e, M

imiv

irus

-lik

e.e.

Indu

ced

from

Hal

omon

as h

alop

hila

str

ains

isol

ated

from

hyp

ersa

line

soils

nea

r A

lican

te, S

pain

.f.

Sedo

m P

onds

are

exp

erim

enta

l pon

ds in

Isr

ael w

here

wat

er fr

om th

e D

ead

Sea

is d

ilute

d w

ith

wat

er fr

om th

e R

ed S

ea.

11

1.5.2 Archaeal halovirusesExcluding this work (see chaper 4.1), ~100 archaeal viruses have been isolated (Atanasovaet al., 2015b, Snyder et al., 2015). These are viruses that infect thermophilic crenarchaea,methanogenic euryarchaea, or halophilic euryarchaea (Dellas et al., 2014, Pietilä et al., 2014,Prangishvili, 2013). About 50 isolated viruses infect haloarchaea from eight different generaof the family Halobacteriaceae (Table 2) (Atanasova et al., 2015a). These virus isolatesoriginate from various hypersaline environments all over the world, and represent thefollowing morphotypes: tailed icosahedral, spindle-shaped (also called fusiform or lemon-shaped), tailless icosahedral, and pleomorphic (Figures 1 C and 2 A, Table 2) (Atanasova etal., 2015a). The Genbank non-redundant (n.r.) database (accessed 23.8.2016) also containsa genome sequence (dsDNA, ~40 kbp) of Halorubrum phage CGΦ46 (NC_021537.1), butno other information is available for this virus. In addition, one complete (environmentalhalophage 1 [EHP-1]) and 42 near-complete viral genomes were retrieved from metagenomicdata obtained from a solar saltern in Santa Pola (Alicante, Spain), and 25 of these weretentatively assigned to infect halophilic archaea Haloquadratum walsbyi (15 viruses),Halorubrum lacusprofundi (3), or nanohaloarchaea (7) (Garcia-Heredia et al., 2012, Santoset al., 2007). Recent methodological advances have also allowed unambiguous identificationof virus-host pairs in hypersaline environments, circumventing a need for cultivation, asdemonstrated by the identification of Nanohaloarchaeal virus 1 (NHV-1) most likelyinfecting a yet unculturable nanohaloarchaeon from Santa Pola saltern (Martínez-García etal., 2014).

12

Tab

le 2

. Arc

haea

l hal

ovir

uses

.

a.V

irus

es w

hose

gen

ome

sequ

ence

is a

vaila

ble

are

in b

old,

and

ast

eris

ks m

ark

viru

ses

who

se s

truc

ture

s ha

ve b

een

dete

rmin

edby

cry

o-E

M o

r cr

yo-e

lect

ron

tom

ogra

phy

and

3-D

imag

e re

cons

truc

tion

. For

full

viru

s na

mes

see

“A

bbre

viat

ed v

irus

nam

es”.

b.Fo

r Φ

H, o

nly

part

ial g

enom

e se

quen

ce is

ava

ilabl

e.c.

For

mor

e in

form

atio

n, s

ee T

able

3.

d.Fo

r m

ore

info

rmat

ion,

see

Tab

le 4

.

Mor

ph

otyp

eV

iru

sesa

Hos

t ge

ner

aR

efer

ence

s

Myo

viru

sH

s1,Φ

Hb ,

S510

0,H

F1,

HF

2, H

RTV

-1,

HR

TV-2

, HR

TV-3

,HR

TV

-5, H

RTV

-6,

HR

TV

-7,H

RT

V-8

, HR

TV-9

, HR

TV-1

0,H

RTV

-11,

HR

TV-1

2,H

ST

V-2

*,H

STV

-3,

HJT

V-1

, HJT

V-2

, HA

TV-1

, HA

TV-2

,HG

TV

-1,

ΦC

h1

Hal

obac

teri

um,

Hal

ofer

ax,

Hal

orub

rum

,H

aloa

rcul

a,H

alog

ranu

m,

Nat

rial

ba

Ata

naso

vaet

al.,

201

2, D

anie

ls a

nd W

ais,

1990

, Nut

tall

and

Dya

ll-Sm

ith,

199

3,Pi

etilä

et a

l., 2

013b

, Sch

nabe

l et a

l.,19

82b,

Senči

lo e

t al.,

201

3, T

ang

et a

l.,20

02, T

orsv

ik a

nd D

unda

s, 1

980,

Tor

svik

and

Dun

das,

197

4, W

itte

et a

l., 1

997

Siph

ovir

usH

h1, H

h3, J

a1, Φ

N, S

45, B

10,B

J1, H

RT

V-4

,H

HT

V-1

, HH

TV

-2,H

CT

V-1

, HC

TV

-2,

HC

TV

-5, H

VT

V-1

*,H

VT

V-2

Hal

obac

teri

um,

Hal

orub

rum

,H

aloa

rcul

a

Ata

naso

vaet

al.,

201

2, D

anie

ls a

ndW

ais,

1984

, Pag

alin

g et

al.,

200

7, P

aulin

g,19

82, P

ieti

lä e

t al.,

201

3b, S

enči

lo e

t al.,

2013

, Tor

svik

, 198

2, V

ogel

sang

-Wen

kean

d O

este

rhel

t, 19

88, W

ais

et a

l., 1

975

Podo

viru

sH

ST

V-1

*H

aloa

rcul

aA

tana

sova

et a

l., 2

012,

Pie

tilä

et a

l.,20

13c

Spin

dle

His

1*H

aloa

rcul

aB

ath

et a

l., 2

006,

Hon

get

al.,

201

5,Pi

etilä

et a

l., 2

013a

Taill

ess

icos

ahed

ralc

SH

1*, H

HIV

-2*,

PH

1,SN

J1H

aloa

rcul

a,H

alor

ubru

m,

Nat

rine

ma

Bam

ford

et a

l., 2

005b

, Gil-

Car

ton

et a

l.,20

15, J

aakk

ola

et a

l., 2

012,

Jää

linoj

a et

al.,

2008

, Por

ter

et a

l., 2

013,

Zha

ng e

tal

., 20

12b

Pleo

mor

phic

dH

HP

V-1

, HH

PV

-2, H

is2,

HR

PV

-1*,

HR

PV

-2, H

RP

V-3

, HR

PV

-6, H

GP

V-1

,S

NJ2

Hal

oarc

ula,

Hal

orub

rum

,H

alog

eom

etri

cum

,N

atri

nem

a

Ata

naso

vaet

al.,

201

2, B

ath

et a

l., 2

006,

Li e

t al.,

201

4, L

iu e

t al.,

201

5, P

ieti

lä e

tal

., 20

09, P

ieti

lä e

t al.,

201

0, P

ieti

lä e

tal

., 20

12, R

oine

et a

l., 2

010,

Senči

lo e

tal

., 20

12

13

1.5.2.1 Tailed icosahedral virusesThe majority of the isolated haloarchaeal viruses are tailed icosahedral resembling themembers of the families Myoviridae (24 viruses) or Siphoviridae (15 viruses) (Atanasova etal., 2015b). One podovirus that infects Haloarcula sinaiiensis (HSTV-1) has also beenisolated (Atanasova et al., 2012, Pietilä et al., 2013c). Although tailed haloarchaeal virusisolates are rather numerous, tailed virus-like particles (VLPs) are considered to constituteonly a minor part of natural viral communities in hypersaline environments (Santos et al.,2012).

Tailed viruses infect haloarchaea belonging to the genera Halobacterium, Haloarcula,Halorubrum, Haloferax, Halogranum, or Natrialba (Table 2), and these viruses (thosestudied for the life cycle) are virulent or temperate. However, no lysin or holin genes (Younget al., 2000) have been identified in the genomes of virulent haloarchaeal tailed viruses, andexact mechanisms of cell lysis are not known. Temperate haloarchaeal tailed viruses arerepresented by only two myoviruses: ΦH infecting Halobacterium salinarum and ΦCh1infecting Natrialba magadii. The ΦH partial genome (L segment) may reside in the hostcytoplasm as a plasmid (Schnabel, 1984), while ΦCh1 integrates into the host genome (Witteet al., 1997). In addition, a few other tailed haloarchaeal viruses may also be temperate, astheir genomes encode putative integrases (Senčilo et al., 2013).

ΦH and ΦCh1 are so far the best studied haloarchaeal tailed viruses in terms of genomefunctions. Molecular studies have showed that ΦH genome is highly unstable and duringinfection rounds, many genetically rearranged variants are produced (Schnabel et al.,1982a). The ΦH genome sequence, albeit not complete, is similar to that of ΦCh1.Noticeably, ΦCh1 is the only archaeal virus known to contain host-derived RNA along withvirus DNA in the virion (Witte et al., 1997). Similarly to ΦH, ΦCh1 also undergoes genomicrearrangements. The rearrangements occur in a certain region of ΦCh1 genome, allowingproduction of different types of tail fibres (Klein et al., 2012). Genomic rearrangements havealso been reported for a few other tailed haloarchaeal viruses (Daniels and Wais, 1998,Pagaling et al., 2007).

The genomes of 17 out of 40 tailed haloarchaeal viruses have been fully sequenced.These are dsDNA molecules, either circularly permuted and terminally redundant(suggesting headful DNA packaging mechanism) or non-permuted with direct terminalrepeats (suggesting replication through concatemeric intermediates) (Senčilo and Roine,2014). The genome sequences are diverse and mosaic, like those of tailed bacteriophages(Krupovič et al., 2010, Pietilä et al., 2013b, Senčilo et al., 2013). Some genes, e.g. onesencoding structural proteins and proteins involved in DNA and RNA metabolism, are alsosimilar to those in bacteriophages, but the majority of predicted genes have no assignedfunctions (Senčilo and Roine, 2014). In terms of sequence similarities observed within tailedhaloarchaeal viruses, nine viruses can be divided into three groups of closely related viruses:HF2-like, HRTV-7-like, and HCTV-1-like (Senčilo and Roine, 2014). Two other viruses(HCTV-2 and HHTV-2) are more distantly related, and six other viruses are singletons(Senčilo and Roine, 2014). Putative proviruses related to haloarchaeal tailed viruses havebeen identified in the genomes of halophilic and/or alkaliphilic, as well as in thermophilicand/or methanogenic euryarchaea (Krupovič et al., 2010, Senčilo et al., 2013).

14

Virion structures have been resolved for Hrr. sodomense myovirus HSTV-2, Har.vallismortis siphovirus HVTV-1, and Har. sinaiiensis podovirus HSTV-1 (Pietilä et al.,2013b, Pietilä et al., 2013c). All three viruses lose their infectivity at low salinity, but virusparticles display no major structural changes, and regain infectivity when high salinity isrestored (Pietilä et al., 2013b, Pietilä et al., 2013c). 3-D reconstruction revealed that HVTV-1 icosahedral capsid is built of MCP capsomers arranged in the lattice with the triangulationnumber (T) T = 13, 5-fold symmetric spikes project from the vertices, and trimeric decorativestructures are located in the center of each MCP hexamer (Pietilä et al., 2013b). In HSTV-2,MCP capsomers are arranged in T = 7 lattice and an additional minor structural protein isinserted between the MCPs, enlarging the capsid (Pietilä et al., 2013b). This was a newmechanism for capsid enlargement, that allows packaging of a larger DNA molecule thanexpected to fit into the icosahedron with T = 7 symmetry (Pietilä et al., 2013b). In HSTV-1,capsomers are arranged in T = 7 lattice, and cone-shaped towers decorate the virion at thecenter of each facet (Pietilä et al., 2013c). Notably, it was shown that HSTV-1 MCP has thecanonical HK97 fold (Pietilä et al., 2013c). Thus, structural analyses allowed placing thisarchaeal virus into the HK97-like structural viral lineage, together with tailedbacteriophages and eukaryotic herpesviruses (see chapter 1.4). Structure prediction andhomology modelling suggest that ΦCh1 may also adopt the HK97-fold, indicating that thelineage may be expanded to include more archaeal viruses (Krupovič et al., 2010).

1.5.2.2 Spindle-shaped virus His1Several archaeal spindle-shaped viruses have been isolated, and the majority of these virusesinfect hyperthermophilic crenarchaea (Krupovič et al., 2014, Pietilä et al., 2013a), and theonly one, His1, infects a halophilic euryarchaeon (Bath and Dyall-Smith, 1998). Notably,spindle-shaped morphotype is unique for archaeal viruses, as no such viruses have beenidentified for eukaryotes or bacteria. Currently, crenarchaeal spindle-shaped viruses areclassified into two families: Fuselloviridae, comprising viruses with one short tail, andBicaudaviridae, comprising viruses with one or two long tails. His1 is the only member ofSalterprovirus genus which is not assigned to any virus family. However, comparativeanalyses have showed that based on the similarities in MCP sequences, all spindle-shapedviruses may be assigned to the two above mentioned families, and His1 has been suggestedto form the Epsilonfusellovirus genus in the Fuselloviridae family (Krupovič et al., 2014).

His1 was isolated from a solar saltern in Victoria, Australia, on Haloarcula hispanica(Bath and Dyall-Smith, 1998). Originating from a highly saline environment, His1, however,tolerates a wide range of salinities (from 50 mM to 4 M NaCl) (Pietilä et al., 2013a). Theinfection is persistent, i.e. virus particles are released from the host cells continuously, whilethe host growth is retarded (Pietilä et al., 2013a, Svirskaitė et al., 2016). His1 genome, whichis a linear ~14.5 kbp long dsDNA molecule, encodes a putative type B DNA polymerase, hasinverted terminal repeats, and terminal proteins attached (Bath et al., 2006). This suggeststhat His1 replicates via protein priming. His1 MCP and a few minor structural proteins forma spindle-shaped virion, but pleomorphism in the virion shape has been observed (Pietilä etal., 2013a). No membrane has been detected in the virion, but the MCP exists in two forms,one of which is modified with a lipid moiety (Pietilä et al., 2013a). It has been suggested thatthe lipid modification enables hydrophobic interactions between the MCPs and thus may

15

account for the virion flexibility (Pietilä et al., 2013a). Cryo-electron tomography revealedthat His1 particles vary in size, but on average are ~92 nm long and ~40 nm wide, having~12 nm long tail-like structures with six tail spikes (Hong et al., 2015). The tail structure wasobserved to be more constant compared to the variable body shape (Hong et al., 2015). Ithas been speculated that His1 may recognize the host and attach to it with its tail spikestructures, and the tail hub may mediate the genome injection, after which spindle-shapedparticles could transform into empty tubes (Hong et al., 2015).

Curiously enough, while His1 is the only isolated spindle-shaped halovirus, culture-independent studies suggest that spindle-shaped viruses are abundant in hypersalineenvironments (Santos et al., 2012). Transmission electron microscopy (TEM) revealedspindle-shaped particles associated with square cells of Haloquadratum walsbyi, one of themost abundant extremely halophilic archaea (Guixa-Boixareu et al., 1996). Although Hqr.walsbyi was discovered already in 1980 by A. E. Walsby, it took more than 20 years to obtainit in pure culture (Burns, 2007, Walsby, 1980). Possibly, spindle-shaped viruses infectingHqr. walsbyi are yet awaiting to be isolated. In addition, in the recent metaviromic studiesof Namib Desert Salt pans, three obtained virus contigs showed sequence similarities to His1genes, including those coding for the MCP and minor structural proteins (Adriaenssens etal., 2016). It has been proposed that these contigs represent the genomes of virusesbelonging to the same virus lineage as His1 (Adriaenssens et al., 2016).

1.5.2.3 Tailless icosahedral virusesBefore this work, only six archaeal tailless icosahedral viruses were known, two of whichinfect thermophilic archaea (Sulfolobus viruses STIV and STIV2), and four infect halophilicarchaea (SH1, HHIV-2, PH1, and SNJ1) (Bamford et al., 2005b, Happonen et al., 2010,Jaakkola et al., 2012, Maaty et al., 2006, Porter et al., 2013, Zhang et al., 2012b). STIV andSTIV2 have been classified to the family Turriviridae, while SH1, HHIV-2, PH1, and SNJ1have been assigned to the family Sphaerolipoviridae together with Thermus phages P23-77and IN93 (Table 3). The Sphaerolipoviridae family includes three genera:Alphasphaerolipovirus (SH1, HHIV-2, PH1), Betasphaerolipovirus (SNJ1), andGammasphaerolipovirus (P23-77, IN93).

16

Tab

le 3

. Mem

bers

of t

he fa

mily

Spha

erol

ipov

irid

ae.

