Next Article in Journal
Editorial for the Special Issue “Parasitic Diseases from Wild Animals with Emphasis on Zoonotic Infections”
Next Article in Special Issue
Single-Cell Technologies to Study Phenotypic Heterogeneity and Bacterial Persisters
Previous Article in Journal
Metagenomic Investigation of a Low Diversity, High Salinity Offshore Oil Reservoir
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cellular Self-Digestion and Persistence in Bacteria

Department of Chemical and Biomolecular Engineering, University of Houston, Houston, TX 77004, USA
*
Author to whom correspondence should be addressed.
Microorganisms 2021, 9(11), 2269; https://doi.org/10.3390/microorganisms9112269
Submission received: 9 October 2021 / Revised: 25 October 2021 / Accepted: 26 October 2021 / Published: 31 October 2021

Abstract

:
Cellular self-digestion is an evolutionarily conserved process occurring in prokaryotic cells that enables survival under stressful conditions by recycling essential energy molecules. Self-digestion, which is triggered by extracellular stress conditions, such as nutrient depletion and overpopulation, induces degradation of intracellular components. This self-inflicted damage renders the bacterium less fit to produce building blocks and resume growth upon exposure to fresh nutrients. However, self-digestion may also provide temporary protection from antibiotics until the self-digestion-mediated damage is repaired. In fact, many persistence mechanisms identified to date may be directly or indirectly related to self-digestion, as these processes are also mediated by many degradative enzymes, including proteases and ribonucleases (RNases). In this review article, we will discuss the potential roles of self-digestion in bacterial persistence.

1. Introduction

Antibiotic failure is a growing concern worldwide [1], and persister cells—a small subpopulation of transiently non-growing drug-tolerant cells within a larger bacterial cell population—significantly contribute to this problem by facilitating the emergence of antibiotic-resistant mutants and the recurrence of microbial infections [2,3,4,5,6]. Because it is not mediated by heritable mutations, the persister state is reversible, and persister formation can occur in response to multiple environmental triggers, including antibiotic treatment [7,8], nutrient depletion [9,10,11], temperature [12], and pH [13,14,15]. A number of pathways have been implicated in persister formation, including the SOS response [7,16,17], the ppGpp-mediated stringent response [10,18], quorum sensing [19,20], and cellular aging [21]. In addition, reactive oxygen species (ROS) [22,23,24], toxin/antitoxin (TA) systems [25,26], and intracellular proteases [15,27] have been involved in this process. Notably, persistence seems to be a conserved phenomenon that has been reported in many cell types, including cancer cells [28,29,30,31]. Persister cells have been identified in almost every pathogenic or nonpathogenic microbial species studied to date, including Escherichia coli, Acinetobacter baumannii, Cyanobacteria, Salmonella Typhimurium, Vibrio cholerae, Xylella fastidiosa, Staphylococcus aureus, Mycobacterium tuberculosis, Candida albicans, and Saccharomyces cerevisiae [31]. Although dormancy is thought to be the prevailing trait that makes these persister phenotypes tolerant to external stresses [32,33,34,35,36], a significant number of studies have shown that persister cells are heterogeneous [37,38,39,40,41,42,43,44,45,46] and can escape cell death pathways through a diverse range of epigenetic mechanisms [10,18,19,47,48,49].
From an evolutionary perspective, self-digestion, known as autophagy in eukaryotes, is an important survival mechanism. This complex intracellular degradation is coordinated by many regulatory proteins and checkpoint kinases and has been well documented in mammalian cells [50,51], although rarely studied in bacteria. Autophagic mechanisms are associated with a diverse range of enzymes, including proteases, nucleases, glycosidases, lipases, and phosphatases, which are essential components of the intracellular degradation machineries [52,53]. Although self-digestion temporarily provides energy to cells in a non-nutritive environment or under stress conditions, this process may result in growth arrest or death due to the degradation of intracellular components. Critically, in some cases, these components are targets of conventional antibiotics. Self-digestion can therefore act as a double-edge sword; while excessive intracellular degradation can eventually result in cell death and the elimination of persisters [27,54], moderate degradation might only cause growth arrest and may render persisters transiently resistant to antibiotics [55]. Thus, mapping this complex network that may mediate persister formation will not only enhance our knowledge of persister cell physiology, but also provide novel antipersister therapeutic approaches.
A number of outstanding reviews on bacterial persister formation/reawakening mechanisms, physiology, evolutionary perspectives, and treatment strategies have been published in the literature to date [31,44,56,57,58,59,60,61,62,63,64,65,66]. Therefore, in this review article, we will particularly focus on the potential links between self-digestion and bacterial persistence. Specifically, we will first discuss the underlying reasons for self-digestion in bacteria and why it is an important survival mechanism, while highlighting the potential degradative mechanisms (including for those protein, ribonucleic acid (RNA), and lipids) that may induce persister cell formation. We will also explore how self-digestion may shape persister cell metabolism. Finally, we will briefly discuss autophagy to highlight the evolutionarily conserved aspects of the relationship between intracellular degradation and drug tolerance.

2. Why Does Cellular Self-Digestion Occur?

Self-digestion in bacteria is a dynamic process that degrades and removes unnecessary or dysfunctional cellular components within the cytoplasm, allowing cells to perform structured deterioration, while recycling key cellular constituents [67,68,69,70,71,72]. Although intracellular degradation continually occurs within bacteria to maintain cellular hemostasis, self-digestion occurs in response to specific stressors, such as starvation and nutrient deprivation [70,73,74]. In nature, most microorganisms are not afforded an abundance of resources needed for growth and reproduction processes [74], and many bacterial species encounter a scarcity of nutrients in their respective ecosystems [75,76]. This lack of nutrients may force cells to enter a quiescent physiological state [77], such as dormancy, to survive in nutrient-limited conditions. However, even in such states, microorganisms may not necessarily be fully dormant during the entire starvation period [74,78], which in nature, can vary from days to years, depending on the ecosystem [74]. For example, microorganisms in salt mines, deep-ocean habitats, ancient rocks, and caves can face starvation periods that may be as lengthy as thousands of years [74,75,76,78,79,80,81,82,83]. Some bacterial species (e.g., Acetonema, Bacillus, Clostridium, Heliobacterium) can survive this prolonged starvation by forming endospores. Sporulation, a tightly regulated, genetically programmed cellular process, is distinct compared to the normal bacterial growth where cells divide by binary fission to generate two identical daughter cells. In contrast, an asymmetric division is observed during sporulation, resulting in generation of two cells within the same cell wall: a small forespore and a large mother cell that engulfs the forespore [84,85,86]. The fully formed spore is released to the environment when the mother cell is completely degraded [84,85,86,87]. We note that persisters and endospores are two distinct phenotypes. While persister cells are often referred to a small subpopulation of non-growing cells in a cell population that can form stochastically or deterministically, endospores are referred to dormant, nonproductive phenotypes produced by certain bacterial species as a result of extreme stress conditions.
Unlike in natural environments, bacteria in the laboratory are provided with ample nutrients to support maximum growth [88]. In such nutrient-rich cultures, the doubling time of some bacteria (e.g., Clostridium perfringens) can be as fast as 10 min during exponential growth [89]. Upon exhaustion of nutrients, however, bacteria enter the so-called stationary phase, wherein growth cessation occurs, although the cells still exhibit certain metabolic activities [67,74,90]. Once bacteria transit from exponential phase to stationary phase, they develop various strategies (e.g., self-digestion) to survive in their nutrient-depleted environment [67,90,91,92,93].
Notably, before initiating any sort of survival response, upon entrance to stationary phase, bacteria undergo several morphological changes. In particular, cells in stationary phase become smaller in size, and rod-shaped bacteria become more spherical in shape. This results from changes to the cell membrane and cell wall. For example, in Escherichia coli entering stationary phase, the cell envelope becomes more rigid, and stress-bearing peptidoglycan layers increase from 0.7–0.8% to 1.4–1.9% of the cell’s dry weight [94]. Moreover, the cell wall becomes more highly cross-linked, and bacteria experience reduced membrane fluidity [67,95].
Reductive division, a process by which cells complete their final rounds of cell division in early-stationary phase, without increasing their biomass, also causes them to become smaller and to adopt a spherical shape [67,96]. This spherical morphology is mostly governed by the RpoS-dependent BolA protein [97,98], which regulates genes encoding for penicillin-binding protein (PBP)5, PBP6, and class C B-lactamases [98]. Overexpression of BolA drastically decreases outer membrane permeability and induces biofilm formation and persistence [99,100]. Aerobic respiration control protein A (ArcA) is another DNA-binding transcriptional regulator that is induced during stationary phase and involved in reductive division [101]. Deletion of arcA results in poor survivability in the absence of exogenous carbon sources [101]. Although, in E. coli and Staphylococcus aureus, levels of persisters in ΔarcA cells are not significantly different than in wild-type (WT) cells [47,102]. ΔarcA cells are unable to undergo reductive division and remain longer in stationary phase [101]. After undergoing reductive division, cells encounter a continuous reduction in cell size due to the degradation of endogenous cellular components [67]. Critically, these cells become highly tolerant to stress conditions, and a number of studies have reported that small [55,103] and aging [21,104] cells in stationary-phase cultures display increased antibiotic tolerance, resulting from their extensive morphological and physiological alterations.

3. What Are the Global Regulators of the Cellular Self-Digestion Network?

Self-digestion is initiated at the beginning of the starvation response, when bacteria begin to degrade their cytoplasmic membrane, cell wall, proteins, RNA, and DNA [67,105,106,107,108]. This process is mediated by a large number of degradative enzymes found in the bacterial cytoplasm, membrane, and periplasm [109]. Critically, despite their prevalence, the regulatory mechanisms that control expression of these molecules at the transcriptional level are largely unknown. Upon nutrient depletion or when cells enter stationary phase, significant changes in the intracellular levels of many global regulators, including DksA, Rpos, and ppGpp, are observed [90,110,111,112,113,114,115,116,117]. However, these regulators have many functions beyond just protecting the cell during starvation. Consequently, they are also induced in response to various stresses, such as oxidative stress, heat/cold shock, osmotic pressure, low pH, ultraviolet (UV)-induced DNA damage, high cell density, and toxic chemicals [115,118,119,120,121,122,123,124,125].
One such regulator is the stationary-phase transcription factor σS (i.e., the alternative sigma factor). Expression of this protein, which is encoded by the rpoS gene, dramatically increases in stationary phase, where it functions to regulate the expression of numerous stress-related genes [90,112,113,114,115,116,117,126]. Strains lacking σS show rapid cessation of growth upon introduction to starvation conditions in E. coli [127]. Notably, although the formation of bacterial persister cells in response to polyamines has been attributed to overexpression of rpoS [128], the effect of rpoS deletion on persistence depends on the experimental conditions and strains being used [129,130,131]. However, RpoS is not the only sigma factor in bacteria; other well-studied sigma factors include σB, σC, σD, and σH in Bacillus subtilis [132,133,134,135], σE, σH, and σS in Pseudomonas aeruginosa [136,137], and σB, σH, and σM in Corynebacterium glutamicum [138]. In E. coli, sigma factors σH and σN are overexpressed during stationary phase and can help cells survive during starvation [139,140]. Similarly, the absence of σE drastically compromises viability of Salmonella cells in stationary phase [141].
In addition to sigma factors, the transcription factor DksA and the alarmone molecule (p)ppGpp, which is synthesized by the RelA and SpoT enzymes, form a global regulator of the stringent response that is activated upon carbon source depletion or amino acid starvation [142,143,144]. Induction of the stringent response has been shown to induce persister formation in bacteria [10,45,145], and numerous research groups have reported reduced persister levels in ΔrelA, ΔspoT, and ΔdksA strains [11,18,146,147,148,149]. However, although molecules such as RpoS, DksA, and ppGpp have been extensively studied in the field of persister research [10,11,18,146,147,148,149], the question of whether they directly regulate persister mechanisms associated with self-digestion has yet to be answered. It is also possible that degradative enzymes may be constitutively expressed, which results in the accumulation of the enzymes in stationary-phase cells; however, this hypothesis needs to be verified.

4. Intracellular Degradation Mechanisms

During self-digestion, cells may initially restrain themselves from degrading essential components that are needed to help them generate energy for survival [74]. Accumulation of glycogen and poly-β-hydroxybutyric acid during exponential growth further ensures bacterial survival during carbon starvation [95,150,151,152]. Thus, bacteria that cannot gather adequate energy-rich molecules, such as glycogen, may rapidly degrade their major cellular components, including RNAs, proteins, and lipids, to generate energy molecules [150,151,152,153,154,155], and this will be further discussed in more detail below.

4.1. RNA Degradation

RNases comprise a group of hydrolytic enzymes that degrade RNA into smaller components [68]. There are two main types of ribonucleases: endoribonucleases (e.g., RNase P, III, BN, HI/II, I, E, G, and LS) and exoribonucleases (e.g., RNase D, T, PH, R and II, and PNPase) [68,156]. Endoribonucleases cleave single-stranded RNAs (ssRNA) or double-stranded RNAs (dsRNA) at internal phosphodiester bonds, whereas exoribonucleases cleave either the 3′ end or 5′ end of an RNA molecule [156,157]. In addition to their ability to degrade RNA, RNases play diverse roles in RNA metabolism, functioning in RNA maturation, quality control, and regulation [68]. RNases from different bacteria are generally conserved; however, some RNases can be species-specific, such as the B. subtilis RNase M5 (5S rRNA maturation), which is not present in E. coli [158].
Degradation of stable RNAs via the action of RNases occurs in response to depletion of nutrient sources during starvation [159]. As ribosomes account for the majority of RNAs in a cell, the RNAs degraded in this process are primarily ribosomal RNAs (rRNAs). These molecules are plentiful in cells and store substantial amounts of nutrients and energy that can be consumed during starvation [160]. For example, approximately 90% of 23S rRNA and 50% of 16S rRNA are degraded in Salmonella strains upon entry into the stationary phase in Luria–Bertani cultures [69]. Transfer RNAs (tRNAs) were found to be more stable in E. coli cells during phosphate starvation in a minimal medium [153]. Further, under starvation conditions, more than 70% of rRNA produced remains unused by ribosomes and is degraded in E. coli [161], suggesting the presence of a conserved molecular mechanism for rRNA degradation. Cells experiencing starvation from specific nutrients, such as carbon [162], nitrogen [163], phosphate [72], and magnesium (II) [164], may digest their rRNAs at different rates [153,160], although the exact extent of rRNA degradation under starvation conditions remains poorly understood. E. coli may digest their ribosomes in a unique manner, and once degradation begins, the 30S ribosomal subunit seems to perish quicker than the 50S subunit [153]. Kaplan and Apirion demonstrated that in starved cells, ribosomal degradation proceeds from polysomes to monosomes to ribosomal subunits [160]. The RNA pieces produced by this process are then further degraded to nucleotides by RNase II and PNPase [160].
Although ribosome dimerization and complex formation with their associated proteins has recently been shown to play a critical role in the resuscitation of rifampicin-induced antibiotic-tolerant cells [63,165,166], a direct correlation between the ability of mutant strains (exhibiting different RNase activities) to recover from starvation and their capacity to degrade RNA has been long established [160]. Specifically, strains that rapidly degrade RNA survive starvation better than more slowly degrading strains [160], suggesting a link between RNase activity and persister formation. RNases associated with type II TA systems, such as MqsR/MqsA [167], MazF/MazE [168], RelE/RelB [169], YoeB/YefM [34], and YafQ/DinJ [170], have been well-studied in the field of persister research. TA systems contain pairs of genes, one of which encodes a stable toxin and another that encodes an unstable antitoxin [171]. Antitoxins, under normal growth conditions, degrade, neutralize, or inhibit the associated toxin molecule [171]. Although the deletion of type II toxin molecules or TA systems, including chpB, mazF, relB/relE, yefM/yoeB, dinJ/yafQ, higB/higA, prlF/yhaV, yafN/O, mqsR/mqsA, and hicA/hicB, does not affect bacterial persistence [172], it is well established that toxins can induce cell cycle arrest by disrupting various cellular processes [171,173,174]. One of the first TA systems to be associated with persistence was the HipA/HipB system. HipA encodes a kinase that can inactivate synthesis of glutamyl tRNA synthetase [175,176], and one HipA mutant, HipA7, was found to show an approximately 100–1000-fold increase in persister levels [25]. In contrast, deletion of the HipA/HipB TA system results in an ~10–100-fold decrease in persister level [169]. Cho and colleagues further showed that rRNAs and tRNAs are primarily degraded in HipA-mediated persister cells, and ribosomes exist in their inactive forms in these cells [177]. MqsR encodes a ribonuclease that interferes with transcription by cleaving mRNA specifically at GCU sites [178], and the MqsR/MqsA TA system is another example of a case where overexpression or deletion of the TA system leads to either an increase or decrease, respectively, in persister formation [167]. Similarly, the RelE/RelB TA system, which has also been shown to aid in persistence, contains an RNase that cleaves mRNA in ribosomal site A, leading to inhibition of translation and growth arrest [179]. The toxin MazF, which is located downstream of the relE gene, also cleaves mRNAs at an ACA sequence at the 5′ end [180]. Although the biological role of MazF/MazE remains a subject of debate, studies have shown it plays a significant role in programmed cell death [181]. A recent study by Harrison et al. further showed that deletion of YafQ from the YafQ/DinJ TA system results in an approximately 2400-fold decrease in cell survival in antibiotic-exposed biofilms [170]. Collectively, these results support a key role for TA systems in bacterial persistence, although recent controversies [172,182,183] indicate that more studies are needed to fully elucidate the connection between TA system and the persistence state.

