A brief review of general stress responses in Bacteria
Bacterial Stress Responses
A review by:
Udaat Mittu-
Integrated PhD in Biology
Principal Investigator: Dr Sunish Kumar Radhakrishnan
1
Abstract:
Antimicrobial resistance is one of the major contributors
to deaths worldwide. Recent data from WHO suggests a
worrying upwards trend in the lethalities caused by
microorganisms which survive and grow in the presence
of antibiotics. The plan of action then is to prevent the
acquisition of resistance by these organisms so as to
save lives. To that effect, what is required is an
understanding of how pathogens escape death in the
face of antimicrobial stress so that we can devise
methods of countering it. Therein lies the reason behind
this review. Antimicrobials induce stress upon bacteria
and bacteria survive by responding to those stresses. It
would serve us well to understand how bacteria sense
the presence of stress and react accordingly. A variety of
responses are exhibited by bacteria to a variety of
stressors, be it temperature, pH, DNA damage etc. So,
section by section, we shall try to uncover the canonical
stress responses of bacteria and then see how
generation of stress responses connects to the ultimate
problem of development of antimicrobial resistance.
The stress responses:
In this section, a brief introduction to the various stress
responses of bacteria which are known to date will be
provided. Information related to the activation and
sensing mechanisms of these stress responses shall be
discussed. Effector functions shall be avoided given the
brief scope of the review however, in another section,
the involvement of some of these responses in
development of persistence, tolerance or resistance to
antimicrobials shall be tackled. With that, let’s start
exploring the stress responses of bacteria.
Temperature related stress responses The Heat shock response - The heat shock response
is induced by increase in the levels of σ32, a
transcription factor coded by rpoH. It is responsible for
activation of various heat shock proteins that allow a
bacterium to survive conditions of high temperature.
Increased temperature causes a transient 15-20 fold
increase in the levels of σ32 that stabilises to 2-3 fold
higher as compared to a bacterium growing at the
optimum temperature. Levels of σ32 are regulated by 3
different mechanisms - (i) By self complementarity of
rpoH mRNA which blocks its ribosome binding site
(RBS) (ii) by binding to DnaK/DnaJ/GrpE chaperones
(iii) by decreasing stability of σ32 at room temperature.
(1)
(i) At non-stress inducing temperatures, the rpoH mRNA
exists in a secondary structure which blocks the RBS
of the mRNA. There are two 5’ proximal regions of
the mRNA which are responsible for its temperature
based regulation called region A and B. These were
detected via rpoH-lacZ reporter fusions which were
subjected to a deletion analysis where small regions
of the rpoH gene were deleted starting from the 3’
end and step by step moving towards the 5’ end. By
doing so, it was observed that regions A and B, when
deleted, cause rpoH mRNA to become temperature
insensitive and express at high levels at 30ºC. A is
responsible for enhancement of translation whereas
region B is responsible for repression of the mRNA at
lower temperatures. These results were verified by
introducing point mutations in these regions in an
independent study. (2,3) These studies show
therefore that the mRNA is a thermosensor and can
be induced by heat. The main mechanism being that
when temperature increase, the rpoH mRNA melts,
hence revealing the RBS. (11)
(ii) DnaK-DnaJ-GrpE point mutations showed increase in
stability of σ32 at low temperatures hence signifying
that this chaperone is closely involved in regulation of
σ32. (4)
(iii)FtsH was one of the first proteases shown to be
involved in causing cleaving of σ32 at room
temperature by showing that purified FtsH can
degrade σ32 in vitro. It was also shown that a specific
region of σ32 is involved in FtsH based regulation by
introducing frameshift mutations and deletions to
show that σ32 with mutations in this specific region,
labelled the ‘RpoH box’, is stable at room
temperature. (5,6) It was later found that mutations in
this region of σ32 affect its binding to both DnaK and
RNAP. Binding to DnaK leads to inhibition while
binding to RNAP leads to stabilisation. (7,8) Thus,
RNAP and DnaK compete for binding to this region of
σ32 and lead to different consequences.
When a bacterium experiences increased temperatures,
levels of unfolded proteins which can bind to
chaperones increases, thus reducing the availability of
DnaK chaperones which shifts the equilibrium towards
RNAP which promotes stabilisation of σ32. This was
proved by showing that at lower temperatures, just by
increasing the amount of unfolded protein, a heat shock
response could be induced. (9,10) The counterbalancing
is performed by cellular proteases such as FtsH which
prevent over accumulation of σ32 (12) and in this
manner, a two to three fold increase in the levels of σ32
is achieved at higher temperatures.
The gram positive bacterium heat shock response is
different. It involves HrcA and CIRCE (controlling
inverted repeat of chaperone expression). HrcA is
repressed by a HrcA repressor which is bound to GroE
chaperone. When levels of misfolded proteins increases
with increased temperature, amount of GroE which can
remain bound to the repressor decreases and as a
result, the activity of the repressor decreases and HrcA
is activated. (13) Formation of non native proteins
activates the CIRCE regulon. (14) Thus, with increase in
temperature, HrcA is activated and a heat shock
response is produced.
The HrcA/CIRCE regulons are present in a diverse
range of organisms from Chlamydia and Spirochaeta
(15,16) which have both dnaK and groE operons, to ⍺
proteobacteria such as Caulobacter crescentus which
has only the groE operon under the control of CIRCE/
HcrA. (16,17) It is also important to mention here that in
Caulobacter crescentus, the heat shock response is
positively controlled by RpoH and controlled negatively
by HcrA. (17,18)
2
Cold shock response - The response occurs on three
fronts - (i) membrane associated changes (ii) nucleic
acid associated responses (iii) Csp regulon.
(I) The membrane response is associated with
adjustment of lipid composition of a membrane.
Unsaturated fatty acid (UFA) composition increases
upon encounter with lower temperatures. UFAs
reduce the melting point of the membrane and thus
maintain the fluid properties of the lipid bilayer. (19)
Points of differences exist between gram negative
and gram positive organisms. In Escherichia coli,
enzymes such as β-ketoacyl-acyl carrier protein
synthase II is expressed at normal temperatures and
is activated when the temperature falls. Upon
activation it produces cis-11-octadecenic acid (a
UFA), using palmitoleic acid as a substrate. (19-21)
In contrast, in gram positive organisms such as
Bacillus subtilis, the response is cold induced rather
than cold activated. Enzyme desaturase is
expressed from des gene when the temperature
falls. The enzyme inserts a double bond into the acyl
chain of membrane phospholipids hence introducing
UFAs into the membrane in response to decrease in
temperature. (22)
(II) Nucleic acids themselves can serve as
thermosensors by becoming negatively supercoiled
at lower temperatures. (23) In both Escherichia coli
and Bacillus subtilis, transient increase in negative
supercoiling was observed as temperatures fell. This
was established by transforming a plasmid inside
Bacillus subtilis, inducing stress on the organism,
isolating the plasmid and then running it on a gel to
see how the plasmid was altered. It was observed
that negative supercoiling increased in response to
a cold shock. Further, positive supercoiling
increased as a response to heat shock. Nucleic
acids also respond by altering their supertwist in
response to a variety of stresses besides
temperature based stresses. (24) It is obvious that
changes in the nature of folding of the nucleic acids
will have large scale ramifications. Existence of twist
sensitive promoters which respond to such changes
supports this idea e.g. recA has a cold shock
inducible twist sensitive promoter. (23)
(III) Csp (cold shock protein) regulon is spear headed by
CspA protein which was identified in Escherichia coli
and since then has been found to have homologs in
more than 60 species including gram-positive,
thermophilic, gram negative, psychotrophic and
mesophilic bacteria. (25) The response is immediate
and is followed by low temperature adaptation which
allows growth at low temperatures. (26) Induction of
cold shock leads to a transient growth lag phase
during which Csp expression increases dramatically.
This was observed in Escherichia coli by performing
two dimensional gel electrophoresis for total cellular
proteins which were labelled with radioactive
methionine. (S35) The protein levels were observed
at varying time points after the temperature
downshift and it was observed that cold shock
proteins are expressed at a higher rate while other
cellular protein expression is arrested for a short
period, which is followed by cold adaptation where
Csp proteins continue to be expressed alongside
other cellular proteins. (25) The Csps expressed are
divided into two classes, where class I represents
proteins induced dramatically upon cold shock and
class II represents proteins which are present at
37ºC and are marginally upregulated during cold
shock response. (27) The Csp regulon comes under
class I (28-30) while proteins such as RecA come
under class II. (31)
Proteins such as peptidyl prolyl isomerase (better known
as a trigger factor) are important for refolding of proteins
which are damaged due to cold temperatures. This
enzyme is cold induced and its over expression in
Escherichia coli leads to increased viability of cells. (32)
The common theme of temperature related stress
responses is that both responses display transient
increased expression followed by temperature adaptive
behaviour which allows the bacterium to grow at stress
inducing temperatures.
Envelope stress response -
Envelope stress response is generated by 2 different
regulons under the control of two orthogonal factors (I)
σE and (II) CpxAR two component system.
(I) σE - This response is activated when mutations or
events which cause protein misfolding in the
envelope or overproduction of mutated envelope
proteins occur. Misfolding of outer membrane
proteins or mutations in the periplasmic proteins
also produces the same response. (33) The signal
generated by misfolded proteins is transduced
across the inner membrane via a membrane
localised factor called RseA. A 25 fold increase in
σE pathway activity is observed when null mutations
are introduced in rseA gene. After these mutations
are made, σE response becomes insensitive to
envelope stress and is always active.
Overexpression of RseA represses the σE pathway.
(34,35) Another member of this is the RseB protein.
Null mutations in the rseB gene cause only a two to
three fold increase in the σE response and over
expression of RseB suppresses the σE response
but not completely, thus it serves as a molecule
which fine tunes the σE stress response. (36,37)
When σE activating signals are received, DegS
mediated degradation of RseA occurs which
activates the σE pathway completely. (38)
(II) CpxAR - system activates when some sort of
perturbation to envelope protein biogenesis occurs.
Cues which can cause activation of this pathway are
altered pH, improper folding of inner membrane
proteins or absence of phosphatidyl ethanolamine
(PE) in mutants, or perturbation of protein folding
and its localisation to the bacterial envelope. (39-41)
This signal for activation is transferred across the
envelope by the sensor histidine kinase (HK) CpxA
and the response regulator (RR) CpxR. (42-44)
CpxA functions as an autokinase, CpxR kinase and
a phosphatase for the CpXR which has been
phosphorylated already. The ratio between CpxA
kinase vs the CpxA phosphatase activity determines
the extent of activation of this pathway. (43) As
CpxR is phosphorylated, it gains the ability to bind to
consensus sequences present upstream of the cpxactivated genes and thus activates the expression of
3
the Cpx mediated regulon. (43,45) Another moiety
known as CpxP exists which when deleted leads to
a constitutive three- to five- fold increase in the
activity of the CpxAR pathway. It serves a similar
function as the RseB molecule in the σE pathway in
that it serves as a method of fine tuning the CpxAR
response. (46,47) CpxP can bind to the sensing
domain of CpxA and maintenance of a CpxA:CpxP
ratio of 1:1 inhibits the auto kinase activity of CpxA
by about 50%. (47,48) This ability allows CpxP to
serve as a repressor molecule for the CpxAR
pathway.
SOS response SOS response occurs in response to DNA damage and
it leads to the activation of greater than 20 genes. (49)
The response consists of two separate components LexA and UmuD. UmuD and LexA share homology in
their carboxy-terminal domains. Both of these proteins
undergo an autodigestion when they are acted upon by
an activated RecA protein which exists on a RecA/
ssDNA nucleoprotein filament where the RecA protein
polymerises. This RecA/ssDNA assembly (RecA*)
serves to activate components under LexA and UmuD.
(50-54) Thus, cellular RecA serves as a sensor for DNA
damage and activates the SOS response.
(I) LexA - The protein LexA serves as an inhibitor to
various components of the SOS response by binding
to their similarity regulatory sequence (SOS box)
which is present close to the promoters of these
genes. The interaction between RecA* and LexA
activates the inherent ability of LexA to autodigest.
(53,54) The autodigestion occurs between the A84G85 residues (55) thus inactivating the repressor
function of the protein and activating all the genes
under its repression. RecBCD enzyme functions in
the case of a double stranded break in the DNA.
This RecBCD complex in tandem with single
stranded binding proteins activates RecA
downstream in the pathway which leads to the same
response. (56) As the level of LexA in the cell
decreases due to autodigestion, various genes
under the repression of LexA get activated.
(II) UmuD - umud and umuc genes exist in an operon
which is under the repression of LexA. (57) Like
LexA, UmuD gets cleaved by RecA*, although at a
slower rate as compared to LexA, but which
nonetheless leads to the formation of UmuD’ that
along with UmuC allows the pathway to perform its
function which is SOS mutagenesis, better known as
translesion synthesis. (50,51,58) It is important to
mention here that uncleaved UmuD in a complex
with UmuC serves an important role in cell cycle
regulation. This complex serves as a checkpoint for
DNA damage and allows repairs to be completed
before the cell can replicate its DNA. (59) UmuD and
UmuD’ in their functional complexes exist as
homodimers which are linked to each other via
Cysteine amino acids present in fixed positions
which form a disulphide bond. (60,61) In the
UmuD2C or UmuD’2C complex, UmuC itself is a
DNA polymerase which can function in translesion
synthesis, however, another polymerase can be
activated by the UmuDC pathway which is the DinB
polymerase. (62-65) dinB gene was found to be
necessary for generating untargeted mutations in
Escherichia coli.(62,63) DinB is an error prone
polymerase which explains why DinB gene is
necessary for mutagenesis. (66) It lacks 3’ to 5’
proofreading exonuclease activity, thus it cannot
proof read the nucleotides it adds. Further, DinB
also has the ability to introduce -1 frameshift
mutations which can lead to large scale changes.