Gen

us

Vir

usa

Isol

atio

nh

ost

Lif

e cy

cle

Isol

atio

nso

urc

eS

tru

ctu

res

reso

lved

Cap

sid

dia

met

er,T

bM

CP

scG

enom

e, G

Bac

c. n

o.d

Ref

eren

ces

Alp

ha-

spha

ero-

lipov

irus

SH

1H

aloa

rcul

ahi

span

ica

Lyti

cSa

line

lake

(Ser

pent

ine

Lake

),A

ustr

alia

Vir

ion

(9.6

Å)

~80

nm

,T

= 2

8,de

xtro

VP7

, VP4

Line

ar d

sDN

A,

30,8

89 b

p,A

Y950

802

Bam

ford

et a

l.,20

05b,

Jääl

inoj

a et

al.,

2008

, Kiv

elä

etal

., 20

06,

Port

er e

t al.,

2005

HH

IV-2

Hal

oarc

ula

hisp

anic

aLy

tic

Sola

r sa

lter

n,(M

argh

erita

di S

avoi

a),

Ital

y

Vir

ion

(13

Å)

~80

nm

,T

= 2

8,de

xtro

VP7

, VP4

Line

ar d

sDN

A,

30,5

78 b

p,JN

9684

79

Gil-

Car

ton

etal

., 20

15,

Jaak

kola

et a

l.,20

12PH

1H

aloa

rcul

ahi

span

ica

Lyti

cSa

line

lake

(Pin

k La

kes)

,A

ustr

alia

~50

nm

VP7

, VP4

Line

ar d

sDN

A,

28,0

72 b

p,K

C25

2997

Port

e ret

al.,

2013

Bet

a-sp

haer

o-lip

ovir

us

SN

J1N

atri

nem

asp

. J7-

2Ly

soge

nic

Nat

rine

ma

sp. J

7-1

70-7

5 nm

PB6,

PB

2C

ircu

lar

dsD

NA

,16

,341

bp,

AY0

4885

0.1

Liu

et a

l., 2

015,

Zhan

g et

al.,

2012

b

Gam

ma-

spha

ero-

lipov

irus

P23

-77

Ther

mus

ther

mop

hilu

sLy

tic

Alk

alin

e ho

tsp

ring

,N

ew Z

eala

nd

Vir

ion

(14

Å),

MC

Ps (

3 Å

),~

78 n

m,

T =

28,

dext

roV

P16,

VP1

7C

ircu

lar

dsD

NA

,17

,036

bp,

GQ

4037

89

Jaat

inen

et a

l.,20

08,

Jala

svuo

ri e

tal

., 20

09IN

93Th

erm

usth

erm

ophi

lus

Lyso

geni

cTh

erm

usaq

uati

cus

TZ2

~13

0nm

VP1

3,V

P14

Cir

cula

rds

DN

A,

19,6

04 b

p,A

B06

3393

Mat

sush

ita

etal

., 19

95,

Mat

sush

ita

and

Yana

se, 2

009

a.Ty

pe s

peci

es n

ames

are

in b

old.

b.

T, tr

iang

ulat

ion

num

ber.

c.Sm

all a

nd la

rge

MC

Ps.

d.

Gen

Ban

k ac

cess

ion

num

bers

.

17

(i) Alphasphaerolipoviruses SH1 (Porter et al., 2005), HHIV-2 (Jaakkola et al., 2012),and PH1 (Porter et al., 2013) are virulent viruses isolated from distant hypersalineenvironments (saline lakes in Australia and a solar saltern in Spain) on Haloarculahispanica (Table 3). These viruses have narrow host ranges. In addition to Har. hispanica,SH1 is able to infect Haloarcula sp. PV7 and Halorubrum sp. CSW 2.09.4 which is closelyrelated to Hrr. sodomense (Jaakkola et al., 2012, Porter et al., 2005). HHIV-2 is also able toinfect Haloarcula sp. PV7 (Atanasova et al., 2012, Jaakkola et al., 2012), and PH1 infectsHalorubrum, sp. CSW 2.09.4 (Porter et al., 2013). It has been shown that in spite of thenarrow host ranges, SH1 and PH1 genomes can be transfected into several additionalarchaeal strains, suggesting that these virus genomes may replicate in a wide range of strains(Porter and Dyall-Smith, 2008, Porter et al., 2013). Although there have been somecontradictory proposals regarding SH1 exit strategy (Porter et al., 2007, Porter et al., 2005),recent studies, which included traditional metrics and biochemical measurements, haveshown that SH1 life cycle is lytic (Svirskaitė et al., 2016). Concurrently with virus release,cell culture density and viable cell counts drop, host cell membrane becomes compromised,cell respiration is declining, and adenosine triphosphate (ATP) leakage from the cells isdetected (Svirskaitė et al., 2016).

The genomes of alphasphaerolipoviruses are linear dsDNA molecules about 30 kbplong having inverted terminal repeats and terminal proteins attached (Bamford et al.,2005b, Jaakkola et al., 2012, Porter and Dyall-Smith, 2008, Porter et al., 2013). Whetherthese viruses replicate via protein-primed system remains to be seen, as no putative protein-primed DNA polymerase gene was found in their sequences. In these viruses, the mostdistinct genes encoding structural proteins are those for putative spikes (see below), whichmost probably serve as host recognition devices (Jaakkola et al., 2012, Jäälinoja et al.,2008). The genomes of alphasphaerolipoviruses have no predicted integrase genes.However, two related putative proviruses have been recognized in the chromosomes ofhalophilic archaea Halobiforma lacisalsi (provirus HaloLacP1) and Haladaptatuspaucihalophilus (provirus HalaPauP1) (Porter et al., 2013).

The virions of alphasphaerolipoviruses are icosahedrons of ~80 nm in diameter(except PH1), and they have an inner lipid membrane composed of lipids selectively acquiredfrom the host cell membrane (Bamford et al., 2005b, Gil-Carton et al., 2015, Jäälinoja et al.,2008). Cryo-EM structures of SH1 and HHIV-2 revealed that these viruses have capsidsarranged in an unusual T = 28 dextro lattice (same T number is also reported for Thermusphage P23-77, see below) (Gil-Carton et al., 2015, Jäälinoja et al., 2008). The capsids arebuilt of pseudo-hexameric capsomers having either two or three towers on the capsomersurface. The capsomers are formed by small and large MCPs, which are called virion proteins(VPs) VP7 (small MCP) and VP4 (large MCP) in both viruses. VP7 has one β-barrel domain,whereas VP4 has two single β-barrel domains, one on the top of the other (Gil-Carton et al.,2015, Jäälinoja et al., 2008). For HHIV-2, it has been proposed that VP4-VP4 homodimersand VP4-VP7 heterodimers make the lattice, which is stabilized by disulphide bridges acrossthe capsomers. The upper domain of VP4 (a β-barrel) forms the capsomer towers (Gil-Carton et al., 2015). The capsomer pseudo-hexameric footprint is conserved in SH1 andHHIV-2, while spike complexes located at 5-fold vertices have different architectures. SH1capsid is decorated with horn-like spikes (Jäälinoja et al., 2008), and HHIV-2 has propeller-

18

like spikes (Gil-Carton et al., 2015). This observation highlights conservation in the majorcapsid building blocks (virus “self” elements) and faster evolution in the host recognitionstructures.

(ii) Betasphaerolipovirus SNJ1 is a temperate virus, which resides as a provirus(plasmid) in Natrinema sp. J7-1, and produces plaques on Natrinema sp. J7-2 (Zhang et al.,2012b). SNJ1 life cycle mode depends on salinity: it is lytic at ≥ 3.4 M NaCl and lysogenic atlower salinity (3 M NaCl) (Mei et al., 2015). SNJ1 virions are 70-75 nm in diameter andcontain an inner lipid membrane (Liu et al., 2015). The genome is a circular DNA moleculeof about 16 kbp long (Zhang et al., 2012b). Little sequence similarity is observed betweenSNJ1 and alphasphaerolipoviruses.

(iii) Archaeal tailless icosahedral haloviruses share structural similarities with thebacterial tailless icosahedral viruses P23-77 and IN93 belonging to the genusGammasphaerolipovirus (Pawlowski et al., 2014). These viruses infect Thermusthermophilus strains (host range is narrow), undergo lytic (P23-77) or lysogenic (IN93) lifecycles, and have circular dsDNA genomes (Jaatinen et al., 2008, Jalasvuori et al., 2009,Matsushita and Yanase, 2009). P23-77 virion and MCPs structures have been resolved(Rissanen et al., 2013). Like in SH1 and HHIV-2, P23-77 icosahedral capsid of 78 nm indiameter is arranged in T = 28 dextro lattice. Small MCP VP16 and large MCP VP17 formpseudohexameric capsomers with two towers, which are located either on the same side orthe opposite corners of the capsomer (Rissanen et al., 2013). In all three viruses, P23-77,SH1, and HHIV-2, the lattice footprint is conserved, while P23-77 has yet another type ofspikes (elongated) which emerge from the 5-fold vertices (Gil-Carton et al., 2015, Rissanenet al., 2013).

Sphaerolipoviruses share virion architectural principles with the viruses belonging tothe PRD1-adenovirus structural lineage (see chapter 1.4). Together with SSIP-1 and severalrelated putative proviruses, sphaerolipoviruses constitute a distinct group of single β-barrelviruses within the lineage.

1.5.2.4 Pleomorphic virusesThe group of pleomorphic haloviruses includes eight viruses isolated from remotehypersaline locations on Haloarcula, Halorubrum, or Halogeometricum strains, and onetemperate virus induced from Natrinema sp. J7-1 (Table 4) (Liu et al., 2015, Pietilä et al.,2016). In addition, several related putative proviruses have been identified in thechromosomes or extrachromosomal elements of halophilic archaea belonging to severaldifferent genera (Liu et al., 2015). Pleomorphic archaeal viruses have been recently classifiedto the new family Pleolipoviridae (Pietilä et al., 2016). These viruses display virions ofspherical or pleomorphic shape, which are in fact vesicles built of lipids (which areunselectively derived from the host membrane) and typically only two structural proteinspecies: membrane-associated protein and spike protein (Pietilä et al., 2010, Pietilä et al.,2012). Cryo-EM reconstruction of Halorubrum pleomorphic virus 1 (HRPV-1) virions hasshown that the spike proteins are distributed irregularly on the virion surface (Pietilä et al.,2010). Pleolipoviruses undergo non-lytic life cycles, presumably entering the host cell byfusion of the virus lipid membrane with the host cell membrane (Pietilä et al., 2016).

19

Assembled virions are released continuously most probably via budding, retarding the hostgrowth, either slightly (as in case of HRPV-1 infection) or quite substantially (as in case ofHis2 infection) (Pietilä et al., 2009, Pietilä et al., 2012, Svirskaitė et al., 2016). The hostrange is typically narrow or limited to the isolation strain (Atanasova et al., 2012, Pietilä etal., 2009, Pietilä et al., 2012). The only temperate pleolipovirus, SNJ2, is able to integrateto and excise from the host chromosome (transfer RNA [tRNA] gene) (Liu et al., 2015).Peculiarly, SNJ2 production is dependent on the presence of SNJ1, the temperate virusresiding in Natrinema sp. J7-1 as a plasmid (Liu et al., 2015) (Table 3, see also chapter1.5.2.3). However, the nature of the interaction between SNJ1 and SNJ2 remains to beclarified. The genome type of pleolipoviruses is not conserved, it may be either circularssDNA or circular/linear dsDNA, depending on the virus (Pietilä et al., 2016). Interestingly,the genomes of HGPV-1, HRPV-3, and SNJ2 are circular dsDNA molecules with single-stranded interruptions, whose function is not yet clear (Liu et al., 2015, Senčilo et al., 2012).Although little sequence similarity is observed between pleolipovirus genomes, they allcontain a conserved set of genes, including those coding for structural proteins and aputative NTPase (Liu et al., 2015, Senčilo et al., 2012). Based on the gene content,pleolipoviruses are grouped into three genera: Alphapleolipovirus, Betapleolipovirus, andGammapleolipovirus (Table 4) (Pietilä et al., 2016).

20

Tab

le 4

. Arc

haea

l ple

omor

phic

vir

uses

in th

e fa

mily

Pleo

lipov

irid

ae.

Gen

us

Vir

usa

Isol

atio

n h

ost

Vir

ion

dia

me

ter

Lif

e cy

cle

Mem

bran

ep

rote

in,

spik

ep

rote

in

Gen

ome,

Gen

eBan

kac

cess

ion

nu

mbe

r

Isol

atio

n s

ourc

eR

efer

ence

s

Alp

ha-

pleo

lipo-

viru

s

HR

PV

-1H

alor

ubru

msp

.PV

6~

41 n

mPe

rsis

tent

VP3

, VP4

Cir

cula

r ss

DN

A,

7,04

8 nt

,FJ

6856

51

Sola

r sa

lter

n,Tr

apan

i, It

aly

Piet

iläet

al.,

200

9,Pi

etilä

et a

l., 2

010,

Piet

ilä e

t al.,

201

2H

RPV

-2H

alor

ubru

msp

.SS5

-4~

54 n

mPe

rsis

tent

VP4

,VP5

Cir

cula

r ss

DN

A,

10,6

56 n

t,JN

8822

64

Sola

r sa

lter

n, S

amut

Sakh

on, T

haila

ndA

tana

sova

et a

l., 2

012,

Piet

ilä e

t al.,

201

2,Se

nčilo

et a

l., 2

012

HR

PV-6

Hal

orub

rum

sp.S

S7-4

~49

nm

Pers

iste

ntV

P4, V

P5C

ircu

lar

ssD

NA

,8,

549

nt,

JN88

2266

Sola

r sa

lter

n,Sa

mut

Sakh

on, T

haila

ndPi

etil ä

et a

l., 2

012,

Senč

ilo e

t al.,

201

2

HH

PV-1

Hal

oarc

ula

hisp

anic

a~

52 n

mPe

rsis

tent

VP3

, VP4

Cir

cula

r ds

DN

A,

8,08

2 bp

,G

U32

1093

Sola

r sa

lter

n,M

argh

erit

a di

Sav

oia,

Ital

y

Roi

neet

al.,

201

0,Se

nčilo

et a

l., 2

012

HH

PV-2

Hal

oarc

ula

hisp

anic

a~

50nm

Pers

iste

ntO

RF3

pro

duct

,O

RF4

pro

duct

Cir

cula

r ss

DN

A,

8,17

6 nt

,K

F056

323

Sola

r sa

lter

n, H

ulu

Isla

nd, L

iaon

ing,

Chi

na

Liet

al.,

201

4

Bet

a-pl

eolip

o-vi

rus

HR

PV

-3H

alor

ubru

msp

.SP3

-3~

67 n

mPe

rsis

tent

VP1

, VP2

Cir

cula

rds

DN

Ab ,

8,77

0 bp

,JN

8822

65

Salt

wat

er,S

edom

Pond

s, I

srae

lA

tana

sova

et a

l., 2

012,

Piet

ilä e

t al.,

201

2,Se

nčilo

et a

l., 2

012

HG

PV-1

Hal

ogeo

met

ricu

msp

. CG

-9~

56 n

mPe

rsis

tent

VP2

and

VP3

,V

P4C

ircu

lar

dsD

NA

b ,9,

694

bp,

JN88

2267

Sola

r sa

lter

n, C

abo

deG

ata,

Spa

in

SNJ2

Nat

rine

ma

sp. J

7-1

70-8

0nm

Lyso

geni

cV

P12,

VP1

3C

ircu

lar

dsD

NA

b,c ,

16,9

92 b

p,A

JVG

0100

0023

Nat

rine

ma

sp. J

7-1

Liu

et a

l., 2

015

Gam

ma-

pleo

lipo-

viru

s

His

2H

aloa

rcul

ahi

span

ica

~71

nm

Pers

iste

ntV

P27,

VP2

8an

d V

P29

Line

ar d

sDN

A,

16,0

67 b

p,A

F191

797

Salin

e la

ke(P

ink

Lake

s), V

icto

ria,

Aus

tral

ia

Bat

het

al.,

200

6,Pi

etilä

et a

l., 2

012

a.Ty

pe s

peci

es a

re in

bol

d.b.

Bet

aple

olip

ovir

us g

enom

es a

re d

sDN

A m

olec

ules

with

sin

gle-

stra

nded

inte

rrup

tion

s.c.

SNJ2

gen

ome

corr

espo

nds

to th

e re

gion

inN

atri

nem

a sp

. J7-

1 ch

rom

osom

e (W

GS

cont

ig04

, nt 1

9,79

2-36

,797

).

21

1.6 Virus-host interactions in hypersaline environmentsThe number of VLPs in hypersaline environments is high (up to 109-101o VLP/ml), increasingalong the salinity gradient (Bettarel et al., 2011, Boujelben et al., 2012, Guixa-Boixareu etal., 1996, Santos et al., 2012). Moreover, in many hypersaline reservoirs, free virusessignificantly outnumber cells (up to 100 VLP/cell) (Santos et al., 2012). In the saltiest ponds(above 25 % total salinity), viruses are usually the only predators controlling prokaryoticgrowth rates and mortality (Guixa-Boixareu et al., 1996, Pedros-Alio et al., 2000). Asrevealed by electron microscopy, the major virus morphotypes in hypersaline environmentsare spindle-shaped, spherical, tailed, and filamentous (Boujelben et al., 2012, Guixa-Boixareu et al., 1996, Sime-Ngando et al., 2011). Spindle-shaped VLPs are often reported asthe most abundant ones, and their abundance is increasing along the salinity gradient,correlating with the abundance of square cells of Hqr. walsbyi (Bettarel et al., 2011, Guixa-Boixareu et al., 1996, Sime-Ngando et al., 2011). Surprisingly, VLPs of very unusual shapeswere observed in hypersaline Lake Retba (Senegal): stars, rods, hooks, and hairpins amongothers (Sime-Ngando et al., 2011). It is, however, questionable whether these unusual shapesrepresent real virions or TEM-related artefacts. In addition, it is important to remember thatmicroscopy examinations, especially those of environmental samples, may not be accurate,because along with virus particles, similar sized extracellular vesicles (Forterre et al., 2013,Soler et al., 2015) or recently discovered nanohaloarchaeal cells may be counted as sphericalVLPs.