4.2. Protein Degradation

Proteases play a vital role in maintaining basal levels of regulatory proteins and removing misfolded and abnormal proteins from bacteria. These proteolytic enzymes can be divided into two groups, based on whether the cleavage position is inside the protein (endopeptidases or proteinases) or at the terminus (exopeptidases or peptidases) [184]. Depending on their cellular location, proteases can also be classified as cytoplasmic (e.g., Lon, ClpAP, ClpXP, HslUV), periplasmic (e.g., Tsp, HtrA, protease III), or membrane proteases (e.g., FtsH, OmpT) [71,185,186,187,188,189,190]. Although there are a few examples of energy-independent proteases, including protease III, VII, HtrA, Mi, and Tsp, the majority of intracellular proteolytic processes operate at the post-translational level and are powered by ATP hydrolysis [71,109]. Specifically, ATP hydrolysis is required to change the conformation of the protease, unfold the substrate, and pass the substrate through the protease active site [191].
Energy-dependent proteases are highly significant in E. coli and are responsible for more than 90% of the proteolytic activity taking place in the cytoplasm [71]. This model organism encodes five different AAA+ (ATPases associated with diverse cellular activities) proteases, Lon, ClpXP, ClpAP, HslUV, and ClpYQ, as well as the essential protease FtsH [192,193,194]. The ATPase and proteolytic domains of these proteins are located at the bacterial cytoplasm; the former is responsible for initiation of substrate degradation by the ATP-dependent unfoldase and translocation of the unfolded protein to the proteolytic domain. Here, it is further broken down into smaller peptides, five to 25 amino acids in length, with the help of peptidase [193,195].
In some cases, protein degradation can occur through a multistep process, with initial cleavage mediated by an ATP-dependent protease (rate-limiting step), followed by digestion via ATP-independent proteases and peptidases, ultimately leading to formation of free amino acids [71,196]. Peptidases display very high levels of activity, and only a small number of intermediate products of proteolysis are found in cells [71]. Proteases, particularly the ATP-dependent proteases, are extremely substrate-specific due to their structural features. They are also much larger in size (up to 750 kDa) than peptidases active in the extracellular medium, such as trypsin, which has a size of less than 50 kDa [188,196]. From a thermodynamic point of view, protein degradation is spontaneous, even for ATP-dependent proteases such as Lon and ClpAP, which can degrade a trace amount of small peptides without the need of ATP [197,198,199].
Lon, the first and most widely studied ATP-dependent protease, is a cytoplasmic serine protease, which is considered to be the primary protease for quality control in E. coli [200]. It is involved in the degradation of misfolded proteins, along with certain major regulatory proteins, such as the cell division regulator, SulA, and the capsule synthesis regulator, RcsA [201,202,203,204,205,206,207]. Lon can play an active role in persister formation, as it degrades several labile antitoxins of type II TA modules, releasing intra-bacterial toxins that cause growth inhibition. For example, the antitoxin RelB is degraded by Lon, decreasing intracellular toxin–antitoxin levels and leading to the accumulation of free RelE toxin, which induces global inhibition of translation [208]. Other antitoxins degraded by Lon include CcdA, HipB, and MazE [208,209,210].
Intriguingly, reduced levels of fluoroquinolone-tolerant persisters were observed in Lon-deficient cells [13,211], although the question of whether this phenotype is dependent on the activity of TA modules is highly debatable [13,211,212,213]. As part of the DNA-damage response, the cell division inhibitor, SulA, is upregulated when the cells are treated with fluoroquinolone antibiotics. Thus, in the absence of Lon, SulA accumulation may also affect persister cell survival [13,211,213,214].
Lon might also not be the only protease involved in TA module activation, as researchers have found that Clp proteases are also capable of degrading several antitoxins, including MazE and DinJ [215,216]. The Clp chaperone–protease family is another major group of ATP-dependent serine peptidases that are responsible for the degradation of a huge number of proteins. ClpAP and ClpXP both contain the proteolytic component, ClpP, which along with the co-factor ClpS, has been found to be required for environmental adaptation and extended viability in stationary phase. ClpS adapter protein specifically inhibits the degradation of ssrA-tagged substrates by ClpAP but directs ClpAP to degrade aggregated proteins and possibly N-end rule substrates [217,218]. In addition, the ClpAP protease is responsible for activation of the ParDE TA system by degrading the ParD antitoxin, resulting in transient growth arrest [219]. On the other hand, acyldepsipeptide antibiotic (ADEP4)-activated ClpP can become highly non-specific and kill growth-inhibited persister cells by degradation over 400 intracellular targets [27]. Altogether, despite the fact that both the Lon and Clp proteases have been extensively studied in the field of persister research, it remains unclear whether other ATP-dependent or energy-independent proteases may also function as critical persister molecules.

4.3. Lipid Degradation

During the transition to stationary phase, the fatty acid degradation regulon is overexpressed and provides a carbon source to starved cells via the digestion of membrane components [220]. FadR is a global regulator of lipid and fatty acid metabolism and acts as a switch between fatty acid β-oxidation and fatty acid biosynthesis [221]. FadR represses several genes and inhibits transcription of the fad genes [221,222], and its activity is modulated by the long-chain acyl-CoA thioester small effector molecule, which binds directly to FadR [221]. This protein complex (FadR-acyl-CoA thioester) cannot bind to the operator sequence in the promoter of the fatty acid degradative genes, leading to fad gene activation [221,223].
In response to carbon starvation, derepression of the FadR regulon results in digestion of membrane lipids, yielding fatty acids that are utilized by acyl-CoA synthetase to generate acyl-CoA [224]. The β-oxidation enzymes encoded by fadBA, fadE, fadFG, and fadH then convert acyl-CoA into acetyl-CoA, which is an important source of energy during starvation [224]. Consistent with these observations, it has been shown that a strain lacking acetyl-CoA dehydrogenase (encoded by fadF) barely survives carbon starvation [224]. Of note, in E. coli, lipid degradation occurs on the outer and inner membrane, as there are no intracellular forms of lipid storage in this bacterium [67].
Notably, cells can also activate emergency derepression pathways independently of FadR in stationary phase to survive carbon starvation [225]. Thus, in stationary phase, the fad genes become active via activity of ppGpp-programmed RNA polymerase together with cyclic adenosine monophosphate (cAMP)-cAMP repressor protein (Crp) complex [225]. However, it has been proposed that under these conditions, medium- or short-chain acyl-CoA is the substrate for β-oxidation, not long-chain acyl-CoA [67,225]. These medium and short chains of acyl-CoA also bind FadR and prevent binding to the operator sequence of the fad genes, leaving them active during the self-digestion process [225]. Critically, although lipid metabolism and fatty acid oxidation are known to be a critical survival mechanism for cancer persisters by providing energy molecules [28,226,227], this remains largely unexplored for prokaryotic persisters.

5. Links between Metabolism and Cellular Self-Digestion

Bacteria must produce a significant amount of energy molecules and building blocks to meet their high metabolic demands. This may lead to metabolic stress that promotes cellular self-digestion. For example, metabolic activity can result in free radical formation via the respiratory chain [228], which can damage proteins, RNAs, DNAs, and lipids, thereby initiating their cellular degradation [229,230].
Although the synthesis of glycolytic enzymes, as well as pyruvate formate lyase, phospho-transacetylase, and acetate kinase, and the subsequent downregulation of enzymes associated with anabolic pathways, has been observed during self-digestion in stationary phase [101,231], persister cells may have unique metabolic mechanisms (Figure 1). In one previous study, we explored the relationship between metabolic activity and persistence using fluorescence-activated cell sorting (FACS) and a redox sensor green dye that measures cytochrome and oxidoreductase rates in the electron transport chain (ETC). With this system, we detected a positive correlation between ETC activity and persistence in stationary-phase cultures [55]. To further determine whether metabolic activity in stationary phase is involved in persister formation, we treated cells with metabolic inhibitors or transferred them to anaerobic conditions in early-stationary phase and measured persister cell levels in late-stationary phase [55,232]. We found that these treatments significantly reduce persister cell levels, confirming an essential role for metabolism in persister cell formation.
Given that the metabolism of non-growing cells primarily derives from the digestion of endogenous cellular components, such as phospholipids, ribosomes, and proteins, during stationary phase [55], we further measured cell size, protein levels, and rRNA integrity in cell cultures with increased (i.e., untreated late-stationary phase) or decreased (i.e., metabolic inhibitor-treated or anaerobic-transferred late-stationary phase) levels of persister cells [55]. We found that untreated late-stationary-phase cells contain significantly more degraded rRNAs and proteins and are markedly smaller than metabolic inhibitor-treated or anaerobically transferred stationary-phase cells [55]. We also determined that deletion of metabolic genes encoding the citric acid (TCA) cycle and ETC enzymes reduces persister levels by preventing digestion of intracellular components, yielding cells that are more vulnerable to cell death when exposed to antibiotics in fresh medium [55,232] (Figure 1).
Although persister cell metabolism is significantly reduced compared to that of exponentially growing cells [13,37,102,103,233], persister cells must still undergo active energy metabolism in order to maintain their adenylate energy charge. Notably, there is evidence that self-digestion mediates the metabolism of persister cells, particularly those formed throughout the stationary phase (Figure 1). In fact, numerous independent studies have shown that persisters can harbor ETC activities [55,234], catabolize certain substrates to generate proton motive force (PMF) [234,235,236], produce energy molecules [237,238], and drive the futile production and degradation of RNA, leading to energy generation and dissipation [239]. Persister cells must also be able to repair antibiotic-induced damage to survive [16,17], and most repair mechanisms (e.g., DNA repair) are strongly ATP-dependent [240,241,242,243]. Further, a number of independent groups have shown that deletion of enzymes associated with the TCA cycle and ETC (e.g., sdhA, sucA, mdh) drastically reduce persister levels, indicating the importance of energy metabolism in persister cell formation and/or survival [21,55,244].
Metabolism involves a highly complex enzymatic network that is controlled by a number of transcriptional regulators, including ArcA, Cra, Crp, DksA, Fnr, Lrp, and Rpos, whose expression levels are drastically altered when cells enter stationary phase [140,246,247,248,249,250,251,252]. Critically, these regulators may be involved in self-digestion and in mediating the persister metabolism. Mutants deficient in arcA, for example, lose their culturability rapidly after a few days in stationary phase [101]. Cyclic AMP (a product of Cya enzyme) along with its receptor protein, Crp, may also play an important role in bacterial persistence. Persister cells in E. coli and S. aureus were previously shown to metabolize specific carbon sources and become susceptible to aminoglycoside (AG), which inhibits protein synthesis [234,235,253]. The aminoglycoside potentiation was further extended to Gram-negative pathogens, Salmonella enterica and Klebsiella pneumoniae, in a subsequent study [236] that also confirms earlier evidence of nascent protein synthesis in persisters [254]. Using the AG-potentiation assay, we found that a panel of carbon sources could not potentiate the AG-mediated killing of persisters derived from Δcrp and Δcya strains, indicating a role for these regulators in this process [47]. AG uptake is a unique, energy-requiring process, requiring both the electrochemical potential and the proton gradient across the cytoplasmic membrane [234]. Thus, the fact that persisters can efficiently metabolize certain substrates (e.g., glucose or glycerol) and generate PMF [234,235] supports the existence of active energy metabolism in these cells. However, we note that existence of an active mechanism does not necessarily imply “an upregulation” in that mechanism. The metabolism of persister cells is still likely to be lower than that of exponentially growing cells, which are metabolically highly active [13,37,102,103,233]. Regardless, this does not refute the proposed metabolic model highlighting the reliance of persister cell survival on energy metabolism. In fact, the addiction of some non-proliferating persister types (e.g., cancer persisters) to oxidative phosphorylation has repeatedly been shown [226,227,255,256,257,258].

