Next Article in Journal
Antimicrobial Resistance and Phylogenetic Relatedness of Salmonella Serovars in Indigenous Poultry and Their Drinking Water Sources in North Central Nigeria
Previous Article in Journal
Fungi-Derived Bioactive Compounds as Potential Therapeutic Agents for Pancreatic Cancer: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Resuscitation Promotion Factor: A Pronounced Bacterial Cytokine in Propelling Bacterial Resuscitation

School of Light Industry Science and Engineering, Beijing Technology and Business University, Beijing 100048, China
*
Author to whom correspondence should be addressed.
Microorganisms 2024, 12(8), 1528; https://doi.org/10.3390/microorganisms12081528
Submission received: 6 July 2024 / Revised: 19 July 2024 / Accepted: 24 July 2024 / Published: 25 July 2024
(This article belongs to the Section Molecular Microbiology and Immunology)

Abstract

:
While confronted with unfavorable growth conditions, bacteria may transform into the dormant state, such as viable but nonculturable (VBNC) state, which is a reversible state characterized by low metabolic activity and lack of division. These dormant cells can be reactivated through the influence of the resuscitation promoting factor (Rpf) family, which are classified as autocrine growth factors and possess peptidoglycan hydrolase activities. To date, with the significant resuscitation or growth promotion ability of Rpf, it has been extensively applied to increasing bacterial diversity and isolating functional microbial species. This review provides a comprehensive analysis of the distribution, mode of action, and functional mechanisms of Rpf proteins in various bacterial species. The aim is to create opportunities for decoding microbial communities and extracting microbial resources from real samples across different research fields.

1. Introduction

Bacteria are masters of survival under unfavorable growth conditions. Many bacteria have developed the capacity to enter a non-replicative state with diminished metabolic activities, enabling them to endure a wide range of adverse environments. Such non-replicating states encompass persisters and viable but not culturable (VBNC) bacteria, as well as dormant spores [1,2]. Despite the diverse forms adopted by these non-replicating cells, they often possess a thicker (e.g., increased cross-linking, shortened glycan strands of peptidoglycan) cell wall in order to enhance their robustness to withstand external stress [3,4,5,6]. In the context of the non-replicative state, we employ the term “dormancy” to denote a reversible condition characterized by reduced metabolic activity, allowing a cell to endure prolonged periods without undergoing division [1,7,8]. It has been proposed that dormant cells have the potential to be reactivated into typical, colony-forming bacteria through a specific resuscitation procedure [3,9,10].
Peptidoglycan is present in almost all bacteria cell walls, providing shape, rigidity, and osmotic stability, which serves as a robust protective barrier for bacteria [11]. All peptidoglycans produced by bacteria share the same basic core structure: glycan backbones consisting of alternating N-acetylmuramic acid and N-acetylglucosamine residues, as well as short peptides extending from the lactyl groups of the N-acetylmuramic acid residues [12]. The composition of peptide stems and cross-links exhibits variation across different bacterial species. In Gram-negative species, such as Escherichia coli, the peptide sequence is L-Ala-D-Glu-m-Dap-D-Ala. Peptides most often connect from the carboxyl group of D-Ala at position 4 of one peptide to the amino group of the m-Dap residue at position 3 of another peptide (3–4 cross-link). In Gram-positive species, such as Micrococcus luteus, the peptide stem is L-Ala-D-Glu-L-Lys-D-Ala, which connects with another one through a 3–4 cross-link with a bridge consisting of a peptide stem [13]. Specifically, in some special Gram-positive bacteria, such as mycobacteria, the glycan backbone comprises N-acetylglucosamine and both N-acetylmuramic acid and N-glycolylmuramic acid moieties, and the stem peptide consists of D-iGln-m-DAP-D-Ala-L-Ala residues with modifications. The cross-link for mycobacterial is a 3–3 cross-link [13]. These peptide bridges cross-link parallel strands to provide a mesh-like structure, for maintaining the integrity of the cell membrane [13]. In dormant cells, peptidoglycan exhibits relatively low reactivity since bacteria stop dividing, while in actively growing cells it displays high turnover [14]. Many dormant cells necessitate the degradation of their robust protective cell wall in order to reinitiate active growth, and various bacteria have developed distinct strategies to accomplish this process. Among them, the so-called “resuscitation promoting factor” (Rpf) has garnered extensive attention.
The Rpf protein was initially identified in M. luteus as a bacterial cytokine that facilitates the revival and proliferation of quiescent or dormant cells [15,16]. The resuscitation-promoting ability of Rpf is profound. It was reported that the presence of Rpf protein at a picomolar concentration could increase the number of viable dormant M. luteus cells by at least 100-fold [16]. As a muralytic enzyme, similar proteins are widely distributed among other high G+C Gram-positive bacteria, including corynebacteria, mycobacteria, streptomycetes, and some fermicutes (which contain Rpf analogues) [17]. Up to now, Rpfs (largely derived from M. luteus) have been extensively employed to promote bacterial growth or resuscitation (not only for M. luteus itself, but also for other species), primarily for enhancing microbial diversity and isolating hard-to-culture species [18,19,20], although their functional mechanisms remain incompletely elucidated.
In this review, different aspects of the Rpf proteins including their basic characteristics, applications, structural properties, enzymatic activities, and functional mechanisms have been comprehensively reviewed, aiming to provide updated and in-depth references for researchers and establish a theoretical foundation for decoding the “dark matter” (the dormant majority) of microbial communities in the real world.

2. Discovery, Diversity, and Application of Rpf

In natural environments, a large proportion of bacteria are in the dormancy state defined as “a reversible state with low metabolic activity, allowing cells to persist for prolonged periods without undergoing division” [2,21]. For nonsporulating bacteria, it may be considered dormancy regarding the VBNC state [7,9,22]. For such instances, a common feature is their inability to generate colonies on their routine culture media, but dormant cells can be converted to normal, colony-forming bacteria following a special procedure of resuscitation involving cultivation in liquid media, which is termed as “resuscitation” [23,24,25]. Initially, in 1985, Colwell et al. first documented the phenomenon that the growing M. luteus entered a quiescent state followed by reactivation of growth [26]. Subsequently, in 1994, Votyakova et al. found that resuscitation and cell growth of starved M. luteus cells with lost culturability were favored by the use of the mixture of fresh lactate medium and supernatant from late-logarithmic-phase M. luteus cultures as the resuscitation medium, thus suggesting that this recovery is due to the excretion of some factor(s) that promoted the transition of cells from a state in which they are incapable of growth and division to the one of normal state [27]. Later, Mukamolva et al. discovered that M. luteus possessed the ability to secrete a factor capable of resuming its growth during its cultivation in the logarithmic growth phase [15]. The protein was then successfully isolated and designated as Rpf [16]. In picomolar concentrations, this protein significantly enhances the viable cell count of dormant M. luteus cultures by at least 100-fold and also promotes the growth of viable cells, as well as other high G+C Gram-positive organisms within Mycobacterium sp. [16]. Hence, the identification of Rpf and elucidation of its activity signifies a significant milestone in the revival culture of dormant bacteria.
Rpf enzymes were originally defined on the basis of their resuscitation-promoting activities in Micrococcus and, later, in Mycobacterium. The remarkable attention garnered by Rpfs of Mycobacterium tuberculosis stems from their significant implications in both the development and treatment of tuberculosis [28,29]. In subsequent research, Mukamolva et al. identified the widespread presence of the gene responsible for Rpf protein production in other high G+C Gram-positive actinomycetes including Corynebacterium sp., Streptomyces sp., Rhodococcus sp., Nocardiopsis sp., Curtobacterium sp., and so on [17,30], thus revealing the existence of the Rpf protein family [16,31]. In addition, other Rpf analogues in certain Gram-negative/-positive species have also been reported to exhibit resuscitation-promoting capabilities. For example, the YeaZ protein in Vibrio parahaemolyticus, V. harveyi, Salmonella typhimurium, and Escherichia coli has been shown to have promoting effects on bacterial VBNC-state recovery [32,33,34,35]; Lmo0186 and Lmo2522 proteins in Listeria monocytogenes are functional equivalents of Rpf, both of which display muralytic activity against crude cell-wall preparations and can reduce lag phase [36]; the recombinant Rpf-like protein from Achromobacter sp. HR2 exhibited muralytic activity and demonstrated significant efficacy in resuscitating the growth of VBNC cells, as well as stimulating the growth of normal cells [37]. Overall, the majority of studies have primarily focused on investigating the resuscitation or growth promotion potential of M. luteus Rpf, owing to its well-established clarity and effectiveness of action over the past few decades [38,39,40]. In recent years, there has been a noticeable emergence of Rpfs from novel species, exhibiting significant effects on their activity and growth/resuscitation ability [10,41]. By employing such methods, the sources of Rpf can be expanded, but further investigation is needed, through which additional characteristics can be elucidated, thereby enhancing their practical applicability.
With regard to applications of Rpf, extensive research has been conducted on the bacterial resuscitation/growth-promotion ability of Rpfs from different species, as well as their functional mechanisms. The promotion effects at the initial stage were specifically targeted toward individual bacteria [28,38,42,43,44]. To date, there has been a focus on the application of Rpf to enhance bacterial diversity and obtain more functional species from various real environmental samples. To elucidate the bacterial community structure, sequence-based techniques such as 16S rRNA sequencing and metagenomic analysis were commonly employed, referred to as non-culture-based methods [20]. Occasionally, these methods were complemented with culturomics analysis to comprehensively illustrate the promotion effects of Rpf. On the one hand, specific bacterial genera were enriched upon Rpf treatment, many of which were associated with bioremediation in polluted environmental samples [18,19,45,46]. On the other hand, the resuscitation ability of Rpf led to the isolation of species with specific functions or novel species. For example, 51 potentially novel bacterial species were isolated from soil [47]; 2 rare actinobacteria were also obtained from soil after Rpf treatment [48]; Shi et al. isolated Achromobacter sp. HR2 that exhibited effective degradation of Aroclor 1242 [49]. Collectively, the application of Rpfs and their positive effects are presented in Table 1.