Culture-independent approaches, such as pulsed-field gel electrophoresis of total viralDNA, as well as metagenomics, metatranscriptomics, and metaproteomics, have beenapplied to assess the diversity and activity of viral communities in hypersalineenvironments. In general, these studies have shown that halovirus communities are diverseand specific for hypersaline environments (Boujelben et al., 2012, Díez et al., 2000,Emerson et al., 2013b, Roux et al., 2016, Sandaa et al., 2003). However, the literaturecontains contradictory reports about (i) temporal and (ii) spatial variation in haloviruscommunities:

(i) Temporal: Metagenomic analyses of water samples from San Diego salterns takenat different time points have shown that the most abundant viruses, as well as the mostabundant microbial species, were stable, while the variation in microdiversity (changing ofgenotypes) was observed (Rodriguez-Brito et al., 2010). Emerson et al. (2013a, 2012)reanalysed the same dataset and suggested that virus community was actually dynamic atthe taxon level and that variation observed at the genotype level in fact reflected shifts at thepopulation level. In a set of studies, Emerson and colleagues analysed metaviromes fromLake Tyrell samples collected over different timescales, and concluded that haloviruspopulations were stable over days, relatively equally stable and dynamic over weeks, anddynamic over months to years (Emerson et al., 2013a, Emerson et al., 2012, Emerson et al.,2013b).

(ii) Spatial: Several studies have suggested that common traits can be found in themetaviromes reconstructed from distant hypersaline environments (Garcia-Heredia et al.,2012, Roux et al., 2016, Santos et al., 2010, Sime-Ngando et al., 2011). For example, commonfeatures were identified in the metaviromes from hypersaline ponds in Senegal, Australia,

22

and Spain, suggesting that virus communities are stable across continents (Roux et al.,2016). However, the usage of different methods for the preparation of environmental DNA,different sequencing techniques, and sequence analyses tools impose challenges for directdata comparisons. The lack of universal virus genes also makes it difficult to makecomparisons across samples. In addition, different conclusions regarding similarity may bedrawn depending on whether nucleotide or protein sequences are compared. For example,in the above mentioned study of Lake Tyrell virus community dynamics, the viromes werecompared at the nucleotide sequence level, with the threshold of 90% identity, andconsidered to be highly dynamic over months to years (Emerson et al., 2012). When thesame datasets were used in a large comparison of ~40 viromes and the similarity wasassessed at the protein sequence level, the Lake Tyrell viromes were closely relatedcompared to the other datasets (Roux et al., 2016). In addition, the stringency of comparisonhas to be taken into account, as demonstrated in the same study of the Lake Tyrell viruscommunity: when the fragment recruitment detection threshold was reduced from 90% to75% nucleotide identity, some previously unseen similarities to the sequenced halovirusesand San Diego metavirome sequences were detected (Emerson et al., 2012).

Typically, a high percentage of environmental virus sequences do not match to thedatabases (Boujelben et al., 2012, Santos et al., 2010, Sime-Ngando et al., 2011). However,some studies have reported sequence similarities between metaviromes and the isolatedhaloviruses. Metagenomic analyses of the samples from hypersaline ponds in Senegalrevealed sequence matches to the isolated haloarchaeal tailed viruses, pleolipoviruses, andHis1, as well as environmental halophages assembled from Santa Pola salterns (Roux et al.,2016). In metaproteomics studies of the virus community in Deep Lake, Antarctica, eightdistinct MCPs were detected, five matching to the known isolated haloarchaeal siphoviruses,two to the putative tailed environmental halophages, EHP-12 and EHP-6, and one to Vibriophage VBM1 (Tschitschko et al., 2015). In addition to the MCPs, a prohead protease wasdetected in the metaproteomic dataset, indicating active virus life cycles in Deep Lake(Tschitschko et al., 2015). The activity of viruses in the environment may also be assessedby metatranscriptomics. For instance, total environmental RNA from Santa Pola salternswas analysed using microarrays constructed based on the previously obtained metaviromefrom the same environment (Santos et al., 2011). This work has shown that halovirus geneswere actively expressed at the sampling time and that the most active viruses were those thattentatively infect haloarchaea and Salinibacter strains (Santos et al., 2011).

Low number of matches between metagenomic sequences and databases indicates thatonly a small part of haloviruses have been characterized. A global sampling of distanthypersaline environments performed in 2012 doubled the number of isolated haloarchaealviruses, and revealed numerous virus-host pairs within and across samples (Atanasova etal., 2012). Although the analysis of environmental sequences allows reconstruction ofcomplete or almost complete virus genomes and tentative identification of virus-host pairs(Garcia-Heredia et al., 2012, Martínez-García et al., 2014, Santos et al., 2007), moleculardetails of virus-host interactions may be studied in depth only in isolated virus-host systems.Obviously, rapidly accumulating sequence data needs functional annotation, and in thisrespect, cultivation-based studies cannot be neglected.

23

Although hypersaline environments, such as solar salterns, may be seen as simplifiedsystems for ecological studies due to low species diversity, high cell abundance, and shortfood chains, there is still little understanding of virus-host interactions in hypersalineeconiches. Multiple factors may affect virus-host interplay. It has been shown that salinityaffects the adsorption of haloviruses to the host cells (Kukkaro and Bamford, 2009, Mei etal., 2015), as well as regulates switches between lytic and lysogenic virus life cycles (Bettarelet al., 2011, Daniels and Wais, 1990, Mei et al., 2015). Another important aspect of virus-host interplay is the ability of host cells to resist virus attacks. Virus defense systems arecommonly found in archaea and include clustered regularly interspaced short palindromicrepeats (CRISPR) and CRISPR-associated proteins (Cas) immune system (Garrett et al.,2011, Sorek et al., 2008), toxin-antitoxin system (Yamaguchi et al., 2011), superinfectionexclusion mechanism (Stolt and Zillig, 1992), and bacteriophage exclusion (BREX)mechanism (Goldfarb et al., 2015). Finally, more spatial and temporal studies in hypersalineenvironments are needed to test the current models of virus-host dynamics, such as Lotka-Volterra “kill the winner” (Thingstad, 2000), constant-diversity dynamics (Rodriguez-Valera et al., 2009), or recently proposed “piggyback the winner” (Knowles et al., 2016). Toconclude, only coupling of various approaches to study virus-host interplay would provide acomplete view on the true virus diversity, as well as temporal and spatial dynamics inhypersaline environments.

24

2. Aims of the StudyIt has been estimated that viruses are present virtually in all types of environments that onecan imagine, being a significant component of the biosphere (Breitbart and Rohwer, 2005,Dolja and Krupovič, 2013, Forterre and Prangishvili, 2013, Rohwer et al., 2009). Virusesdisplay huge genetic diversity, but more modest variety of virion shapes (Krupovič andBamford, 2011). One way to comprehend the diversity in the virosphere is to percept virionas a “vertically-inherited” viral “self” (Abrescia et al., 2010, Krupovič and Bamford, 2010).It has been suggested that only a limited number of virion architectures exist, and deepevolutionary links have been found between viruses that infect hosts from all three domainsof cellular life (Abrescia et al., 2012, Bamford et al., 2002, Bamford, 2003, Oksanen et al.,2012). The key to understand ultimate viral diversity, virus functions and complexevolutionary relations of different virus groups is to isolate new virus-host systems and tocharacterize them in detail. Hypersaline environments, where archaea dominate, are a richsource of new viruses, and only very little is known about archaeal viruses. This thesis isdevoted to the isolation of new haloarchaeal viruses in order to probe their morphologicaldiversity. Virus-host interactions in a hypersaline environment are also addressed. Inaddition, comprehensive descriptions of two new isolates are provided.

The specific aims of this work were:

- To isolate new archaeal viruses from high salinity environments and to determine theirmorphotypes;

- To reveal virus-host interactions in high salinity environments using culture-basedapproach;

- To describe certain isolated viruses in molecular detail, including characterization ofvirus infection cycle and virion components.

25

3. Materials and MethodsSalt samples used in this study included saltwater and salt crystals collected from a solarsaltern in Samut Sakhon (Thailand) in 2009 and 2010 (Table S1 in I). Fifty one halophilicprokaryotic strains isolated from the salt samples (Tables 1 and S2 in I), as well as elevenculture collection strains (Table 1 in I) were used here. The studied viruses included 45isolates listed in publication I (Tables 2 and S3 in I) and Haloarcula hispanica pleomorphicvirus 3 (HHPV3) described in publication III. Methods and bioinformatics tools used in thiswork are listed in Table 5 and Table 6, respectively. Virus genome sequences were depositedin the GenBank database. For more detailed information about the methods, please see thepublications I-III and references therein.

Table 5. Laboratory methods used in this study.

Method Used inCultivation of halophilic prokaryotes and their viruses I, II, IIIPlaque assay I, II, IIIIsolation of halophilic prokaryotes from salt samples IIsolation of archaeal viruses from salt samples I, IIISpot-on-lawn test for virus infectivity I, IIIVirus adsorption assay II, IIIVirus life cycle analysis II, IIIElectrochemical and ATP measurements during virus infection IINegative staining and transmission electron microscopy I, II, IIIThin-section TEM IIPurification of virus particles by NaCl-polyethylene glycol 6000precipitation, rate zonal and equilibrium centrifugation

I, II, III

Bradford assay for measuring protein concentration I, IISodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) I, II, IIICoomassie blue staining of gels I, II, IIIPr0-Q Emerald, SYPRO-Ruby, and Sudan Black B staining of gels II, IIIIdentification of proteins by N-terminal sequencing IIIdentification of proteins by mass spectrometry II, IIIExtraction of lipids and thin-layer chromatography (TLC) II, IIIDNA extraction and enzymatic digestion II, IIIAgarose gel electrophoresis and ethidium bromide staining II, IIIPolymerase chain reaction (PCR) I, IIISanger sequencing I, II, IIIIllumina sequencing III

26

Table 6. Bioinformatic tools used in this study.

Type of analysis Tool(s) Used inPhylogenetic analysis Molecular Evolutionary Genetics

Analysis (MEGA)I, II, III

Sequence alignment Multiple Sequence Comparison byLog-Expectation (MUSCLE)

I, II, III

EMBOSS Stretcher, EMBOSS Needle II, IIIVisualization, transfer, andalignment of sequences

Geneious v. 6.1.6 by Biomatters I, II, III

Prediction of open reading frames(ORFs)

GenMarkS, Glimmer II, III

Prediction of replication origin Ori-Finder 2 IIICalculating of guanine-cytosinecontent

Science Buddies “Genomics %G~CContent Calculator”

II, III

Calculating of isoelectric point andmolecular mass

Compute pI/MW tool of ExPASy tools II, III

Prediction of tRNA sequences ARAGORN v1.2.36 IIDetection of repeats Tandem Repeats Finder, EMBOSS

EinvertedII, III

Homology detection Basic local alignment search tool(BLAST)

II, III

Detection of conserved domains NCBI Conserved Domain Search II, IIIInterProScan at EMBL-EBI III

Prediction of transmembranehelices

TMHMM v 2.0 II, III

Prediction of coiled-coil regions COILS tool II, IIIGraphical visualization Circos IIPrediction of signal peptides Signal P v 4.1 III

27

4. Results and Discussion4.1 New archaeal viruses isolated from high salinityIn order to isolate new archaeal viruses, salt samples (Table SI in I) were collected from asolar saltern in Samut Sakhon, Thailand in 2009 and 2010. From these samples, halophilicprokaryotes were isolated: 36 archaeal and 15 bacterial strains (Table 1 in I). Based on theirpartial 16S rRNA gene sequences, the strains were classified at genus level. Bacterial strainsbelonged to four genera of two phyla: Proteobacteria (Rhodovibrio) and Firmicutes(Halobacillus, Pontibacillus, and Sedimibacillus). The isolated archaea were assigned to tengenera of Halobacteriaceae family, with Halorubrum and Haloarcula strains being themost numerous (10 and 6, respectively). The strains of these genera are commonly reportedas prevalent in culture-based samplings of hypersaline environments (Birbir et al., 2007,Boujelben et al., 2014, Burns et al., 2004). Using the isolated archaeal strains, as well as ourculture collection strains (Table 1 in I), we isolated 46 archaeal viruses from the abovementioned salt samples (Tables 2 and S3 in I; publication III, see chapter 4.4), for 37 ofwhich purification protocols were developed and virion morphotype was determined byTEM. The 37 purified viruses differed one from another by plaque morphology (Figures S1in I and S1 in III), virion morphotype (Figures 2 in I, 5 in II, and 3 in III), sensitivity of virusinfectivity to chloroform (Table 2 in I, and chapter 4.4), host ranges (Tables S4 and S5 in I,and chapter 4.4), and structural protein patterns of purified virus particles (Figures S2 in I,2 in II, and 4 in III) and thus were considered unique. These isolates were named accordingto their isolation host and morphology (Table 2 in I, and chapter 4.4). The isolated virusesdisplayed four different morphotypes: myovirus (27 isolates), siphovirus (4), pleomorphic(5) and tailless icosahedral (1) (Figures 2 in I, 5 in II, and 3 in III).

Among 37 isolated viruses with known morphotypes (I, III), one pleomorphic virus(HRPV-6) had been described in detail elsewhere (see chapter 1.5.2.4) (Pietilä et al., 2012,Senčilo et al., 2012). Thus, in this thesis, only 36 isolated viruses are counted as new oneswhen compared to the number of all haloarchaeal viruses reported in the currently availableliterature (see chapter 1.5.2.4). However, HRPV-6, which had been isolated from the samelocation during the same sampling, was included in virus-host interactions tests, so I willdiscuss the interactions of all 37 isolates in chapter 4.2.

The isolation of new viruses has increased the total number of isolated archaealhaloviruses from 54 to 90 (Figures 2 A and B). Noticeably, the morphotype distributionobserved in this study is highly similar to that in the previous global sampling of distanthypersaline environments (Atanasova et al., 2012), with tailed icosahedral virus isolatesbeing the most numerous. Obviously, the isolated viruses represented only a part of naturalviral diversity in the studied environment, as a culture-based approach is limited to thestrains that can be cultivated on artificial growth media and viruses that form detectableplaques on host lawns. However, this does not diminish a possible role of the isolated virusesin the dynamics of halophilic microbial communities (see chapter 4.2 for virus-hostinteractions). So far, culture-dependent samplings of viral diversity in hypersalineenvironments yielded a relatively low number of different virus morphotypes (Atanasova etal., 2015a), supporting the idea that all viruses can be grouped into a limited number of

28

structural lineages (Bamford et al., 2002, Oksanen et al., 2012). Future isolations of virus-host pairs from hypersaline environments will show if haloarchaeal viruses with virions ofunusual shape, e.g. possibly resembling those of thermophilic crenarchaeal viruses(Prangishvili et al., 2016) or the particles observed in TEM examinations of environmentalsamples (Sime-Ngando et al., 2011), also exist.

Figure 2. Distribution of morphotypes of isolated haloarchaeal viruses (A) excluding and (B)including current study. Total numbers of isolates of a given morphotype are indicated in brackets.

4.2 Virus-host interactionsTo reveal virus-host interactions, all isolated viruses were tested against all isolatedhalophilic prokaryotes (except Halorubrum sp. SS13-13 and Halogranum sp. SS13-5 whichdid not form a dense lawn) and additional archaeal culture collection strains. For 37 viruseswith known morphotypes, we detected 269 specific interactions across genera and samples,all archaea-specific (Tables S4 and S5 in I; publication III, see chapter 4.4). In addition, 73interactions were observed for nine virus isolates of unknown morphotype (Table S3 in I).Further, I discuss the interactions of the 37 unique virus isolates (Table 2 in I and publicationIII).

Originally, all the viruses were isolated on Halorubrum or Haloarcula strains, butcross-testing showed that strains from other genera (Halobacterium, Halobellus, andHaloterrigena) were also susceptible to the isolated viruses. However, no viruses wereisolated for Halolamina, Halogranum, Haloferax, Halogeometricum, and Natrinemastrains. Myoviruses had the broadest host ranges (from 3 to 14 hosts, 9 on average), andsome Halorubrum strains were susceptible to many viruses (up to 26 different viruses).These two components accounted for the majority (~75%) of the interactions that wereobserved. Viruses of the other morphotypes had narrower host ranges.

