Skip to main content
Applied and Environmental Microbiology logoLink to Applied and Environmental Microbiology
. 2018 Aug 1;84(16):e00859-18. doi: 10.1128/AEM.00859-18

Microevolution of Streptococcus agalactiae ST-261 from Australia Indicates Dissemination via Imported Tilapia and Ongoing Adaptation to Marine Hosts or Environment

Minami Kawasaki a, Jerome Delamare-Deboutteville a, Rachel O Bowater b, Mark J Walker c, Scott Beatson c, Nouri L Ben Zakour c,*, Andrew C Barnes a,
Editor: Johanna Björkrothd
PMCID: PMC6070751  PMID: 29915111

Streptococcus agalactiae (GBS) is a significant pathogen of humans and animals. Some lineages have become adapted to particular hosts, and serotype Ib is highly specialized to fish. Here, we show that this lineage is likely to have been distributed widely by the global trade in tilapia for aquaculture, with probable introduction into Australia in the 1970s and subsequent dissemination in wild fish populations. We report here the variability in the polysaccharide capsule among this lineage but identify a cohort of common surface proteins that may be a focus of future vaccine development to reduce the biosecurity risk in international fish trade.

KEYWORDS: Streptococcus agalactiae, evolution, genome analysis, epidemiology, fish, aquaculture

ABSTRACT

Streptococcus agalactiae (group B Streptococcus [GBS]) causes disease in a wide range of animals. The serotype Ib lineage is highly adapted to aquatic hosts, exhibiting substantial genome reduction compared with terrestrial conspecifics. Here, we sequence genomes from 40 GBS isolates, including 25 isolates from wild fish and captive stingrays in Australia, six local veterinary or human clinical isolates, and nine isolates from farmed tilapia in Honduras, and compared them with 42 genomes from public databases. Phylogenetic analysis based on nonrecombinant core-genome single nucleotide polymorphisms (SNPs) indicated that aquatic serotype Ib isolates from Queensland were distantly related to local veterinary and human clinical isolates. In contrast, Australian aquatic isolates are most closely related to a tilapia isolate from Israel, differing by only 63 core-genome SNPs. A consensus minimum spanning tree based on core-genome SNPs indicates the dissemination of sequence type 261 (ST-261) from an ancestral tilapia strain, which is congruent with several introductions of tilapia into Australia from Israel during the 1970s and 1980s. Pangenome analysis identified 1,440 genes as core, with the majority being dispensable or strain specific, with non-protein-coding intergenic regions (IGRs) divided among core and strain-specific genes. Aquatic serotype Ib strains have lost many virulence factors during adaptation, but six adhesins were well conserved across the aquatic isolates and might be critical for virulence in fish and for targets in vaccine development. The close relationship among recent ST-261 isolates from Ghana, the United States, and China with the Israeli tilapia isolate from 1988 implicates the global trade in tilapia seed for aquaculture in the widespread dissemination of serotype Ib fish-adapted GBS.

IMPORTANCE Streptococcus agalactiae (GBS) is a significant pathogen of humans and animals. Some lineages have become adapted to particular hosts, and serotype Ib is highly specialized to fish. Here, we show that this lineage is likely to have been distributed widely by the global trade in tilapia for aquaculture, with probable introduction into Australia in the 1970s and subsequent dissemination in wild fish populations. We report here the variability in the polysaccharide capsule among this lineage but identify a cohort of common surface proteins that may be a focus of future vaccine development to reduce the biosecurity risk in international fish trade.

INTRODUCTION

Streptococcus agalactiae, or Lancefield group B Streptococcus (GBS), is a commensal and occasionally pathogenic bacterium with a very diverse host range. A common commensal in the urogenital tracts of humans, GBS is also a leading cause of morbidity in newborns causing meningitis, septicemia, and pneumonia (14). S. agalactiae can cause septicemic infections in cattle, domestic dogs and cats, camels, reptiles, and amphibians (58). In fish, disease outbreaks caused by S. agalactiae have substantial impact on the aquaculture industry, particularly on the production of warm freshwater species, such as tilapia (Oreochromis spp.) (912). Most outbreaks to date in freshwater farmed fish have resulted from infection by highly adapted strains of GBS with genomes that are 10 to 15% smaller than their terrestrial conspecifics (13). Unusually, S. agalactiae also causes significant mortality in wild aquatic animals, including grouper, stingrays, and mullet (5), suggesting further adaptation to marine and freshwater aquatic hosts.

Microevolution within a bacterial species can be driven by host or environmental adaptation (13, 14), permitting an inference of the epidemiology of disease outbreaks and how pathogens may have transferred within and between geographic regions (1416). This requires an analysis of factors that evolve at a sufficiently rapid pace to be informative over relatively short timespans. In GBS, capsular serotyping either with antibodies or by molecular serotyping (i.e., sequencing of the capsular operon) has become a widely used method of typing for population studies (1720), and currently, S. agalactiae can be divided into 10 capsular serotypes (Ia, Ib, and II to IX) (18, 21). Determining capsular serotypes is also critical for vaccine formulation, since capsular polysaccharide (CPS) is highly immunogenic, and antibodies against CPS can confer excellent protection against infections by the homologous CPS serotype (20, 22, 23). Further typing resolution is provided by multilocus sequence typing (MLST), a method that has been employed to great effect to conduct global population studies of isolates based on genetic variations among relatively slowly evolving housekeeping genes (17). Combining molecular serotyping and MLST in the analysis of S. agalactiae revealed that the majority of isolates associated with aquatic environments and hosts fall within serotypes Ia and Ib, in which Ia isolates belong to sequence type 7 (ST-7) in clonal complex 7 (CC-7) and ST-103 in CC-103 (12, 18, 2428). Serotype Ib strains isolated in Central and South America are ST-260 and ST-552 in CC-552 (27, 29), and strains isolated in Australia, Israel, Belgium, China, Ghana, the United States, and Southeast Asia belong to ST-261 (5, 6, 9, 13, 2931). Serotype III is commonly causative of disease in humans but has also been isolated from fish in Thailand, China, and recently, Brazil (26, 3234).

While capsular serotyping and MLST have been useful in inferring the origins and dispersal of GBS subtypes, they do not display sufficient resolution to explore the spread and evolution within individual sequence types, nor can they reflect the complete genetic diversity of S. agalactiae (35). The rapid fall in the cost of whole-genome sequencing coupled with multiplexing and rapid development of open-source bioinformatics tools has permitted a much deeper analysis of evolution, host adaptation, and epidemiological modeling within a single bacterial species (36), including those from aquatic hosts (14, 15). Bacteria, such as S. agalactiae, that can colonize multiple host species often have greater genomic intraspecies diversity (37). In GBS, two major evolutionary trends have been implicated in rapid adaptation to new hosts, namely, the acquisition of new genes by lateral gene transfer and genome reduction via gene loss integral to host specialization (13, 38). For example, S. agalactiae Ia strains GD201008-001 and ZQ0910, isolated from tilapia in China, carry a 10-kb genomic island (GI) which is absent from their closely related human isolate A909. Moreover, this 10-kb GI bears many similarities with the Streptococcus anginosus SK52/DSM 20563 genome sequence, suggesting possible transfer from S. anginosus to GBS, with implications for virulence in tilapia (13, 39). During fish host adaptation, serotype Ib strains have undergone reductive evolution, resulting in 10 to 25% of their genomes being lost compared to terrestrial S. agalactiae isolates and serotype Ia piscine strains (13). The evolution of S. agalactiae by genome reduction is an ongoing process, with a high number of pseudogenes present in GBS genomes from aquatic sources (13).

The evolution of S. agalactiae and adaptation to aquatic hosts is an incomplete and ongoing process; consequently, sequencing the genomes from a few isolates is insufficient to understand the full potential genetic diversity of S. agalactiae as a species (35). The pangenome or supragenome of a bacterial species defines the full complement of genes, or the union of all the gene sets, within the species (35). This pangenome is subdivided into its core genome, which includes all the genes that are present in all the strains of the same bacterial species and must therefore be responsible for essential biological functions to allow the species to survive, and the accessory genome, containing species-specific genes that are unique to single strains or constrained to a cohort of strains within the species; these genes contribute to the diversity makeup of the species. The pangenome of a species resolves the true genomic diversity of that species and permits the identification of gene cohorts that are essential to the species as a whole, along with gene complements in the accessory genome that permit host or habitat specialization (35). Moreover, by identifying potential antigens within the pangenome that are conserved across all strains that infect a particular host type, vaccine targets can be specified that are likely to cross-protect (35, 40). Indeed, the first multicomponent protein-containing universal vaccine against human S. agalactiae was developed using a pangenome reverse-vaccinology approach by analyzing eight human isolates to predict putative antigens that were conserved among those strains (41). Some antigens in this vaccine are in the accessory genome; consequently, it is important to analyze as much as possible of the dispensable genome for vaccine development (35, 41).

The S. agalactiae pangenome is now well advanced but still “open” (i.e., new genes continue to be added with more sequenced genomes) and geographically constrained (35, 40). In the present study, we sequenced the genomes of new aquatic S. agalactiae strains isolated from tilapia in Honduras and from wild and captive marine fish in Australia. We infer the potential epidemiological distribution of ST-261 in aquatic hosts in Australia and show continuing adaptation to saltwater fish. Moreover, we identify conserved surface proteins across ST-260 and ST-261 that may have potential for incorporation into vaccines for aquaculture of important food fish species, such as tilapia and grouper.

RESULTS AND DISCUSSION

GBS isolates from marine fish and rays in Queensland and tilapia in Honduras belong to differing host-adapted lineages.