3. Structural Characteristics of Rpf Protein Family

Do the Rpfs, with their diversity, exhibit similar functionality? Given the close correlation between protein function and structure, a phylogenetic analysis was firstly performed on Rpfs from major bacterial genera (Figure 1), based on their primary structural characteristics. The protein sequences of Rpfs exhibit significant divergence across different genera (Figure 1). Among these, RpfA and RpfD from Streptomyces sp. exhibit the highest degree of similarity, forming a distinct clade with M. luteus, implying that these proteins demonstrate relatively greater sequence homology (Figure 1). Additionally, it is a prevalent occurrence for bacteria to contain multiple Rpf proteins, as observed in species such as Corynebacterium glutamicum, M. tuberculosis, and Streptomyces sp. The protein sequences of Rpfs from a specific strain can exhibit significant variation, as seen in the case of Rpf1 and Rpf2 in C. glutamicum, Rpf B/C/E in Streptomyces sp. SID7813, and RpfA-E in M. tuberculosis (Figure 1). This suggests that different Rpfs within a specific microorganism may possess distinct spatial structures, thereby potentially playing diverse roles in promoting bacterial resuscitation/growth.
While the primary structures of Rpf from different species exhibit significant variation, their active domains share some common characteristics. The majority of Rpfs possess a conserved domain with specific functional attributes. Structural prediction of this domain suggests that it primarily consists of a lysozyme-like domain (c-type lysozyme fold), LysM domain, and G5 domain (Table 2). The presence of a lysozyme-like domain suggests that Rpfs may have lysozyme-like activity or even bind saccharides [62]. The LysM modules consist of 43–50 amino acids that adopt a highly conserved βααβ-fold, with particularly high sequence conservation in the first 16 residues [63]. Prokaryotic LysM modules bind peptidoglycan, the main component of the bacterial cell wall [64]. Similarly, the G5 domain may confer localization or substrate specificity on the proteins, which is typically present in proteins involved in bacterial cell-wall metabolism [65]. Taking these factors into consideration, it is possible that Rpf could potentially stimulate bacterial resuscitation or growth by facilitating the hydrolysis of cell-wall peptidoglycan.
Experimental investigations into the active domains of Rpfs have provided additional validation for the predicted outcomes. Cohen-Gonsaud et al. reported on the solution structure of the Rpf domain from M. tuberculosis Rv1009 (RpfB) solved by heteronuclear multidimensional Nuclear Magnetic Resonance (NMR) [66]. Results revealed that the Rpf domain is a compact hybrid of the soluble lytic transglycosidase and c-type lysozyme folds, both of which cleave peptidoglycan. A few crucial features such as the catalytic glutamate are conserved, which is critical for resuscitation activity [66]. It has also been found in a study on the structure of the conservative domain RpfB from M. tuberculosis by NMR that the conservative domain has a spatial organization similar to the tertiary structure of lysozyme C [67,68]. The conservative domain of Rpf is characterized by six α-helices and one β-sheet, which are spatially arranged in a manner identical to the α-helical regions of lysozyme C and certain lytic enzymes involved in bacterial cell-wall metabolism [67,68]. Subsequently, X-ray crystallographic analysis revealed that the catalytic domain of RpfB retains all essential catalytic residues for lysozyme activity and exhibits six peptidoglycan-binding sites, akin to various types of lysozymes. As a result, the catalytic domain was characterized as a “mini-lysozyme” [69]. Additionally, the 3D structural information of Rpfs is also presented in Table 2, which displays high diversity. Hence, structural studies of Rpf unequivocally demonstrate their classification as peptidoglycan hydrolases, implying that potential enzymatic activity may be crucial for the observed physiological impacts.

4. Enzymatic Activity of Rpf

Based on the structural characteristics of Rpfs, their catalytic domains demonstrate their similarity to lysozyme and to lytic transglycosylases—the groups of enzymes that cleave the β-1,4-glycosidic bond between N-acetylmuramic acid and N-acetylglucosamine. In order to elucidate the functional mode of the proteins, experimental assays were performed to assess their enzymatic activities.
As a founding member of the Rpf family, M. luteus Rpf was proved to be a muralytic enzyme, as revealed by its ability to cause lysis of E. coli, release fluorescent material from fluorescamine-labeled cell walls of M. luteus and hydrolyze the artificial lysozyme substrate of 4-methylumbelliferyl-β-D-N,N′,N″-triacetylchitotrioside while expressed and secreted into the periplasm [70]. It has been demonstrated that substitution of the putative catalytic glutamate in Rpf protein from M. luteus with glutamine (E54Q) resulted in only a partial loss of activity [38,42,68], while substitution with alanine (E54A) or lysine (E54K) effectively suppressed the enzymatic activity [70]. The activity was significantly inhibited in a mutant Rpf with two substituted cysteine residues that could potentially contribute to the formation of a functionally crucial intramolecular disulfide bond, but the substitution of either Cys53 or Cys114 separately resulted in only a partial decrease in activity [66,70].
In Rhodococcus sp., the RpfB protein of Rhodococcus sp. (GX12401) had the highest homology with M. tuberculosis RpfB [40]. When 4-methylumbelliferyl-β-D-N,N′,N″-triacetylchitoside was employed as the enzyme substrate for assessing lysozyme activity, it was observed that the recombinant protein RpfB exhibited robust stability and enzymatic efficacy [40]. In the case of R. erythropolis KB1, Rpf-1 exhibited muralytic and weak protease activities, with the former being responsible for its growth promotion and resuscitation effects [71]. Moreover, substitution of the conserved amino acid residue Gln69 resulted in a significant reduction in muralytic activity, highlighting the pivotal role of Gln69 in enzyme function [71].
For Streptomyces coelicolor, Rpfs (A–E) were found to promote spore germination and are critical for growth under nutrient-limiting conditions, and in vitro biochemical assays revealed various levels of peptidoglycan cleavage capabilities for each Rpf [72]. The two domains of LysM and LytM, typically existing in lysozyme and lytic transglycosylase, respectively, share the same substrate specificity [73]. However, lysozyme catalyzes a hydrolysis reaction that requires water to break their cognate glycosidic bonds, and lytic transglycosylase cleaves peptidoglycan with the concomitant formation of an intramolecular 1,6-anhydromuramoyl product [13,73]. Sexton et al. (2020) conducted an assessment of the activity of all five Rpfs from S. coelicolor (RpfA–E) by hydrolyzing M. luteus peptidoglycan suspended in 18O-labeled water to specify their activities. Upon identification of the products, the results indicated that RpfA/B/C/E each functioned as lytic transglycosylases rather than muramidases or β-N-acetylglucosaminidases [74]. Additionally, a mutagenesis assay revealed that their LysM and LytM domains enhanced the enzyme activity of Rpf [74].