Viruses isolated from Samut Sakhon saltern were able to infect the strains isolatedfrom the same location (from both 2009 and 2010 samples), as well as culture collectionstrains originating from geographically distant hypersaline environments (Tables S4 and S5in I and publication III). For example, Halorubrum sodomense isolated from the Dead Seawaters (Israel) (Oren, 1983) could be infected by 18 myoviruses, and “Haloarcula

29

californiaea” isolated from coastal salt deposits in Baja California (Mexico) (Javor et al.,1982) was a host for 12 viruses (11 myoviruses and one icosahedral virus). These diverseinteractions indicate a constant interplay between haloarchaea and their viruses (especiallymyoviruses) extending over time, genus, and location “barriers” and are in line with theprevious observations of global distribution of microbes and viruses in hypersalineenvironments (Atanasova et al., 2012, Dyall-Smith et al., 2011, Oh et al., 2010, Roux et al.,2016).

To gain more insights into viruses inhabiting hypersaline environments, here, Iprovide a detailed molecular characterization of two isolates obtained from Samut Sakhonsaltern: Haloarcula californiae icosahedral virus 1 (HCIV-1, see chapter 4.3) andHaloarcula hispanica pleomorphic virus 3 (HHPV3, see chapter 4.4).

4.3 Molecular characterization of HCIV-1

4.3.1 HCIV-1 infection cycleAlthough HCIV-1 had been isolated from salt crystals (I), it tolerated a wide range ofsalinities. No major changes in infectivity were detected in a range of 3 to 23 % total salinityfor a 24-h period tested (Figure S1 in II). The ability to tolerate high and low ionic strengthmay be advantageous in such environments as a solar saltern, where salinity is changing dueto natural rainfalls and water evaporation.

HCIV-1 was isolated on Har. californiae, but could also infect Har. japonica withapproximately same efficiency of plating, and with much lower efficiency, Har. hispanicaand Halorubrum sp. SS7-4 (I). It seems that a narrow host range is typical for all currentlyknown haloarchaeal tailless icosahedral viruses (see chapter 1.5.2.3). The adsorption ofHCIV-1 particles to Har. californiae cells was relatively slow, but efficient: at ~1.5 h postinfection (p.i.), ~45 % of virus particles adsorbed to the cells, and by ~5 h p.i, the maximumof 80% adsorption was reached (Figure S2A in II). The adsorption rate constant was 5.7 ×10-11 ml/min (calculated for the first h p.i.), which is in a range typical for halophilic archaealviruses (Jaakkola et al., 2012, Kukkaro and Bamford, 2009, Svirskaitė et al., 2016). At 2.5 hp.i., tail-like structures between Har. californiae cells and icosahedral virus particles werevisualized by TEM (Figures 1 B and C in II). To my knowledge, similar structures have notbeen reported for other tailless archaeal viruses, but were described for tailless icosahedralbacteriophages, e.g. PRD1 (Peralta et al., 2013) and ΦX174 (Sun et al., 2014), as the “devices”for genome delivery.

The single-step growth curve of HCIV-1 infection showed that at 10-12 h p.i., theturbidity of the infected culture started to decrease (Figures 1 A and G in II), while thenumber of free progeny viruses was rapidly increasing (Figure 1G in II). At the same time,cell debris were accumulating, as revealed by light (Figure S2 B in II) and electronmicroscopy (Figure 1 E in II). In addition, the leakage of ATP from the cells and binding ofa lipophilic indicator anion (phenyldicarbaundecaborane) to the cells occurred also at ~10 hp.i., indicating that cell integrity was compromised (Figures 1 F and H in II). Moreover, thelevel of oxygen consumption was significantly decreasing compared to the uninfectedculture (Figure 1 I in II). All this suggested that assembled HCIV-1 particles exit cells via cell

30

lysis. The same type of study approach has been applied to SH1, showing that the virus isvirulent (Svirskaitė et al., 2016). However, in SH1 infection, cell lysis occurs earlier (at ~6 hp.i.), which probably reflects specificity of virus-host interactions. We could not identify anylysis-related genes in HCIV-1 genome (see below). However, this is not surprising takinginto account that no lysis-related genes similar to those of bacteriophages are found in thegenomes of archaeal viruses. Moreover, the only currently described archaeal virus exitstrategy is that of viruses STIV and Sulfolobus islandicus rod-shaped virus 2 (SIRV2)(Snyder et al., 2013). These viruses exit Sulfolobus cells through pyramide-like openings,not observed for any other virus-host systems (Snyder et al., 2013).

4.3.2 HCIV-1 virion organizationAn optimized multistep purification protocol was used to obtain highly purified HCIV-1particles (Table S1 and Figure 2 in II). The major structural components of HCIV-1 virionwere determined (Figures 2-5 in II). TEM of highly purified particles revealed that theicosahedral HCIV-1 virion was ~70 nm in diameter and contained an inner layer(membrane) (Figure 4). The virion consisted of at least 12 protein species (Figures 2 and 4A in II), which were designated according to their amino acid sequence similarities (seebelow) to the corresponding proteins of SH1, PH1, and HHIV-2. Notably, the overallstructural protein profiles were similar, but not identical in HCIV-1, SH1, and HHIV-2(Figure 4 A in II). HCIV-1 lipid profile consisted of three major phospholipid species:phosphatidylglycerol, phosphatidylglycerophosphate methyl ester, andphosphatidylglycerosulfate (Figure 3 in II). HCIV-1 and Har. californiae lipid compositionsdiffered (Figure 3 in II), suggesting that HCIV-1 acquires lipids selectively from the host cellmembrane. A selective acquisition of the host phospholipids has been previously reportedfor other tailless inner membrane-containing icosahedral dsDNA viruses, e.g. SH1 (Bamfordet al., 2005b), HHIV-2 (Gil-Carton et al., 2015), PRD1 (Laurinavičius et al., 2004b), Bam35(Laurinavičius et al., 2004b), as well as for the enveloped tailless icosahedral dsRNAbacteriophage Φ6 (Laurinavičius et al., 2004a).

Based on the sequence similarities with SH1 and HHIV-2, for which virion structureshave been resolved (Gil-Carton et al., 2015, Jäälinoja et al., 2008), we propose the followingvirion model for HCIV-1. HCIV-1 proteins VP4 and VP7 are the MCPs, while proteins VP2,VP3 and VP4 form the vertex complex, and proteins VP10 and VP12 are associated with theinner membrane (Figure 4). Based on the similar sizes of SH1, HHIV-2, and HCIV-1 virions,HCIV-1 capsid lattice could also be arranged in T = 28 (Gil-Carton et al., 2015, Jäälinoja etal., 2008). Moreover, HCIV-1 virion assembly model might mimic that described for HHIV-2 (Gil-Carton et al., 2015) (see chapter 1.5.2.3). We suggest that the small MCP VP7, whichhas one single β-barrel domain, and the large MCP VP4, which has two single β-barreldomains, form pseudo-hexameric capsomers in HCIV-1 capsid. It has been proposed that inHHIV-2, disulphide bridges formed between cysteine residues in the MCPs stabilize thecapsid lattice (Gil-Carton et al., 2015). However, in HCIV-1, only VP7 has the conservedcysteine residue. It remains to be elucidated how the MCPs are linked in HCIV-1 capsid andwhat mechanisms are used to stabilize the lattice.

31

The HCIV-1 genome is a linear dsDNA molecule of 31,314 bp long (Genbank acc. no.KT809302) (II). The genome was predicted to contain 47 open reading frames (ORFs), tenof which were confirmed to be genes encoding structural proteins (Figure 4 C and Table S2in II). A BLASTx search against the NCBI n.r. database found that 43 out of the 47 ORFs hadmatches to haloarchaeal viral sequences, and the majority of these hits were toalphasphaerolipoviruses SH1, PH1, and HHIV-2 (Table S2 in II). The overall gene syntenyand sequence similarities suggested that these viruses are closely related (Figure 3, FigureS5 in III). In addition, a BLAST search yielded hits to haloarchaeal sequences (Table S5 inII), including previously known alphasphaerolipovirus-related proviruses HaloLakP1 andHalaPauP1 (Pawlowski et al., 2014, Porter et al., 2013). We also identified a new putativeHCIV-1-related provirus in Haladaptatus cibarius (NZ_JDTH01000002.1, nt 488,665-496,650) and designated it as HalaCibP1 (Table S5 in II).

Figure 3. The genomes of HCIV-1 and alphasphaerolipoviruses SH1, PH1, and HHIV-2. ORFs/genesare shown as grey arrows, and homologous ones encoding (putative) structural proteins are in thesame colours. Inverted terminal repeats (ITRs) are shown as light blue arrows. ORFs/genes productsare indicated next to the coding regions. The structural protein names (VPs) are the same for all fourviruses, except spike proteins (asterisk): VP3 and VP6 for SH1, PH1, and HCIV-1, but VP16 andVP17 for HHIV-2. Amino acid similarities (%) between proteins VP12, MCPs and putative ATPasesare shown in between the genomes.

HCIV-1 putative protein 13 is predicted to be an ATPase, as it contains canonicalWalker A and B motifs (Walker et al., 1982), as well as the P9/A32-specific motif found inATPases of the viruses belonging to the PRD1-adenovirus structural lineage (Strömsten etal., 2005). The core capsid proteins, MCPs and VP12, as well as putative ATPases, are themost conserved proteins in HCIV-1 and alphasphaerolipoviruses (Figure 3), which is in linewith the idea of conserved viral “self” elements. On the contrary, vertex complex proteins,most likely involved in host cell recognition, have the lowest level of similarity (II) (Jaakkolaet al., 2012). HCIV-1 putative spike proteins, VP3 and VP6, are similar to those of SH1, notHHIV-2, and thus expected to have a horn-like shape, as in SH1 (Jaakkola et al., 2012,Jäälinoja et al., 2008).

Based on all of these characteristics, HCIV-1 belongs to Alphasphaerolipovirus genusof the Sphaerolipoviridae family, which comprises icosahedral dsDNA viruses infectingextremophilic archaea or bacteria (Pawlowski et al., 2014). The phylogenetic analyses of

32

MCPs and ATPase sequences of sphaerolipoviruses, as well as related putative proviruses,indicated that HCIV-1 clusters with the group of alphasphaerolipoviruses (Figure 6 in II).Our proposal to classify HCIV-1 as a new species (Haloarcula virus HCIV1) in the genusAlphasphaerolipovirus has been recently approved by ICTV. In respect to the structure-based virus classification, the conserved structural elements responsible for virionorganization in HCIV-1 indicate that this virus belongs to the “single β-barrel” subgroup ofPRD1-adenovirus structural lineage (Gil-Carton et al., 2015, Oksanen et al., 2012).Amazingly, four of the five currently known haloarchaeal tailless icosahedral viruses appearto be closely related, despite being isolated from geographically distant locations (Figure 4).In terms of taxonomy, eukaryotic, bacterial, or archaeal viruses that are distributed world-wide in moderate and extreme environments and which are currently assigned to a numberof different families may be related by the principles of virion architecture and thus mayconstitute just one structural lineage (Abrescia et al., 2012, Oksanen et al., 2012). Forexample, PRD1-adenovirus lineage unites the viruses currently classified to the familiesTectiviridae (PRD1, Bam35), Adenoviridae (adenovirus), Corticoviridae (PM2),Turriviridae (STIV and STIV2), and Sphaerolipoviridae (HCIV-1, SH1, PH1, HHIV-2, SNJ1,P23-77) (Abrescia et al., 2012).

Figure 4. HCIV-1 virion schematic model and global distribution of alphasphaerolipoviruses(geographic locations from which viruses were isolated are indicated). Map source: Wikimediacommons.

33

4.4 Molecular characterization of HHPV3

4.4.1 HHPV3 infection cycleHHPV3 was isolated from salt crystals on Har. hispanica and could not infect any otherhaloarchaeal strains of the 47 tested ones (III). The adsorption of HHPV3 particles to Har.hispanica cells reached ~75 % in 2 h (Figure 1 A in III), and the adsorption rate constant was2.4×10-12 ml/min (calculated for the first 30 min p.i.). Starting at ~6 h p.i., the titer of freeprogeny viruses was increasing and reached the maximum of ~8×1010 plaque forming units(PFU)/ml by ~13 h p.i. (Figure 1 B in III). Approximately at the same time (~12 h p.i.), theturbidity of the infected culture stopped increasing and remained at the same level,indicating growth retardation, while the uninfected culture continued growing. Starting at~6 h p.i., the number of viable cells was decreasing, and by ~12 h p.i. the difference in viablecell counts between the infected and uninfected cultures was about two orders of magnitude(Figure 1 C in III). Thus, although HHPV3 life cycle seems to be non-lytic, the infectionresults in significant growth retardation. Very narrow host ranges, and similar non-lyticinfection cycles with various degrees of host growth retardation have been reported forpleolipoviruses (see chapter 1.5.2.4), which presumably enter the cell via fusion of the virionand the cell membrane and exit the cell via budding (Pietilä et al., 2016, Svirskaitė et al.,2016).

4.4.2 HHPV3 infectivity at different salinitiesTesting the stability of HHPV3 infectivity in different salt solutions (Figures 2 and S3 in III)revealed that CaCl2 and NaCl are critical for the maintenance of HHPV3 infectivity.Noticeably, no major changes in HHPV3 infectivity were observed when incubated in abuffer containing 3 to 5.5 M NaCl over a 24-h period (Figure 2 A in III). The solution of 5.5M NaCl was saturated and contained salt crystals. The fact that HHPV3 was isolated fromsalt crystals attests to its ability to survive in saturated salt. However, at 1.5 M NaCl, HHPV3demonstrated a peculiar, irreversible, pH-dependent loss of infectivity (Figure 2 A in III).NaCl may affect the virus isoelectric point, as shown for some bacteriophages (Langlet et al.,2008). Possibly, changes in HHPV3 virion net charge resulted in clustering of virus particlesat certain pH.

Here, we also tested the effect of NaCl concentration on the infectivity of thepleolipoviruses HRPV-1 (Pietilä et al., 2009, Pietilä et al., 2010, Pietilä et al., 2012),Halogeometricum pleomorphic virus 1 (HGPV-1), and Halorubrum pleomorphic virus 3(HRPV-3) (Atanasova et al., 2012, Pietilä et al., 2012, Senčilo et al., 2012). For HRPV-1, theconcentration of 2-3 M NaCl was optimal, while HGPV-1 and HRPV-3 demonstrated nosignificant infectivity changes when incubated in their optimal buffers containing from 0 to5.5 M (saturated) NaCl. Such maintenance of high infectivity over a full range of NaClconcentrations observed in betapleolipoviruses HGPV-1 and HRPV-3 indicates their highresistance to natural salinity changes that occur in hypersaline environments.Unfortunately, very little is known about virus survival strategies in extreme conditions. Itwould be intriguing to see how virus particles can cope with osmotic stress, lacking anymeans to actively regulate their interactions with the surrounding environment. Theproperties of a pleolipoviral virion, a flexible membrane vesicle decorated by spike proteins

34

(Pietilä et al., 2010, Pietilä et al., 2012), may be a key to survival. However, more insightsinto virion structural organization are needed to understand what features underlie thedifferences observed in the adaptability of different pleolipoviruses to salinity changes.

4.4.3 HHPV3 virion organizationHighly purified HHPV3 particles were obtained with optimized purification protocols(Figure 3 A and Table S3 in III) and further used for the analyses of virion components.HHPV3 particles are spherical, but slightly pleomorphic, ~50 nm in diameter (Figure 3 B inIII). This type of pleomorphic morphology is characteristic to a group of haloarchaealpleolipoviruses (Pietilä et al., 2016). HHPV3 has two major viral structural protein species,which are not glycosylated, and a lipid fraction (Figure 4 in III). TLC of the lipids extractedfrom HHPV3 particles or Har. hispanica cells (Bamford et al., 2005b) showed that the virusand the host had the same lipid species in approximately same proportions (Figure 5 in III),indicating a non-selective mode of lipid acquisition by HHPV3 from the host membrane.Pleolipoviruses are considered to acquire lipids non-selectively, although only His2 and itshost, Har. hispanica, display identical lipid profiles, while in lipid compositions of the othervirus-host pairs minor differences are observed, with virus lipids lacking one or more hostlipid species (Liu et al., 2015, Pietilä et al., 2012). Thus, some degree of selectivity might becharacteristic to certain pleolipoviruses.

The HHPV3 genome is a circular 11,648 bp-long dsDNA molecule (Genbank acc. no.KX344510) (Figures S4 A and B in III). Fragmentation patterns observed after HHPV3 DNAwas treated with Mung Bean endonuclease and Bal-31 exonuclease (Figures S4 C and D inIII) suggest that the dsDNA molecule might have nicks or ssDNA regions. This type ofgenome organization (discontinuous dsDNA) has been previously reported forbetapleolipoviruses HGPV-1, HRPV-3, and SNJ2 (Liu et al., 2015, Senčilo et al., 2012). InHRPV-3 and HGPV-1, the discontinuities are ssDNA regions (Senčilo et al., 2012), while inSNJ2, both nicks and ssDNA regions are found (Liu et al., 2015). In HRPV-3, theinterruptions are located mainly on the coding strand, while in SNJ2, the gaps aredistributed on both strands. Both in HRPV-3 and SNJ2, the discontinuities are present incoding and in intergenic regions, and a specific motif GCCCA often precedes them (Liu etal., 2015, Senčilo et al., 2012). This motif can also be found in HHPV3 genome on bothstrands (Figure 6 A in III). Site-specific nicks in genomic DNA have also been reported forsome bacteriophages, e.g. T5 (Abelson and Thomas, 1966) and T7 (Khan et al., 1995), buttheir biological function is unknown. The question of how the nicks/gaps are introduced andwhat role they may have in virus replication remains to be answered. Liu and colleagues(2015) have speculated that the nicks and ssDNA regions in SNJ2 and related viruses mightinduce DNA damage responses from the host, which could promote virus replication.

