Next Article in Journal
Ultrasound- and Molecular Sieves-Assisted Synthesis, Molecular Docking and Antifungal Evaluation of 5-(4-(Benzyloxy)-substituted phenyl)-3-((phenylamino)methyl)-1,3,4-oxadiazole-2(3H)-thiones
Next Article in Special Issue
Inhalable Antitubercular Therapy Mediated by Locust Bean Gum Microparticles
Previous Article in Journal
The Homocoupling Reaction of Aromatic Terminal Alkynes by a Highly Active Palladium(II)/AgNO3 Cocatalyst in Aqueous Media Under Aerobic Conditions
Previous Article in Special Issue
The Current Case of Quinolones: Synthetic Approaches and Antibacterial Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Self-resistance in Streptomyces, with Special Reference to β-Lactam Antibiotics

1
HO Bio Institute, 33-9, Yushima-2, Bunkyo-ku, Tokyo 113-0034, Japan
2
Department of Biochemistry, Meiji Pharmaceutical University, 522-1, Noshio-2, Kiyose, Tokyo 204-8588, Japan
Molecules 2016, 21(5), 605; https://doi.org/10.3390/molecules21050605
Submission received: 23 March 2016 / Revised: 26 April 2016 / Accepted: 29 April 2016 / Published: 10 May 2016

Abstract

:
Antibiotic resistance is one of the most serious public health problems. Among bacterial resistance, β-lactam antibiotic resistance is the most prevailing and threatening area. Antibiotic resistance is thought to originate in antibiotic-producing bacteria such as Streptomyces. In this review, β-lactamases and penicillin-binding proteins (PBPs) in Streptomyces are explored mainly by phylogenetic analyses from the viewpoint of self-resistance. Although PBPs are more important than β-lactamases in self-resistance, phylogenetically diverse β-lactamases exist in Streptomyces. While class A β-lactamases are mostly detected in their enzyme activity, over two to five times more classes B and C β-lactamase genes are identified at the whole genomic level. These genes can subsequently be transferred to pathogenic bacteria. As for PBPs, two pairs of low affinity PBPs protect Streptomyces from the attack of self-producing and other environmental β-lactam antibiotics. PBPs with PASTA domains are detectable only in class A PBPs in Actinobacteria with the exception of Streptomyces. None of the Streptomyces has PBPs with PASTA domains. However, one of class B PBPs without PASTA domain and a serine/threonine protein kinase with four PASTA domains are located in adjacent positions in most Streptomyces. These class B type PBPs are involved in the spore wall synthesizing complex and probably in self-resistance. Lastly, this paper emphasizes that the resistance mechanisms in Streptomyces are very hard to deal with, despite great efforts in finding new antibiotics.

Graphical Abstract

1. Introduction

The first antibiotic, penicillin, was discovered in 1929 by Fleming [1] and rediscovered in 1940 by Chain et al. [2] and in 1941 by Abraham et al. [3]. Surprisingly, even before the rediscovery of penicillin, a bacterial penicillinase (β-lactamase) was reported [4] and only few years later, penicillin-resistant Staphylococcus aureus [5,6] and Streptococcus pneumoniae [7] were already described, suggesting strongly that the resistant trait was intrinsically present in their genomes of these bacteria and, in addition, penicillin-inactivating substances (β-lactamases?) were speculated to be involved in the resistance [8]. Gram-negative bacteria such as Neisseria gonorrhoeae [9], Neisseria meningitides [10] and Escherichia coli [11] were intrinsically resistant. The same phenomenon was also observed with streptomycin [12]. That the resistance traits were inherently present in the genomes of Streptomyces was supported by the fact that the percentage production of β-lactamases in newly isolated Streptomyces strains and those isolated thirty years ago from soil was not so different with each other at the time in 1978 [13]. If the selective pressure of the use of β-lactam antibiotics would affect the resistant trait, the relative frequencies of β-lactamase producing Streptomyces strains newly isolated from soil would be much more, indicating that the selective pressure at that time is not so strong as at the present time. Recently, however, D’Costa et al. demonstrated that genes conferring resistance to tetracyclines, vancomycin, aminoglycosides, β-lactams and macrolides existed in 30,000-year old Beringian permafrost and that the amino acid sequences of the β-lactamases showed identities between 53% and 84% with those of known determinants, suggesting that antibiotic resistance is a natural phenomenon that predates the modern selective pressure of clinical antibiotic use [14].
Drug resistance is one of the greatest challenges in modern medicine. Among drug resistance, antibiotic resistance is the most prevailing and threatening topic in public health. According to the information of the Centers for Disease Control and Prevention (CDC), at least 2 million people become infected with bacteria that are resistant to antibiotics and at least 23,000 people die each year as a direct result of these infections in the United States each year [15], and in 2009 the Antibiotic Resistance Genes Database (ARDB) contained resistance information for 13,293 genes [16]. These resistance genes are hypothesized to originate in the antibiotic-producing bacteria such as Streptomyces [17,18,19]. Surprisingly, however, when Dantas et al. cultured clonal isolates from 11 different soils that could use one of 18 different antibiotics as their sole carbon source, phylogenetic profiling of the clonal isolates revealed a diverse set of species consisting of orders Burkholderiales (41%), Pseudomonadales (24%) and Actinomycetales (7%) and others and, indicated that pathogenic microbes can more readily use resistance genes originating from bacteria subsisting on antibiotics than the resistance gene from more distantly related antibiotic producer bacteria [20]. Burkholderiales and Pseudomonadales belong to the proteobacteria and are known to function as scavengers, capable of using a large variety of single carbon sources as food. However, horizontal gene transfer is possible from proteobacteria such as Burkholderiales and Pseudomonadales to Actinomycetales including antibiotic-producing Streptomyces in soil and then to pathogenic bacteria. From the result of a metagenomic analysis to isolate antibiotic resistance genes from soil, Riesenfeld et al. concluded that soil bacteria are a reservoir of antibiotic resistance genes with greater diversity than previously recognized [21]. However, D’Costa et al. described that soil dwelling bacteria produce and encounter a myriad of antibiotics and are a reservoir of resistance determinants (resistome) [22], although they used strains in soil samples collected from diverse locations such as urban, agricultural field, and forest but resembled actinomycetes both morphologically and microscopically [23]. Using multidrug-resistant proteobacteria from the soil, Forsberg et al. isolated 110 resistance genes. Of the 110 resistance genes, 18 had 100% amino acid identity to entries in GenBank and another 32 were highly similar. Interestingly, of 110 resistance genes, 55 were β-lactamase genes highly divergent from those of the antibiotic-producing Streptomyces. From these results they indicated ancient evolutionary relationships between β-lactamases from the soil bacteria and those of the antibiotic-producing bacteria [24]. However, β-lactamases from proteobacteria belong mainly to class C β-lactamases, while those from Streptomyces are class A β-lactamases [19]. So, it is not surprising that these two β-lactamases are only distantly related as described below. In any way, it is more reasonable to speculate that the resistance genes originated in antibiotic producing microorganisms and these resistance genes were developed from natural sources to clinic, human, animal, and other environments by transduction, transformation and conjugation to make the enormous diversity of antibiotic resistance genes and no barriers among the microbial compartments in the microbial world at the present time [25,26,27,28,29].
The phylum Actinobacteria constitute one of the largest phyla within the bacteria [30,31]. Streptomyces species belonging to the Actinobacteria dwell in the soil and are high guanine + cytosine (G + C)-content Gram positive bacteria. They are characterized by their peculiar morphology to undergo complex cellular differentiation like filamentous fungi [32] and by their production of a wide variety of secondary metabolites including β-lactam antibiotics [33,34] such as thienamycin [35] and clavulanic acid [36]. However, unlike penicillin- and cephalosporin-producing fungi, Streptomyces species are prokaryotic microorganisms so they must protect themselves from the attack of antibiotics to avoid suicide, that is, they have to have self-resistance mechanisms. In addition, β-lactam antibiotics have been most widely used for chemotherapy of infectious diseases even about 90 years after penicillin’s discovery, even though the chemical structures were extensively modified. β-Lactam antibiotic resistance is caused mainly by two mechanisms: antibiotic-degrading enzymes, β-lactamases and modification of target sites, penicillin-binding proteins (PBPs). Streptomyces species are known to be highly resistant to benzyl penicillin, although they are Gram-positive bacteria, so it is interesting to know the self-resistance mechanisms in Streptomyces and the relationships between PBPs, β-lactams and β-lactamases [19,37]. This review discusses the self-resistance in Streptomyces, focusing primarily on β-lactam antibiotics.

2. Self-resistance mechanisms

As described above, the self-resistance in antibiotic-producing bacteria such as Streptomyces is at least one of the origins of multi-drug resistance prevailing and threatening in various environments at the present time. Major biochemical mechanisms of drug resistance are detoxication or inactivation of the drug, changing the target site, blocking the transport of the drug into the cell and the efflux pump system [19,38]. These mechanisms are true with the self-resistance mechanisms or, bacteria use very similar or the same mechanisms for self-resistance in the antibiotic-producing bacteria and antibiotic resistance in pathogenic bacteria. While the major resistance mechanism in pathogenic bacteria is β-lactamases, that in Streptomyces is supposed to be penicillin-binding proteins [19]. A number of excellent review papers and databases have been published on antibacterial multidrug resistance and self-resistance [19,39,40,41,42,43].

2.1. β-Lactamases

β-Lactamases are the enzymes which catalyze the hydrolysis of the β-lactam antibiotics to produce antimicrobiologically inactive compounds. Because of this, β-lactamases are responsible for β-lactam resistance in many pathogenic bacteria [44]. These enzymes, however, are also produced by nonpathogenic bacteria such as Streptomyces [13,45], Bacillus [46] and the cyanobacteria [47,48]. In addition, recent whole genome analyses expand the β-lactamase world to β-lactamase superfamily proteins [49,50], β-lactamase domain-containing proteins [51] and RNA-metabolizing metallo-β-lactamases [52,53]. Consequently, β-lactamase-producing organisms include a thermobacterium (Meiothermus ruber DSM 1279) (GenBank accession number, ADD27888, the same hereafter, class C β-lactamase), a green sulfur bacterium (Pelodictyon phaeoclathratiforme BU-1, ACF43212, a metallo-β-lactamase), a green non-sulfur bacterium (Herpetosiphon aurantiacus DSM 785, ABX07048, a metallo-β-lactamase), a methylotrophic bacterium [54] (Methylovorus glucosetrophus SIP3-4, ACT50532, a metallo-β-lactamase), a fungus (Schizosaccharomyces pombe, CAA93219, a tRNA processing endonuclease, a metallo-β-lactamase) and even a plant [55] (Arabidopsis thaliana, AAC49866, a glyoxalase II mitochondrial enzyme) and mammals [56,57] (Rattus norvegicus and Homo sapiens, EDL84248 “a serine β-lactamase-like protein” and NP_116246 “a serine β-lactamase-like mitochondria protein”, respectively). Furthermore, an alkalophilic and halotolerant Gram-positive bacterium Oceanobacillus iheyensis isolated from the bottom of the Pacific Ocean at a depth of 1050 m [58] and Pseudomonas fluorescens isolated from a remote Antarctic coastal area [59] also produce β-lactamases. Moreover, screening of metagenomic libraries from soil bacterial communities for clones that confer antibiotic resistance gave new types of β-lactamases such as oligomeric [60] and bifunctional β-lactamases [61,62]. A bifunctional β-lactamase was also isolated from a pathogen Mycobacterium tuberculosis H37Rv, an Actinobacterium [63,64]. Therefore, various kinds of β-lactamases or β-lactamase superfamily are distributed in a variety of organisms, but in this review β-lactamases are described in a more classical sense.
β-Lactamases are grouped into four classes on the basis of their amino acid sequences [65,66,67,68,69,70,71,72]. This classification reflects more or less the substrate specificity: class A (penicillin-hydrolyzing), class C (cephalosporin-hydrolyzing) and class D (oxacillin-hydrolyzing) β-lactamases are active site serine enzymes and class B β-lactamases require one or two zinc ions for their activity and are called as metallo-β-lactamases [69,73,74,75,76]. A previous phylogenetic study indicated that class A β-lactamases could be divided into three subgroups [77]. β-Lactamases are not essential bacterial proteins in themselves, but are speculated to have evolved from the essential PBPs in some β-lactam-producing bacteria such as Streptomyces or related bacteria, because these bacteria have to have some self-protective strategies against β-lactam antibiotics [19,48]. PBPs are involved in peptidoglycan biosynthesis in prokaryotic microorganisms. Therefore, it is interesting to know the relationships of β-lactam antibiotics, β-lactamases and PBPs in Streptomyces, β-lactam-producing microorganisms. In addition, the relationship of β-lactamases in Streptomyces and those in pathogenic bacteria should be clarified. In this section, the phylogenetic relationships were analyzed among β-lactamases in Streptomyces and a couple of pathogenic bacteria, whose genome sequences were wholly clarified.
Previously, I investigated the characteristics of 59 β-lactamase candidate sequences among 529 putative β-lactamase and PBP sequences in 19 Actinobacteria [78,79], which were collected from the NCBI database [80]. There were some characteristic features in the 59 candidate sequences. First, β-lactamase candidates were not distributed uniformly in the strains: while Saccharomonospora viridis, Streptomyces avermitilis, Streptomyces scabiei and Streptosporangium roseum possess as many as five, twelve, five and five candidate genes, respectively, Mycobacterium leprae, Nocardioides sp. JS614 and Rhodococcus erythropolis had no candidate genes. In general, Streptomyces species produced β-lactamase candidates at higher rates. However, the numbers of candidate sequences were not necessarily related to the taxonomic classification. For example, although Streptosporangium and Nocardioides were closely related taxonomically [81], the numbers of β-lactamase candidates in Streptosporangium were five but those in Nocardioides were zero. Even in the same genus, M. tuberculosis produced four but M. leprae produced none. β-Lactamases detected by their enzymatic activity also distributed independently of actinobacterial taxonomy [82]. These results indicate strongly that β-lactamases evolve and are produced independent of taxonomy.
Second, most of the β-lactamases detected so far by their enzymatic activity in Gram-positive bacteria belong to class A and a few belong to class B. However, this fact was not the case with whole genomic level.
Third, it is intriguing that Thermomonospora curvata, a high G + C Gram-positive soil thermophilic Actinobacteria, produced two class B and one class D β-lactamase candidates but no class A and class C enzymes. Usually, Gram-positive bacteria produce class A and class B β-lactamases, but not class C and class D. In addition, T. curvata DSM43183 Tcur2040 β-lactamase is closely related to Gram-negative class D β-lactamases, especially OXA-5 and OXA-27 type β-lactamase [79], although T. curvata is a Gram positive bacteria. Hereafter, genetic analyses are extended to β-lactamases of other 12 Streptomyces species listed in Table 1.