(63,67)
Thus, the SOS response which occurs in response to
DNA damage acts to both repair the DNA, perform bet
hedging strategies such as using an error prone
polymerase to fish for mutations and it also holds
influence over the cell cycle such that it can pause the
cell cycle to ensure proper repair of DNA before the
bacterium proceeds to replicate the DNA.
Oxidative stress response Oxidative stress response is regulated by the OxyR and
SoxRS regulons. The OxyR locus consists of genes
which are induced by H2O2 (68) while the SoxRS locus
consists of genes which are induced by the presence of
O2•- moeity. (69,70)
(I) OxyR - OxyR based sensing occurs via the
formation of a disulphide bond between C199 and
C208 which occurs due to oxidation of OxyR. This
was established via a two pronged approach.
Mutation of either of the two cysteine residues led to
the development of an OxyR protein which was
unable to sense the presence of H2O2 in the
medium. Further, a mass spectrometric analysis was
performed for both the reduced and oxidised forms
of OxyR. Using that data, it was established that the
oxidised form of OxyR had a disulphide bond while
the reduced OxyR had thiol groups at the cysteine
residues. Hence, the mechanism of sensing of H2O2
was deduced. OxyR must be oxidised directly for it
to perform its function. (71) It was shown using
synthetically designed oligonucleotides which were
selected for using electrophoretic mobility shift
assays followed by purification and sequencing via
Sanger’s dideoxy sequencing method that the
oxidised and reduced forms of OxyR bind to
different consensus sequences. The consensus
sequences for both the oxidised and reduced forms
of OxyR were found to be - ATAGnt elements, four of
them present on major grooves for oxidised form,
while for the reduced form, the sequence was two
ATAGnt elements which were on major grooves
separated by one helical turn of the DNA. (72) Thus,
oxidation of OxyR would lead to its activation and
also change its preferred consensus sequence,
hence activating the OxyR regulon.
(II) SoxR - SoxR recognises the O2•- moeity via iron
sulphur clusters.The oxidation of SoxR iron sulphur
cluster from [2Fe-2S]1+ to [2Fe-2S]2+ is how it is
activated. This mechanism was proved by
performing dithionite induced reduction of Fe-SoxR
protein and it was shown that the SoxR was able to
regain its ability to activate transcription after
autoxidation. (73) This finding was bolstered by the
fact that a mutant SoxR which was constitutively
4
active was found to exist in its oxidised form even in
the absence of any stress. (74) It is after its
activation that SoxR can induce the transcription of
genes which exist under the SoxRS regulon.
OxyR and SoxRS exist in a wide variety of organisms.
OxyR homologs have been detected in organisms such
as Erwinia carotovora (first oxyR homolog to be found
by performing DNA hybridisations), Mycobacterium
species, and even Xanthomonas campestris. (75,76,77)
SoxR homologs have been detected in Pseudomonas
aeruginosa, Bordetella pertussis, Vibrio cholerae and
many more bacteria. (78)
OxyR and SoxRS perform many more functions besides
just providing a defense against reactive oxygen
species. OxyR has been shown to provide resistance
against HOCl and organic solvents and even some
reactive nitrogen species. SoxRS has been deemed
important in response to antibiotic related stresses,
where some articles have went on to say that SoxRS
activation could be a comprehensive response against
xenobiotic agents rather than just a defense against O2•-.
(78)
Stringent response -
The stringent response is activated in response to
nutritional starvation and a multitude of other stresses
such as osmotic stress and even exposure to
antimicrobial compounds and xenobiotics. The major
player in the stringent response is the alarmone
(p)ppGpp which is synthesised by 2 categories of
enzymes - (i) RelA/SpoT homologue proteins (RSH) (ii)
small alarmone synthetases (SAS) and small alarmone
hydrolases (79,80)
(I) RSH genes - The RSH gene in Escherichia coli is
RelA and it is under the control of 4 promoters. P3
and P4 of σ54 and P1 and P2 of relA itself. The
homolog for RelA in gram positive bacterium is a
RelA/SpoT enzyme which contains both synthetase
and hydrolase domains. (79) RelA is activated by
different molecular players under different types of
nutrient starvation conditions or even different types
of stresses. CRP can activate relA under carbon
stress, σ54 can activate P3 and P4 of relA when
nitrogen starvation conditions are encountered. HNS can activate relA when heat shock stress is
encountered and rpoS, which codes for σS can
activate relA when osmotic stresses are
encountered. (81-83)
(II) SAS genes - SAS genes are differentially regulated
during growth. This was discovered when sequence
similarity searches were performed in Bacillus
subtilis which was believed to have only one RelA/
SpoT homologue. Two genes, named yjbM and
ywaC, were identified and after deducing their amino
acid sequence, it was found that both of these genes
had similarity to RelA/SpoT (p)ppGpp synthetase
domains. Northern blot analyses revealed that these
genes were differentially expressed during different
phases of growth. (84) Multiple genes such as this
exist in a variety of organisms. For example, relP
exists in Bacillus subtilis which has similar functions.
(79) RelP is under the control of σM induced regulon
which is responsible for generating a response to
envelope stress caused by alkaline shock or by
stress induced by the antimicrobial vancomycin. (85)
Besides these genes, (p)ppGpp production can be
affected by additional control of the alarmone
synthesising enzymes themselves. (p)ppGpp
synthesising enzymes are known to have differing
preferences for their substrate. Some enzymes such as
the RelA of Escherichia coli show a greater preference
for GDP whereas the Rel homologue of Mycobacterium
tuberculosis has a greater preference for GTP. This was
found to occur due to presence of either an EXDD or a
RXKD domain in the substrate binding domain for these
enzymes. By performing a (p)ppGpp synthesis assay
with enzymes with each of these domains, it was found
that an EXDD domain preferred GDP as a substrate and
RXKD was partial towards GTP. (86) In this manner,
substrate preferences can serve as a control on the
activity of these enzymes. Beyond that, the product
(p)ppGpp itself when produced, can promote the
transcription of relA as seen in Escherichia coli, thus
creating a positive feedback loop. (87) However, even
this positive feedback loop can be product selective. It
was observed in Bacillus subtilis that pppGpp can
positively regulate the rel homologue (RelQ), but ppGpp
fails to do so. This occurs because RelQ must form a
tetramer with two pppGpp molecules which occurs via
the interaction of pppGpp molecules to the allosteric
binding site of the enzyme. ppGpp fails to do so. This
was revealed by isolating the tetramer and discerning its
crystal structure. (88)
Thus, (p)ppGpp production is controlled by a variety of
mechanisms and it is understandable why it is so given
that it is involved in large scale responses to a variety of
stresses. These functions shall be discussed in the next
section where we will take a look at second messengers
in bacteria.
General Stress response -
General stress response is under two different master
regulators in Gram positive and Gram negative bacteria.
Gram negative bacteria are under the control of rpoS
gene coded σS factor (Escherichia coli), whereas the
Gram positive bacteria are under the control of sigB
gene coded σB. (Bacillus subtilis) We shall cover in brief
the regulation of these factors in this section and their
functions in the next section where we look at the
master regulators in stress responses.
σS in Escherichia coli σS is a master regulator of the stationary phase of
growth in Escherichia coli and governs a large number
of stationary phase genes. (78) σS controls a different
set of genes than σ70 since a deletion of the rpoS gene
completely removes the expression of σS induced
genes and σ70 is unable to induce the same genes. So,
unlike factors like heat shock factor σ32 which
accumulates to a certain level and competes against
σ70 to change the global expression pattern, σS
controls a different set of genes entirely. (89,90) σS is
regulated tightly in different stages of growth. Its cellular
levels are almost non-existent during the log phase of
growth, however, as various stresses start to
accumulate the levels of σS start to increase. Thus,
conditions encountered during stationary phase of
growth such as nutrient starvation, altered pH of the
5
medium or increased cell density tend to induce
translation of the rpoS mRNA. (89,91-97) Since this
factor responds to so many different kinds of stresses, it
is regulated at multiple levels.
The major mRNA transcript for σS is generated by
rpoSp. The rpoS mRNA is expressed at low levels
during log phase of growth, but as the cells move into
lag phase, the levels of this mRNA starts to increase.
(94,98-101) It was also found that a second messenger
cAMP is a repressor of rpoS transcription as both crp
and cya loss of function mutations increased the levels
of rpoS transcription which was observed using a
rpoS::lacZ fusion. (94) Another second messenger
ppGpp was found to positively affect rpoS transcription.
(92,102) Regulation of rpoS goes beyond just
transcription. Suspicions arose regarding this when
different reports were received for rpoS::lacZ fusions for
transcription and translation. (94,99,103) It was then
discovered that the protein Hfq is required for rpoS
mRNA translation. A small RNA was also found to be
involved in the control of rpoS translation. It turned out to
be DsrA which showed upregulation of rpoS translation
when lower temperatures were experienced by the
bacterium. (104-106)
Another level of control of σS is at the post translational
level. σS can undergo proteolysis. The half life of σS
was estimated to be between 1 to 4 minutes in log
phase cells, but as the cells progress into lag phase, the
stability of σS increased steadily. (91,94) The proteins
which are essential for this are ClpX and RssB. RssB
can bind directly to σS and it acts as a chauffeur which
presents σS to ClpXP for proteolysis. Phosphorylation of
RssB increases the affinity of RssB to σS leading to the
conclusion that environmental cues can alter σS
proteolysis by modulating the binding of RssB to σS.
(91,107,108)
Thus, a master regulator like σS is regulated at multiple
levels and responds to a large variety of stresses, hence
its categorisation as a general stress response. A
general rule of thumb for this response is that it is
generated when the stress increases gradually as
compared to one which can be immediately lethal to the
bacterium. Besides Escherichia coli, σS has been
reported in Pseudomonas species and Legionella
pneumophila. (78)
σB in Bacillus subtilis σB factor is expressed by the sigB gene. It was found by
a large number of independent studies that σB is a
stationary phase master regulator and experiences
increase in intracellular levels in response to a variety of
stresses. Thus, like σS of Escherichia coli, σB is a
general stress response regulator. (109-112) Unlike σS
however, σB has a redundant transcription factor, but
even though a redundant factor exists, making a sigB
null mutant rendered the organism sensitive to stresses
such as ph alterations, heat shock, osmotic and
oxidative stress. (78) Given that σB is a master
regulator, it is obvious that it is under strict regulation.
The main repressor of σB is RsbW protein which
remains bound to σB and it is the release of the sigma
factor from RsbW which allows the induction of general
stress response related genes. RsbW itself is under the
control of RsbV which controls the state of
phosphorylation of RsbW. In normal cells, RsbV itself is
phosphorylated at a conserved serine residue and it
cannot interact with RsbW. So, RsbW is free to interact
with σB and prevent its function. But, when a cell is
under stress, RsbV is dephosphorylated by different
enzymes, which allows RsbV to interact with RsbW and
dephosphorylate it. This releases σB from RsbW and a
stress response is generated. (110,113-115) Thus, the
activation of a stress response relies on the
dephosphorylation of RsbV. Two phosphatases have
been detected for RsbV activation - RsbP, which is
activated when the cell experiences energy stress, and
RsbU, which is activated when the cell experiences
environmental stress. (116-118) Thus, a partner switch
mechanism where a dephosphorylated RsbV forces
RsbW to switch partners is what governs the activation
of σB.
Many other organisms possess a σB homologue such
as Listeria monocytogenes, Staphylococcus aureus and
Mycobacterium tuberculosis. (Mycobacterium has a σF
factor which performs in a manner similar to σB)
(119-121)
Antimicrobial stress response -
Antimicrobial stress responses can be induced by
intended or off-target effects of antimicrobials. The most
general response of bacteria to antimicrobial stress is to
generate persister cells. Persistence is a state of
existence of bacteria where upon being exposed to an
antimicrobial, only a specific subset of the population
survives. What is important to note is that these
persistent cells do not experience an increase in their
minimum inhibitory concentration for that specific
antimicrobial. Thus, persistence is a population level
phenomenon. Regardless of what antimicrobial is used,
persistent cells can recover when the antimicrobial is
removed. (122,123)
Oxidative stress response, rpoS mediated general
stress response and Stringent response are the major
stress responses that have been implicated in
generation of persistence. (123) It has been known for
some time now that antimicrobials can have off-target
effects. A well known example is of β-lactam antibiotics
which can cause production of reactive oxygen species
via a reaction known as the Fenton reaction which goes
2+
3+
.
as follows - Fe + H O
Fe + HO + OH2
2
The generation of ROS can lead to the oxidation of
OxyR and SoxR which activates the oxidative stress
response. Further, general stress responses such as
RpoS mediated stress response can be activated due to
oxidative stress as alluded to earlier in the review. (124)
Activation of the stringent response by antimicrobials
also leads to large scale changes in transcription in
Bacteria. Changes such as reduction in replication rate
of bacteria and reducing the elongation rate of DNA
caused by the stringent response can serve as
mechanisms of persistence. (125) Beyond persistence,
Sturgeon et al went to show that the stringent response
can induce acquisition of resistance in a population by
promoting horizontal gene transfer. Using a PintI1 βgalactosidase reporter (where IntI1 codes for the
integron integrase which allows acquisition or exchange
of antibiotic resistance genes) they observed that
6
deletion of RelA gene caused a reduction in the levels of
IntI1 expression in a biofilm, thus indicating that the
stringent response contributes to the activation integron
integrase and thus, can serve as a method of acquisition
of antimicrobial resistance. (126)
Another response of bacteria to antimicrobial stress can
be a bet hedging strategy, termed ‘Hypermutation’.