6. Importance of Autophagy in Drug Tolerance

Given the similarities between prokaryotic and eukaryotic persisters [259], it is appealing to draw parallels between bacterial and mammalian persister cells and to expect that knowledge gained from one will enrich our understanding of the other [55]. Although the “persister” term has recently been used in cancer research [28,29,30], drug tolerance mechanisms in tumor cells that are not mediated by heritable mutations have long been known in the field. Indeed, a lot of mechanisms associated with bacterial persisters have been already identified in drug-tolerant cancer cells, such as enhanced efflux activities (e.g., higher rate of drug export from the cell), enhanced repair activities (e.g., efficient repair of drug-induced cell damage), active bypass pathways (e.g., alternative pathways to avoid the drug target), or altered cell metabolism (e.g., a distinct set of active metabolic reactions) [260,261,262].
There are also similarities between cellular self-digestion and autophagy. However, due to the lack of organelles and complex structures in bacteria, a compartmentalized autophagy is not observed in these cells. Autophagy in eukaryotes promotes cell survival and generates energy and intermediate molecules for vital anabolic processes by degrading cellular components in lysosomes [263]. Within these lysosomes, proteins, DNAs, RNAs, polysaccharides, and lipids are hydrolyzed by a diverse range of degradative enzymes [264]. Lysozymes contain approximately 50 different hydrolytic enzymes [265], including the Lon protease [266], which was first identified in bacteria and is known to be a crucial persistence molecule, as discussed above.
Depending on the mechanism that mediates transport of cytosolic cargo to lysosomes, eukaryotic autophagy can be categorized into three major types: macroautophagy, microautophagy, and chaperone-mediated autophagy (CMA) [267]. Although, macroautophagy and microautophagy are similar in terms of initiation, termination, and capacity for the sequestration of large structures, they are distinct pathways [268]. The formation of autophagosomes is a hallmark of macroautophagy and consists of several distinct steps (i.e., nucleation, elongation, and closure of the double-membraned vesicle) [269]; these are mediated by a cluster of proteins known as autophagy-related proteins (ATGs). Once the autophagosome is formed, it fuses with a lysosome, and a single-membrane vesicle (i.e., the autophagic or Cvt body) is released into the lumen. In contrast, during microautophagy, cargos are directly engulfed and taken up by the lysosomal membrane as result of local deformation and rearrangement of the membrane, allowing the cytosolic cargos to be degraded by vacuolar hydrolases and enzymes within the lysosome [264]. The resulting macromolecules such as amino acids or fatty acids from the lysosomal degradation are recycled back into the cytosol via membrane permeases to be used in anabolic processes. In CMA, the degradation of soluble cytosolic proteins in lysosomes is highly selective. Substrate selection in CMA is regulated by cytosolic chaperones that recognize pentapeptide motifs in the amino acid sequence of the substrate proteins [270]. In this type of autophagic process, substrates are not engulfed, but, instead, are translocated across the lysosomal membrane in a receptor-mediated manner [271].
In eukaryotes, the autophagy regulatory network is highly complex and tightly connected to redundant signaling pathways, some of which are related to the cell cycle and proliferation, including the mammalian target of rapamycin kinase (TOR), nuclear factor of kappa light polypeptide gene enhancer (NFKB/NF-κB), mitogen-activated protein kinase (MAPK), and tumor protein p53 (TP53) cascades [272]. Notably, autophagy may play a critical role in cancer persistence, as there is evidence for reciprocal interactions between autophagy and cell cycle arrest, the hallmark of cancer persistence [273]. Cancer persisters can escape cell death pathways (e.g., apoptosis) by inactivating their cell-proliferation signaling pathways during treatment. In fact, many targeted therapeutics may induce cell dormancy by directly inhibiting cell-proliferation signaling pathways, whereas chemotherapeutics may indirectly stimulate growth arrest by activating stress signaling pathways [274,275,276,277]. Critically, the mechanisms associated with cell growth arrest may be mediated by the same signaling pathways that are involved in autophagy [273].
The degree of stress (e.g., starvation) that a cell is under can determine if autophagy will be used as a means for survival [263] or programmed cell death [278]. However, autophagy is not only a defense mechanism for starvation, but a necessary function for molecular recycling and maintaining the homeostasis of non-starved cells [263]. Knockout of ATGs has been shown to cause severe developmental problems in mice, including abnormalities at the cellular level [279], obesity [280], lung dysfunction [281], tumorigenesis [282,283], and death [283]. Critically, dependence of tumor cell survival and growth on basal autophagy has also been demonstrated via ATG knockout [284], and numerous observations support a key role for autophagy in cancer growth and survival. In one instance, it was shown that deletion of ATG7 leads to metabolic and proliferative problems in cancer, causing cancer cells to become more sensitive to starvation and more dependent on glutamine [285]. RAS-mutated cancers were also found to exhibit upregulated autophagy, leading to sustained TCA cycle metabolism [286], increased levels of glycolysis [287], and enhanced tumorigenesis [286,288].
In addition to providing support for cancer growth and survival under stress, autophagy can help cancer cells resist treatment. Increasing levels of autophagic flux have been correlated with higher cancer cell survival rates and shortened patient survival times in melanoma [289]. The specific mechanisms that mediate this phenomenon have not been fully elucidated; however, autophagy appears to play various protective roles in cancer, depending on the cancer type and the treatment method. For example, in BRAF-mutated cancers, BRAF inhibition leads to endoplasmic reticulum (ER) stress, which subsequently increases autophagic activity, protecting the cells from apoptosis and maintaining mitochondrial activity [290]. Autophagy also makes mTOR-mutated cancer cells tolerant to mTOR inhibitors by eliminating receptor-interacting protein kinases (RIPKs), which promote necroptosis when autophagy is inhibited [291]. Overall, both autophagy and self-digestion seem to be evolutionarily conserved mechanisms that allow organisms to survive undesirable environmental conditions. Both processes can be activated in response to extracellular stress conditions such as nutrient deprivation; maintain cellular energy homeostasis; may use similar degradative enzymes (such as Lon); and can lead to growth arrest, and therefore tolerance to both antibiotics (for bacteria) and chemotherapeutics (for cancer cells).

7. Concluding Remarks

Persistence and self-digestion (or autophagy), which appear to be evolutionarily conserved phenomena, are observed in many prokaryotic and eukaryotic cell types. These processes allow organisms to survive in undesirable environmental conditions, leading to formation of persister cells. Critically, mapping a self-digestion-mediated persistence mechanism from its initial exogenous stress trigger, through its signal transduction, to the source of antibiotic tolerance, may provide us an opportunity to categorize previously identified mechanisms within one complex network. Such a strategy may also uncover novel antipersister therapeutic approaches, as inhibition of intracellular degradation is known to reduce persister formation [55,234,245]. Conversely, stimulating self-digestion might also be useful as an alternative antipersister strategy, due to the fact that enhanced intracellular degradation can also be detrimental to persister cells [27,54]. However, the study of cellular self-digestion mechanisms is challenging, and a number of critical questions remain to be addressed, some of which are as follows:
(i)
If a proposed mechanism is essential for persister formation and survival, genetically perturbing the mechanism should eliminate persisters or reduce their levels; however, this method may not be ideal for redundant systems. Self-digestion-mediated persistence is potentially a collective effect of many different degradative enzymes, which makes it difficult to test using conventional methods. One way to investigate these mechanisms is to perform single-cell analysis. With the use of antibiotic treatments and fluorescent reporters for degradative enzymes, a correlation between persistence and the enzymes expression levels can be performed in cell populations where self-digestion is significantly upregulated (e.g., late-stationary-phase cultures).
(ii)
Although Kim et al. claim that persisters and “viable but non-culturable” (VBNC) cells represent the same phenotypes [54], the VBNC state is thought to be a transitory phase on the spectrum between persistence and cell death [292,293]. While persister cells can exit from persistence state (stochastically or deterministically) and colonize, the resuscitation of VBNC cells is rarely observed [35,235,293,294]. In fact, bacteria associated with asymptomatic infections may be in a non-replicating or slowly replicating state and cannot be easily cultured in vitro [295,296,297], and this “viable but non-culturable” state observed in pathogenic bacteria has long been known [296]. Further, a number of independent groups have shown antibiotic-treated cultures contain many more VBNC cells than persisters [35,37,235,298]. If persisters and VBNC cells represent two distinct phenotypes on the live- and dead-cell spectrum, then, a threshold level of intracellular degradation may play a critical role in the phenotypic switch between persistence and the VBNC state, which remains to be validated.
(iii)
Persister metabolism is a controversial topic, reflecting the complexity and diversity of persister cell formation, survival, and resuscitation mechanisms, as well as the influence of culture conditions [31,61,299]. Although persisters are mostly non-growing cells [33,300,301,302], and their metabolism is generally lower than that of exponentially growing cells [13,37,102,103,233], these phenotypes might be at a metabolic steady state, providing energy molecules necessary for their survival [237,239]. Although it is well established that autophagy plays a crucial role in the metabolism of drug-tolerant cancer cells, it remains to be determined whether this is also true for bacterial persisters.
(iv)
The levels of global regulators, such as Rpos, ppGpp, and cAMP/Crp, are significantly altered in cells during their transition to stationary phase [110,116,117,126,144,147]. However, we still do not know if these molecules regulate expression of degradative enzymes, as the promoters of many genes encoding degradative enzymes are not well characterized [303]. Constitutive expression of degradative enzymes may result in their accumulation in stationary phase, which would make intracellular degradation more apparent in stationary-phase cells, where cell growth and protein synthesis are minimal. However, this has yet to be validated.
(v)
Recently, several groups have uncovered a correlation between protein aggregation and bacterial persistence, although protein aggregation seems to be associated with the VBNC phenotype [12,304,305,306,307]. While these results may contradict with self-digestion-mediated persister mechanisms at first glance, it is well known that protein aggregation can induce autophagy in mammalian cells [308,309,310]. Whether a similar phenomenon is also present in bacterial cells is yet to be determined.

