ArticlePDF AvailableLiterature Review

Phage single-gene lysis: Finding the weak spot in the bacterial cell wall

Authors:
  • Armata Pharmaceuticals

Abstract and Figures

In general, the last step in the vegetative cycle of bacterial viruses, or bacteriophages, is lysis of the host. Double-stranded DNA phages require multiple lysis proteins, including at least one enzyme that degrades the cell wall (peptidoglycan). In contrast, the lytic ssDNA and ssRNA phages have a single lysis protein that achieves cell lysis without enzymatically degrading the peptidoglycan (PG). Here, we review four “Single-gene lysis” or Sgl proteins. Three of the Sgls block bacterial cell wall synthesis by binding to and inhibiting several enzymes in the PG precursor pathway. The target of the fourth Sgl, L from bacteriophage MS2, is still unknown, but we review evidence indicating that it is likely a protein involved in maintaining cell wall integrity. Even though only a few phage genomes are available to date, the ssRNA Leviviridae are a rich source of novel Sgls, which may facilitate further unraveling of bacterial cell wall biosynthesis and discovering new antibacterial agents.
Content may be subject to copyright.
1
Phage single-gene lysis: Finding the weak spot in the bacterial cell wall
Karthik Chamakura1,2 and Ry Young1,2*
From the 1Department of Biochemistry and Biophysics, Texas A&M University, College Station TX
77843-2128; 2Center for Phage Technology, Texas A&M AgriLife Research, Texas A&M University
College Station TX 77843-2128
Running title: Phage Sgl proteins target peptidoglycan biosynthesis
*To whom correspondence should be addressed: Ry Young: Department of Biochemistry and
Biophysics, Texas A&M University, College Station TX 77843-2128; ryland@tamu.edu; Tel.(979)845-
2087.
Keywords: Single-gene lysis, phage, peptidoglycan biosynthesis, MurA, MraY, MurJ, flippase,
autolysin, chaperone, antibiotic
_____________________________________________________________________________________
ABSTRACT
In general, the last step in the vegetative cycle of
bacterial viruses, or bacteriophages, is lysis of the
host. Double-stranded DNA phages require
multiple lysis proteins, including at least one
enzyme that degrades the cell wall (peptidoglycan).
In contrast, the lytic ssDNA and ssRNA phages
have a single lysis protein that achieves cell lysis
without enzymatically degrading the peptidoglycan
(PG). Here, we review four “Single-gene lysis” or
Sgl proteins. Three of the Sgls block bacterial cell
wall synthesis by binding to and inhibiting several
enzymes in the PG precursor pathway. The target
of the fourth Sgl, L from bacteriophage MS2, is still
unknown, but we review evidence indicating that it
is likely a protein involved in maintaining cell wall
integrity. Even though only a few phage genomes
are available to date, the ssRNA Leviviridae are a
rich source of novel Sgls, which may facilitate
further unraveling of bacterial cell wall
biosynthesis and discovering new antibacterial
agents.
________________________________________
INTRODUCTION TO SMALL LYTIC
PHAGES AND “SINGLE-GENE LYSIS”
By definition, the lytic bacteriophages
encode proteins for disruption of the host envelope.
The large dsDNA phages, the Caudovirales, have
multiple lysis proteins, including holins,
endolysins, and spanins, targeting the cytoplasmic
or inner membrane (IM), peptidoglycan (PG) and
outer membrane (OM), respectively, as well as
multiple proteins that regulate the lysis process
(1,2). In contrast, the small lytic phages of Gram-
negative hosts, comprising the ssDNA
(Microviridae) and ssRNA (Leviviridae), achieve
host lysis by a single gene, encoding a protein
lacking any PG degrading activity (3). This review
exclusively focuses on these Single-Gene Lysis
(Sgl) proteins of small lytic phages.
Bacterial cell wall structure and biosynthesis
An exploration of Sgl mechanisms requires
a brief review of the structure of the Gram-negative
cell wall and its biosynthesis. The key to the
structure and shape of the cell envelope is the PG
layer, consisting of 2-3 layers of glycan strands
made up of repeating disaccharide units of
MurNAc-pentapeptide and GlcNAc, cross-linked
by peptide bridges between pentapeptide side
chains of MurNAcs of adjacent strands (Fig. 1a) (4-
6). The PG has considerable tensile strength (3-300
MPa), allowing the cell to tolerate high internal
osmotic pressures (3-10 atm) while maintaining
shape (7, 8). The entire PG of a cell can be isolated
as a single complex polymer, the sacculus, studies
of which have revealed that the glycan chains run
almost perpendicular to the long axis of the cell (9).
In most Gram-negative bacteria, the PG layer is
covalently linked through >105 peptide linkages to
the C-terminal Lys residue of the major lipoprotein,
Lpp, the PG-linked Lpp is anchored almost
exclusively in the inner leaflet of OM (5, 10, 11).
Biosynthesis of the PG can be divided into
cytoplasmic, membrane, and periplasmic stages.
There are 7 enzymatic steps in the cytoplasm,
http://www.jbc.org/cgi/doi/10.1074/jbc.TM118.001773The latest version is at
JBC Papers in Press. Published on November 12, 2018 as Manuscript TM118.001773
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
2
beginning with the transfer of an enolpyruvyl
moiety from PEP to UDP-GlcNAc, catalyzed by
MurA (Supp. Fig. 1) (12). After reduction of the
enolpyruvyl moiety by MurB to create UDP-
MurNAc, the next five enzymes are involved in
adding amino acids that form the pentapeptide [L-
Ala D-Glu m-Dap D-Ala D-Ala] to the lactyl group,
resulting in the final soluble intermediate UDP-
MurNAc-pep5 (Fig. 1a) (5, 13). The first
membrane-linked step in PG synthesis begins with
the transfer of this sugar nucleotide pentapeptide to
the lipid carrier undecaprenyl phosphate (C55-P or
UndP). This reaction is catalyzed on the
cytoplasmic face of the IM by the integral
membrane protein MraY to generate a
monosaccharide-lipid compound, Lipid I (Fig. 1a
and Supp. Fig. 2a). MurG then catalyzes the
addition of a second sugar moiety (UDP-GlcNAc),
resulting in the final precursor, Lipid II (Fig. 1a).
The last step of the membrane phase is the flipping
of Lipid II so that its disaccharide pentapeptide
moiety is on the periplasmic face of the membrane.
The enzyme “flippase” that effects this transfer has
been controversial until very recently. Although
FtsW was shown to flip lipid II in vitro (14,15),
several lines of evidence now support MurJ as
being the essential lipid II flippase (16-21). The
extracellular steps of PG biosynthesis utilize the
energy stored in the phosphodiester-muramic acid
bond of the flipped Lipid II and in the D-Ala-D-Ala
peptide bond to drive the glycosyl-transferase (GT)
and cross-linking reactions respectively (13). These
steps are carried out by mono- or bi- functional
penicillin binding proteins (PBPs) and recently,
RodA, a member of the SEDS superfamily, was
shown to catalyze GT reactions (22-24).
THE FIRST SGL: PROTEIN E FROM
MICROVIRUS PHIX174
Famous phage, famous gene
X174 is the founding member and genetic
paradigm of the Microviridae, which are nearly as
widespread as the Caudovirales (25). It was the
first gDNA to be completely sequenced and also to
be synthesized in vitro (26, 27). The 10 genes
include three that are embedded out-of-frame in
essential cistrons (Fig. 1b). One of these embedded
genes is E, encoding the Sgl protein in the +1
reading frame of the essential scaffolding gene D
(Fig 1b). E is famous not only for being the first
embedded gene to be discovered but also the first
gene to be subjected to site-directed mutagenesis
(26, 28). More important here is the fact that it is
the only DNA-virus Sgl and, it was the first Sgl
gene for which the lytic mechanism was firmly
established. The methods for working out its
functional pathway have been replicated for all the
other Sgl systems and thus will be reviewed here in
some detail.
Genetics clarifies E function
The E gene was cloned into a medium copy
expression plasmid and shown to support lysis after
induction (29, 30). Despite this early focus and the
availability of the cloned E gene, the lysis
mechanism of E remained controversial for two
decades (3). Early transmission electron
microscopy studies showed that cells infected with
X174 lysed as a result of septal blebs in dividing
cells, generating a morphology that was remarkably
similar to penicillin-mediated lysis (31, 32).
Reproduction of this morphology after induction of
the cloned E led to the general model that E
interfered with PG biosynthesis (30, 33). However,
based on physiological and scanning-EM studies,
other groups proposed that E functioned either by
the activation of unspecified autolytic functions
(34-36) or by the formation of polymeric
“transmembrane tunnels” that opened the
cytoplasm directly to the medium (37). This
profusion of models painted a confusing picture for
the mechanism of E lysis, primarily because all
lacked genetic evidence.
Mutational and deletion analysis of E
revealed that the lytic function requires only the
first 34 residues; lytic function was retained without
the C-terminal 57 residues, a highly basic, Pro-rich
segment, as long it was replaced by a stable
cytoplasmic domain, even -galactosidase (38, 39).
The first genetic approach to Sgl function, to be
repeated successfully for several other Sgl systems,
was done by selecting spontaneous host mutants
resistant to plasmid-borne expression of E (38).
The first E-resistance mutations turned out to be in
a single locus, designated as slyD (“sensitivity to
lysis”), which encodes a FKBP-type PPIase
(peptidylprolyl-cis-trans-isomerase) (40). Indeed,
X174 infections of slyD knockout mutants
proceed normally in every respect except lysis
never occurs and virions hyper-accumulate.
Purified SlyD was shown to accelerate folding of
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
3
proteins limited by cis-trans peptidylprolyl
isomerization steps but its function in vivo is not
known (41). However, E was found to be highly
unstable in a slyD background, suggesting that
SlyD has a chaperone role, probably in folding the
C-terminal domain (CTD), which is rich in Pro
residues (42).
To continue pursuit of the E target, bypass
suppressor mutations were isolated as rare plaque-
formers on a slyD lawn (42). These Epos (plates on
slyD) were found to be missense alleles at the N-
terminus of E, each of which resulted in a ~10X
increase in the biosynthesis rate, thereby
compensating for the proteolytic instability of E.
This proved to be a key technological advance.
Spontaneous “Eps” host mutants (E-pos sensitivity)
that were resistant to induction of the plasmid-
borne Epos allele were selected and cross-streaked
for sensitivity to the phage. Of ~2000 survivors, all
but three retained X174-sensitivity and
presumably were defective in E expression or
plasmid copy number. Subsequent genetic
mapping and sequencing revealed that the
mutations mapped to residues (P172 and F288) in
TMDs 5 and 9 of MraY (Fig. 3). Labelling
experiments with [3H]mDAP showed that E
blocked cell wall synthesis ~20 min before lysis,
and TLC chromatography revealed depletion of
lipid-linked label and the accumulation of UDP-
GlcNAc, confirming the inhibition of MraY (43).