2.1.1. Class A β-lactamases

The amino acid sequences of putative 25 class A β-lactamases together with a reference sequence (TEM-1, Klebsiella penumoniae plasmid, YP_005351445) were aligned by using the Muscle and T-coffee programs [83,84]. The results showed that all the sequences in Table 2 possessed main residues involved in the catalytic reaction (Table 3), except SCAT_1418, SCAT_4581, SRIM_07318, and F750_2387 (Table S1). These four sequences are defective in S70XXK73 and S130DN132. Furthermore, SCAT_1418, SCAT_4581 and F750_2387 do not carry K234T/SG236, indicating that these candidates do not function as β-lactamases (Table 2).
Although some other sequences are also deficient or possess incompletely identical signature amino acid sequences, they are supposed to operate as β-lactamases even if the catalytic efficiency is not perfect. GWG166 in SSCG_00160, GWG166 in F750_6336, IWE166 in B446_02285, IWE166 in B466_33010, RVI166 in STSU_15177, RGI166 in SCAT_0807, HYE166 in SCAT_5692, QYE166 in SSCG_00130, and RWS166 in F750_2387 are examples. In fact, blast analysis with SSCG_00160 as a query, putative β-lactamases from Saccharomonospora viridis (KHF44103), Amycolicicoccus subflavus DQS3-9A1 (AEF33802) and Rhodococcus sp. B7740 (AJW42765) were identified.
In addition blast analysis using SCAT_1418 as a query, metallo-dependent hydrolase of Streptomyces violaceusniger (WP_014060235) and metallo-dependent hydrolase of Streptomyces himastatinicus (WP_00971837) were identified. Actually, SCAT_1418 possesses some class B β-lactamase signature residues. Likewise, SCAT_4581 is highly similar to Zn-dependent hydrolases from Actinobacteria, so that they may be class B β-lactamases.
SRIM_07318 is too short to function as a β-lactamase, but may be a part of class A β-lactamase, as its amino acid sequence of the C-terminal part is very similar to that of class A β-lactamases from Streptomyces albus (WP_060729068) and Rhodococcus rhodnii (WP_037260801). F750_2387 is a large membrane protein and belongs to metallo-β-lactamase superfamily. The fact that SCAT_1418, SCAT_4581 and F750_2387 belong to metallo-β-lactamases but not to class A reflects on large molecular distances between these β-lactamases and others such as TEM-1 and SHV-1 (Table S2). Interestingly, the amino acid sequence of SRIM_10531 is completely identical with that of Streptomyces cellulosae (D12653). This is one example of the horizontal gene transfer. Another interesting thing is that the amino acid sequences of B446_02285 and B446_33010 in class A β-lactamases, and B446_02210 and B446_33085, and B446_02000 and B446_33290 in class B β-lactamases, and B446_01720 and B446_33570 in β-lactamase-like sequences and B446_01315 and B446_33975 in class C β-lactamases are also completely identical, indicating that about 7.7 Mb region including these genes were duplicated in the chromosome in the past. In fact, two transposases exist at both sides in reverse direction from 7,984,885-7,985,573n and 287,352-288,040n in Streptomyces collinus genome, respectively, suggesting that these transposases acted in the transposition or the duplication.
A phylogenetic tree of putative 20 class A β-lactamases constructed with those of S. aureus (M15526), P. aeruginosa (Q03170), K. pneumoniae (TEM-1, YP_005351445), K. pneumoniae (SHV-11, X98101), and E. coli (SHV-1, AF148850) as reference sequences disclosed that although these reference sequences are originated from quite different species, they form only one cluster (cluster C in Figure 1). In contrast, β-lactamases of Streptomyces are scattered in various branches, indicating that only a few β-lactamase genes of Streptomyces species have been transferred to pathogenic bacteria at the present time.
However, it is possible that other β-lactamase genes of Streptomyces species may transfer to pathogenic bacteria in the near future, although many threatening class A genes may be evolved through mutation but not transferred to pathogenic bacteria, as it is seen in TEM-type and SHV-type β-lactamases [89]. Actually, molecular distances determined by the protdist program [90] between most class A β-lactamases of Streptomyces and those of pathogenic bacteria (Table S2) are not so large as in class B and class C β-lactamases, indicating that these class A β-lactamases are related closely with each other and with those of pathogenic bacteria than class B and class C enzymes. It is speculated, therefore, that the horizontal gene transfer of class A β-lactamase genes is easier than for class B and class C β-lactamases. The β-lactamases of Streptomyces can be classified into blue-dextran/NADP+-binding and non-binding groups [91,92]. Blue-dextran/NADP+-non-binding β-lactamases belong to cluster A and those of binding β-lactamases go to cluster B in the phylogenetic tree (Figure 1). However, the role of NADP+-binding of β-lactamases remains to be clarified.
As for the relationship between class A β-lactamases and the β-lactam biosynthetic gene cluster, SSCG_00130 (corresponding to SCLAV_4216, partial stop and SCAT_5692) is located in the terminal region of the cephamycin biosynthetic gene cluster and SSCG_00160 (corresponding to SCLAV_4187) is located within the clavulanic acid biosynthetic gene cluster in S. clavuligerus (Figure S1). These two clusters exist in the adjacent position. SSCG_01467 (corresponding to SCLAV_2436) is not related to β-lactam biosynthetic genes [78]. Similarly, SCAT_5692 is present in the terminal region of the cephamycin biosynthetic gene cluster in S. cattleya (Figure S2). However, a gene corresponding to SSCG_00160 in S. clavuligerus is missing in S. cattleya in accord with the absence of clavulanic acid biosynthesis. Therefore, it is suggested strongly that class A β-lactamases and β-lactam biosynthesis are closely related with each other. However, it remains to be clarified whether these β-lactamases are involved in the self-resistance in the Streptomyces species. SSCG_00160 (SCLAV_4187) is also thought to be involved in clavulanic acid biosynthesis [93,94].

2.1.2. Class B β-lactamases

On the basis of phylogenetic and sequences analyses, 63 class B β-lactamase candidates were selected in 12 Streptomyces species. The amino acid sequence alignment using the Muscle and T-coffee programs [83,84] revealed that all the sequences contain the conserved active site residues (HXHXD120) except SSCG_05130, SHJG_8335, SCAT_4145, and SRIM_16845 (Table 4 and Table S3; In the case of AAA22276_Bacillus cereus reference sequence, the active site residues are shifted to downstream in Table S3).
However, blast analyses using these sequences as queries clarified that they are closely related to αββα metallo β-lactamase (MBL) fold hydrolases from Streptomyces such as S. roseochrmogenus (WP_023553317), S. silaceus (WP_055701146), S. sclerotialus (WP_030612498), S. corchorusii (WP_059262406), S. hygroscopicus (WP_058083348), S. achromogenes (WP_030611171), S. bingchenggensis (WP_014181511), S. alboniger (WP_055536123), S. rimosus (WP_033031918), S. albus (WP_060732298), S. monomycini (WP_030021421) and others, indicating that these sequences belong to the metallo β-lactamase superfamily. As in the case of class A, while the class B β-lactamases of pathogenic bacteria form compact clusters in a phylogenetic tree, those from Streptomyces are dispersed quite extensively (Figure 2).
Moreover, comparing to Class A and class C β-lactamases, molecular distances calculated by using the protdist program [90] are mostly quite large between those of Streptomyces species with each other and those with pathogenic bacteria (Table S4), suggesting that the class B β-lactamase enzymes are assemblies of highly diverged proteins. Recently, new type of metallo-β-lactamases such as New Delhi metallo-β-lactamase (NDM) have emerged and swiftly spread worldwide. It is possible that various metallo-β-lactamases of Streptomyces are also transferred to pathogenic bacteria and pose a similar public health threat.

2.1.3. Class C β-lactamases

The amino acid sequence alignment of putative 94 class C β-lactamases is shown in Table S5. From this alignment, the presence or absence of the active site residues was inferred and the results was shown in Table 5. All the sequences possess active site residues shown in Table 3, with exceptions of SHJG_8639 (YP_006249777), SHJG_2828 (YP_006243975), SSCG_03668 (ZP_05006222), SRIM_31085 (ELQ79325), BN159_7932 (YP_007526438), SSQG_02491 (ZP_07303604), and SSQG_00225 (ZP_07301338). However, the N-terminal sequences of SHJG_2828, SSCG_03668, SRIM_31085, SSQG_02491 and SSQG_00225 are missing, so the numbers of amino acid residues are too short to compare. In fact, blast analysis shows that SHJG_8639 is similar to Zn-dependent hydrolases (l-ascorbic acid metabolism, UlaG) of Streptomyces corchorusii (WP_059265188), S. bingchenggensis (WP_043485943) , and Nonomuraea candida (WP_043633066); SHJG_2828 is similar to Zn-dependent hydrolases (glyoxylase, GloB) of multiple species of Streptomyces (WP_037824258, WP_037629085 and others), Streptomyces fulvoviolaceus (WP_03061340) , and S. avermitilis (WP_010988349); SSCG_03668 is related to serine hydrolases of Streptomyces aureus (WP_037619730), Actinokineospora spheciospongiae (WP_035277635) and Micromonospora parva (WP_030334201); SRIM_31085 is similar to PBPs of S. rimosus (WP_030645517), S. albus (WP_060729102) and S. yokosukanensis (KUN09412); BN_7932 is associated with serine hydrolases of Streptomyces venezuelae (WP_055639726), Streptomyces vietnamensis (WP_041132293), and Streptomyces antibioticus (WP_059193915); SSQS_02491 is similar to serine hydrolases of Streptomyces olindensis (KDN73989), Streptomyces acidiscabies (WP_059045061), and Streptomyces torulosus (WP_055716632); and SSQG_00225 is related to PBPs of Streptomyces chartreusis (WP_010033556), Streptomyces iakyrus (WP_051814832), and Streptomyces pactum (WP_055421147). Therefore, all of these sequences are members of the β-lactamase superfamily. The amino acid numbers of SHJG_2828, SSCG_03668, SRIM_31085, SSQG_02491 and SSQG_00225 are too short to function as β-lactamases, and the C-terminal residues in SSCG_03303 are missing.
A phylogenetic tree was constructed with β-lactamases of Rhodobacter sphaeroides (YP_355265), Mycobacterium smegmatis (NC_018289), Acinetobacter baumannii (CAB77444), Aeromonas caviae (AF462690_1), and E. coli (ABM69263 and NP_418574) as reference sequences (Figure 3). Although six reference sequences are selected from various different species [95], they form only one cluster as described previously [77,78]. On the other hand, β-lactamases of Streptomyces are dispersed into numerous clusters, indicating that they originate from different kinds of proteins. A molecular distance analysis using the protdist program [90] confirmed this observation (Table S6). The molecular distances between Streptomyces β-lactamases and that of pathogenic bacteria are very large, but those of β-lactamases of Streptomyces are also large with each other, so if we continue to develop and use new, newer and the newest β-lactam antibiotics at the present rate, a dreadful microbial revolution may occur near future.

2.2. Penicillin-Binding Proteins (PBPs)

The bacterial cell wall peptidoglycan is a three-dimensional, covalently closed, net-like mesh called sacculus in which glycan strands are cross-linked by peptide chains. It maintains cell shape and provides mechanical strength to resist osmotic pressure [96,97]. The peptidoglycan chains are composed of alternating β-1,4-linked N-acetylglucosamine and N-acetylmuramic acid residues. The carboxyl groups of the N-acetylmuramic acid residues are involved in amide bond to terminal l-alanine residues of the peptide units l-alanyl-γ-d-glutamyl-l-diaminoacyl-d-alanine. The glycan chain is biosynthesized by catalysis of glycosyltransferases and the cross-linkage of the peptide chains between two adjacent glycan chains is catalyzed by transpeptidases [87,98]. The transpeptidases, also called penicillin-binding proteins (PBPs), were initially identified as their ability to bind penicillins [99,100,101]. Depending on the structure and the catalytic activity of their N-terminal domain, they are classified into class A, class B and class C PBPs [87,96,101,102]. Both class A and class B PBPs have the transpeptidase activity in the C-terminal region. However, the N-terminal domain in class A PBPs is responsible for their glycosyltransferase activity, while the glycosyltransferase domain is lacking in class B PBPs. Instead, the PBP dimer domain (Pfam03717) is present at the N-terminal region in class B PBPs. However the detailed functional role of the dimer domain is yet to be fully explicated. Class C PBPs are also called low molecular weight PBPs and, having the carboxypeptidase activity, are responsible for the maturation and recycling of the peptidoglycan [96], but they are not essential and are excluded from further study. There are many excellent review paper on PBPs [87,96,97,98,100,101,102], and the comprehensive review on the PBPs in Actinobacteria in general was already published [103], so this review describes the PBPs in Streptomyces, especially emphasizing their roles in self-resistance.

2.2.1. Class A PBPs

The numbers and types of putative PBPs in 24 Streptomyces are summarized in Table 6. The amino acid sequences of putative 100 class A PBPs in 24 Streptomyces species together with a reference sequence (Q04704/CVP62105, S. pneumoniae) were aligned by the Muscle and T-coffee programs (Table S7). All sequences analyzed possess active site S465XXK468 residues, excluding SAV_4423, SBI_09068, STRVI_3845, SCAB_64431, SZN_18682, DC74_7647, SSQG_02328, SSFG_02387, SAV_7219, SHJG_3853, SRIM_08328, SBI_03076, and STRVI_2314. However, active site residues S524XN546, and K716TG718 are mostly conserved. In fact, blast analysis of STRVI_3845 as a query showed that glycosyltransferases of S. iranensis (WP_044567878), S. hygroscopicus (WP_060948597) and S. bingchenggensis (WP_014181633) are identified and, that of SSQG_02328 as a query, PBPs of S. afghaniensis (WP_037667886), S. iakyrus (WP_033306427), and S. caeruleatus (KUN93289) are identified, and that of SHJG_3853 as a query, PBPs of S. antibioticus (WP_053211715), S. reticuli (CUW28219), and S. achromogenes (WP_030604457) are identified. In addition, the glycosyltransferase and transpeptidase domains are preserved in these sequences, indicating that these sequences can function as PBPs. A phylogenetic tree was constructed by using SCO4049 (putative penicillin acylase) as outgroup (Figure 4). The class A PBPs are clearly divided to three groups: A, B and C. SAV_4423 is a lone sequence. The PBPs in these groups are obviously corresponding to those in clusters A, B and C in the amino acid alignment (Table S7).
According to the penicillin-binding cores of PBPs, class A PBPs are classified into two subclasses in Gram-negative bacteria and three subclasses in Gram-positive bacteria [104]: A1 (whose prototype is Escherichia coli PBP1A, PBPA_ECOLI), A2 (whose prototype is E. coli PBP1B, PBPB_ECOLI), A3 (whose prototype is Streptococcus pneumoniae 1A, PBPA_STRPN), A4 (whose prototype is S. pneumoniae 2A, AJ002292), and A5 (whose prototype is S. pneumoniae 1B, AJ002291). Interesting enough, all class A PBPs from Streptomyces analyzed in this paper form a completely different cluster in the phylogenetic tree from these five subclasses (Figure 4), where E values (number of alignments expected by chance) between PBPs in subclasses A1 to A5 and those in Streptomyces are in the range of 3.4 × 10–16 to 4.9 × 10–31, indicating low similarities. These results suggests strongly that the gene transfer and/or gene conversion occurred rarely between PBPs in Streptomyces and those in Gram-positive and Gram-negative bacteria. However, the molecular distances are not so large except PBP of Helicobacter pylori and PBPs of Streptomyces, indicating that these PBPs are not so different as speculated by the phylogenetic tree with each other (Table S8), and it is possible that the horizontal gene transfer of these genes may take place from Streptomyces to pathogenic bacteria near future.

2.2.2. Class B PBPs

One hundred sixty one class B PBPs in 24 Streptomyces listed in Table 6 were analyzed. Intriguingly, more than half of Streptomyces species carry two successive class B PBPs in tandem. For example, XNR_2096 and XNR_2097 (E value is 5.4 × 10–58, the same hereafter), SAV_3603 and SAV_3604 (3 × 10–68), SCLAV_4178 and SCLAV_4180 (1.5 × 10–11), SMCF_7795 and SMCF_7796 (3.3 × 10–60), SCO3156 and SCO3157 (2.8 × 10–47), BN159_5121 and BN159_5122 (1.4 × 10–62), SSFG_04216 and SSFG_04217 (1.3 × 10–56), SSRG_03705 and SSRG_03706 (2.7 × 10–62), SHJG_4627 and SHJG_4628 (2.8 × 10–69), SLIV_21910 and SLIV_21915 (4.3 × 10–58), SCAB_53611 and SCAB_53621 (6.8 × 10–53), SSEG_00010 and SSEG_00011 (3.8 × 10–63), and SSQG_03242 and SSQG_03243 (5.5 × 10–65). These amino acid sequences are not only very similar to each other, but also all the sequences belong to the same cluster in a phylogenetic tree constructed by using SCO4049 as outgroup (cluster C in Figure 5). In fact, the amino-acid sequence identity and similarity of PBPs in cluster C in the phylogenetic tree are in the range of 49.2%–51.8% and 71.8%–77.8%, respectively. Moreover, the nucleotide sequences of each pair are arrayed in the same direction, indicating that they were duplicated and transferred to each other very recently. Interestingly enough, the pair of a cephamycin and clavulanic acid producer, that is, S. clavuligerus SCLAV_4179 and SCLAV_4180 is an exception.
SCLAV_4179 belong to cluster B and SCLAV_4180 pertain to cluster A in the phylogenetic tree and the similarity of the amino-acid sequences is very low (E-value is 1.5 × 10–11). This character may be related to β-lactam production and these two PBPs seem to behave independently, because these two sequences are aligned in the reverse direction, i.e., SCLAV_4180 (pbpA) acts together with the clavulanic acid biosynthetic gene cluster, while SCLAV_4179 (pbp2) goes on with SCLAV_4178 (cytochrome P450, partial start) and SCLAV_4177 (RNA polymerase) or independently [78]. Moreover, PBP SCLAV_4179 in S. clavuligerus is reported to have a low affinity to β-lactam antibiotics and is essential to the growth [105]. Therefore, it is assumed that PBP SCLAV_4179 is involved in the self-resistance. SCLAV_4198 (class BPBP, pcbR) is located in the border between cephamycin and clavulanic acid biosynthetic gene clusters but behaves independently from these clusters, considering the gene direction. In S. cattleya which produces only cephamycin but not clavulanic acid, genes corresponding to SCLAV_4180 and SCLAV_4198 are deleted, but a gene (SCAT_5676) corresponding to SCLAV_4179 and responsible for self-resistance is retained [78]. This fact supports the above ideas (Figure S2).
Related to self-resistance, Ogawara and Horikawa reported nearly 40 years ago that β-lactam-producing Streptomyces species such as S. clavuligerus possessed PBPs of very low affinity to benzylpenicillin [106]. Later, two PBPs in S. clavuligerus, that is, SCLAV_4179 and SCLAV_4198 were reported to have low affinity to penicillins [105,107]. A mutant disrupted in SCLAV_4198 gene exhibited a significant decrease in its resistance to benzylpenicillin and cephalosporins [107]. Moreover, a probe containing SCLAV_4198 hybridized to genomic DNAs from β-lactam producers, S. jumonjinensis NRRL 5741, S. griseus NRRL 3851 and S. lipmanii NRRL 3584, suggesting that SCLAV_4198-like sequences and SCLAV_4198-mediated resistance mechanisms are likely to be present in these β-lactam-producing species. Likewise, most Streptomyces species have two PBPs in SCLAV_4198-belonging cluster C in the tree.
These characters are moderately similar with each other: SRIM_04191 and SRIM_06646 (1.4 × 10–59), SCLAV_4198 and SCLAV_2276 (8.1 × 10–27), F750_6320 and F750_2998 (5.6 × 10–29), STRVI_1135 and STRVI_3190 (1.2 × 10–61), and SBI_04376 and SBI_06233 (7.7 × 10–36) are examples. Together with the fact that tandem two genes in most Streptomyces species are associated with cluster C as described above, these results clearly indicate that most Streptomyces species are firmly defended by the presence of two low-affinity PBPs. Moreover, cluster B PBPs, to which SCLAV_4179 belong, reconfirm the self-resistance [103], that is, two pairs of low affinity PBPs (e.g., SCLAV_4198 and SCLAV_2276 in cluster C in Figure 5, and SCLAV_4179 and SCLAV_1774 in cluster B in Figure 5) guard Streptomyces from the attack of β-lactam antibiotics.

3. PASTA Domains: Connector between Penicillin-Binding Proteins and Protein Kinases

Protein phosphorylation was first described as a major regulatory mechanism in eukaryotes [108,109]. Later, the regulation by phosphorylation with serine/threonine protein kinases (STPKs) in particular has been expanded to prokaryotes [110,111,112,113]. While investigating the S. coelicolor homolog of PknB, a STPK, Yeats et al. identified a definitive domain in its C-terminus. This domain is also found in C-terminus of high molecular weight PBPs, so that it is termed as the PASTA domain (penicillin-binding protein and serine/threonine kinase-associated domain) [114]. PASTA domains are not found in eukaryotes and are restricted to firmicutes and Actinobacteria. Interestingly, some PASTA domains in PBPs such as PBP2x of S. pneumoniae bind peptidoglycan and β-lactam antibiotics [115], but others such as PonA2 of M. tuberculosis do not [116], indicating that PASTA domains in PBPs operate in a species-specific manner. On the other hand, PASTA domains in STPKs bind peptidoglycan and β-lactam antibiotics, and STPKs with PASTA domains are reported to play important roles in the regulation of bacterial cell division, and morphogenesis [111,112,114,117,118]. Previously, I described the distribution of PASTA domains in Actinobacteria [119], so in this review I discuss the interaction of PASTA domains in PBPs and STPK from the point of view of self-resistance.
In sharp contrast to other bacteria such as B. subtilis, Clostridium perfringens and S. pneumonia, PBPs with PASTA domains are detected only in class A PBPs but not in class B PBPs of Actinobacteria. In addition, none of Streptomyces species has PBPs with PASTA domain. Class A PBPs have both transglycosylase and transpeptidase domains, whereas class B PBPs have only transpeptidase domain. Therefore, it is interesting to know the interaction between transglycosylase and PASTA domains in Actinobacteria. However, it is not interpreted yet. Related to this fact, one of class B PBPs without PASTA domain and a STPK with four PASTA domains are located in adjacent position in most Streptomyces species (Table 6). For example, DC74_4185 and DC74_4186, XNR_3038 and XNR_3037, SAV_4339 and SAV_4338, SBI_05407 and SBI_05406, SCO3847 and SCO3848, and others. These PBPs belong to cluster Ac in a phylogenetic tree (Figure 5). The genomic structures in PBP-STPK regions of S. coelicolor, S. clavuligerus, and S. griseus are very similar and suggest that STPK, PBP and FtsW/RodA family protein form an operon (Table 7). In S. coelicolor, this family of PBP (i.e., SCO3847) is reported to form the Streptomyces spore wall synthesizing complex (SSSC) [120,121]. Although they are supposed to be transported to peptidoglycan biosynthesis machinery through the PASTA domains of STPK (i.e., SCO3848) and be involved in the peptidoglycan biosynthesis and/or morphogenesis coupled with PBPs and FtsW/RodA family proteins, Jones et al. reported that SCO3848 (PknB) regulates the timing of development and TCA cycle favoring antibiotic production together with two forkhead-associated proteins (FHA, SCO3843 and SCO3844) [122]. In addition, SCO3848 (PknB) are supposed to play coordinately with the protein phosphatase (i.e., SCO3845) inside the operon. It is very interesting to know how self-resistance is implicated in this reaction, because the identity and the similarity of PBPs of this group (cluster Ac in Figure 5) to SCLAV_4180 are in the range of 40%–91%, and 70%–95%, respectively and those of SCLAV_2947 (a low affinity PBP) and SCLAV_4180 are 53.3% and 80.2% in 497 amino acid residues, respectively (E value is 1.7 × 10–102). The indication that clusters Aa, Ab, and Ac PBPs in Figure 5 are close together with each other is confirmed by the molecular distance determination (Table S9). Cluster A PBPs in Figure 5 are also closely related with PBPs of pathogenic bacteria (Table S9). Unfortunately, however, the role of cluster A PBPs in self-resistance remains to be elucidated.

