You are on page 1of 13

Review

Cite This: ACS Infect. Dis. 2019, 5, 1505−1517 pubs.acs.org/journal/aidcbc

(p)ppGpp and the Stringent Response: An Emerging Threat to


Antibiotic Therapy
Joanne K. Hobbs* and Alisdair B. Boraston*
Department of Biochemistry and Microbiology, University of Victoria, 3800 Finnerty Road, Victoria, BC V8P 5C2, Canada

ABSTRACT: In 1969, Cashel and Gallant first observed the presence


of (p)ppGppthe signaling molecule of the stringent responsein
starved bacterial cells. Fifty years later, (p)ppGpp and the stringent
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

response have emerged as essential master regulators of not only the


bacterial response to stress but also almost all aspects of bacterial
physiology, virulence, and immune evasion. More worryingly, a wealth
Downloaded via UNIV OF SAO PAULO on March 22, 2020 at 00:57:14 (UTC).

of data now indicate that (p)ppGpp and stringent response activation


pose a serious threat to the efficacy and clinical success of
antimicrobial therapy. Here, we focus on the central role that
(p)ppGpp and the stringent response play in the phenomenon of
antibiotic tolerance, as well as the acquisition, development, and
expression of antibiotic resistance. We review these consequences of
stringent response activation in relation to the main proteins involved
in (p)ppGpp production and control, in particular the complex interplay between monofunctional and bifunctional long RelA/
SpoT homologues (RSHs) and small alarmone synthetases (SASs). We also review the growing evidence to suggest that there
are multiple other indirect pathways of stringent response induction that can affect antibiotic efficacy. Finally, we summarize
recent studies that indicate the in vivo and clinical impact of (p)ppGpp production on antibiotic treatment outcomes. We
conclude by reviewing the progress to date in the search for novel therapeutics that target the stringent response.
KEYWORDS: stringent response, (p)ppGpp, antibiotic tolerance, antibiotic resistance, RelA/SpoT homologues,
small alarmone synthetase

viability5,15,16and that bacterial phyla, classes, and even


T his year marks the 50th anniversary of the discovery of
guanosine 5′-diphosphate 3′-diphosphate (ppGpp) and
guanosine 5′-triphosphate 3-diphosphate (pppGpp)collec-
genera differ in (i) the mechanisms through which (p)ppGpp
affects transcription and translation,14 and (ii) the config-
tively known as (p)ppGpp or “magic spot”in starved uration and number of (p)ppGpp-producing enzymes they
bacterial cells.1 These “alarmones” are the effectors of the possess.17
stringent response, a universal stress response that is classically The RelA/SpoT homologue (RSH) superfamily contains
induced by amino acid deprivation.2 Since its discovery, two types of characterized and functionally important enzyme
(p)ppGpp has emerged as a master regulator of almost all that synthesize and/or hydrolyze (p)ppGpp: long RSHs and
aspects of bacterial physiology, including growth rate control, small alarmone synthetases (SASs).17 Small alarmone hydro-
phase transition, sporulation, motility, competence, biofilm lases (SAHs) also exist but are largely uncharacterized and are
formation, toxin production, and a multitude of other virulence sporadically distributed among bacteria; therefore, they will
associations.3−5 There is also a growing body of evidence that not be discussed further here. Long RSHs contain both a
implicates (p)ppGpp and the stringent response in antibiotic (p)ppGpp hydrolase domain and a (p)ppGpp synthetase
tolerance6−10 and resistance.11−13 domain in the N-terminal half of the protein, while the C-
In the classical description of stringent response activation in terminal half is involved in regulation and ribosome binding
Escherichia coli, amino acid starvation results in the presence of
(Figure 1).18 They can be further divided into three groups
uncharged tRNAs in the A site of ribosomes. The stalled
according to activity and bacterial distribution, and herein, we
ribosomes are sensed by the protein RelA, which responds by
will be using the RelA/SpoT/Rel naming system.17,18 RelA
synthesizing (p)ppGpp from GTP/GDP and ATP; an
accompanying hydrolase protein, SpoT, degrades the proteins are found in γ- and β-proteobacteria, such as E. coli
(p)ppGpp once the amino acid deprivation has been alleviated. and Pseudomonas aeruginosa. The absence of a conserved
The overall effect of (p)ppGpp during the stringent response is HDXXED motif in the hydrolase domain active site renders
the downregulated transcription of most metabolic genes and this domain inactive.18 Therefore, RelA is a monofunctional
the upregulation of genes associated with amino acid long RSH. The same classes of Proteobacteria also possess a
biosynthesis and stress responses.2 We now know that
(p)ppGpp production can be induced by a number of different Received: May 31, 2019
stimuli3,5,14on top of the basal levels required for Published: July 9, 2019

© 2019 American Chemical Society 1505 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

otherwise be lethal, without exhibiting an increase in MIC.


Historically, tolerance was often referred to as “phenotypic
resistance” or “non-inherited resistance”;26−28 however, it is
now known that tolerance can also be genotypic.29,30 The
terms tolerance and persistence are also used interchangeably
in the literature31 but are actually distinct phenomena.
Tolerance can be detected in time-kill assays as a slower rate
of killing that applies to the whole population (Figure 2).
Figure 1. Domain structure of common RelA/SpoT homologue
(RSH) proteins. Long RSHs are composed of two regions: an N-
terminal enzymatic region and a C-terminal regulatory region. The
enzymatic region contains hydrolase and synthetase domains that
hydrolyze and synthesize (p)ppGpp, respectively. The regulatory
region contains a ThrRS, GTPase, and SpoT (TGS) domain, a
conserved helical domain (H), a zinc-finger or conserved cysteine
domain (ZFD/CC), and an RNA recognition motif or aspartate
kinase, chorismate, and TyrA (ACT) domain (RRM/ACT).18 In
RelA proteins, the hydrolase domain is inactive (indicated by *), and
in SpoT proteins, the synthetase domain exhibits only weak activity
(indicated by #). The small alarmone synthetases RelV, RelP, and
RelQ consist of only a synthetase domain that bears little sequence
identity to the synthetase domain of long RSHs.3
Figure 2. Schematic depiction of time-kill results obtained with
second long RSH, SpoT. SpoT proteins exhibit both susceptible, tolerant, persistent and resistant strains. Tolerance can be
detected in time-kill assays as a slower rate of killing and greater
synthetase and hydrolase activity; however, the synthetase
minimum duration for killing (MDK), for example, 99% of the
function is weak and activated by different nutritional stimuli population compared with a susceptible strain, despite having the
to that of RelA.5 Finally, the most widely distributed long RSH same minimum inhibitory concentration (MIC). Persistence mani-
(and RSH in general) is Rel.17 Rel (also sometimes referred to fests as a similar initial rate of killing and MDK99 to that of a
as Rsh or RelA) is a truly bifunctional RSH, exhibiting efficient susceptible strain (and the same MIC), followed by the emergence of
synthetase and hydrolase activity.19 a small persistent subpopulation that dies more slowly. A resistant
Rel proteins are found in many important human pathogen- strain exhibits an MIC greatly above that of a susceptible strain and is
containing phyla, including the Firmicutes (e.g., Staphylococcus resistant to killing in a time-kill assay (at antibiotic concentrations that
aureus, Enterococcus spp.) and Actinobacteria (e.g., Mycobacte- kill the other strains).
rium tuberculosis). The two antagonistic enzymatic functions of
Rel are thought to be controlled through conformational shifts Persistence, however, is a property displayed by a small
and reciprocal regulation,20 with the synthetase-off/hydrolase- subpopulation of bacteria. The initial rate of killing for a
on conformation being the resting state of the protein.21 persistent strain is similar to that of its susceptible counterpart
Of the 12 subgroups of SAS that have been bioinformatically and applies to the majority of the population; however, a
identified, only three have been functionally characterized: subpopulation (typically <1%) emerges that can survive
RelP, RelQ, and RelV.17 SASs consist of an isolated synthetase exposure to the antibiotic for longer.25 Therefore, tolerance
domain that bears little sequence identity to the synthetase and persistence can be distinguished from each other in a time-
domain of long RSHs.3 Due to the toxicity of high (p)ppGpp kill assay (Figure 2).32 The distinction between tolerance and
levels,22 SASs are only found in bacteria that also possess Rel persistence is also complicated by the use of the term
or SpoT.17 SASs have also been identified in bacteriophage.23 “persistent” to describe difficult-to-treat infections. While
The vast majority of Firmicutes members possess at least one resistance is a well-known and recognized cause of antibiotic
SAS, and important human pathogen-containing genera, like treatment failure, tolerance is a poorly understood antibiotic
Staphylococcus and Clostridium, contain both RelP and RelQ.17 escape strategy and greatly underappreciated contributor to
RelV is only found in a small subset of γ-proteobacteria, persistent and relapsing infections.25,29,33
including Vibrio spp. While Rel/RelA is the main contributor In the past decade, the clinical significance of the stringent
to (p)ppGpp production during the stringent response, these response has been realized through the isolation of two clinical
accessory SASs are thought to play an important role in isolates (S. aureus and Enterococcus faecium) bearing mutations
bacterial homeostasis.5 in Rel that led to elevated (p)ppGpp levels.6,34 These isolates
In addition to their role in bacterial virulence,3 induction of were both associated with persistent infections that required
the stringent response and elevated (p)ppGpp levels have been prolonged treatment with multiple antibiotics to clear. Here,
implicated in antibiotic resistance,11−13 tolerance,6−10 and we review the wealth of evidence that now implicates the
persistence.24 In this Review, we will employ the definitions of stringent response and (p)ppGpp in the expression of
Brauner et al.25 for these three phenomena. Antibiotic antibiotic tolerance and resistance. We also provide an
resistance describes the inheritable ability of bacteria to grow overview of recent attempts to design and discover inhibitors
of the stringent response.


at high concentrations of an antibiotic for an indefinite period
of time. It can be readily detected as a minimum inhibitory
concentration (MIC) above that of a susceptible strain (and THE STRINGENT RESPONSE AND ANTIBIOTIC
above the resistance breakpoint, where one exists). In contrast, TOLERANCE
antibiotic tolerance describes the ability of bacteria to survive The molecular mechanisms behind antibiotic tolerance are
transient exposure to concentrations of antibiotic that would varied and not fully understood;26,29 however, the overarching
1506 DOI: 10.1021/acsinfecdis.9b00204
ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

