Next Article in Journal
Linking Groundwater to Surface Discharge Ecosystems: Archaeal, Bacterial, and Eukaryotic Community Diversity and Structure in Quebec (Canada)
Next Article in Special Issue
Microbial Risks Caused by Livestock Excrement: Current Research Status and Prospects
Previous Article in Journal
Seroepidemiological and Molecular Survey for the Detection of SARS-CoV-2 Infection among Children in Iran, September 2020 to June 2021: 1-Year Cross-Sectional Study
Previous Article in Special Issue
Biological Degradation of Plastics and Microplastics: A Recent Perspective on Associated Mechanisms and Influencing Factors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Phylogenetic Revisit to a Review on Predatory Bacteria

Graduate School of Integrated Sciences for Life, Hiroshima University, 1-4-4 Kagamiyama, Higashihiroshima 739-8528, Japan
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(7), 1673; https://doi.org/10.3390/microorganisms11071673
Submission received: 29 May 2023 / Revised: 22 June 2023 / Accepted: 26 June 2023 / Published: 27 June 2023
(This article belongs to the Special Issue Latest Review Papers in Environmental Microbiology 2023)

Abstract

:
Predatory bacteria, along with the biology of their predatory behavior, have attracted interest in terms of their ecological significance and industrial applications, a trend that has been even more pronounced since the comprehensive review in 2016. This mini-review does not cover research trends, such as the role of outer membrane vesicles in myxobacterial predation, but provides an overview of the classification and newly described taxa of predatory bacteria since 2016, particularly with regard to phylogenetic aspects. Among them, it is noteworthy that in 2020 there was a major phylogenetic reorganization that the taxa hosting Bdellovibrio and Myxococcus, formerly classified as Deltaproteobacteria, were proposed as the new phyla Bdellovibrionota and Myxococcota, respectively. Predatory bacteria have been reported from other phyla, especially from the candidate divisions. Predatory bacteria that prey on cyanobacteria and predatory cyanobacteria that prey on Chlorella have also been found. These are also covered in this mini-review, and trans-phylum phylogenetic trees are presented.

1. Introduction

As comprehensively reviewed by Pérez et al. (2016) [1], predatory bacteria are a group of prokaryotes that can actively hunt and consume other bacteria as their food source. By doing so, they can alter the abundance and diversity of the prey bacteria and thus influence the overall structure of the microbial community. In addition to predatory bacteria, protists and bacteriophages can also have significant impacts on the biomass, structure, and function of microbial communities, though their impacts differ in size, prey specificity, and hunting tactics [2]. Among their interwoven interactions, this mini-review, as an update of Pérez et al. (2016) [1], focuses on predatory bacteria with reference to phylogenetic aspects, particularly after the proposal in 2020 of the new phyla Bdellovibrionota and Myxococcota, which show distinct hunting strategies of predation [3].
Pérez et al. (2016) [1] reviewed the hunting strategies of predators of the order Bdellovibrionales, which physically attach to prey cells with flagella-based motility and penetrate into the periplasm of the prey cells, and the order Myxococcales, which are known for a “group attack” with gliding motility, the secretion of lytic enzymes, and the release of antibiotics. Pérez et al. (2016) [1] also reviewed the genomes, transcriptomes, and comparative genomics of predators, including the idea of the “predatome”, i.e., the protein families in phenotypes of predatory bacteria [4]. Through detailed analysis of the predation-related proteins and the encoding genes, predatory properties are predicted for the clades whose predations are not yet known in the phyla Bdellovibrionota and Myxococcota [3]. Moreover, detailed analyses on the correlation between antibiotics biosynthesis and predation indicate that myxobacteria may be prioritized for the discovery of unexplored natural products [5,6,7].
After Pérez et al. (2016) [1], ecological significances and industrial applications of predatory bacteria have been increasingly studied. For example, a study on the potential use of predatory bacteria as alternatives to antibiotics showed that intrarectal inoculations of Bdellovibrio bacteriovorus and Micavibrio aeruginosavorus lead to beneficial and adverse changes, respectively, in rat gut microflora, indicating a top-down control [8]. A large-scale field study using stable isotopes 18O and 13C demonstrated that activities of obligate predators are increased by substrates added to preys, indicating a bottom-up trophic control [9]. A high-resolution microscopic study revealed the submillimeter-scale changes in Vibrio cholerae biofilms attacked by Bdellovibrio bacteriovorus [10]. A recent review evaluates that potential uses of Bdellovibrio and like organisms (BALOs) in medical, agricultural, biotechnological, and environmental applications are achievable and should be pursued [11].
This mini-review aims to update Pérez et al. (2016) [1] based on these publications along with previous ones and to present a comprehensive, trans-phylum phylogenetic tree of predatory bacteria. A PubMed search using the simple phrase “predatory bacteria” (double quotations are needed to combine “predatory” and “bacteria” as a solid phrase) resulted in a total of 137 publications in and after 2016, with a peak of 24 hits in 2021 (accessed on 16 May 2023; https://pubmed.ncbi.nlm.nih.gov/?term=predatory+bacteria&filter=years.2016-2023). The number decreases to 21 from 137 when the search word “phylogenetics” is added. One of the 21 hits reports a genomically characterized nonpredatory nature of an isolate related to the predatory genus Herpetosiphon in the phylum Chloroflexota [12], which represents a possibility of a predatome-based search for predatory bacteria in diverse taxa.

2. Phylogenetics of Predatory Bacteria in and after 2016

2.1. Trans-Phylum Phylogenetics in and after 2016

Trans-phylum phylogenetic analysis of predatory bacteria was conducted in some of the PubMed-searched publications in and after 2016. In metagenomics of the intertidal soils along the Peruvian coastline, about 0.5% of 16S rRNA gene sequences are ascribed to predatory bacteria of multiple phyla, and 30 antibiotic-producing strains are kept in cultures with Escherichia coli XL1 Blue or Pheobacter inhibens DSM17395 as prey [13]. In the V4–V5 microbiomics of the Chesapeake Bay sediment, amplicons are affiliated with the predatory genus Haliangium and the class Polyangia (phylum Myxococcota), the genus-level clade of the phylum Bdellovibrionota, and the Bradymonadales clade of the phylum Desulfobacterota, along with the abundant amplicons affiliated with Ca. Electrothrix of the phylum Desulfobacteriota [14]. A metatranscriptomic study combined with a set of curated genomes revealed increased and decreased expressions of proteases by the members of the phyla Bacteroidota-Actinobacteriota and Myxococcota, respectively [15].

2.2. Phylogenetics on the Members of BALOs, or the Phylum Bdellovibrionota, in and after 2016

Most of the PubMed-searched publications on predatory bacterial phylogenetics deal with selected taxa such as BALOs in the phylum Bdellovibrionota, for example, with a co-occurrence microbial network, particularly with Myxococcales of the phylum Myxococcota [16]. Two major BALO clades, i.e., the Bacteriovorax and Bdellovibrio clades, are targeted to decipher their roles and interplays in the microbiomes of wastewater treatment plants [17]. Comparative genomics of ten BALO genomes and one genome of a marine Bacteoidota predator (Saprospira grandis) allowed the genome-mining of 18 putative “predatomes” [18]. Comparative genomics of 152 BALO genomes from databases and 5 metagenome-assembled genomes (MAGs) from the Mariana Trench deep-sea water suggested that the chitinase-possessing members of Oligoflexia, the third class of the phylum Bdellovibrionota, may also act as predatory bacteria [19].
Phylogenetic aspects have also been viewed from the prey side. A wide prey range of a marine, prey-generalist BALO, Halobacteriovorax marinus, is discussed in evolutionary reference to horizontal gene transfer [20], although other Halobacteriovorax isolates (no phylogenetic information available) are rather prey-specific to Vibrio parahaemolyticus, a seafood-associated pathogen, and not to other seafood-associated pathogens, such as Vibrio vulnificus, Vibrio alginolyticus, Escherichia coli O157:H7, and Salmonella enterica serovar Typhimurium DT104 [21]. Bdellovibrio bacteriovorus is also known as prey-specific to Gram-negative bacteria by burrowing through the outer membrane and peptidoglycan cell wall and entering the periplasmic space of the prey cells [22].
Prey-specificity or prey-generality may be correlated to the alpha diversity of the BALO-associated microbiomes as suggested by a correlation study based on nine microbiomes from animals and environments [23], although this correlation is currently a correlation, not strong enough to be a prediction. Among the BALOs, alpha and beta diversities of the Halobacteriovorax clades in an estuary microbiome may be driven by inputs of pollutants, such as dissolved inorganic phosphorus and NH4+-N, as implicated by an analysis of the 676F-1193R region of the 16S rRNA gene sequence [24], though they are probably indirectly driven via bottom-up trophic control [9].
Detailed genotypic and phenotypic comparisons have also been performed for a soil isolate Bdellovibrio sp. NC01 and the type strain Bdellovibrio bacteriovorus HD100, indicating that reduced predatory activities of the isolate NC01 are probably due to the absence of ten genes in its genome [25].

2.3. Phylogenetics on the Members of the Predatory Phylum Myxococcota in and after 2016

A microbiomic study at a wastewater treatment plant using 13C-labeled Escherichia coli ESS5 and Pseudomonas putida ESE1 as preys discovered the dominance of Haliangium and the mle1–27 clade of the phylum Myxococcota, contrary to the conventional Bdellovibrionota-dominant view, which is further confirmed by analyzing the global datasets [26] from the Global Water Microbiome Consortium [27]. A large-scale study on agricultural soil microbiomes revealed that, out of all 6151 bacterial operational taxonomic units (OTUs), 242 are the myxobacterial OTUs affiliated with the phylum Myxococcota and that the alpha and beta diversities of the myxobacterial communities are more sensitive to geography (location and climate) than fertilization [28].

3. Description of New Taxa and Characterization of New Strains

3.1. Candidate Divisons

Co-cultured with the methanogenic archaeon, Methanosaeta, which is probably the globally prevalent methane producer [29], an epibiotic bacterial strain OP3 LiM that preys on the archaeon Methanosaeta has been characterized by genomics and proteomics as well as fluorescence and electron microscopy. The strain OP3 represents the candidate genus and a species, “Ca. Velamenicoccus archaeovorus” gen. nov., sp. nov., affiliated with the phylum “Ca. Omnitrophica” (candidate division OP3) [30].
Two new predatory species from formerly candidate divisions have also been reported. One is the bacterial strain TM7x HMT-952, affiliated with the formerly candidate division TM7 (currently the phylum “Ca. Saccharibacteria”). TM7x is epibiotic to Actinomyces odontolyticus subspecies actinosynbacter XH001 [31,32,33] and proposed to be designated as “Ca. Nanosynbacter lyticus” TM7x [34,35].
Another one is the ultra-small epibiont to the gammaproteobacterial photoautotroph Chromatium minus, inhabiting the karstic lakes in northeastern Spain, that was tentatively named Vampirococcus in 1986 [36]. In the same article, a Gram-negative, facultatively anaerobic epsilonproteobacterium, Daptobacter, was also mentioned but has been less studied and unvalidated. In 2013, Chromatium (Halochromatium)-like bacteria and associated ultra-small (550 nm × 220 nm) epibionts, which morphologically matched to the Vampirococcus reported in 1986, were sampled in a hypersaline lake in northeastern Spain and subject to a two-member consortium mini-metagenomic analysis, yielding a genome similar to the genomes affiliated with the superphylum “Ca. Patescibacteria” (the candidate phyla radiation, CPR) or “Ca. Absconditabacteria” (formerly SR1). Based on the putative Vampirococcus genome, “Ca. Vampirococcus lugosii” has been proposed [37].