The HHPV3 genome was predicted to contain 17 ORFs, two of which were confirmedto be genes (gene 1 and gene 3) encoding two major structural protein species, VP1 and VP3(Figures 4 and 6, Table S6 in III). A set of HHPV3 ORFs/genes has sequence similarity,albeit not high, to the conserved cluster of genes observed in pleolipoviruses (Table S6 in III,Figure 5). Moreover, HHPV3 and other pleolipovirus genomes are collinear (Figure 5).Based on the sequence similarities with HRPV-1, the type species of the family

35

Pleolipoviridae (Pietilä et al., 2016), HHPV3 VP1 is predicted to be a membrane-associatedprotein and VP3 is most probably a spike protein. Putative protein 6 contains a domaincharacteristic to NTPases (Table S7 in III) and displays similarity to the previously detectedputative NTPases of pleolipoviruses (Figure 6 C in III, Figure 5). Based on the observed genesynteny, HHPV3 can be assigned to the group of betapleolipoviruses (Figure 5).

Figure 5. The genomes of HHPV3 and pleolipoviruses: (A) alphapleolipoviruses, (B)betapleolipoviruses, (C) gammapleolipovirus. ORFs/genes are shown as grey arrows, andhomologous ones are in the same colours. In HHPV3, the products of genes 1 and 3, i.e. VP1 andVP3, are indicated. (D) Schematic model of HHPV3 virion.

HHPV3-related sequence elements are present in many haloarchaea (Figures 6 D andS5). Some of these elements have been previously reported as betapleolipovirus-relatedputative proviruses, all flanked by an integrase and tRNA genes in haloarchaealchromosomes (Liu et al., 2015, Pietilä et al., 2016). The highest sequence similarities wereobserved between HHPV3 and putative proviruses in Har. hispanica and Har. marismortui(Figure 6 D in III). In addition to the known ones (Bath et al., 2006, Chen et al., 2014, Dyall-Smith et al., 2011, Liu et al., 2015, Pietilä et al., 2009, Roine et al., 2010, Senčilo et al., 2012),eight more putative pleolipovirus-related proviruses were identified in this work (Figure S5

36

in III). The fact that the known pleolipoviruses were isolated from distant locations and thatrelated proviruses are commonly found in the genomes of archaea originating from varioushypersaline environments suggests the extreme abundance of pleolipoviruses in nature. Aninteresting example of pleolipovirus-related proviruses is the provirus that was identified inthe genome of ‘Halanaeroarchaeum sulfurireducens’ M27-SA2, isolated from deep-sea (~3km depth) salt-saturated anoxic lake Medee (Messina et al., 2016). In this prophage, thepleolipovirus-like ORFs are interrupted by a single CRISPR array and eight associated cas-genes. It seems that the system might have a role in abolishing cellular anti-phagemechanisms, but additional data is needed to understand its functions (Messina et al.,2016).

Based on the cumulative data, HHPV3 virion is suggested to be a lipid vesicle, whichcontains one membrane-associated protein (VP1), and the vesicle surface is decorated by aspike protein (VP3) (Figure 5 D). HHPV3 is proposed to be a member of Betapleolipovirusgenus of the family Pleolipoviridae, which comprises haloarchaeal pleomorphic virusesdistributed worldwide. The discovery of pleolipoviruses further highlights the observationthat similar virion types can be found all across the globe (Liu et al., 2015, Pietilä et al., 2016,Roine et al., 2010, Senčilo et al., 2012).

37

5. Conclusions and Future ProspectsIn this work, 37 viruses were obtained from the samples collected from Samut Sakhonsaltern in 2009 and 2010. These isolates have significantly increased the total number ofisolated archaeal viruses. All new viruses displayed the previously known morphotypes.Moreover, the overall morphotype distribution observed in the known haloarchaeal virusesremains approximately the same when the viruses isolated here are added. Thus far, culture-dependent surveys of the hypersaline environments have revealed only a handful of differentvirus morphotypes (Atanasova et al., 2015a, Atanasova et al., 2015b). Whether this reflectsthe natural diversity in these extreme environments remains to be revealed by future studies,coupling culture-dependent and -independent methods. However, only the culture-basedapproach provides an opportunity for a comprehensive characterization of viruses inmolecular detail. The isolates obtained here are a valuable source for further more detailedresearch, including genome sequencing and annotation, as well as structural studies.

Multiple virus-host interactions were observed between the isolated viruses andhalophilic archaeal strains derived from the same location or from distant hypersalineenvironments. The interactions occurred across time and genus barriers, suggesting that thestudied viruses and haloarchaeal strains constitute a dynamic part of the natural microbiotaof hypersaline environments. Although tailed viruses are considered to be rare in highsalinity environments (Santos et al., 2012), numerous interactions observed betweenmyoviruses and a group of Halorubrum strains indicate that these tailed viruses may havea significant input in natural virus-host interplay.

A more in-depth characterization was performed for two new viruses with differentmorphotypes, tailless icosahedral virus HCIV-1 and pleomorphic virus HHPV3. Based onmany lines of evidence, the closest relatives for HCIV-1 are the other haloarchaeal taillessicosahedral viruses, SH1, PH1, and HHIV-2. Thus, HCIV-1 is classified as a new species inAlphasphaerolipovirus genus of the family Sphaerolipoviridae, which comprises taillessicosahedral dsDNA viruses that infect extremophilic archaea or bacteria. The conservedelements responsible for HCIV-1 virion architecture suggest that the virus belongs to thePRD1-adenovirus structural lineage together with other sphaerolipoviruses. HHPV3 issuggested to belong to the group of pleolipoviruses, and is proposed to be classified as a newspecies in Betapleolipovirus genus of the family Pleolipoviridae.

Due to the astronomical number of viruses in the biosphere, the probability to isolatesame viruses all over the globe seems to be low (Hendrix, 2002, Saren et al., 2005, Suttle,2007). However, the cases of HCIV-1 and HHPV3, together with the examples from othervirus groups (Saren et al., 2005), suggest the global distribution and resemblance of distinctvirion architectures. The results obtained in this work are in line with the idea that theenormous virus diversity resides at sequence level, while the number of different virionarchitectures is limited (Krupovič and Bamford, 2011). So far, striking structural similaritieshave been observed between various viruses classified into numerous families which are notassigned to any higher taxa (Abrescia et al., 2012). In this respect, structure-basedclassification is potentially a highly efficient approach to resolve current challenges of virustaxonomy. More sampling and detailed studies on the currently available virus isolates will

38

reveal how many ways there are to build an infectious virus particle and what evolutionarypathways underlie current virus diversity.

39

6. AcknowledgementsThis work was carried out during the years 2012-2016 at the University of Helsinki, Instituteof Biotechnology and Department of Biosciences, Programme on Molecular Virology headedby Academy Professor Dennis Bamford and supervised by Docent Hanna Oksanen, DocentMaija Pietilä, and PhD Nina Atanasova.

I owe my deepest gratitude to my supervisor, Hanna, for expert advice, constantsupport, and guidance in everything related to our projects, especially in project planningand academic writing. You have always had time for my questions, and I learned so muchthrough our discussions. Thank you for your trust, patience, and kind attitude.

With great pleasure, I thank Dennis for welcoming me for PhD studies in his researchgroup. Thank you for taking care of all of us, for your wisdom, kindness, and sense ofhumour. Your immense knowledge and experience have been of great support for this thesiswork. I also deeply appreciate your work as a member of my thesis advisory committee.

I am deeply thankful to my co-supervisors, Maija and Nina. Maija, thank you forguiding me in learning laboratory techniques and in data analyses. Your excellentorganizational skills and clarity of scientific thinking have been inspiring for me. Nina, thankyou for your passion in hunting for new viruses and for outstanding courage and persistencyin taming the “odd” ones. Thank you also for our lively discussions over a wide range oftopics!

I am grateful to Docent Sari Timonen for being a member of my thesis advisorycommittee. Thank you for sincere interest, enthusiasm, and helpful advice in my researchand studies. Thank you for such inspiring and fun meetings!

I express special gratitude to the pre-examiners of this thesis, Professor KennethStedman and Associate Professor Josefa Antón. Thank you for reading my thesis, its criticalassessment, and valuable comments.

I sincerely appreciate the financial support, which made this work possible, providedby the Centre for International Mobility CIMO, the Viikki Doctoral Programme in MolecularBiosciences (VGSB), the Doctoral Programme in Microbiology and Biotechnology (MBDP),the Ella and Georg Ehrnrooth Foundation, and the Alfred Kordelin Foundation. I alsoacknowledge the support of employees and the use of experimental resources of EU ESFRIInstruct Centre for Virus and Macromolecular Complex Production (ICVIR) at theUniversity of Helsinki, Finland.

I share the credit of my work with all my co-authors. I thank PhD Andrius Buivydas forhis contribution to the discovery of new archaeal viruses. I am grateful to Docent JanneRavantti for guiding me in the (exciting) world of Bioinformatics, for general computersupport, and for challenging me with “suomi-perjantai”. I am thankful to Julija Svirskaitefor the enthusiasm in fighting against electrodes (and winning them!) and being mycompanion in long virus life cycle experiments. Thank you also for your friendship!

My gratitude is extended to all of our excellent technicians, who have contributed somuch to the efficiency and productivity of our research laboratory. I express specialgratitude to Helin Veskiväli for being a highly skilled assistant for the work presented in this

40

thesis. I also thank Sari Korhonen and Riitta Tarkiainen for their helpful advice in laboratorypractices.

I would like to acknowledge all members of the Programme on Molecular Virology forcreating friendly and relaxed atmosphere. I especially thank Irma Orientaite, AlesiaRomanovskaya, and Daria Starkova for interesting conversations over lunch/tea. I alsoexpress my appreciation to Outi Lyytinen for being a great companion in travelling and for(patiently) teaching me Finnish.

I would like to give special thanks to my friends, Luiza, Könul, and Konstantin. Despitebusy schedules and geographical distances, I have felt your presence and moral support.Thank you for all the bright and joyful moments and our conversations full of laugh!

Finally, I express heartfelt gratitude to my beloved husband Sergey and my dearparents, Alexey and Galina, for their endless love and absolute support. Your presence in mylife fills it with happiness and joy. Without you, it would be hardly feasible to accomplishthis work.

Tatiana

October 2016

41

7. ReferencesAalto, A.P., Bitto, D., Ravantti, J.J., Bamford, D.H., Huiskonen, J.T., Oksanen, H.M., 2012. Snapshot of virus

evolution in hypersaline environments from the characterization of a membrane-containing Salisaetaicosahedral phage 1. Proc Natl Acad Sci U S A 109, 18, 7079-7084.

Abelson, J., Thomas, C., 1966. The anatomy of the T5 bacteriophage DNA molecule. J Mol Biol 18, 2, 262-288.

Abrescia, N., Grimes, J.M., Fry, E.E., Ravantti, J.J., Bamford, D.H., Stuart, D.I., 2010. What does it take tomake a virus: the concept of the viral “self". In: Stockley, P.G., Twarock, R. (editors). Emerging Topics inPhysical Virology, Imperial College Press, London, pp. 35-58.

Abrescia, N.G., Bamford, D.H., Grimes, J.M., Stuart, D.I., 2012. Structure unifies the viral universe. AnnuRev Biochem 81, 795-822.

Abrescia, N.G., Grimes, J.M., Kivelä, H.M., Assenberg, R., Sutton, G.C., Butcher, S.J., Bamford, J.K.H.,Bamford, D.H., Stuart, D.I., 2008. Insights into virus evolution and membrane biogenesis from thestructure of the marine lipid-containing bacteriophage PM2. Mol Cell 31, 5, 749-761.

Acharya, R., Fry, E., Stuart, D., Fox, G., Rowlands, D., Brown, F., 1989. The three-dimensional structure offoot-and-mouth disease virus at 2.9 Å resolution. Nature 337, 6209, 709-716.

Adriaenssens, E.M., van Zyl, L.J., Cowan, D.A., Trindade, M.I., 2016. Metaviromics of Namib desert saltpans: a novel lineage of haloarchaeal salterproviruses and a rich source of ssDNA viruses. Viruses 8, 1, 14.

Aherfi, S., Colson, P., La Scola, B., Raoult, D., 2016. Giant viruses of amoebas: an update. Front Microbiol 7,349.

Akita, F., Chong, K.T., Tanaka, H., Yamashita, E., Miyazaki, N., Nakaishi, Y., Suzuki, M., Namba, K., Ono, Y.,Tsukihara, T., Nakagawa, A., 2007. The crystal structure of a virus-like particle from thehyperthermophilic archaeon Pyrococcus furiosus provides insight into the evolution of viruses. J Mol Biol368, 5, 1469-1483.

Al-Mailem, D., Sorkhoh, N., Al-Awadhi, H., Eliyas, M., Radwan, S., 2010. Biodegradation of crude oil andpure hydrocarbons by extreme halophilic archaea from hypersaline coasts of the Arabian Gulf.Extremophiles 14, 3, 321-328.

Andrei, A., Banciu, H.L., Oren, A., 2012. Living with salt: metabolic and phylogenetic diversity of archaeainhabiting saline ecosystems. FEMS Microbiol Lett 330, 1, 1-9.

Atanasova, N.S., Bamford, D.H., Oksanen, H.M., 2015a. Haloarchaeal virus morphotypes. Biochimie 118,333-343.

Atanasova, N.S., Oksanen, H.M., Bamford, D.H., 2015b. Haloviruses of archaea, bacteria, and eukaryotes.Curr Opin Microbiol 25, 40-48.

Atanasova, N.S., Roine, E., Oren, A., Bamford, D.H., Oksanen, H.M., 2012. Global network of specific virus-host interactions in hypersaline environments. Environ Microbiol 14, 2, 426-440.

Athappilly, F.K., Murali, R., Rux, J.J., Cai, Z., Burnett, R.M., 1994. The refined crystal structure of hexon, themajor coat protein of adenovirus type 2, at 2.9 Å resolution. J Mol Biol 242, 4, 430-455.

Bahar, M.W., Graham, S.C., Stuart, D.I., Grimes, J.M., 2011. Insights into the evolution of a complex virusfrom the crystal structure of vaccinia virus D13. Structure 19, 7, 1011-1020.

Baker, M.L., Jiang, W., Rixon, F.J., Chiu, W., 2005. Common ancestry of herpesviruses and tailed DNAbacteriophages. J Virol 79, 23, 14967-14970.

Baltimore, D., 1971. Expression of animal virus genomes. Bacteriol Rev 35, 3, 235-241.

Bamford, D.H., 2003. Do viruses form lineages across different domains of life? Res Microbiol 154, 4, 231-236.

Bamford, D.H., Burnett, R.M., Stuart, D.I., 2002. Evolution of viral structure. Theor Popul Biol 61, 4, 461-470.

Bamford, D.H., Grimes, J.M., Stuart, D.I., 2005a. What does structure tell us about virus evolution? CurrOpin Struct Biol 15, 6, 655-663.

42

Bamford, D.H., Ravantti, J.J., Rönnholm, G., Laurinavičius, S., Kukkaro, P., Dyall-Smith, M., Somerharju, P.,Kalkkinen, N., Bamford, J.K., 2005b. Constituents of SH1, a novel lipid-containing virus infecting thehalophilic euryarchaeon Haloarcula hispanica. J Virol 79, 14, 9097-9107.

Bath, C., Cukalac, T., Porter, K., Dyall-Smith, M.L., 2006. His1 and His2 are distantly related, spindle-shapedhaloviruses belonging to the novel virus group, Salterprovirus. Virology 350, 1, 228-239.

Bath, C., Dyall-Smith, M.L., 1998. His1, an archaeal virus of the Fuselloviridae family that infects Haloarculahispanica. J Virol 72, 11, 9392-9395.

Benson, S.D., Bamford, J.K., Bamford, D.H., Burnett, R.M., 1999. Viral evolution revealed by bacteriophagePRD1 and human adenovirus coat protein structures. Cell 98, 6, 825-833.

Benson, S.D., Bamford, J.K.H., Bamford, D.H., Burnett, R.M., 2004. Does common architecture reveal a virallineage spanning all three domains of life? Mol Cell 16, 5, 673-685.

Bettarel, Y., Bouvier, T., Bouvier, C., Carré, C., Desnues, A., Domaizon, I., Jacquet, S., Robin, A., Sime-Ngando, T., 2011. Ecological traits of planktonic viruses and prokaryotes along a full-salinity gradient.FEMS Microbiol Ecol 76, 2, 360-372.

Birbir, M., Calli, B., Mertoglu, B., Bardavid, R.E., Oren, A., Ogmen, M.N., Ogan, A., 2007. Extremelyhalophilic Archaea from Tuz Lake, Turkey, and the adjacent Kaldirim and Kayacik salterns. J MicrobiolBiotechnol 23, 3, 309-316.

Boetius, A., Joye, S., 2009. Thriving in salt. Science 324, 5934, 1523-1525.

Bolhuis, H., Cretoiu, M.S., Stal, L.J., 2014. Molecular ecology of microbial mats. FEMS Microbiol Ecol 90, 2,335-350.

Boughalmi, M., Saadi, H., Pagnier, I., Colson, P., Fournous, G., Raoult, D., La Scola, B., 2013. High-throughput isolation of giant viruses of the Mimiviridae and Marseilleviridae families in the Tunisianenvironment. Environ Microbiol 15, 7, 2000-2007.

Boujelben, I., Gomariz, M., Martínez-García, M., Santos, F., Peña, A., López, C., Antón, J., Maalej, S., 2012.Spatial and seasonal prokaryotic community dynamics in ponds of increasing salinity of Sfax solar salternin Tunisia. Antonie Van Leeuwenhoek 101, 4, 845-857.

Boujelben, I., Martínez-García, M., van Pelt, J., Maalej, S., 2014. Diversity of cultivable halophilic archaeaand bacteria from superficial hypersaline sediments of Tunisian solar salterns. Antonie Van Leeuwenhoek106, 4, 675-692.

Bowers, K.J., Wiegel, J., 2011. Temperature and pH optima of extremely halophilic archaea: a mini-review.Extremophiles 15, 2, 119-128.

Breitbart, M., Rohwer, F., 2005. Here a virus, there a virus, everywhere the same virus? Trends Microbiol 13,6, 278-284.

Burns, D.G., Janssen, P.H., Itoh, T., Kamekura, M., Li, Z., Jensen, G., Rodrìguez-Valera, F., Bolhuis, H., Dyall-Smith, M.L. ,2007. Haloquadratum walsbyi gen. nov., sp. nov., the square haloarchaeaon of Walsby,isolated from saltern crystallizers in Australia and Spain. Int J Syst Evol Micr 57, 387-392.

Burns, D.G., Camakaris, H.M., Janssen, P.H., Dyall-Smith, M.L., 2004. Combined use of cultivation-dependent and cultivation-independent methods indicates that members of most haloarchaeal groups inan Australian crystallizer pond are cultivable. Appl Environ Microbiol 70, 9, 5258-5265.

Calvo, C., de la Paz, A. G., Bejar, V., Quesada, E., Ramos-Cormenzana, A., 1988. Isolation andcharacterization of phage F9-11 from a lysogenic Deleya halophila strain. Curr Microbiol 17, 1, 49-53.

Calvo, C., de La Paz, A. G., Martínez-Checa, F., Caba, M.A., 1994. Behaviour of two D. halophilabacteriophages with respect to salt concentrations and other environmental factors. Toxicol EnvironChemi 43, 1-2, 85-93.

Chen, S., Wang, C., Xu, J.P., Yang, Z.L., 2014. Molecular characterization of pHRDV1, a new virus-likemobile genetic element closely related to pleomorphic viruses in haloarchaea. Extremophiles 18, 2, 195-206.

Claverie, J.M., Abergel, C., 2010. Mimivirus: the emerging paradox of quasi-autonomous viruses. TrendsGenet 26, 10, 431-437.

Daniels, L.L., Wais, A.C., 1998. Virulence in phage populations infecting Halobacterium cutirubrum. FEMSMicrobiol Ecol 25, 2, 129-134.

43

Daniels, L.L., Wais, A.C., 1984. Restriction and modification of halophage S45 in Halobacterium. CurrMicrobiol 10, 3, 133-136.

Daniels, L.L., Wais, A.C., 1990. Ecophysiology of bacteriophage S5100 infecting Halobacterium cutirubrum.Appl Environ Microbiol 56, 11, 3605-3608.

DasSarma, S., DasSarma, P., 2012. Halophiles. In: eLS. John Wiley & Sons Ltd, Chichester.http://www.els.net [doi: 10.1002/9780470015902.a0000394.pub3]

DasSarma, S., DasSarma, P., 2015. Halophiles and their enzymes: negativity put to good use. Curr OpinMicrobiol 25, 120-126.

Dellas, N., Snyder, J.C., Bolduc, B., Young, M.J., 2014. Archaeal viruses: diversity, replication, and structure.Annu Rev Virol 1, 1, 399-426.

Díez, B., Antón, J., Guixa-Boixereu, N., Pedrós-Alió, C., Rodríguez-Valera, F., 2000. Pulsed-field gelelectrophoresis analysis of virus assemblages present in a hypersaline environment. Int Microbiol 3, 3,159-164.

Dokland, T., McKenna, R., Ilag, L.L., Bowman, B.R., Incardona, N.L., Fane, B.A., Rossmann, M.G., 1997.Structure of a viral procapsid with molecular scaffolding. Nature 389, 6648, 308-313.

Dolja, V.V., Krupovič, M., 2013. Accelerating expansion of the viral universe. Curr Opin Virol 3, 5, 542-545.

Dupré, J., Guttinger, S., 2016. Viruses as living processes. Stud Hist Philos Biol Biomed Sci pii: S1369-8486(16)30003-6. doi: 10.1016/j.shpsc.2016.02.010. [Epub ahead of print]

Dyall-Smith, M.L., Pfeiffer, F., Klee, K., Palm, P., Gross, K., Schuster, S.C., Rampp, M., Oesterhelt, D., 2011.Haloquadratum walsbyi: limited diversity in a global pond. PLoS One 6, 6, e20968.

Edbeib, M.F., Wahab, R.A., Huyop, F., 2016. Halophiles: biology, adaptation, and their role indecontamination of hypersaline environments. J Microbiol Biotechnol 32, 8, 1-23.

Emerson, J.B., Andrade, K., Thomas, B.C., Norman, A., Allen, E.E., Heidelberg, K.B., Banfield, J.F., 2013a.Virus-host and CRISPR dynamics in archaea-dominated hypersaline Lake Tyrrell, Victoria, Australia.Archaea 2013, article ID 370871, 12 p.

Emerson, J.B., Thomas, B.C., Andrade, K., Allen, E.E., Heidelberg, K.B., Banfield, J.F., 2012. Dynamic viralpopulations in hypersaline systems as revealed by metagenomic assembly. Appl Environ Microbiol 78, 17,6309-6320.

Emerson, J.B., Thomas, B.C., Andrade, K., Heidelberg, K.B., Banfield, J.F., 2013b. New approaches indicateconstant viral diversity despite shifts in assemblage structure in an Australian hypersaline lake. ApplEnviron Microbiol 79, 21, 6755-6764.

Finsterbusch, T., Mankertz, A., 2009. Porcine circoviruses - small but powerful. Virus Res 143, 2, 177-183.

Fokine, A., Leiman, P.G., Shneider, M.M., Ahvazi, B., Boeshans, K.M., Steven, A.C., Black, L.W.,Mesyanzhinov, V.V., Rossmann, M.G., 2005. Structural and functional similarities between the capsidproteins of bacteriophages T4 and HK97 point to a common ancestry. Proc Natl Acad Sci U S A 102, 20,7163-7168.

Forterre, P., Soler, N., Krupovič, M., Marguet, E., Ackermann, H., 2013. Fake virus particles generated byfluorescence microscopy. Trends Microbiol 21, 1, 1-5.

Forterre, P., 2002. The origin of DNA genomes and DNA replication proteins. Curr Opin Microbiol 5, 5, 525-532.

Forterre, P., Prangishvili, D., 2013. The major role of viruses in cellular evolution: facts and hypotheses. CurrOpin Virol 3, 5, 558-565.

Forterre, P., Prangishvili, D., 2009. The great billion-year war between ribosome- and capsid-encodingorganisms (cells and viruses) as the major source of evolutionary novelties. Ann N Y Acad Sci 1178, 65-77.

Fu, C.Q., Zhao, Q., Li, Z.Y., Wang, Y.X., Zhang, S.Y., Lai, Y.H., Xiao, W., Cui, X.L., 2016. A novel Halomonasventosae-specific virulent halovirus isolated from the Qiaohou salt mine in Yunnan, Southwest China.Extremophiles 20, 1, 101-110.

Garcia-Heredia, I., Martin-Cuadrado, A., Mojica, F.J., Santos, F., Mira, A., Antón, J., Rodriguez-Valera, F.,2012. Reconstructing viral genomes from the environment using fosmid clones: the case ofhaloviruses. PLoS One 7, 3, e33802.

44

Garrett, R.A., Vestergaard, G., Shah, S.A., 2011. Archaeal CRISPR-based immune systems: exchangeablefunctional modules. Trends Microbiol 19, 11, 549-556.

Ghai, R., Pašić, L., Fernández, A.B., Martin-Cuadrado, A., Mizuno, C.M., McMahon, K.D., Papke, R.T.,Stepanauskas, R., Rodriguez-Brito, B., Rohwer, F., 2011. New abundant microbial groups in aquatichypersaline environments. Scientific Reports 1, article number 135 (2011).

Gil-Carton, D., Jaakkola, S.T., Charro, D., Peralta, B., Castano-Diez, D., Oksanen, H.M., Bamford, D.H.,Abrescia, N.G., 2015. Insight into the assembly of viruses with vertical single beta-barrel major capsidproteins. Structure 23, 10, 1866-1877.

Goel, U., Kauri, T., Kushner, D.J., Ackermann, H., 1996. A moderately halophilic Vibrio from a Spanishsaltern and its lytic bacteriophage. Can J Microbiol 42, 10, 1015-1023.

Gogvadze, E., Buzdin, A., 2009. Retroelements and their impact on genome evolution and functioning. CellMol Life Sci 66, 23, 3727-3742.

Goldfarb, T., Sberro, H., Weinstock, E., Cohen, O., Doron, S., Charpak-Amikam, Y., Afik, S., Ofir, G., Sorek,R., 2015. BREX is a novel phage resistance system widespread in microbial genomes. EMBO J 34, 2, 169-183.

Gowen, B., Bamford, J.K., Bamford, D.H., Fuller, S.D., 2003. The tailless icosahedral membrane virus PRD1localizes the proteins involved in genome packaging and injection at a unique vertex. J Virol 77, 14, 7863-7871.

Grant, S., Grant, W.D., Jones, B.E., Kato, C., Li, L., 1999. Novel archaeal phylotypes from an East Africanalkaline saltern. Extremophiles 3, 2, 139-145.

Grimes, J.M., Burroughs, J.N., Gouet, P., Diprose, J.M., Malby, R., Zientara, S., Mertens, P.P., Stuart, D.I.,1998. The atomic structure of the bluetongue virus core. Nature 395, 6701, 470-478.

Guixa-Boixareu, N., Calderón-Paz, J., Heldal, M., Bratbak, G., Pedrós-Alió, C., 1996. Viral lysis andbacterivory as prokaryotic loss factors along a salinity gradient. Aquat Microb Ecol 11, 3, 215-227.

Gupta, R.S., Naushad, S., Baker, S. 2015. Phylogenomic analyses and molecular signatures for the classHalobacteria and its two major clades: a proposal for division of the class Halobacteria into an emendedorder Halobacteriales and two new orders, Haloferacales ord. nov. and Natrialbales ord. nov.,containing the novel families Haloferacaceae fam. nov. and Natrialbaceae fam. nov. Int J Syst EvolMicrobiol 65, 3, 1050–1069

Gupta, R.S., Naushad, S., Fabros, R., Adeolu, M. 2016. A phylogenomic reappraisal of family-level divisionswithin the class Halobacteria: proposal to divide the order Halobacteriales into the familiesHalobacteriaceae, Haloarculaceae fam. nov., and Halococcaceae fam. nov., and the order Haloferacalesinto the families, Haloferacaceae and Halorubraceae fam nov. Antonie van Leeuwenhoek, 109, 4, 565-587.

Hambly, E., Suttle, C.A., 2005. The viriosphere, diversity, and genetic exchange within phage communities.Curr Opin Microbiol 8, 4, 444-450.

Happonen, L.J., Redder, P., Peng, X., Reigstad, L.J., Prangishvili, D., Butcher, S.J., 2010. Familialrelationships in hyperthermo- and acidophilic archaeal viruses. J Virol 84, 9, 4747-4754.

Häring, M., Vestergaard, G., Rachel, R., Chen, L., Garrett, R.A., Prangishvili, D., 2005a. Virology:independent virus development outside a host. Nature 436, 7054, 1101-1102.

Häring, M., Rachel, R., Peng, X., Garrett, R.A., Prangishvili, D., 2005b. Viral diversity in hot springs ofPozzuoli, Italy, and characterization of a unique archaeal virus, Acidianus bottle-shaped virus, from a newfamily, the Ampullaviridae. J Virol 79, 15, 9904-9911.

Hendrix, R.W., 2002. Bacteriophages: evolution of the majority. Theor Popul Biol 61, 4, 471-480.

Hogle, J.M., Chow, M., Filman, D.J., 1985. Three-dimensional structure of poliovirus at 2.9 Å resolution.Science 229, 4720, 1358-1365.

Hong, C., Oksanen, H., Liu, X., Jakana, J., Bamford, D., 2014. A structural model of the genome packagingprocess in a membrane-containing double-stranded DNA viruses. PLoS Biol 12, 12, e1002024.

Hong, C., Pietilä, M.K., Fu, C.J., Schmid, M.F., Bamford, D.H., Chiu, W., 2015. Lemon-shaped halo archaealvirus His1 with uniform tail but variable capsid structure. Proc Natl Acad Sci U S A 112, 8, 2449-2454.

45

Huiskonen, J.T., de Haas, F., Bubeck, D., Bamford, D.H., Fuller, S.D., Butcher, S.J., 2006. Structure of thebacteriophage ϕ6 nucleocapsid suggests a mechanism for sequential RNA packaging. Structure 14, 6,1039-1048.

Jaakkola, S.T., Penttinen, R.K., Vilen, S.T., Jalasvuori, M., Rönnholm, G., Bamford, J.K.H., Bamford, D.H.,Oksanen, H.M., 2012. Closely related archaeal Haloarcula hispanica icosahedral viruses HHIV-2 and SH1have nonhomologous genes encoding host recognition functions. J Virol 86, 9, 4734-4742.

Jäälinoja, H.T., Roine, E., Laurinmäki, P., Kivelä, H.M., Bamford, D.H., Butcher, S.J., 2008. Structure andhost-cell interaction of SH1, a membrane-containing, halophilic euryarchaeal virus. Proc Natl Acad Sci US A 105, 23, 8008-8013.

Jaatinen, S.T., Happonen, L.J., Laurinmäki, P., Butcher, S.J., Bamford, D.H., 2008. Biochemical andstructural characterisation of membrane-containing icosahedral dsDNA bacteriophages infectingthermophilic Thermus thermophilus. Virology 379, 1, 10-19.

Jalasvuori, M., Jaatinen, S.T., Laurinavičius, S., Ahola-Iivarinen, E., Kalkkinen, N., Bamford, D.H., Bamford,J.K.H., 2009. The closest relatives of icosahedral viruses of thermophilic bacteria are among viruses andplasmids of the halophilic archaea. J Virol 83, 18, 9388-9397.

Javor, B., Requadt, C., Stoeckenius, W., 1982. Box-shaped halophilic bacteria. J Bacteriol 151, 3, 1532-1542.

Jiang, W., Li, Z., Zhang, Z., Baker, M.L., Prevelige, P.E.,Jr, Chiu, W., 2003. Coat protein fold and maturationtransition of bacteriophage P22 seen at subnanometer resolutions. Nat Struct Biol 10, 2, 131-135.

Kauri, T., Ackermann, H., Goel, U., Kushner, D.J., 1991. A bacteriophage of a moderately halophilicbacterium. Arch Microbiol 156, 6, 435-438.

Khan, S.A., Hayes, S.J., Wright, E.T., Watson, R.H., Serwer, P., 1995. Specific single-stranded breaks inmature bacteriophage T7 DNA. Virology 211, 1, 329-331.

Kivelä, H.M., Roine, E., Kukkaro, P., Laurinavičius, S., Somerharju, P., Bamford, D.H., 2006. Quantitativedissociation of archaeal virus SH1 reveals distinct capsid proteins and a lipid core. Virology 356, 1-2, 4-11.

Klein, R., Rossler, N., Iro, M., Scholz, H., Witte, A., 2012. Haloarchaeal myovirus φCh1 harbours a phasevariation system for the production of protein variants with distinct cell surface adhesion specificities.Mol Microbiol 83, 1, 137-150.

Knowles, B., Silveira, C.B., Bailey, B.A., Barott, K., Cantu, V.A., Cobian-Guemes, A.G., Coutinho, F.H.,Dinsdale, E.A., Felts, B., Furby, K.A., George, E.E., Green, K.T., Gregoracci, G.B., Haas, A.F., Haggerty,J.M., Hester, E.R., Hisakawa, N., Kelly, L.W., Lim, Y.W., Little, M., Luque, A., McDole-Somera, T.,McNair, K., de Oliveira, L.S., Quistad, S.D., Robinett, N.L., Sala, E., Salamon, P., Sanchez, S.E., Sandin, S.,Silva, G.G., Smith, J., Sullivan, C., Thompson, C., Vermeij, M.J., Youle, M., Young, C., Zgliczynski, B.,Brainard, R., Edwards, R.A., Nulton, J., Thompson, F., Rohwer, F., 2016. Lytic to temperate switching ofviral communities. Nature 531, 7595, 466-470.

Koonin, E.V., Dolja, V.V., 2013. A virocentric perspective on the evolution of life. Curr Opin Virol 3, 5, 546-557.

Koonin, E.V., Starokadomskyy, P., 2016. Are viruses alive? The replicator paradigm sheds decisive light on anold but misguided question. Stud Hist Philos Biol Biomed Sci pii: S1369-8486(16)30010-3. doi:10.1016/j.shpsc.2016.02.016. [Epub ahead of print]

Kostyrka, G., 2016. What roles for viruses in origin of life scenarios? Stud Hist Philos Biol Biomed Sci pii:S1369-8486(16)30007-3. doi: 10.1016/j.shpsc.2016.02.014. [Epub ahead of print]

Krupovič, M., Forterre, P., Bamford, D.H., 2010. Comparative analysis of the mosaic genomes of tailedarchaeal viruses and proviruses suggests common themes for virion architecture and assembly with tailedviruses of bacteria. J Mol Biol 397, 1, 144-160.

Krupovič, M., Bamford, D.H., 2011. Double-stranded DNA viruses: 20 families and only five differentarchitectural principles for virion assembly. Curr Opin Virol 1, 2, 118-124.

Krupovič, M., Bamford, D.H., 2010. Order to the viral universe. J Virol 84, 24, 12476-12479.

Krupovič, M., Quemin, E.R., Bamford, D.H., Forterre, P., Prangishvili, D., 2014. Unification of the globallydistributed spindle-shaped viruses of the Archaea. J Virol 88, 4, 2354-2358.

Kuhlbrandt, W., 2014. Cryo-EM enters a new era. eLife 3, e03678.

46

Kukkaro, P., Bamford, D.H., 2009. Virus-host interactions in environments with a wide range of ionicstrengths. Environ Microbiol Rep 1, 1, 71-77.

Kushner, D.J., Kamekura, M., 1988. Physiology of halophilic eubacteria. In: Rodríguez-Varela, F.(editor). Halophilic Bacteria. CRC Press; Boca Raton, FL, USA, pp. 109–138.

Langlet, J., Gaboriaud, F., Duval, J.F., Gantzer, C., 2008. Aggregation and surface properties of F-specificRNA phages: implication for membrane filtration processes. Water Res 42, 10, 2769-2777.

Laurinavičius, S., Käkelä, R., Bamford, D.H., Somerharju, P., 2004a. The origin of phospholipids of theenveloped bacteriophage phi6. Virology 326, 1, 182-190.

Laurinavičius, S., Käkelä, R., Somerharju, P., Bamford, D.H., 2004b. Phospholipid molecular species profilesof tectiviruses infecting Gram-negative and Gram-positive hosts. Virology 322, 2, 328-336.

Leuko, S., Rettberg, P., Pontifex, A.L., Burns, B.P., 2014. On the response of halophilic archaea to spaceconditions. Life 4, 1, 66-76.

Li, M., Wang, R., Zhao, D., Xiang, H., 2014. Adaptation of the Haloarcula hispanica CRISPR-Cas system to apurified virus strictly requires a priming process. Nucleic Acids Res 42, 4, 2483-2492.

Liu, Y., Wang, J., Liu, Y., Wang, Y., Zhang, Z., Oksanen, H.M., Bamford, D.H., Chen, X., 2015. Identificationand characterization of SNJ2, the first temperate pleolipovirus integrating into the genome of the SNJ1-lysogenic archaeal strain. Mol Microbiol 98, 6, 1002-1020.

Lokossou, A.G., Toudic, C., Barbeau, B., 2014. Implication of human endogenous retrovirus envelopeproteins in placental functions. Viruses 6, 11, 4609-4627.

Luk, A.W., Williams, T.J., Erdmann, S., Papke, R.T., Cavicchioli, R., 2014. Viruses of Haloarchaea. Life 4, 4,681-715.

Lwoff, A., Tournier, P., 1966. The classification of viruses. Annu Rev Microbiol 20, 45-74.

Maaty, W.S., Ortmann, A.C., Dlakic, M., Schulstad, K., Hilmer, J.K., Liepold, L., Weidenheft, B., Khayat, R.,Douglas, T., Young, M.J., Bothner, B., 2006. Characterization of the archaeal thermophile Sulfolobusturreted icosahedral virus validates an evolutionary link among double-stranded DNA viruses from alldomains of life. J Virol 80, 15, 7625-7635.

Mann, N.H., 2003. Phages of the marine cyanobacterial picophytoplankton. FEMS Microbiol Rev 27, 1, 17-34.

Martínez-García, M., Santos, F., Moreno-Paz, M., Parro, V., Antón, J., 2014. Unveiling viral-host interactionswithin the 'microbial dark matter'. Nat Commun 5, 4542.

Matsushita, I., Yamashita, N., and Yokota, A., 1995. Isolation and characterization of bacteriophage inducedfrom a new isolate of Thermus aquaticus. Microbial Cult Coll 11, 2, 133-138.

Matsushita, I., Yanase, H., 2009. The genomic structure of Thermus bacteriophage ϕIN93. J Biochem 146, 6,775-785.

McHugh, C.A., Fontana, J., Nemecek, D., Cheng, N., Aksyuk, A.A., Heymann, J.B., Winkler, D.C., Lam, A.S.,Wall, J.S., Steven, A.C., Hoiczyk, E., 2014. A virus capsid-like nanocompartment that stores iron andprotects bacteria from oxidative stress. EMBO J 33, 17, 1896-1911.

Mei, Y., He, C., Huang, Y., Liu, Y., Zhang, Z., Chen, X., Shen, P., 2015. Salinity regulation of the interaction ofhalovirus SNJ1 with its host and alteration of the halovirus replication strategy to adapt to the variableecosystem. PLoS One 10, 4, e0123874.

Messina, E., Sorokin, D.Y., Kublanov, I.V., Toshchakov, S., Lopatina, A., Arcadi, E., Smedile, F., La Spada, G.,La Cono, V., Yakimov, M.M., 2016. Complete genome sequence of ‘Halanaeroarchaeum sulfurireducens’M27-SA2, a sulfur-reducing and acetate-oxidizing haloarchaeon from the deep-sea hypersaline anoxiclake Medee. Stand Genomic Sci 11, 1, 1.

Morais, M.C., Choi, K.H., Koti, J.S., Chipman, P.R., Anderson, D.L., Rossmann, M.G., 2005. Conservation ofthe capsid structure in tailed dsDNA bacteriophages: the pseudoatomic structure of ϕ29. Mol Cell 18, 2,149-159.

Morgan, G.J., 2016. What is a virus species? Radical pluralism in viral taxonomy. Stud Hist Philos BiolBiomed Sci pii: S1369-8486(16)30004-8. doi: 10.1016/j.shpsc.2016.02.009. [Epub ahead of print]

47

Naitow, H., Tang, J., Canady, M., Wickner, R.B., Johnson, J.E., 2002. L-A virus at 3.4 A resolution revealsparticle architecture and mRNA decapping mechanism. Nat Struct Biol 9, 10, 725-728.

Nakagawa, A., Miyazaki, N., Taka, J., Naitow, H., Ogawa, A., Fujimoto, Z., Mizuno, H., Higashi, T.,Watanabe, Y., Omura, T., Cheng, R.H., Tsukihara, T., 2003. The atomic structure of rice dwarf virusreveals the self-assembly mechanism of component proteins. Structure 11, 10, 1227-1238.

Nandhagopal, N., Simpson, A.A., Gurnon, J.R., Yan, X., Baker, T.S., Graves, M.V., Van Etten, J.L.,Rossmann, M.G., 2002. The structure and evolution of the major capsid protein of a large, lipid-containing DNA virus. Proc Natl Acad Sci U S A 99, 23, 14758-14763.

Narasingarao, P., Podell, S., Ugalde, J.A., Brochier-Armanet, C., Emerson, J.B., Brocks, J.J., Heidelberg,K.B., Banfield, J.F., Allen, E.E., 2012. De novo metagenomic assembly reveals abundant novel majorlineage of Archaea in hypersaline microbial communities. ISME J 6, 1, 81-93.

Nuttall, S.D., Dyall-Smith, M.L., 1993. HF1 and HF2: novel bacteriophages of halophilic archaea. Virology197, 2, 678-684.

Oh, D., Porter, K., Russ, B., Burns, D., Dyall-Smith, M., 2010. Diversity of Haloquadratum and otherhaloarchaea in three, geographically distant, Australian saltern crystallizer ponds. Extremophiles 14, 2,161-169.

Oksanen, H.M., Pietilä, M.K., Senčilo, A., Atanasova, N.S., Roine, E., Bamford, D.H., 2012. Virus universe:can it be constructed from a limited number of viral architectures? In: Witzany, G. (editor). Viruses:Essential Agents of Life. Dordrecht: Springer Science+Business Media, pp. 83-105.

Oren, A., 2015. Halophilic microbial communities and their environments. Curr Opin Biotechnol 33, 119-124.

Oren, A., 2014a. The ecology of Dunaliella in high-salt environments. J Biol Res (Thessalon) 21, 23.

Oren, A., 2014b. Taxonomy of halophilic Archaea: current status and future challenges. Extremophiles 18, 5,825-834.

Oren, A., 2010. The dying Dead Sea: the microbiology of an increasingly extreme environment. Lakes &Reservoirs: Research & Management 15, 3, 215-222.

Oren, A., 2002. Halophilic microorganisms and their environments. Springer Science+Business MediaDordrecht, Netherlands, 575 p.

Oren, A., 1983. Halobacterium sodomense sp. nov., a Dead Sea halobacterium with an extremely highmagnesium requirement. Int J Syst Bacteriol 33, 2, 381-386.

Pagaling, E., Haigh, R.D., Grant, W.D., Cowan, D.A., Jones, B.E., Ma, Y., Ventosa, A., Heaphy, S., 2007.Sequence analysis of an archaeal virus isolated from a hypersaline lake in Inner Mongolia, China. BMCGenomics 8, 1, 1.

Pauling, C., 1982. Bacteriophages of Halobacterium halobium: isolated from fermented fish sauce andprimary characterization. Can J Microbiol 28, 8, 916-921.

Pawlowski, A., Rissanen, I., Bamford, J.K.H., Krupovič, M., Jalasvuori, M., 2014. Gammasphaerolipovirus, anewly proposed bacteriophage genus, unifies viruses of halophilic archaea and thermophilic bacteriawithin the novel family Sphaerolipoviridae. Arch Virol 159, 6, 1541-1554.

Pedros-Alio, C., Calderon-Paz, J.I., MacLean, M.H., Medina, G., Marrase, C., Gasol, J.M., Guixa-Boixereu, N.,2000. The microbial food web along salinity gradients. FEMS Microbiol Ecol 32, 2, 143-155.

Pennazio, S., 2003. The contribution of plant biology to the concept of virus (1886-1917). Riv Biol 96, 2, 241-260.

Peralta, B., Gil-Carton, D., Castano-Diez, D., Bertin, A., Boulogne, C., Oksanen, H.M., Bamford, D.H.,Abrescia, N.G., 2013. Mechanism of membranous tunnelling nanotube formation in viral genomedelivery. PLoS Biol 11, 9, e1001667.

Pfeifer, F., 2015. Haloarchaea and the formation of gas vesicles. Life (Basel) 5, 1, 385-402.

Pietilä, M.K., Atanasova, N.S., Oksanen, H.M., Bamford, D.H., 2013a. Modified coat protein forms theflexible spindle-shaped virion of haloarchaeal virus His1. Environ Microbiol 15, 6, 1674-1686.

Pietilä, M.K., Roine, E., Paulin, L., Kalkkinen, N., Bamford, D.H., 2009. An ssDNA virus infecting archaea: anew lineage of viruses with a membrane envelope. Mol Microbiol 72, 2, 307-319.

48

Pietilä, M.K., Atanasova, N.S., Manole, V., Liljeroos, L., Butcher, S.J., Oksanen, H.M., Bamford, D.H., 2012.Virion architecture unifies globally distributed pleolipoviruses infecting halophilic archaea. J Virol 86, 9,5067-5079.

Pietilä, M.K., Demina, T.A., Atanasova, N.S., Oksanen, H.M., Bamford, D.H., 2014. Archaeal viruses andbacteriophages: comparisons and contrasts. Trends Microbiol 22, 6, 334-344.

Pietilä, M.K., Laurinavičius, S., Sund, J., Roine, E., Bamford, D.H., 2010. The single-stranded DNA genomeof novel archaeal virus Halorubrum pleomorphic virus 1 is enclosed in the envelope decorated withglycoprotein spikes. J Virol 84, 2, 788-798.

Pietilä, M.K., Laurinmäki, P., Russell, D.A., Ko, C.C., Jacobs-Sera, D., Butcher, S.J., Bamford, D.H., Hendrix,R.W., 2013b. Insights into head-tailed viruses infecting extremely halophilic archaea. J Virol 87, 6, 3248-3260.

Pietilä, M.K., Laurinmäki, P., Russell, D.A., Ko, C.C., Jacobs-Sera, D., Hendrix, R.W., Bamford, D.H.,Butcher, S.J., 2013c. Structure of the archaeal head-tailed virus HSTV-1 completes the HK97 fold story.Proc Natl Acad Sci U S A 110, 26, 10604-10609.

Pietilä, M.K., Roine, E., Senčilo, A., Bamford, D.H., Oksanen, H.M., 2016. Pleolipoviridae, a newly proposedfamily comprising archaeal pleomorphic viruses with single-stranded or double-stranded DNA genomes.Arch Virol 161, 1, 249-56.

Porter, K., Russ, B.E., Dyall-Smith, M.L., 2007. Virus–host interactions in salt lakes. Curr Opin Microbiol 10,4, 418-424.

Porter, K., Dyall-Smith, M.L., 2008. Transfection of haloarchaea by the DNAs of spindle and roundhaloviruses and the use of transposon mutagenesis to identify non-essential regions. Mol Microbiol 70, 5,1236-1245.

Porter, K., Kukkaro, P., Bamford, J.K.H., Bath, C., Kivelä, H.M., Dyall-Smith, M.L., Bamford, D.H., 2005.SH1: A novel, spherical halovirus isolated from an Australian hypersaline lake. Virology 335, 1, 22-33.

Porter, K., Tang, S.L., Chen, C.P., Chiang, P.W., Hong, M.J., Dyall-Smith, M., 2013. PH1: an archaeovirus ofHaloarcula hispanica related to SH1 and HHIV-2. Archaea 2013, 456318.

Pradeu, T., Kostyrka, G., Dupre, J., 2016. Understanding viruses: philosophical investigations. Stud HistPhilos Biol Biomed Sci pii: S1369-8486(16)30002-4. doi: 10.1016/j.shpsc.2016.02.008. [Epub ahead ofprint]

Prangishvili, D., Basta, T., Garrett, R.A., Krupovič, M., 2016. Viruses of the Archaea. In: eLS. John Wiley &Sons Ltd, Chichester. http://www.els.net [doi: 10.1002/9780470015902.a0000774.pub3]

Prangishvili, D., 2013. The wonderful world of archaeal viruses. Annu Rev Microbiol 67, 565-585.

Prangishvili, D., Vestergaard, G., Haring, M., Aramayo, R., Basta, T., Rachel, R., Garrett, R.A., 2006.Structural and genomic properties of the hyperthermophilic archaeal virus ATV with an extracellularstage of the reproductive cycle. J Mol Biol 359, 5, 1203-1216.

Ravantti, J., Bamford, D., Stuart, D.I., 2013. Automatic comparison and classification of protein structures. JStruct Biol 183, 1, 47-56.

Reinisch, K.M., Nibert, M.L., Harrison, S.C., 2000. Structure of the reovirus core at 3.6 Å resolution. Nature404, 6781, 960-967.

Rice, G., Tang, L., Stedman, K., Roberto, F., Spuhler, J., Gillitzer, E., Johnson, J.E., Douglas, T., Young, M.,2004. The structure of a thermophilic archaeal virus shows a double-stranded DNA viral capsid type thatspans all domains of life. Proc Natl Acad Sci U S A 101, 20, 7716-7720.

Rissanen, I., Grimes, J.M., Pawlowski, A., Mäntynen, S., Harlos, K., Bamford, J.K.H., Stuart, D.I., 2013.Bacteriophage P23-77 capsid protein structures reveal the archetype of an ancient branch from a majorvirus lineage. Structure 21, 5, 718-726.