Table 1. Associations between the Stringent Response and Antibiotic Tolerance


Organism (and RSH Genotype or induction (p)ppGpp
Category of induction proteins) method level Tolerance phenotype
Nutritional/chemical Escherichia coli (RelA, Isoleucine deprivation Higha Tolerance to ampicillin39
SpoT)
Escherichia coli (RelA, Lysine deprivation Higha Tolerance to most penicillins, cephalosporins and carbapenems38
SpoT)
Escherichia coli (RelA, Mupirocin treatment High41 Tolerance to ampicillin and norfloxacin41
SpoT)
Pseudomonas SHX treatment High7 Tolerance to ofloxacin7
aeruginosa (RelA,
SpoT)
Vibrio cholerae (RelA, SHX treatment High40 Tolerance to tetracycline, erythromycin and chloramphenicol40
SpoT, RelV)
Staphylococcus aureus Mupirocin treatment Higha Tolerance to vancomycin42
(Rel, RelP, RelQ)
Genetic manipulation of Escherichia coli (RelA, Overexpression of RelA High7 Tolerance to ampicillin,44 mecillinam46,47 and microcin J2548
one or more RSH SpoT)
Escherichia coli (RelA, ΔrelA Lowa Reduced tolerance to penicillin under amino acid deprivation37
SpoT)
Loss of tolerance to ampicillin39,41
Loss of tolerance to ampicillin, D-cycloserine and moenomycin
under amino acid deprivation43
Tolerance to ampicillin during regrowth in minimal medium with
glucose following prolonged stationary phase56
Escherichia coli (RelA, ΔrelA ΔspoT Nonea Loss of tolerance to ofloxacin under amino acid deprivation at
SpoT) 24 h51
Increased killing by ampicillin (in LB medium)41
Increased killing by ofloxacin and tobramycin in a biofilm7
Pseudomonas ΔrelA Low49 Increased killing by ofloxacin49
aeruginosa (RelA,
SpoT)
Pseudomonas ΔrelAΔspoT None7 Loss of tolerance to gentamicin, ofloxacin and meropenem in
aeruginosa (RelA, stationary phase50
SpoT)
Increased killing by ofloxacin,7,52 meropenem,7 colistin,7 and
gentamicin7 in a biofilm
Increased killing by ofloxacin in stationary phase7
Vibrio cholerae (RelA, ΔrelAΔspoT High40 Tolerance to tetracycline, erythromycin and chloramphenicol40
SpoT, RelV)
Vibrio cholerae (RelA, ΔrelAΔspoTΔrelV None40 Loss of tolerance to tetracycline, erythromycin and
SpoT, RelV) chloramphenicol in stationary phase40
Enterococcus faecalis Δrel High8 Tolerance to vancomycin8
(Rel, RelQ)
Enterococcus faecalis ΔrelQ Low8 Increased killing by vancomycin8
(Rel, RelQ)
Enterococcus faecalis ΔrelΔrelQ None8 Increased killing by vancomycin,8 norfloxacin,15 and ampicillin15
(Rel, RelQ)
Staphylococcus aureus ΔrelPΔrelQ Lowa Increased killing by ampicillin and vancomycin22
(Rel, RelP, RelQ)
Staphylococcus aureus ΔrelP or ΔrelQ Normala No change in killing by ampicillin or vancomycin22
(Rel, RelP, RelQ)
Staphylococcus aureus ΔrelΔrelPΔrelQ Nonea Increased killing by ampicillin and vancomycin22
(Rel, RelP, RelQ)
Mycobacterium Δrel None53 Loss of tolerance to isoniazid under nutrient starvation54
tuberculosis (Rel)
Staphylococcus aureus Gln360Stop mutation in Higha Tolerance to ciprofloxacin57
(Rel, RelP, RelQ) rel
Staphylococcus aureus Leu61Phe mutation in rel Higha Tolerance to daptomycin58
(Rel, RelP, RelQ)
Enterococcus faecium Leu152Phe mutation in rel High6 Tolerance to vancomycin, linezolid and daptomycin in a biofilm6
(Rel, RelQ)
Genetic manipulation of Escherichia coli (RelA, ΔdksA Normala Increased killing by fluoroquinolones, streptomycin and
an accessory protein SpoT) ampicillin59
Pseudomonas ΔdksA Moderateb49 Tolerance to ciprofloxacin and ofloxacin in stationary phase49
aeruginosa (RelA,
SpoT)
Vibrio cholerae (RelA, ΔdksA Normal40 Loss of tolerance to tetracycline, erythromycin, and
SpoT, RelV) chloramphenicol in stationary phase40
Staphylococcus aureus ΔrsgA or rsgA mutation Normala Tolerance to β-lactams, vancomycin, and ciprofloxacin9
(Rel, RelP, RelQ)

1507 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

Table 1. continued
Organism (and RSH Genotype or induction (p)ppGpp
Category of induction proteins) method level Tolerance phenotype
Escherichia coli (RelA, Overexpression of HipA High10,60 Tolerance to carbenicillin,10 ampicillin,10,61 ciprofloxacin,31,61
SpoT) norfloxacin,60 cefotaxime,62 ofloxacin,62 and mitomycin C62
Escherichia coli (RelA, Overexpression of HipA in Lowa Loss of tolerance to ampicillin conferred by HipA
SpoT) ΔrelA overexpression10
Escherichia coli (RelA, Overexpression of HipA Normala Reduced tolerance to cefotaxime, ofloxacin and mitomycin C
SpoT) mutants conferred by wildtype HipA62
Escherichia coli (RelA, argS, alaS, pheS, leuS, thrS, High46 Tolerance to mecillinam,46,63 ciprofloxacin,57,63 rifampicin,63
SpoT) glnS or aspS mutation chloramphenicol,63 ampicillin,63 and trimethoprim63
Escherichia coli (RelA, argS, alaS, leuS or aspS Low46 Loss of tolerance to mecillinam46 or ciprofloxacin63 conferred by
SpoT) mutation and ΔrelA an aminoacyl tRNA synthetase mutation
Staphylococcus aureus ileS mutation Higha Tolerance to vancomycin, imipenem, and daptomycin42
(Rel, RelP, RelQ)
a
Not experimentally determined; theoretical level based on genotype and/or induction method. bDetermined during the stationary phase of
growth. Abbreviations: RelA/SpoT homologue (RSH), serine hydroxamate (SHX)

Figure 3. Induction and control of the stringent response in γ/β-proteobacteria and most other classes of bacteria. The cascade of (p)ppGpp
production, interaction, and degradation is shown. Step 1: Deprivation of one of more amino acid leads to an uncharged tRNA entering the A site
of the ribosome. Step 2: The stalled ribosome is sensed by RelA (2a) or Rel (2b), which responds by synthesizing (p)ppGpp. Step 3: (p)ppGpp
acts to downregulate transcription/translation either by binding to RNA polymerase (RNAP) in conjunction with DksA (3a), or binding to, and
inhibiting, GTPases such as RsgA that are required for proper ribosome assembly (3b). Many other mechanisms of (p)ppGpp-mediated
transcriptional/translational control exist; only those that have been experimentally associated with antibiotic tolerance are shown. Step 4: When
the amino acid deprivation has been alleviated, SpoT (4a) or Rel (4b) degrades (p)ppGpp; Rel catalyzes (p)ppGpp hydrolysis off of the ribosome.
Additional contributors to (p)ppGpp production and stringent response control include: (i) tetrameric small alarmone synthetases and (ii) tRNA
charging enzymes. Most members of the Firmicutes and Vibrio spp. possess accessory synthetases that synthesize (p)ppGpp; SpoT also possesses
weak (p)ppGpp synthetase activity (shown in gray). Inactivation or reduced activity of tRNA charging enzymes can indirectly induce the stringent
response. The protein kinase HipA phosphorylates the glutamyl-tRNA synthetase GltX, thereby inactivating it and preventing the charging of
glutamyl-tRNAs. Mutations in aminoacyl tRNA synthetase genes (such as ileS) can reduce their activity and lead to stringent response activation in
a similar manner. Boxes indicate enzymes and processes that are specific to that given species, genus, or phylum. It should be noted that stringent
response-activating mutations in aminoacyl tRNA synthetase genes and homologues of RsgA are also found in γ/β-proteobacteria.

feature is slow growth. Known mechanisms of slow growth that metabolism and growth rate leads to multidrug tolerance.35,36
lead to tolerance include defects in electron transport or Classical activation of the stringent response leads to the
thymidine biosynthesis, as seen in small colony variants shutdown of almost all metabolic processes and a state of
(SCVs), and biofilms.28 As the majority of clinically relevant quiescence in cells. In the clinical setting, however, it is now
antibiotics target active metabolic processes, a reduction in clear that the stringent response is not binary in its control and
1508 DOI: 10.1021/acsinfecdis.9b00204
ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

can be partially activated.6,34 This results in cells that exhibit of gene knockouts on (p)ppGpp levels is more complicated.
reduced, rather than abolished, metabolism and growth.5 As Double knockouts of relA and spoT in E. coli and P. aeruginosa
such, these cells are still able to proliferate and cause are unable to synthesize (p)ppGpp, even when stimulated;
infections,6,34 but benefit from the antibiotic tolerance afforded therefore, they exhibit a loss of tolerance or increased
by slow growth.33,35,36 A variety of different methods to sensitivity to killing by an antibiotic in the stationary
modulate levels of (p)ppGpp in different bacterial species have phase,50 under amino acid deprivation,51 or in a biofilm7,52
been employed to investigate the association between the (Table 1). In contrast, a ΔrelAΔspoT mutant of Vibrio cholerae
stringent response and antibiotic tolerance. They can be exhibits tolerance to inhibitors of protein synthesis.40 This
separated into three categories: chemical/nutritional, genetic discordance is due to the presence of RelV in V. cholerae, which
manipulation of RSHs, and genetic manipulation of an in the absence of SpoT produces unregulated amounts of
accessory protein (Table 1). (p)ppGpp (Figure 3). As expected, the tolerance phenotype of
Early Studies with E. coli. The first association between V. cholerae is abolished in a triple ΔrelAΔspoTΔrelV mutant.
the stringent response and antibiotic tolerance was made In bacteria that possess a bifunctional Rel, deletion of the rel
almost 40 years ago. Following the observation that amino acid gene has varying effects. The M. tuberculosis genome encodes
starvation renders E. coli tolerant to the lytic effects of for only a single RSH, Rel; therefore, a rel knockout mutant
penicillin, this tolerance was found to be largely dependent on possesses no detectable (p)ppGpp53 and does not exhibit the
the presence of RelA.37 Amino acid deprivation is now a usual tolerance to isoniazid induced by starvation conditions.54
commonly employed method of stringent response induction In comparison, deletion of rel in Enterococcus faecalis has the
in tolerance studies, and it can be achieved either through the same ultimate effect as deletion of both relA and spoT in V.
use of defined medium that lacks one or more amino acid,37,38 cholerae (i.e., high [p]ppGpp)8 because E. faecalis also
the addition of excess valine leading to isoleucine starvation,39 possesses an accessory SAS, RelQ. Production of (p)ppGpp
or through the inclusion of an amino acid analog such as serine by RelQ in the absence of a hydrolase results in tolerance to
hydroxamate (SHX).7,40 The effects of isoleucine starvation the cell wall-active antibiotic vancomycin.8 This is despite the
can also be mimicked by the addition of mupirocin fact that in response to mupirocin the rel mutant does not
(pseudomonic acid),41,42 which is an isoleucyl tRNA mount the canonical stringent transcriptional response or
synthetase inhibitor. Following the initial association between appear to accumulate large amounts of (p)ppGpp.8,55 There-
amino acid starvation, RelA, and penicillin tolerance, the fore, the authors concluded that (p)ppGpp-mediated tolerance
stringent response was linked to the tolerance of starved E. coli in E. faecalis occurs at relatively low (p)ppGpp concentrations
to other inhibitors of cell wall biosynthesis.38,43 Reversal of this and below those associated with typical induction of the
tolerance in relA+ strains, via apparent “relaxation” of the stringent response. In agreement with the proposed
stringent response with a protein synthesis inhibitor, also importance of basal, RelQ-mediated (p)ppGpp levels for
provided support for the role of RelA in multidrug tolerance.44 tolerance in this bacterium, a ΔrelQ mutant is more readily
However, the use of amino acid deprivation to induce killed by vancomycin under nonstressed conditions than the
(p)ppGpp production was not considered an ideal model wildtype.8 As expected, a double ΔrelΔrelQ strain possesses no
system as it causes global effects outside of increased (p)ppGpp and is killed faster than the wildtype and single
(p)ppGpp.45 Therefore, Rodionov and Ishiguro used over- mutants by multiple antibiotics.15
expression of RelA via a tac promoter and variable inducer S. aureus represents one of the most complex systems of
concentrations to demonstrate the direct relationship between (p)ppGpp control as it contains Rel and two accessory SASs
(p)ppGpp and ampicillin tolerance.44 A similar genetic system (RelP and RelQ; Table 1). Like RelQ in E. faecalis, RelP and
was also subsequently used to demonstrate the relationship RelQ have been strongly implicated in antibiotic tolerance in S.
between (p)ppGpp concentration and tolerance to the β- aureus, specifically tolerance to cell wall biosynthesis inhibitors.
lactam mecillinam and the peptide antibiotic microcin Transcription of relP and relQ, but not rel, is induced by
J25.46−48 In fact, Joseleau-Petit et al. identified a threshold exposure to ampicillin and vancomycin. Furthermore, a double
concentration of (p)ppGpp above which tolerance was ΔrelPΔrelQ mutant is more readily killed by these agents (but
exhibited.47 Later studies have since revealed that the tolerance not ciprofloxacin, a fluoroquinolone DNA synthesis inhibitor)
conferred by elevated (p)ppGpp levels extends beyond cell than the wildtype.22 Interestingly, single deletion mutants of
wall synthesis inhibitors to antibiotics with diverse mechanisms relP or relQ exhibit similar time-kill kinetics to the wildtype. In
of action; e.g., refs 7 and 49. These early experiments with E. contrast to E. faecalis, a rel deletion mutant of S. aureus could
coli formed the basis of our understanding of the relationship not be generated and characterized in terms of tolerance due to
between the stringent response and tolerance; however, the the essentiality of the hydrolase domain.22 This essentiality was
complement of RelA/SpoT RSHs found in E. coli and other γ- shown to apply only when both RelP and RelQ were present.
proteobacteria is not typical of the vast majority of human Therefore, the authors concluded that a single SAS (in
pathogens and bacteria in general. addition to rel) is sufficient to confer tolerance to cell wall-
Complex Control of (p)ppGpp Levels. With the advent active antibiotics but does not lead to a toxic accumulation of
of the genomic era, the full repertoire of bacterial RSH proteins (p)ppGpp. This raises the question of why S. aureus, and many
was realized and with it the complexity of (p)ppGpp control in other bacteria, possess multiple SASs.
different pathogenic bacteria (Figure 3). By far the most Indirect Induction and Control of the Stringent
common method employed in the study of the stringent Response. Variations in the complement of RSHs are not
response and tolerance is genetic manipulation of one or more the only differences in stringent response induction and
RSH (Table 1). In β- and γ-proteobacteria, elevated (p)ppGpp control between bacterial species that affect tolerance (Figure
levels can be achieved through the overexpression of 3). Broadly speaking, the transcriptional and translational
RelA.44,46−48 In bacteria that possess the bifunctional RSH effects of (p)ppGpp occur via two distinct mechanisms in
Rel, or a combination of Rel and SASs, deciphering the effects Proteobacteria and other bacterial phyla.5 In Proteobacteria,
1509 DOI: 10.1021/acsinfecdis.9b00204
ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

(p)ppGpp binds to RNA polymerase to bring about the (HipB) and confer high level persistence. Deletion of relA led
transcriptional changes that are the hallmark of the stringent to a considerable reduction in the high persistence phenotype
response.45 This binding, and the downstream effects conferred by hipA7, and deletion of both relA and spoT
associated with induction of the stringent response, are returned the persistence rate to that of the wildtype hipA
critically dependent on the transcription factor DksA (Figure strain.66
3).64 The effect of dksA deletion on (p)ppGpp-mediated HipA induces the stringent response by inactivating the
antibiotic tolerance has been investigated in E. coli, P. tRNA synthetase GltX. Nonsynonymous mutations in the
aeruginosa, and V. cholerae (Table 1). Deletion of dksA in E. genes encoding for other tRNA synthetases can also reduce
coli led to an increased rate of killing by ampicillin, two their activity, activate the stringent response, and confer
fluoroquinolones, and a protein synthesis inhibitor compared tolerance.42,46,57,63 Numerous mutations that confer tolerance
with the wildtype.59 A similar effect was observed in V. to the β-lactam mecillinam or ciprofloxacin in E. coli were
cholerae, whereby a ΔdksA mutant exhibited a loss of tolerance, mapped to argS, alaS, and leuS, which encode for arginyl-,
usually conferred by the stationary phase, to three different alanyl-, and leucyl-tRNA synthetase, respectively.46,63 Other
protein synthesis inhibitors.40 These findings are in agreement tolerant mutants were found to bear mutations in the genes
with the inability of ΔdksA mutants to facilitate the encoding for different tRNA synthetases but not characterized
downstream effects of (p)ppGpp production64 and thus further.46,63 The biochemical effects of the argS, alaS, and leuS
tolerance. Paradoxically, a dksA deletion mutant of P. mutations on tRNA synthetase activity were not tested.
aeruginosa exhibited a slightly elevated (p)ppGpp level However, the mutants exhibited slow growth and/or high
compared with the wildtype and significant tolerance to (p)ppGpp levels; therefore, it is likely that the mutations
fluoroquinolones in the stationary phase.49 In this instance, the reduced tRNA synthetase activity leading to the presence of
authors noted that the ΔdksA mutant exhibited higher uncharged tRNAs. The tolerance conferred by the tRNA
antibiotic tolerance than a ΔspoT mutant, despite containing synthetase mutations was also found to be either partially or
considerably less (p)ppGpp. Therefore, they speculated that entirely dependent on the presence of relA. Similar effects of
DksA may negatively regulate the expression of a putative gene tRNA synthetase mutations have been observed in S.
involved in tolerance in this bacterium. aureus.42,57 Nonsynonymous mutations in leuS and the ileS
In most bacterial phyla, (p)ppGpp does not bind to RNA gene encoding for isoleucyl-tRNA synthetase have been
polymerase, and its effects on transcription and translation identified in slow-growing, ciprofloxacin- or vancomycin-
occur indirectly through the modulation of intracellular GTP tolerant mutants. These mutants exhibited the classic hallmarks
levels.5,14,16 There are a multitude of mechanisms through of tolerance: an MIC similar to the parental strain and far
which (p)ppGpp affects GTP synthesis, including inhibition of below the resistance breakpoint, but a >2-fold higher minimum
the GTPase RsgA.9 RsgA binds to the 30S ribosomal subunit duration for killing (MDK)32 required to kill 99.9% of the
and is thought to play a chaperoning role in proper ribosome population. In agreement with the multidrug tolerance
assembly.65 Corrigan et al. have shown that (p)ppGpp binds to phenotype observed with other methods of stringent response
RsgA and inhibits its GTPase activity, thereby disrupting induction, the ileS mutations also conferred tolerance to the
ribosome assembly and leading to slow growth.9 Deletion of membrane-damaging antibiotic daptomycin69 and the β-lactam
imipenem.42


rsgA, or introduction of a mutated version with reduced
GTPase, mimics the effects of (p)ppGpp in S. aureus. Both the
deletion and mutated strains exhibit tolerance to β-lactams, THE STRINGENT RESPONSE AND ANTIBIOTIC
vancomycin, and ciprofloxacin (Table 1). Given that RsgA is RESISTANCE
only one of many binding partners of (p)ppGpp, there are a While tolerance is emerging as an important contributor to
great number of other indirect mechanisms through which antibiotic treatment failure in its own right, it also acts as a
(p)ppGpp-mediated tolerance could be achieved. precursor to the development of resistance.30,70 Antibiotic
DksA and RsgA represent crucial facilitators in the resistance represents an ever-growing threat to our ability to
downstream effects of stringent response activation once cure bacterial infections, and methicillin-resistant S. aureus
(p)ppGpp has been synthesized in response to uncharged (MRSA) is of particular concern.71 In recent years, (p)ppGpp
tRNAs. Several indirect mechanisms of stringent response production has been linked with the complex expression of
induction that lead to the occurrence of uncharged tRNAs methicillin resistance in S. aureus, as well as the ability of
have also been associated with tolerance. HipA is a protein bacteria to acquire resistance genes or mutate antibiotic
kinase found in E. coli, and some other Gram-negative targets.
bacteria,66 that phosphorylates the glutamyl-tRNA synthetase (p)ppGpp and the Expression of β-Lactam Resist-
enzyme (GltX; Figure 3).31,60 This phosphorylation inhibits ance. Methicillin resistance is a term used to refer to
aminoacylation, resulting in the presence of uncharged resistance to the β-lactam family of antibiotics, and it is
tRNAGlu and induction of the stringent response. Over- typically conferred by the presence of the mecA gene (either
expression of wildtype HipA in E. coli results in tolerance to chromosomal or on a plasmid), which encodes for the
β-lactams and fluoroquinolones (Table 1).10,31,60−62,67 Some alternate, low-affinity penicillin-binding protein PBP2A.71 In
studies have referred to this phenotype as “persistence,” but most clinical isolates of MRSA, β-lactam resistance is expressed
using the definitions of Brauner et al.,25 we conclude this to be heterogeneously; the population as a whole exhibits relatively
tolerance. It should, however, be noted that HipA was low level resistance, with an MIC barely above the
originally associated with bona f ide persistence.68 HipA- susceptibility breakpoint, but there exist subpopulations that
mediated persistence has been directly linked to induction of exhibit very high levels of resistance.72 The stringent response
the stringent response through studies with the hipA7 allele.66 has been implicated in the mechanism of heterogeneous β-
This allele contains two amino acid changes that enable HipA lactam resistance, specifically in the transition from hetero- to
to be expressed nontoxically without its antitoxin partner homogeneous resistance. Mwangi et al. used whole genome
1510 DOI: 10.1021/acsinfecdis.9b00204
ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

sequencing (WGS) to compare two populations (low and high restored to the parental level by the presence of mupirocin
resistance) within an engineered heterogeneous MRSA and (presumably through induction of rel-mediated (p)ppGpp
identified two rel mutations in the highly resistant cells.12 synthesis). Therefore, it appears that high level β-lactam
These were the only mutations found genome-wide. The resistance can be induced either through Rel or RelQ, but not
primary mutation introduced a premature stop codon at the RelP.
end of the synthetase domain, thereby generating a truncated Alternate routes of stringent response induction have also
Rel that lacks the regulatory region (Figure 1). The second rel been linked to the expression of β-lactam resistance. WGS of
mutation was located in the now-deleted region. Subsequently, six naturally occurring heteroresistant MRSA strains revealed
a further two truncating mutations have been identified in the many mutations associated with the highly resistant sub-
regulatory region of Rel in highly resistant subpopulations of populations, including nonsynonymous mutations in five
clinical MRSA isolates,73,74 and yet another truncation different tRNA synthetase genes.73 As reviewed earlier,
mutation has been linked to ciprofloxacin tolerance.57 The mutations in these (or analogous) genes have already been
original truncated Rel mutant exhibited homogeneous high- shown to lead to elevated (p)ppGpp levels. Other mutations
level β-lactam resistance, slow growth, and elevated (p)ppGpp associated with homogeneous, high-level β-lactam resistance
consistent with partial activation of the stringent response.12 were identified in genes related to guanine metabolism,
There are three ways in which deletion of the regulatory transcription, and translation, and hypothesized to potentially
domain of Rel could function to activate the stringent activate the stringent response indirectly.73,74,77 Mutations in
response: promote synthetase activity, reduce hydrolase several of these genes have also been proposed to play a role in
activity, or a combination of both. Biochemical and whole- the natural homogeneous β-lactam resistance profile of MRSA
cell studies with S. aureus have shown that the synthetase COL, a strain that exhibits slow growth and elevated
function of Rel is negatively affected by the regulatory region (p)ppGpp.78
and truncation increases synthetase activity; however, the Mutation Rates and Resistance Acquisition under the
increase in (p)ppGpp synthesis is not enough to counteract the Stringent Response. There are two mechanisms through
dominant hydrolase activity of Rel (which is apparently which bacteria can become resistant to an antibiotic:
unaffected by the regulatory domain).21 Therefore, the endogenous mutation or the acquisition of foreign DNA.82
molecular details behind stringent response activation via Rel Experimental evidence indicates a role for the stringent
truncation is unknown. Nevertheless, the ability of stringent response in both of these mechanisms. Stress responses, like
response induction to convert heterogeneous MRSA isolates to the stringent response, have long been associated with
homogeneous, highly resistant forms has now been demon- increased bacterial mutability.83 Deletion of relA/rel from E.
strated in many clonal types through the simple addition of coli or Bacillus subtilis leads to lower reversion (i.e., mutation)
mupirocin or SHX to the growth medium.11,12,73,75−78 These rates in a variety of amino acid auxotrophs, and (p)ppGpp
stringent response-activated strains exhibit higher levels of concentration has been shown to be positively correlated with
PBP2A than their “relaxed” heterogeneous counterparts. the mutation rate in these genes.84−86 However, this
Interestingly, the resistance expression profile of MRSA strains (p)ppGpp-mediated rise in mutation rate is thought to be
that do not possess an alternate PBP like PBP2A, and are due to the increase in transcription rate of these genes,84,85 and
instead resistant to β-lactams by virtue of mutations in the the vast majority of genes in the bacterial genome are actually
native PBP, does not change in the presence of mupirocin.11,78 downregulated under the stringent response.2 Nevertheless,
Given the complexity of (p)ppGpp control and the different deletion of relA and spoT has been reported to suppress the
routes of stringent response induction reviewed here in emergence of fluoroquinolone-resistant colonies in stationary
relation to tolerance, it is not surprising that heterogeneous phase P. aeruginosa.7 Furthermore, activation of the stringent
expression of β-lactam resistance has also been associated with response via the hipA7 allele has recently been shown to
mutations in genes other than rel. A partial revertant of the increase the emergence of resistant colonies in the presence of
original Rel truncation mutant was found to bear a non- four different antibiotics.70 This effect is thought to be linked
synonymous mutation in relQ.12 This mutant exhibited an to the tolerance/persistence phenotype of the hipA7 strain.
intermediate level of β-lactam resistance and increased colony Tolerance mutations are known to precede resistance
size. The relQ mutation can only be assumed to have reduced mutations in E. coli exposed to ampicillin.30 Interestingly, the
basal production of (p)ppGpp to partially counter the effect of main tolerance mutations observed in this study occurred in
the Rel truncation. Similarly, several relQ truncation mutations genes associated with indirect stringent response activation
have been reported to reverse small colony size and high-level (metG; methionyl-tRNA synthetase) and/or linked to high
vancomycin resistance in slow-growing S. aureus.79,80 A more level β-lactam resistance (prsA).73,74,78 A nonsynonymous rel
detailed study into the effects of RelP and RelQ on β-lactam mutation was also identified in a laboratory strain of B. subtilis
resistance and mecA expression was recently performed.81 that had been passaged in the presence of daptomycin until it
Deletion of relQ was found to greatly reduce the level of β- developed resistance.87 The rel mutation itself did not confer
lactam resistance through reduced mecA expression, while daptomycin resistance, and we can only speculate that it may
deletion of relP led to higher mecA expression than the have arisen prior to, and potentially encouraged the emergence
wildtype (in response to β-lactam exposure) and an increased of, the bona f ide resistance mutation. The B. subtilis rel
MIC. Promoter analysis of both relQ and relP revealed that the mutation is distinct from one recently reported in several S.
relQ promoter is 5-fold stronger than that controlling relP and aureus cultures following serial passage in the presence of
is significantly induced by deletion of relP; however, the relP daptomycin (Table 1).58 The S. aureus mutation conferred
promoter is not responsive to relQ deletion. Therefore, tolerance to daptomycin; any potential effect on subsequent
deletion of relP is more than compensated for by increased resistance development was not investigated.
expression of relQ, leading to the increased resistance level In terms of the role of the stringent response in resistance
observed. In the case of the ΔrelQ mutant, resistance could be acquisition, the presence of relA and spoT has been implicated
1511 DOI: 10.1021/acsinfecdis.9b00204
ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

in the increased expression of an E. coli integron integrase in persistent bacteremia, was found by WGS to bear a Leu152Phe
biofilms.13 Integrases catalyze the insertion of incoming gene mutation in the hydrolase domain of Rel. This mutation arose
cassettes, such as resistance cassettes, into integrons that can after only 3 days of antibiotic therapy and was the only
then express them. Biofilms are associated with high levels of consistent mutation identified during the 28-day infection (the
resistance gene transfer, and E. coli biofilms exhibit elevated parental isolate already possessed numerous resistance genes/
expression of a class 1 integron integrase compared with mutations). The rel mutation was shown to confer elevated
planktonic growth.13 Deletion of relA and spoT abrogated this (p)ppGpp levels but no significant tolerance to daptomycin or
effect and returned biofilm integrase expression to planktonic linezolid in planktonic culture. However, in biofilm culture the
levels (in the absence of any effect on biofilm density). The rel mutant could not be eradicated by high concentrations of
explicit effect of relA-spoT deletion on the transfer of resistance daptomycin, linezolid, or vancomycin (unlike the wildtype).
genes in biofilms, and other bacterial populations, remains to This study is the first example of the clinical significance of the
be tested. stringent response in relation to antibiotic therapy.
In Vivo Effects of (p)ppGpp on Antibiotic Efficacy. It is It should be noted that, while there are only two reports to
clear that the stringent response and (p)ppGpp are important date of clinical isolates bearing stringent response-activating
contributors to antibiotic tolerance and resistance in vitro. mutations, the stringent response is almost certainly activated
There is also now growing evidence of the in vivo importance under many infection states and a contributor to persistent
of these relationships. Here, we use the term “in vitro” to refer infections. Bacteria experience a range of nutrient and other
to studies of bacteria in isolation (i.e., in test tube stresses when transitioning through the human body,88 and the
experiments), and the term “in vivo” to describe animal difficulty in treating many susceptible (by MIC) infections is
model studies. Murine infection models of M. tuberculosis and strongly suspected to lie in the presence of slow-growing,
P. aeruginosa have been used with RSH knockouts to tolerant populations of bacteria.33,89 In fact, several studies
demonstrate the contribution of the stringent response to have demonstrated the heterogeneity of bacterial growth rates
antibiotic tolerance in vivo. During the chronic phase of within host tissues, as well as the positive correlation between
infection, wildtype M. tuberculosis is naturally tolerant to the growth rate and antibiotic killing.90,91 We suspect that, with
antibiotic isoniazid; a Δrel mutant was able to establish a further research, the clinical significance of the stringent
chronic infection in mice but was susceptible to significant response will prove to have been greatly underappreciated.
killing by isoniazid.54 Similarly, the fluoroquinolone ofloxacin
was relatively ineffective in mice infected with stationary phase
P. aeruginosa, but survival rates in mice infected with a
■ THE SEARCH FOR STRINGENT RESPONSE
INHIBITORS
ΔrelAΔspoT mutant and treated with ofloxacin were The overwhelming evidence reviewed here that links
significantly higher.7 Deletion of relA and spoT also increased (p)ppGpp and the stringent response to antibiotic tolerance
the efficacy of ofloxacin in a murine biofilm model. These and resistance, as well as virulence,3 has led to growing interest
studies provide a direct link between the stringent response in the development of stringent response inhibitors. There are
and antibiotic efficacy in vivo. currently two approaches being taken: (i) inhibition of
By far, most of the studies reviewed here have exploited gene (p)ppGpp synthesis by Rel/RelA using substrate analogs and
knockouts or methods of specific nutritional deprivation to (ii) indirect abrogation of (p)ppGpp synthesis and accumu-
induce/prevent (p)ppGpp production and study the effects on lation through the use of protein synthesis inhibitors or
antibiotic efficacy. However, does this provide a biologically- cationic peptides.
and clinically relevant model of stringent response activation Design and Discovery of (p)ppGpp Synthetase
and its implications for therapy? In 2010, Gao et al. published Inhibitors. The first reported inhibitors of (p)ppGpp
the first report of a clinical isolate in which the stringent synthetase activity were a series of ppGpp analogs, the most
response had been partially activated.34 This MRSA isolate was potent of which bore bis-phosphonate substitutions at
associated with a case of persistent and recurrent bacteremia positions 5′ and 3′ of the ribose ring in place of the
that required >100 days of antibiotic treatment to cure. During pyrophosphate moieties.92 Further development of ppGpp
the course of treatment, bacterial culture revealed that the analogs led to relacin (Figure 4).93 Relacin is a 2′-
original isolate had evolved a SCV phenotype. WGS of the deoxyguanosine-based analog of ppGpp in which the
parental and SCV isolates identified four nonsynonymous pyrophosphate moieties have been replaced with glycyl-glycine
mutations, including one in the rel gene. This Phe128Tyr dipeptides linked to the sugar ring by a carbamate bridge. It
mutation occurred in the hydrolase domain of Rel and was rationally designed based on the structure of Rel from
conferred elevated (p)ppGpp levels, reduced growth rate, and Streptococcus dysgalactiae,20 and modeling indicates that it
small colonies, presumably due to reduced hydrolase activity. occupies much of the synthetase active site and forms a range
This discovery provided the first evidence of the clinical of hydrogen bonds and hydrophobic interactions. In vitro and
relevance of stringent response activation. In addition to the rel in vivo data suggest that relacin inhibits (p)ppGpp synthesis by
mutation, the SCV also possessed three mutations associated both Rel and RelA, but millimolar concentrations are required.
with antibiotic resistance or reduced susceptibility. For two of At the cellular level, relacin led to the death of B. subtilis (and
these mutations, it is not known whether they arose before or other bacteria) with an estimated IC50 of 200 μM and
after the rel mutation; however, the mutation that led to prevented spore and biofilm formation. Prior to bacterial
reduced susceptibility to linezolid occurred after the death, relacin also induced a prolonged exponential phase, a
emergence of the SCV phenotype. The rel mutation was not feature that has been observed previously in a spoT knockout
investigated for its ability to confer antibiotic tolerance. of Helicobacter pylori.94 Further evidence for the specificity of
Recently, a second clinical isolate bearing a stringent relacin was provided by the resistance of a B. subtilis rel
response-activating mutation has been identified.6 An isolate knockout to the bactericidal effect of relacin. Given the
of Enterococcus faecium, again associated with a case of structural differences between long RSHs and SASs,20,95,96 it is
1512 DOI: 10.1021/acsinfecdis.9b00204
ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

testing revealed all of these to be nonspecific antibacterials.