3.2. Phyla Bdelovibrionota and Myxococcota

The 16S rRNA gene sequences of a soil bacterium strain LBG001 and the type strain Bdellovibrio bacteriovorus HD100 share a 97% similarity, which does not discriminate based on a traditional 97% threshold [38] but does discriminate based on an updated view [39]. The genomic features of LBG001, i.e., its average nucleotide identity, average amino acid identity, and digital DNA–DNA hybridization values with other Bdellovibrio members, are as low as <79%, <72% and <17%, respectively, enough to be discriminated, and thus LBG001 is described as the new species Bdellovibrio reynosensis [40], whose specific epithet indicates its place of origin, Reynosa, a Mexican city on the southern bank of the Rio Grande.
One of the five new species of the genus Myxococcus described in 2020 is also named after its place of origin, i.e., a settlement on the island of Anglesey in North Wales, UK. The settlement’s name appears in the specific epithet as Myxococcus llanfairpwllgwyngyllgogerychwyrndrobwllllantysiliogogogochensis, whose 16S rRNA gene sequence is not explicitly available from its shotgun-sequenced genome [41].
Two new predatory strains isolated from a freshwater pond, Bacteriovorax stolpii HI3 (phylum Bdelovibrionota) and Myxococcus sp. MH1 (phylum Myxococcota), show a wide range of preys, including Escherichia coli HB101 and 52 environmental strains consisting of 8 Gram-positive strains affiliated with the phyla Actinomycetota and Bacillota and 44 Gram-negative strains affiliated with the phyla Bacteroidota and Pseudomonadota. While Myxococcus sp. MH1 preys on all the 53 prey strains, Bacteriovorax stolpii HI3 preys on 25 environmental Gram-negative strains as well as E. coli HB101 [42].
The term “myxobacteria” (or “slime bacteria”) has been ambiguously defined but may be regarded as the members of the orders Myxococcales and Polyangia in the phylum Myxococcota [43]. Omics studies, including “predatomics” of myxobacteria based on 163 publicly available genomes and 24 newly added genomes, were reviewed in 2021 [44]. More extensive review on myxobacteria with >400 references has recently been published [43].

3.3. Phylum Planctomycetota

The phylum Planctomycetota, along with Verrucomicrobiota and Chlamydiota, composes the superphylum PVC, whose member is hypothesized to be an ancestor of the proto-eukaryotic cell via symbiogenesis with an archaeon [45,46]. Within the phylum Planctomycetota, a bacterium that exhibits eukaryote-like phagocytosis predation was discovered from the seawater of Palau. The bacterium is described as “Ca. Uab amorphum”, whose genus is named after a giant of Palauan mythology, indicating its size as giant (4.5–7.8 μm × 2.8–5.5 μm), and it is capable of engulfment or phagocytosis [47].

3.4. Order Bradymonadales in the Phylum Desulfobacteriota

In addition to the two major taxa of predatory bacteria, i.e., the phyla Bdelovibrionota and Myxococcota, the order Bradymonadales may represent a third type of predation. The order Bradymonadales was formerly ascribed to the class Deltaproteobacteria of the phylum Proteobacteria but is currently affiliated with the phylum Desulfobacteriota [3]. The bacteria of this order, or Bradymonabacteria, show a unique predation type that is different from the so-called “obligate” and “facultative” as defined in Pérez et al. (2016), [1] which are characterized as “completely prey-dependent” and “prey-independent” [48]. The predation type of Bradymonabacteria is characterized as “facultatively prey-dependent” and termed lately as “facultative” [48], and the previous term “facultative” from Pérez et al. (2016) [1] was newly replaced with “opportunistic”.
Within the order Bradymonadales, a new genus and new species has been proposed for the wide-ranging predatory strain YN101 from the sediment of a marine solar saltern, which is described as Persicimonas caeni gen. nov., sp. nov. in the family Bradymonadaceae [49].
Another new genus has also been proposed during the descriptions of two new species, Lujinxingia litoralis gen. nov. sp. nov. and Lujinxingia sediminis sp. nov., in the family Bradymonadaceae [50]. The genus name Lujinxingia is named after a Chinese microbiologist, Jin-Xing Lu, or Lu Jin-Xian in the Chinese name order.
A detailed comparison between Lujinxingia sediminis and a related strain, V1718, from a Chinese island, has led to the proposal of two new families, in addition to the family Bradymonadaceae, within the order Bradymonadales. During the description of the strain V1718 as Microvenator marinus gen. nov., sp. nov., two new families, Microvenatoraceae fam. nov. and Lujinxingiaceae fam. nov., were proposed. The genus Lujinxingia has been reclassified to the latter [51].

3.5. Order Herpetosiphonales in the Phylum Chloroflexota

Predatory species of the phylum Chloroflexota are known in the genus Herpetosiphon established in 1968 during the description of the filamentous gliding bacterium, Herpetosiphon aurantiacus gen. et sp. n. (=gen. nov., sp. nov.) [52]. One of the original strains used for the description was isolated from the slimy coating of Chara sp., a charophyte green alga, growing in Birch Lake, Minnesota, which implies predation on the slime-forming bacteria. The genus Herpetosiphon is placed in the family Herpetosiphonaceae within the order Herpetosiphonales, whose detailed descriptions were given by Gupta et al. (2013) [53].
An environmental strain, Hp g472 (DSM 52871), was isolated from a sandy soil in the beach of Poel island, Germany, and described as Herpetosiphon gulosus sp. nov., whose specific epithet derives from its gluttony, i.e., strong predatory activity [54].
Another environmental strain, CA052B, represents the fifth species of the genus Herpetosiphonaceae, described as Herpetosiphon llansteffanense sp. nov., whose specific epithet derives from the name of the origin, i.e., soil from the edge of a stream near the village of Llansteffan in Wales, UK [55]. This species demonstrates efficient predatory activity against a diverse array of prey microbial species using a “wolf pack” strategy that is possibly mediated by the secreted outer membrane vesicles containing a variety of hydrolytic enzymes [56].

3.6. Family Chitinophagaceae in the Phylum Bacteroidota

Biological soil crusts, or biocrusts, are regarded as promising mitigants for arid and semiarid lands. Filamentous cyanobacteria such as Microcoleus spp. play major roles in biocrust formation by stabilizing soil particles [57] and performing nitrogen fixation as well as carbon fixation via photosynthesis [58]. Biocrusts are applied to the restoration of damaged or disturbed lands [59]. The inoculation of “nursery grown” biocrusts have been attempted; however, inoculated biocrusts are often devastated by pathogen-like agents, leaving “plaques” [60]. From the diseased biocrusts, a cyanobacteria predator was enriched with Microcoleus vaginatus (PCC 9802), characterized by microscopy and physical–biochemical tests, genome-sequenced, and described as a new species, Ca. Cyanoraptor togatus. sp. nova., in the family Chitinophagaceae within the phylum Bacteroidota [61]. Bethany et al. (2022) [61] estimate that predation by Cyanoraptor reduces photosynthetic production of cyanobacteria by >10%.

3.7. Class Melainabacteria in the Phylum Cyanobacteria

Cyanobacteria can be predators, although they are preyed by predatory bacteria, such as Ca. Cyanoraptor togatus. However, the preys are not bacteria but eukaryotic microalgae such as Chlorella. The known predatory cyanobacterium is Vampirovibrio chlorellavorus [62] in the class Melainabacteria [63]. The members of the class Melanibacteria are affiliated with the phylum Cyanobacteria and characterized by nonphotosynthetic metabolisms and dark habitats, such as the human gut and groundwater [64]. Affiliation of the predatory Vampirovibrio chlorellavorus with the class Melanibacteria agrees with the nonphotosynthetic nature of this taxon. The genome of Vampirovibrio chlorellavorus was reconstructed in 2009 from the old co-culture with Chlorella vulgaris deposited in 1978 and dealt with in an academic article in 2015 [65]. There have been established strategies to defend commercially produced Chlorella from predation by Vampirovibrio chlorellavorus [66].

4. Phylogenetic Tree of Predatory Bacteria

A total of 136 sequences of predatory bacterial 16S rRNA genes were collected from 12 phyla (including candidate phyla) of Actinobacteriota, Bacteroidota, Bdellovibrionota, Chloroflexota, Cyanobacteria, Desulfobacteriota, Myxococcota, Ca. Omnitrophica (OP3), Ca. Patescibacteria (CPR) or Ca. Absconditabacteria (SR1), Planctomycetota, Pseudomonadota, and Ca. Saccharibacteria (TM7). The available 16S rRNA gene sequence of “Ca. Vampirococcus lugosii” (accession number MW286273, 1071 bp) [37] was the shortest among the collected sequences, and the phylogenetic trees with and without “Ca. V. lugosii” were constructed, along with the 35 reference sequences from current and former bacterial phyla. The sequences were aligned online with MEGA11 (https://www.megasoftware.net/; accessed on 20 May 2023) [67], and the phylogenetic trees based on the maximum likelihood method were drawn online with iTOL v6 (https://itol.embl.de/; accessed on 20 May 2023) [68] (Figure 1 and Figure S1). Figure 1 shows the phylogenetic tree based on the sequences of about 1.6 kb after alignment excluding the shortest 1071 bp sequence of “Ca. V. lugosii”. The 1.6 kb length, instead of the generally cited 1.5 kb, resulted from the alignment of 135 (136 minus 1, Ca. V. lugosii) full-length and near-full-length sequences that contain “gaps”.
Figure S1 displays the tree based on sequences of about 0.6 kb including “Ca. V. lugosii”. Information about the used 16S rRNA sequences of predatory bacteria [4,18,20,31,37,42,47,49,50,61,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120] are listed in Table S1 along with the hunting strategies of the corresponding predators. The reference 16S rRNA gene sequences from 35 representative, current, and former bacterial phyla [121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154] are listed in Table S2.
The tree based on “1.6 kb-long” 16S rRNA gene sequences (Figure 1) shows (1) rather nonstreamlined phylogeny and predation strategies in the Bdellovibrionota (orange) and Pseudomonadota (pale pink) cluster from about 3:15 to 5:25 when Figure 1 is seen as the disk display of a 12 h clock and (2) streamlined phylogeny and predation strategies in the Myxococcata cluster (red), except the betaproteobacterial “AF005994 Aristabacter necator” [113] at 10:30. Interestingly, “AF005994 Aristabacter necator” was very deep-branched at about 10 o’clock on the “0.6 kb-long 16S-tree”, as shown in Figure S1.
Figure S1 (0.6 kb long 16S tree) includes the shortest sequence of “MW286273 Vampirococcus lugosii” [37] at about 8:30, neighbored with “OM390184 Nanosynbacter lyticus” [31]. Different from Figure 1, Figure S1 placed the nonpredatory “NR_042149 Fibrobacter succinogenes” [140] clustered with “AB540021 Oligoflexus tunisiensis” [128], which is predicted to be predatory [3] but as yet unconfirmed (Nakai, pers. comm.). These inconsistencies may be a hint for hunting novel predatory bacteria.
Another notable irregularity is “CP075895 Chitinophagaceae bacterium” at 9:10 in Figure 1 and 9:05 in Figure S1. This bacterium, Ca. Cyanoraptor togatus LGM1 [61], is the only known obligatory predator within the phylum Bacteroidota and is the only known endobiotic invader outside the Bdellovibrionota (orange)–Pseudomonadota (pale pink) cluster from a little before 3:00 to about 5:25. It is also the first obligatory, intracellular predator of cyanobacteria.
Only some of the predatory species of the phyla Bdelovibrionota and Myxococcota are shown in Figure 1 and Figure S1, which would have exhibited more significant proportions of the phyla if all the predatory species were included. However, the importance of the phyla in the phylogeny of predatory bacteria is already explicit in the current Figure 1 and Figure S1 with only selected species.
Figure 1. Trans-phylum phylogenetic tree of 135 sequences of predatory bacterial 16S rRNA genes listed in Table S1 [4,18,20,31,37,42,47,49,50,61,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120], except [37] and 35 sequences from representative, current, and former bacterial phyla listed in Table S2 [121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154]. Predation properties are indicated by the symbols as follows: , obligate, epibiotic; , obligate, endobiotic or direct invasion; , opportunistic, epibiotic; , opportunistic, group attack; , facultative, epibiotic; and, ■, facultative, unknown.
Figure 1. Trans-phylum phylogenetic tree of 135 sequences of predatory bacterial 16S rRNA genes listed in Table S1 [4,18,20,31,37,42,47,49,50,61,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120], except [37] and 35 sequences from representative, current, and former bacterial phyla listed in Table S2 [121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154]. Predation properties are indicated by the symbols as follows: , obligate, epibiotic; , obligate, endobiotic or direct invasion; , opportunistic, epibiotic; , opportunistic, group attack; , facultative, epibiotic; and, ■, facultative, unknown.
Microorganisms 11 01673 g001aMicroorganisms 11 01673 g001b
A myxobacterial species, Sorangium cellulosum, is not a typical epibiotic or endobiotic predator but a cellulolytic consumer as suggested by its genome [109], and its strong “lytic” activity may be regarded as predatory [155]. No such “lytic predators” are included in Figure 1 and Figure S1.

