FEMS Microbiology Reviews, fux015, 41, 2017, 549–574 doi: 10.1093/femsre/fux015 Advance Access Publication Date: 11 April 2017 Review Article REVIEW ARTICLE Regulation of heat-shock genes in bacteria: from signal sensing to gene expression output Davide Roncarati∗ and Vincenzo Scarlato Department of Pharmacy and Biotechnology (FaBiT), University of Bologna, 40126 Bologna, Italy ∗ Corresponding author: Department of Pharmacy and Biotechnology (FaBiT), University of Bologna, via Selmi, 3, 40126 Bologna, Italy. Tel: (+39) 051 2099320; Fax: (+39) 051 2094286; E-mail: davide.roncarati@unibo.it One sentence summary: The authors provide an updated review of the strategies employed by many important bacteria to finely regulate transcription of heat-shock genes, including a comprehensive description of the several ways in which environmental cues are perceived and processed. Editor: Franz Narberhaus ABSTRACT The heat-shock response is a mechanism of cellular protection against sudden adverse environmental growth conditions and results in the prompt production of various heat-shock proteins. In bacteria, specific sensory biomolecules sense temperature fluctuations and transduce intercellular signals that coordinate gene expression outputs. Sensory biomolecules, also known as thermosensors, include nucleic acids (DNA or RNA) and proteins. Once a stress signal is perceived, it is transduced to invoke specific molecular mechanisms controlling transcription of genes coding for heat-shock proteins. Transcriptional regulation of heat-shock genes can be under either positive or negative control mediated by dedicated regulatory proteins. Positive regulation exploits specific alternative sigma factors to redirect the RNA polymerase enzyme to a subset of selected promoters, while negative regulation is mediated by transcriptional repressors. Interestingly, while various bacteria adopt either exclusively positive or negative mechanisms, in some microorganisms these two opposite strategies coexist, establishing complex networks regulating heat-shock genes. Here, we comprehensively summarize molecular mechanisms that microorganisms have adopted to finely control transcription of heat-shock genes. Keywords: heat-shock response; transcriptional regulation; alternative σ factors; HspR repressor; HrcA repressor; CtsR repressor INTRODUCTION The heat-shock response is a protective mechanism that is crucial for bacterial survival and adaptation to hostile environmental conditions. This response is apparently universal as it has been observed in every bacterial species investigated. It consists of a set of well-coordinated responses and processes, mostly involving the strictly regulated production of various heat-shock proteins. These, in turn, mainly include molecular chaperones and proteases, whose intracellular abundance rapidly increases upon exposure to a variety of environmental stresses. The principal function of heat-shock proteins is to assist protein folding. They are also involved in rescue or degradation of denatured proteins and deleterious misfolded aggregates. Upon sudden temperature increase, heat-shock proteins transiently accumulate in the cell. Once the organism has adapted to the new temperature, the amount of heat-shock proteins decreases to a steady-state level that is frequently greater than the initial basal level in order to assist bacterial growth under non-optimal environmental conditions. Besides their roles in protecting cellular proteins from environmental insults and in maintaining cellular homeostasis, some heat-shock proteins are important virulence factors, while others appear to affect pathogenesis indirectly. Although the heat-shock response is universally conserved in both prokaryotes and eukaryotes, the basic molecular mechanisms governing the regulation of heat-shock genes differ considerably among Received: 30 November 2016; Accepted: 14 March 2017 C FEMS 2017. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com 549 Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 550 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 bacterial species. In particular, bacteria have evolved different regulatory strategies that combine transcriptional and posttranscriptional mechanisms to rapidly induce heat-shock protein synthesis only when required. In order to rapidly respond to environmental cues and to ensure expression of heat-shock proteins only when appropriate, bacteria employ specific sensory biomolecules that are able to perceive temperature fluctuations and translate them into coordinated gene expression outputs. In this respect, a wide repertoire of thermal sensors has evolved among microorganisms. Specifically, bacteria can sense temperature changes using nucleic acids (DNA or RNA) or proteins. In the present review article, we provide an updated, comprehensive description of the strategies employed by many important microorganisms to finely regulate transcription of genes coding for heat-shock proteins. In the first part of the article, we survey the molecular mechanisms of regulation (e.g. positive, negative, simple and complex) that guarantee the proper transcriptional output once stress signals are detected. In the second part, we provide a comprehensive description of the several ways in which environmental cues are perceived and processed. The heat-shock response: role and principal mediators The heat-shock response was first described by Ritossa as a phenomenon observed in the polytene chromosomes of salivary glands of Drosophila melanogaster flies when moved from their normal growth temperature to 37◦ C (Ritossa 1962). Triggered by environmental stress insults, the heat-shock response consists of the induction or upregulation of synthesis of a group of proteins that help the cell to survive under conditions that would normally be lethal. Actually, several stress conditions are able to elicit heat-shock response, such as temperature variations, osmotic changes, desiccation, antibiotics, solvents or heavy metals. However, the heat-shock response triggered by a sudden temperature increase has been widely used as a model system for studying the impact of stress on biological systems. Environmental stress conditions primarily affect cellular protein homeostasis (maintenance of an overall balance of folded proteins), posing a serious threat to their integrity. Heatshock proteins are involved in several processes in bacterial cells, including assisting the folding of newly synthesized proteins, preventing aggregation of proteins under stress conditions and recovering proteins that have been partially or completely unfolded by stresses such as a sudden temperature increase. During stress, in fact, spontaneous folding of newly synthetized proteins is inefficient and error-prone, and a predominant fraction of nascent polypeptides undergoes a chaperone-mediated folding process (Mogk et al. 1999). Also, a large fraction of already folded proteins gets partially or completely denatured and becomes prone to form deleterious aggregates. Moreover, the rates of transcription and translation decrease because the stability of the RNA polymerase subunits and of the translation factor EF-G are sensitive to temperature increase (Mogk et al. 1999). Some heat-shock proteins are also required during normal growth condition and they are abundant under all metabolic conditions. GroEL and DnaK, the bacterial representatives of the two major chaperone families Hsp60 and Hsp70, respectively, play a key role in protein folding even during nonstressed growth conditions, although their action becomes more important during stress. They bind hydrophobic surfaces of unfolded proteins and, together with their co-chaperones (GroES and DnaJ-GrpE) and ATP hydrolysis, they promote the acquisition of proper folding by the substrate polypeptides through different mechanisms that reflect their different architectures. While GroEL monomers assemble into a cylindrical complex to form two heptameric rings that enclose the entire substrate protein inside a large cavity, giving the polypeptide a chance to fold without interacting with any other protein, the DnaK chaperone is mostly monomeric, exerting its action by binding short surface-exposed hydrophobic amino-acid sequences (Xu, Horwich and Sigler 1997; Mogk et al. 1999). It is worth mentioning that the names groEL and groES should be used for Escherichia coli genes only, because this nomenclature was adopted following the observation that mutation of these genes prevented the plating of several bacteriophages (Georgopoulos et al. 1972; Takano and Kakefuda 1972). An alternative used nomenclature has been proposed, in which homologs of GroEL are called Cpn60, and those of GroES are named Cpn10 (Coates, Shinnick and Ellis 1993). As both nomenclatures are still widely used, in this review we decided to adopt the nomenclature used for E. coli, GroEL and GroES. Another group of heat-shock proteins expressed by the bacterial cell and upregulated during stress is composed by proteases. These proteins are responsible essentially for the removal of damaged polypeptides from stressed cells. Some proteases are multicomponent systems, in which a catalytic subunit (e.g. ClpP and HslV) associates to substrate recognition subunits (ClpA or ClpX for ClpP and HslU for HslV), which are co-chaperones able to remodel substrate polypeptides upon ATP hydrolysis and deliver them to proteolytic degradation (Missiakas et al. 1996; Wawrzynow, Banecki and Zylicz 1996). While these proteases assemble into complex ring-shaped structures, other members of this group of heat-shock proteins combine on a single polypeptide both chaperone and protease activities (e.g. Lon, FtsH and the periplasmic serine protease DegP). The small heat-shock proteins constitute a highly heterogeneous group of proteins whose expression is induced upon stress challenges. Their principal role is to bind and protect unfolded proteins, by holding them in a conformation not subjected to degradation, until they will be efficiently refolded by ATP-driven chaperones (Matuszewska et al. 2005). Finally, some heat-shock proteins have been proposed as important virulence factors, while others appear to affect pathogenesis indirectly. In this respect, there are several examples in which molecular chaperones seem to have taken on diverse functions other than just protein folding. For example, GroEL1 of Mycobacterium smegmatis, a GroEL paralog with no heat-shockrelated functions, is involved in mycolic acid synthesis and biofilm formation (Ojha et al. 2005). Moreover, the GroES homolog of Helicobacter pylori, besides its prototypical role as cochaperonin, plays a role in the storage and trafficking of Ni2+ ions, a crucial virulence determinant for this human gastric pathogen (de Reuse, Vinella and Cavazza 2013). On the other hand, there is evidence that molecular chaperones in many cases might act as direct virulence factors. Even though it still remains a controversial concept, it has been proposed that several bacterial species use cell surface GroEL and DnaK chaperones as adhesins during the interaction with host cells (reviewed in Henderson and Martin 2011). Moreover, chaperones proteins have been shown to have cell-to-cell signaling properties and, in several cases, are able to induce the synthesis of proinflammatory cytokines and promote apoptosis (Henderson and Martin 2011). Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 551 REGULATORY MECHANISMS OF HEAT-SHOCK GENES In bacteria, regulation of heat-shock genes can be positive or negative. Positive regulation exploits alternative sigma factors to specifically direct the transcriptional machinery to a subset of selected promoters, while negative regulation is orchestrated by transcriptional repressors. Interestingly, while several bacteria adopt exclusively positive or negative mechanisms, in some microorganisms these opposite strategies coexist, establishing a complex regulatory network of heat-shock genes. Positive transcriptional regulation of heat-shock genes Positive transcriptional regulation of heat-shock genes relies on the use of a dedicated alternative σ factor, the subunit of the RNA polymerase that confers promoter recognition specificity to the transcription enzyme. In general, this transcriptional control mechanism exploits the ability of condition-specific alternative σ factors to outcompete the housekeeping σ subunit normally associated with RNA polymerase and redirects the enzyme to a selected subset of heat-shock gene promoters, thereby reprogramming cellular transcription. The strategy of transiently increasing transcription of heat-shock genes being dependent upon temperature stress through a dedicated σ factor is widely used among bacteria, although the model organism Escherichia coli has long served as the paradigm for this kind of regulation. Escherichia coli heat-shock sigma-32 (σ32 ) The heat-shock response in E. coli is governed by the htpR gene product, whose high amino acid sequence similarity with the housekeeping σ70 was previously described (Landick et al. 1984). The 32-kDa HtpR polypeptide was purified from crude extracts and combined in vitro with RNA polymerase ‘core’ enzyme, consisting of the four subunits α2, β and β’only (a preparation of the RNA polymerase enzyme that retains catalytic activity but is unable to initiate de novo transcription in vitro). This in vitro transcription assay showed that HtpR was able to confer specificity to RNA polymerase in recognizing heat-shock promoters, without requiring the vegetative sigma factor, σ70 . Following the demonstration that HtpR was a σ factor able to promote transcription initiation at heat-shock promoters, its gene was renamed rpoH (in analogy to the already characterized rpoD gene, coding for the vegetative σ70 factor) and the gene product was called sigma-32 (σ32 , sometimes also called σH ) (Grossman, Erickson and Gross 1984). As stated above, to exert its function, σ32 interacts with RNA polymerase core enzyme and directs the transcription machinery to specific promoters. The vegetative σ factor and almost all alternative σ factors constitute a homologous set of proteins (except σ54 , which belongs to a separate family), the σ70 family of proteins, with several regions of sequence conservation. The extensive σ-core interaction in the formation of the RNA polymerase holoenzyme (the enzymatic form with α2ββ’σ subunit composition) involves the same regions and even equivalent residues in σ32 and σ70 (Sharp et al. 1999). What specifically targets the transcription machinery to heat-shock genes is the recognition of particular promoter elements by σ32 DNA-binding domains (DBD) (Kumar et al. 1995). Several genome-wide studies that addressed the definition of heat-shock promoter elements confirmed that those recognized by σ32 are different from those bound by RNA polymerase associated with the housekeeping σ70 . Specifically, while the –35 box consensus sequence for σ32 binding, TTGAAA, is highly similar to the σ70 TTGACA consensus, the σ32 specific CCCCATNT -10 box shows no similarities to the σ70 TATAAT -10 consensus hexamer (Nonaka et al. 2006; Wade et al. 2006; Koo et al. 2009). Accordingly, a mutational analysis of residues belonging to σ32 regions 2.4 and 4.2 (conserved regions among the various σ factors belonging to the σ70 family, which are involved in the recognition of –10 and –35 core promoter elements, respectively) showed that the recognition of the –35 box is mediated by similar amino acids clustered in regions 4.2 of σ32 and σ70 , while the interaction with the downstream –10 element involves different residues (Kourennaia, Tsujikawa and Dehaseth 2005). The rapid, transient transcriptional increase of heat-shock genes upon shift to higher temperature is directly linked to the amount of active σ32 in the cell. Normally present in very small amounts and with a half-life of just a few minutes at 37◦ C, the σ32 level rapidly and transiently increases during heat-shock as a result of changes in its rate of synthesis and stability (Grossman et al 1987; Lesley, Thompson and Burgess 1987; Straus, Walter and Gross 1987). Even though the cellular level of σ32 depends mostly on post-transcriptional regulatory mechanisms, the control of rpoH transcription appears to be quite complex. The region upstream of the rpoH coding sequence harbors five promoters (named P1 and P3 through P6) recognized by different σ factors and responsive to several stimuli (Fig. 1, left panel). Specifically, P4 and P5 are transcribed by σ70 -associated RNA polymerase and P3 and P6 are bound by σ24 and σ54 , respectively, while P1 can be transcribed by σ70 or σS depending on the growth phase (Erickson et al. 1987; Erickson and Gross 1989; Wang and Kaguni 1989a; Nagai et al. 1990; Missiakas and Raina 1998; Pallen 1999; Janaszak et al. 2009). In addition, the P6 promoter is catabolite sensitive and its transcription is dependent on cAMP levels and on the CRP activator (Nagai et al. 1990), while cAMP-CRP together with CytR anti-activator form a repression complex overlapping P3, P4 and P5 (Kallipolitis and Valentin-Hansen 1998). Also DnaA, the DNA-binding protein involved in initiation of chromosomal DNA replication, takes part in the regulation of rpoH by repressing the P3 and P4 promoters (Wang and Kaguni 1989b) (Fig. 1, left panel). This complex regulatory network is responsive to various environmental signals and conditions, providing the proper amount of rpoH transcript and, accordingly, an extensive adaptability of the stress response. The hypothesis that the transient increase of cellular σ32 amount following heat stress occurs at the translational level has been proposed according to independent observations. First, the striking increase of the rate of synthesis of σ32 upon heat-shock does not correlate with a significant increase of rpoH transcription (Erickson et al. 1987). Second, to understand the contribution of transcriptional and/or post-transcriptional mechanisms to the expression of σ32 , the activity of a reporter gene (lacZ encoding the β-galactosidase enzyme) was monitored using transcriptional and translational fusions with the 5 region of rpoH. Results from these analyses showed that temperature regulated increase of σ32 was not dependent on the rpoH promoter, but was controlled by the translational signals of RpoH mRNA, suggesting a mechanism of translational control for σ32 during heat-shock (Straus, Walter and Gross 1987). Several studies have identified the regions of rpoH mRNA important for proper σ32 regulation, suggesting a mechanism involving the mRNA secondary structure formed by the 5 coding sequence of rpoH close to the translation start codon and, thereby, affecting translation efficiency (Kamath-Loeb and Gross 1991; Nagai, Yuzawa and Yura 1991; Yuzawa et al. 1993). All these findings converged in the current model of temperature-mediated translational regulation of σ32 . Specifically, during normal growth Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 552 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 Figure 1. Transcriptional and translational control of σ32 expression. Left panel: schematic representation of the region upstream of the rpoH coding sequence (depicted by a gray block arrow) containing five promoters (indicated by bent arrows and named P1 and P3 to P6) recognized by different σ factors (indicated underneath each promoter name). Transcription from some of these promoters is also regulated by transcriptional regulators: the arrowhead denotes positive regulation, while hammerheads indicate negative regulation. Central panel: schematic representation of the RpoH mRNA secondary structure at physiological temperature. Transcript sequences of crucial importance for the efficient initiation of translation close to the ribosome-binding site and the AUG start codon are sequestered in double-stranded structures, which leads to inefficient ribosome binding (gray dashed circle indicates the portion of the transcript containing these translation initiation sequences). Right panel: enlarged view of the RpoH mRNA region involved in the initiation of translation. While during normal growth conditions ribosome binding is hindered by secondary structure formation, at high temperature the partial melting of the mRNA secondary structure enhances ribosome entry and translational initiation. Black lines indicate Watson–Crick base pairing. conditions, the rpoH mRNA folds in a complex secondary structure (schematically represented in Fig. 1, central panel) that hides sequences crucial for the efficient initiation of translation surrounding the ribosome-binding site and the AUG start codon. Conversely (Fig. 1, right panel), at high temperature the partial melting of the mRNA secondary structure enhances ribosome entry and translational initiation without the involvement of additional cellular components (Morita et al. 1999). rpoH represents the first described example of a class of temperature-sensing RNA sequences known as RNA thermometers (Kortmann and Narberhaus 2012), which provide a rapid post-transcriptional mechanism to control the synthesis of heat-shock proteins (discussed below). The temperature-mediated positive regulation of σ32 synthesis through structured mRNA combines with the complex control of the heat-shock σ factor activity and stability, mediated by the concerted action of chaperones and proteases. The finding that E. coli strains carrying mutations in the dnaK, dnaJ and grpE genes were characterized by enhanced expression of heat-shock proteins at low temperature and by a prolonged transcriptional response upon heat challenge suggested a role for the DnaK-DnaJ-GrpE machinery as a negative regulator of heat-shock response (Straus, Walter and Gross 1990). The key role of the DnaK system in the regulation of heat-shock gene expression was further characterized, and a direct physical ATP-dependent interaction between σ32 and DnaK, DnaJ and GrpE chaperones was demonstrated both in vitro and in vivo (Gamer, Bujard and Bukau 1992; Liberek et al. 1992). In addition, the σ32 functionality is modulated by another major heat-shock protein, the chaperonin GroEL, as its depletion determined enhanced transcription of heat-shock genes, while its overexpression established the opposite effect (Kanemori, Mori and Yura 1994; Guisbert et al. 2004; Patra et al. 2015). The activity of σ32 appears to be dependent on its direct interaction with free chaperones (Guisbert et al. 2004). When unfolded proteins accumulate in the cell, they bind to and sequester chaperones, allowing σ32 to dissociate as a stable protein. In other words, it is widely accepted that the degree of sequestration of the DnaKJE/GroESL chaperones through binding to misfolded proteins determines the level of transcription of heat-shock genes, linking in this way the transcriptional response to the cellular folding state of proteins (Tomoyasu et al. 1998). The effect of chaperones on σ32 stability is directly linked to specific cellular proteases involved in its degradation. In particular, it was shown that an E. coli strain depleted of the membrane-bound ATP-dependent FtsH metalloprotease (formerly named HflB) accumulated σ32 and induced the heat-shock response. Furthermore, the half-life of the highly unstable σ factor increased by a factor of up to 12-fold in mutants with reduced FtsH function (Herman et al. 1995). Moreover, in vitro assays with purified components confirmed the direct association and the FtsH-mediated processing of σ32 (Tomoyasu et al. 1995). Also, additional cellular proteases, such as the ATP-dependent protease HslVU (also known as ClpQY), seem to play a role in the heat-shock response by modulating in vivo the turnover of σ32 (Kanemori et al. 1997). All the experimental data converged to a model for the homeostatic control of σ32 concentration in the cell in response to environmental stimuli. Normally present in low amounts and highly unstable, upon temperature upshift, σ32 levels rapidly increase as a consequence of enhanced mRNA translation and transient stabilization (positive feedforward mechanism). This induction phase is followed by a recovery/shut-off phase, mediated by chaperones and proteases (known as negative feedback loops), in which the activity and the stability of σ32 rapidly decrease to a new steady-state level. Even though the role of the DnaK and GroE chaperone systems in modulating σ32 activity and in promoting proteasemediated degradation was well documented in vivo, some contradictory observations had for many years hampered its recapitulation in vitro and the detailed mechanism remained elusive. Specifically, several studies, focused on the identification and the characterization of mutants with altered homeostatic regulation (known as ‘dysregulation mutants’), led to the definition of a short region of σ32 (named homeostatic control region) that appeared to be important for both stability and activity of σ32 (Obrist and Narberhaus 2005; Yura et al. 2007). Escherichia coli cells expressing heat-shock σ factor with mutations in the homeostatic control region showed increased σ32 activity and stability, Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 553 Figure 2. Representation of the current model for activity and stability control of σ32 , depicting its interaction with chaperones, its transit to the inner membrane and its FtsH protease-mediated degradation. Gray arrowheads denote negative modulation. consistent with its inferred role in the feedback control mechanism (Obrist and Narberhaus 2005; Yura et al. 2007). Surprisingly, purified dysregulation mutant σ32 isoforms showed normal binding to chaperones in in vitro interaction assays and were not inhibited by chaperones when assayed by in vitro transcription experiments (Yura et al. 2007; Suzuki et al. 2012). These discrepancies suggested that the negative feedback control of σ32 mediated by DnaK and GroE chaperones must be more complex than previously believed and that there must be at least an additional step involving some other cellular components besides chaperone binding. This missing step was recently characterized by Lim et al. (2013). In particular, they showed that σ32 is associated to the inner membrane instead of being distributed throughout the cytoplasm, as previously thought. This membrane association is mediated by the signal recognition particle (SRP) in combination with the SRP receptor (SR). This cotranslational protein targeting machinery (SRP-SR), normally acting only on periplasmic and membrane proteins, specifically targets σ32 to the inner membrane, which is crucial for proper homeostatic control upon heat-shock. The homeostatic control region of σ32 is directly involved in this process. Using in vivo crosslinking assays, it was recently shown that the homeostatic control region of σ32 encompasses the binding site recognized by a subunit of SRP (Miyazaki et al. 2016). These latter findings allowed for revision of the model for activity and stability control of σ32 as shown in Fig. 2. It is worth noting that membrane localization of σ32 may allow the heat-shock regulator to monitor protein folding status both in the cytoplasm and in the membrane, linking cytoplasmic and membrane protein damage to the heatshock response. The regulation of gene expression through the use of alternative sigma factors guarantees a global transcriptional reprogramming of the bacterial transcriptome in response to changing environmental conditions. In this light, it is not surprising that σ32 has a broad regulon, controlling the transcription of hundreds of genes upon temperature increase. Several genomewide studies focusing on the characterization of the σ32 regulon monitored gene expression changes upon inducible overexpression of σ32 using microarrays (Zhao, Liu and Burgess 2005; Nonaka et al. 2006). Besides known genes encoding proteases and chaperones involved in protein homeostasis, these analyses allowed the identification of σ32 -dependent targets involved in other key cellular processes. These processes include protection of both DNA and RNA, enhancement of transcription and translation at high temperature, as well as genes coding for other transcriptional regulators, thus propagating the response. Moreover, the finding that several σ32 -dependent genes have important roles in membrane protection and functionality suggested for the first time an intimate link between σ32 regulation and membrane homeostasis (Nonaka et al. 2006). Combining chromatin immunoprecipitation with microarrays (ChIP-to-chip), Wade et al. (2006) identified in vivo several new σ32 promoters and compared their locations with the positions of σ70 promoters determined in a separate study. Notably, they demonstrated that the majority of σ32 promoters overlaps with σ70 promoters, with promoter elements typical of one σ factor interspersed within the specific elements of the other σ, and that they can be transcribed, both in vivo and in vitro, by the RNA polymerase associated with σ32 or σ70 (Wade et al. 2006). Even though these data confirmed previous observations on rrnBP1 and gapA promoters (Newlands et al. 1993; Charpentier and Branlant 1994), the functional overlap between σ32 and σ70 was found to be much more extensive than expected. This promoter organization allows basal σ70 -dependent expression of many genes under normal conditions of growth to be boosted by σ32 upon temperature challenge. Considering the huge reprogramming of E. coli transcriptome imposed by σ32 , a strategy that directly or indirectly targets this pleiotropic regulator might be exploited by any invading organism to manipulate host response. In fact, the first J-domain protein (i.e. a protein that contain a J-domain able to stimulate DnaK activity) encoded by a bacteriophage has recently been isolated (Perrody et al. 2012). Notably, this protein, named Rki, is expressed by the phage RB43 early after infection of E. coli and, by specifically interacting and interfering with the DnaK chaperone function, can stabilize σ32 , thereby promoting accumulation of heat-shock proteins. The authors proposed that σ32 stabilization by Rki could help phage growth, likely promoting middle- and late-gene transcription and/or facilitating the proper folding of phage proteins as a consequence of the GroESL chaperonin accumulation (Perrody et al. 2012). Escherichia coli heat-shock sigma-E (σE ) Besides σ32 , the master regulator of heat-shock response, in E. coli a second heat-shock regulon is governed by another Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 554 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 Figure 3. Schematic representation of the mechanism of regulation of σE activity. During normal growth conditions (A), σE is sequestered to the inner membrane by the concerted action of the antisigma factor RseA and the periplasmic co-antisigma factor RseB. As a consequence, genes belonging to the σE regulon are transcriptionally inactive. Upon stress (B), RseA is completely degraded by a cascade of regulated proteolytic events in which DegS and RseP membraneassociated proteases are primarily involved. Following the complete degradation of RseA by cytoplasmic proteases, σE is available for interaction with the RNA polymerase enzyme and the transcription of σEspecific promoters is rapidly activated. alternative sigma factor named σE (also known as σ24 ). This 24-kDa alternative σ factor, encoded by the leading gene of the tetracistronic rpoE-rseA-rseB-rseC operon, was initially isolated as the factor responsible for the transcription of rpoH P3 promoter (as reported in Fig. 1, left panel) with distinct promoter specificity from the known sigma factors (Mecsas et al. 1993). It is a member of the extracytoplasmic function subfamily of sigma factors that function as effector molecules responding to extracytoplasmic stimuli such as cell wall and membrane stress and the oxidation state (Missiakas and Raina 1998; Helmann 2002). Specifically, E. coli σE is heat inducible and responsive to stress in the extracytoplasmic compartment of the bacterial cell (like, for example, accumulation of misfolded proteins translocated across the cytoplasmic membrane into the periplasmic space) and its activity is governed by a complex mechanism based on regulated proteolysis (Mecsas et al. 1993). Under normal growth conditions, σE is tethered to the membrane and kept inactive by the antisigma factor RseA (Fig. 3A). RseB and RseC, co-expressed in the same operon together with σE and RseA, also take part in σE regulation. Specifically, while the mechanism of positive regulation exerted by the inner-membrane-associated RseC protein on σE activity still remains to be fully characterized (Missiakas et al. 1997), it was demonstrated that RseB is located in the periplasmic space and acts as a co-antisigma factor when a direct interaction with RseA is established (Cezairliyan and Sauer 2007). Upon stress, RseA is completely degraded by the sequential action of different proteases and σE is free to interact with the RNA polymerase core enzyme to direct transcription of a specific subset of promoters characterized by peculiar –10 and –35 consensus elements (Rhodius et al. 2006; Ades 2008) (Fig. 3B). In detail, upon stress signal perception, the first event of the regulatory cascade is the activation of DegS, a membrane-associated protease that cleaves the periplasmic domain of RseA only upon dissociation of RseB from the antisigma factor (Walsh et al. 2003; Grigorova et al. 2004; Lima et al. 2013). A second membrane-associated protease named RseP further digests RseA in its transmembrane region, releasing in the cytoplasm a truncated C-terminal portion of the antisigma factor still associated to σE (Kanehara, Ito and Akiyama 2002, 2003). The last step of the σE activation process involves other cellular proteases such as ClpP/X-A, Lon and/or HslUV that completely degrade the remaining undigested part of RseA, freeing σE in the cytoplasm (Chaba et al. 2007)(Fig. 3B). This membrane localization of σE , a strategy also employed to maintain other alternative σ factors in their inactive form when their activity is not necessary, differs from the membrane-localized state of σ32 described above, a mechanism used for the dynamic regulation of the principal E. coli regulator. The σE regulon consists of tens of genes, whose function is almost completely devoted to folding of proteins in the cell envelope and to biosynthesis/transport of lipopolysaccharide (LPS), a component of the outer membrane of Gram-negative bacteria (Dartigalongue, Missiakas and Raina 2001; Rhodius et al. 2006). In fact, in addition to several genes involved in LPS biosynthesis, some σE -regulated genes encode periplasm-localized proteins that act on misfolded polypeptides (such as, for example, the periplasmic protease HtrA/DegP), while other regulon members code for cytoplasmic proteins involved in coordinating environmental stimuli with expression of the σE regulon (Dartigalongue, Missiakas and Raina 2001; Rhodius et al. 2006). However, as in the case of σ32 described above, an extensive functional overlap has been observed between σE and the housekeeping σ70 . In fact, a substantial fraction of σE -transcribed promoters can also be bound by σ70 , demonstrating that in E. coli the wide overlap of alternative σ factors with σ70 is not limited to σ32 (Wade et al. 2006). Negative transcriptional regulation of heat-shock genes In addition to positive regulation by means of specialized σ factors, another strategy to control heat-shock genes transcription relies on dedicated repressors. Under normal laboratory growth conditions, these DNA-binding regulators bind specific operators and repress transcription of heat-shock genes, while upon heat stress, the transcription of these genes is rapidly induced. Promoters driving transcription of genes controlled by heat-shock repressors harbor –10 and –35 elements recognized and bound by RNA polymerase associated with the vegetative σ factor. The first indications of the existence of alternative mechanisms of transcriptional regulation that are different from heat-shock σ factors came from studies of Bacillus subtilis stress response. In this bacterium, it was initially thought that the σ28 containing RNA polymerase controlled the transcription of heatshock genes because of its overlapping promoter specificity with Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 555 E. coli σ32 -containing RNA polymerase (Briat, Gilman and Chamberlin 1985). Subsequent studies demonstrated that σ28 was dedicated to transcription of flagellar genes and to some genes coding for chemotaxis effectors and that different B. subtilis chaperone genes, groESL and dnaK among others, were preceded by vegetative promoter sequences (Li and Wang 1992; Wetzstein et al. 1992). Afterwards, Chang et al. (1994) demonstrated that transcription of groESL was dependent on the vegetative sigma factor, σ43 , indicating that other mechanisms of heat-shock gene regulation, different from positive strategies based on alternative σ factors, must exist. In the subsequent years, several heatshock repressor proteins were identified and their mechanisms of regulation via interaction with conserved DNA elements located in the promoters of heat-shock genes were characterized. A common feature of heat-shock repressors is, besides the negative regulation of target genes, that they repress their own promoter. This regulatory scheme, known as negative autoregulation, is widely found among transcriptional repressors (negative autoregulation has been observed for about 50% of E. coli repressors) and guarantees fast response kinetics upon signal perception (Alon 2007). Consequently, upon heat stress and subsequent rapid induction of the controlled promoters, repressor concentration rapidly increases and reaches a threshold amount necessary for the re-establishment of the repressed state. In this way, the induction phase that rapidly appears upon heat stress is immediately followed by a shut-off phase of transcription. Currently, it appears well established that negative transcriptional regulation through dedicated repressors constitutes an important and widespread mechanism employed by bacteria to control heat-shock gene expression, not necessarily as an alternative to positive regulation described above. HrcA repressor The observation that a conserved sequence element is embodied in the promoter regions controlling the transcription of class I heat-shock genes of B. subtilis (consisting of two operons, the bicistronic groES-groEL and the heptacistronic dnaK operon) paved the way to the identification and characterization of the HrcA repressor, one of the most widespread heat-shock negative regulators in bacteria. Specifically, an inverted repeat (IR) consisting of 9 bp separated by a 9-bp spacer was identified in the DNA region close to the transcription start site of dnaK and groESL genes (Schmidt et al. 1992; Wetzstein et al. 1992). The role of this IR as a negative cis-element in the control of chaperone expression was demonstrated by introducing mutational changes within the upstream, downstream or both halves of the IR preceding dnaK and observing increased expression at low temperature and a reduction in the stimulation of the operon after heat-shock (Zuber and Schumann 1994). Considering that the conserved IR was observed in many different bacterial species, always within promoters controlling chaperone genes, it was named CIRCE for controlling inverted repeat of chaperone expression. Even though the possibility was initially considered that the CIRCE sequence could act alone through the formation of a temperaturedependent secondary structure, the demonstration of involvement of a DNA-binding protein in CIRCE-dependent negative regulation followed soon after. In particular, the inactivation of orf39, the first gene of the B. subtilis dnaK operon, led to the upregulation of dnaK and groESL even at physiological growth temperature (Schulz, Tzschaschel and Schumann 1995). Moreover, the characterization of regulatory mutants constitutively overexpressing groE and dnaK at low temperature revealed that all mutations map within orf39. The finding that cell extracts of E. coli overproducing Orf39 were able to specifically retard DNA probes harboring a CIRCE sequence in electrophoretic mobility shift assays (EMSA) clearly suggested that Orf39 was the negative regulator of chaperones expression in B. subtilis, likely acting as a CIRCE-specific DNA-binding repressor (Yuan and Wong 1995b). Similar observations were made for Caulobacter crescentus, where disruption of an orf39 homologous gene led to the conclusion that it encoded a negative regulator and, thus, it was named HrcA for heat-shock regulation