Relacin was also put through this screen as a benchmark, but
failed to inhibit the growth of the ΔrelPΔrelQ mutant in either
growth medium at 2 mM. A more recent set of ppGpp analogs
also had no effect on growth in this system, despite exhibiting
IC50 values as low as 50 μM against E. coli RelA.98
An alternate high-throughput screening approach for
inhibitors of Rel that proved more successful was recently
reported.54 This screen employed a C-terminal truncated
version of Rel from M. tuberculosis and a novel (p)ppGpp
synthetase assay based on detection of released AMP. From a
library of 2 million compounds, 791 hits were obtained and
further screened in Rel-specific whole-cell assays, leading to the
identification of compound X9 as the lead candidate (Figure
4). X9 exhibited an IC50 of ∼15 μM against purified Rel and 2
μM against nutrient-starved M. tuberculosis in vitro. At 4 μM, it
Figure 4. Substrates and chemical inhibitors of RelA/Rel. Shown are also enhanced the susceptibility of nutrient-starved M.
the chemical structures of natural, rationally designed and screened tuberculosis to killing by isoniazid. This synergy (and the
inhibitors of RelA/Rel identified to date. RelA/rel substrates (boxed) activity of X9 alone against M. tuberculosis) could be abolished
ppGpp and GDP are shown for comparison. Relacin is a rationally by overexpression of Rel, thereby providing further evidence
designed analog of ppGpp, whereas vitamin C is a natural inhibitor that Rel is the target of X9. The molecular details of how X9
that resembles GDP. Compounds AC and AB are designed acetylated binds to, and inhibits, Rel is not known. Nevertheless, X9
and acetylated-benzoylated derivatives of guanosine, respectively. displays the most potent activity of any Rel inhibitor to date.
Compound X9 is the lead candidate inhibitor from a high-throughput
Indirect Strategies To Prevent or Reverse Stringent
screen against Rel.
Response Induction. In addition to targeting the synthetase
activity of Rel/RelA directly, indirect methods aimed at
not surprising that relacin and other related ppGpp analogs preventing the synthesis and/or accumulation of (p)ppGpp
have no inhibitory effect against RelQ from E. faecalis.97,98 have been investigated. It was first noted more than 20 years
Several modifications to relacin have since improved its ago that the protein synthesis inhibitor chloramphenicol
potency. Replacement of the glycyl-glycine dipeptides with prevents (p)ppGpp accumulation (and concomitant ampicillin
glutamyl-glutamine dipeptides led to compound 2d.99 This tolerance) in E. coli.44 More recently, the protein synthesis
compound inhibited the (p)ppGpp synthetase activity of E. coli inhibitor thiostrepton was shown to inhibit the synthetase
RelA with greater potency than relacin; however, millimolar function of RelA in a reconstituted biochemical system.41
concentrations were still required. Based on the studies of Furthermore, three different protein synthesis inhibitors
relacin and other ppGpp analogs, functionalization of the (chloramphenicol, thiostrepton, and tetracycline) inhibited
amine group at the C-2 position of the guanine moiety in (p)ppGpp accumulation in E. coli and B. subtilis exposed to
guanosine led to compounds AC and AB (Figure 4).100 These mupirocin.41 The general ability of protein synthesis inhibitors
compounds exhibited IC50 values in the range of 40 μM against to abrogate (p)ppGpp accumulation is thought to stem from
a mycobacterial Rel. Compound AB was found to be the more interference in the association between Rel/RelA and stalled
potent of the two agents, but both compounds had a negative ribosomes (specifically the unacylated A site) and/or the
impact on long-term survival and biofilm formation in increase in tRNA aminoacylation that results from the
Mycobacterium smegmatis. Interestingly, Vitamin C exhibits inhibition of translation. Unfortunately, neither chloramphe-
potent bactericidal activity against mycobacterial species and nicol or tetracycline was able to resensitize mupirocin-treated
has been proposed as a possible scaffold for future Rel cultures to ampicillin.41
inhibitors by virtue of its structural analogy with GDP (Figure An alternate strategy aimed at reversing stringent response
4).101 Vitamin C itself binds with weak affinity to Rel from M. activation is the promotion of (p)ppGpp degradation.
smegmatis, and millimolar concentrations are required to (p)ppGpp and the stringent response have been linked to
inhibit (p)ppGpp synthesis in vitro. biofilm formation in a range of pathogenic bacteria,3 and
An auxotrophy-based high-throughput screen has been biofilms are often used as an experimental model of an
developed in an attempt to identify novel Rel inhibitors.102 activated stringent response.7,52 A number of antibiofilm
This screen employs a ΔrelPΔrelQ knockout of B. subtilis so cationic peptides have been identified and proposed to act
that effects of any inhibitors on (p)ppGpp production by Rel via disruption of the stringent response.104,105 These peptides
are not masked by SAS-mediated compensation. The screen display synergy with conventional antibiotics against in vitro
relies on a key phenotypic difference between cells that do and biofilms and in animal models of infection.105,106 1018 is a
do not possess the ability to synthesize (p)ppGpp, namely, small, synthetic, L-amino acid peptide derived from a bovine
differential growth when starved of lysine or valine. Wildtype host defense peptide. It exhibits potent activity against Gram-
and (p)ppGpp null mutants can grow equally well in the positive and Gram-negative bacteria in biofilms, but not in
absence of lysine, but growth of a (p)ppGpp null mutant is planktonic culture, and overproduction of (p)ppGpp (either
strictly dependent on valine.103 Therefore, specific inhibition of through overexpression of RelA or addition of SHX) leads to
Rel can be detected as inhibition of growth of the ΔrelPΔrelQ resistance to 1018.104 Furthermore, 1018 prevents (p)ppGpp
mutant on valine but not lysine. A library of 17,500 accumulation in a range of bacteria and appears to eliminate it
compounds were screened using this approach, and five postaccumulation in P. aeruginosa. A distinct and more potent
potential inhibitors were identified. Unfortunately, further D-amino acid peptide, DJK-5, exhibits similar activity against

1513 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

(p)ppGpp.105 The mechanism of action of these peptides to clear that there is a vast number of genes that could induce
prevent accumulation, and promote degradation, of (p)ppGpp (p)ppGpp production if mutated. This will likely make genetic
is proposed to occur via a direct interaction with (p)ppGpp.104 identification of stringent response activation in clinical isolates
In coprecipitation and NMR spectroscopy experiments, 1018 more complicated. Furthermore, current methods of
exhibited greater binding to ppGpp than other nucleotides. (p)ppGpp quantitation (e.g., thin layer chromatography with
32
The disappearance of (p)ppGpp from 1018- and DJK-5- P-labeled material) are laborious and not suited to screening
treated cells (under detection conditions that dissociate of clinical isolates. An alternative route to identifying stringent
peptide-ppGpp complexes) led the authors to suggest that response activation would be through the detection of its
these peptides do not sequester (p)ppGpp, but instead effects (e.g., tolerance). The current gold standard for
promote its degradation by the cell.104 A mechanism for how detection of tolerance is the time-kill assay. We strongly
these peptides promote (p)ppGpp degradation has not been advocate the use of time-kills to distinguish between tolerance
proposed. and persistence in the literature; however, these assays are not
It should be noted that the (p)ppGpp-specific activity of well suited to high-throughput clinical use. A number of
1018 has been questioned by Andresen et al.102,107 First, 1018 recently developed agar-based assays for tolerance57,108 may
was put through the auxotrophy screen for Rel inhibitors and facilitate the identification of tolerance (related to the stringent
found to inhibit growth of the ΔrelPΔrelQ mutant under lysine response or otherwise) among clinical isolates. Overall, more
or valine deprivation with similar potency. 1018 also exhibited research is required to assess the extent of the problem posed
similar potency against a (p)ppGpp null mutant of B. subtilis, by (p)ppGpp and the stringent response in the clinic.


which lacks any (p)ppGpp. In a separate study, a ΔrelAΔspoT
mutant of E. coli was actually slightly sensitized to, rather than AUTHOR INFORMATION
protected from, 1018 inhibition compared with the wild-
type.107 Second, under different experimental conditions to Corresponding Authors
those employed previously, planktonic and biofilm cultures of *Phone: 250-721-7084. Fax: 250-721-8855. E-mail: jhobbs@
P. aeruginosa were seemingly equally inhibited by 1018 (and a uvic.ca.
control peptide in which the amino acid sequence of 1018 had *Phone: 250-472-4168. Fax: 250-721-8855. E-mail: boraston@
been reversed). Therefore, the authors concluded that the uvic.ca.
antimicrobial effect of 1018 and related peptides does not rely ORCID
on specific recognition of (p)ppGpp. This area requires further Alisdair B. Boraston: 0000-0001-6417-0592
investigation; however, it may be, as suggested by Andresen et
al.,107 that peptide exposure stimulates indirect degradation of Notes
(p)ppGpp in much the same way that certain protein synthesis The authors declare no competing financial interest.
inhibitors are known to “relax” the stringent response.41,44

■ CONCLUSIONS
■ ACKNOWLEDGMENTS
The writing of this review article was supported by a grant
As reviewed here, there is now overwhelming evidence that from the British Columbia Lung Association awarded to J.K.H.
implicates (p)ppGpp and activation of the stringent response and A.B.B., and a Canadian Institutes of Health Research
in the ability of bacteria to evade and resist the action of many operating grant awarded to A.B.B. (PJT 159786).


antibiotics. This ability is not only mediated by the canonical
long RSHs, but also accessory SASs and the wider network of ABBREVIATIONS
potential stringent response inducers (e.g., tRNA synthetases,
HipA) and downstream facilitators (e.g., DksA, RsgA). ppGpp guanosine 5′-diphosphate 3′-diphosphate; pppGpp
Furthermore, it should be noted that antibiotic tolerance has guanosine 5′-triphosphate 3-diphosphate; MDK minimum
also been linked to (p)ppGpp production outside of the duration for killing; MIC minimum inhibitory concentration;
classical stringent response.15,22 This recent appreciation of the MRSA methicillin-resistant Staphylococcus aureus; PBP pen-
complexities and extent of (p)ppGpp control leads to two icillin-binding protein; RSH RelA/SpoT homologue; SAH
considerations in relation to antibiotic efficacy and treatment small alarmone hydrolases; SAS small alarmone synthetase;
strategies. First, attempts to target (p)ppGpp production with SCV small colony variant; SHX serine hydroxamate; WGS
novel therapeutics need to consider the RSH makeup of the whole genome sequencing
target organism(s). As we have seen with deletion mutants
(which mimic the effects of enzyme inhibition to some extent),
abrogating the activity of one RSH can actually lead to
■ REFERENCES
(1) Cashel, M., and Gallant, J. (1969) Two compounds implicated in
stringent response activation via the remaining RSHs8,40,81 the function of the RC gene of Escherichia coli. Nature 221 (5183),
(Table 1). In terms of the ongoing search for inhibitors of the 838−841.
bifunctional Rel enzyme, which is the most widespread RSH, (2) Cashel, M., Gentry, D., Hernandez, V., and Vinella, D. (1996)
what has yet to be addressed is whether inhibition of The stringent response. In Escherichia coli and Salmonella: Cellular and
synthetase activity also inhibits hydrolase function (e.g., by Molecular Biology (Neidhardt, F. C., Curtiss, R. I., Ingraham, J. L., Lin,
“locking” Rel in the synthetase-on/hydrolase-off conforma- E. C. C., Low, K. B., Magasanik, B., Reznikoff, W. S., Riley, M.,
Schaechter, M., and Umbarger, H. E., Eds.) pp 1458−1496, ASM
tion). In many pathogenic bacteria this would lead to an Press, Washington, D.C.
overproduction of (p)ppGpp by SASs. As such, therapeutic (3) Dalebroux, Z. D., Svensson, S. L., Gaynor, E. C., and Swanson,
strategies that promote the degradation of (p)ppGpp (as M. S. (2010) ppGpp conjures bacterial virulence. Microbiol. Mol. Biol.
proposed for cationic peptides) may be more useful. Second, Rev. 74 (2), 171−199.
the two stringent response-activated clinical isolates reported (4) Boutte, C. C., and Crosson, S. (2013) Bacterial lifestyle shapes
to date both possessed mutations in rel; however, it is now stringent response activation. Trends Microbiol. 21 (4), 174−180.

1514 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

(5) Gaca, A. O., Colomer-Winter, C., and Lemos, J. A. (2015) Many and Wolz, C. (2018) Regulation of the opposing (p)ppGpp
means to a common end: the intricacies of (p)ppGpp metabolism and synthetase and hydrolase activities in a bifunctional RelA/SpoT
its control of bacterial homeostasis. J. Bacteriol. 197 (7), 1146−1156. homologue from Staphylococcus aureus. PLoS Genet. 14 (7),
(6) Honsa, E. S., Cooper, V. S., Mhaissen, M. N., Frank, M., Shaker, No. e1007514.
J., Iverson, A., Rubnitz, J., Hayden, R. T., Lee, R. E., Rock, C. O., (22) Geiger, T., Kästle, B., Gratani, F. L., Goerke, C., and Wolz, C.
Tuomanen, E. I., Wolf, J., and Rosch, J. W. (2017) RelA mutant (2014) Two small (p)ppGpp synthases in Staphylococcus aureus
Enterococcus faecium with multiantibiotic tolerance arising in an mediate tolerance against cell envelope stress conditions. J. Bacteriol.
immunocompromised host. mBio 8 (1), No. e02124-16. 196 (4), 894−902.
(7) Nguyen, D., Joshi-Datar, A., Lepine, F., Bauerle, E., Olakanmi, (23) Jimmy, S., Saha, C. K., Stavropoulos, C., Garcia-Pino, A.,
O., Beer, K., McKay, G., Siehnel, R., Schafhauser, J., Wang, Y., Hauryliuk, V., and Atkinson, G. C. (2019) Discovery of small
Britigan, B. E., and Singh, P. K. (2011) Active starvation responses alarmone synthetases and their inhibitors as toxin-antitoxin loci.
mediate antibiotic tolerance in biofilms and nutrient-limited bacteria. bioRxiv, DOI: 10.1101/575399.
Science 334 (6058), 982−986. (24) Maisonneuve, E., and Gerdes, K. (2014) Molecular
(8) Abranches, J., Martinez, A. R., Kajfasz, J. K., Chávez, V., Garsin, mechanisms underlying bacterial persisters. Cell 157 (3), 539−548.
D. A., and Lemos, J. A. (2009) The molecular alarmone (p)ppGpp (25) Brauner, A., Fridman, O., Gefen, O., and Balaban, N. Q. (2016)
mediates stress responses, vancomycin tolerance, and virulence in Distinguishing between resistance, tolerance and persistence to
Enterococcus faecalis. J. Bacteriol. 191 (7), 2248−2256. antibiotic treatment. Nat. Rev. Microbiol. 14 (5), 320−330.
(9) Corrigan, R. M., Bellows, L. E., Wood, A., and Gründling, A. (26) Corona, F., and Martinez, J. L. (2013) Phenotypic resistance to
(2016) ppGpp negatively impacts ribosome assembly affecting growth antibiotics. Antibiotics (Basel, Switz.) 2 (2), 237−255.
and antimicrobial tolerance in Gram-positive bacteria. Proc. Natl. (27) Levin, B. R., and Rozen, D. E. (2006) Non-inherited antibiotic
Acad. Sci. U. S. A. 113 (12), E1710−E1719. resistance. Nat. Rev. Microbiol. 4 (7), 556−562.
(10) Bokinsky, G., Baidoo, E. E. K., Akella, S., Burd, H., Weaver, D., (28) Kahl, B. C., Becker, K., and Löffler, B. (2016) Clinical
Alonso-Gutierrez, J., García-Martín, H., Lee, T. S., and Keasling, J. D. significance and pathogenesis of staphylococcal small colony variants
(2013) HipA-triggered growth arrest and β-lactam tolerance in in persistent infections. Clin. Microbiol. Rev. 29 (2), 401−427.
Escherichia coli are mediated by relA-dependent ppGpp synthesis. J. (29) Meylan, S., Andrews, I. W., and Collins, J. J. (2018) Targeting
Bacteriol. 195 (14), 3173−3182. antibiotic tolerance, pathogen by pathogen. Cell 172 (6), 1228−1238.
(11) Aedo, S., and Tomasz, A. (2016) Role of the stringent stress (30) Levin-Reisman, I., Ronin, I., Gefen, O., Braniss, I., Shoresh, N.,
response in the antibiotic resistance phenotype of methicillin-resistant and Balaban, N. Q. (2017) Antibiotic tolerance facilitates the
Staphylococcus aureus. Antimicrob. Agents Chemother. 60 (4), 2311− evolution of resistance. Science 355 (6327), 826−830.
2317. (31) Germain, E., Castro-Roa, D., Zenkin, N., and Gerdes, K. (2013)
(12) Mwangi, M. M., Kim, C., Chung, M., Tsai, J., Vijayadamodar, Molecular mechanism of bacterial persistence by HipA. Mol. Cell 52
G., Benitez, M., Jarvie, T. P., Du, L., and Tomasz, A. (2013) Whole- (2), 248−254.
genome sequencing reveals a link between β-lactam resistance and (32) Brauner, A., Shoresh, N., Fridman, O., and Balaban, N. Q.
synthetases of the alarmone (p)ppGpp in Staphylococcus aureus. (2017) An experimental framework for quantifying bacterial
Microb. Drug Resist. 19 (3), 153−159.
tolerance. Biophys. J. 112 (12), 2664−2671.
(13) Strugeon, E., Tilloy, V., Ploy, M.-C., and Re, S. Da. (2016) The
(33) Tuomanen, E., Durack, D. T., and Tomasz, A. (1986)
stringent response promotes antibiotic resistance dissemination by
Antibiotic tolerance among clinical isolates of bacteria. Antimicrob.
regulating integron integrase expression in biofilms. mBio 7 (4),
Agents Chemother. 30 (4), 521−527.
No. e00868-16.
(34) Gao, W., Chua, K., Davies, J. K., Newton, H. J., Seemann, T.,
(14) Hauryliuk, V., Atkinson, G. C., Murakami, K. S., Tenson, T.,
Harrison, P. F., Holmes, N. E., Rhee, H.-W., Hong, J.-I., Hartland, E.
and Gerdes, K. (2015) Recent functional insights into the role of
(p)ppGpp in bacterial physiology. Nat. Rev. Microbiol. 13 (5), 298− L., Stinear, T. P., and Howden, B. P. (2010) Two novel point
309. mutations in clinical Staphylococcus aureus reduce linezolid suscept-
(15) Gaca, A. O., Kajfasz, J. K., Miller, J. H., Liu, K., Wang, J. D., ibility and switch on the stringent response to promote persistent
Abranches, J., and Lemos, J. A. (2013) Basal levels of (p)ppGpp in infection. PLoS Pathog. 6, No. e1000944.
Enterococcus faecalis: the magic beyond the stringent response. mBio 4 (35) Tuomanen, E., Cozens, R., Tosch, W., Zak, O., and Tomasz, A.
(5), No. e00646-13. (1986) The rate of killing of Escherichia coli by β-lactam antibiotics is
(16) Kriel, A., Bittner, A. N., Kim, S. H., Liu, K., Tehranchi, A. K., strictly proportional to the rate of bacterial growth. Microbiology 132
Zou, W. Y., Rendon, S., Chen, R., Tu, B. P., and Wang, J. D. (2012) (5), 1297−1304.
Direct regulation of GTP homeostasis by (p)ppGpp: a critical (36) Eng, R. H., Padberg, F. T., Smith, S. M., Tan, E. N., and
component of viability and stress resistance. Mol. Cell 48 (2), 231− Cherubin, C. E. (1991) Bactericidal effects of antibiotics on slowly
241. growing and nongrowing bacteria. Antimicrob. Agents Chemother. 35
(17) Atkinson, G. C., Tenson, T., and Hauryliuk, V. (2011) The (9), 1824−1828.
RelA/SpoT homolog (RSH) superfamily: distribution and functional (37) Goodell, W., and Tomasz, A. (1980) Alteration of Escherichia
evolution of ppGpp synthetases and hydrolases across the tree of life. coli murein during amino acid starvation. J. Bacteriol. 144 (3), 1009−
PLoS One 6 (8), No. e23479. 1016.
(18) Irving, S. E., and Corrigan, R. M. (2018) Triggering the (38) Tuomanen, E., and Tomasz, A. (1986) Induction of autolysis in
stringent response: signals responsible for activating (p)ppGpp nongrowing Escherichia coli. J. Bacteriol. 167 (3), 1077−1080.
synthesis in bacteria. Microbiology 164 (3), 268−276. (39) Rodionov, D. G., Pisabarro, A. G., de Pedro, M. A., Kusser, W.,
(19) Mechold, U., Murphy, H., Brown, L., and Cashel, M. (2002) and Ishiguro, E. E. (1995) Beta-lactam-induced bacteriolysis of amino
Intramolecular regulation of the opposing (p)ppGpp catalytic acid-deprived Escherichia coli is dependent on phospholipid synthesis.
activities of Rel(Seq), the Rel/Spo enzyme from Streptococcus J. Bacteriol. 177 (4), 992−997.
equisimilis. J. Bacteriol. 184 (11), 2878−2888. (40) Kim, H. Y., Go, J., Lee, K.-M., Oh, Y. T., and Yoon, S. S. (2018)
(20) Hogg, T., Mechold, U., Malke, H., Cashel, M., and Hilgenfeld Guanosine tetra- and pentaphosphate increase antibiotic tolerance by
(2004) Conformational antagonism between opposing active sites in a reducing reactive oxygen species production in Vibrio cholerae. J. Biol.
bifunctional RelA/SpoT homolog modulates (p)ppGpp metabolism Chem. 293 (15), 5679−5694.
during the stringent response. Cell 117 (1), 57−68. (41) Kudrin, P., Varik, V., Oliveira, S. R. A., Beljantseva, J., Del Peso
(21) Gratani, F. L., Horvatek, P., Geiger, T., Borisova, M., Mayer, C., Santos, T., Dzhygyr, I., Rejman, D., Cava, F., Tenson, T., and
Grin, I., Wagner, S., Steinchen, W., Bange, G., Velic, A., Maček, B., Hauryliuk, V. (2017) Subinhibitory concentrations of bacteriostatic

1515 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

antibiotics induce relA-dependent and relA-independent tolerance to Rose, W. E. (2018) Daptomycin selects for genetic and phenotypic
β-lactams. Antimicrob. Agents Chemother. 61 (4), No. e02173-16. adaptations leading to antibiotic tolerance in MRSA. J. Antimicrob.
(42) Singh, M., Matsuo, M., Sasaki, T., Morimoto, Y., Hishinuma, T., Chemother. 73 (8), 2030−2033.
and Hiramatsu, K. (2016) In vitro tolerance of drug-naive Staph- (59) Hansen, S., Lewis, K., and Vulić, M. (2008) Role of global
ylococcusaureus strain FDA209P to vancomycin. Antimicrob. Agents regulators and nucleotide metabolism in antibiotic tolerance in
Chemother. 61 (2), No. e01154-16. Escherichia coli. Antimicrob. Agents Chemother. 52 (8), 2718−2726.
(43) Kusser, W., and Ishiguro, E. E. (1985) Involvement of the relA (60) Kaspy, I., Rotem, E., Weiss, N., Ronin, I., Balaban, N. Q., and
gene in the autolysis of Escherichia coli induced by inhibitors of Glaser, G. (2013) HipA-mediated antibiotic persistence via
peptidoglycan biosynthesis. J. Bacteriol. 164 (2), 861−865. phosphorylation of the glutamyl-tRNA-synthetase. Nat. Commun. 4
(44) Rodionov, D. G., and Ishiguro, E. E. (1995) Direct correlation (1), 3001.
between overproduction of guanosine 3′,5′-bispyrophosphate (61) Falla, T. J., and Chopra, I. (1998) Joint tolerance to beta-lactam
(ppGpp) and penicillin tolerance in Escherichia coli. J. Bacteriol. 177 and fluoroquinolone antibiotics in Escherichia coli results from
(15), 4224−4229. overexpression of hipA. Antimicrob. Agents Chemother. 42 (12),
(45) Traxler, M. F., Summers, S. M., Nguyen, H.-T., Zacharia, V. M., 3282−3284.
Hightower, G. A., Smith, J. T., and Conway, T. (2008) The global, (62) Correia, F. F., D’Onofrio, A., Rejtar, T., Li, L., Karger, B. L.,
ppGpp-mediated stringent response to amino acid starvation in Makarova, K., Koonin, E. V., and Lewis, K. (2006) Kinase activity of
Escherichia coli. Mol. Microbiol. 68 (5), 1128−1148. overexpressed HipA is required for growth arrest and multidrug
(46) Vinella, D., D’Ari, R., Jaffé, A., and Bouloc, P. (1992) Penicillin tolerance in Escherichia coli. J. Bacteriol. 188 (24), 8360−8367.
binding protein 2 is dispensable in Escherichia coli when ppGpp (63) Garoff, L., Huseby, D. L., Praski Alzrigat, L., and Hughes, D.
synthesis is induced. EMBO J. 11 (4), 1493−1501. (2018) Effect of aminoacyl-tRNA synthetase mutations on suscept-
(47) Joseleau-Petit, D., Thévenet, D., and D’Ari, R. (1994) ppGpp ibility to ciprofloxacin in Escherichia coli. J. Antimicrob. Chemother. 73
concentration, growth without PBP2 activity, and growth-rate control (12), 3285−3292.
in Escherichia coli. Mol. Microbiol. 13 (5), 911−917. (64) Paul, B. J., Barker, M. M., Ross, W., Schneider, D. A., Webb, C.,
(48) Pomares, M. F., Vincent, P. A., Farías, R. N., and Salomón, R. Foster, J. W., and Gourse, R. L. (2004) DksA: a critical component of
A. (2008) Protective action of ppGpp in microcin J25-sensitive the transcription initiation machinery that potentiates the regulation
strains. J. Bacteriol. 190 (12), 4328−4334. of rRNA promoters by ppGpp and the initiating NTP. Cell 118 (3),
(49) Viducic, D., Ono, T., Murakami, K., Susilowati, H., Kayama, S., 311−322.
Hirota, K., and Miyake, Y. (2006) Functional analysis of spoT, relA (65) Jomaa, A., Stewart, G., Martín-Benito, J., Zielke, R., Campbell,
and dksA genes on quinolone tolerance in Pseudomonas aeruginosa T. L., Maddock, J. R., Brown, E. D., and Ortega, J. (2011)
under nongrowing condition. Microbiol. Immunol. 50 (4), 349−357. Understanding ribosome assembly: the structure of in vivo assembled
(50) Martins, D., McKay, G., Sampathkumar, G., Khakimova, M., immature 30S subunits revealed by cryo-electron microscopy. RNA 17
English, A. M., and Nguyen, D. (2018) Superoxide dismutase activity (4), 697−709.
confers (p)ppGpp-mediated antibiotic tolerance to stationary-phase (66) Korch, S. B., Henderson, T. A., and Hill, T. M. (2003)
Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. U. S. A. 115 (39), Characterization of the hipa7 allele of Escherichia coli and evidence
9797−9802. that high persistence is governed by (p)ppGpp synthesis. Mol.
(51) Fung, D. K. C., Chan, E. W. C., Chin, M. L., and Chan, R. C. Y. Microbiol. 50 (4), 1199−1213.
(2010) Delineation of a bacterial starvation stress response network (67) Germain, E., Roghanian, M., Gerdes, K., and Maisonneuve, E.
which can mediate antibiotic tolerance development. Antimicrob. (2015) Stochastic induction of persister cells by HipA through
Agents Chemother. 54 (3), 1082−1093. (p)ppGpp-mediated activation of mRNA endonucleases. Proc. Natl.
(52) Khakimova, M., Ahlgren, H. G., Harrison, J. J., English, A. M., Acad. Sci. U. S. A. 112 (16), 5171−5176.
and Nguyen, D. (2013) The stringent response controls catalases in (68) Moyed, H. S., and Bertrand, K. P. (1983) hipA, a newly
Pseudomonas aeruginosa and is required for hydrogen peroxide and recognized gene of Escherichia coliK-12 that affects frequency of
antibiotic tolerance. J. Bacteriol. 195 (9), 2011−2020. persistence after inhibition of murein synthesis. J. Bacteriol. 155 (2),
(53) Primm, T. P., Andersen, S. J., Mizrahi, V., Avarbock, D., Rubin, 768−775.
H., and Barry, C. E. (2000) The stringent response of Mycobacterium (69) Hobbs, J. K., Miller, K., O’Neill, A. J., and Chopra, I. (2008)
tuberculosis is required for long-term survival. J. Bacteriol. 182 (17), Consequences of daptomycin-mediated membrane damage in Staph-
4889−4898. ylococcus aureus. J. Antimicrob. Chemother. 62 (5), 1003−1008.
(54) Dutta, N. K., Klinkenberg, L. G., Vazquez, M.-J., Segura-Carro, (70) Windels, E. M., Michiels, J. E., Fauvart, M., Wenseleers, T., Van
D., Colmenarejo, G., Ramon, F., Rodriguez-Miquel, B., Mata-Cantero, den Bergh, B., and Michiels, J. (2019) Bacterial persistence promotes
L., Porras-De Francisco, E., Chuang, Y.-M., Rubin, H., Lee, J. J., Eoh, the evolution of antibiotic resistance by increasing survival and
H., Bader, J. S., Perez-Herran, E., Mendoza-Losana, A., and mutation rates. ISME J. 13 (5), 1239−1251.
Karakousis, P. C. (2019) Inhibiting the stringent response blocks (71) Gajdács, M. (2019) The continuing threat of methicillin-
Mycobacterium tuberculosis entry into quiescence and reduces resistant Staphylococcus aureus. Antibiotics 8 (2), 52.
persistence. Sci. Adv. 5 (3), No. eaav2104. (72) Tomasz, A., Nachman, S., and Leaf, H. (1991) Stable classes of
(55) Gaca, A. O., Abranches, J., Kajfasz, J. K., and Lemos, J. A. phenotypic expression in methicillin-resistant clinical isolates of
(2012) Global transcriptional analysis of the stringent response in staphylococci. Antimicrob. Agents Chemother. 35 (1), 124−129.
Enterococcus faecalis. Microbiology 158 (Pt8), 1994−2004. (73) Dordel, J., Kim, C., Chung, M., Pardos de la Gándara, M.,
(56) Varik, V., Oliveira, S. R. A., Hauryliuk, V., and Tenson, T. Holden, M. T. J., Parkhill, J., de Lencastre, H., Bentley, S. D., and
(2016) Composition of the outgrowth medium modulates wake-up Tomasz, A. (2014) Novel determinants of antibiotic resistance:
kinetics and ampicillin sensitivity of stringent and relaxed Escherichia identification of mutated loci in highly methicillin-resistant sub-
coli. Sci. Rep. 6 (1), 22308. populations of methicillin-resistant Staphylococcus aureus. mBio 5 (2),
(57) Matsuo, M., Hiramatsu, M., Singh, M., Sasaki, T., Hishinuma, No. e01000.
T., Yamamoto, N., Morimoto, Y., Kirikae, T., and Hiramatsu, K. (74) Pardos de la Gándara, M., Borges, V., Chung, M., Milheiriço,
(2019) Genetic and transcriptomic analyses of ciprofloxacin-tolerant C., Gomes, J. P., de Lencastre, H., and Tomasz, A. (2018) Genetic
Staphylococcus aureus isolated by the replica plating tolerance isolation determinants of high-level oxacillin resistance in methicillin-
system (REPTIS). Antimicrob. Agents Chemother. 63 (2), No. e02019- resistantStaphylococcus aureus. Antimicrob. Agents Chemother. 62 (6),
18. No. e00206-18.
(58) Berti, A. D., Shukla, N., Rottier, A. D., McCrone, J. S., Turner, (75) Kim, C., Mwangi, M., Chung, M., Milheirço, C., de Lencastre,
H. M., Monk, I. R., Baines, S. L., Howden, B. P., Proctor, R. A., and H., and Tomasz, A. (2013) The mechanism of heterogeneous beta-

1516 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517
ACS Infectious Diseases Review

lactam resistance in MRSA: key role of the stringent stress response. (93) Wexselblatt, E., Oppenheimer-Shaanan, Y., Kaspy, I., London,
PLoS One 8 (12), No. e82814. N., Schueler-Furman, O., Yavin, E., Glaser, G., Katzhendler, J., and
(76) Chung, M., Kim, C. K., Conceiçaõ , T., Aires-De-Sousa, M., de Ben-Yehuda, S. (2012) Relacin, a novel antibacterial agent targeting
Lencastre, H., and Tomasz, A. (2016) Heterogeneous oxacillin- the stringent response. PLoS Pathog. 8 (9), No. e1002925.
resistant phenotypes and production of PBP2A by oxacillin- (94) Zhou, Y. N., Coleman, W. G., Yang, Z., Yang, Y., Hodgson, N.,
susceptible/mecA-positive MRSA strains from Africa. J. Antimicrob. Chen, F., and Jin, D. J. (2008) Regulation of cell growth during serum
Chemother. 71 (10), 2804−2809. starvation and bacterial survival in macrophages by the bifunctional
(77) Milheiriço, C., de Lencastre, H., and Tomasz, A. (2017) Full- enzyme SpoT in Helicobacter pylori. J. Bacteriol. 190 (24), 8025−8032.
genome sequencing identifies in the genetic background several (95) Manav, M. C., Beljantseva, J., Bojer, M. S., Tenson, T., Ingmer,
determinants that modulate the resistance phenotype in methicillin- H., Hauryliuk, V., and Brodersen, D. E. (2018) Structural basis for
resistant Staphylococcus aureus strains carrying the novel mecC gene. (p)ppGpp synthesis by the Staphylococcus aureus small alarmone
Antimicrob. Agents Chemother. 61 (3), No. e02500-16. synthetase RelP. J. Biol. Chem. 293 (9), 3254−3264.
(78) Kim, C. K., Milheiriço, C., de Lencastre, H., and Tomasz, A. (96) Steinchen, W., Schuhmacher, J. S., Altegoer, F., Fage, C. D.,
(2017) Antibiotic resistance as a stress response: recovery of high- Srinivasan, V., Linne, U., Marahiel, M. A., and Bange, G. (2015)
level oxacillin resistance in methicillin-resistant Staphylococcus aureus Catalytic mechanism and allosteric regulation of an oligomeric
“auxiliary” ( fem) mutants by induction of the stringent stress (p)ppGpp synthetase by an alarmone. Proc. Natl. Acad. Sci. U. S. A.
response. Antimicrob. Agents Chemother. 61 (8), No. e00313-17. 112 (43), 13348−13353.
(79) Matsuo, M., Hishinuma, T., Katayama, Y., and Hiramatsu, K. (97) Gaca, A. O., Kudrin, P., Colomer-Winter, C., Beljantseva, J.,
(2015) A mutation of RNA polymerase β’ subunit (rpoC) converts Liu, K., Anderson, B., Wang, J. D., Rejman, D., Potrykus, K., Cashel,
M., Hauryliuk, V., and Lemos, J. A. (2015) From (p)ppGpp to
heterogeneously vancomycin-intermediate Staphylococcus aureus
(pp)pGpp: characterization of regulatory effects of pGpp synthesized
(hVISA) into “slow VISA. Antimicrob. Agents Chemother. 59 (7),
by the small alarmone synthetase of Enterococcus faecalis. J. Bacteriol.
4215−4225.
197 (18), 2908−2919.
(80) Matsuo, M., Yamamoto, N., Hishinuma, T., and Hiramatsu, K.
(98) Beljantseva, J., Kudrin, P., Jimmy, S., Ehn, M., Pohl, R., Varik,
(2019) Identification of a novel gene associated with high-level β- V., Tozawa, Y., Shingler, V., Tenson, T., Rejman, D., and Hauryliuk,
lactam resistance in heterogeneous vancomycin-intermediate Staph- K. (2017) Molecular mutagenesis of ppgpp: turning a rela activator
ylococcus aureus strain Mu3 and methicillin-resistant S. aureus strain into an inhibitor. Sci. Rep. 7 (1), 41839.
N315. Antimicrob. Agents Chemother. 63 (2), No. e00712-18. (99) Wexselblatt, E., Kaspy, I., Glaser, G., Katzhendler, J., and Yavin,
(81) Bhawini, A., Pandey, P., Dubey, A. P., Zehra, A., Nath, G., and E. (2013) Design, synthesis and structure−activity relationship of
Mishra, M. N. (2019) RelQ mediates the expression of β-lactam novel relacin analogs as inhibitors of Rel proteins. Eur. J. Med. Chem.
resistance in methicillin-resistant Staphylococcus aureus. Front. Micro- 70, 497−504.
biol. 10, 339. (100) Syal, K., Flentie, K., Bhardwaj, N., Maiti, K., Jayaraman, N.,
(82) Munita, J. M., and Arias, C. A. (2016) Mechanisms of antibiotic Stallings, C. L., and Chatterji, D. (2017) Synthetic (p)ppGpp
resistance. Microbiol. Spectr. 4 (2), 481. analogue is an inhibitor of stringent response in mycobacteria.
(83) Foster, P. L. (2007) Stress-induced mutagenesis in bacteria. Antimicrob. Agents Chemother. 61 (6), No. e00443-17.
Crit. Rev. Biochem. Mol. Biol. 42 (5), 373−397. (101) Syal, K., Bhardwaj, N., and Chatterji, D. (2017) Vitamin C
(84) Wright, B. E. (1996) The effect of the stringent response on targets (p)ppGpp synthesis leading to stalling of long-term survival
mutation rates in Escherichia coliK-12. Mol. Microbiol. 19 (2), 213− and biofilm formation in Mycobacterium smegmatis. FEMS Microbiol.
219. Lett. 364 (1), No. fnw282.
(85) Wright, B. E., and Minnick, M. F. (1997) Reversion rates in a (102) Andresen, L., Varik, V., Tozawa, Y., Jimmy, S., Lindberg, S.,
leuB auxotroph of Escherichia coliK-12 correlate with ppGpp levels Tenson, T., and Hauryliuk, V. (2016) Auxotrophy-based high
during exponential growth. Microbiology 143 (3), 847−854. throughput screening assay for the identification of Bacillus subtilis
(86) Rudner, R., Murray, A., and Huda, N. (1999) Is there a link stringent response inhibitors. Sci. Rep. 6 (1), 35824.
between mutation rates and the stringent response in Bacillus subtilis? (103) Kriel, A., Brinsmade, S. R., Tse, J. L., Tehranchi, A. K., Bittner,
Ann. N. Y. Acad. Sci. 870, 418−422. A. N., Sonenshein, A. L., and Wang, J. D. (2014) GTP dysregulation
(87) Hachmann, A.-B., Sevim, E., Gaballa, A., Popham, D. L., in Bacillussubtilis cells lacking (p)ppGpp results in phenotypic amino
Antelmann, H., and Helmann, J. D. (2011) Reduction in membrane acid auxotrophy and failure to adapt to nutrient downshift and
phosphatidylglycerol content leads to daptomycin resistance in regulate biosynthesis genes. J. Bacteriol. 196 (1), 189−201.
Bacillus subtilis. Antimicrob. Agents Chemother. 55 (9), 4326−4337. (104) de la Fuente-Núñez, C., Reffuveille, F., Haney, E. F., Straus, S.
(88) Fang, F. C., Frawley, E. R., Tapscott, T., and Vázquez-Torres, K., and Hancock, R. E. W. (2014) Broad-spectrum anti-biofilm
A. (2016) Bacterial stress responses during host infection. Cell Host peptide that targets a cellular stress response. PLoS Pathog. 10 (5),
Microbe 20, 133−143. No. e1004152.
(89) Handwerger, S., and Tomasz, A. (1985) Antibiotic tolerance (105) de la Fuente-Núñez, C., Reffuveille, F., Mansour, S. C.,
among clinical isolates of bacteria. Annu. Rev. Pharmacol. Toxicol. 25 Reckseidler-Zenteno, S. L., Hernández, D., Brackman, G., Coenye, T.,
(1), 349−380. and Hancock, R. E. W. (2015) D-enantiomeric peptides that eradicate
(90) Claudi, B., Spröte, P., Chirkova, A., Personnic, N., Zankl, J., wild-type and multidrug-resistant biofilms and protect against lethal
Schürmann, N., Schmidt, A., and Bumann, D. (2014) Phenotypic Pseudomonas aeruginosa infections. Chem. Biol. 22 (2), 196−205.
variation of Salmonella in host tissues delays eradication by (106) Pletzer, D., Mansour, S. C., and Hancock, R. E. W. (2018)
antimicrobial chemotherapy. Cell 158 (4), 722−733. Synergy between conventional antibiotics and anti-biofilm peptides in
(91) Kopf, S. H., Sessions, A. L., Cowley, E. S., Reyes, C., Van a murine, sub-cutaneous abscess model caused by recalcitrant
Sambeek, L., Hu, Y., Orphan, V. J., Kato, R., and Newman, D. K. ESKAPE pathogens. PLoS Pathog. 14 (6), No. e1007084.
(2016) Trace incorporation of heavy water reveals slow and (107) Andresen, L., Tenson, T., and Hauryliuk, V. (2016) Cationic
heterogeneous pathogen growth rates in cystic fibrosis sputum. bactericidal peptide 1018 does not specifically target the stringent
response alarmone (p)ppGpp. Sci. Rep. 6 (1), 36549.
Proc. Natl. Acad. Sci. U. S. A. 113 (2), E110−E116.
(108) Gefen, O., Chekol, B., Strahilevitz, J., and Balaban, N. Q.
(92) Wexselblatt, E., Katzhendler, J., Saleem-Batcha, R., Hansen, G.,
(2017) TDtest: easy detection of bacterial tolerance and persistence
Hilgenfeld, R., Glaser, G., and Vidavski, R. R. (2010) ppGpp
in clinical isolates by a modified disk-diffusion assay. Sci. Rep. 7,
analogues inhibit synthetase activity of Rel proteins from Gram-
41284.
negative and Gram-positive bacteria. Bioorg. Med. Chem. 18 (12),
4485−4497.

1517 DOI: 10.1021/acsinfecdis.9b00204


ACS Infect. Dis. 2019, 5, 1505−1517

You might also like