5. Future Perspective

According to the mitochondrial endosymbiotic theory, the origin of mitochondria is regarded as the endosymbiosis of an ancestral alphaproteobacterial linage [156]. A popular idea about the mechanisms of mitochondria acquisition is the engulfment or phagocytosis of proto-mitochondria by proto-eukaryotes [157]. Asgard archaea are regarded as a paraphyletic lineage of proto-eukaryotes [158]. However, Asgard archaea are genomically predicted to be nonphagocytotic [159], although a cultured Asgard archaeon “Ca. Prometheoarchaeum syntrophicum” is hypothesized to entangle–engulf–endogenize aerobic bacteria as metabolic partners like proto-mitochondria [160]. In any case, the endogenized proto-mitochondria have to stay “undigested”, and a new idea arose from the study of predatory bacteria. That is, the proto-mitochondrion was not undigested food but an “attenuated predator” [79,161,162]. More detailed studies on the known alphaproteobacterial predators, such as Ensifer adhaerens [111,112,163], as well as the finding of novel alphaproteobacterial predators, will shed more light on the relationship between the origins of mitochondria and predatory bacteria.
Cell sizes of predatory bacteria would be another perspective. Predatory bacteria of the candidate phylum radiation (CPR) represent ultra-small predators. Flat and stacked cells of 500–600 nm diameters and 200–250 nm heights are observed with the epibiotic predator or “sucker”, Ca. Vampirococcus lugosii (phylum Ca. Absconditabacteria or candidate division SR1), of photosynthetic anoxic bacteria [37]. Coccoid cells of 200–300 nm diameters are seen for the facultative epibiotic predator Ca. Velamenicoccus archeaovorus (phylum Ca. Omnitrophica or OP3) of Bacteria and Archaea [30,110]. Coccoid cells of 200–300 nm diameters along with rod cells are found for the epibiont predator Ca. Nanosynbacter lyticus (phylum Ca. Saccharibacteria or TM7) of human oral Actinomyces odontolyticus [31,32]. A recent transcriptomic study suggests that the N. lyticusA. odontolyticus relationship is symbiotic rather than predatory [164]. The occurrences of these small predators (symbionts, parasites, or pathogens) support the idea that the ultra-small cell sizes of the CPR bacteria are associated with the small genome sizes, leading to dependence on other prokaryotes with fully (or more) functional genomes and larger cells [165].
Cells of Bdelovibrio, the obligate invading/attenuating predators, are also as small as 0.2–0.5 μm × 0.5–2.5 μm (200–300 nm × 500–2500 nm), larger than the ultra-small CPR bacteria but significantly smaller than the prey cells [22]. In contrast to these small-sized predators, the flat, round, or oval cells of the engulfing bacterium Ca. Uab amorphum (phylum Planctomycetota) are as large as ~4–5 μm in diameter, reaching 10 μm in diameter after the engulfment of bacterial preys [47]. Spherical to ovoid cells of Gemmata obscuriglobus, the not-bacteria-but-protein engulfing bacterium [166] within the phylum Planctomycetota, are in the moderate size range of 1.4–3.0 μm × 0.5–3.0 μm [167]. The type species of the phylum Planctomycetota is Planctomyces bekefii, which represents a rare example of as-yet-uncultivated bacteria with validly published names but no known identity [168]. The isolation and cultivation of more Planctomycetota species may assist in elucidating the origin and function of engulfment (endocytosis, phagocytosis) for predation [169,170].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/microorganisms11071673/s1, Figure S1: Trans-phylum phylogenetic tree based on 136 (about 0.6 kb) sequences of predatory bacterial 16S rRNA genes listed in Table S1 and 35 sequences from representative, current, and former bacterial phyla listed in Table S2; Table S1: List of 136 sources of 16S rRNA gene sequences and hunting strategies of predatory bacteria; Table S2: List of 35 sources of reference 16S rRNA gene sequences from representative, current, and former bacterial phyla; Data S1: FASTA of a total of 170 (about 1.6 kb) 16S rRNA gene sequences of predatory bacteria for Figure 1. Data S2: FASTA of 171 (about 0.6 kb) reference 16S rRNA gene sequences for Figure S1; Data S3: 170 (about 1.6 kb, aligned) 16S rRNA gene sequences for Figure 1; Data S4: 171 (about 0.6 kb, aligned) 16S rRNA gene sequences for Figure S1; Data S5: 170 (about 1.6 kb, trimmed) 16S rRNA gene sequences for Figure 1; Data S6: 171 (about 0.6 kb, trimmed) 16S rRNA gene sequences for Figure S1; Data S7: 170 (about 1.6 kb, for iTOL) 16S rRNA gene sequences for Figure 1; Data S8: 171 (about 0.6 kb, for iTOL) 16S rRNA gene sequences for Figure S1.

Author Contributions

Conceptualization and writing, T.N.; visualization, S.K. and R.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All the source information of the used DNA sequences are available in Tables S1 and S2. The processed DNA sequences for the construction of the phylogenetic trees (Figure 1 and Figure S1) are available in Supplementary Data S1 to S8.

Acknowledgments

Ryosuke Nakai at National Institute of Advanced Industrial Science and Technology, Japan, kindly reviewed and edited the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pérez, J.; Moraleda-Munoz, A.; Marcos-Torres, F.J.; Muñoz-Dorado, J. Bacterial predation: 75 years and counting! Environ. Microbiol. 2016, 18, 766–779. [Google Scholar] [CrossRef] [PubMed]
  2. Johnke, J.; Cohen, Y.; de Leeuw, M.; Kushmaro, A.; Jurkevitch, E.; Chatzinotas, A. Multiple micro-predators controlling bacterial communities in the environment. Curr. Opin. Biotechnol. 2014, 27, 185–190. [Google Scholar] [CrossRef] [PubMed]
  3. Waite, D.W.; Chuvochina, M.; Pelikan, C.; Parks, D.H.; Yilmaz, P.; Wagner, M.; Loy, A.; Naganuma, T.; Nakai, R.; Whitman, W.B.; et al. Proposal to reclassify the proteobacterial classes Deltaproteobacteria and Oligoflexia, and the phylum Thermodesulfobacteria into four phyla reflecting major functional capabilities. Int. J. Syst. Evol. Microbiol. 2020, 70, 5972–6016. [Google Scholar] [CrossRef] [PubMed]
  4. Pasternak, Z.; Pietrokovski, S.; Rotem, O.; Gophna, U.; Lurie-Weinberger, M.N.; Jurkevitch, E. By their genes ye shall know them: Genomic signatures of predatory bacteria. ISME J. 2012, 7, 756–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Korp, J.; Gurovic, M.S.V.; Nett, M. Antibiotics from predatory bacteria. Beilstein J. Org. Chem. 2016, 12, 594–607. [Google Scholar] [CrossRef] [Green Version]
  6. Gregory, K.; Salvador, L.A.; Akbar, S.; Adaikpoh, B.I.; Stevens, D.C. Survey of Biosynthetic Gene Clusters from Sequenced Myxobacteria Reveals Unexplored Biosynthetic Potential. Microorganisms 2019, 7, 181. [Google Scholar] [CrossRef] [Green Version]
  7. Ahearne, A.; Albataineh, H.; Dowd, S.E.; Stevens, D.C. Assessment of Evolutionary Relationships for Prioritization of Myxobacteria for Natural Product Discovery. Microorganisms 2021, 9, 1376. [Google Scholar] [CrossRef]
  8. Shatzkes, K.; Tang, C.; Singleton, E.; Shukla, S.; Zuena, M.; Gupta, S.; Dharani, S.; Rinaggio, J.; Connell, N.D.; Kadouri, D.E. Effect of predatory bacteria on the gut bacterial microbiota in rats. Sci. Rep. 2017, 7, srep43483. [Google Scholar] [CrossRef] [Green Version]
  9. Hungate, B.A.; Marks, J.C.; Power, M.E.; Schwartz, E.; van Groenigen, K.J.; Blazewicz, S.J.; Chuckran, P.; Dijkstra, P.; Finley, B.K.; Firestone, M.K.; et al. The Functional Significance of Bacterial Predators. mBio 2021, 12, e00466-21. [Google Scholar] [CrossRef]
  10. Wucher, B.R.; Elsayed, M.; Adelman, J.S.; Kadouri, D.E.; Nadell, C.D. Bacterial predation transforms the landscape and community assembly of biofilms. Curr. Biol. 2021, 31, 2643–2651.e3. [Google Scholar] [CrossRef]
  11. Mookherjee, A.; Jurkevitch, E. Interactions between Bdellovibrio and like organisms and bacteria in biofilms: Beyond predator–prey dynamics. Environ. Microbiol. 2021, 24, 998–1011. [Google Scholar] [CrossRef]
  12. Ward, L.M.; Fischer, W.W.; McGlynn, S.E. Candidatus Anthektikosiphon siderophilum OHK22, a New Member of the Chloroflexi Family Herpetosiphonaceae from Oku-okuhachikurou Onsen. Microbes Environ. 2020, 35, ME20030. [Google Scholar] [CrossRef]
  13. Linares-Otoya, L.; Linares-Otoya, V.; Armas-Mantilla, L.; Blanco-Olano, C.; Crüsemann, M.; Ganoza-Yupanqui, M.L.; Campos-Florian, J.; König, G.M.; Schäberle, T.F. Diversity and Antimicrobial Potential of Predatory Bacteria from the Peruvian Coastline. Mar. Drugs 2017, 15, 308. [Google Scholar] [CrossRef] [Green Version]
  14. Liau, P.; Kim, C.; Saxton, M.A.; Malkin, S.Y. Microbial succession in a marine sediment: Inferring interspecific microbial interactions with marine cable bacteria. Environ. Microbiol. 2022, 24, 6348–6364. [Google Scholar] [CrossRef]
  15. Sieradzki, E.T.; Nuccio, E.E.; Pett-Ridge, J.; Firestone, M.K. Expression of macromolecular organic nitrogen degrading enzymes identifies potential mediators of soil organic N availability to an annual grass. ISME J. 2023, 17, 967–975. [Google Scholar] [CrossRef]
  16. Ezzedine, J.A.; Jacas, L.; Desdevises, Y.; Jacquet, S. Bdellovibrio and like organisms in Lake Geneva: An unseen elephant in the room? Front. Microbiol. 2020, 11, 98. [Google Scholar] [CrossRef]
  17. Cohen, Y.; Pasternak, Z.; Müller, S.; Hübschmann, T.; Schattenberg, F.; Sivakala, K.K.; Abed-Rabbo, A.; Chatzinotas, A.; Jurkevitch, E. Community and single cell analyses reveal complex predatory interactions between bacteria in high diversity systems. Nat. Commun. 2021, 12, 1–13. [Google Scholar] [CrossRef]
  18. Li, N.; Wang, K.; Williams, H.N.; Sun, J.; Ding, C.; Leng, X.; Dong, K. Analysis of gene gain and loss in the evolution of predatory bacteria. Gene 2017, 598, 63–70. [Google Scholar] [CrossRef]
  19. Li, Q.-M.; Zhou, Y.-L.; Wei, Z.-F.; Wang, Y. Phylogenomic Insights into Distribution and Adaptation of Bdellovibrionota in Marine Waters. Microorganisms 2021, 9, 757. [Google Scholar] [CrossRef]
  20. Enos, B.G.; Anthony, M.K.; DeGiorgis, J.A.; Williams, L.E. Prey Range and Genome Evolution of Halobacteriovorax marinus Predatory Bacteria from an Estuary. Msphere 2018, 3, e00508-17. [Google Scholar] [CrossRef] [Green Version]
  21. Richards, G.P.; Fay, J.P.; Uknalis, J.; Olanya, O.M.; Watson, M.A. Purification and Host Specificity of Predatory Halobacteriovorax Isolates from Seawater. Appl. Environ. Microbiol. 2016, 82, 922–927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Lovering, A.L.; Sockett, R.E. Microbe Profile: Bdellovibrio bacteriovorus: A specialized bacterial predator of bacteria. Microbiology 2021, 167, 001043. [Google Scholar] [CrossRef] [PubMed]
  23. Johnke, J.; Fraune, S.; Bosch, T.C.G.; Hentschel, U.; Schulenburg, H. Bdellovibrio and Like Organisms Are Predictors of Microbiome Diversity in Distinct Host Groups. Microb. Ecol. 2020, 79, 252–257. [Google Scholar] [CrossRef] [PubMed]
  24. Zhao, H.; Yang, S.; Qin, X.; Huang, J.; Huang, H.; Li, W.; Jiang, G.; Tang, J.; Dong, K.; Li, N. Disentangling the Ecological Processes and Driving Forces Shaping the Seasonal Pattern of Halobacteriovorax Communities in a Subtropical Estuary. Microb. Ecol. 2023, 1–12. [Google Scholar] [CrossRef]
  25. Williams, L.E.; Cullen, N.; DeGiorgis, J.A.; Martinez, K.J.; Mellone, J.; Oser, M.; Wang, J.; Zhang, Y. Variation in genome content and predatory phenotypes between Bdellovibrio sp. NC01 isolated from soil and B. bacteriovorus type strain HD100. Microbiology 2019, 165, 1315–1330. [Google Scholar] [CrossRef] [Green Version]
  26. Zhang, L.; Huang, X.; Zhou, J.; Ju, F. Active predation, phylogenetic diversity, and global prevalence of myxobacteria in wastewater treatment plants. ISME J. 2023, 17, 671–681. [Google Scholar] [CrossRef]
  27. Wu, L.; Ning, D.; Zhang, B.; Li, Y.; Zhang, P.; Shan, X.; Zhang, Q.; Brown, M.R.; Li, Z.; Van Nostrand, J.D.; et al. Global diversity and biogeography of bacterial communities in wastewater treatment plants. Nat. Microbiol. 2019, 4, 1183–1195. [Google Scholar] [CrossRef] [Green Version]
  28. Dai, W.; Liu, Y.; Yao, D.; Wang, N.; Ye, X.; Cui, Z.; Wang, H. Phylogenetic diversity of stochasticity-dominated predatory myxobacterial community drives multi-nutrient cycling in typical farmland soils. Sci. Total. Environ. 2023, 871, 161680. [Google Scholar] [CrossRef]
  29. Smith, K.S.; Ingram-Smith, C. Methanosaeta, the forgotten methanogen? Trends Microbiol. 2007, 15, 150–155. [Google Scholar] [CrossRef]
  30. Kizina, J.; Jordan, S.F.; Martens, G.A.; Lonsing, A.; Probian, C.; Kolovou, A.; Santarella-Mellwig, R.; Rhiel, E.; Littmann, S.; Markert, S.; et al. Methanosaeta and “Candidatus Velamenicoccus archaeovorus”. Appl. Environ. Microbiol. 2022, 88, e0240721. [Google Scholar] [CrossRef]
  31. He, X.; McLean, J.S.; Edlund, A.; Yooseph, S.; Hall, A.P.; Liu, S.-Y.; Dorrestein, P.C.; Esquenazi, E.; Hunter, R.C.; Cheng, G.; et al. Cultivation of a human-associated TM7 phylotype reveals a reduced genome and epibiotic parasitic lifestyle. Proc. Natl. Acad. Sci. USA 2014, 112, 244–249. [Google Scholar] [CrossRef] [Green Version]
  32. Bor, B.; Poweleit, N.; Bois, J.S.; Cen, L.; Bedree, J.K.; Zhou, Z.H.; Gunsalus, R.P.; Lux, R.; McLean, J.S.; He, X.; et al. Phenotypic and Physiological Characterization of the Epibiotic Interaction Between TM7x and Its Basibiont Actinomyces. Microb. Ecol. 2016, 71, 243–255. [Google Scholar] [CrossRef] [Green Version]
  33. McLean, J.S.; Liu, Q.; Bor, B.; Bedree, J.K.; Cen, L.; Watling, M.; To, T.T.; Bumgarner, R.E.; He, X.; Shi, W. Draft Genome Sequence of Actinomyces odontolyticus subsp. actinosynbacter Strain XH001, the Basibiont of an Oral TM7 Epibiont. Genome Announc. 2016, 4, e01685-15. [Google Scholar] [CrossRef] [Green Version]
  34. McLean, J.S.; Bor, B.; Kerns, K.A.; Liu, Q.; To, T.T.; Solden, L.; Hendrickson, E.L.; Wrighton, K.; Shi, W.; He, X. Acquisition and Adaptation of Ultra-small Parasitic Reduced Genome Bacteria to Mammalian Hosts. Cell Rep. 2018, 32, 107939. [Google Scholar] [CrossRef]
  35. Bedree, J.K.; Bor, B.; Cen, L.; Edlund, A.; Lux, R.; McLean, J.S.; Shi, W.; He, X. Quorum Sensing Modulates the Epibiotic-Parasitic Relationship Between Actinomyces odontolyticus and Its Saccharibacteria epibiont, a Nanosynbacter lyticus Strain, TM7x. Front. Microbiol. 2018, 9, 2049. [Google Scholar] [CrossRef] [Green Version]