The average size of the draft genomes of isolates from fish and rays in Queensland was 1,801,022 bp, containing 1,881 genes on average, while the mean genome size of strains from Honduran tilapia was 1,801,133 bp, with an average of 1,869 genes, consistent with the small genomes associated with the host-adapted aquatic lineage (13) (Tables 1 and 2). Molecular serotyping indicated that all Queensland and Honduran aquatic strains belong to serotype Ib, but Queensland isolates belong to sequence type 261 (ST-261), while Honduran strains belong to ST-260. Both ST-260 and ST-261 have previously been identified as infecting aquatic animals. ST-260 isolates have been isolated from tilapia in Brazil and Costa Rica and occupy clonal complex 552 (CC552), along with ST-552 and ST-553 strains also isolated from tilapia in Latin America (27), ST-261 has been isolated from tilapia in the United States, China, Ghana, and Israel (9, 13, 30).

TABLE 1.

S. agalactiae isolates and sequences used in this study

Isolate Origina Yr Host Serotype STb Accession no. Assembly level
QMA0264 QLD, AU 2008 Epinephelus lanceolatus Ib 261 QGSK00000000 Contig
QMA0266 QLD, AU 2008 E. lanceolatus Ib 261 QGSJ00000000 Contig
QMA0267 QLD, AU 2008 E. lanceolatus Ib 261 QGSI00000000 Contig
QMA0268 QLD, AU 2009 Pomadasys kaakan Ib 261 QGSH00000000 Contig
QMA0271 QLD, AU 2009 Arius thalassinus Ib 261 CP029632 Complete
QMA0273 QLD, AU 2009 A. thalassinus Ib 261 QGSG00000000 Contig
QMA0274 QLD, AU 2009 Liza vaigiensis Ib 261 QGSF00000000 Contig
QMA0275 QLD, AU 2009 Aptychotrema rostrata Ib 261 QGSE00000000 Contig
QMA0276 QLD, AU 2009 Himantura granulata Ib 261 QGSD00000000 Contig
QMA0277 QLD, AU 2009 Dasyatis fluviorum Ib 261 QGSC00000000 Contig
QMA0280 QLD, AU 2010 E. lanceolatus Ib 261 QGSB00000000 Contig
QMA0281 QLD, AU 2010 E. lanceolatus Ib 261 QGSA00000000 Contig
QMA0284 QLD, AU 2010 E. lanceolatus Ib 261 QGRZ00000000 Contig
QMA0285 QLD, AU 2010 E. lanceolatus Ib 261 QGRY00000000 Contig
QMA0287 QLD, AU 2010 P. kaakan Ib 261 QGRX00000000 Contig
QMA0290 QLD, AU 2010 A. thalassinus Ib 261 QGRW00000000 Contig
QMA0292 QLD, AU 2010 A. rostrata Ib 261 QGRV00000000 Contig
QMA0294 QLD, AU 2010 E. lanceolatus Ib 261 QGRU00000000 Contig
QMA0320 QLD, AU 2010 Dasyatis fluviorum Ib 261 QGRT00000000 Contig
QMA0321 QLD, AU 2010 D. fluviorum Ib 261 QGRS00000000 Contig
QMA0323 QLD, AU 2010 D. fluviorum Ib 261 QGRR00000000 Contig
QMA0326 QLD, AU 2010 D. fluviorum Ib 261 QGRQ00000000 Contig
QMA0347 QLD, AU 2010 D. fluviorum Ib 261 QGRT00000000 Contig
QMA0368 QLD, AU 2010 E. lanceolatus Ib 261 QGRO00000000 Contig
QMA0369 QLD, AU 2011 E. lanceolatus Ib 261 QGRN00000000 Contig
QMA0485 Honduras 2014 Oreochromis niloticus Ib 260 QGRG00000000 Contig
QMA0487 Honduras 2014 O. niloticus Ib 260 QGRF00000000 Contig
QMA0488 Honduras 2014 O. niloticus Ib 260 QGRE00000000 Contig
QMA0489 Honduras 2014 O. niloticus Ib 260 QGRD00000000 Contig
QMA0494 Honduras 2014 O. niloticus Ib 260 QHHT00000000 Contig
QMA0495 Honduras 2014 O. niloticus Ib 260 QGRC00000000 Contig
QMA0496 Honduras 2014 O. niloticus Ib 260 QGRB00000000 Contig
QMA0497 Honduras 2014 O. niloticus Ib 260 QGRA00000000 Contig
QMA0499 Honduras 2014 O. niloticus Ib 260 QGQZ00000000 Contig
QMA0355 QLD, AU 2011 Homo sapiens Ia 23 QGRM00000000 Contig
QMA0357 QLD, AU 2011 H. sapiens Ia 23 QGRL00000000 Contig
QMA0336 NT, AU 2005 Crocodylus porosus Ia 23 QGRK00000000 Contig
QMA0300 QLD, AU 2008 Canis lapis familiaris V 1 QGRJ00000000 Contig
QMA0303 QLD, AU 2009 Felis catus V 1 QGRI00000000 Contig
QMA0306 QLD, AU 2005 Bos taurus V 1 QGRH00000000 Contig
GS16-0008 Ghana 2016 O. niloticus Ib 261 SRX2698682 Contig
GS16-0031 Ghana 2016 O. niloticus Ib 261 SRX2698681 Contig
GS16-0035 Ghana 2016 O. niloticus Ib 261 SRX2698680 Contig
GS16-0046 Ghana 2016 O. niloticus Ib 261 SRX2698679 Contig
ND2-22 Israel 1988 O. niloticus Ib 261 FO393392 Complete
138P USA 2007 O. niloticus Ib 261 CP007482 Complete
138spar USA 2011 O. niloticus Ib 261 CP007565.1 Complete
GX026 China 2011 O. niloticus Ib 261 CP011328.1 Complete
S13 Brazil 2015 O. niloticus Ib 552 CP018623.1 Complete
S25 Brazil 2015 O. niloticus Ib 552 CP015976.1 Complete
SA20 Brazil O. niloticus Ib 552c CP003919.2 Complete
GD201008-001 China 2010 O. niloticus Ia 7 NC_018646.1 Complete
HN016 China 2011 O. niloticus Ia 7 NZ_CP011325.1 Complete
WC1535 China 2015 O. niloticus Ia 7 NZ_CP016501.1 Complete
A909 USA H. sapiens Ia 7 NC_007432.1 Complete
GBS85147 Brazil H. sapiens Ia 103 NZ_CP010319.1 Complete
Sag37 China 2014 H. sapiens Ibd 12 NZ_CP019978.1 Complete
GBS1-NY USA 2012 H. sapiens II 22 NZ_CP007570.1 Complete
GBS2_NM USA 2012 H. sapiens II 22 NZ_CP007571.1 Complete
GBS6 USA 2009 H. sapiens II 22 NZ_CP007572.1 Complete
FDAARGOS_254 USA 2014 H. sapiens IId 22 CP020449.1 Complete
COH1 USA H. sapiens III 17 NZ_HG939456.1 Complete
NEM316 France H. sapiens III 23 NC_004368.1 Complete
CU_GBS_08 Hong Kong 2008 H. sapiens III 283 NZ_CP010874.1 Complete
CU_GBS_98 Hong Kong 1998 H. sapiens III 283 NZ_CP010875.1 Complete
NGBS128 Canada 2010 H. sapiens III 17 NZ_CP012480.1 Complete
SG-M1 Singapore 2015 H. sapiens III 283 NZ_CP012419.2 Complete
H002 China 2011 H. sapiens III 736 NZ_CP011329.1 Complete
Sag158 China 2014 H. sapiens IIId 19 NZ_CP019979.1 Complete
NGBS061 Canada 2010 H. sapiens IV 459 NZ_CP007631.1 Complete
NGBS572 Canada 2012 H. sapiens IV 452 NZ_CP007632.1 Complete
2603V/R USA H. sapiens V 110 NC_004116.1 Complete
CNCTC10/84 USA H. sapiens V 26 NZ_CP006910.1 Complete
NGBS357 Canada 2011 H. sapiens V 1 NZ_CP012503.1 Complete
SS1 USA 1992 H. sapiens V 1 NZ_CP010867.1 Complete
GBS-M002 China 2014 H. sapiens VId 1 NZ_CP013908.1 Complete
SA111 Portugal 2013 H. sapiens IId 61e NZ_LT545678.1 Complete
FWL1402 China 2014 Hoplobatrachus chinensis IIId 739e NZ_CP016391.1 Complete
09mas018883 Sweden B. taurus Vd 1 NC_021485.1 Complete
GBS ST-1 USA 2015 C. lapis familiaris V 1 NZ_CP013202.1 Complete
ILRI005 Kenya Camelus dromedarius Vd 609 NC_021486.1 Complete
ILRI112 Kenya C. dromedarius VId 617 HF952106.1 Complete
a

QLD, Queensland; AU, Australia; NT, Northern Territory.

b

ST, sequence type.

c

Gap in glcK.

d

Serotype was detected with Kaptive.

e

ST was determined via the Center for Genomic Epidemiology.

TABLE 2.

Genome assembly statistics

Strain BioSample no. Sequence yield (bp) No. of contigs Genome size (bp) GC content (%) N50 (bp) Coverage (×)
QMA0264 SAMN07998566 383,903,016 26 1,800,255 35.27 181,173 213
QMA0266 SAMN07998567 383,903,016 23 1,799,604 35.26 181,173 213
QMA0267 SAMN07998568 822,647,364 22 1,805,121 35.33 181,785 456
QMA0268 SAMN07998569 541,821,504 22 1,796,834 35.25 130,961 302
QMA0271 SAMN07998570 488,654,460 1 1,802,470 35.28 1,802,470 271
QMA0273 SAMN07998571 488,596,752 26 1,799,223 35.26 181,173 272
QMA0274 SAMN07998572 666,932,868 33 1,795,949 35.23 91,064 371
QMA0275 SAMN07998573 444,271,968 28 1,798,598 35.26 129,867 247
QMA0276 SAMN07998574 595,443,828 30 1,801,706 35.28 101,410 330
QMA0277 SAMN07998575 532,021,392 30 1,794,975 35.23 125,723 296
QMA0280 SAMN07998576 507,086,244 26 1,798,991 35.26 130,981 282
QMA0281 SAMN07998577 947,517,396 28 1,795,694 35.26 130,961 528
QMA0284 SAMN07998578 870,597,252 27 1,804,787 35.3 129,871 482
QMA0285 SAMN07998579 946,715,952 25 1,798,637 35.26 181,149 526
QMA0287 SAMN07998580 744,380,280 27 1,799,347 35.27 130,969 414
QMA0290 SAMN07998581 1,059,786,168 24 1,804,000 35.3 181,153 587
QMA0292 SAMN07998582 383,835,312 22 1,800,208 35.26 181,172 213
QMA0294 SAMN07998583 807,742,572 24 1,800,272 35.26 181,172 449
QMA0300 SAMN07998584 703,661,784 33 2,109,161 35.25 115,466 334
QMA0303 SAMN07998585 949,167,156 42 2,123,226 35.05 94,602 447
QMA0306 SAMN07998586 745,047,660 106 2,165,975 35.52 40,135 344
QMA0320 SAMN07998587 445,338,936 30 1,804,060 35.3 181,347 247
QMA0321 SAMN07998588 406,041,552 22 1,800,073 35.26 181,173 226
QMA0323 SAMN07998589 487,351,368 22 1,800,031 35.26 181,173 271
QMA0326 SAMN07998590 847,046,508 22 1,800,125 35.26 181,173 471
QMA0336 SAMN07998591 437,035,536 37 2,022,886 35.26 87,595 216
QMA0347 SAMN07998592 942,292,680 21 1,795,434 35.23 181,173 525
QMA0355 SAMN07998593 903,514,248 66 1,996,200 35.31 63,524 453
QMA0357 SAMN07998594 564,675,720 35 2,004,249 35.23 92,563 282
QMA0368 SAMN07998595 506,471,448 26 1,803,560 35.3 104,282 281
QMA0369 SAMN07998596 1,056,496,560 28 1,804,404 35.3 130,985 586
QMA0485 SAMN07998597 490,192,752 28 1,803,278 35.27 109,741 272
QMA0487 SAMN07998598 467,765,088 28 1,800,682 35.25 109,741 260
QMA0488 SAMN07998599 356,791,512 29 1,800,616 35.25 109,741 198
QMA0489 SAMN07998600 432,449,472 28 1,801,015 35.25 109,741 240
QMA0494 SAMN07998601 363,434,736 37 1,792,992 35.28 71,935 203
QMA0495 SAMN07998602 412,230,672 28 1,803,007 35.27 109,741 229
QMA0496 SAMN07998603 499,707,096 29 1,801,879 35.27 109,737 277
QMA0497 SAMN07998604 364,009,464 28 1,800,976 35.25 109,741 202
QMA0499 SAMN07998605 428,636,208 28 1,799,426 35.24 109,741 238

The aquatic isolates analyzed herein extend the known host range of GBS ST-261 to rays and marine finfish and expand the environmental distribution from freshwater to marine habitats, in addition to those previously reported (5, 6, 42). GBS strains isolated from humans and terrestrial animals in Queensland and Northern Territory, Australia, were also sequenced and have larger genomes of 2,072,596 bp, comprising 2,067 genes on average, suggesting that recent possible local transfer from terrestrial origin to Australian fish is highly unlikely, although probable transfer between humans and fish has been reported for ST-7 GBS elsewhere (8, 39, 43). Nonhuman mammalian strains from Australia sequenced in this study, QMA0300 and QMA0303, belong to ST-1 serotype V, and QMA0306 belongs to ST-67 serotype III, whereas human isolates QMA0355 and QMA0357 and crocodile strain QMA0336 belong to ST-23 serotype Ia. Indeed, the high sequence identity between the human ST-23 serotype Ia isolates and those from farm-raised crocodiles supports the idea of probable human transfer to these animals, as previously implied (44).

To identify the possible origin of the ST-261 isolates from marine fish in Australia, a phylogenetic tree was constructed by maximum likelihood from 29,689 nonrecombinant core-genome SNPs and short indel positions derived by the alignment of whole-genome sequences of 25 Queensland fish isolates, 9 Honduran isolates, 6 Queensland terrestrial isolates, and 42 genomes from public databases (Fig. 1A). Two distinct groups were resolved, the first entirely composed of aquatic isolates (including serotype Ib isolates from ST-552, ST-260, and ST-261), while the second major group comprised various isolates of terrestrial origin and some fish isolates from ST-7 serotype Ia that may have infected fish via transfer from terrestrial sources (Fig. 1A). The serotype Ib aquatic group branched into three distinct lineages based on ST (Fig. 1A). One lineage comprised all Honduran strains of ST-260, which were derived from a lineage comprising ST-552 isolates from Brazil (Fig. 1A). This is supportive of previous observations in which an extended typing system based on MLST, virulence genes, and serotype indicated geographic endemism within fish isolates from differing regions of Brazil (27). The isolates belonging to ST-261 from the United States, Israel, Ghana, China, and Queensland clustered together (Fig. 1A). The second major division, containing serotype Ia fish strains along with terrestrial isolates, was more complex, but the isolates largely clustered in line with serotype and ST (Fig. 1A). The aquatic serotype Ia isolates clustered with human isolates of ST-7 (Fig. 1A). These fish isolates have acquired a 10-kb genomic island, putatively from S. anginosus, in contrast to their human ST7 serotype Ia relatives (39). Lineage 10 comprised serotype III ST-17 human isolates (Fig. 1A). Serotype Ia strains were divided among several additional ST groups, where three Australian serotype Ia ST-23 strains (QMA0336, QMA0355, and QMA0357) isolated from humans and a saltwater crocodile clustered together with strains of serotype III ST-23 (NEM316) and serotype IV ST-452 (NGBS572), also of human origin (Fig. 1A). This lineage appears to be derived from NEM316, which is a frequent cause of late-onset disease in human infants (45). A further serotype Ia isolate was located in ST-103 and clustered together with two strains of serotype V from ST-609 and ST-617 isolated from camels (46) (Fig. 1A). Two newly sequenced Australian terrestrial strains, QMA0300 isolated from a dog and QMA0303 isolated from a cat, clustered with other serotype V ST-1 strains isolated from human, cattle, and dog hosts (Fig. 1A). This lineage also contained NGBS061 serotype IV ST-495 and GBS-M002 serotype VI ST-1 (Fig. 1A). ST-1 emerged as a significant cause of infection and disease in humans during the 1990s but was recently inferred to have evolved from strains causing mastitis in cattle in the 1970s (47). Moreover, QMA0306 ST-1 serotype V from cattle in Queensland was closely related to SA111 ST-61 serotype II, which represents a host-adapted lineage of S. agalactiae that is dominant in cattle in Europe (48) (Fig. 1A).

FIG 1.

FIG 1

(A) Maximum likelihood phylogeny of 82 S. agalactiae strains. The tree was inferred from alignment of 6,050 nonrecombinant core-genome SNPs. Branch length was adjusted for ascertainment bias using Felsenstein's correction implemented in RAxML (76). Nodes are supported by 1,000 bootstrap replicates. The tree was rooted using S. pyogenes M1 GAS (RefSeq accession number NC_002737.2) as an outgroup. (B) Minimum spanning tree showing relationship among ST-261 serotype Ib GBS isolates based on a distance matrix derived from nonrecombinant core-genome SNP alignment. Edge labels indicate the number of nonrecombinant core-genome SNPs between adjacent strains. The consensus network was computed in MSTgold (78).

Our phylogeny based on whole-genome SNPs does not support a recent direct transfer of GBS from Australian human clinical or terrestrial animal sources to marine fish and stingrays, in spite of close proximity of many of the wild fish cases to human habitation (6). Consequently, we refined our analyses to the ST-261 aquatic host-adapted lineage to attempt to infer a possible route of introduction and subsequent evolution in Australian marine fish. A consensus minimum spanning tree based on a distance matrix composed of all core-genome SNPs derived from alignment of the ST-261 serotype Ib strains revealed a likely original introduction via tilapia from Israel, with only 63 core-genome SNPs separating an Australian stingray isolate from the type strain of Streptococcus difficile (reassigned as S. agalactiae serotype Ib [49]), isolated from tilapia in Israel in 1988 (Fig. 1B) (50). Tilapia were imported on a number of occasions during the 1970s and 1980s from Israel into North Queensland around Cairns and Townsville, and a number of strains and hybrids have since colonized rivers and creeks throughout Queensland (51). Globally, the aquatic ST-261 lineage appears to have been transferred through human movements of tilapia for aquaculture and other purposes over the last several decades. The U.S. and Chinese tilapia isolates also appear to derive from the early ND2-22 isolate, as do recent isolates from tilapia in Ghana (Fig. 1B). Indeed, phylogenetic analysis by maximum likelihood of draft whole-genome alignments suggests that the Ghanaian and Chinese isolates share a recent common ancestor that is derived from ND2-22, with only 60 SNPs separating ND2-22 and the Ghanaian strains (9). We identified only 36 core-genome SNPs separating the Ghanaian isolates from ND2-22, but this reflects the smaller core genome in our study as a result of the high number of GBS isolates analyzed (40 isolates in the present study compared with 9 isolates in the previous study) (9).

The minimum spanning tree implicates continued adaptation of the ST-261 lineage postintroduction and suggests that grouper (marine Teleostei family) may have been infected via estuary stingrays (Dasyatidae family) (Fig. 1B). Stingrays are occasional prey for adult giant Queensland grouper, and stingray barbs have been found in the gut of grouper postmortem (R. O. Bowater, unpublished data). ST-261 GBS has also caused mortality in captive stingrays in South East Queensland, translocated from Cairns in North Queensland (42).

The S. agalactiae pangenome comprises a small core of protein-coding genes and is open.

To further elucidate adaptation among the fish-pathogenic GBS types, a pangenome was built from 39 complete genomes retrieved from GenBank and using our curated ST-261 grouper isolate QMA0271 as a high-quality reference seed genome. The resulting pangenome was 4,074,275 bp (Fig. 2A). All-versus-all BLAST analysis of the genomes implemented in BRIG clearly indicated the substantial reduction in genome size among the fish-pathogenic ST-261 cohort, as previously reported for a limited number of ST-261 isolates (13). Here, we find that ST-260 and ST-552 fish-pathogenic sequence types within serotype Ib are similarly reduced and that conservation among the serotype Ib strains is high (Fig. 2A). In total, 4,603 protein-coding genes were predicted in the S. agalactiae pangenome using Roary (Fig. 2B), which is consistent with previous research (39). The number of core genes was 1,440 (representing 35% of the pangenome), while previous studies reported 1,202 to 1,267 genes in the pangenome (39, 40). These differences may result from the methods being used to examine the pangenome, the difference in number of sequences being used to create the pangenome, and finally, the use of draft sequences in the analysis (39). A majority of protein-coding genes found in the pangenome belonged to both the dispensable and strain-specific genes. This could be a result of the inclusion of a high number of serotype Ib strain sequences, which were all significantly smaller (approximately 1.8 Mbp) than those of other isolates. Liu et al. (39) demonstrated that removing Ib piscine isolates from their analysis resulted in an increase in the number of core genes.

FIG 2.

FIG 2

The pangenome of S. agalactiae. (A) BLASTN-based sequence comparison of 82 S. agalactiae genomes against the S. agalactiae pangenome as reference constructed with BRIG 0.95 (80). Rings from the innermost to the outermost show GC content and GC skew of the pangenome reference, and then sequence similarities of each of the 82 strains listed in the legend; from top to bottom, rings are colored according to origin, with fish isolates belonging to serotype Ib ST-261, ST-260, and ST-552 in blue, fish strains belonging to serotype Ia in purple, and terrestrial strains in red. The outermost ring (black) represents the reference pangenome. (B) Proportion of protein-coding genes in the core, soft core, shell, and cloud of the pangenome of 82 S. agalactiae isolates determined with Roary. (C) Histogram indicates the frequency of genes (protein-coding) in red and IGRs (non-protein-coding intergenic regions) in blue-green from 82 S. agalactiae genomes analyzed by Piggy (52).

Frequency analysis of IGRs in S. agalactiae pangenome showed that the number of IGRs shared across all strains was smaller than core protein-coding genes, whereas strain-specific IGRs were much higher than protein-coding strain-specific genes (Fig. 2C). IGR analysis with Piggy excludes IGRs that are less than 30 bp, which may result in fewer IGRs than protein-coding genes in core regions (52). Most IGRs identified in the pangenome belonged to either core genes or strain-specific genes (Fig. 2C), in line with previous findings in Staphylococcus aureus ST-22 and Escherichia coli ST-131, where similar distributions of IGRs were detected (52). The gradients of the accumulation curves for both protein-coding genes and IGRs are still strongly positive; therefore, the S. agalactiae pangenome remains open.

To determine whether strain-specific IGR sequences may be derived from genes (or pseudogenes) that have undergone erosion, we queried the gene sequences used to construct the pangenome by BLASTn using 1,539 strain-specific representative IGR sequence clusters (output from Piggy). A total of 1,862 hits were retrieved from 683/1,539 representative IGR sequence clusters (see Table S2 in the supplemental material) supportive of the gene origins of many IGRs. Overrepresentation of sequences from the QMA0271 reference genome among the hits (309/1,862 from 39 genomes employed in the pangenome) was expected, as it was employed as the seed for the pangenome assembly. Of the 241 unique hits among genes from QMA0271, 39 hits fell in genes that we annotated as pseudogenes (Table S2). Pseudogenes are predicted to be rapidly lost from the genome, evidenced by the lack of conserved pseudogenes across multiple strains of the same species (53, 54). The relatively high presence of pseudogenes or remnants thereof within the serotype Ib ST-261 QMA0271 is supportive of ongoing adaptation of this lineage to the aquatic host and habitat (13). As pseudogenes are often not annotated as such in database assemblies (54, 55), further manual inspection of the 849 unique BLAST hits was performed in Excel and by reference to the pangenome. Pseudogenes may arise through frameshift SNPs, resulting in early termination or interruption by insertion elements (55), and 239 of the 849 unique BLAST hits were identified as insertion element (IS) transposases and a further four as or phage/prophage proteins, suggesting possible gene disruptions by insertion (Table S2). Hypothetical proteins comprised 215 of the 849 unique sequences identified by BLAST of strain-specific IGR through the pangenome, while genes annotated as transcriptional regulators comprised 48 sequences (Table S2). These results lend preliminary support to the hypothesis that much of the IGR of bacteria comprises gene remnants and is worthy of in-depth exploration.

Aquatic serotype Ib isolates have a reduced repertoire of virulence factors.

Almost all genes classed as adhesins were involved in immune evasion and host invasion, and most of the toxin-related genes found in terrestrial isolates were absent from serotype Ib aquatic isolates (Fig. 3). Rosinski-Chupin et al. (13) reported that ∼60% of the virulence genes found in human strains were present in a serotype Ib GBS strain from fish. We found that the CAMP factor gene cfa-cfb was present in all strains, including serotype Ib ST-260, ST-261, and ST-552 isolates, but the CAMP reaction was previously reported to be negative for ST-260 and ST-261 strains (13, 28). These authors identified that the CAMP factor gene in ST-261 is disrupted, while the gene in ST-260 is unaltered, but the level of gene expression may be too low to be detected by the test (13). Most of the genes in the cyl locus have been lost from serotype Ib strains, and only cylB, an ABC ATP binding cassette transporter, was present in ST-260 and ST-552 (Fig. 3). The cyl locus is responsible for hemolytic activity and the production of pigment via cotranscription of cylF and cylL (56). In ST-261, the cyl gene cluster is replaced by a genomic island, resulting in a loss of hemolytic activity (13). Of particular relevance to virulence and antigenic diversity, the transmembrane immunoglobulin A-binding C protein beta-antigen gene cba was absent in all aquatic isolates (Fig. 3). This gene has been reported in type Ib and Ia GBS strains previously (57), is implicated in virulence in neonates, and is upregulated in the GBS serotype Ia strain A909 in response to human serum (5759). In contrast, capsule-related genes were largely conserved and associated with serotype (Fig. 3), supporting the major role of capsule in virulence of fish-pathogenic streptococci (60, 61).

FIG 3.

FIG 3

Virulence gene presence and absence in S. agalactiae. Genes were identified from VFDB to create an S. agalactiae database for comparison of 82 strains by BLAST using SeqFindr, with default settings (95% identity cutoff).

Although serotype Ib aquatic isolates have lost the majority of the virulence factors found in terrestrial isolates, most contained six sequences that were identified as probable adhesins by homology (Fig. 3). These ST-261 adhesins (named 0337, 0626, 0856, 1185, 1196, and 1648, based on positions in the annotated ST-261 genome from QMA0271) were fully conserved among aquatic serotype Ib ST-261 isolates (Fig. 3). Moreover, ST-261 adhesin 0856 was only present in ST-261 and two serotype II isolates, GBS1-NY and SA111 (Fig. 3). ST-261 adhesins 1185 and 1648 were shown to be unique to aquatic serotype Ib isolates, being absent from Ia fish isolates and all terrestrial strains (Fig. 3). The analysis indicated that ST-261 adhesins 0337, 0626, and 1196 were well conserved across most of the isolates, regardless of origin, but 1196 was the only adhesin-like gene present in all strains analyzed (Fig. 3). ST-261 adhesin 0626 was also ubiquitous but contained deletions in the Sag37 isolate (Fig. 3). Some terrestrial strains, such as QMA0355, QMA0357, QMA0336, and NGBS572, lacked adhesin 0337 (Fig. 3).

As capsular polysaccharide is a requirement for full virulence in several fish-pathogenic streptococci (60, 61) and is the major antigen in GBS (6264), further analysis of the serotype Ib cps operon was conducted. Within the serotype Ib lineage, the capsular polysaccharide operon was well conserved (Fig. 4). However, QMA0281 from grouper had a deletion at position 341 in cpsB, encompassing cpsC, cpsD, and cpsE (Fig. 4), resulting in a chimeric open reading frame (ORF). Moreover, an insertion in a TA repeat at position 557 in cpsH resulted in an early stop codon, marginally reducing gene size (Fig. 4). QMA0368 had a TT insertion at nucleotide position 745 in cpsE, resulting in the insertion of an early stop codon and truncation of the gene (Fig. 4). Deletion of this region that includes the priming glycosyltransferase for capsular biosynthesis results in a loss of capsule, attenuated virulence, and modified pathology in the fish pathogen Streptococcus iniae (65). Buoyant density analysis in Percoll indicated that QMA0281 is also deficient in capsule (not shown). cps operon SNPs among the Australian serotype Ib isolates were rare (Fig. 4). Indeed, only 1 SNP in neuA separated QMA0271 from the 1988 tilapia isolate from Israel, suggesting very little evolution of the cps operon since the introduction of the lineage (Fig. 4). This may reflect a well-adapted capsule for colonization-naive hosts, as the Australian isolates were from wild fish and captive stingrays recently after capture and transport, thus placing little selective pressure for novel capsular sequence types. Immunity in fish drives the evolution of novel capsular sequence types in S. iniae, but these reported cases were all in high-intensity aquaculture with occasional use of autogenous vaccination and an opportunity for the development of cps-specific immunity (65). This may explain the relatively high number of SNPs in genes encoding sugar- and sialic acid-modifying enzymes among the isolates from tilapia farmed in Honduras relative to the other isolates examined, as autogenous vaccination is occasionally used on farms where isolates were sourced and may apply selective pressure favoring modified polysaccharide capsule.

FIG 4.

FIG 4

The capsular polysaccharide (cps) operon of the ST-261 lineage. The operon was identified in GenBank files manually and then compared by BLAST using Easyfig (91).

Conclusions.

The clade of aquatic S. agalactiae serotype Ib, including ST-260, ST-261, and ST-552, is a highly adapted fish pathogen with a substantially reduced genome compared to all other serotypes from terrestrial mammalian, reptile, and fish hosts. These variants were originally identified as S. difficile due to their impoverished growth on laboratory media and hence difficulty of isolation from diseased fish (50). Isolates from fish that were previously identified as S. difficile were assigned to serotype Ib GBS a few years later (49), but the recent discovery that serotype Ib isolates have substantially smaller genomes than those of other GBS serotypes (13) explains the marked phenotypic difference that merited early phenotypic assignation of these strains to a separate species. Other serotypes have been isolated from fish, notably serotype Ia and serotype III, but these seem to arise from terrestrial transfer rather than being retained among the aquatic host population (28). In contrast, serotype Ib has only been isolated from fish and stingrays, appears to be well adapted, and is likely to have been transferred internationally via trade in domesticated tilapia, evidenced here by the very close relationship (only a few SNPs) between a strain isolated from tilapia in 1988 in Israel and those found in fish and stingrays in Australia since 2008 and in tilapia in the United States, China, and Ghana. The ST-261 lineage in Australia likely arrived with several introductions of tilapia in the 1970s and 1980s. Tilapia are classed as a noxious pest in Australia, and their import was banned, but not before several lines became established throughout tropical and subtropical freshwater habitats in Queensland (51). Although ST-261 serotype Ib GBS has not been isolated from farm fish in Queensland in spite of the proximity of freshwater farms to tilapia-infested creeks, this clade of GBS is a substantial problem in farmed tilapia globally. The cohort of putative adhesins identified here to be conserved throughout all fish-pathogenic serotypes (including Ia and III, in addition to Ib) may be promising candidates for cost-effective cross-serotype protective vaccines for aquaculture and are worthy of future research.

MATERIALS AND METHODS

Bacterial strains and culture conditions.

Forty S. agalactiae isolates comprising strains collected from fish in Honduras and Australia, along with reptiles, humans, and other terrestrial mammals from Australia, were chosen for sequencing (Table 1). Of these 40 isolates, 25 strains were collected from several species of fish in Queensland, Australia, two human clinical strains were from Queensland, Australia, one strain was isolated from saltwater crocodile (Crocodylus porosus) in the Northern Territory, Australia, and three isolates were collected from domestic animals, including cats, dogs, and cattle in Australia. Additionally, nine isolates originating from disease in farmed tilapia in Honduras were sequenced during this study. All isolates were maintained at −80°C in Todd-Hewitt broth (THB; Oxoid) containing 25% glycerol as frozen stock. The isolates were recovered from stock on Columbia agar containing 5% sheep blood (Oxoid) for 24 h at 28°C. For liquid culture, the isolates were grown in THB for 18 h with low agitation at 28°C.

DNA extraction and sequencing.

Genomic DNA (gDNA) was extracted from 10-ml early stationary-phase THB cultures with the DNeasy minikit (Qiagen), according to the manufacturer's instructions. The quantity of extracted DNA was measured by Qubit fluorimetry (Invitrogen), and the quality was checked by agarose gel electrophoresis. To confirm the purity of the gDNA, the 16S rRNA gene was amplified by PCR using universal primers 27F and 1492R (66), and the PCR products were sent to Australian Genome Research Facility (AGRF, Brisbane, Australia) for Sanger sequencing. The 16S amplicon sequences were assembled in Sequencher version 5.2.2 and analyzed by BLAST. Once identity and purity were confirmed, Nextera XT paired-end libraries were generated using gDNA from each isolate and sequenced on the Illumina HiSeq 2000 platform system (AGRF, Melbourne, Australia).

De novo assembly and annotation.

Illumina sequencing yielded between 5,288,952 and 12,577,340 read pairs for each strain. Read quality control and contaminant screening were performed using FastQC (67). Reads were trimmed using the clip function in Nesoni (https://github.com/Victorian-Bioinformatics-Consortium/nesoni) and then assembled de novo with SPAdes assembler version 3.11 (68). Assemblies were quality checked using Quast version 4.6 (69). The assemblies of fish isolates in Queensland comprised about 1.8 Mbp of assembled sequence, while terrestrial strains comprised 2 Mbp. The assembled contigs for all Queensland strains were reordered against an internal curated reference genome from S. agalactiae strain QMA0271, using the Mauve contig ordering tool (70). Automated annotation was performed using Prokka 1.12 (71).

Molecular serotyping and MLST.

Reference sequences for the nine CPS serotypes (Table 3) (21) were retrieved from GenBank to generate a database for the prediction of capsular serotype from the draft genomes with Kaptive, using default settings (72). To determine multilocus sequence types (MLST), all draft assemblies were analyzed using the Center for Genomic Epidemiology web-tools MLST version 1.8, using an S. agalactiae configuration (see https://cge.cbs.dtu.dk//services/MLST/) (73).

TABLE 3.

Reference sequences used for molecular capsule serotyping

Serotype Accession no. Size (bp) Reference
Ia AB028896.2 25,021 85
Ib AB050723.1 9,987 86
II EF990365.1 12,864 87
III AF163833.1 17,276 88
IV AF355776.1 17,596 89
V AF349539.1 18,239 89
VI AF337958.1 16,448 89
VII AY376403.1 14,202 90
VIII AY375363.1 12,637 90

Phylogenetic analysis.

To estimate approximate phylogenetic relationships among our strains and other isolates with whole-genome sequences available in GenBank (Table 1), a core-genome single nucleotide polymorphism (SNP)-based phylogenetic tree was constructed. Whole-genome sequences of Queensland and Honduras tilapia isolates, terrestrial isolates, and the genomes obtained from GenBank were aligned with Parsnp in the Harvest Tools suite version 1.2 (74). The genome of Streptococcus pyogenes M1 (group A Streptococcus [GAS]) (RefSeq accession number NC_002737.2) was also included as an outgroup for tree rooting. Hypothetical recombination sites in the core-genome alignment were detected and filtered out with Gubbins (75). Maximum likelihood phylogenetic trees were inferred with RAxML version 8.2.9 (76) based on nonrecombinant core-genome SNPs under a general time-reversible (GTR) nucleotide substitution model, with 1,000 bootstrap replicates. Ascertainment bias associated with analysis of only variable sites was accounted for using Felsenstein's correction implemented in RAxML (76). The resulting tree was exported and rooted, nodes with low bootstrap support were collapsed with Dendroscope, and the resulting tree/cladogram was annotated with Evolview version 2 (77).

In order to infer possible phylogenetic relationships within the aquatic ST-261 clade of GBS, minimum spanning trees were constructed in MSTgold 2.4 (78) from a distance matrix based on core-genome SNPs derived by the alignment of 27 ST-261 genomes in Geneious version 9.1 (Biomatters, Inc.). A consensus tree was constructed based on inference of 2,400 trees, and only those edges occurring in greater than 50% of trees were included in the consensus.

Pangenome analysis.

A reference pangenome was built with the GView server (79) using our curated genome of the Queensland ST-261 serotype Ib grouper isolate QMA0271 as a seed and 38 complete genomes from public databases added sequentially (Table 1). Only complete genomes were included into the reference pangenome to avoid incomplete genes associated with the high fragmentation of draft sequences. For visualization, draft and complete genomes were compared with the reference pangenome using BRIG 0.95 (80). To investigate core and accessory protein-coding genes, sequences from all strains were analyzed with Roary (81) using default settings, and gene presence-absence tables are provided in Table S1. Since Roary only considers protein-coding sequences, we used Piggy (52) to identify non-protein-coding intergenic regions (IGRs) in each strain, which comprise about 15% of the GBS genomes. The 1,539 strain-specific IGR sequences identified in Piggy were then extracted from the representative merged clusters of IGR sequences and used as query sequences in a local BLASTn analysis (E value 1e−10) against all the coding sequences in the pangenome. The output from the BLAST search was tabulated and analyzed further in Microsoft Excel (Table S2).

Identification and comparison of virulence factors.

Virulence factor screening was performed using SeqFindR (82) by comparing the assembled genomes of all strains in this study along with the genomes available through GenBank (Table 1) against a list of 51 S. agalactiae-specific virulence factor sequences collected from the Virulence Factors Database (VFDB) (83), complemented by six additional sequences identified in the sequenced ST-261 S. agalactiae strain ND2-22 (Table 1).

Analysis of effects of SNPs in ST-261 clade.

Putative effects of SNPs among and between the ST-261 clade were determined using SnpEff version 4.3p (84). First, a new S. agalactiae database was constructed from the GenBank-formatted curated reference genome for QMA0271, in accordance with the manual (http://snpeff.sourceforge.net/SnpEff_manual.html#databases). Then, a VCF file generated from curated SNPs generated by an alignment of the complete genomes of 27 ST-261 isolates in Geneious version 9.1 was annotated for SNP effect with SnpEff using the database from QMA0271 as a reference (see the supplemental material).

Accession number(s).

The accession numbers for genome assemblies used in this study are presented with strain metadata in Table 1. The assembly statistics and BioSample numbers are in Table 2.

Supplementary Material

Supplemental material

ACKNOWLEDGMENTS

This work was funded by the Fisheries Research and Development Corporation Aquatic Animal Health Subprogram Project 2010/34 awarded to R.O.B. and A.C.B., and the Australian Research Council Linkage Program LP130100242 with cosponsorship by Elanco Australia awarded to A.C.B., M.J.W., and S.B.

Footnotes

Supplemental material for this article may be found at https://doi.org/10.1128/AEM.00859-18.

REFERENCES

  • 1.Brochet M, Couvé E, Zouine M, Vallaeys T, Rusniok C, Lamy M-C, Buchrieser C, Trieu-Cuot P, Kunst F, Poyart C, Glaser P. 2006. Genomic diversity and evolution within the species Streptococcus agalactiae. Microbes Infect 8:1227–1243. doi: 10.1016/j.micinf.2005.11.010. [DOI] [PubMed] [Google Scholar]
  • 2.Edwards MS, Gonik B. 2013. Preventing the broad spectrum of perinatal morbidity and mortality through group B streptococcal vaccination. Vaccine 31:D66–D71. doi: 10.1016/j.vaccine.2012.11.046. [DOI] [PubMed] [Google Scholar]
  • 3.Six A, Bellais S, Bouaboud A, Fouet A, Gabriel C, Tazi A, Dramsi S, Trieu-Cuot P, Poyart C. 2015. Srr2, a multifaceted adhesin expressed by ST-17 hypervirulent group B Streptococcus involved in binding to both fibrinogen and plasminogen. Mol Microbiol 97:1209–1222. doi: 10.1111/mmi.13097. [DOI] [PubMed] [Google Scholar]
  • 4.Doumith M, Mushtaq S, Martin V, Chaudhry A, Adkin R, Coelho J, Chalker V, MacGowan A, Woodford N, Livermore DM, BSAC Resistance Surveillance Standing Committee. 2017. Genomic sequences of Streptococcus agalactiae with high-level gentamicin resistance, collected in the BSAC bacteraemia surveillance. J Antimicrob Chemother 72:2704–2707. doi: 10.1093/jac/dkx207. [DOI] [PubMed] [Google Scholar]
  • 5.Delamare-Deboutteville J, Bowater R, Condon K, Reynolds A, Fisk A, Aviles F, Barnes AC. 2015. Infection and pathology in Queensland grouper, Epinephelus lanceolatus, (Bloch), caused by exposure to Streptococcus agalactiae via different routes. J Fish Dis 38:1021–1035. doi: 10.1111/jfd.12273. [DOI] [PubMed] [Google Scholar]
  • 6.Bowater RO, Forbes-Faulkner J, Anderson IG, Condon K, Robinson B, Kong F, Gilbert GL, Reynolds A, Hyland S, McPherson G, Brien JO, Blyde D. 2012. Natural outbreak of Streptococcus agalactiae (GBS) infection in wild giant Queensland grouper, Epinephelus lanceolatus (Bloch), and other wild fish in northern Queensland, Australia. J Fish Dis 35:173–186. doi: 10.1111/j.1365-2761.2011.01332.x. [DOI] [PubMed] [Google Scholar]
  • 7.Elliott JA, Facklam RR, Richter CB. 1990. Whole-cell protein patterns of nonhemolytic group B, type Ib, streptococci isolated from humans, mice, cattle, frogs, and fish. J Clin Microbiol 28:628–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Evans JJ, Klesius PH, Gilbert PM, Shoemaker CA, Al Sarawi MA, Landsberg J, Duremdez R, Al Marzouk A, Al Zenki S. 2002. Characterization of β-haemolytic group B Streptococcus agalactiae in cultured seabream, Sparus auratus L., and wild mullet, Liza klunzingeri (Day), in Kuwait. J Fish Dis 25:505–513. doi: 10.1046/j.1365-2761.2002.00392.x. [DOI] [Google Scholar]
  • 9.Verner-Jeffreys D, Wallis T, Cano Cejas I, Ryder D, Haydon D, Domazoro JF, Dontwi J, Field T, Adjei-Boteng D, Wood G, Bean T, Feist SW. 2017. Streptococcus agalactiae multilocus sequence type 261 is associated with mortalities in the emerging Ghanaian tilapia industry. J Fish Dis 41:175–179. doi: 10.1111/jfd.12681. [DOI] [PubMed] [Google Scholar]
  • 10.Li K, Liu L, Clausen JH, Lu M, Dalsgaard A. 2016. Management measures to control diseases reported by tilapia (Oreochromis spp.) and whiteleg shrimp (Litopenaeus vannamei) farmers in Guangdong, China. Aquaculture 457:91–99. doi: 10.1016/j.aquaculture.2016.02.008. [DOI] [Google Scholar]
  • 11.Chen M, Li L-P, Wang R, Liang W-W, Huang Y, Li J, Lei A-Y, Huang W-Y, Gan X. 2012. PCR detection and PFGE genotype analyses of streptococcal clinical isolates from tilapia in China. Vet Microbiol 159:526–530. doi: 10.1016/j.vetmic.2012.04.035. [DOI] [PubMed] [Google Scholar]
  • 12.Zhang D, Li A, Guo Y, Zhang Q, Chen X, Gong X. 2013. Molecular characterization of Streptococcus agalactiae in diseased farmed tilapia in China. Aquaculture 412:64–69. doi: 10.1016/j.aquaculture.2013.07.014. [DOI] [Google Scholar]
  • 13.Rosinski-Chupin I, Sauvage E, Mairey B, Mangenot S, Ma L, Da Cunha V, Rusniok C, Bouchier C, Barbe V, Glaser P. 2013. Reductive evolution in Streptococcus agalactiae and the emergence of a host adapted lineage. BMC Genomics 14:252. doi: 10.1186/1471-2164-14-252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Brynildsrud O, Feil EJ, Bohlin J, Castillo-Ramirez S, Colquhoun D, McCarthy U, Matejusova IM, Rhodes LD, Wiens GD, Verner-Jeffreys DW. 2014. Microevolution of Renibacterium salmoninarum: evidence for intercontinental dissemination associated with fish movements. ISME J 8:746–756. doi: 10.1038/ismej.2013.186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Barnes AC, Delamare-Deboutteville J, Brosnahan C, Gudkovs N, Morrison RN, Carson J. 2016. Whole genome analysis of Yersinia ruckeri isolated over 27 years in Australia and New Zealand reveals geographical endemism over multiple lineages and recent evolution under host selection. Microb Genom 2:e000095. doi: 10.1099/mgen.0.000095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Sabat AJ, Budimir A, Nashev D, Sá-Leão R, Van Dijl JM, Laurent F, Grundmann H, Friedrich AW, ESCMID Study Group of Epidemiological Markers (ESGEM). 2013. Overview of molecular typing methods for outbreak detection and epidemiological surveillance. Euro Surveill 18:20380. doi: 10.2807/ese.18.04.20380-en. [DOI] [PubMed] [Google Scholar]
  • 17.Jones N, Bohnsack JF, Takahashi S, Oliver KA, Chan M-S, Kunst F, Glaser P, Rusniok C, Crook DWM, Harding RM, Bisharat N, Spratt BG. 2003. Multilocus sequence typing system for group B streptococcus. J Clin Microbiol 41:2530–2536. doi: 10.1128/JCM.41.6.2530-2536.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Carvalho-Castro GA, Silva JR, Paiva LV, Custódio DAC, Moreira RO, Mian GF, Prado IA, Chalfun-Junior A, Costa GM. 2017. Molecular epidemiology of Streptococcus agalactiae isolated from mastitis in Brazilian dairy herds. Braz J Microbiol 48:551–559. doi: 10.1016/j.bjm.2017.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Kong F, Gowan S, Martin D, James G, Gilbert GL. 2002. Serotype identification of group B streptococci by PCR and sequencing. J Clin Microbiol 40:216–226. doi: 10.1128/JCM.40.1.216-226.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kapatai G, Patel D, Efstratiou A, Chalker VJ. 2017. Comparison of molecular serotyping approaches of Streptococcus agalactiae from genomic sequences. BMC Genomics 18:429. doi: 10.1186/s12864-017-3820-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Sheppard AE, Vaughan A, Jones N, Turner P, Turner C, Efstratiou A, Patel D, Modernising Medical Microbiology Informatics Group, Walker AS, Berkley JA, Crook DW, Seale AC. 2016. Capsular typing method for Streptococcus agalactiae using whole-genome sequence data. J Clin Microbiol 54:1388–1390. doi: 10.1128/JCM.03142-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Eldar A, Shapiro O, Bejerano Y, Bercovier H. 1995. Vaccination with whole-cell vaccine and bacterial protein extract protects tilapia against Streptococcus difficile meningoencephalitis. Vaccine 13:867–870. doi: 10.1016/0264-410X(94)00067-W. [DOI] [PubMed] [Google Scholar]
  • 23.Evans JJ, Klesius PH, Shoemaker CA. 2004. Efficacy of Streptococcus agalactiae (group B) vaccine in tilapia (Oreochromis niloticus) by intraperitoneal and bath immersion administration. Vaccine 22:3769–3773. doi: 10.1016/j.vaccine.2004.03.012. [DOI] [PubMed] [Google Scholar]
  • 24.Chong S, Wong W, Lee W, Tan Z, Tay Y, Teo X, Chee L, Fernandez C. 2017. Streptococcus agalactiae outbreaks in cultured golden pomfret, Trachinotus blochii (Lacépède), in Singapore. J Fish Dis 40:971–974. doi: 10.1111/jfd.12570. [DOI] [PubMed] [Google Scholar]
  • 25.Sun J, Fang W, Ke B, He D, Liang Y, Ning D, Tan H, Peng H, Wang Y, Ma Y, Ke C, Deng X. 2016. Inapparent Streptococcus agalactiae infection in adult/commercial tilapia. Sci Rep 6:26319. doi: 10.1038/srep26319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Kannika K, Pisuttharachai D, Srisapoome P, Wongtavatchai J, Kondo H, Hirono I, Unajak S, Areechon N. 2017. Molecular serotyping, virulence gene profiling and pathogenicity of Streptococcus agalactiae isolated from tilapia farms in Thailand by multiplex PCR. J Appl Microbiol 122:1497–1507. doi: 10.1111/jam.13447. [DOI] [PubMed] [Google Scholar]
  • 27.Godoy D, Carvalho-Castro G, Leal C, Pereira U, Leite R, Figueiredo H. 2013. Genetic diversity and new genotyping scheme for fish pathogenic Streptococcus agalactiae. Lett Appl Microbiol 57:476–483. doi: 10.1111/lam.12138. [DOI] [PubMed] [Google Scholar]
  • 28.Evans JJ, Bohnsack JF, Klesius PH, Whiting AA, Garcia JC, Shoemaker CA, Takahashi S. 2008. Phylogenetic relationships among Streptococcus agalactiae isolated from piscine, dolphin, bovine and human sources: a dolphin and piscine lineage associated with a fish epidemic in Kuwait is also associated with human neonatal infections in Japan. J Med Microbiol 57:1369–1376. doi: 10.1099/jmm.0.47815-0. [DOI] [PubMed] [Google Scholar]
  • 29.Delannoy CMJ, Crumlish M, Fontaine MC, Pollock J, Foster G, Dagleish MP, Turnbull JF, Zadoks RN. 2013. Human Streptococcus agalactiae strains in aquatic mammals and fish. BMC Microbiol 13:41. doi: 10.1186/1471-2180-13-41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Pridgeon JW, Zhang D. 2014. Complete genome sequence of a virulent streptococcus agalactiae strain, 138P, isolated from diseased Nile tilapia. Genome Announc 2:e00295-. doi: 10.1128/genomeA.00295-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Lusiastuti AM, Textor M, Seeger H, Akineden Ö, Zschöck M. 2014. The occurrence of Streptococcus agalactiae sequence type 261 from fish disease outbreaks of tilapia Oreochromis niloticus in Indonesia. Aquacult Res 45:1260–1263. doi: 10.1111/are.12069. [DOI] [Google Scholar]
  • 32.Suanyuk N, Kong F, Ko D, Gilbert GL, Supamattaya K. 2008. Occurrence of rare genotypes of Streptococcus agalactiae in cultured red tilapia Oreochromis sp. and Nile tilapia O. niloticus in Thailand–relationship to human isolates? Aquaculture 284:35–40. doi: 10.1016/j.aquaculture.2008.07.034. [DOI] [Google Scholar]
  • 33.Li L, Wang R, Liang W, Gan X, Huang T, Huang Y, Li J, Shi Y, Chen M, Luo H. 2013. Rare serotype occurrence and PFGE genotypic diversity of Streptococcus agalactiae isolated from tilapia in China. Vet Microbiol 167:719–724. doi: 10.1016/j.vetmic.2013.09.001. [DOI] [PubMed] [Google Scholar]
  • 34.Chideroli RT, Amoroso N, Mainardi RM, Suphoronski SA, de Padua SB, Alfieri AF, Alfieri AA, Mosela M, Moralez ATP, de Oliveira AG, Zanolo R, Di Santis GW, Pereira UP. 2017. Emergence of a new multidrug-resistant and highly virulent serotype of Streptococcus agalactiae in fish farms from Brazil. Aquaculture 479:45–51. doi: 10.1016/j.aquaculture.2017.05.013. [DOI] [Google Scholar]
  • 35.Tettelin H, Masignani V, Cieslewicz MJ, Donati C, Medini D, Ward NL, Angiuoli SV, Crabtree J, Jones AL, Durkin AS, Deboy RT, Davidsen TM, Mora M, Scarselli M, Margarit y Ros I, Peterson JD, Hauser CR, Sundaram JP, Nelson WC, Madupu R, Brinkac LM, Dodson RJ, Rosovitz MJ, Sullivan SA, Daugherty SC, Haft DH, Selengut J, Gwinn ML, Zhou L, Zafar N, Khouri H, Radune D, Dimitrov G, Watkins K, O'Connor KJ, Smith S, Utterback TR, White O, Rubens CE, Grandi G, Madoff LC, Kasper DL, Telford JL, Wessels MR, Rappuoli R, Fraser CM. 2005. Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: implications for the microbial “pan-genome.” Proc Natl Acad Sci U S A 102:13950–13955. doi: 10.1073/pnas.0506758102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Francis AR, Tanaka MM. 2012. Evolution of variation in presence and absence of genes in bacterial pathways. BMC Evol Biol 12:55. doi: 10.1186/1471-2148-12-55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Puymège A, Bertin S, Guédon G, Payot S. 2015. Analysis of Streptococcus agalactiae pan-genome for prevalence, diversity and functionality of integrative and conjugative or mobilizable elements integrated in the tRNALys CTT gene. Mol Genet Genomics 290:1727–1740. doi: 10.1007/s00438-015-1031-9. [DOI] [PubMed] [Google Scholar]
  • 38.Richards VP, Palmer SR, Pavinski Bitar PD, Qin X, Weinstock GM, Highlander SK, Town CD, Burne RA, Stanhope MJ. 2014. Phylogenomics and the dynamic genome evolution of the genus Streptococcus. Genome Biol Evol 6:741–753. doi: 10.1093/gbe/evu048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Liu G, Zhang W, Lu C. 2013. Comparative genomics analysis of Streptococcus agalactiae reveals that isolates from cultured tilapia in China are closely related to the human strain A909. BMC Genomics 14:775. doi: 10.1186/1471-2164-14-775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.He E-M, Chen C-W, Guo Y, Hsu M-H, Zhang L, Chen H-L, Zhao G-P, Chiu C-H, Zhou Y. 2017. The genome of serotype VI Streptococcus agalactiae serotype VI and comparative analysis. Gene 597:59–65. doi: 10.1016/j.gene.2016.10.030. [DOI] [PubMed] [Google Scholar]
  • 41.Maione D, Margarit I, Rinaudo CD, Masignani V, Mora M, Scarselli M, Tettelin H, Brettoni C, Iacobini ET, Rosini R, D'Agostino N, Miorin L, Buccato S, Mariani M, Galli G, Nogarotto R, Nardi-Dei V, Vegni F, Fraser C, Mancuso G, Teti G, Madoff LC, Paoletti LC, Rappuoli R, Kasper DL, Telford JL, Grandi G. 2005. Identification of a universal group B streptococcus vaccine by multiple genome screen. Science 309:148–150. doi: 10.1126/science.1109869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Bowater RO, Dennis MM, Blyde D, Stone B, Barnes AC, Delamare-Deboutteville J, Horton MA, White M, Condon K, Jones R. 2017. Epizootics of Streptococcus agalactiae infection in captive rays from Queensland, Australia. J Fish Dis 41:223–232. doi: 10.1111/jfd.12701. [DOI] [PubMed] [Google Scholar]
  • 43.Evans JJ, Klesius PH, Pasnik DJ, Bohnsack JF. 2009. Human Streptococcus agalactiae isolate in Nile tilapia (Oreochromis niloticus). Emerg Infect Dis 15:774–776. doi: 10.3201/eid1505.080222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Bishop EJ, Shilton C, Benedict S, Kong F, Gilbert GL, Gal D, Godoy D, Spratt BG, Currie BJ. 2007. Necrotizing fasciitis in captive juvenile Crocodylus porosus caused by Streptococcus agalactiae: an outbreak and review of the animal and human literature. Epidemiol Infect 135:1248–1255. doi: 10.1017/S0950268807008515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Herbert MA, Beveridge CJ, McCormick D, Aten E, Jones N, Snyder LA, Saunders NJ. 2005. Genetic islands of Streptococcus agalactiae strains NEM316 and 2603VR and their presence in other group B streptococcal strains. BMC Microbiol 5:31. doi: 10.1186/1471-2180-5-31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Fischer A, Liljander A, Kaspar H, Muriuki C, Fuxelius HH, Bongcam-Rudloff E, de Villiers EP, Huber CA, Frey J, Daubenberger C, Bishop R, Younan M, Jores J. 2013. Camel Streptococcus agalactiae populations are associated with specific disease complexes and acquired the tetracycline resistance gene tetM via a Tn916-like element. Vet Res 44:86. doi: 10.1186/1297-9716-44-86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Flores AR, Galloway-Pena J, Sahasrabhojane P, Saldana M, Yao H, Su X, Ajami NJ, Holder ME, Petrosino JF, Thompson E, Margarit YRI, Rosini R, Grandi G, Horstmann N, Teatero S, McGeer A, Fittipaldi N, Rappuoli R, Baker CJ, Shelburne SA. 2015. Sequence type 1 group B Streptococcus, an emerging cause of invasive disease in adults, evolves by small genetic changes. Proc Natl Acad Sci U S A 112:6431–6436. doi: 10.1073/pnas.1504725112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Almeida A, Alves-Barroco C, Sauvage E, Bexiga R, Albuquerque P, Tavares F, Santos-Sanches I, Glaser P. 2016. Persistence of a dominant bovine lineage of group B Streptococcus reveals genomic signatures of host adaptation. Environ Microbiol 18:4216–4229. doi: 10.1111/1462-2920.13550. [DOI] [PubMed] [Google Scholar]
  • 49.Vandamme P, Devriese LA, Pot B, Kersters K, Melin P. 1997. Streptococcus difficile is a nonhemolytic group B, type Ib Streptococcus. Int J Syst Bacteriol 47:81–85. doi: 10.1099/00207713-47-1-81. [DOI] [PubMed] [Google Scholar]
  • 50.Eldar A, Bejerano Y, Bercovier H. 1994. Streptococcus shiloi and Streptococcus difficile: two new streptococcal species causing a meningoencephalitis in fish. Curr Microbiol 28:139–143. doi: 10.1007/BF01571054. [DOI] [Google Scholar]
  • 51.Mather PB, Arthington AH. 1991. An assessment of genetic differentiation among feral Australian tilapia populations. Austral J Mar Freshwater Res 42:721–728. doi: 10.1071/MF9910721. [DOI] [Google Scholar]
  • 52.Thorpe HA, Bayliss SC, Sheppard SK, Feil EJ. 2017. Piggy: a rapid, large-scale pan-genome analysis tool for intergenic regions in bacteria. bioRxiv doi: 10.1101/179515. [DOI] [PMC free article] [PubMed]
  • 53.Kuo CH, Ochman H. 2010. The extinction dynamics of bacterial pseudogenes. PLoS Genet 6:e1001050. doi: 10.1371/journal.pgen.1001050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Lerat E, Ochman H. 2005. Recognizing the pseudogenes in bacterial genomes. Nucleic Acids Res 33:3125–3132. doi: 10.1093/nar/gki631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Lerat E, Ochman H. 2004. Psi-Phi: exploring the outer limits of bacterial pseudogenes. Genome Res 14:2273–2278. doi: 10.1101/gr.2925604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Spellerberg B, Martin S, Brandt C, Lutticken R. 2000. The cyl genes of Streptococcus agalactiae are involved in the production of pigment. FEMS Microbiol Lett 188:125–128. doi: 10.1111/j.1574-6968.2000.tb09182.x. [DOI] [PubMed] [Google Scholar]