Author Contributions

All authors have been involved in writing, reviewing, and editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by R01-AI143643 Award from the National Institute of Allergy and Infectious Diseases of the National Institute of Health, and University of Houston start-up grant.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ventola, C.L. The Antibiotic Resistance Crisis: Part 1: Causes and Threats. Pharm. Ther. 2015, 40, 277. [Google Scholar]
  2. Levin-Reisman, I.; Ronin, I.; Gefen, O.; Braniss, I.; Shoresh, N.; Balaban, N.Q. Antibiotic tolerance facilitates the evolution of resistance. Science 2017, 355, 826–830. [Google Scholar] [CrossRef]
  3. Barrett, T.C.; Mok, W.W.K.; Murawski, A.M.; Brynildsen, M.P. Enhanced antibiotic resistance development from fluoroquinolone persisters after a single exposure to antibiotic. Nat. Commun. 2019, 10, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bakkeren, E.; Diard, M.; Hardt, W.-D. Evolutionary causes and consequences of bacterial antibiotic persistence. Nat. Rev. Microbiol. 2020, 18, 479–490. [Google Scholar] [CrossRef]
  5. Windels, E.M.; Michiels, J.E.; Fauvart, M.; Wenseleers, T.; Van den Bergh, B.; Michiels, J. Bacterial persistence promotes the evolution of antibiotic resistance by increasing survival and mutation rates. ISME J. 2019, 13, 1239–1251. [Google Scholar] [CrossRef] [PubMed]
  6. Levin-Reisman, I.; Brauner, A.; Ronin, I.; Balaban, N.Q. Epistasis between antibiotic tolerance, persistence, and resistance mutations. Proc. Natl. Acad. Sci. USA 2019, 116, 14734–14739. [Google Scholar] [CrossRef] [Green Version]
  7. Dörr, T.; Lewis, K.; Vulić, M. SOS Response Induces Persistence to Fluoroquinolones in Escherichia coli. PLOS Genet. 2009, 5, e1000760. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Dörr, T.; Vulić, M.; Lewis, K. Ciprofloxacin Causes Persister Formation by Inducing the TisB toxin in Escherichia coli. PLOS Biol. 2010, 8, e1000317. [Google Scholar] [CrossRef] [Green Version]
  9. Nguyen, D.; Joshi-Datar, A.; Lepine, F.; Bauerle, E.; Olakanmi, O.; Beer, K.; McKay, G.; Siehnel, R.; Schafhauser, J.; Wang, Y.; et al. Active starvation responses mediate antibiotic tolerance in biofilms and nutrient-limited bacteria. Science 2011, 334, 982–986. [Google Scholar] [CrossRef] [Green Version]
  10. Amato, S.M.; Orman, M.A.; Brynildsen, M.P. Metabolic Control of Persister Formation in Escherichia coli. Mol. Cell 2013, 50, 475–487. [Google Scholar] [CrossRef] [Green Version]
  11. Fung, D.K.C.; Chan, E.W.C.; Chin, M.L.; Chan, R.C.Y. Delineation of a bacterial starvation stress response network which can mediate antibiotic tolerance development. Antimicrob. Agents Chemother. 2010, 54, 1082–1093. [Google Scholar] [CrossRef] [Green Version]
  12. Pu, Y.; Li, Y.; Jin, X.; Tian, T.; Ma, Q.; Zhao, Z.; Lin, S.Y.; Chen, Z.; Li, B.; Yao, G.; et al. ATP-Dependent Dynamic Protein Aggregation Regulates Bacterial Dormancy Depth Critical for Antibiotic Tolerance. Mol. Cell 2019, 73, 143–156.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Shan, Y.; Gandt, A.B.; Rowe, S.E.; Deisinger, J.P.; Conlon, B.P.; Lewis, K. ATP-Dependent persister formation in Escherichia coli. MBio 2017, 8. [Google Scholar] [CrossRef] [Green Version]
  14. Karki, P.; Mohiuddin, S.G.; Kavousi, P.; Orman, M.A. Investigating the effects of osmolytes and environmental pH on bacterial persisters. Antimicrob. Agents Chemother. 2020, 64, e02393-19. [Google Scholar] [CrossRef]
  15. Helaine, S.; Cheverton, A.M.; Watson, K.G.; Faure, L.M.; Matthews, S.A.; Holden, D.W. Internalization of salmonella by macrophages induces formation of nonreplicating persisters. Science 2014, 343, 204–208. [Google Scholar] [CrossRef]
  16. Goormaghtigh, F.; Van Melderen, L. Single-cell imaging and characterization of Escherichia coli persister cells to ofloxacin in exponential cultures. Sci. Adv. 2019, 5, eaav9462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Völzing, K.G.; Brynildsen, M.P. Stationary-phase persisters to ofloxacin sustain DNA damage and require repair systems only during recovery. MBio 2015, 6, e00731-15. [Google Scholar] [CrossRef] [Green Version]
  18. Hansen, S.; Lewis, K.; Vulić, M. Role of global regulators and nucleotide metabolism in antibiotic tolerance in Escherichia coli. Antimicrob. Agents Chemother. 2008, 52, 2718–2726. [Google Scholar] [CrossRef] [Green Version]
  19. Vega, N.M.; Allison, K.R.; Khalil, A.S.; Collins, J.J. Signaling-mediated bacterial persister formation. Nat. Chem. Biol. 2012, 8, 431–433. [Google Scholar] [CrossRef] [PubMed]
  20. Ng, W.-L.; Bassler, B.L. Bacterial Quorum-Sensing Network Architectures. Annu. Rev. Genet. 2009, 43, 197–222. [Google Scholar] [CrossRef] [Green Version]
  21. Luidalepp, H.; Jõers, A.; Kaldalu, N.; Tenson, T. Age of inoculum strongly influences persister frequency and can mask effects of mutations implicated in altered persistence. J. Bacteriol. 2011, 193, 3598–3605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Grant, S.S.; Kaufmann, B.B.; Chand, N.S.; Haseley, N.; Hung, D.T. Eradication of bacterial persisters with antibiotic-generated hydroxyl radicals. Proc. Natl. Acad. Sci. USA 2012, 109, 12147–12152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Cirillo, S.L.G.; Subbian, S.; Chen, B.; Weisbrod, T.R.; Jacobs, W.R.; Cirillo, J.D. Protection of Mycobacterium tuberculosis from reactive oxygen species conferred by the mel2 locus impacts persistence and dissemination. Infect. Immun. 2009, 77, 2557–2567. [Google Scholar] [CrossRef] [Green Version]
  24. Wu, Y.; Vulić, M.; Keren, I.; Lewis, K. Role of oxidative stress in persister tolerance. Antimicrob. Agents Chemother. 2012, 56, 4922–4926. [Google Scholar] [CrossRef] [Green Version]
  25. Moyed, H.S.; Bertrand, K.P. hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. J. Bacteriol. 1983, 155, 768–775. [Google Scholar] [CrossRef] [Green Version]
  26. Schumacher, M.A.; Piro, K.M.; Xu, W.; Hansen, S.; Lewis, K.; Brennan, R.G. Molecular mechanisms of HipA-mediated multidrug tolerance and its neutralization by HipB. Science 2009, 323, 396–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Conlon, B.P.; Nakayasu, E.S.; Fleck, L.E.; LaFleur, M.D.; Isabella, V.M.; Coleman, K.; Leonard, S.N.; Smith, R.D.; Adkins, J.N.; Lewis, K. Activated ClpP kills persisters and eradicates a chronic biofilm infection. Nature 2013, 503, 365–370. [Google Scholar] [CrossRef] [Green Version]
  28. Hangauer, M.J.; Viswanathan, V.S.; Ryan, M.J.; Bole, D.; Eaton, J.K.; Matov, A.; Galeas, J.; Dhruv, H.D.; Berens, M.E.; Schreiber, S.L.; et al. Drug-tolerant persister cancer cells are vulnerable to GPX4 inhibition. Nature 2017, 551, 247–250. [Google Scholar] [CrossRef] [Green Version]
  29. Sharma, S.V.; Lee, D.Y.; Li, B.; Quinlan, M.P.; Takahashi, F.; Maheswaran, S.; McDermott, U.; Azizian, N.; Zou, L.; Fischbach, M.A.; et al. A Chromatin-Mediated Reversible Drug-Tolerant State in Cancer Cell Subpopulations. Cell 2010, 141, 69–80. [Google Scholar] [CrossRef] [Green Version]
  30. Ramirez, M.; Rajaram, S.; Steininger, R.J.; Osipchuk, D.; Roth, M.A.; Morinishi, L.S.; Evans, L.; Ji, W.; Hsu, C.-H.; Thurley, K.; et al. Diverse drug-resistance mechanisms can emerge from drug-tolerant cancer persister cells. Nat. Commun. 2016, 7, 1–8. [Google Scholar] [CrossRef] [PubMed]
  31. Van den Bergh, B.; Fauvart, M.; Michiels, J. Formation, physiology, ecology, evolution and clinical importance of bacterial persisters. FEMS Microbiol. Rev. 2017, 41, 219–251. [Google Scholar] [CrossRef]
  32. Wood, T.K.; Knabel, S.J.; Kwan, B.W. Bacterial persister cell formation and dormancy. Appl. Environ. Microbiol. 2013, 79, 7116–7121. [Google Scholar] [CrossRef] [Green Version]
  33. Balaban, N.Q.; Merrin, J.; Chait, R.; Kowalik, L.; Leibler, S. Bacterial persistence as a phenotypic switch. Science 2004, 305, 1622–1625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Shah, D.; Zhang, Z.; Khodursky, A.B.; Kaldalu, N.; Kurg, K.; Lewis, K. Persisters: A distinct physiological state of E. coli. BMC Microbiol. 2006, 6, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Roostalu, J.; Jõers, A.; Luidalepp, H.; Kaldalu, N.; Tenson, T. Cell division in Escherichia colicultures monitored at single cell resolution. BMC Microbiol. 2008, 8, 1–14. [Google Scholar] [CrossRef] [Green Version]
  36. Kwan, B.W.; Valenta, J.A.; Benedik, M.J.; Wood, T.K. Arrested protein synthesis increases persister-like cell formation. Antimicrob. Agents Chemother. 2013, 57, 1468–1473. [Google Scholar] [CrossRef] [Green Version]
  37. Orman, M.A.; Brynildsen, M.P. Dormancy is not necessary or sufficient for bacterial persistence. Antimicrob. Agents Chemother. 2013, 57, 3230–3239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Kussell, E.; Kishony, R.; Balaban, N.Q.; Leibler, S. Bacterial PersistenceA Model of Survival in Changing Environments. Genetics 2005, 169, 1807–1814. [Google Scholar] [CrossRef] [Green Version]
  39. Gefen, O.; Balaban, N.Q. The importance of being persistent: Heterogeneity of bacterial populations under antibiotic stress. FEMS Microbiol. Rev. 2009, 33, 704–717. [Google Scholar] [CrossRef]
  40. Kussell, E.; Leibler, S. Ecology: Phenotypic diversity, population growth, and information in fluctuating environments. Science 2005, 309, 2075–2078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Mettetal, J.T.; Muzzey, D.; Pedraza, J.M.; Ozbudak, E.M.; van Oudenaarden, A. Predicting stochastic gene expression dynamics in single cells. Proc. Natl. Acad. Sci. USA 2006, 103, 7304–7309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Henry, T.C.; Brynildsen, M.P. Development of Persister-FACSeq: A method to massively parallelize quantification of persister physiology and its heterogeneity. Sci. Rep. 2016, 6, 1–17. [Google Scholar] [CrossRef] [PubMed]
  43. Rotem, E.; Loinger, A.; Ronin, I.; Levin-Reisman, I.; Gabay, C.; Shoresh, N.; Biham, O.; Balaban, N.Q. Regulation of phenotypic variability by a threshold-based mechanism underlies bacterial persistence. Proc. Natl. Acad. Sci. USA 2010, 107, 12541–12546. [Google Scholar] [CrossRef] [Green Version]
  44. Allison, K.R.; Brynildsen, M.P.; Collins, J.J. Heterogeneous bacterial persisters and engineering approaches to eliminate them. Curr. Opin. Microbiol. 2011, 14, 593–598. [Google Scholar] [CrossRef] [Green Version]
  45. Amato, S.M.; Brynildsen, M.P. Persister Heterogeneity Arising from a Single Metabolic Stress. Curr. Biol. 2015, 25, 2090–2098. [Google Scholar] [CrossRef] [Green Version]
  46. Barth, V.C., Jr.; Rodrigues, B.Á.; Bonatto, G.D.; Gallo, S.W.; Pagnussatti, V.E.; Ferreira, C.A.S.; de Oliveira, S.D. Heterogeneous Persister Cells Formation in Acinetobacter baumannii. PLoS ONE 2013, 8, e84361. [Google Scholar] [CrossRef]
  47. Mok, W.W.K.; Orman, M.A.; Brynildsen, M.P. Impacts of global transcriptional regulators on persister metabolism. Antimicrob. Agents Chemother. 2015, 59, 2713–2719. [Google Scholar] [CrossRef] [Green Version]
  48. Leung, V.; Lévesque, C.M. A stress-inducible quorum-sensing peptide mediates the formation of persister cells with noninherited multidrug tolerance. J. Bacteriol. 2012, 194, 2265–2274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Wakamoto, Y.; Dhar, N.; Chait, R.; Schneider, K.; Signorino-Gelo, F.; Leibler, S.; McKinney, J.D. Dynamic persistence of antibiotic-stressed mycobacteria. Science 2013, 339, 91–95. [Google Scholar] [CrossRef]
  50. Klionsky, D.J. The molecular machinery of autophagy: Unanswered questions. J. Cell Sci. 2005, 118, 7–18. [Google Scholar] [CrossRef] [Green Version]
  51. Noda, T.; Ohsumi, Y. Tor, a Phosphatidylinositol Kinase Homologue, Controls Autophagy in Yeast. J. Biol. Chem. 1998, 273, 3963–3966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kaminskyy, V.; Zhivotovsky, B. Proteases in autophagy. Biochim. Biophys. Acta Proteins Proteom. 2012, 1824, 44–50. [Google Scholar] [CrossRef] [PubMed]
  53. Rabinowitz, J.D.; White, E. Autophagy and metabolism. Science 2010, 330, 1344–1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Kim, J.-S.; Chowdhury, N.; Yamasaki, R.; Wood, T.K. Viable but non-culturable and persistence describe the same bacterial stress state. Environ. Microbiol. 2018, 20, 2038–2048. [Google Scholar] [CrossRef]
  55. Orman, M.A.; Brynildsen, M.P. Inhibition of stationary phase respiration impairs persister formation in E. coli. Nat. Commun. 2015, 6, 1–13. [Google Scholar] [CrossRef] [Green Version]
  56. Lewis, K. Persister cells. Annu. Rev. Microbiol. 2010, 64, 357–372. [Google Scholar] [CrossRef]
  57. Lewis, K. Persister cells, dormancy and infectious disease. Nat. Rev. Microbiol. 2006, 5, 48–56. [Google Scholar] [CrossRef]
  58. Fisher, R.A.; Gollan, B.; Helaine, S. Persistent bacterial infections and persister cells. Nat. Rev. Microbiol. 2017, 15, 453–464. [Google Scholar] [CrossRef] [PubMed]
  59. Wilmaerts, D.; Windels, E.M.; Verstraeten, N.; Michiels, J. General Mechanisms Leading to Persister Formation and Awakening. Trends Genet. 2019, 35, 401–411. [Google Scholar] [CrossRef]
  60. Balaban, N.Q.; Helaine, S.; Lewis, K.; Ackermann, M.; Aldridge, B.; Andersson, D.I.; Brynildsen, M.P.; Bumann, D.; Camilli, A.; Collins, J.J.; et al. Definitions and guidelines for research on antibiotic persistence. Nat. Rev. Microbiol. 2019, 17, 441–448. [Google Scholar] [CrossRef] [Green Version]
  61. Amato, S.M.; Fazen, C.H.; Henry, T.C.; Mok, W.W.K.; Orman, M.A.; Sandvik, E.L.; Volzing, K.G.; Brynildsen, M.P. The role of metabolism in bacterial persistence. Front. Microbiol. 2014, 5, 70. [Google Scholar] [CrossRef] [Green Version]
  62. Harms, A.; Maisonneuve, E.; Gerdes, K. Mechanisms of bacterial persistence during stress and antibiotic exposure. Science 2016, 354, aaf4268. [Google Scholar] [CrossRef] [PubMed]
  63. Wood, T.K.; Song, S.; Yamasaki, R. Ribosome dependence of persister cell formation and resuscitation. J. Microbiol. 2019, 57, 213–219. [Google Scholar] [CrossRef] [PubMed]
  64. Defraine, V.; Fauvart, M.; Michiels, J. Fighting bacterial persistence: Current and emerging anti-persister strategies and therapeutics. Drug Resist. Updat. 2018, 38, 12–26. [Google Scholar] [CrossRef] [PubMed]
  65. Kester, J.C.; Fortune, S.M. Persisters and beyond: Mechanisms of phenotypic drug resistance and drug tolerance in bacteria. Crit. Rev. Biochem. Mol. Biol. 2014, 49, 91–101. [Google Scholar] [CrossRef]
  66. Kaldalu, N.; Hauryliuk, V.; Tenson, T. Persisters—as elusive as ever. Appl. Microbiol. Biotechnol. 2016, 100, 6545–6553. [Google Scholar] [CrossRef] [Green Version]
  67. Nyström, T. Stationary-phase physiology. Annu. Rev. Microbiol. 2004, 58, 161–181. [Google Scholar] [CrossRef]
  68. Bechhofer, D.H.; Deutscher, M.P. Bacterial ribonucleases and their roles in RNA metabolism. Crit. Rev. Biochem. Mol. Biol. 2019, 54, 242. [Google Scholar] [CrossRef]
  69. Hsu, D.; Shih, L.M.; Zee, Y.C. Degradation of rRNA in Salmonella strains: A novel mechanism to regulate the concentrations of rRNA and ribosomes. J. Bacteriol. 1994, 176, 4761–4765. [Google Scholar] [CrossRef] [Green Version]
  70. Maruyama, H.; Ono, M.; Mizuno, D. Ribosome degradation and the degradation products in starved Escherichia coli: III. Ribosomal RNA degradation during the complete deprivation of nutrients. Biochim. Biophys. Acta Nucleic Acids Protein Synth. 1970, 199, 176–183. [Google Scholar] [CrossRef]
  71. Maurizi, M.R. Proteases and protein degradation in Escherichia coli. Experientia 1992, 48, 178–201. [Google Scholar] [CrossRef]
  72. Maruyama, H.; Mizuno, D. Ribosome degradation and the degradation products in starved Escherichia coli: I. Comparison of the degradation rate and of the nucleotide pool between Escherichia coli B and Q-13 strains in phosphate deficiency. Biochim. Biophys. Acta Nucleic Acids Protein Synth. 1970, 199, 159–165. [Google Scholar] [CrossRef]
  73. Watson, S.P.; Clements, M.O.; Foster, S.J. Characterization of the Starvation-Survival Response of Staphylococcus aureus. J. Bacteriol. 1998, 180, 1750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Morita, R.Y. The starvation-survival state of microorganisms in nature and its relationship to the bioavailable energy. Experientia 1990, 46, 813–817. [Google Scholar] [CrossRef]
  75. Burgess, G. Bacteria in Oligotrophic Environments: Starvation Survival Lifestyle. World J. Microbiol. Biotechnol. 1997, 14, 305. [Google Scholar] [CrossRef]
  76. Sebastián, M.; Estrany, M.; Ruiz-González, C.; Forn, I.; Sala, M.M.; Gasol, J.M.; Marrasé, C. High Growth Potential of Long-Term Starved Deep Ocean Opportunistic Heterotrophic Bacteria. Front. Microbiol. 2019, 10, 760. [Google Scholar] [CrossRef] [PubMed]
  77. Kolter, R.; Siegele, D.A.; Tormo, A. The stationary phase of the bacterial life cycle. Annu. Rev. Microbiol. 2003, 47, 855–874. [Google Scholar] [CrossRef]
  78. Hoppe, H.-G. Determination and properties of actively metabolizing heterotrophic bacteria in the sea, investigated by means of micro-autoradiography. Mar. Biol. 1976, 36, 291–302. [Google Scholar] [CrossRef]
  79. Reiser, R.; Tasch, P. Investigation of the viability of osmophile bacteria of great geological age. Trans. Kans. Acad. Sci. 1960, 63, 31–34. [Google Scholar] [CrossRef]
  80. Tabor, P.S.; Ohwada, K.; Colwell, R.R. Filterable marine bacteria found in the deep sea: Distribution, taxonomy, and response to starvation. Microb. Ecol. 1981, 7, 67–83. [Google Scholar] [CrossRef]
  81. MA, H.; MT, M. Distribution of ultramicrobacteria in a gulf coast estuary and induction of ultramicrobacteria. Microb. Ecol. 1987, 14, 113–127. [Google Scholar] [CrossRef]
  82. Lipman, C.B. Living Microörganisms in Ancient Rocks. J. Bacteriol. 1931, 22, 183. [Google Scholar] [CrossRef] [Green Version]
  83. Vreeland, R.H.; Rosenzweig, W.D.; Powers, D.W. Isolation of a 250 million-year-old halotolerant bacterium from a primary salt crystal. Nature 2000, 407, 897–900. [Google Scholar] [CrossRef] [PubMed]
  84. de Hoon, M.J.; Eichenberger, P.; Vitkup, D. Hierarchical evolution of the bacterial sporulation network. Curr. Biol. 2010, 20, R735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Stragier, P.; Losick, R. Molecular Genetics of Sporulation in Bacillus Subtilis. Annu. Rev. Genet. 2003, 30, 297–341. [Google Scholar] [CrossRef]
  86. Kay, D.; Warren, S.C. Sporulation in Bacillus subtilis. Morphological changes. Biochem. J. 1968, 109, 819–824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Piggot, P.J.; Coote, J.G. Genetic aspects of bacterial endospore formation. Bacteriol. Rev. 1976, 40, 908. [Google Scholar] [CrossRef]
  88. Maier, R.M.; Pepper, I.L. Bacterial Growth. Environ. Microbiol. Third Ed. 2015, 37–56. [Google Scholar] [CrossRef]
  89. Li, J.; Paredes-Sabja, D.; Sarker, M.R.; McClane, B.A. Clostridium perfringens Sporulation and Sporulation-Associated Toxin Production. Microbiol. Spectr. 2016, 4. [Google Scholar] [CrossRef] [Green Version]
  90. Navarro Llorens, J.M.; Tormo, A.; Martínez-García, E. Stationary phase in gram-negative bacteria. FEMS Microbiol. Rev. 2010, 34, 476–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Nystrom, T.; Flardh, K.; Kjelleberg, S. Responses to multiple-nutrient starvation in marine Vibrio sp. strain CCUG 15956. J. Bacteriol. 1990, 172, 7085–7097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. NH, A.; T, N.; S, K. Starvation-induced modulations in binding protein-dependent glucose transport by the marine Vibrio sp. S14. FEMS Microbiol. Lett. 1990, 58, 205–210. [Google Scholar] [CrossRef]
  93. Britos, L.; Abeliuk, E.; Taverner, T.; Lipton, M.; McAdams, H.; Shapiro, L. Regulatory Response to Carbon Starvation in Caulobacter crescentus. PLoS ONE 2011, 6, e18179. [Google Scholar] [CrossRef]
  94. Mengin-Lecreulx, D.; Van Heijenoort, J. Effect of growth conditions on peptidoglycan content and cytoplasmic steps of its biosynthesis in Escherichia coli. J. Bacteriol. 1985, 163, 208–212. [Google Scholar] [CrossRef] [Green Version]
  95. Jaishankar, J.; Srivastava, P. Molecular Basis of Stationary Phase Survival and Applications. Front. Microbiol. 2017, 8, 2000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Pletnev, P.; Osterman, I.; Sergiev, P.; Bogdanov, A.; Dontsova, O. Survival guide: Escherichia coli in the stationary phase. Acta Naturae 2015, 7, 22. [Google Scholar] [CrossRef] [PubMed]
  97. Lange, R.; Hengge-Aronis, R. Identification of a central regulator of stationary-phase gene expression in Escherichia coli. Mol. Microbiol. 1991, 5, 49–59. [Google Scholar] [CrossRef] [PubMed]
  98. Santos, J.M.; Lobo, M.; Matos, A.P.A.; de Pedro, M.A.; Arraiano, C.M. The gene bolA regulates dacA (PBP5), dacC (PBP6) and ampC (AmpC), promoting normal morphology in Escherichia coli. Mol. Microbiol. 2002, 45, 1729–1740. [Google Scholar] [CrossRef]
  99. Freire, P.; Vieira, H.L.A.; Furtado, A.R.; de Pedro, M.A.; Arraiano, C.M. Effect of the morphogene bolA on the permeability of the Escherichia coli outer membrane. FEMS Microbiol. Lett. 2006, 260, 106–111. [Google Scholar] [CrossRef]
  100. Guinote, I.B. Functional Studies on BolA and Related Genes: Increasing the Understanding of a Protein with Pleiotropic Effects ProQuest. Available online: https://0-www-proquest-com.brum.beds.ac.uk/docview/1924948863?pq-origsite=gscholar&fromopenview=true (accessed on 27 September 2021).
  101. Nyström, T.; Larsson, C.; Gustafsson, L. Bacterial defense against aging: Role of the Escherichia coli ArcA regulator in gene expression, readjusted energy flux and survival during stasis. EMBO J. 1996, 15, 3219–3228. [Google Scholar] [CrossRef]
  102. Conlon, B.P.; Rowe, S.E.; Gandt, A.B.; Nuxoll, A.S.; Donegan, N.P.; Zalis, E.A.; Clair, G.; Adkins, J.N.; Cheung, A.L.; Lewis, K. Persister formation in Staphylococcus aureus is associated with ATP depletion. Nat. Microbiol. 2016, 1, 1–7. [Google Scholar] [CrossRef] [Green Version]
  103. Manuse, S.; Shan, Y.; Canas-Duarte, S.J.; Bakshi, S.; Sun, W.-S.; Mori, H.; Paulsson, J.; Lewis, K. Bacterial persisters are a stochastically formed subpopulation of low-energy cells. PLoS Biol. 2021, 19, e3001194. [Google Scholar] [CrossRef]
  104. Knudsen, G.M.; Ng, Y.; Gram, L. Survival of bactericidal antibiotic treatment by a persister subpopulation of Listeria monocytogenes. Appl. Environ. Microbiol. 2013, 79, 7390–7397. [Google Scholar] [CrossRef] [Green Version]
  105. Malmcrona-Friberg, K.; Tunlid, A.; Mårdén, P.; Kjelleberg, S.; Odham, G. Chemical changes in cell envelope and poly-β-hydroxybutyrate during short term starvation of a marine bacterial isolate. Arch. Microbiol. 1986, 144, 340–345. [Google Scholar] [CrossRef]
  106. Hood, M.A.; Guckert, J.B.; White, D.C.; Deck, F. Effect of nutrient deprivation on lipid, carbohydrate, DNA, RNA, and protein levels in Vibrio cholerae. Appl. Environ. Microbiol. 1986, 52, 788–793. [Google Scholar] [CrossRef] [Green Version]
  107. Guckert, J.B.; Antworth, C.P.; Nichols, P.D.; White, D.C. Phospholipid, ester-linked fatty acid profiles as reproducible assays for changes in prokaryotic community structure of estuarine sediments. FEMS Microbiol. Ecol. 1985, 1, 147–158. [Google Scholar] [CrossRef]
  108. Mårdén, P.; Tunlid, A.; Malmcrona-Friberg, K.; Odham, G.; Kjelleberg, S. Physiological and morphological changes during short term starvation of marine bacterial islates. Arch. Microbiol. 1985, 142, 326–332. [Google Scholar] [CrossRef]
  109. Gottesman, S. Proteases and their targets in Escherichia coli. Annu. Rev. Genet. 1996, 30, 465–506. [Google Scholar] [CrossRef] [PubMed]
  110. Cavanagh, A.T.; Chandrangsu, P.; Wassarman, K.M. 6S RNA regulation of relA alters ppGpp levels in early stationary phase. Microbiology 2010, 156, 3791. [Google Scholar] [CrossRef] [Green Version]
  111. Gentry, D.R.; Hernandez, V.J.; Nguyen, L.H.; Jensen, D.B.; Cashel, M. Synthesis of the stationary-phase sigma factor σ(s) is positively regulated by ppGpp. J. Bacteriol. 1993, 175, 7982–7989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Lacour, S.; Landini, P. σS-dependent gene expression at the onset of stationary phase in Escherichia coli: Function of σS-dependent genes and identification of their promoter sequences. J. Bacteriol. 2004, 186, 7186–7195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Weber, H.; Polen, T.; Heuveling, J.; Wendisch, V.F.; Hengge, R. Genome-wide analysis of the general stress response network in Escherichia coli: σS-dependent genes, promoters, and sigma factor selectivity. J. Bacteriol. 2005, 187, 1591–1603. [Google Scholar] [CrossRef] [Green Version]
  114. Hengge-Aronis, R. Signal Transduction and Regulatory Mechanisms Involved in Control of the σ S (RpoS) Subunit of RNA Polymerase. Microbiol. Mol. Biol. Rev. 2002, 66, 373–395. [Google Scholar] [CrossRef] [Green Version]
  115. Lange, R.; Hengge-Aronis, R. The cellular concentration of the sigma S subunit of RNA polymerase in Escherichia coli is controlled at the levels of transcription, translation, and protein stability. Genes Dev. 1994, 8, 1600–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Hirsch, M.; Elliott, T. Stationary-phase regulation of RpoS translation in Escherichia coli. J. Bacteriol. 2005, 187, 7204–7213. [Google Scholar] [CrossRef] [Green Version]
  117. Dong, T.; Schellhorn, H.E. Global effect of RpoS on gene expression in pathogenic Escherichia coli O157:H7 strain EDL933. BMC Genomics 2009, 10, 1–17. [Google Scholar] [CrossRef] [Green Version]
  118. Martínez-García, E.; Tormo, A.; Navarro-Llorens, J. Further studies on RpoS in enterobacteria: Identification of rpoS in Enterobacter cloacae and Kluyvera cryocrescens. Arch. Microbiol. 2001, 175, 395–404. [Google Scholar] [CrossRef]
  119. Hengge-Aronis, R. Back to log phase: σS as a global regulator in the osmotic control of gene expression in Escherichia coli. Mol. Microbiol. 1996, 21, 887–893. [Google Scholar] [CrossRef] [PubMed]
  120. Mandel, M.J.; Silhavy, T.J. Starvation for different nutrients in Escherichia coli results in differential modulation of RpoS levels and stability. J. Bacteriol. 2005, 187, 434–442. [Google Scholar] [CrossRef] [Green Version]
  121. Muffler, A.; Traulsen, D.D.; Lange, R.; Hengge-Aronis, R. Posttranscriptional osmotic regulation of the σs subunit of RNA polymerase in Escherichia coli. J. Bacteriol. 1996, 178, 1607–1613. [Google Scholar] [CrossRef] [Green Version]
  122. Bearson, S.M.D.; Benjamin, W.H.; Swords, W.E.; Foster, J.W. Acid shock induction of RpoS is mediated by the mouse virulence gene mviA of Salmonella typhimurium. J. Bacteriol. 1996, 178, 2572–2579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Heuveling, J.; Possling, A.; Hengge, R. A role for Lon protease in the control of the acid resistance genes of Escherichia coli. Mol. Microbiol. 2008, 69, 534–547. [Google Scholar] [CrossRef] [PubMed]
  124. Muffler, A.; Barth, M.; Marschall, C.; Hengge-Aronis, R. 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. J. Bacteriol. 1997, 179, 445–452. [Google Scholar] [CrossRef] [Green Version]
  125. Merrikh, H.; Ferrazzoli, A.E.; Bougdour, A.; Olivier-Mason, A.; Lovett, S.T. A DNA damage response in Escherichia coli involving the alternative sigma factor, RpoS. Proc. Natl. Acad. Sci. USA 2009, 106, 611–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Hengge-Aronis, R. Survival of hunger and stress: The role of rpoS in early stationary phase gene regulation in E. coli. Cell 1993, 72, 165–168. [Google Scholar] [CrossRef]
  127. Notley-McRobb, L.; King, T.; Ferenci, T. rpoS mutations and loss of general stress resistance in Escherichia coli populations as a consequence of conflict between competing stress responses. J. Bacteriol. 2002, 184, 806–811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Tkachenko, A.G.; Kashevarova, N.M.; Tyuleneva, E.A.; Shumkov, M.S. Stationary-phase genes upregulated by polyamines are responsible for the formation of Escherichia coli persister cells tolerant to netilmicin. FEMS Microbiol. Lett. 2017, 364, 84. [Google Scholar] [CrossRef] [PubMed]
  129. Liu, S.; Wu, N.; Zhang, S.; Yuan, Y.; Zhang, W.; Zhang, Y. Variable Persister Gene Interactions with (p)ppGpp for Persister Formation in Escherichia coli. Front. Microbiol. 2017, 8, 1795. [Google Scholar] [CrossRef]
  130. Wu, N.; He, L.; Cui, P.; Wang, W.; Yuan, Y.; Liu, S.; Xu, T.; Zhang, S.; Wu, J.; Zhang, W.; et al. Ranking of persister genes in the same Escherichia coli genetic background demonstrates varying importance of individual persister genes in tolerance to different antibiotics. Front. Microbiol. 2015, 6, 1003. [Google Scholar] [CrossRef] [Green Version]
  131. Hong, S.H.; Wang, X.; O’Connor, H.F.; Benedik, M.J.; Wood, T.K. Bacterial persistence increases as environmental fitness decreases. Microb. Biotechnol. 2012, 5, 509–522. [Google Scholar] [CrossRef]