5. Functional Mechanism of Rpf

5.1. Rpf Functions with Its Partner RipA

The structural characteristics and biochemical activities described above suggest that the enzymatic hydrolysis of peptidoglycan, facilitated by Rpf, may play a crucial role in promoting bacterial growth and resuscitation. Numerous enzymes are involved in the synthesis and degradation of peptidoglycan, working in conjunction with other enzymes [75]. Through screening in a yeast two-hybrid system, Hett et al. discovered that the function of RpfB from M. tuberculosis is associated with an endopeptidase RipA (resuscitation-promoting factor-interacting protein) [76]. RipA is a secreted, cell-associated protein, which localizes to the septa of the bacterial cell and is indeed a peptidoglycan hydrolase that is capable of digesting cell-wall material [76]. Deletion of the gene encoding RipA was also demonstrated which resulted in a reduction in growth rate and the development of anomalous morphology, such as branching and formation of chains, in M. tuberculosis and M. smegmatis cells [77]. These findings suggest that RipA is likely to play a crucial role during the final stage of cell division [78]. It is common for certain bacterial species to contain multiple Rpf proteins; however, not all of these Rpfs have been thoroughly investigated. Therefore, the subsequent sections will primarily focus on the already studied Rpfs. In terms of enzymatic activities of the two proteins, RpfB is known to cleave the glycoside bonds of the cell-wall peptidoglycan [66], while RipA is predicted to function as an endopeptidase that cleaves the stem peptide [D-iGlu-meso-diaminopimelic acid (Dap)] [79,80]. The interaction between RipA and RpfB at the septum of dividing cells has been shown to positively regulate peptidoglycan hydrolysis, as evidenced by the synergistic degradation of PG in vitro by the RipA–RpfB complex (Figure 2a) [76,79]. RpfE can also interact with RipA [75].

5.2. Rpf May Participate in the Remodeling of the Bacterial Cell Wall to Stimulate Growth or Resuscitation

In harsh environments, bacteria often modify their cell-wall structure as a means of adaptation. For example, actinomycetes tend to produce dormant exospores with thick cell walls to withstand adverse factors, resulting in difficulty germinating and growing under laboratory conditions [70,79]; VBNC-state bacteria may encounter a series of peptidoglycan modifications such as high degree of cross-linking, an increase in DAP–DAP cross-linking, and a shortening of the average length of glycan strands, in comparison with dividing cells, resulting in a more robust cell-wall structure [5,6,81]. In such conditions, a return to active growth requires the breakdown of the thick protective cell wall facilitated by Rpf proteins. The mechanism of action of Rpf remains a topic of debate. The preceding sections have indicated that the Rpf and RipA complex has the potential to function during peptidoglycan hydrolysis. Consequently, some researchers thought that the growth promotion or resuscitation stimulation capability of Rpf may be attributed to the remodeling of the bacterial cell wall. It was emphasized that Rpf activity in S. coelicolor was not correlated with peptidoglycan-responsive Ser/Thr kinases for cell signaling, which thus suggested that in Streptomyces, Rpfs had a structural rather than signaling function during spore germination [72,74]. Some Rpfs exhibit lytic transglycosylase activity, which facilitates the cleavage of the glycan backbone to accommodate the insertion of new peptidoglycan units during growth and is implicated in cell-wall remodeling processes such as envelope spanning (Figure 2b) [82,83].
As proteins related to the hydrolysis of cell-wall peptidoglycan, the remodeling of peptidoglycan by Rpfs must be rigorously regulated in order to maintain the structural integrity of the cell wall. In M. tuberculosis, the activity of RpfB–RipA is negatively impacted by its interactions with penicillin binding protein 1 (PBP1) [76]. The promotion of germination and growth of Streptomyces coelicolor is attributed to a unique mode of enzymatic autoregulation, which is mediated by a domain (DUF348) located within the N-terminus of the protein. Removal of this domain resulted in significantly enhanced peptidoglycan cleavage capacity. Therefore, it is speculated that the association between RipA and RpfB may induce a conformational change in RpfB, thereby alleviating the DUF348-mediated inhibition of enzyme activity [72].
Collectively, Rpf may function as a hydrolase, leading to restricted hydrolysis of the modified peptidoglycan in dormant cells. This process may promote cell-wall synthesis and growth, finally stimulating the initiation of the division process for resuscitation.

5.3. Rpf-Released Peptidoglycan Fragments Function as Signaling Molecules to Stimulate Growth/Resuscitation

The structure of a mycobacterial peptidoglycan synergistic hydrolysis product under the influence of RpfB-RipA comlex was identified to be N-acetylglucosaminyl-β(1→4)-N-glycolyl-1,6-anhydromuramoyl-L-alanyl-D-isoglutamate (anhydroMurNAc) [75]. It was further revealed that the disaccharide-dipeptide fragments represent the minimal structural unit capable of reviving dormant mycobacterial cells [75,84].
Whether Rpf activity promotes resuscitation by simply remodeling the dormant cell wall or by generating peptidoglycan fragments (muropeptides) that function as signaling molecules has been the subject of much debate. Regarding the stimulation mechanism of the peptidoglycan fragments, some researchers argue that bacteria can be revived from a non-growing state through the sensing of specific peptidoglycan fragments by signaling pathways [84]. For example, bacteria always apply eukaryotic-like serine/threonine protein kinases (STPKs) and associated serine/threonine protein phosphatase systems to bind and respond to peptidoglycan [85]. STPKs are monomeric transmembrane polypeptides with an extracellular sensor domain and an intracellular cytoplasmic protein kinase domain, with which phosphorylates target proteins upon recognition of a specific signal, leading to a coordinated cellular response [83]. The typical serine/threonine protein kinase found in Bacillus subtilis, PrkC, which has an external domain consisting of three repeat PASTA (penicillin-binding protein and serine/threonine kinase-associated) domain motifs, has been implicated in the regulation of growth through its modulation of sporulation and biofilm formation [86]. During active growth, it was shown that disaccharide-dipeptide fragments serve as potent germinants for B. subtilis spores, requiring the involvement of PrkC in peptidoglycan binding and the initiation of a signal cascade leading to spore germination [87].
In mycobacteria, PknB is the homologue of PrkC, which is the essential STPK [88]. PknB shares a high degree of structural similarity to other STPKs, with an intracellular kinase domain as well as an extracellular sensing domain containing four PASTA motifs, suggesting that this protein is also able to bind peptidoglycan [89]. Therefore, in conjunction with the biochemical role of Rpfs in the modification of cell-wall structure, it is conceivable that Rpfs and other enzymes synergistically function to generate disaccharide-dipeptide fragments as signaling molecules, which can bind to PknB or other PrkC homologues to trigger resuscitation in a manner akin to spore germination in B. subtilis [87]. However, several components of this mechanism are currently based on hypotheses, and further investigation is required to obtain more direct evidence in future studies.

6. Prospections

Here in this review, the diversity, application, structural characteristics, and functional mechanisms of Rpfs are comprehensively reviewed. Rpfs have garnered significant interest for their ability to enhance bacterial diversity and isolate “hidden” bacterial species, due to their functional feature of exerting a substantial effect at low concentrations. To date, the focus of Rpf treatment has primarily been on environmental samples. Broadening the research scope to include additional areas such as the food industry and agriculture may potentially open up a novel avenue, although it is essential to thoroughly assess any associated risks. Moreover, another important task for the future is the elucidation of the nature and action preference of the Rpfs, as well as the mechanisms of their action.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC), grant number 32302245.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wood, T.K.; Knabel, S.J.; Kwan, B.W. Bacterial persister cell formation and dormancy. Appl. Environ. Microbiol. 2013, 79, 7116–7121. [Google Scholar] [CrossRef] [PubMed]
  2. Oliver, J.D. Recent findings on the viable but nonculturable state in pathogenic bacteria. FEMS Microbiol. Rev. 2010, 34, 415–425. [Google Scholar] [CrossRef] [PubMed]
  3. Anuchin, A.M.; Mulyukin, A.L.; Suzina, N.E.; Duda, V.I.; El-Registan, G.I.; Kaprelyants, A.S. Dormant forms of Mycobacterium smegmatis with distinct morphology. Microbiology 2009, 155, 1071–1079. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, X.; Li, Y.; Bai, K.; Gu, M.; Xu, X.; Jiang, N.; Chen, Y.; Li, J.; Luo, L. Class A penicillin-binding protein c is responsible for stress response by regulation of peptidoglycan assembly in Clavibacter michiganensis. Microbiol. Spectr. 2022, 10, e01816-22. [Google Scholar] [CrossRef] [PubMed]
  5. Signoretto, C.; Lleò, M.D.M.; Canepari, P. Modification of the peptidoglycan of Escherichia coli in the viable but nonculturable state. Curr. Microbiol. 2002, 44, 125–131. [Google Scholar] [CrossRef]
  6. Dong, K.; Pan, H.; Yang, D.; Rao, L.; Zhao, L.; Wang, Y.; Liao, X. Induction, detection, formation, and resuscitation of viable but non-culturable state microorganisms. CRFSFS 2020, 19, 149–183. [Google Scholar] [CrossRef] [PubMed]
  7. Ayrapetyan, M.; Williams, T.; Oliver, J.D. Relationship between the viable but nonculturable state and antibiotic persister cells. J. Bacteriol. 2018, 200, 10–1128. [Google Scholar] [CrossRef]
  8. Shleeva, M.O.; Kudykina, Y.K.; Vostroknutova, G.N.; Suzina, N.E.; Mulyukin, A.L.; Kaprelyants, A.S. Dormant ovoid cells of Mycobacterium tuberculosis are formed in response to gradual external acidification. Tuberculosis 2011, 91, 146–154. [Google Scholar] [CrossRef]
  9. Oliver, J.D. The viable but nonculturable state in bacteria. J. Microbiol. 2005, 43, 93–100. [Google Scholar]
  10. Pan, H.; Wang, J.; Sun, Z.; Shi, R.; Ren, Q. Resuscitation of baijiu pit mud bacteria based on Rpf protein of Umezawaea beigongshangensis. Arch. Microbiol. 2024, 206, 184. [Google Scholar] [CrossRef]
  11. Schleifer, K.H.; Kandler, O.J.B.R. Peptidoglycan types of bacterial cell walls and their taxonomic implications. Bacteriol. Rev. 1972, 36, 407–477. [Google Scholar] [CrossRef] [PubMed]
  12. Jankute, M.; Cox, J.A.; Harrison, J.; Besra, G.S. Assembly of the mycobacterial cell wall. Annu. Rev. Microbiol. 2015, 69, 405–423. [Google Scholar] [CrossRef] [PubMed]
  13. Vollmer, W.; Blanot, D.; De Pedro, M.A. Peptidoglycan structure and architecture. FEMS Microbiol. Rev. 2008, 32, 149–167. [Google Scholar] [CrossRef]
  14. Cava, F.; de Pedro, M.A. Peptidoglycan plasticity in bacteria: Emerging variability of the murein sacculus and their associated biological functions. Curr. Opin. Microbiol. 2014, 18, 46–53. [Google Scholar] [CrossRef] [PubMed]
  15. Mukamolova, G.V.; Kaprelyants, A.S.; Kell, D.B. Secretion of an antibacterial factor during resuscitation of dormant cells in Micrococcus luteus cultures held in an extended stationary phase. Antonie Van Leeuwenhoek 1995, 67, 289–295. [Google Scholar] [CrossRef] [PubMed]
  16. Mukamolova, G.V.; Kaprelyants, A.S.; Young, D.I.; Young, M.; Kell, D.B. A bacterial cytokine. Proc. Natl. Acad. Sci. USA 1998, 95, 8916–8921. [Google Scholar] [CrossRef]
  17. Su, X.; Chen, X.; Hu, J.; Shen, C.; Ding, L. Exploring the potential environmental functions of viable but non-culturable bacteria. World J. Microbiol. Biotechnol. 2013, 29, 2213–2218. [Google Scholar] [CrossRef]
  18. Cai, J.; Liu, J.; Pan, A.; Liu, J.; Wang, Y.; Liu, J.; Sun, F.; Lin, H.; Chen, J.; Su, X. Effective decolorization of anthraquinone dye reactive blue 19 using immobilized Bacillus sp. JF4 isolated by resuscitation-promoting factor strategy. Water Sci. Technol. 2020, 81, 1159–1169. [Google Scholar] [CrossRef] [PubMed]
  19. Bodor, A.; Bounedjoum, N.; Feigl, G.; Duzs, A.; Laczi, K.; Szilágyi, A.; Rákhely, G.; Perei, K. Exploitation of extracellular organic matter from Micrococcus luteus to enhance ex situ bioremediation of soils polluted with used lubricants. J. Hazard. Mater. 2021, 417, 125996. [Google Scholar] [CrossRef]
  20. Guzman, J.; Raval, D.; Hauck, D.; Titz, A.; Poehlein, A.; Degenkolb, T.; Daniel, R.; Vilcinskas, A. The resuscitation-promoting factor (Rpf) from Micrococcus luteus and its putative reaction product 1, 6-anhydro-MurNAc increase culturability of environmental bacteria. Access Microbiol. 2023, 5, 000647-v4. [Google Scholar] [CrossRef]
  21. Ayrapetyan, M.; Williams, T.C.; Oliver, J.D. Bridging the gap between viable but non-culturable and antibiotic persistent bacteria. Trends Microbiol. 2015, 23, 7–13. [Google Scholar] [CrossRef] [PubMed]
  22. Pinto, D.; Santos, M.A.; Chambel, L. Thirty years of viable but nonculturable state research: Unsolved molecular mechanisms. Crit. Rev. Microbiol. 2015, 41, 61–76. [Google Scholar] [CrossRef] [PubMed]
  23. Whitesides, M.D.; Oliver, J.D. Resuscitation of Vibrio vulnificus from the viable but nonculturable state. Appl. Environ. Microbiol. 1997, 63, 1002–1005. [Google Scholar] [CrossRef]
  24. Mizunoe, Y.; Wai, S.N.; Ishikawa, T.; Takade, A.; Yoshida, S.I. Resuscitation of viable but nonculturable cells of Vibrio parahaemolyticus induced at low temperature under starvation. FEMS Microbiol. Lett. 2000, 186, 115–120. [Google Scholar] [CrossRef]
  25. Pinto, D.; Almeida, V.; Almeida Santos, M.; Chambel, L. Resuscitation of Escherichia coli VBNC cells depends on a variety of environmental or chemical stimuli. J. Appl. Microbiol. 2011, 110, 1601–1611. [Google Scholar] [CrossRef]
  26. Colwell, R.R.; Brayton, P.R.; Grimes, D.J.; Roszak, D.B.; Huq, S.A.; Palmer, L.M. Viable but non-culturable Vibrio cholerae and related pathogens in the environment: Implications for release of genetically engineered microorganisms. Nat. Biotechnol. 1985, 3, 817–820. [Google Scholar] [CrossRef]
  27. Votyakova, T.V.; Kaprelyants, A.S.; Kell, D.B. Influence of viable cells on the resuscitation of dormant cells in Micrococcus luteus cultures held in an extended stationary phase: The population effect. Appl. Environ. Microbiol. 1994, 60, 3284–3291. [Google Scholar] [CrossRef]
  28. Mukamolova, G.V.; Turapov, O.A.; Young, D.I.; Kaprelyants, A.S.; Kell, D.B.; Young, M. A family of autocrine growth factors in Mycobacterium tuberculosis. Mol. Microbiol. 2002, 46, 623–635. [Google Scholar] [CrossRef]
  29. Cole, S.T.; Brosch, R.; Parkhill, J.; Garnier, T.; Churcher, C.; Harris, D.; Gordon, S.V.; Eiglmeier, K.; Gas, S.; Barry, C.E., 3rd; et al. Deciphering the biology of Mycobacterium tuberculosis from the complete genome sequence. Nature 1998, 393, 537–544. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Liu, J.; Cao, M.; Zhang, Y.; Zhang, X. Role of resuscitation promoting factor-like protein from Nocardiopsis halophila. Microorganisms 2023, 11, 485. [Google Scholar] [CrossRef]
  31. Mukamolova, G.V.; Kormer, S.S.; Kell, D.B.; Kaprelyants, A.S. Stimulation of the multiplication of Micrococcus luteus by an autocrine growth factor. Arch. Microbiol. 1999, 172, 9–14. [Google Scholar] [CrossRef]
  32. Panutdaporn, N.; Kawamoto, K.; Asakura, H.; Makino, S.I. Resuscitation of the viable but non-culturable state of Salmonella enterica serovar Oranienburg by recombinant resuscitation-promoting factor derived from Salmonella Typhimurium strain LT2. Int. J. Food Microbiol. 2006, 106, 241–247. [Google Scholar] [CrossRef]
  33. Aydin, I.; Saijo-Hamano, Y.; Namba, K.; Thomas, C.; Roujeinikova, A. Structural analysis of the essential resuscitation promoting factor YeaZ suggests a mechanism of nucleotide regulation through dimer reorganization. PLoS ONE 2011, 6, e23245. [Google Scholar] [CrossRef]
  34. Handford, J.I.; Ize, B.; Buchanan, G.; Butland, G.P.; Greenblatt, J.; Emili, A.; Palmer, T. Conserved network of proteins essential for bacterial viability. J. Bacteriol. 2009, 191, 4732–4749. [Google Scholar] [CrossRef]
  35. Li, Y.; Chen, J.; Zhao, M.; Yang, Z.; Yue, L.; Zhang, X. Promoting resuscitation of viable but nonculturable cells of Vibrio harveyi by a resuscitation-promoting factor-like protein YeaZ. J. Appl. Microbiol. 2017, 122, 338–346. [Google Scholar] [CrossRef] [PubMed]
  36. Pinto, D.; Sao-Jose, C.; Santos, M.A.; Chambel, L. Characterization of two resuscitation promoting factors of Listeria monocytogenes. Microbiology 2013, 159, 1390–1401. [Google Scholar] [CrossRef] [PubMed]
  37. Shi, J.; Zhou, X.; Zhang, S.; Sun, F.; Shen, C.; Su, X. Unveiling the distribution characteristics of rpf-like genes and indigenous resuscitation promoting factor production in PCB-contaminated soils. J. Environ. Manag. 2024, 357, 120803. [Google Scholar] [CrossRef] [PubMed]
  38. Biketov, S.; Mukamolova, G.V.; Potapov, V.; Gilenkov, E.; Vostroknutova, G.; Kell, D.B.; Young, M.; Kaprelyants, A.S. Culturability of Mycobacterium tuberculosis cells isolated from murine macrophages: A bacterial growth factor promotes recovery. FEMS Immunol. Med. Microbiol. 2000, 29, 233–240. [Google Scholar] [CrossRef]
  39. Aktas, D.; Bagirova, M.; Allahverdiyev, A.M.; Abamor, E.S.; Safarov, T.; Kocazeybek, B.S. Resuscitation of the Helicobacter pylori coccoid forms by resuscitation promoter factor obtained from Micrococcus luteus. Curr. Microbiol. 2020, 77, 2093–2103. [Google Scholar] [CrossRef]
  40. Branysova, T.; Petru, N.; Marin, M.A.L.; Solcova, M.; Demnerova, K.; Stiborova, H. Uncovering the microbial diversity of Czech Republic archives: A study of metabolically active airborne microbes. Heliyon 2024, 10, e27930. [Google Scholar] [CrossRef]
  41. Gong, X.; Lu, H.; Wu, J.; Zhou, Y.; Yang, L.; Wang, Y.; Shen, N.; Jiang, M. Enzymatic properties and biological activity of resuscitation-promoting factor B of Rhodococcus sp. (GX12401). Front. Microbiol. 2022, 13, 965843. [Google Scholar] [CrossRef]
  42. Shleeva, M.O.; Bagramyan, K.; Telkov, M.V.; Mukamolova, G.V.; Young, M.; Kell, D.B.; Kaprelyants, A.S. Formation and resuscitation of ‘non-culturable’cells of Rhodococcus rhodochrous and Mycobacterium tuberculosis in prolonged stationary phase. Microbiology 2002, 465, 133242. [Google Scholar]
  43. Schroeckh, V.; Martin, K. Resuscitation-promoting factors: Distribution among actinobacteria, synthesis during life-cycle and biological activity. Antonie Van Leeuwenhoek 2006, 89, 359–365. [Google Scholar] [CrossRef]
  44. Hartmann, M.; Barsch, A.; Niehaus, K.; Pühler, A.; Tauch, A.; Kalinowski, J. The glycosylated cell surface protein Rpf2, containing a resuscitation-promoting factor motif, is involved in intercellular communication of Corynebacterium glutamicum. Arch. Microbiol. 2004, 182, 299–312. [Google Scholar] [CrossRef]
  45. Liu, Y.; Su, X.; Lu, L.; Ding, L.; Shen, C. A novel approach to enhance biological nutrient removal using a culture supernatant from Micrococcus luteus containing resuscitation-promoting factor (Rpf) in SBR process. Environ. Sci. Pollut. Res. Int. 2016, 23, 4498–4508. [Google Scholar] [CrossRef] [PubMed]
  46. Su, X.; Wang, Y.; Xue, B.; Zhang, Y.; Mei, R.; Zhang, Y.; Hashmi, M.; Lin, H.; Chen, J.; Sun, F. Resuscitation of functional bacterial community for enhancing biodegradation of phenol under high salinity conditions based on Rpf. Bioresour. Technol. 2018, 261, 394–402. [Google Scholar] [CrossRef]
  47. Lopez Marin, M.A.; Strejcek, M.; Junkova, P.; Suman, J.; Santrucek, J.; Uhlik, O. Exploring the potential of Micrococcus luteus culture supernatant with resuscitation-promoting factor for enhancing the culturability of soil bacteria. Front. Microbiol. 2021, 12, 685263. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, Y.; Shi, J.; Tang, L.; Zhang, Y.; Zhang, Y.; Wang, X.; Zhang, X. Evaluation of Rpf protein of Micrococcus luteus for cultivation of soil actinobacteria. Syst. Appl. Microbiol. 2021, 44, 126234. [Google Scholar] [CrossRef] [PubMed]
  49. Shi, J.; Yang, Y.; Zhang, S.; Lin, Q.; Sun, F.; Lin, H.; Shen, C.; Su, X. New insights into survival strategies and PCB bioremediation potential of resuscitated strain Achromobacter sp. HR2 under combined stress conditions. J. Hazard. Mater. 2024, 465, 133242. [Google Scholar] [CrossRef]
  50. Mukamolova, G.V.; Turapov, O.A.; Kazarian, K.; Telkov, M.; Kaprelyants, A.S.; Kell, D.B.; Young, M. The rpf gene of Micrococcus luteus encodes an essential secreted growth factor. Mol. Microbiol. 2002, 46, 611–621. [Google Scholar] [CrossRef]
  51. Su, X.; Shen, H.; Yao, X.; Ding, L.; Yu, C.; Shen, C. A novel approach to stimulate the biphenyl-degrading potential of bacterial community from PCBs-contaminated soil of e-waste recycling sites. Bioresour. Technol. 2013, 146, 27–34. [Google Scholar] [CrossRef] [PubMed]
  52. Li, Z.; Zhang, Y.; Wang, Y.; Mei, R.; Zhang, Y.; Hashmi, M.Z.; Lin, H.; Su, X. A new approach of Rpf addition to explore bacterial consortium for enhanced phenol degradation under high salinity conditions. Curr. Microbiol. 2018, 75, 1046–1054. [Google Scholar] [CrossRef] [PubMed]