Hypermutation is a phenotype of bacteria where upon
exposure to antibiotics, the inherent rate of acquisition of
de novo mutations in bacteria increases in what is called
‘Stress induced mutagenesis’. Until a suitable adaptation
is made, this increased rate of mutation persists, and
when the adaptation is complete, normal rates of
mutation are restored via anti-mutator mechanisms. An
obvious mechanism of producing such mutations exists
in the SOS response which depends on the activation of
an error prone DinB polymerase which can induce such
mutations. There also exist permanently hypermutable
strains which contain mutator alleles which have
mutations in genes responsible for making DNA repairs
such as mutS and mutY both of which are involved in
mutability repair of DNA. Such temporary hypermutators
and permanent mutators have been isolated from
clinically relevant pathogenic strains such as
Streptococcus pneumoniae and Escherichia coli.
(127,128)
Thus, a wide variety of responses can be generated in
response to antimicrobial based stresses, ranging from
passive persistence to active mechanisms which can
generate resistance.
With this, the general stress responses of bacteria have
been covered. It is time then to take a closer look at the
molecular players which activate the effector enzymes
and molecules that produce these responses.
The molecular players:
First, we shall learn about the major master regulators
which cause widespread changes in bacterial
metabolism during induction of stress, followed by the
myriad of second messenger molecules which have
been shown to cause large scale alterations themselves.
In favour of the limited scope, the major effectors under
these master regulators will be covered and only their
large scale effects will be mentioned.
The master regulators OxyR and SoxR The proteins under the OxyR and SoxR regulons do not
overlap much even though both of them respond to
oxidative stress. One of the proteins that is activated by
both OxyR and SoxR is the fur gene which expresses a
repressor of Fe3+ ion uptake, hence protecting the cell
from OH• radicals. (127) Beyond that, the proteins under
the two regulons vary, but most of them are antioxidants.
Some examples of antioxidants under OxyR are katG
(hydroperoxidase I), ahpCF (alkyl hydroperoxide
reductase), gorA (glutathione reductase) and grxA
(glutaredoxin). (68) OxyR also activates OxyS which is a
small RNA which protects the cell against mutagenesis
and also acts as a repressor for rpoS and fhlA. In
contrast, SoxR activates a different set of genes. SodA
(superoxide dismutase) represents the major part of the
antioxidant arsenal of SoxR regulon. Other members of
the regulon include nfo (DNA repair enzyme) and O2•resistant enzymes such as acnA product aconitase.
(69,70,128) Besides, ROS, SoxR is also involved in
protecting the bacterium against nitro compounds by
expressing nitroreductase A. SoxR also activates efflux
pumps via the activation of acrAB gene coded drug
efflux pump and tolC encoded OMP. (129-131) Thus,
both OxyR and SoxR induce a wide variety of genes.
However, there are some genes which protect the
bacterium against oxidative damage which are clearly
not under the control of OxyR and SoxR. Some of these
genes are katE which expresses hydroperoxidase II,
recA, mutY and mutM which are involved in DNA
damage control and the sodB and sodC genes which
code for superoxide dismutases themselves. (132-135)
There is also some overlap of the SoxR and OxyR
regulons with the σS regulon. KatG and gorA (OxyR
induced) and acnA (SoxR) are some of the genes which
are activated by σS. σS has been shown to be an
inducer of SodC. (135-137)
σS and σBSince the responses in the general stress response are
diverse, we shall tackle them section by section. σS
response is directly involved in oxidative stress
response as alluded to earlier in this review. The search
for σS induced oxidative stress effectors began when it
was discovered that rpoS null mutants were sensitive to
hydrogen peroxide. (98) All the genes conferring
oxidative stress resistance under the control of σS have
already been covered in the OxyR and SoxR regulon
section.
A large part of the general response is spearheaded by
the production of trehalose by the oprAB operon, which
serves as a stress resistor against osmotic, desiccation
and heat shock stresses. (138-140) The ecp-htrE operon
also provides similar protections to the bacterium. (141)
The σS response can also code for GABA antiporter by
inducing gadC and gadB which allows a bacterium to
tolerate altered pH. (142) Besides just protection against
stress, the general stress response can also induce
apoptosis of a bacterium in a population such that the
nutrients from the dead cell can support the cells nearby.
The ecnAB operon can cause such a response. (143)
Pathogenic bacteria experience decreased virulence
when the rpoS is repressed. Further, specific cases
have been registered where σS is required for a
pathogen to survive in the host or remain virulent. An
example of that would be the esp genes which are
required by enteropathogenic and enterohaemorrhagic
Escherichia coli to express a type III toxin/antitoxin
system. (91,144)
Like σS, σB also penetrates various stress responses. It
induces class II genes of the heat shock response and
some class III genes such as clpP and clpC. Further, it
controls genes such as KatB and KatX which are
catalases that are expressed during starvation and
oxidative stress response. Beyond that, σB also
supports the osmotic stress response by inducing opuE
gene and protects against desiccation by expressing
GsiB. (145-147) An important function of σB response is
causing drug efflux by activating the bmr operon genes
7
bmrU, bmr and bmrR which code for an efflux
transporter that removes a diverse range of drugs such
as Rifampin and Streptomycin which have been shown
to induce σB response. (148-150)
Second messengers Second messengers are molecules which transfer
signals from a receptor to a target. These molecules
diffuse rapidly in the cytoplasm of the bacterium and
help in mounting a large scale response to various
environmental cues. Here we shall cover the major
second messenger molecules which are involved in
stress based responses in bacteria.
(p)ppGpp (p)ppGpp was discovered by Cashel and Gallant as a
‘magic spot’ on an autoradiogram of a starved
Escherichia coli cell. They exposed Escherichia coli cells
to radioactive phosphorus (P32) and compared acid
soluble substrates separated via thin layer
chromatography and saw that a spot was visible
between GTP and the origin only for cells which were
amino acid starved. This magic spot was later found to
be Guanosine nucleotides pppGpp and ppGpp. (151)
Since then, various studies have revealed that these
molecules perform functions way beyond just the
stringent response and apparently have their hand in
almost all pockets of the bacterial system. The functions
of (p)ppGpp can be divided into three parts (I) Reprogamming transcription - (p)ppGpp along
with DksA can bind directly to RNA polymerase and
can cause allosteric changes in the RNAP which
affect the stability of the RNAP-promoter complex.
(152,153) By causing changes in the RNAPpromoter complex stability, large scale alterations in
transcription are achieved. Besides that, two
independent studies by Corrigan et al and Zhang et
al showed that (p)ppGpp can interact with various
ribosome-associated GTPases such as Era in both
gram positive and gram negative bacteria.
(154,155) By doing so, (p)ppGpp can affect the
assembly of ribosomes themselves as these
GTPases are essential assembly factors in bacteria.
(156) Beyond that, (p)ppGpp can control the cellular
pools of GTP by inhibiting enzymes such as HprT
and Gmk that contribute to GTP biosynthesis, which
leads to decreased levels of GTP available in the
cell. As a result, reduction in GTP levels can affect
processes such as elongation of DNA replication.
(157) This allows for the perfect segue to the next
point.
(II) Stalling DNA replication - (p)ppGpp molecules can
stall DNA replication in two ways, directly by
inhibiting initiation and indirectly by reducing the
cellular levels of GTP and inhibiting elongation.
(157) DnaG is the target enzyme for (p)ppGpp and
by binding to it, the process of RNA primer
synthesis which is necessary for DNA replication is
inhibited. As alluded to earlier, the process of
elongation is indirectly inhibited by decrease in
cellular levels of GTP. The result is a dramatic
decrease in the growth and multiplication of
bacteria. Whats intriguing about this stalling is that
conventionally, stalled replication forks recruit RecA
for resolution, however, when replication is stalled
by (p)ppGpp, no such recruitment occurs and thus
DNA replication comes to a complete halt. (158)
(III) Affecting various pathways - Beyond the above
two processes, (p)ppGpp is a known contributor to
bacterial persistence, tolerance and resistance. It
was shown in Escherichia coli that a mutation in the
hipA gene resulted in generation of persistence at a
high rate. In an independent study, (p)ppGpp was
found to be necessary for the generation of this
persistence phenotype. (159,160) Studies with
multiple bacteria has also implicated (p)ppGpp in
generation of tolerance. Khakimova et al proved in
Pseudomonas aeruginosa that (p)ppGpp is
necessary for generation of tolerance to both
antibiotics and oxidative stress. They generated
ΔrelAΔspoT cells and observed that cellular
catalases such as katA and katB underperformed. A
related decrease in tolerance was observed when
the bacterium was challenged with colistin and
ofloxacin. (161) Resistance and (p)ppGpp go hand
in hand as well. Earlier in this review, it was
mentioned that the stringent response can induce
horizontal gene transfer as a strategy to acquire a
resistance conferring gene. Beyond that, some
studies on Methicillin-resistant Staphylococcus
aureus demonstrated that induction of (p)ppGpp
can increase the resistance of this bacterium to βlactam antibiotics. (162,163)
Besides the prototypical alarmones, recently pGpp and
(p)ppApp have been discovered. pGpp was discovered
in Bacillus anthracis and using a differential radial
capillary action ligand assay, it was unearthed that pGpp
binds to different targets as compared to (p)ppGpp an
example of which is that pGpp interacts with effectors
involved in purine biosynthesis. (164) In stark contrast,
(p)ppApp was discovered when a type VI toxin-antitoxin
was studied in Pseudomonas aeruginosa. The toxin
Tas1 when invading other bacteria led to the
overproduction of (p)ppApp which depleted the cellular
pools of the target cells, leading to their death. (165)
Cyclic dinucleotides c-di-GMP and c-di-AMP are two molecules which play
extensive roles in bacterial stress responses.
c-di-GMP -The proteins that interact with c-di-GMPs
have c-di-GMP interacting domains. Some of the
important proteins with these domains are Class I and
Class II c-di-GMP riboswitches, proteins with a PilZ
domain and some proteins which have either a GGDEF
or an EAL domain. (166-168) c-di-GMP is responsible
for shifting the lifestyle of the bacterium towards the
formation of a biofilm by binding to YcgR. This YcgR-cdi-GMP complex then interacts with flagellar proteins
FliG and FliM and inhibits the motility of a bacterium
(which was assessed in the study by checking the
motility of the bacterium on soft agar after removing a
known repressor of c-di-GMP, YhjH). (169) Other
effector functions of c-di-GMP include controlling
development, an example of which is Caulobacter
crescentus, a bacterium which has two life stages, a
swarmer and a stalked stage. Some of the proteins
involved in the transition of the bacterium from swarmer
8
to stalked stage are PleD and DgcB, both of which are
diguanylate cyclases i.e. the enzymes which synthesise
c-di-GMP. Another piece of evidence for c-di-GMP as an
effector molecule is that the levels of c-di-GMP are lower
in swarmers as compared to stalked cells. (170)
This second messenger is also a major player in
bacterial virulence. An important example of this is
Clostridium difficile which depends on c-di-GMP to
switch from a motile state to an adhesion state where it
expresses multiple adhesins such that it can attach to
the intestinal wall of the host. (171) On top of that, the
second messenger is also involved in the regulation of
toxin production by this pathogen. c-di-GMP was found
to inhibit a gene called sigD, a gene which when
ectopically expressed, increased the toxin production.
(172) c-di-GMP was also found to be involved in
protection against oxidative stress in Mycobacterium
smegmatis in which it was directly responsible for the
phosphorylation of devR operon which when deleted,
made the bacterium sensitive to oxidative damage.
(173)
c-di-AMP - c-di-AMP on the other hand has only a few
recognised interactions, even fewer that are relevant to
virulence and survival. (174) DarR of Mycobacterium
smegmatis is known to interact with c-di-AMP and is
also involved in cold shock response. (175) A study
showed that deletion of a gene (pdeA) in Streptococcus
mutans increased the amount of c-di-AMP which
promoted biofilm formation. (176) Much is still unknown
about the functions of this second messenger so it is a
ripe area for scientific exploration.
With the molecular players covered, we shall now move
on to the final section in which we look at how stress
response allow bacteria to survive antimicrobial
challenge.
Bacteria Fighting back:
In this part of the review, information regarding how
stress responses manifest in the form antimicrobial
persistence, tolerance and resistance will be covered.
Again, given the limited scope, only mechanisms which
are under the control of stress responses will be
covered.
Contribution of stress responses in
development of antimicrobial persistent,
tolerant and resistant bacterium Antimicrobial persistence, tolerance and resistance are
all phenotypes which allow survival of bacteria when
exposed to antimicrobials.
A variety of mechanisms can lead to the development of
these phenotypes. Bacteria which have not been
exposed to an antimicrobial can already be in a persister
or tolerant state by the induction of stress responses as
part of a biofilm which can cause both nutritional and
oxygen deprivation which can in turn lead to the
activation of stringent response and oxygen starvation,
both of which inhibit the active growth. This can make a
bacterium persistent, tolerant or in some cases, even
resistant to an antimicrobial. It was observed in
Pseudomonas aeruginosa that cells which were largely
deprived of oxygen via using oxygen electrodes. When
these biofilms were exposed to antimicrobials such as
tobramycin and ciprofloxacin, cells in early stage
biofilms (4h) were almost completely wiped out as
compared to that of comparatively elder biofilms (48 h)
in which a much larger portion of cells survived.
(177,178) Nutrient starvation can lead to the activation
of stringent response and its effector, the (p)ppGpp
alarmone, which also works to inhibit bacterial growth.
Obviously then, in such a state a bacterium would
become tolerant to antimicrobials as was seen in
Escherichia coli which was shown to be tolerant to βlactam antibiotic exposure when the stringent response
was active. (179,180) Further, (p)ppGpp renders
penicillin binding protein 2 redundant and thus causes
resistance to mecillinam. (181)
Oxidative stress has been implicated as the go to
method of killing by bactericidal antibiotics, many of
which generate OH• radicals that cause large scale
damage to the target bacterium. (182-184) It has been
shown that deletion of genes which respond to oxidative
stress can make a bacterium susceptible to
antimicrobials. (185)
Envelope stress has been registered as a culprit in this
process as well. When this response is active, especially
CpxAR, cells experience reduced susceptibility to a
variety of antimicrobials such as β-lactams,
aminoglycosides and even cationic antimicrobial
peptides. (185,186) This largely occurs because of the
activation of drug efflux pumps that reduce the
intracellular concentration of antimicrobials by removing
them actively. (185,187,188) In a comprehensive study
on two component systems (TCS), it was proved that
overexpression of 15 of the TCSs led to multi drug
resistance in Escherichia coliI. Some examples of such
TCSs are baeR, soxR, ompR and RcsB, all of which led
to upregulation of drug efflux related genes and thus led
to immediate increase in the MIC of the bacterium. (188)
The general stress response regulated by σS and σB is
in on the action as well. A mutation in the sigB gene led
to Bacillus subtilis becoming susceptible to rifampin.
(189) Its overexpression led to resistance to β-lactam
antibiotics in Staphylococcus aureus. (190)
σE was shown to restore some tolerance to ofloxacin in
Pseudomonas aeruginosa when it was overexpressed in
a ΔLasIR strain (suppresses quorum sensing and makes
the bacterium highly sensitive to ofloxacin). (191)
Even beyond all of these, stress induced mutagenesis
(hypermutation) is a major source of development of
antimicrobial resistant mutations which are heritable.
This follows a set pattern where antimicrobial stress
leads to DNA damage which activates SOS response,
along with σS based general and σE based envelope
stress response, all of which lead to the use of an error
prone DNA polymerase for repair (DinB - SOS; DNA Pol
II and IV - σS) which leads to a rapid and transient
increase in mutation rate which increases the probability
of achieving a mutation which can make the bacterium
resistant to the antimicrobial responsible for the stress.
(192-194)
Hence, stress responses to a large degree contribute to
the development of persistent, tolerant and resistant
phenotypes across a large variety of bacteria.
9
Specific contribution of oxidative stress
induced by antimicrobials to development of
resistance It has been known for a while that bactericidal drugs can
kill a bacterium in ways other than just the intended
method. A study in 2007 by Kohanski et al argued that
ROS mediated killing is a common mechanism of action
in all bactericidal antimicrobials. (195) A mechanism for
how ROS could induce cell death was known by that
time. Imlay et al had already described in 1988 that
hydroxyl radicals can kill a bacterium by causing
oxidative damage, affecting Fe-S clusters and even
disrupting the TCA cycle, all of which combined can
overwhelm a bacterium and lead to its death. (196)
Studies arguing Kohanski's hypothesis arrived quickly
which grew bacteria in anoxic conditions and tested the
lethality of antimicrobials and found that even in the
absence of ROS, bacteria were still killed, which was the
anti-thesis of Kohanski’s idea. (197) In any case, what is
important that ROS can be generated by antimicrobials
and can contribute to both bacterial death and
development of resistance. Multiple studies made the
argument that antimicrobials employ multiple
mechanisms by which they cause cell death, of which,
ROS generation is significant. (198,199) A recent article
by Hong et al confirmed this significance by measuring
cell death occurring in bacteria after the antibiotic stress
was removed. To prove that it was ROS which was
responsible for bacterial death, they supplemented the
growth medium with bipyridyl which prevents the
generation of ROS. The supplementation by bipyridyl led
to a 30-fold increase in the survival rate of the bacterium
as compared to the control, all in the absence of the
original stressor, nalidixic acid. (200) Thus, ROS
generation plays a role in bactericidal activity. Beyond
just the oxidative damage, ROS can lead to DNA
damage and can lead to double stranded DNA breaks. It
is a major inducer of the SOS response, which itself
employs RecBCD for repair of double stranded break
repair which can be unsuccessful and hence lead to
lethality. (201,202)
How does all of this contribute to the development of
antimicrobial resistance? Oxidative stress can activate
oxidative stress response, the general stress response
and SOS response as mentioned in the review already.
These stress responses themselves in the previous
section have been implicated in multiple mechanisms of
generation of persistence, tolerance and resistance by
using strategies such as drug efflux pumps and by
inducing hypermutation via the use of error prone DNA
polymerases. It is obvious then how oxidative stress
response could be connected to generation of
antimicrobial resistance. There is a glaring lack of
studies which indicate this correlation however, so this
still remains a hypothesis. It leaves a hole in the
literature which needs to be plugged.
The epigenetic perspective on generation of
antimicrobial resistance -
were too high to just occur via spontaneous mutations,
bolstered by the information that said resistance can
revert and the bacterium can become sensitive, which is
the tell tale sign of an epigenetic alteration. It was shown
in a study by Adam et al that low level exposure to
antibiotics led to change in the expression of multiple
genes which led to increase in antibiotic resistance and
that the survival rate increase by about 5 fold when Dam
methylase, an enzyme which methylates the adenine of
GATC, was upregulated. (203) This phenomenon was
named adaptive resistance. Since then, some studies
have tried to look at the epigenetic basis of acquisition of
antimicrobial resistance. A study by Cohen et al showed
that removal of Dam methylase from Escherichia coli
made it sensitive to the action of β-lactam antibiotics.
(204) There is a lack of studies which explore the impact
of epigenetics and its contribution to rapid development
of resistance, and thus, there lies another area where
research could be conducted.
Conclusion and Future perspectives:
The aim of this review was to understand how a
bacterium survives in a world with so many stresses and
fluctuations. Section by section, we explored the
canonical stress responses, followed by the master
regulators and effectors employed in the generation of
said responses followed by an obstacle which will
become more daunting in modern times, the problem of
bacterial persistence, tolerance and resistance. We have
achieved some level of understanding of how bacteria
respond to stress and how those stress responses allow
bacteria to escape antimicrobials. Having said that,
there are still a lot of questions which remain
unanswered and thus, a lot of work remains to be done.
Questions ranging from how (p)ppGpp synthesis works
in its entirety to how oxidative stress can contribute to
the generation of antimicrobial resistance remain
unanswered. Beyond that, an entire field of epigenetic
changes in bacteria remains somewhat unexplored even
though there is some evidence which suggests that
epigenetics might be involved in the rapid escape of
bacteria from the wrath of antimicrobials.
A huge question which the literature shows no instance
of addressing is, how exactly is the presence of
antimicrobials sensed? What characteristic of
antimicrobials awakens the bacterium which then begins
to retaliate? What are the molecules involved in such a
sensory process? Answering these questions could
revolutionise the field of drug development as it could
provide us with a roadmap of what pitfalls to avoid when
making a drug and thus kill bacteria while they are none
the wiser.
With all said and done, it is imperative to reiterate the
fact that research in acquisition of antimicrobial
resistance is of extreme importance. With antimicrobial
resistant pathogen related deaths trending upwards, a
hypothetical clock is ticking and we must endeavour to
have all the answers and a plan of action ready before
the alarm rings.
Epigenetic inheritance was birthed as an idea when it
was observed that the rates of acquisition of resistance
10
Lab report:
Background:
In Caulobacter crescentus, licensing of flagellation occurs via a regulator called MadA (motility and division
regulator) which binds directly to FlhA in the flagellar basal body and releases a protein called FliX which binds to
FlbD. This FliX-FlbD interaction allows the hierarchical process of flagellar biogenesis to move from class-II
components to class-III components. (See schematic below) This is supported by the fact that in ΔmadA cells, the
levels of class-II components (such as FliF and FliX) increases whereas the levels of class-III components (such as
FlgE and FlgI) decreases. It is believed that the tethering of FliX to FlhA is what keeps it inactive and that its
subsequent removal from FlhA due to the binding of MadA to FlhA introduces a conformational change in FliX which
activates it and allows it to activate FlbD and thus, the genes downstream. Until now however, the conformational
change which leads to the activation of FliX, hence leading to its binding to FlbD remains unknown. (205) The aim of
the project is to elucidate this conformational change that occurs in FliX which leads to its activation.
Regulation of Flix-FlbD activity by MadA during flagellar assembly.
Reprinted from ‘An organelle-tethering mechanism couples flagellation to cell division in
bacteria’ by Siwach et al, Developmental Cell, 2021
Objectives of the project:
1. To develop a mutant FliX which can activate FlbD without the intervention of MadA.
2. To isolate and crystallise the FliX mutant and study its structure.
Developing a FliX mutant:
A WT FliX gene was ligated to pmT335 high copy number plasmid which has a Gentamicin resistance marker.
The plasmid is passaged through Escherichia coli XL red cells which are ΔmutHLS (error prone replication due to
impaired mismatch repair.) By doing so, random mutations shall be introduced in the plasmid due to the error
prone DNA replication.
11
These plasmids harbouring random mutations would be transformed into ΔmadA and ΔmadAΔfliX Caulobacter
crescentus cells.
The cells carrying these mutant plasmids shall then be screened for motility by performing a soft agar motility
assay and looking for colonies which show the development of swarmer cells.
In case such a colony is observed, the plasmid would then be isolated from it, transformed into ΔmadA and
ΔmadAΔfliX and observed under the microscope to look for motile cells. If motility is observed, the plasmids are to
be sent for sequencing.
References:-.
Schumann, W. (2016). Regulation of bacterial heat shock stimulons. In Cell Stress and Chaperones (Vol. 21, Issue 6, pp.
959–968). Cell Stress and Chaperones. https://doi.org/10.1007/s--z
Nagai, H., Yuzawa, H., & Yura, T. (1991). Interplay of two cis-acting mRNA regions in translational control of σ32 synthesis
during the heat shock response of Escherichia coli. Proceedings of the National Academy of Sciences of the United
States of America, 88(23),-. https://doi.org/10.1073/pnas-
Yuzawa, H., Nagai, H., Mori, H., & Yura, T. (1993). Heat induction of θ32 synthesis mediated by mRNA secondary
structure: A primary step of the heat shock response in Escherichia coli. Nucleic Acids Research, 21(23),-.
https://doi.org/10.1093/nar/-
Straus, D., Walter, W., & Gross, C. A. (1990). DnaK, DnaJ, and GrpE heat shock proteins negatively regulate heat shock
gene expression by controlling the synthesis and stability of σ32. Genes and Development, 4(12 A),-. https://
doi.org/10.1101/gad.4.12a.2202
Nagai, H., Yuzawa, H., Kanemori, M., & Yura, T. (1994). A distinct segment of the σ32 polypeptide is involved in DnaKmediated negative control of the heat shock response in Escherichia coli. Proceedings of the National Academy of
Sciences of the United States of America, 91(22),-. https://doi.org/10.1073/pnas-
Nakahigashi, K., Yanagi, H., & Yura, T. (1995). Isolation and sequence analysis of rpoH genes encoding σ32 homologs
from gram negative bacteria: Conserved mRNA and protein segments for heat shock regulation. Nucleic Acids Research,
23(21),-. /pmc/articles/PMC307394/?report=abstract
McCarty, J. S., Rüdiger, S., Schönfeld, H. J., Schneider-Mergener, J., Nakahigashi, K., Yura, T., & Bukau, B. (1996).
Regulatory region C of the E. coli heat shock transcription factor, σ32, constitutes a DnaK binding site and is conserved
among eubacteria. Journal of Molecular Biology, 256(5), 829–837. https://doi.org/10.1006/jmbi-
Joo, D. M., Nolte, A., Calendar, R., Zhou, Y. N., & Jin, D. J. (1998). Multiple regions on the Escherichia coli heat shock
transcription factor σ32 determine core RNA polymerase binding specificity. Journal of Bacteriology, 180(5),-.
https://doi.org/10.1128/jb-
Bukau, B. (1993). Regulation of the Escherichia coli heat shock response. Molecular Microbiology, 9(4), 671–680. https://
doi.org/10.1111/j-.tb01727.x
Goff, S. A., & Goldberg, A. L. (1985). Production of abnormal proteins in E. coli stimulates transcription of ion and other
heat shock genes. Cell, 41(2), 587–595. https://doi.org/10.1016/S-
Morita, M. T., Tanaka, Y., Kodama, T. S., Kyogoku, Y., Yanagi, H., & Yura, T. (1999). Translational induction of heat shock
transcription factor σ32: Evidence for a built-in RNA thermosensor. Genes and Development, 13(6), 655–665. https://
doi.org/10.1101/gad.13.6.655
Kanemori, M., Yanagi, H., & Yura, T. (1999). Marked instability of the σ32 heat shock transcription factor at high
temperature. Implications for heat shock regulation. Journal of Biological Chemistry, 274(31),-. https://
doi.org/10.1074/jbc-
Mogk, A., Homuth, G., Scholz, C., Kim, L., Schmid, F. X., & Schumann, W. (1997). The GroE chaperonin machine is a
major modulator of the CIRCE heat shock regulon of Bacillus subtilis. EMBO Journal, 16(15),-. https://doi.org/
10.1093/emboj/-
Mogk, A., Völker, A., Engelmann, S., Hecker, M., Schumann, W., & Völker, U. (1998). Nonnative proteins induce
expression of the Bacillus subtilis CIRCE regulon. Journal of Bacteriology, 180(11),-. https://doi.org/10.1128/
jb-
Hecker, M., Schumann, W., & Völker, U. (1996). Heat-shock and general stress response in Bacillus subtilis. In Molecular
Microbiology (Vol. 19, Issue 3, pp. 417–428). Blackwell Publishing Ltd. https://doi.org/10.1046/j-.x
Segal, G., & Ron, E. Z. (1996). Regulation and organization of the groE and dnaK operons in Eubacteria. FEMS
Microbiology Letters, 138(1), 1–10. https://doi.org/10.1111/j-.tb08126.x
Roberts, R. C., Toochinda, C., Avedissian, M., Baldini, R. L., Gomes, S. L., & Shapiro, L. (1996). Identification of a
Caulobacter crescentus operon encoding hrcA, involved in negatively regulating heat-inducible transcription, and the
chaperone gene grpE. Journal of Bacteriology, 178(7),-. https://doi.org/10.1128/jb-
12
‐
1.