  132. Boylan, S.A.; Thomas, M.D.; Price, C.W. Genetic method to identify regulons controlled by nonessential elements: Isolation of a gene dependent on alternate transcription factor sigma B of Bacillus subtilis. J. Bacteriol. 1991, 173, 7856. [Google Scholar] [CrossRef] [Green Version]
  133. Johnson, W.C.; Moran, C.P.; Losick, R. Two RNA polymerase sigma factors from Bacillus subtilis discriminate between overlapping promoters for a developmentally regulated gene. Nature 1983, 302, 800–804. [Google Scholar] [CrossRef]
  134. Helmann, J.D. Alternative sigma factors and the regulation of flagellar gene expression. Mol. Microbiol. 1991, 5, 2875–2882. [Google Scholar] [CrossRef]
  135. Predich, M.; Nair, G.; Smith, I. Bacillus subtilis early sporulation genes-kinA, spo0F, and spo0A are transcribed by the RNA polymerase containing σ(H). J. Bacteriol. 1992, 174, 2771–2778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Potvin, E.; Sanschagrin, F.; Levesque, R.C. Sigma factors in Pseudomonas aeruginosa. FEMS Microbiol. Rev. 2008, 32, 38–55. [Google Scholar] [CrossRef] [Green Version]
  137. Raiger-Iustman, L.J.; Ruiz, J.A. The alternative sigma factor, σS, affects polyhydroxyalkanoate metabolism in Pseudomonas putida. FEMS Microbiol. Lett. 2008, 284, 218–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Pátek, M.; Nešvera, J. Promoters and Plasmid Vectors of Corynebacterium glutamicum. Corynebacterium Glutamicum 2013, 23, 51–88. [Google Scholar] [CrossRef]
  139. Jishage, M.; Ishihama, A. Regulation of RNA polymerase sigma subunit synthesis in Escherichia coli: Intracellular levels of sigma 70 and sigma 38. J. Bacteriol. 1995, 177, 6832–6835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Jishage, M.; Iwata, A.; Ueda, S.; Ishihama, A. Regulation of RNA polymerase sigma subunit synthesis in Escherichia coli: Intracellular levels of four species of sigma subunit under various growth conditions. J. Bacteriol. 1996, 178, 5447–5451. [Google Scholar] [CrossRef] [Green Version]
  141. Testerman, T.L.; Vazquez-Torres, A.; Xu, Y.; Jones-Carson, J.; Libby, S.J.; Fang, F.C. The alternative sigma factor σE controls antioxidant defences required for Salmonella virulence and stationary-phase survival. Mol. Microbiol. 2002, 43, 771–782. [Google Scholar] [CrossRef]
  142. Srivatsan, A.; Wang, J.D. Control of bacterial transcription, translation and replication by (p)ppGpp. Curr. Opin. Microbiol. 2008, 11, 100–105. [Google Scholar] [CrossRef]
  143. Potrykus, K.; Cashel, M. (p)ppGpp: Still Magical? Annu. Rev. Microbiol. 2008, 62, 35–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Hauryliuk, V.; Atkinson, G.C.; Murakami, K.S.; Tenson, T.; Gerdes, K. Recent functional insights into the role of (p)ppGpp in bacterial physiology. Nat. Rev. Microbiol. 2015, 13, 298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Amato, S.M.; Brynildsen, M.P. Nutrient Transitions Are a Source of Persisters in Escherichia coli Biofilms. PLoS ONE 2014, 9, e93110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Viducic, D.; Ono, T.; Murakami, K.; Susilowati, H.; Kayama, S.; Hirota, K.; Miyake, Y. Functional Analysis of spoT, relA and dksA Genes on Quinolone Tolerance in Pseudomonas aeruginosa under Nongrowing Condition. Microbiol. Immunol. 2006, 50, 349–357. [Google Scholar] [CrossRef]
  147. Abranches, J.; Martinez, A.R.; Kajfasz, J.K.; Chavez, V.; Garsin, D.A.; Lemos, J.A. The Molecular Alarmone (p)ppGpp Mediates Stress Responses, Vancomycin Tolerance, and Virulence in Enterococcus faecalis. J. Bacteriol. 2009, 191, 2248–2256. [Google Scholar] [CrossRef] [Green Version]
  148. Chowdhury, N.; Kwan, B.W.; Wood, T.K. Persistence Increases in the Absence of the Alarmone Guanosine Tetraphosphate by Reducing Cell Growth. Sci. Rep. 2016, 6, 1–9. [Google Scholar] [CrossRef]
  149. Gaca, A.O.; Kajfasz, J.K.; Miller, J.H.; Liu, K.; Wang, J.D.; Abranches, J.; Lemos, J.A. Basal levels of (p)ppGpp in Enterococcus faecalis: The magic beyond the stringent response. MBio 2013, 4, e00646-13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Dawes, E.; Ribbons, D. Studies on the endogenous metabolism of Escherichia coli. Biochem. J. 1965, 95, 332–343. [Google Scholar] [CrossRef] [Green Version]
  151. G, S.; NE, G. Role and oxidation pathway of poly-beta-hydroxybutyric acid in Micrococcus halodenitrificans. Can. J. Microbiol. 1962, 8, 255–269. [Google Scholar] [CrossRef]
  152. van Houte, J.; Jansen, H.M. Role of Glycogen in Survival of Streptococcus mitis. J. Bacteriol. 1970, 101, 1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Davis, B.D.; Luger, S.M.; Tai, P.C. Role of ribosome degradation in the death of starved Escherichia coli cells. J. Bacteriol. 1986, 166, 439–445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Matin, A. Physiology, molecular biology and applications of the bacterial starvation response Physiology, zyxwvu molecular biology and applications of the bacterial starvation response. J. Appl. Bacteriol. Symp. Ser. 1992, 73, 49–57. [Google Scholar] [CrossRef]
  155. Matin, A.; Veldhuis, C.; Stegeman, V.; Veenhuis, M. Selective advantage of a Spirillum sp. in a carbon-limited environment. Accumulation of poly-beta-hydroxybutyric acid and its role in starvation. J. Gen. Microbiol. 1979, 112, 349–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Yang, W. Nucleases: Diversity of Structure, Function and Mechanism. Q. Rev. Biophys. 2011, 44, 1. [Google Scholar] [CrossRef] [PubMed]
  157. Nicholson, A.W. Ribonuclease III mechanisms of double-stranded RNA cleavage. Wiley Interdiscip. Rev. RNA 2014, 5, 31. [Google Scholar] [CrossRef] [PubMed]
  158. Sogin, M.L.; Pace, N.R. In vitro maturation of precursors of 5S ribosomal RNA from Bacillus subtilis. Nature 1974, 252, 598–600. [Google Scholar] [CrossRef] [PubMed]
  159. Deutscher, M.P. Degradation of RNA in bacteria: Comparison of mRNA and stable RNA. Nucleic Acids Res. 2006, 34, 659–666. [Google Scholar] [CrossRef] [PubMed]
  160. Kaplan, R.; Apirion, D. The fate of ribosomes in Escherichia coli cells starved for a carbon source. J. Biol. Chem. 1975, 250, 1854–1863. [Google Scholar] [CrossRef]
  161. Gausing, K. Regulation of ribosome production in Escherichia coli: Synthesis and stability of ribosomal RNA and of ribosomal protein messenger RNA at different growth rates. J. Mol. Biol. 1977, 115, 335–354. [Google Scholar] [CrossRef]
  162. Jacobson, A.; Gillespie, D. Metabolic events occurring during recovery from prolonged glucose starvation in Escherichia coli. J. Bacteriol. 1968, 95, 1030–1039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Ben-Hamida, F.; Schlessinger, D. Synthesis and breakdown of ribonucleic acid in Escherichia coli starving for nitrogen. Biochim. Biophys. Acta -Nucleic Acids Protein Synth. 1966, 119, 183–191. [Google Scholar] [CrossRef]
  164. McCarthy, B.J. The effects of magnesium starvation on the ribosome content of Escherichia coli. Biochim. Biophys. Acta -Spec. Sect. Nucleic Acids Relat. Subj. 1962, 55, 880–889. [Google Scholar] [CrossRef]
  165. Kim, J.-S.; Yamasaki, R.; Song, S.; Zhang, W.; Wood, T.K. Single cell observations show persister cells wake based on ribosome content. Environ. Microbiol. 2018, 20, 2085–2098. [Google Scholar] [CrossRef] [PubMed]
  166. Song, S.; Wood, T.K. ppGpp ribosome dimerization model for bacterial persister formation and resuscitation. Biochem. Biophys. Res. Commun. 2020, 523, 281–286. [Google Scholar] [CrossRef]
  167. Kim, Y.; Wood, T.K. Toxins Hha and CspD and small RNA regulator Hfq are involved in persister cell formation through MqsR in Escherichia coli. Biochem. Biophys. Res. Commun. 2010, 391, 209–213. [Google Scholar] [CrossRef] [Green Version]
  168. Tripathi, A.; Dewan, P.C.; Siddique, S.A.; Varadarajan, R. MazF-induced Growth Inhibition and Persister Generation in Escherichia coli. J. Biol. Chem. 2014, 289, 4191–4205. [Google Scholar] [CrossRef] [Green Version]
  169. Keren, I.; Shah, D.; Spoering, A.; Kaldalu, N.; Lewis, K. Specialized persister cells and the mechanism of multidrug tolerance in Escherichia coli. J. Bacteriol. 2004, 186, 8172–8180. [Google Scholar] [CrossRef] [Green Version]
  170. Harrison, J.J.; Wade, W.D.; Akierman, S.; Vacchi-Suzzi, C.; Stremick, C.A.; Turner, R.J.; Ceri, H. The chromosomal toxin gene yafQ is a determinant of multidrug tolerance for Escherichia coli growing in a biofilm. Antimicrob. Agents Chemother. 2009, 53, 2253–2258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Page, R.; Peti, W. Toxin-antitoxin systems in bacterial growth arrest and persistence. Nat. Chem. Biol. 2016, 12, 208–214. [Google Scholar] [CrossRef]
  172. Goormaghtigh, F.; Fraikin, N.; Putrinš, M.; Hallaert, T.; Hauryliuk, V.; Garcia-Pino, A.; Sjödin, A.; Kasvandik, S.; Udekwu, K.; Tenson, T.; et al. Reassessing the role of type II toxin-antitoxin systems in formation of Escherichia coli type II persister cells. MBio 2018, 9. [Google Scholar] [CrossRef] [Green Version]
  173. Wang, X.; Wood, T.K. Toxin-antitoxin systems influence biofilm and persister cell formation and the general stress response. Appl. Environ. Microbiol. 2011, 77, 5577–5583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Helaine, S.; Kugelberg, E. Bacterial persisters: Formation, eradication, and experimental systems. Trends Microbiol. 2014, 22, 417–424. [Google Scholar] [CrossRef] [PubMed]
  175. Germain, E.; Castro-Roa, D.; Zenkin, N.; Gerdes, K. Molecular Mechanism of Bacterial Persistence by HipA. Mol. Cell 2013, 52, 248–254. [Google Scholar] [CrossRef] [PubMed]
  176. Kaspy, I.; Rotem, E.; Weiss, N.; Ronin, I.; Balaban, N.Q.; Glaser, G. HipA-mediated antibiotic persistence via phosphorylation of the glutamyl-tRNA-synthetase. Nat. Commun. 2013, 4, 1–7. [Google Scholar] [CrossRef] [Green Version]
  177. Cho, J.; Rogers, J.; Kearns, M.; Leslie, M.; Hartson, S.D.; Wilson, K.S. Escherichia coli persister cells suppress translation by selectively disassembling and degrading their ribosomes. Mol. Microbiol. 2015, 95, 352–364. [Google Scholar] [CrossRef]
  178. Yamaguchi, Y.; Park, J.H.; Inouye, M. MqsR, a Crucial Regulator for Quorum Sensing and Biofilm Formation, Is a GCU-specific mRNA Interferase in Escherichia coli. J. Biol. Chem. 2009, 284, 28746–28753. [Google Scholar] [CrossRef] [Green Version]
  179. Pedersen, K.; Zavialov, A.V.; Pavlov, M.Y.; Elf, J.; Gerdes, K.; Ehrenberg, M. The Bacterial Toxin RelE Displays Codon-Specific Cleavage of mRNAs in the Ribosomal A Site. Cell 2003, 112, 131–140. [Google Scholar] [CrossRef] [Green Version]
  180. Zhang, Y.; Zhang, J.; Hoeflich, K.P.; Ikura, M.; Qing, G.; Inouye, M. MazF Cleaves Cellular mRNAs Specifically at ACA to Block Protein Synthesis in Escherichia coli. Mol. Cell 2003, 12, 913–923. [Google Scholar] [CrossRef]
  181. Engelberg-Kulka, H.; Hazan, R.; Amitai, S. mazEF: A chromosomal toxin-antitoxin module that triggers programmed cell death in bacteria. J. Cell Sci. 2005, 118, 4327–4332. [Google Scholar] [CrossRef] [Green Version]
  182. Holden, D.W.; Errington, J. Type II toxin-antitoxin systems and persister cells. MBio 2018, 9. [Google Scholar] [CrossRef] [Green Version]
  183. Kim, J.-S.; Wood, T.K. Persistent Persister Misperceptions. Front. Microbiol. 2016, 7, 2134. [Google Scholar] [CrossRef] [PubMed]
  184. McDonald, J.K. An overview of protease specificity and catalytic mechanisms: Aspects related to nomenclature and classification. Histochem. J. 1985, 17, 773–785. [Google Scholar] [CrossRef]
  185. Weichart, D.; Querfurth, N.; Dreger, M.; Hengge-Aronis, R. Global role for ClpP-containing proteases in stationary-phase adaptation of Escherichia coli. J. Bacteriol. 2003, 185, 115–125. [Google Scholar] [CrossRef] [Green Version]
  186. Tomoyasu, T.; Gamer, J.; Bukau, B.; Kanemori, M.; Mori, H.; Rutman, A.J.; Oppenheim, A.B.; Yura, T.; Yamanaka, K.; Niki, H. Escherichia coli FtsH is a membrane-bound, ATP-dependent protease which degrades the heat-shock transcription factor sigma 32. EMBO J. 1995, 14, 2551–2560. [Google Scholar] [CrossRef]
  187. Spiers, A.; Lamb, H.K.; Cocklin, S.; Wheeler, K.A.; Budworth, J.; Dodds, A.L.; Pallen, M.J.; Maskell, D.J.; Charles, I.G.; Hawkins, A.R. PDZ Domains Facilitate Binding of High Temperature Requirement Protease A (HtrA) and Tail-specific Protease (Tsp) to Heterologous Substrates through Recognition of the Small Stable RNA A (ssrA)-encoded Peptide. J. Biol. Chem. 2002, 277, 39443–39449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Rozkov, A.; Enfors, S.-O. Analysis and Control of Proteolysis of Recombinant Proteins in Escherichia coli. Adv. Biochem. Eng. Biotechnol. 2004, 89, 163–195. [Google Scholar] [CrossRef]
  189. Hwang, B.Y.; Varadarajan, N.; Li, H.; Rodriguez, S.; Iverson, B.L.; Georgiou, G. Substrate specificity of the Escherichia coli outer membrane protease OmpP. J. Bacteriol. 2007, 189, 522–530. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Keiler, K.C.; Silber, K.R.; Sauer, R.T.; Downard, K.M.; Papayannopoulos, I.A.; Biemann, K. C-terminal specific protein degradation: Activity and substrate specificity of the Tsp protease. Protein Sci. 1995, 4, 1507–1515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Zwickl, P.; Baumeister, W.; Steven, A. Dis-assembly lines: The proteasome and related ATPase-assisted proteases. Curr. Opin. Struct. Biol. 2000, 10, 242–250. [Google Scholar] [CrossRef]
  192. Bittner, L.-M.; Arends, J.; Narberhaus, F. When, how and why? Regulated proteolysis by the essential FtsH protease in Escherichia coli. Biol. Chem. 2017, 398, 625–635. [Google Scholar] [CrossRef] [PubMed]
  193. Sauer, R.T.; Bolon, D.N.; Burton, B.M.; Burton, R.E.; Flynn, J.M.; Grant, R.A.; Hersch, G.L.; Joshi, S.A.; Kenniston, J.A.; Levchenko, I.; et al. Sculpting the Proteome with AAA+ Proteases and Disassembly Machines. Cell 2004, 119, 9–18. [Google Scholar] [CrossRef] [Green Version]
  194. Gur, E.; Sauer, R.T. Recognition of misfolded proteins by Lon, a AAA+ protease. Genes Dev. 2008, 22, 2267. [Google Scholar] [CrossRef] [Green Version]
  195. RT, S.; TA, B. AAA+ proteases: ATP-fueled machines of protein destruction. Annu. Rev. Biochem. 2011, 80, 587–612. [Google Scholar] [CrossRef]
  196. GOLDBERG, A.L. The mechanism and functions of ATP-dependent proteases in bacterial and animal cells. Eur. J. Biochem. 1992, 203, 9–23. [Google Scholar] [CrossRef] [PubMed]
