Sie sind auf Seite 1von 27

FEMS Microbiology Reviews 29 (2005) 303–329

www.fems-microbiology.org

q
Genomics of the Bacillus cereus group of organisms
David A. Rasko a, Michael R. Altherr b, Cliff S. Han b, Jacques Ravel a,*

a
The Institute for Genomic Research, 9712 Medical Center Drive, Rockville, MD 20850, USA
b
Los Alamos National Laboratory, P.O. Box 1663, Los Alamos, NM 87545, USA

Received 5 November 2004; received in revised form 22 December 2004; accepted 23 December 2004

First published online 28 January 2005

Abstract

Members of the Bacillus cereus group of organisms include Bacillus cereus, Bacillus anthracis and Bacillus thuringiensis. Collec-
tively, these organisms represent microbes of high economic, medical and biodefense importance. Given this significance, this group
contains the highest number of closely related fully sequenced genomes, giving the unique opportunity for thorough comparative
genomic analyses. Much of the disease and host specificity of members of this group can be attributed to their plasmids, which vary
in size and number. Chromosomes exhibit a high level of synteny and protein similarity, with limited differences in gene content,
questioning the speciation of the group members. Genomic data have spurred functional studies that combined microarrays and
proteomics. Recent advances are reviewed in this article and highlight the advantages of genomic approaches to the study of this
important group of bacteria.
 2005 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.

Keywords: Bacillus cereus; Bacillus anthracis; Genome; Plasmid

Contents

1. Introduction – The Bacillus cereus group of organisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304


1.1. Bacillus thuringiensis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
1.2. Bacillus cereus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
1.3. B. anthracis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
2. Molecular identification of members of the B. cereus group of organisms: One species on the basis of genetic evidence? . . . 305
3. Genome sequencing projects completed and in progress. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
4. B. anthracis genomic analysis for molecular and forensic epidemiological purposes . . . . . . . . . . . . . . . . . . . . . . . . . . 309
5. Comparison of the chromosome sequences of the B. cereus group of organisms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
5.1. Non-pathogenic B. cereus ATCC 14579 and B. cereus ATCC 10987 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309
5.2. The pathogenic members of the B. cereus group of organism, B. cereus G9241, B. cereus Zebra Killer, and
B. thuringiensis 97-27 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312
6. The B. cereus group of organisms plasmid sequence comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
6.1. General features of the B. cereus group plasmids identified in genome sequencing. . . . . . . . . . . . . . . . . . . . . . . 313
6.2. Replication and mobility mechanisms of the B. cereus group plasmids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317

q
Edited by Mark J. Pallen.
*
Corresponding author. Tel.: +1 301 795 7884; fax: +1 301 838 0208.
E-mail address: jravel@tigr.org (J. Ravel).

0168-6445/$22.00  2005 Federation of European Microbiological Societies. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.femsre.2004.12.005
304 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

7. The phage of the B. cereus group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320


7.1. Genomic phage content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
7.2. Lytic induction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
7.3. Exploitation of phage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
8. Functional genomics of the B. cereus group of organisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
8.1. Transcriptional analysis using microarrays. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
8.2. Proteomics studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
9. Summary: One species on the basis of genomic evidence? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325

1. Introduction – The Bacillus cereus group of organisms death of the insect the bacterium is released into the soil
where B. thuringiensis is a ubiquitous inhabitant. In this
The Bacillus cereus group of organisms contains environment and under favorable nutrient condition the
Bacillus thuringiensis, Bacillus anthracis and Bacillus cer- spores could germinate and grow. The Cry genes encode
eus (sensu stricto). This group of Gram-positive spore- the crystal toxins and are usually located on large, trans-
formers forms a highly homogeneous subdivision of missible plasmids. The presence of Cry protein crystals
the genus Bacillus. Demonstration of their high genetic in the spore is speculated to give B. thuringiensis an
relatedness has contributed to the suggestion that advantage in the soil environment upon sporulation [7]
B. anthracis, B. cereus and B. thuringiensis are members over B. cereus, B. thuringiensis is phenotypically distin-
of a single species, B. cereus sensu lato [1–3]. Tradition- guished from B. cereus only by the formation of intracel-
ally, these organisms have been differentiated based on lular protein crystals during sporulation. Overall,
their phenotypic characteristics, including pathogenic genetic studies have shown that B. cereus and B. thurin-
potential. Bacillus mycoides, Bacillus pseudomycoides giensis are essentially identical [8]. In addition, like B.
and Bacillus weihenstephanensis are also members of cereus, B. thuringiensis could be considered an opportu-
the B. cereus group of organisms. However, as no geno- nistic pathogen in animals and human [9–11].
mic data are available, these species have not been in-
cluded in this review. 1.2. Bacillus cereus

1.1. Bacillus thuringiensis B. cereus is ubiquitous in nature and an opportunistic


pathogen, often associated with two forms of human
B. thuringiensis has long been regarded as an insect food poisoning, characterized by either diarrhea and
pathogen alone. The insecticidal spectrum varies within abdominal distress or nausea and vomiting. In healthy
the 82 different serotypes reported [4], and affects insects individuals but mostly in individuals with certain under-
primarily from the orders Lepidoptera, Diptera and lying conditions such as, immunocompromised patients,
Coleoptera. There are also reports of B. thuringiensis iso- or patients recovering from surgery, B. cereus has been
lates active against mosquitoes that are vectors for dis- known to cause a variety of infections, including:
ease, such as malaria and yellow fever [5]. B. endophthalmitis, bacteremia, septicemia, endocarditis,
thuringiensis spore preparations have been successfully salpingitis, cutaenous infections, pneumonia and menin-
commercialized as biopesticides. The spores are associ- gitis [12,13]. B. cereus is found as a contaminant in many
ated with large crystal protein inclusions, which can food products, including dairy products. However, its
make up to 25% of the dry weight of the spore prepara- primary ecological niche is the soil environment. It is
tions. The crystals are aggregates of a large protein also commonly found as part of the gut microflora of
(about 130–140 kDa) that is actually a protoxin. Upon invertebrates, not only as spores but also as growing
ingestion by insect larvae, the protoxin crystals solubi- vegetative cells [14]. By definition, B. cereus is acrystallif-
lize in the mid-gut, where it is cleaved by a gut protease erous, but a B. cereus strain carrying a functional cry
to produce an active toxin (d-endotoxins) of about gene is considered as a B. thuringiensis strain [15]. No
60 kDa. The toxin binds to the mid-gut epithelial cells, virulence factors specific to B. cereus have been identi-
creating pores in the cell membranes. As a result, the fied, and proteins thought to be specific to B. cereus have
gut is rapidly immobilized and the epithelial cells lyse. recently been found in B. thuringiensis isolates [6].
The insect larva stops feeding and often dies from lethal
septicemia [6]. Little is known about the ecology of B. 1.3. B. anthracis
thuringiensis and conflicting reports are reviewed by Jen-
sen et al. [7]. The B. thuringiensis natural environment is B. anthracis is the etiological agent of anthrax, an
thought to be the insect host intestinal system. Upon acute fatal disease found primarily among herbivores,
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 305

but in fact all mammals are susceptible. In recent years, were able to consistently distinguish B. cereus from B.
it has become best known for its use as a biological thuringiensis, but all showed a high level of genetic diver-
weapon [16,17]. The bacterium is endemic or hyper-en- sity between these two phylogenetically interspersed
demic in Africa, Asia and South America [15] but can species.
be found in most of the world. B. anthracis spores are In addition, these data confirmed that B. anthracis is
highly resistant to heat, ultraviolet and ionizing radia- a monomorphic species, in which the overall diversity
tions, pressure and a variety of chemical agents and among isolates cannot be accurately distinguished using
are thought to survive in the environment for decades these techniques. Suppression subtractive hybridization
and possibly centuries [18]. Despite the fact that B. was also used to catalogue some of the unique genetic
anthracis shares the same ecological niche as B. cereus content of B. anthracis [35]. Ninety-three genomic differ-
and B. thuringiensis, and can be readily isolated from ences were discovered that could potentially be used as
contaminated soil samples, it is still unclear if its life cy- ‘‘signatures’’ to discriminate B. anthracis from its closest
cle includes a vegetative stage outside the host [19]. relatives but not between B. anthracis strains. The
Spores are ingested by herbivores and germinate within inability to separate B. anthracis isolates from one an-
the host to produce vegetative cells, which multiply and other led the Keim Genetics Laboratory to develop a ro-
produce virulence factors, ultimately killing the host. bust B. anthracis molecular typing system based on
Upon death, large numbers of bacilli are released into variable number of tandem repeats (VNTR). A multi-
the environment (i.e., soil), and sporulate upon contact ple-locus VNTR analysis (MLVA) using 8 then 15 loci
with air, completing the life cycle. efficiently cluster B. anthracis in two major groups (A
B. anthracis can be differentiated from B. cereus and and B), with the A group widely distributed worldwide,
B. thuringiensis though microbiological and biochemical while the B group has a limited geographic distribution
tests. B. anthracis isolates are non-hemolytic, non- [36].
motile, penicillin-sensitive, susceptible to c-phage, and These molecular typing methods have yet to answer
produce a poly-c-D-glutamic acid capsule [20]. Like whether B. cereus, B. thuringiensis and B. anthracis be-
B. thuringiensis, the ability of B. anthracis to cause a dis- long to one unique species and are variants of that
ease is primarily attributed to its plasmid content. Fully species. The simplest markers used for phenotypic dif-
virulent strains of B. anthracis carry two large plasmids, ferentiation of these species are all encoded on plas-
pXO1 (181 kb) and pXO2 (96 kb), which encode the mids (i.e., the pXO2 encoded capsule gene cluster)
machinery necessary to produce and regulate the an- [19]. The fact that plasmids can easily be transferred
thrax virulence factors, the tripartite toxin and the cap- or lost makes these criteria unacceptable for typing
sule, respectively [21,22]. purposes. For example, it is believed that a plasmid-
less isolate of B. anthracis is indistinguishable from
B. cereus [37]. In addition, the isolation of B. cereus
2. Molecular identification of members of the B. cereus isolates containing the B. anthracis plasmid pXO1,
group of organisms: One species on the basis of genetic with a novel capsule and associated with an illness
evidence? resembling anthrax [20], makes this classification inap-
propriate. B. anthracis appears to be a highly success-
The classification and taxonomic separation of ful variant of the B. cereus group, with a large
B. anthracis, B. cereus and B. thuringiensis has long been number of strains that have been isolated, simply be-
cause for controversy among bacteriologists, and distin- cause it is so easy to identify cases of classical an-
guishing these species is rather difficult even with mod- thrax, which in turn is a readily recognizable disease
ern molecular tools. The close relationship among the [38]. Is it possible that this ease of isolation has biased
different members of the B. cereus group of organisms our view of this group and thus B. anthracis could
has been established through the use of molecular tech- represent an over-sampled B. cereus? RepPCR analysis
niques such as DNA–DNA hybridization [23] and more placed B. anthracis as an independent lineage in a B.
recently by comparing 16S or 23S rRNA sequences or cereus cluster. Perhaps all these species should be clas-
16S–23S rRNA spacer regions [2,24–26]. However, these sified as B. cereus and subsequently differentiated by
methods could not adequately differentiate members of their plasmid content, as suggested previously [3]?
this group. There are numerous reports of attempts at Analyses of large culture collections of B. anthracis,
developing molecular typing systems that would dis- B. cereus and B. thuringiensis isolates by AFLP and
criminate between B. cereus, B. thuringiensis and MLST [20,28,31] have identified a class of organisms
B. anthracis. These included multilocus enzyme electro- containing toxigenic B. cereus and B. thuringiensis that
phoresis (MEE) [3,8,27], multilocus sequence typing are closely related to B. anthracis. These isolates were
(MLST) [28–30], fluorescent amplified fragment length phylogenetically distinct from environmental B. cereus
polymorphism analysis (AFLP) [31–33] and rep-PCR and B. thuringiensis [31] and might represent the closest
fingerprinting [34] among others. None of these studies ancestor to B. anthracis.
306 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

It is with this view of the genetic diversity of the lent isolate of the Ames strain was obtained from the
B. cereus group that we have entered the genomic era. Defense Evaluation and Research Facility at Porton
This knowledge has driven our selection of organisms Down, UK [43], where it was cured of plasmids pXO1
to sequence and genomic-scale tools have been can be and pXO2 by incubation at 43 C and by novobiocin
used to attempt to settle these outstanding questions. treatment, respectively [40]. The attenuated B. anthracis
Ames Porton Down strain was completely sequenced
and released in 2003. The original B. anthracis Ames
3. Genome sequencing projects completed and in progress Ancestor strain, from which both the Florida and Por-
ton Down isolates were derived, was sequenced to com-
Like many other bacteria, the B. cereus group of pletion, and is the only fully virulent strain of B.
organisms has benefited from the genomic revolution anthracis completely sequenced to date. This strain
that started in 1995 with the publication of the first was acquired in 1981 by researchers at the United States
microbial genome sequence, Haemophilus influenzae Army Medical Research Institute for Infectious Dis-
[39]. Access to the genome sequence of a representative eases (USAMRIID) from the Texas A&M Veterinary
of each member of the group opens up the fields of Medicine and Diagnostic Laboratory Bacteriology
transcriptomics and proteomics, and furthers our under- Department. The strain was originally isolated from a
standing of the mechanisms driving the specific pathoge- dead cow in Sarita, TX. An original culture received
nicity of these organisms. In addition, armed with the in 1981 and stored at USAMRIID since, was the source
genome sequence and a better understanding of the biol- of genomic DNA for sequencing. This strain is the pro-
ogy of B. anthracis, the generation of better and safer genitor to all the Ames strains used in laboratories
vaccine and anti-microbial drugs is possible. As de- around the world, hence it is an important reference.
scribed above, plasmid content confers some of the phe- A set of six diverse isolates of B. anthracis was selected
notypic traits used to distinguish B. anthracis, B. cereus for genome sequencing to draft coverage (8·). These iso-
and B. thuringiensis. Comparative genomics of members lates were carefully chosen to represent as much genetic
within the group can address the question – are these diversity as possible, as previously established through
organisms separate species, or the same species carrying MLVA typing [36]. Three representatives of the A group
different plasmids? (B. anthracis North Western America (AAER00000000),
The first B. anthracis genome sequencing project was B. anthracis Vollum (AAEP00000000), B. anthracis Aus-
underway prior to the Autumn 2001 mail anthrax at- tralia 94 (AAES00000000) – Table 1), two of the B group
tacks [40,41]. This bioterrorism event has had an enor- (B. anthracis Kruger B (AAEQ00000000), B. anthracis
mous impact on the number and choice of projects CNEVA (AAEN00000000) – Table 1) and one isolate
undertaken as genomic tools were applied to microbial belonging to the newly discovered C group (B. anthracis
forensics and biodefense/biopreparedness became a dri- A1055 (AAEO00000000)) were sequenced and have been
ver for a number of funding agencies. recently released to GenBank (Table 1). Comparative
To date, the genome sequence of 15 isolates from the analysis of these genomes to the B. anthracis Ames
B. cereus group of organisms are available in public dat- Ancestor genome has provided much insight into the epi-
abases (Table 1) and others are underway. Conse- demiology and ecology of B. anthracis [42]. Lastly, the
quently, this group of bacteria provides one of the sequence of B. anthracis Sterne 34F2 (pXO1+, pXO2 ),
richest collections of near neighbor sequences and will a strain used in laboratories around the world, and cured
likely profoundly impact future systematics efforts. B. of plasmid pXO2 by Sterne in 1937, was also released in
anthracis genome projects account for two thirds of GenBank (AE017225). It is avirulent but maintains an-
these projects (Table 1). These 10 projects are aimed at thrax toxin production and is the derivative of most live-
establishing a better understanding of the genetic struc- stock vaccines in use throughout the world [44].
ture of the natural B. anthracis population, through the B. cereus ATCC 14579 was selected for whole gen-
discovery of novel polymorphisms, and possibly of ome sequencing as it is non-pathogenic and is the Type
genes unique to certain strains. Single nucleotide poly- strain for the B. cereus species [45,46]. The genome se-
morphisms (SNPs) were used to unravel the molecular quence of this B. cereus ‘‘background’’ isolate was com-
evolutionary history of the group, and were the basis pared to the B. anthracis Ames Florida sequence in an
of a robust and finely detailed typing scheme for B. attempt to provide a basis for whole-genome-based phy-
anthracis isolates [42]. The first draft genome to be pub- logenetic analysis [45]. This analysis was complemented
lished was that of the B. anthracis Ames strain isolated by the completion of B. cereus ATCC 10987, a non-
from the victim of the bioterror attack in Florida [42]. lethal dairy isolate phylogenetically more closely related
The choice was obviously motivated by the need to com- to B. anthracis than B. cereus ATCC 14579 [47]. While
pare the draft sequence of this isolate to that of the the genome sequence of B. cereus ATCC 14579 did re-
Ames strain in progress at the time [40] at The Institute veal a small previously unknown extrachromosomal lin-
for Genomic Research (TIGR). The latter, a non-viru- ear molecule [45], B. cereus ATCC 10987 contained a
Table 1
Strains sequenced and compared in this study
Strain Alternate Source MLVA Genotype Plasmid Coveragec GenBank Reference
designation genotype group genotype Accession No.
B. anthracis Ames – Texas, Cow; plasmids cured GT62 A3.b pXO1 , pXO2 Complete AE016879 [40]
B. anthracis Ames Ancestor A0581 Texas, Cow; plasmids included GT62 A3.b pXO1+, pXO2+ Complete AE017334,
AE017336 (pXO1),
AE017335 (pXO2)
B. anthracis Ames Florida A2012 Patient, Florida, USA GT62 A3.b pXO1+, pXO2+ 12X AAAC01000001, [41]

D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329


AE011190 (pXO1),
AE011191 (pXO2)
B. anthracis Kruger B Kruger National Park, GT87 B1 pXO1+, pXO2+ 12X AAEQ00000000
South Africa
B. anthracis Western A0071 Bison; Canada GT3 A1.a pXO1+, pXO2+ 12X AAER00000000
North America
B. anthracis CNEVA-9066 A0402 Patient, Southern France GT79 B2 pXO1+, pXO2+ 12X AAEN00000000
B. anthracis A0155 – Louisiana, USA N/Ab C pXO1+, pXO2+ 12X AAEO00000000
B. anthracis Australia 94 A0039 Victoria Province, Australia GT55 A3.a pXO1+, pXO2+ 12X AAES00000000
B. anthracis Vollum A4088 – GT79 A4 pXO1+, pXO2+ 12X AAEP0000000
B. anthracis Sternea – pXO2-deficient; GT62 A3.b pXO1+, pXO2 Complete AE017225 Brettin, et al. Unpublished, 2004
the basis for animal
vaccines throughout
the world
d
B. cereus ATCC 10987 – Dairy – – pXO1 , pXO2 Complete AE017194 [47]
B. cereus ATCC 14579a – Dairy – – pXO1 , pXO2 e
Complete AE016877 [45]
f
B. cereus ZK – Zebra, Africa – – pXO1 , pXO2 Complete CP000001 Brettin, et al. Unpublished, 2004
B. cereus G9241 – Louisiana, USA – – pXO1+, pXO2 g
12X AAEK00000000 [20]
f
B. thuringiensis serovar – Severe human tissue – – pXO1 , pXO2 Complete AE017355 Brettin, et al. Unpublished, 2004
konkukian str. 97-27a necrosis, cf. [10]
a
Not sequenced at The Institute for Genomic Research, B. anthracis Sterne and Bacillus thuringiensis serovar konkukian str. 97-27 were sequenced at The Department of Energy Joint Genome
Institute, Bacillus cereus ATCC 14579 was sequenced by Integrated Genomics.
b
Not typed using genotyping schema to date.
c
Complete genomes, one contig no gaps. From GenBank – The Sterne strain of Bacillus anthracis is lacking the plasmid pXO2, resulting in an avirulent phenotype that maintains the main anthrax
toxin. Prevention of anthrax in cattle can be accomplished by vaccination with living spores of the Sterne strain. Most livestock vaccines in use throughout the world are derivatives of the live spore
vaccine formulated by Sterne in 1937 and still use descendants of his strain 34F2.
d
B. cereus ATCC 10987 harbors a large plasmid, pBc10987 (GenBank Accession No. AE017195) that is similar to pXO1, however it lacks the virulence portion that encodes the tripartite lethal
toxin and associated regulatory proteins.
e
B. cereus ATCC 14579 contains a 15-kb linear plasmid that appears to be a phage, however it does not appear to contain any large circular plasmids.
f
No plasmids were released with this genomic sequence into GenBank, however plasmids present (Table 2).
g
B. cereus G9241 contains a plasmid that is 99.6% identical to the pXO1 plasmid from B. anthracis, however it lacks the pXO2 plasmid. There is an additional plasmid that encodes a capsule
biosynthesis operon.

307
308 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

large circular plasmid with homology to the B. anthracis and unlike most B. thuringiensis strains, it is closely re-
plasmid pXO1 [47]. lated to B. anthracis based on phylogenetic analysis using
Another group of isolates has been selected for AFLP [31]. The genome sequences of these three isolates
sequencing, as they were not identifiable as B. anthracis open the possibility of developing a better understanding
using classical biochemical and microbiological tests as of their pathogenic potential when compared to B.
detailed above. These isolates were shown to be highly anthracis. These sequences may also provide a snapshot
virulent and were thought to be the cause of a disease of the closest phylogenetic ancestors to B. anthracis.
resembling anthrax in both humans and animals. The However, these strains are the exception rather than
first of these atypical isolates to be sequenced was B. cer- the rule, and may instead constitute another successful
eus G9241 [20], reportedly isolated from the sputum and evolutionary lineage of B. cereus. Despite its importance
blood of a welder with life-threatening pneumonia, as a biopesticide, B. thuringiensis as a non-pathogenic
whose clinical presentation resembled those of reported species is under-represented in the collection of whole
for 10 bioterrorism-associated inhalation anthrax pa- genome sequences. A single genome, B. thuringiensis
tients from 2001 [17]. Interestingly, B. cereus G9241 har- subsp. israelensis, has been sequenced to 8· and assem-
bors a plasmid with 99.6% identity to B. anthracis pXO1. bled in more than 800 contigs (www.integratedgenom-
Although homologues of pXO2-encoded capsule genes ics.com), but is not available to the scientific
were not found, a polysaccharide capsule cluster is en- community for analysis. However, the sequence of the
coded on a second previously unidentified plasmid. B. crystal toxin-encoding plasmid pBtoxis has been recently
cereus Zebra Killer (ZK) was isolated from a swab of published [48] and showed some similarity with B.
the carcass of a dead zebra suspected of having died of anthracis plasmid pXO1.
anthrax in the Etosha National Park, Namibia. This iso- The isolates for which whole genome sequences are
late is phylogenetically the closest completely sequenced now available do not properly represent the diversity
isolate to B. anthracis when typed by either AFLP or observed in the B. cereus group of organisms using
MLST [31] (Fig. 1). B. thuringiensis 97-27 (subsp. konku- methods such as MLST (Fig. 1). The choice of strains
kian (serotype H34)) was originally isolated from a case sequenced so far reflects a bias driven by a need to
of severe human tissue necrosis in a 28 year old otherwise understand rare pathogenic traits in some of these spe-
healthy male patient [10]. Biochemical tests and the pres- cies. Two non-pathogenic B. cereus are available to the
ence of crystals in sporulation culture established the iso- public and one non-pathogenic B. thuringiensis isolate
late as B. thuringiensis. The apparent pathogenic has been sequenced, while 13 pathogenic isolates have
properties of this isolate is unusual for B. thuringiensis been sequenced. The high number of B. anthracis strains

Fig. 1. Multi-locus sequence typing (MLST) neighbor-joining phylogenetic tree of representatives of the B. cereus group of organisms. All B.
anthracis isolates (ST1-3) colored red are represented as a single point on the tree. B. cereus isolates are colored black and B. thuringiensis are colored
in blue. ‘‘ST’’ following each strain designation indicates the sequence type as listed in the PubMLST database [29]. Isolates that have been sequenced
are boxed in gray.
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 309

sequenced has taken the whole genome sequencing ap- A sub-set of 990 SNPs were then typed against a panel of
proach to the limits of its resolution and has led to a 26 diverse B. anthracis strains using a high-throughput
phylogenetic resolution of the group never achieved assay based on the SnaPshot primer extension protocol
for any other bacterial species [42]. (Applied Biosystems). This data demonstrated that pre-
cise phylogenetic topologies could be achieved, yielding
accurate information on internodal distances. In addi-
4. B. anthracis genomic analysis for molecular and tion, using appropriate B. cereus outgroups, the authors
forensic epidemiological purposes were able to determine the evolutionary root of the B.
anthracis clades. The root lies closer to a newly described
The disciplines of emerging infectious disease and group C than either of the two previously described A
forensic microbiology have shifted from using biological and B groups. Unequal evolutionary rates were observed
phenotypes as markers to using more reliable and quan- among the sequenced isolates that could be correlated
tifiable molecular markers, such as SNPs [49]. B. anthra- with ecological parameters and strain attributes, such
cis, as shown above, has received considerable attention as host availability in its natural environment. The young
because of its demonstrated use as a biological weapon evolutionary characteristics of the B. anthracis group
and the difficulties associated with forensic tracking of were confirmed by the rarity of SNPs and low overall
this genetically homogenous group of organisms homoplasy. Armed with this new set of markers, it is
[36,42]. In response to the threat inherent to B. anthra- now possible to select ‘‘canonical SNPs’’ for identifying
cis, vast resources have been allocated to develop diag- long branches or key phylogenetic positions [51]. A
nostic characters for this species through genome molecular typing strategy that maximizes the use of
sequencing. Soon after the Autumn 2001 attacks [50], markers discovered through whole-genome comparison
a draft genome sequence of a B. anthracis Ames Florida was recently established for B. anthracis [51]. PHRANA
strain was generated, and compared to the nearly com- (Progressive hierarchical resolving assays using nucleic
plete genome sequence of the plasmid-less B. anthracis acids) is a nested approach that employs canonical SNPs,
Ames Porton strain [41]. This analysis revealed 60 new MLVA (15 loci), and simple nucleotides repeats (SNRs).
markers that included SNPs, indels and tandem repeats. PHRANA makes use of both the low resolving power of
Only four differences were discovered between the main canonical SNPs and the high resolving power of SNRs, in
chromosomes of the Florida and Porton isolates (two a sequential manner. This approach allows for high-reso-
SNPs and two short indels). lution and accurate representation of the relationships
For the first time polymorphic markers discovered between B. anthracis isolates by minimizing the effect of
through comparative genome sequence analysis were homoplasy using a stepwise approach [51].
used to test a collection of anthrax isolates and were able
to divide these samples into distinct subgroups [41]. This
study was critical in establishing that infectious disease 5. Comparison of the chromosome sequences of the
outbreaks, both naturally occurring and maliciously re- B. cereus group of organisms
leased (bioterrorism), could be investigated through
whole genome-based analysis. More importantly, the 5.1. Non-pathogenic B. cereus ATCC 14579 and
study introduced statistical models allowing scientists B. cereus ATCC 10987
to discriminate true polymorphism from random
sequencing error. This was critical as in light of the rapid The genome sequences of two non-pathogenic B. cer-
generation of B. anthracis draft genome sequences. A eus isolates (ATCC 14579 and ATCC 10987) have been
draft genome is by definition unfinished, and the poor completed and compared to that of B. anthracis [45,47].
quality of portions of a draft sequence is of little use To better understand the degree of relatedness of mem-
for polymorphism discovery. bers of the B. cereus group, one should compare the
The set of polymorphisms found when these two chromosome gene content and genetic structure. These
strains were compared was not sufficient for typing unse- organisms demonstrate a wide range of phenotypes
quenced strains. It was speculated that sequencing other and pathological effects, however these effects are often
distantly related B. anthracis strains might yield addi- derived from factors encoded on extra-chromosomal
tional polymorphisms and increase the resolution of the elements, such as the large toxin encoding plasmid
method. A recent study applied this approach to a larger pXO1 in B. anthracis. Sequence analysis of nine chro-
number of B. anthracis draft genomes and discussed the mosomal genes has suggested that B. anthracis, B. cereus
phylogenetic discovery bias that resulted from using and B. thuringiensis should be considered as one species
SNPs extracted from whole-genome sequence compari- [3]. Access to the complete genome sequence of both B.
sons [42]. More than 3500 rare high quality SNPs were cereus and B. anthracis isolates gave the opportunity to
discovered when the genome of five B. anthracis draft revisit the question and explore the genetic content that
genomes were compared to the closed B. anthracis Ames. governs some of these specific phenotypes.
310 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

A detailed comparative analysis of the genomes of Using a B. anthracis DNA microarray, comparative
B. cereus and B. anthracis, revealed a small subset of genome hybridization (CGH) was performed with 19 di-
genes unique to either species, most of which are anno- verse isolates from B. cereus group of organisms, includ-
tated as hypothetical [47]. The majority of these genes ing B. weihenstephanensis [40]. CGH confirmed the
are located at the terminus of replication, indicating that overall similarity of chromosomal genes among this
genome plasticity mostly occurs in that region, as previ- group of close relatives. Interestingly, the set of core
ously observed for other microbial groups [52]. These genes conserved across each members of the group is lo-
data suggest a history of insertion and/or deletion in cated around the origin of replication, This study also
the evolution of the B. cereus group [47]. It was observed revealed six regions unique to the B. anthracis chromo-
that, in many cases, genes found at a specific position in some, comprising four phages regions and IS110 related
one genome were replaced with others at the correspond- insertion elements. Not surprisingly, the presence of
ing loci in another. These regions often were the result of non-toxin related pXO1 gene homologues were detected
insertion/deletion events of mobile genetic element such in half the 19 strains tested and is further evidence for
as phages, transposons, IS elements or metabolic adapta- the mobility of pXO1 genes within the B. cereus group
tions. For example, a nine-gene urease gene cluster was (see section below). In contrast, there were few pXO2
identified in the genome of B. cereus ATCC 10987 that gene homologues found in the set of isolates tested.
replaces the B. anthracis/B. cereus ATCC 14579 genes Chromosomally encoded genes that may contribute
that encodes for hypothetical proteins, blasticidin S to pathogenicity, including hemolysins, phospholipases
deaminase and S layer protein. Additionally, a xylose and iron acquisition systems were identified in the gen-
utilization operon was found to replace functions such omes of B. anthracis. These include genes similar to
as nitrate reduction, nitrite reduction and molybdopterin those known to boost viral infectivity by degrading the
synthesis. The lack of nitrate and nitrite reduction ap- mucin layer of insect guts, and genes that contribute
pears to have been compensated by the acquisition of to the virulence of the Gram-positive pathogen, Listeria
the urease operon, allowing the bacterium to use ammo- monocytogenes [40]. These genes and some newly identi-
nia as a nitrogen source derived from urea. In addition, fied surface proteins constitute potential targets for drug
B. cereus ATCC 10987 contains a unique set of genes and vaccine development.
responsible for the transport and utilization of tagatose. Analysis of the metabolic potential of both B. anthra-
This particular strain has been isolated during a study on cis and B. cereus shows that these organisms have an ex-
cheese spoilage, where this carbohydrate can be found. panded capacity for amino-acid and peptide utilization,
This replacement might represent another example of and hence are equipped for a life in a protein rich envi-
metabolic adaptations to the carbohydrate-rich environ- ronment. This lifestyle was emphasized by the presence
ment of milk [47,135]. A cluster encoding for the arginine of six amino-acid efflux systems, which prevent accumu-
deiminase pathway was identified in both B. cereus lation of amino acids to bacteriostatic concentrations
ATCC 10987 and ATCC 14579 that, like in Streptococ- during growth on peptides [40]. In addition, complex
cus pyogenes, might enable B. cereus to survive in acidic carbohydrate degradation appears to be favored by
conditions [45]. On the other hand, it is thought that these organisms over small sugars. They indeed lack cat-
ammonium inhibits receptor-mediated internalization abolic capacity for the utilization of mannose, arabinose
of the lethal toxin, hence deletion of the arginine deimin- and rhamnose among others, but seem to be capable of
ase pathway in B. anthracis might be the result of an evo- cleavage of extracellular chitin and chitosan [40]. These
lutionary adaptation to a pathogenic lifestyle [45]. metabolic capabilities suggest that the most recent
These specific evolutionary events represent niche ancestor to the B. cereus group of organisms might have
specific adaptations, but do not eliminate the fact that been able to thrive on animals or insects [53].
they are highly similar and syntenic (conserved gene or- Comparative analyses suggest that major differences
der). An analysis using normalized blast scores to com- between members of the B. cereus group might repre-
pare the proteomes of B. cereus G9241, B. cereus ATCC sent fine alteration in gene expression rather than the
10987 to that of B. anthracis, demonstrated how syn- level of sequence divergence or gene content
tenic these three genomes are [47,135]. Fig. 2 represents [40,45,47]. PlcR is a pleiotropic transcriptional regula-
a similar analysis with an added level of color-coded tor that upregulates the expression of more than 100
protein similarity and shows the high level of gene sim- genes in B. cereus through binding to an upstream pal-
ilarity and gene order (Figs. 2(a) and (b)), with no inver- indromic motif [45,54,55]. PlcR activity has been
sions or genetic rearrangements of large genome shown to be regulated by the presence of a secreted
segments. This analysis also shows that B. cereus and reimported pentapetide produced from the pro-
G9241 is more closely related to B. anthracis than to cessing of the PapR protein C-terminus [56]. The papR
B. cereus ATCC 10987 (Fig. 2(c)). These bioinformatic gene is itself upregulated by PlcR, forming a quorum
results are mirrored in the increased pathogenicity of sensing-like system [56,57]. Proteomic-based compara-
B. cereus G9241. tive analysis of B. cereus ATCC 14579 and a DPlcR
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 311

Fig. 2. Blast score ratio analysis of B. cereus G9241, B. cereus ATCC 10987 and B. anthracis Ames Ancestor. Blast score ratios are obtained by
dividing the Blast score for the most similar query peptide versus the reference peptide by the Blast score for the reference peptide against itself. The
Blast score ratio can then be graphically represented as a synteny plot of two genomes using the genomic location of the best hit in each genome as
the coordinates on the Cartesian plane [135]. These synteny plots demonstrate that pathogenic and non-pathogenic B. cereus group members have a
high degree of similarity as indicated by the color of each peptide, represented by a dot, and conserved gene order (synteny) represented by the dots
forming a line with a slope of 1. (a) Synteny plot of B. anthracis vs. B. cereus G9241. (b) Synteny plot of B. anthracis vs. B. cereus ATCC 10987. The
scale at the top of the figure indicates the color and Blast score ratio correlation represented in (a) and (b). A Blast score ratio of less than 0.4 is not
considered significant. (c) Blast score ratio analysis of B. cereus G9241 and B. cereus ATCC 10987 using B. anthracis Ames Ancestor as the reference.
The Blast score ratio for each of the query genomes can also be used to obtain an overall idea of similarity between the genomes compared (red –
highly conserved in all three genomes; orange – unique to the genome used as the reference; green – shared between B. anthracis Ames Ancestor and
B. cereus G9241; blue – shared between B. anthracis Ames Ancestor and B. cereus ATCC 10987).
312 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

variant of the same strain by two-dimensional gel elec- to B. anthracis inferred using AFLP [31] or MLST
trophoresis has identified collagenase, phospholipases, ([28] and Fig. 1).
hemolysins, proteases and enterotoxins genes that are In common with other members of the B. cereus
PlcR upregulated [54]. Some of these activities are con- group, B. cereus G9241, B. cereus ZK, and B. thuringien-
sidered virulence factors in B. cereus and all of their sis 97-27, share a set of genes that are associated with
associated genes displayed a PlcR binding box in the virulence, such as non-hemolytic enterotoxins, chan-
promoter region. Even though these genes are also nel-forming type III hemolysins, a perfringolysin O (lis-
found in B. anthracis, a nonsense point mutation in teriolysin O), phospholipases C, and a family of
the PlcR gene is responsible for an abolition or dra- extracellular proteases [60]. Interestingly, the hemolytic
matic reduction in their expression such that produc- B. cereus ZK does not contain the hemolysin BL (hbl)
tion of lecithinase, protease and hemolysins is operon, which is a primary factor in diarrheal B. cereus
undetectable [40]. A few genes encoding cytotoxin K food poisoning [61]. In contrast, B. cereus G9241 con-
and non-hemolytic enterotoxin C subunit, amongst tains four hemolysins (A, BL, II and III), and has also
others, have only been found in B. cereus [47]. The been shown to be phenotypically hemolytic [16]. In line
acquisition of the toxin encoding pXO1 plasmid and with other B. cereus and B. thuringiensis isolates, these
its regulator AtxA has been shown to be incompatible strains contain a plcR gene encoding for a full-length,
with the chromosomally encoded PlcR [58]. Interest- potentially functional protein and a similar set of genes
ingly, a recent study showed that a fused PlcR–PapR putatively upregulated by this protein, with a few excep-
construct was able to restore strong hemolytic activities tions. The histidine protein kinase (BC3528) homolo-
when introduced in B. anthracis [59], indicating that gous to the sporulation kinase KinB is a member of
these genes, while not upregulated, are still fully func- the PlcR regulatory network in B. cereus ATCC 14579
tional and have not undergone evolutionary decay. [45] and B. cereus G9241 [20], but is absent in B. cereus
The loss of regulation of these chromosomally encoded ATCC 10987 [47], B. anthracis, B. thuringiensis 97-27
genes might represent another example of an adaptive and B. cereus ZK.
response of B. anthracis to a plasmid driven advanta- The genomes of these pathogenic members of the B.
geous pathogenic lifestyle [40] and because these genes cereus group also contain metabolic adaptations similar
are still fully functional, it confirms the relatively to those found in the non-pathogenic group. For exam-
young evolutionary age of B. anthracis. ple, like B. cereus ATCC 10987, B. cereus ZK contains a
14 kb gene replacement consisting of a gene cluster
5.2. The pathogenic members of the B. cereus group of responsible for the transport and utilization of tagatose
organism, B. cereus G9241, B. cereus Zebra Killer, and [47]. Similarly, B. thuringiensis 97-27, but not B. cereus
B. thuringiensis 97-27 G9241 or B. cereus ZK, is able to utilize arginine
through the arginine deiminase-dependent pathway,
The genome sequences of one B. thuringiensis and which is also found in B. cereus ATCC 14579 and B. cer-
two B. cereus isolates that were associated with severe eus ATCC 10987 [45,47]. These metabolic adaptations
disease have been sequenced in order to gain insight into certainly reflect the environment in which each of the
their pathogenic potential. While one can only infer the organisms thrive, and does not appear to be link to a
pathogenicity of B. cereus ZK based on its isolation pathogenic characteristic of any of these isolates.
from a dead zebra, the opportunistic pathogenicity of Interestingly, and relevant to its pathogenic potential,
B. thuringiensis 97-27 has been confirmed by infection a putative polysaccharide capsule cluster similar to the
of both immunosuppressed and immunocompetent mice one present on B. cereus ATCC 14579 is also present
[9–11]. On the other hand, the pathogenic characteristics on B. thuringiensis 97-27. While a distantly related poly-
of B. cereus G9241 are supported by the severe anthrax- saccharide capsule cluster is found in the same genomic
like clinical presentation it caused. This isolate has been location in B. cereus ATCC 10987 no polysaccharide
shown to be lethal in A/J mice challenged by intraperito- capsule cluster is found on the genomes of either B.
neal injections [20]. anthracis or B. cereus ZK. Unfortunately, no experimen-
The genome sequences of these isolates give us the tal data are available that demonstrate B. cereus ZK or
opportunity for comparison to the non-pathogenic B. B. thuringiensis 97-27 encapsulation in vivo or in vitro.
cereus isolate and B. anthracis. Structurally, these gen- In contrast, B. cereus G9241 has been shown to produce
omes show a high level of synteny and protein identity a capsule by staining with India ink [20] and a novel
to B. anthracis and non-pathogenic B. cereus isolates. polysaccharide capsule gene cluster has been identified
Overall, the proteome of B. thuringiensis 97-27, B. cereus on one of its plasmids. However, no other identifiable
ZK and B. cereus G9241 show a higher similarity to that capsule biosynthetic gene cluster has been identified in
of B. anthracis than to that of the non-pathogenic B. cer- the genome sequence.
eus ATCC 14579. These observations are in agreement Comparative analysis of the genome of these patho-
with the phylogenetic position of these isolates relative genic isolates do not conclusively provide the informa-
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 313

tion necessary to assign a particular set of genes to their ever, the function of the other plasmids in the group
virulent phenotype. Many of the differences are shared is relatively unknown. However, one could expect that
with other non-pathogenic counterparts. It appears that these plasmids encode the peptides responsible for the
B. cereus ZK shares the most similarity with B. anthra- specific phenotypes that differentiate the members of
cis, justifying AFLP typing data as the closest phyloge- this group.
netically isolate to B. anthracis [31]. These isolates all
contain small and/or large plasmids, for which sequence
is now available. A detailed analysis of these plasmids 6.1. General features of the B. cereus group plasmids
might provide a better insight into the pathogenicity of identified in genome sequencing
their host isolates.
6.1.1. B. anthracis plasmids
6.1.1.1. pXO1. The 181-kb plasmid pXO1 encodes
6. The B. cereus group of organisms plasmid sequence genes for the production and regulation of the tripar-
comparisons tite lethal toxin. The three portions of the lethal toxin
are produced separately and assembled. The eukaryotic
Historically, pathogenic potential and diverse host membrane-interacting protective antigen (PA) is ex-
range have defined the members of the B. cereus group. pressed as a pre-protein and proteolytically processed
The small, usually metabolic-based differences found in into its active form, known to elicit a protective im-
the chromosomal content cannot account for the spec- mune response against anthrax [64]. The toxin has
trum of disease and host range observed in this group. two additional components – the lethal factor (LF)
However, much of the disease and host specificity in and the edema factor (EF), which are responsible for
this group can be attributed to plasmid content. The the proteolytic cleavage of several mitogen-activated
plasmids associated with the B. cereus group of organ- protein kinases (MAPKKs) and convert intracellular
isms have broad size range (5–200 kb) and vary in ATP into cAMP, respectively. The biological activities
number ([62] and Table 2). Questions about their role of these subunits are not fully characterized but result
in the life cycles of the species that harbor them re- in aberrant signaling inside the macrophages and fluid
main. The most striking example of the plasmid con- accumulation in the lung [65]. All three subunits are re-
tent affecting host range and pathogenesis occurs in quired for active toxin production and activity (for re-
B. anthracis. B. anthracis harbors two plasmids, one views see [66–70]). The three structural genes, cya (EF),
that encodes the tripartite lethal toxin complex, lef (LF) and pagA (PA), are under the control of at
pXO1 [22] and the other, which contains the biosyn- least two regulatory elements, AtxA and PagR [71],
thetic genes for the poly-c-D-glutamic acid capsule and have been shown to be expressed early in the
[21,63]. While each plasmid is a distinct entity, it has growth of B. anthracis [72]. These genes are encoded
been shown that loss of either one results in an atten- within a 44.5-kb pathogenicity island that is transposi-
uated B. anthracis isolate. The role of the B. anthracis tionally active [19]. However, its inversion does not
plasmids in pathogenesis is exquisitely known, how- appear to affect virulence [41].

Table 2
Comparison of plasmid sequences from plasmids identified in genome sequencing projects
Plasmid name Organism Strain Size (bp) G+C (%) Replicon Replication Number Functional Hypothetical
system of CDS CDS CDSa
pXO1 B. anthracis Allb 181,677 33 Unknown Unknown 204 48 156
pXO2 B. anthracis Allb 94,830 33 pAMb1 Theta 104 26 78
pBc10987 B. cereus ATCC 10987 208,369 33 Unknown Unknown 242 91 151
pBClin15 B. cereus ATCC 14579 15,100 38 Phage Phage 21 4 17
pBCXO1c B. cereus G9241 191,110F 33 Unknown Unknown 177 61 116
pBC218d B. cereus G9241 218,094F 32 Unknown Unknown 185 116 69
pZK5 B. cereus Zebra Killer 5108 31 Group VII RCR 5 2 3
pZK8 B. cereus Zebra Killer 8191 32 Group VII RCR 8 4 4
pZK9 B. cereus Zebra Killer 9150 31 Group VII RCR 10 5 5
pZK54 B. cereus Zebra Killer 53,501 32 pAMb1 Theta 58 27 31
pZK467 B. cereus Zebra Killer 466,370 33 Unknown Unknown 465 228 237
pBT9727 B. thuringiensis 9727 77,122 33 pAMb1 Theta 80 18 62
a
Hypothetical includes the proteins with no know function including CDS with hypothetical, conserved hypothetical and Bacillus cereus group-
specific protein designations.
b
B. anthracis pXO1 from GenBank Accession No. AE017336. B. anthracis pXO2 from GenBank Accession No. AE017335.
c
GenBank Accession No. AAEK01000020, AAEK01000036 and AAEK01000038 (not currently closed molecules).
d
GenBank Accession No. AAEK01000004.
314 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

Within the isolates of B. anthracis for which the gen- [21] and encodes 104 genes for which 78 do not have
ome sequence is available, the pXO1 sequences are functional assignments (78/104, 75%). In contrast to
highly similar. Apart from a small number of SNPs, pXO1, the pXO2 replication region has been identified
SNRs and VNTRs, no large insertions or deletions have and, recently, a 60-nucleotide region essential for the ini-
been observed [41]. To date only one B. anthracis isolate tiation of plasmid replication has been characterized
has been identified without the pXO1 plasmid [42]. This [80]. pXO2 is a theta replicating plasmid similar to the
isolate represents a new lineage in the B. anthracis prototypical pAMb1 plasmid from Enterococcus fea-
group. However, little information is available for this calis. Detailed analysis of the replication machinery is
isolate and it is possible that pXO1 might have been lost provided below. Several studies have indicated that
during laboratory passages. pXO2 sequences are not widely distributed among clo-
The level of sequence coverage obtained for pXO1 sely related species based on CGH, hybridization and
from whole-genome shotgun sequencing projects can PCR experiments [40,63]. Only one isolate, B. thuringi-
be used to estimate the molecular ratio of plasmid to ensis AW06, showed any significant sequence similarity
chromosome [41]. For pXO1, the copy number has been to pXO2 in previous studies, however the B. thuringien-
suggested to be 2–3 copies per chromosome copy ([41] sis plasmid pBT9727 described in this work has signifi-
and this study). This estimate appears to be much higher cant similarity to pXO2 (see below).
than indicated by other studies [73], and certainly repre-
sents a snapshot of the dynamic plasmid content in the 6.1.2. B. cereus plasmids
bacterium at the time of genomic DNA preparation. In contrast to the conserved plasmid content ob-
The origin of replication of pXO1 has never been served among B. anthracis, B. cereus isolates contain a
conclusively identified [22,74]. It has been suggested that diverse range of plasmids – no strains have yet been
it lies within an 11-kb region discovered by subcloning identified with identical plasmid content. B. cereus plas-
[22,75], however this region does not contain any genes mids vary from 5- to almost 500 kb in size and only a
similar to other known plasmid replication systems. This limited number have been implicated in pathogenesis
is a recurring theme among the large plasmids of the (Table 2). A subset do not encode for any obvious phe-
B. cereus group and is addressed further below. While notypes and may be considered cryptic. B. cereus gen-
many studies have focused on the toxin and its regula- ome projects have identified a large number of
tors, five years after the complete sequence of pXO1 plasmids, and their analysis has revealed a number of
was first published [22], 148 of the 204 potential coding conserved regions within the large plasmids group.
sequences (72.5%) remain functionally unidentified. De- The smaller plasmids, in contrast, show greater similar-
tailed transcriptional mapping studies should be able to ity to B. thuringiensis smaller plasmids.
refine the annotation and in the process identify novel
genes involved in the pathogenicity of the organism in 6.1.2.1. pBc10987. A single 208-kb plasmid, named
the future. pBc10987, was identified from the non-pathogenic iso-
late B. cereus ATCC 10987 [47]. pBc10987 shows surpris-
6.1.1.2. pXO2. pXO2 (96 kb) encodes for the synthesis ing similarity to the plasmid pXO1 of B. anthracis (40%
and degradation of another well-known virulence factor, nucleotide identity), however it lacks the pathogenicity
the poly-c-D-glutamic acid capsule [76]. This capsular island (PI) containing the genes that encode for the tri-
material is thought to protect the vegetative bacterial partite lethal toxin and its associated regulators [47]. In
cells during transit through the macrophage, hence lieu of the pXO1 PI, pBc10987 contains genes potentially
evading the host immune response and increasing sys- involved in adaptation to either an environmental or
temic sepsis [77]. The poly-c-D-glutamic acid capsule is pathogenic lifestyle. Environmental adaptations of
produced by a three-gene operon capABC under the pBc10987 include a copper requiring tyrosinase, arsenite
control of a number of regulators including the toxin resistance and its associated regulators as well as an
regulator AtxA encoded on pXO1 [71], and the pXO2 amino acid transport system. The pathogenic adapta-
encoded regulators AcpA and AcpB. Degradation of tions comprise of two potential novel toxins [22,47]. Un-
the capsular material, by CapD [78], encoded on like the PI in pXO1, this region in pBc10987 is not
pXO2, results in high and low molecular weight forms flanked by mobile genetic elements, suggesting that it
of the capsule, both of which essential for infection has not been lost from pBc10987 [22,47]. These findings
[79]. B. anthracis Sterne lacks pXO2, and is commonly suggest that in conjunction with the chromosome, B. cer-
used around the world as an anthrax veterinary vaccine. eus group plasmids contribute to the metabolic fitness of
Very little is known about the molecular mechanism for the species.
the transport of the capsular material to the surface of In addition to these potential metabolic or patho-
the bacterial cell. genic adaptations, a number of aspects of plasmid biol-
The first complete sequence of pXO2 was obtained ogy and chromosome-plasmid crosstalk are conserved
from a B. anthracis Pasteur strain (pXO1 , pXO2+) between pXO1 and pBc10987. Like pXO1, the replica-
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 315

tion machinery of pBc10987 has not been identified virulence genes, protective antigen (99.7% amino acid
using bioinformatics tools, however the similar plasmid identity), lethal factor (99% amino acid identity) and
copy number (1–3 copies per cell) suggests that the edema factor (96% amino acid identity), as well as the
replication mechanism is conserved between the two. known virulence regulatory proteins AtxA (100% amino
Divergent copies of abrB, a pleiotropic transition state acid identity) and PagR (98.6% amino acid identity)
regulator, are present on pXO1 and pBc10987, as well [20]. Price et al. [91] have previously used minor varia-
as the chromosomes of B. anthracis and B. cereus ATCC tions in the protective antigen (PA) protein as a classifi-
10987. AbrB has been shown in B. subtilis to modulate cation system. Using this system, pBCXO1-encoded PA
the switch between biofilm formation and sporulation is most similar to genotype V, often associated with the
[81], as well as regulating competence [82,83]. In western North America diversity group of B. anthracis
B. anthracis, AbrB has been shown to negatively regu- [20]. ELISA experiments demonstrated that PA was
late toxin production [84,85]. While pBc10987 and present in the supernatant of a stationary phase culture,
pXO1 AbrB proteins are similar, they differ significantly however it is unclear which protective antigen subunit
at their N-terminus, in that the pXO1 protein is 27 ami- was recognized by this assay [20]. This high level of iden-
no acids shorter rendering it potentially inactive [84]. tity between pXO1 and pBCXO1 suggests that both iso-
The role of the pBc10987 encoded AbrB protein is cur- lates acquired their plasmid from a common ancestor or
rently unknown, but it is speculated that it acts as a reg- that transfer occurred into B. cereus G9241 from B.
ulator of the genes replacing the pXO1 pathogenicity anthracis. The successful transfer of pXO1 from B.
island [47]. anthracis to a close relative with the aid of a mobilizing
Besides AbrB, there are two more examples of possi- plasmid [92] supports the latter.
ble genetic exchange between the chromosome and the The regulators, PlcR and AtxA, have been shown to
plasmid of the B. cereus ATCC 10987. Identical copies be functionally incompatible in B. anthracis. The intro-
of a transposable element similar to the S. aureus duction of pXO1-encoded AtxA is thought to have led
Tn554 [86] are present on both the chromosome and to the selection for the plcR nonsense mutation in
plasmid. There are four genes associated with the B. anthracis [58]. Interestingly, B. cereus G9241 appears
Tn554-like element and one, bclA, is of interest. In B. to encode fully functional copies of both PlcR and
anthracis, multiple functions have been attributed to AtxA. However, PlcR contains a region of lower simi-
BclA. It is the major spore surface antigenic protein larity to B. anthracis PlcR that is thought to compensate
[87,88] and has been shown to play a role in the depth for the presence of AtxA or represent another poten-
determination of the spore coat [89]. Divergent proteins tially inactive form of PlcR. It is also hypothesized that
of the same family as BlcA have been found on the chro- the pBCXO1 plasmid may have been acquired recently
mosomes of B. anthracis and some plasmids of B. thur- and the incompatibility has not had time to generate a
ingiensis [90]. Interestingly, the bclA gene in B. anthracis nonsense mutation. In light of the discovery of this
is limited to the chromosome, whereas in B. cereus important organism, functional studies of B. cereus
ATCC 10987 similar copies of bclA are present on the G9241 regulatory pathways must be performed to deve-
chromosome and the plasmid. The presence of similar lop a better understanding of this evolutionary event.
genes on other plasmids suggests that genetic exchange B. cereus G9241 contains a second plasmid, pBC218
occurred between plasmids and chromosomes in the (218 kb), with limited similarity to a region in
B. cereus group of organisms. pZK467 (Fig. 3(b)). Intriguingly, pBC218 carries a sec-
ond copy of atxA, however the gene product is only
6.1.2.2. pBCXO1 and pBC218. B. cereus G9241 has 78% identical to the B. anthracis homologue [20]. In
been associated with an illness resembling inhalation addition, pBC218 encodes some of the known B. anthra-
anthrax [20]. Initial CGH studies revealed that B. cis virulence factors [20]. Genes encoding for a homo-
cereus G9241 contained genes with a high degree of logue of the protective antigen peptide (60% amino
hybridization to pXO1, including all three anthrax acid identity) and the lethal factor (36% amino acid
toxin genes, however, no hybridization was observed identity) are also found on pBC218, however the plas-
to pXO2. Unexpectedly, a 191-kb plasmid with a mid does not contain a homologue of the edema factor.
high degree of similarity and synteny to B. anthracis The lethal factor is significantly truncated when com-
pXO1 was found in B. cereus G9241 [20]. In addi- pared to that of B. anthracis pXO1 and is probably
tion, B. cereus G9241 contains a second 218-kb plas- not functional. In contrast, the pBC218-encoded protec-
mid, previously unidentified, that encodes a novel tive antigen peptide is identical to B. anthracis at all 10
polysaccharide capsule biosynthetic cluster. These dominant negative amino acid residues identified by
plasmids are expected to contribute greatly to the ob- Mourez et al. [93]. Another 33 residues were identified
served anthrax-like clinical presentation. that resulted in decreased activity when substituted by
The pXO1-like plasmid, named pBCXO1, is 99.6% cysteine, of which 27 are identical in pBC218-encoded
identical to pXO1. This similarity extends to the pXO1 protective antigen, with the remaining six being
316 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

viously thought that B. cereus, as a species in the group,


did not produce capsule [6]. Further examination of the
pBC218 sequence revealed a putative polysaccharide
biosynthetic operon, representing the only capsule bio-
synthetic cluster identified in the genome of B. cereus
G9241. This gene cluster is thought to be responsible
for the capsule produced by this isolate. The lack of
any other capsule suggests that this plasmid-encoded
polysaccharide capsule might compensate for the lack
of poly-c-D-glutamic acid capsule and help B. cereus
G9241 to evade the host immune system.
The plasmid content of this isolate represents a com-
posite of B. cereus and B. anthracis. It also raises inter-
esting questions regarding the regulatory interaction
between AtxA and PlcR, the assembly of the toxin
and the role of the novel capsule in pathogenesis.

6.1.2.3. pBClin15. A 15-kb linear plasmid, pBClin15,


was identified by genome sequencing of B. cereus ATCC
14579 [45]. This plasmid contains none of the typical
replication machinery associated with small plasmids
identified in either B. cereus or B. thuringiensis, but does
contain genes that suggest that it may represent a novel
phage. Recently, it has been reported that pBClin15
shows some level of similarity to the Bam35 phage of
B. thuringiensis var. alesti strain 35 [94,95]. Additional
comparisons and experimental data suggest that
Bam35 can exist both as a prophage and a free phage
[95]. In contrast to Bam35, pBClin15 lacks terminal in-
Fig. 3. Identification of regions with proteomic and syntenic similar- verted repeats that may allow protein-primed replica-
ities by comparisons of selected B. cereus group plasmids. Blast score tion, an observation that led to the suggestion that
ratio analyses were used to compare the putative plasmid proteomes as
described in Fig. 2. (a) Synteny plot of pBT9727 vs pXO2. The
pBClin15 is a degenerate phage [95].
pBT9727 plasmid contains a similar backbone structure to pXO2. The
replication proteins are conserved in plasmid location and peptide 6.1.3. B. cereus Zebra Killer (ZK) plasmids
sequence however pBT9727 lacks the poly-c-D-glutamic acid capsule The plasmids from B. cereus ZK represent a novel set
biosynthetic genes, indicated by the gray box. An additional unique of plasmids carried by an isolate that causes a disease
region of 10 kb is identified in pBT9727 by the gray box (22–33 kb).
48 of the 80 – pBT9727 peptides have homologues in pXO2.
resembling anthrax in animals (P.C.B. Turnbull, per-
(b) Comparison of pBC218 and pZK467 demonstrates that these sonal communication). Five distinct plasmids were iden-
plasmids have limited similarity with the exception of one region, tified in the genome of B. cereus ZK, representing the
highlighted in the gray box. This region is conserved in both gene largest number of plasmids harbored by a strain of
content and gene order. B. cereus so far. Interestingly, the three small plasmids
(less than 10 kb, Table 2) pZK5 (CP000041), pZK8
(CP000043) and pZK9 (CP000044) are more similar to
conservative substitutions [20,93]. The presence of these known B. thuringiensis plasmids than any previously se-
additional toxin subunits is raising some interesting quenced B. cereus plasmids, whereas the larger two plas-
questions about their putative function in the pathogen- mids, pZK54 (CP000042) and pZK467 (CP000040),
esis of this organism. Could the pBC218-encoded PA share characteristics with pXO2 and pBC218, respec-
form a complex with the pBCXO1 lethal factor and ede- tively. The three small plasmids are predicted to repli-
ma factor to form what would appear to be novel func- cate through a rolling-circle mechanism; their
tional anthrax toxin? How do these fragments interact functional annotation is mostly limited to the identifica-
and are they coordinately regulated? tion of replication and mobilization proteins. pZK54
India ink staining and microbiological analyses dem- (54 kb) encodes a pXO2-like putative theta replication
onstrated that B. cereus G9241 is encapsulated and that system, however no identifiable replication system was
capsule production is not regulated by increased CO2 found in pZK467 (466 kb), as in pXO1 [17,76].
concentrations like the B. anthracis pXO2-encoded pZK467 and pBC218 share an approximately 40-kb re-
poly-c-D-glutamic acid capsule. Interestingly, It was pre- gion where synteny is retained, however, no functional
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 317

assignments exist in this region (Fig. 3(b)). pZK467 is genes potentially involved in the biosynthesis and export
the only known plasmid with similarity to pBC218. A of a cyclic peptide antibiotic similar to Enterococcus fae-
large number of transposase genes or mobile elements calis AS-48 [48]. One can speculate that this peptide
have been identified in these two large plasmids. These could provide B. thuringiensis with a competitive advan-
elements may facilitate gene exchange between plasmids tage in the environment or act as a signalling molecule.
as well as between plasmids and the chromosome. None
of the B. cereus ZK plasmids encode homologues of 6.1.4.2. pBT9727. B. thuringiensis 97-27 has been shown
known virulence factors in B. anthracis, B. cereus or B. to produce crystal protein in sporulated culture by di-
thuringiensis [6]. Sequence analysis of B. cereus ZK plas- rect microscopic examination [54]. Detailed sequence
mids does not improve our understanding of the patho- analysis of pBT9727 (CP000047), the sole plasmid
genicity of this organism, which currently remains found in B. thuringiensis 97-27, did not reveal any
unclear. Analysis of the relative copy number of these genes encoding for the Cry toxin with similarity to al-
plasmids obtained from sequence analysis indicates that ready known B. thuringiensis toxin genes. Interestingly,
each has a copy number of 1–2, except for pZK5 which pBT9727 shows similarity to B. anthracis pXO2. Com-
is estimated at 0.6. These data might suggest that parison of the predicted coding regions of the two plas-
pZK5 is unstable under the laboratory culture condi- mids revealed that pBT9727 shares 89% (82/92) of its
tions employed and is in the process of being lost. putative coding sequences with pXO2. Their replication
proteins are almost identical and the predicted origin
6.1.4. B. thuringiensis plasmids of replication is well conserved (Fig. 4, cf. [80]). The le-
Phenotypic characterization of B. thuringiensis is of- vel of protein similarity, combined with the conserva-
ten based on the presence of plasmid-encoded large tion of gene order, suggests that these plasmids might
Cry protein inclusions of the d-endotoxin [48]. Recently, have diverged recently. Like pBC10987 and pXO1;
a study by Andrup et al. [90] described six B. thuringien- pBT9727 and pXO2 share a common backbone (Fig.
sis plasmids (less than 20 kb) that did not carry the cry 3(a)). The pXO2 region encoding for the poly-c-D-glu-
gene. It is possible to group B. thuringiensis plasmids tamic acid capsule biosynthetic genes is replaced on
based on the similarity of their replication and mobiliza- pBT9727 with genes encoding hypothetical proteins
tion machinery, however this system did not extend to and putative mobile elements. This replacement sug-
the larger B. thuringiensis or B. cereus plasmids. gests that pBT9727 might have evolved to fulfil other
functions than providing this isolate with capsule bio-
6.1.4.1. pBtoxis. In addition to the four known Cry and synthetic genes.
two known Cyt toxins, the 128-kb circular plasmid,
pBtoxis encodes a third Cyt-type sequence with an addi- 6.2. Replication and mobility mechanisms of the B. cereus
tional C-terminal domain previously unseen in such pro- group plasmids
teins [48]. B. thuringiensis subsp. israelensis carries
pBtoxis, and its toxin crystals have been demonstrated 6.2.1. Unidentified replicons
to be one of the most toxic combinations tested [96]. The following plasmids have no identified replication
However, it is unclear if this toxicity is related to the spe- system: pXO1, pBCXO1, pBC10987, pBC218, pBtoxis
cific combination of toxins or to toxin interaction with and pZK467 (Table 2). A number of hypotheses have
other plasmid encoded features. GC skew analysis indi- been put forward to account for this lack of clearly iden-
cated a putative origin of replication, however no repli- tifiable replication machinery, including that these plas-
cation proteins has been identified, similar to the other mids carry a novel replication mechanism [47]. Attempts
large B. cereus group plasmids (Table 2). The coding se- have been made to identify the regions involved in the
quences adjacent to this region in pBt001 showed >78% replication machinery of pXO1 using subcloning. Kas-
amino acid identity to pXO1-49, and is located near a par and Robertson [75] identified an 11 kb region that
similar putative replication origin on pXO1, also pre- is thought to play a role in pXO1 replication. However,
dicted by GC skew analysis. However, experimental this region does not encode genes with similarity to any
functional evidence is not available. other plasmid replication system known. One gene, a
In addition to the toxin genes, pBtoxis encodes a type I DNA topoisomerase was identified by compari-
number of genes that are thought to enhance crystal for- son of pBC10987 (BCEA0140, [47]) and pXO1
mation and subsequent cell viability by acting as chaper- (BXA0213, [41]). This protein appears to be conserved
ones. Interestingly, like pBC10987 and pXO1, pBtoxis in the following large B. cereus plasmids with an uniden-
encodes peptides that are involved in host sporulation tified replication system pBc10987, pOX1 and pBCXO1,
and germination [21,47]. The exact role of these proteins and is also found in pZK467 but where it is thought to
is unknown, however in B. anthracis the lack of these be non-functional as it is in three gene fragments
genes on pXO1 is detrimental [97]. One of the more sur- (pZK467_319-321). No homologue to this peptide is
prising findings in pBtoxis is the presence of a set of present in pBC218 or pBtoxis. In pBc10987, this type I
318 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

Fig. 4. Identification of the origin of replication structure in theta replicating plasmids of B. cereus group. (a) Alignment of the 60 nucleotide region
identified by Tinsley et al. [80] as the origin of replication. The gray box highlights a conserved region containing the origin, which is identified by the
arrow. Other conserved nucleotides in the region are indicated with an ‘‘*’’. The plasmids are from Bacillus anthracis pXO2 (GenBank Accession No.
AE017335), B. thuringiensis pBT9727 (GenBank Accession No. AE017335), Bacillus thuringiensis pAW63 (GenBank Accession No. AJ011655) and
Enterococcus faecalis pAMb1 (GenBank Accession No. AF007787). (b) The pXO2 replication region. The B. anthracis pXO2 iteron-binding region is
between the copy number control gene parA and the replication gene repS. Arrows on the nucleotide sequence identify the iteron-binding repeat,
‘‘ATGTGTAA’’. There are 10 direct repeats in the forward orientation and three in the reverse orientation in pXO2. There are also other degenerate
repeats in the region that are not indicated. The structure of the replication region including a similar repeated region is present in B. thuringiensis
pBT9727.

DNA topoisomerase may associate with a putative Many of the genes found on these plasmids have sig-
DNA polymerase III b-subunit thought to increase the nificant similarity with chromosomally encoded genes
processivity of replication [98]. Unfortunately, this sec- from other members of the B. cereus group, B. subtilis,
ond gene is not present in any of the other large B. cer- Bacillus halodurans and other low G+C Gram-positive
eus group plasmids. It may participate in pBc10987 species such as Listeria and Staphylococcus. Addition-
replication but does not appear to be essential for the ally, the GC skew, gene organization and gene orienta-
replication of all members of this group of plasmids. An- tion bias of these large plasmids appear more similar
other plasmid-borne replication gene candidate is a to that of chromosomes. In line with the chromosomal
host-factor-like protein (BCEA0146 – pBc10987; reduction theory, one can speculate that the replication
BXA0206 – pXO1; pZK467_0115 – pZK467, ORFBT116 machinery for these plasmids is actually chromosom-
– pBtoxis). These peptides have nucleotide binding do- ally encoded and only the actual origin of replication
mains and are similar in all large B. cereus group plas- is present on the plasmids themselves. Recent work
mids lacking an identifiable replication system. While has identified a chromosomally encoded helicase in
it is unlikely that this peptide accounts for the entire rep- the B. cereus group of organisms similar to that of
lication machinery, it is a conserved plasmid-encoded Staphylococcus aureus plasmid pT181 [99,100]. This
peptide. helicase, PcrA, functions as a nickase and can initiate
One alternate hypothesis is that B. cereus plasmids replication of pT181. The B. anthracis PcrA has been
with no identifiable replication system are actually shown to function as a helicase and initiate replication
the result of reduction of an ancestral chromosome. of pT181. It is thought that the B. cereus homologues
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 319

could perform similar function on plasmids that do


not have a clearly recognizable replication system.
Additionally, such a system would explain why the
replication regions of these plasmids have gone
undiscovered.

6.2.2. Theta replicons


Based on similarity to other replication proteins
identified from plasmids in the B. cereus group and
other Gram-positive organisms, the origin of
replication was readily identified in pXO2 [101,102].
While well characterized in B. thuringiensis plasmid
pAW63, the pXO2 origin of replication was not func-
tionally characterized until recently [80]. Tinsley et al.
[80] demonstrated that the minimal replicon comprises
solely of the RepS peptide (BXB0039) and the origin
of replication. In addition, the functional origin of
replication was shown to be limited to a 60 bp region
to which RepS specifically binds (Fig. 4(a)). Further-
more, RepS was shown to bind to the single stranded
form of this region, providing further evidence that
RepS is responsible for the initiation of replication.
pBT9727 origin of replication (59/60 nucleotides) and
RepS are similar (91% amino acid identity) to that
of pXO2 indicating that pBT9727 belongs to the theta
replicating plasmid family. On the other hand, plas-
mid pZK54 is tenuously categorized as a theta repli-
cating plasmid based on similarity of the replication
associated protein, pZK54_001. pZK54 lacks the Fig. 5. Comparison and clustering of B. cereus group plasmids based
on replication and mobilization proteins. Replication or mobilization
mobilization protein and the 60 nt origin of replica- proteins, as identified by annotation, were aligned and compared using
tion identified by Tinsley et al. [80], suggesting that CLUSTALW. The unrooted neighbor-joining phylogenetic trees were
a putative novel mechanism might be involved in generated amd displayed with Tree View. Theta-replicating plasmid
pZK54 maintenance and mobility. proteins are within the blue ovals, Rolling circle replication plasmids of
All theta replication proteins from B. cereus group group IV are in the green ovals and only two members of the group III
family replication proteins could be identified for inclusion.
plasmids share a significant level of similarity and cluster (a) Replication proteins. (b) Mobilization proteins.
together phylogenetically suggesting a common ances-
tral origin (Fig. 5(a)).
6.2.4. Plasmid mobility
While the B. anthracis plasmids have not been di-
6.2.3. Rolling circle replicons rectly shown to be self-transmissible, previous reports
All B. cereus group small plasmids (<10 kb) appear to have demonstrated that some of them can be mobilized
replicate through a rolling circle mechanism. Sequence with the help of conjugative plasmids [74]. B. thuringien-
analysis and comparison of all B. cereus replication pro- sis subsp. israeliensis pXO16 is an example of such a
teins, including theta and rolling circle, allow for cluster- conjugative plasmid [74,103]. Interestingly, no transfer
ing of these plasmids into distinct groups (Fig. 5). This or mobilization regions are identifiable in the sequence
comparison includes type sequences for each plasmid of B. cereus large plasmids. In contrast, the mobilization
group as determined by Andrup et al. [90]. All small proteins encoded on the smaller B. cereus ZK and other
rolling circle plasmid replication proteins clustered B. thuringiensis plasmids suggests that they may be self-
together and are clearly distinct from theta replication mobilizable but, lack the ability to create pores to trans-
protein sequences. pZK8 and pZK9 both clustered with fer themselves to the recipient cells. It is possible that the
group VII plasmids as defined previously [90]. pZK5 is combination of the mobilization capabilities of one plas-
also a group VII member but might represent an outlier mid and an unidentified membrane associated transfer
to this group. Classification of the small rolling circle system from another plasmid will allow transfer of
replicating plasmids by this sequence analysis is a robust both/all plasmids from the donor cell to a willing recipi-
and effective way to identify the mode of replication for ent cell. The B. cereus large plasmids harbor candidate
such plasmids. pore formation genes, such as the TraD/G conjugation
320 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

proteins, that could form a pore through a membrane of bacteriophage is beginning to be appreciated (for re-
complex however their functionality has not been dem- cent reviews, see [109,110]).
onstrated experimentally. Sequencing of mobilizable B. Genes carried by bacteriophage encode proteins that
cereus group plasmids such as pOX11 [92], pXO12 [92] modulate lysogenic conversion of the host and may pro-
or pXO16 [74,104] would advance our understanding vide a selective advantage [111]. The prophage may re-
of plasmid transfer mechanisms in this group of organ- main competent to enter a lytic cycle in response to
isms. A number of limited studies have been undertaken appropriate induction signals or may remain quiescent
that demonstrate that B. cereus group plasmid transfer for generations. Over time, the prophage may acquire
is not affected by DNase, involves membrane interaction mutations that preclude their re-entry into a lytic cycle
[74] and in some cases employs an Ôaggregation sub- of growth, while genes originally carried by the virus
stanceÕ [104], all which suggest a conjugative transfer can remain active in the host extending the lysogens
mechanism. niche over non-lysogenic isogenic strains. In fact, there
Based on sequence analysis of the plasmids of the are several examples of phage-derived sequences repre-
B. cereus group of organisms, it is evident in some cases senting in excess of 10% of a sequenced bacterial gen-
that the significant differences observed in pathogenicity ome [109]. Phage can also facilitate horizontal gene
and host range are often dictated by genes carried on transfer and promote genomic rearrangements, a factor
plasmids. As suggested for Yersina pestis [105], where that has contributed to the emergence of bacterial
pathogenicity and host range is determined by the plas- pathogens (for review, see [112]). This is most clearly
mid content of an isolate, the designation of ‘‘plasmid- demonstrated in a comparison of the genomes of two
ovar’’ would be applicable to the B. cereus group of E. coli stains, the benign K12 and the virulent
organisms, such that B. anthracis would be B. cereus pla- O157:H7 where significant genetic differences between
smidovar anthracis. these two strains can be attributed to differences in pro-
phage content [113].
While the best-characterized phage in the Gram-
7. The phage of the B. cereus group positive Bacillus genera are from B. subtilis, similar
prophages have been identified in all members of the
Besides plasmids, bacteriophages are another impor- B. cereus group. We recognize that the identification
tant source of gene flow in bacteria. Bacteriophages are of prophage in bacterial genomes is non-trivial
viruses that infect bacteria. Bacteriophage can be either [109,110]. Consequently, the current estimates of the
lytic, redirecting cellular processes for the sole purpose prophage contribution to the genomic content of the
of producing additional virus progeny, or temperate, members of the B. cereus group should be regarded as
insetting themselves into the host bacterium genome conservative and tentative.
where they are transmitted ‘‘benignly’’ from generation
to generation in concert with a host replicon. Bacterio- 7.1. Genomic phage content
phages have an extremely high degree of host specificity
and may integrate as a prophage into a unique site or Each of the 10 B. anthracis genome sequences contains
multiple locations in the host genome. This specificity four prophages located in the same genomic location.
has been exploited as typing feature [106] and proposed These prophages have been designated LambdaBa01-04
as a possible therapeutic intervention in the treatment of [40]. The B. anthracis prophages appear to be unique to
bacterial diseases [107,108]. As an increasing number of B. anthracis, however, prophage LambdaBa01 is at least
bacterial genomes are being sequenced, more prophage partially present in B. cereus ZK (Fig. 6). Interestingly,
are being identified and the distribution and diversity while the gene content appears to be conserved the

Fig. 6. Blast similarities represented using Blast score ratios and visualized with TIGR MEV (multiple expression viewer). Each vertical line
represents a unique gene in one of the predicted B. anthracis prophage and each horizontal column represents a different isolate. Yellow regions
indicate similarity, whereas the black regions are dissimilar. This figure demonstrates that only B. cereus Zebra Killer contains one of the phage from
B. anthracis. The B. anthracis phage are not generally conserved among this group and neither are the B. cereus or B. thuringiensis phage (data not
shown).
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 321

relative genomic location is different. Whether these dif- 7.3.1. Use as a typing tool for B. anthracis (Gamma and
ferences represent a subtle variation in species-specific Cherry phage)
phage integration sites, or provide an evolutionary in- While there are four prophages in each of the B.
sight into the relationship between B. cereus ZK and anthracis genomes sequenced to date, clinical laborato-
the B. anthracis are important questions that have not ries utilize, in combination with a number of biochemi-
yet been resolved. cal tests, cell lysis with Gamma phage as a phenotypic
Three prophages have been identified in the genome characteristic in the identification of B. anthracis [106].
sequence B. cereus ATCC 10987, while B. cereus Sensitivity to Gamma phage remains highly specific to
ATCC 14579 contains six putative integrated pro- B. anthracis, however B. cereus ATCC 4342 is also sen-
phages and a linear non-integrated phage, designated sitive to Gamma phage. As more isolates closely related
pBClin15. B. cereus G9241 contains a not well-charac- to B. anthracis are being isolated with pathogenic char-
terized cryptic phage of 29,886 bp (pBClin29, [20]) that acteristics, Gamma phage sensitivity among B. cereus
encodes phage-like proteins and a plasmid replicon isolates will become increasingly common and will re-
similar to B. anthracis pXO2. Interestingly, the B. thur- duce the discriminating power of this assay to differenti-
ingiensis 97-27 genome appears to contain as many as ate B. anthracis from B. cereus.
10 prophages on the chromosome and two prophage
on the pBT9727 plasmid. Including the B. anthracis 7.3.2. Use of phage lysis protein as a biological control
LambaBa01-like prophage, B. cereus ZK genome may While the complete sequence for Gamma phage is
have as many as 27 chromosomally encoded pro- not yet available, its lysis protein, PlyG, has been
phages, based on the occurrence of distributed phage cloned, sequenced and characterized [115]. The lytic
genes, and an additional nine in plasmid pZK467, activity appears to be restricted to B. anthracis, how-
and four in pZK54. None of the putative prophage ever the specificity of this activity has not been fully
integration sites in B. thuringiensis 97-27 or B. cereus elucidated. It has been suggested that the lysin pro-
ZK appear to be associated with tRNA genes as is of- teins from the lysogenic phages of B. anthracis could
ten the case [114]. These high numbers of prophages be utilized as a method for biological control. The
may be an overestimate as some consist of only a specificity for B. anthracis of such proteins has been
few genes with some degree of similarity to other previously demonstrated [115] and used for clinical
phages. typing of B. anthracis (see above). This activity may
also be applied to specifically lyse B. anthracis in other
7.2. Lytic induction situations, such as treatment of infections. The possi-
bility of using these proteins as biological control
An important test of prophage competence is the mechanisms holds promise, but cross reactivity with
ability to initiate a lytic cycle by induction. We have certain B. cereus strains could be a problem [115]
demonstrated the competence of several prophage to and, more importantly, the lytic activity of these pro-
induce a lytic cycle using a standard mitomycin C teins is limited to vegetative cells. These proteins do
treatment from a number of B. cereus and B. thuringi- not affect B. anthracis spores, the infectious particles.
ensis strains (Rasko, unpublished). Not surprisingly, it A strategy has been suggested that employs the use
was shown that the host-range of these phages is re- of spore germinant solutions to induce the outgrowth
stricted to the species or strain from which they were of vegetative forms in combination with a lysin solu-
obtained. However, there is one notable exception, tion to destroy emergent vegetative cells. The rapid
strain B. cereus ATCC 4342 can be used to propagate concerted conversion from spores to vegetative forms
the Gamma and Cherry phages that were previously (less than 10 min) [116] suggests that this approach
thought to be limited to B. anthracis strains [106,115]. could be applied successfully. Both the germinant
The factors responsible for the expansion of the phage and lysin solution should be harmless enough to both
range are currently under investigation. Finally, it is man and machine to use not only in the case of a bio-
interesting to note that typing with MLST places B. terror attack, but also as prophylaxis for the treatment
cereus ATCC 4342 in the phylogenetic clade that con- of troops and machines returning from anthrax con-
tains B. anthracis and may represent a close evolution- taminated areas. Additionally, since the phage lysin
ary relationship (Fig. 1). proteins are thought to attack the basic building
blocks of the cell wall, it would be unlikely that resis-
7.3. Exploitation of phage tance to these methods will develop as rapidly as anti-
biotic resistance. Interestingly, no equivalent system
Given the high degree of specificity and limited host using phages has been proposed or examined for B.
range of bacteriophage, it seems quite reasonable to cereus or B. thuringiensis, certainly due to the fact that
use them and their encoded activities as a means of a phage would be specific to one strain but not for the
strain identification and infection control. entire group.
322 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

8. Functional genomics of the B. cereus group of [40]. B. cereus isolates from clinical sources contained
organisms more pXO1 and/or pXO2 genes than non-clinical iso-
lates, suggesting a pathogenic role for plasmid-encoded
Since the publication of the first B. cereus group genes. While this technique does not allow the determi-
organism genomes [40,45], a significant number of stud- nation of gene order, it confirms the presence of
ies have taken advantage of the sequence to generate plasmid-like genes in these species. Molecular examina-
functional data. Not surprisingly B. anthracis has been tion of some of these clinical isolates has revealed the
the focus of much of this work. The mail anthrax at- presence of large plasmids (>200 kb) that are currently
tacks of 2001 provided an additional impetus for re- under investigation (Rasko, unpublished data). Using
search as well as an infusion of funding for the the ratios obtained for chromosomal genes, it was pos-
development of microarray and proteomic research sible to reconstruct the phylogenetic relationship of
programs. these 19 isolates – the clustering obtained was consis-
tent with that obtained using other methods [3,8,27–
8.1. Transcriptional analysis using microarrays 29,32,33].
In addition to CGH experiments, this array has also
8.1.1. B. cereus/B. thuringiensis used for expression analysis in two separate studies. The
The availability of B. cereus specific arrays by a lim- expression of plasmid-encoded virulence factors and
ited number of companies (NimbleGen and Qiagen) has genes that are involved with their regulation were ana-
not yet generated any published data. Additionally, a lyzed. Comparison of expression patterns for wild type
commercially available B. thuringiensis specific array and B. anthracis AtxA knockout mutants demonstrated
has not been advertised. While these species-specific ar- that AtxA is the dominant regulatory protein and ap-
rays are useful, most information would derive from an pears to be a master virulence regulator in B. anthracis
array that contains the B. cereus core genomic elements [71]. AtxA was shown to regulate AcpA and AcpB,
as well as the unique genes from a number of species or which were previously thought to regulate pXO2-
isolates. The advantage of building such a chimeric ar- encoded capsule biosynthesis.
ray would be to capture the level of conservation of B. anthracis gene expression patterns were analyzed
any interrogated isolate in relation to the core genotype, as the bacterium progressed through logarithmic phase
but it would also detect the level of divergence and pos- and entered sporulation [72]. In contrast to the previ-
sible gene transfer between isolates/species in compara- ous study, Liu et al. did not make use of mutational
tive genomic hybridization (CGH) experiments. In analysis to examine the regulation but rather took a
addition to obtaining a metric for the level of conserva- less invasive method to examine gene expression as
tion, the chimeric array will allow representation of mul- growth progressed to unravel the regulatory cascade
tiple strains, isolates or species on a single array without associated with spore formation. Five distinct waves
the need for redundant coding sequences to be repre- of gene expression containing 36% of all predicted
sented, thus saving space, time, effort and money. B. anthracis genes were observed. The five waves
roughly corresponded to early, mid or late logarithmic
8.1.2. B. anthracis growth, stationary phase growth and spore formation.
A B. anthracis Ames microarray (both 70-mer oligo- The data revealed that the sporulation program in B.
nucleotide and amplicon-based) is available through anthracis is similar to the well-characterized B. subtilis
the NIAID funded Pathogen Functional Genomics Re- system [119]. However, B. anthracis sigma factor A
source Center (http://pfgrc.tigr.org/). Other arrays have (SigA) expression patterns deviate from those of its
been produced and used for species typing [117] and B. subtilis counterpart. sigA is induced during the final
resequencing [118]. Read et al. [40] utilized an ampli- stages of sporulation in B. anthracis whereas in B. sub-
con-based array as a comparative CGH tool to interro- tilis SigA is produced during vegetative growth [119].
gate the genomic content of 19 diverse B. cereus group Additional examination of the expression data demon-
organisms. This diverse set included a number of the B. strated that the plasmid-encoded virulence factors were
cereus strains from clinical sources, mostly periodontal expressed only in early logarithmic phase. The true
infections from Norway [3,28,62] as well as strains power of this study is that the expression data were
from environmental sources and common laboratory combined with a proteomic analysis of the sporulating
strains including B. cereus ATCC 14579 and B. cereus cultures [72].
ATCC 10987 [45,47]. This study revealed a high degree
of genomic similarity among these 19 B. cereus isolates. 8.2. Proteomics studies
In addition, it indicates the presence of pXO1 gene
homologues in half of the 19 strains examined, consis- 8.2.1. B. anthracis
tent with other studies [63]. In contrast, very few The genome sequence and its predicted proteome
homologues of pXO2 genes were found to hybridize have allowed for high-throughput proteomic analyses
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 323

where proteins are enzymatically digested and the result- genic targets. Five of the eight candidates were
ing peptides analyzed with two-dimensional liquid chro- preselected through bioinformatic analysis of the gen-
matography coupled with tandem mass spectrometry ome sequence of B. anthracis [124]. This study highlights
analysis. Peptide molecular weights are then matched that a combination of genomic data, bioinformatics
bioinformatically to specific proteins. Combining this analysis and proteomics can contribute to novel anthrax
approach with expression analysis in a single study high- vaccine development.
lights the complementary nature of genomic and proteo-
mic analyses. Liu et al. [72] applied such an approach to 8.2.2. B. cereus
the study of spore formation and were able to rapidly Unlike B. anthracis, proteomic analyses were applied
document high-resolution temporal changes in gene to B. cereus vegetative cells and not spores. B. cereus
expression associated with spore formation, while pro- PlcR mutant strains were compared to the wild type
teomic analysis provided a detailed snapshot of the pro- using protein 2D gel electrophoresis [54]. As expected,
tein content and relative abundance in the spore. the inactivation of plcR in most cases abolished or sig-
Surprisingly, genes that were upregulated in the final nificantly reduced the expression of PlcR-regulated pep-
stages of spore formation were rarely identified by the tides. These peptides included a number of virulence
proteomic analysis as being constituents of the spore. factors known to be under the control of PlcR, such
Of the 873 genes identified to be upregulated in the final as collagenase, hemolysins, proteases and toxins. Those
stages preceding sporulation, only 173 were identified directly regulated by PlcR were most significantly af-
proteomically as being spore components [72]. This fected, whereas those indirectly regulated were reduced
suggests that the sporulating cells obtain proteins from in expression levels.
pre-existing peptide pools to form the spore. If the Detailed 2D gel electrophoresis proteomic analysis
microarray and proteomic data had not been combined of B. cereus biofilm establishment and maintenance
into a single study, a number of incorrect conclusions as on glass wool has been reported [125,126]. In the
to the temporal origin of the proteins in the spore could dairy industry, B. cereus biofilms inside storage tanks
have been made. and associated piping can create problems for product
While the study by Liu et al. [72] demonstrated the sterility and longevity. Using a proteomic approach,
power of using proteomic analysis in a high-throughput B. cereus isolates in biofilms have been shown to ex-
mode, other proteomic studies are focusing on rapid and press at least 10 additional proteins and lack seven
accurate speciation of B. anthracis or B. cereus based on proteins during adhesion to a solid surface and estab-
the proteomic content of their spore employing a com- lishment of the biofilm [125]. These peptides may play
bined 2D-gel electrophoresis and tandem mass spec- important roles in the maintenance of the biofilm and/
trometry approach [120]. These studies have been or represent metabolic changes triggered when the
successful in differentiating spores and vegetative forms bacterium switches from a planktonic to a sessile
of B. anthracis from its close relatives [120,121]. In addi- lifestyle.
tion to examining and differentiating the Bacillus spe-
cies, these proteomic studies have identified potential 8.2.3. B. thuringiensis
new spore coat proteins not previously annotated as While there are few published B. thuringiensis proteo-
such or known to reside in the membrane or spore mic studies, significant progress have been made in the
surface [122]. identification of sporulation regulatory pathways using
Anthrax vaccine preparations have also been charac- MALDI-TOF (matrix-assisted laser desorption/ioniza-
terized through proteomic analysis. A recent study using tion time of flight) analysis [127]. These pathways are
2D gel-electrophoresis [123] confirmed that the major a driver of B. thuringiensis biology, as the spore is the
constituent of the B. anthracis vaccine was the protective bacterial form used as a biopesticide. Other proteomic
antigen, but identified a number of minor contaminating studies have focused on the identification of the Cry tox-
products. These minor constituents included proteolytic in receptor in the Manduca sexta midgut using 1D and
cleavage products of the protective antigen, lethal factor 2D gel electrophoresis [128]. These different steps are
and edema factor as well as other putative virulence fac- essential in the life cycle of B. thuringiensis – proteomics
tors such as EA1, Sap and S-layer proteins. Addition- could have a major impact in developing a better under-
ally, cell constituents could be identified such as standing of bacterial virulence mechanisms, both by
60 kDa heat-shock protein, enolase, fructose-bisphos- studying the host and the bacterium itself.
phate aldolase and nucleoside diphosphate kinase [123]. Overall, proteomics and microarray studies have
Proteomics can also be applied to examine the spore opened the opportunity to leverage the genomic informa-
or bacterial surface for novel vaccine candidates. 2D gel tion of the members of this group. These studies have
electrophoresis of the spore proteome probed with sera enabled a better understanding of spore formation
from immunized animals identified eight in vivo immu- regulatory cascades, identified unknown regulatory ele-
nogens, six of which were previously reported as anti- ments and pathways, and have started to examine the
324 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

interactions of the B. cereus group of organisms with role can be associated with any of the harbored plas-
their respective environments. mids. Interestingly, clustering of these plasmids based
on the similarity of their replication systems is possi-
ble, offering an alternative method of systematic
9. Summary: One species on the basis of genomic classification.
evidence? Certainly our analyses demonstrate a bias toward
significant pathogens of the group. However, not all
Attempts at bacterial systematics began long before members of the group are highly virulent. While 15
the discovery of DNA as the hereditable material. In genome sequences are publicly available for a diverse
fact, the seminal discovery demonstrating that DNA set of B. cereus group organisms, genome data analy-
was the genetic material, through the transformation sis has not delivered in developing a better under-
of avirulent pneumococci to the virulent form [129] standing of the mechanisms that contribute to, and
hinted at the importance of horizontal gene transfer as limit the acquisition of virulence. Access to such infor-
a mechanism to increase genetic diversity and thereby mation will help to further refine the definition of a
niche expansion by a microbial species. Bacteria were species.
originally classified largely on the basis of phenotype, Technological advances in functional genomics and
morphology, ecology and/or associated disease state. proteomics are giving scientists the opportunity to lever-
For example, the bacteria that are the subject of this re- age genome sequence data. From the analysis presented
view were named Bacillus for their rod-shape, cereus for here and those of others, it is apparent that gene content
presumably an association with cereal crops, and might play a decreased role in the diversity of phenotype
anthracis as the cause of the disease anthrax. The criteria and pathogenicity observed in the B. cereus group. Subtle
for the ÔspeciesÕ designation have been widely debated, changes to regulatory networks may be responsible for
however the tenet of reproductive isolation is largely the range of phenotypic traits displayed by the B. cereus
unassailable. This principle seemed inviolate in asexually group members. Functional studies and genetic linkage
reproducing bacteria until the demonstration of ‘‘fertil- experiments, combined with proteomic analysis should
ity’’ in Escherichia coli in the early 1950s [130,131]. provide a better understanding of these regulatory path-
The subsequent discovery that antibiotic resistance ways and their genetic basis. In addition, using these tech-
could be disseminated between different species within nologies combined with access to the genomes of the
the Enterobacteriacae via plasmids [132] caused recon- human, pathogenic and non-pathogenic members of
sideration of the basis of bacterial classification. the group, it is now possible to study the interaction of
Classical bacterial systematics is now being chal- the bacterium with its host. This may allow the emer-
lenged by discoveries being promulgated by the gence of a detailed molecular understanding of the path-
genomics revolution. The defining characteristics of a ogenicity of this group of organisms that will benefit the
species must be grounded in its genetic/genomic archi- development of treatment and prevention measures.
tecture. In the B. cereus group of organisms, virulence The evolution of B. anthracis as a highly successful
and pathogenicity appear to be promiscuous and pathogen causing a readily identifiable disease has led
spread with plasmids. As reviewed in this paper, it to its over-representation in culture collections of iso-
is clear that the bacterial chromosomes of the se- lates from the B. cereus group. Thus, should B. anthracis
quenced members of this group are extremely similar. be considered an oversampled B. cereus? Based on geno-
These chromosomes show a high level of synteny and mic analyses of representatives of the group, these iso-
protein identity, a combination never observed be- lates carry different plasmids in a similar genetic
tween different bacterial species. Furthermore, there background. Only subtle differences in gene content
is evidence for a shared set of core putative virulence and protein similarities are observed when the chromo-
factors between different pathogenic and non-patho- somes of members of the group are compared. While it
genic members of the group. Very few chromosomal is true that B. anthracis can be readily differentiated
genes or sets of genes are unique to one species. These from B. cereus based on biochemical tests [33], a Ôprob-
regions often represent metabolic adaptations, and in lemÕ still exists for borderline isolates such as the patho-
many cases are found in another species with different genic B. cereus G9241 [20], where these tests fail to
phenotypic and pathogenic traits. Their presence can- recognize the pathogenic potential of this isolate. The
not be associated with a specific subset of organisms. limited definition of the nature of B. anthracis empha-
Conversely, much of the disease and host specificity in sizes the challenges faced by public health officials when
this group can be attributed to plasmid content (i.e., confronted with non-B. anthracis pathogenic isolates,
pXO1 and pXO2 in B. anthracis). The pXO1-like plas- and their inability to identify them as such [20]. The gen-
mid in B. cereus G9241 has been hypothesized to be ome of B. cereus G9241 would be indistinguishable from
responsible for the anthrax-like clinical presentation these of any other B. cereus isolates, if it did not contain
caused by this isolate. However, in other isolates no a homologue of B. anthracis pXO1. Classification of the
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 325

members of the B. cereus group of organisms has long [5] Orduz, S., Restrepo, W., Patino, M.M. and Rojas, W. (1995)
been the source of controversy. In 1952, based on the Transfer of toxin genes to alternate bacterial hosts for
mosquito control. Memorias Do Instituto Oswaldo Cruz 90,
observation that an isolate of B. anthracis had lost both 97–107.
virulence, and its plasmids, and was indistinguishable [6] Schnepf, E., Crickmore, N., Van Rie, J., Lereclus, D., Baum, J.,
from B. cereus in term of pathogenicity, Smith et al. Feitelson, J., Zeigler, D.R. and Dean, D.H. (1998) Bacillus
[133] concluded ‘‘B. anthracis is taxonomically a patho- thuringiensis and its pesticidal crystal proteins. Microbiology and
genic variety of B. cereus’’. The bacterium was later Molecular Biology Reviews 62, 775–790.
[7] Jensen, G.B., Hansen, B.M., Eilenberg, J. and Mahillon, J.
listed as B. cereus var anthracis [134], a nomenclature (2003) The hidden lifestyles of Bacillus cereus and relatives.
questioned ever since [37,38]. While genome sequence Environmental Microbiology 5, 631–640.
analysis has not elucidated the pathogenic potential of [8] Helgason, E., Caugant, D.A., Lecadet, M.M., Chen, Y.H.,
some isolates of the group, it has offered unprecedented Mahillon, J., Lovgren, A., Hegna, I., Kvaloy, K. and Kolstø,
insights into the core of their making. Based on the A.B. (1998) Genetic diversity of Bacillus cereus and B. thurin-
giensis isolates from natural sources. Current Microbiology 37,
chromosome genomic comparison reviewed in this pa- 80–87.
per, it is not possible to distinguish members of the B. [9] Hernandez, E., Ramisse, F., Cruel, T., le Vagueresse, R. and
cereus group from one another. Should they be consid- Cavallo, J.D. (1999) Bacillus thuringiensis serotype H34 isolated
ered one species based on genomic evidence? It is recog- from human and insecticidal strains serotypes 3a3b and H14 can
nized that for economical and social reasons, changing lead to death of immunocompetent mice after pulmonary
infection. FEMS Immunology Medical Microbiology 24, 43–47.
the current nomenclature would be quite a challenge. [10] Hernandez, E., Ramisse, F., Ducoureau, J.P., Cruel, T. and
It is hoped that this paper may represent a starting Cavallo, J.D. (1998) Bacillus thuringiensis subsp. konkukian
point for discussion and further novel studies through (serotype H34) superinfection: case report and experimental
focusing on the similarities and distinctions that contrib- evidence of pathogenicity in immunosuppressed mice. Journal of
ute to the nature of this group of bacteria. Clinical Microbiology 36, 2138–2139.
[11] Hernandez, E., Ramisse, F., Gros, P. and Cavallo, J. (2000)
Superinfection by Bacillus thuringiensis H34 or 3a3b can lead to
death in mice infected with the influenza A virus. FEMS
Immunology Medical Microbiology 29, 177–181.
Acknowledgments [12] Logan, N.A. and Turnbull, P.C. (1999) (Murray, P.R., Ed.),
Manual of Clinical Microbiology, pp. 357–369. American
D.A.R. and J.R. are supported with Federal funds Society for Microbiology, Washington, DC.
from the National Institute of Allergy and Infectious [13] Drobniewski, F.A. (1993) Bacillus cereus and related species.
Disease, National Institutes of Health, under Contract Clinical Microbiology Reviews 6, 324–338.
[14] Margulis, L., Jorgensen, J.Z., Dolan, S., Kolchinsky, R., Rainey,
No. N01-AI15447. M.R.A. and C.S.H. are supported F.A. and Lo, S.C. (1998) The arthromitus stage of Bacillus
by the US Department of Energy Contract W-7405- cereus: intestinal symbionts of animals. Proceedings of the
ENG-36 and LAUR#04-8482. We are grateful to Garry National Academy of Sciences of the United States of America
Myers, Thomas Brettin and Paul Jackson for their com- 95, 1236–1241.
ments during the preparation of this manuscript and to [15] A.B. Kolstø, D. Lereclus, M. Mock, Genome structure and
evolution of the Bacillus cereus group, in: Pathogenicity Islands
Gary Xie for his technical assistance. and the Evolution of Pathogenic Microbes, vol. 2, 2002, pp. 95–
108.
[16] Jernigan, D.B., Raghunathan, P.L., Bell, B.P., Brechner, R.,
Bresnitz, E.A., Butler, J.C., Cetron, M., Cohen, M., Doyle, T.,
References Fischer, M., Greene, C., Griffith, K.S., Guarner, J., Hadler, J.L.,
Hayslett, J.A., Meyer, R., Petersen, L.R., Phillips, M., Pinner,
[1] Chen, M.L. and Tsen, H.Y. (2002) Discrimination of R., Popovic, T., Quinn, C.P., Reefhuis, J., Reissman, D.,
Bacillus cereus and Bacillus thuringiensis with 16S rRNA Rosenstein, N., Schuchat, A., Shieh, W.J., Siegal, L., Swerdlow,
and gyrB gene based PCR primers and sequencing of D.L., Tenover, F.C., Traeger, M., Ward, J.W., Weisfuse, I.,
their annealing sites. Journal of Appl. Microbio. 92, 912– Wiersma, S., Yeskey, K., Zaki, S., Ashford, D.A., Perkins, B.A.,
919. Ostroff, S., Hughes, J., Fleming, D., Koplan, J.P. and Gerber-
[2] Daffonchio, D., Cherif, A. and Borin, S. (2000) Homoduplex ding, J.L. (2002) Investigation of bioterrorism-related anthrax,
and heteroduplex polymorphisms of the amplified ribosomal United States, 2001. Epidemiologic Findings 8, 1019–1028.
16S–23S internal transcribed spacers describe genetic relation- [17] Jernigan, J.A., Stephens, D.S., Ashford, D.A., Omenaca, C.,
ships in the ‘‘Bacillus cereus group’’. Applied and Environmental Topiel, M.S., Galbraith, M., Tapper, M., Fisk, T.L., Zaki, S.,
Microbiology 66, 5460–5468. Popovic, T., Meyer, R.F., Quinn, C.P., Harper, S.A., Fridkin,
[3] Helgason, E., Økstad, O.A., Caugant, D.A., Johansen, H.A., S.K., Sejvar, J.J., Shepard, C.W., McConnell, M., Guarner, J.,
Fouet, A., Mock, M., Hegna, I. and Kolstø, A.B. (2000) Bacillus Shieh, W.J., Malecki, J.M., Gerberding, J.L., Hughes, J.M. and
anthracis, Bacillus cereus, and Bacillus thuringiensis – One species Perkins, B.A. (2001) Bioterrorism-related inhalational anthrax:
on the basis of genetic evidence. Applied and Environmental the first 10 cases reported in the United States. Emerging
Microbiology 66, 2627–2630. Infectious Diseases 7, 933–944.
[4] Lecadet, M.M., Frachon, E., Dumanoir, V.C., Ripouteau, H., [18] Turnbull, P.C.B. (2002) Introduction: anthrax history, disease
Hamon, S., Laurent, P. and Thiery, I. (1999) Updating the H- and ecology. Anthrax 271, 1–19.
antigen classification of Bacillus thuringiensis. Journal of Applied [19] Mock, M. and Fouet, A. (2001) Anthrax. Annual Review of
Microbiology 86, 660–672. Microbiology 55, 647–671.
326 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

[20] Hoffmaster, A.R., Ravel, J., Rasko, D.A., Chapman, G.D., fragment length polymorphism analysis of Norwegian Bacillus
Chute, M.D., Marston, C.K., De, B.K., Sacchi, C.T., Fitzgerald, cereus and Bacillus thuringiensis soil isolates. Applied and
C., Mayer, L.W., Maiden, M.C., Priest, F.G., Barker, M., Jiang, Environmental Microbiology 67, 4863–4873.
L., Cer, R.Z., Rilstone, J., Peterson, S.N., Weyant, R.S., [34] Cherif, A., Brusetti, L., Borin, S., Rizzi, A., Boudabous, A.,
Galloway, D.R., Read, T.D., Popovic, T. and Fraser, C.M. Khyami-Horani, H. and Daffonchio, D. (2003) Genetic
(2004) Identification of anthrax toxin genes in a Bacillus cereus relationship in the ÔBacillus cereus groupÕ by rep-PCR
associated with an illness resembling inhalation anthrax. Pro- fingerprinting and sequencing of a Bacillus anthracis-specific
ceedings of the National Academy of Sciences of the United rep-PCR fragment. Journal of Applied Microbiology 94,
States of America 101, 8449–8454. 1108–1119.
[21] Okinaka, R., Cloud, K., Hampton, O., Hoffmaster, A., Hill, K., [35] Radnedge, L., Agron, P.G., Hill, K.K., Jackson, P.J., Ticknor,
Keim, P., Koehler, T., Lamke, G., Kumano, S., Manter, D., L.O., Keim, P. and Andersen, G.L. (2003) Genome differences
Martinez, Y., Ricke, D., Svensson, R. and Jackson, P. (1999) that distinguish Bacillus anthracis from Bacillus cereus and
Sequence, assembly and analysis of pXO1 and pXO2. Journal of Bacillus thuringiensis. Applied and Environmental Microbiology
Applied Microbiology 87, 261–262. 69, 2755–2764.
[22] Okinaka, R.T., Cloud, K., Hampton, O., Hoffmaster, A.R., Hill, [36] Keim, P., Price, L.B., Klevytska, A.M., Smith, K.L., Schupp,
K.K., Keim, P., Koehler, T.M., Lamke, G., Kumano, S., J.M., Okinaka, R., Jackson, P.J. and Hugh-Jones, M.E. (2000)
Mahillon, J., Manter, D., Martinez, Y., Ricke, D., Svensson, R. Multiple-locus variable-number tandem repeat analysis reveals
and Jackson, P.J. (1999) Sequence and organization of pXO1, genetic relationships within Bacillus anthracis. Journal of Bac-
the large Bacillus anthracis plasmid harboring the anthrax toxin teriology 182, 2928–2936.
genes. Journal of Bacteriology 181, 6509–6515. [37] Turnbull, P.C., Hutson, R.A., Ward, M.J., Jones, M.N., QUinn,
[23] Priest, F.G. (1981) DNA homology in the genus Bacillus In: The C.P., Finnie, N.J., Duggleby, C.J., Kramer, J.M. and Melling, J.
Aerobic Endospore-forming Bacteria (Berkeley, R.C.W. and (1992) Bacillus anthracis but not always anthrax. Journal of
Goodfellow, M., Eds.), pp. 35–57. Academic Press, London. Applied Bacteriology 72, 21–28.
[24] Ash, C. and Collins, M.D. (1992) Comparative analysis of 23S [38] Turnbull, P.C. (1999) Definitive identification of Bacillus anthra-
ribosomal RNA gene sequences of Bacillus anthracis and emetic cis – a review. Journal of Applied Microbiology 87, 237–240.
Bacillus cereus determined by PCR direct sequencing. FEMS [39] Fleischmann, R.D., Adams, M.D., White, O., Clayton, R.A.,
Microbiology Letters 94, 75–80. Kirkness, E.F., Kerlavage, A.R., Bult, C.J., Tomb, J.F.,
[25] Ash, C., Farrow, J.A.E., Dorsch, M., Stackebrandt, E. and Dougherty, B.A. and Merrick, J.M. (1995) Whole-genome
Collins, M.D. (1991) Comparative analysis of Bacillus anthracis, random sequencing and assembly of Haemophilus influenzae
Bacillus cereus, and related species on the basis of reverse Rd. Science 269, 496–512.
transcriptase sequencing of 16S ribosomal RNA. International [40] Read, T.D., Peterson, S.N., Tourasse, N., Baillie, L.W., Paulsen,
Journal of Systematic Bacteriology 41, 343–346. I.T., Nelson, K.E., Tettelin, H., Fouts, D.E., Eisen, J.A., Gill,
[26] Bavykin, S.G., Lysov, Y.P., Zakhariev, V., Kelly, J.J., Jackman, S.R., Holtzapple, E.K., Økstad, O.A., Helgason, E., Rilstone, J.,
J., Stahl, D.A. and Cherni, A. (2004) Use of 16S rRNA, 23S Wu, M., Kolonay, J.F., Beanan, M.J., Dodson, R.J., Brinkac,
rRNA, and gyrB gene sequence analysis to determine phyloge- L.M., Gwinn, M., DeBoy, R.T., Madpu, R., Daugherty, S.C.,
netic relationships of Bacillus cereus group microorganisms. Durkin, A.S., Haft, D.H., Nelson, W.C., Peterson, J.D., Pop,
Journal of Clinical Microbiology 42, 3711–3730. M., Khouri, H.M., Radune, D., Benton, J.L., Mahamoud, Y.,
[27] Carlson, C.R., Caugant, D.A. and Kolstø, A.B. (1994) Geno- Jiang, L.X., Hance, I.R., Weidman, J.F., Berry, K.J., Plaut,
typic diversity among Bacillus cereus and Bacillus thuringiensis R.D., Wolf, A.M., Watkins, K.L., Nierman, W.C., Hazen, A.,
strains. Applied and Environmental Microbiology 60, 1719– Cline, R., Redmond, C., Thwaite, J.E., White, O., Salzberg, S.L.,
1725. Thomason, B., Friedlander, A.M., Koehler, T.M., Hanna, P.C.,
[28] Helgason, E., Tourasse, N.J., Meisal, R., Caugant, D.A. and Kolstø, A.B. and Fraser, C.M. (2003) The genome sequence of
Kolstø, A.B. (2004) Multilocus sequence typing scheme for Bacillus anthracis Ames and comparison to closely related
bacteria of the Bacillus cereus group. Applied and Environmen- bacteria. Nature 423, 81–86.
tal Microbiology 70, 191–201. [41] Read, T.D., Salzberg, S.L., Pop, M., Shumway, M., Umayam,
[29] Priest, F.G., Barker, M., Baillie, L.W., Holmes, E.C. and L., Jiang, L.X., Holtzapple, E., Busch, J.D., Smith, K.L.,
Maiden, M.C. (2004) Population structure and evolution of the Schupp, J.M., Solomon, D., Keim, P. and Fraser, C.M. (2002)
Bacillus cereus group. Journal of Bacteriology 186, 7959–7970. Comparative genome sequencing for discovery of novel poly-
[30] Ko, K.S., Kim, J.W., Kim, J.M., Kim, W., Chung, S.I., Kim, I.J. morphisms in Bacillus anthracis. Science 296, 2028–2033.
and Kook, Y.H. (2004) Population structure of the Bacillus [42] Pearson, T., Busch, J.D., Ravel, J., Read, T.D., Rhoton, S.D.,
cereus group as determined by sequence analysis of six house- UÕRen, J.M., Simonson, T.S., Kachur, S.M., Leadem, R.R.,
keeping genes and the plcR gene. Infection and Immunity 72, Cardon, M.L., Van Ert, M.N., Huynh, L.Y., Fraser, C.M. and
5253–5261. Keim, P. (2004) Phylogenetic discovery bias in Bacillus anthracis
[31] Hill, K.K., Ticknor, L.O., Okinaka, R.T., Asay, M., Blair, H., using single-nucleotide polymorphisms from whole-genome
Bliss, K.A., Laker, M., Pardington, P.E., Richardson, A.P., sequencing. Proceedings of the National Academy of Sciences
Tonks, M., Beecher, D.J., Kemp, J.D., Kolstø, A.B., Wong, of the United States of America 101, 13536–13541.
A.C.L., Keim, P. and Jackson, P.J. (2004) Fluorescent amplified [43] Baillie, L. and Read, T.D. (2001) Bacillus anthracis, a bug with
fragment length polymorphism analysis of Bacillus anthracis, attitude!. Current Opinions in Microbiology 4, 78–81.
Bacillus cereus, and Bacillus thuringiensis isolates. Applied and [44] Turnbull, P.C. (1991) Anthrax vaccines: past, present and future.
Environmental Microbiology 70, 1068–1080. Vaccine 9, 533–539.
[32] Keim, P., Kalif, A., Schupp, J., Hill, K., Travis, S.E., Richmond, [45] Ivanova, N., Sorokin, A., Anderson, I., Galleron, N., Candelon,
K., Adair, D.M., Hugh-Jones, M., Kuske, C.R. and Jackson, P. B., Kapatral, V., Bhattacharyya, A., Reznik, G., Mikhailova,
(1997) Molecular evolution and diversity in Bacillus anthracis as N., Lapidus, A., Chu, L., Mazur, M., Goltsman, E., Larsen, N.,
detected by amplified fragment length polymorphism markers. DÕSouza, M., Walunas, T., Grechkin, Y., Pusch, G., Haselkorn,
Journal of Bacteriology 179, 818–824. R., Fonstein, M., Ehrlich, S.D., Overbeek, R. and Kyrpides, N.
[33] Ticknor, L.O., Kolstø, A.B., Hill, K.K., Keim, P., Laker, M.T., (2003) Genome sequence of Bacillus cereus and comparative
Tonks, M. and Jackson, P.J. (2001) Fluorescent amplified analysis with Bacillus anthracis. Nature 423, 87–91.
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 327

[46] Sneath, P.H.A. (1986) Endospore-forming Gram-positive rods [62] Helgason, E., Caugant, D.A., Olsen, I. and Kolstø, A.B. (2000)
and cocci (Sneath, P.H.A., Mair, N.S., Sharpe, M.E. and Holt, Genetic structure of population of Bacillus cereus and Bacillus
J.G., Eds.), BergeyÕs Manual of Systematic Bacteriology, vol. 2, thuringiensis isolates associated with periodontitis and other
p. 1131. Williams & Wilkins, Blatimore. human infections. Journal of Clinical Microbiology 38, 1615–
[47] Rasko, D.A., Ravel, J., Økstad, O.A., Helgason, E., Cer, R.Z., 1622.
Jiang, L., Shores, K.A., Fouts, D.E., Tourasse, N.J., Angiuoli, [63] Pannucci, J., Okinaka, R.T., Williams, E., Sabin, R., Ticknor,
S.V., Kolonay, J., Nelson, W.C., Kolstø, A.B., Fraser, C.M. and L.O. and Kuske, C.R. (2002) DNA sequence conservation
Read, T.D. (2004) The genome sequence of Bacillus cereus between the Bacillus anthracis pXO2 plasmid and genomic
ATCC 10987 reveals metabolic adaptations and a large plasmid sequence from closely related bacteria. BMC Genomics 3, 34.
related to Bacillus anthracis pXO1. Nucleic Acids Research 32, [64] Gladstone, G.P. (1946) Immunity to anthrax – protective antigen
977–988. present in cell-free culture filtrates. British Journal of Experi-
[48] Berry, C., OÕNeil, S., Ben-Dov, E., Jones, A.F., Murphy, L., mental Pathology 27, 394–418.
Quail, M.A., Holden, M.T.G., Harris, D., Zaritsky, A. and [65] Lacy, D.B. and Collier, R.J. (2002) Structure and function of
Parkhill, J. (2002) Complete sequence and organization of anthrax toxin. Anthrax 271, 61–85.
pBtoxis, the toxin-coding plasmid of Bacillus thuringiensis subsp [66] Friedlander, A.M. (1990) The anthrax toxins In: Trafficking of
israelensis. Applied and Environmental Microbiology 68, 5082– Bacterial Toxins (Saelinger, C.B., Ed.), pp. 121–138. CRC Press,
5095. Boca Raton, FL.
[49] Cummings, C.A. and Relman, D.A. (2002) Genomics and [67] Leppla, S.H. (1995) Anthrax toxins In: Bacterial Toxins and
microbiology. Microbial forensics – ‘‘cross-examining patho- Virulence Factors in Disease (Moss, J., Iglewski, B., Vaughn, M.
gens’’. Science 296, 1976–1979. and Tu, A.T., Eds.), pp. 543–572. Dekker, New York, NY.
[50] CDC, Update: investigation of anthrax associated with [68] Mogridge, J., Mourez, M. and Collier, R.J. (2001) Involvement
intentional exposure and interim public health guidelines, of domain 3 in oligomerization by the protective antigen moiety
October, Morbidity and Mortality Weekly Reports 50 (2001) of anthrax toxin. Journal of Bacteriology 183, 2111–2116.
889–893. [69] Pannifer, A.D., Wong, T.Y., Schwarzenbacher, R., Renatus, M.,
[51] Keim, P., Van Ert, M.N., Pearson, T., Vogler, A.J., Huynh, L.Y. Petosa, C., Bienkowska, J., Lacy, D.B., Collier, R.J., Park, S.,
and Wagner, D.M. (2004) Anthrax molecular epidemiology and Leppla, S.H., Hanna, P. and Liddington, R.C. (2001) Crystal
forensics: using the appropriate marker for different evolutionary structure of the anthrax lethal factor. Nature 414, 229–233.
scales. Infection, Genetics and Evolution 4, 205–213. [70] Singh, Y., Khanna, H., Chopra, A.P. and Mehra, V. (2001) A
[52] Suyama, M. and Bork, P. (2001) Evolution of prokaryotic gene dominant negative mutant of Bacillus anthracis protective
order: genome rearrangements in closely related species. Trends antigen inhibits anthrax toxin action in vivo. Journal of
in Genetics 17, 10–13. Biological Chemistry 276, 22090–22094.
[53] Parkhill, J. and Berry, C. (2003) Genomics: relative pathogenic [71] Bourgogne, A., Drysdale, M., Hilsenbeck, S.G., Peterson, S.N.
values. Nature 423, 23–25. and Koehler, T.M. (2003) Global effects of virulence gene
[54] Gohar, M., Økstad, O.A., Gilois, N., Sanchis, V., Kolstø, A.B. regulators in a Bacillus anthracis strain with both virulence
and Lereclus, D. (2002) Two-dimensional electrophoresis anal- plasmids. Infection and Immunity 71, 2736–2743.
ysis of the extracellular proteome of Bacillus cereus reveals the [72] Liu, H., Bergman, N.H., Thomason, B., Shallom, S., Hazen, A.,
importance of the PlcR regulon. Proteomics 2, 784–791. Crossno, J., Rasko, D.A., Ravel, J., Read, T.D., Peterson, S.N.,
[55] Lereclus, D., Agaisse, H., Gominet, M., Salamitou, S. and Yates 3rd, J. and Hanna, P.C. (2004) Formation and compo-
Sanchis, V. (1996) Identification of a Bacillus thuringiensis gene sition of the Bacillus anthracis endospore. Journal of Bacteriol-
that positively regulates transcription of the phosphatidylinosi- ogy 186, 164–178.
tol-specific phospholipase C gene at the onset of the stationary [73] Coker, P.R., Smith, K.L., Fellows, P.F., Rybachuck, G.,
phase. Journal of Bacteriology 178, 2749–2756. Kousoulas, K.G. and Hugh-Jones, M.E. (2003) Bacillus anthra-
[56] Slamti, L. and Lereclus, D. (2002) A cell–cell signaling peptide cis virulence in Guinea pigs vaccinated with anthrax vaccine
activates the PlcR virulence regulon in bacteria of the Bacillus adsorbed is linked to plasmid quantities and clonality. Journal of
cereus group. EMBO Journal 21, 4550–4559. Clinical Microbiology 41, 1212–1218.
[57] Slamti, L., Perchat, S., Gominet, M., Vilas-Boas, G., Fouet, A., [74] Andrup, L., Jorgensen, O., Wilcks, A., Smidt, L. and Jensen,
Mock, M., Sanchis, V., Chaufaux, J., Gohar, M. and Lereclus, G.B. (1996) Mobilization of ‘‘nonmobilizable’’ plasmids by the
D. (2004) Distinct mutations in PlcR explain why some strains of aggregation-mediated conjugation system of Bacillus thuringien-
the Bacillus cereus group are nonhemolytic. Journal of Bacteri- sis. Plasmid 36, 75–85.
ology 186, 3531–3538. [75] Kaspar, R.L. and Robertson, D.L. (1987) Purification and
[58] Mignot, T., Mock, M., Robichon, D., Landier, A., Lereclus, D. physical analysis of Bacillus anthracis plasmids pXO1 and pXO2.
and Fouet, A. (2001) The incompatibility between the PlcR- and Biochemistry Biophysical Research Communications 149, 362–
AtxA-controlled regulons may have selected a nonsense muta- 368.
tion in Bacillus anthracis. Molecular Microbiology 42, 1189– [76] Uchida, I., Sekizaki, T., Hashimoto, K. and Terakado, N. (1985)
1198. Association of the encapsulation of Bacillus anthracis with a 60
[59] Pomerantsev, A.P., Pomerantseva, O.M. and Leppla, S.H. megadalton plasmid. Journal of General Microbiology 131 (Pt.
(2004) A spontaneous translational fusion of Bacillus cereus 2), 363–367.
PlcR and PapR activates transcription of PlcR-dependent genes [77] Makino, S.I., Uchida, I., Terakado, N., Sasakawa, C. and
in Bacillus anthracis via binding with a specific palindromic Yoshikawa, M. (1989) Molecular characterization and protein
sequence. Infection and Immunity 72, 5814–5823. analysis of the cap region, which is essential for encapsulation in
[60] Økstad, O.A., Gominet, M., Purnelle, B., Rose, M., Lereclus, D. Bacillus anthracis. Journal of Bacteriology 171, 722–730.
and Kolstø, A.B. (1999) Sequence analysis of three Bacillus [78] Uchida, I., Makino, S., Sasakawa, C., Yoshikawa, M., Sugim-
cereus loci carrying PIcR-regulated genes encoding degradative oto, C. and Terakado, N. (1993) Identification of a novel gene,
enzymes and enterotoxin. Microbiology 145, 3129–3138. dep, associated with depolymerization of the capsular polymer in
[61] Beecher, D.J., Schoeni, J.L. and Wong, A.C.L. (1995) Entero- Bacillus anthracis. Molecular Microbiology 9, 487–496.
toxic activity of hemolysin BL from Bacillus cereus. Infection [79] Makino, S., Watarai, M., Cheun, H.I., Shirahata, T. and
and Immunity 63, 4423–4428. Uchida, I. (2002) Effect of the lower molecular capsule released
328 D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329

from the cell surface of Bacillus anthracis on the pathogenesis of toxin crystal to mosquitocidal activity of Bacillus thuringiensis
anthrax. Journal of Infectious Diseases 186, 227–233. subsp. israeliensis. FEMS Microbiology Letters 131, 249–2554.
[80] Tinsley, E., Naqvi, A., Bourgogne, A., Koehler, T.M. and Khan, [97] Guidi-Rontani, C., Pereira, Y., Ruffie, S., Sirard, J.C., Weber-
S.A. (2004) Isolation of a minireplicon of the virulence plasmid Levy, M. and Mock, M. (1999) Identification and characteriza-
pXO2 of Bacillus anthracis and characterization of the plasmid- tion of a germination operon on the virulence plasmid pXO1 of
encoded RepS replication protein. Journal of Bacteriology 186, Bacillus anthracis. Molecular Microbiology 33, 407–414.
2717–2723. [98] Ason, B., Handayani, R., Williams, C.R., Bertram, J.G.,
[81] Hamon, M.A. and Lazazzera, B.A. (2001) The sporulation Hingorani, M.M., OÕDonnell, M., Goodman, M.F. and Bloom,
transcription factor Spo0A is required for biofilm development L.B. (2003) Mechanism of loading the Escherichia coli DNA
in Bacillus subtilis. Molecular Microbiology 42, 1199–1209. polymerase III beta sliding clamp on DNA. Bonafide primer/
[82] Hamoen, L.W., Kausche, D., Marahiel, M.A., van Sinderen, D., templates preferentially trigger the gamma complex to hydrolyze
Venema, G. and Serror, P. (2003) The Bacillus subtilis transition ATP and load the clamp. Journal of Biological Chemistry 278,
state regulator AbrB binds to the 35 promoter region of comK. 10033–10040.
FEMS Microbiology Letters 218, 299–304. [99] Anand, S.P., Mitra, P., Naqvi, A. and Khan, S.A. (2004) Bacillus
[83] Hamoen, L.W., Venema, G. and Kuipers, O.P. (2003) Control- anthracis and Bacillus cereus PcrA helicases can support DNA
ling competence in Bacillus subtilis: shared use of regulators. unwinding and in vitro rolling-circle replication of plasmid
Microbiology 149, 9–17. pT181 of Staphylococcus aureus. Journal of Bacteriology 186,
[84] Koehler, T.M. (2002) Bacillus anthracis genetics and virulence 2195–2199.
gene regulation. Current Topics in Microbiology Immunology [100] Naqvi, A., Tinsley, E. and Khan, S.A. (2003) Purification and
271, 143–164. characterization of the PcrA helicase of Bacillus anthracis.
[85] Saile, E. and Koehler, T.M. (2002) Control of anthrax toxin gene Journal of Bacteriology 185, 6633–6639.
expression by the transition state regulator abrB. Journal of [101] Wilcks, A., Jayaswal, N., Lereclus, D. and Andrup, L. (1998)
Bacteriology 184, 370–380. Characterization of plasmid pAW63, a second self-transmissible
[86] Bastos, M.C. and Murphy, E. (1988) Transposon Tn554 encodes plasmid in Bacillus thuringiensis subsp. kurstaki HD73. Micro-
three products required for transposition. EMBO Journal 7, biology 144, 1263–1270.
2935–2941. [102] Wilcks, A., Smidt, L., Økstad, O.A., Kolstø, A.B., Mahillon,
[87] Steichen, C., Chen, P., Kearney, J.F. and Turnbough, C.L. J. and Andrup, L. (1999) Replication mechanism and
(2003) Identification of the immunodominant protein and other sequence analysis of the replicon of pAW63, a conjugative
proteins of the Bacillus anthracis exosporium. Journal of plasmid from Bacillus thuringiensis. Journal of Bacteriology
Bacteriology 185, 1903–1910. 181, 3193–3200.
[88] Sylvestre, P., Couture-Tosi, E. and Mock, M. (2003) Polymor- [103] Andrup, L., Smidt, L., Andersen, K. and Boe, L. (1998) Kinetics
phism in the collagen-like region of the Bacillus anthracis BclA of conjugative transfer: a study of the plasmid pXO16 from
protein leads to variation in exosporium filament length. Journal Bacillus thuringiensis subsp. israelensis. Plasmid 40, 30–43.
of Bacteriology 185, 1555–1563. [104] Jensen, G.B., Andrup, L., Wilcks, A., Smidt, L. and Poulsen,
[89] Sylvestre, P., Couture-Tosi, E. and Mock, M. (2002) A collagen- O.M. (1996) The aggregation-mediated conjugation system of
like surface glycoprotein is a structural component of the Bacillus thuringiensis subsp. israelensis: host range and kinetics of
Bacillus anthracis exosporium. Molecular Microbiology 45, 169– transfer. Current Microbiology 33, 228–236.
178. [105] Anisimov, A.P., Lindler, L.E. and Pier, G.B. (2004) Intraspecific
[90] Andrup, L., Jensen, G.B., Wilcks, A., Smidt, L., Hoflack, L. and diversity of Yersinia pestis. Clinical Microbiology Reviews 17,
Mahillon, J. (2003) The patchwork nature of rolling-circle 434–464.
plasmids: comparison of six plasmids from two distinct Bacillus [106] Brown, E.R. and Cherry, W.B. (1955) Specific identification of
thuringiensis serotypes. Plasmid 49, 205–232. Bacillus anthracis by means of a variant bacteriophage. Journal
[91] Price, L.B., Hugh-Jones, M., Jackson, P.J. and Keim, P. (1999) of Infectious Diseases 96, 34–39.
Genetic diversity of protective antigen genes of Bacillus anthra- [107] Stone, R. (2002) Bacteriophage therapy: StalinÕs forgotten cure.
cis. Journal of Bacteriology 181, 2358–2362. Science 298, 728–731.
[92] Battisti, L., Green, B.D. and Thorne, C.B. (1985) Mating system [108] Stone, R. (2002) Bacteriophage therapy: food and agriculture:
for plasmid transfer of plasmids among Bacillus anthracis, testing grounds for phage therapy. Science 298, 730.
Baciluus, cereus and Bacillus thuringiensis. Journal of Bacteriol- [109] Canchaya, C., Proux, C., Fournous, G., Bruttin, A. and
ogy 162, 543–550. Brussow, H. (2003) Prophage genomics. Microbiology and
[93] Mourez, M., Yan, M., Lacy, D.B., Dillon, L., Bentsen, L., Molecular Biology Reviews 67, 238–276.
Marpoe, A., Maurin, C., Hotze, E., Wigelsworth, D., Pimental, [110] Casjens, S. (2003) Prophages and bacterial genomics: what have
R.A., Ballard, J.D., Collier, R.J. and Tweten, R.K. (2003) we learned so far?. Molecular Microbiology 49, 277–300.
Mapping dominant-negative mutations of anthrax protective [111] Desiere, F., McShan, W.M., van Sinderen, D., Ferretti, J.J. and
antigen by scanning mutagenesis. Proceedings of the National Brussow, H. (2001) Comparative genomics reveals close genetic
Academy of Sciences of the United States of America 100, relationships between phages from dairy bacteria and pathogenic
13803–13808. streptococci: evolutionary implications for prophage–host inter-
[94] Ackermann, H.W., Roy, R., Martin, M., Murthy, M.R. and actions. Virology 288, 325–341.
Smirnoff, W.A. (1978) Partial characterization of a cubic [112] Brussow, H., Canchaya, C. and Hardt, W.D. (2004) Phages and
Bacillus phage. Canadian Journal of Microbiology 24, 986– the evolution of bacterial pathogens: from genomic rearrange-
993. ments to lysogenic conversion. Microbiology and Molecular
[95] Stromsten, N.J., Benson, S.D., Burnett, R.M., Bamford, D.H. Biology Reviews 68, 560.
and Bamford, J.K. (2003) The Bacillus thuringiensis linear [113] Perna, N.T., Plunkett 3rd, G., Burland, V., Mau, B., Glasner,
double-stranded DNA phage Bam35, which is highly similar to J.D., Rose, D.J., Mayhew, G.F., Evans, P.S., Gregor, J.,
the Bacillus cereus linear plasmid pBClin15, has a prophage Kirkpatrick, H.A., Posfai, G., Hackett, J., Klink, S., Boutin,
state. Journal of Bacteriology 185, 6985–69859. A., Shao, Y., Miller, L., Grotbeck, E.J., Davis, N.W., Lim, A.,
[96] Crickmore, N., Bone, E.J., Williams, J.A. and Ellar, D.J. (1995) Dimalanta, E.T., Potamousis, K.D., Apodaca, J., Ananthar-
Contributions of the indiviual components of the delta-endo- aman, T.S., Lin, J., Yen, G., Schwartz, D.C., Welch, R.A. and
D.A. Rasko et al. / FEMS Microbiology Reviews 29 (2005) 303–329 329

Blattner, T. () Genome sequence of enterohaemorrhagic Esch- [125] Oosthuizen, M.C., Steyn, B., Lindsay, D., Brozel, V.S. and von
erichia coli O157:H7. Nature 409, 529–533. Holy, A. (2001) Novel method for the proteomic investigation of
[114] Campbell, A.M. (1992) Chromosomal insertion sites for phages a dairy-associated Bacillus cereus biofilm. FEMS Microbiology
and plasmids. Journal of Bacteriology 174, 7495–7499. Letters 194, 47–51.
[115] Schuch, R., Nelson, D. and Fischetti, V.A. (2002) A bacteriolytic [126] Oosthuizen, M.C., Steyn, B., Theron, J., Cosette, P., Lindsay,
agent that detects and kills Bacillus anthracis. Nature 418, 884– D., von Holy, A. and Brozel, V.S. (2002) Proteomic analysis
889. reveals differential protein expression by Bacillus cereus during
[116] Ireland, J.A. and Hanna, P.C. (2002) Macrophage-enhanced biofilm formation. Applied and Environmental Microbiology 68,
germination of Bacillus anthracis endospores requires gerS. 2770–2780.
Infection and Immunity 70, 5870–5872. [127] Chen, F.C., Shen, L.F., Tsai, M.C. and Chak, K.F. (2003) The
[117] Volokhov, D., Pomerantsev, A., Kivovich, V., Rasooly, A. and IspA proteaseÕs involvement in the regulation of the sporulation
Chizhikov, V. (2004) Identification of Bacillus anthracis by process of Bacillus thuringiensis is revealed by proteomic
multiprobe microarray hybridization. Diagnostic Microbiology analysis. Biochemical and Biophysical Research Communica-
and Infectious Disease 49, 163–171. tions 312, 708–715.
[118] Zwick, M.E., Macaffee, F., Cutler, D.J., Read, T.D., Ravel, J., [128] McNall, R.J. and Adang, M.J. (2003) Identification of novel
Bowman, G.R., Galloway, D.R. and Mateczum, A. (2004) Bacillus thuringiensis Cry1Ac binding proteins in Manduca sexta
Microarray-based resequencing of multiple Bacillus anthracis midgut through proteomic analysis. Insect Biochemistry and
isolates. Genome Biology 6, R10. Molecular Biology 33, 999–1010.
[119] Kroos, L. and Yu, Y.T.N. (2000) Regulation of sigma factor [129] Avery, O.T., MacLeod, C.M. and McCarty, M. (1944)
activity during Bacillus subtilis development. Current Opinion in Studies on the chemical nature of the substance inducing
Microbiology 3, 553–560. transformation of pneumococcal types. Inductions of trans-
[120] Warscheid, B. and Fenselau, C. (2004) A targeted proteomics formation by a desoxyribonucleic acid fraction isolated from
approach to the rapid identification of bacterial cell mixtures by pneumococcus type III. Journal of Experimental Medicine 79,
matrix-assisted laser desorption/ionization mass spectrometry. 137–158.
Proteomics 4, 2877–2892. [130] Hayes, W. (1953) Observations on a transmissible agent deter-
[121] Warscheid, B. and Fenselau, C. (2003) Characterization of mining sexual differentiation in Bacti. coli. Journal of General
Bacillus spore species and their mixtures using postsource decay Microbiology 8, 72–88.
with a curved-field reflectron. Analytical Chemistry 75, 5618– [131] Lederberg, J., Cavalli, L.L. and Lederberg, E.M. (1952) Sex
5627. compatibility in E. coli. Genetics 37, 720–730.
[122] Lai, E.M., Phadke, N.D., Kachman, M.T., Giorno, R., Vazquez, [132] Watanabe, T. (1963) Infectious heredity of multiple drug
S., Vazquez, J.A., Maddock, J.R. and Driks, A. (2003) Proteo- resistance in bacteria. Bacteriological Reviews 27, 87–115.
mic analysis of the spore coats of Bacillus subtilis and Bacillus [133] N.R. Smith, R.E. Gordon, F.E. Clark, Aerobic sporeforming
anthracis. Journal of Bacteriology 185, 1443–1454. bacteria, US Department of Agriculture, Agriculture Mono-
[123] Whiting, G.C., Rijpkema, S., Adams, T. and Corbel, M.J. (2004) graph No. 16, US Government Printing Office, Washington, DC,
Characterisation of adsorbed anthrax vaccine by two-dimen- 1952.
sional gel electrophoresis. Vaccine 22, 4245–4251. [134] R.E. Gordon, W.C. Haynes, C.H.-N. Pang, The genus Bacillus,
[124] Ariel, N., Zvi, A., Makarova, K.S., Chitlaru, T., Elhanany, E., Aagriculture Handbook No. 427, United States Department of
Velan, B., Cohen, S., Friedlander, A.M. and Shafferman, A. Argiculture, Agricultural Research Service, Washington, DC,
(2003) Genome-based bioinformatic selection of chromosomal 1973.
Bacillus anthracis putative vaccine candidates coupled with [135] Rasko, D.A., Myers, G.S. and Ravel, J. (2005) Visualization of
proteomic identification of surface-associated antigens. Infection comparative genomic analyses by BLAST score ratio BMC.
and Immunity 71, 4563–4579. Bioinformatics 6, 2.

Das könnte Ihnen auch gefallen