4. Relationship between β-Lactam Biosynthetic Gene, β-Lactamase and PBP: Additional Comments

The relationship between β-lactam biosynthetic gene clusters, β-lactamases, and PBPs of S. clavuligerus and S. cattleya was shown in Figure S1 [33,123,124,125,126,127]. The genetic organization of clavulanic acid biosynthetic gene cluster of S. clavuligerus was published [93,128]. Comparison of cephamycin gene clusters in S. clavuligerus and S. cattleya indicates clearly that clavulanic acid gene cluster from SCLAV_4180 to SCLAC_4198 in S. clavuligerus was inserted between SCAT_5676 and SCAT_5678 of S. cattleya (Figure S2). SCAT_5677 is missing. Interestingly, both sides of the clavulanic acid gene cluster are occupied by PBPs (SCLAV_4180 and SCLAV_4198), indicating that these PBPs behave together with clavulanic acid/cephamycin gene cluster and are involved in the self-resistance of S. clavuligerus. Similar gene constructs for cephamycin biosynthesis are observed in Nocardia lactamdurans, Lysobacter lactamgenus, Penicillium chrysogenum, and Acremonium chrysogenum [129,130,131,132]. However, β-lactamase and PBP genes can be detected neither in β-lactam biosynthetic gene clusters nor in whole genomes of P. chrysogenum, and A. chrysogenum. On the other hand, the organizations of clavulanic acid gene clusters of three Streptomyces species, S. clavuligerus, S. flavogirseus, and S. viridis are also similar with each other. Unfortunately, however, relationship between β-lactam biosynthesis, β-lactamases, and PBPs remains to be clarified, although β-lactamase was reported to be expressed during the active growth phase, prior to the formation of β-lactam antibiotics [129]. Recently, an interesting paper on taxonomy, physiology, and natural products of Actinobacteria was published [133].

5. Conclusions

Although antibiotics still play a key role for the prevention of microbial infections as remaining treasures from the twentieth century, antibiotic resistance is prevailing and putting us in a critical situation. Furthermore, it is said that the post-antibiotic era is coming soon [134]. As described in this paper, the β-lactamases in antibiotic-producing Streptomyces are diverse in their characteristics, and the PBPs are multiplexed in their guard systems. In addition, as antibiotic resistance mechanisms are supposed to originate and evolve in their producing microbes, and be transferred to pathogenic bacteria by transformation, transduction, transfection and/or conjugation, the public health crisis will be getting worse and worse. β-Lactamases and PBPs are two major resistance mechanisms in pathogenic bacteria against β-lactam antibiotics, the most frequently used antibiotics for infectious diseases at the present time. Moreover, from about 40 years ago, the rate of discovery of new antibiotics has declined rapidly and many large pharmaceutical companies have abandoned research and development on antibiotics [135]. To avoid this situation, therefore, it is urgently needed to make the public and the government recognize the situation. In addition, new antibiotics should be screened by using various technologies such as activation of cryptic gene clusters for antibiotic biosynthesis, metagenomics mining, combinatorial biosynthesis, target-directed computer-aided chemical synthesis, systems biology, genetic manipulation, omics, and so on [136,137,138,139,140,141]. Furthermore, combination therapies with monoclonal antibodies and vaccines are also required. When these requirements are attained, “New Golden Age” is possible [141]. However, it is worth mentioning again that the resistance mechanisms in Streptomyces species are very hard to deal with as described in this paper.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/21/5/605/s1.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Fleming, A. On the antibacterial action of cultures of penicillium, with special reference to their use in the isolation of B. influenzae. Br. J. Exp. Pathol. 1929, 10, 226–236. [Google Scholar] [CrossRef]
  2. Chain, E.; Florey, H.W.; Gardner, A.D.; Heatley, N.G.; Jennings, M.A.; Orr-Ewing, J.; Sanders, A.G. Penicillin as a chemotherapeutic agent. Lancet 1940, 236, 226–228. [Google Scholar] [CrossRef]
  3. Abraham, E.P.; Chain, E.; Fletcher, C.M.; Florey, H.W.; Gardner, A.D.; Heatley, N.G.; Jennings, M.A. Further observations in penicillin. Lancet 1941, 238, 177–189. [Google Scholar] [CrossRef]
  4. Abraham, E.P.; Chain, E. An enzyme from bacteria able to destroy penicillin. Rev. Infect. Dis. 1940, 10, 677–678. [Google Scholar] [CrossRef]
  5. Plough, H.H. Penicillin resistance of Staphylococcus aureus and its clinical implications. Am. J. Clin. Pathol. 1945, 15, 446–451. [Google Scholar] [CrossRef] [PubMed]
  6. Bondi, A., Jr.; Dietz, C.C. Penicillin resistant staphylococci. Proc. Soc. Exp. Biol. Med. 1945, 60, 55–58. [Google Scholar] [CrossRef] [PubMed]
  7. Eriksen, K.R. Studies on induced resistance to penicillin in pneumococcus type I. Acta Pathol. Microbiol. Scand. 1945, 22, 398–405. [Google Scholar] [CrossRef] [PubMed]
  8. Gots, J.S. Production of extracellular penicillin-inactivating substances associated with penicillin resistance in Staphylococcus aureus. Proc. Soc. Exp. Biol. Med. 1945, 60, 165–168. [Google Scholar] [CrossRef] [PubMed]
  9. Miller, C.P.; Bohnhoff, M. Studies on the action of penicillin; development of penicillin resistance by gonococcus. Proc. Soc. Exp. Biol. Med. 1945, 60, 354–356. [Google Scholar] [CrossRef] [PubMed]
  10. Miller, C.P.; Bohnhoff, M. Studies on action of penicillin; virulence of penicillin resistant strains of meningococcus. Proc. Soc. Exp. Biol. Med. 1945, 60, 356–362. [Google Scholar] [CrossRef] [PubMed]
  11. Shwartzman, G. Studies on the nature of resistance of Gram-negative bacilli to penicillin. J. Exp. Med. 1946, 83, 65–88. [Google Scholar] [CrossRef] [PubMed]
  12. Waksman, S.A.; Reilly, H.C.; Schatz, A. Strain specificity and production of antibiotic substances. V. Strain resistance of bacteria to antibiotic substances, especially streptomycin. Proc. Natl. Acad. Sci. USA 1945, 31, 157–164. [Google Scholar] [CrossRef] [PubMed]
  13. Ogawara, H.; Horikawa, S.; Shimada-Miyoshi, S.; Yasuzawa, K. Production and property of beta-lactamases in Streptomyces: Comparison of the strains isolated newly and thirty years ago. Antimicrob. Agents Chemother. 1978, 13, 865–870. [Google Scholar] [CrossRef] [PubMed]
  14. D’Costa, V.M.; King, C.E.; Kalan, L.; Morar, M.; Sung, W.W.; Schwarz, C.; Froese, D.; Zazula, G.; Calmels, F.; Debruyne, R.; et al. Antibiotic resistance is ancient. Nature 2011, 477, 457–461. [Google Scholar] [CrossRef] [PubMed]
  15. About antimicrobial resistance. Available online: http://www.cdc.gov/drugresistance/ (accessed on 20 March 2016).
  16. Liu, B.; Pop, M. ARDB—Antibiotic resistance genes database. Nucleic Acids Res. 2009, 37, D443–D447. Available online: http://ardb.cbcb.umd.edu/ (accessed on 20 March 2016). [Google Scholar] [CrossRef] [PubMed]
  17. Walker, M.S.; Walker, J.B. Streptomycin biosynthesis and metabolism. Enzymatic phosphorylation of dihydrostreptobiosamine moieties of dihydrostreptomycin-(streptidino) phosphate and dihydrostreptomycin by Streptomyces extracts. J. Biol. Chem. 1970, 245, 6683–6689. [Google Scholar] [PubMed]
  18. Benveniste, R.; Davies, J. Aminoglycoside antibiotic-inactivating enzymes in actinomycetes similar to those present in clinical isolates of antibiotic-resistant bacteria. Proc. Natl. Acad. Sci. USA 1973, 70, 2276–2280. [Google Scholar] [CrossRef] [PubMed]
  19. Ogawara, H. Antibiotic resistance in pathogenic and producing bacteria, with special reference to β-lactam antibiotics. Microbiol. Rev. 1981, 45, 591–619. [Google Scholar] [PubMed]
  20. Dantas, G.; Sommer, M.O.; Oluwasegun, R.D.; Church, G.M. Bacteria subsisting on antibiotics. Science 2008, 320, 100–103. [Google Scholar] [CrossRef] [PubMed]
  21. Riesenfeld, C.S.; Goodman, R.M.; Handelsman, J. Uncultured soil bacteria are a reservoir of new antibiotic resistance genes. Environ. Microbiol. 2004, 6, 981–989. [Google Scholar] [CrossRef] [PubMed]
  22. Wright, G.D. The antibiotic resistome: The nexus of chemical and genetic diversity. Nat. Rev. Microbiol. 2007, 5, 175–186. [Google Scholar] [CrossRef] [PubMed]
  23. D’Costa, V.M.; McGrann, K.M.; Hughes, D.W.; Wright, G.D. Sampling the antibiotic resistome. Science 2006, 311, 374–377. [Google Scholar] [CrossRef] [PubMed]
  24. Forsberg, K.J.; Reyes, A.; Wang, B.; Selleck, E.M.; Sommer, M.O.; Dantas, G. The shared antibiotic resistome of soil bacteria and human pathogens. Science 2012, 337, 1107–1111. [Google Scholar] [CrossRef] [PubMed]
  25. Mole, B. MRSA: Farming up trouble. Nature 2013, 499, 398–400. [Google Scholar] [CrossRef] [PubMed]
  26. Martinez, J.L. Antibiotics and antibiotic resistance genes in natural environments. Science 2008, 321, 365–367. [Google Scholar] [CrossRef] [PubMed]
  27. Aminov, R.I. Horizontal gene exchange in environmental microbiota. Front. Microl. 2011, 2. [Google Scholar] [CrossRef] [PubMed]
  28. Perry, J.A.; Wright, G.D. The antibiotic resistance “mobilome”: Searching for the link between environment and clinic. Front. Microbiol. 2013, 4. [Google Scholar] [CrossRef] [PubMed]
  29. Francino, M.P. Antibiotics and the human gut microbiome: Dysbioses and accumulation of resistance. Front. Micribiol. 2016, 6. [Google Scholar] [CrossRef] [PubMed]
  30. Gao, B.; Gupta, R.S. Phylogenetic framework and molecular signatures for the main clades of the phylum Actinobacteria. Microbiol. Mol. Biol. Rev. 2012, 76, 66–112. [Google Scholar] [CrossRef] [PubMed]
  31. Ventura, M.; Canchaya, C.; Tauch, A.; Chandra, G.; Fitzgerald, G.F.; Chater, K.F.; van Sinderen, D. Genomics of Actinobacteria: Tracing the evolutionary history of an ancient phylum. Microbiol. Mol. Biol. Rev. 2007, 71, 495–548. [Google Scholar] [CrossRef] [PubMed]
  32. Chater, K.F.; Biró, S.; Lee, K.J.; Palmer, T.; Schrempf, H. The complex extracellular biology of Streptomyces. FEMS Microbiol. Rev. 2010, 34, 171–198. [Google Scholar] [CrossRef] [PubMed]
  33. Liras, P.; Martin, J.F. Gene clusters for β-lactam antibiotics and control of their expression: Why have clusters evolved, and from where did they originate? Int. Microbiol. 2006, 9, 9–19. [Google Scholar] [PubMed]
  34. Liu, G.; Chater, K.F.; Chandra, G.; Niu, G.; Tan, H. Molecular regulation of antibiotic biosynthesis in streptomyces. Microbiol. Mol. Biol. Rev. 2013, 77, 112–143. [Google Scholar] [CrossRef] [PubMed]
  35. Kahan, J.S.; Kahan, F.M.; Goegelman, R.; Currie, S.A.; Jackson, M.; Stapley, E.; Miller, T.W.; Miller, A.K.; Hendlin, D.; Mochales, S.; et al. Thienamycin, a new beta-lactam antibiotic. I. Discovery, taxonomy, isolation and physical properties. J. Antibiot. 1979, 32, 1–12. [Google Scholar] [CrossRef] [PubMed]
  36. Reading, C.; Cole, M. Clavulanic acid: A beta-lactamase-inhiting beta-lactam from Streptomyces clavuligerus. Antimicrob. Agents Chemother. 1977, 11, 852–857. [Google Scholar] [CrossRef] [PubMed]
  37. Koch, A.L. Penicillin binding proteins, β-lactams, and lactamases: Offensives, attacks, and defensive countermeasures. Crit. Rev. Microbiol. 2000, 26, 205–220. [Google Scholar] [CrossRef] [PubMed]
  38. Davis, B.D.; Maas, W.K. Analysis of the biochemical mechanism of drug resistance in certain bacterial mutants. Proc. Natl. Acad. Sci. USA 1952, 38, 775–785. [Google Scholar] [CrossRef] [PubMed]
  39. The Comprehensive Antibiotic Resistance Database. Available online: http://arpcard.mcmaster.ca/ (accessed on 20 March 2016).
  40. King, D.T.; Sobhanifar, S.; Strynadka, N.C. One ring to rule them all: Current trends in combating bacterial resistance to the β-lactams. Protein Sci. 2016. [Google Scholar] [CrossRef] [PubMed]
  41. Cundliffe, E.; Demain, A.L. Avoidance of suicide in antibiotic-producing microbes. J. Ind. Microbiol. Biotechnol. 2010, 37, 643–672. [Google Scholar] [CrossRef] [PubMed]
  42. McArthur, A.G.; Waglechner, N.; Nizam, F.; Yan, A.; Azad, M.A.; Baylay, A.J.; Bhullar, K.; Canova, M.J.; De Pascale, G.; Ejim, L.; et al. The comprehensive antibiotic resistance database. Antimicrob. Agents Chemother. 2013, 57, 3348–3357. [Google Scholar] [CrossRef] [PubMed]
  43. Gibson, M.K.; Forsberg, K.J.; Dantas, G. Improved annotation of antibiotic resistance determinants reveals microbial resistomes cluster by ecology. ISME J. 2015, 9, 207–216. [Google Scholar] [CrossRef] [PubMed]
  44. Richmond, M.H.; Sykes, R.B. The β-lactamases of gram-negative bacteria and their possible physiological role. Adv. Microb. Physiol. 1973, 9, 31–88. [Google Scholar] [PubMed]
  45. Ogawara, H. Production and property of β-lactamases in Streptomyces. Antimicrob. Agents Chemother. 1975, 8, 402–408. [Google Scholar] [CrossRef] [PubMed]
  46. Abraham, E.P.; Waley, S.G. Beta-Lactamases; Hamilton-Miller, J.M.T., Smith, J.T., Eds.; Academic Press: London, UK, 1979; p. 338. [Google Scholar]
  47. Kushner, D.J.; Breuil, C. Penicillinase (β-lactamase) formation by blue-green algae. Arch. Microbiol. 1977, 112, 219–223. [Google Scholar] [CrossRef] [PubMed]
  48. Urbach, C.; Fastrez, J.; Soumillion, P. A new family of cyanobacterial penicillin-binding proteins. A missing link in the evolution of class A β-lactamases. J. Biol. Chem. 2008, 283, 32516–32526. [Google Scholar] [CrossRef] [PubMed]
  49. Podzelinska, K.; He, S.-M.; Wathier, M.; Yakunin, A.; Proudfoot, M.; Hove-Jensen, B.; Zechel, D.L.; Jia, Z. Structure of PhnP, a phosphodiesterase of the carbon-phosphorus lyase pathway for phosphonate degradation. J. Biol. Chem. 2009, 284, 17216–17226. [Google Scholar] [CrossRef] [PubMed]
  50. Sattler, S.A.; Wang, X.; Lewis, K.M.; DeHan, P.J.; Park, C.M.; Xin, Y.; Liu, H.; Xian, M.; Xun, L.; Kang, C. Characterizations of two bacterial persulfide dioxygenases of the metallo-β-lactamase superfamily. J. Biol. Chem. 2015, 290, 18914–18923. [Google Scholar] [CrossRef] [PubMed]
  51. Daiyasu, H.; Osaka, K.; Ishino, Y.; Toh, H. Expansion of the zinc metallo-hydrolase family of the β-lactamase fold. FEBS Lett. 2001, 503, 1–6. [Google Scholar] [CrossRef]
  52. Dominski, Z. Nucleases of the metallo-β-lactamase family and their role in DNA and RNA metabolism. Crit. Rev. Biochem. Mol. Biol. 2007, 42, 67–93. [Google Scholar] [CrossRef] [PubMed]
  53. Vogel, A.; Schilling, O.; Späth, B.; Marchfelder, A. The tRNase Z family of proteins: Physiological functions, substrate specificity and structural properties. Biol. Chem. 2005, 386, 1253–1264. [Google Scholar] [CrossRef] [PubMed]
  54. Lapidus, A.; Clum, A.; Labutti, K.; Kaluzhnaya, M.G.; Lim, S.; Beck, D.A.; del Glavina Rio, T.; Nolan, M.; Mavromatis, K.; Huntemann, M.; et al. Genomes of three methylotrophs from a single niche reveal the genetic and metabolic divergence of the Methylophilaceae. J. Bacteriol. 2011, 193, 3757–3764. [Google Scholar] [CrossRef] [PubMed]
  55. Limphong, P.; Nimako, G.; Thomas, P.W.; Fast, W.; Makaroff, C.A.; Crowder, M.W. Arabidopsis thaliana mitochondrial glyoxalase 2-1 exhibits β-lactamase activity. Biochemistry 2009, 48, 8491–8493. [Google Scholar] [CrossRef] [PubMed]
  56. Florea, L.; Di Francesco, V.; Miller, J.; Turner, R.; Yao, A.; Harris, M.; Walenz, B.; Mobarry, C.; Merkulov, G.V.; Charlab, R.; et al. Gene and alternative splicing annotation with AIR. Genome Res. 2005, 15, 54–66. [Google Scholar] [CrossRef] [PubMed]
  57. Smith, T.S.; Southan, C.; Ellington, K.; Campbell, D.; Tew, D.G.; Debouck, C. Identification, genomic organization, and mRNA expression of LACTB, encoding a serine β-lactamase-like protein with an amino-terminal transmembrane domain. Genomics 2001, 78, 12–14. [Google Scholar] [CrossRef] [PubMed]
  58. Toth, M.; Smith, C.; Frase, H.; Mobashery, S.; Vakulenko, S. An antibiotic-resistance enzyme from a deep-sea bacterium. J. Am. Chem. Soc. 2010, 132, 816–823. [Google Scholar] [CrossRef] [PubMed]
  59. Michaux, C.; Massant, J.; Kerff, F.; Frère, J.-M.; Docquier, J.D.; Vandenberghe, I.; Samyn, B.; Pierrard, A.; Feller, G.; Charlier, P.; et al. Crystal structure of a cold-adapted class C β-lactamase. FEBS J. 2008, 275, 1687–1697. [Google Scholar] [CrossRef] [PubMed]
  60. Kim, S.-B.; Joo, S.; Yoon, S.; Kim, S.; Moon, J.; Ryu, Y.; Kim, K.K.; Kim, T.D. Purification, crystallization and preliminary crystallographic analysis of Est-Y29: A novel oligomeric β-lactamase. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2009, 65, 310–312. [Google Scholar] [CrossRef] [PubMed]
  61. Allen, H.K.; Moe, L.; Rodbumrer, J.; Gaarder, A.; Handelsman, J. Functional metagenomics reveals diverse β-lactamases in a remote Alaskan soil. ISME J. 2009, 3, 243–251. [Google Scholar] [CrossRef] [PubMed]
  62. Donato, J.J.; Moe, L.A.; Converse, B.J.; Smart, K.D.; Berklein, F.; McManus, P.S.; Handelsman, J. Metagenomic analysis of apple orchard soil reveals antibiotic resistance genes encoding predicted bifunctional proteins. Appl. Environ. Microbiol. 2010, 76, 4396–4401. [Google Scholar] [CrossRef] [PubMed]
  63. Sun, L.; Zhang, L.; Zhang, H.; He, Z.G. Characterization of a bifunctional β-lactamase/ribonuclease and its interaction with a chaperone-like protein in the pathogen Mycobacterium tuberculosis H37Rv. Biochemistry 2011, 76, 350–358. [Google Scholar] [CrossRef] [PubMed]
  64. Singh, G.; Kumar, A.; Arya, S.; Gupta, U.D.; Singh, K.; Kaur, J. Characterization of a novel esterase Rv1497 of Mycobacterium tuberculosis H37Rv demonstrating β-lactamase activity. Enzyme Microb. Technol. 2016, 82, 180–190. [Google Scholar] [CrossRef] [PubMed]
  65. Ambler, R.P. The structure of beta-lactamases. Philos. Trans. R. Soc. Lond. B Biol. Sci. 1980, 289, 321–331. [Google Scholar] [CrossRef] [PubMed]
  66. Bush, K.; Jacoby, G.A. Updated functional classification of β-lactamases. Antimicrob. Agents Chemother. 2010, 54, 969–976. [Google Scholar] [CrossRef] [PubMed]
  67. Jaurin, B.; Grundström, T. ampC cephalosporinase of Escherichia coli K-12 has a different evolutionary origin from that of beta-lactamases of the penicillinase type. Proc. Natl. Acad. Sci. USA 1981, 78, 4897–4901. [Google Scholar] [CrossRef] [PubMed]
  68. Ouellette, M.; Bissonnette, L.; Roy, P.H. Precise insertion of antibiotic resistance determinants into Tn21-like transposons: Nucleotide sequence of the OXA-1 β-lactamase gene. Proc. Natl. Acad. Sci. USA 1987, 84, 7378–7382. [Google Scholar] [CrossRef] [PubMed]
  69. Palzkill, T. Metallo-β-lactamase structure and function. Ann. N. Y. Acad. Sci. 2013, 1277, 91–104. [Google Scholar] [CrossRef] [PubMed]
  70. Frère, J.M.; Galleni, M.; Bush, K.; Dideberg, O. Is it necessary to change the classification of β-lactamases? J. Antimicrob. Chemother. 2005, 55, 1051–1053. [Google Scholar] [CrossRef] [PubMed]
  71. Bush, K.; Jacoby, G.A.; Medeiros, A.A. A functional classification scheme for β-lactamases and its correlation with molecular structure. Antimicrob. Agents Chemother. 1995, 39, 1211–1233. [Google Scholar] [CrossRef] [PubMed]
  72. Bradford, P.A. Extended-spectrum β-lactamases in the 21st century: Characterization, epidemiology, and detection of this important resistance threat. Clin. Microbiol. Rev. 2001, 14, 933–951. [Google Scholar] [CrossRef] [PubMed]
  73. Pettinati, I.; Brem, J.; Lee, S.Y.; McHugh, P.J.; Schofield, C.J. The chemical biology of human metallo-β-lactamase fold proteins. Trends Biochem. Sci. 2016, 41, 338–355. [Google Scholar] [CrossRef] [PubMed]
  74. Bebrone, C. Metallo-β-lactamases (classification, activity, genetic organization, structure, zinc coordination) and their superfamily. Biochem. Pharmacol. 2007, 74, 1686–1701. [Google Scholar] [CrossRef] [PubMed]
  75. Garau, G.; Di Guilmi, A.M.; Hall, B.G. Structure-based phylogeny of the metallo-β-lactamases. Antimicrob. Agents Chemother. 2005, 49, 2778–2784. [Google Scholar] [CrossRef] [PubMed]
  76. Walsh, T.R.; Toleman, M.A.; Poirel, L.; Nordmann, P. Metallo-β-lactamases: The quiet before the storm? Clin. Microbiol. Rev. 2005, 18, 306–325. [Google Scholar] [CrossRef] [PubMed]
  77. Ogawara, H. Phylogenetic tree and sequence similarity of β-lactamases. Mol. Phylogenet. Evol. 1993, 2, 97–111. [Google Scholar] [CrossRef] [PubMed]
  78. Ogawara, H. Self-resistance to β-lactam antibiotics in Streptomyces (in Japanese). Bull. Meiji Pharmaceut. Univ. 2014, 43, 1–20. [Google Scholar]
  79. Ogawara, H. Molecular phylogenetics of β-lactamases in Actinobacteria. Bull. Meiji Pharmaceut. Univ. 2013, 42, 1–18. [Google Scholar]
  80. BioProject. Available online: http://www.ncbi.nlm.nih.gov/guide/all/ (accessed on 10 September 2011).
  81. Stackebrandt, E.; Rainey, F.A.; Ward-Rainey, N.L. Proposal for a New Hierarchic Classification System, Actinobacteria classis nov. Int. J. Syst. Bacteriol. 1997, 47, 479–491. [Google Scholar] [CrossRef]
  82. Ogawara, H.; Kawamura, N.; Kudo, T.; Suzuki, K.I.; Nakase, T. Distribution of β-lactamases in actinomycetes. Antimicrob. Agents Chemother. 1999, 43, 3014–3017. [Google Scholar] [PubMed]
  83. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [PubMed]
  84. Notredame, C.; Higgins, D.G.; Heringa, J. T-Coffee: A novel method for fast and accurate multiple sequence alignment. J. Mol. Biol. 2000, 302, 205–217. [Google Scholar] [CrossRef] [PubMed]
  85. Galleni, M.; Lamotte-Brasseur, J.; Rossolini, G.M.; Spencer, J.; Dideberg, O.; Frère, J.-M. Metallo-β-lactamases Working Group. Standard numbering scheme for class B beta-lactamases. Antimicrob. Agents Chemother. 2001, 45, 660–663. [Google Scholar] [CrossRef] [PubMed]
  86. Lobkovsky, E.; Moews, P.C.; Liu, H.; Zhao, H.; Frere, J.-M.; Knox, J.R. Evolution of an enzyme activity: Crystallographic structure at 2-A resolution of cephalosporinase from the ampC gene of Enterobacter cloacae P99 and comparison with a class A penicillinase. Proc. Natl. Acad. Sci. USA 1993, 90, 11257–11261. [Google Scholar] [CrossRef] [PubMed]
  87. Goffin, C.; Ghuysen, J.M. Biochemistry and comparative genomics of SxxK superfamily acyltransferases offer a clue to the mycobacterial paradox: Presence of penicillin-susceptible target proteins versus lack of efficiency of penicillin as therapeutic agent. Microbiol. Mol. Biol. Rev. 2002, 66, 702–738. [Google Scholar] [CrossRef] [PubMed]
  88. Larkin, M.A.; Blackshields, G.; Brown, N.P.; Chenna, R.; McGettigan, P.A.; McWilliam, H.; Valentin, F.; Wallace, I.M.; Wilm, A.; Lopez, R.; et al. Clustal W and Clustal X version 2.0. Bioinformatics 2007, 23, 2947–2948. [Google Scholar] [CrossRef] [PubMed]
  89. β-Lactamase Classification and Amino Acid Sequences for TEM, SHV and OXA Extended-Spectrum and Inhibitor Resistant Enzymes. Available online: http://www.lahey.org/studies/ (accessed on 20 March 2016).
  90. Protdist-Program to Compute Distance Matrix from Protein Sequences. Available online: http://evolution.genetics.washington.edu/phylip/doc/protdist.html (accessed on 20 March 2016).
  91. Urabe, H.; Toyama, K.; Ogawara, H. Cloning from Streptomyces cellulosae of the gene encoding beta-lactamase, a blue-dextran binding protein. J. Antibiot. 1990, 43, 1483–1488. [Google Scholar] [CrossRef] [PubMed]
  92. Ogawara, H. Sequence of a gene encoding β-lactamase from Streptomyces cellulosae. Gene 1993, 124, 111–114. [Google Scholar] [CrossRef]
  93. Jensen, S.E.; Paradkar, A.S.; Mosher, R.H.; Anders, C.; Beatty, P.H.; Brumlik, M.J.; Griffin, A.; Barton, B. Five additional genes are involved in clavulanic acid biosynthesis in Streptomyces clavuligerus. Antimicrob. Agents Chemother. 2004, 48, 192–202. [Google Scholar] [CrossRef] [PubMed]
  94. Li, R.; Khaleeli, N.; Townsend, C.A. Expansion of the clavulanic acid gene cluster: Identification and in vivo functional analysis of three new genes required for biosynthesis of clavulanic acid by Streptomyces clavuligerus. J. Bacteriol. 2000, 182, 4087–4095. [Google Scholar] [CrossRef] [PubMed]
  95. Jacoby, G.A. AmpC β-lactamases. Clin. Microbiol. Rev. 2009, 22, 161–182. [Google Scholar] [CrossRef] [PubMed]
  96. Macheboeuf, P.; Contreras-Martel, C.; Job, V.; Dideberg, O.; Dessen, A. Penicillin binding proteins: Key players in bacterial cell cycle and drug resistance processes. FEMS Microbiol. Rev. 2006, 30, 673–691. [Google Scholar] [CrossRef] [PubMed]
  97. Typas, A.; Banzhaf, M.; Gross, C.A.; Vollmer, W. From the regulation of peptidoglycan synthesis to bacterial growth and morphology. Nat. Rev. Microbiol. 2012, 10, 123–136. [Google Scholar] [CrossRef] [PubMed]
  98. Van Heijenoort, J. Lipid intermediates in the biosynthesis of bacterial peptidoglycan. Microbiol. Mol. Biol. Rev. 2007, 71, 620–635. [Google Scholar] [CrossRef] [PubMed]
  99. Suginaka, H.; Blumberg, P.M.; Strominger, J.L. Multiple penicillin-binding components in Bacillus subtilis, Bacillus cereus, Staphylococcus aureus, and Escherichia coli. J. Biol. Chem. 1972, 247, 5279–5288. [Google Scholar] [PubMed]
  100. Spratt, B.G. Distinct penicillin binding proteins involved in the division, elongation, and shape of Escherichia coli K12. Proc. Natl. Acad. Sci. USA 1975, 72, 2999–3003. [Google Scholar] [CrossRef] [PubMed]
  101. Pratt, R.F. Substrate specificity of bacterial DD-peptidases (penicillin-binding proteins). Cell. Mol. Life Sci. 2008, 65, 2138–2155. [Google Scholar] [CrossRef] [PubMed]
  102. Sauvage, E.; Kerff, F.; Terrak, M.; Ayala, J.A.; Charlier, P. The penicillin-binding proteins: Structure and role in peptidoglycan biosynthesis. FEMS Microbiol. Rev. 2008, 32, 234–258. [Google Scholar] [CrossRef] [PubMed]
  103. Ogawara, H. Penicillin-binding proteins in Actinobacteria. J. Antib. 2015, 68, 223–245. [Google Scholar] [CrossRef] [PubMed]
  104. Goffin, C.; Ghuysen, J.M. Multimodular penicillin-binding proteins: An enigmatic family of orthologs and paralogs. Microbiol. Mol. Biol. Rev. 1998, 62, 1079–1093. [Google Scholar] [PubMed]
  105. Ishida, K.; Hung, T.V.; Liou, K.; Lee, H.C.; Shin, C.H.; Sohng, J.K. Characterization of pbpA and pbp2 encoding penicillin-binding proteins located on the downstream of clavulanic acid gene cluster in Streptomyces clavuligerus. Biotechnol. Lett. 2006, 28, 409–417. [Google Scholar] [CrossRef] [PubMed]
  106. Ogawara, H.; Horikawa, S. Penicillin-binding proteins of Streptomyces cacaoi, Streptomyces olivaceus, and Streptomyces clavuligerus. Antimicrob. Agents Chemother. 1980, 17, 1–7. [Google Scholar] [CrossRef] [PubMed]
  107. Paradkar, A.S.; Aidoo, K.A.; Wong, A.; Jensen, S.E. Molecular analysis of a β-lactam resistance gene encoded within the cephamycin gene cluster of Streptomyces clavuligerus. J. Bacteriol. 1996, 178, 6266–6274. [Google Scholar] [PubMed]
  108. Hanks, S.K.; Hunter, T. Protein kinases 6. The eukaryotic protein kinase superfamily: Kinase (catalytic) domain structure and classification. FASEB J. 1995, 9, 576–596. [Google Scholar] [PubMed]
  109. Hunter, T. Tyrosine phosphorylation: Thirty years and counting. Curr Opin. Cell Biol. 2009, 21, 140–146. [Google Scholar] [CrossRef] [PubMed]
  110. Muñoz-Dorado, J.; Inouye, S.; Inouye, M. A gene encoding a protein serine/threonine kinase is required for normal development of M. xanthus, a gram-negative bacterium. Cell 1991, 67, 995–1006. [Google Scholar] [CrossRef]
  111. Pereira, S.F.; Goss, L.; Dworkin, J. Eukaryote-like serine/threonine kinases and phosphatases in bacteria. Microbiol. Mol. Biol. Rev. 2011, 75, 192–212. [Google Scholar] [CrossRef] [PubMed]
  112. Manuse, S.; Fleurie, A.; Zucchini, L.; Lesterlin, C.; Grangeasse, C. Role of eukaryotic-like serine/threonine kinases in bacterial cell division and morphogenesis. FEMS Microbiol. Rev. 2016, 40, 41–56. [Google Scholar] [CrossRef] [PubMed]
  113. Ulrich, L.E.; Koonin, E.V.; Zhulin, I.B. One-component systems dominate signal transduction in prokaryotes. Trends. Microbiol. 2005, 13, 52–56. [Google Scholar] [CrossRef] [PubMed]
  114. Yeats, C.; Finn, R.D.; Bateman, A. The PASTA domain: A β-lactam-binding domain. Trends Biochem. Sci. 2002, 27, 438–440. [Google Scholar] [CrossRef]
  115. Gordon, E.; Mouz, N.; Duée, E.; Dideberg, O. The crystal structure of the penicillin-binding protein 2x from Streptococcus pneumoniae and its acyl-enzyme form: Implication in drug resistance. J. Mol. Biol. 2000, 299, 477–485. [Google Scholar] [CrossRef] [PubMed]
  116. Calvanese, L.; Falcigno, L.; Maglione, C.; Marasco, D.; Ruggiero, A.; Squeglia, F.; Berisio, R.; D’Auria, G. Structural and binding properties of the PASTA domain of PonA2, a key penicillin binding protein from Mycobacterium tuberculosis. Biopolymers 2014, 101, 712–719. [Google Scholar] [CrossRef] [PubMed]
  117. Ruggiero, A.; De Simone, P.; Smaldone, G.; Squeglia, F.; Berisio, R. Bacterial cell division regulation by Ser/Thr kinases: A structural perspective. Curr. Protein Pept. Sci. 2012, 13, 756–766. [Google Scholar] [CrossRef] [PubMed]
  118. Peters, K.; Schweizer, I.; Beilharz, K.; Stahlmann, C.; Veening, J.W.; Hakenbeck, R.; Denapaite, D. Streptococcus pneumoniae PBP2x mid-cell localization requires the C-terminal PASTA domains and is essential for cell shape maintenance. Mol. Microbiol. 2014, 92, 733–755. [Google Scholar] [CrossRef] [PubMed]
  119. Ogawara, H. Distribution of PASTA domains in penicillin-binding proteins and serine/threonine kinases of Actinobacteria. J. Antibiot. 2016. [Google Scholar] [CrossRef] [PubMed]
  120. McCormick, J.R.; Flärdh, K. Signals and regulators that govern Streptomyces development. FEMS Micribiol. Rev. 2012, 36, 206–231. [Google Scholar] [CrossRef] [PubMed]
  121. Ladwig, N.; Franz-Wachtel, M.; Hezel, F.; Soufi, B.; Macek, B.; Wohlleben, W.; Muth, G. Control of morphological differentiation of Streptomyces coelicolor A3(2) by phosphorylation of MreC and PBP2. PLoS ONE 2015, 10, e0125425. [Google Scholar] [CrossRef] [PubMed]
  122. Jones, G.; Del Sol, R.; Dudley, E.; Dyson, P. Forkhead-associated proteins genetically linked to the serine/threonine kinase PknB regulate carbon flux towards antibiotic biosynthesis in Streptomyces coelicolor. Microb. Biotechnol. 2011, 4, 263–274. [Google Scholar] [CrossRef] [PubMed]
  123. Pérez-Llarena, F.; Martín, J.F.; Galleni, M.; Coque, J.J.; Fuente, J.L.; Frère, J.M.; Liras, P. The bla gene of the cephamycin cluster of Streptomyces clavuligerus encodes a class A β-lactamase of low enzymatic activity. J. Bacteriol. 1997, 179, 6028–6034. [Google Scholar]
  124. Streptomyces clavuligerus ATCC 27064 chromosome, whole genome shotgun sequence. Available online: http://www.ncbi.nlm.nih.gov/nuccore/294323433?report=graph (accessed on 20 March 2016).
  125. Streptomyces cattleya NRRL 8057 = DSM 46488, complete genome. Available online: http://www.ncbi.nlm.nih.gov/nuccore/357397620?report=graph (accessed on 20 March 2016).
  126. Jensen, S.E. Biosynthesis of clavam metabolites. J. Ind. Microbiol. Biotechnol. 2012, 39, 1407–1419. [Google Scholar] [CrossRef] [PubMed]
  127. Paradkar, A. Clavulanic acid production by Streptomyces clavuligerus: Biogenesis, regulation and strain improvement. J. Antibiot. 2013, 66, 411–420. [Google Scholar] [CrossRef] [PubMed]
  128. Mellado, E.; Lorenzana, L.M.; Rodríguez-Sáiz, M.; Díez, B.; Liras, P.; Barredo, J.L. The clavulanic acid biosynthetic cluster of Streptomyces clavuligerus: Genetic organization of the region upstream of the car gene. Microbiology 2002, 148, 1427–1438. [Google Scholar]
  129. Coque, J.J.; Liras, P.; Martín, J.F. Genes for a β-lactamase, a penicillin-binding protein and a transmembrane protein are clustered with the cephamycin biosynthetic genes in Nocardia lactamdurans. EMBO J. 1993, 12, 631–639. [Google Scholar] [PubMed]
  130. Kimura, H.; Izawa, M.; Sumino, Y. Molecular analysis of the gene cluster involved in cephalosporin biosynthesis from Lysobacter lactamgenus YK90. Appl. Microbiol. Biotechnol. 1996, 44, 589–596. [Google Scholar] [CrossRef] [PubMed]
  131. Martin, J.F. Clusters of genes for the biosynthesis of antibiotics: Regulatory genes and overproduction of pharmaceuticals. J. Ind. Microbiol. 1992, 9, 73–90. [Google Scholar] [CrossRef] [PubMed]
  132. Brakhage, A.A.; Al-Abdallah, Q.; Tüncher, A.; Spröte, P. Evolution of β-lactam biosynthesis genes and recruitment of trans-acting factors. Phytochemistry 2005, 66, 1200–1210. [Google Scholar] [CrossRef] [PubMed]
  133. Barka, E.A.; Vatsa, P.; Sanchez, L.; Gaveau-Vaillant, N.; Jacquard, C.; Klenk, H.P.; Clément, C.; Ouhdouch, Y.; van Wezel, G.P. Taxonomy, Physiology, and Natural Products of Actinobacteria. Microbiol. Mol. Biol. Rev. 2016, 80. [Google Scholar] [CrossRef] [PubMed]
  134. Alanis, A.J. Resistance to antibiotics: Are we in the post-antibiotic era? Arch. Med. Res. 2005, 36, 697–705. [Google Scholar] [CrossRef] [PubMed]
  135. Cooper, M.A.; Shlaes, D. Fix the antibiotics pipeline. Nature 2014, 472. [Google Scholar] [CrossRef] [PubMed]
  136. Rokem, J.S.; Lantz, A.E.; Nielsen, J. Systems biology of antibiotic production by microorganisms. Nat. Prod. Rep. 2007, 24, 1262–1287. [Google Scholar] [CrossRef] [PubMed]
  137. Bull, A.T.; Stach, J.E. Marine actinobacteria: New opportunities for natural product search and discovery. Trends Microbiol. 2007, 15, 491–499. [Google Scholar] [CrossRef] [PubMed]
  138. Demain, A.L. Importance of microbial natural products and the need to revitalize their discovery. J. Ind. Microbiol. Biotechnol. 2014, 41, 185–201. [Google Scholar] [CrossRef] [PubMed]
  139. Bachmann, B.O.; Van Lanen, S.G.; Baltz, R.H. Microbial genome mining for accelerated natural products discovery: Is a renaissance in the making? J. Ind. Microbiol. Biotechnol. 2014, 41, 175–184. [Google Scholar] [CrossRef] [PubMed]
  140. Katz, L.; Baltz, R.H. Natural product discovery: Past, present, and future. J. Ind. Microbiol. Biotechnol. 2016, 43, 155–176. [Google Scholar] [CrossRef] [PubMed]
  141. Baltz, R.H. Genetic manipulation of secondary metabolite biosynthesis for improved production in Streptomyces and other actinomycetes. J. Ind. Microbiol. Biotechnol. 2016, 43, 343–370. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Phylogenetic tree of putative 20 class A β-lactamases on the basis of amino acid sequences with those of Staphylococcus aureus (M15526), Pseudomonas aeruginosa (Q03170), Klebsiella pneumoniae (TEM-1, YP_005351445), Klebsiella pneumoniae (SHV-11, X98101), and Escherichia coli (SHV-1, AF148850) as reference sequences. The tree was constructed by using ClustalX 2 [88] and E. coli AmpC (ABM69263) as outgroup. Clusters A and B show blue-dextran/NADP+-non-binding and binding β-lactamases from Streptomyces, respectively and, cluster C shows those from pathogenic bacteria used as reference sequences. The bootstrap probabilities are shown at branching nodes.
Figure 1. Phylogenetic tree of putative 20 class A β-lactamases on the basis of amino acid sequences with those of Staphylococcus aureus (M15526), Pseudomonas aeruginosa (Q03170), Klebsiella pneumoniae (TEM-1, YP_005351445), Klebsiella pneumoniae (SHV-11, X98101), and Escherichia coli (SHV-1, AF148850) as reference sequences. The tree was constructed by using ClustalX 2 [88] and E. coli AmpC (ABM69263) as outgroup. Clusters A and B show blue-dextran/NADP+-non-binding and binding β-lactamases from Streptomyces, respectively and, cluster C shows those from pathogenic bacteria used as reference sequences. The bootstrap probabilities are shown at branching nodes.
Molecules 21 00605 g001
Figure 2. Phylogenetic tree of putative 63 class B β-lactamases on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and Escherichia coli glyoxalase II (CDZ19109) as outgroup. β-Lactamases from Fluoribacter gormanii (CAB96921), Elizabethkingia meningoseptica GOB-1 (AF090141_1), Aeromonas hydrophila (CAA40386), Bacillus cereus (AAA22276), Acinetobacter pittii IMP_1 (ADI87504), and Bacteroides fragilis (AAA22904) were used as reference sequences. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S4). The bootstrap probabilities are shown at branching nodes.
Figure 2. Phylogenetic tree of putative 63 class B β-lactamases on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and Escherichia coli glyoxalase II (CDZ19109) as outgroup. β-Lactamases from Fluoribacter gormanii (CAB96921), Elizabethkingia meningoseptica GOB-1 (AF090141_1), Aeromonas hydrophila (CAA40386), Bacillus cereus (AAA22276), Acinetobacter pittii IMP_1 (ADI87504), and Bacteroides fragilis (AAA22904) were used as reference sequences. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S4). The bootstrap probabilities are shown at branching nodes.
Molecules 21 00605 g002
Figure 3. Phylogenetic tree of putative 94 class C β-lactamases on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and Streptomyces cacaoi class A β-lactamase (D90201) as outgroup. β-lactamases of Rhodobacter sphaeroides (YP_355265), Mycobacterium smegmatis (NC_018289), Acinetobacter baumannii (CAB77444), Aeromonas caviae (AF462690_1), and E. coli (ABM69263 and NP_418574) were used as reference sequences. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S6). The bootstrap probabilities are shown at branching nodes.
Figure 3. Phylogenetic tree of putative 94 class C β-lactamases on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and Streptomyces cacaoi class A β-lactamase (D90201) as outgroup. β-lactamases of Rhodobacter sphaeroides (YP_355265), Mycobacterium smegmatis (NC_018289), Acinetobacter baumannii (CAB77444), Aeromonas caviae (AF462690_1), and E. coli (ABM69263 and NP_418574) were used as reference sequences. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S6). The bootstrap probabilities are shown at branching nodes.
Molecules 21 00605 g003
Figure 4. Phylogenetic tree of putative 100 class A PBPs on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and S. coelicolor putative penicillin acylase (SCO4049, CAC32320) as outgroup. E. coli PBP1A (subclass A1, PBPA_ECOLI), E. coli PBP1B (subclass A2, PBPB_ECOLI), S. pneumoniae PBP1A (subclass A3, PBPA_STRPN), S. pneumoniae PBP2A (subclass A4, AJ002292), and S. pneumoniae PBP1B (subclass A5, AJ002291) were used as reference sequences. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S8). The PBPs in groups A, B, and C are corresponding to those in clusters A, B, and C in the amino acid alignment shown in Table S7. The bootstrap probabilities are shown at branching nodes.
Figure 4. Phylogenetic tree of putative 100 class A PBPs on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and S. coelicolor putative penicillin acylase (SCO4049, CAC32320) as outgroup. E. coli PBP1A (subclass A1, PBPA_ECOLI), E. coli PBP1B (subclass A2, PBPB_ECOLI), S. pneumoniae PBP1A (subclass A3, PBPA_STRPN), S. pneumoniae PBP2A (subclass A4, AJ002292), and S. pneumoniae PBP1B (subclass A5, AJ002291) were used as reference sequences. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S8). The PBPs in groups A, B, and C are corresponding to those in clusters A, B, and C in the amino acid alignment shown in Table S7. The bootstrap probabilities are shown at branching nodes.
Molecules 21 00605 g004
Figure 5. Phylogenetic tree of putative 161 class B PBPs on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and S. coelicolor putative penicillin acylase (SCO4049, CAC32320) as outgroup. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S9). The bootstrap probabilities are shown at branching nodes. The phylogenetic tree were arbitrary divided to three groups, A, B, and C.
Figure 5. Phylogenetic tree of putative 161 class B PBPs on the basis of amino acid sequences. The tree was constructed by using ClustalX 2 [88] and S. coelicolor putative penicillin acylase (SCO4049, CAC32320) as outgroup. The sequences marked with yellow boxes were used for the calculation of molecular distances (Table S9). The bootstrap probabilities are shown at branching nodes. The phylogenetic tree were arbitrary divided to three groups, A, B, and C.
Molecules 21 00605 g005
Table 1. Classification of β-lactamases in 12 Streptomyces species.
Table 1. Classification of β-lactamases in 12 Streptomyces species.
BacteriaPrefixClass AClass BClass CClass D
Streptomyces albulus CCRC11814K53001315, 35835 *18626, 25018, 2524002767, 10843, 40281none
Streptomyces cattleya NRRL805SCAT0807, 1418, 4581, 56920258, 1598, 4145, 4517, 48221363, 1488, 5541, 5637none
Streptomyces clavuligerus ATCC27064SSCG00130, 00160, 0146700563, 00791, 02181, 04638, 05032, 05130, 05295, 0557101647, 03303, 03668, 04090, 04120, 04299none
Streptomyces collinus Tu365B44602285, 3301002000, 02210, 05525, 13210, 29385, 33085, 3329001315, 03940, 11745, 11965, 12445, 13750, 19000, 22405, 33975none
Streptomyces davawensis JCM4913BN15957900265, 0293, 4295, 5380, 5740, 6071, 7079,7978, 82911677, 1822, 3452, 3645, 4558, 5112, 5953, 5970, 6145, 7932, 8163none
Streptomyces hygroscopicus subsp. jinggangensis 5008SHJG8264, 82660444, 1156, 1422, 2523, 3996, 4046, 83350734, 0750, 1734, 2053, 2219, 2828, 3769, 3873, 4357, 4919, 5160, 8639none
Streptomyces rimosus subsp. rimosus ATCC10970SRIM07618, 1053108043, 16115, 16845, 19739, 21549, 23716, 2642200410, 02396, 04151, 04583, 10671, 11591, 15850, 24376, 28551, 31085, 32256, 37671 none
Streptomyces scabiei 87.22SCAB387312518126931, 46171, 52151, 55861none
Streptomyces sp. PAMC26508F7502387, 6336, 68230075, 23740558, 2251, 3609, 5886none
Streptomyces tsukubaensis NRRL18488STSU15167, 15177, 3233007338, 08188, 24671, 3162000260, 00675, 00895, 15744, 17713, 20672, 23056, 33145, none
Streptomyces venezuelae ATCC 10712SVEN00890710, 1591, 2335, 2660, 72910306, 0457, 1972, 2198, 3165, 3520, 3691, 5322, 6030, 6060, 7092none
Streptomyces viridochromogenes DSM 40736SSQG0112702569, 02955, 05776, 05938, 0640900225, 00387, 00417, 02214, 02353, 02491, 02679, 03888, 04096, 04848, 06609none
*: Prefix K530-RS; Accession Number: WP_020929156 and WP_037636700, respectively.
Table 2. Presence or absence of active site residues in class A β-lactamases.
Table 2. Presence or absence of active site residues in class A β-lactamases.
β-LactamaseSXXK73SDN132RW/YE166D179KT/SG236
YP_004910945_SCAT_1418XXXX
YP_004914072_SCAT_4581XXXX
ZP_05002833_SSCG_00160
YP_007863245_F750_6336
YP_006249403_SHJG_8264XE
YP_008384100_B446_02285X
YP_008390217_B446_33010X
EIF91547_STSU_15177X
YP_004910346_SCAT_0807X
YP_004915182_SCAT_5692HYE
ZP_05002803_SSCG_00130QYE
WP_037636700_K530_RS35835
EIF91545_STSU_15167
ELQ83985_SRIM_07318XX
WP_020929156_K530_RS01315
YP_006249405_SHJG_8266
ELQ83341_SRIM_10531
YP_003489505_SCAB_38731
YP_007524296_BN159_5790
ZP_05003887_SSCG_01467
YP_006875636_SVEN_0089
ZP_07302240_SSQG_01127
YP_007859362_F750_2387XXEX
EIF88144_STSU_32330XX
YP_007863731_F750_6823
○: presence; X: absence; △: probable; other symbols are for amino acid residues.
Table 3. The list of active site residues discussed in this paper.
Table 3. The list of active site residues discussed in this paper.
Class A β-lactamasesS70XXK73S130DN132R164W/YE166D179K234T/SG236
Class B β-lactamases:H116XHXD120G123H196H263
Class C β-lactamasesS64XXK67Y150XN152 K315TG317S/A318
Class D β-lactamasesS67TFK70S115XVY142GW154K205TGW221XXG
Class A PBPsS465XXK468S524XN526 K716TG718
Class B PBPs:S307XXK310S356XN358 S356XN358
The amino acid numbering follows Ambler [65] for classes A and D, Galleni et al. [85] for class B, and Lobkovsky et al. [86] for class C β-lactamases, and Goffin and Ghuysen [87] for PBPs.
Table 4. Presence or absence of active site residues in class B β-lactamases.
Table 4. Presence or absence of active site residues in class B β-lactamases.
β-LactamaseHXHXD120G123H196H263
EOT02442_K530_18626DXX
YP_006878205_SVEN_2660DXX
YP_007523886_BN159_5380DXX
ZP_07304068_SSQG_02955DXX
ZP_05007878_SSCG_05130
YP_006249473_SHJG_8335
YP_004913644_SCAT_4145
ELQ82101_SRIM_16845
ZP_07307522_SSQG_06409EXR
AAA22276_Bacillus_cereus
YP_008384085_B446_02210X
YP_008390232_B446_33085X
ZP_05004854_SSCG_02181
YP_004909803_SCAT_0258
ELQ82279_SRIM_16115
ZP_05007361_SSCG_04638LR
YP_007524577_BN159_6071L
YP_007857070_F750_0075Y
ZP_05003465_SSCG_00791E
ZP_05007967_SSCG_05295
ELQ83811_SRIM_08043
YP_004911125_SCAT_1598
ZP_07303682_SSQG_02569
YP_007524246_BN159_5740
YP_008386271_B446_13210
YP_006245191_SHJG_4046
YP_007859349_F750_2374
YP_006877880_SVEN_2335
ZP_05007774_SSCG_05032
EIF89665_STSU_24671
YP_007518799_BN159_0293N
YP_006242571_SHJG_1422W
ELQ80751_SRIM_23716W
YP_006242306_SHJG_1156WX
ELQ81190_SRIM_21549WT
YP_006241594_SHJG_0444WD
YP_006245141_SHJG_3996WX
YP_008384043_B446_02000S
YP_008389494_B446_29385GA
ZP_07307051_SSQG_05938GA
YP_004914313_SCAT_4822GA
YP_004914015_SCAT_4517GT
YP_003488184_SCAB_25181AGS
ZP_07306889_SSQG_05776GS
ELQ81507_SRIM_19739GS
ZP_05008219_SSCG_05571GS
EIF92931_STSU_08188GS
YP_007526484_BN159_7978
YP_007526796_BN159_8291
EOT01146_K530_25240X
EIF93163_STSU_07338X
ELQ80212_SRIM_26422S
YP_008384742_B446_05525
YP_006243671_SHJG_2523
EIF88273_STSU_31620A
YP_006876256_SVEN_0710A
○: presence; X: absence; △: probable; other symbols are for amino acid residues.
Table 5. Presence or absence of active site residues in class C β-lactamases.
Table 5. Presence or absence of active site residues in class C β-lactamases.
β-LactamaseSXXK70YXN152KTG317
YP_006249777_SHJG_8639XX
YP_006243975_SHJG_2828XX
YP_006242882_SHJG_1734X
AGS69568_B446_13750
YP_006879065_SVEN_3520HTG
ZP_05006222_SSCG_03668XXX
AmpC_Escherichia_coli
ELQ82327_SRIM_15850YSSHDG
ELQ79325_SRIM_31085XXX
YP_003491153_SCAB_55861YSGHDG
EIF89980_STSU_23056_391
YP_006878710_SVEN_3165XRAG
YP_007520183_BN159_1677HTG
YP_007526438_BN159_7932XYSHX
YP_006882637_SVEN_7092YSHX
YP_003488355_SCAB_26931HDG
ELQ85060_SRIM_02396
YP_006880867_SVEN_5322
ZP_07305209_SSQG_04096
YP_004915128_SCAT_5637YNG
YP_006241884_SHJG_0734YNG
YP_006881575_SVEN_6030HNG
YP_008384429_B446_03940HNG
YP_006243201_SHJG_2053HNG
EIF94426_STSU_00675HNG
ZP_05005976_SSCG_03303X
YP_006246060_SHJG_4919YRG
YP_006877518_SVEN_1972HGG
ELQ78022_SRIM_37671HGG
YP_008388104_B446_22405CSNRSG
ZP_07303604_SSQG_02491XHDG
ELQ79790_SRIM_28551HNG
YP_007522151_BN159_3645HSG
ZP_07305961_SSQG_04848HNG
ZP_05006696_SSCG_04120HSG
ZP_07301500_SSQG_00387KSG
ZP_07303792_SSQG_02679HGG
EIF87980_STSU_33145YHAX
YP_007521958_BN159_3452YHSX
YP_007524459_BN159_5953YHA
ELQ84672_SRIM_04151YHAHGG
YP_007523618_BN159_5112YHAHPG/RGG
YP_006245502_SHJG_4357YHGHTG/RGG
ZP_05006916_SSCG_04299YHGHTG/RGG
EIF91419_STSU_15744YHGHTG/RGG
YP_004915032_SCAT_5541YSVX
YP_007520328_BN159_1822YSVX
ZP_07307722_SSQG_06609YSV
EOT05599_K530_02767YNVX
EIF94507_STSU_00260YHVX
EIF90453_STSU_20672YNT
YP_008383910_B446_01315YDTRVG
YP_008390406_B446_33975YDTRVG
YP_006241900_SHJG_0750YGTRYG
ELQ85328_SRIM_00410HSG
ZP_05006666_SSCG_04090HSG
EIF91041_STSU_17713HTG
YP_003490218_SCAB_46171HGG
ZP_07305001_SSQG_03888HTG
YP_006246301_SHJG_5160HSG
YP_007523064_BN159_4558HSG
YP_008387425_B446_19000HSG
YP_006879236_SVEN_3691HTG
YP_007860556_F750_3609HSG
YP_003490794_SCAB_52151
ELQ83369_SRIM_10671HNG
YP_008385982_B446_11745HDG
ZP_07301338_SSQG_00225XHSG
ELQ79081_SRIM_32256HSG
YP_007862798_F750_5886HSG
ELQ80643_SRIM_24376X
YP_006876003_SVEN_0457YTDHFG
YP_006881605_SVEN_6060YTDHYG
YP_004911015_SCAT_1488HFG
ZP_05004320_SSCG_01647HFG
YP_006877743_SVEN_2198HFG
EOT04000_K530_10843HFG
ELQ84558_SRIM_04583HFG
YP_007859229_F750_2251HFG
YP_007524476_BN159_5970HFG
ZP_07303466_SSQG_02353HFG
YP_006245018_SHJG_3873HFG
YP_008386122_B446_12445HFG
YP_007857553_F750_0558YSDHTG
YP_006243367_SHJG_2219YSDHTG
YP_007526668_BN159_8163YSDHTG
ZP_07301530_SSQG_00417YSDHTG
EIF94367_STSU_00895YSDHTG
YP_006875852_SVEN_0306YSDHTG
YP_004910890_SCAT_1363YSDHTG
ELQ83067_SRIM_11591YSDHTG
ZP_07303327_SSQG_02214YSDHTG
YP_007524651_BN159_6145YSDHTG
YP_008386026_B446_11965YSDHTG
YP_006244914_SHJG_3769YSDHTG
○: presence; X: absence; △: probable; other symbols are for amino acid residues.
Table 6. The numbers and types of putative PBP and protein kinase genes.
Table 6. The numbers and types of putative PBP and protein kinase genes.
BacteriaPrefixGenome Size (Mb)Class A PBPClass B PBPPBP with PASTA Domain *No of Protein KinaseProtein Kinase with PASTA Domain *
Streptomyces albulus NK660DC749.373326, 4231, 4477, 5204, 76473128, 3135, 3598, 4185, 5459none402595(4), 4186(4)
Streptomyces albus J1074XNR6.841770, 2736, 2983, 41271496, 2096, 2097, 3038, 4337, 4789none253037(4), 3064(1), 4768(4)
Streptomyces avermitilis MA-4680SAV9.123225, 4294, 4423, 4583, 5179, 72192952, 3603, 3604, 4339, 5458, 6116, 6387none334338(4), 4371(1), 6092(4)
Streptomyces bingchenggensis BCW-1SBI11.94 03076, 04174, 05361, 05810, 06697, 0906802283, 04376, 05407, 06233, 07119, 07873none6705406(4), 07851(4)
Streptomyces cattleya NRRL8057SCAT8.091929, 2889, 3140, 39060768, 1207, 1730, 1901, 3088, 4153, 5676none181232(4), 3053(1), 3089(4)
Streptomyces clavuligerus ATCC27064SCLAV8.562006, 2887, 39421087, 1301, 1774, 2276, 2947, 4154, 4179, 4180, 4198none331326(4), 2946(4), 2991(1)
Streptomyces coelicoflavus ZG0656SMCF8.481708, 75953764, 4686, 7469, 7795, 7796, 8190, 8286, 8884none226490 (4), 8300 (1),8885 (4)
Streptomyces coelicolor A3(2)SCO9.052897, 3580, 3901, 50391875, 2090, 2608, 3156, 3157, 3771, 3847, 4013, 5301none372110(4), 3821(1), 3848(4)
Streptomyces collinus Tu365B4468.3815140, 19060, 2358009755, 10960, 13820, 16355, 19320, 24955none3911080(4), 19315(4), 19450(1)
Streptomyces davawensis JCM4913BN1599.563357, 4150, 4546, 53913075, 4478, 5121, 5122, 5684, 6352, 6632none354396(1), 4479(4), 6328(4)
Streptomyces ghanaensis ATCC 14672SSFG8.5102387, 02608, 03635, 0447902394, 03587, 04216, 04217, 04765, 05266none3503552(1), 03588(4)
Streptomyces griseoflavus Tu4000SSRG8.0502182, 02879, 03203, 0396101957, 03076, 03158, 03705, 03706, 04177, 04634, 04850none3503115(1), 03159(4), 04610(4)
Streptomyces griseus subsp. griseus NBRC 13350SGR8.552494, 3341, 3679, 46472203, 3726, 4232, 4340, 4934, 5621none303725(4), 5391(4)
Streptomyces hygroscopicus subsp. jinggangensis 5008SHJG10.383853, 4373, 5171, 5432, 61363336, 4100, 4627, 4628, 5219, 6411none433594(4), 5218(4), 5252(1)
Streptomyces lividans TK24SLIV8.3513190, 18760, 20390, 2320511865, 18345, 19035, 19450, 21910, 21915, 24635, 28340none3519030(4), 19175(1), 27155(4)
Streptomyces rimosus subsp. rimosus ATCC10970SRIM9.500295, 08328, 13873, 2268900065, 04191, 06646, 15770, 26297, 31850, 31885none4000070(4), 07563(1), 24996(4)
Streptomyces scabiei 87.22SCAB10.15 33601, 41401, 56801, 6443110101, 29591, 45551, 53611, 53621, 60051, 70631none430931(1), 45201(1), 67711(4)
Streptomyces sp. PAMC26508F7507.632719, 3337, 3596, 45801743, 2434, 2998, 3546, 4834, 6320none271975(4), 3512(1), 3547(4)
Streptomyces sp. SirexAA-ESACTE7.41 2371, 3027, 3329, 42911307, 1519, 2029, 2618, 2701, 3283, 4532none321542(4), 3284(3)
Streptomyces sviceus ATCC 29083SSEG9.3101073, 07525, 03439, 0416400010, 00011, 00733, 01896, 09019, 09517, none4002705(4), 06024(4)
Streptomyces venezuelae ATCC 10712SVEN8.23 2646, 3350, 3677, 47051522, 1745, 2386, 2985, 3631, 4995none371769(4), 3592(1), 3632(4)
Streptomyces violaceusniger Tu4113STRVI11.141350, 2314, 3845, 8252, 90050275, 1135, 3190, 7171, 7897, 7904none380274(4), 7194(4)
Streptomyces viridochromogenes DSM 40736SSQG8.6502328, 02941, 03901, 04279, 0511301781, 02628, 03242, 03243, 03958, 05348none3702054(4), 03956(4), 03996(1)
Streptomyces zinciresistens K42SZN8.2206389, 16730, 18682, 2849302952, 10458, 13352, 17932, 18819, 22026none3908009(4), 17937(4), 21616(1)
* The number in parenthesis is the numbers of PASTA domains.
Table 7. Comparison of genomic structures in PBP-STPK region of three Streptomyces species.
Table 7. Comparison of genomic structures in PBP-STPK region of three Streptomyces species.
Gene No.Protein NameNo. of aaGene No.Protein NameNo. of aaGene No.Protein NameNo. of aa
SCO3844hypothetical protein172aa+ *SCLAV_2950hypothetical protein132aa−SGR_3729hypothetical protein173aa−
SCO3845protein phosphatase515aa+SCLAV_2949protein phosphatase507aa−SGR_3728protein phosphatase499aa−
SCO3846FtsW/RodA family protein479aa+SCLAV_2948FtsW/RodA family protein470aa−SGR_3727FtsW/RodA family protein466aa−
SCO3847PBP490aa+SCLAV_2947PBP484aa−SGR_3726PBP485aa−
SCO3848STPK673aa+SCLAV_2946STPK682aa−SGR_3725STPK664aa−
SCO3849hypothetical protein253aa−SCLAV_2945hypothetical protein231aa+SGR_3724hypothetical protein245aa+
SCO3850hypothetical protein352aa−SCLAV_2944hypothetical protein461aa+SGR_3723hypothetical protein474aa+
SCO3851glutamine amidotransferase212aa−SCLAV_2943glutamine amidotransferase212aa+SGR_3722glutamine amidotransferase212aa+
SCO3852hypothetical protein69aa−SCLAV_2942hypothetical protein68aa+SGR_3721hypothetical protein62aa+
*: +: plus strand, −: minus strand.

Share and Cite

MDPI and ACS Style

Ogawara, H. Self-resistance in Streptomyces, with Special Reference to β-Lactam Antibiotics. Molecules 2016, 21, 605. https://doi.org/10.3390/molecules21050605

AMA Style

Ogawara H. Self-resistance in Streptomyces, with Special Reference to β-Lactam Antibiotics. Molecules. 2016; 21(5):605. https://doi.org/10.3390/molecules21050605

Chicago/Turabian Style

Ogawara, Hiroshi. 2016. "Self-resistance in Streptomyces, with Special Reference to β-Lactam Antibiotics" Molecules 21, no. 5: 605. https://doi.org/10.3390/molecules21050605

Article Metrics

Back to TopTop