  36. Guerrero, R.; Pedros-Alio, C.; Esteve, I.; Mas, J.; Chase, D.; Margulis, L. Predatory prokaryotes: Predation and primary consumption evolved in bacteria. Proc. Natl. Acad. Sci. USA 1986, 83, 2138–2142. [Google Scholar] [CrossRef] [Green Version]
  37. Moreira, D.; Zivanovic, Y.; López-Archilla, A.I.; Iniesto, M.; López-García, P. Reductive evolution and unique predatory mode in the CPR bacterium Vampirococcus lugosii. Nat. Commun. 2021, 12, 1–11. [Google Scholar] [CrossRef]
  38. Stackebrandt, E.; Goebel, B.M. Taxonomic Note: A Place for DNA-DNA Reassociation and 16S rRNA Sequence Analysis in the Present Species Definition in Bacteriology. Int. J. Syst. Evol. Microbiol. 1994, 44, 846–849. [Google Scholar] [CrossRef] [Green Version]
  39. Edgar, R.C. Updating the 97% identity threshold for 16S ribosomal RNA OTUs. Bioinformatics 2018, 34, 2371–2375. [Google Scholar] [CrossRef] [Green Version]
  40. Ajao, Y.O.; Rodríguez-Luna, I.C.; Elufisan, T.O.; Sánchez-Varela, A.; Cortés-Espinosa, D.V.; Camilli, A.; Guo, X. Bdellovibrio reynosensis sp. nov., from a Mexico soil sample. Int. J. Syst. Evol. Microbiol. 2022, 72, 005608. [Google Scholar] [CrossRef]
  41. Chambers, J.; Sparks, N.; Sydney, N.; Livingstone, P.G.; Cookson, A.R.; Whitworth, D.E. Comparative genomics and pan-genomics of the Myxococcaceae, including a description of five novel species: Myxococcus eversor sp. nov., Myxococcus llanfairpwllgwyngyllgogerychwyrndrobwllllantysiliogogogochensis sp. nov., Myxococcus vastator sp. nov., Pyxidicoccus caerfyrddinensis sp. nov. and Pyxidicoccus trucidator sp. nov". Genome Biol. Evol. 2020, 12, 2289–2302. [Google Scholar] [CrossRef] [PubMed]
  42. Inoue, D.; Hiroshima, N.; Nakamura, S.; Ishizawa, H.; Ike, M. Characterization of Two Novel Predatory Bacteria, Bacteriovorax stolpii HI3 and Myxococcus sp. MH1, Isolated from a Freshwater Pond: Prey Range, and Predatory Dynamics and Efficiency. Microorganisms 2022, 10, 1816. [Google Scholar] [CrossRef] [PubMed]
  43. Saggu, S.K.; Nath, A.; Kumar, S. Myxobacteria: Biology and bioactive secondary metabolites. Res. Microbiol. 2023; in press. [Google Scholar] [CrossRef] [PubMed]
  44. Whitworth, D.E.; Sydney, N.; Radford, E.J. Myxobacterial Genomics and Post-Genomics: A Review of Genome Biology, Genome Sequences and Related ‘Omics Studies. Microorganisms 2021, 9, 2143. [Google Scholar] [CrossRef] [PubMed]
  45. Forterre, P. A new fusion hypothesis for the origin of Eukarya: Better than previous ones, but probably also wrong. Res. Microbiol. 2011, 162, 77–91. [Google Scholar] [CrossRef]
  46. Baum, D.A.; Baum, B. An inside-out origin for the eukaryotic cell. BMC Biol. 2014, 12, 76. [Google Scholar] [CrossRef] [Green Version]
  47. Shiratori, T.; Suzuki, S.; Kakizawa, Y.; Ishida, K.-I. Phagocytosis-like cell engulfment by a planctomycete bacterium. Nat. Commun. 2019, 10, 5529. [Google Scholar] [CrossRef] [Green Version]
  48. Mu, D.-S.; Wang, S.; Liang, Q.-Y.; Du, Z.-Z.; Tian, R.; Ouyang, Y.; Wang, X.-P.; Zhou, A.; Gong, Y.; Chen, G.-J.; et al. Bradymonabacteria, a novel bacterial predator group with versatile survival strategies in saline environments. Microbiome 2020, 8, 1–15. [Google Scholar] [CrossRef]
  49. Wang, S.; Mu, D.; Du, Z.-J. Persicimonas caeni gen. nov., sp. nov., the Representative of a Novel Wide-Ranging Predatory Taxon in Bradymonadales. Front. Microbiol. 2020, 11, 698. [Google Scholar] [CrossRef] [Green Version]
  50. Guo, L.-Y.; Li, C.-M.; Wang, S.; Mu, D.-S.; Du, Z.-J. Lujinxingia litoralis gen. nov., sp. nov. and Lujinxingia sediminis sp. nov., two new representatives in the order Bradymonadales. Int. J. Syst. Evol. Microbiol. 2019, 69, 2767–2774. [Google Scholar] [CrossRef]
  51. Wang, S.; Chen, G.-J.; Du, Z.-J. Microvenator marinus gen. nov., sp. nov., isolated from marine sediment, and description of Microvenatoraceae fam. nov. and Lujinxingiaceae fam. nov. Int. J. Syst. Evol. Microbiol. 2022, 72, 005380. [Google Scholar] [CrossRef]
  52. Holt, J.G.; Lewin, R.A. Herpetosiphon aurantiacus gen. et sp. n., a new filamentous gliding organism. J. Bacteriol. 1968, 95, 2407–2408. [Google Scholar] [CrossRef] [Green Version]
  53. Gupta, R.S.; Chander, P.; George, S. Phylogenetic framework and molecular signatures for the class Chloroflexi and its different clades; proposal for division of the class Chloroflexi class. nov. into the suborder Chloroflexineae subord. nov., consisting of the emended family Oscillochloridaceae and the family Chloroflexaceae fam. nov., and the suborder Roseiflexineae subord. nov., containing the family Roseiflexaceae fam. nov. Antonie Van Leeuwenhoek 2012, 103, 99–119. [Google Scholar] [CrossRef]
  54. Pan, X.; Kage, H.; Martin, K.; Nett, M. Herpetosiphon gulosus sp. nov., a filamentous predatory bacterium isolated from sandy soil and Herpetosiphon giganteus sp. nov., nom. rev. Int. J. Syst. Evol. Microbiol. 2017, 67, 2476–2481. [Google Scholar] [CrossRef]
  55. Livingstone, P.G.; Morphew, R.M.; Cookson, A.R.; Whitworth, D.E. Genome Analysis, Metabolic Potential, and Predatory Capabilities of Herpetosiphon llansteffanense sp. nov. Appl. Environ. Microbiol. 2018, 84, e01040-18. [Google Scholar] [CrossRef] [Green Version]
  56. Jurkevitch, E. Predatory Behaviors in Bacteria—Diversity and Transitions. Microbe Mag. 2007, 2, 67–73. [Google Scholar] [CrossRef]
  57. Garcia-Pichel, F.; Wojciechowski, M.F. The Evolution of a Capacity to Build Supra-Cellular Ropes Enabled Filamentous Cyanobacteria to Colonize Highly Erodible Substrates. PLoS ONE 2009, 4, e7801. [Google Scholar] [CrossRef]
  58. Elbert, W.; Weber, B.; Burrows, S.; Steinkamp, J.; Büdel, B.; Andreae, M.O.; Pöschl, U. Contribution of cryptogamic covers to the global cycles of carbon and nitrogen. Nat. Geosci. 2012, 5, 459–462. [Google Scholar] [CrossRef]
  59. St Clair, L.L.; Johansen, J.R.; Webb, B.L. Rapid stabilization of fire-disturbed sites using a soil crust slurry: Inoculation studies. Reclam. Reveg. Res. 1986, 4, 261–269. [Google Scholar]
  60. Bethany, J.; Giraldo-Silva, A.; Nelson, C.; Barger, N.N.; Garcia-Pichel, F. Optimizing the Production of Nursery-Based Biological Soil Crusts for Restoration of Arid Land Soils. Appl. Environ. Microbiol. 2019, 85, e00735-19. [Google Scholar] [CrossRef] [Green Version]
  61. Bethany, J.; Johnson, S.L.; Garcia-Pichel, F. High impact of bacterial predation on cyanobacteria in soil biocrusts. Nat. Commun. 2022, 13, 1–10. [Google Scholar] [CrossRef] [PubMed]
  62. Gomov, B.V.; Mamkaeva, K. Proposal of a new genus Vampirovibrio for chlorellavorus bacteria previously assigned to Bdellovibrio. Mikrobiologia 1980, 49, 165–167. [Google Scholar]
  63. Soo, R.M.; Skennerton, C.T.; Sekiguchi, Y.; Imelfort, M.; Paech, S.J.; Dennis, P.G.; Steen, J.A.; Parks, D.H.; Tyson, G.W.; Hugenholtz, P. An Expanded Genomic Representation of the Phylum Cyanobacteria. Genome Biol. Evol. 2014, 6, 1031–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Di Rienzi, S.C.; Sharon, I.; Wrighton, K.C.; Koren, O.; Hug, L.A.; Thomas, B.C.; Goodrich, J.K.; Bell, J.T.; Spector, T.D.; Banfield, J.F.; et al. The human gut and groundwater harbor non-photosynthetic bacteria belonging to a new candidate phylum sibling to Cyanobacteria. eLife 2013, 2, e01102. [Google Scholar] [CrossRef]
  65. Soo, R.M.; Woodcroft, B.J.; Parks, D.H.; Tyson, G.W.; Hugenholtz, P. Back from the dead; the curious tale of the predatory cyanobacterium Vampirovibrio chlorellavorus. PeerJ 2015, 3, e968. [Google Scholar] [CrossRef] [Green Version]
  66. Ganuza, E.; Sellers, C.E.; Bennett, B.W.; Lyons, E.M.; Carney, L.T. A Novel Treatment Protects Chlorella at Commercial Scale from the Predatory Bacterium Vampirovibrio chlorellavorus. Front. Microbiol. 2016, 7, 848. [Google Scholar] [CrossRef] [Green Version]
  67. Tamura, K.; Stecher, G.; Kumar, S. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol. Biol. Evol. 2021, 38, 3022–3027. [Google Scholar] [CrossRef]
  68. Letunic, I.; Bork, P. Interactive Tree Of Life (iTOL) v5: An online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef]
  69. Bentley, S.D.; Chater, K.F.; Cerdeño-Tárraga, A.-M.; Challis, G.L.; Thomson, N.R.; James, K.D.; Harris, D.E.; Quail, M.A.; Kieser, H.; Harper, D.; et al. Complete genome sequence of the model actinomycete Streptomyces coelicolor A3(2). Nature 2002, 417, 141–147. [Google Scholar] [CrossRef]
  70. Ibrahimi, M.; Korichi, W.; Hafidi, M.; Lemee, L.; Ouhdouch, Y.; Loqman, S. Marine Actinobacteria: Screening for Predation Leads to the Discovery of Potential New Drugs against Multidrug-Resistant Bacteria. Antibiotics 2020, 9, 91. [Google Scholar] [CrossRef] [Green Version]
  71. Mahler, L.; Wink, K.; Beulig, R.J.; Scherlach, K.; Tovar, M.; Zang, E.; Martin, K.; Hertweck, C.; Belder, D.; Roth, M. Detection of antibiotics synthetized in microfluidic picolitre-droplets by various actinobacteria. Sci. Rep. 2018, 8, 13087. [Google Scholar] [CrossRef]
  72. Schwudke, D.; Strauch, E.; Krueger, M.; Appel, B. Taxonomic Studies of Predatory Bdellovibrios Based on 16S rRNA Analysis, Ribotyping and the hit Locus and Characterization of Isolates from the Gut of Animals. Syst. Appl. Microbiol. 2001, 24, 385–394. [Google Scholar] [CrossRef]
  73. Crossman, L.C.; Chen, H.; Cerdeño-Tárraga, A.-M.; Brooks, K.; Quail, M.A.; Pineiro, S.A.; Hobley, L.; Sockett, R.E.; Bentley, S.D.; Parkhill, J.; et al. A small predatory core genome in the divergent marine Bacteriovorax marinus SJ and the terrestrial Bdellovibrio bacteriovorus. ISME J. 2012, 7, 148–160. [Google Scholar] [CrossRef] [Green Version]
  74. Pineiro, S.A.; Stine, O.C.; Chauhan, A.; Steyert, S.R.; Smith, R.; Williams, H.N. Global survey of diversity among environmental saltwater Bacteriovoracaceae. Environ. Microbiol. 2007, 9, 2441–2450. [Google Scholar] [CrossRef]
  75. Baer, M.L.; Ravel, J.; Chun, J.; Hill, R.T.; Williams, H.N. A proposal for the reclassification of Bdellovibrio stolpii and Bdellovibrio starrii into a new genus, Bacteriovorax gen. nov. as Bacteriovorax stolpii comb. nov. and Bacteriovorax starrii comb. nov., respectively. Int. J. Syst. Evol. Microbiol. 2000, 50, 219–224. [Google Scholar] [CrossRef] [Green Version]
  76. Jurkevitch, E.; Minz, D.; Ramati, B.; Barel, G. Prey Range Characterization, Ribotyping, and Diversity of Soil and Rhizosphere Bdellovibrio spp. Isolated on Phytopathogenic Bacteria. Appl. Environ. Microbiol. 2000, 66, 2365–2371. [Google Scholar] [CrossRef] [Green Version]
  77. Rendulic, S.; Jagtap, P.; Rosinus, A.; Eppinger, M.; Baar, C.; Lanz, C.; Keller, H.; Lambert, C.; Evans, K.J.; Goesmann, A.; et al. A Predator Unmasked: Life Cycle of Bdellovibrio bacteriovorus from a Genomic Perspective. Science 2004, 303, 689–692. [Google Scholar] [CrossRef]
  78. Koval, S.F.; Hynes, S.H.; Flannagan, R.S.; Pasternak, Z.; Davidov, Y.; Jurkevitch, E. Bdellovibrio exovorus sp. nov., a novel predator of Caulobacter crescentus. Int. J. Syst. Evol. Microbiol. 2013, 63, 146–151. [Google Scholar] [CrossRef] [Green Version]
  79. Davidov, Y.; Huchon, D.; Koval, S.F.; Jurkevitch, E. A new ?-proteobacterial clade of Bdellovibrio-like predators: Implications for the mitochondrial endosymbiotic theory. Environ. Microbiol. 2006, 8, 2179–2188. [Google Scholar] [CrossRef]
  80. Wang, Z.; Kadouri, D.E.; Wu, M. Genomic insights into an obligate epibiotic bacterial predator: Micavibrio aeruginosavorus ARL-13. BMC Genom. 2011, 12, 453. [Google Scholar] [CrossRef] [Green Version]
  81. Gherna, R.; Woese, C. A Partial Phylogenetic Analysis of the “Flavobacter-Bacteroides” Phylum: Basis for Taxonomic Restructuring. Syst. Appl. Microbiol. 1992, 15, 513–521. [Google Scholar] [CrossRef] [PubMed]
  82. Meincke, L.; Copeland, A.; Lapidus, A.; Lucas, S.; Berry, K.W.; Del Rio, T.G.; Hammon, N.; Dalin, E.; Tice, H.; Pitluck, S.; et al. Complete genome sequence of Polynucleobacter necessarius subsp. asymbioticus type strain (QLW-P1DMWA-1T). Stand. Genom. Sci. 2012, 6, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Xie, G.; Bruce, D.C.; Challacombe, J.F.; Chertkov, O.; Detter, J.C.; Gilna, P.; Han, C.S.; Lucas, S.; Misra, M.; Myers, G.L.; et al. Genome Sequence of the Cellulolytic Gliding Bacterium Cytophaga hutchinsonii. Appl. Environ. Microbiol. 2007, 73, 3536–3546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Yoon, J.; Oku, N.; Park, S.; Kasai, H.; Yokota, A. Porifericola rhodea gen. nov., sp. nov., a new member of the phylum Bacteroidetes isolated by the bait-streaked agar technique. Antonie van Leeuwenhoek 2011, 100, 145–153. [Google Scholar] [CrossRef] [PubMed]
  85. Filippini, M.; Svercel, M.; Laczko, E.; Kaech, A.; Ziegler, U.; Bagheri, H.C. Fibrella aestuarina gen. nov., sp. nov., a filamentous bacterium of the family Cytophagaceae isolated from a tidal flat, and emended description of the genus Rudanella Weon et al. 2008. Int. J. Syst. Evol. Microbiol. 2011, 61, 184–189. [Google Scholar] [CrossRef] [Green Version]
  86. McBride, M.J.; Xie, G.; Martens, E.C.; Lapidus, A.; Henrissat, B.; Rhodes, R.G.; Goltsman, E.; Wang, W.; Xu, J.; Hunnicutt, D.W.; et al. Novel Features of the Polysaccharide-Digesting Gliding Bacterium Flavobacterium johnsoniae as Revealed by Genome Sequence Analysis. Appl. Environ. Microbiol. 2009, 75, 6864–6875. [Google Scholar] [CrossRef] [Green Version]
  87. Banning, E.C.; Casciotti, K.L.; Kujawinski, E.B. Novel strains isolated from a coastal aquifer suggest a predatory role for flavobacteria. FEMS Microbiol. Ecol. 2010, 73, 254–270. [Google Scholar] [CrossRef] [Green Version]
  88. Piñeiro-Vidal, M.; Riaza, A.; Santos, Y. Tenacibaculum discolor sp. nov. and Tenacibaculum gallaicum sp. nov., isolated from sole (Solea senegalensis) and turbot (Psetta maxima) culture systems. Int. J. Syst. Evol. Microbiol. 2008, 58, 21–25. [Google Scholar] [CrossRef]
  89. Suzuki, M.; Nakagawa, Y.; Harayama, S.; Yamamoto, S. Phylogenetic analysis and taxonomic study of marine Cytophaga-like bacteria: Proposal for Tenacibaculum gen. nov. with Tenacibaculum maritimum comb. nov. and Tenacibaculum ovolyticum comb. nov., and description of Tenacibaculum mesophilum sp. nov. and Tenacibaculum amylolyticum sp. nov. Int. J. Syst. Evol. Microbiol. 2001, 51, 1639–1652. [Google Scholar] [CrossRef]
  90. Kiss, H.; Nett, M.; Domin, N.; Martin, K.; Maresca, J.A.; Copeland, A.; Lapidus, A.; Lucas, S.; Berry, K.W.; Del Rio, T.G.; et al. Complete genome sequence of the filamentous gliding predatory bacterium Herpetosiphon aurantiacus type strain (114-95T). Stand. Genom. Sci. 2011, 5, 356–370. [Google Scholar] [CrossRef] [Green Version]
  91. Wang, Z.-J.; Liu, Q.-Q.; Zhao, L.-H.; Du, Z.-J.; Chen, G.-J. Bradymonas sediminis gen. nov., sp. nov., isolated from coastal sediment, and description of Bradymonadaceae fam. nov. and Bradymonadales ord. nov. Int. J. Syst. Evol. Microbiol. 2015, 65, 1542–1549. [Google Scholar] [CrossRef] [Green Version]
  92. Gong, Y.; Ping, X.-Y.; Zeng, C.-H.; Wang, S.-X.; Zhou, Y.; Wang, M.-Y.; Mu, D.-S.; Du, Z.-J. Predation capacity of Bradymonabacteria, a recently discovered group in the order Bradymonadales, isolated from marine sediments. Arch. Microbiol. 2022, 204, 1–11. [Google Scholar] [CrossRef]
  93. Spröer, C.; Reichenbach, H.; Stackebrandt, E. The correlation between morphological and phylogenetic classification of myxobacteria. Int. J. Syst. Evol. Microbiol. 1999, 49, 1255–1262. [Google Scholar] [CrossRef]
  94. Shimkets, L.; Woese, C.R. A phylogenetic analysis of the myxobacteria: Basis for their classification. Proc. Natl. Acad. Sci. USA 1992, 89, 9459–9463. [Google Scholar] [CrossRef] [Green Version]
  95. Huntley, S.; Hamann, N.; Wegener-Feldbrugge, S.; Treuner-Lange, A.; Kube, M.; Reinhardt, R.; Klages, S.; Muller, R.; Ronning, C.M.; Nierman, W.C.; et al. Comparative Genomic Analysis of Fruiting Body Formation in Myxococcales. Mol. Biol. Evol. 2010, 28, 1083–1097. [Google Scholar] [CrossRef] [Green Version]
  96. Iizuka, T.; Jojima, Y.; Fudou, R.; Yamanaka, S. Isolation of myxobacteria from the marine environment. FEMS Microbiol. Lett. 1998, 169, 317–322. [Google Scholar] [CrossRef]
  97. Huntley, S.; Zhang, Y.; Treuner-Lange, A.; Kneip, S.; Sensen, C.W.; Søgaard-Andersen, L. Complete Genome Sequence of the Fruiting Myxobacterium Corallococcus coralloides DSM 2259. J. Bacteriol. 2012, 194, 3012–3013. [Google Scholar] [CrossRef] [Green Version]
  98. Livingstone, P.; Morphew, R.; Whitworth, D.E. Myxobacteria Are Able to Prey Broadly upon Clinically-Relevant Pathogens, Exhibiting a Prey Range Which Cannot Be Explained by Phylogeny. Front. Microbiol. 2017, 8, 1593. [Google Scholar] [CrossRef] [Green Version]
  99. Schieferdecker, S.; Exner, T.E.; Gross, H.; Roth, M.; Nett, M. New myxothiazols from the predatory bacterium Myxococcus fulvus. J. Antibiot. 2014, 67, 519–525. [Google Scholar] [CrossRef] [Green Version]
  100. Goldman, B.S.; Nierman, W.C.; Kaiser, D.; Slater, S.C.; Durkin, A.S.; Eisen, J.A.; Ronning, C.M.; Barbazuk, W.B.; Blanchard, M.; Field, C.; et al. Evolution of sensory complexity recorded in a myxobacterial genome. Proc. Natl. Acad. Sci. USA 2006, 103, 15200–15205. [Google Scholar] [CrossRef] [Green Version]
  101. Jain, R.; Habermann, B.H.; Mignot, T. Complete Genome Assembly of Myxococcus xanthus Strain DZ2 Using Long High-Fidelity (HiFi) Reads Generated with PacBio Technology. Genome Announc. 2021, 10, e0053021. [Google Scholar] [CrossRef] [PubMed]
  102. Oyaizu, H.; Woese, C. Phylogenetic Relationships Among the Sulfate Respiring Bacteria, Myxobacteria and Purple Bacteria. Syst. Appl. Microbiol. 1985, 6, 257–263. [Google Scholar] [CrossRef]
  103. Vos, M.; Velicer, G.J. Genetic Population Structure of the Soil Bacterium Myxococcus xanthus at the Centimeter Scale. Appl. Environ. Microbiol. 2006, 72, 3615–3625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Garcia, R.; Gerth, K.; Stadler, M.; Dogma, I.J.; Müller, R. Expanded phylogeny of myxobacteria and evidence for cultivation of the ‘unculturables’. Mol. Phylogenetics Evol. 2010, 57, 878–887. [Google Scholar] [CrossRef]
  105. Iizuka, T.; Jojima, Y.; Fudou, R.; Hiraishi, A.; Ahn, J.-W.; Yamanaka, S. Plesiocystis pacifica gen. nov., sp. nov., a marine myxobacterium that contains dihydrogenated menaquinone, isolated from the Pacific coasts of Japan. Int. J. Syst. Evol. Microbiol. 2003, 53, 189–195. [Google Scholar] [CrossRef] [Green Version]
