ArticlePDF AvailableLiterature Review

Prangishvili D, Forterre P, Garrett RA.. Viruses of the Archaea: a unifying view. Nat Rev Microbiol 4: 837-848

Authors:

Abstract

DNA viruses of the Archaea have highly diverse and often exceptionally complex morphotypes. Many have been isolated from geothermally heated hot environments, raising intriguing questions about their origins, and contradicting the widespread notion of limited biodiversity in extreme environments. Here, we provide a unifying view on archaeal viruses, and present them as a particular assemblage that is fundamentally different in morphotype and genome from the DNA viruses of the other two domains of life, the Bacteria and Eukarya.
The discovery of the Archaea was a significant break-
through in the recent history of biology. Whereas cell-
ultrastructure studies had initially suggested a division
of living organisms into eukaryotes and prokaryotes,
molecular sequence analyses — pioneered by Carl Woese
in the 1970s — revealed the existence of three different
classes of ribosomal RNAs and ribosomes in cellular
organisms. This discovery led to the replacement of the
prokaryote/eukaryote dichotomy by a trinity of domains,
the Archaea, Bacteria and Eukarya
1
. Subsequently, ribos-
omal RNA sequence comparisons led to the division
of the archaeal domain into two main kingdoms, the
Crenarchaeota and the Euryarchaeota
2
.
Over the past three decades we have accrued a broad
knowledge of the biological diversity of the Archaea.
This includes an outline of their physiology, biochem-
istry and molecular biology, and many insights into
their evolutionary relationships with the Bacteria and
the Eukarya
3
. Although Archaea resemble Bacteria
in their cellular ultrastructure and genome organization,
their DNA replication, transcription and translation
machineries show many similarities to their eukaryotic
counterparts. In addition, other features seem to have
either arisen, or have been exclusively conserved, within
the archaeal domain (for example, ether-linked mem-
branes). In this context, archaeal viruses are particularly
interesting.
The first two archaeal viruses that were isolated
visually resembled bacteriophage T4 and other mem-
bers of the family Myoviridae, with icosahedral heads,
contractile helical tails and linear, double-stranded (ds)
DNA genomes
4,5
. Subsequently, a few head-tail archaeal
viruses were reported with non-contractile tails similar
to lambdoid bacteriophages of the family Siphoviridae
6,7
.
Based on these initial studies, it was inferred, albeit erro-
neously, that archaeal viruses constituted a variety of the
ubiquitous head-tail bacteriophages.
Recently however, this view has changed radically.
Electron-microscopy investigations of samples collected
from natural environments that contain predominantly
archaea, and the enrichment cultures derived therefrom,
revealed that the head-tail phenotype is rare among the
archaeal viruses (for reviews, see
REFS 8,9). In fact, cul-
tured archaeal viruses, which to date all have dsDNA
genomes, exhibit a range of virion morphotypes, most
of which have not been observed before for any dsDNA
virus. There are exceptional forms, including
fusiforms,
droplet and bottle shapes, and linear and spherical virions,
with more complex virions combining features of these
different forms. Moreover, genome-sequence analyses
have demonstrated that most of the archaeal viruses are
unrelated to other known viruses and suggest that they
might have different, and possibly multiple, evolutionary
origins
10
.
Below we present the archaeal viruses, grouped accord-
ing to the gross features of their virion morphotypes, and
describe their genomic properties and relationships with
the host cell. Moreover, the evolutionary relationships
of the archaeal viruses to those of the Bacteria and Eukarya
are discussed. The taxonomic assignments used have
either been approved, or are pending, at the International
Committee on Taxonomy of Viruses (
ICTV).
Fusiform viruses
Viruses with fusiform virions, single or two-tailed, are
common in, and exclusive to, the Archaea
11–22
. They
constitute a large fraction of the known archaeal viruses
and are associated with a broad range of hosts repre-
senting the three main phenotypes of cultured archaeal
species, the
hyperthermophiles, extreme halophiles and
anaerobic methane-producers from the Euryarchaeota
and Crenarchaeota kingdoms
(TABLE 1). Moreover,
electron-microscopy analysis revealed an abundance
of fusiform-virus-like particles in habitats where
these archaea predominate, for example in hot, acidic
springs
23–25
and in hypersaline waters
26,27
.
The isolated fusiform viruses are diverse in their
structural and genomic characteristics, and they
*Molecular Biology of the
Gene in Extremophiles Unit,
Institut Pasteur, rue du
Docteur Roux 25, F-75724
Paris Cedex 15, France.
Danish Archaea Centre,
Institute of Molecular Biology,
Copenhagen University,
Sølvgade 83H, DK-1307
Copenhagen K, Denmark.
Correspondence to D.P.
e-mail: prangish@pasteur.fr
doi:10.1038/nrmicro1527
Fusiform
An organism that is spindle-
shaped: wider in the middle
and tapering towards the ends.
Hyperthermophile
An organism that has an
optimal growth temperature
above 80°C.
Extreme halophile
An organism that requires
extremely high levels of sodium
chloride for growth.
Viruses of the Archaea: a unifying view
David Prangishvili*, Patrick Forterre* and Roger A. Garrett
DNA viruses of the Archaea have highly diverse and often exceptionally complex
morphotypes. Many have been isolated from geothermally heated hot environments, raising
intriguing questions about their origins, and contradicting the widespread notion of limited
biodiversity in extreme environments. Here, we provide a unifying view on archaeal viruses,
and present them as a particular assemblage that is fundamentally different in morphotype
and genome from the DNA viruses of the other two domains of life, the Bacteria and Eukarya.
REVIEWS
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 4
|
NOVEMBER 2006
|
837
© 2006 Nature Publishing Group
Positively supercoiled DNA
A DNA molecule in which the
number of topological links
between the two strands is
superior to the number of
turns.
Lysogen
A bacterium or archaeon that
contains a viral genome
integrated into the
chromosome.
have been classified into the Fuselloviridae family,
the proposed ‘Bicaudaviridae’ family and the genus
Salterprovirus, although some remain unclassified. Most
have a circular genome and carry an integrase gene
that can facilitate integration into host chromosomes
28
.
The exceptions are the haloarchaeal salterproviruses
Haloarcula hispanica virus 1 (
His1) and Haloarcula his-
panica virus (
His2), which have linear genomes lacking
integrase genes
11
.
The Fusellov iridae family. Known members of the
Fuselloviridae infect the hyperthermophilic crenar-
chaeon Sulfolobus
(TABLE 1), and their replication and
release leave the host cell intact. The virions, which
are in the size range 55–60 × 80–100 nm, have short
tails which are uniform in size and carry thin terminal
fibres
15,16,18–20
(FIG. 1a). The best studied is the type spe-
cies of the family, the Sulfolobus spindle-shaped virus
1 (SSV1). Its circular genome is
positively supercoiled
29
and during infection it integrates into a tRNA gene in
the host chromosome, producing a partitioned integrase
gene while the tRNA gene remains intact
28
. UV irradia-
tion or mytomycin treatment induce viral replication
and temporarily inhibit the growth of
lysogens without
causing their lysis
15
.
The transcription pattern of the SSV1 genome is
relatively simple, with most transcripts being produced
constitutively (for a review, see
REF. 30). However, follow-
ing UV irradiation, upregulation of some constitutive
transcripts is observed, together with the appearance of
a short RNA molecule that might facilitate the initiation
of DNA replication.
The ‘Bicaudaviridae’ family. The sole member
of this proposed family, the Acidianus two-tailed
virus (ATV)
14,17
, is also the only known virus of the
acidophilic, hyperthermophilic archaea that is capable
of host lysis. Its reproductive cycle has some unique
features
14,17
. Virions are extruded from host cells as
tail-less, fusiform particles, which then develop long
tails at each pointed end at temperatures above 75°C,
close to the temperature of the natural habitat of the
host
(FIG. 1b). This major, extracellular morphological
development is independent of the host cells or any
energy sources, and its molecular mechanism remains
unclear. The tails consist of tubes, which terminate in
an anchor-like structure, and contain a periodic fila-
mentous structure
(FIG. 2). One function of the elon-
gated, flexible tails might be to enhance the probability
of virion adsorption to a new host cell.
There is circumstantial evidence that the newly
discovered process of extracellular tail development of
ATV might be shared by other fusiform viruses of the
hyperthermophilic archaea. In growth cultures of
uncharacterized Acidianus species, fusiform particles
have often been observed with one or two tails, which
can differ in length
23
. This could reflect different stages
of extracellular morphogenesis.
The Salterprovirus genus. The two distantly related
viruses His1 and His2, which have been assigned to
this genus, infect strains of the extremely halophilic
genus Haloarcula of the Euryarchaeota
11,12
. Both
viruses are lytic and the linear genomes show no
sequence similarity to the circular genomes of the
other fusiform viruses. The virions of these viruses
are similar in size (44 × 67–77 nm) but, despite
pronounced morphological similarities, their major
structural proteins are not orthologous. Each of the
viruses encodes a DNA polymerase which might be
primed by proteins attached to the termini of their
linear genomes.
Table 1 | Archaeal viruses with exceptional morphologies
Family and genus Species Archaeal
kingdom
Genome
details*
References
Fusiform viruses
Fuselloviridae, Fusellovirus Sulfolobus spindle-shaped virus 1 (SSV1) Cr c, 15.5 15, 16, 18
Sulfolobus spindle-shaped virus 2 (SSV2) Cr c, 14.8 19
Sulfolobus spindle-shaped virus - Yellowstone 1 (SSV-Y1) Cr c, 16.5 20
Sulfolobus spindle-shaped virus - Kamchatka 1 (SSV-K1) Cr c, 17.4 20
Fuselloviridae, Salterprovirus Haloarcula hispanica virus 1 (His1) Eu ln, 14.5 11, 12
Haloarcula hispanica virus 2 (His2) Eu ln, 16.1 11, 12
Bicaudaviridae, Bicaudavirus’
Acidianus two-tailed virus (ATV) Cr c, 62.7 14, 17
Unclassified Sulfolobus tengchongensis spindle-shaped virus 1 (STSV1) Cr c, 75.3 22
Pyrococcus abyssi virus 1 (PAV1) Eu c, 18.0 13
Methanococcus voltae virus-like particle (?) Eu c, 23.0 21
Bottle- and droplet-shaped viruses
Ampullaviridae, Ampullavirus
Acidianus bottle-shaped virus (ABV) Cr ln, 23.9 31
Guttaviridae, Guttavirus Sulfolobus neozealandicus droplet-shaped virus (SNDV) Cr c, 20.0 32
*Genome details shown are the form of the genome and the size in kb.
Taxonomic proposal pending at the International Committee on Taxonomy of Viruses.
c, covalently closed circular; Cr, Crenarchaeota; Eu, Euryarchaeota; ln, linear.
REVIEWS
838
|
NOVEMBER 2006
|
VOLUME 4 www.nature.com/reviews/micro
© 2006 Nature Publishing Group
d
a
b
c
Unclassified fusiform viruses. The virion (107 ×
230 nm) of the still unclassified Sulfolobus tengchon-
gensis spindle-shaped virus 1 (
STSV1) of the hyper-
thermophilic crenarchaeon Sulfolobus is the largest
of the known fusiform viruses and has a short tail,
the length of which varies in the range 0–133 nm
22
.
Although variable tail length could result from the
extracellular tail-development process described above
for ATV, the genome shows minimal similarity in
gene content to either ATV or other fusiform viruses.
Exceptionally, it carries open reading frames (ORFs)
encoding gene products implicated in DNA modifica-
tion and has a complex pattern of DNA methylation.
A putative origin of replication has been identified,
which carries multiple AT-rich repeats.
Fusiform virus-like particles with short tails, named
Pyrococcus abyssi virus 1 (PAV1), have also been reported
to be produced by a member of the Euryarchaeota, the
hyperthermophile
Pyrococcus abyssi
13
. Although the
putative virus resembles crenarchaeal fuselloviruses in
size, morphology and genome structure, none of the
annotated encoded proteins exhibit sequence similarity
to proteins encoded by SSV1 and its relatives (C. Geslin,
personal communication). Another fusiform virus of
the Euryarchaeota might be represented by pleomorphic
particles produced by the methanogen
Methanococcus
voltae
21
, containing circular dsDNA which is also found
integrated in the host chromosome. However, owing to
the distortion of particles by purification procedures, the
report of the morphotype of these particles is contro-
versial and it is difficult to distinguish whether they are
lemon-shaped or oblate.
Bottle-shaped and droplet-shaped viruses
The virions of two other archaeal viruses, the Acidianus
bottle-shaped virus (ABV) and the Sulfolobus neozealan-
dicus droplet-shaped virus (SNDV), have morphological
features that are so unique that each of the viruses has
been assigned to a new family (
FIG. 1 c,d; TABLE 1).
The ABV virion has a complex form resembling a
bottle (230-nm long, 4–75-nm wide). The virus infects
the hyperthermophilic Acidianus genus and has been
assigned to the new family Ampullaviridae’
31
. The
virion has no elements of icosahedral or helical sym-
metry
(FIG. 1c) and differs in its basic architecture from
any known virus. The envelope encases a cone-shaped
core formed by a torroidally supercoiled nucleoprotein
filament. A disc is present at the broader end, to which
20 (±2) short, thick filaments are attached
(FIG. 1c). Their
function remains unclear, while the virion seems to
adsorb to the host cell through its narrow end. As for the
genomes of salterproviruses, the ABV genome seems to
be replicated by a virus-encoded DNA polymerase that
is primed by a protein attached to the genomic termini.
The virus genome also encodes a putative RNA molecule
with notable secondary structural similarity to the RNA
molecule that has been implicated in DNA packaging of
the bacteriophage φ29 and its relatives (X. Peng, personal
communication).
The virus SNDV of the hyperthermophilic genus
Sulfolobus exhibits a complex droplet-shaped virion
(90 × 180 nm) and is the sole member of the family
Guttaviridae
32
. The droplet carries multiple long, thin
fibres that are attached at its apex, and the surface seems
to be helically ribbed
(FIG. 1d). The circular dsDNA
genome is extensively modified, probably by methylation,
and has not been sequenced.
Linear viruses
Linear particles are the main virion type found in ter-
restrial hot environments (>80°C) where crenarchaea
of the genera Sulfolobus, Acidianus and Thermoproteus
predominate
23–25,31
. All linear viruses isolated from
these environments to date infect members of these
genera. They have dsDNA genomes, a property not
previously observed for any linear virus, and there-
fore they have been classified into two new families:
the stiff, rod-like Rudiviridae
33–35
, and the flexible,
filamentous Lipothrixviridae
36–40
. Originally, the dis-
crimination of the families was based mainly on differ-
ences in virion structure and this was later supported
by comparative genomics. Nevertheless, in contrast to
members of the other archaeal viral families for which
genome sequences are available, a substantial frac-
tion of orthologous genes, including some encoding
glycosyl transferases and transcriptional regulators,
are shared by the rudiviruses and lipothrixviruses
10
.
For example, of the 45 predicted genes of the rudivirus
Sulfolobus islandicus rod-shaped virus 1 (
SIRV1), nine
share orthologues with the lipothrixvirus Sulfolobus
islandicus filamentous virus (
SIFV). These observa-
tions indicate that it is possible that the known linear
archaeal viruses share a relatively recent common
ancestor
10
. Therefore, we propose to classify the families
Figure 1 | Electron micrographs of archaeal viruses with exceptional
morphologies. a | Sulfolobus spindle-shaped virus 1 (SSV1) (inset) and its extrusion
from the host cell. b | The extracellularly developed Acidianus two tailed virus (ATV)
(inset) and its extrusion from the host cell. c | Acidianus bottle-shaped virus (ABV).
d | Sulfolobus neozealandicus droplet-shaped virus (SNDV). All images are negatively
stained with uranyl acetate, except for part b, which was platinum-shadowed. Scale
bars represent 100 nm. Parts a and d are courtesy of W. Zillig. Part b is reproduced
from
Nature REF. 14 © (2005) Macmillan Publishers Ltd. Part c is reproduced with
permission from
REF. 31 © (2005) American Society for Microbiology.
REVIEWS
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 4
|
NOVEMBER 2006
|
839
© 2006 Nature Publishing Group
>75°C
Rudiviridae and Lipothrixviridae into a new viral order,
the ‘Ligamenvirales’ (from the Latin ligamen, for string,
thread).
Members of the ‘Ligamenvirales’ have linear dsDNA
genomes and show similarities in their relationships
with their host cells. In contrast to archaeal viruses with
circular genomes, no integrated viral genomes, or
fragments thereof, have been detected in host chromo-
somes. Consistent with this observation, genomes of
the ligamenviruses lack an integrase gene. It seems that
these viruses persist stably in host cells and that their
replication is not induced by stress factors such as UV
irradiation or mytomycin C treatment. The only known
exception is the lytic lipothrixvirus Thermoproteus tenax
virus 1 (TTV1) of the neutrophilic hyperthermophile
genus Thermoproteus
39
.
The linear genomes of members of theLigamenvirales’
have diverse terminal repeat structures ranging from
the large (~1.5–2 kb) inverted terminal repeats (ITRs)
of the rudiviruses to the shorter (~0.5–1 kb) ITRs of the
lipothrixviruses. Many ITRs have conserved sequence
motifs including, for example, short, regularly spaced,
direct repeats in the rudiviruses and the lipothrixvirus
SIFV
34
. Remarkably, the multiple repeat pattern at the
termini of the genome of the lipothrixvirus Acidianus
filamentous virus 1 (
AFV1) resemble the telomeric ends
of the linear chromosomes of eukaryotes
37
.
The two DNA strands of the rudiviral genomes are
covalently linked, 5-3, generating small terminal loops
at each end
41
, however there is no evidence so far for
similar structures being present in the lipothrixviral
genomes
36
.
The Rudiviridae family. Virions belonging to this fam-
ily do not have an envelope and vary considerably in
length (the average size is 23 × 610–900 nm), with the
length proportional to the size of the linear dsDNA
genomes for the three characterized rudiviruses
33,35
(
FIG. 3a; TABLE 2). The virions contain a tube-like super-
helix formed by linear dsDNA and copies of a single,
glycosylated, basic DNA-binding protein. Plugs that are
approximately 50 nm in length are located at each end
of the tube-like structure and each carries three short
tail fibres (
FIG. 3a).
SIRV1 replicates producing head-to-head and tail-to-
tail linked replicative intermediates, which are probably
resolved by the Holliday junction resolvase that is com-
mon to all rudiviruses
34,42
. The formation of these inter-
mediates is consistent with a self-priming mechanism of
replication, similar to that proposed for some members
of a group of the larger eukaryal nucleocytoplasmic
DNA viruses (NCLDV)
43
.
Consistent with the rudiviral–host relationships
being relatively unsophisticated, in vivo studies have
demonstrated a simple transcription pattern for the viral
genomes, with few genes exhibiting temporal regula-
tion
44
. A host-encoded transcription activator was shown
to be involved in the regulation of viral transcription
45
.
The Lipothrixviridae family. The flexible, filamentous
virions of members of this family are surrounded by
envelopes containing lipids that have been acquired
from the host. They show considerable diversity in their
terminal structures, which, in combination with differ-
ing properties of their linear genomes and presumed
differing replication mechanisms, has provided a basis
for their classification into four genera.
The virion of the α lipothrixvirus TTV1 (400 nm
× 38 nm) has a lipid bilayer envelope which encases a
helical core consisting of DNA covered by multimers of
two DNA-binding proteins
39
. The β lipothrixvirus SIFV
(1,950 nm × 24 nm) differs in that the virion termini
taper, ending in mop-like structures to which six tail
fibres are attached
36
. This structure unfolds like a spi-
der’s legs, before attaching to receptors on the host cell
membrane. For the γ lipothrixvirus AFV1 (900 × 24 nm),
the virion termini carry exceptional claw-like structures
(FIG. 3b), which clamp onto viral receptors located on host
cell pili and maintain a firm contact
37
. The virion of the
δ lipothrixvirus AFV2 (1100 × 24 nm) has a complex
collar at the termini with two sets of attached filaments,
resembling a bottle brush
38
(FIG. 3c).
Spherical viruses
The known species of archaeal spherical viruses exhibit
virions of two main types (
FIG. 3 d,e; TABLE 2). The
spherical virion (diameter, 100 nm) of the Pyrobaculum
spherical virus (
PSV), a member of the proposed family
Globuloviridae, is enveloped and contains multimers
of a 33-kDa protein
46
(FIG. 3d). The lipid-containing
envelope encases a nucleoprotein core which has a
superhelical arrangement. PSV infects species of the
anaerobic and hyperthermophilic crenarchaeal genera
Pyrobaculum and Thermoproteus. The linear genome
carries 190-bp ITRs, and encodes a putative viroid-
like RNA molecule of unknown function (X. Peng,
personal communication). The Thermoproteus tenax
spherical virus 1 (
TTSV1), with similar morphological
Figure 2 | Extracellular tail development of Acidianus two-tailed virus (ATV).
Cryo-electron micrographs of tailless and two-tailed virions. The lower panels show
sections through the three-dimensional reconstructions of different portions of the
negatively stained tails obtained by electron tomography. Scale bars represent 50 nm.
Modified with permission from
REF.17 © (2006) Elsevier.
REVIEWS
840
|
NOVEMBER 2006
|
VOLUME 4 www.nature.com/reviews/micro
© 2006 Nature Publishing Group
a
b
d
e
c
Mesophilic
An organism that grows best in
a temperature range between
20°C and 45°C.
and genomic properties to PSV, should probably also be
assigned to this family
47
.
The two other known spherical viruses, Sulfolobus
turreted icosahedral virus (
STIV)
48,49
and Haloarcula
hispanica virus (
SH1) of the haloarchaeal genera
Haloarcula and Halorubrum
50,51
, reveal some mor-
phological similarities. The virions of both are non-
tailed icosahedra with an internal lipid layer
(FIG. 3e).
Therefore, they share an architectural principle with
virions of the bacterial Tectiv iridae family. Moreover,
the crystal structure of the 37-kDa major capsid pro-
tein of STIV is closely similar to those of the major
capsid proteins of the bacterial tectivirus PRD1 and
the eukaryal phycodnavirus Paramecium bursaria
Chlorella virus 1 (
PBCV1), suggesting a common
ancestry, although the protein sequences show no sig-
nificant similarity
52
. Image reconstruction of the STIV
virion revealed a unique virus architecture including
complex, turret-like projections extending from each
of the vertices
48
.
SH1 is a lytic virus with a high burst-size of approxi-
mately 200 virus particles per cell, and the release mech-
anism is based on cell disruption
51
. By contrast, STIV is
non-lytic and persists stably in the host cell. In line with
differences in morphology and virus-host interactions,
the genomes of the two viruses reveal no evidence of
homologous gene content. Furthermore, although the
STIV genome is circular, that of SH1 is linear, carrying
309-bp ITRs.
Head-tail viruses
In addition to the icosahedral viruses, archaea repli-
cate another virion morphotype that is common to
bacterial viruses. These are the head-tail phages, with
non-enveloped virions carrying icosahedral heads and
helical tails
(FIG. 4). All are associated with the kingdom
Euryarchaeota and they exclusively infect extreme halo-
philes or methanogens that are either
mesophilic or mod-
erately thermophilic. Sixteen head-tail phages have been
reported, which exhibit diverse sizes of heads, tails and
linear dsDNA genomes, and they have been assigned to
the bacteriophage families Myoviridae and Siphoviridae
(for reviews, see
REFS 8,10,53). Most head-tail phages
have not been studied beyond a basic description and
are currently unavailable in laboratory collections. The
best characterized are two pairs of closely related myo-
viruses, ΦH
54–57
and ΦCh1 (REFS 58–60), and HF1 and
HF2
(REFS 61–64), which infect haloarchaea, and the
siphovirus ΨM1 and its deletion mutant ΨM2, which
infect Methanothermobacter
65–67
(TABLE 3). In addition to
their phage-like virion structures, these viruses resem-
ble dsDNA bacteriophages in their genome content (see
below), and in possessing mosaic genomes that have
undergone extensive genetic exchange
68,69
.
Although so far viral diversity has been studied
in only a few archaea-rich habitats, it is becoming
increasingly clear that head-tail particles might be
a rare virion morphotype in environments where
archaea dominate. For example, in samples from high-
temperature hydrothermal sites, many morphotypes
are observed but rarely head-tail forms
23–25,70
. Moreover,
in hypersaline waters where haloarchaea predominate,
spindle-shaped, spherical and star-shaped virus-like
particles are most common
26,27
. Therefore, the abun-
dance of head-tail haloarchaeal viruses among those
isolated almost certainly reflects a major bias in the
approaches used for their detection and isolation.
These were based mainly on their ability to induce host-
cell lysis and produce plaques on lawns of a limited
diversity of host cells
53
.
The ecology of the archaeal viruses
Despite the widespread presence of archaea on our planet,
specific screening for archaeal viruses so far has only been
done in extreme hydrothermal and hypersaline environ-
ments. The available results suggest that the composition
of viral communities reflects that of their hosts and is
similar at different geographical locations with compa-
rable environmental conditions. Therefore, overlapping
subsets of known morphotypes of hyperthermophilic
viruses have been observed at geothermal environments
in Iceland, Eastern Russia (the Kamchatka peninsula), the
Naples region of Italy and Yellowstone National Park in
the USA, and in deep-sea hydrothermal vents
23–25, 31,70
.
Viruses from the same family which have been isolated
from different geographical sites and infect closely related
hosts often show strong sequence conservation. Therefore,
approximately half of the genes of PSV and TTSV1, which
have been assigned to the ‘Globuloviridae with origins in
the USA and Indonesia, respectively, are orthologous
10
. A
comparable degree of similarity occurs for members of the
Figure 3 | Electron micrographs of linear and spherical viruses of the archaea. The
figure shows: a | Sulfolobus islandicus rod-shaped virus 1 (SIRV1); b | Acidianus filamentous
virus 1 (AFV1); and c | Acidianus filamentous virus 2 (AFV2) (with terminal structures
shown in insets); d | Pyrobaculum spherical virus (PSV); e | Haloarcula hispanica virus
(SH1). All negatively stained with uranyl acetate. Scale bars represent 100 nm in main
images, and 50 nm in insets. Part a is courtesy of W. Zillig. Part b is reproduced with
permission from
REF.37 © (2003) Elsevier. Part c is reproduced with permission from
REF 38 © (2005) American Society for Microbiology. Part d is reproduced with
permission from
REF.46 © (2004) Elsevier. Part e is reproduced with permission from
REF.51 © (2005) Elsevier.
REVIEWS
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 4
|
NOVEMBER 2006
|
841
© 2006 Nature Publishing Group
Ribbon-helix-helix
(RHH). A structural motif
consisting of four helices in an
open array of two hairpins.
Helix-turn-helix
(HTH). A structural motif
common in DNA-binding
proteins; typically the second
helix fits into the DNA major
groove.
Fuselloviridae isolated on three different continents
20,71
.
These genomes contain a fairly conserved region, or
conserved sets of genes, and a more variable region
19,20
.
Some of this genetic diversity might arise by intergenomic
recombination between related species.
The high similarity in the gene content of viral spe-
cies from the same family that are separated by large
geographical distances is in contrast to the minimal
similarity observed for species from different families
that coexist in the same local microbial community.
The most extreme example of this is provided by four
viruses that were isolated from an acidic hot spring
(pH2, 85°C) at Pozzuoli, near Naples, Italy. The rod-
shaped ARV1, filamentous AFV2, bottle-shaped ABV
and two-tailed ATV, representing four different viral
families, coexist in the same habitat and can even
replicate in the same host strains. Despite this, these
viruses show no detectable signs of intergenomic
genetic exchange.
Acidothermophilic aquatic environments, from
which the most unusual viral morphotypes have been
isolated, contain a concentration of viruses that is sig-
nificantly lower than the concentration observed in
other analysed ecosystems
72,73
. This could be caused by
several factors including, for example, the limited stabil-
ity of virions at high temperatures and low pH, or by the
capacity of most known viruses from such environments
to persist stably in the host cell rather than cause lysis.
The latter strategy apparently reduces the possibility of
direct exposure of a virus to the harsh environmental
conditions. ATV, the only known lytic virus from an
acidic hot spring, finalizes its replication cycle extra-
cellularly under the environmental conditions that are
favourable for host growth, and this might contribute to
its survival strategy.
Archaeal virus genomics
The genes of archaeal viruses generally yield few sig-
nificant matches to sequences in the public sequence
databases
10
. The exceptions are the several homolo-
gous matches between genes of the euryarchaeal
head-tail viruses ΦCh1, ΨM1/2, HF1 and HF2,
and those of the head-tail bacteriophages from the
families Myoviridae and Siphoviridae
58,63,64,67
. These
viruses carry homologous genes encoding, for exam-
ple, proteins involved in capsid and tail formation and
virion assembly, as well as transcriptional regulators,
ATPases and nucleases.
For non-head-tail viruses, most predicted gene
products lack recognizable functions and homologues
in extant databases, other than in closely related
archaeal viruses. Identified functions are confined
to a few proteins involved in glycosylation (glycosyl-
transferases
74
), DNA replication, recombination and
integration (a Holliday junction resolvase
42
, DNA
polymerase (
REF.11 and X. Peng, personal communica-
tion), RecB endonuclease and integrase
28
), nucleotide
metabolism (a dUTPase
75
and thymidylate synthase),
diverse transcriptional regulatory proteins, mainly
of the
ribbon-helix-helix (RHH) and helix-turn-helix
(HTH) types, and functionally diverse ATPases. In a
recent re-evaluation and comparative study of all the
annotated genomes of crenarchaeal viruses, working
close to the limits of statistical significance, several
additional sequence matches were identified, in par-
ticular matches to small transcriptional regulatory
proteins. The functional subgroups of, for example,
AAA+ ATPases (ATPases associated with various cel-
lular activities), and DNA replication enzymes, were
classified in more detail
10
. Nevertheless, the current
uniqueness of most gene products of these viruses
Table 2 | Archaeal linear and spherical viruses
Family and genus Species Archaeal
kingdom
Genome
details*
References
Linear viruses
Rudiviridae, Rudivirus Sulfolobus islandicus rod-shaped virus 1 (SIRV1) Cr ln, 32.3 33
Sulfolobus islandicus rod-shaped virus 1 (SIRV2) Cr ln, 35.5 33
Acidianus rod-shaped virus 1 (ARV1) Cr ln, 24.7 35
Lipothrixviridae, α-lipothrixvirus
Thermoproteus tenax virus 1 (TTV1)
§
Cr ln, 15.9 39, 40
Lipothrixviridae, β-lipothrixvirus
Sulfolobus islandicus filamentous virus (SIFV) Cr ln, 40.9 36
Thermoproteus tenax virus 2 (TTV2)
§
Cr ND 39
Thermoproteus tenax virus 3 (TT3)
§
Cr ND 39
Lipothrixviridae, γ-lipothrixvirus
Acidianus filamentous virus 1 (AFV1) Cr ln, 21.1 37
Lipothrixviridae, δ-lipotrixvirus
Acidianus filamentous virus 2 (AFV2) Cr ln, 31.8 38
Spherical viruses
Globuloviridae
, ‘Globulovirus Pyrobaculum spherical virus (PSV) Cr ln, 28.3 46
Thermoproteus tenax spherical virus 1 (TTSV1) Cr ln, 20.9 47
Unclassified Haloarcula hispanica virus (SH1) Eu ln, 30.9 50, 51
Sulfolobus turreted icosahedral virus (STIV) Cr ln, 17.7 48
*Genome details show the form of the genome and the size in kb.
Currently not available in laboratory collections.
§
Taxonomic proposal pending at the
International Committee on Taxonomy of Viruses. c, covalently closed circular; Cr, Crenarchaeota; Eu, Euryarchaeota; ln, linear; ND, not determined.
REVIEWS
842
|
NOVEMBER 2006
|
VOLUME 4 www.nature.com/reviews/micro
© 2006 Nature Publishing Group
ab
von Willebrand factor A
motif
A structural motif that has
been implicated in the
formation of diverse types of
specific protein–protein
interaction and cell adhesion.
was verified, strengthening the idea that crenarchaeal
viruses are unrelated to any other viruses
10
. These
general properties and conclusions also extend to the
non head-tail euryarchaeal viruses, including the fusi-
form viruses His1, His2 and PAV, and the spherical
virus SH1, which have been sequenced and analysed
more recently (
REFS 11,50 and C. Geslin, personal
communication).
Few homologous genes are shared between
members of different virus families (except for the
Ligamenviridae’) and there is no evidence so far for
intergenomic recombination occurring between virus
families. Nevertheless, comparative studies on fusello-
virus genomes suggest that they carry one region in
which the gene content varies
19,20,76
, and evidence for
intergenomic recombination has recently been observed
among genomes of the lipothrixviruses (G. Vestergaard,
personal communication).
Exceptionally, the bicaudavirus ATV carries four
transposase genes, possibly present as insertion
sequence (IS) elements, which are similar in sequence to
the transposase genes in Sulfolobus spp. chromosomes
and conjugative plasmids
17
, suggesting that some interg-
enomic exchange has occurred
(FIG. 5a). ATV is excep-
tional in that it encodes six large proteins, in the size
range of 600–1,940 amino acids, which are rich in repeat
structures, with some exhibiting extensive predicted
coiled-coil regions, as well as regions of low complexity
sequence, one of which carries a
Von Willebrand factor
A motif
. At least some of these proteins seem to con-
tribute to the extracellular development of the tail-like
appendages
14,17
.
Cellular defence or regulatory mechanisms
Almost all archaeal genomes contain large clusters
of regularly spaced repeat elements (known as short
regularly spaced repeats (SRSR) or clustered regularly
interspaced short palindrome repeats (CRISPR)
77
) that
can constitute more than 1% of the whole genome
78
.
Much smaller clusters are found in about 50% of
bacterial genomes
78
. For a given archaeon cluster,
although the sequence of the repeats and size of
the spacer sequences tend to be highly conserved, the
spacer sequences themselves are generally unique
78
.
Initially, evidence was presented for the clusters
resembling centromere structures and being involved
in chromosomal segregation or plasmid partitioning
77
.
Although this hypothesis has not been refuted, recent
evidence from the same authors demonstrated that a
few of the spacer sequences present in these clusters
are highly similar to sequences found in the genomes
of archaeal rudiviruses or conjugative plasmids
79
. This
result was reinforced on the completion of the ATV
viral genome sequence, which revealed nine posi-
tive viral matches with spacer sequences in clusters
of the
Sulfolobus solfataricus P2 genome
78
(FIG. 5). As
the repeat clusters produce larger transcripts that are
probably processed into smaller fragments approach-
ing the size of the spacer sequences, they can poten-
tially target and inactivate viral messenger RNAs or
genes
78,80,81
. Therefore, it has been suggested that the
spacer sequences might provide the basis for a cellular
defence mechanism against future invasion by related
viruses or plasmids
79
. Bioinformatics evidence suggests
that the group of genes, designated cas for CRISPR-
associated, are physically associated and co-functional
with some clusters
82
. They might be involved in adding
new repeat-spacer units to the clusters by interact-
ing with the flanking sequences that adjoin actively
Figure 4 | Electron micrographs of head-tail viruses of
archaea and bacteria. a | The haloarchaeal virus φH1.
b | The bacteriophage P2. Both are negatively stained with
uranyl acetate. Scale bars represent 100 nm. Part a is
courtesy of W. Zillig. Part b is reproduced with permission
from http://www.biochem.wisc.edu/inman/empics.
Table 3 | Archaeal head-tail viruses*
Family and genus Species Host Archaeal
kingdom
Genome
details
References
Myoviridae, ‘φH-like viruses φH
Halobacterium salinarum Eu ln, 59.0 54–56
φCh1
Natrialba magadii Eu ln, 58.5 58,60
Unassigned species in the
family
HF1 Haloferax volcanii and
Halobacterium salinarum
Eu ln, 75.9 61,63
HF2 Halorubrum coriense Eu ln, 77.7 61–64
Siphoviridae, ‘ ψM1-like viruses ψM1/2
Methanothermobacter marburgensis Eu ln, 30.4/26.1 65–67
*Only those with sequenced genomes are listed.
Genome details show the form of the genome and the size in kb. Eu, Euryarchaeota; ln, linear.
REVIEWS
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 4
|
NOVEMBER 2006
|
843
© 2006 Nature Publishing Group
61
192
545
326b
618
892 (2×)
710
529
145
+145 +326b–618 –127 –61
127
ATV
62,730 bp
a
b
24bp identical repeats 41bp non-ATV spacers Spacers matching sequences
from corresponding ATV ORFs
Small interfering RNA
Non-coding RNAs (around 22
nucleotides long) derived from
the processing of long double-
stranded RNA during RNA
interference. They direct the
destruction of mRNA targets
that have the same sequence.
Micro RNA
A short (21–22-nucleotide)
RNA silencing trigger that is
processed from short stem–
loop precursors that are
encoded in the genomes of
metazoans and certain viruses.
growing clusters, possibly involving a retrotransposi-
tion mechanism
78,83
. The whole apparatus could be
evolutionarily related to the eukaryotic
small interfering
RNA
(siRNA) and micro RNA (miRNA) systems
83
.
Unresolved questions include why the clusters
are so extensive in archaea and why so many spacer
matches to ATV are found? One possible explana-
tion for the latter is that whereas most non-head-tail
archaeal viruses enjoy a benign relationship with
the host, and persist stably in the cell at a low copy
number without causing its lysis, a high level of ATV
replication results in host lysis and so perhaps the lytic
ATV induces a stronger host response.
Some insight into how archaeal viruses overcome cel-
lular defence mechanisms of the host is provided by the
rudivirus SIRV1. When passed through diverse, closely
related host strains of Sulfolobus islandicus, SIRV1 under-
goes major genomic changes, including transpositions,
duplications and deletions, all concentrated in six main
regions of the viral genome and often resulting in an
alteration in gene size
84
. Many of the changes are caused
by insertions or deletions of 12 bp, although the mecha-
nism by which this occurs is still not fully understood
84
.
However, the changes seem to reflect the adaptation of
the virus to the new host and could be a reaction to the
RNAs being produced from spacers of the chromosomal
repeat clusters.
Another mechanism of adaptation to cellular
defences might be exemplified by the protein TPX,
which is expressed from the α lipothrixvirus TTV1.
The genome carries two low-complexity sequence
regions consisting of the repetitive sequence CCNACN
(where N is any nucleotide) one of which is located
within the tpx gene and the other in a non-coding
region. Frequent recombination seems to occur
between these two regions, which generates multiple
variants of the TPX protein in the cell
85
.
Evolutionary considerations
Despite a modest number of isolated species, the
morphological diversity of the dsDNA viruses of
the Archaea strongly surpasses that of the numerous
known species of dsDNA viruses of the Bacteria
86
.
According to the recent report of the ICTV, 96% of
the 452 known species of bacterial dsDNA viruses are
non-enveloped head-tail bacteriophages, assigned to
three families of the order Caudovirales (Myoviridae,
Siphoviridae and Podoviridae), whereas only about
4% are non-enveloped tailless icosahedra or envel-
oped pleomorphic spheres
87
. As discussed above, in
addition to the Caudovirales and tailless icosahedral
viruses, archaea replicate a plethora of viruses with
morphotypes that are not encountered among dsDNA
viruses of either bacteria or eukarya
(FIG. 6). The diver-
sity of archaeal viruses in hot habitats is especially
striking compared with the relative uniformity of the
viral landscape in aquatic environments at moderate
and low temperatures, which is dominated by head-
tail viruses
88
. The origin and nature of this biodiver-
sity raises intriguing evolutionary questions. Possibly,
such diversity was once common in all environments
but was later reduced by the successful expansion of
bacteria and their phages in biotopes with moderate
and low temperatures, whereas hot environments still
remain a refuge for multiple unusual viral forms.
The occurrence of the Caudovirales among
archaeal viruses was earlier interpreted as evidence
for these viruses predating the divergence of the
Archaea and Bacteria, as it was believed that viruses
were unable to spread across domain boundaries
owing to the pronounced differences in the molecu-
lar biology and lifestyles of the two domains
89
.
However, it is remarkable that, in known cases,
caudoviruses infect archaea carrying a high percent-
age of genes of bacterial origin, in particular certain
mesophilic or moderately thermophilic haloarchaea
or methanoarchaea
90,91
. Apparently, these bacte-
rial genes have adapted successfully to an archaeal
cellular context. Furthermore, extremely halophilic
Figure 5 | Genome structure of Acidianus two-tailed virus (ATV). a | Map of the
circular ATV genome with open reading frames (ORFs) represented by arrows and
labelled according to the number of amino acids in the predicted protein. The ORFs
are colour coded as follows: grey, have homologues in other crenarchaeal viral
genomes; green, have homologues in archaeal conjugative plasmids; blue,
transposases; purple, genes for structural proteins. For the numbered ORFs,
fragments of identical, or near identical, sequences are present as CRISPR spacers
in the chromosome of Sulfolobus solfataricus P2. b | Repeat-spacer elements from a
section of the 96-repeat-spacer CRISPR cluster ‘d’ of S. solfataricus P2. The colour-
coded elements are: yellow, 24-bp identical repeats; grey, ~41-bp non-ATV spacers;
red, spacers matching sequences from corresponding ATV ORFs (+ and – indicate
the same and opposite direction, respectively, relative to the ORF mRNA). Part a is
modified with permission from
REF. 17 © (2006) Elsevier. Part b is modified with
permission from
REF.78.
REVIEWS
844
|
NOVEMBER 2006
|
VOLUME 4 www.nature.com/reviews/micro
© 2006 Nature Publishing Group
Myoviridae
Salterprovirus
Unclassified (Fuselloviridae?)
Fuselloviridae
Nimaviridae
Baculoviridae
Ascoviridae
Polydnaviridae
Herpesviridae
Adenoviridae
Polyomaviridae, Papillomaviridae
Poxviridae
Asfarviridae
Mimiviridae
Iridoviridae
Phycodnaviridae
STIV
SH1
Rudiviridae
Lipothrixviridae
Guttaviridae
‘Globuloviridae’
‘Ampullaviridae’
‘Bicaudaviridae’
Siphoviridae
Myoviridae
Siphoviridae
Podoviridae
Plasmaviridae
Corticoviridae
Tectiviridae
Bacteria Archaea
?
Eukarya
Euryarchaea Crenarchaea
Last universal common
ancestor
The progenitor from which all
current life is thought to have
evolved.
bacteria have been discovered, which could have
facilitated the adaptation of bacteriophages to rep-
licate in the haloarchaea
92
. Therefore, it now seems
likely that the Caudovirales first entered archaea by
interdomain spreading. Such an origin is also consist-
ent with archaeal caudaviruses carrying many genes
with homologues in either bacterial chromosomes or
plasmids.
Assuming that archaeal Caudovirales do, indeed,
originate from bacteria, then we are faced with the
intriguing perspective that each of the three domains
of life was originally characterized by a unique set of
associated dsDNA viruses
(FIG. 6). Could the associa-
tion of a specific set of viruses with each domain mean
that the old preconception of viruses originating from
their hosts is correct? Probably not, because there are
no apparent exclusive evolutionary links between the
basic viral components, the capsid and replication
apparatus found in one domain and the cellular com-
ponents of that domain. On the contrary, recent struc-
tural studies have shown that some viruses present in
different domains could have evolved from entities (or
a common gene pool) that predated the divergence of
the domains
93
. For instance, the icosahedral archaeal
viruses STIV and SH1 share a basic architectural
principle and/or a capsid protein with similar struc-
tures (demonstrated for STIV) with several bacte-
rial (Tectiviridae) and eukaryotic (Phycodnaviridae,
Adenoviridae) icosahedral viruses
52
.
One possible explanation for the existence of three
different ‘virospheres, each associated with a specific
domain, is that these virospheres were selected when
the domains first arose. Therefore, the first evolving
organisms of each separate domain could have already
been infected by different subsets of viruses from the
ancestral virosphere, which predated the
last universal
common ancestor
, LUCA. Although the descendants
of LUCA might have rapidly expanded on the planet,
the lineages leading to the three ancestors and their
associated viruses might well have originated later in
distinct areas and remained geographically isolated for
a long time period. Therefore, viruses might already
have been specialized to co-evolve with cells of one
domain and were unable to efficiently infect cells of
another domain when the cells started to share the
same environment. The mechanism and the tim-
ing of the speciation of the three domains remains
unknown, however, a recent hypothesis suggests that
DNA viruses could have played a direct role in their
formation by introducing DNA genomes into ancient
RNA cells
(BOX 1). Such a model would be compatible
with each domain being associated with its own viral
landscape if one assumes that three ancestral RNA cells
were infected by three different sets of DNA viruses at
the time of the RNA–DNA transition
(BOX 1).
The present lack of known archaeal RNA viruses
also raises an intriguing evolutionary issue. The strate-
gies recently used to detect and isolate archaeal viruses,
Figure 6 | The families of dsDNA viruses associated with the three domains of life. The virus families listed are
approved by the International Committee on Taxonomy of viruses, and the schematic representation of virions (not drawn
to scale) are presented as in
REF. 87. Proposed families are shown in inverted commas. SH1, Haloarcula hispanica virus 1;
STIV, Sulfolobus turreted icosahedral virus.
REVIEWS
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 4
|
NOVEMBER 2006
|
845
© 2006 Nature Publishing Group
E
LUCA
EV
Eukarya
Eukaryotic viruses
BBacteria
Bacterial
viruses
Archaeal
viruses
Archaea
BV
A
AV
?
Modern
virosphere
Ancient virosphere
Ancestors of
modern viruses
Modern
virosphere
at least from hydrothermal environments
72
, might have
been expected to yield RNA viruses, although one can-
not yet exclude a methodological problem. A possible
explanation for their absence could be that they were
already absent from the virosphere associated with the
common archaeal ancestor. Given that RNA is much
more labile than DNA at high temperature, this infer-
ence would concur with recent phylogenetic analyses
based on the available sequences of the translation
and transcription apparatus, which indicate that the
archaeal ancestor was a hyperthermophile
94
.
The discovery and exploration of the fascinating
archaeal DNA viruses has paralleled, and contributed
Box 1 | The origins of viruses
Three hypotheses have been
proposed to explain the origin
of viruses: (
a) they originated
in a pre-cellular world (the
‘virus first’ hypothesis); (
b)
they originated by a reduction
from parasitic cells; or (
c) they
originated from fragments of
cellular genetic material that
escaped from cellular control
and became parasites (the
escape hypothesis). Originally,
these hypotheses were
proposed in the framework of
the prokaryote/eukaryote
dichotomy. Just as the
erroneous concept of the
prokaryotes became the
paradigm for considering
bacterial evolution, the
escape hypothesis became the
paradigm for explaining the origin of viruses.
In its classical version, the escape hypothesis maintained that bacteriophages originated from bacterial genomes and
eukaryotic viruses from eukaryotic genomes. Amazingly, although most archaeal viruses are completely unrelated to
bacterial viruses, they are still classified as ‘bacteriophages’ in this outdated framework. For example, archaeal viruses are
illustrated under the heading “Families and genera infecting bacteria” in the latest edition of Virus Taxonomy:
Classification and Nomenclature of Viruses
(REF. 87). This occurs presumably because the archaeal domain is still not
recognized by some biologists.
The problem of virus origin remained deadlocked until recently, when progress in the molecular description
of viral proteins caused many scientists to realize that viruses form a world of their own, and that it is futile to
continue to speculate on their origin in the framework of the discredited prokaryote/eukaryote dichotomy.
The discovery that viruses which are associated with different domains can share similar, and apparently
homologous, features strongly suggests that viruses are ancient and that they predated the last universal
cellular ancestor (LUCA)
93,95
.
The early hypotheses for viral origin have now been re-evaluated in the context of this new framework
95
.
Currently, the main debate is between those who suggest a long period of acellular evolution (up to the actual
emergence of archaea and bacteria) and those who favour an early appearance of cells. Those who suggest the
former have revived the virus-first hypothesis, hypothesis (
a) above. For instance, Koonin and Martin recently
suggested that viruses emerged from an assemblage of self-replicating elements thriving in a hydrothermal vent,
using inorganic compartments as their hosts
96
. For some of those who favour an early emergence of cellular
organisms, viruses are considered to have originated from RNA–protein-based cells, either by reduction from
parasitic RNA cells or from genetic material that escaped from the genomes of these cells (variants of hypotheses (
b)
and (
c) above)
95
. A major question mark is the evolutionary relationship between DNA viruses and RNA viruses: did
DNA viruses originate from RNA viruses or from primitive DNA cells, or both
95,97
? In one hypothesis, DNA itself is
considered a viral invention, that is, DNA first appeared in viruses and was later transferred to cells
98
. In this model,
three such independent transfers could have initiated the three modern DNA lineages and the modern virospheres
associated with them
99
.
The figure shows a model of the formation of the three modern virospheres, coincident with the formation of the
three domains of life.
to, a recent upsurge of interest in the evolution of
viruses in general
(BOX 1). Although the known
archaeal viruses reveal an exceptional degree of diver-
sity with regard to both morphotypes and genomes,
this might still be an underestimation as they have
only been isolated from a limited number of host spe-
cies and taxa (mainly from the orders Sulfolobales and
Halobacteriales). Therefore, screening for virus–host
systems in other phylogenetic taxa of the archaeal
domain is an important research priority that prom-
ises to provide new insights into important questions
concerning the nature and origin of the modern
virosphere.
REVIEWS
846
|
NOVEMBER 2006
|
VOLUME 4 www.nature.com/reviews/micro
© 2006 Nature Publishing Group
1. Woese, C. R. & Fox, G. E. Phylogenetic structure of the
prokaryotic domain: the primary kingdoms. Proc. Natl
Acad. Sci. USA 74, 5088–5090 (1977).
2. Woese, C. R., Kandler, O. & Wheelis, M. L. Towards a
natural system of organisms: proposal for the domains
Archaea, Bacteria, and Eucarya. Proc. Natl Acad. Sci.
USA 87, 4576–4579 (1990).
3. Garrett, R. A. & Klenk, H.-P. (eds) Archaea. Evolution,
Physiology and Molecular Biology (Blackwell
Publishing, Oxford, 2006).
This is a recent and comprehensive book that
covers the ecology, physiology, molecular biology
and evolution of the Archaea.
4. Torsvik, T. & Dundas, I. D. Bacteriophage of
Halobacterium salinarium. Nature 248, 680–681
(1974).
5. Wais, A. C., Kon, M., MacDonald, R. E. & Stollar, B. D.
Salt-dependent bacteriophage infecting
Halobacterium cutirubrum and H. halobium. Nature
256, 314–315 (1975).
6. Pauling, C. Bacteriophages of Halobacterium halobium:
isolated from fermented fish sauce and primary
characterization. Can. J. Microbiol. 28, 916–921
(1982).
7. Rohrmann, G. F., Cheney, R. & Pauling, C.
Bacteriophages of Halobacterium halobium: virion
DNAs and proteins. Can. J. Microbiol. 29, 627–629
(1983).
8. Dyall-Smith, M., Tang, S. L. & Bath, C. Haloarchaeal
viruses: how diverse are they? Res. Microbiol. 154,
309–313 (2003).
9. Prangishvili, D. Evolutionary insights from studies on
viruses of hyperthermophilic archaea. Res. Microbiol.
154, 289–294 (2003).
10. Prangishvili, D., Garrett, R. A. & Koonin, E. V.
Evolutionary genomics of archaeal viruses: Unique
viral genomes in the third domain of life. Virus Res.
117 , 52–67 (2006).
Describes a detailed analysis of proteins encoded
by archaeal viruses, with an emphasis on the
comparative genomics of crenarchaeal viruses.
11. Bath, C., Cukalac, T., Porter, K. & Dyall-Smith, M. L.
His1 and His2 are distantly related, spindle-shaped
haloviruses belonging to the novel virus group,
Salterprovirus. Virology 350, 228–239 (2006).
12. Bath, C. & Dyall-Smith, M. L. His1, an archaeal virus
of the Fuselloviridae family that infects Haloarcula
hispanica. J. Virol. 72, 9392–9395 (1998).
13.
Geslin, C. et al. PAV1, the first virus-like particle
isolated from a hyperthermophilic euryarchaeote,
‘Pyrococcus abyssi’. J. Bacteriol. 185, 3888–3894
(2003).
14. Haring, M. et al. Virology: independent virus
development outside a host. Nature 436, 1101–1102
(2005).
15. Martin, A. et al. SAV1, a temperate u.v.-inducible DNA
virus-like particle from the archaebacterium Sulfolobus
acidocaldarius isolate B12. EMBO J. 3, 2165–2168
(1984).
16. Palm, P. et al. Complete nucleotide sequence of the
virus SSV1 of the archaebacterium Sulfolobus
shibatae. Virology 185, 242–250 (1991).
17. Prangishvili, D. et al. Structural and genomic
properties of the hyperthermophilic archaeal virus ATV
with an extracellular stage of the reproductive cycle.
J. Mol. Biol. 359, 1203–1216 (2006).
18. Schleper, C., Kubo, K. & Zillig, W. The particle SSV1
from the extremely thermophilic archaeon Sulfolobus
is a virus: demonstration of infectivity and of
transfection with viral DNA. Proc. Natl Acad. Sci. USA
89, 7645–7649 (1992).
19. Stedman, K. M. et al. Relationships between
fuselloviruses infecting the extremely thermophilic
archaeon Sulfolobus: SSV1 and SSV2. Res. Microbiol.
154, 295–302 (2003).
20. Wiedenheft, B. et al. Comparative genomic analysis of
hyperthermophilic archaeal Fuselloviridae viruses.
J. Virol. 78, 1954–1961 (2004).
21. Wood, A. G., Whitman, W. B. & Konisky, J. Isolation
and characterization of an archaebacterial virus-like
particle from Methanococcus voltae A3. J. Bacteriol.
171, 93–98 (1989).
22. Xiang, X. et al. Sulfolobus tengchongensis spindle-
shaped virus STSV1: virus–host interactions and
genomic features. J. Virol. 79, 8677–8686 (2005).
23. Rachel, R. et al. Remarkable morphological diversity
of viruses and virus-like particles in hot terrestrial
environments. Arch. Virol. 147, 2419–2429 (2002).
24. Rice, G. et al. Viruses from extreme thermal
environments. Proc. Natl Acad. Sci. USA 98,
13341–13345 (2001).
25. Zillig, W. et al. Screening for Sulfolobales, their
plasmids, and their viruses in Icelandic solfataras.
Syst. Appl. Microbiol. 16, 609–628 (1994).
References 23, 24 and 25 describe the diversity of
viral morphotypes observed in samples from hot
springs in the United States and Iceland.
26. Guixa-Boixareu, N., Calderon-Paz, J. I., Heldal, M.,
Bratbak, G., Pedros-Alio, C. Viral lysis and bacterivory
as prokaryotic loss factors along a salinity gradient.
Aquat. Microb. Ecol. 11, 215–227 (1996).
27. Oren, A., Bratbak, G. & Heldal, M. Occurrence of
virus-like particles in the Dead Sea. Extremophiles 1,
143–149 (1997).
Describes virus-like particles in samples from the
hypersaline Dead Sea.
28. Muskhelishvili, G., Palm, P. & Zillig, W. SSV1-encoded
site-specific recombination system in Sulfolobus
shibatae. Mol. Gen. Genet. 237, 334–342 (1993).
29. Nadal, M., Mirambeau, G., Forterre, P., Reiter, W.-D.
& Duguet, M. Positively supercoiled DNA in a virus-
like-particle of an archaebacterium. Nature 321,
256–258 (1986).
30. Zillig, W. et al. in The Biochemistry of Archaea
(Archaebacteria) (ed. Kates, M. et al.) 367–391
(Elsevier Science Publishers B. V., 1993).
31. Haring, M., Rachel, R., Peng, X., Garrett, R. A. &
Prangishvili, D. Viral diversity in hot springs of
Pozzuoli, Italy, and characterization of a unique
archaeal virus, Acidianus bottle-shaped virus, from a
new family, the Ampullaviridae. J. Virol. 79,
9904–9911 (2005).
Describes the viral diversity in a single acidic, hot
spring in southern Italy.
32. Arnold, H. P., Ziese, U. & Zillig, W. SNDV, a novel virus
of the extremely thermophilic and acidophilic
archaeon Sulfolobus. Virology 272, 409–416 (2000).
33. Prangishvili, D. et al. A novel virus family, the
Rudiviridae: Structure, virus-host interactions and
genome variability of the Sulfolobus viruses SIRV1 and
SIRV2. Genetics
152, 1387–1396 (1999).
34. Peng, X. et al. Sequences and replication of genomes
of the archaeal rudiviruses SIRV1 and SIRV2:
relationships to the archaeal lipothrixvirus SIFV and
some eukaryal viruses. Virology 291, 226–234 (2001).
35. Vestergaard, G. et al. A novel rudivirus, ARV1, of the
hyperthermophilic archaeal genus Acidianus. Virology
336, 83–92 (2005).
36. Arnold, H. P. et al. A novel lipothrixvirus, SIFV, of the
extremely thermophilic crenarchaeon Sulfolobus.
Virology 267, 252–266 (2000).
37. Bettstetter, M., Peng, X., Garrett, R. A. &
Prangishvili, D. AFV1, a novel virus infecting
hyperthermophilic archaea of the genus Acidianus.
Virology 315, 68–79 (2003).
38. Haring, M. et al. Structure and genome organization
of AFV2, a novel archaeal lipothrixvirus with unusual
terminal and core structures. J. Bacteriol. 187,
3855–3858 (2005).
39. Janekovic, D. et al. TTV1, TTV2, TTV3, a family of
viruses of the extremely thermophilic, anaerobic
sulfur-reducing archaebacterium Thermoproteus
tenax. Mol. Gen. Genet. 192, 39–45 (1983).
40. Neumann, H., Schwass, V., Eckerskorn, C. & Zillig, W.
Identification and characterization of the genes
encoding three structural proteins of the
Thermoproteus tenax virus TTV1. Mol. Gen. Genet.
217, 105–110 (1989).
41. Blum, H., Zillig, W., Mallok, S., Domdey, H. &
Prangishvili, D. The genome of the archaeal virus
SIRV1 has features in common with genomes of
eukaryal viruses. Virology 281, 6–9 (2001).
42. Birkenbihl, R. P., Neef, K., Prangishvili, D. &
Kemper, B. Holliday junction resolving enzymes of
archaeal viruses SIRV1 and SIRV2. J. Mol. Biol. 309,
1067–1076 (2001).
43. Moss, B. in Fields Virology (eds Fields, B. N., Knipe,
D. M. & Howley, P. M.) 2637–2671 (Lippincott-Raven,
Philadelphia, 1996).
44. Kessler, A., Brinkman, A. B., van der Oost, J. &
Prangishvili, D. Transcription of the rod-shaped viruses
SIRV1 and SIRV2 of the hyperthermophilic archaeon
Sulfolobus. J. Bacteriol. 186, 7745–7753 (2004).
45. Kessler, A. et al. A novel archaeal regulatory protein,
Sta1, activates transcription from viral promoters.
Nucleic Acids Res. 14 Sep 2006 (doi:10.1093/nar/
gk1502).
Reports on the strategy of an archaeal virus to
co-opt a host-cell regulator to promote the
transcription of some of its genes.
46. Haring, M. et al. Morphology and genome
organization of the virus PSV of the hyperthermophilic
archaeal genera Pyrobaculum and Thermoproteus: a
novel virus family, the Globuloviridae. Virology 323,
233–242 (2004).
47. Ahn, D. G. et al. TTSV1, a new virus-like particle
isolated from the hyperthermophilic crenarchaeote
Thermoproteus tenax. Virology 351, 280–290 (2006).
48. Rice, G. et al. The structure of a thermophilic archaeal
virus shows a double-stranded DNA viral capsid type
that spans all domains of life. Proc. Natl Acad. Sci.
USA 101, 7716–7720 (2004).
49. Maaty, W. S. et al. Characterization of the archaeal
thermophile Sulfolobus turreted icosahedral virus
validates an evolutionary link among double-stranded
DNA viruses from all domains of life. J. Virol. 80,
7625–7635 (2006).
50. Bamford, D. H. et al. Constituents of SH1, a novel
lipid-containing virus infecting the halophilic
euryarchaeon Haloarcula hispanica. J. Virol. 79,
9097–9107 (2005).
51. Porter, K. et al. SH1: A novel, spherical halovirus
isolated from an Australian hypersaline lake. Virology
335, 22–33 (2005).
52. Khayat, R. et al. Structure of an archaeal virus capsid
protein reveals a common ancestry to eukaryotic and
bacterial viruses. Proc. Natl Acad. Sci. USA 102,
18944–18949 (2005).
Shows that the major capsid protein of STIV is
highly similar in structure to capsid proteins of
some bacterial and eukaryal viruses.
53. Porter, K., Dyall-Smith, M. L. in Methods in
Microbiology — Extremophiles (eds Reiney, F. A. &
Oren, A.) 681–702 (Elsevier, London, 2006).
Reviews cultivation methods for the haloarchaea
and their viruses.
54. Gropp, F., Grampp, B., Stolt, P., Palm, P. & Zillig, W.
The immunity-conferring plasmid p phi HL from the
Halobacterium salinarium phage phiH: nucleotide
sequence and transcription. Virology 190, 45–54
(1992).
55. Schnabel, H., Palm, P., Dick, K. & Grampp, B.
Sequence analysis of the insertion element ISH1.8
and of associated structural changes in the genome
of phage phiH of the archaebacterium Halobacterium
halobium. EMBO J. 3, 1717–1722 (1984).
56. Schnabel, H. et al. Halobacterium halobium phage
phiH. EMBO J. 1, 87–92 (1982).
57. Stolt, P. & Zillig, W. Transcription of the halophage
phiH repressor gene is abolished by transcription from
an inversely oriented lytic promoter. FEBS Lett. 344,
125–128 (1994).
58. Klein, R. et al. Natrialba magadii virus phiCh1: first
complete nucleotide sequence and functional
organization of a virus infecting a haloalkaliphilic
archaeon. Mol. Microbiol. 45, 851–863 (2002).
59. Rossler, N., Klein, R., Scholz, H. & Witte, A. Inversion
within the haloalkaliphilic virus phi Ch1 DNA results in
differential expression of structural proteins. Mol.
Microbiol. 52, 413–426 (2004).
60. Witte, A. et al. Characterization of Natronobacterium
magadii phage phi Ch1, a unique archaeal phage
containing DNA and RNA. Mol. Microbiol. 23,
603–616 (1997).
61. Nuttall, S. D. & Dyall-Smith, M. L. HF1 and HF2: novel
bacteriophages of halophilic archaea. Virology 197,
678–684 (1993).
62. Nuttall, S. D. & Dyall-Smith, M. L. Halophage HF2:
genome organization and replication strategy. J. Virol.
69, 2322–2327 (1995).
63. Tang, S. L., Nuttall, S. & Dyall-Smith, M. Haloviruses
HF1 and HF2: evidence for a recent and large
recombination event. J. Bacteriol. 186, 28102817
(2004).
64. Tang, S. L. et al. HF2: a double-stranded DNA tailed
haloarchaeal virus with a mosaic genome. Mol.
Microbiol. 44, 283–296 (2002).
65. Jordan, M., Meile, L. & Leisinger, T. Organization of
Methanobacterium thermoautotrophicum
bacteriophage psi M1 DNA. Mol. Gen. Genet. 220,
161–164 (1989).
66. Meile, L., Jenal, U., Studer, D., Jordan, M., Leisinger,
T. Characterization of psiM1, a virulent phage of
Methanobacterium thermoautotrophicum Marburg.
Arch. Microbiol. 152, 105–110 (1989).
67. Pfister, P., Wasserfallen, A., Stettler, R. & Leisinger, T.
Molecular analysis of Methanobacterium phage
psiM2. Mol. Microbiol. 30, 233–244 (1998).
68. Hendrix, R. W., Smith, M. C., Burns, R. N., Ford, M. E.
& Hatfull, G. F. Evolutionary relationships among
diverse bacteriophages and prophages: all the world’s
a phage. Proc. Natl Acad. Sci. USA 96, 2192–2197
(1999).
REVIEWS
NATURE REVIEWS
|
MICROBIOLOGY VOLUME 4
|
NOVEMBER 2006
|
847
© 2006 Nature Publishing Group
69. Pedulla, M. L. et al. Origins of highly mosaic
mycobacteriophage genomes. Cell 113, 171–182
(2003).
70. Geslin, C., Le Romancer, M., Gaillard, M., Erauso, G.
& Prieur, D. Observation of virus-like particles in high
temperature enrichment cultures from deep-sea
hydrothermal vents. Res. Microbiol. 154, 303–307
(2003).
71. Stedman, K. M., Clore, A. & Combet-Blac, Y. in
SGM Symposium 66: Prokaryotic Diversity —
Mechanisms and Significance. (eds Logan, N. M.,
Lappin-Scott, H. M. & Oyston, P. C. F.) 131–144
(Cambridge University Press, 2006).
72. Prangishvili, D. in Methods in Microbiology —-
Extremophiles (eds Reiney, F. A. & Oren, A.) 331–348
(Elsevier, London, 2006).
Reviews the methods of isolating archaeal virus–
host systems from hydrothermal environments.
73. Ortmann, A. C., Wiedenheft, B., Douglas, T. & Young, M.
Hot crenarchaeal viruses reveal deep evolutionary
connections. Nature Rev. Microbiol. 4, 520–528 (2006).
74. Larson, E. T., Reiter, D., Young, M. & Lawrence, C. M.
Structure of A197 from Sulfolobus turreted
icosahedral virus: a crenarchaeal viral
glycosyltransferase exhibiting the GT-A fold. J. Virol.
80, 7636–7644 (2006).
75. Prangishvili, D. et al. Biochemical and phylogenetic
characterization of the dUTPase from the archaeal
virus SIRV. J. Biol. Chem. 273, 6024–6029 (1998).
76. Snyder, J. C. et al. Effects of culturing on the
population structure of a hyperthermophilic virus.
Microb. Ecol. 48, 561–566 (2004).
77. Mojica, F. J., Ferrer, C., Juez, G. & Rodriguez-Valera, F.
Long stretches of short tandem repeats are present in
the largest replicons of the archaea Haloferax
mediterranei and Haloferax volcanii and could be
involved in replicon partitioning. Mol. Microbiol. 17,
85–93 (1995).
78. Lillestol, R. K., Redder, P., Garrett, R. A. & Brugger, K.
A putative viral defence mechanism in archaeal cells.
Archaea 2, 59–72 (2006).
79. Mojica, F. J., Diez-Villasenor, C., Garcia-Martinez, J. &
Soria, E. Intervening sequences of regularly spaced
prokaryotic repeats derive from foreign genetic
elements. J. Mol. Evol. 60, 174–182 (2005).
80. Tang, T. H. et al. Identification of 86 candidates for
small non-messenger RNAs from the archaeon
Archaeoglobus fulgidus. Proc. Natl Acad. Sci. USA 99,
7536–7541 (2002).
81. Tang, T. H.
et al. Identification of novel non-coding
RNAs as potential antisense regulators in the
archaeon Sulfolobus solfataricus. Mol. Microbiol. 55,
469–481 (2005).
82. Jansen, R., van Embden, J. D., Gaastra, W. & Schouls,
L. M. Identification of a novel family of sequence
repeats among prokaryotes. Omics 6, 23–33 (2002).
83. Makarova, K. S., Grishin, N. V., Shabalina, S. A., Wolf,
Y. I. & Koonin, E. V. A putative RNA-interference-based
immune system in prokaryotes: computational analysis
of the predicted enzymatic machinery, functional
analogies with eukaryotic RNAi, and hypothetical
mechanisms of action. Biol. Direct 1, 7 (2006).
84. Peng, X., Kessler, A., Phan, H., Garrett, R. A. &
Prangishvili, D. Multiple variants of the archaeal DNA
rudivirus SIRV1 in a single host and a novel
mechanism of genomic variation. Mol. Microbiol. 54,
366–375 (2004).
85. Neumann, H. & Zillig, W. The TTV1-encoded viral
protein TPX: primary structure of the gene and the
protein. Nucleic Acids Res. 18, 195 (1990).
86. Ackermann, H. W. Frequency of morphological phage
descriptions in the year 2000. Brief review. Arch.
Virol. 146, 843–857 (2001).
87. Fauquet, C. M., Mayo, M. A., Maniloff, J.,
Desselberger, U. & Ball, L. A. (eds) Virus Taxonomy:
Classification and Nomenclature of Viruses (Elsevier,
Amsterdam, 2005).
88. Wommack, K. E. & Colwell, R. R. Virioplankton: viruses
in aquatic ecosystems. Microbiol. Mol. Biol. Rev. 64,
69–114 (2000).
89. Zillig, W. et al. Viruses, plasmids and other genetic
elements of thermophilic and hyperthermophilic
Archaea. FEMS Microbiol. Rev. 18, 225–236 (1996).
90. Koonin, E. V., Makarova, K. S. & Aravind, L. Horizontal
gene transfer in prokaryotes: quantification and
classification. Annu. Rev. Microbiol. 55, 709–742
(2001).
91. Wiezer, A, & Merkl, R. A comparative categorization of
gene flux in diverse microbial species. Genomics 86,
462–475 (2005).
92. Mongodin, E. F. et al. The genome of Salinibacter
ruber: convergence and gene exchange among
hyperhalophilic bacteria and archaea. Proc. Natl Acad.
Sci. USA 102, 18147–18152 (2005).
93. Bamford, D. H., Grimes, J. M. & Stuart, D. I. What
does structure tell us about virus evolution? Curr.
Opin. Struct. Biol.
15, 655–663 (2005).
Discusses the potential for structural analyses of
virion architecture and coat protein topology to
provide insights into viral evolution.
94. Brochier, C., Forterre, P. & Gribaldo, S. An emerging
phylogenetic core of Archaea: phylogenies of
transcription and translation machineries converge
following addition of new genome sequences. BMC
Evol. Biol. 5, 36 (2005).
95. Forterre, P. The origin of viruses and their possible
roles in major evolutionary transitions. Virus Res. 11 7,
5–16 (2006).
Presents a hypothesis that viruses played a crucial
role in the invention of DNA and the development
of its replication mechanisms, and in the formation
of the three domains of life.
96. Koonin, E. V. & Martin, W. On the origin of genomes
and cells within inorganic compartments. Trends
Genet. 21, 647–654 (2005).
97. Claverie, J. M. Viruses take center stage in cellular
evolution. Genome Biol. 7, 110 (2006).
98. Forterre, P. The two ages of the RNA world, and the
transition to the DNA world: a story of viruses and
cells. Biochimie 87, 793–803 (2005).
99. Forterre, P. Three RNA cells for ribosomal lineages
and three DNA viruses to replicate their genomes:
a hypothesis for the origin of cellular domain.
Proc. Natl Acad. Sci. USA 103, 3669–3674 (2006).
Acknowledgements
We would like to thank all the members of our laboratories who
have contributed to this work, and to thank colleagues who
have made their unpublished results available to us.
Competing interests statement
The authors declare no competing financial interests.
DATABASES
The following terms in this article are linked online to:
Entrez Genome: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=genome
AFV1 | His1 | His2 | PBCV1 | PSV | SH1 | SIFV | SIRV1 | STIV |
STSV1 | TTSV1
Entrez Genome Project:
http://www.ncbi.nlm.nih.gov/
entrez/query.fcgi?db=genomeprj
Methanococcus voltae | Pyrococcus abyssi | Sulfolobus
solfataricus P2
FURTHER INFORMATION
David Prangishvili’s homepage: http://www.pasteur.fr/
recherche/unites/bmge-archaea/david%20Prangishvili/pran
gishvili%20personal%20page.htm
ICTV: http://www.ncbi.nlm.nih.gov/ICTVdb/index.htm
The Sulfolobus Database: http://www.sulfolobus.org/cbin/
mutagen.pl
Access to this links box is available online.
REVIEWS
848
|
NOVEMBER 2006
|
VOLUME 4 www.nature.com/reviews/micro
© 2006 Nature Publishing Group
Reproducedwithpermissionofthecopyrightowner.Furtherreproductionprohibitedwithoutpermission.
... Genetically, the viruses of archaea are highly divergent from those of the other domains and from each other (Prangishvili et al., 2006;Dellas et al., 2014). Still, the few homologs discovered so far have provided powerful insights into the mechanisms, functions, and origins of viral systems across the domains (Birkenbihl et al., 2001;Blum et al., 2001;Peng et al., 2001;Khayat et al., 2005;Krupovic et al., 2018). ...
Article
Full-text available
Viruses have played a central role in the evolution and ecology of cellular life since it first arose. Investigations into viral molecular biology and ecological dynamics have propelled abundant progress in our understanding of living systems, including genetic inheritance, cellular signaling and trafficking, and organismal development. As well, the discovery of viral lineages that infect members of all three domains suggest that these lineages originated at the earliest stages of biological evolution. Research into these viruses is helping to elucidate the conditions under which life arose, and the dynamics that directed its early development. Archaeal viruses have only recently become a subject of intense study, but investigations have already produced intriguing and exciting results. STIV was originally discovered in Yellowstone National Park and has been the focus of concentrated research. Through this research, a viral genetic system was created, a novel lysis mechanism was discovered, and the interaction of the virus with cellular ESCRT machinery was revealed. This review will summarize the discoveries within this group of viruses and will also discuss future work.
... In addition, the term "virus" was restricted to those infecting eukaryotes, while those infecting bacteria were called "phages", mirroring the dichotomic classification of life that became prominent in the 1960s between eukaryotes and prokaryotes [5]. Despite the characterization of the actual tripartite division of life in three cellular domains (Eukarya, Archaea, Bacteria) [6], each associated with a specific portion of the virosphere [7], this nomenclature remained ...
Article
Full-text available
Our perception of viruses has been drastically evolving since the inception of the field of virology over a century ago. In particular, the discovery of giant viruses from the Nucleocytoviricota phylum marked a pivotal moment. Their previously concealed diversity and abundance unearthed an unprecedented complexity in the virus world, a complexity that called for new definitions and concepts. These giant viruses underscore the intricate interactions that unfold over time between viruses and their hosts, and are themselves suspected to have played a significant role as a driving force in the evolution of eukaryotes since the dawn of this cellular domain. Whether they possess exceptional relationships with their hosts or whether they unveil the actual depths of evolutionary connections between viruses and cells otherwise hidden in smaller viruses, the attraction giant viruses exert on the scientific community and beyond continues to grow. Yet, they still hold surprises. Indeed, the recent identification of mirusviruses connects giant viruses to herpesviruses, each belonging to distinct viral realms. This discovery substantially broadens the evolutionary landscape of Nucleocytoviricota. Undoubtedly, the years to come will reveal their share of surprises.
... Their staggering number is so great to simply overshadow all other life forms we know of, combined, making them the most abundant in all biomes and ecosystems [128] . Their biodiversity is equally great as viruses are able to infect host species throughout all major branches of the cellular tree of life, from archea to bacteria to eukarya, irrespective of their environmental niches [129] . Given their extremely variable size, morphologies, and genetic options for inheritance across generations, including adaptation and evolvability, viruses can strongly impact hosts affecting all levels of their biological organization, from host-genome composition to internal ecosystem function [130] . ...
... This allows the possibility to genetically modify them to express peptides or proteins on their double-stranded or single-double-stranded genome and the shape of their capsid protein. Despite the great variation, 95% of them are non-enveloped, dsDNA, tailed phages [30]. In the majority of vaccine development cases, the capsid of dsDNA-tailed phages is used to express a viral antigen or a portion of an antigen, most typically via non-covalent linkages [31]. ...
Article
Full-text available
Phages are highly ubiquitous biological agents, which means they are ideal tools for molecular biology and recombinant DNA technology. The development of a phage display technology was a turning point in the design of phage-based vaccines. Phages are now recognized as universal adjuvant-free nanovaccine platforms. Phages are well-suited for vaccine design owing to their high stability in harsh conditions and simple and inexpensive large-scale production. The aim of this review is to summarize the overall breadth of the antiviral therapeutic perspective of phages contributing to the development of phage-based vaccines for COVID-19. We show that phage vaccines induce a strong and specific humoral response by targeted phage particles carrying the epitopes of SARS-CoV-2. Further, the engineering of the T4 bacteriophage by CRISPR (clustered regularly in-terspaced short palindromic repeats) presents phage vaccines as a valuable platform with potential capabilities of genetic plasticity, intrinsic immunogenicity, and stability.
... In particular, when considering capsids from viruses that infect hyperthermophilic Archaea members, considerable attention has been paid to their extremely high stability and morphological diversity (Uldahl and Peng 2013). Viruses infecting hyperthermophilic Archaea members have remarkable diversity in terms of virion structures, which, along with icosahedral and filamentous particles, include virions resembling bottles, spindles, droplets, and coils, which are not observed among eukaryotic viruses and bacteriophages (Prangishvili et al. 2006). In addition, these viruses exhibit high stability against extreme conditions, such as high temperature and acidic environments, similar to the natural habitat of hyperthermophilic archaeal hosts. ...
Article
Full-text available
Virus capsid proteins have various applications in diverse fields such as biotechnology, electronics, and medicine. In this study, the major capsid protein of bacilliform clavavitus APBV1, which infects the hyperthermophilic archaeon Aeropyrum pernix, was successfully expressed in Escherichia coli. The gene product was expressed as a histidine-tagged protein in E. coli and purified to homogeneity using single-step nickel affinity chromatography. The purified recombinant protein self-assembled to form bacilliform virus-like particles at room temperature. The particles exhibited tolerance against high concentrations of organic solvents and protein denaturants. In addition, we succeeded in fabricating functional nanoparticles with amine functional groups on the surface of ORF6-81 nanoparticles. These robust protein nanoparticles can potentially be used as a scaffold in nanotechnological applications.
... van The wide spectrum of archaeal cell shapes and sizes schematically visualized in scale. Representative microscopy images are reprinted from several studies on archaea including: Archaeal viruses 115 ...
Article
Full-text available
The past decade has revealed the diversity and ubiquity of archaea in nature, with a growing number of studies highlighting their importance in ecology, biotechnology and even human health. Myriad lineages have been discovered, which expanded the phylogenetic breadth of archaea and revealed their central role in the evolutionary origins of eukaryotes. These discoveries, coupled with advances that enable the culturing and live imaging of archaeal cells under extreme environments, have underpinned a better understanding of their biology. In this Review we focus on the shape, internal organization and surface structures that are characteristic of archaeal cells as well as membrane remodelling, cell growth and division. We also highlight some of the technical challenges faced and discuss how new and improved technologies will help address many of the key unanswered questions.
... These viruses can be divided into eukaryotic viruses, which mostly infect human cells; prokaryotic viruses, otherwise known as bacteriophages or phages (Zarate et al. 2018); and archaeal viruses (Matijašić et al. 2020) (Table 1). Most archaeal viruses isolated so far have been doublestranded DNA (dsDNA) viruses (Prangishvili, Forterre and Garrett 2006). The interactions between archaeal viruses and their hosts as well as their impact on human health are still largely unexplored (Matijašić et al. 2020), and thus, will not be further discussed in this review. ...
Article
Full-text available
The human gut virome is comprised of diverse commensal and pathogenic viruses. The colonization by these viruses begins right after birth through vaginal delivery, then continues through breastfeeding, and broader environmental exposure. Their constant interaction with their bacterial hosts in the body shapes not only our microbiomes but us. In addition, these viruses interact with the immune cells, trigger a broad range of immune responses, and influence different metabolic pathways. Besides its key role in regulating the human gut homeostasis, the intestinal virome contributes to disease development in distant organs, both directly and indirectly. In this review, we will describe the changes in the gut virome through life, health, and disease, followed by discussing the interactions between the virome, the microbiome, and the human host as well as providing an overview of their contribution to gut disease and disease of distant organs.
... By contrast, the CK culture displayed long-lasting slow growth kinetics and continued to exhibit high-titer virus particles. The DNA content distributions of the CK culture were also similar to those before the transfer, indicating that SMV1 maintains a stable infectious status in the cells instead of inducing cell lysis, in line with previous studies on SMV1 and other Sulfolobus viruses (Guo et al., 2019;Liu et al., 2021a;Prangishvili et al., 2006;Uldahl et al., 2016). Further, if the initial viral titer is much lower than the cell number, an Abi response would not arrest culture growth. ...
Article
Argonaute (Ago) proteins are widespread nucleic-acid-guided enzymes that recognize targets through complementary base pairing. Although, in eukaryotes, Agos are involved in RNA silencing, the functions of prokaryotic Agos (pAgos) remain largely unknown. In particular, a clade of truncated and catalytically inactive pAgos (short pAgos) lacks characterization. Here, we reveal that a short pAgo protein in the archaeon Sulfolobus islandicus, together with its two genetically associated proteins, Aga1 and Aga2, provide robust antiviral protection via abortive infection. Aga2 is a toxic transmembrane effector that binds anionic phospholipids via a basic pocket, resulting in membrane depolarization and cell killing. Ago and Aga1 form a stable complex that exhibits nucleic-acid-directed nucleic-acid-recognition ability and directly interacts with Aga2, pointing to an immune sensing mechanism. Together, our results highlight the cooperation between pAgos and their widespread associated proteins, suggesting an uncharted diversity of pAgo-derived immune systems.
Chapter
Thirty thousand years back, when humans were still primitives, the viruses were more life-like; they could regenerate themselves.
Article
Full-text available
We estimated prokaryotic mortality due to viruses and bacterivores through salinity gradients in 2 solar salterns. In each saltern system, successive ponds provided steady state environments with a range of salinities from 37 to 372‰. Prokaryotic and viral abundance increased with salinity, reaching about 108 prokaryotic cells ml-1 and 109 virus-like particles (VLP) ml-1 at salinities higher than 250‰. Prokaryotic doubling times became longer than 2 d above 250‰ salinity until the end of the gradient. Bacterivory accounted for all the production at lower salinities but it was found to be zero at the highest salinities. The percentage of visibly infected cells was not different among the ponds where infected cells could be detected and it was always lower than 4%. From the percentage of infected cells and using conversion factors from the literature we estimated rates of prokaryotic mortality due to viral lysis about 0.6 to 2 × 106 prokaryotes ml-1 were lysed daily by the viruses in the salterns. This number represented a low percentage of prokaryotic abundance and production compared to the prokaryotic losses due to bacterivores (0.2 to 4 × 107 bacteria ml-1 d-1). However, viral production reached values higher than 108 VLP ml-1 d-1 above 250‰ salinity, due to the large burst size (200 viruses cell-1) found in a particular morphotype of prokaryotes, the sguare archaea. These archaea represented more than 25% of the prokaryotic assemblage above 250‰ salinity. At this point they became the prokaryotic morphotype with the largest percentage of infected cells (1 to 10% of square archaea with visible phages inside). A lemon-shaped virus (similar to one described for some other groups of archaea) was found infecting square archaea, its abundance increased in the saltiest ponds together with that of the square archaea. In this system viruses did not exert a strong control over the prokaryotic abundance and growth rate.
Article
A virus infecting the hyperthermophilic archaeon Sulfolobus tengchongensis has been isolated from a field sample from Tengchong, China, and characterized. The virus, denoted STSV1 (Sulfolobas tengchongensis spindle-shaped virus 1), has the morphology of a spindle (230 by 107 nm) with a tail of variable length (68 nm on average) at one end and is the largest of the known spindle-shaped viruses. After infecting its host, the virus multiplied rapidly to high titers (> 10(10) PFU/ml). Replication of the virus retarded host growth but did not cause lysis of the host cells. STSV1 did not integrate into the host chromosome and existed in a carrier state. The STSV1 DNA was modified in an unusual fashion, presumably by virally encoded modification systems. STSV1 harbors a double-stranded DNA genome of 75,294 bp, which shares no significant sequence similarity to those of fuselloviruses. The viral genome contains a total of 74 open reading frames (ORFs), among which 14 have a putative function. Five ORFs encode viral structural proteins, including a putative coat protein of high abundance. The products of the other nine ORFs are probably involved in polysaccharide biosynthesis, nucleotide metabolism, and DNA modification. The viral genome divides into two nearly equal halves of opposite gene orientation. This observation as well as a GC-skew analysis point to the presence of a putative viral origin of replication in the 1.4-kb intergenic region between ORF1 and ORF74. Both morphological and genomic features identify STSV1 as a novel virus infecting the genus Sulfolobus.
Article
A novel archaeal virus, His1, was isolated from hypersaline waters in southeastern Australia. It was lytic, grew only on Haloarcula hispanica (titers of up to 10(11) PFU/ml), and displayed a lemon-shaped morphology (74 by 44 nm) previously reported only for a virus of the extreme thermophiles (SSV1). The density of His1 was approximately 1.28 g/ml. similar to that of SSV1 (1.24 g/ml). Purified particles were resistant to low salt concentrations. The genome was linear, double-stranded DNA of 14.9 kb, similar to the genome of SSV1 (15.5 kb), Morphologically, this isolate clearly belongs to the recently proposed Fuselloviridae family of archaeal viruses. It is the first member of this family from the extremely halophilic archaea, and its host, H. hispanica, can be readily manipulated genetically.
Article
We have sampled acidic springs, water and mud holes of 14 major solfataric fields of Iceland in order to isolate both heterotrophic and autotrophic members of the order Sulfolobales and to find affiliated genetic elements e.g. plasmids and viruses. The diverstiy of 120 isolates was analysed by comparing DNA restriction fragment patterns. The 44 heterotrophic isolates belonged to only two types, the solfataricus (S) type and the islandicus (I) type as judged by the restriction patterns of their DNAs. None of the heterotrophic isolates was able to grow autotrophically. The more than 70 autotrophic isolates belonged to only three types, the Desulfurolobus (D) type, the closely related A type and, rarely, the B type. None of the autotrophic isolates was able to grow heterotrophically. Primary heterotrophic colonies often gave rise to obligately autotrophic isolates when submitted to autotrophic selection, probably because they constituted synthrophic associatons. Inversely, primary autotrophic colonies sometimes yielded obligately heterotrophic isolates when submitted to heterotrophic selection.
Article
A novel archaeal bacteriophage, Phi Ch1, was isolated from a haloalkalophilic archaeon Natronobacterium magadii upon spontaneous lysis. The phage-cured strain N. magadii(L13) was used to demonstrate infectivity of phage Phi Ch1. The turbid-plaque morphology and the fact that N. magadii cells isolated from plaques were able to produce phage indicated that Phi Ch1 is a temperate phage. The phage morphology resembles other members of Myoviridae-infecting Halobacterium species. In solution below 2 M NaCl, the phage lost its morphological stability and infectivity. One- and two-dimensional SDS-PAGE of phage particles revealed at least four major and five minor proteins with molecular masses ranging from 15 to 80 kDa and acidic isoelectric points. Southern blot analysis of chromosomal DNA of a lysogenic N. magadii strain showed that Phi Chl exists as a chromosomally integrated prophage. The phage particles contain both double-stranded, linear DNA (approx. 55 kbp) as well as several RNA species (80-700 nucleotides). Hybridization of labelled RNA fragments to total DNA from N. magadii and Phi Chl showed that the virion-associated RNA is host encoded. Part of the phage DNA population is modified and restriction analysis revealed evidence for adenine methylation. Phage Phi Chl is the first virus described for the genus Natronobacterium, and the first phage containing DNA and RNA in mature phage particles.
Article
Over 5100 bacterial viruses have been examined in the electron microscope since 1959. About 4950 phages (96%) are tailed and only 186 phages (3.6%), are cubic, filamentous, or pleomorphic. Phages belong to 13 virus families and occur in over 140 bacterial genera. Phages are listed by morphotypes and host genera. Siphoviridae or phages with long, noncontractile tails comprise 61% of tailed phages. The distribution of phages in different bacterial phylogenetic divisions is shown.
Article
The methanogenic archaeon Methanobacterium thermoautotrophicum Marburg is infected by the double-stranded DNA phage ΨM2. The complete phage genome sequence of 26 111 bp was established. Thirty-one open reading frames (orfs), all of them organized in the same direction of transcription, were identified. On the basis of comparison of the deduced amino acid sequences to known proteins and by searching for conserved motifs, putative functions were assigned to the products of six orfs. These included three proteins involved in packaging DNA into the capsid, two putative phage structural proteins and a protein related to the Int family of site-specific recombinases. Analysis of the N-terminal amino acid sequences of three phage-encoded proteins led to the identification of two genes encoding structural proteins and of peiP, the structural gene of pseudomurein endoisopeptidase. This enzyme is involved in the lysis of host cells, and it appears to belong to a novel enzyme family. peiP was overexpressed in Escherichia coli, and its product was shown to catalyse the in vitro lysis of M. thermoautotrophicum cells. Comparison of the phage ΨM2 DNA sequence with parts of the sequence of the wild-type phage ΨM1 suggests that ΨM2 is a deletion derivative, which formed by homologous recombination between two copies of a direct repeat.
Chapter
This chapter discusses available information on transcription in archaea as compared to bacteria and eucarya. Immunochemical analysis and later sequencing of cloned genes revealed similarities between archaeal, bacterial and eucaryal nuclear ribo nucleic acid polymerase (RNAP) components. Several sequences of genes of large components of bacterial RNAPs and of each of the three eucaryal RNAPs are available for sequence comparison with the archaeal homologs. The similarity between the archaeal and these eucaryal sequences is indeed often higher than that between components of the two different eucaryal RNAPs themselves. The striking deviation of sequences of many archaeal and bacterial genes seems inconsistent with the correspondence between archaea and bacteria in genome size and gene stock and in the composition and order of genes in gene clusters. The strongly contrasting apparent disorganization of eucaryal genomes could have resulted from a different mode of evolution initiated by the fusion—namely, by inventive addition rather than economized adaptation to a changing environment.
Article
The double-stranded (ds)DNA virus φCh1 infects the haloalkaliphilic archaeon Natrialba magadii. The complete DNA sequence of 58 498 bp of the temperate virus was established, and the probable functions of 21 of 98 φCh1-encoded open reading frames (ORFs) have been assigned. This knowledge has been used to propose functional modules each required for specific functions during virus development. The φCh1 DNA is terminally redundant and circularly permuted and therefore appears to be packaged by the so-called headful mechanism. The presence of ORFs encoding homologues of proteins involved in plasmid replication as well as experimental evidence indicate a plasmid-mediated replication strategy of the virus. Results from nanosequencing of virion components suggest covalent cross-linking of monomers of at least one of the structural proteins during virus maturation. A comparison of the φCh1 genome with the partly sequenced genome of Halobacterium salinarum virus φH revealed a close relationship between the two viruses, although their host organisms live in distinct environments with respect to the different pH values required for growth.