  53. Su, X.; Xue, B.; Wang, Y.; Hashmi, M.Z.; Lin, H.; Chen, J.; Mei, R.; Wang, Z.; Sun, F. Bacterial community shifts evaluation in the sediments of Puyang River and its nitrogen removal capabilities exploration by resuscitation promoting factor. Ecotoxicol. Environ. Saf. 2019, 179, 188–197. [Google Scholar] [CrossRef] [PubMed]
  54. Ye, Z.; Li, H.; Jia, Y.; Fan, J.; Wan, J.; Guo, L.; Su, X.; Zhang, Y.; Wu, W.; Shen, C. Supplementing resuscitation-promoting factor (Rpf) enhanced biodegradation of polychlorinated biphenyls (PCBs) by Rhodococcus biphenylivorans strain TG9T. Environ. Pollut. 2020, 263, 114488. [Google Scholar] [CrossRef] [PubMed]
  55. Kuo, V.; Lehmkuhl, B.K.; Lennon, J.T. Resuscitation of the microbial seed bank alters plant-soil interactions. Mol. Ecol. 2021, 30, 2905–2914. [Google Scholar] [CrossRef] [PubMed]
  56. Su, X.; Li, S.; Xie, M.; Tao, L.; Zhou, Y.; Xiao, Y.; Lin, H.; Chen, J.; Sun, F. Enhancement of polychlorinated biphenyl biodegradation by resuscitation promoting factor (Rpf) and Rpf-responsive bacterial community. Chemosphere 2021, 263, 128283. [Google Scholar] [CrossRef]
  57. Han, Z.; Hou, A.; Cai, X.; Xie, M.; Sun, F.; Shen, C.; Lin, H.; Yu, H.; Su, X. Unlocking the potential of resuscitation-promoting factor for enhancing anaerobic microbial dechlorination of polychlorinated biphenyls. Sci. Total Environ. 2023, 897, 165440. [Google Scholar] [CrossRef] [PubMed]
  58. Su, X.; Xie, M.; Han, Z.; Xiao, Y.; Wang, R.; Shen, C.; Hashmi, M.; Sun, F. Resuscitation-promoting factor accelerates enrichment of highly active tetrachloroethene/polychlorinated biphenyl-dechlorinating cultures. Appl. Environ. Microbiol. 2023, 89, e0195122. [Google Scholar] [CrossRef]
  59. Gao, H.; Bai, Y.; Xue, Y.; Wang, L.; Fan, A.; Ding, X.; Xu, Z. Expression, purification, and characterization of soluble RpfD with high bioactivity as a recombinant protein in Mycobacterium vaccae. Protein Expr. Purif. 2007, 55, 112–118. [Google Scholar] [CrossRef]
  60. Puspita, I.D.; Uehara, M.; Katayama, T.; Kikuchi, Y.; Kitagawa, W.; Kamagata, Y.; Asano, K.; Nakatsu, C.H.; Tanaka, M. Resuscitation promoting factor (Rpf) from Tomitella biformata AHU 1821T promotes growth and resuscitates non-dividing cells. Microbes Environ. 2013, 28, 58–64. [Google Scholar] [CrossRef]
  61. Puspita, I.D.; Kitagawa, W.; Kamagata, Y.; Tanaka, M.; Nakatsu, C.H. Increase in bacterial colony formation from a permafrost ice wedge dosed with a Tomitella biformata recombinant resuscitation-promoting factor protein. Microbes Environ. 2015, 30, 151–156. [Google Scholar] [CrossRef]
  62. Cohen-Gonsaud, M.; Keep, N.H.; Davies, A.P.; Ward, J.; Henderson, B.; Labesse, G. Resuscitation-promoting factors possess a lysozyme-like domain. Trends Biochem. Sci. 2004, 29, 7–10. [Google Scholar] [CrossRef] [PubMed]
  63. Mesnage, S.; Dellarole, M.; Baxter, N.J.; Rouget, J.B.; Dimitrov, J.D.; Wang, N.; Fujimoto, Y.; Hounslow, A.M.; Lacroix-Desmazes, S.; Fukase, K.; et al. Molecular basis for bacterial peptidoglycan recognition by LysM domains. Nat. Commun. 2014, 5, 4269. [Google Scholar] [CrossRef] [PubMed]
  64. Bateman, A.; Bycroft, M. The structure of a LysM domain from E. coli membrane-bound lytic murein transglycosylase D (MltD). J. Mol. Biol. 2000, 299, 1113–1119. [Google Scholar] [CrossRef]
  65. Bateman, A.; Holden, M.T.; Yeats, C. The G5 domain: A potential N-acetylglucosamine recognition domain involved in biofilm formation. Bioinformatics 2005, 21, 1301–1303. [Google Scholar] [CrossRef]
  66. Cohen-Gonsaud, M.; Barthe, P.; Bagnéris, C.; Henderson, B.; Ward, J.; Roumestand, C.; Keep, N.H. The structure of a resuscitation-promoting factor domain from Mycobacterium tuberculosis shows homology to lysozymes. Nat. Struct. Mol. Biol. 2005, 12, 270–273. [Google Scholar] [CrossRef] [PubMed]
  67. Ruggiero, A.; Tizzano, B.; Geerlof, A.; Pedone, E.; Pedone, C.; Wilmanns, M.; Berisio, R. Expression, purification, crystallization and preliminary X-ray crystallographic analysis of a resuscitation-promoting factor from Mycobacterium tuberculosis. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2007, 63, 870–873. [Google Scholar] [CrossRef]
  68. Maione, V.; Ruggiero, A.; Russo, L.; De Simone, A.; Pedone, P.V.; Malgieri, G.; Berisio, R.; Isernia, C. NMR structure and dynamics of the resuscitation promoting factor RpfC catalytic domain. PLoS ONE 2015, 10, e0142807. [Google Scholar] [CrossRef]
  69. Ruggiero, A.; Tizzano, B.; Pedone, E.; Pedone, C.; Wilmanns, M.; Berisio, R. Crystal structure of the resuscitation-promoting factor (DeltaDUF)RpfB from M. tuberculosis. J. Mol. Biol. 2009, 385, 153–162. [Google Scholar] [CrossRef]
  70. Mukamolova, G.V.; Murzin, A.G.; Salina, E.G.; Demina, G.R.; Kell, D.B.; Kaprelyants, A.S.; Young, M. Muralytic activity of Micrococcus luteus Rpf and its relationship to physiological activity in promoting bacterial growth and resuscitation. Mol. Microbiol. 2006, 59, 84–98. [Google Scholar] [CrossRef]
  71. Luo, D.; Chen, J.; Xie, G.; Yue, L.; Wang, Y. Enzyme characterization and biological activities of a resuscitation promoting factor from an oil degrading bacterium Rhodococcus erythropolis KB1. PeerJ. 2019, 7, e6951. [Google Scholar] [CrossRef]
  72. Sexton, D.L.; St-Onge, R.J.; Haiser, H.J.; Yousef, M.R.; Brady, L.; Gao, C.; Leonard, J.; Elliot, M.A. Resuscitation-promoting factors are cell wall-lytic enzymes with important roles in the germination and growth of Streptomyces coelicolor. J. Bacteriol. 2015, 197, 848–860. [Google Scholar] [CrossRef] [PubMed]
  73. Scheurwater, E.; Reid, C.W.; Clarke, A.J. Lytic transglycosylases: Bacterial space-making autolysins. Int. J. Biochem. Cell Biol. 2008, 40, 586–591. [Google Scholar] [CrossRef]
  74. Sexton, D.L.; Herlihey, F.A.; Brott, A.S.; Crisante, D.A.; Shepherdson, E.; Clarke, A.J.; Elliot, M.A. Roles of LysM and LytM domains in resuscitation-promoting factor (Rpf) activity and Rpf-mediated peptidoglycan cleavage and dormant spore reactivation. J. Biol. Chem. 2020, 295, 9171–9182. [Google Scholar] [CrossRef]
  75. Nikitushkin, V.D.; Demina, G.R.; Shleeva, M.O.; Guryanova, S.V.; Ruggiero, A.; Berisio, R.; Kaprelyants, A.S. A product of RpfB and RipA joint enzymatic action promotes the resuscitation of dormant mycobacteria. FEBS J. 2015, 282, 2500–2511. [Google Scholar] [CrossRef] [PubMed]
  76. Hett, E.C.; Chao, M.C.; Steyn, A.J.; Fortune, S.M.; Deng, L.L.; Rubin, E.J. A partner for the resuscitation-promoting factors of Mycobacterium tuberculosis. Mol. Microbiol. 2007, 66, 658–668. [Google Scholar] [CrossRef] [PubMed]
  77. Ruggiero, A.; Marasco, D.; Squeglia, F.; Soldini, S.; Pedone, E.; Pedone, C.; Berisio, R. Structure and functional regulation of RipA, a mycobacterial enzyme essential for daughter cell separation. Structure 2010, 18, 1184–1190. [Google Scholar] [CrossRef]
  78. Nikitushkin, V.D.; Demina, G.R.; Kaprelyants, A.S. Rpf proteins are the factors of reactivation of the dormant forms of actinobacteria. Biochemistry 2016, 81, 1719–1734. [Google Scholar] [CrossRef] [PubMed]
  79. Hett, E.C.; Chao, M.C.; Deng, L.L.; Rubin, E.J. A mycobacterial enzyme essential for cell division synergizes with resuscitation-promoting factor. PLoS Pathog. 2008, 4, e1000001. [Google Scholar] [CrossRef]
  80. Squeglia, F.; Ruggiero, A.; Romano, M.; Vitagliano, L.; Berisio, R. Mutational and structural study of RipA, a key enzyme in Mycobacterium tuberculosis cell division: Evidence for the L-to-D inversion of configuration of the catalytic cysteine. Acta Crystallogr. D Biol. Crystallogr. 2014, 70, 2295–2300. [Google Scholar] [CrossRef]
  81. Signoretto, C.; del Mar Lleò, M.; Tafi, M.C.; Canepari, P. Cell wall chemical composition of Enterococcus faecalis in the viable but nonculturable state. Appl. Environ. Microbiol. 2000, 66, 1953–1959. [Google Scholar] [CrossRef] [PubMed]
  82. Vollmer, W.; Joris, B.; Charlier, P.; Foster, S. Bacterial peptidoglycan (murein) hydrolases. FEMS Microbiol. Rev. 2008, 32, 259–286. [Google Scholar] [CrossRef]
  83. Kana, B.D.; Mizrahi, V. Resuscitation-promoting factors as lytic enzymes for bacterial growth and signaling. FEMS Immunol. Med. Microbiol. 2010, 58, 39–50. [Google Scholar] [CrossRef]
  84. Nikitushkin, V.D.; Demina, G.R.; Shleeva, M.O.; Kaprelyants, A.S. Peptidoglycan fragments stimulate resuscitation of “non-culturable” mycobacteria. Antonie Van Leeuwenhoek 2013, 103, 37–46. [Google Scholar] [CrossRef]
  85. Shi, L.; Potts, M.; Kennelly, P.J. The serine, threonine, and/or tyrosine-specific protein kinases and protein phosphatases of prokaryotic organisms: A family portrait. FEMS Microbiol. Rev. 1998, 22, 229–253. [Google Scholar] [CrossRef]
  86. Madec, E.; Laszkiewicz, A.; Iwanicki, A.; Obuchowski, M.; Séror, S. Characterization of a membrane-linked Ser/Thr protein kinase in Bacillus subtilis, implicated in developmental processes. Mol. Microbiol. 2002, 46, 571–586. [Google Scholar] [CrossRef] [PubMed]
  87. Shah, I.M.; Laaberki, M.H.; Popham, D.L.; Dworkin, J. A eukaryotic-like Ser/Thr kinase signals bacteria to exit dormancy in response to peptidoglycan fragments. Cell 2008, 135, 486–496. [Google Scholar] [CrossRef] [PubMed]
  88. Fernandez, P.; Saint-Joanis, B.; Barilone, N.; Jackson, M.; Gicquel, B.; Cole, S.T.; Alzari, P.M. The Ser/Thr protein kinase PknB is essential for sustaining mycobacterial growth. J. Bacteriol. 2006, 188, 7778–7784. [Google Scholar] [CrossRef]
  89. Ortiz-Lombardıa, M.; Pompeo, F.; Boitel, B.; Alzari, P.M. Crystal structure of the catalytic domain of the PknB serine/threonine kinase from Mycobacterium tuberculosis. J. Biol. Chem. 2003, 278, 13094–13100. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic analysis of Rpfs from different bacteria. The phylogenetic tree was conducted based on amino acid sequences of Rpfs via Mega 11 software using the neighbor-joining method. Bootstrap values were based on 1000 replications. Bar, 0.2 means substitutions per nucleotide position.
Figure 1. Phylogenetic analysis of Rpfs from different bacteria. The phylogenetic tree was conducted based on amino acid sequences of Rpfs via Mega 11 software using the neighbor-joining method. Bootstrap values were based on 1000 replications. Bar, 0.2 means substitutions per nucleotide position.
Microorganisms 12 01528 g001
Figure 2. Proposed role of Rpf in remodeling of peptidoglycan and production of muropeptide signaling molecules to promote resuscitation. (a) Rpf functions with its partner RipA. Rpf is predicted to cleave the β-1,4 glycosidic bond between NAM and NAG in the glycan backbone, in a manner similar to lytic transglycosidases or lysozymes, whereas RipA cleaves bonds in the peptide stems. (b) Remodeling of bacterial cell wall to stimulate growth or resuscitation by Rpf. Rpf and RipA work synergistically to break down the thick protective cell wall for revival, which is followed by accommodation of the insertion of new peptidoglycan units. (c) Signaling molecules of peptidoglycan fragments released by Rpf-stimulating resuscitation. Muropeptide signaling molecules, the Rpf and RipA hydrolysis product, can bind to the extracellular sensor domain of STPKs including PrkC, PknB and then be phosphorylated to trigger resuscitation. All the arrows (black, red and white) represent the direction of the flow.
Figure 2. Proposed role of Rpf in remodeling of peptidoglycan and production of muropeptide signaling molecules to promote resuscitation. (a) Rpf functions with its partner RipA. Rpf is predicted to cleave the β-1,4 glycosidic bond between NAM and NAG in the glycan backbone, in a manner similar to lytic transglycosidases or lysozymes, whereas RipA cleaves bonds in the peptide stems. (b) Remodeling of bacterial cell wall to stimulate growth or resuscitation by Rpf. Rpf and RipA work synergistically to break down the thick protective cell wall for revival, which is followed by accommodation of the insertion of new peptidoglycan units. (c) Signaling molecules of peptidoglycan fragments released by Rpf-stimulating resuscitation. Muropeptide signaling molecules, the Rpf and RipA hydrolysis product, can bind to the extracellular sensor domain of STPKs including PrkC, PknB and then be phosphorylated to trigger resuscitation. All the arrows (black, red and white) represent the direction of the flow.
Microorganisms 12 01528 g002
Table 1. The utilization of Rpf proteins from diverse sources in bacterial resuscitation.
Table 1. The utilization of Rpf proteins from diverse sources in bacterial resuscitation.
Bacterial Origin of RpfExperimental SamplePromotive Effects
M. luteusMycobacterium tuberculosisRpf stimulates M. tuberculosis viability measured by the MPN method [35].
M. luteusPronounced influence on the true lag phase and cell growth on lactate minimal medium [50].
Polychlorinated
biphenyls-contaminated soils
The enrichment culture produced by the addition of SRpf significantly enhanced biphenyl degradation efficiency, cell growth, and bacterial diversity [51].
Activated sludgeThe phyla of Proteobacteria and Actinobacteria, which are closely related to biological nutrient removal, were greatly abundant after SRpf addition [45].
Activated sludgeResuscitated functional bacterial community for enhancing
biodegradation of phenol linked mainly to the genus Corynebacterium and the genera Proteiniphilum and Petrimonas [46].
Activated sludgeTwo strains from genera Bacillus and Corynebacterium could enhance phenol degradation under high salinity conditions [52].
Sediments from a nitrogen-polluted riverFrom URS and LRS, a total of thirteen strains with heterotrophic nitrification ability were resuscitated by Rpf addition, which belonged to genera Bacillus, Pseudomonas, Stenotrophomonas, and Acinetobacter [53].
Helicobacter pyloriShowed high resuscitation effect on kanamycin-induced H. pylori coccoid forms [36].
MSM medium with reactive
blue 19
Immobilized Bacillus sp. JF4 isolated by Rpf could effectively decolorize anthraquinone dye RB 19 [18].
Liquid culture with aroclor 1242Exogenous Rpf stimulated endogenous rpf expression in Rhodococcus biphenylivorans TG9T, leading to the resuscitation of VBNC TG9T cells, which finally contributed to its polychlorinated biphenyls-degrading capacity [54].
Soils polluted with used lubricant oils (ULOs)Bioremediation of ULOs was significantly enhanced, which was attributable both to the reactivation of hydrocarbonoclastic bacterial genera (e.g., Pseudomonas, Comamonas, Stenotrophomonas, and Gordonia) [19].
Soil containing field mustardrRpf reduced soil respiration, decreased bacterial abundance, and increased fungal abundance [55].
SoilA total of 51 potentially novel bacterial species were isolated after SRpf treatment [47].
Polychlorinated biphenyl (PCB)-contaminated soilsThe rRpf-responsive population was mostly represented by Sphingomonas and Pseudomonas, which are most likely the key PCB degraders [56].
SoilTwo rare actinobacteria were obtained in the experimental group supplemented with rRpf protein [48].
AirIncreased bacterial culturability by up to 30% when synergistically employed Reasoner’s 2A agar (R2A) and R2A + Rpf media [37].
Cockroach gut and soilPromoted the cultivation of a diverse set of bacteria and in particular certain clades of the phyla Actinomycetota and Bacillota [20].
Anaerobic mineral salt medium with PCBsEnhanced PCB dechlorination 1.2–14.1 times higher than control cultures, and Desulfobacterota and Bacteroidetes were Rpf-responsive non-dechlorinators [57].
PCB-contaminated soilsThe dechlorinators, including Dehalococcoides in Chloroflexi and Desulfitobacterium in Firmicutes, were greatly enriched
via rRpf amendment [58].
Rhodococcus rhodochrousR. rhodochrousMPN count was close to the total cell count by incorporating supernatant taken from growing cultures [42].
R. rhodochrous, R. fascians, M. luteusGrowth was promoted while incubated with culture supernatant [43].
Rhodococcus sp. (GX12401)Rhodococcus
VBNC cells
Rhodococcus VBNC cells were resuscitated with 1 pM rRpf treatment, which increased by 18% compared with the control [40].
R. fasciansR. rhodochrous, R. fascians, M. luteusGrowth was promoted while incubated with culture supernatant [43].
Mycobacterium tuberculosisM. bovisActively growing cells of M. bovis (BCG) did not respond to these proteins, whereas bacteria exposed to a prolonged stationary phase did [38].
M. tuberculosisMPN count value was further increased by one log using supernatant from an actively growing culture [42].
RpfD could efficiently stimulate the resuscitation of
M. tuberculosis H37Ra, and the rabbit anti-RpfD serum
could completely inhibit this effect [59].
Corynebacterium glutamicumC. glutamicumThe growth of C. glutamicum was impaired in Δrpf1 and Δrpf2 strains; a significant shortening of the apparent lag phases was observed after addition of culture supernatants [44].
Tomitella biformata AHU 1821TT. biformataAmounts of 2 and 20 μg/mL recombinant Rpf protein resuscitated non-dividing T. biformata cells [60].
Melted surface sterilized ice wedge samplesSamples treated by 1250 pM rRpf had a 14-fold higher number of CFUs on agar plates after 8 d incubation [61].
Nocardiopsis halophilaNocardiopsis halophilaResuscitation effect on the VBNC N. halophila was most evident when 0.1 mg/mL of the rRpf protein was added to the medium [28].
Listeria monocytogeneslmo0186 lmo2522 double mutant of L. monocytogenesLmo0186 and Lmo2522 (equivalent to Rpf) stimulated the growth of lmo0186 lmo2522 double mutant that was growth-arrested [33].
SRpf: culture supernatants from M. luteus containing Rpf. rRpf: recombinant Rpf.
Table 2. Rpf structural diversity in bacteria.
Table 2. Rpf structural diversity in bacteria.
SpeciesProteinProtein Length/AADomain3D Structure
M. luteusRpf
(O86308)
223172–220: LysM domainMicroorganisms 12 01528 i001
M. tuberculosisRpfA
(P9WG31)
40741–114: Transglycosylase-like domainMicroorganisms 12 01528 i002
RpfB
(P9WG29)
362192–272: G5 domainMicroorganisms 12 01528 i003
RpfC
(O07747)
17670–140: Transglycosylase-like domainMicroorganisms 12 01528 i004
RpfD
(P9WG27)
15450–121: Transglycosylase-like domainMicroorganisms 12 01528 i005
RpfE
(O53177)
17298–168: Transglycosylase-like domain;
25–84: PRK12438 super family
Microorganisms 12 01528 i006
Corynebacterium glutamicumRpf1
(Q6M6W5)
19337–111: Resuscitation-promoting factor core lysozyme-like;
117–190: Rpf1 C-terminal domain
Microorganisms 12 01528 i007
Rpf2
(Q6M6N7)
374210–290: G5 domainMicroorganisms 12 01528 i008
Streptomyces sp. SID7813RpfA
(A0A7K2UFD5)
24445–117: Transglycosylase-like domain
195–243: LysM domain
Microorganisms 12 01528 i009
RpfB
(A0A7K2UEW9)
458183–211: DUF348 domain
291–371: G5 domain
Microorganisms 12 01528 i010
RpfC
(A0A7K2UEQ7)
341285–334: LysMMicroorganisms 12 01528 i011
RpfD
(A0A7K2U8R1)
439165–214: LysMMicroorganisms 12 01528 i012
RpfE
(A0A7K2US99)
12148–121: Transglycosylase-like domainMicroorganisms 12 01528 i013
Rhodococcus fasciansRpf
(A0A260UFL0)
375205–285: G5 domainMicroorganisms 12 01528 i014
Curtobacterium sp. 8I-2Rpf
(A0A653LDC6)
234184–232: LysM domainMicroorganisms 12 01528 i015
Pseudonocardia autotrophicaRpf
(A0A1Y2MXK4)
13457–127: Rpf core lysozyme-like domainMicroorganisms 12 01528 i016
Oerskovia enterophilaRpf2
(A0A163S4N4)
434239–320: G5 domainMicroorganisms 12 01528 i017
Nocardiopsis gilvaRpf
(A0A223S313)
206126–200: Rpf core lysozyme-like domainMicroorganisms 12 01528 i018
AA: amino acid. LysM: the domain recognizes polysaccharides containing N-acetylglucosamine (GlcNAc) residues including peptidoglycan, an essential component of the bacterial cell wall. G5: the domain may confer localization or substrate specificity on the proteins in which it is found. The numbers listed in the domain column represent amino acid positions of the domains. Information on protein name, length, domain, and 3D structure was found in the Uniprot database (https://www.uniprot.org/, accessed on 31 May 2024). The codes under the protein name are accession numbers of the Rpfs in Uniprot.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, X.; Ren, Q.; Sun, Z.; Wu, Y.; Pan, H. Resuscitation Promotion Factor: A Pronounced Bacterial Cytokine in Propelling Bacterial Resuscitation. Microorganisms 2024, 12, 1528. https://doi.org/10.3390/microorganisms12081528

AMA Style

Li X, Ren Q, Sun Z, Wu Y, Pan H. Resuscitation Promotion Factor: A Pronounced Bacterial Cytokine in Propelling Bacterial Resuscitation. Microorganisms. 2024; 12(8):1528. https://doi.org/10.3390/microorganisms12081528

Chicago/Turabian Style

Li, Xinxin, Qing Ren, Zhanbin Sun, Yanan Wu, and Hanxu Pan. 2024. "Resuscitation Promotion Factor: A Pronounced Bacterial Cytokine in Propelling Bacterial Resuscitation" Microorganisms 12, no. 8: 1528. https://doi.org/10.3390/microorganisms12081528

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

Article Metrics

Back to TopTop