  197. Gottesman, S.; Maurizi, M.R. Regulation by proteolysis: Energy-dependent proteases and their targets. Microbiol. Rev. 1992, 56, 592. [Google Scholar] [CrossRef] [PubMed]
  198. Baker, T.A.; Sauer, R.T. ClpXP, an ATP-powered unfolding and protein-degradation machine. Biochim. Biophys. Acta -Mol. Cell Res. 2012, 1823, 15–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Rozkov, A. Control of Proteolysis of Recombinant Proteins in Escherichia coli. Ph.D. Thesis, Kungl Tekniska Högskolan, Stockholm, Sweden, 2001. [Google Scholar]
  200. Mizusawa, S.; Gottesman, S. Protein degradation in Escherichia coli: The lon gene controls the stability of sulA protein. Proc. Natl. Acad. Sci. USA 1983, 80, 358–362. [Google Scholar] [CrossRef] [Green Version]
  201. Laskowska, E.; Kuczyńska-Wiśnik, D.; Skórko-Glonek, J.; Taylor, A. Degradation by proteases Lon, Clp and HtrA, of Escherichia coli proteins aggregated in vivo by heat shock; HtrA protease action in vivo and in vitro. Mol. Microbiol. 1996, 22, 555–571. [Google Scholar] [CrossRef]
  202. Chung, C.H.; Goldberg, A.L. The product of the lon (capR) gene in Escherichia coli is the ATP-dependent protease, protease La. Proc. Natl. Acad. Sci. USA 1981, 78, 4931–4935. [Google Scholar] [CrossRef] [Green Version]
  203. Dopazo, A.; Tormo, A.; Aldea, M.; Vicente, M. Structural inhibition and reactivation of Escherichia coli septation by elements of the SOS and TER pathways. J. Bacteriol. 1987, 169, 1772–1776. [Google Scholar] [CrossRef] [Green Version]
  204. Torres-Cabassa, A.S.; Gottesman, S. Capsule synthesis in Escherichia coli K-12 is regulated by proteolysis. J. Bacteriol. 1987, 169, 981–989. [Google Scholar] [CrossRef] [Green Version]
  205. Schoemaker, J.M.; Gayda, R.C.; Markovitz, A. Regulation of cell division in Escherichia coli: SOS induction and cellular location of the SulA protein, a key to lon-associated filamentation and death. J. Bacteriol. 1984, 158, 551–561. [Google Scholar] [CrossRef] [Green Version]
  206. Higashitani, A.; Ishii, Y.; Kato, Y.; Horiuchi, K. Functional dissection of a cell-division inhibitor, SulA, of Escherichia coli and its negative regulation by Lon. Mol. Gen. Genet. MGG 1997, 254, 351–357. [Google Scholar] [CrossRef] [PubMed]
  207. Aertsen, A.; Michiels, C.W. SulA-dependent hypersensitivity to high pressure and hyperfilamentation after high-pressure treatment of Escherichia coli lon mutants. Res. Microbiol. 2005, 156, 233–237. [Google Scholar] [CrossRef] [PubMed]
  208. Christensen, S.K.; Maenhaut-Michel, G.; Mine, N.; Gottesman, S.; Gerdes, K.; Van Melderen, L. Overproduction of the Lon protease triggers inhibition of translation in Escherichia coli: Involvement of the yefM-yoeB toxin-antitoxin system. Mol. Microbiol. 2004, 51, 1705–1717. [Google Scholar] [CrossRef]
  209. van Melderen, L.; Thi, M.H.D.; Lecchi, P.; Gottesman, S.; Couturier, M.; Maurizi, M.R. ATP-dependent Degradation of CcdA by Lon Protease: Effects of Secondary Structure and Heterologous Subunit Interactions. J. Biol. Chem. 1996, 271, 27730–27738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Hansen, S.; Vulić, M.; Min, J.; Yen, T.-J.; Schumacher, M.A.; Brennan, R.G.; Lewis, K. Regulation of the Escherichia coli HipBA Toxin-Antitoxin System by Proteolysis. PLoS ONE 2012, 7, e39185. [Google Scholar] [CrossRef]
  211. Theodore, A.; Lewis, K.; Vulić, M. Tolerance of Escherichia coli to Fluoroquinolone Antibiotics Depends on Specific Components of the SOS Response Pathway. Genetics 2013, 195, 1265–1276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Ramisetty, B.C.M.; Ghosh, D.; Roy Chowdhury, M.; Santhosh, R.S. What Is the Link between Stringent Response, Endoribonuclease Encoding Type II Toxin–Antitoxin Systems and Persistence? Front. Microbiol. 2016, 7, 1882. [Google Scholar] [CrossRef] [Green Version]
  213. Mohiuddin, S.G.; Massahi, A.; Orman, M.A. Lon Deletion Impairs Persister Cell Resuscitation in Escherichia coli. bioRxiv 2021. [Google Scholar] [CrossRef]
  214. Harms, A.; Fino, C.; Sørensen, M.A.; Semsey, S.; Gerdes, K. Prophages and growth dynamics confound experimental results with antibiotic-tolerant persister cells. MBio 2017, 8. [Google Scholar] [CrossRef] [Green Version]
  215. Prysak, M.H.; Mozdzierz, C.J.; Cook, A.M.; Zhu, L.; Zhang, Y.; Inouye, M.; Woychik, N.A. Bacterial toxin YafQ is an endoribonuclease that associates with the ribosome and blocks translation elongation through sequence-specific and frame-dependent mRNA cleavage. Mol. Microbiol. 2009, 71, 1071–1087. [Google Scholar] [CrossRef]
  216. Aizenman, E.; Engelberg-Kulka, H.; Glaser, G. An Escherichia coli chromosomal “addiction module” regulated by guanosine [corrected] 3’,5’-bispyrophosphate: A model for programmed bacterial cell death. Proc. Natl. Acad. Sci. USA 1996, 93, 6059–6063. [Google Scholar] [CrossRef] [Green Version]
  217. Erbse, A.; Schmidt, R.; Bornemann, T.; Schneider-Mergener, J.; Mogk, A.; Zahn, R.; Dougan, D.A.; Bukau, B. ClpS is an essential component of the N-end rule pathway in Escherichia coli. Nature 2006, 439, 753–756. [Google Scholar] [CrossRef] [PubMed]
  218. Dougan, D.A.; Reid, B.G.; Horwich, A.L.; Bukau, B. ClpS, a Substrate Modulator of the ClpAP Machine. Mol. Cell 2002, 9, 673–683. [Google Scholar] [CrossRef]
  219. Dubiel, A.; Wegrzyn, K.; Kupinski, A.P.; Konieczny, I. ClpAP protease is a universal factor that activates the parDE toxin-antitoxin system from a broad host range RK2 plasmid. Sci. Rep. 2018, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  220. Farewell, A.; Diez, A.A.; DiRusso, C.C.; Nyström, T. Role of the Escherichia coli FadR regulator in stasis survival and growth phase-dependent expression of the uspA, fad, and fab genes. J. Bacteriol. 1996, 178, 6443–6450. [Google Scholar] [CrossRef] [Green Version]
  221. DiRusso, C.C.; Nyström, T. The fats of Escherichia coli during infancy and old age: Regulation by global regulators, alarmones and lipid intermediates. Mol. Microbiol. 1998, 27, 1–8. [Google Scholar] [CrossRef]
  222. Jimenez-Diaz, L.; Caballero, A.; Segura, A. Regulation of Fatty Acids Degradation in Bacteria. Aerob. Util. Hydrocarb. Oils Lipids 2019, 751–771. [Google Scholar] [CrossRef]
  223. John, E.; Cronan, J.; Subrahmanyam, S. FadR, transcriptional co-ordination of metabolic expediency. Mol. Microbiol. 1998, 29, 937–943. [Google Scholar] [CrossRef]
  224. Spector, M.P.; DiRusso, C.C.; Pallen, M.J.; del Portillo, F.G.; Dougan, G.; Finlay, B.B. The medium-/long-chain fatty acyl-CoA dehydrogenase (fadF) gene of Salmonella typhimurium is a phase 1 starvation-stress response (SSR) locus. Microbiology 1999, 145, 15–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Kvint, K.; Hosbond, C.; Farewell, A.; Nybroe, O.; Nyström, T. Emergency derepression: Stringency allows RNA polymerase to override negative control by an active repressor. Mol. Microbiol. 2000, 35, 435–443. [Google Scholar] [CrossRef] [Green Version]
  226. Shen, S.; Faouzi, S.; Souquere, S.; Pierron, R.; Scoazec, J.-Y.; Correspondence, C.R.; Roy, S.; Routier, E.; Libenciuc, C.; André, F.; et al. Melanoma Persister Cells Are Tolerant to BRAF/MEK Inhibitors via ACOX1-Mediated Fatty Acid Oxidation. CellReports 2020, 33, 108421. [Google Scholar] [CrossRef]
  227. Karki, P.; Angardi, V.; Mier, J.C.; Orman, M.A. A Transient Metabolic State In Melanoma Persister Cells Mediated By Chemotherapeutic Treatments. bioRxiv 2021. [Google Scholar] [CrossRef]
  228. Imlay, J.A.; Fridovich, I. Assay of metabolic superoxide production in Escherichia coli. J. Biol. Chem. 1991, 266, 6957–6965. [Google Scholar] [CrossRef]
  229. Imlay, J.A.; Linn, S. DNA damage and oxygen radical toxicity. Sci. Sci. 1988, 240, 1302–1309. [Google Scholar] [CrossRef] [Green Version]
  230. Wolff, S.P.; Garner, A.; Dean, R.T. Free radicals, lipids and protein degradation. Trends Biochem. Sci. 1986, 11, 27–31. [Google Scholar] [CrossRef]
  231. Nyström, T. The glucose-starvation stimulon of Escherichia coli: Induced and repressed synthesis of enzymes of central metabolic pathways and role of acetyl phosphate in gene expression and starvation survival. Mol. Microbiol. 1994, 12, 833–843. [Google Scholar] [CrossRef] [PubMed]
  232. Mohiuddin, S.G.; Hoang, T.; Saba, A.; Karki, P.; Orman, M.A. Identifying Metabolic Inhibitors to Reduce Bacterial Persistence. Front. Microbiol. 2020, 11, 472. [Google Scholar] [CrossRef] [Green Version]
  233. Cameron, D.R.; Shan, Y.; Zalis, E.A.; Isabella, V.; Lewis, K. A genetic determinant of persister cell formation in bacterial pathogens. J. Bacteriol. 2018, 200. [Google Scholar] [CrossRef] [Green Version]
  234. Allison, K.R.; Brynildsen, M.P.; Collins, J.J. Metabolite-enabled eradication of bacterial persisters by aminoglycosides. Nature 2011, 473, 216–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Orman, M.A.; Brynildsen, M.P. Establishment of a method to rapidly assay bacterial persister metabolism. Antimicrob. Agents Chemother. 2013, 57, 4398–4409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Rosenberg, C.R.; Fang, X.; Allison, K.R. Potentiating aminoglycoside antibiotics to reduce their toxic side effects. PLoS ONE 2020, 15, e0237948. [Google Scholar] [CrossRef] [PubMed]
  237. Bokinsky, G.; Baidoo, E.E.K.; Akella, S.; Burd, H.; Weaver, D.; Alonso-Gutierrez, J.; García-Martín, H.; Lee, T.S.; Keasling, J.D. Hipa-triggered growth arrest and β-lactam tolerance in escherichia coli are mediated by RelA-dependent ppGpp synthesis. J. Bacteriol. 2013, 195, 3173–3182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Radzikowski, J.L.; Vedelaar, S.; Siegel, D.; Ortega, Á.D.; Schmidt, A.; Heinemann, M. Bacterial persistence is an active σS stress response to metabolic flux limitation. Mol. Syst. Biol. 2016, 12, 882. [Google Scholar] [CrossRef]
  239. Mok, W.W.K.; Park, J.O.; Rabinowitz, J.D.; Brynildsen, M.P. RNA futile cycling in model persisters derived from mazF accumulation. MBio 2015, 6. [Google Scholar] [CrossRef] [Green Version]
  240. Fu, H.; Le, S.; Chen, H.; Muniyappa, K.; Yan, J. Force and ATP hydrolysis dependent regulation of RecA nucleoprotein filament by single-stranded DNA binding protein. Nucleic Acids Res. 2013, 41, 924–932. [Google Scholar] [CrossRef] [Green Version]
  241. McEntee, K.; Weinstock, G.M.; Lehman, I.R. Initiation of general recombination catalyzed in vitro by the recA protein of Escherichia coli. Proc. Natl. Acad. Sci. USA 1979, 76, 2615–2619. [Google Scholar] [CrossRef] [Green Version]
  242. Shibata, T.; DasGupta, C.; Cunningham, R.P.; Radding, C.M. Purified Escherichia coli recA protein catalyzes homologous pairing of superhelical DNA and single-stranded fragments. Proc. Natl. Acad. Sci. USA 1979, 76, 1638–1642. [Google Scholar] [CrossRef] [Green Version]
  243. Cox, M.M.; Lehman, I.R. recA protein-promoted DNA strand exchange. Stable complexes of recA protein and single-stranded DNA formed in the presence of ATP and single-stranded DNA binding protein. J. Biol. Chem. 1982, 257, 8523–8532. [Google Scholar] [CrossRef]
  244. Ma, C.; Sim, S.; Shi, W.; Du, L.; Xing, D.; Zhang, Y. Energy production genes sucB and ubiF are involved in persister survival and tolerance to multiple antibiotics and stresses in Escherichia coli. FEMS Microbiol. Lett. 2010, 303, 33–40. [Google Scholar] [CrossRef] [Green Version]
  245. Orman, M.A.; Brynildsen, M.P. Persister formation in Escherichia coli can be inhibited by treatment with nitric oxide. Free Radic. Biol. Med. 2016, 93, 145–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Grainger, D.C.; Aiba, H.; Hurd, D.; Browning, D.F.; Busby, S.J.W. Transcription factor distribution in Escherichia coli: Studies with FNR protein. Nucleic Acids Res. 2007, 35, 269–278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Zinser, E.R.; Kolter, R. Prolonged stationary-phase incubation selects for lrp mutations in Escherichia coli K-12. J. Bacteriol. 2000, 182, 4361–4365. [Google Scholar] [CrossRef] [Green Version]
  248. Tani, T.H.; Khodursky, A.; Blumenthal, R.M.; Brown, P.O.; Matthews, R.G. Adaptation to famine: A family of stationary-phase genes revealed by microarray analysis. Proc. Natl. Acad. Sci. USA 2002, 99, 13471–13476. [Google Scholar] [CrossRef] [Green Version]
  249. Brown, L.; Gentry, D.; Elliott, T.; Cashel, M. DksA affects ppGpp induction of RpoS at a translational level. J. Bacteriol. 2002, 184, 4455–4465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  250. Paul, B.J.; Barker, M.M.; Ross, W.; Schneider, D.A.; Webb, C.; Foster, J.W.; Gourse, R.L. DksA: A Critical Component of the Transcription Initiation Machinery that Potentiates the Regulation of rRNA Promoters by ppGpp and the Initiating NTP. Cell 2004, 118, 311–322. [Google Scholar] [CrossRef] [Green Version]
  251. Franchini, A.G.; Ihssen, J.; Egli, T. Effect of Global Regulators RpoS and Cyclic-AMP/CRP on the Catabolome and Transcriptome of Escherichia coli K12 during Carbon- and Energy-Limited Growth. PLoS ONE 2015, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Mika, F.; Hengge, R. A two-component phosphotransfer network involving ArcB, ArcA, and RssB coordinates synthesis and proteolysis of σS (RpoS) in E. coli. Genes Dev. 2005, 19, 2770–2781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Orman, M.A.; Mok, W.W.K.; Brynildsen, M.P. Aminoglycoside-Enabled Elucidation of Bacterial Persister Metabolism. Curr. Protoc. Microbiol. 2015, 36, 17.9.1–17.9.4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Gefen, O.; Gabay, C.; Mumcuoglu, M.; Engel, G.; Balaban, N.Q. Single-cell protein induction dynamics reveals a period of vulnerability to antibiotics in persister bacteria. Proc. Natl. Acad. Sci. USA 2008, 105, 6145–6149. [Google Scholar] [CrossRef] [Green Version]
  255. Viale, A.; Pettazzoni, P.; Lyssiotis, C.A.; Ying, H.; Sánchez, N.; Marchesini, M.; Carugo, A.; Green, T.; Seth, S.; Giuliani, V.; et al. Oncogene ablation-resistant pancreatic cancer cells depend on mitochondrial function. Nature 2014, 514, 628–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Echeverria, G.V.; Ge, Z.; Seth, S.; Zhang, X.; Jeter-Jones, S.; Zhou, X.; Cai, S.; Tu, Y.; McCoy, A.; Peoples, M.; et al. Resistance to neoadjuvant chemotherapy in triple-negative breast cancer mediated by a reversible drug-tolerant state. Sci. Transl. Med. 2019, 11. [Google Scholar] [CrossRef]
  257. Raha, D.; Wilson, T.R.; Peng, J.; Peterson, D.; Yue, P.; Evangelista, M.; Wilson, C.; Merchant, M.; Settleman, J. The cancer stem cell marker aldehyde dehydrogenase is required to maintain a drug-tolerant tumor cell subpopulation. Cancer Res. 2014, 74, 3579–3590. [Google Scholar] [CrossRef] [Green Version]
  258. Roesch, A.; Fukunaga-Kalabis, M.; Schmidt, E.C.; Zabierowski, S.E.; Brafford, P.A.; Vultur, A.; Basu, D.; Gimotty, P.; Vogt, T.; Herlyn, M. A Temporarily Distinct Subpopulation of Slow-Cycling Melanoma Cells Is Required for Continuous Tumor Growth. Cell 2010, 141, 583–594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Dawson, C.C.; Intapa, C.; Jabra-Rizk, M.A. “Persisters”: Survival at the Cellular Level. PLOS Pathog. 2011, 7, e1002121. [Google Scholar] [CrossRef] [Green Version]
  260. Borst, P. Cancer drug pan-resistance: Pumps, cancer stem cells, quiescence, epithelial to mesenchymal transition, blocked cell death pathways, persisters or what? Open Biol. 2012, 2. [Google Scholar] [CrossRef] [Green Version]
  261. Baguley, B.C. Multiple Drug Resistance Mechanisms in Cancer. Mol. Biotechnol. 2010, 46, 308–316. [Google Scholar] [CrossRef] [PubMed]
  262. Redmond, K.M.; Wilson, T.R.; Johnston, P.G.; Longley, D.B. Resistance mechanisms to cancer chemotherapy. Front. Biosci. 2008, 13, 5138–5154. [Google Scholar] [CrossRef] [Green Version]
  263. Dikic, I.; Elazar, Z. Mechanism and medical implications of mammalian autophagy. Nat. Rev. Mol. Cell Biol. 2018, 19, 349–364. [Google Scholar] [CrossRef]
  264. Mizushima, N. Autophagy: Process and function. Genes Dev. 2007, 21, 2861–2873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Cooper, G.M. Lysosomes. In The Cell: A Molecular Approach, 2rd ed.; Sinauer Associates: Sunderland, MA, USA, 2000. [Google Scholar]
  266. Zientara-Rytter, K.; Subramani, S. Autophagic degradation of peroxisomes in mammals. Biochem. Soc. Trans. 2016, 44, 431–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Höhn, A.; Tramutola, A.; Cascella, R. Proteostasis Failure in Neurodegenerative Diseases: Focus on Oxidative Stress. Oxid. Med. Cell. Longev. 2020, 2020. [Google Scholar] [CrossRef] [Green Version]
  268. Yorimitsu, T.; Klionsky, D.J. Autophagy: Molecular machinery for self-eating. Cell Death Differ. 2005, 12, 1542–1552. [Google Scholar] [CrossRef] [Green Version]
  269. Wesselborg, S.; Stork, B. Autophagy signal transduction by ATG proteins: From hierarchies to networks. Cell. Mol. Life Sci. 2015, 72, 4721–4757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Fred Dice, J. Peptide sequences that target cytosolic proteins for lysosomal proteolysis. Trends Biochem. Sci. 1990, 15, 305–309. [Google Scholar] [CrossRef]
  271. Dice, J.F. Chaperone-Mediated Autophagy. Autophagy 2007, 3, 295–299. [Google Scholar] [CrossRef] [Green Version]
  272. Feng, Y.; Yao, Z.; Klionsky, D.J. How to control self-digestion: Transcriptional, post-transcriptional, and post-translational regulation of autophagy. Trends Cell Biol. 2015, 25, 354–363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Mathiassen, S.G.; De Zio, D.; Cecconi, F. Autophagy and the Cell Cycle: A Complex Landscape. Front. Oncol. 2017, 7, 51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Aqbi, H.F.; Butler, S.E.; Keim, R.; Idowu, M.O.; Manjili, M.H. Chemotherapy-induced tumor dormancy and relapse. J. Immunol. 2017, 198. [Google Scholar]
  275. Kun, E.; Tsang, Y.T.M.; Ng, C.W.; Gershenson, D.M.; Wong, K.K. MEK inhibitor resistance mechanisms and recent developments in combination trials. Cancer Treat. Rev. 2021, 92, 102137. [Google Scholar] [CrossRef] [PubMed]
  276. Kurppa, K.J.; Liu, Y.; To, C.; Zhang, T.; Fan, M.; Vajdi, A.; Knelson, E.H.; Xie, Y.; Lim, K.; Cejas, P.; et al. Treatment-Induced Tumor Dormancy through YAP-Mediated Transcriptional Reprogramming of the Apoptotic Pathway. Cancer Cell 2020, 37, 104–122.e12. [Google Scholar] [CrossRef]
  277. Wang, L.; Peng, Q.; Yin, N.; Xie, Y.; Xu, J.; Chen, A.; Yi, J.; Tang, J.; Xiang, J. Chromatin accessibility regulates chemotherapy-induced dormancy and reactivation. Mol. Ther. -Nucleic Acids 2021, 26, 269–279. [Google Scholar] [CrossRef]
  278. Tsujimoto, Y.; Shimizu, S. Another way to die: Autophagic programmed cell death. Cell Death Differ. 2005, 12, 1528–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  279. Komatsu, M.; Waguri, S.; Ueno, T.; Iwata, J.; Murata, S.; Tanida, I.; Ezaki, J.; Mizushima, N.; Ohsumi, Y.; Uchiyama, Y.; et al. Impairment of starvation-induced and constitutive autophagy in Atg7-deficient mice. J. Cell Biol. 2005, 169, 425–434. [Google Scholar] [CrossRef] [PubMed]
  280. Malhotra, R.; Warne, J.P.; Salas, E.; Xu, A.W.; Debnath, J. Loss of Atg12, but not Atg5, in pro-opiomelanocortin neurons exacerbates diet-induced obesity. Autophagy 2015, 11, 145–154. [Google Scholar] [CrossRef]
  281. Cheong, H.; Wu, J.; Gonzales, L.K.; Guttentag, S.H.; Thompson, C.B.; Lindsten, T. Analysis of a lung defect in autophagy-deficient mouse strains. Autophagy 2014, 10, 45–56. [Google Scholar] [CrossRef] [Green Version]
  282. Takamura, A.; Komatsu, M.; Hara, T.; Sakamoto, A.; Kishi, C.; Waguri, S.; Eishi, Y.; Hino, O.; Tanaka, K.; Mizushima, N. Autophagy-deficient mice develop multiple liver tumors. Genes Dev. 2011, 25, 795–800. [Google Scholar] [CrossRef] [Green Version]
  283. Yue, Z.; Jin, S.; Yang, C.; Levine, A.J.; Heintz, N. Beclin 1, an autophagy gene essential for early embryonic development, is a haploinsufficient tumor suppressor. Proc. Natl. Acad. Sci. USA 2003, 100, 15077–15082. [Google Scholar] [CrossRef] [Green Version]
  284. Degenhardt, K.; Mathew, R.; Beaudoin, B.; Bray, K.; Anderson, D.; Chen, G.; Mukherjee, C.; Shi, Y.; Gélinas, C.; Fan, Y.; et al. Autophagy promotes tumor cell survival and restricts necrosis, inflammation, and tumorigenesis. Cancer Cell 2006, 10, 51–64. [Google Scholar] [CrossRef] [Green Version]
  285. Strohecker, A.M.; Guo, J.Y.; Karsli-Uzunbas, G.; Price, S.M.; Chen, G.J.; Mathew, R.; McMahon, M.; White, E. Autophagy Sustains Mitochondrial Glutamine Metabolism and Growth of BrafV600E–Driven Lung Tumors. Cancer Discov. 2013, 3, 1272–1285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Guo, J.Y.; Chen, H.-Y.; Mathew, R.; Fan, J.; Strohecker, A.M.; Karsli-Uzunbas, G.; Kamphorst, J.J.; Chen, G.; Lemons, J.M.S.; Karantza, V.; et al. Activated Ras requires autophagy to maintain oxidative metabolism and tumorigenesis. Genes Dev. 2011, 25, 460–470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Lock, R.; Roy, S.; Kenific, C.M.; Su, J.S.; Salas, E.; Ronen, S.M.; Debnath, J. Autophagy facilitates glycolysis during Ras-mediated oncogenic transformation. Mol. Biol. Cell. 2010, 22, 165–178. [Google Scholar] [CrossRef] [PubMed]
  288. Yang, S.; Wang, X.; Contino, G.; Liesa, M.; Sahin, E.; Ying, H.; Bause, A.; Li, Y.; Stommel, J.M.; Dell’Antonio, G.; et al. Pancreatic cancers require autophagy for tumor growth. Genes Dev. 2011, 25, 717–729. [Google Scholar] [CrossRef] [Green Version]
  289. Ma, X.-H.; Piao, S.; Wang, D.; Mcafee, Q.W.; Nathanson, K.L.; Lum, J.J.; Li, L.Z.; Amaravadi, R.K. Measurements of Tumor Cell Autophagy Predict Invasiveness, Resistance to Chemotherapy, and Survival in Melanoma. Clin. Cancer Res. 2011, 17, 3478–3489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  290. Ma, X.-H.; Piao, S.-F.; Dey, S.; Mcafee, Q.; Karakousis, G.; Villanueva, J.; Hart, L.S.; Levi, S.; Hu, J.; Zhang, G.; et al. Targeting ER stress–induced autophagy overcomes BRAF inhibitor resistance in melanoma. J. Clin. Investig. 2014, 124, 1406–1417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Bray, K.; Mathew, R.; Lau, A.; Kamphorst, J.J.; Fan, J.; Chen, J.; Chen, H.-Y.; Ghavami, A.; Stein, M.; DiPaola, R.S.; et al. Autophagy Suppresses RIP Kinase-Dependent Necrosis Enabling Survival to mTOR Inhibition. PLoS ONE 2012, 7, e41831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Bollen, C.; Dewachter, L.; Michiels, J. Protein Aggregation as a Bacterial Strategy to Survive Antibiotic Treatment. Front. Mol. Biosci. 2021, 8, 259. [Google Scholar] [CrossRef]
  293. Ayrapetyan, M.; Williams, T.C.; Oliver, J.D. Bridging the gap between viable but non-culturable and antibiotic persistent bacteria. Trends Microbiol. 2015, 23, 7–13. [Google Scholar] [CrossRef] [PubMed]
  294. Ayrapetyan, M.; Williams, T.C.; Baxter, R.; Oliver, J.D. Viable but nonculturable and persister cells coexist stochastically and are induced by human serum. Infect. Immun. 2015, 83, 4194–4203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Grant, S.S.; Hung, D.T. Persistent bacterial infections, antibiotic tolerance, and the oxidative stress response. Virulence 2013, 4, 273–283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Li, L.; Mendis, N.; Trigui, H.; Oliver, J.D.; Faucher, S.P. The importance of the viable but non-culturable state in human bacterial pathogens. Front. Microbiol. 2014, 5, 258. [Google Scholar] [CrossRef] [Green Version]
  297. McCune, R.M.; Feldmann, F.M.; McDermott, W. Microbial Persistence II. Characteristics of the Sterile State of Tubercle bacilli. J. Exp. Med. 1966, 123, 469–486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Bamford, R.A.; Smith, A.; Metz, J.; Glover, G.; Titball, R.W.; Pagliara, S. Investigating the physiology of viable but non-culturable bacteria by microfluidics and time-lapse microscopy. BMC Biol. 2017, 15, 1–12. [Google Scholar] [CrossRef]
  299. Prax, M.; Bertram, R. Metabolic aspects of bacterial persisters. Front. Cell. Infect. Microbiol. 2014, 4, 148. [Google Scholar] [CrossRef] [PubMed]
  300. Bigger, J.W. Treatment of Staphyloeoeeal Infections with Penicillin by Intermittent Sterilisation. Lancet 1944, 244, 497–500. [Google Scholar] [CrossRef]
  301. Scherrer, R.; Moyed, H.S. Conditional impairment of cell division and altered lethality in hipA mutants of Escherichia coli K-12. J. Bacteriol. 1988, 170, 3321–3326. [Google Scholar] [CrossRef] [Green Version]
  302. Keren, I.; Kaldalu, N.; Spoering, A.; Wang, Y.; Lewis, K. Persister cells and tolerance to antimicrobials. FEMS Microbiol. Lett. 2004, 230, 13–18. [Google Scholar] [CrossRef] [Green Version]
  303. Keseler, I.M.; Mackie, A.; Santos-Zavaleta, A.; Billington, R.; Bonavides-Martínez, C.; Caspi, R.; Fulcher, C.; Gama-Castro, S.; Kothari, A.; Krummenacker, M.; et al. The EcoCyc database: Reflecting new knowledge about Escherichia coli K-12. Nucleic Acids Res. 2017, 45, D543–D550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Goode, O.; Smith, A.; Łapińska, U.; Bamford, R.; Kahveci, Z.; Glover, G.; Attrill, E.; Carr, A.; Metz, J.; Pagliara, S. Heterologous Protein Expression Favors the Formation of Protein Aggregates in Persister and Viable but Nonculturable Bacteria. ACS Infect. Dis. 2021, 7, 1848–1858. [Google Scholar] [CrossRef] [PubMed]
  305. Leszczynska, D.; Matuszewska, E.; Kuczynska-Wisnik, D.; Furmanek-Blaszk, B.; Laskowska, E. The Formation of Persister Cells in Stationary-Phase Cultures of Escherichia coli Is Associated with the Aggregation of Endogenous Proteins. PLoS ONE 2013, 8, e54737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Dewachter, L.; Fauvart, M.; Michiels, J. Bacterial Heterogeneity and Antibiotic Survival: Understanding and Combatting Persistence and Heteroresistance. Mol. Cell 2019, 76, 255–267. [Google Scholar] [CrossRef]
  307. Dewachter, L.; Bollen, C.; Wilmaerts, D.; Louwagie, E.; Herpels, P.; Matthay, P.; Khodaparast, L.; Khodaparast, L.; Rousseau, F.; Schymkowitz, J.; et al. The Dynamic Transition of Persistence toward the Viable but Nonculturable State during Stationary Phase Is Driven by Protein Aggregation. MBio 2021, 12. [Google Scholar] [CrossRef]
  308. Tyedmers, J.; Mogk, A.; Bukau, B. Cellular strategies for controlling protein aggregation. Nat. Rev. Mol. Cell Biol. 2010, 11, 777–788. [Google Scholar] [CrossRef] [PubMed]
  309. Tannous, P.; Zhu, H.; Nemchenko, A.; Berry, J.M.; Johnstone, J.L.; Shelton, J.M.; Francis, J.; Miller, J.; Rothermel, B.A.; Hill, J.A. Intracellular Protein Aggregation Is a Proximal Trigger of Cardiomyocyte Autophagy. Circulation 2008, 117, 3070–3078. [Google Scholar] [CrossRef] [Green Version]
  310. Filimonenko, M.; Isakson, P.; Finley, K.D.; Anderson, M.; Jeong, H.; Melia, T.J.; Bartlett, B.J.; Myers, K.M.; Birkeland, H.C.G.; Lamark, T.; et al. The Selective Macroautophagic Degradation of Aggregated Proteins Requires the PI3P-Binding Protein Alfy. Mol. Cell 2010, 38, 265–279. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Self-digestion mediated stationary-phase metabolism in bacteria. Self-digestion enables cells to transiently tolerate starvation conditions by recycling essential energy molecules. Perturbing the proposed metabolic mechanism genetically (deleting TCA cycle enzymes) [55], chemically (chlorpromazine (CPZ) [232], potassium cyanide (KCN) [55], and nitric oxide (NO) [245] treatments), and environmentally (removing O2) [55] can reduce persister formation during the stationary phase.
Figure 1. Self-digestion mediated stationary-phase metabolism in bacteria. Self-digestion enables cells to transiently tolerate starvation conditions by recycling essential energy molecules. Perturbing the proposed metabolic mechanism genetically (deleting TCA cycle enzymes) [55], chemically (chlorpromazine (CPZ) [232], potassium cyanide (KCN) [55], and nitric oxide (NO) [245] treatments), and environmentally (removing O2) [55] can reduce persister formation during the stationary phase.
Microorganisms 09 02269 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mohiuddin, S.G.; Ghosh, S.; Ngo, H.G.; Sensenbach, S.; Karki, P.; Dewangan, N.K.; Angardi, V.; Orman, M.A. Cellular Self-Digestion and Persistence in Bacteria. Microorganisms 2021, 9, 2269. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms9112269

AMA Style

Mohiuddin SG, Ghosh S, Ngo HG, Sensenbach S, Karki P, Dewangan NK, Angardi V, Orman MA. Cellular Self-Digestion and Persistence in Bacteria. Microorganisms. 2021; 9(11):2269. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms9112269

Chicago/Turabian Style

Mohiuddin, Sayed Golam, Sreyashi Ghosh, Han G. Ngo, Shayne Sensenbach, Prashant Karki, Narendra K. Dewangan, Vahideh Angardi, and Mehmet A. Orman. 2021. "Cellular Self-Digestion and Persistence in Bacteria" Microorganisms 9, no. 11: 2269. https://0-doi-org.brum.beds.ac.uk/10.3390/microorganisms9112269

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop