Вы находитесь на странице: 1из 16

REVIEWS

The enigmatic archaeal virosphere


David Prangishvili1, Dennis H. Bamford2, Patrick Forterre1, Jaime Iranzo3,
Eugene V. Koonin3 and Mart Krupovic1
Abstract | One of the most prominent features of archaea is the extraordinary diversity of their
DNA viruses. Many archaeal viruses differ substantially in morphology from bacterial and eukaryotic
viruses and represent unique virus families. The distinct nature of archaeal viruses also extends to
the gene composition and architectures of their genomes and the properties of the proteins that
they encode. Environmental research has revealed prominent roles of archaeal viruses in
influencing microbial communities in ocean ecosystems, and recent metagenomic studies have
uncovered new groups of archaeal viruses that infect extremophiles and mesophiles in diverse
habitats. In this Review, we summarize recent advances in our understanding of the genomic and
morphological diversity of archaeal viruses and the molecular biology of their life cycles and virus–
host interactions, including interactions with archaeal CRISPR–Cas systems. We also examine the
potential origins and evolution of archaeal viruses and discuss their place in the global virosphere.

Thermophilic The year 2017 marks the 40th anniversary of the dis- the epipelagic zone and the mesopelagic zone of the ocean
Requiring high temperatures covery of archaea1. The third domain of life, which at have shown that ~10% of the most abundant viruses in
for optimal growth. the time of its discovery included only a few species, these zones are associated with archaea9. Diverse archaeal
has been extensively populated in recent years with viruses have also been detected in benthic deep-sea
Acidophilic
Thriving under highly acidic
numerous thermophilic, acidophilic, alkaliphilic, h­ alophilic, ­ecosystems, where their turnover is as fast as 2–3 days10.
conditions. ­ ethanogenic and ammonia-oxidizing archaea.
m Virus predation is one of the primary causes of
Moreover, the exploration of microbial diversity through microbial mortality, with major consequences for
Alkaliphilic culture-independent approaches has substantially global biogeochemistry. It was recently shown that in
Thriving under highly alkaline
expanded our understanding of archaeal diversity and oceanic surface sediments across 1,000–10,000 m water
conditions.
uncovered numerous species that may represent many depths, viral infection has a higher impact on archaea
Halophilic new phyla and orders of the Archaea2. than on bacteria and mainly affects members of the
Requiring high levels of sodium Culture-independent approaches have radically Thaumarchaeota. Moreover, in the top 50 cm of ocean
chloride for growth. changed the perception of the typical habitats of archaea. sediment, virus-induced lysis of archaea was estimated
The original notion that archaea thrive only in environ- to account for up to one-third of the total microbial bio-
ments with extreme conditions was challenged and over- mass that is killed each year, resulting in the release of
turned by the discovery of archaea in a diverse range ~0.3–0.5 gigatonnes of carbon per year 10. This realiza-
of terrestrial and aquatic environments and even within tion has changed the perception of archaeal viruses from
1
Department of Microbiology, and on the human body 3. Remarkably, in deep-sea eco- interesting curiosities of the virosphere to prominent
Institut Pasteur, 25 rue du Dr systems, which constitute ~90% of the global biosphere, players in the biosphere.
Roux, Paris 75015, France. the abundance of archaea is comparable to that of bac- In addition to having important roles in the function-
2
Department of Biosciences, teria4, and archaeal species that oxidize ammonium to ing of deep-sea ecosystems and global biogeochemical
University of Helsinki,
Helsinki 00014, Finland.
nitrate (members of the phylum Thaumarchaeota) rep- cycles, archaeal viruses attract attention owing to their
3
National Center for resent one of the most abundant cell types in oceans5. remarkable diversity, unique morphologies and ability
Biotechnology Information, The high abundance of archaea in deep-sea ecosystems to withstand extreme environments. Known archaeal
National Library of Medicine, and their metabolic diversity 6 suggest that they have viruses have been isolated from terrestrial and marine
Bethesda, Maryland 20894,
a substantial impact on global nitrogen and carbon thermal environments with temperatures exceeding
USA.
cycles. Moreover, it has been demonstrated that deep- 80 °C and hypersaline lakes with nearly saturating con-
Correspondence to D.P. 
and M.K.
sea benthic archaea have an important role in recycling centrations of sodium chloride. The viruses that were iso-
david.prangishvili@pasteur.fr; ­sedimentary organic compounds7,8. lated from thermal environments infect hyperthermophilic
krupovic@pasteur.fr Concurrently, the abundance of archaeal viruses in members of the orders Sulfolobales, Desulfurococcales
doi:10.1038/nrmicro.2017.125 the oceans has also been documented. Metagenomic and Thermoproteales from the phylum Crenarchaeota
Published online 10 Nov 2017 analyses of double-stranded DNA (dsDNA) viruses in as well as the order Thermococcales from the phylum

724 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Box 1 | Metagenomics of archaeal viruses


Many metagenomic studies have considerably expanded our perception of the global diversity of archaeal viruses.
These studies were carried out on a variety of environmental samples from extreme habitats that are dominated by
archaea, such as terrestrial hot springs141, hypersaline lakes140 and salt pans142, but also from various oceanic sites9,10,143–145
from which archaeal viruses had not been previously isolated. Several complete or near-complete genome sequences
have been assembled for hyperthermophilic archaeal viruses of the families Rudiviridae23,141,146, Fuselloviridae147,
Lipothrixviridae141, Bicaudaviridae141 and Ampullaviridae141 as well as hyperhalophilic viruses of the order
Caudovirales148,149 and the genus Salterprovirus142. Two new genomes of uncultivated ampullaviruses are of particular
value141 because the ampullavirus Acidianus bottle-shaped virus (ABV) was the sole member of the family for over a
decade83. Approximately 50 complete genomes of head–tailed haloviruses have been cloned and sequenced148–150,
providing the first genomic insights into viruses of Haloquadratum walsbyi, a squared archaeon that is found in
hypersaline environments. These uncultivated viruses have been further investigated using metatranscriptomics,
which indicated dynamic virus–host interactions in hypersaline settings151. Furthermore, several genomes have been
assembled for hyperthermophilic archaeal viruses that cannot be assigned to existing families141,152, indicating that our
knowledge of viral diversity in extreme geothermal environments remains limited. In particular, a positive-strand RNA
virus genome that is distantly related to eukaryotic RNA viruses has been assembled from metagenomic sequences from
hot springs in Yellowstone National Park, USA, that are dominated by the crenarchaeon Sulfolobus solfataricus, but the
actual host of this has not been confirmed153.
Perhaps the greatest advance that has been facilitated by metagenomic studies has been the discovery of the viral
Methanogenic diversity that is associated with ubiquitous, environmentally important archaea, many of which remain uncultivated.
Producing methane as a Most notably, over 50 genomes have been assembled for viruses, putatively called Magroviruses, associated with Marine
metabolic by-product in anoxic Group II Euryarchaeota, some of the most abundant microorganisms in ocean surface waters144,145. Magroviruses have
conditions. large genomes of ~100 kb; they are most closely related to members of the Caudovirales that infect halophilic archaea
and have similar-sized genomes12. Magroviruses encode a nearly complete DNA replication apparatus; overall,
Benthic
archaeal viruses appear to follow the general trend that is observed among double-stranded DNA (dsDNA) viruses,
Related to the ecological
region at the lowest sea level,
whereby viruses with larger genomes approach self-sufficiency for genome replication154.
including the sediment surface Thaumarchaeota is another group of archaea that is ubiquitous in aquatic and terrestrial environments. Except for
and some subsurface layers. a single putative provirus53, no viruses have been described for this group of archaea. However, two putative
thaumarchaeal virus genomes of the order Caudovirales have been obtained through single-cell genomics and fosmid
Epipelagic zone sequencing approaches155,156. Uncultured viruses have also been described for nano-sized archaea known as Archaeal
The illuminated zone at the Richmond Mine acidophilic nanoorganisms (ARMAN)157,158. Notably, one of these viruses, ANMV1, also a member of the
surface of the sea where Caudovirales, has been found to encode a diversity-generating retroelement, which uses mutagenic reverse transcription
enough light is available for and retrohoming with the potential to generate 1018 variants of the tail fibre ligand-binding domain158.
photosynthesis.
Although the uncultivated viruses that are described above remain unclassified, the International Committee on
Mesopelagic zone
Taxonomy of Viruses (ICTV) has recently issued a recommendation to classify viruses that are known solely by their
The zone close to the sea genome sequence159. Owing to this change, the number of recognized species of archaeal viruses is expected to
surface in which light increase substantially in the future.
penetrates but is insufficient
for photosynthesis.

Hyperthermophilic Euryarchaeota, whereas viruses from saline waters infect last archaeal common ancestor13. Alternatively, archaeal
Having an optimal growth hyperhalophilic members of the class Halobacteria, also RNA viruses may exist but have yet to be discovered
temperature at or above from the Euryarchaeota (see Supplementary information (BOX 1). Owing to the unique features of their virions and
80 °C.
S1 (table) for the complete list of characterized archaeal genomes, characterized archaeal viruses are ­classified
Fosmid sequencing viruses and their taxonomic classification). In addition, a into 17 virus families (TABLE 1).
Sequencing of large DNA number of uncultured archaeal viruses have been found In this Review, we summarize the current knowledge
fragments cloned into a fosmid. by using m­ etagenomics but remain unclassified (BOX 1). of archaeal virus morphotypes and structures, their
All archaeal viruses that have been isolated to date genome architectures, their mechanisms of genome
Mutagenic reverse
transcription and have either dsDNA or single-stranded DNA (ssDNA) replication and virion egress and their interactions with
retrohoming genomes. Their genomes are generally small and range CRISPR–Cas systems, and we discuss their evolution
Targeted replacement of a in size from 5.2 kb for the clavavirus Aeropyrum pernix and potential origins.
variable repeat coding region bacilliform virus 1 (APBV1)11, which is among the small-
within a gene with a sequence
derived from reverse
est known dsDNA genomes, to 144 kb for Halogranum Virion morphology
transcription of a cognate tailed virus 1 (HGTV1), a head–tailed virus that infects Archaeal viruses can be broadly divided into two cate-
non-coding template repeat. Halogranum spp.12. The dsDNA genomes are either cir- gories: archaea-specific viruses that have no structural
cular or linear. Archaeal viruses with linear genomes or genetic counterparts among bacterial or eukaryotic
Hyperhalophilic
employ different strategies for the protection and repli- viruses and cosmopolitan viruses that possess struc-
Requiring extremely high levels
of sodium chloride for growth. cation of their genome ends, including covalently closed tural and genetic features that are similar to those of
hairpins and proteins that are covalently attached to the ­bacterial and eukaryotic viruses.
Last archaeal common termini. Considering the low stability of long RNA mol­
ancestor ecules at high temperature, at least in the extracellular Archaea-specific viruses
The most recent population
of organisms from which all
environment, it has been suggested that the absence of Nearly all known archaea-specific viruses infect hyper-
extant archaea have a common RNA viruses among known archaeal viruses could be thermophiles of the phylum Crenarchaeota. They have
descent. due to the postulated hyperthermophilic lifestyle of the a range of unique morphologies that have never been

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 725


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Capsid observed among viruses that infect bacterial or eukary- Bicaudaviridae are generally highly pleomorphic and
The protein shell that encloses otic cells, such as shapes that resemble bottles, spindles,
have tails of different lengths at one or both pointed ends
the genetic material of the droplets and coils, in addition to morphologies c­ ommon of the spindle-shaped body 18–23. For the type species of
virus. in the virosphere, such as spheres and filaments. the Bicaudaviridae, Acidianus two-tailed virus (ATV), the
Convergent evolution
tails, which were shown to develop extracellularly, consist
The independent evolution of Viruses with unique morphologies. Arguably, archaeal of helically arranged globular subunits24,25.
similar features in species viruses with the most unusual morphologies are mem- Another unusual virion morphology is observed
of different lineages. bers of the family Ampullaviridae. The champagne-­ in members of the family Guttaviridae. Guttaviruses
bottle-like shape of their virions is determined by the resemble spindles in which one of the two pointed ends
extraordinary manner in which the linear dsDNA has been rounded, rendering them droplet-shaped26,27
genome is condensed by the capsid proteins into a cone- (FIG. 1a). Notably, guttaviruses share several genes in
shaped inner core and encased with the lipid-­containing common with fuselloviruses28, suggesting that the two
envelope14 (FIG. 1a). The uncommon coil-like virion families of viruses are evolutionarily related.
shape of members of the family Spiraviridae is also
determined by a special method of genome packing 15; Viruses with common morphologies. Despite their
the circular nucleoprotein filament formed from ssDNA simi­lar shape, filamentous archaeal viruses with dsDNA
and capsid proteins is condensed into a rope-like struc- genomes are unrelated to filamentous bacterial and
ture that is further ­c ondensed into a ­higher-order eukaryotic viruses. These filamentous archaeal viruses
helix (FIG. 1a). belong to the families Rudiviridae, Lipothrixviridae,
Some of the most widespread and abundant Tristromaviridae and Clavaviridae. The tubular, rigid
archaea-specific viruses have spindle-shaped viri- and non-enveloped virions of the Rudiviridae are
ons. Viruses that have this morphology belong to formed by the condensation of the linear dsDNA
the Fuselloviridae or the Bicaudaviridae 16 (FIG.  1a) . through the binding of multiple copies of the single
Members of these two families share little sequence MCP that adopts an unusual four helix-bundle fold.
similarity, and their virions consist of unrelated major The ends of the viruses have three thin fibres that
capsid proteins (MCPs), suggesting that their shared are involved in host cell recognition29–32 (FIG. 1a). The
morphology is the result of convergent evolution. The flexible enveloped virions of Lipothrixviridae contain
virions of the Fuselloviridae have a bundle of filaments a nucleoprotein core that is formed by linear dsDNA
at one of the two pointed ends and show a certain and multiple copies of two MCPs, which are structur-
degree of pleo­morph­icity 17. The mature virions of the ally similar to each other and to the rudivirus MCP33–35.

Table 1 | Representative viruses of the Archaea


Family Species Host Genome topology Refs
and length (bp)
Ampullaviridae Acidianus bottle-shaped virus Acidianus convivator L 23,900 14
Bicaudaviridae Acidianus two-tailed virus A. convivator C 62,730 25
Spiraviridae Aeropyrum coil-shaped virus Aeropyrum pernix C* 24,893 nt 15
Fuselloviridae Sulfolobus spindle-shaped virus 1 Sulfolobus shibatae C 15,465 162
Guttaviridae Sulfolobus neozealandicus Sulfolobus neozealandicus NA 26
droplet-shaped virus
Aeropyrum pernix ovoid virus 1‡ A. pernix C 13,769 27
Rudiviridae Sulfolobus islandicus rod-shaped virus 2 Sulfolobus islandicus L 35,450 29
Lipothrixviridae Acidianus filamentous virus 1 Acidianus hospitalis L 21,080 78
Tristromaviridae Pyrobaculum filamentous virus 1 Pyrobaculum arsenaticum L 17,714 38
Clavaviridae Aeropyrum pernix bacilliform virus 1 A. pernix C 5,278 11
Globuloviridae Pyrobaculum spherical virus Pyrobaculum sp. D11 L 28,337 160
Portogloboviridae Sulfolobus polyhedral virus 1 S. shibatae L 20,222 39
Pleolipoviridae Halorubrum pleomorphic virus 1 Halorubrum spp. C* 7,048 nt 44
Myoviridae Halorubrum sodomense tailed virus 2 Halorubrum sodomense L 68,187 49
Siphoviridae Haloarcula vallismortis tailed virus 1 Haloarcula vallismortis L 102,32 49
Podoviridae Haloarcula sinaiiensis tailed virus 1 Haloarcula sinaiiensis L 32,189 50
Sphaerolipoviridae Haloarcula hispanica virus SH1 Haloarcula hispanica L 30,898 163
Turriviridae Sulfolobus turreted icosahedral virus Sulfolobus solfataricus C 17,663 164
Listed are archaeal viruses shown in FIG. 1. All genomes — except those marked with asterisks — are double-stranded DNA.
C, covalently closed circular DNA; L, linear DNA; NA, not available. *Single-stranded DNA.‡The only member of the family with a
sequenced genome.

726 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

The structures at the ends of filamentous virions differ long, non-contractile tails, whereas the sheath-covered
among family members36; for instance, the structures of tails of myoviruses are contractile and podoviruses
Acidianus filamentous virus 1 (AFV1) ends are claw-like have short tails. Members of all three families of caudo­
in appearance (FIG. 1a). The high structural similarity of viruses have been isolated from archaea49,50. Although
the MCPs, as well as the fairly high numbers of homolo­ the majority of these isolates infect hyperhalophiles,
gous genes (see below), indicate a common ances- head–tailed viruses have also been isolated from
try of the non-enveloped Rudiviridae and enveloped methano­gens51. Furthermore, related proviruses have
Lipothrixviridae, and the two families have ­therefore been identified in the genomes of a wide range of eury­
been classified together in the order Ligamenvirales 37. archaea, including members of the classes Halobacteria,
The filamentous virions of the Tristromaviridae Methanomicrobia, Archaeoglobi, Methanobacteria
contain three MCPs, of which two condense the linear and Methanococci 52 , as well as members of the
dsDNA into a helical nucleoprotein core and the third phylum Thaumarchaeota53.
forms a protein sheath over this nucleoprotein core Cosmopolitan, tailless icosahedral viruses of
and mediates its interaction with the lipid-containing archaea are related to bacterial and eukaryotic icosa-
­envelope of the virion. Both ends of the virion are dec- hedral viruses with capsid proteins that have a double
orated with bundles of thin filaments38 (FIG. 1a). The or a single jelly-roll fold. These viruses are classified
bacilliform Clavaviridae are non-enveloped and encode into two families, Turriviridae and Sphaerolipoviridae,
a single MCP; the ends of the virion are asymmetric — respectively. The latter family consists of three genera,
one is rounded, and the other is slightly pointed11. The of which two include viruses that infect halophilic
virion does not have any terminal filaments, and it is archaea, whereas viruses from the other genus infect
unclear how it interacts with the host cell (FIG. 1a). bacteria54. By contrast, turriviruses have been found to
Spherical archaea-specific viruses are classified infect only crenarchaea from the genus Sulfolobus 55. The
into the families Globuloviridae and Portogloboviridae. overall virion organization is similar in the two groups
Virions of the Globuloviridae have a lipid-­containing of viruses: the icosahedral protein capsid encases a
envelope and a superhelical nucleoprotein core protein­aceous membrane vesicle that contains a dsDNA
that contains linear dsDNA. In the virions of the genome, which is either linear or circular 54–56. The main
Portogloboviridae, the circular dsDNA genome is con- difference between viruses in the two families is that
densed by capsid proteins into a spherical nucleopro- the turrivirus capsids are built from a single MCP55,57,
tein coil, which forms an inner core of the virion. The whereas sphaerolipoviruses have two p ­ aralogous
spherical inner core is surrounded by a lipid membrane MCPs54,56.
and further encased by an outer icosahedral protein The morphological diversity of archaeal viruses is
shell39 (FIG. 1a). remarkable, especially when compared with bacterial
Viruses of the family Pleolipoviridae have envel- viruses, which have a limited range of morphologies,
oped pleomorphic virions that resemble membrane despite their enormous genetic variability. The vast
vesicles that are widely produced by archaea40 (FIG. 1a). morphological landscape of the archaeal virosphere
These viruses lack a distinct nucleocapsid that is typi- is shaped by only a small number of species, which
cal of enveloped viruses; instead, the naked genome is represent 2% of the >6,000 species of bacterial viruses
packaged into a membrane vesicle that contains two that have been characterized 46. Moreover, no new
membrane-spanning viral proteins41,42. Remarkably, morphologies of bacterial viruses have been identified
pleolipovirus particles can contain either linear dsDNA since the 1970s, despite the isolation of hundreds of
or circular ssDNA or dsDNA genomes43–45. new species. By contrast, the variety of archaea-­specific
viral morphotypes is continuously expanding with the
Cosmopolitan viruses description of viruses from new environments and
The cosmopolitan viruses of the archaeal virosphere hosts. The evolutionary factors that drive this strik-
include head–tailed viruses — viruses with icosahedral ing diversification of the archaeal virion shape are not
capsids (heads) and helical appendages (tails) attached yet understood.
to them — of the order Caudovirales and tailless
icosa­hedral viruses of the families Sphaerolipoviridae Virion structures of archaeal viruses
and Turriviridae (FIG. 1b). All of these viruses, except The remarkable morphological diversity of archaeal
for turri­v iruses, infect members of the phylum viruses raises important questions regarding their ori-
Euryarchaeota; the biological basis for such host speci­ gins and their relationship to viruses that infect organ-
ficity remains unknown and difficult to understand, isms from other domains of life. Several recent studies
as some of the Euryarchaeota inhabit extreme habitats tried to address these questions by using cryo-electron
similar to those of Crenarchaeota. microscopy (cryo‑EM) to determine the structures of
Proviruses
Viral genomes integrated into Archaeal members of the order Caudovirales several archaeal viruses (FIGS 2,3). Structural studies have
the host chromosome. are morphologically indistinguishable from head– revealed previously unexpected evolutionary relation-
tailed bacterial viruses 46–48. The three families of ships between viruses, which could not be predicted
Jelly-roll fold the Caudovirales — Siphoviridae, Myoviridae and using genome sequence data alone, and have facilitated
A structural protein fold
composed of eight β‑strands
Podoviridae — were originally classified according to our understanding of the molecular details that under-
arranged in two antiparallel the morphology of their tails, which serve as cell rec- lie the ability of archaeal viruses to withstand extreme
four-stranded β‑sheets. ognition and penetration devices. Siphoviruses have environments.

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 727


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Homology modelling Structures of archaeal viruses from all three families have the HK97‑like fold50, a prevalent MCP topology
The construction of an of the Caudovirales were determined using cryo‑EM. that is found in all tailed bacteriophages and in eukar-
atomic-resolution model of the The capsids of the myovirus Haloarcula vallismortis yotic herpes­viruses (FIG. 2); these findings indicate that
protein from its amino acid tailed virus 1 (HVTV1) and the siphovirus Halorubrum the lineage of viruses with the HK97‑like fold in their
sequence and an experimental
three-dimensional structure of
sodomense tailed virus 2 (HSTV2) were reconstructed MCPs spans all three domains of life58–61.
a related homologous protein. to ~10 Å resolution and the podovirus HSTV1 capsid to Cryo‑EM structures have also been determined
8.9 Å resolution49,50. The higher-resolution reconstruc- for viruses that belong to two groups of tailless archaeal
tion of the HSTV1 capsid was sufficient to identify viruses with icosahedral capsids, namely, the Turriviridae
the polypeptide backbone in the structure of the MCP, and Sphaerolipoviridae. The near-atomic structure of the
and it validated earlier hypotheses on the evolutionary turrivirus Sulfolobus turreted icosahedral virus (STIV)
relationship between viruses that infect hosts from the virion55 shows that the virion is constructed from several
three domains of life. Homology modelling and sequence capsid proteins, most of which have the jelly-roll fold. The
analyses52 revealed that MCPs of tailed archaeal viruses MCP has the double jelly-roll fold, which is widespread

a Archaea-specific viruses
Ampullaviridae (ABV) Bicaudaviridae (ATV)

Fuselloviridae Guttaviridae Tristromaviridae Clavaviridae Globuloviridae


Spiraviridae (ACV) (SSV1) (SNDV) (PFV1) (APBV1) (PSV)

Lipothrixviridae (AFV1) Rudiviridae (SIRV2) Portogloboviridae (SPV1)

Pleolipoviridae (HRPV1)

b Cosmopolitan archaeal viruses Sphaerolipoviridae


Myoviridae (HSTV2) Siphoviridae (HVTV1) Podoviridae (HSTV1) (SH1) Turriviridae (STIV)

Nature Reviews | Microbiology


728 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro
©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

A‑form in dsDNA viruses from all three domains of life57,62–64, never been observed in a biological entity. However,
One of the three major forms suggesting an evolutionary relationship between STIV the parameters of DNA packaging are different in the
of double-stranded DNA, with and these other viruses (FIG. 2). Unlike members of the two virions: the twist of A‑form DNA in SIRV2 is
a 23 Å helical diameter and Turriviridae, sphaerolipoviruses encode two MCPs, both 11.2 bp/turn, whereas in AFV1, it is 10.8 bp/turn.
11 bp per helix turn.
of which have the single jelly-roll fold and form homo­ In the structures of the SIRV2 and AFV1 virions, the
dimers and heterodimers that are the building blocks nucleoprotein helix is composed of asymmetric units,
of the icosahedral capsid65,66 (FIG. 2). The two MCPs are each of which contains two molecules of the MCP:
oriented in the capsid lattice similarly to the charac- a homodimer in the case of SIRV2 and a heterodimer in
teristic orientation in viruses with MCPs that have the the case of AFV1. The helices from neighbouring protein
double jelly-roll fold, such as turriviruses55. Furthermore, molecules are packed in an antiparallel, interdigitated
members of Sphaerolipoviridae and Turriviridae encode arrangement and wrap around the DNA. The amino
homologous genome-­packaging ATPases and are thus terminus of the SIRV2 MCP, which is unstructured in
considered to share a common ancestor 67,68. solution, is folded in the virion into a helix-turn-helix
The hypothesis that the unusual morphologies of structure and interacts with the DNA phosphate back-
archaeal viruses reflect unknown forms of virion organ- bone through conserved polar and hydrophobic residues.
ization is supported by three-dimensional reconstruc- In addition to the hydrophobic protein–­protein inter­
tions of filamentous virions of the rudivirus Sulfolobus actions across helical turns, this arrangement ensures that
islandicus rod-shaped virus 2 (SIRV2) and the lipothrix- the SIRV2 DNA is completely encapsidated by protein
virus AFV1 at near-atomic resolution33,69 (FIG. 3). In both and inaccessible to solvent (FIG. 3b), thereby maintaining
virions, the dsDNA was found in the A‑form, which has its integrity under extreme conditions. By contrast, in
the AFV1 virion, the protein–protein interactions across
helical turns are absent, resulting in looser packaging
◀ Figure 1 | Electron micrographs of archaeal viruses. a | Archaea-specific viruses. Virus of the DNA, a thinner filament and greater flexibility of
families and species are indicated. Acidianus filamentous virus 1 (AFV1), the inset shows the virion compared to SIRV2 (FIG. 3d). Consequently, the
the terminal structure. Acidianus two-tailed virus (ATV), the arrows indicate virion tails lipid envelope of AFV1 seems to protect the viral DNA in
(left), which undergo extracellular development (right). Pyrobaculum spherical virus (PSV),
highly acidic environments (FIG. 3e).
the arrows indicate spherical protrusions. Negative stain with uranyl acetate, except for
Aeropyrum coil-shaped virus (ACV), Halorubrum pleomorphic virus 1 (HRPV1), Sulfolobus In the structure of the AFV1 virion, the lipid enve-
islandicus rod-shaped virus 2 (SIRV2) and Sulfolobus polyhedral virus 1 (SPV1), which are lope is only 20–25 Å in thickness69, half the thickness of
in the vitreous ice. Bars, 100 nm. b | Cosmopolitan archaeal viruses. Halorubrum archaeal membranes70. The major lipids that are selec-
sodomense tailed virus 2 (HSTV2), arrow and arrowhead point to the non-contracted and tively incorporated into the AFV1 membrane  are
contracted tails, respectively; Haloarcula hispanica virus SH1, open arrows point to the glycerol dibiphytanyl glycerol tetraethers that lack a
claw-like spikes present at the five folds of the icosahedral capsid. Bars, 100 nm. cyclopentane moiety; as shown by molecular model-
ABV, Acidianus bottle-shaped virus; APBV1, Aeropyrum pernix bacilliform virus 1; ling, they adopt a U‑shaped, ‘horseshoe’ conformation
ATV, Acidianus two-tailed virus; HSTV1, Haloarcula sinaiiensis tailed virus 1; HVTV1, in the virion membrane (FIG. 3e,f). In addition to the
Haloarcula vallismortis tailed virus 1; PFV1, Pyrobaculum filamentous virus 1; SNDV canonical lipid bilayer of the Bacteria and the Eukarya
Sulfolobus neozealandicus droplet-shaped virus; SSV1, Sulfolobus spindle-shaped virus 1;
and the archaeal lipid monolayer, the viral horseshoe
STIV, Sulfolobus turreted icosahedral virus. Part a (ABV) is adapted with permission from
American Society for Microbiology (REF. 14): Häring, M. et al. Viral diversity in hot springs membrane represents a third, previously unobserved
of Pozzuoli, Italy, and characterization of a unique archaeal virus, Acidianus bottle-shaped type of biological membrane. Notably, the lipid envelope
virus, from a new family, the Ampullaviridae. J. Virol. 79, 9904–9911 (2005) http://dx.doi. of the fusellovirus (SSV1) also appears to be thinner than
org/10.1128/JVI.79.15.9904-9911.2005. Part a (ATV) left-panel is courtesy of R. Aramayo, the host cytoplasmic membrane and has a lipid composi-
and the right-panel is adapted with permission from REF. 25, Elsevier. Part a (ACV) is tion that closely resembles that of AFV1 (REFS 69,71,72),
adapted with permission from REF. 15, Proceedings of the National Academy of Sciences. suggesting that the U‑shaped conformation is a wide-
Part a (APBV1) is adapted with permission from REF. 11, Elsevier. Part a (PSV) is adapted spread ­feature of archaeal virus membranes Notably,
with permission from REF. 160, Elsevier. Part a (AFV1) is adapted with permission from the majority of viruses that infect hyperthermophiles
REF. 78, Elsevier. Part a (PFV1) is adapted with permission from REF. 38, Proceedings of the
are enveloped, suggesting that lipid membranes might
National Academy of Sciences. Part a (SSV1) is adapted with permission from REF. 161,
confer thermostability and provide a mechanism for
Springer. Part a (HRPV1) is adapted with permisson from the American Society for
Microbiology (REF. 41): Pietilä, M. K. et al. Virion architecture unifies globally distributed these viruses to enter into and exit from the host cell73.
pleolipoviruses infecting halophilic archaea. J. Virol. 86, 5067–5079 (2012) http://dx.doi. The structures of spindle-shaped archaeal viruses
org/10.1128/JVI.06915-11. Part a (SPV1) is adapted with permission from the American SSV1 and Haloarcula hispanica virus 1 (His1), which
Society for Microbiology (REF. 39): Liu, Y. et al. A novel type of polyhedral viruses infecting infect hyperthermophilic-acidophilic and hyper­
hyperthermophilic archaea. J. Virol. 91, e00589-17 (2017) http://dx.doi.org/10.1128/ halophilic hosts, respectively, have also been recon-
JVI.00589-17. Part a (SNDV) is adapted with permission from REF. 26, Elsevier. Part a (SIRV2) structed74,75. Owing to the inherent flexibility of SSV1
is adapted with permission from REF. 32, Biochemical Society Transactions http://dx.doi. and His1 virions, only fairly low-resolution models
org/10.1042/BST20120313. Part b (HSTV1) is adapted with permission from the American of entire virions could be obtained (~32 Å and ~57 Å,
Society for Microbiology (REF. 49): Pietilä, M. K. et al. Insights into head-tailed viruses respectively)76,77. Nevertheless, these models revealed
infecting extremely halophilic archaea. J. Virol. 87, 3248–3260 (2013) http://dx.doi.org/
interesting features of these virions; whereas the body
10.1128/JVI.03397-12; and from REF. 50, Proceedings of the National Academy of
Sciences. Part b (HSTV2) is adapted with permission from the American Society for of the His1 and SSV1 virions appears smooth and
Microbiology (REF. 49): Pietilä, M. K. et al. Insights into head-tailed viruses infecting lacks major features, the terminal structures that dec-
extremely halophilic archaea. J. Virol. 87, 3248–3260 (2013) http://dx.doi.org/10.1128/ orate one of the two pointed ends of the virions, and
JVI.03397-12. Part b (SH1) is adapted with permission from REF. 65, Proceedings of the are presumably involved in viral attachment, display
National Academy of Sciences. Part b (STIV) is courtesy of M. Young. six-fold symmetry.

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 729


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

PBCV1 HCMV (’floor domain’)

Eukarya

Adenoviridae, Lavidaviridae, Herpesviridae


NCLDV, Polintons

PRD1 HK97
P23-77

Bacteria

Myoviridae, Podoviridae,
Sphaerolipoviridae Corticoviridae, Tectiviridae Siphoviridae

STIV HSTV1
HCIV1

Archaea

Myoviridae, Podoviridae,
Sphaerolipoviridae Turriviridae Siphoviridae

SJR fold DJR fold HK97 fold

Figure 2 | Major capsid proteins of cosmopolitan archaeal viruses. Homologous viral proteins with
Nature the HK97‑like
Reviews fold,
| Microbiology
double jelly-roll (DJR) fold and single jelly-roll (SJR) fold are vertically aligned and arranged horizontally according to the
domain of life to which their hosts are classified. Names of families of bacterial, archaeal and eukaryotic viruses sharing
homologous capsid proteins are listed under the corresponding structures, and virus names are provided above the
structural models. The structures are coloured according to the secondary structure elements: α‑helices, red; β‑strands,
blue; and random coil, grey. The two major capsid proteins of bacterial and archaeal sphaerolipoviruses are boxed (dashed
lines). Only the ‘floor’ domain61 of the human cytomegalovirus (HCMV) capsid protein is shown. The X‑ray structures of the
capsid proteins of HCIV1 and HSTV1 are not available and are represented with homology-based models50,56. HCIV1,
Haloarcula californiae icosahedral virus 1; HK97, Escherichia virus HK97; HSTV1, Haloarcula sinaiiensis tailed virus 1;
NCLDV, nucleo-cytoplasmic large DNA viruses; P23‑77, Thermus virus P23‑77. PBCV1, Paramecium bursaria Chlorella
virus 1; PRD1, Pseudomonas virus PRD1; STIV, Sulfolobus turreted icosahedral virus. RCSB Protein Data Bank (PDB)
accession numbers for the major capsid protein structures: PBCV1, PDB entry 1J5Q; HCMV, PDB entry 5VKU; P23‑77 PDB
entry 3ZMN (left) and 3ZN4 (right); PRD1 PDB entry 1HX6; HK97 PDB entry 1OHG; STIV PDB entry 1BBD.

Virus–host interactions in archaea primary stage of virus–host interactions29,78,79. This mode


The morphological diversity of archaeal viruses is of interaction has been confirmed for the rudivirus
matched by the range of mechanisms that are employed SIRV2, which initiates infection by binding to the tips
by these viruses to interact with their hosts. In this sec- of the pilus-like appendages on Sulfolobus islandicus cells
tion, we outline our current understanding of virus–host through three terminal fibres on the virion79 (FIG. 4a).
interactions during archaeal viral life cycles, including The adsorption of SIRV2 is remarkably rapid: the viri-
attachment and virus entry, genome replication, and ons irreversibly bind to the appendages within seconds
virion morphogenesis and egress. and reach the cell surface within 1 minute (FIG. 4a). The
cell surface proteins and type IV secretion proteins that
Attachment and virus entry are involved in the SIRV2 entry have been identified80,
Information on the early stages of infection by but the mechanisms of virion translocation and DNA
archaea-specific viruses is scarce. Filamentous viruses ­injection remain unknown.
that are bound through their ends to cellular appendages The spindle-shaped virions of Fuselloviridae and
have been frequently observed using electron micros- Bicaudaviridae do not attach to cellular appendages.
copy, which suggests that such contacts represent the Instead, they are often observed bound to cellular

730 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Capsid Terminal fibres Nucleoprotein core ‘Claws’ fragments or membrane-derived vesicles, which implies
that these viruses interact with receptors on the host cell
surface to gain entry into cells. For a putative member
Lipid envelope of the Bicaudaviridae, Sulfolobus monocaudavirus 1
b (SMV1), this primary interaction has been suggested to
be followed by fusion between the virion envelope and
the cell membrane21. The fusion mechanism for entry
is also probably employed by members of the family
Pleolipoviridae44.
Similarly to bacterial viruses of the Caudovirales,
the first stages of infection by archaeal head–tailed
viruses appear to be mediated by the tail. For the myo-
virus φCh1, which infects the haloalkaliphilic archaeon
Natrialba magadii, the role of virion tail fibres in primary
c adsorption has been experimentally demonstrated81. The
C N N genome of φCh1 contains an invertible region that con-
C
tains genes that encode the tail fibre proteins separated
by an invertase gene. The inversion leads to exchange
of the carboxy-termini of the tail fibre proteins, thereby
creating protein variants with distinct cell surface
C
­adhesion specificities81.
C
N N Genome replication and packaging
The mechanisms of genome replication have been
d
experimentally studied for only a small number of
archaeal viruses, and in many cases, the mode of rep-
lication has been inferred from the presence of genome
replication-associated genes in viral genomes. Archaeal
viruses of the order Caudovirales, which have the lar­
gest known genomes among archaeal viruses, encode

Figure 3 | Virion organization of filamentous viruses


SIRV2 and AFV1. a | Schematic representations of
e Sulfolobus islandicus rod-shaped virus 2 (SIRV2) (left) and
Acidianus filamentous virus 1 (AFV1) (right). b | SIRV2
encapsidates the A‑form DNA. The left panel shows two
turns of the virion nucleoprotein superhelix exposing a
stretch of the viral A‑form DNA (in gold). The right panel
shows a surface representation of how three dimers of the
SIRV2 capsid protein bind to a stretch of the viral DNA,
making it inaccessible to the solvent. c | Comparison of the
homodimer of the SIRV2 capsid protein (left) with the
heterodimer formed from two paralogous capsid proteins
of AFV1 (right). The two monomers of SIRV2 capsid protein
are shown in magenta and green, whereas the two capsid
proteins of AFV1 are shown in red and yellow. d | Differences
in the packing of the nucleoprotein superhelix in SIRV2 (left)
and AFV1 (right). One homodimer and one heterodimer of
f the capsid proteins are shown for SIRV2 and AFV1,
GDGT-0 respectively. The proteins are coloured as in panel c. e | SIRV2
HO
(left) and AFV1 (right) virions viewed along the long virion
O O
axis. Capsid proteins are shown in green. The outer ring
O O surrounding the nucleoprotein core in AFV1 represents the
OH lipid envelope. f | The structure of the main lipid component
of the AFV1 envelope, glycerol dibiphytanyl glycerol
tetraether (GDGT) lipid (GDGT‑0; top). Schematic
representation (bottom left) and model (bottom right) of
GDGT‑0 in the horseshoe conformation. The hydrophilic
headgroups are shown in red. Part b (right-panel) is adapted
with permission from REF. 33, American Association for the
Advancement of Science. Part c is reproduced with
permission from REF. 69, eLife. Part d (right-panel) and part e
are adapted with permission from REF. 69, eLife.

Nature Reviews | Microbiology


NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 731
©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a 1

2
3

S-layer
b

4 1
UV 2
Cytoplasm
3
Appendages

5a
Cell membrane
VAP
4b
6a
4a
5b

6b

Figure 4 | Life cycles of lytic and temperate archaeal viruses. a | Steps of the interaction of the rudivirus Sulfolobus
islandicus rod-shaped virus 2 (SIRV2) with the host cell: step 1 and step 2: progression of theNature
virion (green) from the tip of a
Reviews | Microbiology
cellular appendage towards the cell surface; step 3: disassembly of the virion at the cell surface and delivery of the viral
genome into the cell interior; step 4: replication of the viral genome in a discrete focus in the cell cytoplasm; step 5a:
assembly of the virions, arranged into bundles containing ~50 viral particles; step 5b: concomitantly with virion assembly,
pyramidal structures form on the cell surface, perforating the S‑layer; step 6a: disintegration of the virion bundles into
separate, mature virions; and step 6b: opening of the pyramidal structures, allowing virion egress. b | Steps of the
induction of the temperate fusellovirus Sulfolobus spindle-shaped virus 1 (SSV1): step 1: UV irradiation leads to
reactivation of the SSV1 provirus (orange), which is integrated in the chromosome of its host Sulfolobus shibatae (black
circle); step 2: replication of the SSV1 genome; step 3: expression of viral proteins and formation of the viral nucleoprotein
(red); step 4a and step 4b: budding of the viral nucleoprotein through the cytoplasmic membrane containing viral capsid
proteins (green), resulting in formation of elongated virus-like particles; and step 5: maturation of the elongated particles
into characteristic spindle-shaped virions and their detachment from the cytoplasmic membrane. The scheme is
complemented with the tomographic surface-rendered volumes of SSV1 particles at different stages of budding.
Red, putative nucleoprotein. Scale bars, 200 nm. The 3D models in part a are reproduced with permission from REF. 103,
Proceedings of the National Academy of Sciences. The 3D models in part b are reproduced with permission from
American Society for Microbiology (REF. 71): Quemin, E. R. et al. Eukaryotic-like virus budding in Archaea. mBio 7,
Invertible region
e01439‑16 (2016) http://dx.doi.org/10.1128/mBio.01439-16.
A genome region that can
excise and reintegrate into the
same genome in inverted
orientation.
some or even most components of the DNA replication respective hosts on several independent occasions82.
Protein-primed DNA machinery, including DNA polymerases, proliferat- Viruses with small genomes (10–20 kb) tend to encode
polymerases ing cell nuclear antigen (PCNA; a DNA sliding clamp either p­ rotein-primed DNA polymerases74,83 or rolling-­circle
DNA polymerases capable of
that acts as a processivity factor for DNA polymerase), ­replication endo­nucleases (RCRE) 40,84, key proteins
the protein-primed initiation
step of DNA elongation. archaeo-eukaryotic primases and replicative helicases. for ssDNA replication in all three domains of life85,86.
Viruses with medium-sized genomes (20–50 kb) typ- It should be noted, however, that for many archaea-­
Rolling-circle replication ically encode only essential proteins that recruit the specific viruses, DNA replication proteins have not
The model of unidirectional host DNA replication machinery. For example, the been found in their genomes, implying that these viruses
DNA replication that can
rapidly synthesize multiple
replicative minichromosome-maintenance (MCM) rely primarily on the host DNA replication machinery
copies of circular ssDNA complex (a heli­case component of the DNA replication or employ heretofore uncharacterized mechanisms of
molecules. fork) has been acquired by different viruses from their genome replication. One such unique mechanism that

732 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

appears to rely on recombination for both initiation and virion of the fusellovirus SSV1 acquires its e­ nvelope
termination of genome replication was observed for the during budding of the nucleoprotein inner core
lipothrixvirus AFV1, although the genes that facilitate through the cytoplasmic membrane of the host cell71
this mechanism are currently unknown87. (FIG. 4b). Virions first emerge as elongated particles that
Most efforts in understanding genome replication are attached to the cell membrane and subsequently
during archaeal virus infection have been directed undergo maturation into characteristic spindle-shaped
towards the rudivirus SIRV2. Similar to some bacterial particles. Virus egress mechanisms strongly resemble
viruses, SIRV2 DNA synthesis is confined to a region the budding of enveloped eukaryotic viruses such as
near the periphery of the infected cells88. Although viral HIV, influenza virus and Ebola virus. Moreover, the
DNA synthesis is performed by one of the four paralo­ formation of a constricted ‘neck’ that facilitates viral
gous host DNA polymerases, namely, DNA polymer- budding is indistinguishable in appearance for SSV1
ase 1 (Dpo1), recruitment of the replication machinery and eukary­otic viruses. Many eukaryotic enveloped
appears to be orchestrated by viral proteins. One of viruses use components of the endosomal sorting com-
the early-expressed viral DNA-binding proteins, gp1 plexes required for transport (ESCRT) machinery for
(REF. 89), interacts with the host-encoded sliding clamp budding 94. The ESCRT pathway catalyses membrane
PCNA and presumably recruits it for the assembly of fission events that are required for the abscission stage
the replisome on the viral DNA90. Furthermore, SIRV2 of cytokinesis and for vesicle biogenesis95. The pres-
encodes several proteins that are thought to be involved ence of homologues of the eukaryotic ESCRT proteins
in DNA replication, recombination and repair, some of in the Sulfolobus spp., where they appear to perform
which have been experimentally characterized. These similar functions96, suggests that the budding of SSV1
proteins include a Holliday junction resolvase, a ssDNA-­ and enveloped ­eukaryotic viruses may rely on similar
binding protein, a ssDNA-annealing ATPase, a Cas4‑like cellular functions.
ssDNA nuclease and a Rep protein that is homologous The morphogenesis of Bicaudaviridae virions is also
to RCRE32. It was suggested that SIRV2 could employ likely to take place at the host cell surface and could
a combination of strand-displacement, rolling-circle and be coupled to ESCRT-dependent virus budding 97. The
strand-coupled genome replication mechanisms, which virion morphogenesis of the type species of the family,
generate multimeric and highly branched intermedi- ATV, is not completed at the cell surface: after being
ates reaching >1,200 kb in size (~34 genome units)91. released, the spindle-shaped virions develop long
However, whether all of these mechanisms occur dur- protrusions from the two pointed ends. High tem-
ing each cycle of viral DNA replication and the specific perature, close to that of the natural environment, is
proteins that are involved remain unclear. a prerequi­site for this morphological transformation,
Genome packaging has not been studied in detail which is accompanied by contraction of the ‘spindle’
for any archaeal virus. However, based on compara­ and the longitudinal association of globular subunits of
tive genomics, archaeal viruses of the Caudovirales unknown identity into helical, hollow tubes24. Virions
order employ virion assembly, maturation and genome of SMV1, one of the several putative family members,
packaging strategies similar to those that are used by can also develop tails in the extracellular environment 21.
bacterial viruses of the order. In particular, all archaeal However, for other bicaudaviruses, such as Sulfolobus
caudoviruses encode homologues of the large terminase tengchongensis spindle-shaped viruses 1 and 2 (STSV1
subunit, an enzyme that packages the viral DNA into and STSV2), no extracellular stage of virion morpho-
preassembled, empty capsids and subsequently processes genesis has been observed; once released from the cell,
the concatameric viral DNA into single-genome-length the virion morphology does not appear to change20.
units52. Similarly, members of the Turriviridae and By contrast, all stages of Rudiviridae and Turriviridae
Sphaerolipoviridae families encode A32‑like packaging virion morphogenesis seem to occur in the cytoplasm
ATPases of the FtsK–HerA superfamily, which is typical of the host cell. Mature virions exit the cells via a spe-
of dsDNA viruses with MCPs that have double or single cial gateway structure that has a seven-fold symmetry,
Holliday junction resolvase jelly-roll folds68. Moreover, the DNA packaging enzymes known as the virus-associated pyramid (VAP). The
A highly specialized struc- of two turriviruses have been shown to have ATPase VAP develops on the surface of infected cells and pro-
ture-selective endonuclease activity in vitro, and the proteins have been identified in trudes through the surface layer (S‑layer), and it opens as
that cleaves four-way DNA virions92,93, as is also the case for bacterial viruses with the ‘flower petals’ at the end of the infection cycle98,99 (FIG. 4a).
intermediates that can form
during DNA replication.
homologous capsid proteins68, emphasizing the evolu- The VAP consists of multiple copies of a 10 kDa viral
tionary connection between these bacterial and archaeal protein100–102. The protein has a transmembrane domain
Strand-displacement viruses. Thus, related viruses that infect hosts from dif- that facilitates its insertion into cellular membranes and
Of genome replication, ferent domains of life encode homologous proteins for can form pyramidal structures in all tested biological
involving the displacement of
virion formation and genome packaging, whereas pro- membranes from all three domains of life103. Pyramidal
a downstream DNA strand
encountered during DNA teins that are involved in genome replication either are structures, albeit of six-fold symmetry, have been
replication. unique or are recruited from the respective hosts. observed on the surface of uncultivated Pyrobaculum
spp. cells that are infected with an unidentified filamen-
Strand-coupled genome Virion morphogenesis and egress tous virus104 as well as on the surface of Pyrobaculum
replication
The model of DNA replication
The morphogenesis of pleomorphic virions of the oguniense105, suggesting that the VAP-based virus egress
that couples leading-strand Pleolipoviridae and of the spindle-shaped virions of mechanism is more common among crenarchaeal
and lagging-strand synthesis. the Fuselloviridae is coupled with virion egress. The viruses than p ­ reviously thought.

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 733


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

The enveloped Tristromaviridae virions also appear conserved genes in archaeal viral genomes encode pro-
to mature in the cytoplasm of the host cell. For exam- teins that are involved in virus–host interactions, for
ple, at the late stages of the viral life cycle, the host cells example, in counteracting host defence systems, such
are densely packed with the Pyrobaculum filamentous as CRISPR–Cas systems (see below).
virus 1 (PFV1) virions (>100 per cell), and their release Structural and comparative genomics studies
coincides with cell lysis38. The mechanism of intra­ clearly indicate that cosmopolitan archaeal viruses
cellular acquisition of the lipid envelope by the virions share common ancestry with specific groups of bac-
remains unclear. Release of the archaeal members of the terial viruses (Sphaerolipoviridae) or bacterial and
order Caudovirales also occurs by cell lysis. Notably, the eukary­otic viruses (Turriviridae and Caudovirales)28,67.
siphovirus ψM2, which infects Methanothermobacter By contrast, archaea-specific viruses have no counter-
marburgensis, encodes pseudomurein endoisopeptidase parts or even strong phylogenetic relationships with
(PeiP)51, which may be analogous to the endolysin com- bacteriophages or eukaryotic viruses, raising impor-
ponent of the holin–endolysin lysis systems that are tant questions regarding their origins. Recently, the
widely encoded by tailed bacteriophages106. Although evolutionary relationships between all dsDNA viruses
PeiP mediates lysis of M. marburgensis cells in vitro107, were examined using the bipartite network approach,
its participation in virion release or, alternatively, virus which traces connections between viral genomes
entry has not been investigated in vivo. through shared gene families113. This analysis grouped
Despite considerable gains in knowledge in recent all bacterial and eukaryotic viruses and the cosmo­
years, our understanding of virus–host interactions in politan archaeal viruses into four distinct supermodules,
archaea remains incomplete. In particular, the molecular whereas archaea-specific viruses remained largely dis-
mechanisms and, in many cases, the key genes that are connected from the global dsDNA virosphere67, despite
involved in different stages of archaeal viral life cycles occasional ‘gene sharing’ with the other modules. The
remain unknown. Future research is required to fill these subsequent detailed dissection of the archaea-specific
gaps in our understanding and to provide novel insights virus network revealed strong modularity, with six dis-
into virus–host co-­evolution and molecular details of tinct modules, each including one or two virus families,
the infection process under extreme e­ nvironmental and four additional virus families disconnected from
conditions. the rest of the archaeal virus network28. Compared to
simi­lar networks of eukaryotic and bacterial viruses, the
Genomics and evolutionary relationships archaea-specific viruses are sparsely connected within
At the onset of archaeal virus research, the information the network, with only a few shared gene families, most
that could be gained from archaeal virus genomes was notably the ribbon–helix–helix (RHH) DNA-binding
limited owing to the high number of uncharacterized proteins and glycosyltransferases, shared by different
and unique genes. The expansion of genomic data- network modules (FIG. 5). In these cases, the prevalence
bases enabled the power of comparative genomics to of the two gene families in viral genomes most likely
be applied to evolutionary studies of archaeal viruses. results from their independent acquisition from hosts
In this section, we outline the current understanding and/or horizontal gene transfer between viruses. The
of the evolutionary relationships between different lack of strong connectivity among the modules sug-
groups of archaeal viruses and with bacteriophages, gests that most of the archaea-specific virus groups are
eukaryotic viruses and non-viral mobile genetic evolutionarily distinct and have evolved independently
elements (MGEs). of one another. The order Ligamenvirales, which
The unique nature of viruses that infect hyper­ includes two fairly distant but clearly related families of
Pseudomurein thermo­philic Crenarchaeota is limited not only to archaea-specific viruses and forms one of the modules
endoisopeptidase their morphology but also to the content of their in the bipartite network, is the only notable exception
(PeiP). An enzyme that cleaves genomes, with ~90% of the genes lacking detect­ discovered so far.
pseudomurein cell-wall sacculi
able homologues (except those from other cren­ From whence did the archaea-specific viruses
of the methanogens.
archaeal viruses) in existing sequence databases28,108. origi­nate, and when did this happen? Why do they
Endolysin The genomes of some archaeal viruses do not have a only infect archaea? There are at least two non-­
A type of peptidoglycan- single gene with a functionally characterized homo- mutually exclusive explanations. Some of the archaea-­
hydrolysing enzyme produced logue108. As the tertiary structure of proteins is more specific virus groups could have emerged during the
by many bacterial viruses
towards the end of the
conserved than genetic sequences, structural genomics early stages of cellular evolution and been retained
lytic cycle. has been employed to uncover potential functions and in the Archaea but lost in the domains Bacteria and
the provenance of viral genes109. The crystal structures Eukarya114. Other archaeal virus groups could have
Structural genomics of many archaeal viral proteins have been resolved; evolved concomitantly with the Archaea or, even
The description of the
however, many of these proteins adopt unusual, novel more recently, within specific archaeal lineages. The
three‑dimensional structure
of every protein encoded by folds, providing limited insight into the function and observed limited gene sharing between different groups
a given genome. origin of these proteins110,111. Unexpectedly, in the case of archaea-specific viruses seems to make the latter pos-
of the fusello­virus SSV1, almost half of the viral genes, sibility particularly plausible. The large diversity of the
Supermodules many of which are functionally uncharacterized, are not archaea-specific viruses and the lack of evolutionary
Clusters of modules of tightly
connected genomes joined
essential for infectivity and could be knocked out with- connections between different groups sharply contrast
through higher-level shared out obvious phenotypic effects112. An intriguing possi- with the extensive exchange of genes among bacterio-
genes. bility is that many of the uncharacterized and poorly phages that also extends to the cosmopolitan archaeal

734 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Archaeal virosphere encode replication initiation endonucleases homolo-


Non-viral mobilome gous to the corresponding proteins encoded by small
plasmids that replicate through rolling-circle repli-
cation, such as the Archaeoglobus profundus plasmid
pGS5 (REF. 118) and the Thermococcus prieurii plas-
Ligamenvirales Clavaviridae mid pTP2 (REF. 119). In some archaeal viruses, for exam-
ple, spindle-shaped Pyrococcus abyssi virus 1 (PAV1)120,
nearly half of the viral genome shares homology with
Tristromaviridae plasmids121. Thus, the replication modules of many
archaea-specific viruses could be derived from non-­
viral MGEs. Combining such modules with the archi-
Ampullaviridae
tectural modules that provide viral structural genes
GTase Caudovirales could have resulted in the emergence of the new viral
lineages. The origins of the major virion proteins are
Bicaudaviridae
unknown for the majority of archaea-specific viruses.
RHH Turriviridae However, many of these proteins have fairly simple
DnaA
Fuselloviridae folds that are widespread in cellular proteomes62. For
example, the major virion proteins of rudiviruses and
bicaudaviruses have distinct helical-bundle folds33,122,
whereas those of spindle-shaped viruses presumably
Sphaerolipoviridae comprise two highly hydrophobic α-helixes16. Thus, it is
Guttaviridae possible that both of these proteins could evolve de novo
into virion components in the context of archaeal virus
Spiraviridae genomes. Moreover, it was recently found that one of
Globuloviridae
the major nucleocapsid proteins of Thermoproteus
Portogloboviridae
Pleolipoviridae tenax virus 1, a member of the family Tristromaviridae,
was exapted from a truncated and inactivated archaeal
Cas4‑like endonuclease123. The similar exaptation of cel-
lular proteins to function as major virion components
appears to be a recurrent theme in virus evolution62.
Figure 5 | Schematic representation of the gene-sharing network among different
Nature Reviews | Microbiology
families of archaeal viruses. Viral families that share genes are grouped in modules that Archaeal viruses and CRISPR–Cas
are highlighted in grey. Network modules were identified by applying a community
CRISPR–Cas systems provide an RNA-based mech-
detection algorithm to the bipartite network of gene sharing in the archaeal virosphere27.
These modules can be interpreted as sharing common evolutionary history. Families
anism to defend against invasive genetic elements
outside grey circles form separate, single-family modules. Viruses from the same module in prokaryotes. The archaeal hyperthermophiles are
share numerous genes (not shown in the figure), but gene sharing across modules is hosts to multiple groups of archaea-specific viruses
limited to three widespread genes, namely, glycosyltransferase of the GT‑B superfamily with unique morphotypes and gene complements
(GTase), ribbon–helix–helix-domain-containing protein (RHH) and an AAA+ superfamily (see above), and they also universally encompass
ATPase homologous to bacterial DnaA. Bidirectional arrows indicate the connections CRISPR–Cas systems, often multiple loci of different
between some families of archaeal viruses and capsidless (non-viral) mobile genetic types within the same genomes 124. The ubiquity of
elements that are discussed in the text. CRISPR–Cas systems in archaeal hyperthermophiles
contrasts with their relative rarity among bacteria,
<40% of which encode a CRISPR–Cas system, accord-
viruses28,67,113,115. The biological basis of this disjointed ing to recent estimates124,125, and implies that CRISPR–
organization of the archaea-specific virosphere remains Cas defence might have, to a large extent, shaped the
enigmatic but could be clarified by further studies of evolution of the viral genomes. The presence of many
archaeal virus–host interactions and a comprehensive genes that counteract host defences is a common fea-
characterization of archaeal virus diversity. ture of viral genomes, as exemplified by the numerous
Notably, several modules in the gene-sharing net- proteins that antagonize the immune system in animal
work of archaeal viruses also include non-viral, cap- poxviruses126 and herpesviruses127 or RNAi suppres-
sidless MGEs 28, namely, plasmids and casposons sors in plant RNA viruses128. Moreover, multiple anti-
(a recently discovered group of integrative MGEs that CRISPR proteins with specificity for different types and
appear to have given rise to the CRISPR–Cas adaptation subtypes of CRISPR–Cas systems have been identified
machinery responsible for the acquisition of new spa­ and studied in several bacteriophages129–131. Similar to
cers from the invading viruses and plasmids)116,117. In all these viruses, archaeal viruses are expected to have
these cases, viruses and non-viral MGEs share genes evolved anti-CRISPR counter defence strategies, but
that encode major genome replication proteins. In par- they have so far not been identified. As many of the
ticular, family 1 casposons encode protein-primed DNA proteins that are encoded by archaeal viruses are small,
polymerases closely related to those found in ampulla­ poorly conserved and lack identifiable domains with
viruses, the salterprovirus His1 and gamma­pleo­lipo­ known activities, similar to counter defence proteins
virus His2 (REF.  116), whereas alphapleolipoviruses in general and anti-CRISPR proteins in particular,

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 735


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Mesophiles there are many candidates to screen. The search for Conclusions and outlook
Organisms that grow best anti-CRISPR proteins in archaeal viruses is undoubt- Viruses of the archaea have remarkable virion architec-
in moderate temperature, edly a promising research direction for understand- tures, genome structures, genes and modes of interaction
typically between 20 and ing CRISPR–Cas immunity and counter defences with their hosts. Recent structural and functional studies
45 °C.
in archaea. have further expanded the repertoire of unique structures
CRISPR spacers It remains unclear why CRISPR–Cas systems are and mechanisms found in this part of the virosphere
Short fragments of viral DNA ubiquitous in archaeal hyperthermophiles. One poten- through unexpected discoveries, such as the A‑form DNA
from previous exposure to tial explanation is that the comparatively low mutation in virions, a new type of membrane in enveloped viruses
the virus, inserted between
rate of viruses that propagate in extreme environments, and the VAPs. Understanding the biological and evolu-
repetitive sequences of the
CRISPR–Cas system. combined with the lower effective population sizes of tionary basis of these unusual features is a major challenge
hyperthermophiles compared to mesophiles, leads to for future research that will require the development of
Protospacers fairly low virus diversity, which would maximize the new model systems for studying virus–host interactions
Fragments of invading mobile benefits of immune memory; that is, the emergence in archaea. Concomitant metagenomic studies have led
genetic element from which
CRISPR spacers are derived.
of virus escape mutants is expected to be slower 132,133. to the discovery of new groups of archaeal viruses. Some
Simply put, owing to the limited opportunity for virus of these, such as the magroviruses, are highly abundant in
Primed adaptation escape, the CRISPR–Cas systems of hyperthermophiles the environment but remained unknown for a long period
A process in which an existing appear to be more efficient in ‘remembering’ past of time because their hosts have not been cultivated.
spacer against a foreign DNA
infections and controlling the respective viruses than Moreover, concurrent environmental and geochemical
promotes rapid and efficient
acquisition of additional those of mesophiles. A recent comprehensive survey of research has revealed that archaeal viruses are important
spacers from the same the CRISPR spacers in bacterial and archaeal genomes players in the biogeochemistry of the biosphere.
foreign DNA. identi­fied a low proportion of protospacers with per- Despite substantial progress in the field of archaeal
fect complementarity in both hyperthermophilic and virology, a comprehensive understanding of the global
mesophilic archaea (~1% in hyperthermophilic archaea diversity of archaeal viruses is still lacking. Indeed, the
compared to the mean value of ~7% among all bacte- viruses that have been characterized thus far infect only a
rial and archaeal phyla). As expected, the majority of limited number of host taxa and have been isolated from
the protospacers mapped to archaeal virus genomes134. specific environments. Future research should focus on
The low proportion of spacers with a detectable proto- the isolation of new archaeal virus–host systems from a
spacer in the viral or plasmid genome could be due to broader range of ecosystems. Of particular interest are
the ability of different CRISPR–Cas systems, particu- viruses that infect currently uncultured archaea, many
larly type I‑A and type III systems that are common in of which are abundant in the environment and have
archaea124, to utilize spacers with multiple mismatches major roles in the biosphere. It is likely that advances
for inter­ference and primed adaptation135–138. However, in this direction will be achieved by combining culture-­
the existence of uncharacterized, local viromes could dependent and culture-independent high-throughput
be a complementary explanation for the scarcity of approaches, such as metagenomics, metaproteomics
spacers that could be matched to known archaeal virus and single-cell genomics. This research is expected to
genomes. In agreement with this hypothesis, it has been provide an unbiased view of the distribution, abundance
shown that CRISPR arrays in archaeal communities and diversity of archaeal viruses, allowing more-­accurate
are enriched in spacers that better align to local viral comparisons between bacterial and archaeal viruses
genomes than foreign viral genomes139,140. and informing construction of evolutionary scenarios.
Given the high prevalence of CRISPR–Cas systems in Another promising line of inquiry, boosted by the devel-
archaea, especially hyperthermophiles, future detailed opment of diverse tools of molecular biology, genetics
studies of the co-evolution of archaeal viruses with their and microscopy, should focus on deciphering the molec-
hosts and the strategies that are exploited by archaeal ular mechanisms that underlie virus–host interactions.
viruses to overcome CRISPR immunity could provide This research should yield crucial insights into the evo-
valuable insights into the mechanisms and evolution of lutionary history of viruses, their adaptation to extreme
prokaryotic adaptive immunity. ­environments and their co‑evolution with their hosts.

1. Woese, C. R. & Fox, G. E. Phylogenetic structure of the 7. Lloyd, K. G. et al. Predominant archaea in marine and isolation of the Aeropyrum pernix bacilliform
prokaryotic domain: the primary kingdoms. Proc. Natl sediments degrade detrital proteins. Nature 496, virus 1, APBV1, the first representative of the
Acad. Sci. USA 74, 5088–5090 (1977). 215–218 (2013). family Clavaviridae. Virology 402, 347–354
2. Hug, L. A. et al. A new view of the tree of life. 8. Takano, Y. et al. Sedimentary membrane lipids (2010).
Nat. Microbiol. 1, 16048 (2016). recycled by deep-sea benthic archaea. Nat. Geosci. 3, 12. Sencilo, A. et al. Snapshot of haloarchaeal tailed
3. Lurie-Weinberger, M. N. & Gophna, U. Archaea in and 858–861 (2010). virus genomes. RNA Biol. 10, 803–816 (2013).
on the human body: health implications and future 9. Roux, S. et al. Ecogenomics and potential 13. Forterre, P. The common ancestor of archaea and
directions. PLoS Pathog. 11, e1004833 (2015). biogeochemical impacts of globally abundant ocean eukarya was not an archaeon. Archaea 2013,
4. Lloyd, K. G., May, M. K., Kevorkian, R. T. & Steen, A. D. viruses. Nature 537, 689–693 (2016). 372396 (2013).
Meta-analysis of quantification methods shows that This study suggests a global impact of archaeal 14. Häring, M., Rachel, R., Peng, X., Garrett, R. A. &
archaea and bacteria have similar abundances in the viruses in the epipelagic and mesopelagic ocean. Prangishvili, D. Viral diversity in hot springs of
subseafloor. Appl. Environ. Microbiol. 79, 7790–7799 10. Danovaro, R. et al. Virus-mediated archaeal Pozzuoli, Italy, and characterization of a unique
(2013). hecatomb in the deep seafloor. Sci. Adv. 2, archaeal virus. Acidianus bottle-shaped virus, from
5. Karner, M. B., DeLong, E. F. & Karl, D. M. Archaeal e1600492 (2016). a new family, the Ampullaviridae. J. Virol. 79,
dominance in the mesopelagic zone of the Pacific This study shows that archaeal viruses have a 9904–9911 (2005).
Ocean. Nature 409, 507–510 (2001). profound role in the functioning of deep-sea 15. Mochizuki, T. et al. Archaeal virus with exceptional
6. Offre, P., Spang, A. & Schleper, C. Archaea in ecosystems and in global biogeochemical cycles. virion architecture and the largest single-stranded
biogeochemical cycles. Annu. Rev. Microbiol. 67, 11. Mochizuki, T. et al. Diversity of viruses of the DNA genome. Proc. Natl Acad. Sci. USA 109,
437–457 (2013). hyperthermophilic archaeal genus Aeropyrum, 13386–13391 (2012).

736 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

16. Krupovic, M., Quemin, E. R., Bamford, D. H., 38. Rensen, E. I. et al. A virus of hyperthermophilic principles for virion assembly. Curr. Opin. Virol. 1,
Forterre, P. & Prangishvili, D. Unification of the archaea with a unique architecture among DNA 118–124 (2011).
globally distributed spindle-shaped viruses of the viruses. Proc. Natl Acad. Sci. USA 113, 2478–2483 59. Suhanovsky, M. M. & Teschke, C. M. Nature’s favorite
Archaea. J. Virol. 88, 2354–2358 (2014). (2016). building block: deciphering folding and capsid
17. Redder, P. et al. Four newly isolated fuselloviruses 39. Liu, Y. et al. A novel type of polyhedral viruses assembly of proteins with the HK97‑fold. Virology
from extreme geothermal environments reveal infecting hyperthermophilic archaea. J. Virol. 91, 479–480, 487–497 (2015).
unusual morphologies and a possible interviral e00589–e00517 (2017). 60. Baker, M. L., Jiang, W., Rixon, F. J. & Chiu, W.
recombination mechanism. Environ. Microbiol. 11, 40. Pietilä, M. K., Roine, E., Sencilo, A., Bamford, D. H. Common ancestry of herpesviruses and tailed DNA
2849–2862 (2009). & Oksanen, H. M. Pleolipoviridae, a newly proposed bacteriophages. J. Virol. 79, 14967–14970 (2005).
18. Häring, M. et al. Virology: independent virus family comprising archaeal pleomorphic viruses with 61. Yu, X., Jih, J., Jiang, J. & Zhou, Z. H. Atomic structure
development outside a host. Nature 436, 1101–1102 single-stranded or double-stranded DNA genomes. of the human cytomegalovirus capsid with its securing
(2005). Arch. Virol. 161, 249–256 (2016). tegument layer of pp150. Science 356, eaam6892
19. Hochstein, R. A., Amenabar, M. J., Munson- 41. Pietilä, M. K. et al. Virion architecture unifies globally (2017).
McGee, J. H., Boyd, E. S. & Young, M. J. Acidianus distributed pleolipoviruses infecting halophilic 62. Krupovic, M. & Koonin, E. V. Multiple origins of viral
tailed spindle virus: a new archaeal large tailed spindle archaea. J. Virol. 86, 5067–5079 (2012). capsid proteins from cellular ancestors. Proc. Natl
virus discovered by culture-independent methods. 42. Pietilä, M. K., Laurinavicius, S., Sund, J., Roine, E. & Acad. Sci. USA 114, E2401–E2410 (2017).
J. Virol. 90, 3458–3468 (2016). Bamford, D. H. The single-stranded DNA genome of 63. Krupovic, M. & Bamford, D. H. Virus evolution: how
A new spindle-shaped virus and its host are novel archaeal virus Halorubrum pleomorphic virus far does the double beta-barrel viral lineage extend?
characterized using culture-independent techniques. 1 is enclosed in the envelope decorated with Nat. Rev. Microbiol. 6, 941–948 (2008).
20. Xiang, X. et al. Sulfolobus tengchongensis spindle- glycoprotein spikes. J. Virol. 84, 788–798 (2010). 64. Sinclair, R. M., Ravantti, J. J. & Bamford, D. H.
shaped virus STSV1: virus-host interactions and 43. Roine, E. et al. New, closely related haloarchaeal viral Nucleic and amino acid sequences support structure-
genomic features. J. Virol. 79, 8677–8686 (2005). elements with different nucleic acid types. J. Virol. 84, based viral classification. J. Virol. 91, e02275‑16
21. Uldahl, K. B. et al. Life cycle characterization of 3682–3689 (2010). (2017).
Sulfolobus monocaudavirus 1, an extremophilic 44. Pietilä, M. K., Roine, E., Paulin, L., Kalkkinen, N. & 65. Jäälinoja, H. T. et al. Structure and host-cell
spindle-shaped virus with extracellular tail Bamford, D. H. An ssDNA virus infecting archaea: interaction of SH1, a membrane-containing, halophilic
development. J. Virol. 90, 5693–5699 (2016). a new lineage of viruses with a membrane envelope. euryarchaeal virus. Proc. Natl Acad. Sci. USA 105,
22. Erdmann, S. et al. A novel single-tailed fusiform Mol. Microbiol. 72, 307–319 (2009). 8008–8013 (2008).
Sulfolobus virus STSV2 infecting model Sulfolobus 45. Sencilo, A., Paulin, L., Kellner, S., Helm, M. & Roine, E. 66. Gil-Carton, D. et al. Insight into the assembly of
species. Extremophiles 18, 51–60 (2014). Related haloarchaeal pleomorphic viruses contain viruses with vertical single β-barrel major capsid
23. Erdmann, S., Le Moine Bauer, S. & Garrett, R. A. different genome types. Nucleic Acids Res. 40, proteins. Structure 23, 1866–1877 (2015).
Inter-viral conflicts that exploit host CRISPR immune 5523–5534 (2012). 67. Iranzo, J., Krupovic, M. & Koonin, E. V. The double-
systems of Sulfolobus. Mol. Microbiol. 91, 900–917 46. Ackermann, H. W. & Prangishvili, D. Prokaryote stranded DNA virosphere as a modular hierarchical
(2014). viruses studied by electron microscopy. Arch. Virol. network of gene sharing. mBio 7, e00978‑16 (2016).
24. Prangishvili, D. The wonderful world of archaeal 157, 1843–1849 (2012). 68. Strömsten, N. J., Bamford, D. H. & Bamford, J. K.
viruses. Annu. Rev. Microbiol. 67, 565–585 (2013). 47. Pietilä, M. K., Demina, T. A., Atanasova, N. S., In vitro DNA packaging of PRD1: a common
25. Prangishvili, D. et al. Structural and genomic Oksanen, H. M. & Bamford, D. H. Archaeal viruses mechanism for internal-membrane viruses. J. Mol. Biol.
properties of the hyperthermophilic archaeal virus ATV and bacteriophages: comparisons and contrasts. 348, 617–629 (2005).
with an extracellular stage of the reproductive cycle. Trends Microbiol. 22, 334–344 (2014). 69. Kasson, P. et al. Model for a novel membrane
J. Mol. Biol. 359, 1203–1216 (2006). 48. Krupovic, M., Prangishvili, D., Hendrix, R. W. & envelope in a filamentous hyperthermophilic virus.
26. Arnold, H. P., Ziese, U. & Zillig, W. SNDV, a novel Bamford, D. H. Genomics of bacterial and archaeal eLife 6, e26268 (2017).
virus of the extremely thermophilic and acidophilic viruses: dynamics within the prokaryotic virosphere. Near-atomic-resolution structure of an enveloped
archaeon Sulfolobus. Virology 272, 409–416 (2000). Microbiol. Mol. Biol. Rev. 75, 610–635 (2011). virion of archaeal virus AFV1 reveals novel
27. Mochizuki, T., Sako, Y. & Prangishvili, D. Provirus 49. Pietilä, M. K. et al. Insights into head-tailed viruses membrane organization not previously observed
induction in hyperthermophilic archaea: infecting extremely halophilic archaea. J. Virol. 87, in viruses or cellular organisms.
characterization of Aeropyrum pernix spindle-shaped 3248–3260 (2013). 70. Valentine, D. L. Adaptations to energy stress dictate
virus 1 and Aeropyrum pernix ovoid virus 1. 50. Pietilä, M. K. et al. Structure of the archaeal head- the ecology and evolution of the Archaea. Nat. Rev.
J. Bacteriol. 193, 5412–5419 (2011). tailed virus HSTV‑1 completes the HK97 fold story. Microbiol. 5, 316–323 (2007).
28. Iranzo, J., Koonin, E. V., Prangishvili, D. & Proc. Natl Acad. Sci. USA 110, 10604–10609 71. Quemin, E. R. et al. Eukaryotic-like virus budding in
Krupovic, M. Bipartite network analysis of the (2013). Archaea. mBio 7, e01439‑16 (2016).
archaeal virosphere: evolutionary connections Cryo‑EM reconstruction of the archaeal podovirus This electron-tomography study shows that
between viruses and capsidless mobile elements. HSTV‑1 capsid reveals a major capsid protein fold enveloped, spindle-shaped viruses of archaea are
J. Virol. 90, 11043–11055 (2016). found in tailed bacteriophages and eukaryotic released from the cell by a budding mechanism
Comprehensive bipartite network analysis of all herpesviruses. highly reminiscent of that used by many enveloped
known archaeal virus genomes reveals evolutionary 51. Pfister, P., Wasserfallen, A., Stettler, R. & Leisinger, T. eukaryotic viruses, such as HIV and influenza virus.
connections between different groups of viruses as Molecular analysis of Methanobacterium phage ψM2. 72. Quemin, E. R. et al. Sulfolobus spindle-shaped virus 1
well as non-viral mobile genetic elements. Mol. Microbiol. 30, 233–244 (1998). contains glycosylated capsid proteins, a cellular
29. Prangishvili, D. et al. A novel virus family, the 52. Krupovic, M., Forterre, P. & Bamford, D. H. chromatin protein, and host-derived lipids. J. Virol.
Rudiviridae: Structure, virus-host interactions and Comparative analysis of the mosaic genomes of tailed 89, 11681–11691 (2015).
genome variability of the sulfolobus viruses SIRV1 archaeal viruses and proviruses suggests common 73. Kristensen, D. M., Saeed, U., Frishman, D. &
and SIRV2. Genetics 152, 1387–1396 (1999). themes for virion architecture and assembly with Koonin, E. V. A census of α-helical membrane proteins
30. Vestergaard, G. et al. A novel rudivirus, ARV1, of the tailed viruses of bacteria. J. Mol. Biol. 397, 144–160 in double-stranded DNA viruses infecting bacteria and
hyperthermophilic archaeal genus Acidianus. Virology (2010). archaea. BMC Bioinformatics 16, 380 (2015).
336, 83–92 (2005). 53. Krupovic, M., Spang, A., Gribaldo, S., Forterre, P. & 74. Bath, C., Cukalac, T., Porter, K. & Dyall-Smith, M. L.
31. Vestergaard, G. et al. Stygiolobus rod-shaped virus Schleper, C. A thaumarchaeal provirus testifies for an His1 and His2 are distantly related, spindle-shaped
and the interplay of crenarchaeal rudiviruses with ancient association of tailed viruses with archaea. haloviruses belonging to the novel virus group,
the CRISPR antiviral system. J. Bacteriol. 190, Biochem. Soc. Trans. 39, 82–88 (2011). Salterprovirus. Virology 350, 228–239 (2006).
6837–6845 (2008). 54. Pawlowski, A., Rissanen, I., Bamford, J. K., 75. Martin, A. et al. SAV 1, a temperate u.v.-inducible
32. Prangishvili, D., Koonin, E. V. & Krupovic, M. Genomics Krupovic, M. & Jalasvuori, M. DNA virus-like particle from the archaebacterium
and biology of Rudiviruses, a model for the study of Gammasphaerolipovirus, a newly proposed Sulfolobus acidocaldarius isolate B12. EMBO J. 3,
virus-host interactions in Archaea. Biochem. Soc. Trans. bacteriophage genus, unifies viruses of halophilic 2165–2168 (1984).
41, 443–450 (2013). archaea and thermophilic bacteria within the novel 76. Hong, C. et al. Lemon-shaped halo archaeal virus His1
33. DiMaio, F. et al. A virus that infects a family Sphaerolipoviridae. Arch. Virol. 159, with uniform tail but variable capsid structure.
hyperthermophile encapsidates A‑form DNA. Science 1541–1554 (2014). Proc. Natl Acad. Sci. USA 112, 2449–2454 (2015).
348, 914–917 (2015). 55. Veesler, D. et al. Atomic structure of the 75 MDa 77. Stedman, K. M., DeYoung, M., Saha, M.,
Near-atomic-resolution structure of the rod-shaped extremophile Sulfolobus turreted icosahedral virus Sherman, M. B. & Morais, M. C. Structural insights
archaeal virus reveals the A‑form of DNA in a determined by CryoEM and X‑ray crystallography. into the architecture of the hyperthermophilic
biological entity and provides clues about the Proc. Natl Acad. Sci. USA 110, 5504–5509 (2013). Fusellovirus SSV1. Virology 474, 105–109 (2015).
thermostability of the viral particle. Near-atomic-resolution structure of the turrivirus This work and the one in REF 76 present cryo‑EM
34. Goulet, A. et al. Acidianus filamentous virus 1 coat STIV reveals fine details of the virion organization reconstructions of spindle-shaped archaeal viruses
proteins display a helical fold spanning the filamentous and strengthens the evolutionary connection that are hyperthermophilic and hyperhalophilic,
archaeal viruses lineage. Proc. Natl Acad. Sci. USA between STIV and bacterial and eukaryotic viruses respectively.
106, 21155–21160 (2009). with double jelly-roll capsid proteins. 78. Bettstetter, M., Peng, X., Garrett, R. A. &
35. Szymczyna, B. R. et al. Synergy of NMR, computation, 56. Demina, T. A. et al. HCIV‑1 and other tailless Prangishvili, D. AFV1, a novel virus infecting
and X‑ray crystallography for structural biology. icosahedral internal membrane-containing viruses of hyperthermophilic archaea of the genus acidianus.
Structure 17, 499–507 (2009). the family Sphaerolipoviridae. Viruses 9, 32 (2017). Virology 315, 68–79 (2003).
36. Pina, M., Bize, A., Forterre, P. & Prangishvili, D. 57. Khayat, R. et al. Structure of an archaeal virus capsid 79. Quemin, E. R. et al. First insights into the entry
The archeoviruses. FEMS Microbiol. Rev. 35, protein reveals a common ancestry to eukaryotic and process of hyperthermophilic archaeal viruses. J. Virol.
1035–1054 (2011). bacterial viruses. Proc. Natl Acad. Sci. USA 102, 87, 13379–13385 (2013).
37. Prangishvili, D. & Krupovic, M. A new proposed 18944–18949 (2005). 80. Deng, L. et al. Unveiling cell surface and type IV
taxon for double-stranded DNA viruses, the order 58. Krupovic, M. & Bamford, D. H. Double-stranded DNA secretion proteins responsible for archaeal rudivirus
“Ligamenvirales”. Arch. Virol. 157, 791–795 (2012). viruses: 20 families and only five different architectural entry. J. Virol. 88, 10264–10268 (2014).

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 737


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

81. Klein, R., Rossler, N., Iro, M., Scholz, H. & Witte, A. 103. Daum, B. et al. Self-assembly of the general 126. Bidgood, S. R. & Mercer, J. Cloak and dagger:
Haloarchaeal myovirus φCh1 harbours a phase membrane-remodeling protein PVAP into sevenfold alternative immune evasion and modulation strategies
variation system for the production of protein variants virus-associated pyramids. Proc. Natl Acad. Sci. USA of poxviruses. Viruses 7, 4800–4825 (2015).
with distinct cell surface adhesion specificities. 111, 3829–3834 (2014). 127. Ressing, M. E. et al. Immune evasion by Epstein-Barr
Mol. Microbiol. 83, 137–150 (2012). This study presents a three-dimensional virus. Curr. Top. Microbiol. Immunol. 391, 355–381
82. Krupovic, M., Gribaldo, S., Bamford, D. H. & reconstruction of the virus-associated pyramids (2015).
Forterre, P. The evolutionary history of archaeal MCM employed by rod-shaped virus SIRV2 during the 128. Zhao, J. H., Hua, C. L., Fang, Y. Y. & Guo, H. S. The
helicases: a case study of vertical evolution combined egress. dual edge of RNA silencing suppressors in the virus-
with hitchhiking of mobile genetic elements. Mol. Biol. 104. Bize, A. et al. Viruses in acidic geothermal environments host interactions. Curr. Opin. Virol. 17, 39–44 (2016).
Evol. 27, 2716–2732 (2010). of the Kamchatka Peninsula. Res. Microbiol. 159, 129. Chowdhury, S. et al. Structure reveals mechanisms of
83. Peng, X., Basta, T., Haring, M., Garrett, R. A. & 358–366 (2008). viral suppressors that intercept a CRISPR RNA-guided
Prangishvili, D. Genome of the Acidianus bottle- 105. Rensen, E., Krupovic, M. & Prangishvili, D. Mysterious surveillance complex. Cell 169, 47–57.e11 (2017).
shaped virus and insights into the replication and hexagonal pyramids on the surface of Pyrobaculum 130. Pawluk, A. et al. Inactivation of CRISPR-Cas systems
packaging mechanisms. Virology 364, 237–243 cells. Biochimie 118, 365–367 (2015). by anti-CRISPR proteins in diverse bacterial species.
(2007). 106. Young, R. Phage lysis: three steps, three choices, one Nat. Microbiol. 1, 16085 (2016).
84. Wang, Y. et al. Identification, characterization, and outcome. J. Microbiol. 52, 243–258 (2014). 131. Sontheimer, E. J. & Davidson, A. R. Inhibition of
application of the replicon region of the halophilic 107. Luo, Y., Pfister, P., Leisinger, T. & Wasserfallen, A. CRISPR-Cas systems by mobile genetic elements.
temperate sphaerolipovirus SNJ1. J. Bacteriol. 198, Pseudomurein endoisopeptidases PeiW and PeiP, Curr. Opin. Microbiol. 37, 120–127 (2017).
1952–1964 (2016). two moderately related members of a novel family of 132. Iranzo, J., Lobkovsky, A. E., Wolf, Y. I. & Koonin, E. V.
85. Krupovic, M. Networks of evolutionary interactions proteases produced in Methanothermobacter strains. Evolutionary dynamics of the prokaryotic adaptive
underlying the polyphyletic origin of ssDNA viruses. FEMS Microbiol. Lett. 208, 47–51 (2002). immunity system CRISPR-Cas in an explicit ecological
Curr. Opin. Virol. 3, 578–586 (2013). 108. Prangishvili, D., Garrett, R. A. & Koonin, E. V. context. J. Bacteriol. 195, 3834–3844 (2013).
86. Chandler, M. et al. Breaking and joining single- Evolutionary genomics of archaeal viruses: unique 133. Weinberger, A. D., Wolf, Y. I., Lobkovsky, A. E.,
stranded DNA: the HUH endonuclease superfamily. viral genomes in the third domain of life. Virus Res. Gilmore, M. S. & Koonin, E. V. Viral diversity threshold
Nat. Rev. Microbiol. 11, 525–538 (2013). 117, 52–67 (2006). for adaptive immunity in prokaryotes. mBio 3,
87. Pina, M. et al. Unique genome replication mechanism 109. Greene, L. H. et al. The CATH domain structure e00456‑12 (2012).
of the archaeal virus AFV1. Mol. Microbiol. 92, database: new protocols and classification levels give a 134. Shmakov, S. A. et al. The CRISPR spacer space is
1313–1325 (2014). more comprehensive resource for exploring evolution. dominated by sequences from the species-specific
88. Martinez-Alvarez, L., Deng, L. & Peng, X. Formation Nucleic Acids Res. 35, D291–D297 (2007). mobilome. mBio 8, e01397‑17 (2017).
of a viral replication focus in Sulfolobus cells infected 110. Krupovic, M., White, M. F., Forterre, P. & 135. Fineran, P. C. et al. Degenerate target sites mediate
by the rudivirus Sulfolobus islandicus rod-shaped Prangishvili, D. Postcards from the edge: structural rapid primed CRISPR adaptation. Proc. Natl Acad. Sci.
virus 2. J. Virol. 91, e00486‑17 (2017). genomics of archaeal viruses. Adv. Virus Res. 82, USA 111, E1629–E1638 (2014).
This study shows that genome replication of the 33–62 (2012). 136. Manica, A., Zebec, Z., Steinkellner, J. & Schleper, C.
rod-shaped virus SIRV2 is confined to discrete foci 111. Dellas, N., Lawrence, C. M. & Young, M. J. A survey Unexpectedly broad target recognition of the CRISPR-
in the cytoplasm, where the host DNA polymerase of protein structures from archaeal viruses. Life 3, mediated virus defence system in the archaeon
and viral replication factors are recruited. 118–130 (2013). Sulfolobus solfataricus. Nucleic Acids Res. 41,
89. Peeters, E. et al. DNA-interacting characteristics of 112. Iverson, E. A., Goodman, D. A., Gorchels, M. E. & 10509–10517 (2013).
the archaeal rudiviral protein SIRV2_Gp1. Viruses 9, Stedman, K. M. Extreme mutation tolerance: nearly 137. Maniv, I., Jiang, W., Bikard, D. & Marraffini, L. A.
190 (2017). half of the archaeal fusellovirus Sulfolobus spindle- Impact of different target sequences on type III
90. Gardner, A. F., Bell, S. D., White, M. F., shaped virus 1 genes are not required for virus CRISPR-Cas immunity. J. Bacteriol. 198, 941–950
Prangishvili, D. & Krupovic, M. Protein-protein function, including the minor capsid protein gene, vp3. (2016).
interactions leading to recruitment of the host DNA J. Virol. 91, e02406‑16 (2017). 138. Mousaei, M., Deng, L., She, Q. & Garrett, R. A. Major
sliding clamp by the hyperthermophilic Sulfolobus 113. Iranzo, J., Krupovic, M. & Koonin, E. V. A network and minor crRNA annealing sites facilitate low
islandicus rod-shaped virus 2. J. Virol. 88, perspective on the virus world. Commun. Integr. Biol. stringency DNA protospacer binding prior to type I‑A
7105–7108 (2014). 10, e1296614 (2017). CRISPR-Cas interference in Sulfolobus. RNA Biol. 13,
91. Martinez-Alvarez, L., Bell, S. D. & Peng, X. Multiple 114. Prangishvili, D. Archaeal viruses: living fossils of the 1166–1173 (2016).
consecutive initiation of replication producing novel ancient virosphere? Ann. NY Acad. Sci. 1341, 35–40 139. Held, N. L. & Whitaker, R. J. Viral biogeography
brush-like intermediates at the termini of linear viral (2015). revealed by signatures in Sulfolobus islandicus
dsDNA genomes with hairpin ends. Nucleic Acids Res. 115. Hendrix, R. W. Bacteriophage genomics. Curr. Opin. genomes. Environ. Microbiol. 11, 457–466 (2009).
44, 8799–8809 (2016). Microbiol. 6, 506–511 (2003). 140. Emerson, J. B. et al. Virus-host and CRISPR dynamics
92. Dellas, N. et al. Structure-based mutagenesis of 116. Krupovic, M., Makarova, K. S., Forterre, P., in Archaea-dominated hypersaline Lake Tyrrell,
Sulfolobus turreted icosahedral virus B204 reveals Prangishvili, D. & Koonin, E. V. Casposons: a new Victoria, Australia. Archaea 2013, 370871 (2013).
essential residues in the virion-associated DNA- superfamily of self-synthesizing DNA transposons at 141. Gudbergsdóttir, S. R., Menzel, P., Krogh, A., Young, M.
packaging ATPase. J. Virol. 90, 2729–2739 (2015). the origin of prokaryotic CRISPR-Cas immunity. & Peng, X. Novel viral genomes identified from six
93. Happonen, L. J. et al. The structure of the NTPase BMC Biol. 12, 36 (2014). metagenomes reveal wide distribution of archaeal
that powers DNA packaging into Sulfolobus turreted 117. Krupovic, M., Béguin, P. & Koonin, E. V. Casposons: viruses and high viral diversity in terrestrial hot
icosahedral virus 2. J. Virol. 87, 8388–8398 (2013). mobile genetic elements that gave rise to the CRISPR- springs. Environ. Microbiol. 18, 863–874 (2016).
94. Hurley, J. H. & Hanson, P. I. Membrane budding and Cas adaptation machinery. Curr. Opin. Microbiol. 38, The study describes several novel
scission by the ESCRT machinery: it’s all in the neck. 36–43 (2017). hyperthermophilic archaeal virus genomes
Nat. Rev. Mol. Cell Biol. 11, 556–566 (2010). 118. Lopez-Garcia, P., Forterre, P., van der Oost, J. & assembled from metagenomic data.
95. Scourfield, E. J. & Martin-Serrano, J. Growing Erauso, G. Plasmid pGS5 from the hyperthermophilic 142. Adriaenssens, E. M., van Zyl, L. J., Cowan, D. A. &
functions of the ESCRT machinery in cell biology and archaeon Archaeoglobus profundus is negatively Trindade, M. I. Metaviromics of Namib Desert salt
viral replication. Biochem. Soc. Trans. 45, 613–634 supercoiled. J. Bacteriol. 182, 4998–5000 (2000). pans: a novel lineage of haloarchaeal salterproviruses
(2017). 119. Gorlas, A., Krupovic, M., Forterre, P. & Geslin, C. Living and a rich source of ssDNA viruses. Viruses 8, 14
96. Makarova, K. S., Yutin, N., Bell, S. D. & Koonin, E. V. side by side with a virus: characterization of two novel (2016).
Evolution of diverse cell division and vesicle formation plasmids from Thermococcus prieurii, a host for the 143. Nigro, O. D. et al. Viruses in the oceanic basement.
systems in Archaea. Nat. Rev. Microbiol. 8, 731–741 spindle-shaped virus TPV1. Appl. Environ. Microbiol. mBio 8, e02129‑16 (2017).
(2010). 79, 3822–3828 (2013). 144. Nishimura, Y. et al. Environmental viral genomes shed
97. Liu, J. et al. Functional assignment of multiple ESCRT- 120. Geslin, C. et al. Analysis of the first genome of a new light on virus-host interactions in the ocean.
III homologs in cell division and budding in Sulfolobus hyperthermophilic marine virus-like particle, PAV1, mSphere 2, e00359‑16 (2017).
islandicus. Mol. Microbiol. 105, 540–553 (2017). isolated from Pyrococcus abyssi. J. Bacteriol. 189, 145. Philosof, A. et al. Novel abundant oceanic viruses of
98. Bize, A. et al. A unique virus release mechanism in 4510–4519 (2007). uncultured marine group II Euryarchaeota identified
the Archaea. Proc. Natl Acad. Sci. USA 106, 121. Krupovic, M., Gonnet, M., Hania, W. B., Forterre, P. & by genome-centric metagenomics. Curr. Biol. 27,
11306–11311 (2009). Erauso, G. Insights into dynamics of mobile genetic 1362–1368 (2017).
99. Brumfield, S. K. et al. Particle assembly and elements in hyperthermophilic environments from five This study, along with REF 144, describes the
ultrastructural features associated with replication new Thermococcus plasmids. PLoS ONE 8, e49044 identification of a novel, diverse group of archaeal
of the lytic archaeal virus Sulfolobus turreted (2013). viruses, called Magroviruses, associated with
icosahedral virus. J. Virol. 83, 5964–5970 (2009). 122. Goulet, A. et al. Getting the best out of long- environmentally abundant, uncultured Marine
100. Quax, T. E. F., Krupovic, M., Lucas, S., Forterre, P. & wavelength X‑rays: de novo chlorine/sulfur SAD Group II Euryarchaeota.
Prangishvili, D. The Sulfolobus rod-shaped virus 2 phasing of a structural protein from ATV. Acta 146. Servin-Garciduenas, L. E., Peng, X., Garrett, R. A. &
encodes a prominent structural component of the Crystallogr. D Biol. Crystallogr. 66, 304–308 (2010). Martinez-Romero, E. Genome sequence of a novel
unique virion release system in Archaea. Virology 123. Krupovic, M., Cvirkaite-Krupovic, V., Prangishvili, D. archaeal rudivirus recovered from a mexican hot
404, 1–4 (2010). & Koonin, E. V. Evolution of an archaeal virus spring. Genome Announc. 1, e00040‑12 (2013).
101. Quax, T. E. et al. Simple and elegant design of a virion nucleocapsid protein from the CRISPR-associated 147. Servin-Garciduenas, L. E., Peng, X., Garrett, R. A. &
egress structure in Archaea. Proc. Natl Acad. Sci. USA Cas4 nuclease. Biol. Direct 10, 65 (2015). Martinez-Romero, E. Genome sequence of a novel
108, 3354–3359 (2011). 124. Makarova, K. S. et al. An updated evolutionary archaeal fusellovirus assembled from the metagenome
102. Snyder, J. C., Brumfield, S. K., Peng, N., She, Q. & classification of CRISPR-Cas systems. of a mexican hot spring. Genome Announc 1,
Young, M. J. Sulfolobus turreted icosahedral virus Nat. Rev. Microbiol. 13, 722–736 (2015). e00164‑13 (2013).
c92 protein responsible for the formation of pyramid- 125. Burstein, D. et al. Major bacterial lineages are 148. Garcia-Heredia, I. et al. Reconstructing viral genomes
like cellular lysis structures. J. Virol. 85, 6287–6292 essentially devoid of CRISPR-Cas viral defence from the environment using fosmid clones: the case of
(2011). systems. Nat. Commun. 7, 10613 (2016). haloviruses. PLoS ONE 7, e33802 (2012).

738 | DECEMBER 2017 | VOLUME 15 www.nature.com/nrmicro


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

149. Santos, F. et al. Metagenomic approach to the study 158. Paul, B. G. et al. Targeted diversity generation by funding grants 283072 and 255342. P.F. is supported by the
of halophages: the environmental halophage 1. intraterrestrial archaea and archaeal viruses. European Research Council under the European Union’s
Environ. Microbiol. 9, 1711–1723 (2007). Nat. Commun. 6, 6585 (2015). Seventh Framework Program (FP/2007-2013)/Project
150. Martinez-Garcia, M., Santos, F., Moreno-Paz, M., 159. Simmonds, P. et al. Consensus statement: virus EVOMOBIL — ERC Grant Agreement no. 340440. J.I. and
Parro, V. & Anton, J. Unveiling viral-host interactions taxonomy in the age of metagenomics. Nat. Rev. E.V.K. are supported by intramural funds of the US
within the ‘microbial dark matter’. Nat. Commun. 5, Microbiol. 15, 161–168 (2017). Department of Health and Human Services (to the National
4542 (2014). 160. Häring, M. et al. Morphology and genome Library of Medicine). M.K. is supported by l’Agence Nationale
151. Santos, F. et al. Metatranscriptomic analysis of organization of the virus PSV of the hyperthermophilic de la Recherche (France) project ENVIRA.
extremely halophilic viral communities. ISME J. 5, archaeal genera Pyrobaculum and Thermoproteus:
1621–1633 (2011). a novel virus family, the Globuloviridae. Virology 323, Author contributions
152. Garrett, R. A. et al. Metagenomic analyses of novel 233–242 (2004). D.P., M.K., D.H.B. and E.V.K. researched data for the article.
viruses and plasmids from a cultured environmental 161. Stedman, K. in The Springer Index of Viruses (eds D.P., M.K., P.F., E.V.K. and J.I. substantially contributed to
sample of hyperthermophilic neutrophiles. Tidona, C. & Darai, G.) 561–566 (Springer-Verlag, discussion of content. D.P., M.K. and E.V.K. wrote the article.
Environ. Microbiol. 12, 2918–2930 (2010). 2011). D.P., M.K. and E.V.K. reviewed and edited the manuscript
153. Bolduc, B. et al. Identification of novel positive-strand 162. Schleper, C., Kubo, K. & Zillig, W. The particle SSV1 before submission.
RNA viruses by metagenomic analysis of archaea- from the extremely thermophilic archaeon Sulfolobus
dominated Yellowstone hot springs. J. Virol. 86, is a virus: demonstration of infectivity and of Competing interests statement
5562–5573 (2012). transfection with viral DNA. Proc. Natl Acad. Sci. USA The authors declare no competing interests.
154. Kazlauskas, D., Krupovic, M. & Venclovas, C. The logic 89, 7645–7649 (1992).
of DNA replication in double-stranded DNA viruses: 163. Porter, K. et al. SH1: A novel, spherical halovirus Publisher’s note
insights from global analysis of viral genomes. isolated from an Australian hypersaline lake. Virology Springer Nature remains neutral with regard to jurisdictional
Nucleic Acids Res. 44, 4551–4564 (2016). 335, 22–33 (2005). claims in published maps and institutional affiliations.
155. Chow, C. E., Winget, D. M., White, R. A., 3 rd, 164. Rice, G. et al. The structure of a thermophilic archaeal
Hallam, S. J. & Suttle, C. A. Combining genomic virus shows a double-stranded DNA viral capsid type
sequencing methods to explore viral diversity and that spans all domains of life. Proc. Natl Acad. Sci. DATABASES
reveal potential virus-host interactions. USA 101, 7716–7720 (2004). RCSB Protein Data Bank:
Front. Microbiol. 6, 265 (2015). http://www.rcsb.org/pdb/home/home.do
156. Labonte, J. M. et al. Single-cell genomics-based Acknowledgements 1J5Q | 5VKU | 3ZMN | 3ZN4 | 1HX6 | 1OHG | 1BBD
analysis of virus-host interactions in marine surface D.P. is supported by l’Agence Nationale de la Recherche
bacterioplankton. ISME J. 9, 2386–2399 (2015). (France) project EXAVIR and by the European Union’s Horizon SUPPLEMENTARY INFORMATION
157. Burstein, D. et al. New CRISPR-Cas systems from 2020 research and innovation programme under grant See online article: S1 (table)
uncultivated microbes. Nature 542, 237–241 agreement 685778, project VIRUS‑X. D.H.B. was supported ALL LINKS ARE ACTIVE IN THE ONLINE PDF
(2017). by the Academy Professors programme, Academy of Finland

NATURE REVIEWS | MICROBIOLOGY VOLUME 15 | DECEMBER 2017 | 739


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

Вам также может понравиться