  106. Zaburannyi, N.; Bunk, B.; Maier, J.; Overmann, J.; Müller, R. Genome Analysis of the Fruiting Body-Forming Myxobacterium Chondromyces crocatus Reveals High Potential for Natural Product Biosynthesis. Appl. Environ. Microbiol. 2016, 82, 1945–1957. [Google Scholar] [CrossRef] [Green Version]
  107. Garcia, R.; Gemperlein, K.; Müller, R. Minicystis rosea gen. nov., sp. nov., a polyunsaturated fatty acid-rich and steroid-producing soil myxobacterium. Int. J. Syst. Evol. Microbiol. 2014, 64, 3733–3742. [Google Scholar] [CrossRef]
  108. Awal, R.P.; Garcia, R.; Müller, R. Racemicystis crocea gen. nov., sp. nov., a soil myxobacterium in the family Polyangiaceae. Int. J. Syst. Evol. Microbiol. 2016, 66, 2389–2395. [Google Scholar] [CrossRef] [Green Version]
  109. Schneiker, S.; Perlova, O.; Kaiser, O.; Gerth, K.; Alici, A.; Altmeyer, M.O.; Bartels, D.; Bekel, T.; Beyer, S.; Bode, E.; et al. Complete genome sequence of the myxobacterium Sorangium cellulosum. Nat. Biotechnol. 2007, 25, 1281–1289. [Google Scholar] [CrossRef]
  110. Rotaru, A.-E. Visualization of Candidate Division OP3 Cocci in Limonene-Degrading Methanogenic Cultures. J. Microbiol. Biotechnol. 2012, 22, 457–461. [Google Scholar] [CrossRef] [Green Version]
  111. Williams, L.E.; Baltrus, D.A.; O’donnell, S.D.; Skelly, T.J.; Martin, M.O. Complete Genome Sequence of the Predatory Bacterium Ensifer adhaerens Casida A. Genome Announc. 2017, 5, e01344-17. [Google Scholar] [CrossRef] [Green Version]
  112. Rudder, S.; Doohan, F.; Creevey, C.J.; Wendt, T.; Mullins, E. Genome sequence of Ensifer adhaerens OV14 provides insights into its ability as a novel vector for the genetic transformation of plant genomes. BMC Genom. 2014, 15, 268. [Google Scholar] [CrossRef] [Green Version]
  113. Cain, C.C.; Lee, D.; Waldo, R.H.; Henry, A.T.; Casida, E.J.; Wani, M.C.; Wall, M.E.; Oberlies, N.H.; Falkinham, J.O. Synergistic Antimicrobial Activity of Metabolites Produced by a Nonobligate Bacterial Predator. Antimicrob. Agents Chemother. 2003, 47, 2113–2117. [Google Scholar] [CrossRef] [Green Version]
  114. Poehlein, A.; Kusian, B.; Friedrich, B.; Daniel, R.; Bowien, B. Complete Genome Sequence of the Type Strain Cupriavidus necator N-1. J. Bacteriol. 2011, 193, 5017. [Google Scholar] [CrossRef] [Green Version]
  115. Saeedi, A.; Cummings, N.J.; McLean, D.; Connerton, I.F.; Connerton, P.L. Venatorbacter cucullus gen. nov sp. nov a novel bacterial predator. Sci. Rep. 2021, 11, 21393. [Google Scholar] [CrossRef]
  116. Zhao, Y.; Jiang, T.; Xu, H.; Xu, G.; Qian, G.; Liu, F. Characterization of Lysobacter spp. strains and their potential use as biocontrol agents against pear anthracnose. Microbiol. Res. 2020, 242, 126624. [Google Scholar] [CrossRef]
  117. Puopolo, G.; Giovannini, O.; Pertot, I. Lysobacter capsici AZ78 can be combined with copper to effectively control Plasmopara viticola on grapevine. Microbiol. Res. 2014, 169, 633–642. [Google Scholar] [CrossRef]
  118. Takami, H.; Toyoda, A.; Uchiyama, I.; Itoh, T.; Takaki, Y.; Arai, W.; Nishi, S.; Kawai, M.; Shin-Ya, K.; Ikeda, H. Complete genome sequence and expression profile of the commercial lytic enzyme producer Lysobacter enzymogenes M497-1. DNA Res. 2017, 24, 169–177. [Google Scholar] [CrossRef] [Green Version]
  119. Aslam, Z.; Yasir, M.; Jeon, C.O.; Chung, Y.R. Lysobacter oryzae sp. nov., isolated from the rhizosphere of rice (Oryza sativa L.). Int. J. Syst. Evol. Microbiol. 2009, 59, 675–680. [Google Scholar] [CrossRef]
  120. Anzai, Y.; Kim, H.; Park, J.-Y.; Wakabayashi, H.; Oyaizu, H. Phylogenetic affiliation of the pseudomonads based on 16S rRNA sequence. Int. J. Syst. Evol. Microbiol. 2000, 50, 1563–1589. [Google Scholar] [CrossRef] [Green Version]
  121. Hiraishi, A.; Kishimoto, N.; Kosako, Y.; Wakao, N.; Tano, T. Phylogenetic position of the menaquinone-containing acidophilic chemo-organotrophAcidobacterium capsulatum. FEMS Microbiol. Lett. 1995, 132, 91–94. [Google Scholar] [CrossRef] [PubMed]
  122. Stackebrandt, E.; Charfreitag, O. Partial 16S rRNA primary structure of five Actinomyces species: Phylogenetic implications and development of an Actinomyces israelii-specific oligonucleotide probe. J. Gen. Microbiol. 1990, 136, 37–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Burggraf, S.; Olsen, G.; Stetter, K.; Woese, C. A Phylogenetic Analysis of Aquifex pyrophilus. Syst. Appl. Microbiol. 1992, 15, 352–356. [Google Scholar] [CrossRef] [PubMed]
  124. Tamaki, H.; Tanaka, Y.; Matsuzawa, H.; Muramatsu, M.; Meng, X.-Y.; Hanada, S.; Mori, K.; Kamagata, Y. Armatimonas rosea gen. nov., sp. nov., of a novel bacterial phylum, Armatimonadetes phyl. nov., formally called the candidate phylum OP10. Int. J. Syst. Evol. Microbiol. 2011, 61, 1442–1447. [Google Scholar] [CrossRef]
  125. Katayama, T.; Nobu, M.K.; Kusada, H.; Meng, X.-Y.; Hosogi, N.; Uematsu, K.; Yoshioka, H.; Kamagata, Y.; Tamaki, H. Isolation of a member of the candidate phylum ‘Atribacteria’ reveals a unique cell membrane structure. Nat. Commun. 2020, 11, 6381. [Google Scholar] [CrossRef]
  126. Setyorini, E.; Kim, Y.-J.; Takenaka, S.; Murakami, S.; Aoki, K. Purification and characterization of a halotolerant intracellular protease fromBacillus subtilis strain FP-133. J. Basic Microbiol. 2006, 46, 294–304. [Google Scholar] [CrossRef] [Green Version]
  127. Urios, L.; Agogué, H.; Lesongeur, F.; Stackebrandt, E.; Lebaron, P. Balneola vulgaris gen. nov., sp. nov., a member of the phylum Bacteroidetes from the north-western Mediterranean Sea. Int. J. Syst. Evol. Microbiol. 2006, 56, 1883–1887. [Google Scholar] [CrossRef] [Green Version]
  128. Nakai, R.; Nishijima, M.; Tazato, N.; Handa, Y.; Karray, F.; Sayadi, S.; Isoda, H.; Naganuma, T. Oligoflexus tunisiensis gen. nov., sp. nov., a Gram-negative, aerobic, filamentous bacterium of a novel proteobacterial lineage, and description of Oligoflexaceae fam. nov., Oligoflexales ord. nov. and Oligoflexia classis nov. Int. J. Syst. Evol. Microbiol. 2014, 64, 3353–3359. [Google Scholar] [CrossRef]
  129. Mori, K.; Yamaguchi, K.; Sakiyama, Y.; Urabe, T.; Suzuki, K.-I. Caldisericum exile gen. nov., sp. nov., an anaerobic, thermophilic, filamentous bacterium of a novel bacterial phylum, Caldiserica phyl. nov., originally called the candidate phylum OP5, and description of Caldisericaceae fam. nov., Caldisericales ord. nov. and Caldisericia classis nov. Int. J. Syst. Evol. Microbiol. 2009, 59, 2894–2898. [Google Scholar] [CrossRef]
  130. Miroshnichenko, M.L.; Kostrikina, N.A.; Chernyh, N.A.; Pimenov, N.V.; Tourova, T.P.; Antipov, A.N.; Spring, S.; Stackebrandt, E.; Bonch-Osmolovskaya, E.A. Caldithrix abyssi gen. nov., sp. nov., a nitrate-reducing, thermophilic, anaerobic bacterium isolated from a Mid-Atlantic Ridge hydrothermal vent, represents a novel bacterial lineage. Int. J. Syst. Evol. Microbiol. 2003, 53, 323–329. [Google Scholar] [CrossRef]
  131. Tmoko, P.; Fukushi, H.; Ochiai, Y.; Yamaguchi, T.; Hirai, K. Phylogenetic Analysis of the Genus Chlamydia Based on 16S rRNA Gene Sequences. Int. J. Syst. Evol. Microbiol. 1997, 47, 425–431. [Google Scholar] [CrossRef] [Green Version]
  132. Alexander, B.; Andersen, J.H.; Cox, R.P.; Imhoff, J.F. Phylogeny of green sulfur bacteria on the basis of gene sequences of 16S rRNA and of the Fenna-Matthews-Olson protein. Arch. Microbiol. 2002, 178, 131–140. [Google Scholar] [CrossRef]
  133. Hanada, S.; Hiraishi, A.; Shimada, K.; Matsuura, K. Chloroflexus aggregans sp. nov., a Filamentous Phototrophic Bacterium Which Forms Dense Cell Aggregates by Active Gliding Movement. Int. J. Syst. Evol. Microbiol. 1995, 45, 676–681. [Google Scholar] [CrossRef] [Green Version]
  134. Macy, J.M.; Nunan, K.; Hagen, K.D.; Dixon, D.R.; Harbour, P.J.; Cahill, M.; Sly, L.I. Chrysiogenes arsenatis gen. nov., sp. nov., a New Arsenate-Respiring Bacterium Isolated from Gold Mine Wastewater. Int. J. Syst. Evol. Microbiol. 1996, 46, 1153–1157. [Google Scholar] [CrossRef] [Green Version]
  135. Alexiev, A.; Coil, D.A.; Badger, J.H.; Enticknap, J.; Ward, N.; Robb, F.T.; Eisen, J.A. Complete Genome Sequence of Coprothermobacter proteolyticus DSM 5265. Genome Announc. 2014, 2, e00470-14. [Google Scholar] [CrossRef] [Green Version]
  136. Greene, A.C.; Patel, B.K.C.; Sheehy, A.J. Deferribacter thermophilus gen. nov., sp. nov., a Novel Thermophilic Manganese- and Iron-Reducing Bacterium Isolated from a Petroleum Reservoir. Int. J. Syst. Evol. Microbiol. 1997, 47, 505–509. [Google Scholar] [CrossRef] [Green Version]
  137. Rainey, F.A.; Nobre, M.F.; Schumann, P.; Stackebrandt, E.; DA Costa, M.S. Phylogenetic Diversity of the Deinococci as Determined by 16S Ribosomal DNA Sequence Comparison. Int. J. Syst. Evol. Microbiol. 1997, 47, 510–514. [Google Scholar] [CrossRef] [Green Version]
  138. Gibbs, M.D.; Reeves, R.A.; Bergquist, P.L. Cloning, sequencing, and expression of a xylanase gene from the extreme thermophile Dictyoglomus thermophilum Rt46B.1 and activity of the enzyme on fiber-bound substrate. Appl. Environ. Microbiol. 1995, 61, 4403–4408. [Google Scholar] [CrossRef] [Green Version]
  139. Geissinger, O.; Herlemann, D.P.R.; Moörschel, E.; Maier, U.G.; Brune, A. The Ultramicrobacterium “Elusimicrobium minutum” gen. nov., sp. nov., the First Cultivated Representative of the Termite Group 1 Phylum. Appl. Environ. Microbiol. 2009, 75, 2831–2840. [Google Scholar] [CrossRef] [Green Version]
  140. Beéra-Maillet, C.; Ribot, Y.; Forano, E. Fiber-Degrading Systems of Different Strains of the Genus Fibrobacter. Appl. Environ. Microbiol. 2004, 70, 2172–2179. [Google Scholar] [CrossRef] [Green Version]
  141. Kapatral, V.; Anderson, I.; Ivanova, N.; Reznik, G.; Los, T.; Lykidis, A.; Bhattacharyya, A.; Bartman, A.; Gardner, W.; Grechkin, G.; et al. Genome Sequence and Analysis of the Oral Bacterium Fusobacterium nucleatum Strain ATCC 25586. J. Bacteriol. 2002, 184, 2005–2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Zhang, H.; Sekiguchi, Y.; Hanada, S.; Hugenholtz, P.; Kim, H.; Kamagata, Y.; Nakamura, K. Gemmatimonas aurantiaca gen. nov., sp. nov., a Gram-negative, aerobic, polyphosphate-accumulating micro-organism, the first cultured representative of the new bacterial phylum Gemmatimonadetes phyl. nov. Int. J. Syst. Evol. Microbiol. 2003, 53, 1155–1163. [Google Scholar] [CrossRef] [PubMed]
  143. Iino, T.; Mori, K.; Uchino, Y.; Nakagawa, T.; Harayama, S.; Suzuki, K.-I. Ignavibacterium album gen. nov., sp. nov., a moderately thermophilic anaerobic bacterium isolated from microbial mats at a terrestrial hot spring and proposal of Ignavibacteria classis nov., for a novel lineage at the periphery of green sulfur bacteria. Int. J. Syst. Evol. Microbiol. 2010, 60, 1376–1382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Cho, J.-C.; Vergin, K.L.; Morris, R.M.; Giovannoni, S.J. Lentisphaera araneosa gen. nov., sp. nov, a transparent exopolymer producing marine bacterium, and the description of a novel bacterial phylum, Lentisphaerae. Environ. Microbiol. 2004, 6, 611–621. [Google Scholar] [CrossRef]
  145. Westberg, J.; Persson, A.; Holmberg, A.; Goesmann, A.; Lundeberg, J.; Johansson, K.-E.; Pettersson, B.; Uhlén, M. The Genome Sequence of Mycoplasma mycoides subsp. mycoides SC Type Strain PG1T, the Causative Agent of Contagious Bovine Pleuropneumonia (CBPP). Genome Res. 2004, 14, 221–227. [Google Scholar] [CrossRef] [Green Version]
  146. Yarza, P.; Spröer, C.; Swiderski, J.; Mrotzek, N.; Spring, S.; Tindall, B.J.; Gronow, S.; Pukall, R.; Klenk, H.-P.; Lang, E.; et al. Sequencing orphan species initiative (SOS): Filling the gaps in the 16S rRNA gene sequence database for all species with validly published names. Syst. Appl. Microbiol. 2013, 36, 69–73. [Google Scholar] [CrossRef] [Green Version]
  147. Keuter, S.; Kruse, M.; Lipski, A.; Spieck, E. Relevance of Nitrospira for nitrite oxidation in a marine recirculation aquaculture system and physiological features of a Nitrospira marina-like isolate. Environ. Microbiol. 2011, 13, 2536–2547. [Google Scholar] [CrossRef]
  148. Ward, N.; Rainey, F.A.; Stackebrandt, E.; Schlesner, H. Unraveling the extent of diversity within the order Planctomycetales. Appl. Environ. Microbiol. 1995, 61, 2270–2275. [Google Scholar] [CrossRef] [Green Version]
  149. Andrésson, O.S.; Fridjónsson, O.H. The sequence of the single 16S rRNA gene of the thermophilic eubacterium Rhodothermus marinus reveals a distant relationship to the group containing Flexibacter, Bacteroides, and Cytophaga species. J. Bacteriol. 1994, 176, 6165–6169. [Google Scholar] [CrossRef] [Green Version]
  150. Allison, M.J.; Mayberry, W.R.; Mcsweeney, C.S.; Stahl, D.A. Synergistes jonesii, gen. nov., sp.nov.: A Rumen Bacterium That Degrades Toxic Pyridinediols. Syst. Appl. Microbiol. 1992, 15, 522–529. [Google Scholar] [CrossRef]
  151. Friedrich, M.W. Phylogenetic Analysis Reveals Multiple Lateral Transfers of Adenosine-5′-Phosphosulfate Reductase Genes among Sulfate-Reducing Microorganisms. J. Bacteriol. 2002, 184, 278–289. [Google Scholar] [CrossRef] [Green Version]
  152. Wu, D.; Raymond, J.; Wu, M.; Chatterji, S.; Ren, Q.; Graham, J.E.; Bryant, D.A.; Robb, F.; Colman, A.; Tallon, L.J.; et al. Complete Genome Sequence of the Aerobic CO-Oxidizing Thermophile Thermomicrobium roseum. PLoS ONE 2009, 4, e4207. [Google Scholar] [CrossRef] [Green Version]
  153. Nelson, K.E.; Clayton, R.A.; Gill, S.R.; Gwinn, M.L.; Dodson, R.J.; Haft, D.H.; Hickey, E.K.; Peterson, J.D.; Nelson, W.C.; Ketchum, K.A.; et al. Evidence for lateral gene transfer between Archaea and Bacteria from genome sequence of Thermotoga maritima. Nature 1999, 399, 323–329. [Google Scholar] [CrossRef]
  154. Ward-Rainey, N.; Rainey, F.A.; Schlesner, H.; Stackebrandt, E. Assignment of hitherto unidentified 16S rDNA species to a main line of descent within the domain Bacteria. Microbiology 1995, 141, 3247–3250. [Google Scholar] [CrossRef] [Green Version]
  155. Phillips, K.E.; Akbar, S.; Stevens, D.C. Concepts and conjectures concerning predatory performance of myxobacteria. Front. Microbiol. 2022, 13. [Google Scholar] [CrossRef]
  156. Martin, W.F.; Garg, S.G.; Zimorski, V. Endosymbiotic theories for eukaryote origin. Philos. Trans. R. Soc. B Biol. Sci. 2015, 370, 20140330. [Google Scholar] [CrossRef] [Green Version]
  157. Dacks, J.B.; Field, M.C.; Buick, R.; Eme, L.; Gribaldo, S.; Roger, A.J.; Brochier-Armanet, C.; Devos, D.P. The changing view of eukaryogenesis–fossils, cells, lineages and how they all come together. J. Cell Sci. 2016, 129, 3695–3703. [Google Scholar] [CrossRef] [Green Version]
  158. Eme, L.; Spang, A.; Lombard, J.; Stairs, C.W.; Ettema, T.J.G. Erratum: Archaea and the origin of eukaryotes. Nat. Rev. Genet. 2017, 16, 120. [Google Scholar] [CrossRef] [Green Version]
  159. Burns, J.A.; Pittis, A.A.; Kim, E. Gene-based predictive models of trophic modes suggest Asgard archaea are not phagocytotic. Nat. Ecol. Evol. 2018, 2, 697–704. [Google Scholar] [CrossRef]
  160. Imachi, H.; Nobu, M.K.; Nakahara, N.; Morono, Y.; Ogawara, M.; Takaki, Y.; Takano, Y.; Uematsu, K.; Ikuta, T.; Ito, M.; et al. Isolation of an archaeon at the prokaryote–eukaryote interface. Nature 2020, 577, 519–525. [Google Scholar] [CrossRef] [Green Version]
  161. Davidov, Y.; Jurkevitch, E. Predation between prokaryotes and the origin of eukaryotes. Bioessays 2009, 31, 748–757. [Google Scholar] [CrossRef] [PubMed]
  162. Bremer, N.; Tria, F.D.K.; Skejo, J.; Garg, S.G.; Martin, W.F. Ancestral state reconstructions trace mitochondria but not phagocytosis to the last eukaryotic common ancestor. Genome Biol. Evol. 2022, 14, evac079. [Google Scholar] [CrossRef] [PubMed]
  163. Casida, L.E. Ensifer adhaerens gen. nov., sp. nov.: A Bacterial Predator of Bacteria in Soil. Int. J. Syst. Evol. Microbiol. 1982, 32, 339–345. [Google Scholar] [CrossRef]
  164. Hendrickson, E.L.; Bor, B.; Kerns, K.A.; Lamont, E.I.; Chang, Y.; Liu, J.; Cen, L.; Schulte, F.; Hardt, M.; Shi, W.; et al. Transcriptome of Epibiont Saccharibacteria Nanosynbacter lyticus Strain TM7x During the Establishment of Symbiosis. J. Bacteriol. 2022. [Google Scholar] [CrossRef] [PubMed]
  165. Tian, R.; Ning, D.; He, Z.; Zhang, P.; Spencer, S.J.; Gao, S.; Shi, W.; Wu, L.; Zhang, Y.; Yang, Y.; et al. Small and mighty: Adaptation of superphylum Patescibacteria to groundwater environment drives their genome simplicity. Microbiome 2020, 8, 1–15. [Google Scholar] [CrossRef] [Green Version]
  166. Lonhienne, T.G.; Sagulenko, E.; Webb, R.I.; Lee, K.C.; Franke, J.; Devos, D.P.; Nouwens, A.; Carroll, B.J.; Fuerst, J.A. Endocytosis-like protein uptake in the bacterium Gemmata obscuriglobus. Proc. Natl. Acad. Sci. USA 2010, 107, 12883–12888. [Google Scholar] [CrossRef] [Green Version]
  167. Franzmann, P.D.; Skerman, V.B.D. Gemmata obscuriglobus, a new genus and species of the budding bacteria. Antonie van Leeuwenhoek 1984, 50, 261–268. [Google Scholar] [CrossRef]
  168. Dedysh, S.N.; Henke, P.; Ivanova, A.A.; Kulichevskaya, I.S.; Philippov, D.A.; Meier-Kolthoff, J.P.; Göker, M.; Huang, S.; Overmann, J. 100-year-old enigma solved: Identification, genomic characterization and biogeography of the yet uncultured Planctomyces bekefii. Environ. Microbiol. 2019, 22, 198–211. [Google Scholar] [CrossRef]
  169. Wiegand, S.; Jogler, M.; Boedeker, C.; Pinto, D.; Vollmers, J.; Rivas-Marín, E.; Kohn, T.; Peeters, S.H.; Heuer, A.; Rast, P.; et al. Cultivation and functional characterization of 79 planctomycetes uncovers their unique biology. Nat. Microbiol. 2019, 5, 126–140. [Google Scholar] [CrossRef]
  170. Kaboré, O.D.; Godreuil, S.; Drancourt, M. Planctomycetes as Host-Associated Bacteria: A Perspective That Holds Promise for Their Future Isolations, by Mimicking Their Native Environmental Niches in Clinical Microbiology Laboratories. Front. Cell. Infect. Microbiol. 2020, 10, 519301. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kamada, S.; Wakabayashi, R.; Naganuma, T. Phylogenetic Revisit to a Review on Predatory Bacteria. Microorganisms 2023, 11, 1673. https://doi.org/10.3390/microorganisms11071673

AMA Style

Kamada S, Wakabayashi R, Naganuma T. Phylogenetic Revisit to a Review on Predatory Bacteria. Microorganisms. 2023; 11(7):1673. https://doi.org/10.3390/microorganisms11071673

Chicago/Turabian Style

Kamada, Saki, Ryoka Wakabayashi, and Takeshi Naganuma. 2023. "Phylogenetic Revisit to a Review on Predatory Bacteria" Microorganisms 11, no. 7: 1673. https://doi.org/10.3390/microorganisms11071673

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

Article Metrics

Back to TopTop