at CIRCE elements (Roberts et al. 1996). Direct interaction between HrcA protein and the CIRCE element was demonstrated in B. subtilis by showing that a recombinant purified HrcA was able to retard a labeled probe containing a CIRCE element in the EMSA assay (Mogk et al. 1997). Detailed biochemical studies aimed at the molecular level dissection of the HrcA-CIRCE interaction have been hampered in many cases by its instability and strong tendency to form insoluble aggregates. For this reason, HrcA binding to target promoters has been characterized in detail only in a limited number of cases. For example, in Helicobacter pylori the development of a dedicated expression and purification protocol allowed demonstration by DNase I footprinting that HrcA binds operator elements similar to the CIRCE sequences overlapping both groESL and dnaK core promoters (Roncarati et al. 2007a,b). On both target promoters, the protected regions cover about 30 bp, suggesting that a dimer of HrcA interacts with the DNA. A similar situation was also observed for Chlamydia trachomatis, where HrcA showed full protection of the CIRCE element and neighboring nucleotides on the dnaK promoter, covering about 40 bp of the DNA probe (Wilson and Tan 2004). In this latter microorganism, in vitro transcription assays of selected promoters carried out in the presence of different concentrations of purified HrcA demonstrated that it was able to specifically repress dnaK transcription in a concentration-dependent manner (Wilson and Tan 2002). The position of the CIRCE element is almost always in a region close to the transcription start site, overlapping the portion of the DNA contacted by the RNA polymerase during initiation of transcription. The mechanism of repression exerted by HrcA upon binding to the CIRCE element relies on the physical occupancy of the core promoter with a consequent impediment of RNA polymerase binding under physiological growth conditions (Fig. 4A). Consistent with this model, when the position of the CIRCE element of the B. subtilis groE promoter was shifted from the original position to + 4 further downstream to position + 25 with respect to the transcription start site, the negative regulatory effect mediated by this cis-acting element was abolished (Yuan and Wong 1995a). HrcA-mediated repression of heat-shock genes is generally responsive to different kinds of environmental stresses. In many cases, when bacterial cells are exposed to stressful conditions such as high temperature, high salt concentration or a condition that provokes accumulation of misfolded proteins in the cytoplasm, the transcription from HrcA-controlled promoters is rapidly derepressed (Schmidt, Hertel and Hammes 1999; Laport et al. 2004). Even though it is well established that CIRCE acts as a specific HrcA-binding site at the DNA level, it has also been observed that the IR affects the stability of the RNA transcript. Specifically, in Rhodobacter capsulatus, the secondary structure imposed by the CIRCE element on the groESL mRNA appears to differentially affect transcript stability at various temperatures (J¨ager, J¨ager and Klug 2004). To date, the crystal structure of the HrcA repressor of the hyperthermophile Thermotoga maritima provides the only detailed structural information for this heat-shock regulator. The 2.2 ˚A resolution crystal structure shows that TmHrcA forms a dimeric structure, and each monomer is composed of a winged helix-turn-helix (HTH) Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 556 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 Figure 4. Negative regulation of heat-shock gene transcription by HrcA, HspR and CtsR repressors. (A) Under physiological growth conditions, HrcA binds the CIRCE conserved IR and represses transcription by hindering RNA polymerase promoter binding (depicted by black arrowhead). GroES-GroEL chaperonin interacts with HrcA and stimulates its DNA-binding activity. This GroE-mediated feedback control of HrcA activity is negatively modulated by accumulation of unfolded proteins during stress. (B) HspR-mediated repression of heat-shock genes at normal growth temperature is due to binding of the regulator to conserved HAIR sequences close to the core promoter, thereby interfering with RNA polymerase binding (indicated by a black arrowhead). In S. coelicolor, DnaK chaperone interacts with HspR and positively modulates its DNA-binding activity. Upon direct interaction between HspR and DnaK, the repressor becomes competent for promoter binding, while DnaK acts as a corepressor of HspR, with no direct contact to DNA. The HspR-DnaK functional interaction is negatively modulated by the accumulation of misfolded proteins in the cytoplasm that titrate away the chaperone. (C) Current model for the mechanism of repression and for the heat-dependent regulation of CtsR repressor activity. During control conditions, CtsR binds to conserved IR sequences and represses transcription of heat-shock genes (depicted by a black arrowhead). Following a sudden increase of temperature, CtsR undergoes a heat-induced structural alteration and loses affinity for DNA (indicated by a gray dashed line). During environmental stress conditions, upon temperature-dependent autophosphorylation, McsB becomes active and, assisted by its co-activator McsA, functions as an adaptor to target CtsR to ClpP/C-mediated proteolytic degradation. Upon dephosphorylation by the cognate phosphatase YwlE, McsB becomes deactivated. N-terminal DBD, a central GAF-like domain (probably involved in dimerization rather than in cGMP binding) and a C-terminal inserted dimerizing domain. However, the conformation of both DBDs in the dimer suggests that the TmHrcA is incompatible with DNA binding and may represent an inactive form of this protein (Liu et al. 2005). The detailed structure of the DBD, combined with results of mutagenesis studies of conserved residues (Hitomi et al. 2003; Wiegert and Schumann 2003), suggests that HrcA binding to DNA should involve the interaction of the recognition helix of the HTH domain with the DNA major groove and the interaction of a loop connecting two α-helices with the DNA backbone. It is worth noting that the amino acid sequence similarity of HrcA proteins from different species is surprisingly low. The overall sequence identity is often <30% with the exception of HrcA proteins from related organisms (Narberhaus 1999). Hence, the molecular features derived from the solved structure described above may be different among the various HrcA regulators. In order to identify additional genes regulated by HrcA besides the typical groE and dnaK genes, genome-wide expression studies have been combined with the search of conserved CIRCE elements and with in vitro DNA-binding studies. What generally emerges from such approaches is that HrcA is involved, directly or indirectly, in the regulation of several genes linked to various cellular processes. For example, a microarray-based transcriptome analysis performed in Listeria monocytogenes revealed that HrcA regulates, albeit indirectly, genes involved in stress response, metabolism, translation and DNA replication (Yuewei et al. 2007). Similar pleiotropic effects of hrcA mutation were observed with analogous approaches in other microorganisms, including H. pylori and Lactobacillus plantarum (Roncarati et al. 2007a; Van Bokhorst-van de Veen et al. 2013). However, the number of genes directly regulated by HrcA appears to be restricted and limited to genes involved in stress response. For example, in the human pathogen Mycoplasma genitalium, the genes encoding the stress proteases Lon and ClpB possess a conserved CIRCE element within their promoter regions, suggesting a HrcA-mediated negative regulation (Musatovova, Dhandayuthapani and Baseman 2006). Moreover, various lines of evidence indicate that the expression of clp genes in several Lactobacillus species is regulated by HrcA, instead of CtsR, the conserved clp gene regulator of low G+C Gram-positive bacteria (Suokko et al. 2005, 2008), while in Streptococcus salivarus these two heat-shock repressors are cooperative in controlling clpP gene transcription (Chastanet and Msadek 2003). HspR repressor The heat-shock transcriptional repressor HspR was first discovered and characterized in the Streptomyces genus. Specifically, it was shown that, in Streptomyces coelicolor, the distal gene of Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 557 the heat-inducible dnaK operon encoded a novel heat-shock protein, similar to the GlnR repressors of Bacillus spp. and to the MerR family of transcriptional regulators. In vitro DNA-binding assays were carried out to demonstrate that HspR was able to bind to three inverted repeat sequences (IR1, IR2 and IR3) in the dnaK promoter region, centred at positions –75, –49 and +4 with respect to the transcriptional start site. This strongly suggests a direct role for HspR in heat-shock gene regulation (Bucca et al. 1995). Genetic and biochemical evidence of HspR regulatory function followed soon after. In particular, it was shown that disruption of the hspR gene led to high-level constitutive transcription of dnaK operon. Moreover, the addition of an E. coli purified recombinant his-tagged HspR to in vitro transcription assays clearly demonstrated repression of dnaK transcription, providing additional evidence for direct involvement of HspR in dnaK regulation (Bucca, Hindle and Smith 1997). Similar observations were made in the same year in S. albus (Grandvalet, Servant and Mazodier 1997), where HspR was shown to bind an IR identical to IR3 of S. coelicolor in the promoter region of the protease gene clpB (Grandvalet, de Cr´ecy-Lagard and Mazodier 1999). Genes similar to hspR were observed in several other bacterial species, and HspR-binding sites were detected upstream of various heatshock genes of these microorganisms. The HspR-binding site was named HAIR (for HspR Associated Inverted Repeats) and a consensus sequence was proposed (Grandvalet, de Cr´ecy-Lagard and Mazodier 1999). So far, HspR repressors have been found and characterized, to various extents, in several bacterial species and it is now well established that this represents, in addition to the widespread HrcA repressor, a widely employed system to negatively control heat-shock genes transcription. The HAIR sequences have been found, in most cases, in the vicinity of the core promoter region, often overlapping the –10 and/or –35 boxes (Bucca et al. 1995; Grandvalet, de Cr´ecy-Lagard and Mazodier 1999; Stewart et al. 2002). This observation suggests that HspR prevents RNA polymerase binding at physiological temperature by a steric hindrance mechanism at the promoter (Fig. 4B). In some cases, however, HspR-HAIR interaction takes place far upstream of the core promoter region, in an atypical position for a transcriptional repressor (Spohn et al. 2004; Schmid et al. 2005). The long distance between the HAIR sequence and the core promoter –10/–35 boxes supports the hypothesis of a complex mechanism for HspR-mediated transcriptional repression, which encompasses more than simple steric occlusion of the promoter. In this respect, dual repression of some heat-shock genes mediated by HspR and HrcA in H. pylori represents a clear example of this mechanism. The distal binding sites of HspR on groESL and dnaK promoters (between positions –43/–120 and –78/–149 with respect to the transcriptional start site, respectively) combine with proximal HrcA/CIRCE interactions, occluding core promoter regions (Roncarati et al. 2007a, discussed in detail below). Another intriguing example of complex HspRmediated regulation has been observed in Mycobacterium tuberculosis. In this intracellular human pathogen, the acr2 gene, encoding a member of the widespread α-crystallin family of molecular chaperones, is transcriptionally controlled by HspR in combination with the response regulator PhoP (Singh et al. 2014). In this case, the heat-shock repressor HspR binds to the HAIR sequence located in the core promoter region (Stewart et al. 2002), while PhoP-binding region maps downstream from the transcriptional start site. Moreover, a direct HspR-PhoP interaction has been demonstrated, as well as the crucial importance of the simultaneous presence of both regulators for heat-shock-dependent in vivo regulation of acr2 (Singh et al. 2014). Accordingly, the current model for acr2 regulation postulates that under normal growth condition both the PhoP and HspR, bound to their target sites, form a higher-order DNA-protein structure that prevents RNA polymerase binding and the direct HspR-PhoP interaction provides additional stability to this complex. Initially identified as the regulator of major chaperone genes such as dnaK in Streptomycetes spp., the identification of additional genes controlled by HspR in several different bacterial species was enabled by the advent of microarray techniques. In S. coelicolor, DNA microarray-based analysis of gene expression in wild-type and hspR-disruption mutant strains led to the identification of 17 genes controlled, directly or indirectly, by HspR (Bucca et al. 2003). This analysis was refined in a subsequent work by the same group, by implementation of a high-density microarray (Bucca et al. 2009). Specifically, gene expression analysis was combined with immunoprecipitation of in vivo HspRbinding sites and hybridization of the recovered DNA fragments on a DNA microarray chip (ChIP-on-chip). From this analysis, it emerged that, in S. coelicolor, HspR directly represses transcription of genes encoding chaperones (dnaK-grpE-dnaJ-hspR operon) and heat-shock proteases (lon and clpB), as well as one ribosomal RNA gene and two tRNA genes, thus broadening, for the first time, the regulatory role of HspR in this microorganism. In H. pylori, to investigate if the regulatory function of HspR is restricted to promoters of chaperone genes or if it is directly involved in controlling additional genes, an in vitro selection of genomic DNA fragments bound by purified HspR protein was developed (Delany et al. 2002). Besides the three previously characterized binding sites in the promoter regions of the multicistronic operons coding for the components of the DnaK and GroE machineries (Spohn and Scarlato 1999), two novel binding sites were identified, located in the 3 region of both speA and tlpB genes, which encode an arginine decarboxylase and a methylaccepting chemotaxis protein, respectively. Moreover, sequence alignment of novel HspR-binding sites highlighted HAIR-like sequences as well as conserved nucleotides extending outside the previously proposed consensus binding sequence. However, parallel macroarray hybridization of cDNA probes deriving from H. pylori wild-type and hspR-mutant strains failed to observe deregulation of these novel target genes, suggesting the existence of a minority of non-canonical binding sites, apparently not associated with regulatory functions (Roncarati et al. 2007a). In the last 15 years, the identification of members of the HspR regulon was also pursued through array-based whole transcriptome analyses in several other unrelated bacterial species including Campylobacter jejuni, M. tubercolosis and Deinococcus radiodurans (Stewart et al. 2002; Andersen et al. 2005; Schmid et al. 2005; Holmes, Penn and Lund 2010). Overall, these studies suggested that HspR directly represses, alone or in combination with other transcriptional regulators, the transcription of a limited set of genes that encode the major cellular chaperones and heat-shock proteases. However, indirect effects of hspR disruption affecting transcript abundance of genes involved in diverse cellular processes not strictly related to heat-shock have been observed. For example, in H. pylori and in the closely related C. jejuni, hspR deletion led to substantially lower expression of many genes coding for proteins involved in cell motility, and accordingly, hspR-mutant strain showed reduced motility (Andersen et al. 2005; Roncarati et al. 2007a). One hypothesis for the intersection of HspR-regulated heat-shock response and flagellar assembly may be that accumulation of chaperone proteins in the mutant strain alters the assembly of the flagellar apparatus and/or increases the activity of dedicated flagellar transcriptional regulators, which in turn establishes negative feedback for the programmed transcription of motility genes. Another Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 558 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 interesting example of HspR-mediated interplay between the heat-shock response and an unrelated central cellular process was observed in Bifidobacterium breve UCC2003 (Zomer et al. 2009). In this microorganism, an SOS response is induced by oxidative stress as well as by heat-shock. Repression of DNA damage repair genes is mediated by the LexA regulator, whose selfcleavage and consequent activation is promoted by the RecA protein. Induction of RecA expression is probably under the control of HspR under heat-shock conditions, an observation that directly links the repressor to a process employed by Bifidobacterium to respond to an excess of DNA damage (Zomer and van Sinderen 2010). CtsR repressor The characterization of B. subtilis heat-shock response revealed the existence of a subgroup of stress-responsive genes whose transcription is governed by two alternative promoters: one recognized by the RNA polymerase complexed with the vegetative sigma factor σA and the other is dependent on the general stress sigma factor σB . The transcription from both the vegetative σA and σB -dependent promoters of this class III subgroup of genes (consisting of trxA, clpC and clpP operons) was induced by various stress signals, even though the induction pattern differed with respect to the particular gene and stress conditions (Kr ¨uger, Msadek and Hecker 1996; Gerth et al. 1998; Scharf et al. 1998). The observation that vegetative promoters of the above-mentioned genes remained stress inducible even on a σB -mutant background, together with the lack of a CIRCE consensus sequence typical of B. subtilis HrcA-regulated class I genes, prompted scientists to search for a novel heat-shock repressor responsible for this regulation. The first gene of the clpC operon, named orf1, was shown to encode a product with a predicted HTH DBD, suggesting a regulatory role for the protein (Kr ¨uger et al. 1997). This hypothesis was confirmed by the observation that transcript amounts of the class III clpC and clpP genes were significantly upregulated in a orf1-mutant strain compared to the wild-type under normal growth conditions, while class I (dnaK) and class II (ctc) genes were similarly expressed between the two strains (Kr ¨uger and Hecker 1998). The experimental confirmation of Orf1 being a direct transcriptional repressor came from EMSA assays carried out on a radiolabeled clpC operon promoter (Kr ¨uger and Hecker 1998). Subsequent studies by Derr´e, Rapoport and Msadek (1999) expanded the analysis of targets directly bound by the repressor, thereafter renamed CtsR for Class three stress gene repressor. Specifically, it was shown that CtsR binds a directly repeated heptanucleotide sequence positioned in close proximity of the transcription start site of clpP and clpC operons (Derr´e, Rapoport and Msadek 1999). Moreover, CtsR binding was shown to take place on five distinct binding sites nearby the clpE transcription start point, also postulating the existence of complex promoter architectures (Derr´e et al. 1999). Both CtsR homologs and target sequences were found upstream of clp and other heat-shock genes of several Gram-positive bacteria, an observation that suggested CtsR as a widespread transcriptional repressor involved in heat-shock regulation. In this respect, it was subsequently shown that, in L. monocytogenes, CtsR regulation is strikingly similar to that in B. subtilis, describing for the first time a stress response regulatory gene in a human pathogen (Nair et al. 2000). Specifically, it was demonstrated that L. monocytogenes CtsR represses the transcription of clpP, clpC, clpE and clpB genes by directly binding to their respective vegetative promoters (Nair et al. 2000; Chastanet et al. 2004). Considering that the positions of the binding sites characterized so far on several promoters of various bacterial species overlap the core promoter region, CtsR should exert its repressive function by sterically occluding RNA polymerase binding to DNA (Fig. 4C). The first indications that the prototypical B. subtilis CtsR could bind DNA as a dimer, evinced from analysis of the binding sites and from some biochemical studies (Derr´e et al. 2000), were further confirmed by the determination of the crystal structure of the repressor bound to a 26 bp DNA derivative of the clpC promoter (Fuhrmann et al. 2009). The crystal structure revealed that each protomer is characterized by an N-terminal winged HTH DBD and a C-terminal dimerization domain composed by four α-helices organized as a four-helix bundle. DNA recognition is based on the insertion of the HTH domain in the major groove and a concomitant interaction of the extended β-hairpin wing with the minor groove. Some crucial residues, located within the HTH domain and involved in DNA binding, were identified and found to be highly conserved in the CtsR protein family (Fuhrmann et al. 2009). The CtsR repressor is historically considered the master regulator of the cellular protein quality control genes of low-GC Gram-positive bacteria. This definition is based on the observation that CtsR typically controls the expression of genes encoding stress proteases mainly involved in degradation of misfolded/denatured polypeptides as well as of deleterious protein aggregates. For example, in the model organism B. subtilis, but also in the human pathogen L. monocytogenes and in the probiotic bacterium Lactococcus lactis, CtsR essentially controls the transcription of several components of the Clp machinery (Derr´e et al. 1999; Nair et al. 2000; Varmanen, Ingmer and Vogensen 2000; Chastanet et al. 2004), while the regulation of the classical molecular chaperones, such as DnaK and GroE, is HrcA dependent. In some cases, however, CtsR is also able to regulate the expression of HrcA-dependent genes. In Staphylococcus aureus, for example, dnaK and groE operons are coregulated by both CtsR and HrcA. These regulators contact adjacent binding sites and act synergistically to maintain low basal levels of expression of both operons in the absence of stress (Chastanet, Fert and Msadek 2003). In some other instances, as in Oenococcus oeni, CtsR solely regulates clp machinery genes and also GroE/DnaK molecular chaperones. This lactic acid bacterium lacks hrcA in its genome and CtsR has taken over the regulation of typical HrcA targets (Grandvalet et al. 2005). Interestingly, some studies have revealed new non-canonical genes regulated by CtsR. These include genes encoding small heat-shock proteins, such as Hsp18 in O. oeni and Hsp1 in Lactobacillus plantarum, the membranebound FtsH protease in L. plantarum and the tatAC genes that code for two minimal translocases responsible for the twin arginine translocation (Tat) pathway in L. monocytogenes (Grandvalet et al. 2005; Hu et al. 2007; Fiocco et al. 2009, 2010). The CtsR master regulator binds the promoters of target genes and represses their transcription under physiological growth conditions. Intriguingly, upon heat-shock, CtsR loses DNA-binding activity and the non-functional protein undergoes regulated proteolytic degradation mediated by the ClpCP protease (Elsholz et al. 2010) (Fig. 4C). The first indications that CtsR is a specific target of ClpPC under stress conditions came from in vivo stability assays performed in B. subtilis wild-type and in clpP or clpC mutant strains (Kr ¨uger et al. 2001). Specifically, Kr ¨uger and colleagues demonstrated that, after inhibiting translation with tetracycline, CtsR was much more labile in the wild-type strain, compared to mutant strains. It also became clear that the products of two genes belonging to the CtsR operon, McsA and McsB, play a crucial role in the proteolytic turnover of the repressor. In particular, mcsB encodes a heat-activated kinase that targets nonfunctional heat-inactivated CtsR, while mcsA codes for an activator of McsB Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 559 (Kirstein et al. 2005, 2007). The current model for CtsR proteolytic degradation, represented in Fig. 4C, postulates that, under normal growth conditions, McsB is kept inactive by interaction with ClpC. Upon heat exposure, McsB is titrated away from ClpC and becomes activated by temperature-dependent autophosphorylation and by interaction with McsA activator. Following this sophisticated dual activation step, McsB acts as an adaptor, targeting non-functional CtsR species and efficiently driving them to degradation. Active McsB adaptor is downregulated by either dephosphorylation by the cognate phosphatase YwlE or by rapid degradation of the phosphorylated form (Elsholz et al. 2010). RheA repressor RheA represents an example of heat-shock transcriptional repressor with a restricted distribution among bacteria, being found only in S. albus. The first indications that S. albus possessed an additional heat-shock transcriptional regulator, in addition to HrcA and HspR, came from the analysis of a DNA region upstream of the hsp18 gene, encoding a heat-inducible protein belonging to the small heat-shock protein family. This open reading frame, named orfY and coding for a 225-amino acid polypeptide, is transcribed from a vegetative promoter in the opposite orientation to hsp18. Moreover, disruption of orfY led to the accumulation of hsp18 transcript, even at low temperature, suggesting a direct or indirect involvement of OrfY in the transcriptional regulation of hsp18 (Servant and Mazodier 1996). Additional biochemical and genetic studies provided evidence that OrfY was the direct repressor of hsp18. Thereafter, OrfY was named RheA for repressor of heat-shock protein eighteen. In particular, expression of S. albus RheA in E. coli determined repression of transcription from the hsp18 promoter fused to lacZ as reporter gene. Moreover, a DNA probe encompassing the hsp18 promoter was specifically retarded by a crude cell extract containing RheA in EMSA assays (Servant, Rapoport and Mazodier 1999). Considering that the transcription start sites of the divergent hsp18 and rheA genes are separated only by 25 nucleotides and that the –10 boxes of the two promoters partially overlap, it was not surprising to discover that RheA also negatively autoregulates its own transcription (Servant, Rapoport and Mazodier 1999). The RheA repressor binds two IRs (a perfect IR TGTCATCN5-GATGACA, flanked by an imperfect IR GTCATC-N5-GACGAC) located in the intergenic region of the divergent rheA and hsp18 core promoters, thereby occluding the access to RNA polymerase and repressing both genes’ transcription under normal growth conditions (Servant, Grandvalet and Mazodier 2000). RheA represents the first described transcriptional repressor of small heatshock proteins, anticipating the similar scenario confirmed in O. oeni and Clostridium acetobutylicum, where hsp18 is transcriptionally repressed by CtsR (Derr´e, Rapoport and Msadek 1999). Feedback regulatory circuits modulate the activity of heat-shock repressors Similar to the complex homeostatic control of σ32 in E. coli described above, feedback regulatory circuits of heat-shock transcriptional repressor activity mediated by chaperones have been observed in several bacterial species. Soon after the identification and characterization of the HrcA repressor as a negative regulator of class I heat-shock genes in B. subtilis, experimental evidence pointed out the role of the GroE chaperonin as a modulator of HrcA regulon (Mogk et al. 1997). In particular, depletion of the intracellular level of GroESL was associated with high expression of the HrcA-controlled dnaK operon at all temperatures. By contrast, GroESL overexpression led to the hyper-repression of transcription from the target promoter. Moreover, EMSA assays showed that the affinity of HrcA binding to a DNA probe containing a CIRCE operator is increased in the presence of GroESL. It is worth noting that the shifted fragment migrated to the same position in the absence or presence of GroEL, indicating a role for the chaperonin in the modulation of HrcA DNA-binding activity rather than functioning as a corepressor (Mogk et al. 1997). Following the formal demonstration of the direct interaction between HrcA and GroE in vitro (Reischl, Wiegert and Schumann 2002), a model was proposed for their functional interplay (Fig. 4A). This model, known as the ‘titration model’, postulates that HrcA is kept in an active conformation by GroESL, competent for binding to CIRCE elements. Upon stress insult, the accumulation of misfolded polypeptides in the cell sequesters and titrates away the chaperonin, which can no longer interact with the HrcA repressor. Without the presence of bound chaperonin, HrcA loses DNA-binding affinity and repression of class I heat-shock genes is relieved. This model is also supported by the finding that addition of ethanol, treatment with puromycin (that causes accumulation of truncated polypeptides in the cytoplasm) and overexpression of substrates of GroESL derepress the HrcA regulon (Mogk et al. 1998). The same feedback mechanism operates in C. trachomatis (Wilson et al. 2005), where the interaction between HrcA and GroE was investigated in more detail. It is worth noting that C. trachomatis harbors in its genome three genes coding for GroEL homologs: groEL1, groEL2 and groEL3. These genes are expressed constitutively throughout the developmental cycle of the bacterium (Karunakaran et al. 2003) and differentially regulated in response to heat-shock. In fact, while transcription of groEL1 is negatively regulated by the HrcA-CIRCE system and induced upon temperature increase, transcription of groEL2 and groEL3 is not responsive to heat-shock nor to HrcA regulation (Karunakaran et al. 2003; Hanson and Tan 2015). In this system, the characteristic additional C-terminal tail of chlamydial HrcA interfered with repressor binding to the CIRCE element in vitro and also with HrcAmediated transcriptional repression in vitro and in vivo (Chen, Wilson and Tan 2011). Intriguingly, the negative effect exerted by the HrcA C-terminal inhibitory region could be counteracted by GroE. Specifically, it was shown that recombinant GroEL was able to enhance HrcA activity (HrcA-binding activity to CIRCE elements in vitro as well as HrcA-mediated repression in vivo) and that this effect was more pronounced on the full-length HrcA rather than on the truncated version (without the C-terminal tail) of the repressor. Intriguingly, the positive effect of recombinant GroEL on HrcA-binding activity to its operator was shown to be ATP independent, suggesting a non-canonical mechanism of action of the chaperonin (Chen, Wilson and Tan 2011). The characterization of the GroESL-mediated feedback control of HrcA repressor activity in several distant bacterial species, like for example in C. crescentus and H. pylori (Susin et al. 2004; Roncarati et al. 2007a), suggests that it may represent a general mechanism to post-transcriptionally control repressor functionality in response to environmental stimuli. The hypothesis that homeostatic control of transcriptional regulators by chaperones might represent a novel common theme in heat-shock gene regulation is confirmed by the characterization of the feedback control of the heat-shock master repressor HspR by DnaK. The DnaK-HspR functional interaction was initially characterized in S. coelicolor. It was shown that the in vitro formation of a stable complex between HspR and its DNA target was DnaK dependent, suggesting that DnaK stimulates HspR DNA-binding activity (Bucca et al. 2000). Surprisingly, the Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 560 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 DnaK-mediated modulation of HspR was shown to be independent from co-chaperones DnaJ and GrpE and required no addition of ATP, thereby excluding a mechanism based on DnaK chaperoning function. In vitro EMSA assays showed the formation of a ternary complex composed of DnaK interacting with HspR bound to DNA. Thus, S. coelicolor DnaK acts as a corepressor of HspR, with no direct contact with DNA (Bucca et al. 2000). This role of DnaK in heat-shock gene regulation was further confirmed both in vitro and in vivo. In particular, it was shown that DnaK was able to stimulate HspR-mediated repression in in vitro transcription assays (Bucca et al. 2000). Moreover, in vivo depletion of cellular DnaK led to high-level expression of HspRrepressed promoters at normal growth temperature (Bucca et al. 2003). All these findings converged in a model of DnaK feedback control of HspR activity that resembles, with some differences, the titration model valid for GroE-HrcA (Fig. 4B). That is, under normal environmental conditions DnaK interacts with HspR and the ternary complex binds to target DNA, resulting in tight transcriptional repression. During heat-shock, DnaK is sequestered by unfolded proteins that accumulate in the cytoplasm and HspR is less efficiently able to repress the transcription of target genes whose expression results induced. More recently, a similar DnaK-mediated feedback loop controlling HspR activity has been investigated and characterized in M. tubercolosis, although with some important differences. In this bacterium, DnaK interacts directly with the HspR-HAIR complex and stimulates repressor DNA-binding activity in an ATP-independent manner (Parijat and Batra 2015). However, in this pathogen HspR modulation does not depend solely on DnaK as in S. coelicolor. Evidence supports an important role played also by the GroE chaperonins GroEL1 and GroEL2 in the HspR activation process (Das Gupta, Bandyopadhyay and Das Gupta 2008; Bandyopadhyay et al. 2012). In the context of feedback regulation of heat-shock repressors, H. pylori represents a peculiar example in which two chaperones modulate the activity of two distinct transcriptional regulators. As described above, H. pylori uses both HrcA and HspR repressors for the regulation of heat-shock response. However, while HrcA activity is feedback regulated by the GroESL chaperonin according to the typical ‘titration model’ proposed for B. subtilis, the DNA-binding activity of HspR is negatively modulated by the heat-shock protein CbpA, rather than by DnaK (Fig. 5). In analogy with the homologous protein of E. coli (Chae et al. 2004; Bird et al. 2006), the HspR-repressed cbpA gene is thought to encode a dual-function protein, working both as a DnaJ-like co-chaperone of DnaK and as a nucleoid-associated protein involved in nucleoid structuring function. In order to characterize a possible post-transcriptional control over HspR by the DnaK chaperone system, surprisingly it was found that, upon direct protein–protein interaction, CbpA alone was able to negatively modulate HspR binding to target promoters in vitro without contacting the DNA, but only when the repressor was not bound to its operators. In addition, overexpression of CbpA led to deregulation of heat-shock response in vivo (Roncarati, Danielli and Scarlato 2011). These findings suggest important considerations and add new perspectives in heat-shock gene regulation. First of all, the effect of CbpA on HspR DNA binding (negative effect) is opposite to the modulation typically exerted by DnaK or GroE on HspR and HrcA (positive effect), implying that the accepted model for chaperone feedback regulation of transcriptional repressors is insufficient to explain CbpA-HspR functional interaction in H. pylori. A possible explanation is that CbpA regulation of HspR-binding activity is adopted by H. pylori to fine-tune the regulatory network shutoff response governing heat-shock response (Roncarati, Danielli and Scarlato 2011). Second, considering the putative functions of CbpA as both a co-chaperone and a nucleoid-associated protein (preliminary results suggest that CbpA of H. pylori possesses both functions; D. Roncarati, personal communication), it would be informative to characterize the functional interplay between HspR and CbpA. In other words, an intriguing hypothesis is that the heat-shock regulator HspR itself, aside from its role in the repression of transcription regulated by CbpA, might influence the co-chaperone and/or nucleoid-associated activity of CbpA by direct protein– protein interaction (Fig. 5). This would be a novel example in which heat-shock gene regulation intersects with distinct cellular functions, as for example the maintenance and regulation of the bacterial nucleoid. Cross-comparison of heat-shock transcriptional repressors Experimental data available so far allow to highlight similarities as well as differences among the major heat-shock transcriptional repressors described in the previous sections. Starting from the mechanism of regulation, they all repress Figure 5. Model for HrcA and HspR-dependent heat-shock gene regulation in H. pylori. Two dedicated transcriptional repressors, HrcA and HspR, directly repress three multicistronic operons containing the major chaperone genes of H. pylori. Specifically, the cbp operon encodes a homolog of the E. coli curved DNA-binding protein A (CbpA), the heat-shock transcriptional repressor HspR and a protein with putative helicase function; the dnaK operon encodes the heat-shock transcriptional repressor HrcA, the co-chaperone GrpE and the DnaK chaperone; the gro operon encodes the chaperonin system GroES-GroEL. Positive regulation of HrcA activity by the GroESL chaperonin in depicted by a gray dashed arrow, while negative modulation of HspR by CbpA is highlighted by a grey dashed hammerhead. The hypothetical negative effect of HspR on CbpA activity is indicated by a grey dashed hammerhead marked with a question mark. Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 561 transcription, under normal growth conditions, by sterically occluding RNA polymerase binding to target promoters. In almost all cases, in fact, these repressors interact with the target promoters by binding conserved sequences that overlap core promoter elements, such as –10 and –35 boxes. A closer inspection of their DNA-binding architecture reveals some differences. While HrcA-regulated promoters typically contain a single CIRCE IR and are bound by a dimeric form of the repressor, other promoters are controlled in a more complex manner by HspR and CtsR repressors. For instance, B. subtilis clpE promoter, but also S. coelicolor and H. pylori dnaK promoters harbor multiple repressors’ binding sites spread all over the 5 control region, suggesting the existence of complex mechanisms of transcriptional regulation. Another common feature shared by all heatshock repressors is their negative autoregulation. Negative autoregulation is a recurrent regulatory pattern, also called network motif, found in ∼40% of the known transcriptional regulators in E. coli (Thieffry et al. 1998). This network motif has been suggested to increase the homeostasis of the autoregulated gene product in the context of stochastic gene expression noise, to speed up the response kinetics upon signal perception and to provide more linear dose responses (Becskei and Serrano 2000; Rosenfeld, Elowitz and Alon 2002; Nevozhay et al. 2009). In the context of heat-shock gene regulation, negative autoregulation ensures tight control of steady-state repressors amount under normal environmental conditions, followed by a prompt induction upon heat-shock of regulators’ gene transcription. Many more peculiarities emerge when considering the posttranscriptional regulation of repressors activity and stability/turnover mediated by chaperones and proteases, respectively. As shown in Fig. 4, while CtsR and RheA activity does not appear to be feedback modulated by molecular chaperones, both HrcA and HspR interplay with distinct and specific chaperone systems, whose actions add an additional level of regulation. On the other hand, the repressors’ turnover that takes place following heat stress has been deeply characterized for CtsR (shown in Fig. 4C), while detailed information for the other heat-shock repressors is still missing. In this respect, Jastrab et al. (2015) have recently identified a novel pathway responsible for the degradation of the heat-shock repressor HspR in M. tubercolosis. Specifically, they showed that the heat-shock repressor is actively degraded by the prokaryotic proteasome encoded by this human pathogen, and that degradation is enhanced by an ATP-independent proteasome activator, called PafE, and by HspR denaturation. Combinations of heat-shock transcriptional regulators The transcriptional regulation of heat-shock genes displays varying degrees of complexity among various microorganisms, reflecting the extreme diversity of genetic regulatory mechanisms in bacteria. In several bacteria, heat-shock transcriptional regulation is managed exclusively by a dedicated σ factor that enables the RNA polymerase enzyme to transcribe specific genes important for the response to temperature stress (Fig. 6A, left panel, and Table 1). In such simple cases, all the thermoresponsive genes are under the control of a single regulator. For example, in Pseudomonas aeruginosa and in Vibrio cholerae, transcriptional regulation of heat-shock genes is governed by an alternative sigma factor specific for heat stress, similar in sequence and function to E. coli σ32 (Allan et al. 1988; Slamti, Livny and Waldor 2007). As described above, the model organism E. coli also adopts exclusively positive regulators for the heat-shock response. Here, two alternative σ factors (σ32 and σE ) are involved in heat-shock regulation, with σ32 playing a major role in sensing cytoplasmic and inner-membrane stimuli, while σE is committed to extracytoplasmic (envelope) stress response (Fig. 6A, left panel, and Table 1). In other bacterial species, the coexistence of positive and negative control mechanisms, together regulating the expression of distinct sets of heat-shock genes, has been observed. For example, in the Gram-positive pathogen L. monocytogenes, there are three groups of heat-shock genes, each one controlled by a different strategy (Fig. 6A, central panel). In detail, genes belonging to class I (encoding members of the major chaperone systems DnaK and GroE) and to class III (encoding chaperones and ATP-dependent Clp proteases) are negatively regulated by HrcA and by the CtsR heat-shock repressor, respectively. In contrast, genes belonging to class II (coding for general stress proteins) are positively controlled by the alternative sigma factor σB (Table 1) (van der Veen et al. 2007). A similar situation, in which positive and negative mechanisms are both employed to control separate sets of genes, has been described in other bacteria such as B. subtilis and Agrobacterium tumefaciens (Table 1) (Nakahigashi et al. 1999; Schumann 2003). Intriguingly, in cyanobacteria, both positive and negative strategies of transcriptional regulation combine to control expression of some heat-shock genes (reviewed by Rajaram, Chaurasia and Apte 2014). For instance, in Synechocystis PCC6803, the groESL operon is negatively regulated by the HrcA/CIRCE system and positively regulated by the alternative sigma factors SigE and SigB (Nakamoto, Suzuki and Kojima 2003; Singh et al. 2006; Kojima and Nakamoto 2007). Furthermore, the histidine kinase Hik34 is involved in the negative regulation of the chaperonin gene (Slabas et al. 2006; ˇCerveny et al. 2015). In addition to transcriptional mechanisms, RNAbased regulation of heat-shock genes has also been reported in cyanobacteria. Expression of the hsp17 gene, coding for a small heat-shock protein also known as HspA, is transcriptionally dependent on the alternative sigma factors SigE and SigB (Singh et al. 2006; Tuominen et al. 2006, 2008), and post-transcriptionally controlled by a temperature-sensitive RNA structure in the 5 -UTR portion of the gene (Table 1) (Kortmann et al. 2011). The regulatory schemes found in bacteria that adopt only negative mechanisms of transcriptional regulation could be very simple. However, combinations of repressors and overlaps of their regulons are not rare and sometimes lead to complex regulatory networks. One of the simplest regulatory scheme is employed, for example, by the human pathogen M. genitalium (Fig. 6A, right panel). This Gram-positive bacterium, which has the smallest genome among self-replicating free living organisms, possesses a single housekeeping σ factor and manages heat-shock gene regulation exclusively through the HrcA repressor (Table 1) (Musatovova, Dhandayuthapani and Baseman 2006). Similarly, a single transcriptional repressor, CtsR, controls heat-shock response in the lactic acid bacterium O. oeni (Grandvalet et al. 2005) (Fig. 6A, right panel, and Table 1). A common manner of combining negative regulators is based on the exploitation of several transcriptional repressors, each specifically controlling a particular set of genes. In B. subtilis and closely related species, the HrcA regulator is in charge of keeping groESL and dnaK genes repressed under normal growth conditions, while Clp proteaseencoding genes are regulated solely by the master repressor CtsR (Table 1) (Chastanet et al. 2001). An extreme example is represented by S. albus, a soil bacterium that uses an arsenal of three different repressors to specifically regulate distinct sets of heatshock genes without any cross-regulation and regulons overlap Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 562 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 Figure 6. Combinations of heat-shock transcriptional regulators and complex regulatory circuits. (A) While in some bacterial species heat-shock transcription is controlled exclusively by positive (left panel) or negative (right panel) mechanisms of regulation, in several cases positive and negative control strategies coexist (central panel). Black arrows indicate positive regulation, while black hammerheads show negative regulation. (B) Examples in which distinct transcriptional repressors are employed to control heat-shock gene expression. In S. albus HrcA, HspR and RheA repressors regulate separate sets of genes, whereas in S. pneumoniae HrcA and CtsR regulons partially overlap. In S. aureus, the HrcA regulon is completely embedded within the CtsR regulon. (C) Schematic representation of the incoherent feedforward loop governing heat-shock gene regulation in S. aureus and H. pylori. A master regulator, shown in red (HspR for H. pylori or CtsR for S. aureus), represses several heat-shock genes. Among these genes, the groESL operon is also repressed by the HrcA regulator, which is, in turn, under the control of the master repressor. (Fig. 6A, right panel, and Fig. 6B). Specifically, while groEL regulation involves HrcA (Grandvalet, Rapoport and Mazodier 1998), dnaK and clpB genes are repressed by HspR (Bucca, Hindle and Smith 1997; Grandvalet, Servant and Mazodier 1997) and the hsp18 gene is transcriptionally controlled by RheA (Table 1) (Servant, Rapoport and Mazodier 1999). However, in several other bacterial species, the regulons of two different heat-shock repressors partially overlap and this results in some genes being simultaneously controlled by more than one regulatory protein, rendering these systems more complicated than the examples described above. HrcA is most commonly involved in those situations, partnering with CtsR in some bacteria, while in some other cases it interacts directly with HspR. A typical example is the dual regulation of groESL by CtsR and HrcA in S. pneumoniae (Fig. 6B). In this microorganism, the groESL 5 region harbors a highly conserved CIRCE operator, which is the specific recognition sequence for HrcA-binding mapping immediately downstream from the transcription start site. Also, there is an upstream CtsR-binding site overlapping the –10 and –35 hexamers (Chastanet et al. 2001). The two binding sites are in close proximity, separated by only 16 bp. This tandem operator arrangement, involving a crucial region for transcription initiation, results in the S. pneumoniae groESL operon being under dual repression by both HrcA and CtsR (Fig. 6B). Accordingly, it was demonstrated that, in a mutant strain lacking CtsR, transcription of groESL was just slightly increased at 37◦ C, consistent with functional HrcA repression (Chastanet et al. 2001). This non-redundant dual repression mechanism, also employed very similarly by S. salivarus for the regulation of groESL and clpP (Table 1) (Chastanet and Msadek 2003), allows transcription of these genes to be maintained at very low levels in the absence of stress by the synergistic action of both CtsR and HrcA and to be fully induced by both regulators upon signal perception. Also, in S. aureus, CtsR and HrcA combine to control the expression of groESL and dnaK, Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 563 Table 1. Regulatory mechanisms controlling heat-shock genes’ transcription in a panel of bacteria. Organism Transcriptional regulator Regulated genes Agrobacterium tumefaciens RpoH groESL, dnaK-dnaJ, grpE, clpP and others HrcA groESL Bacillus subtilis HrcA Class I: groESL, dnaK-dnaJ, grpE, and others σB Class II: genes coding for general stress proteins CtsR Class III: clpP, clpE, ctsR-mcsA-mcsB-clpC and others Unknown CssS/CssR Class IV: htpG Unknown Class V: htrA, htrB and others Class VI: ftsH, clpX and others Escherichia coli σ32 (σH , RpoH) groESL, dnaK-dnaJ, grpE, ibpA and others σE (σ24 ) degP, clpX, lon and others Helicobacter pylori HspR cbpA-hspR-helicase, groESL, hrcA-grpE-dnaK HrcA groESL, hrcA-grpE-dnaK Listeria monocytogenes HrcA Class I: groESL, dnaK, and others σB Class II: genes coding for general stress proteins CtsR Class III: clpP, clpB, hslU and others Mycoplasma genitalium HrcA dnaK, lon, clpB Oenococcus oeni HrcA groESL, dnaK, clpP, hsp18 and others Pseudomonas aeruginosa σH (RpoH) groESL, dnaK, and others Staphylococcus aureus HrcA hrcA-dnaK, groESL CtsR clpP, clpC, clpB, hrcA-dnaK, groESL σB genes coding for general stress proteins Streptococcus pneumoniae HrcA groESL, dnaK-dnaJ, grpE CtsR groESL, clpP, clpC, clpE Streptococcus salivarus HrcA clpP, groESL, dnaK CtsR clpP, groESL, other clp genes Streptomyces albus HrcA groESL1, groEL2, dnaJ2 HspR dnaK, clpB RheA hsp18 Synechocystis PCC6803 HrcA groESL, groEL-2, dnaK2 and others SigB groESL, groEL-2, dnaK2, hspA, htpG and others SigE groESL, groEL-2, dnaK2, hspA, htpG and others HiK34 groESL, dnaK2, htpG Vibrio cholerae σ32 (RpoH) groESL, dnaK-dnaJ, lon, clpB, and others but they interact in an even more peculiar and complex way (Fig. 6B). In particular, S. aureus CtsR alone represses the transcription of several clp genes, but also combines with HrcA to control expression mediated by the dnaK and the groESL operons. Moreover, since hrcA is the first gene of the dnaK operon, this results in the HrcA regulon being entirely embedded within the CtsR regulon (Table 1) (Chastanet, Fert and Msadek 2003). As in S. pneumoniae, CtsR- and HrcA-specific operators are organized in tandem and overlap the region adjacent to the transcription start site, allowing non-redundant dual repression during normal growth. Strikingly, an almost identical regulatory network has been described in the distantly related Gram-negative human pathogen H. pylori, where HrcA combines with the HspR repressor (Fig. 5). Here, HspR alone represses transcription of its own tricistronic operon (Spohn and Scarlato 1999), while, in combination with HrcA, it controls the expression of groESL and hrcA-grpE-dnaK operons (Spohn et al. 2004). Hence, similarly as for S. aureus, this results in the H. pylori HrcA regulon being embedded within the regulon of a second master heat-shock repressor (i.e. HspR instead of CtsR). A detailed analysis of in vitro DNA binding by HrcA and HspR revealed that their operators are arranged in tandem, consistent with the typical architecture for dually regulated promoters, but that the position of the HspR operator maps to far upstream from the core promoter to an atypical position for a repressor (Roncarati et al. 2007a). Moreover, even though both repressors are required for regulation (a single hspR or hrcA mutation leads to complete derepression of the dually controlled promoters), binding in vitro to their respective adjacent operators occurs in an independent, noncooperative manner. An interesting explanation for the peculiar interaction of heat-shock repressors found in H. pylori and S. aureus can be inferred by considering a logical scheme involving both regulators (Fig. 6C). In both organisms, there is a master regulator (HspR or CtsR) which directly regulates another regulator (HrcA) and a target gene (groESL), which, in turn, is regulated by both HrcA and the master regulator (HspR or CtsR). Moreover, considering that all three regulatory interactions are repressive, these circuits seem to represent rare examples of incoherent type-2 feed forward loops (Alon 2007; Danielli, Amore and Scarlato 2010). This peculiar network motif is employed to modulate the dynamic behavior of the circuit, greatly speeding up the transcriptional response of target genes upon input stress signals. Some experimental observations of H. pylori sustain this hypothesis. In fact, transcription of groESL was rapidly induced (2 min) upon heat-shock in a wild-type genetic background, while in a mutant strain in which the binding site of HspR on the groESL promoter was deleted (and, so, the direct connection between HspR and groESL was interrupted), the derepression of groESL transcription was observed only 60 min after temperature challenge (Spohn et al. 2004). Even though further experimental Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 564 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 characterizations are needed to dissect the dynamic properties of regulatory circuits such as those described above, they could reflect specific evolutionary adaptations to particular needs encountered by bacteria in their specific niches. The examples described above provide a complex picture of heat-shock genes’ regulation and raise interesting questions about their role and evolution. Even though, in many cases, clear experimental evidences are still missing, some speculations can be put forward. For instance, the presence in the same organism of diverse regulatory mechanisms involved in the control of distinct as well as overlapping sets of genes might have evolved to allow bacteria a proper response to distinct stresses. Alternatively, such regulatory systems might account for a ‘differential response’ in terms of both stress severity and differential pattern of gene expression in response to a stress insult. The latter aspect has been observed in S. albus, where different sets of heat-shock genes are controlled by different repressors (Fig. 6A). Specifically, two different regulatory patterns have been observed, consisting of constitutive synthesis of GroEL and Hsp18 (regulated by HrcA and RheA, respectively) at high temperature and transient heat-shock-induced synthesis of HspRregulated ClpB and DnaK proteins (Mazodier et al. 1991). A heatshock intensity-dependent response (differential response) of some regulatory systems or individual genes has been documented in different bacterial species, including M. tuberculosis and Corynebacterium glutamicum (Young and Garbe 1991; Engels et al. 2004). The use of diverse regulatory systems by the same organism to control different heat-shock genes becomes particularly relevant and interesting when it involves the differential regulation of multiple copies of genes coding for the GroEL chaperonins. It has been observed that a significant proportion of bacterial genomes contains two or more chaperonin genes and, in several instances, these genes appear to be differentially regulated (Lund 2009). The current notion is that, where present, these multiple chaperonins encoded by duplicated groE genes have evolved a degree of subfunctionalization and the differential regulation observed may reflect the various contexts in which they are needed or the ecological niche they have to cope with. For example, in Bradyrhizobium japonicum, a nitrogen-fixing and roots-nodulating bacterium belonging to the Alphaproteobacteria group, the seven chaperonin genes present in its genome show complex pattern of transcriptional regulation, not limited to heat-shock response. At least one of these chaperonin genes has been shown to be transcriptionally regulated by the NifA activator and by σ54 , which are both involved in the regulation of nitrogen fixation genes under limited oxygen conditions (Fischer 1994). Genetic analyses provide additional evidences for a link between the regulation of some chaperonins, root nodulation and nitrogen fixation, even though the specificity of chaperonin novel functions appeared not absolute (Lund 2009). A similar scenario of differential regulation of multiple chaperonin genes linked to the evolution of chaperonins’ specialized functions has also been proposed for other bacterial groups, including Cyanobacteria and Chlamydia (Lund 2009). MECHANISMS OF HEAT SENSING The ability of bacteria to rapidly respond to sudden temperature increase depends on heat-sensing mechanisms that integrate environmental cues to activate appropriate response pathways. To date, various mechanisms of thermoregulation have been described in bacteria and they involve nearly all classes of biomolecules including lipids, proteins and nucleic acids (i.e. both DNA and RNA). All of these classes can act as thermosensors that detect changes in the environmental temperature and initiate relevant cellular responses. Heat-sensing mechanisms can be direct, by which the temperature directly affects the activity of the sensing biomolecule, or can be indirect, by which the consequences of a sudden temperature increase (for example, the accumulation of misfolded proteins in the cytoplasm) are detected. Even though temperature is a ubiquitous signal that influences several cellular pathways, this chapter focuses mainly on sensing mechanisms that trigger heat-shock gene expression, and just a few examples of thermosensors involved in the regulation of virulence determinants will be described (extensively reviewed by Klinkert and Narberhaus 2009; Shapiro and Cowen 2012). Temperature sensing through RNA The most rapid way of changing gene expression in response to temperature variations is based on a cis-regulatory element that is part of the mRNA encoding the heat-shock protein to be regulated. This mechanism of thermoregulation guarantees a very fast response upon signal perception because temperature affects the translation efficiency of both the intracellular pool of mRNA molecules as well as transcription already in progress. The general principle is based on the formation of zipper-like, temperature-sensitive secondary structures that characterize the mRNAs subjected to this kind of regulation (Fig. 7A). In particular, at physiological temperature the 5 region of this category of mRNA forms a structure that hinders sequence elements crucial for the initiation of translation, such as the ribosomebinding site (also known as the Shine-Dalgarno sequence) and the translation start codon. When such sequence elements are involved in a secondary structure, the recognition and binding of the transcript by the ribosome is hampered, thereby negatively affecting translation of the mRNA. Upon a temperature increase, the secondary structure goes through a rearrangement or partial melting. As a consequence, ribosomes can easily gain access to the 5 mRNA region and translation is enhanced. Several temperature-sensing RNA sequences, also known as RNA thermometers, have been discovered and characterized in some details in the last two decades and were recently reviewed in a comprehensive and detailed article by Kortmann and Narberhaus (2012). The first RNA thermometer was discovered in E. coli and regulates the expression of the heat-shock alternative sigma factor σ32 , already discussed above. It represents one of the most complex RNA thermometers known so far, with an extended secondary structure that is not confined only to the RpoH mRNA 5 region (Fig. 1A). Specifically, although in the characterized secondary structure the Shine-Dalgarno sequence is partially exposed, translation of RpoH mRNA is hindered at low temperature by an intramolecular pairing between the 5 untranslated region and a portion of the coding sequence immediately downstream the AUG start codon (Kortmann and Narberhaus 2012). However, much simpler RNA thermometers exist, as exemplified by the one controlling the expression of the Bradyrhizobium japonicum hspA gene. This cis-regulatory sequence constitutes the founding member of the most abundant class of RNA thermometers called ROSE for repression of heat-shock genes expression. ROSE elements are typically involved in the regulation of small heat-shock proteins and they have a length ranging from 60 to about 120 nucleotides. ROSE elements were noticed as a conserved DNA element preceding several heatshock genes in various rhizobia. Initially, it was speculated that they act at the DNA level and their regulatory mechanism was Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 565 Figure 7. Mechanisms of heat sensing. (A) Sensing through RNA involves temperature-sensitive secondary structures in the 5 region of the mRNA subjected to this kind of regulation. During normal growth conditions, a secondary structure forms and sequence elements crucial for efficient initiation of translation are masked and poorly accessible for ribosome binding. Upon temperature increase, a structural rearrangement or resolution of the structure exposes such sequence elements and translation is enhanced. Symbols: SD, Shine-Dalgarno sequence element; AUG, translation start codon. (B) Temperature sensing through DNA: model of temperaturedependent regulation of virF transcription in S. flexneri. The DNA region flanked by the H-NS binding sites assumes a curvature at low temperature that allows contacts between H-NS dimers bound to separate binding sites and the formation of a repressive nucleoprotein complex. At higher temperature, this curvature weakens and the promoter is more accessible to RNA polymerase for virF gene transcription. (C) Transcriptional regulators as intrinsic heat sensors. Intrinsic heat-sensing repressors are competent for DNA binding at permissive temperature (green oval), and transcription of heat-shock genes is repressed. Upon heat challenge, a temperatureinduced structural transition (shown by a green rectangle) lead to a decrease of repressor DNA-binding activity and transcription is derepressed. (D) Indirect heatsensing mechanism mediated by chaperones. The DNA-binding activity of transcriptional regulators is modulated by chaperones. During normal growth conditions, the chaperone interacts with the heat-shock regulator and exerts its regulatory function. Several positively modulated repressors (such as HrcA and HspR in some bacterial species) gain DNA-binding activity upon chaperone interaction, while in other instances the interaction with the chaperone results in the sequestration of the activator (for example, E. coli σ32 ). Following heat stress, chaperones are sequestered by misfolded proteins that accumulate in the cytoplasm and transcriptional regulators are released and their DNA-binding activity results are positively or negatively affected. thought to be dependent on the binding of a repressor protein (Narberhaus et al. 1998). Noting that it was possible to predict a similar secondary structure for all 15 known ROSE elements, a post-transcriptional mechanism was proposed and verified by a detailed mutational analysis of a ROSE-hspA-lacZ translational fusion (Nocker et al. 2001). Subsequent extensive molecular and structural characterizations revealed some key features of such kinds of RNA thermometers (Chowdhury et al. 2003, 2006). Interestingly, it was shown that ROSE-based RNA thermometers comprise two to four stem loops and that they can gradually respond to temperature variation and provide differential levels of expression regulation according to the severity of heat stress. Moreover, the dissection of the crucial intramolecular interactions involved in secondary structure formation revealed that, even though ROSE elements are characterized by a conserved short stretch of nucleotides (U(U/C)GCU), non-canonical base pairs also seem to be crucial for fine differential temperature sensing. Additional examples of RNA thermometers containing ROSE or ROSE-like elements have also been found in other bacterial species such as Caulobacter, Salmonella and Escherichia coli. In the latter species, a ROSE-like RNA thermometer controls the temperature-dependent expression of the small heatshock protein IbpA (coding for an inclusion body binding protein), whose transcription is σ32 dependent (Nocker et al. 2001; Waldminghaus et al. 2009). Another class of RNA thermometers is based on a short stretch of four conserved uridines that base pair with the AGGA nucleotide sequence constituting the ShineDalgarno sequence. This element, called fourU, was initially Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 566 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 characterized in Salmonella enterica, where it controls the temperature-dependent expression of the agsA gene that encodes for a small heat-shock protein (Waldminghaus et al. 2007). Contrary to ROSE elements, the formation of the doublestranded RNA secondary structure in fourU elements is governed solely by canonical base pairing (Kortmann and Narberhaus 2012). Moreover, a peculiar feature of the fourU elements, derived during a detailed structural investigation of the agsA thermometer, is that the melting process in response to temperature fluctuation is dependent on the concentration of Mg2+ ions, due to the presence of an Mg2+ -binding site placed directly on the fourU element (Rinnenthal et al. 2011). It is worth noting that thermoregulation based on RNA-sensing structures has not evolved exclusively with respect to heat-shock genes. Several pathogenic bacteria, which are subjected to shifts in temperature during their life cycle (from environmental temperature of 28◦ C–30◦ C to body temperature of 37◦ C or even higher during infection), can modulate the expression of some virulence-associated genes in response to temperature variations using RNA thermometers (Grosso-Becera, Serv´ın-Gonz´alez and Sober´on-Ch´avez 2015). For example, in the obligate commensal of the human nasopharynx Neisseria meningitidis, RNA thermometers were found in the 5 region of three genes involved in the meningococcal resistance against immune killing. Specifically, cssR, fHbp and lst genes (the first coding for an enzyme involved in the exopolysaccharide biosynthesis, the second coding for the host complement regulator factor H-binding protein and the third encoding a protein necessary for LPS sialylation) harbor RNA thermometers adjacent to the ShineDalgarno sequence that enhance their expression at 42◦ C, a temperature that is reached in the nasopharynx during Neisseria and viral co-infections (Loh et al. 2013; Barnwal et al. 2016). Temperature sensing through DNA In some cases, temperature variations can be directly sensed by the DNA of the bacterial cell. Several physiological metabolic pathways of a microorganism are influenced by the external conditions experienced in its ecological niche. In turn, the metabolic state of the cell is influenced, including the ADP to ATP ratio. As a consequence, enzymes that require ATP as a cofactor, such as DNA gyrase, will be affected. This enzyme, whose activity is strictly linked to the topological state of the DNA, is dependent on ATP and inhibited by ADP (so change in the ATP:ADP ratio influences DNA gyrase activity). For this reason, external fluctuating conditions that affect metabolic processes, including osmotic and heat-shock, can ultimately influence the global level of DNA supercoiling (Hsieh, Burger and Drlica 1991; Dorman and Corcoran 2009). Considering that DNA supercoiling can influence gene transcription, DNA can be considered as a thermosensor of environmental temperature change acting through variations of the global topological state of the chromosome in response to external stimuli. One of the primary parameters of DNA topology that responds to temperature changes is plasmid supercoiling. In particular, it was demonstrated both in mesophilic and hyperthermophilic bacteria that sudden temperature variations lead to transient changes in plasmid DNA topology and this effect, in turn, has a significant impact on transcription efficiency (L´opez-Garc´ıa and Forterre 1997). In some other cases, however, local DNA structures mediate temperature sensing and affect transcription of neighboring genes. Some DNA sequences identified in E. coli and in several other bacteria are characterized by a sequence specific topological conformation that is able to assume various conformations in response to temperature variation. When these DNA regions are proximal to promoters, the local conformational variation induced by heatshock is transduced into a modulation of RNA polymerase binding efficiency, thereby affecting the transcription of the downstream genes (Nickerson and Achberger 1995). An example is represented by the region upstream of the plc gene, encoding the phospholipase C (PLC) in Clostridium perfrigens (Katayama et al. 1999). In this case, the three phased A-tracts that precede the PLC encoding gene confer a strongly bent conformation to this DNA sequence. By using in vitro transcription assays, it was demonstrated that the stimulatory effect on the promoter activity of the A-tract sequence was temperature dependent, probably due to changes in the bending angle upon temperature fluctuations. Besides the direct activation of transcription mediated by bent DNA through the facilitation of RNA polymerase binding, temperature-dependent local DNA curvatures can indirectly regulate the efficiency of transcription by affecting the interaction of proteins with a regulatory role on gene expression. In this respect, nucleoid-associated proteins such as IHF, Fis, HU and H-NS have been shown to be involved in this kind of process. One of the best studied examples concerns the temperaturedependent regulation of virF transcription in Shigella flexneri, a pathogenic bacterium able to invade human intestinal epithelium (Fig. 7B). The AraC-like VirF regulator, a protein that triggers the activation of several genes with invasive functions, must be expressed only after the shift from the outside environment to the host. It has been demonstrated that the transcription of the virF promoter is kept repressed at non-permissive temperatures (below 32◦ C) by the nucleoid-associated protein HNS (Colonna et al. 1995). Moreover, it was shown that the virF promoter harbors two H-NS binding sites separated by an intrinsically bent region, whose existence was predicted in silico and demonstrated by in vitro experimental results (Falconi et al. 1998; Prosseda et al. 2004). Intriguingly, upon temperature increase, this DNA region undergoes an abrupt structural transition (at ∼32◦ C) that affects H-NS accessibility to target DNA sites and represses the virF promoter (Prosseda et al. 2004). A similar interplay between temperature-dependent DNA bending and H-NS binding was demonstrated to modulate hemolysin gene expression in the model organism E. coli (Madrid et al. 2002) and to enhance the expression of a type III secretion system above 30◦ C in S. enterica (Duong et al. 2007). To summarize, even though several examples of temperature-dependent regulation of gene expression point to the role of DNA as a sensor biomolecule, it appears that this mechanism of heat sensing is much more pertinent to regulating the expression of virulence genes rather than in the heat-shock response. Sensing through proteins Bacteria can also sense and respond to temperature fluctuations in their ecological niches by using proteins as heat sensors. Different classes of proteins, including kinases, heat-shock transcriptional repressors and chaperones, have been characterized as sensors of temperature changes (Klinkert and Narberhaus 2009). However, the two latter categories of proteins are primarily employed to transduce heat stress signals that trigger a transcriptional heat-shock response. Intrinsic heatsensing transcriptional repressors are able to directly modulate the transcription of target genes in response to temperature fluctuation. Specifically, they are competent for promoter binding and repress gene transcription only at physiological temperature. Upon heat-shock, these thermosensing repressors undergo Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 567 a conformational change that lowers relative binding affinity for their operators. As a consequence, target gene transcription becomes derepressed (Fig. 7C). With respect to the previous point, CtsR, the global repressor of heat-shock genes of Bacillus subtilis and other low-GC Gram-positive bacteria, is one of the best characterized examples (Elsholz et al. 2010). In vitro DNAbinding assays carried out at different temperatures showed that CtsR-binding activity to the clpC promoter drastically drops under heat-shock conditions (50◦ C), compared to normal growth temperature at 37◦ C. Interestingly, the temperature-dependent loss of DNA-binding activity of CtsR was demonstrated to be reversible. In fact, when CtsR was first incubated at a nonpermissive temperature and then at 37◦ C, the regulator regained DNA-binding activity. In the case of CtsR, the temperaturedependent conformational change responsible for the loss of DNA-binding activity was shown to be limited to a short glycinerich loop region within the DBD, constituting the precise functional site for heat sensing (Elsholz et al. 2010). The Streptomyces albus heat-shock repressor RheA described above was shown to be capable of sensing temperature variations in the absence of other factors, behaving similarly to CtsR. Upon heat challenge, RheA loses DNA-binding activity and the transition from active to inactive form was related to a temperature-induced reversible conformational change (Servant, Grandvalet and Mazodier 2000). A similar paradigm of temperature-dependent DNA-binding capacity was also clearly demonstrated for the virulence-associated RovA regulator of Yersinia pestis and for the TlpA regulator of S. enterica serovar Typhimurium (Hurme et al. 1997; Herbst et al. 2009). In the latter example, in particular, temperature reversibly affects the DNA-binding competent coiledcoil domains of the oligomerized state of the protein, thereby affecting DNA-binding ability. Another widespread heat-shock repressor, HrcA, can act, in some cases, as an intrinsic protein thermometer. The observation that the thermal denaturation profiles of B. subtilis and of B. thermoglucosidasius HrcACIRCE complexes (assayed by light scattering) were highly consistent with the different growth temperatures of the two microorganisms led to the proposal that the HrcA repressor might have a role as a thermosensor (Hitomi et al. 2003). A direct role as thermosensor was recently demonstrated for the HrcA repressor of Helicobacter pylori. Specifically, in vitro DNaseI footprinting assay results showed that HrcA-mediated DNA binding is strictly temperature dependent (Roncarati, Danielli and Scarlato 2014). When H. pylori HrcA was exposed to temperatures above 37◦ C (physiological growth temperature), it dramatically lowered binding affinity to CIRCE operators and became essentially inactive. Intriguingly, the loss of binding activity upon heat-shock treatment appeared to be irreversible in vitro, likely as a consequence of the major structural change as the temperature exceeded 40◦ C. However, the situation is different in vivo, where HrcA is able to recover its repressive function after the temperature challenge (Roncarati, Danielli and Scarlato 2014). Furthermore, it was demonstrated that the GroESL chaperonin plays a key role and that it is able to restore HrcA-binding activity lost upon heat challenge. The working model for HrcA-mediated heat sensing in H. pylori postulates that, after a sudden increase of temperature, heat-mediated inactivation of the regulator leads to promoter derepression and increased transcription of the heatshock gene promoters. Then, following the induction step, at least a fraction of the denatured repressor is refolded by GroESL and can take part, together with the newly synthetized HrcA, in transcriptional regulation of the target genes. Considering the key role of GroESL in regulating HrcA-binding activity upon heat sensing, this functional interaction could be exploited by H. pylori to adjust the transcriptional response under various stress conditions. That is to say that, depending on the severity of a particular stress stimulus, the amount of GroE chaperonin available for the functional interaction with HrcA will change, influencing the restoration of target gene regulation. A different situation was observed for Chlamydia trachomatis HrcAdependent heat-shock gene regulation. In this obligate intracellular human pathogen, HrcA binds to CIRCE operators within groESL and dnaK promoters and represses their transcription under normal growth conditions. Heat-shock gene transcription is then induced upon a sudden temperature increase. However, it was shown that HrcA-binding activity to CIRCE was not affected by elevated temperature in vitro (Wilson and Tan 2004). Recently, Hanson and Tan (2015) developed a method based on ChIP to study heat-shock gene regulation during an intracellular infection, and they were able to monitor in vivo HrcA binding to target promoters at various temperatures. This approach allowed demonstration that, upon a sudden temperature increase, the in vivo induction of heat-shock gene transcription was due to a decrease of HrcA binding to groESL and dnaK promoters. The discrepancy between the in vitro and in vivo results for HrcA temperature sensitivity suggests that elevated temperature is not sufficient per se to modulate chlamydial HrcA-binding capacity and implies that additional in vivo factors must be involved in the regulation of the HrcA-mediated heat-shock response. Based on previous observations of a functional interaction between HrcA and GroE (Wilson et al. 2005), it has been proposed that the chaperonin is the actual heat sensor of the regulatory circuit. The latter example represents a well-characterized mechanism employed in several cases that link environmental stress signals to a corresponding transcriptional response. This signal-sensing system does not directly detect the temperature increase; rather, the consequences of heat-shock on cellular proteins are sensed. Specifically, this strategy, represented in Fig. 7D, is based on the direct interaction between heat-shock transcriptional regulators and the major cellular chaperones. DNA-binding activity of transcriptional regulators is positively or negatively regulated by chaperones that are available for interaction during normal growth conditions. Upon sudden temperature upshift, chaperones are titrated away by denatured proteins that accumulate in the cytoplasm and transcriptional regulators are released. As described above (Figs 2 and Fig. 4A and B), the activity of HrcA and HspR repressors of several bacterial species, but also of the heat-shock σ32 of E. coli, is regulated in this way by the principal cellular chaperones such as the GroE chaperonin and DnaKDnaJ-GrpE system (Straus, Walter and Gross 1990; Babst, Hennecke and Fischer 1996; Mogk et al. 1997; Bucca et al. 2000; Reischl, Wiegert and Schumann 2002; Wilson et al. 2005). According to this indirect heat-sensing mechanism, chaperones, and not transcriptional regulators, are the actual ‘stress sensors’. An interesting aspect emerging from the examples of the sensing mechanisms described above is that, in several cases, the regulation of the heat-shock response involves both direct temperature sensing (through RNA thermometers and intrinsic heat-sensing transcriptional regulators) and indirect chaperone-mediated sensing of unfolded proteins. These composed systems allow bacteria to respond differentially, depending on severity and type of environmental stresses. For example, RNA thermometers can detect minor temperature variations allowing the system to immediately respond, before cellular processes are altered. In this respect, E. coli σ32 regulation represents a perfect example. Here temperature-regulated translation of RpoH mRNA combines with two feedback loops that Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 568 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 control σ32 activity and stability. An interesting implication of these combined systems of signal sensing is that cells might discriminate heat stress from any other type of stress. Consequently, bacteria might differentially induce the expression of specific set of genes, thus increasing specificity of stress response. CONCLUSIONS This review highlights the importance of a sudden response of various bacterial species to temperature changes and summarizes the key mechanisms that diverse bacteria adopt to counteract imminent cellular damage. This highly conserved defense mechanism against environmental stress, known as the heatshock response, appears to be controlled by a plethora of different strategies in bacteria. Control of heat-shock gene transcription is successfully and efficiently achieved by the use of specialized regulatory proteins exerting their positive or negative action on the initiation of transcription by the RNA polymerase enzyme. Strikingly, positive and negative mechanisms of regulation, alone or in combination, are employed by microorganisms to rapidly reprogram gene transcription upon stress perception. Additionally, control of gene transcription can be combined with posttranscriptional mechanisms of protein ex- pression. The model organism Escherichia coli has long served as the central paradigm for understanding the mechanism of stress induced activation of heat-shock gene transcription based on the use of alternative sigma factors. After decades of intense research on the heat-shock sigma factor σ32 , it appears that a complex multilayered regulatory cascade orchestrates σ32 homeostasis, in which transcriptional and translational control combines with chaperone mediated modulation of σ32 stability. Moreover, the recent discovery of inner membrane localization of the heat-shock sigma factor mediated by the SRP-SR co-translational targeting system adds another key part to this intricate story. However, further efforts are needed to understand several mechanistic aspects of the transit of σ32 to the membrane. Negative regulation of heat-shock gene transcription through dedicated repressors has been studied in detail in several bacterial species. Different aspects of the major heat-shock repressors have been characterized, such as their DNA-binding mechanism, the characterization of their regulon and their biochemical properties (such as, for example, their heat-sensing capabilities). In this respect, it is worth noting that the way in which the master heat-shock repressor HspR senses fluctuations of environmental temperature remains elusive. In fact, while for HrcA, CtsR and RheA several studies pinpointed the direct or indirect (chaperone-mediated) mechanisms of heat sensing, the HspR repressor appears to be a heat stable protein and its interplay with the DnaK chaperone complex could be experimentally demonstrated in only a few cases. Further studies should be carried out to understand how temperature can modulate the DNA-binding activity of HspR. In several cases, heat-shock regulatory circuits appear to be built on very complex structures in which mechanisms of transcriptional control combine with several feedback loops, whose effects on the properties of the regulatory systems are hardly predictable. In this respect, the question arises as to whether the use of such complex regulation strategies in the heat-shock systems is the result of evolutionary accidents or it derives from specifically designed solutions to different requirements in the field of heat-shock response. Rigorous analyses have been carried out with systems biology approaches and computational simulations of E. coli heat-shock response (El-Samad et al. 2005; Kurata et al. 2006). As described above, the E. coli heat-shock response involves complicated interactions, where both the activity and stability of the alternative sigma factor σ32 are feedback controlled by chaperones and proteases, respectively. From these studies, it emerges that the use of feedback loops gives valuable properties to the system, in terms in primis of robustness, defined as the capability of a system to operate reliably when its physical parameters vary within their expected ranges. Furthermore, the use of FtsH protease-mediated degradation feedback appears to provoke a faster response to a heat stimulation and to limit the effects of biochemical noise (El-Samad et al. 2005; Kurata et al. 2006). Further similar studies will undoubtedly be crucial for the investigation and characterization of other complex heat-shock regulatory schemes. Another intriguing aspect of heat-shock gene regulation that deserves further characterization concerns the combination of different transcriptional regulators that establish highly complex regulatory schemes. In particular, the cases of Staphylococcus aureus and Helicobacter pylori, for which two repressors are employed and the regulon of one repressor is embedded within the regulon of the other repressor, represent rare examples of regulatory networks whose functional implications remain to be defined. The recent development and application of the ‘omics’ techniques will allow further detailed characterization and understanding of the heat-shock regulons specifically governed by various heat-shock regulators. Functional genomics studies are expected to deepen and expand our knowledge of the impact of heat stress on the bacterial transcriptome and of the regulatory mechanisms that control such responses. FUNDING This work was supported by Grants from the Italian Ministry of Education, University and Research (2010P3S8BR 003) and by a grant from the University of Bologna. Conflict of interest. None declared. REFERENCES Ades SE. Regulation by destruction: design of the σE envelope stress response. Curr Opin Microbiol 2008;11:535–40. Allan B, Linseman M, MacDonald LA et al. Heat shock response of Pseudomonas aeruginosa. J Bacteriol 1988;170:3668–74. Alon U. Network motifs: theory and experimental approaches. Nat Rev Genet 2007;8:450–61. Andersen MT, Brøndsted L, Pearson BM et al. Diverse roles for HspR in Campylobacter jejuni revealed by the proteome, transcriptome and phenotypic characterization of an hspR mutant. Microbiology 2005;151:905–15. Babst M, Hennecke H, Fischer HM. Two different mechanisms are involved in the heat-shock regulation of chaperonin gene expression in Bradyrhizobium japonicum. Mol Microbiol 1996;19:827–39. Bandyopadhyay B, Das Gupta T, Roy D et al. DnaK dependence of the mycobacterial stress responsive regulator HspR is mediated through its hydrophobic C-terminal tail. J Bacteriol 2012;194:4688–97. Barnwal RP, Loh E, Godin KS et al. Structure and mechanism of a molecular rheostat, an RNA thermometer that modulates immune evasion by Neisseria meningitidis. Nucleic Acids Res 2016;44:9426–37. Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 569 Becskei A, Serrano L. Engineering stability in gene networks by autoregulation. Nature 2000;405:590–3. Bird JG, Sharma S, Roshwalb SC et al. Functional analysis of CbpA, a DnaJ homolog and nucleoid-associated DNA-binding protein. J Biol Chem 2006;281:34349–56. Briat JF, Gilman MJ, Chamberlin MJ. Bacillus subtilis σ28 and Escherichia coli σ32 (htpR) are minor factors that display an overlapping promoter specificity. J Biol Chem 1985;260:2038–41. Bucca G, Brassington AME, Hotchkiss G et al. Negative feedback regulation of dnaK, clpB and lon expression by the DnaK chaperone machine in Streptomyces coelicolor, identified by transcriptome and in vivo DnaK-depletion analysis. Mol Microbiol 2003;50:153–66. Bucca G, Brassington AME, Sch¨onfeld HJ et al. The HspR regulon of Streptomyces coelicolor: a role for the DnaK chaperone as a transcriptional co-repressor. Mol Microbiol 2000;38:1093–103. Bucca G, Ferina G, Puglia AM et al. The dnaK operon of Streptomyces coelicolor encodes a novel heat-shock protein which binds to the promoter region of the operon. Mol Microbiol 1995;17:663–74. Bucca G, Hindle Z, Smith CP. Regulation of the dnaK operon of Streptomyces coelicolor A3(2) is governed by HspR, an autoregulatory repressor protein. J Bacteriol 1997;179:5999–6004. Bucca G, Laing E, Vassilis M et al. Development and application of versatile high density microarrays for genome-wide analysis of Streptomyces coelicolor: characterization of the HspR regulon. Genome Biol 2009;10:R5. ˇCerven´y J, Sinetova MA, Zavˇrel T et al. Mechanisms of high temperature resistance of Synechocystis sp. PCC 6803: an impact of Histidine Kinase 34. Life (Basel) 2015;5:676–99. Cezairliyan BO, Sauer RT. Inhibition of regulated proteolysis by RseB. P Natl Acad Sci USA 2007;104:3771–6. Chaba R, Grigorova IL, Flynn JM et al. Design principles of the proteolytic cascade governing the sigmaE-mediated envelope stress response in Escherichia coli: keys to graded, buffered, and rapid signal transduction. Gene Dev 2007;21:124–36. Chae C, Sharma S, Hoskins JR et al. CbpA, a DnaJ homolog, is a DnaK co-chaperone, and its activity is modulated by CbpM. J Biol Chem 2004;279:33147–53. Chang BY, Chen KY, Wen YD et al. The response of a Bacillus subtilis temperature-sensitive sigA mutant to heat stress. J Bacteriol 1994;176:3102–10. Charpentier B, Branlant C. The Escherichia coli gapA gene is transcribed by the vegetative RNA polymerase holoenzyme E Sigma 70 and by the heat shock RNA polymerase E Sigma 32. J Bacteriol 1994;176:830–9. Chastanet A, Derre I, Nair S et al. clpB, a novel member of the Listeria monocytogenes CtsR regulon, is involved in virulence but not in general stress tolerance. J Bacteriol 2004;186: 1165–74. Chastanet A, Fert J, Msadek T. Comparative genomics reveal novel heat shock regulatory mechanisms in Staphylococcus aureus and other Gram-positive bacteria. Mol Microbiol 2003;47:1061–73. Chastanet A, Msadek T. ClpP of Streptococcus salivarius is a novel member of the dually regulated class of stress response genes in gram-positive bacteria. J Bacteriol 2003;185:683–7. Chastanet A, Prudhomme M, Claverys JP et al. Regulation of Streptococcus pneumoniae clp genes and their role in competence development and stress survival. J Bacteriol 2001;183: 7295–307. Chen AL, Wilson AC, Tan M. A Chlamydia-specific C-terminal region of the stress response regulator HrcA modulates its repressor activity. J Bacteriol 2011;193:6733–41. Chowdhury S, Maris C, Allain FH et al. Molecular basis for temperature sensing by an RNA thermometer. EMBO J 2006;25:2487–97. Chowdhury S, Ragaz C, Kreuger E et al. Temperature-controlled structural alterations of an RNA thermometer. J Biol Chem 2003;278:47915–21. Coates AR, Shinnick TM, Ellis RJ. Chaperonin nomenclature. Mol Microbiol 1993;8:787. Colonna B, Casalino M, Fradiani PA et al. H-NS regulation of virulence gene expression in enteroinvasive Escherichia coli harboring the virulence plasmid integrated into the host chromosome. J Bacteriol 1995;177:4703–12. Danielli A, Amore G, Scarlato V. Built shallow to maintain homeostasis and persistent infection: insight into the transcriptional regulatory network of the gastric human pathogen Helicobacter pylori. PLoS Pathog 2010;6:e1000938. Dartigalongue C, Missiakas D, Raina S. Characterization of the Escherichia coli sigma E regulon. J Biol Chem 2001;276: 20866–75. Das Gupta T, Bandyopadhyay B, Das Gupta SK. Modulation of DNA-binding activity of Mycobacterium tuberculosis HspR by chaperones. Microbiology 2008;154:484–90. Delany I, Spohn G, Rappuoli R et al. In vitro selection of high affinity HspR-binding sites within the genome of Helicobacter pylori. Gene 2002;283:63–9. de Reuse H, Vinella D, Cavazza C. Common themes and unique proteins for the uptake and trafficking of nickel, a metal essential for the virulence of Helicobacter pylori. Front Cell Infect Microbiol 2013;3:94. Derr´e I, Rapoport G, Devine K et al. ClpE, a novel type of HSP100 ATPase, is part of the CtsR heat shock regulon of Bacillus subtilis. Mol Microbiol 1999;32:581–93. Derr´e I, Rapoport G, Msadek T. CtsR, a novel regulator of stress and heat shock response, controls clp and molecular chaperone gene expression in gram-positive bacteria. Mol Microbiol 1999;31:117–31. Derr´e I, Rapoport G, Msadek T. The CtsR regulator of stress response is active as a dimer and specifically degraded in vivo at 37 degrees C. Mol Microbiol 2000;38:335–47. Dorman CJ, Corcoran CP. Bacterial DNA topology and infectious disease. Nucleic Acids Res 2009;37:672–8. Duong N, Osborne S, Bustamante VH et al. Thermosensing coordinates a cis-regulatory module for transcriptional activation of the intracellular virulence system in Salmonella enterica serovar Typhimurium. J Biol Chem 2007;282: 34077–84. El-Samad H, Kurata H, Doyle JC et al. Surviving heat shock: control strategies for robustness and performance. P Natl Acad Sci USA 2005;102:2736–41. Elsholz AK, Michalik S, Z ¨uhlke D et al. CtsR, the Gram-positive master regulator of protein quality control, feels the heat. EMBO J 2010;29:3621–9. Engels S, Schweitzer JE, Ludwig C et al. clpC and clpP1P2 gene expression in Corynebacterium glutamicum is controlled by a regulatory network involving the transcriptional regulators ClgR and HspR as well as the ECF sigma factor sigmaH. Mol Microbiol 2004;52:285–302. Erickson JW, Gross CA. Identification of the σE subunit of Escherichia coli RNA polymerase: a second alternative sigma involved in high temperature gene expression. Gene Dev 1989;3:1462–71. Erickson JW, Vaughn V, Walter WA et al. Regulation of the promoters and transcripts of rpoH, the Escherichia coli heat shock regulatory gene. Gene Dev 1987;1:419–32. Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 570 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 Falconi M, Colonna B, Prosseda G et al. Thermoregulation of Shigella and Escherichia coli EIEC pathogenicity. A temperature-dependent structural transition of DNA modulates accessibility of virF promoter to transcriptional repressor H-NS. EMBO J 1998;17:7033–43. Fiocco D, Capozzi V, Collins M et al. Characterization of the CtsR stress response regulon in Lactobacillus plantarum. J Bacteriol 2010;192:896–900. Fiocco D, Collins M, Muscariello L et al. The Lactobacillus plantarum ftsH gene is a novel member of the CtsR stress response regulon. J Bacteriol 2009;191:1688–94. Fischer HM. Genetic regulation of nitrogen fixation in rhizobia. Microbiol Rev 1994;58:352–86. Fuhrmann J, Schmidt A, Spiess S et al. McsB is a protein arginine kinase that phosphorylates and inhibits the heat shock regulator CtsR. Science 2009;324:1323–7. Gamer J, Bujard H, Bukau B. Physical interaction between heat shock proteins DnaK, DnaJ, and GrpE and the bacterial heat shock transcription factor sigma 32. Cell 1992;69:833–42. Georgopoulos CP, Hendrix RW, Kaiser AD et al. Role of the host cell in bacteriophage morphogenesis: effects of a bacterial mutation on T4 head assembly. Nat New Biol 1972;239:38–41. Gerth U, Kr ¨uger E, Derr`e I et al. Stress induction of the Bacillus subtilis clpP gene encoding a homologue of the proteolytic component of the Clp protease and the involvement of ClpP and ClpX in stress tolerance. Mol Microbiol 1998;28:787–802. Grandvalet C, Coucheney F, Beltramo C et al. CtsR is the master regulator of stress response gene expression in Oenococcus oeni. J Bacteriol 2005;187:5614–23. Grandvalet C, de Cr´ecy-Lagard V, Mazodier P. The ClpB ATPase of Streptomyces albus G belongs to the HspR heat shock regulon. Mol Microbiol 1999;31:521–32. Grandvalet C, Rapoport G, Mazodier P. hrcA, encoding the repressor of the groEL genes in Streptomyces albus G, is associated with a second dnaJ gene. J Bacteriol 1998;180:5129–34. Grandvalet C, Servant P, Mazodier P. Disruption of hspR, the repressor gene of the dnaK operon in Streptomyces albus G. Mol Microbiol 1997;23:77–84. Grigorova IL, Chaba R, Zhong HJ et al. Fine-tuning of the Escherichia coli σE envelope stress response relies on multiple mechanisms to inhibit signal-independent proteolysis of the trans- membrane anti-sigma factor, RseA. Gene Dev 2004;18:2686–97. Grossman AD, Erickson JW, Gross CA. The htpR gene product of E. coli is a sigma factor for heat-shock promoters. Cell 1984;38:383–90. Grossman AD, Straus DB, Walter WA et al. Sigma 32 synthesis can regulate the synthesis of heat shock proteins in Escherichia coli. Gene Dev 1987;1:179–84. Grosso-Becera MV, Serv´ın-Gonz´alez L, Sober´on-Ch´avez G. RNA structures are involved in the thermoregulation of bacterial virulence-associated traits. Trends Microbiol 2015;23: 509–18. Guisbert E, Herman C, Lu CZ et al. A chaperone network controls the heat shock response in E. coli. Gene Dev 2004;18:2812–21. Hanson BR, Tan M. Transcriptional regulation of the Chlamydia heat shock stress response in an intracellular infection. Mol Microbiol 2015;97:1158–67. Helmann JD. The extracytoplasmic function (ECF) sigma factors. Adv Microb Physiol 2002;46:47–110. Henderson B, Martin A. Bacterial virulence in the moonlight: multitasking bacterial moonlighting proteins are virulence determinants in infectious disease. Infect Immun 2011;79:3476–91. Herbst K, Bujara M, Heroven AK et al. Intrinsic thermal sensing controls proteolysis of Yersinia virulence regulator RovA. PLoS Pathog 2009;5:e1000435. Herman C, Th´evenet D, D’Ari R et al. Degradation of sigma 32, the heat shock regulator in Escherichia coli, is governed by HflB. P Natl Acad Sci USA 1995;92:3516–20. Hitomi M, Nishimura H, Tsujimoto Y et al. Identification of a helix-turn-helix motif of Bacillus thermoglucosidasius HrcA essential for binding to the CIRCE element and thermostability of the HrcA-CIRCE complex, indicating a role as a thermosensor. J Bacteriol 2003;185:381–5. Holmes CW, Penn CW, Lund PA. The hrcA and hspR regulons of Campylobacter jejuni. Microbiology 2010;156:158–66. Hsieh LS, Burger RM, Drlica K. Bacterial DNA supercoiling and [ATP]/[ADP]. Changes associated with a transition to anaerobic growth. J Mol Biol 1991;219:443–50. Hurme R, Berndt KD, Normark SJ et al. A proteinaceous gene regulatory thermometer in Salmonella. Cell 1997;90:55–64. Hu Y, Raengpradub S, Schwab U et al. Phenotypic and transcriptomic analyses demonstrate interactions between the transcriptional regulators CtsR and Sigma B in Listeria monocytogenes. Appl Environ Microb 2007;73:7967–80. Janaszak A, Nadratowska-Wesolowska B, Konopa G et al. The P1 promoter of the Escherichia coli rpoH gene is utilized by sigma 70-RNAP or sigma s-RNAP depending on growth phase. FEMS Microbiol Lett 2009;291:65–72. J¨ager S, J¨ager A, Klug G. CIRCE is not involved in heat-dependent transcription of groESL but in stabilization of the mRNA 5 -end in Rhodobacter capsulatus. Nucleic Acids Res 2004;32: 386–96. Jastrab JB, Wang T, Murphy JP et al. An adenosine triphosphateindependent proteasome activator contributes to the virulence of Mycobacterium tuberculosis. P Natl Acad Sci USA 2015;112:E1763–72. Kallipolitis BH, Valentin-Hansen P. Transcription of rpoH, encoding the Escherichia coli heat-shock regulator sigma32, is negatively controlled by the cAMP-CRP/CytR nucleoprotein complex. Mol Microbiol 1998;29:1091–9. Kamath-Loeb AS, Gross CA. Translational regulation of sigma 32 synthesis: requirement for an internal control element. J Bacteriol 1991;173:3904–6. Kanehara K, Ito K, Akiyama Y. YaeL (EcfE) activates the σE pathway of stress response through a site-2 cleavage of anti-σE , RseA. Gene Dev 2002;16:2147–55. Kanehara K, Ito K, Akiyama K. YaeL proteoylsis of RseA is controlled by the PDZ domain of YaeL and a Gln-rich region of RseA. EMBO J 2003;22:6389–98. Kanemori M, Mori H, Yura T. Effects of reduced levels of GroE chaperones on protein metabolism: enhanced synthesis of heat shock proteins during steady-state growth of Escherichia coli. J Bacteriol 1994;176:4235–42. Kanemori M, Nishihara K, Yanagi H et al. Synergistic roles of HslVU and other ATP-dependent proteases in controlling in vivo turnover of sigma32 and abnormal proteins in Escherichia coli. J Bacteriol 1997;179:7219–25. Karunakaran KP, Noguchi Y, Read TD et al. Molecular analysis of the multiple GroEL proteins of Chlamydiae. J Bacteriol 2003;185:1958–66. Katayama S, Matsushita O, Jung CM et al. Promoter upstream bent DNA activates the transcription of the Clostridium perfringens phospholipase C gene in a low temperaturedependent manner. EMBO J 1999;18:3442–50. Kirstein J, Dougan DA, Gerth U et al. The tyrosine kinase McsB is a regulated adaptor protein for ClpCP. EMBO J 2007;26:2061–70. Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 571 Kirstein J, Z ¨uhlke D, Gerth U et al. A tyrosine kinase and its activator control the activity of the CtsR heat shock repressor in B. subtilis. EMBO J 2005;24:3435–45. Klinkert B, Narberhaus F. Microbial thermosensors. Cell Mol Life Sci 2009;66:2661–76. Kojima K, Nakamoto H. A novel light- and heat-responsive regulation of the groE transcription in the absence of HrcA or CIRCE in cyanobacteria. FEBS Lett 2007;581:1871–80. Koo BM, Rhodius VA, Campbell EA et al. Dissection of recognition determinants of Escherichia coli sigma32 suggests a composite -10 region with an ’extended -10 motif and a core -10 element. Mol Microbiol 2009;72:815–29. Kortmann J, Narberhaus F. Bacterial RNA thermometers: molecular zippers and switches. Nat Rev Microbiol 2012;10:255–65. Kortmann J, Sczodrok S, Rinnenthal J et al. Translation on demand by a simple RNA-based thermosensor. Nucleic Acids Res 2011;39:2855–68. Kourennaia OV, Tsujikawa L, Dehaseth PL. Mutational analysis of Escherichia coli heat shock transcription factor sigma 32 reveals similarities with sigma 70 in recognition of the -35 promoter element and differences in promoter DNA melting and -10 recognition. J Bacteriol 2005;187:6762–9. Kr ¨uger E, Hecker M. The first gene of the Bacillus subtilis clpC operon, ctsR, encodes a negative regulator of its own operon and other class III heat shock genes. J Bacteriol 1998;180: 6681–8. Kr ¨uger E, Msadek T, Hecker M. Alternate promoters direct stressinduced transcription of the Bacillus subtilis clpC operon. Mol Microbiol 1996;20:713–23. Kr ¨uger E, Msadek T, Ohlmeier S et al. The Bacillus subtilis clpC operon encodes DNA repair and competence proteins. Microbiology 1997;143 (Pt 4):1309–16. Kr ¨uger E, Z ¨uhlke D, Witt E et al. Clp-mediated proteolysis in Gram-positive bacteria is autoregulated by the stability of a repressor. EMBO J 2001;20:852–63. Kumar A, Grimes B, Logan M et al. A hybrid sigma subunit directs RNA polymerase to a hybrid promoter in Escherichia coli. J Mol Biol 1995;246:563–71. Kurata H, El-Samad H, Iwasaki R et al. Module-based analysis of robustness tradeoffs in the heat shock response system. PLoS Comput Biol 2006;2:e59. Landick R, Vaughn V, Lau ET et al. Nucleotide sequence of the heat shock regulatory gene of E. coli suggests its protein product may be a transcription factor. Cell 1984;38:175–82. Laport MS, Lemos JA, Bastos Md Mdo C et al. Transcriptional analysis of the groE and dnaK heat-shock operons of Enterococcus faecalis. Res Microbiol 2004;155:252–8. Lesley SA, Thompson NE, Burgess RR. Studies of the role of the Escherichia coli heat shock regulatory protein sigma 32 by the use of monoclonal antibodies. J Biol Chem 1987;262:5404–7. Li M, Wang SL. Cloning and characterization of the groESL operon from Bacillus subtilis. J Bacteriol 1992;174:3981–92. Liberek K, Galitski TP, Zylicz M et al. The DnaK chaperone modulates the heat shock response of Escherichia coli by binding to the sigma 32 transcription factor. P Natl Acad Sci USA 1992;89:3516–20. Lim B, Miyazaki R, Neher S et al. Heat shock transcription factor σ32 co-opts the signal recognition particle to regulate protein homeostasis in E. coli. PLoS Biol 2013;11:e1001735. Lima S, Guo MS, Chaba R et al. Dual molecular signals mediate the bacterial response to outer-membrane stress. Science 2013;340:837–41. Liu J, Huang C, Shin DH et al. Crystal structure of a heat-inducible transcriptional repressor HrcA from Thermotoga maritima: structural insight into DNA binding and dimerization. J Mol Biol 2005;350:987–96. Loh E, Kugelberg E, Tracy A et al. Temperature triggers immune evasion by Neisseria meningitidis. Nature 2013;502:237–40. L´opez-Garc´ıa P, Forterre P. DNA topology in hyperthermophilic archaea: reference states and their variation with growth phase, growth temperature, and temperature stresses. Mol Microbiol 1997;23:1267–79. Lund PA. Multiple chaperonins in bacteria–why so many? FEMS Microbiol Rev 2009;33:785–800. Madrid C, Nieto JM, Paytubi S et al. Temperature- and H-NS-dependent regulation of a plasmid-encoded virulence operon expressing Escherichia coli hemolysin. J Bacteriol 2002;184:5058–66. Matuszewska M, Kuczy ´nska-Wi´snik D, Laskowska E et al. The small heat shock protein IbpA of Escherichia coli cooperates with IbpB in stabilization of thermally aggregated proteins in a disaggregation competent state. J Biol Chem 2005;280:12292–8. Mazodier P, Guglielmi G, Davies J et al. Characterization of the groEL-like genes in Streptomyces albus. J Bacteriol 1991;173:7382–6. Mecsas J, Rouviere PE, Erickson JW et al. The activity of sigma E, an Escherichia coli heat inducible sigma-factor, is modulated by expression of outer membrane proteins. Genes Dev 1993;7(12B):2618–28. Missiakas D, Mayer MP, Lemaire M et al. Modulation of the Escherichia coli σE (RpoE) heat- shock transcription-factor activity by the RseA, RseB and RseC proteins. Mol Microbiol 1997;24:355–71. Missiakas D, Raina S. The extracytoplasmic function sigma factors: role and regulation. Mol Microbiol 1998;28:1059–66. Missiakas D, Schwager F, Betton JM et al. Identification and characterization of HsIV HsIU (ClpQ ClpY) proteins involved in overall proteolysis of misfolded proteins in Escherichia coli. EMBO J 1996;15:6899–909. Miyazaki R, Yura T, Suzuki T et al. A novel SRP recognition sequence in the homeostatic control region of heat shock transcription factor σ(32). Sci Rep 2016;6:24147. Mogk A, Homuth G, Scholz C et al. The GroE chaperonin machine is a major modulator of the CIRCE heat shock regulon of Bacillus subtilis. EMBO J 1997;16:4579–90. Mogk A, Tomoyasu T, Goloubinoff P et al. Identification of thermolabile Escherichia coli proteins: prevention and reversion of aggregation by DnaK and ClpB. EMBO J 1999;18:6934–49. Mogk A, V¨olker A, Engelmann S et al. Nonnative proteins induce expression of the Bacillus subtilis CIRCE regulon. J Bacteriol 1998;180:2895–900. Morita MT, Tanaka Y, Kodama TS et al. Translational induction of heat shock transcription factor sigma32: evidence for a builtin RNA thermosensor. Gene Dev 1999;13:655–65. Musatovova O, Dhandayuthapani S, Baseman JB. Transcriptional heat shock response in the smallest known self-replicating cell, Mycoplasma genitalium. J Bacteriol 2006;188:2845–55. Nair S, Derr´e I, Msadek T et al. CtsR controls class III heat shock gene expression in the human pathogen Listeria monocytogenes. Mol Microbiol 2000;35:800–11. Nakahigashi K, Ron EZ, Yanagi H et al. Differential and independent roles of a sigma(32) homolog (RpoH) and an HrcA repressor in the heat shock response of Agrobacterium tumefaciens. J Bacteriol 1999;181:7509–15. Nakamoto H, Suzuki M, Kojima K. Targeted inactivation of the hrcA repressor gene in cyanobacteria. FEBS Lett 2003;549: 57–62. Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 572 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 Nagai H, Yano R, Erickson JW et al. Transcriptional regulation of the heat shock regulatory gene rpoH in Escherichia coli: involvement of a novel catabolite-sensitive promoter. J Bacteriol 1990;172:2710–7. Nagai H, Yuzawa H, Yura T. Interplay of two cis-acting mRNA regions in translational control of sigma 32 synthesis during the heat shock response of Escherichia coli. P Natl Acad Sci USA 1991;88:10515–9. Narberhaus F. Negative regulation of bacterial heat shock genes. Mol Microbiol 1999;31:1–8. Narberhaus F, K¨aser R, Nocker A et al. A novel DNA element that controls bacterial heat shock gene expression. Mol Microbiol 1998;28:315–23. Nevozhay D, Adams RM, Murphy KF et al. Negative autoregulation linearizes the dose-response and suppresses the heterogeneity of gene expression. P Natl Acad Sci USA 2009;106:5123–8. Newlands JT, Gaal T, Mecsas J et al. Transcription of the Escherichia coli rrnB P1 promoter by the heat shock RNA polymerase (E sigma 32) in vitro. J Bacteriol 1993;175:661–8. Nickerson CA, Achberger EC. Role of curved DNA in binding of Escherichia coli RNA polymerase to promoters. J Bacteriol 1995;177:5756–61. Nocker A, Hausherr T, Balsiger S et al. A mRNA-based thermosensor controls expression of rhizobial heat shock genes. Nucleic Acids Res 2001;29:4800–7. Nonaka G, Blankschien M, Herman C et al. Regulon and promoter analysis of the E. coli heat-shock factor, sigma32, reveals a multifaceted cellular response to heat stress. Gene Dev 2006;20:1776–89. Obrist M, Narberhaus F. Identification of a turnover element in region 2.1 of Escherichia coli sigma32 by a bacterial one-hybrid approach. J Bacteriol 2005;187:3807–13. Ojha A, Anand M, Bhatt A et al. GroEL1: a dedicated chaperone involved in mycolic acid biosynthesis during biofilm formation in mycobacteria. Cell 2005;123:861–73. Pallen M. RpoN-dependent transcription of rpoH? Mol Microbiol 1999;31:393. Parijat P, Batra JK. Role of DnaK in HspR-HAIR interaction of Mycobacterium tuberculosis. IUBMB Life 2015;67:816–27. Patra M, Roy SS, Dasgupta R et al. GroEL to DnaK chaperone network behind the stability modulation of σ(32) at physiological temperature in Escherichia coli. FEBS Lett 2015;589: 4047–52. Perrody E, Cirinesi AM, Desplats C et al. A bacteriophage-encoded J-domain protein interacts with the DnaK/Hsp70 chaperone and stabilizes the heat-shock factor σ32 of Escherichia coli. PLoS Genet 2012;8:e1003037. Prosseda G, Falconi M, Giangrossi M et al. The virF promoter in Shigella: more than just a curved DNA stretch. Mol Microbiol 2004;51:523–37. Rajaram H, Chaurasia AK, Apte SK. Cyanobacterial heat-shock response: role and regulation of molecular chaperones. Microbiology 2014;160:647–58. Reischl S, Wiegert T, Schumann W. Isolation and analysis of mutant alleles of the Bacillus subtilis HrcA repressor with reduced dependency on GroE function. J Biol Chem 2002;277: 32659–67. Rhodius VA, Suh WC, Nonaka G et al. Conserved and variable functions of the sigmaE stress response in related genomes. PLoS Biol 2006;4:e2. Rinnenthal J, Klinkert B, Narberhaus F et al. Modulation of the stability of the Salmonella fourU-type RNA thermometer. Nucleic Acids Res 2011;39:8258–70. Ritossa F. A new puffing pattern induced by temperature shock and DNP in Drosophila. Experientia 1962;18:571–73. Roberts RC, Toochinda C, Avedissian M et al. Identification of a Caulobacter crescentus operon encoding hrcA, involved in negatively regulating heat-inducible transcription, and the chaperone gene grpE. J Bacteriol 1996;178:1829–41. Roncarati D, Danielli A, Scarlato V. CbpA acts as a modulator of HspR repressor DNA binding activity in Helicobacter pylori. J Bacteriol 2011;193:5629–36. Roncarati D, Danielli A, Scarlato V. The HrcA repressor is the thermosensor of the heat-shock regulatory circuit in the human pathogen Helicobacter pylori. Mol Microbiol 2014;92: 910–20. Roncarati D, Danielli A, Spohn G et al. Transcriptional regulation of stress response and motility functions in Helicobacter pylori is mediated by HspR and HrcA. J Bacteriol 2007a;189:7234–43. Roncarati D, Spohn G, Tango N et al. Expression, purification and characterization of the membrane-associated HrcA repressor protein of Helicobacter pylori. Protein Expr Purif 2007b;51:267–75. Rosenfeld N, Elowitz MB, Alon U. Negative autoregulation speeds the response times of transcription networks. J Mol Biol 2002;323:785–93. Scharf C, Riethdorf S, Ernst H et al. Thioredoxin is an essential protein induced by multiple stresses in Bacillus subtilis. J Bacteriol 1998;180:1869–77. Schmid AK, Howell HA, Battista JR et al. HspR is a global negative regulator of heat shock gene expression in Deinococcus radiodurans. Mol Microbiol 2005;55:1579–90. Schmidt A, Schiesswohl M, Volker U et al. Cloning, sequencing, mapping, and transcriptional analysis of the groESL operon from Bacillus subtilis. J Bacteriol 1992;174:3993–99. Schmidt G, Hertel C, Hammes WP. Molecular characterisation of the dnaK operon of Lactobacillus sakei LTH681. Syst Appl Microbiol 1999;22:321–8. Schulz A, Tzschaschel B, Schumann W. Isolation and analysis of mutants of the dnaK operon of Bacillus subtilis. Mol Microbiol 1995;15:421–9. Schumann W. The Bacillus subtilis heat shock stimulon. Cell Stress Chaperones 2003;8:207–17. Servant P, Grandvalet C, Mazodier P. The RheA repressor is the thermosensor of the HSP18 heat shock response in Streptomyces albus. P Natl Acad Sci USA 2000;97:3538–43. Servant P, Mazodier P. Heat induction of hsp18 gene expression in Streptomyces albus G: transcriptional and posttranscriptional regulation. J Bacteriol 1996;178:7031–6. Servant P, Rapoport G, Mazodier P. RheA, the repressor of hsp18 in Streptomyces albus G. Microbiology 1999;145:2385–91. Shapiro RS, Cowen LE. Thermal control of microbial development and virulence: molecular mechanisms of microbial temperature sensing. MBio 2012;3: e00238–12. Sharp MM, Chan CL, Lu CZ et al. The interface of sigma with core RNA polymerase is extensive, conserved, and functionally specialized. Gene Dev 1999;13:3015–26. Singh AK, Summerfield TC, Li H et al. The heat shock response in the cyanobacterium Synechocystis sp. Strain PCC 6803 and regulation of gene expression by HrcA and SigB. Arch Microbiol 2006;186:273–86. Singh R, Anil Kumar V, Das AK et al. A transcriptional corepressor regulatory circuit controlling the heat-shock response of Mycobacterium tuberculosis. Mol Microbiol 2014;94:450–65. Slabas AR, Suzuki I, Murata N et al. Proteomic analysis of the heat shock response in Synechocystis PCC6803 and a thermally Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 Roncarati and Scarlato 573 tolerant knockout strain lacking the histidine kinase 34 gene. Proteomics 2006;6:845–64. Slamti L, Livny J, Waldor MK. Global gene expression and phenotypic analysis of a Vibrio cholerae rpoH deletion mutant. J Bacteriol 2007;189:351–62. Spohn G, Danielli A, Roncarati D et al. Dual control of Helicobacter pylori heat shock gene transcription by HspR and HrcA. J Bacteriol 2004;186:2956–65. Spohn G, Scarlato V. The autoregulatory HspR repressor protein governs chaperone gene transcription in Helicobacter pylori. Mol Microbiol 1999;34:663–74. Stewart G, Wernisch L, Stabler R et al. Dissection of the heatshock response in Mycobacterium tuberculosis using mutants and microarrays. Microbiology 2002;148:3129–38. Straus D, Walter W, Gross CA. DnaK, DnaJ, and GrpE heat shock proteins negatively regulate heat shock gene expression by controlling the synthesis and stability of sigma 32. Gene Dev 1990;4:2202–9. Straus DB, Walter WA, Gross CA. The heat shock response of E. coli is regulated by changes in the concentration of sigma 32. Nature 1987;329:348–51. Suokko A, Poutanen M, Savijoki K et al. ClpL is essential for induction of thermotolerance and is potentially part of the HrcA regulon in Lactobacillus gasseri. Proteomics 2008;8: 1029–41. Suokko A, Savijoki K, Malinen E et al. Characterization of a mobile clpL gene from Lactobacillus rhamnosus. Appl Environ Microb 2005;71:2061–9. Susin MF, Perez HR, Baldini RL et al. Functional and structural analysis of HrcA repressor protein from Caulobacter crescentus. J Bacteriol 2004;186:6759–67. Suzuki H, Ikeda A, Tsuchimoto S et al. Synergistic binding of DnaJ and DnaK chaperones to heat shock transcription factor σ32 ensures its characteristic high metabolic instability: implications for heat shock protein 70 (Hsp70)-Hsp40 mode of function. J Biol Chem 2012;287:19275–83. Takano T, Kakefuda T. Involvement of a bacterial factor in morphogenesis of bacteriophage capsid. Nat New Biol 1972;239:34–7. Thieffry D, Huerta AM, P´erez-Rueda E et al. From specific gene regulation to genomic networks: a global analysis of transcriptional regulation in Escherichia coli. Bioessays 1998;20:433–40. Tomoyasu T, Gamer J, Bukau B et al. Escherichia coli FtsH is a membrane-bound, ATP-dependent protease which degrades the heat-shock transcription factor sigma 32. EMBO J 1995;14:2551–60. Tomoyasu T, Ogura T, Tatsuta T et al. Levels of DnaK and DnaJ provide tight control of heat shock gene expression and protein repair in Escherichia coli. Mol Microbiol 1998;30: 567–81. Tuominen I, Pollari M, Aguirre von Wobeser E et al. Sigma factor SigC is required for heat acclimation of the cyanobacterium Synechocystis sp. strain PCC 6803. FEBS Lett 2008;582: 346–50. Tuominen I, Pollari M, Tyystj¨arvi E et al. The SigB sigma factor mediates high-temperature responses in the cyanobacterium Synechocystis sp. PCC6803. FEBS Lett 2006;580: 319–23. van Bokhorst-van de Veen H, Bongers RS, Wels M et al. Transcriptome signatures of class I and III stress response deregulation in Lactobacillus plantarum reveal pleiotropic adaptation. Microb Cell Fact 2013;12:112. van der Veen S, Hain T, Wouters JA et al. The heat-shock response of Listeria monocytogenes comprises genes involved in heat shock, cell division, cell wall synthesis, and the SOS response. Microbiology 2007;153:3593–607. Varmanen P, Ingmer H, Vogensen FK. ctsR of Lactococcus lactis encodes a negative regulator of clp gene expression. Microbiology 2000;146:1447–55. Wade JT, Castro Roa D, Grainger DC et al. Extensive functional overlap between sigma factors in Escherichia coli. Nat Struct Mol Biol 2006;13:806–14. Waldminghaus T, Heidrich N, Brantl S et al. FourU: a novel type of RNA thermometer in Salmonella. Mol Microbiol 2007;65: 413–24. Waldminghaus T, Gaubig LC, Klinkert B et al. The Escherichia coli ibpA thermometer is comprised of stable and unstable structural elements. RNA Biol 2009;6:455–63. Walsh NP, Alba BM, Bose B et al. OMP peptide signals initiate the envelope-stress response by activating DegS protease via relief of inhibition mediated by its PDZ domain. Cell 2003;113:61–71. Wang Q, Kaguni JM. A novel sigma factor is involved in expression of the rpoH gene of Escherichia coli. J Bacteriol 1989a;171:4248–53. Wang QP, Kaguni JM. dnaA protein regulates transcriptions of the rpoH gene of Escherichia coli. J Biol Chem 1989b;264:7338–44. Wawrzynow A, Banecki B, Zylicz M. The Clp ATPases define a novel class of molecular chaperones. Mol Microbiol 1996;21:895–9. Wetzstein M, Volker U, Dedio J et al. Cloning, sequencing and molecular analysis of the dnaK locus from Bacillus subtilis. J Bacteriol 1992;174:3300–10. Wiegert T, Schumann W. Analysis of a DNA-binding motif of the Bacillus subtilis HrcA repressor protein. FEMS Microbiol Lett 2003;223:101–6. Wilson AC, Tan M. Functional analysis of the heat shock regulator HrcA of Chlamydia trachomatis. J Bacteriol 2002;184: 6566–71. Wilson AC, Tan M. Stress response gene regulation in Chlamydia is dependent on HrcA-CIRCE interactions. J Bacteriol 2004;186:3384–91. Wilson AC, Wu CC, IIIYates JR et al. Chlamydial GroEL autoregulates its own expression through direct interactions with the HrcA repressor protein. J Bacteriol 2005;187:7535–42. Xu Z, Horwich AL, Sigler PB. The crystal structure of the asymmetric GroEL-GroES-(ADP)7 chaperonin complex. Nature 1997;388:741–50. Young DB, Garbe TR. Heat shock proteins and antigens of Mycobacterium tuberculosis. Infect Immun 1991;59:3086–93. Yuan G, Wong SL. Regulation of groE expression in Bacillus subtilis: the involvement of the sA-like promoter and the roles of the inverted repeat sequence (CIRCE). J Bacteriol 1995a;177:5427–33. Yuan G, Wong SL. Isolation and characterization of Bacillus subtilis groE regulatory mutants: evidence for orf39 in the dnaK operon as a repressor gene in regulating the expression of both groE and dnaK. J Bacteriol 1995b;177: 6462–8. Yuewei H, Haley FO, Raengpradub S et al. Transcriptomic and phenotypic analyses suggest a network between the transcriptional regulators HrcA and sigmaB in Listeria monocytogenes. Appl Environ Microb 2007;73:7981–91. Yura T, Guisbert E, Poritz M et al. Analysis of sigma32 mutants defective in chaperone-mediated feedback control reveals Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021 574 FEMS Microbiology Reviews, 2017, Vol. 41, No. 4 unexpected complexity of the heat shock response. P Natl Acad Sci USA 2007;104:17638–43. Yuzawa H, Nagai H, Mori H et al. Heat induction of sigma 32 synthesis mediated by mRNA secondary structure: a primary step of the heat shock response in Escherichia coli. Nucleic Acids Res 1993;21:5449–55. Zhao K, Liu M, Burgess RR. The global transcriptional response of Escherichia coli to induced sigma 32 protein involves sigma 32 regulon activation followed by inactivation and degradation of sigma 32 in vivo. J Biol Chem 2005;280:17758–68. Zomer A, Fernandez M, Kearney B et al. An interactive regulatory network controls stress response in Bifidobacterium breve UCC2003. J Bacteriol 2009;191:7039–49. Zomer A, van Sinderen D. Intertwinement of stress response regulons in Bifidobacterium breve UCC2003. Gut Microbes 2010;1:100–2. Zuber U, Schumann W. CIRCE, a novel heat shock element involved in regulation of heat shock operon dnaK of Bacillus subtilis. J Bacteriol 1994;176: 1359–63. Downloadedfromhttps://academic.oup.com/femsre/article/41/4/549/3574666bygueston14September2021