18.
Baldini, R. L., Avedissian, M., & Gomes, S. L. (1998). The CIRCE element and its putative repressor control cell cycle
expression of the Caulobacter crescentus groESL operon. Journal of Bacteriology, 180(7),-. https://doi.org/
10.1128/jb-. Sinensky, M. (1974). Homeoviscous adaptation: a homeostatic process that regulates the viscosity of membrane lipids in
Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America, 71(2), 522–525.
https://doi.org/10.1073/pnas-. Garwin, J. L., & Cronan, J. E. (1980). Thermal modulation of fatty acid synthesis in Escherichia coli does not involve de
novo enzyme synthesis. Journal of Bacteriology, 141(3),-. https://doi.org/10.1128/jb-. Garwin, J. L., Klages, A. L., & Cronan, J. E. (1980). β-Ketoacyl-acyl carrier protein synthase II of Escherichia coli.
Evidence for function in the thermal regulation of fatty acid synthesis. Journal of Biological Chemistry, 255(8),-.
https://doi.org/10.1016/s-. Aguilar, P. S., Cronan, J. E., & De Mendoza, D. (1998). A Bacillus subtilis gene induced by cold shock encodes a
membrane phospholipid desaturase. Journal of Bacteriology, 180(8),-. https://doi.org/10.1128/
jb-. Wang, J. Y, & Syvanen, M. (1992). DNA twist as a transcriptional sensor for environmental changes. In Molecular
Microbiology (Vol. 6, Issue 14, pp-). Mol Microbiol. https://doi.org/10.1111/j-.tb01358.x
24. Krispin, O., & Allmansberger, R. (1995). Changes in DNA supertwist as a response of Bacillus subtilis towards different
kinds of stress. FEMS Microbiology Letters, 134(2–3), 129–135. https://doi.org/10.1111/j-.tb07926.x
25. Berger, F., Normand, P., & Potier, P. (1997). capA, a cspA-like gene that encodes a cold acclimation protein in the
psychrotrophic bacterium Arthrobacter globiformis SI55. Journal of Bacteriology, 179(18),-. https://doi.org/
10.1128/jb-. Panoff, J.-M., Corroler, D., Thammavongs, B., & Boutibonnes, P. (1997). Differentiation between Cold Shock Proteins and
Cold Acclimation Proteins in a Mesophilic Gram-Positive Bacterium, Enterococcus faecalis JH2-2. In JOURNAL OF
BACTERIOLOGY (Vol. 179, Issue 13). http://jb.asm.org/
27. Thieringer, H. A., Jones, P. G., & Inouye, M. (1998). Cold shock and adaptation. BioEssays, 20(1), 49–57. https://doi.org/
10.1002/(SICI-:1<49::AID-BIES8>3.0.CO;2-N
28. Goldstein, J., Pollitt, N. S., & Inouye, M. (1990). Major cold shock protein of Escherichia coli. Proceedings of the National
Academy of Sciences of the United States of America, 87(1), 283–287. https://doi.org/10.1073/pnas-. Lee, S. J., Xie, A., Jiang, W., Etchegaray, J. P, Jones, P. G., & Inouye, M. (1994). Family of the major cold shock protein,
CspA (CS7.4), of Escherichia coli, whose members show a high sequence similarity with the eukaryotic Y box binding
proteins. Molecular Microbiology, 11(5), 833–839. https://doi.org/10.1111/j-.tb00361.x
30. Nakashima, K., Kanamaru, K., Mizuno, T., & Horikoshi, K. (1996). A novel member of the cspA family of genes that is
induced by cold shock in Escherichia coli. Journal of Bacteriology, 178(10),-. https://doi.org/10.1128/
jb-. Walker, G. C. (1984). Mutagenesis and inducible responses to deoxyribonucleic acid damage in Escherichia coli. In
Microbiological Reviews (Vol. 48, Issue 1, pp. 60–93). American Society for Microbiology (ASM). https://doi.org/10.1128/
mmbr-. Graumann, P., Schröder, K., Schmid, R., & Marahiel, M. A. (1996). Cold shock stress-induced proteins in Bacillus subtilis.
Journal of Bacteriology, 178(15),-. https://doi.org/10.1128/jb-. Missiakas, D., Betton, J. M., & Raina, S. (1996). New components of protein folding in extracytoplasmic compartments of
escherichia coli SurA, FkpA and Skp/OmpH. Molecular Microbiology, 21(4), 871–884. https://doi.org/10.1046/
j-.x
34. Missiakas, D., Betton, J. M., & Raina, S. (1996). New components of protein folding in extracytoplasmic compartments of
escherichia coli SurA, FkpA and Skp/OmpH. Molecular Microbiology, 21(4), 871–884. https://doi.org/10.1046/
j-.x
35. De Las Peñas, A., Connolly, L., & Gross, C. A. (1997). The σ(E)-mediated response to extracytoplasmic stress in
Escherichia coli is transduced by RseA and RseB, two negative regulators of σ(E). Molecular Microbiology, 24(2), 373–
385. https://doi.org/10.1046/j-.x
36.Missiakas, D., Mayer, M. P., Lemaire, M., Georgopoulos, C., & Raina, S. (1997). Modulation of the Escherichia coli σ(E)
(RpoE) heat-shock transcription-factor activity by the RseA, RseB and RseC proteins. Molecular Microbiology, 24(2), 355–
371. https://doi.org/10.1046/j-.x
37.Raivio, T. L., & Silhavy, T. J. (1999). The σ(E) and Cpx regulatory pathways: Overlapping but distinct envelope stress
responses. Current Opinion in Microbiology, 2(2), 159–165. https://doi.org/10.1016/S-.Rouvière, P. E., De Las Peñas, A., Mecsas, J., Lu, C. Z., Rudd, K. E., & Gross, C. A. (1995). rpoE, the gene encoding the
second heat-shock sigma factor, sigma E, in Escherichia coli. The EMBO Journal, 14(5),-. https://doi.org/10.1002/
j-.tb07084.x
39.Ades, S. E., Connolly, L. E., Alba, B. M., & Gross, C. A. (1999). The Escherichia coli σ(E)-dependent extracytoplasmic stress
response is controlled by the regulated proteolysis of an anti-σ factor. Genes and Development, 13(18),-. https://
doi.org/10.1101/gad-.Danese, P. N., & Silhavy, T. J. (1998). CpxP, a stress-combative member of the Cpx regulon. Journal of Bacteriology, 180(4),
831–839. https://doi.org/10.1128/jb-.Mileykovskaya, E., & Dowhan, W. (1997). The Cpx two-component signal transduction pathway is activated in Escherichia
coli mutant strains lacking phosphatidylethanolamine. Journal of Bacteriology, 179(4),-. https://doi.org/10.1128/
jb-.Danese, P. N., Oliver, G. R., Barr, K., Bowman, G. D., Rick, P. D., & Silhavy, T. J. (1998). Accumulation of the enterobacterial
common antigen lipid II biosynthetic intermediate stimulates degP transcription in Escherichia coli. Journal of Bacteriology,
180(22),-. https://doi.org/10.1128/jb-.Jianming, D., Shiro, luchi, Hoi-Shan, K., Zhe, L., & Lin, E. C. C. (1993). The deduced amino-acid sequence of the cloned
cpxR gene suggests the protein is the cognate regulator for the membrane sensor, CpxA, in a two-component signal
transduction system of Escherichia coli. Gene, 136(1–2), 227–230. https://doi.org/10.1016/--J
‐
‐
‐
‐
13
44.Raivio, T. L., & Silhavy, T. J. (1997). Transduction of envelope stress in Escherichia coli by the Cpx two- component system.
Journal of Bacteriology, 179(24),-. https://doi.org/10.1128/jb-.Weber, R. F., & Silverman, P. M. (1988). The Cpx proteins of Escherichia coli K12. Structure of the CpxA polypeptide as an
inner membrane component. Journal of Molecular Biology, 203(2), 467–478. https://doi.org/10.1016/-.Pogliano, J., Lynch, A. S., Belin, D., Lin, E. C. C., & Beckwith, J. (1997). Regulation of Escherichia coli cell envelope proteins
involved in protein folding and degradation by the Cpx two-component system. Genes and Development, 11(9),-.
https://doi.org/10.1101/gad-.Raivio, T. L., Popkin, D. L., & Silhavy, T. J. (1999). The Cpx envelope stress response is controlled by amplification and
feedback inhibition. Journal of Bacteriology, 181(17),-. https://doi.org/10.1128/jb-.Fleischer, R., Heermann, R., Jung, K., & Hunke, S. (2007). Purification, reconstitution, and characterization of the CpxRAP
envelope stress system of Escherichia coli. Journal of Biological Chemistry, 282(12),-. https://doi.org/10.1074/
jbc.M-.Friedberg, E. C., Walker, G. C., Siede, W., Wood, R. D., Schultz, R. A., & Ellenberger, T. (2005). DNA Repair and
Mutagenesis. In DNA Repair and Mutagenesis. ASM Press. https://doi.org/10.1128/-.Burckhardt, S. E., Woodgate, R., Scheuermann, R. H., & Echols, H. (1988). UmuD mutagenesis protein of Escherichia coli:
Overproduction, purification, and cleavage by RecA. Proceedings of the National Academy of Sciences of the United States
of America, 85(6),-. https://doi.org/10.1073/pnas-.Shinagawa, H., Iwasaki, H., Kato, T., & Nakata, A. (1988). RecA protein-dependent cleavage of UmuD protein and SOS
mutagenesis. Proceedings of the National Academy of Sciences of the United States of America, 85(6),-. https://
doi.org/10.1073/pnas-.Sassanfar, M., & Roberts, J. W. (1990). Nature of the SOS-inducing signal in Escherichia coli. The involvement of DNA
replication. Journal of Molecular Biology, 212(1), 79–96. https://doi.org/10.1016/-.Little, J. W. (1984). Autodigestion of lexA and phage λ repressors. Proceedings of the National Academy of Sciences of the
United States of America, 81(3 I),-. https://doi.org/10.1073/pnas-.Little, J. W. (1993). LexA cleavage and other self-processing reactions. In Journal of Bacteriology (Vol. 175, Issue 16, pp-). J Bacteriol. https://doi.org/10.1128/jb-.Horii, T., Ogawa, T., Nakatani, T., Hase, T., Matsubara, H., & Ogawa, H. (1981). Regulation of SOS functions: Purification of
E. coli LexA protein and determination of its specific site cleaved by the RecA protein. Cell, 27(3 PART 2), 515–522. https://
doi.org/10.1016/-.Anderson, D. G., & Kowalczykowski, S. C. (1998). Reconstitution of an SOS response pathway: Derepression of
transcription in response to DNA breaks. Cell, 95(7), 975–979. https://doi.org/10.1016/S-.Bagg, A., Kenyon, C. J., & Walker, G. C. (1981). Inducibility of a gene product required for UV and chemical mutagenesis in
Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America, 78(9 II). https://doi.org/
10.1073/pnas-.Nohmi, T., Battista, J. R., Dodson, L. A., & Walker, G. C. (1988). RecA-mediated cleavage activates UmuD for mutagenesis:
Mechanistic relationship between transcriptional derepression and posttranslational activation. Proceedings of the National
Academy of Sciences of the United States of America, 85(6),-. https://doi.org/10.1073/pnas-.Opperman, T., Murli, S., Smith, B. T., & Walker, G. C. (1999). A model for a umuDC-dependent prokaryotic DNA damage
checkpoint. Proceedings of the National Academy of Sciences of the United States of America, 96(16),-. https://
doi.org/10.1073/pnas-.Guzzo, A., Lee, M. H., Oda, K., & Walker, G. C. (1996). Analysis of the region between amino acids 30 and 42 of intact
UmuD by a monocysteine approach. Journal of Bacteriology, 178(24),-. https://doi.org/10.1128/
jb-.Lee, M. H., Ohta, T., & Walker, G. C. (1994). A monocysteine approach for probing the structure and interactions of the
UmuD protein. Journal of Bacteriology, 176(16),-. https://doi.org/10.1128/jb-.Brotcorne-Lannoye, A., & Maenhaut-Michel, G. (1986). Role of RecA protein in untargeted UV mutagenesis of bacteriophage
λ: Evidence for the requirement for the dinB gene. Proceedings of the National Academy of Sciences of the United States of
America, 83(11),-. https://doi.org/10.1073/pnas-.Kim, S. R., Maenhaut-Michel, G., Yamada, M., Yamamoto, Y., Matsui, K., Sofuni, T., Nohmi, T., & Ohmori, H. (1997). Multiple
pathways for SOS-induced mutagenesis in Escherichia coli: An overexpression of dinB/dinP results in strongly enhancing
mutagenesis in the absence of any exogenous treatment to damage DNA. Proceedings of the National Academy of Sciences
of the United States of America, 94(25),-. https://doi.org/10.1073/pnas-.Reuven, N. B., Arad, G., Maor-Shoshani, A., & Livneh, Z. (1999). The mutagenesis protein UmuC is a DNA polymerase
activated by UmuD’, RecA, and SSB and is specialized for translesion replication. Journal of Biological Chemistry, 274(45),-. https://doi.org/10.1074/jbc-.Tang, M., Shen, X., Frank, E. G., O’Donnell, M., Woodgate, R., & Goodman, M. F. (1999). UmuD′2C is an error-prone DNA
polymerase, Escherichia coli pol V. Proceedings of the National Academy of Sciences of the United States of America,
96(16),-. https://doi.org/10.1073/pnas-.Wagner, J., Gruz, P., Kim, S.-R., Yamada, M., Matsui, K., Fuchs, R. P. ., & Nohmi, T. (1999). The dinB Gene Encodes a Novel
E. coli DNA Polymerase, DNA Pol IV, Involved in Mutagenesis. Molecular Cell, 4(2), 281–286. https://doi.org/10.1016/
S-.Wood, R. D., & Hutchinson, F. (1984). Non-targeted mutagenesis of unirradiated lambda phage in Escherichia coli host cells
irradiated with ultraviolet light. Journal of Molecular Biology, 173(3), 293–305. https://doi.org/10.1016/-.Christman, M. F., Morgan, R. W., Jacobson, F. S., & Ames, B. N. (1985). Positive control of a regulon for defenses against
oxidative stress and some heat-shock proteins in Salmonella typhimurium. Cell, 41(3), 753–762. https://doi.org/10.1016/
S-.Greenberg, J. T., Monach, P., Chou, J. H., Josephy, P. D., & Demple, B. (1990). Positive control of a global antioxidant
defense regulon activated by superoxide-generating agents in Escherichia coli. Proceedings of the National Academy of
Sciences of the United States of America, 87(16),-. https://doi.org/10.1073/pnas-
14
70.Tsaneva, I. R., & Weiss, B. (1990). soxR, a locus governing a superoxide response regulon in Escherichia coli K-12. Journal
of Bacteriology, 172(8),-. https://doi.org/10.1128/jb-.Zheng, M., Åslund, F., & Storz, G. (1998). Activation of the OxyR transcription factor by reversible disulfide bond formation.
Science,-),-. https://doi.org/10.1126/science-.Toledano, M. B., Kullik, I., Trinh, F., Baird, P. T., Schneider, T. D., & Storz, G. (1994). Redox-dependent shift of OxyR-DNA
contacts along an extended DNA-binding site: A mechanism for differential promoter selection. Cell, 78(5), 897–909. https://
doi.org/10.1016/S-.Gaudu, P., & Weiss, B. (1996). SoxR, a [2Fe-2S] transcription factor, is active only in its oxidized form. Proceedings of the
National Academy of Sciences of the United States of America, 93(19),-. https://doi.org/10.1073/
pnas-.Hidalgo, E., Ding, H., & Demple, B. (1997). Redox signal transduction: Mutations shifting [2Fe-2S] centers of the SoxR
sensor-regulator to the oxidized form. Cell, 88(1), 121–129. https://doi.org/10.1016/S-.Calcutt, M. J., Lewis, M. S., & Eisenstark, A. (1998). The oxyR gene from Erwinia carotovora : cloning, sequence analysis
and expression in Escherichia coli . FEMS Microbiology Letters, 167(2), 295–301. https://doi.org/10.1111/
j-.tb13242.x
76.Loprasert, S., Atichartpongkun, S., Whangsuk, W., & Mongkolsuk, S. (1997). Isolation and analysis of the Xanthomonas alkyl
hydroperoxide reductase gene and the peroxide sensor regulator genes ahpC and ahpF-oxyR-orfX. Journal of Bacteriology,
179(12),-. https://doi.org/10.1128/jb-.Deretic, V., Song, J., & Pagán-Ramos, E. (1997). Loss of oxyR in Mycobacterium tuberculosis. In Trends in Microbiology
(Vol. 5, Issue 9, pp. 367–372). Elsevier Current Trends. https://doi.org/10.1016/S-X-.Kennedy Shriver Institute of Child Health, E., Development Bethesda, H., & Hengge, R. (2011). BACTERIAL STRESS
RESPONSES Gisela Storz Second Edition. http://estore.asm.org
79.Irving, S. E., & Corrigan, R. M. (2018). Triggering the stringent response: Signals responsible for activating (p)ppGpp
synthesis in bacteria. In Microbiology (United Kingdom) (Vol. 164, Issue 3, pp. 268–276). Microbiology Society. https://
doi.org/10.1099/mic-.Liu, K., Bittner, A. N., & Wang, J. D. (2015). Diversity in (p)ppGpp metabolism and effectors. In Current Opinion in
Microbiology (Vol. 24, pp. 72–79). Elsevier Ltd. https://doi.org/10.1016/j.mib-.Metzger, S., Dror, I., Aizenman, E., … G. S.-J. of B., & 1988, undefined. (n.d.). The nucleotide sequence and characterization
of the relA gene of Escherichia coli. Elsevier. Retrieved May 2, 2021, from https://www.sciencedirect.com/science/article/pii/
S-.Nakagawa, A., Oshima, T., & Mori, H. (2006). Identification and characterization of a second, inducible promoter of relA in
Escherichia coli. In Genes Genet. Syst (Vol. 81). https://www.jstage.jst.go.jp/article/ggs/81/5/81_5_299/_article/-char/ja/
83.Brown, D. R., Barton, G., Pan, Z., Buck, M., & Wigneshweraraj, S. (2014). Nitrogen stress response and stringent response
are coupled in Escherichia coli. Nature Communications, 5. https://doi.org/10.1038/ncomms5115
84.Nanamiya, H., Kasai, K., Nozawa, A., Yun, C.-S., Narisawa, T., Murakami, K., Natori, Y., Kawamura, F., & Tozawa, Y. (2007).
Identification and functional analysis of novel (p)ppGpp synthetase genes in Bacillus subtilis. Molecular Microbiology, 67(2),
291–304. https://doi.org/10.1111/j-.x
85.Eiamphungporn, W., & Helmann, J. D. (2008). The Bacillus subtilis σM regulon and its contribution to cell envelope stress
responses. Molecular Microbiology, 67(4), 830–848. https://doi.org/10.1111/j-.x
86.Sajish, M., Kalayil, S., Verma, S. K., Nandicoori, V. K., & Prakash, B. (2009). The significance of EXDD and RXKD motif
conservation in Rel proteins. Journal of Biological Chemistry, 284(14),-. https://doi.org/10.1074/jbc.M-.Shyp, V., Tankov, S., Ermakov, A., Kudrin, P., English, B. P., Ehrenberg, M., Tenson, T., Elf, J., & Hauryliuk, V. (2012). Positive
allosteric feedback regulation of the stringent response enzyme RelA by its product. EMBO Reports, 13(9), 835–839. https://
doi.org/10.1038/embor-.Steinchen, W., Schuhmacher, J. S., Altegoer, F., Fage, C. D., Srinivasan, V., Linne, U., Marahiel, M. A., & Bange, G. (2015).
Catalytic mechanism and allosteric regulation of an oligomeric (p)ppGpp synthetase by an alarmone. Proceedings of the
National Academy of Sciences of the United States of America, 112(43),-. https://doi.org/10.1073/
pnas-.Jishage, M., Iwata, A., Ueda, S., & Ishihama, A. (1996). Regulation of RNA polymerase sigma subunit synthesis in
Escherichia coli: Intracellular levels of four species of sigma subunit under various growth conditions. Journal of Bacteriology,
178(18),-. https://doi.org/10.1128/jb-.Hengge-Aronis, R. (1999). Interplay of global regulators and cell physiology in the general stress response of Escherichia
coli. Current Opinion in Microbiology, 2(2), 148–152. https://doi.org/10.1016/S-.Bearson, S. M. D., Benjamin, W. H., Swords, W. E., & Foster, J. W. (1996). Acid shock induction of RpoS is mediated by the
mouse virulence gene mviA of Salmonella typhimurium. Journal of Bacteriology, 178(9),-. https://doi.org/10.1128/
jb-.Gentry, D. R., Hernandez, V. J., Nguyen, L. H., Jensen, D. B., & Cashel, M. (1993). Synthesis of the stationary-phase sigma
factor σ(s) is positively regulated by ppGpp. Journal of Bacteriology, 175(24),-. https://doi.org/10.1128/
jb-.Jishage, M., & Ishihama, A. (1995). Regulation of RNA polymerase sigma subunit synthesis in Escherichia coli: Intracellular
levels of σ70 and σ38. Journal of Bacteriology, 177(23),-. https://doi.org/10.1128/jb-.Lange, R., & Hengge-Aronis, R. (1994). The cellular concentration of the σ(S) subunit of RNA polymerase in Escherichia coli
is controlled at the levels of transcription, translation, and protein stability. Genes and Development, 8(13),-.
https://doi.org/10.1101/gad-.Muffler, A., Barth, M., Marschall, C., & Hengge-Aronis, R. (1997). Heat shock regulation of σ(S) turnover: A role for DnaK and
relationship between stress responses mediated by σ(S) and σ32 in Escherichia coli. Journal of Bacteriology, 179(2), 445–
452. https://doi.org/10.1128/jb-.Muffler, A., Traulsen, D. D., Lange, R., & Hengge-Aronis, R. (1996). Posttranscriptional osmotic regulation of the σs subunit
of RNA polymerase in Escherichia coli. Journal of Bacteriology, 178(6),-. https://doi.org/10.1128/
jb-
15
97.Sledjeski, D. D., Gupta, A., & Gottesman, S. (1996). The small RNA, DsrA, is essential for the low temperature expression of
RpoS during exponential growth in Escherichia coli. The EMBO Journal, 15(15),-. https://doi.org/10.1002/
j-.tb00773.x
98.Lange, R., & Hengge Aronis, R. (1991). Identification of a central regulator of stationary phase gene expression in
Escherichia coli. Molecular Microbiology, 5(1), 49–59. https://doi.org/10.1111/j-.tb01825.x
99.McCann, M. P., Fraley, C. D., & Matin, A. (1993). The putative σ factor KatF is regulated posttranscriptionally during carbon
starvation. Journal of Bacteriology, 175(7),-. https://doi.org/10.1128/jb-.Mulvey, M. R., Switala, J., Borys, A., & Loewen, P. C. (1990). Regulation of transcription of katE and katF in Escherichia
coli. Journal of Bacteriology, 172(12),-. https://doi.org/10.1128/jb-.Schellhorn, H. E., & Stones, V. L. (1992). Regulation of katF and katE in Escherichia coli K-12 by weak acids. Journal of
Bacteriology, 174(14),-. https://doi.org/10.1128/jb-.Lange, R., Fischer, D., & Hengge-Aronis, R. (1995). Identification of transcriptional start sites and the role of ppGpp in the
expression of rpoS, the structural gene for the σ(s) subunit of RNA polymerase in Escherichia coli. Journal of Bacteriology,
177(16),-. https://doi.org/10.1128/jb-.Loewen, P. C., Von Ossowski, I., Switala, J., & Mulvey, M. R. (1993). KatF (σ(S)) synthesis in Escherichia coli is subject to
posttranscriptional regulation. In Journal of Bacteriology (Vol. 175, Issue 7, pp-). American Society for
Microbiology (ASM). https://doi.org/10.1128/jb-.Brown, L., & Elliott, T. (1996). Efficient translation of the RpoS sigma factor in Salmonella typhimurium requires host factor I,
an RNA-binding protein encoded by the hfq gene. Journal of Bacteriology, 178(13),-. https://doi.org/10.1128/
jb-.Muffler, A., Fischer, D., & Hengge-Aronis, R. (1996). The RNA-binding protein HF-I, known as a host factor for phage Qβ
RNA replication, is essential for rpoS translation in Escherichia coli. Genes and Development, 10(9),-. https://
doi.org/10.1101/gad-.Muffler, A., Traulsen, D. D., Fischer, D., Lange, R., & Hengge-Aronis, R. (1997). The RNA-binding protein HF-I plays a
global regulatory role which is largely, but not exclusively, due to its role in expression of the σ(S) subunit of RNA polymerase
in Escherichia coli. Journal of Bacteriology, 179(1), 297–300. https://doi.org/10.1128/jb-.Schweder, T., Lee, K. H. O., Lomovskaya, O., & Matin, A. (1996). Regulation of Escherichia coli starvation sigma factor (σs)
by ClpXP protease. Journal of Bacteriology, 178(2), 470–476. https://doi.org/10.1128/jb-.Becker, G., Klauck, E., & Hengge-Aronis, R. (1999). Regulation of RpoS proteolysis in Escherichia coli: The response
regulator RssB is a recognition factor that interacts with the turnover element in RpoS. Proceedings of the National Academy
of Sciences of the United States of America, 96(11),-. https://doi.org/10.1073/pnas-.Boylan, S. A., Redfield, A. R., Brody, M. S., & Price, C. W. (1993). Stress-induced activation of the σ(B) transcription factor
of Bacillus subtilis. Journal of Bacteriology, 175(24),-. https://doi.org/10.1128/jb-.Benson, A. K., & Haldenwang, W. G. (1992). Characterization of a regulatory network that controls σ(B) expression in
Bacillus subtilis. Journal of Bacteriology, 174(3), 749–757. https://doi.org/10.1128/jb-.Benson, A. K., & Haldenwang, W. G. (1993). Regulation of σ(B) levels and activity in Bacillus subtilis. Journal of
Bacteriology, 175(8),-. https://doi.org/10.1128/jb-.Boylan, S. A., Rutherford, A., Thomas, S. M., & Price, C. W. (1992). Activation of Bacillus subtilis transcription factor σ(B) by
a regulatory pathway responsive to stationary-phase signals. Journal of Bacteriology, 174(11),-. https://doi.org/
10.1128/jb-.Alper, S., Dufour, A., Garsin, D. A., Duncan, L., & Losick, R. (1996). Role of adenosine nucleotides in the regulation of a
stress-response transcription factor in Bacillus subtilis. Journal of Molecular Biology, 260(2), 165–177. https://doi.org/
10.1006/jmbi-.Benson, A. K., & Haldenwang, W. G. (1993). Bacillus subtilis σB is regulated by a binding protein (RsbW) that blocks its
association with core RNA polymerase. Proceedings of the National Academy of Sciences of the United States of America,
90(6),-. https://doi.org/10.1073/pnas-.Dufour, A., & Haldenwang, W. G. (1994). Interactions between a Bacillus subtilis anti-σ factor (RsbW) and its antagonist
(RsbV). Journal of Bacteriology, 176(7),-. https://doi.org/10.1128/jb-.Vijay, K., Brody, M. S., Fredlund, E., & Price, C. W. (2000). A PP2C phosphatase containing a PAS domain is required to
convey signals of energy stress to the σ(B) transcription factor of Bacillus subtilis. Molecular Microbiology, 35(1), 180–188.
https://doi.org/10.1046/j-.x
117.Voelker, U., Voelker, A., Maul, B., Hecker, M., Dufour, A., & Haldenwang, W. G. (1995). Separate mechanisms activate σ(B)
of Bacillus subtilis in response to environmental and metabolic stresses. Journal of Bacteriology, 177(13),-.
https://doi.org/10.1128/jb-.Yang, X., Kang, C. M., Brody, M. S., & Price, C. W. (1996). Opposing pairs of serine protein kinases and phosphatases
transmit signals of environmental stress to activate a bacterial transcription factor. Genes and Development, 10(18),-. https://doi.org/10.1101/gad-.Becker, L. A., Çetin, M. S., Hutkins, R. W., & Benson, A. K. (1998). Identification of the gene encoding the alternative sigma
factor σ(B) from Listera monocytogenes and its role in osmotolerance. Journal of Bacteriology, 180(17),-. https://
doi.org/10.1128/jb-.Kullik, I., & Giachino, P. (1997). The alternative sigma factor σ(B) in Staphylococcus aureus: Regulation of the sigB operon
in response to growth phase and heat shock. Archives of Microbiology, 167(2–3), 151–159. https://doi.org/10.1007/
s-.Demaio, J., Zhang, Y., Ko, C., Young, D. B., & Bishai, W. R. (1996). A stationary-phase stress-response sigma factor from
Mycobacterium tuberculosis. Proceedings of the National Academy of Sciences of the United States of America, 93(7),-. https://doi.org/10.1073/pnas-.Brauner, A., Fridman, O., Gefen, O., & Balaban, N. Q. (2016). Distinguishing between resistance, tolerance and persistence
to antibiotic treatment. In Nature Reviews Microbiology (Vol. 14, Issue 5, pp. 320–330). Nature Publishing Group. https://
doi.org/10.1038/nrmicro.2016.34
‐
‐
16
123.Eisenreich, W., Rudel, T., Heesemann, J., & Goebel, W. (2021). Persistence of Intracellular Bacterial Pathogens—With a
Focus on the Metabolic Perspective. In Frontiers in Cellular and Infection Microbiology (Vol. 10). Frontiers Media S.A. https://
doi.org/10.3389/fcimb-.Van Acker, H., & Coenye, T. (2017). The Role of Reactive Oxygen Species in Antibiotic-Mediated Killing of Bacteria. Trends
in Microbiology, 25(6), 456–466. https://doi.org/10.1016/j.tim-.Das, B., & Bhadra, R. K. (2020). (p)ppGpp Metabolism and Antimicrobial Resistance in Bacterial Pathogens. In Frontiers in
Microbiology (Vol. 11, p. 563944). Frontiers Media S.A. https://doi.org/10.3389/fmicb-.Strugeon, E., Tilloy, V., Ploy, M. C., & Da Re, S. (2016). The stringent response promotes antibiotic resistance
dissemination by regulating integron integrase expression in biofilms. MBio, 7(4). https://doi.org/10.1128/mBio-.Zheng, M., Doan, B., Schneider, T. D., & Storz, G. (1999). OxyR and SoxRS regulation of fur. Journal of Bacteriology,
181(15),-. https://doi.org/10.1128/jb-.Gruer, M. J., & Guest, J. R. (1994). Two genetically-distinct and differentially-regulated aconitases (AcnA and AcnB) in
Escherichia coli. Microbiology, 140(10),-. https://doi.org/10.1099/-.Liochev, S. I., Hausladen, A., & Fridovich, I. (1999). Nitroreductase A is regulated as a member of the soxRS regulon of
Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America, 96(7),-.
https://doi.org/10.1073/pnas-.Aono, R., Tsukagoshi, N., & Yamamoto, M. (1998). Involvement of outer membrane protein TolC, a possible member of the
mar-sox regulon, in maintenance and improvement of organic solvent tolerance of Escherichia coli K-12. Journal of
Bacteriology, 180(4), 938–944. https://doi.org/10.1128/jb-.Ma, D., Alberti, M., Lynch, C., Nikaido, H., & Hearst, J. E. (1996). The local repressor AcrR plays a modulating role in the
regulation of acrAB genes of Escherichia coli by global stress signals. Molecular Microbiology, 19(1), 101–112. https://
doi.org/10.1046/j-.x
132.Imlay, J. A., & Linn, S. (1988). DNA damage and oxygen radical toxicity. Science Science,-),-. https://
doi.org/10.1126/science-.Demple, B., & Harrison, L. (1994). Repair of Oxidative Damage to DNA: Enzymology and Biology. Annual Review of
Biochemistry, 63(1), 915–948. https://doi.org/10.1146/annurev.bi-.Fridavich, I. (1995). Superoxide radical and superoxide dismutases. In Annual Review of Biochemistry (Vol. 64, pp. 97–
112). Annual Reviews Inc. https://doi.org/10.1146/annurev.bi-.Strohmeier Gort, A., Ferber, D. M., & Imlay, J. A. (1999). The regulation and role of the periplasmic copper, zinc superoxide
dismutase of Escherichia coli. Molecular Microbiology, 32(1), 179–191. https://doi.org/10.1046/j-.x
136.Loewen, P. C., Hu, B., Strutinsky, J., & Sparling, R. (1998). Regulation in the rpoS regulon of Escherichia coli. Canadian
Journal of Microbiology, 44(8), 707–717. https://doi.org/10.1139/cjm-44-8-707Cunningham, L., Gruer, M. J., & Guest, J. R.
(1997). Transcriptional regulation of the aconitase genes (acnA and acnB) of Escherichia coli. Microbiology, 143(12),-. https://doi.org/10.1099/-.Cunningham, L., Gruer, M. J., & Guest, J. R. (1997). Transcriptional regulation of the aconitase genes (acnA and acnB) of
Escherichia coli. Microbiology, 143(12),-. https://doi.org/10.1099/-.Strom, A. R., & Kaasen, I. (1993). Trehalose metabolism in Escherichia coli: stress protection and stress regulation of gene
expression. In Molecular Microbiology (Vol. 8, Issue 2, pp. 205–210). Mol Microbiol. https://doi.org/10.1111/
j-.tb01564.x
139.Hengge-Aronis, R., Klein, W., Lange, R., Rimmele, M., & Boos, W. (1991). Trehalose synthesis genes are controlled by the
putative sigma factor encoded by rpoS and are involved in stationary-phase thermotolerance in Escherichia coli. Journal of
Bacteriology, 173(24),-. https://doi.org/10.1128/jb-.Welsh, D. T., & Herbert, R. A. (1999). Osmotically induced intracellular trehalose, but not glycine betaine accumulation
promotes desiccation tolerance in Escherichia coli . FEMS Microbiology Letters, 174(1), 57–63. https://doi.org/10.1111/
j-.tb13549.x
141.Raina, S., Missiakas, D., Baird, L., Kumar, S., & Georgopoulos, C. (1993). Identification and transcriptional analysis of the
Escherichia coli htrE operon which is homologous to pap and related pilin operons. Journal of Bacteriology, 175(16),-. https://doi.org/10.1128/jb-.Waterman, S. R., & Small, P. L. C. (1996). Identification of σ(S)-dependent genes associated with the stationary-phase
acid-resistance phenotype of Shigella flexneri. Molecular Microbiology, 21(5), 925–940. https://doi.org/10.1046/
j-.x
143.Bishop, R. E., Leskiw, B. K., Hodges, R. S., Kay, C. M., & Weiner, J. H. (1998). The entericidin locus of Escherichia coli and
its implications for programmed bacterial cell death. Journal of Molecular Biology, 280(4), 583–596. https://doi.org/10.1006/
jmbi-.Beltrametti, F., Kresse, A. U., & Guzmán, C. A. (1999). Transcriptional regulation of the esp genes of enterohemorrhagic
Escherichia coli. Journal of Bacteriology, 181(11),-. https://doi.org/10.1128/jb-.Petersohn, A., Engelmann, S., Setlow, P., & Hecker, M. (1999). The katX gene of Bacillus subtilis is under dual control of
σ(B) and σ(F). Molecular and General Genetics, 262(1), 173–179. https://doi.org/10.1007/s-.Engelmann, S., Lindner, C., & Hecker, M. (1995). Cloning, nucleotide sequence, and regulation of katE encoding a σ(B)dependent catalase in Bacillus subtilis. Journal of Bacteriology, 177(19),-. https://doi.org/10.1128/
jb-.Spiegelhalter, F., & Bremer, E. (1998). Osmoregulation of the opuE proline transport gene from Bacillus subtilis :
contributions of the sigma A and sigma B dependent stress responsive promoters. Molecular Microbiology, 29(1), 285–296.
https://doi.org/10.1046/j-.x
148.Ahmed, M., Borsch, C. M., Taylor, S. S., Vázquez-Laslop, N., & Neyfakh, A. A. (1994). A protein that activates expression of
a multidrug efflux transporter upon binding the transporter substrates. Journal of Biological Chemistry, 269(45),-. https://doi.org/10.1016/s-
‐
‐
‐
17
149.Neyfakh, A. A., Bidnenko, V. E., & Lan Bo Chen. (1991). Efflux-mediated multidrug resistance in Bacillus subtilis: Similarities
and dissimilarities with the mammalian system. Proceedings of the National Academy of Sciences of the United States of
America, 88(11),-. https://doi.org/10.1073/pnas-.Michele, T. M., Ko, C., & Bishai, W. R. (1999). Exposure to antibiotics induces expression of the Mycobacterium
tuberculosis sigF gene: Implications for chemotherapy against mycobacterial persistors. Antimicrobial Agents and
Chemotherapy, 43(2), 218–225. https://doi.org/10.1128/aac-.Cashel, M., & Gallant, J. (1969). Two compounds implicated in the function of the RC gene of escherichia coli. Nature,-), 838–841. https://doi.org/10.1038/221838a0
152.Ross, W., Vrentas, C. E., Sanchez-Vazquez, P., Gaal, T., & Gourse, R. L. (2013). The magic spot: A ppGpp binding site on
E. coli RNA polymerase responsible for regulation of transcription initiation. Molecular Cell, 50(3), 420–429. https://doi.org/
10.1016/j.molcel-.Kanjee, U., Ogata, K., & Houry, W. A. (2012). Direct binding targets of the stringent response alarmone (p)ppGpp.
Molecular Microbiology, 85(6),-. https://doi.org/10.1111/j-.x
154.Zhang, Y., Zborníková, E., Rejman, D., & Gerdes, K. (2018). Novel (p)ppGpp binding and metabolizing proteins of
Escherichia coli. MBio, 9(2). https://doi.org/10.1128/mBio-.Corrigan, R. M., Bellows, L. E., Wood, A., & Gründling, A. (2016). PpGpp negatively impacts ribosome assembly affecting
growth and antimicrobial tolerance in Grampositive bacteria. Proceedings of the National Academy of Sciences of the United
States of America, 113(12), E1710–E1719. https://doi.org/10.1073/pnas-.Karbstein, K. (2007). Role of GTPases in ribosome assembly. Biopolymers, 87(1), 1–11. https://doi.org/10.1002/bip-.Denapoli, J., Tehranchi, A. K., & Wang, J. D. (2013). Dose-dependent reduction of replication elongation rate by (p)ppGpp
in Escherichia coli and Bacillus subtilis. Molecular Microbiology, 88(1), 93–104. https://doi.org/10.1111/mmi-.Wang, J. D., Sanders, G. M., & Grossman, A. D. (2007). Nutritional Control of Elongation of DNA Replication by (p)ppGpp.
Cell, 128(5), 865–875. https://doi.org/10.1016/j.cell-.Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L., & Leibler, S. (2004). Bacterial persistence as a phenotypic switch.
Science,-),-. https://doi.org/10.1126/science-.Korch, S. B., Henderson, T. A., & Hill, T. M. (2003). Characterization of the hipA7 allele of Escherichia coli and evidence that
high persistence is governed by (p)ppGpp synthesis. Molecular Microbiology, 50(4),-. https://doi.org/10.1046/
j-.x
161.Khakimova, M., Ahlgren, H. G., Harrison, J. J., English, A. M., & Nguyen, D. (2013). The stringent response controls
catalases in Pseudomonas aeruginosa and is required for hydrogen peroxide and antibiotic tolerance. Journal of
Bacteriology, 195(9),-. https://doi.org/10.1128/JB-.Kim, C., Mwangi, M., Chung, M., Milheirco, C., De Lencastre, H., & Tomasz, A. (2013). The mechanism of heterogeneous
beta-lactam resistance in MRSA: Key role of the stringent stress response. PLoS ONE, 8(12). https://doi.org/10.1371/
journal.pone-.Dordel, J., Kim, C., Chung, M., de la Gándara, M. P., Holden, M. T. J., Parkhill, J., de Lencastre, H., Bentley, S. D., &
Tomasz, A. (2014). Novel determinants of antibiotic resistance: Identification of mutated Loci in Highly methicillin-resistant
subpopulations of methicillin-resistant Staphylococcus aureus. MBio, 5(2). https://doi.org/10.1128/mBio-.Yang, J., Anderson, B. W., Turdiev, A., Turdiev, H., Stevenson, D. M., Amador-Noguez, D., Lee, V. T., & Wang, J. D. (2020).
The nucleotide pGpp acts as a third alarmone in Bacillus, with functions distinct from those of (p) ppGpp. Nature
Communications, 11(1), 1–11. https://doi.org/10.1038/s-.Ahmad, S., Wang, B., Walker, M. D., Tran, H. K. R., Stogios, P. J., Savchenko, A., Grant, R. A., McArthur, A. G., Laub, M. T.,
& Whitney, J. C. (2019). An interbacterial toxin inhibits target cell growth by synthesizing (p)ppApp. Nature,-), 674–
678. https://doi.org/10.1038/s-.Kulshina, N., Baird, N. J., & Ferré-D’Amaré, A. R. (2009). Recognition of the bacterial second messenger cyclic diguanylate
by its cognate riboswitch. Nature Structural and Molecular Biology, 16(12),-. https://doi.org/10.1038/nsmb-.Smith, K. D., Shanahan, C. A., Moore, E. L., Simon, A. C., & Strobel, S. A. (2011). Structural basis of differential ligand
recognition by two classes of bis-(3′-5′)-cyclic dimeric guanosine monophosphate-binding riboswitches. Proceedings of the
National Academy of Sciences of the United States of America, 108(19),-. https://doi.org/10.1073/
pnas-.Jenal, U., Reinders, A., & Lori, C. (2017). Cyclic di-GMP: Second messenger extraordinaire. In Nature Reviews
Microbiology (Vol. 15, Issue 5, pp. 271–284). Nature Publishing Group. https://doi.org/10.1038/nrmicro-.Paul, K., Nieto, V., Carlquist, W. C., Blair, D. F., & Harshey, R. M. (2010). The c-di-GMP Binding Protein YcgR Controls
Flagellar Motor Direction and Speed to Affect Chemotaxis by a “Backstop Brake” Mechanism. Molecular Cell, 38(1), 128–
139. https://doi.org/10.1016/j.molcel-.Abel, S., Bucher, T., Nicollier, M., Hug, I., Kaever, V., Abel zur Wiesch, P., & Jenal, U. (2013). Bi-modal Distribution of the
Second Messenger c-di-GMP Controls Cell Fate and Asymmetry during the Caulobacter Cell Cycle. PLoS Genetics, 9(9),
e-. https://doi.org/10.1371/journal.pgen-.Purcell, E. B., McKee, R. W., McBride, S. M., Waters, C. M., & Tamayo, R. (2012). Cyclic diguanylate inversely regulates
motility and aggregation in clostridium difficile. Journal of Bacteriology, 194(13),-. https://doi.org/10.1128/
JB-.McKee, R. W., Mangalea, M. R., Purcell, E. B., Borchardt, E. K., & Tamayo, R. (2013). The second messenger cyclic DiGMP regulates Clostridium difficile toxin production by controlling expression of sigD. Journal of Bacteriology, 195(22),-. https://doi.org/10.1128/JB-.Hu, Q., Zhang, J., Chen, Y., Hu, L., Li, W., & He, Z. G. (2019). Cyclic di-GMP co-activates the two-component transcriptional
regulator DevR in Mycobacterium smegmatis in response to oxidative stress. Journal of Biological Chemistry, 294(34),-. https://doi.org/10.1074/jbc.RA-.Fahmi, T., Port, G. C., & Cho, K. H. (2017). c-di-AMP: An Essential Molecule in the Signaling Pathways that Regulate the
Viability and Virulence of Gram-Positive Bacteria. Genes, 8(8). https://doi.org/10.3390/genes-.Zhang, L., Li, W., & He, Z. G. (2013). DarR, a TetR-like transcriptional factor, is a cyclic di-AMP-responsive repressor in
Mycobacterium smegmatis. Journal of Biological Chemistry, 288(5),-. https://doi.org/10.1074/jbc.M-
18
176.Peng, X., Zhang, Y., Bai, G., Zhou, X., & Wu, H. (2016). Cyclic di-AMP mediates biofilm formation. Molecular Microbiology,
99(5), 945–959. https://doi.org/10.1111/mmi-.Borriello, G., Werner, E., Roe, F., Kim, A. M., Ehrlich, G. D., & Stewart, P. S. (2004). Oxygen limitation contributes to
antibiotic tolerance of Pseudomonas aeruginosa in biofilms. Antimicrobial Agents and Chemotherapy, 48(7),-.
https://doi.org/10.1128/AAC-.Borriello, G., Richards, L., Ehrlich, G. D., & Stewart, P. S. (2006). Arginine or nitrate enhances antibiotic susceptibility of
Pseudomonas aeruginosa in biofilms. Antimicrobial Agents and Chemotherapy, 50(1), 382–384. https://doi.org/10.1128/
AAC-.Kusser, W., & Ishiguro, E. E. (1985). Involvement of the relA gene in the autolysis of Escherichia coli induced by inhibitors of
peptidoglycan biosynthesis. Journal of Bacteriology, 164(2), 861–865. https://doi.org/10.1128/jb-.Kusser, W., & Ishiguro, E. E. (1987). Suppression of mutations conferring penicillin tolerance by interference with the
stringent control mechanism of Escherichia coli. Journal of Bacteriology, 169(9),-. https://doi.org/10.1128/
jb-.Vinella, D., D’Ari, R., & Bouloc, P. (1992). Penicillin binding protein 2 is dispensable in Escherichia coli when ppGpp
synthesis is induced. EMBO Journal, 11(4),-. https://doi.org/10.1002/j-.tb05194.x
182.Dwyer, D. J., Kohanski, M. A., & Collins, J. J. (2009). Role of reactive oxygen species in antibiotic action and resistance. In
Current Opinion in Microbiology (Vol. 12, Issue 5, pp. 482–489). Curr Opin Microbiol. https://doi.org/10.1016/
j.mib-.Kohanski, M. A., Dwyer, D. J., Wierzbowski, J., Cottarel, G., & Collins, J. J. (2008). Mistranslation of Membrane Proteins
and Two-Component System Activation Trigger Antibiotic-Mediated Cell Death. Cell, 135(4), 679–690. https://doi.org/
10.1016/j.cell-.Kolodkin-Gal, I., Sat, B., Keshet, A., & Engelberg-Kulka, H. (2008). The communication factor EDF and the toxin-antitoxin
module mazEF determine the mode of action of antibiotics. PLoS Biology, 6(12),-. https://doi.org/10.1371/
journal.pbio-.Kuczyńska-Wiśnik, D., Matuszewska, E., Furmanek-Blaszk, B., Leszczyńska, D., Grudowska, A., Szczepaniak, P., &
Laskowska, E. (2010). Antibiotics promoting oxidative stress inhibit formation of Escherichia coli biofilm via indole signalling.
Research in Microbiology, 161(10), 847–853. https://doi.org/10.1016/j.resmic-.Nishino, K., Yamasaki, S., Hayashi-Nishino, M., & Yamaguchi, A. (2010). Effect of NlpE overproduction on multidrug
resistance in Escherichia coli. Antimicrobial Agents and Chemotherapy, 54(5),-. https://doi.org/10.1128/
AAC-.Weatherspoon-Griffin, N., Zhao, G., Kong, W., Kong, Y., Morigen, Andrews-Polymenis, H., McClelland, M., & Shi, Y. (2011).
The CpxR/CpxA two-component system up-regulates two tat-dependent peptidoglycan amidases to confer bacterial
resistance to antimicrobial peptide. Journal of Biological Chemistry, 286(7),-. https://doi.org/10.1074/
jbc.M-.Hirakawa, H., Nishino, K., Hirata, T., & Yamaguchi, A. (2003). Comprehensive studies of drug resistance mediated by
overexpression of response regulators of two-component signal transduction systems in Escherichia coli. Journal of
Bacteriology, 185(6),-. https://doi.org/10.1128/JB-.Bandow, J. E., Brötz, H., & Hecker, M. (2002). Bacillus subtilis tolerance of moderate concentrations of rifampin involves the
σB-dependent general and multiple stress response. Journal of Bacteriology, 184(2), 459–467. https://doi.org/10.1128/
JB-.Morikawa, K., Maruyama, A., Inose, Y., Higashide, M., Hayashi, H., & Ohta, T. (2001). Overexpression of sigma factor, σB,
urges Staphylococcus aureus to thicken the cell wall and to resist β-lactams. Biochemical and Biophysical Research
Communications, 288(2), 385–389. https://doi.org/10.1006/bbrc-.Kayama, S., Murakami, K., Ono, T., Ushimaru, M., Yamamoto, A., Hirota, K., & Miyake, Y. (2009). The role of rpoS gene and
quorum-sensing system in ofloxacin tolerance in Pseudomonas aeruginosa. FEMS Microbiology Letters, 298(2), 184–192.
https://doi.org/10.1111/j-.x
192.Galhardo, R. S., Do, R., Yamada, M., Friedberg, E. C., Hastings, P. J., Nohmi, T., & Rosenberg, S. M. (2009). DinB
upregulation is the sole role of the SOS response in stress-induced mutagenesis in Escherichia coli. Genetics, 182(1), 55–
68. https://doi.org/10.1534/genetics-.Frisch, R. L., Su, Y., Thornton, P. C., Gibson, J. L., Rosenberg, S. M., & Hastings, P. J. (2010). Separate DNA Pol II- and Pol
IV-dependent pathways of stress-induced mutation during double-strand-break repair in escherichia coli are controlled by
RpoS. Journal of Bacteriology, 192(18),-. https://doi.org/10.1128/JB-.Gibson, J. L., Lombardo, M. J., Thornton, P. C., Hu, K. H., Galhardo, R. S., Beadle, B., Habib, A., Magner, D. B., Frost, L.
S., Herman, C., Hastings, P. J., & Rosenberg, S. M. (2010). The σe stress response is required for stress-induced mutation
and amplification in Escherichia coli. Molecular Microbiology, 77(2), 415–430. https://doi.org/10.1111/
j-.x
195.Kohanski, M. A., Dwyer, D. J., Hayete, B., Lawrence, C. A., & Collins, J. J. (2007). A Common Mechanism of Cellular Death
Induced by Bactericidal Antibiotics. Cell, 130(5), 797–810. https://doi.org/10.1016/j.cell-.Imlay, J. A., Chin, S. M., & Linn, S. (1988). Toxic DNA damage by hydrogen peroxide through the fenton reaction in vivo and
in vitro. Science,-), 640–642. https://doi.org/10.1126/science-.Liu, Y., & Imlay, J. A. (2013). Cell death from antibiotics without the involvement of reactive oxygen species. Science,-),-. https://doi.org/10.1126/science-.Dwyer, D. J., Collins, J. J., & Walker, G. C. (2015). Unraveling the physiological complexities of antibiotic lethality. In Annual
Review of Pharmacology and Toxicology (Vol. 55, pp. 313–332). Annual Reviews Inc. https://doi.org/10.1146/annurevpharmtox-.Yang, J. H., Bening, S. C., & Collins, J. J. (2017). Antibiotic efficacy — context matters. In Current Opinion in Microbiology
(Vol. 39, pp. 73–80). Elsevier Ltd. https://doi.org/10.1016/j.mib-.Hong, Y., Zeng, J., Wang, X., Drlica, K., & Zhao, X. (2019). Post-stress bacterial cell death mediated by reactive oxygen
species. Proceedings of the National Academy of Sciences of the United States of America, 116(20),-. https://
doi.org/10.1073/pnas-
19
201.Baquero, F., & Levin, B. R. (2021). Proximate and ultimate causes of the bactericidal action of antibiotics. In Nature
Reviews Microbiology (Vol. 19, Issue 2, pp. 123–132). Nature Research. https://doi.org/10.1038/s-.Singleton, M. R., Dillingham, M. S., Gaudier, M., Kowalczykowski, S. C., & Wigley, D. B. (2004). Crystal structure of
RecBCD enzyme reveals a machine for processing DNA breaks. Nature,-), 187–193. https://doi.org/10.1038/
nature-.Adam, M., Murali, B., Glenn, N. O., & Potter, S. S. (2008). Epigenetic inheritance based evolution of antibiotic resistance in
bacteria. BMC Evolutionary Biology, 8(1). https://doi.org/10.1186/-.Cohen, N. R., Ross, C. A., Jain, S., Shapiro, R. S., Gutierrez, A., Belenky, P., Li, H., & Collins, J. J. (2016). A role for the
bacterial GATC methylome in antibiotic stress survival. Nature Genetics, 48(5), 581–586. https://doi.org/10.1038/ng-.Siwach, M., Kumar, L., Palani, S., Muraleedharan, S., Panis, G., Fumeaux, C., Mony, B. M., Sanyal, S., Viollier, P. H., &
Radhakrishnan, S. K. (2021). An organelle-tethering mechanism couples flagellation to cell division in bacteria.
Developmental Cell. https://doi.org/10.1016/j.devcel-
20