Based on the mutational data and membrane
localization of both proteins, Bernhardt and
colleagues proposed that E interacted with MraY
through TMD-TMD interactions that were
disrupted in the mutant alleles (Supp. Fig. 2b) (43,
44).
in vitro analysis of E-mediated inhibition of
MraY
Initial attempts to over-express full-length
E failed due to its inherent lethality, but by doing
inductions in the presence of the heterologous
MraY from B. subtilis, a His-tagged full-length E
protein was purified with a yield of 27 g of EHis6
per liter of culture (45). Using this as an
immunoblot standard, these workers determined
that E was produced at ~500 molecules/cell at the
time of lysis, in agreement with estimates from
radiolabeling and E-LacZ enzyme assays (38, 46).
An assay based on UDP-MurNAc-pentapeptide
DNS, a fluorescent analogue of UDP-MurNAc-
pentapeptide, and phytol-P, a 20 carbon analogue
of C55-P, was developed to determine kinetic
parameters in the presence and absence of E. For
both substrates, the addition of purified E had no
effect on the apparent Km but reduced the Vmax,
indicating that E is a noncompetitive inhibitor of
MraY (45). This was further supported by the
observation that overexpression of either wild-type
or the catalytically inactive allele of mraYD267N
protects cells from E-mediated lysis, presumably by
titrating out E (43, 47).
The E-MraY interaction
Some details of the interaction between E
and MraY have been determined by assessing lysis
kinetics in bulk culture in the context of mutant E
alleles, both from selections and site-directed
mutagenesis, and both wt and catalytically inactive
MraY (43-45, 47). Despite near saturating
mutagenesis, mutations in only five mraY positions
yielded E-insensitive phenotypes: one in TMD5
(G186S), two in TMD9 (F288L and V291M) and
two (P170L; L172) the periplasmic loop above
TMD5 (47). By testing these alleles in context of a
catalytically inactive MraYD267N for the ability to
protect against E lysis in an mraY+/mraYD267N
merodiploid, the alleles were grouped into 3 classes
of apparent affinity for E in vivo, with G186S and
V291M ranking with the wt at the highest affinity,
F288L and the B. subtilis version of MraY the
lowest, and the P170L and P172 alleles
intermediate (47). This classification was reflected
in the in vitro inhibition assays (45). Based on the
crystal structure of MraY from Aquifex aeolicus,
the TMD5 and TMD9 sites are clustered on the
same outside face of the molecule within the
periplasmic leaflet of the IM (Fig. 2 ) (48). TMD9
is split into two helical fragments: the N-terminal
9a, within which both E-insensitivity mutations
map, and the C-terminal 9b, which does not pack in
the helical bundle but is splayed out nearly laterally
in the bilayer. The loop mutations map to a beta
hairpin structure, with one directly above TMD5
(L172) and the other just above TMD9 (P170).
Assuming these mutations reflect critical contacts
in the E binding site, the spatial arrangement
indicates that the N-terminal segment of the TMD
of E makes those contacts. This notion is consistent
with the results from the Clemons group, who
constructed a large set of alanine and leucine
substitutions as well as truncation alleles, focused
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
4
on the TMD of E (49). This collection was analyzed
by induction in vivo, monitoring the kinetics of
lysis and the accumulation of the E protein in the
membrane fraction. Mutations that affected lysis
without affecting accumulation mapped to a single
face of the TMD, suggesting that this face interacts
with the two TMDs of MraY that housed the E-
resistant changes. Also, the TMD as a whole
appears to be relatively insensitive to changes in its
C-terminal segment, which can be replaced by a
homopolymer of Leu residues. Moreover, Phe
substitutions at Leu19 and Leu23, both of which are
on the same helix face as the inactivating mutations
and one of which (L19F) corresponds to an original
Epos mutation, enhance lytic function without
enhancing membrane insertion. Using IMAC-
chromatography and Western blotting, complexes
of E and MraY could be demonstrated. However,
the efficiency of MraY recovery did not correlate
well with lytic function of the E allele used,
indicating that simple binding of E to MraY is not
sufficient for inhibition. Overall, the details of how
E binds MraY and interferes with Lipid I synthesis
will likely require a structure of the E-MraY
complex.
OVERVIEW OF SGL GENES IN SSRNA
PHAGES
The Sgl proteins undoubtedly evolved
because of the extremely constrained size of the
phage genome; compared to the Microviridae, this
constraint is even worse for the ssRNA phages (the
Leviviridae, or the leviviruses), which have ~4 kb
gRNAs and only three core genes (Fig. 1b).
Lacking tail structures, leviviruses exploit
retractable pili to initiate infection. By far the best
studied leviviruses are MS2 and Q, both specific
for the canonical F conjugation pilus; many other
F-specific leviviruses that are related to these two
paradigms have been studied (50-55). However,
seven other distinct leviviruses are known, each
targeting a different retractable pilus (Supp. Table
1) (56-61). The Sgl genes have been identified in
eight of the nine distinct leviviruses by showing
that, like for E, induction of a plasmid-borne clone
is necessary and sufficient for lysis. Importantly,
all cause disruption of the cell wall by accessing
different cellular targets, suggesting that in each
case, a new Sgl was evolved after radiation to a new
retractable pilus, no doubt facilitated by the
extremely high mutation rate of the RNA-
dependent replicase (62). This means that even
with the low total genomic database of less than
50,000 bases of unique Leviviridae genomes, there
are multiple Sgl systems which might be exploited
for probing the biosynthesis and dynamic
homeostasis of the cell wall. In the following, the
Sgl systems of Q, M and MS2 will be reviewed.
The order is not chronological but makes sense in
that the targets of the first two Sgl proteins have
been identified, whereas the MS2 Sgl system is an
enduring mystery.
THE “PROTEIN ANTIBIOTIC”: A2
FROM QBETA
Finding rats
The A2 protein has multiple functions
during Q infection: it functions in virion assembly
(63), is bound to and provides protection for the
gRNA against RNase degradation (64), mediates
interaction with the F-pilus, and is internalized into
the host cytoplasm along with the genomic RNA
(63-65). Remarkably, A2 also functions as the Sgl
protein. The lytic activity of A2 was first
demonstrated by Winter and Gold in 1983, who
showed that induction of A2 cloned on a medium-
copy plasmid is necessary and sufficient to cause
lysis (66).
To identify the target of A2, the same method was
used as for X174 E, selecting for host mutants that
survived induction of a plasmid-borne A2 gene,
followed by screening survivors by cross-streaking
with Q phage and the mutants that passed
selection/screen were designated as rat (resistant to
A-two) mutants (67). Genetics and sequence
analyses revealed a single missense change in
murA, L138Q. As with E, the incorporation of [3H]
mDAP label into PG is blocked at least 20 min
before the onset of lysis in cells induced for A2.
Biochemical analysis of the sugar nucleotide pool
from A2-inhibited cells revealed that UDP-GlcNAc
was elevated, confirming MurA as the target. In
vitro inhibition of MurA by purified A2 could not
be demonstrated, mainly because overexpressed A2
was insoluble. However, in what seems to be the
only instance of using virions for enzyme
inhibition, it was shown that the catalytic activity of
MurA, but not MurAL138Q, in crude extracts could
be blocked by the addition of highly purified Q
virions. Later experiments done with purified
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
5
MurA and Q particles confirmed these results
(68). Based on the turnover number of MurA and
the number of purified virions needed to block its
enzymatic activity, the MurA-A2 dissociation
constant was determined to be ~10 nM (67).
The A2-MurA interaction
In addition to in vitro inhibition assays,
direct protein-protein interaction between A2-MurA
and A2 -MurAL138Q was also demonstrated by yeast-
two-hybrid analysis, with the latter pair displaying
a weaker signal, suggesting that the mutant allele
weakens A2 binding (68). To probe the interaction
of MurA with A2 in vitro, the well-characterized
catalytic pathway of MurA was exploited. The
MurA reaction is well ordered, with UDP-GlcNAc
binding in the catalytic cleft associated with a
dramatic shift from open to closed conformation,
which then allows PEP binding and catalysis.
Interaction studies were done using a soluble MBP-
A2 fusion protein and various forms of MurA,
including the original rat allele, MurAL138Q, and
MurAD305A, which is disabled for catalysis but not
substrate binding, and with various combinations of
the substrates and the suicide inhibitor fosfomycin.
The results clearly showed that A2 preferentially
binds to the UDP-GlcNAc-liganded, closed form
MurA, preventing PEP from binding. Details of the
binding surface were obtained by site-directed
substitutions of amino acids in the area around
L138, yielding a cluster of new rat alleles that
blocked Q plaque formation and clearly defined
an interaction surface surrounding the catalytic
loop, including the catalytic domain, CTD and the
catalytic loop (Supp. Fig. 3).
A2 differs significantly from the Mat
proteins of MS2 and related Leviviridae, especially
in the N-terminus; deletion analysis confirmed that
the lytic function is fully defined in the first 180
residues. To map the interaction domain, A2por
(plates on rat) suppressor alleles were isolated and
mapped to three positions (L28, D52 and E125) in
the N-terminal domain (69). However, none of the
por alleles were lytic when cloned and induced in
the rat1 host. Immunoblot analysis revealed that
A2por mutant levels increased much more rapidly
than the parental A2 during infection, resulting in
early lysis and reduced yield of progeny virions in
the wt host. Inspection of the sequence around the
por sites confirmed that the mutations disrupt
significant RNA structures that repress translational
initiation in the viral RNA, thus bypassing the
reduced A2-MurAL138Q affinity by increasing the
quantity of the phage protein.
The interaction interface was recently
resolved in asymmetric cryo-EM structures of Q
particles in complex with UDP-NAG-liganded
MurA or fosfomycin-liganded MurA (63) (Fig.
3ab). The cryo-EM structures validated the
interaction interface on MurA inferred from the
various rat alleles and also confirmed that the NTD
of A2 is in contact with MurA (Fig. 3c, d, and e).
LYSM: NEW TARGET AND SETTLING A
DEBATE
The lysis gene (lysM) of phage M has
evolved completely embedded in the +1 reading
frame of the rep gene and it encodes a 37 amino
acid protein with a single TMD (58). The functional
LysM-eGFP fusion suggests an N-out and C-in
membrane topology (70). Early insights into the
molecular mechanism of LysM lethality came from
the observations that lysis proceeded through septal
catastrophes, like A2 and E, suggesting that LysM
might be an inhibitor of cell wall biosynthesis (31,
67, 70). The identity of MurJ as the molecular target
of LysM was revealed in multi-copy suppression
experiments, where plasmids carrying murJ in
random fragments of E. coli genome suppressed
lysM lethality. Furthermore, isolation of 9
spontaneous lysM- resistant mutants that mapped to
two of the fourteen TMDs of MurJ suggested a
possible interaction interface and a plausible
mechanism (Fig. 4). Additionally, it was shown that
LysM was specific to MurJ and the cells can be
rescued from LysM lethality by the expression of
heterologous lipid II flippase Amj from B. subtilis.
To address the conformational state in
which LysM binds to MurJ, a substituted-cysteine
accessibility method (SCAM) was used (70, 71). In
the presence of LysM, SCAM analysis showed that
5 TMD positions showed altered SCAM patterns,
which suggested that LysM binding locks MurJ into
one of the two conformations proposed to constitute
the lipid II-flipping cycle. The SCAM labeling
pattern is consistent with MurJ being locked in a
“periplasmic open” conformation, which would
lead to an accumulation of lipid-linked PG
precursors in the inner leaflet of IM and a
corresponding decrease on the periplasmic side.
Both predictions were confirmed by in vivo
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
6
flippase assays, strongly indicating that LysM
blocks MurJ’s activity (70). Moreover, in the
presence of the LysM-resistant murJ alleles, the
precursor levels were restored to normal. The fact
that LysM only targets MurJ and causes lysis
strongly suggests that MurJ is the only active lipid
II flippase in E. coli. However, the interpretation of
the LysM-resistant murJ alleles and the interaction
interface would greatly benefit from the structure of
the MurJ-LysM complex.
`
MS2 LYSIS: TO L AND BACK
L: the first autolysin?
The L gene was not recognized as a gene in
MS2 until the isolation of a plaque-forming
defective nonsense mutant that belonged to a
complementation group distinct from mat, coat and
rep (72). Subsequent radioactive labeling
experiments established L as the fourth gene of
MS2, encoding a 75 aa polypeptide (73). As had
been done with E and A2, a plasmid clone of L was
shown to cause lysis after induction and the L
protein was shown to be associated with the
membrane fraction (73, 74). Opposite to E, it is the
39 residue CTD of L that accounts for membrane-
localization and lytic function, with the N-terminal
36 highly basic residues shown to be dispensable
for lysis (Supp. Fig. 4) (75). A key experiment was
that, after induction, net murein synthesis, as
assessed by [3H] mDAP into SDS-insoluble
material, was unaffected prior to the onset of lysis
(76). This clearly differentiates L action from the E,
A2 and LysM Sgl proteins, all of which cause
cessation of cell wall synthesis by interrupting the
supply of Lipid II to the PG machinery (43, 67, 70).
To these workers, the most significant finding was
that induced L lysis was severely compromised in
acidic (pH 5.5) conditions, despite normal
accumulation of L, raising a compelling analogy to
penicillin-induced autolysis, which is also blocked
under these conditions (77). This led to a general
model in which L effected lysis by inducing
autolysis, although the precise definition thereof
was not provided. In immuno-EM experiments, L
was shown to preferentially localize to apparent
zones of adhesion between the IM and OM (78).
This association with adhesion zones was
emphasized by the fact that L lysis is also
compromised in cells that lacked the periplasmic
osmoprotectant membrane-derived-
oligosaccharide (MDO) (79). These cells were
shown to have many fewer adhesion zones and a
much wider periplasm, and in this case L appeared
to be subject to degradation. Furthermore, a
synthetic polypeptide corresponding to the C-
terminal 25 aa of L was shown to permeabilize both
liposomes and inverted membrane vesicles, leading
the authors to invoke induction of autolysis after
membrane permeabilization (80). However, these
experiments lacked a negative polypeptide control
and the experiments were done at peptide to vesicle
ratio in excess of 1000; moreover, permeabilization
and depolarization does not result in rapid autolysis
in E. coli, so the physiological relevance of these
experiments is questionable.
Back to L: genetic and molecular analysis
The consignment of MS2 L to the role of
phage-encoded autolysin seemed to end further
interest in its function, despite the likelihood that a
critical component of cell wall homeostasis was
targeted. However, over the next decades, a few
new Leviviridae were characterized, many of which
shared the same genetic architecture as MS2,
despite no significant nucleotide sequence
similarity (60, 61) (Supp. Table 1). This included
not only new leviviruses specific for the F pilus but
also against the conjugational pilus of several R-
factor plasmids, and the polar pilus of Pseudomonas
(Supp. Fig. 5) (57, 58, 60, 61). We noticed that the
L proteins, although unrelated in terms of sequence,
shared an apparent domain organization with L:
domain 1, N-terminal, highly charged; domain 2,
very hydrophobic and lacking charged residues;
domain 3, a central Leu-Ser dipeptide; and domain
4, a variable CTD (81). The conserved architecture
suggested that L-like Sgl systems widespread
among Gram-negative bacteria were all targeting
the same host function.
Two genetics-based approaches were
mounted, the first aimed at identifying host factors,
using, as before, inducible plasmid-based clones of
L (81, 82). To avoid mutations that reduced copy
number or L transcription, a blue/white reporter
plasmid was constructed and from hundreds of
colonies surviving L induction from this construct,
two blue colonies were identified and designated as
ill (insensitive to L lysis) mutants (82).
Surprisingly, the ill mutations mapped to dnaJ,
which encodes a widely-conserved chaperone
involved in the heat shock response (83). Analysis
revealed that, in both, a P330Q missense change in
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
7
dnaJ accounted for the Ill phenotype and abolished
MS2 plaque-formation, with both phage and
survival phenotypes recessive. The Pro330 residue
is the most conserved residue in the CTD of DnaJ,
which is clearly a conserved segment although its
function is unclear. The P330Q change was found
to preserve the heat shock function of DnaJ but
abrogates the ability of DnaJ to form complexes
with L. The L suppressors, designated as Lodj
(overcomes dnaJ) alleles, were isolated as mutants
that allowed lysis in dnaJP330Q background; these
proved to be deletions of the dispensable NTD of L.
Isogenic inductions of the parental and Lodj alleles
revealed that lysis was much earlier with the
truncations. These results led to a model in which
the NTD of L has a regulatory, lysis-delay function
that blocks the interaction of L with its target; in this
model, DnaJ is required for relief of this steric block
(Supp. Fig. 6).
To identify key functional elements of L
itself, a nearly-saturated mutational analysis of L
generated a collection of 103 alleles with single
codon changes conferring absolute lysis defects
(Supp. Fig. 4) (81). The mutational distribution
validated the proposed four domain structure of the
L Sgl proteins. Domain 1, comprising the
dispensable, highly basic region, gave rise to only
one non-lytic allele (Q33H). Domain 3 containing
the LS dipeptide motif and the adjacent segments of
Domains 2 and 4 had the most missense changes
conferring non-lytic character. All of the missense
alleles tested were genetically recessive and
generated membrane-associated products of
parental size. In addition, several of the
inactivating missense changes (i.e., L44V, F47L,
F47Y, S49T, F51L and L56F) were conservative,
suggesting the L protein makes specific heterotypic
protein-protein contacts in the membrane.
Taken together, the isolation of dnaJP330Q
and mutational analysis of L both suggest that L
targets a host membrane protein, the interaction is
through the mutationally sensitive residues in
domains 2, 3, and 4, and that, like SlyD and E, a
host chaperone is involved in regulating L function.
Obviously, further investigation into the host
factors involved in L lysis are needed to understand
the mechanistic details of L function.
WHAT’S NEXT?
The premise of this review was that the
study of the Sgl systems of small lytic phages
would be interesting and lead to a better
understanding of the bacterial cell wall biosynthesis
and homeostasis. To summarize what has been
discussed, four Sgl systems have been studied in
depth, one from the microvirus X174 and three
from the Leviviridae. In three cases, the Sgl
proteins turned out to be “protein antibiotics”,
specific inhibitors of different enzymes of the
highly conserved PG biosynthesis pathway (3, 70,
84). Consideration of the available genetic and
biochemical data has already increased our
understanding of how these important enzymes
function, and in the case of LysM, settled the
identification of the flippase that exports Lipid II to
the periplasm. The next level of understanding will
come from detailed structural information about the
Sgl-enzyme inhibition complexes. The fourth case,
the L protein of MS2, has not yet been fully
characterized but appears to be the prototype of a
Sgl type that has evolved multiple times in
Leviviridae infecting a wide range of Gram-
negative bacteria (81). Genetic and biochemical
evidence was cited showing that L does not inhibit
any of the steps that lead to externalized lipid II and
its incorporation into existing PG and suggests that
it targets a host protein. There is no conceivable
answer to the L target mystery that would not be
important, possibly identifying conserved proteins
that are essential for proper coordination or
localization of cell wall synthesis machineries or
are involved in the control of powerful autolytic
enzymes.
Even with the L story still incomplete, this
seems like a pretty good haul of information from
the study of four small genes, starting with very
“low-tech”, old-fashioned and simple genetic
selections. The shocking thing is that this wealth of
molecular information is derived from the study of
only nine distinct Leviviridae (Supp. Table 1),
comprising a total of <50 kb of total genomic
information. Despite the low number, these 9
phages segregate into five different phage-type
based on where the Sgl evolved. Listed from 5’ to
3’ of the gRNA, they are: AP205 (5’ of mat); Q
(mat = A2), MS2 (overlapping end of coat and
beginning of rep), phiCb5 (middle of rep), and M
(near 3’ end of rep), (Fig. 2b). The diverse location
of the Sgl genes suggests that they have evolved
more than once and probably as a late addition to
the genome after speciation to different pili or hosts
(58, 61). (This includes all the L-like genes; none
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
8
of the 6 L-type Sgls has any detectable sequence
identity, other than the LS dipeptide
sequence.) Given the diversity of Sgl systems and
the existence of multiple protein targets in PG
biosynthesis and maintenance, it is not difficult to
imagine the existence of Sgl inhibitors for every
known step in PG biosynthesis and, if L is any
indicator, possibly uncharacterized components
critical for dynamic cell wall homeostasis.
Taken together, it seems obvious that it
would be useful to identify new Sgl genes, and the
old-fashioned “phage hunt” is a reliable
approach. RNA phage hunts have so far been done
for five conjugational pili, resulting in two protein
antibiotic Sgls (A2 and LysM) and four unrelated L-
type Sgls (LMS2, LHGAL, LC1, and LPRR1). Only three non-
conjugational pili (Caulobacter, Acinetobacter and
Pseudomonas) have been targeted, resulting in two
L-type Sgls and one, Lys of Caulobacter phage
phiCB5, which does not have an L-type domain
structure but does have a single N-terminal TMD,
resembling both E and LysM. Considering the
existence of many more retractable pili systems,
there is a clear rationale to conduct RNA phage
hunts in many other systems with retractable pili,
especially in pathogenic bacteria.
Metagenomics is also having an impact. A
recent survey of publicly available RNA-inclusive
metagenomes and RNA virome studies of
invertebrate species led to the identification of ~200
new ssRNA phage genomes (85, 86). Although
most of the new leviviral genomes are partial, ~80
are either complete or nearly so, with all three core
genes annotated. Only one (AVE017) of these
~200 genomes had an annotated Sgl gene, being a
close relative (38% sequence identity) of MS2
L(85). Given their small size, predilection for being
embedded in the core genes, and extreme sequence
diversity at the protein level, finding Sgl candidates
in these genomes poses unique challenges to the
traditional gene annotation tools. Moreover,
currently there is no direct way to sort out these
leviviral genomes to particular bacterial host, or
more specifically, to a particular retractable pilus.
Nevertheless, the promise of more intriguing Sgl
proteins targeting novel components of the bacterial
cell wall machinery surely makes our current effort,
which involves identifying potential Sgl ORFs and
characterizing them one by one for the ability to
support lysis after induction of synthetic clones,
worth doing.
ACKNOWLEDGEMENTS
This work was supported by Public Health Service grant GM27099 and by the Center for Phage Technology
at Texas A&M University, jointly sponsored by Texas A&M AgriLife. The content is solely the
responsibility of the authors and does not necessarily represent the official views of the National Institutes
of Health. The following figures and figure legends were reproduced verbatim from published references:
Fig. 1, Fig. 3, Fig. 4, Supplementary figures 3, 4, 5, and 6. We thank past and current members of the Young
laboratory for their advice. We also thank Ing-Nang Wang, Kaspar Tars and Rene Olsthoorn for providing
ssRNA phage resources and expertise. The clerical assistance of Daisy Wilbert is highly appreciated.
CONFLICT OF INTEREST
The authors declare no conflict of interest.
REFERENCES
1. Young, R. (2013) Phage lysis: do we have the hole story yet? Curr Opin Microbiol 16, 790-797
2. Young, R. (2014) Phage lysis: three steps, three choices, one outcome. J Microbiol 52, 243-258
3. Bernhardt, T. G., Wang, I. N., Struck, D. K., and Young, R. (2002) Breaking free: "protein
antibiotics" and phage lysis. Res Microbiol 153, 493-501
4. Vollmer, W., Blanot, D., and De Pedro, M. A. (2008) Peptidoglycan structure and architecture.
FEMS Microbiol Rev 32, 149-167
5. Typas, A., Banzhaf, M., Gross, C. A., and Vollmer, W. (2012) From the regulation of
peptidoglycan synthesis to bacterial growth and morphology. Nat Rev Microbiol 10, 123-136
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
9
6. Silhavy, T. J., Kahne, D., and Walker, S. (2010) The Bacterial Cell Envelope. Cold Spring Harb
Perspect Biol 2, a000414
7. Thwaites, J. J., and Mendelson, N. H. (1989) Mechanical properties of peptidoglycan as
determined from bacterial thread. Int J Biol Macromol 11, 201-206
8. Stock, J. B., Rauch, B., and Roseman, S. (1977) Periplasmic space in Salmonella typhimurium
and Escherichia coli. J Biol Chem 252, 7850-7861
9. Gan, L., Chen, S., and Jensen, G. J. (2008) Molecular organization of Gram-negative
peptidoglycan. Proc Natl Acad Sci U S A 105, 18953-18957
10. Braun, V. (1975) Covalent lipoprotein from the outer membrane of Escherichia coli. Biochim
Biophys Acta 415, 335-377
11. Cowles, C. E., Li, Y., Semmelhack, M. F., Cristea, I. M., and Silhavy, T. J. (2011) The free and
bound forms of Lpp occupy distinct subcellular locations in Escherichia coli. Mol Microbiol 79,
1168-1181
12. Brown, E. D., Vivas, E. I., Walsh, C. T., and Kolter, R. (1995) MurA (MurZ), the enzyme that
catalyzes the first committed step in peptidoglycan biosynthesis, is essential in Escherichia coli. J
Bacteriol 177, 4194-4197
13. Lovering, A. L., Safadi, S. S., and Strynadka, N. C. (2012) Structural perspective of
peptidoglycan biosynthesis and assembly. Annu Rev Biochem 81, 451-478
14. Mohammadi, T., van Dam, V., Sijbrandi, R., Vernet, T., Zapun, A., Bouhss, A., Diepeveen-de
Bruin, M., Nguyen-Disteche, M., de Kruijff, B., and Breukink, E. (2011) Identification of FtsW
as a transporter of lipid-linked cell wall precursors across the membrane. EMBO J 30, 1425-1432
15. Mohammadi, T., Sijbrandi, R., Lutters, M., Verheul, J., Martin, N. I., den Blaauwen, T., de
Kruijff, B., and Breukink, E. (2014) Specificity of the transport of lipid II by FtsW in Escherichia
coli. J Biol Chem 289, 14707-14718
16. Ruiz, N. (2008) Bioinformatics identification of MurJ (MviN) as the peptidoglycan lipid II
flippase in Escherichia coli. Proc Natl Acad Sci U S A 105, 15553-15557
17. Sham, L. T., Butler, E. K., Lebar, M. D., Kahne, D., Bernhardt, T. G., and Ruiz, N. (2014)
Bacterial cell wall. MurJ is the flippase of lipid-linked precursors for peptidoglycan biogenesis.
Science 345, 220-222
18. Meeske, A. J., Sham, L. T., Kimsey, H., Koo, B. M., Gross, C. A., Bernhardt, T. G., and Rudner,
D. Z. (2015) MurJ and a novel lipid II flippase are required for cell wall biogenesis in Bacillus
subtilis. Proc Natl Acad Sci U S A 112, 6437-6442
19. Kuk, A. C., Mashalidis, E. H., and Lee, S. Y. (2017) Crystal structure of the MOP flippase MurJ
in an inward-facing conformation. Nat Struct Mol Biol 24, 171-176
20. Bolla, J. R., Sauer, J. B., Wu, D., Mehmood, S., Allison, T. M., and Robinson, C. V. (2018)
Direct observation of the influence of cardiolipin and antibiotics on lipid II binding to MurJ. Nat
Chem 10, 363-371
21. Zheng, S., Sham, L. T., Rubino, F. A., Brock, K. P., Robins, W. P., Mekalanos, J. J., Marks, D.
S., Bernhardt, T. G., and Kruse, A. C. (2018) Structure and mutagenic analysis of the lipid II
flippase MurJ from Escherichia coli. Proc Natl Acad Sci U S A 115, 6709-6714
22. Meeske, A. J., Riley, E. P., Robins, W. P., Uehara, T., Mekalanos, J. J., Kahne, D., Walker, S.,
Kruse, A. C., Bernhardt, T. G., and Rudner, D. Z. (2016) SEDS proteins are a widespread family
of bacterial cell wall polymerases. Nature 537, 634-638
23. Cho, H., Wivagg, C. N., Kapoor, M., Barry, Z., Rohs, P. D., Suh, H., Marto, J. A., Garner, E. C.,
and Bernhardt, T. G. (2016) Bacterial cell wall biogenesis is mediated by SEDS and PBP
polymerase families functioning semi-autonomously. Nat Microbiol, 16172
24. Sjodt, M., Brock, K., Dobihal, G., Rohs, P. D. A., Green, A. G., Hopf, T. A., Meeske, A. J.,
Srisuknimit, V., Kahne, D., Walker, S., Marks, D. S., Bernhardt, T. G., Rudner, D. Z., and Kruse,
A. C. (2018) Structure of the peptidoglycan polymerase RodA resolved by evolutionary coupling
analysis. Nature 556, 118-121
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
10
25. Kim, M. S., Park, E. J., Roh, S. W., and Bae, J. W. (2011) Diversity and abundance of single-
stranded DNA viruses in human feces. Appl Environ Microbiol 77, 8062-8070
26. Sanger, F., Air, G. M., Barrell, B. G., Brown, N. L., Coulson, A. R., Fiddes, J. C., Hutchison, C.
A., III, Slocombe, P. M., and Smith, M. (1977) Nucleotide sequence of bacteriophage X174
DNA. Nature 265, 687-695
27. Smith, H. O., Hutchison, C. A., 3rd, Pfannkoch, C., and Venter, J. C. (2003) Generating a
synthetic genome by whole genome assembly:X174 bacteriophage from synthetic
oligonucleotides. Proc Natl Acad Sci U S A 100, 15440-15445
28. Hutchison, C. A., 3rd, Phillips, S., Edgell, M. H., Gillam, S., Jahnke, P., and Smith, M. (1978)
Mutagenesis at a specific position in a DNA sequence. J Biol Chem 253, 6551-6560
29. Henrich, B., Lubitz, W., and Plapp, R. (1982) Lysis of Escherichia coli by induction of cloned
X174 genes. Mol Gen Genet 185, 493-497
30. Young, K. D., and Young, R. (1982) Lytic action of cloned X174 gene E. J Virol 44, 993-1002
31. Bradley, D. E., Dewar, C. A., and Robertson, D. (1969) Structural changes in Escherichia coli
infected with a X174 type bacteriophage. J Gen Virol 5, 113-121
32. Hahn, F. E., and Ciak, J. (1957) Penicillin-induced lysis of Escherichia coli. Science 125, 119-
120
33. Bläsi, U., Henrich, B., and Lubitz, W. (1985) Lysis of Escherichia coli by cloned X174 gene E
depends on its expression. J Gen Microbiol 131, 1107-1114
34. Lubitz, W., and Plapp, R. (1983) Stimulation of autolysis by adsorption of bacteriophage X174
to isolated cell walls. Curr Microbiol 8, 63-65
35. Bläsi, U., Halfmann, G., and Lubitz, W. (1984) Induction of autolysis of Escherichia coli by
X174 gene E product. in Microbial Cell Wall Synthesis and Autolysis (Nombela, C. ed.),
Elsevier Science Publishers, New York. pp 213-218
36. Lubitz, W., Halfmann, G., and Plapp, R. (1984) Lysis of Escherichia coli after infection with
X174 depends on the regulation of the cellular autolytic system. J Gen Microbiol 130, 1079-
1087
37. Witte, A., Wanner, G., Bläsi, U., Halfmann, G., Szostak, M., and Lubitz, W. (1990) Endogenous
transmembrane tunnel formation mediated by X174 lysis protein E. J Bacteriol 172, 4109-4114
38. Maratea, D., Young, K., and Young, R. (1985) Deletion and fusion analysis of theX174 lysis
gene E. Gene 40, 39-46
39. Buckley, K. J., and Hayashi, M. (1986) Lytic activity localized to membrane-spanning region of
X174 E protein. Mol Gen Genet 204, 120-125
40. Roof, W. D., Horne, S. M., Young, K. D., and Young, R. (1994) slyD, a host gene required for
X174 lysis, is related to the FK506-binding protein family of peptidyl-prolyl cis-trans-
isomerases. J Biol Chem 269, 2902-2910
41. Kay, J. E. (1996) Structure-function relationships in the FK506-binding protein (FKBP) family of
peptidylprolyl cis-trans isomerases. Biochem J 314, 361-385
42. Bernhardt, T. G., Roof, W. D., and Young, R. (2002) The Escherichia coli FKBP-type PPIase
SlyD is required for the stabilization of the E lysis protein of bacteriophage X174. Mol
Microbiol 45, 99-108
43. Bernhardt, T. G., Struck, D. K., and Young, R. (2001) The lysis protein E of X174 is a specific
inhibitor of the MraY-catalyzed step in peptidoglycan synthesis. J Biol Chem 276, 6093-6097
44. Bernhardt, T. G., Roof, W. D., and Young, R. (2000) Genetic evidence that the bacteriophage
X174 lysis protein inhibits cell wall synthesis. Proc Natl Acad Sci U S A 97, 4297-4302
45. Zheng, Y., Struck, D. K., and Young, R. (2009) Purification and functional characterization of
X174 lysis protein E. Biochemistry 48, 4999-5006
46. Pollock, T. J., Tessman, E. S., and Tessman, I. (1978) Identification of lysis protein E of
bacteriophage X174. J Virol 28, 408-410
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
11
47. Zheng, Y., Struck, D. K., Bernhardt, T. G., and Young, R. (2008) Genetic analysis of MraY
inhibition by the X174 protein E. Genetics 180, 1459-1466
48. Chung, B. C., Zhao, J., Gillespie, R. A., Kwon, D. Y., Guan, Z., Hong, J., Zhou, P., and Lee, S.
Y. (2013) Crystal structure of MraY, an essential membrane enzyme for bacterial cell wall
synthesis. Science 341, 1012-1016
49. Tanaka, S., and Clemons, W. M., Jr. (2012) Minimal requirements for inhibition of MraY by lysis
protein E from bacteriophage X174. Mol Microbiol 85, 975-985
50. Inokuchi, Y., Jacobson, A. B., Hirose, T., Inayama, S., and Hirashima, A. (1988) Analysis of the
complete nucleotide sequence of the group IV RNA coliphage SP. Nucleic Acids Res. 16, 6205-
6221
51. Inokuchi, Y., Takahashi, R., Hirose, T., Inayama, S., Jacobson, A. B., and Hirashima, A. (1986)
The complete nucleotide sequence of the group II RNA coliphage GA. J Biochem (Tokyo) 99,
1169-1180
52. Inokuchi, Y., Hirashima, A., and Watanabe, I. (1982) Comparison of the nucleotide sequences at
the 3'-terminal region of RNAs from RNA coliphages. J Mol Biol 158, 711-730
53. Stewart, J. R., Vinje, J., Oudejans, S. J., Scott, G. I., and Sobsey, M. D. (2006) Sequence
variation among group III F-specific RNA coliphages from water samples and swine lagoons.
Appl Environ Microbiol 72, 1226-1230
54. Friedman, S. D., Cooper, E. M., Casanova, L., Sobsey, M. D., and Genthner, F. J. (2009) A
reverse transcription-PCR assay to distinguish the four genogroups of male-specific (F+) RNA
coliphages. J Virol Methods 159, 47-52
55. Adhin, M. R., Hirashima, A., and van Duin, J. (1989) Nucleotide sequence from the ssRNA
bacteriophage JP34 resolves the discrepancy between serological and biophysical classification.
Virol. 170, 238-242
56. Kazaks, A., Voronkova, T., Rumnieks, J., Dishlers, A., and Tars, K. (2011) Genome structure of
Caulobacter phage phiCb5. J Virol 85, 4628-4631
57. Ruokoranta, T. M., Grahn, A. M., Ravantti, J. J., Poranen, M. M., and Bamford, D. H. (2006)
Complete genome sequence of the broad host range single-stranded RNA phage PRR1 places it in
the Levivirus genus with characteristics shared with Alloleviviruses. J Virol 80, 9326-9330
58. Rumnieks, J., and Tars, K. (2012) Diversity of pili-specific bacteriophages: genome sequence of
IncM plasmid-dependent RNA phage M. BMC Microbiol 12, 277
59. Klovins, J., Overbeek, G. P., van den Worm, S. H., Ackermann, H. W., and van Duin, J. (2002)
Nucleotide sequence of a ssRNA phage from Acinetobacter: kinship to coliphages. J Gen Virol
83, 1523-1533
60. Olsthoorn, R. C., Garde, G., Dayhuff, T., Atkins, J. F., and Van Duin, J. (1995) Nucleotide
sequence of a single-stranded RNA phage from Pseudomonas aeruginosa: kinship to coliphages
and conservation of regulatory RNA structures. Virology 206, 611-625
61. Kannoly, S., Shao, Y., and Wang, I. N. (2012) Rethinking the evolution of single-stranded RNA
(ssRNA) bacteriophages based on genomic sequences and characterizations of two R-plasmid-
dependent ssRNA phages, C-1 and Hgal1. J Bacteriol 194, 5073-5079
62. Domingo, E., and Holland, J. J. (1997) RNA virus mutations and fitness for survival. Annu Rev
Microbiol 51, 151-178
63. Cui, Z., Gorzelnik, K. V., Chang, J. Y., Langlais, C., Jakana, J., Young, R., and Zhang, J. (2017)
Structures of Q virions, virus-like particles, and the Q-MurA complex reveal internal coat
proteins and the mechanism of host lysis. Proc Natl Acad Sci U S A 114, 11697-11702
64. Weber, K., and Konigsberg, W. (1975) Proteins of the RNA phages. in RNA phages (Zinder, N.
D. ed.), Cold Spring Harbor Laboratory, Cold Spring Harbor NY. pp 51-84
65. Kozak, M., and Nathans, D. (1971) Fate of maturation protein during infection by coliphage
MS2. Nat.New Biol. 234, 209-211
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
12
66. Winter, R. B., and Gold, L. (1983) Overproduction of bacteriophage Q maturation (A2) protein
leads to cell lysis. Cell 33, 877-885
67. Bernhardt, T. G., Wang, I. N., Struck, D. K., and Young, R. (2001) A protein antibiotic in the
phage Q virion: diversity in lysis targets. Science 292, 2326-2329
68. Reed, C. A., Langlais, C., Kuznetsov, V., and Young, R. (2012) Inhibitory mechanism of the Q
lysis protein A2. Mol Microbiol 86, 836-844
69. Reed, C. A., Langlais, C., Wang, I. N., and Young, R. (2013) A2 expression and assembly
regulates lysis in Q infections. Microbiology 159, 507-514
70. Chamakura, K. R., Sham, L. T., Davis, R. M., Min, L., Cho, H., Ruiz, N., Bernhardt, T. G., and
Young, R. (2017) A viral protein antibiotic inhibits lipid II flippase activity. Nat Microbiol 2,
1480-1484
71. Butler, E. K., Davis, R. M., Bari, V., Nicholson, P. A., and Ruiz, N. (2013) Structure-function
analysis of MurJ reveals a solvent-exposed cavity containing residues essential for peptidoglycan
biogenesis in Escherichia coli. J Bacteriol 195, 4639-4649
72. Model, P., Webster, R. E., and Zinder, N. D. (1979) Characterization of Op3, a lysis-defective
mutant of bacteriophage f2. Cell 18, 235-244
73. Beremand, M. N., and Blumenthal, T. (1979) Overlapping genes in RNA phage: a new protein
implicated in lysis. Cell 18, 257-266
74. Coleman, J., Inouye, M., and Atkins, J. (1983) Bacteriophage MS2 lysis protein does not require
coat protein to mediate cell lysis. J Bacteriol 153, 1098-1100
75. Berkhout, B., de Smit, M. H., Spanjaard, R. A., Blom, T., and van Duin, J. (1985) The amino
terminal half of the MS2-coded lysis protein is dispensable for function: implications for our
understanding of coding region overlaps. EMBO J 4, 3315-3320
76. Holtje, J. V., and van Duin, J. (1984) MS2 phage induced lysis of E. coli depends upon the
activity of the bacterial autolysins. in Microbial Cell Wall Synthesis and Autolysis (Nombela, C.
ed.), Elsevier Science Publishers, New York. pp 195-199
77. Walderich, B., Ursinus-Wosner, A., van Duin, J., and Holtje, J. V. (1988) Induction of the
autolytic system of Escherichia coli by specific insertion of bacteriophage MS2 lysis protein into
the bacterial cell envelope. J Bacteriol 170, 5027-5033
78. Walderich, B., and Holtje, J. V. (1989) Specific localization of the lysis protein of bacteriophage
MS2 in membrane adhesion sites of Escherichia coli. J Bacteriol 171, 3331-3336
79. Holtje, J. V., Fiedler, W., Rotering, H., Walderich, B., and van Duin, J. (1988) Lysis induction of
Escherichia coli by the cloned lysis protein of the phage MS2 depends on the presence of
osmoregulatory membrane-derived oligosaccharides. J Biol Chem 263, 3539-3541
80. Goessens, W. H. F., Driessen, A. J. M., Wilschut, J., and van Duin, J. (1988) A synthetic peptide
corresponding to the C-terminal 25 residues of phage MS2-coded lysis protein dissipates the
proton-motive force in Escherichia coli membrane vesicles by generating hydrophilic pores.
EMBO J 7, 867-873
81. Chamakura, K. R., Edwards, G. B., and Young, R. (2017) Mutational analysis of the MS2 lysis
protein L. Microbiology 163, 961-969
82. Chamakura, K. R., Tran, J. S., and Young, R. (2017) MS2 lysis of Escherichia coli depends on
host chaperone DnaJ. J Bacteriol 199
83. Qiu, X. B., Shao, Y. M., Miao, S., and Wang, L. (2006) The diversity of the DnaJ/Hsp40 family,
the crucial partners for Hsp70 chaperones. Cell Mol Life Sci 63, 2560-2570
84. Hatfull, G. F. (2001) Microbiology. The great escape. Science 292, 2263-2264
85. Krishnamurthy, S. R., Janowski, A. B., Zhao, G., Barouch, D., and Wang, D. (2016)
Hyperexpansion of RNA Bacteriophage Diversity. PLoS Biol 14, e1002409
86. Shi, M., Lin, X. D., Tian, J. H., Chen, L. J., Chen, X., Li, C. X., Qin, X. C., Li, J., Cao, J. P.,
Eden, J. S., Buchmann, J., Wang, W., Xu, J., Holmes, E. C., and Zhang, Y. Z. (2016) Redefining
the invertebrate RNA virosphere. Nature 540, 539-543
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
13
FIGURE LEGENDS
Figure 1. The PG biosynthesis pathway and the Sgl system.
(A) The PG precursor pathway, from cytoplasmic UDP-GlcNAc to periplasmic lipid II, with known targets
of 'protein antibiotics' indicated. (B) Genome organization in ssDNA ( X174) and ssRNA (Qβ, AP205,
MS2, PhiCb5, and M) phages. In ssRNA phages, mat encodes the maturation protein responsible for
adsorption to the receptor pilus, coat encodes the capsid protein, and rep encodes the replicase. In Qβ, the
mat gene is named A2 and has the additional function of inducing host lysis.
Figure 2. The E-resistant E. coli mraY alleles mapped on Aquifex aeolicus MraYAA.
The E. coli mraY alleles resistant to E are mapped on the structure of MraY from Aquifex aeolicus .(MraYAA)
(PDB 4J72). The structure is shown from the face that forms the putative E-MraY interface and the domains
of the interface are colored as follows; the periplasmic beta hairpin (yellow), the loop connecting PB with
TMD5 (purple), TMD5 (magenta), and TMD9 (9a and 9b) (grey). The homologous E-resistant residues in
MraYAA are shown as spheres and the catalytically important aspartate (red) and histidine (blue) residues
are shown as stick projections.
Figure 3. The structure of Qβ bound to MurA
(A) The structure of bound to MurA and it is colored as follows coat proteins (blue), A2 (hot pink),
gRNA (yellow) and MurA (orange). (B) A 90° turn and cutaway view of shows MurA bound to the
maturation protein with same color scheme as (A) except in case of coat proteins (radially colored from
light blue to blue) and extra protein density (green). (C) The ribbon model of A2 bound to MurA with
uridine diphosphate N-acetylglucosamine (UDP-GlcNAc) in the active site (cornflower blue). (D) The
ribbon model of MurA viewed from the MurAA2 interface. The point mutations that render MurA resistant
to A2 are labeled and shown as red stick models. Locations of the catalytic loop and the UDP-GlcNAc are
indicated by black arrows. (E) Ribbon model of A2 as viewed from the MurAA2 interface. The region
interacting with MurA, encompassing the N-terminal β-sheet region of residues 30120, is outlined by a
black lasso. The N and C termini are indicated by black arrows.
Figure 4. LysM-resistance mutations map to TMD2 and TMD7 of MurJ.
The amino-acid changes in E. coli MurJ (MurJEC) resulting in LysM-resistance were mapped onto the
structure of MurJ from Thermosipho africanus (MurJTA) (PDB 5T77). (A) The cytoplasmic-open
conformation. (B), A model of the periplasmic-open conformation (19). The TMDs that line the central
hydrophilic cavity are colored: TMD2 (light blue) and TMD7 (magenta). Lateral view (left) and periplasmic
view (right). The changes in MurJEC and homologous amino acids in MurJTA are shown on the right.
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
14
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
15
Figure 2
TMD9a TMD5
TMD9b
PB
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
16
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
17
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
Karthik Chamakura and Ry Young
Phage single-gene lysis: Finding the weak spot in the bacterial cell wall
published online November 12, 2018J. Biol. Chem.
10.1074/jbc.TM118.001773Access the most updated version of this article at doi:
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
at Texas A&M University - Medical Sciences Library on November 12, 2018http://www.jbc.org/Downloaded from
... The inhibition of peptidoglycan synthesis triggers the uncontrolled activity of autolysins, proposing a possible mechanism of single-gene-mediated cell lysis [2,4]. However, although it is under the same regulatory control mechanisms as penicillin-induced lysis, L protein is not involved in the inhibition of peptidoglycan synthesis [14]. ...
... ssRNA phage-induced cell lysis is slow compared to many other phage systems [14]. Additions of chloroform, lysozyme, or both were tested for their ability to enhance the lysis process. ...
... The results of our experiments suggest that the PRR1 lysis protein plays the role of spanin, the OM permeabilizing protein, in dsDNA phage-induced lysis. An increase in OM permeability is the earliest event observed in PRR1 infection, as well as in the case of the phage MS2 [14]. The peptidoglycan of the infected cells could be digested by the cellular autolysins after the PRR1 lysis protein-induced loss of control of these enzymes in the infected cells. ...
Article
Full-text available
The phage PRR1 belongs to the Leviviridae family, a group of ssRNA bacteriophages that infect Gram-negative bacteria. The variety of host cells is determined by the specificity of PRR1 to a pilus encoded by a broad host range of IncP-type plasmids that confer multiple types of antibiotic resistance to the host. Using P. aeruginosa strain PAO1 as a host, we analyzed the PRR1 infection cycle, focusing on cell lysis. PRR1 infection renders P. aeruginosa cells sensitive to lysozyme approximately 20 min before the start of a drop in suspension turbidity. At the same time, infected cells start to accumulate lipophilic anions. The on-line monitoring of the entire infection cycle showed that single-gene-mediated lysis strongly depends on the host cells’ physiological state. The blockage of respiration or a reduction in the intracellular ATP concentration during the infection resulted in the inhibition of lysis. The same effect was observed when the synthesis of PRR1 lysis protein was induced in an E. coli expression system. In addition, lysis was strongly dependent on the level of aeration. Dissolved oxygen concentrations sufficient to support cell growth did not ensure efficient lysis, and a coupling between cell lysis initiation and aeration level was observed. However, the duration of the drop in suspension turbidity did not depend on the level of aeration.
... By contrast, small lytic phages with single-strand RNA and DNA chromosomes lack the genomic space for an MGL system. Instead, these phages, the Microviridae and the ssRNA Leviviricetes, liberate their progeny by inducing lysis through the expression of a single gene (3,4). These genes have been collectively named sgl (single-gene lysis; product Sgl). ...
... As noted above, it has been reported that, for lytic inductions of L, the rate of PG synthesis, as measured by pulse labeling with 3 [H]mDAP, is unaffected up to the time of lysis (24). In contrast, for both E and A 2 , net PG synthesis, as measured by the continuous incorporation of the same label, is halted at least 20 min before the lytic event (23,41). ...
... These results raised the possibility that type I Sgls can be distinguished from type II using PG labeling. The simplest method is to follow the incorporation of 3 [H]mDAP into material that is insoluble in boiling SDS, i.e., the sacculi. However, 3 [H]mDAP labeling requires a minimal medium, which is done with glucose as the carbon source. ...
Article
Full-text available
Single-strand RNA (ssRNA) and single-strand DNA phages elicit host lysis using a single gene, in each case designated as sgl. Of the 11 identified Sgls, three have been shown to be specific inhibitors of different steps in the pathway that supplies lipid II to the peptidoglycan (PG) biosynthesis machinery. These Sgls have been called “protein antibiotics” because the lytic event is a septal catastrophe indistinguishable from that caused by cell wall antibiotics. Here, we designate these as type I Sgls. In this formalism, the other eight Sgls are assigned to type II, the best-studied of which is protein L of the paradigm F-specific ssRNA phage MS2. Comparisons have suggested that type II Sgls have four sequence elements distinguished by hydrophobic and polar character. Environmental metatranscriptomics has revealed thousands of new ssRNA phage genomes, each of which presumably has an Sgl. Here, we describe methods to distinguish type I and type II Sgls. Using phase contrast microscopy, we show that both classes of Sgls cause the formation of blebs prior to lysis, but the location of the blebs differs significantly. In addition, we show that L and other type II Sgls do not inhibit the net synthesis of PG, as measured by radio-labeling of PG. Finally, we provide direct evidence that the Sgl from Pseudomonas phage PP7 is a type I Sgl, in support of a recent report based on a genetic selection. This shows that the putative four-element sequence structure suggested for L is not a reliable discriminator for the operational characterization of Sgls. IMPORTANCE The ssRNA phage world has recently undergone a metagenomic expansion upward of a thousandfold. Each genome likely carries at least one single-gene lysis (sgl) cistron encoding a protein that single-handedly induces host autolysis. Here, we initiate an approach to segregate the Sgls into operational types based on physiological analysis, as a first step toward the alluring goal of finding many new ways to induce bacterial death and the attendant expectations for new antibiotic development.
... In most cases, dsDNA phages encode distinct VAPGHs and endolysins, whereas dsRNA phages feature a single lytic protein that serves as both VAPGH and endolysin [140]. In contrast, ssDNA and ssRNA phages employ a single-gene lysis (Sgl) protein-an impactful degradative protein that induces cytolysis without enzymatically breaking down PGs [141,142]. cases, dsDNA phages encode distinct VAPGHs and endolysins, whereas dsRNA phages feature a single lytic protein that serves as both VAPGH and endolysin [140]. In contrast, ssDNA and ssRNA phages employ a single-gene lysis (Sgl) protein-an impactful degradative protein that induces cytolysis without enzymatically breaking down PGs [141,142]. ...
... cases, dsDNA phages encode distinct VAPGHs and endolysins, whereas dsRNA phages feature a single lytic protein that serves as both VAPGH and endolysin [140]. In contrast, ssDNA and ssRNA phages employ a single-gene lysis (Sgl) protein-an impactful degradative protein that induces cytolysis without enzymatically breaking down PGs [141,142]. (1) Utilization of phage-derived components as antibacterial agents. For instance, the degradation of bacterial cell wall by dsRNA phage lytic enzyme [143] or inhibition of cell wall synthesis using singlegene lysis (Slg) protein from ssDNA and ssRNA phages [141,142,144]. ...
... (1) Utilization of phage-derived components as antibacterial agents. For instance, the degradation of bacterial cell wall by dsRNA phage lytic enzyme [143] or inhibition of cell wall synthesis using singlegene lysis (Slg) protein from ssDNA and ssRNA phages [141,142,144]. (2) Phage cocktail employing natural ssDNA and RNA phages as antimicrobial therapeutics against AMR bacteria and phageresistant mutants [145,146]. ...
Article
Full-text available
RNA and single-stranded DNA (ssDNA) phages make up an understudied subset of bacteriophages that have been rapidly expanding in the last decade thanks to advancements in metaviromics. Since their discovery, applications of genetic engineering to ssDNA and RNA phages have revealed their immense potential for diverse applications in healthcare and biotechnology. In this review, we explore the past and present applications of this underexplored group of phages, particularly their current usage as therapeutic agents against multidrug-resistant bacteria. We also discuss engineering techniques such as recombinant expression, CRISPR/Cas-based genome editing, and synthetic rebooting of phage-like particles for their role in tailoring phages for disease treatment, imaging, biomaterial development, and delivery systems. Recent breakthroughs in RNA phage engineering techniques are especially highlighted. We conclude with a perspective on challenges and future prospects, emphasizing the untapped diversity of ssDNA and RNA phages and their potential to revolutionize biotechnology and medicine.
... By contrast, small lytic phages with single-strand RNA and DNA chromosomes lack the genomic space for an MGL system. Instead, these phages, the Microviridae and the ssRNA Leviviricetes, liberate their progeny by inducing lysis through expression of a single gene (3,4). These genes have been collectively named sgl (single gene lysis; product Sgl). ...
... The position and character of the sgls are particularly striking in the ssRNA phages, which have three core genes that encode virion and replication proteins (Fig. 1A). In contrast to the core genes, which are always in the same order and approximately the same size, the location and size of the sgl is highly variable, embedded out of frame in one of the three core genes or positioned in an intergenic space (3). This strongly indicates that there is an independent evolutionary path for each sgl. ...
... Its Sgl character was unambiguously demonstrated by its ability to cause autolysis by induction from plasmid expression vectors (7)(8)(9). Until recently the Sgls of only 9 other plaqueforming phages have been identified by the same functional criterion (3). This includes two Sgls that have been extensively studied: E, the Sgl of the canonical ssDNA Microvirus X174, and A2, the dual function maturation protein/Sgl of Q, another well-studied Fspecific ssRNA phage of Escherichia coli. ...
Preprint
Full-text available
Until recently only 11 distinct Sgls ( si ngle g ene lysis proteins) have been experimentally identified. Of these, three have been shown to be specific inhibitors of different steps in the pathway that supplies Lipid II to the peptidoglycan (PG) biosynthesis machinery: Qβ A 2 inhibits MurA, ϕX174 E inhibits MraY, and Lys from coliphage M inhibits MurJ. These Sgls have been called “protein antibiotics” because the lytic event is a septal catastrophe indistinguishable from that caused by cell wall antibiotics. Here we propose to designate these as members of type I Sgls, to distinguish them from another Sgl, the L protein of the paradigm ssRNA phage MS2. Although none of the other distinct Sgls have significant sequence similarity to L, alignments suggested the presence of four domains distinguished by hydrophobic and polar character. The simplest notion is that these other Sgls have the same autolytic mechanism and, based on this, constitute type II. Although the number of experimentally confirmed Sgls has not changed, recent environmental metagenomes and metatranscriptomes have revealed thousands of new ssRNA phage genomes, each of which presumably has at least one Sgl gene. Here we report on methods to distinguish type I and type II Sgls. Using phase-contrast microscopy, we show that both classes of Sgls cause the formation of blebs prior to lysis, but the location of the blebs differs significantly. In addition, we show that L and other type II Sgls do not inhibit net synthesis of PG, as measured by incorporation of ³ [H]-diaminopimelic acid. Finally, we provide support for the unexpected finding by Adler and colleagues that the Sgl from Pseudomonas phage PP7 is a type I Sgl, as determined by the two methods. This shows that the sharing the putative 4-domain structure suggested for L is not a reliable discriminator for operational characterization of Sgls. Overall, this study establishes new ways to rapidly classify novel Sgls and thus may facilitate the identification of new cell envelope targets that will help generate new antibiotics.
... Previous electron microscopy studies using immunogold staining localized MS2-L in bacterial cell membranes and to a major part within membrane adhesion sites [5] . An early postulation was the oligomerization of the MS2-L transmembrane domain [6] . Furthermore, lysis as a result of inducing the bacterial autolytic system was discussed [1,7] . ...
... Expression of the C-terminal MS2-L transmembrane domain is sufficient to induce bacterial lysis and leakage in liposomes as well as in E. coli membrane vesicles. This was shown with synthetic peptides covering the C-terminal 25 amino acids of MS2-L [3,6] . A second and potentially simultaneous mechanism is the activation of the bacterial autolytic system by MS2-L [1,29] . ...
... By combining both mechanisms, a two-step process was suggested where membrane interaction of MS2-L first results in membrane depolarization. This triggers the activation of autolytic enzymes, which then locally generate large holes in the murein sacculus [6] . The biological relevance for such a dual action of MS2-L could be a more efficient release of the approx. ...
Article
Full-text available
Background: The peptide MS2-L represents toxins of the ssRNA Leviviridae phage family and consists of a predicted N-terminal soluble domain followed by a transmembrane domain. MS2-L mediates bacterial cell lysis through the formation of large lesions in the cell envelope, but further details of this mechanism as a prerequisite for applied bioengineering studies are lacking. The chaperone DnaJ is proposed to modulate MS2-L activity, whereas other cellular targets of MS2-L are unknown. Methods: Here, we provide a combined in vitro and in vivo overexpression approach to reveal molecular insights into MS2-L action and its interaction with DnaJ. Full-length MS2-L and truncated derivatives were synthesized cell-free and co-translationally inserted into nanodiscs or solubilized in detergent micelles. By native liquid bead ion desorption mass spectrometry, we demonstrate that MS2-L assembles into high oligomeric states after membrane insertion. Results: Oligomerization is directed by the transmembrane domain and is impaired in detergent environments. Studies with truncated MS2-L derivatives provide evidence that the soluble domain acts as a modulator of oligomer formation. DnaJ strongly interacts with MS2-L in membranes as well as in detergent environments. However, this interaction affects neither the MS2-L membrane insertion efficiency nor its oligomerization in nanodisc membranes. In accordance with the in vitro data, the assembly of MS2-L derivatives into large membrane located clusters was monitored by overexpression of corresponding fusions with fluorescent monitors in E. coli cells. Analysis by cryo-electron microscopy indicates that lesion formation is initiated in the outer membrane, followed by disruption of the peptidoglycan layer and disintegration of the inner membrane. Conclusion: MS2-L forms oligomeric complexes similar to the related phage toxin ΦX174-E. The oligomeric interface of both peptides is located within their transmembrane domains. We propose a potential function of the higher-order assembly of small phage toxins in membrane disintegration and cell lysis.
... Based on the catalytic domain, five classes of lytic enzymes are distinguished, i.e., glucosaminidase (cut bonds between murein amino sugars), muramidase, amidase (cut amide bonds between the glycan and the peptide bridge), endopeptidase (cut peptide bridges) and lytic transglycosylase [67,68]. Spanins are phage lysis proteins required to disrupt the outer membrane in the final step of Gram-negative host lysis [69]. Among the lytic proteins, an internal virion protein with endolysin domain was annotated, which forms the inner core of the virion and a channel in the host cell envelope, enabling the introduction of phage DNA [70]. ...
Article
Full-text available
Due to the high microbiological contamination of raw food materials and the increase in the incidence of multidrug-resistant bacteria, new methods of ensuring microbiological food safety are being sought. One solution may be to use bacteriophages (so-called phages) as natural bacterial enemies. Therefore, the aim of this study was the biological and genomic characterization of three newly isolated Serratia-and Enterobacter-specific virulent bacteriophages as potential candidates for food biocontrol. Serratia phage KKP_3708 (vB_Sli-IAFB_3708), Serratia phage KKP_3709 (vB_Sma-IAFB_3709), and Enterobacter phage KKP_3711 (vB_Ecl-IAFB_3711) were isolated from municipal sewage against Serratia liquefaciens strain KKP 3654, Serratia marcescens strain KKP 3687, and Enterobacter cloacae strain KKP 3684, respectively. The effect of phage addition at different multiplicity of infection (MOI) rates on the growth kinetics of the bacterial hosts was determined using a Bioscreen C Pro growth analyzer. The phages retained high activity in a wide temperature range (from −20 • C to 60 • C) and active acidity values (pH from 3 to 12). Based on transmission electron microscopy (TEM) imaging and whole-genome sequencing (WGS), the isolated bacteriophages belong to the tailed bacteriophages from the Caudoviricetes class. Genomic analysis revealed that the phages have linear double-stranded DNA of size 40,461 bp (Serratia phage KKP_3708), 67,890 bp (Serratia phage KKP_3709), and 113,711 bp (Enterobacter phage KKP_3711). No virulence, toxins, or antibiotic resistance genes were detected in the phage genomes. The lack of lysogenic markers indicates that all three bacteriophages may be potential candidates for food biocontrol.
... Engineering strategies and challenges of endolysin as an antibacterial agent against Gram-negative bacteria Tianyu Zheng 1 | Can Zhang 2 lysis stage, phages with ssDNA/RNA genome, such as filamentous phages, encode a lysis effector to inhibit the peptidoglycan (PG) biosynthesis of host bacteria (Chamakura & Young, 2019), while most dsDNA phages utilize endolysin to specifically hydrolyse the bacterial PG layer of host bacteria, thereby releasing their progeny. Endolysin has many different names and is summarized elsewhere (Vázquez & Briers, 2023). ...
Article
Full-text available
Bacteriophage endolysin is a novel antibacterial agent that has attracted much attention in the prevention and control of drug‐resistant bacteria due to its unique mechanism of hydrolysing peptidoglycans. Although endolysin exhibits excellent bactericidal effects on Gram‐positive bacteria, the presence of the outer membrane of Gram‐negative bacteria makes it difficult to lyse them extracellularly, thus limiting their application field. To enhance the extracellular activity of endolysin and facilitate its crossing through the outer membrane of Gram‐negative bacteria, researchers have adopted physical, chemical, and molecular methods. This review summarizes the characterization of endolysin targeting Gram‐negative bacteria, strategies for endolysin modification, and the challenges and future of engineering endolysin against Gram‐negative bacteria in clinical applications, to promote the application of endolysin in the prevention and control of Gram‐negative bacteria.
... Endolysins, known as peptidoglycan hydrolases, are a class of enzymes encoded by genes of bacteriophages that degrade the bacterial cell wall at the end of the lytic cycle [94,95]. In Gram-negative hosts, there is often a third protein whose action is required for complete cell lysis, namely spanin, which breaks down the last barrier-the outer membrane [97,98]. Spanins are lysing proteins, essential for disrupting the outer membrane of the bacterial host at the final stage of bacterial lysis [96,99]. ...
Article
Full-text available
The spoilage of juices by Alicyclobacillus spp. remains a serious problem in industry and leads to economic losses. Compounds such as guaiacol and halophenols, which are produced by Alicyclobacillus, create undesirable flavors and odors and, thus, decrease the quality of juices. The inactivation of Alicyclobacillus spp. constitutes a challenge because it is resistant to environmental factors, such as high temperatures, and active acidity. However, the use of bacteriophages seems to be a promising approach. In this study, we aimed to isolate and comprehensively characterize a novel bacteriophage targeting Alicyclobacillus spp. The Alicyclobacillus phage strain KKP 3916 was isolated from orchard soil against the Alicyclobacillus acidoterrestris strain KKP 3133. The bacterial host's range and the effect of phage addition at different rates of multiplicity of infections (MOIs) on the host's growth kinetics were determined using a Bioscreen C Pro growth analyzer. The Alicyclobacillus phage strain KKP 3916, retained its activity in a wide range of temperatures (from 4 • C to 30 • C) and active acidity values (pH from 3 to 11). At 70 • C, the activity of the phage decreased by 99.9%. In turn, at 80 • C, no activity against the bacterial host was observed. Thirty minutes of exposure to UV reduced the activity of the phages by almost 99.99%. Based on transmission-electron microscopy (TEM) and whole-genome sequencing (WGS) analyses, the Alicyclobacillus phage strain KKP 3916 was classified as a tailed bacteriophage. The genomic sequencing revealed that the newly isolated phage had linear double-stranded DNA (dsDNA) with sizes of 120 bp and 131 bp and 40.3% G+C content. Of the 204 predicted proteins, 134 were of unknown function, while the remainder were annotated as structural, replication, and lysis proteins. No genes associated with antibiotic resistance were found in the genome of the newly isolated phage. However, several regions, including four associated with integration into the bacterial host genome and excisionase, were identified, which indicates the temperate (lysogenic) life cycle of the bacteriophage. Due to the risk of its potential involvement in horizontal gene transfer, this phage is not an appropriate candidate for further research on its use in food biocontrol. To the best of our knowledge, this is the first article on the isolation and whole-genome analysis of the Alicyclobacillus-specific phage.
Article
Phage therapy, a promising alternative to combat multidrug-resistant bacterial infections , harnesses the lytic cycle of bacteriophages to target and eliminate bacteria. Key players in this process are the phage lysis proteins, including holin, endolysin, and spanin, which work syner-gistically to disrupt the bacterial cell wall and induce lysis. Understanding the structure and function of these proteins is crucial for the development of effective therapies. Recombinant versions of these proteins have been engineered to enhance their stability and efficacy. Recent progress in the field has led to the approval of bacteriophage-based therapeutics as drugs, paving the way for their clinical use. These proteins can be combined in phage cocktails or combined with antibiotics to enhance their activity against bacterial biofilms, a common cause of treatment failure. Animal studies and clinical trials are being conducted to evaluate the safety and efficacy of phage therapy in humans. Overall, phage therapy holds great potential as a valuable tool in the fight against mul-tidrug-resistant bacteria, offering hope for the future of infectious disease treatment.
Article
Picobirnaviruses (PBVs) are double-stranded RNA viruses frequently detected in human and animal enteric viromes. Associations of PBVs with enteric graft-versus-host disease and type I diabetes during pregnancy have been established. Since their discovery in 1988, PBVs have been generally assumed to be animal-infecting viruses despite the lack of culture system, animal model, or detection in animal cells or tissues. Recent studies have proposed that bacteria or fungi could be the hosts of PBVs based on genomic analysis. Here, we functionally demonstrate that multiple PBVs of different genome organizations encode bacterial lysins that lyse Escherichia coli. Such genes are typically encoded only by bacteriophages supporting the model that PBVs infect bacterial hosts. Recognition of PBVs as RNA phages in the human gut would completely shift models of how PBVs could impact human health. In addition, expanding the RNA phage world beyond the two recognized clades to three clades has implications for our understanding of the evolution of RNA viruses.
Article
Full-text available
The shape, elongation, division and sporulation (SEDS) proteins are a large family of ubiquitous and essential transmembrane enzymes with critical roles in bacterial cell wall biology. The exact function of SEDS proteins was for a long time poorly understood, but recent work has revealed that the prototypical SEDS family member RodA is a peptidoglycan polymerase-a role previously attributed exclusively to members of the penicillin-binding protein family. This discovery has made RodA and other SEDS proteins promising targets for the development of next-generation antibiotics. However, little is known regarding the molecular basis of SEDS activity, and no structural data are available for RodA or any homologue thereof. Here we report the crystal structure of Thermus thermophilus RodA at a resolution of 2.9 Å, determined using evolutionary covariance-based fold prediction to enable molecular replacement. The structure reveals a ten-pass transmembrane fold with large extracellular loops, one of which is partially disordered. The protein contains a highly conserved cavity in the transmembrane domain, reminiscent of ligand-binding sites in transmembrane receptors. Mutagenesis experiments in Bacillus subtilis and Escherichia coli show that perturbation of this cavity abolishes RodA function both in vitro and in vivo, indicating that this cavity is catalytically essential. These results provide a framework for understanding bacterial cell wall synthesis and SEDS protein function.
Article
Full-text available
Yersinia pseudotuberculosis is a Gram-negative bacterium and zoonotic pathogen responsible for a wide range of diseases, ranging from mild diarrhea, enterocolitis, lymphatic adenitis to persistent local inflammation. TheY. pseudotuberculosisinvasin D (InvD) molecule belongs to the invasin (InvA)-type autotransporter proteins, but its structure and function remain unknown. In this study, we present the first crystal structure of InvD, analyzed its expression and function in a murine infection model, and identified its target molecule in the host. We found that InvD is induced at 37°C and expressed in vivo2-4 days after infection, indicating that InvD is a virulence factor. During infection, InvD was expressed in all parts of the intestinal tract, but not in deeper lymphoid tissues. The crystal structure of the C-terminal adhesion domain of InvD revealed a distinct Ig-related fold, that, apart from the canonical β-sheets, comprises various modifications of and insertions into the Ig-core structure. We identified the Fab fragment of host-derived IgG/IgA antibodies as the target of the adhesion domain. Phage display panning and flow cytometry data further revealed that InvD exhibits a preferential binding specificity toward antibodies with VH3/VK1 variable domains and that it is specifically recruited to a subset of B cells. This finding suggests that InvD modulates Ig functions in the intestine and affects direct interactions with a subset of cell surface-exposed B-cell receptors. In summary, our results provide extensive insights into the structure of InvD and its specific interaction with the target molecule in the host.
Article
Full-text available
Translocation of lipid II across the cytoplasmic membrane is essential in peptidoglycan biogenesis. Although most steps are understood, identifying the lipid II flippase has yielded conflicting results, and the lipid II binding properties of two candidate flippases—MurJ and FtsW—remain largely unknown. Here we apply native mass spectrometry to both proteins and characterize lipid II binding. We observed lower levels of lipid II binding to FtsW compared to MurJ, consistent with MurJ having a higher affinity. Site-directed mutagenesis of MurJ suggests that mutations at A29 and D269 attenuate lipid II binding to MurJ, whereas chemical modification of A29 eliminates binding. The antibiotic ramoplanin dissociates lipid II from MurJ, whereas vancomycin binds to form a stable complex with MurJ:lipid II. Furthermore, we reveal cardiolipins associate with MurJ but not FtsW, and exogenous cardiolipins reduce lipid II binding to MurJ. These observations provide insights into determinants of lipid II binding to MurJ and suggest roles for endogenous lipids in regulating substrate binding.
Article
Full-text available
Significance Host lysis and virion assembly are essential processes during the infection cycle of single-stranded RNA (ssRNA) viruses. Using single-particle cryoelectron microscopy, we visualized how the ssRNA virus, Qβ, uses its single-molecule “tail protein,” A 2 , to inhibit MurA, a bacterial enzyme essential for cell wall biosynthesis, leading to lysis of the host cell. We also revealed an extra coat protein dimer, which instead of being a part of the viral capsid, is sequestered within the virion, binding to an RNA hairpin from a five-way junction in the genomic RNA. The same five-way junction also presents hairpins to bind A 2 and other coat protein dimers in the capsid, potentially supporting a nucleation event for virion assembly.
Article
Full-text available
Small single-stranded nucleic acid phages effect lysis by expressing a single protein, the amurin, lacking muralytic enzymatic activity. Three amurins have been shown to act like 'protein antibiotics' by inhibiting cell-wall biosynthesis. However, the L lysis protein of the canonical ssRNA phage MS2, a 75 aa polypeptide, causes lysis by an unknown mechanism without affecting net peptidoglycan synthesis. To identify residues important for lytic function, randomly mutagenized alleles of L were generated, cloned into an inducible plasmid and the transformants were selected on agar containing the inducer. From a total of 396 clones, 67 were unique single base-pair changes that rendered L non-functional, of which 44 were missense mutants and 23 were nonsense mutants. Most of the non-functional missense alleles that accumulated in levels comparable to the wild-type allele are localized in the C-terminal half of L, clustered in and around an LS dipeptide sequence. The LS motif was used to align L genes from ssRNA phages lacking any sequence similarity to MS2 or to each other. This alignment revealed a conserved domain structure, in terms of charge, hydrophobic character and predicted helical content. None of the missense mutants affected membrane-association of L. Several of the L mutations in the central domains were highly conservative and recessive, suggesting a defect in a heterotypic protein-protein interaction, rather than in direct disruption of the bilayer structure, as had been previously proposed for L.
Article
Full-text available
The L protein of the single-stranded RNA phage MS2 causes lysis of Escherichia coli without inducing bacteriolytic activity or inhibiting net peptidoglycan (PG) synthesis. To find host genes required for L-mediated lysis, spontaneous Ill (insensitivity to L lysis) mutants were selected as survivors of L expression and shown to have a missense change of the highly conserved proline (P330Q) in the C-terminal domain of DnaJ. In the dnaJP330Q mutant host, L-mediated lysis is completely blocked at 30°C without affecting the intracellular levels of L. At higher temperatures (37°C and 42°C), both lysis and L accumulation are delayed. The lysis block at 30°C in the dnaJP330Q mutant was recessive and could be suppressed by L overcomes dnaJ (Lodj) alleles selected for restoration of lysis. All three Lodj alleles lack the highly basic N-terminal half of the lysis protein and cause lysis ∼20 min earlier than full-length L. DnaJ was found to form a complex with full-length L. This complex was abrogated by the P330Q mutation and was absent with the Lodj truncations. These results suggest that, in the absence of interaction with DnaJ, the N-terminal domain of L interferes with its ability to bind to its unknown target. The lysis retardation and DnaJ chaperone dependency conferred by the nonessential, highly basic N-terminal domain of L resembles the SlyD chaperone dependency conferred by the highly basic C-terminal domain of the E lysis protein of ϕX174, suggesting a common theme where single-gene lysis can be modulated by host factors influenced by physiological conditions. IMPORTANCE Small single-stranded nucleic acid lytic phages (Microviridae and Leviviridae) lyse their host by expressing a single “protein antibiotic.” The protein antibiotics from two out of three prototypic small lytic viruses have been shown to inhibit two different steps in the conserved PG biosynthesis pathway. However, the molecular basis of lysis caused by L, the lysis protein of the third prototypic virus, MS2, is unknown. The significance of our research lies in the identification of DnaJ as a chaperone in the MS2 L lysis pathway and the identification of the minimal lytic domain of MS2 L. Additionally, our research highlights the importance of the highly conserved P330 residue in the C-terminal domain of DnaJ for specific protein interactions.
Article
Full-text available
Peptidoglycan (PG) protects bacteria from osmotic lysis, and its biogenesis is a key antibiotic target. A central step in PG biosynthesis is flipping of the lipid-linked PG precursor lipid II across the cytoplasmic membrane for subsequent incorporation into PG. MurJ, part of the multidrug/oligosaccharidyl-lipid/polysaccharide (MOP) transporter superfamily, was recently shown to carry out this process. However, understanding of how MurJ flips lipid II, and of how MOP transporters operate in general, remains limited due to a lack of structural information. Here we present a crystal structure of MurJ from Thermosipho africanus in an inward-facing conformation at 2.0-Å resolution. A hydrophobic groove is formed by two C-terminal transmembrane helices, which leads into a large central cavity that is mostly cationic. Our studies not only provide the first structural glimpse of MurJ but also suggest that alternating access is important for MurJ function, which may be applicable to other MOP superfamily transporters.
Article
Full-text available
Current knowledge of RNA virus biodiversity is both biased and fragmentary, reflecting a focus on culturable or disease-causing agents. Here we profile the transcriptomes of over 220 invertebrate species sampled across nine animal phyla and report the discovery of 1,445 RNA viruses, including some that are sufficiently divergent to comprise new families. The identified viruses fill major gaps in the RNA virus phylogeny and reveal an evolutionary history that is characterized by both host switching and co-divergence. The invertebrate virome also reveals remarkable genomic flexibility that includes frequent recombination, lateral gene transfer among viruses and hosts, gene gain and loss, and complex genomic rearrangements. Together, these data present a view of the RNA virosphere that is more phylogenetically and genomically diverse than that depicted in current classification schemes and provide a more solid foundation for studies in virus ecology and evolution.
Article
Significance A peptidoglycan cell wall provides bacteria with protection from environmental stresses, and interfering with assembly of the cell wall is among the most effective strategies for antibiotic development. To build a cell wall, bacteria first synthesize lipid II on the inner leaflet of their membrane and then flip it across to the outer leaflet, where it is used to make peptidoglycan. Here, we report the structure of the lipid II flippase MurJ from Escherichia coli , and we use high-throughput mutagenesis to identify functionally important regions of the protein. Together with evolutionary covariation analysis, these data show that MurJ must exist in at least two discrete conformational states, providing a framework for understanding lipid II flipping.
Article
For bacteriophage infections, the cell walls of bacteria, consisting of a single highly polymeric molecule of peptidoglycan (PG), pose a major problem for the release of progeny virions. Phage lysis proteins that overcome this barrier can point the way to new antibacterial strategies (1) , especially small lytic single-stranded DNA (the microviruses) and RNA phages (the leviviruses) that effect host lysis using a single non-enzymatic protein (2) . Previously, the A2 protein of levivirus Qβ and the E protein of the microvirus ϕX174 were shown to be 'protein antibiotics' that inhibit the MurA and MraY steps of the PG synthesis pathway (2-4) . Here, we investigated the mechanism of action of an unrelated lysis protein, Lys(M), of the Escherichia coli levivirus M (5) . We show that Lys(M) inhibits the translocation of the final lipid-linked PG precursor called lipid II across the cytoplasmic membrane by interfering with the activity of MurJ. The finding that Lys(M) inhibits a distinct step in the PG synthesis pathway from the A2 and E proteins indicates that small phages, particularly the single-stranded RNA (ssRNA) leviviruses, have a previously unappreciated capacity for evolving novel inhibitors of PG biogenesis despite their limited coding potential.Lys(M), the lysis protein of the Escherichia coli levivirus M, represents a new 'protein antibiotic' that interferes with the synthesis of peptidoglycan by inhibiting lipid II flipping.