Rodriguez-Brito, B., Li, L., Wegley, L., Furlan, M., Angly, F., Breitbart, M., Buchanan, J., Desnues, C.,Dinsdale, E., Edwards, R., 2010. Viral and microbial community dynamics in four aquatic environments.ISME J 4, 6, 739-751.

Rodriguez-Valera, F., Martin-Cuadrado, A., Rodriguez-Brito, B., Pašić, L., Thingstad, T.F., Rohwer, F., Mira,A., 2009. Explaining microbial population genomics through phage predation. Nat Rev Microbiol 7, 11,828-836.

49

Rohwer, F., Prangishvili, D., Lindell, D., 2009. Roles of viruses in the environment. Environ Microbiol 11, 11,2771-2774.

Roine, E., Kukkaro, P., Paulin, L., Laurinavičius, S., Domanska, A., Somerharju, P., Bamford, D.H., 2010.New, closely related haloarchaeal viral elements with different nucleic acid types. J Virol 84, 7, 3682-3689.

Rosario, K., Breitbart, M., 2011. Exploring the viral world through metagenomics. Curr Opin Virol 1, 4, 289-297.

Rossmann, M.G., Arnold, E., Erickson, J.W., Frankenberger, E.A., Griffith, J.P., Hecht, H.J., Johnson, J.E.,Kamer, G., Luo, M., Mosser, A.G., 1985. Structure of a human common cold virus and functionalrelationship to other picornaviruses. Nature 317, 6033, 145-153.

Roux, S., Enault, F., Ravet, V., Colombet, J., Bettarel, Y., Auguet, J., Bouvier, T., Lucas-Staat, S., Vellet, A.,Prangishvili, D., 2016. Analysis of metagenomic data reveals common features of halophilic viralcommunities across continents. Environ Microbiol 18, 889–903.

Rux, J.J., Kuser, P.R., Burnett, R.M., 2003. Structural and phylogenetic analysis of adenovirus hexons by useof high-resolution X-ray crystallographic, molecular modeling, and sequence-based methods. J Virol 77,17, 9553-9566.

Sandaa, R., Skjoldal, E.F., Bratbak, G., 2003. Virioplankton community structure along a salinity gradient ina solar saltern. Extremophiles 7, 5, 347-351.

Santos, F., Meyerdierks, A., Peña, A., Rosselló-Mora, R., Amann, R., Antón, J., 2007. Metagenomic approachto the study of halophages: the environmental halophage 1. Environ Microbiol 9, 7, 1711-1723.

Santos, F., Moreno-Paz, M., Meseguer, I., López, C., Rossello-Mora, R., Parro, V., Antón, J., 2011.Metatranscriptomic analysis of extremely halophilic viral communities. ISME J 5, 10, 1621-1633.

Santos, F., Yarza, P., Parro, V., Briones, C., Antón, J., 2010. The metavirome of a hypersaline environment.Environ Microbiol 12, 11, 2965-2976.

Santos, F., Yarza, P., Parro, V., Meseguer, I., Rosselló-Móra, R., Antón, J., 2012. Culture-independentapproaches for studying viruses from hypersaline environments. Appl Environ Microbiol 78, 6, 1635-1643.

Saren, A., Ravantti, J.J., Benson, S.D., Burnett, R.M., Paulin, L., Bamford, D.H., Bamford, J.K., 2005. Asnapshot of viral evolution from genome analysis of the Tectiviridae family. J Mol Biol 350, 3, 427-440.

Schnabel, H., Schramm, E., Schnabel, R., Zillig, W., 1982a. Structural variability in the genome of phage ϕHof Halobacterium halobium. Mol Gen Genet 188, 3, 370-377.

Schnabel, H., 1984. An immune strain of Halobacterium halobium carries the invertible L segment ofphage ΦH as a plasmid. Proc Natl Acad Sci U S A 81, 4, 1017-1020.

Schnabel, H., Zillig, W., Pfaffle, M., Schnabel, R., Michel, H., Delius, H., 1982b. Halobacterium halobiumphage φH. EMBO J 1, 1, 87-92.

Seaman, P.F., Day, M.J., 2007. Isolation and characterization of a bacteriophage with an unusually largegenome from the Great Salt Plains National Wildlife Refuge, Oklahoma, USA. FEMS Microbiol Ecol 60, 1,1-13.

Senčilo, A., Jacobs-Sera, D., Russell, D.A., Ko, C., Bowman, C.A., Atanasova, N.S., Österlund, E., Oksanen,H.M., Bamford, D.H., Hatfull, G.F., 2013. Snapshot of haloarchaeal tailed virus genomes. RNA Biol 10, 5,803-816.

Senčilo, A., Roine, E., 2014. A glimpse of the genomic diversity of haloarchaeal tailed viruses. Front Microbiol12, 5, 84.

Senčilo, A., Paulin, L., Kellner, S., Helm, M., Roine, E., 2012. Related haloarchaeal pleomorphic virusescontain different genome types. Nucleic Acids Res 40, 12, 5523-5534.

Shen, P.S., Domek, M.J., Sanz-Garcia, E., Makaju, A., Taylor, R.M., Hoggan, R., Culumber, M.D., Oberg, C.J.,Breakwell, D.P., Prince, J.T., Belnap, D.M., 2012. Sequence and structural characterization of Great SaltLake bacteriophage CW02, a member of the T7-like supergroup. J Virol 86, 15, 7907-7917.

50

Sime-Ngando, T., Lucas, S., Robin, A., Tucker, K.P., Colombet, J., Bettarel, Y., Desmond, E., Gribaldo, S.,Forterre, P., Breitbart, M., 2011. Diversity of virus–host systems in hypersaline Lake Retba, Senegal.Environ Microbiol 13, 8, 1956-1972.

Snyder, J.C., Bolduc, B., Young, M.J., 2015. 40 years of archaeal virology: expanding viral diversity. Virology479, 369-378.

Snyder, J.C., Brumfield, S.K., Kerchner, K.M., Quax, T.E., Prangishvili, D., Young, M.J., 2013. Insights into aviral lytic pathway from an archaeal virus-host system. J Virol 87, 4, 2186-2192.

Soler, N., Krupovič, M., Marguet, E., Forterre, P., 2015. Membrane vesicles in natural environments: a majorchallenge in viral ecology. ISME J 9, 4, 793-796.

Sorek, R., Kunin, V., Hugenholtz, P., 2008. CRISPR—a widespread system that provides acquired resistanceagainst phages in bacteria and archaea. Nat Rev Microbiol. 6, 3, 181-186.

Stan-Lotter, H., Fendrihan, S., 2015. Halophilic archaea: life with desiccation, radiation and oligotrophy overgeological times. Life 5, 3, 1487-1496.

Stolt, P., Zillig, W., 1992. In vivo studies on the effects of immunity genes on early lytic transcription in theHalobacterium salinarium phage ϕH. Molec Gen Genet 235, 2-3, 197-204.

Strand, M.R., Burke, G.R., 2012. Polydnaviruses as symbionts and gene delivery systems. PLoS Pathog 8, 7,e1002757.

Strömsten, N.J., Bamford, D.H., Bamford, J.K.H., 2005. In vitro DNA packaging of PRD1: a commonmechanism for internal-membrane viruses. J Mol Biol 348, 3, 617-629.

Sun, L., Young, L.N., Zhang, X., Boudko, S.P., Fokine, A., Zbornik, E., Roznowski, A.P., Molineux, I.J.,Rossmann, M.G., Fane, B.A., 2014. Icosahedral bacteriophage ΦX174 forms a tail for DNA transportduring infection. Nature 505, 7483, 432-435.

Suttle, C.A., 2007. Marine viruses - major players in the global ecosystem. Nat Rev Microbiol 5, 10, 801-812.

Svirskaitė, J., Oksanen, H.M., Daugelavičius, R., Bamford, D.H., 2016. Monitoring physiological changes inhaloarchaeal cell during virus release. Viruses 8, 3, 59.

Tang, S.L., Nuttall, S., Ngui, K., Fisher, C., Lopez, P., Dyall-Smith, M., 2002. HF2: a double-stranded DNAtailed haloarchaeal virus with a mosaic genome. Mol Microbiol 44, 1, 283-296.

Thingstad, T.F., 2000. Elements of a theory for the mechanisms controlling abundance, diversity, andbiogeochemical role of lytic bacterial viruses in aquatic systems. Limnol Oceanogr 45, 6, 1320-1328.

Torsvik, T., 1982. Characterization of four bacteriophages for Halobacterium, with special emphasis onphage Hs1. In: Kandler, O. (editor). Archaebacteria. Gustav Fischer, Stuttgart, p. 351.

Torsvik, T., Dundas, I.D., 1980. Persisting phage infection in Halobacterium salinarium str. 1. J Gen Virol47, 1, 29-36.

Torsvik, T., Dundas, I.D., 1974. Bacteriophage of Halobacterium salinarium. Nature 248, 5450, 680-681.

Tschitschko, B., Williams, T.J., Allen, M.A., Páez-Espino, D., Kyrpides, N., Zhong, L., Raftery, M.J.,Cavicchioli, R., 2015. Antarctic archaea–virus interactions: metaproteome-led analysis of invasion,evasion and adaptation. ISME J 9, 9, 2094-2107.

Uchida, K., Kanbe, C., 1993. Occurrence of bacteriophages lytic for Pediococcus halophilus, a halophiliclactic-acid bacterium, in soy sauce fermentation. J Gen Appl Microbiol 39, 4, 429-437.

Uldahl, K.B., Jensen, S.B., Bhoobalan-Chitty, Y., Martinez-Alvarez, L., Papathanasiou, P., Peng, X., 2016. Lifecycle characterization of Sulfolobus monocaudavirus 1, an extremophilic spindle-shaped virus withextracellular tail development. J Virol 90, 12, 5693-5699.

Ventosa, A., Fernández, A.B., León, M.J., Sánchez-Porro, C., Rodriguez-Valera, F., 2014. The Santa Polasaltern as a model for studying the microbiota of hypersaline environments. Extremophiles 18, 5, 811-824.

Ventosa, A., Márquez, M.C., Sánchez-Porro, C., Rafael, R., 2012. Taxonomy of halophilic archaea andbacteria. In: Vreeland, R.H. (editor). Advances in understanding the biology of halophilicmicroorganisms. Springer, pp. 59-80.

Vogelsang-Wenke, H., Oesterhelt, D., 1988. Isolation of a halobacterial phage with a fully cytosine-methylated genome. Molec Gen Genet 211, 3, 407-414.

51

Vreeland, R.H., Straight, S., Krammes, J., Dougherty, K., Rosenzweig, W.D., Kamekura, M., 2002.Halosimplex carlsbadense gen. nov., sp. nov., a unique halophilic archaeon, with three 16S rRNA genes,that grows only in defined medium with glycerol and acetate or pyruvate. Extremophiles 6, 6, 445-452.

Wais, A.C., Kon, M., MacDonald, R.E., Stollar, B.D., 1975. Salt-dependent bacteriophage infectingHalobacterium cutirubrum and H. halobium. Nature 256, 5515, 314-315.

Walker, J.E., Saraste, M., Runswick, M.J., Gay, N.J., 1982. Distantly related sequences in the alpha- andbeta-subunits of ATP synthase, myosin, kinases and other ATP-requiring enzymes and a commonnucleotide binding fold. EMBO J 1, 8, 945-951.

Walsby, A.E., 1980. A square bacterium. Nature 283, 69 – 71.

Wikoff, W.R., Liljas, L., Duda, R.L., Tsuruta, H., Hendrix, R.W., Johnson, J.E., 2000. Topologically linkedprotein rings in the bacteriophage HK97 capsid. Science 289, 5487, 2129-2133.

Witte, A., Baranyi, U., Klein, R., Sulzner, M., Luo, C., Wanner, G., Kruger, D.H., Lubitz, W., 1997.Characterization of Natronobacterium magadii phage ΦCh1, a unique archaeal phage containing DNAand RNA. Mol Microbiol 23, 3, 603-616.

Yamaguchi, Y., Park, J., Inouye, M., 2011. Toxin-antitoxin systems in bacteria and archaea. Annu Rev Genet45, 61-79.

Yau, S., Lauro, F.M., DeMaere, M.Z., Brown, M.V., Thomas, T., Raftery, M.J., Andrews-Pfannkoch, C., Lewis,M., Hoffman, J.M., Gibson, J.A., Cavicchioli, R., 2011. Virophage control of Antarctic algal host-virusdynamics. Proc Natl Acad Sci U S A 108, 15, 6163-6168.

Yin, Y., Fischer, D., 2008. Identification and investigation of ORFans in the viral world. BMC Genomics 9,24-2164-9-24.

Young, R., Wang, N., Roof, W.D., 2000. Phages will out: strategies of host cell lysis. Trends Microbiol 8, 3,120-128.

Zhang, X., Sun, S., Xiang, Y., Wong, J., Klose, T., Raoult, D., Rossmann, M.G., 2012a. Structure of Sputnik, avirophage, at 3.5-Å resolution. Proc Natl Acad Sci U S A 109, 45, 18431-18436.

Zhang, Z., Liu, Y., Wang, S., Yang, D., Cheng, Y., Hu, J., Chen, J., Mei, Y., Shen, P., Bamford, D.H., Chen, X.,2012b. Temperate membrane-containing halophilic archaeal virus SNJ1 has a circular dsDNA genomeidentical to that of plasmid pHH205. Virology 434, 2, 233-241.

Zhaxybayeva, O., Stepanauskas, R., Mohan, N.R., Papke, R.T., 2013. Cell sorting analysis of geographicallyseparated hypersaline environments. Extremophiles 17, 2, 265-275.

INSTITUTE OF BIOTECHNOLOGY AND DIVISION OF GENERAL MICROBIOLOGYDEPARTMENT OF BIOSCIENCESFACULTY OF BIOLOGICAL AND ENVIRONMENTAL SCIENCESDOCTORAL PROGRAMME IN MICROBIOLOGY AND BIOTECHNOLOGY UNIVERSITY OF HELSINKI

Molecular Characterization of New Archaeal Viruses from High Salinity Environments

TATIANA DEMINA

dissertationes schola doctoralis scientiae circumiectalis, alimentariae, biologicae. universitatis helsinkiensis 26/2016

26/2016

Helsinki 2016 ISSN 2342-5423 ISBN 978-951-51-2706-8

Recent Publications in this Series

9/2016 Satu OlkkolaAntimicrobial Resistance and Its Mechanisms among Campylobacter coli and Campylobacter upsaliensis with a Special Focus on Streptomycin10/2016 Windi Indra MuziasariImpact of Fish Farming on Antibiotic Resistome and Mobile Elements in Baltic Sea Sediment11/2016 Kari Kylä-NikkiläGenetic Engineering of Lactic Acid Bacteria to Produce Optically Pure Lactic Acid and to Develop a Novel Cell Immobilization Method Suitable for Industrial Fermentations12/2016 Jane Etegeneng Besong epse NdikaMolecular Insights into a Putative Potyvirus RNA Encapsidation Pathway and Potyvirus Particles as Enzyme Nano-Carriers13/2016 Lijuan YanBacterial Community Dynamics and Perennial Crop Growth in Motor Oil-Contaminated Soil in a Boreal Climate14/2016 Pia RasinkangasNew Insights into the Biogenesis of Lactobacillus rhamnosus GG Pili and the in vivo Effects of Pili15/2016 Johanna RytiojaEnzymatic Plant Cell Wall Degradation by the White Rot Fungus Dichomitus squalens16/2016 Elli KoskelaGenetic and Environmental Control of Flowering in Wild and Cultivated Strawberries 17/2016 Riikka KylväjäStaphylococcus aureus and Lactobacillus crispatus: Adhesive Characteristics of Two Gram-Positive Bacterial Species18/2016 Hanna AarnosPhotochemical Transformation of Dissolved Organic Matter in Aquatic Environment – From a Boreal Lake and Baltic Sea to Global Coastal Ocean19/2016 Riikka PuntilaTrophic Interactions and Impacts of Non-Indigenous Species in the Baltic Sea Coastal Ecosystems20/2016 Jaana KuuskeriGenomics and Systematics of the White-Rot Fungus Phlebia radiata: Special Emphasis on Wood-Promoted Transcriptome and Proteome21/2016 Qiao ShiSynthesis and Structural Characterization of Glucooligosaccharides and Dextran from Weissella confusa Dextransucrases22/2016 Tapani JokiniemiEnergy Efficiency in Grain Preservation23/2016 Outi NyholmVirulence Variety and Hybrid Strains of Diarrheagenic Escherichia coli in Finland and Burkina Faso24/2016 Anu RiikonenSome Ecosystem Service Aspects of Young Street Tree Plantings25/2016 Jing ChengThe Development of Intestinal Microbiota in Childhood and Host-Microbe Interactions in Pediatric Celiac Diseases

YE

BT

AT

IAN

A D

EM

INA

Molecu

lar Ch

aracterization of N

ew A

rchaeal V

iruses from

High

Salin

ity En

vironm

ents