  • 57.Nagano N, Nagano Y, Taguchi F. 2002. High expression of a C protein beta antigen gene among invasive strains from certain clonally related groups of type Ia and Ib group B streptococci. Infect Immun 70:4643–4649. doi: 10.1128/IAI.70.8.4643-4649.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Yang Q, Zhang M, Harrington DJ, Black GW, Sutcliffe IC. 2011. A proteomic investigation of Streptococcus agalactiae reveals that human serum induces the C protein beta antigen and arginine deiminase. Microbes Infect 13:757–760. doi: 10.1016/j.micinf.2011.03.001. [DOI] [PubMed] [Google Scholar]
  • 59.Yang Q, Zhang M, Harrington DJ, Black GW, Sutcliffe IC. 2010. A proteomic investigation of Streptococcus agalactiae grown under conditions associated with neonatal exposure reveals the upregulation of the putative virulence factor C protein beta antigen. Int J Med Microbiol 300:331–337. doi: 10.1016/j.ijmm.2010.01.001. [DOI] [PubMed] [Google Scholar]
  • 60.Locke JB, Colvin KM, Datta AK, Patel SK, Naidu NN, Neely MN, Nizet V, Buchanan JT. 2007. Streptococcus iniae capsule impairs phagocytic clearance and contributes to virulence in fish. J Bacteriol 189:1279–1287. doi: 10.1128/JB.01175-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Delannoy CMJ, Zadoks RN, Crumlish M, Rodgers D, Lainson FA, Ferguson HW, Turnbull J, Fontaine MC. 2016. Genomic comparison of virulent and non-virulent Streptococcus agalactiae in fish. J Fish Dis 39:13–29. doi: 10.1111/jfd.12319. [DOI] [PubMed] [Google Scholar]
  • 62.Yao KH, Poulsen K, Maione D, Rinaudo CD, Baldassarri L, Telford JL, Sorensen UBS, DEVANI Study Group, Kilian M. 2013. Capsular gene typing of Streptococcus agalactiae compared to serotyping by latex agglutination. J Clin Microbiol 51:503–507. doi: 10.1128/JCM.02417-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Pasnik DJ, Evans JJ, Panangala VS, Klesius PH, Shelby RA, Shoemaker CA. 2005. Antigenicity of Streptococcus agalactiae extracellular products and vaccine efficacy. J Fish Dis 28:205–212. doi: 10.1111/j.1365-2761.2005.00619.x. [DOI] [PubMed] [Google Scholar]
  • 64.Seib KL, Zhao X, Rappuoli R. 2012. Developing vaccines in the era of genomics: a decade of reverse vaccinology. Clin Microbiol Infect 18(Suppl 5):109–116. doi: 10.1111/j.1469-0691.2012.03939.x. [DOI] [PubMed] [Google Scholar]
  • 65.Millard CM, Baiano JC, Chan C, Yuen B, Aviles F, Landos M, Chong RS, Benedict S, Barnes AC. 2012. Evolution of the capsular operon of Streptococcus iniae in response to vaccination. Appl Environ Microbiol 78:8219–8226. doi: 10.1128/AEM.02216-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Amann RI, Ludwig W, Schleifer KH. 1995. Phylogenetic identification and in situ detection of individual microbial cells without cultivation. Microbiol Rev 59:143–169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Andrews S. 2010. FastQC: a quality control tool for high throughput sequence data. https://www.bioinformatics.babraham.ac.uk/projects/fastqc/.
  • 68.Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, Lesin VM, Nikolenko SI, Pham S, Prjibelski AD, Pyshkin AV, Sirotkin AV, Vyahhi N, Tesler G, Alekseyev MA, Pevzner PA. 2012. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol 19:455–477. doi: 10.1089/cmb.2012.0021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Gurevich A, Saveliev V, Vyahhi N, Tesler G. 2013. QUAST: quality assessment tool for genome assemblies. Bioinformatics 29:1072–1075. doi: 10.1093/bioinformatics/btt086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Rissman AI, Mau B, Biehl BS, Darling AE, Glasner JD, Perna NT. 2009. Reordering contigs of draft genomes using the Mauve aligner. Bioinformatics 25:2071–2073. doi: 10.1093/bioinformatics/btp356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Seemann T. 2014. Prokka: rapid prokaryotic genome annotation. Bioinformatics 30:2068–2069. doi: 10.1093/bioinformatics/btu153. [DOI] [PubMed] [Google Scholar]
  • 72.Wyres KL, Wick RR, Gorrie C, Jenney A, Follador R, Thomson NR, Holt KE. 2016. Identification of Klebsiella capsule synthesis loci from whole genome data. Microb Genom 2:e000102. doi: 10.1099/mgen.0.000102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Larsen MV, Cosentino S, Rasmussen S, Friis C, Hasman H, Marvig RL, Jelsbak L, Sicheritz-Pontén T, Ussery DW, Aarestrup FM, Lund O. 2012. Multilocus sequence typing of total-genome-sequenced bacteria. J Clin Microbiol 50:1355–1361. doi: 10.1128/JCM.06094-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Treangen TJ, Ondov BD, Koren S, Phillippy AM. 2014. The Harvest suite for rapid core-genome alignment and visualization of thousands of intraspecific microbial genomes. Genome Biol 15:524. doi: 10.1186/s13059-014-0524-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Croucher NJ, Page AJ, Connor TR, Delaney AJ, Keane JA, Bentley SD, Parkhill J, Harris SR. 2015. Rapid phylogenetic analysis of large samples of recombinant bacterial whole genome sequences using Gubbins. Nucleic Acids Res 43:e15. doi: 10.1093/nar/gku1196. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Stamatakis A. 2014. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30:1312–1313. doi: 10.1093/bioinformatics/btu033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.He Z, Zhang H, Gao S, Lercher MJ, Chen W-H, Hu S. 2016. Evolview v2: an online visualization and management tool for customized and annotated phylogenetic trees. Nucleic Acids Res 44:W236–W241. doi: 10.1093/nar/gkw370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Salipante SJ, Hall BG. 2011. Inadequacies of minimum spanning trees in molecular epidemiology. J Clin Microbiol 49:3568–3575. doi: 10.1128/JCM.00919-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Petkau A, Stuart-Edwards M, Stothard P, Van Domselaar G. 2010. Interactive microbial genome visualization with GView. Bioinformatics 26:3125–3126. doi: 10.1093/bioinformatics/btq588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Alikhan N-F, Petty NK, Ben Zakour NL, Beatson SA. 2011. BLAST Ring Image Generator (BRIG): simple prokaryote genome comparisons. BMC Genomics 12:402. doi: 10.1186/1471-2164-12-402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Page AJ, Cummins CA, Hunt M, Wong VK, Reuter S, Holden MT, Fookes M, Falush D, Keane JA, Parkhill J. 2015. Roary: rapid large-scale prokaryote pan genome analysis. Bioinformatics 31:3691–3693. doi: 10.1093/bioinformatics/btv421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Stanton-Cook M, Ben Zakour N, Alikhan N, Beatson S. 2013. SeqFindR. https://github.com/mscook/SeqFindR.
  • 83.Chen L, Yang J, Yu J, Yao Z, Sun L, Shen Y, Jin Q. 2005. VFDB: a reference database for bacterial virulence factors. Nucleic Acids Res 33:D325–D328. doi: 10.1093/nar/gki008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Cingolani P, Platts A, Wang le L, Coon M, Nguyen T, Wang L, Land SJ, Lu X, Ruden DM. 2012. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly (Austin) 6:80–92. doi: 10.4161/fly.19695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Yamamoto S, Miyake K, Koike Y, Watanabe M, Machida Y, Ohta M, Iijima S. 1999. Molecular characterization of type-specific capsular polysaccharide biosynthesis genes of Streptococcus agalactiae type Ia. J Bacteriol 181:5176–5184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Watanabe M, Miyake K, Yanae K, Kataoka Y, Koizumi S, Endo T, Ozaki A, Iijima S. 2002. Molecular characterization of a novel beta1,3-galactosyltransferase for capsular polysaccharide synthesis by Streptococcus agalactiae type Ib. J Biochem 131:183–191. doi: 10.1093/oxfordjournals.jbchem.a003086. [DOI] [PubMed] [Google Scholar]
  • 87.Martins ER, Melo-Cristino J, Ramirez M. 2007. Reevaluating the serotype II capsular locus of Streptococcus agalactiae. J Clin Microbiol 45:3384–3386. doi: 10.1128/JCM.01296-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Chaffin DO, Beres SB, Yim HH, Rubens CE. 2000. The serotype of type Ia and III group B streptococci is determined by the polymerase gene within the polycistronic capsule operon. J Bacteriol 182:4466–4477. doi: 10.1128/JB.182.16.4466-4477.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Chaffin DO, McKinnon K, Rubens CE. 2002. CpsK of Streptococcus agalactiae exhibits alpha2,3-sialyltransferase activity in Haemophilus ducreyi. Mol Microbiol 45:109–122. doi: 10.1046/j.1365-2958.2002.02988.x. [DOI] [PubMed] [Google Scholar]
  • 90.Cieslewicz MJ, Chaffin D, Glusman G, Kasper D, Madan A, Rodrigues S, Fahey J, Wessels MR, Rubens CE. 2005. Structural and genetic diversity of group B streptococcus capsular polysaccharides. Infect Immun 73:3096–3103. doi: 10.1128/IAI.73.5.3096-3103.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Sullivan MJ, Petty NK, Beatson SA. 2011. Easyfig: a genome comparison visualizer. Bioinformatics 27:1009–1010. doi: 10.1093/bioinformatics/btr039. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplemental material

Articles from Applied and Environmental Microbiology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES