Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Much of what we know regarding bacterial spore structure and function has been learned from studies of the genetically well-characterized bacterium Bacillus subtilis. Molecular aspects of spore structure, assembly, and function are well defined. However, certain bacteria produce spores with an outer spore layer, the exosporium, which is not present on B. subtilis spores. Our understanding of the composition and biological functions of the exosporium layer is much more limited than that of other aspects of the spore. Because the bacterial spore surface is important for the spore's interactions with the environment, as well as being the site of interaction of the spore with the host's innate immune system in the case of spore-forming bacterial pathogens, the exosporium is worthy of continued investigation. Recent exosporium studies have focused largely on members of the Bacillus cereus family, principally Bacillus anthracis and Bacillus cereus. Our understanding of the composition of the exosporium, the pathway of its assembly, and its role in spore biology is now coming into sharper focus. This review expands on a 2007 review of spore surface layers which provided an excellent conceptual framework of exosporium structure and function (A. O. Henriques and C. P. Moran, Jr., Annu Rev Microbiol 61:555-588, 2007, http://dx.doi.org/10.1146/annurev.micro.61.080706.093224). That review began a process of considering outer spore layers as an integrated, multilayered structure rather than simply regarding the outer spore components as independent parts.

Free full text 


Logo of mmbrLink to Publisher's site
Microbiol Mol Biol Rev. 2015 Dec; 79(4): 437–457.
Published online 2015 Oct 28. https://doi.org/10.1128/MMBR.00050-15
PMCID: PMC4651027
PMID: 26512126

The Exosporium Layer of Bacterial Spores: a Connection to the Environment and the Infected Host

SUMMARY

Much of what we know regarding bacterial spore structure and function has been learned from studies of the genetically well-characterized bacterium Bacillus subtilis. Molecular aspects of spore structure, assembly, and function are well defined. However, certain bacteria produce spores with an outer spore layer, the exosporium, which is not present on B. subtilis spores. Our understanding of the composition and biological functions of the exosporium layer is much more limited than that of other aspects of the spore. Because the bacterial spore surface is important for the spore's interactions with the environment, as well as being the site of interaction of the spore with the host's innate immune system in the case of spore-forming bacterial pathogens, the exosporium is worthy of continued investigation. Recent exosporium studies have focused largely on members of the Bacillus cereus family, principally Bacillus anthracis and Bacillus cereus. Our understanding of the composition of the exosporium, the pathway of its assembly, and its role in spore biology is now coming into sharper focus. This review expands on a 2007 review of spore surface layers which provided an excellent conceptual framework of exosporium structure and function (A. O. Henriques and C. P. Moran, Jr., Annu Rev Microbiol 61:555–588, 2007, http://dx.doi.org/10.1146/annurev.micro.61.080706.093224). That review began a process of considering outer spore layers as an integrated, multilayered structure rather than simply regarding the outer spore components as independent parts.

INTRODUCTION

Many soil-dwelling bacteria of the phylum Firmicutes utilize sporulation as a survival mechanism whereby they cease vegetative growth and form an endospore inside the mother cell cytoplasm in times of unfavorable growth conditions, such as nutrient limitation. The spore is a survival rather than a replicative mechanism, as only a single spore develops within the cytoplasm of the sporulating cell (an endospore within a mother cell). The spore is released from the mother cell by lysis, and the mature spore exists as a dormant form of the cell that lacks metabolic activity (1, 2). Spores can persist in the environment for long periods.

Spore components include a core consisting of a cell with a dehydrated cytoplasm bearing high concentrations of calcium dipicolinate (Fig. 1). The core is enveloped in a cortex layer consisting of a modified peptidoglycan. External to the cortex is a proteinaceous shell referred to as the spore coat, often consisting of morphologically distinct inner and outer coat layers. The Bacillus subtilis coat comprises about 70 different proteins (3). Spores are highly refractive to environmental insults, such as desiccation, UV irradiation, heat, and exposure to organic solvents. When favorable growth conditions return, spores are stimulated to undergo a process of germination and outgrowth whereby the dormant spore takes up water, releases calcium ions, resumes metabolic activity, and degrades the cortex and the vegetative cell emerges from the spore coat shell. Spore germination involves signaling from small-molecule germinants that diffuse across the outer spore layers and bind to receptors in the cytoplasmic membrane (4).

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030001.jpg

Transmission electron micrograph of a B. anthracis spore.

Endospore-producing bacteria have substantial impacts on human and animal health. Some are human and/or animal pathogens, whereas others affect humans more indirectly, such as insect pathogens or agents of food or product spoilage. Bacillus anthracis is an important zoonotic pathogen that causes anthrax, a fatal septicemia in ruminants. Human infections result in cutaneous, gastrointestinal, or pulmonary anthrax, depending on the site of spore entry into the patient. Gastrointestinal and pulmonary anthraces are associated with high mortality rates, and anthrax spores are a concern as a weapon of bioterrorism (5). Bacillus cereus is a common cause of food poisoning (6). It is also an invasive pathogen responsible for systemic infections, including endophthalmitis, bloodstream infections, endocarditis, and respiratory tract infections (7). Severe and lethal B. cereus infections have been reported for newborn infants, especially among those born prematurely (8, 9). B. cereus infections in newborns include involvement of the central nervous system or respiratory tract, primary bacteremia, and sepsis. Bacillus thuringiensis, a species closely related to B. anthracis and B. cereus, is an insect pathogen widely utilized in agriculture as a bioinsecticide (10). Sporulating cells of B. thuringiensis produce a crystal toxin that is poisonous to insects in the orders Lepidoptera, Coleoptera, and Diptera (11).

The anaerobic spore former Clostridium difficile is the causal agent of nosocomial antibiotic-associated diarrhea and pseudomembranous colitis and is an important emerging public health concern (12,14). Infection with C. difficile recently surpassed methicillin-resistant Staphylococcus aureus infection as the most common hospital-acquired infection in the United States (15). It is estimated that nosocomial C. difficile infections increase the cost of hospitalization 4-fold compared with matched cohorts, and these infections have been estimated to result in $4.8 billion in excess costs in U.S. acute-care facilities (12). Additionally, 20 to 27% of all C. difficile cases are community associated, with an incidence of 20 to 30 per 100,000 population (16). Clostridium perfringens is one of the most common causes of human foodborne illness, with an estimated million cases of foodborne illness each year in the United States (17, 18). C. perfringens also causes infections of skin and soft tissue, gastroenteritis, gas gangrene, necrotizing enteritis, liver abscess, bacteremia, endophthalmitis, and septic shock with acute hemolysis (19,21). Clostridia are also the second most common cause of anaerobic bacteremia, and the 30-day mortality rate for infected patients is 27 to 44% (19).

Botulism results from exposure to a potent neurotoxin produced by the spore-forming bacterium Clostridium botulinum (22,24). There are five main kinds of botulism. Foodborne botulism is caused by eating foods that contain the botulinum toxin. Canned (especially home-canned) and vacuum-packed foods, which provide the anaerobic environment necessary for contaminating spores to germinate and the vegetative cells to produce the toxin, are the foods usually implicated. Wound botulism is caused by wound contamination with C. botulinum spores, and currently the majority of these cases are associated with black-tar heroin injection. Infant botulism is caused by consumption of the spores of the botulinum bacteria, which then grow in the intestines and release toxin in children under the age of 1 year. Adult intestinal toxemia (adult intestinal colonization) botulism is a very rare form of botulism that occurs among adults by the same route as infant botulism. Iatrogenic botulism can occur from accidental overdose of botulinum toxin in medical or cosmetic product formulations. In the United States, an average of 145 cases of botulism are reported each year, with approximately 15% being foodborne, 65% being infant botulism, and 20% being wound associated (22).

The spore-forming process of B. subtilis is the best studied of such processes (25). Early in sporulation, the cell begins to divide asymmetrically, and the division septum undergoes remodeling, resulting in engulfment of the smaller cell compartment. This results in the formation of a smaller, double-membrane-bound forespore compartment within the mother cell. During spore maturation, the mother cell directs the synthesis and assembly of the outermost spore layers, including spore coat proteins and, depending on the bacterial species, additional outer layers, including the exosporium.

THE EXOSPORIUM LAYER OF BACTERIAL SPORES

Under higher magnification, every spore appears to be of egg-like shape and embedded within a ball-shaped glass-like mass looking like a light, narrow ring surrounding the spores.

—Robert Koch (26)

In his classic 1876 paper on the etiology of anthrax, Robert Koch first described what is now known as the exosporium layer of the Bacillus anthracis spore (26). The term “exosporium” was later coined by Flügge in 1886 (27). Bacterial endospores can be placed into two categories: those containing a distinct exosporium and those lacking this structure. The best studied of the former group are those of B. anthracis and B. cereus. Bacillus subtilis spores are the best-studied example of the exosporium-less category. Much of what we know regarding spore assembly has been learned from genetic and biochemical studies with B. subtilis. Consequently, there is a wealth of knowledge regarding initiation of sporulation through spore coat assembly. Because B. subtilis lacks an exosporium, information regarding exosporial composition and its assembly process is considerably more limited.

The exosporium of the B. cereus family of bacteria consists of a basal layer surrounded by an external nap of hairlike projections (28,30) (Fig. 1). The collagen-like glycoprotein BclA (Bacillus collagen-like protein A) is the principal component of this hairlike nap and is an immunodominant spore antigen (31,33). The exosporium basal layer is a shell composed of a number of different proteins (33,35).

Advances in electron microscopic imaging techniques have led to spectacular improvements in our understanding of the exosporium architecture of spores from the B. cereus family (36,39). The basic structures of the exosporium are similar for B. anthracis, B. cereus, and B. thuringiensis. The balloon-like exosporium is deformable and irregularly shaped and possesses numerous folds and creases. The exosporium is separated from the spore coat by extensive distances (for example, 500 nm in B. anthracis [39]) yet lies in close proximity to the spore coat at other sites around the spore. The region between the basal layer of the exosporium and the outer spore coat is referred to as the interspace (35). The basal layer of the B. anthracis exosporium is approximately 12 to 16 nm thick and appears to be comprised of two, approximately 5-nm-thick sublayers (39). The basal layer has a crystalline structural organization with a 6-fold symmetry and a periodic spacing of 7 nm. The external face of this structure consists of a series of hexagonally packed concave cups in a honeycomb pattern, with the open ends oriented toward the external environment (Fig. 2) (38). The spore core-facing side of this structure consists of a honeycomb pattern of hexameric crowns with a diameter of ~7.5 to 9 nm (Fig. 2). The open ends of the crowns abut one another to form channels between the two faces of the basal layer. The channels have a diameter of ~20 to 34 Å (36, 38). A channel of this size is sufficiently large to permit entry of small-molecule germinants, such as alanine or inosine, but is too small to permit diffusion of larger molecules, such as proteins. The measured channel size is consistent with the results of spore permeation studies (40). This channel arrangement confers the semipermeable barrier properties of the exosporium.

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030002.jpg

Schematic diagram illustrating a model for the exosporium of the B. cereus family. The exosporium basal layer consists of a two-dimensional lattice of cups opening out to the environment on the convex outer face of the layer. There are channels penetrating the basal layer. The collagen-like fibrils of the BclA nap are shown schematically attached to the convex face, terminating in the globular tumor necrosis factor-like C-terminal domain. (Reprinted from reference 38 with permission of the publisher.)

The BclA nap filaments of the B. anthracis spore span 14 to 70 nm (32, 39) and are attached to the outer surface of the basal layer. The length of the filaments is a function of the number of collagen-like triplet amino acid repeats in the BclA protein (32). The hairlike nap covers the entire surface of the exosporium, with an interfilament spacing of approximately 7 nm, the same spacing periodicity as the basal layer crystalline array. The trimeric BclA filaments are oriented with their N termini at the basal layer and their C-terminal domains appearing as a 6-nm knob at the end of the outward-extending filaments.

The exosporium is anchored to the spore coat layer via a number of protein-protein interactions. These have been identified through the isolation of mutants that fail to stably attach the exosporium and from which, as a consequence, the exosporium is released as sheets separate from the mature spores upon lysis of the mother cells. Proteins implicated in exosporium attachment include the spore coat protein CotE and the exosporium- or interspace-localized proteins CotY, ExsA, ExsB, ExsY, and ExsM (1, 41,45). Whether these proteins participate directly in exosporium attachment or indirectly, through proper positioning of proteins actually involved in the attachment, has not been established. CotE is likely to be the spore coat-associated part of the connector between the spore coat and the exosporium. ExsA and/or ExsB may also be part of this connection chain.

THE SPORE CRUST LAYER

B. subtilis lacks an obvious exosporium layer and has the outer spore coat layer as its surface. This is consistent with the finding that known B. cereus family exosporium genes are not present in the B. subtilis genome. However, the surfaces of B. subtilis spores may not be that dissimilar from those of B. cereus family spores. Treatment of B. subtilis spores with urea and mercaptoethanol resulted in the detachment of a layer from the underlying outer coat material that resembled the basal membrane of the exosporium of B. anthracis (46, 47). Use of a ruthenium red staining technique which has been used for visualization of carbohydrate capsules by electron microscopy resulted in the visualization of filaments on the surfaces of B. subtilis spores (Fig. 3) (48). Spore surface filaments are present and are possibly glycoprotein in nature, although this conclusion is drawn with caution, as ruthenium red staining is not totally glycoprotein specific. Spore surface-specific carbohydrates exist in B. subtilis (49). These carbohydrates display structural similarity to those of the B. cereus family spores (49, 50). Production and/or attachment of polysaccharides to the B. subtilis spore surface was shown to be spsM dependent (51). This determinant is the site of SPβ prophage integration, and functional SpsM is produced only following excision of the prophage during sporulation (51). Spores from spsM-negative mutants had altered surface properties, including increased adherence to glass and enhanced aggregation in aqueous environments. The presence of spore surface glycoprotein filaments raises the possibility that functions contributed by the BclA nap on the B. cereus family spores are provided by other surface glycoproteins in certain exosporium-less spores.

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030003.jpg

Electron micrographs of spores from sporulated cultures. (A) B. anthracis stained with ruthenium red. 1, glycoprotein nap; 2, basal membrane. (B) B. anthracis without ruthenium red staining. 1, glycoprotein nap; 2, basal membrane. (C) B. subtilis stained with ruthenium red. 1, glycoprotein nap; 3, outer coat; 4, inner coat. (D) B. subtilis without ruthenium red staining. 3, outer coat; 4, inner coat. (Reprinted from reference 48 with permission of the publisher [copyright Elsevier 2004].)

Additional support for the presence of B. subtilis spore surface glycoproteins came from a study of the cge locus. Mutations in the cgeAB and cgeCDE operons of B. subtilis resulted in spores with altered surface properties (spores tended to clump, pelleted more tightly, and adhered more strongly to plastic and glass surfaces), and it was speculated that this was due to changes in glycosylation of surface proteins (52). In an analysis of spore coat proteins involving fluorescent protein fusions, the B. subtilis spore coat was predicted to be comprised of four distinct layers, including an outermost glycoprotein layer, named the “crust,” that is separated from the outer coat by a small space (53). Genetic and localization studies have shown that cgeA and genes in the cotVWXYZ locus are involved in spore crust formation (53, 54). Interestingly, the CotZ protein, which plays a role in the assembly of the crust layer of B. subtilis, displays sequence similarity to the ExsY protein (35% identity and 46% similarity over 141 residues), a protein involved in exosporium assembly in B. anthracis (42). B. subtilis CotE, CotV, CotY, and CotW coat proteins expressed in Escherichia coli can self-assemble intracellularly into stable complexes (55). These proteins form one-dimensional fibers, two-dimensional sheets, and three-dimensional stacks. With CotY, the mode of assembly resembles that of the exosporium basal layer described for B. anthracis and B. cereus (55). CotY is a component of the B. subtilis crust layer and the B. cereus group exosporium layer. The propensity of these Cot proteins to self-assemble may contribute to the spore coat, crust layer, or exosporium assembly processes.

The questions of whether the crust is functionally equivalent to the surface of the exosporium and whether crust-like layers are found on the outer spore coat in exosporium-containing spores remain unanswered.

THE BOTTLE CAP MODEL OF EXOSPORIUM STRUCTURE

In a study of an exsY mutant of B. anthracis, Turnbough and coworkers discovered that the exosporium layer is not a uniform structure but is comprised of two distinct entities of differing compositions (42, 56). The “bottle cap” part of the exosporium is synthesized first, originating at the mother cell central pole of the spore. After assembly of the bottle cap portion of the exosporium reaches the cap-noncap junction, the noncap portion (~75%) of the exosporium is assembled (Fig. 4). The noncap portion fails to assemble in an exsY-null mutant, yielding spores with a loosely attached, cap-only exosporium that is readily lost when the spore is released from the mother cell (42). The protein compositions of the cap and noncap regions of the exosporium differ. CotY is a cap-specific protein, whereas BclB, ExsY, and Alr are found in the noncap portion of the exosporium (56,58). BclA is found on the exosporium surface in both cap and noncap sites, whereas BxpB is found in larger amounts, but not exclusively, in the noncap portion of the exosporium (58). The junction between the cap and noncap regions may play an important role in exosporium-containing spores during the germination process. Steichen et al. (56) demonstrated that during spore germination and outgrowth, the emerging vegetative cell escapes from its exosporium shell through the displaced cap, suggesting that the cap is designed to pop off the exosporium to permit the emergence of the outgrowing cell. Mutants of B. anthracis lacking the BclB collagen-like protein have an exosporium that exhibits damage and loss of the exosporium at one pole (59). Furthermore, the distribution of exosporium proteins, such as CotY and BxpB, is altered in this mutant, suggesting that BclB is required for correct exosporium assembly. The normally cap-associated CotY protein is distributed more completely around the spore, and BxpB is largely restricted to the noncap pole in bclB-null spores. Localization studies with the BclB-negative spores suggested that without BclB, the cap encompasses >75% of the exosporium's circumference, instead of the usual ~25%. BclB, a non-cap-region exosporium protein, may participate in the transition in exosporium composition that must take place as the assembly of the cap stops and the noncap assembly begins during the discontinuous assembly of the exosporium (56, 58). It has been postulated that in a BclB-null strain, the transition zone between the cap and noncap sections of the exosporium is either displaced to the opposite pole of the spore or enlarged to encompass the midsection of the spore (58).

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030004.jpg

(Top) Illustration of a mature spore in a sporulating cell of B. anthracis. The bottle cap portion of the exosporium, at the mother cell central pole of the spore, is depicted in blue, and the noncap exosporium is shown in red. (Bottom) Three distinct exosporium synthesis events which originate at the mother cell central pole of the spore. (1) The first event is the exosporium scaffold layer portion of the basal layer synthesis involving CotY and ExsY. (2) This is followed closely by outer basal layer assembly, which includes BxpB and the BclA nap layer. When the BxpB layer is then approximately 75% complete, cleavage of the BclA N terminus after amino acid 19 (arrow 3) begins at the mother cell central spore pole, and this wave of cleavage follows the assembly around the spore periphery (57).

ASSEMBLY OF THE EXOSPORIUM

Genetic studies on spore assembly in B. subtilis have ordered the steps in spore production based on the effects of particular mutations on downstream processes (60). Analysis of mutants involving exosporium genes of B. anthracis and B. cereus are more limited but do shed some light on the overall exosporium biosynthetic process. The B. cereus exosporium first appears in the sporulating mother cell as a small lamellar structure adjacent to but distinct from the outer forespore membrane (61). Electron microscopy and protein localization studies have shown that synthesis of the exosporium initiates at the mother cell central spore pole, the region of the exosporium that will become the bottle cap (35, 56, 57, 62). Exosporium assembly progresses by continued deposition of proteins at the leading edge of this nascent exosporium, terminating at the opposite spore pole. However, there is a transition in the incorporation of proteins as synthesis reaches the end of the cap region (approximately 25% of the spore circumference), as the composition of the noncap region of the exosporium differs from that of the bottle cap region. As discussed above, there are proteins involved in anchoring the exosporium to the spore coat. The exact nature of the spore coat anchor has yet to be defined. However, the spore coat protein CotE is involved in this process (43). In CotE-negative spores, the exosporium assembles in sheets which are not anchored to the spore, and mature spores have only fragments of exosporium associated with them. However, the exosporium produced in this mutant possesses the BclA nap layer, a late addition in the exosporium assembly process. Thus, anchoring of the exosporium to the underlying spore coat layer is not required for the exosporium assembly process to proceed.

Known exosporium basal layer proteins include ExsY, CotY, BxpB (also known as ExsFA), and ExsFB. The former two proteins are involved in an earlier stage of exosporium synthesis than the latter pair. Loss of ExsY in B. anthracis results in the production of only the cap region of the exosporium, with no noncap region of the exosporium evident in either late-stage sporulating cells or released mature spores (42). The exosporium cap region is not firmly anchored to the spore and is lost, presumably due to fluid shear forces, when spores are prepared in a liquid medium (42, 44). B. cereus mutants devoid of both ExsY and CotY lack an exosporium layer (42, 44), whereas loss of ExsY results in spores with only fragments of exosporium-like material adherent to the spores. The absence of exosporium caps on B. cereus exsY mutant spores was likely the result of shear forces in place during preparation of the spores (44). B. cereus single cotY mutants produce an intact exosporium. With the B. anthracis exsY mutant, no fragments of apparent exosporium remain attached to the released spore as observed with B. cereus. ExsY appears to be essential for noncap exosporium assembly, whereas its paralog CotY is not required for production of an intact exosporium (44). The cap region exosporium produced in the B. anthracis exsY mutant possesses the BclA nap layer (42). Since BclA incorporation occurs later in the exosporium biosynthesis process, this indicates that later-stage assembly events occur in the cap independently of production of the noncap region.

BxpB (ExsFA) is an exosporium basal layer protein that, along with its paralog ExsFB, is required for assembly of the BclA nap layer. Loss of both BxpB and ExsFB is required for complete loss of BclA. Without BxpB, only about 25% of BclA is incorporated onto the spore surface. Loss of ExsFB results in only a very modest reduction in BclA spore levels (63, 64). Mutants lacking both BxpB and ExsFA produce an intact exosporium encircling the spore. However, this structure is fragile, is easily damaged, and is lost from the spores over time (63). This observation suggests that BxpB and ExsFB are assembled into the growing exosporium at a later stage than that involving ExsY and CotY, since loss of the latter pair results in no visible exosporium structure. Thus, the exosporium is likely assembled in two or more distinct layers, consistent with the two layers of exosporium basal layer evident by electron microscopy, as described above. CotY (at the cap) and ExsY (at the noncap region) are incorporated into an early scaffold layer (the sac sublayer of Henriques and Moran [1]). Subsequently, BxpB and ExsFB are added in a layer that is likely to be positioned externally relative to the ExsY/CotY-containing scaffold layer (Fig. 4).

Much of what is known regarding exosporium synthesis concerns incorporation of the nap glycoprotein BclA and the basal layer protein BxpB into the developing exosporium of B. anthracis. Both proteins appear in the mother cell cytoplasm approximately 4 to 5 h into sporulation (in broth cultures shaken at 37°C). The majority of each of these proteins first appears in the form of higher-molecular-weight complexes rather than monomers, indicating that stable complex formation occurs in the mother cell cytoplasm well in advance of placement of these proteins at the site of the developing exosporium. Appearance of BclA and BxpB in the sporulating cell initiates an hour prior to evidence of initial assembly on the developing spore (57). The BxpB-containing and BclA-containing complexes observed in Western blots are similar in size, suggesting that interactions between these two proteins may occur. The protein-protein interactions in these complexes are stable to boiling in the presence of 1% SDS and 8 M urea and in the presence of reducing agents.

An N-terminal targeting sequence of BclA and certain other collagen-like proteins is responsible for positioning these proteins at the site of the developing exosporium (57, 58, 65,67). No such localization domain is present in the BxpB sequence. Therefore, my colleagues and I have proposed a model whereby BxpB is delivered to the spore by virtue of forming a complex with BclA and possibly, to a lesser extent, with other exosporium-targeting motif sequences (58). In keeping with this, we observed a reduction in the amount of BxpB extractable from spores of a BclA-negative strain, a finding similar to that observed by others (65, 66, 68). However, this model is disputed by Rodenburg et al. (39), based on their analysis of B. anthracis spore structure by high-resolution electron microscopy. They propose that loss of BclA has no effect on BxpB spore incorporation. The assignment of these structures as BxpB, however, was not validated either biochemically or immunologically, so the data remain speculative. How their results can be reconciled with the reduction in BxpB levels in spores in which BclA failed to localize, as previously reported by Tan and Turnbough (65), was not addressed in their more recent report.

Exosporium assembly initiates at the mother cell central pole of the spore (the cap region of the exosporium). Assembly begins with the incorporation of an as yet undefined set of proteins (but likely including CotY and ExsY) which provides the scaffold for assembly of the basal layer. After scaffold assembly has initiated, the assembly of the BclA- and BxpB-containing complexes occurs at this mother cell central spore pole and follows scaffold assembly around the spore (57, 66). When positioning of BclA-containing complexes has occurred over approximately 75% of the spore, proteolytic cleavage of the BclA N terminus after amino acid 19 occurs, beginning at the cap end of the exosporium, and the N-terminal peptide is released into the mother cell cytoplasm (57). This cleavage event may be associated with covalent attachment of BclA to an exosporium basal layer protein, thus anchoring the nap glycoprotein to the exosporium surface. Late in the assembly process, after BclA has been incorporated onto the spore surface, glycosylation of BclA occurs (66).

BACILLUS EXOSPORIUM PROTEINS

Exosporium proteins have principally been defined for strains of B. anthracis and B. cereus. Except where noted below, exosporium proteins of the B. cereus family (B. anthracis, B. cereus, and B. thuringiensis), other than collagen-like proteins, tend not to be found, at least as conserved homologs, in other species of Bacillus or in Clostridia. Therefore, an absence of these proteins in genomic sequences does not rule out the presence of an exosporium. The exosporium-associated Cot proteins display sequence similarity with spore coat proteins of other spore formers, and their locations within spores in the non-B. cereus-family spore-forming bacteria have not been defined. The chromosomal locations of putative exosporium determinants in B. anthracis are shown in Fig. 5. The determinants map to multiple sites in the genome rather than being clustered in a single or small number of loci. The proteins indicated below have been shown to be either exosporium associated or possibly so, in cases where the actual location has not been determined precisely. The BAS numbers indicate the Sterne strain coding sequence (CDS) numbers (NCBI accession numbers NC_005945.1 and AE017225.1).

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030005.jpg

B. anthracis Sterne chromosome with map positions of exosporium determinants.

Alr (BAS0238)

Alanine racemase is an enzyme that can convert l-alanine (a germinant for B. anthracis) to d-alanine (an inhibitor of germination) (56, 69). The presence of Alr in the spore may function to inhibit premature germination of the spore under suboptimal growth conditions, such as in the intracellular macrophage environment. It may also prevent premature germination during sporulation (69, 70). The distribution of Alr in the exosporium of B. anthracis is not uniform. The enzyme is found only in the noncap region of the spore, which comprises ~75% of the spore and excludes the pole at which exosporium synthesis initiates (69).

Arginase

B. anthracis endospores exhibit arginase activity (71). This activity is reduced when the exosporium is removed by sonication and retained in exosporium preparations (71). The spore-associated arginase possibly competes with host cell nitric oxide synthase (NOS 2) for its l-arginine substrate (72). As macrophage-generated nitric oxide is important for microbial killing, the presence of arginase in internalized B. anthracis spores may function to decrease intracellular arginine levels and thus reduce the concentration of this substrate for host NOS 2. This may enhance the bacterium's ability to survive following germination in this intracellular compartment during the initial stages of infection. The arginase enzyme has not yet been localized to a specific site within the exosporium or interspace layer of the spore. There are two proteins annotated as arginases encoded in the B. anthracis Sterne genome (BAS0155 and BAS2260). Of these, only BAS0155 has an expression profile consistent with a potential spore-associated protein (73), and the open reading frame is preceded by a putative σK promoter sequence.

BAS5303

The BAS5303 protein was identified as an exosporium component in a proteomic study of the B. anthracis spore (74). Inclusion of this 133-residue protein of unknown function enhanced the efficacy of an anti-protective antigen (anti-PA) vaccine in a mouse model of infection, and anti-BAS5303 antibodies were found to increase phagocytic uptake of spores (68). Although the location of this protein has not been mapped precisely, it was reported to likely lie below the BclA nap layer on the exosporium, based on increased exposure to antibodies in the absence of BclA (68). The bas5303 determinant is monocistronic and expressed at a very late stage of sporulation (73). The open reading frame is preceded by two putative σK promoter elements, consistent with this expression profile.

BclA (BAS1130)

BclA was the first spore surface glycoprotein discovered in B. anthracis and is the prominent protein component of the exosporium nap layer (31, 33). The filaments of the nap are composed of this trimeric collagen-like protein. BclA is comprised of three domains. The N-terminal domain (NTD) contains the targeting and attachment domains whereby the protein is apparently covalently attached to the basal layer of the exosporium (57, 65, 66, 75). Stable attachment requires proteolytic cleavage of the NTD between residues serine 19 and alanine 20. The NTD possesses a targeting sequence which is required for the positioning of BclA around the developing exosporium and the attachment domain containing the proteolytic cleavage site (Fig. 6) (31, 57, 65, 66). The targeting sequence is conserved in a subset of other exosporium-associated collagen-like proteins, whereas the attachment domain, which lies upstream of the targeting sequence, is not identifiable in these proteins (Fig. 6) (57).

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030006.jpg

N-terminal amino acid sequences of B. anthracis collagen-like proteins possessing the exosporium assembly domain. The proteolytic cleavage site of BclA is denoted by an arrow. Attachment motif sequences are shown in red, and localization motif sequences are shown in green.

The NTD is followed by a collagen repeat region consisting of GX1X2 (primarily GPT) triplet amino acid repeats. These repeats are the primary attachment point for rhamnose-oligosaccharides within BclA (76). The number of glycine-containing triplet amino acid repeats in BclA varies among strains (17 to 91 triplet repeats), and this variation is responsible for the different lengths of the hairlike nap observed on spores of different B. anthracis strains (32, 33). The 134-residue carboxy-terminal domain is organized into an all-β structure with a jelly roll topology (77, 78). It is structurally similar (but dissimilar in sequence) to the globular domain of the human C1q complement protein, which belongs to the tumor necrosis factor-like family of proteins. The trimeric globular head formed by the C-terminal domain is thought to be the nucleation site for the formation of the collagen-like triple helix, as in several mammalian collagens (77).

The polysaccharide composition of BclA includes l-rhamnose, N-acetylgalactosamine, 3-O-methyl rhamnose, and anthrose [2-O-methyl-4-(3-hydroxy-3-methylbutamido)-4,6-dideoxyglucose; found at the nonreducing terminus of a tetrasaccharide] (50, 76). Daubenspeck et al. (76) identified a 715-Da tetrasaccharide [2-O-methyl-4-(3-hydroxy-3-methylbutamido)-4,6-dideoxy-β-d-glucopyranosyl-(1→3)-α-l-rhamnopyranosyl-(1→3)-α-l-rhamnopyranosyl-(1→2)-l-rhamnopyranose] and a 324-Da disaccharide attached to BclA, likely through N-acetylgalactosamine residues at the reducing ends, making the polysaccharides actually pentasaccharides and trisaccharides. Multiple copies of the likely pentasaccharides are O-linked to the collagen-like region of BclA, apparently through threonine residues present in the many GXX repeats in this region. The attachment site of the trisaccharides was predicted to be outside the collagen-like repeat region. The BclA glycoprotein is present in strains of B. cereus and B. thuringiensis, but the presence of anthrose appears to be specific only to BclA from B. anthracis (76). However, antisera raised against anthrose cross-reacted with spores of certain B. cereus strains (79). The C-terminal domain of BclA from B. cereus ATCC 14579 was reported to be glycosylated, with 2-O-methyl-rhamnose and 2,4-O-methyl-rhamnose substitutions found in the C-terminal domain (80).

BclA has also been shown to be an immunodominant protein on the B. anthracis spore surface and contributes to the overall spore hydrophobicity (33, 81). This immunogenicity provides a potential benefit for anthrax vaccine development, as immunization with protective antigen (PA) and BclA protected mice and guinea pigs from challenge with B. anthracis Sterne (pXO1+/pXO2) or Ames strains better than immunization with PA alone (68, 81,83). Loss of BclA has little, if any, effect on sporulation rates of B. anthracis. However, germination of BclA-null spores was reported to occur more rapidly than that of wild-type spores in vitro (84).

The BclA protein has a predicted molecular weight (MW) of 37,000, but it first appears as higher-molecular-weight forms in sporulating cells (57, 66, 67). BclA extracted from spores is found in a heterogeneous band of complexes centered with an apparent MW of >250,000. The higher-molecular-weight forms are likely the result of trimerization of this collagen-like protein, its glycosylation, and its capacity to form covalent and noncovalent but SDS- and urea-resistant complexes with other exosporium proteins.

The bclA determinant (bas1130) is carried within a cluster of genes (bas1127 to bas1145) which is organized as shown in Fig. 7. It possesses a putative promoter element recognized by the σK-bearing form of RNA polymerase, resulting in its expression approximately 3 h into sporulation in the mother cell compartment of the sporulating cell. The bclA determinant is flanked by genes predicted to be involved in glycosylation of this, and perhaps other, spore proteins. The rmlABCD genes are predicted to encode enzymes for rhamnose production, and inactivation of the rmlA determinant resulted in a loss of rhamnose incorporation into spores (85). rmlA mutant spores exhibited decreased binding to macrophages, suggesting a role for the carbohydrate moiety of the BclA glycoprotein in host cell interactions (85). Genes known to be associated with glycosylation of BclA are not found at a single locus on the B. anthracis genome. The anthrose component of the BclA glycoprotein is synthesized by the products encoded within a four-gene operon (bas3322 to bas3319) not linked to bclA (86,88). Furthermore, the bas5304 determinant was shown to be the epimerase required to produce N-acetylgalactosamine for BclA oligosaccharide biosynthesis (86).

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030007.jpg

bclA locus of the B. anthracis Sterne chromosome. Exosporium determinants are shown in black, rhamnose biosynthetic genes important for glycosylation of BclA are shown in light blue, the gene for the spore coat protein CotO is shown in dark blue, putative exosporium determinants are shown in brown, and genes encoding proteins which may be involved in glycosylation of BclA, and perhaps other spore collagen-like proteins, are shown in red (putative methyltransferases) and purple (putative glycosyltransferases).

BclB (BAS2281)

The second B. anthracis spore collagen-like glycoprotein to be identified, BclB, contains a GXT tandem repeat domain (89). 3-O-Methyl rhamnose, galactosamine, and rhamnose were found to be components of this glycoprotein. The BclB protein has sequence similarities to the ExsH and ExsJ proteins of B. cereus and B. thuringiensis. The BclB amino acid sequence at the N terminus is more divergent between B. anthracis and the ExsH/J proteins of the other two members of the B. cereus family. BclB possesses an N-terminal exosporium targeting motif similar to that of BclA, although no site corresponding to the N-terminal proteolytic cleavage site of BclA is found in the BclB amino acid sequence (57). BclB was determined to be present in the exosporium of B. anthracis spores by immunogold transmission electron microscopy (59). Immunofluorescence and flow cytometry studies indicated that BclB is surface exposed on spores, but a higher reactivity in spores lacking the BclA nap suggests that BclB is found principally beneath the BclA spore nap layer. The distribution of BclB around spores is not uniform, with this glycoprotein found in the 75% of the spore corresponding to the noncap region (58). BclB assembly into spores is prevented in bxpB-null mutant spores, suggesting that the BxpB basal layer protein is directly or indirectly required for BclB assembly. Spores from a ΔbclB mutant strain exhibited ruptures of the exosporium layer, usually at one pole of the elliptical spore (59). As a consequence of this exosporium instability, a population of exosporium-free spores is present in a sample of bclB-null spores. Loss of BclB also reduced the volume of the interspace region of the spores (58).

An absence of the BclB protein was found to have a pronounced effect on the distribution of other exosporium proteins (58). The normally cap-associated CotY protein was distributed more completely around the spore. BxpB is found in greater quantities in the noncap region of the exosporium than in the cap, but in the absence of BclB, the distribution of BxpB was largely restricted to the noncap pole.

BetA (BclF; BAS3290)

BetA (also referred to as BclF) is a collagen-like protein with the same overall architecture as that of BclA and BclB, with an N-terminal exosporium targeting motif followed by a collagen-like repeat sequence and then a C-terminal domain (Fig. 6) (90, 91). Expressed from an apparent σK promoter, its transcript level peaks later than those of bclA and bclB (73). BetA is found only in the B. cereus family of bacteria and possesses a small collagen-like repeat region which varies in length in different strains, from 20 repeats in B. thuringiensis Al Hakam to 11 repeats in B. anthracis Sterne and Bacillus mycoides. The differences in the collagen-like repeat region sizes among strains are primarily in the number of GIGITGPTGVTGXT motif repeats (1 motif in B. anthracis Sterne and Ames and 5 motif repeats in B. thuringiensis Al Hakam and B. cereus ATCC 14579). BetA resides in the exosporium basal layer, likely underneath BclA. BetA assembles at the spore surface at around hour 5 of sporulation, and its assembly is BxpB dependent, a feature common to the spore collagen-like glycoproteins with the N-terminal exosporium targeting motif (57, 58, 91). A Bacillus megaterium protein (for example, the sequence under GenBank accession number WP_047934722) displays 28% identity and 51% similarity over a 139-residue sequence with the B. anthracis BetA protein. Whether this protein is associated with the spore or exosporium in this species has not been determined.

BxpA (BAS2008)

BxpA is a glutamine- and proline-rich protein first identified in a proteomic analysis of the B. anthracis exosporium (34). BxpA contains repeated amino acid sequences of 5, 40, and 14 amino acids, and the three longest repeats are in tandem (33). BxpA is found in members of the B. cereus family, although sequence variation, including internal truncations, exists. The protein found in mature spores may be proteolytically processed, with only the C-terminal portion being spore associated (33, 92). Immunogold labeling suggested that most of the BxpA in the spore resides in the cortex layer, with no detectable amounts present in the exosporium (92). The YabG and Tgl proteases appear to be involved in BxpA processing during spore assembly, and the proteolytic cleavage may be associated with protein-cross-linking events (92).

BxpB (ExsFA; BAS1144)

The BxpB (also known as ExsFA) protein of the B. cereus family is an exosporium basal layer protein (63, 64). Spores from mutants lacking BxpB show a substantial reduction in BclA content, and the residual BclA is found predominantly at the cap region of the exosporium (35). Complete loss of BclA requires an additional mutation to inactivate ExsFB, an identically sized protein that exhibits 78% sequence identity with BxpB. Loss of both of the ExsF proteins results in the production of a structurally unstable exosporium layer which is readily lost from the spores (63). BxpB is synthesized in the mother cell compartment of sporulating cells and is immediately found in higher-molecular-weight complexes that are refractive to disruption by boiling in SDS in the presence of a reducing agent. BxpB is important, either directly or indirectly, for the incorporation of a number of exosporium proteins into the spore, including BclA, BclB, and BetA (57,59, 65, 66, 91). It is generally believed that the proteolytic processing of the N terminus of BclA results in covalent attachment of BclA to BxpB, although this has not been demonstrated definitively. However, the N-terminal domains of BclA and BxpB are sufficiently close in sporulating cells that fluorescent fusion proteins are capable of fluorescence resonance energy transfer (FRET) (66). Mutant spores of B. anthracis lacking BxpB were reported to have a defect in an early step of germination (35). In contrast, Steichen et al. (64) found that BxpB-negative spores displayed faster germination and outgrowth kinetics.

BxpC (BAS5053)

BxpC was identified as a 15.3-kDa putative exosporium protein of B. anthracis by Steichen et al. (33). BxpC has 40% identity and 62% similarity to ExsM. The main difference between ExsM and BxpC is in the C terminus, which is positively charged in ExsM but not in BxpC (93). BxpC was present in exosporium-deficient spores of an exsB mutant, suggesting that it may not actually reside on the exosporium, but perhaps in the interspace or outer spore coat (45).

CotE (BAS3619)

CotE is an outer spore coat protein that is required for anchoring of the exosporium layer to the spore coat. Mutants lacking CotE produce a normally appearing exosporium in strips that do not properly assemble around the spore (35). Released mature spores lack the exosporium layer, except for small fragments which remain adherent to the spore. Full-length CotE has not been detected in the exosporium layer, but a 12-kDa N-terminal fragment of CotE was identified in the exosporia of B. anthracis and B. cereus (94). CotE of B. subtilis is a morphogenetic protein in spore coat assembly and affects outer spore coat assembly (95). Thus, its impact on exosporium attachment may be indirect, through defects in outer coat assembly. Henriques and Moran hypothesized that CotE may bridge the nascent coat to the exosporium during early exosporium synthesis and subsequently be cleaved to allow the exosporium to spatially separate from the coat to form the characteristic interspace (1).

CotY (BAS1145)

CotY was identified as an exosporium protein in a proteomic study of spores of B. anthracis, but it is also found in other members of the B. cereus family (34). The cotY gene maps to a cluster of exosporium determinants adjacent to the bclA locus (Fig. 7). CotY and ExsY are exosporium proteins that share substantial sequence similarity. Although their N-terminal sequences (~30 residues) are dissimilar, the rest of their sequences are highly conserved (~93% identical). The CotY protein localizes to the exosporium cap region (58).

ExsA (BAS4324)

Spores of an exsA mutant of B. cereus exhibited decreased hydrophobicity and increased permeability to lysozyme and were blocked at a late stage of germination (41). Truncation mutations resulted in defects in anchoring of the exosporium as well as outer spore coat layers. The location of ExsA within the spore has not been determined. Although it is named as a putative exosporium protein and involved in proper exosporium assembly, ExsA is not likely a true exosporium component. It possesses a conserved N-terminal cortex binding domain which suggests that it resides in the inner coat layer of the spore (41). ExsA displays sequence similarity to the B. subtilis SafA spore protein that functions in spore coat assembly. The B. anthracis Sterne protein is 85% identical to the B. cereus protein.

ExsB (BAS1898)

First identified in proteomic studies of B. anthracis and B. cereus exosporium proteins, ExsB is a polypeptide of unusual amino acid composition, with 20 of its 186 residues (in strain Sterne) being cysteine (45, 94). The protein is highly phosphorylated on threonine residues. Loss of ExsB results in unstable exosporium attachment, with loss of the exosporium layer in spores prepared in liquid medium and only fragments of exosporium present on spores prepared on solid medium (45). Aronson et al. (96) reported that this protein is located principally in the inner coat/cortex region of the spore, with little found in exosporium preparations. Thus, they named this protein Cotγ. The amount of Cotγ was found to be dependent on culture conditions (pH and temperature). Reduced levels of Cotγ resulted in spores lacking an exosporium or with an exosporium closely abutting the spore coat, and the resultant Cotγ-negative spores were less hydrophobic and germinated more rapidly than wild-type spores (96).

The exsB determinant has been shown to be transcribed by the σK-bearing form of RNA polymerase in B. thuringiensis, and its expression is additionally regulated by the GerE transcriptional factor (97).

ExsC (BAS2697)

ExsC was identified in a proteomic study of B. anthracis and B. cereus exosporium proteins (94). The ExsC protein of B. cereus migrates on SDS-PAGE gels at the size predicted for a dimer. The B. anthracis Sterne protein is 67% identical and 82% similar to the B. cereus protein.

ExsFB (BAS2303)

ExsFB is a protein with striking amino acid sequence similarity to ExsFA/BxpB. Mutants lacking ExsFB exhibit a slight reduction in BclA spore content (63). Double mutants lacking both BxpB and ExsFB exhibit a complete loss of the BclA nap layer and have an altered exosporium composition. The residual exosporium present in these double mutants is very fragile and is shed upon storage of the spores. Based upon fluorescent reporter localization studies, it was postulated that ExsFB may be found in the interspace region of the spore (35). My colleagues and I found that ExsFB initially is found at the cap pole of the spore (66). Levels of ExsFB were elevated in bxpB-null and bclA-null mutant spores, with the highest concentration at the cap and smaller amounts around the noncap region of the exosporium.

ExsH

ExsH was identified as a collagen-like spore surface glycoprotein of B. cereus and B. thuringiensis (94, 98, 99). B. cereus ExsH has a C-terminal BclB C-terminal domain (exospore_TM), defined as a C-terminal region in a number of proteins that have extensive collagen-like triple helix repeat regions. The ExsH C-terminal BclB C-terminal domain is predicted by TmHMM (100) to have four transmembrane helices. The N terminus of this protein, upstream of the collagen-like repeats, is more extensive than those of BclA and BclB.

ExsJ

ExsJ was identified as a major glycoprotein in the B. cereus exosporium (94). ExsJ contains a collagen-like GXX repeat sequence and is 81% sequence identical to ExsH, with more of the variation at the N termini of these proteins, upstream of the collagen-like repeats. B. cereus ExsJ has a C-terminal BclB C-terminal domain (exospore_TM) predicted to have four transmembrane helices. The N-terminal domains of the ExsH and ExsJ proteins, upstream of the collagen-like repeats, are longer than those of BclA and BclB. ExsH and ExsJ exhibit some sequence similarity in the N-terminal domains, with 89 of the first 133 residues (67%) being identical.

ExsK (BAS2377)

ExsK was found to be present in at least two distinct locations within the B. anthracis spore, namely, on the surface beneath the BclA nap and in a more interior location, beneath the exosporium basal layer (101). ExsK is incorporated into the spore first at the mother cell proximal pole forming the cap and later becomes distributed around the rest of the spore. The amount of ExsK remains higher at the cap. However, in a mutant lacking the basal layer protein BxpB, ExsK localized to only one spore pole (101). In spores lacking BclA, ExsK fails to assemble into high-molecular-weight species as it does in wild-type spores. Mutant spores lacking ExsK were found to germinate faster and more extensively than their wild-type counterparts (101).

ExsM (BAS2174)

ExsM, a basic protein with a molecular mass of 15.4 kDa, was first identified in a proteomic study of B. cereus (94). B. cereus exsM-null mutant spores were reported to be smaller and rounder than wild-type spores, and 77% of the mutant spores exhibited a double-exosporium arrangement, with both exosporia possessing the BclA nap layer (93). The ExsM-deficient spores were reported to be more resistant to lysozyme and germinated at a higher rate, but they exhibited a delayed outgrowth (93). Mutational loss of B. anthracis exsM produced less pronounced phenotypes with regard to spore size, yielded a lower percentage of duplicate exosporia (30%), and the exosporium was more fragile than that of the wild type (93).

ExsY (BAS1141)

ExsY is an exosporium protein that has been characterized in both B. anthracis and B. cereus (34, 42, 44). A B. anthracis ΔexsY mutant produces an exosporium only at the mother cell central pole of the spore (the cap), while the remainder of the exosporium fails to develop (42). This cap structure is lost from spores prepared in liquid medium but can be detected on spores prepared on solid medium. The cap site exosporium has a BclA nap but shows evidence of a disorganized basal layer. Boydston et al. (42) also reported that the exosporium sublayers (below the BxpB-containing basal layer) were more fragile and easily lost from the spores in the ΔexsY mutant. B. cereus spores (prepared in liquid medium) from a mutant strain containing an insertionally inactivated exsY determinant failed to produce an organized exosporium (44). The surfaces of the released spores were decorated with what appeared to be exosporium fragments (44).

GerQ (YwdL; BAS5242)

GerQ is an exosporium protein first characterized in B. cereus and B. thuringiensis (102, 103). Spores lacking GerQ produce a more fragile exosporium. Immunogold labeling revealed that GerQ is found on the inner surface of the exosporium. Mutant spores lacking GerQ had a reduced rate of germination in response to alanine plus inosine and failed to germinate in response to the germinant receptor-independent calcium dipicolinate (104). GerQ/YwdL is required for retention of CwlJ (a germination-specific cortex lytic enzyme) in the B. subtilis spore (104).

IunA (BAS4961)

IunA is a putative inosine-uridine-preferring nucleoside hydrolase identified as a B. anthracis exosporium protein (K. A. Spreng and G. C. Stewart, unpublished data). It is surface exposed but is more strongly recognized by anti-IunA antibodies in the absence of the BclA nap.

IunH (BAS2693)

IunH is an inosine-uridine-preferring nucleoside hydrolase first identified as a B. anthracis exosporium protein by Lai et al. (106) and Redmond et al. (34). The IunH protein, similar to inosine-uridine-preferring nucleoside hydrolases, is found in the spore, possibly in the interspace region (35).

Superoxide Dismutase Sod15 (BAS1378)

Superoxide dismutases were identified in exosporium preparations by Steichen et al. (33) and Liu et al. (74). Superoxide dismutase molecules within the spore may afford B. anthracis protection against oxidative stress and enhance pathogenicity in animal models of infection (107). The presence of multiple superoxide dismutase genes may provide redundancy needed to ensure protection of germinating spores in animal hosts. The sod15 determinant is transcribed in the late exponential and sporulation phases, consistent with its spore localization (108).

Superoxide Dismutase SodA1 (BAS4177)

SodA1 was identified in exosporium preparations by Steichen et al. (33) and Liu et al. (74). Its contributions to oxidative stress and virulence were examined by Cybulski et al. (107). A strain lacking all four superoxide dismutases encoded in the B. anthracis genome was found to be attenuated for virulence upon intranasal challenge of mice (107).

Cataloging of exosporium proteins remains incomplete. My colleagues and I employed a genetic screen to identify genes whose proteins are important for proper incorporation of BclA into the exosporium (109). Several determinants were identified which either localize to the spore (based on gene fusion studies) or are required for full BclA assembly. These may represent either exosporium proteins or possibly nonspore proteins that participate in exosporium assembly.

FUNCTIONS OF THE EXOSPORIUM IN THE ENVIRONMENT

Spores present in soil environments constitute the source of animal infections by B. anthracis and other soil-associated pathogens. Furthermore, it has been shown that under the right conditions, B. anthracis spores can germinate on and around plant roots and that genetic exchange between strains of Bacillus is possible (110). Because of the spore mode of infection, it is important for spores to persist at their site of deposition and not to get washed too deeply into the soil and thus be inaccessible to grazing ruminants in amounts constituting an infectious dose (111). The exosporium affects the ability of B. anthracis spores to bind to different soil types. Wild-type spores of B. anthracis adhere more strongly than exosporium-deficient spores to soils with high organic material and calcium levels, i.e., soils which have been shown to favor the persistence of this pathogen (112). The presence of BclA on the surfaces of spores resulted in better retention of spores in a porous medium of silica sand (113).

The exosporium layer and the BclA glycoprotein, in spores possessing it, contribute to the overall hydrophobicity of the spore (80, 84, 99). The hydrophobic character contributes to the binding properties of the spore in the environment. Binding to food preparation surfaces is important to the foodborne pathogen B. cereus. Spore adherence to surfaces, such as stainless steel, contributes to the environmental persistence of the spores and ultimately to infection of foods by this bacterium.

Studies have attributed roles for the exosporium (114, 115) or BclA (78, 99) in binding to stainless steel surfaces. In contrast to the results obtained with B. anthracis, a study with a strain of B. cereus (ATCC 14579) indicated an increase in spore hydrophobicity in bclA deletion mutants, and this coincided with a reduction in spore binding to stainless steel (99). However, overall spore hydrophobicity does not adequately explain the differences in stainless steel adhesion properties. Other reports have concluded that the overall hydrophobicity of the spore and the presence of surface appendages by these spores have the greatest impacts on stainless steel binding (116,118).

FUNCTIONS OF THE EXOSPORIUM IN AN INFECTED HOST

The majority of studies concerned with spore infectious processes have focused on the zoonotic pathogen B. anthracis. This bacterium has an unusual host-pathogen interaction. The bacterium, in order to disseminate as many infectious spores into the soil as possible, has evolved mechanisms to gain access to the host's bloodstream, where it replicates to large numbers and kills the host through elaboration of toxins. The sporulation process does not occur to any significant extent in the infected host until after the animal dies and postmortem carcass decomposition provides the oxygen-replete environment needed to induce sporulation. The large bacterial numbers then translate to large numbers of spores released into the soil around the decomposing carcass. To accomplish this infectious process, the infecting spores must transit from the site of infection to the draining lymph node, germinate to the vegetative bacterial form, and initiate invasion of the bloodstream and rapid replication. The exosporium, in particular the surface-exposed BclA protein, participates in many of the initial spore-host interactions.

Studies with animal models of disease and tissue culture-based infections have provided clues to the nature of the initial events in the infectious process. These studies indicate that when spores are introduced into an animal host, they are taken up by macrophages and dendritic cells (119,123). Spores in the lung are phagocytosed by alveolar macrophages or dendritic cells, which initiate the processes leading to pulmonary anthrax (120, 122, 123). For pulmonary infections, it has also been shown that B. anthracis spores are capable of entering epithelial cells of the lung and crossing this cell layer without apparent disruption of the barrier integrity (124, 125). The gastrointestinal form of anthrax ensues when spores are ingested and migrate to the intestinal tract, where they pass across the gut epithelial cell layer without germination. Once outside the intestines, the spores are taken up by the macrophages of the Peyer's patches (119). In either form of anthrax, the phagocytes migrate to the draining lymph nodes. The spores germinate en route within the macrophage or dendritic cell and begin to replicate and produce toxins. At the lymph node, the bacterium-containing phagocytes lyse, releasing the bacteria. Invasion of the bloodstream, replication, toxin elaboration, and death of the susceptible host ensue.

Uptake of spores by phagocytic host cells has been shown to involve interactions between BclA and the host integrin Mac-1 (CR3) (126). Spores lacking the BclA glycoprotein do not specifically target these cells but bind more generally to epithelial cells (127). CD14 was shown to bind to rhamnose residues of BclA and to act as a coreceptor for spore binding by Mac-1 (128). Oliva et al. proposed a model whereby binding of rhamnose to CD14 triggers signaling through Toll-like receptor 2 (TLR-2), leading to activation of phosphatidylinositide 3-kinases and converting Mac-1 into a more active receptor for B. anthracis spores (128). Survival studies using C57BL/6 and CD11b−/− mice revealed that CD11b−/− mice lacking Mac-1 were more resistant to infection with wild-type but not BclA-deficient spores (126, 128).

These results suggest that the spore surface BclA fibers may act to promote uptake by professional phagocytes and to inhibit nonspecific interactions between B. anthracis spores and nonprofessional phagocytic cells. Gu et al. reported that activation of the classical complement pathway is a primary mechanism of spore phagocytosis by murine macrophages (129). Phagocytic uptake of spores was found to be reduced significantly in the absence of complement component C1q or C3. C1q recruitment to the spore surface was shown to be mediated by BclA, which can directly bind C1q in a dose-dependent and saturable manner (129, 130). Furthermore, C1q acted as a bridging molecule between BclA and integrin α2β1 to mediate B. anthracis spore entry into epithelial cells in a complement activation-independent manner. A binding protein for the globular head domains of complement component C1q, designated gC1qR, was also reported to be involved in at least the initial stages of B. cereus spore attachment and/or entry (131).

The above findings suggest a central role for BclA in B. anthracis spore infection of hosts. However, loss of BclA has been shown to have no effect on virulence of B. anthracis in animal models of infection, as measured by 50% lethal dose (LD50) values (84, 132). Additionally, the mean time to death following subcutaneous or intranasal inoculation of ΔbclA mutant spores in an A/J mouse model of infection was shorter than that with wild-type spores (84). Oliva et al. found that C57BL/6 mice were more susceptible to ΔbclA spores than to wild-type spores (the LD50 of ΔbclA spores was approximately 5-fold lower than that of wild-type Sterne spores in a subcutaneous challenge model, and the mean time to death was shorter with the BclA-lacking spores) (126). They also reported similar results with A/J mice after intrathecal spore administration. A better understanding of the early events in the infectious process is needed before we can explain the discrepancy in results reported for animal disease models and why loss of the BclA spore surface glycoprotein either has a minimal impact on the lethal dose or creates an increase in virulence, depending on the animal model of infection utilized.

Spores of B. anthracis can persist in the lungs of infected hosts following aerosol exposure. For example, viable spores were found in the lungs of rhesus macaques at least 42 days following aerosol exposure (133,135) and in mouse lungs for more than 25 to 30 days (136, 137). Spores of B. anthracis were significantly better at persisting in the lung than spores of a non-exosporium-possessing B. subtilis strain, and some of the persisting spores were found within epithelial cells (138). Treatment of mice with an Src family kinase inhibitor significantly reduced B. anthracis dissemination from the lung to distal organs and prolonged the median survival time for mice compared to those for the untreated control group (139). There appears to be a signaling pathway specifically required for spore entry into epithelial cells, and these studies provided evidence suggesting that this pathway is important for dissemination and virulence in vivo (139).

Another possible mechanism whereby spores persist in the lungs involves innate immune system evasion. Interference with the innate immune system may initially occur upon introduction of spores into a susceptible host. Plasminogen efficiently binds to spores in an exosporium- and BclA-dependent manner (140). Plasminogen-bound spores are capable of exhibiting antiopsonic properties by cleaving C3b molecules, resulting in a decrease in macrophage phagocytosis, possibly contributing to a longer persistence of free spores in infected lung tissue.

Perhaps as a consequence of the decreased hydrophobicity of BclA-deficient spores, the extracellular matrix proteins laminin and fibronectin bound significantly better to ΔbclA mutant spores than to wild-type spores (84). BclA-negative spores were also found to adhere better to epithelial cells, fibroblasts, and endothelial cells but not macrophages. Spores lacking the exosporium layer were found to elicit a higher cytokine response (beta interferon, interleukin-1β [IL-1β], tumor necrosis factor alpha [TNF-α], and IL-6) in mouse macrophages than that in response to exosporium-bearing spores (141). This suggests a role for the exosporium in avoidance of innate immune signaling during the early stages of spore infection.

Studies have suggested that the timing of spore germination within phagocytic cells is regulated to prevent premature germination of spores. The timing of germination is thought to be critical to the outcome of the spore-macrophage interaction (70). The exosporium is involved in the initial spore interactions with the intracellular compartment of phagocytes following spore uptake. Enzymes are present in the outer spore layers, apparently to prevent premature germination of the internalized spores and to provide protection to the emerging vegetative cell following germination of the spore (34). These spore-associated enzymes include alanine racemase to convert l-alanine (a germinant) to d-alanine (an inhibitor of l-alanine-induced germination) and at least two, and possibly more, inosine-preferring nucleoside hydrolases that may function to inactivate inosine (another germinant species) (35, 69, 70, 101, 142).

The B. anthracis exosporium contains an arginase that metabolizes l-arginine to l-ornithine and urea, thereby reducing the levels of arginine, the substrate for the host cell nitric oxide synthase (71, 72). This decreases the amount of NO radicals produced in spore-infected macrophages. Counterintuitively, B. anthracis spores have their own NO synthase. Spores lacking this enzyme were attenuated for virulence in an A/J mouse model of infection and were more susceptible to killing when germinating within macrophages (143). Resistance to macrophage killing is thought to be dependent on NO-mediated activation of bacterial catalase and suppression of the Fenton reaction. Additionally, a potential exosporium-associated virulence factor is superoxide dismutase, which may provide protection to the germinating spores against reactive oxygen species-mediated defense systems of phagocytic cells (107, 108). These spore-associated enzymes may protect the newly emerging vegetative cells at a time when they are very sensitive to killing by reactive oxygen species, until toxin production by the newly emerging vegetative bacteria can ensue and ultimately kill the host cell.

CONTRIBUTIONS OF EXOSPORIUM PROTEINS TO PROTECTIVE IMMUNITY

The first demonstration of an effective vaccine against anthrax in animals was conducted by Louis Pasteur at the famous public experiment at Pouilly-le-Fort, in May 1881, to demonstrate his concept of vaccination. The vaccine consisted of B. anthracis spores that were chemically attenuated for virulence. The anthrax vaccine widely used in veterinary medicine today is a live spore vaccine and involves the use of the Sterne strain, which lacks the pXO2 plasmid and thus cannot produce the bacterial cell capsule, an important virulence factor for this bacterium. This vaccine was originally developed at the Onderstepoort Laboratory, Pretoria, South Africa. Human vaccines against anthrax, in contrast, are cell-free preparations whose efficacy is considered to be due primarily to antibodies generated against the protective antigen component of the anthrax lethal and edema toxins. These include the American anthrax vaccine adsorbed (AVA; BioThrax) and the British anthrax vaccine precipitated (AVP). Animal studies of the efficacy of the human vaccine found that protection against the most virulent strains of B. anthracis required the addition of spore antigens to the vaccination regimen (144,147). Improved vaccine protection was observed when the added spore antigens included inactivated whole spores (147,150), live attenuated spores (146, 151,154), BclA (81, 83, 148, 155), BxpB (68, 148), and BAS5303 (68, 148). Antibodies directed against whole spores or, specifically, BclA have been shown to be inhibitory to spore germination in vitro (148, 156). Studies involving passive immunization with spore-specific antisera have produced mixed results. Protection has been reported in some studies but could not be demonstrated in others (148, 157,159). The distinct difference between the unambiguous benefit obtained with active immunization with PA and spore components and the less clear-cut results obtained via passive immunization involving anti-spore-component antibodies underscores the likely involvement of cellular immunity in protection (159).

BACTERIA WITH EXOSPORIUM-CONTAINING SPORES

Several endospore-producing bacteria have been reported to possess exosporia. Examples are listed in Table 1. In addition, genes similar to bclA of B. anthracis are present in the genome sequences of Clostridium autoethanogenum DSM 10061, Clostridium beijerinckii NCIMB 8052, Clostridium ljungdahlii DSM 13528, Clostridium saccharolyticum WM1, and Clostridium ramosum.

TABLE 1

Exosporium-producing bacteria

OrganismCommentReference(s)
Bacillus anthracis160, 161
Bacillus badiusSpore surface appendages present161, 162
Bacillus cereusSpore surface appendages present61
Bacillus circulansSpore surface appendages present162
Bacillus firmusSpore surface appendages present161
Bacillus fusiformis162
Bacillus lentusSpore surface appendages present161, 162
Bacillus megateriumStrain-variable feature163
Bacillus mycoides162, 164, 165
Bacillus niacini162
Bacillus pseudomycoides162
Bacillus pumilus162
Bacillus thiaminolyticusSpore surface appendages present161
Bacillus thuringiensisSpore surface appendages present166
Bacillus vedderi162
Brevibacillus brevisSpore surface appendages present161
Brevibacillus laterosporus162, 167
Clostridium bifermentansSpore surface appendages present168
Clostridium botulinumTubular surface appendages168,172
Multilayered exosporium basal layer
Has nap; interspace layered filaments
Clostridium difficileStrain-variable feature?173, 174
No prominent interspace with most strains (perhaps a crust layer?)
Clostridium pasteurianumExosporium incompletely envelopes the spore (has a terminal opening)175
Clostridium septicumMultilayered176
Clostridium sordelliiSpore surface appendages present168
Clostridium sporogenesSpore surface appendages present168
Lysinibacillus odysseyi162, 177
Lysinibacillus sphaericusThe insecticidal binary toxin is located in the spore interspace adherent to the internal surface of the exosporium178, 179
Paenibacillus alveiSpore surface appendages present162
Pasteuria penetransPeripheral fibers for spore attachment within the interspace180
Rummeliibacillus pycnus162
Viridibacillus neideiSpore surface appendages present162

Bacillus megaterium

There appears to be strain heterogeneity with regard to exosporium production in B. megaterium. An exosporium was visible in spores of more than half of 36 strains examined (163). The B. megaterium exosporium was described in detail and a hairlike nap layer seen on the spore surface in an electron microscopy study by Beaman et al. (181). Hexagonal periodicity was observed with the exosporium basal layer. In addition, longer fibrillar material with a laminar appearance radiated outward from the spore surface. Planar inclusions were evident in the interspace region when spores were visualized by transmission electron microscopy. A novel morphogenetic spore coat protein, BMQ_0737, was found to be important for the proper assembly of the outer spore coat and exosporium layers of the B. megaterium QM B1551 spores (182). A pole-localized nap layer of strain QM B1551 is composed of the BclA1 protein, and this polar nap localization is independent of BxpB (183). A plasmid-free derivative of strain QM B1551 was found to lack the exosporium layer, and the spores were less hydrophobic (183). Of the seven resident plasmids, only pBM600 and possibly pBM500 are required for exosporium production in sporulating cells of this strain. Plasmid-carried orthologues of B. cereus family exosporium genes were identified. Two bclA-like determinants are carried on pBM500 and were designated bclA1 and bclA2. An orthologue of the B. cereus family BxpB/ExsFA exosporium basal layer protein is encoded on plasmid pBM600. Fluorescent protein fusion studies indicated that the BclA1 protein localizes around the developing forespore, but fluorescence was not detected on the surfaces of released spores (183).

Clostridium botulinum

In an early study, spores of C. botulinum type E were found to possess a fragile exosporium layer with tubular appendages (with various lengths of up to 0.56 μm and a 200-Å diameter, with a lumen of 80 Å; a hemispherical cap composed of 40-Å spherical subunits is present at the end of the appendage) present in the interspace region of the spore (171). In a later report involving a type A strain, spores were reported to possess exosporia which were multilayered in appearance, containing up to 15 layers (172). The layers were uniform in appearance and were 3 nm thick, with a center-to-center distance of 7 nm.

Clostridium difficile

Several articles have been published that refer to an exosporium layer of C. difficile (recently proposed to be reclassified as Peptoclostridium difficile) (184). In only one publication is there electron microscopic evidence of a possible traditional exosporium layer with a defined basal layer and interspace region (185). In this report, as in others, micrographs indicate an external crust-like layer (such as that shown in Fig. 8) or fragments of material adherent to the spore coat in strains of C. difficile (13, 174, 185, 186). Joshi et al. also indicated a pronounced variation in spore hydrophobicity among strains, with the more hydrophobic strains possessing a more organized exosporium (173). These studies suggest that possession of an exosporium may not be a feature of C. difficile spores or may be a strain-variable trait in this bacterium. An alternative possibility is that the exosporium layer is extremely fragile and readily lost from spores (13). The more likely possibility is that C. difficile spores lack a true exosporium layer (as proposed previously [13]) and possess a crust layer, albeit one with a pronounced hairlike nap. The bacterium exhibits strain variability in the appearance of the crust layer nap. A proteomic study did not reveal exosporium proteins similar to the known B. cereus family exosporium proteins, except for the presence of collagen-like proteins (187). However, the lack of B. cereus family exosporium homologs does not rule out the existence of C. difficile exosporium structures, since genomic analyses of known exosporium producers outside the B. cereus family suggest that exosporium proteins differ markedly in various species.

An external file that holds a picture, illustration, etc.
Object name is zmr0041524030008.jpg

Transmission electron micrographs of C. difficile strain ATCC 700057 spores. The spores were stained with ruthenium red, a carbohydrate stain to identify the glycoprotein layer of the spore surface. The spores feature a prominent cortex region (the white area surrounding the core). An interspace region, the area between the exosporium and the outer spore coat, is not an obvious feature in C. difficile spores. The ruthenium red-stained putative glycoprotein layer is denoted by arrows.

Three genes encoding collagen-like glycoproteins have been identified in the C. difficile genome (bclA1, bclA2, and bclA3), and the encoded proteins are expressed during sporulation and localize to the spore, with BclA1 likely present in the outermost “exosporium/crust” layer of the spore (13, 188). BclA1, unlike BclA of B. anthracis, was reported to be poorly immunogenic in this study. ClosTron-generated mutants of the bclA1 and bclA2 determinants produced spores with altered surfaces, including release of “sheets of coat-like material” from the spores (188). BclA3 was identified as a glycosylated protein in extracts of C. difficile spores, and the glycosyltransferase gene sgtA, carried upstream of bclA3, is at least partly responsible for glycosylation of the spore surface (189).

A cysteine-rich protein (CdeC) localized to the outermost spore layer was found to be important for the formation of the nap layer of the C. difficile spore and might be the anchoring protein linking the outermost layer to the spore coat (190). The CotA protein was shown to be important for proper assembly of the outermost spore layer, and the CotA, CotE (a bifunctional peroxiredoxin reductase and chitinase), CotF, and SodA (superoxide dismutase) proteins were found to be associated with the spore surface (185). The C. difficile bclA1, bclA2, bclA3, cdeC, cotA, cotCB, cotD, and cotE genes were determined to be dependent on σK for expression; this is the same sigma factor involved in exosporium gene expression in the B. cereus family of organisms (191).

A proteomic analysis of proteins extracted from the outermost layers of C. difficile strain 630 identified 184 distinct proteins (192). Proteins were extracted from spores by use of trypsin treatment or sonication methods. In this report, Flag fusion studies indicated an “exosporium” location for the BclA1, BclA2, BclA3, CdeA, CdeB, CdeC, and CdeM proteins.

The outermost layer of C. difficile spores has been reported to interact with a receptor or receptors on intestinal epithelial cells (193) and may have an adverse effect on macrophage phagosomal membranes (194). Proteolytic treatment of spores decreased adherence to the cells. Removal of the outermost spore layer(s) was also reported to reduce adherence to colonic cells, decrease overall spore hydrophobicity, and increase spore colony-forming frequencies in vitro (13, 173). The more pronounced exosporium-like layers may correlate with increased spore hydrophobicity and better adherence to gut epithelial cell monolayers and environmental stainless steel surfaces (174). Loss of each of the BclA1–3 glycoproteins resulted in decreased spore hydrophobicity. Relative to the wild-type parent strain, mutant spores lacking BclA1 were shown to have a 2-log higher infectious dose requirement in a mouse model of infection and displayed reduced virulence in a hamster infection model (188). BclA1 is thought to be important in the initial colonization of spores in the gut epithelium of the infected host.

Clostridium sporogenes

Spores of the anaerobic bacterium C. sporogenes possess an unusual exosporium feature. The exosporium is reported to have a protrusion, or “sporiduct,” at one pole of the spore, which in some spores becomes an opening or aperture in the exosporium (195, 196). Functions ascribed to the exosporium include roles in germination, outgrowth, and attachment (195, 197). A recent report indicates that the germination and outgrowth process for C. sporogenes strain ATCC 15579 involves rupture of the spore coat at a site adjacent to the exosporium aperture (196). The emerging cell exits the spore coat and exosporium through these aligned openings.

Lysinibacillus sphaericus (formerly Bacillus sphaericus)

A lamellar exosporium with a hexagonal arrangement of subunits was reported in an electron microscopic study of sporulation in the mosquitocidal bacterium L. sphaericus (178). Within the spore interspace, the insecticidal parasporal crystal is enclosed in a mesh-like parasporal envelope (198). A collagen-like protein, BclS, was identified as being part of filamentous structures in the interspace region of the spores (199). Heat resistance and germination defects were noted in spores lacking BclS.

Pasteuria penetrans

Pasteuria penetrans is a parasite of root-knot nematodes whose spores are the infectious form. The spores have been commercialized as a biological control agent (200). Pasteuria endospores have additional structures within the interspace. The perisporium largely consists of peripheral fibers that are present in the interspace and attached to the spore body. These fibers are involved in attachment of the spores to the nematode cuticle (180). After attachment to host cuticle cells, the exosporium is lost, exposing the peripheral fibers to the cuticles of nematode juveniles. Binding of spores to the host cuticle has been proposed to involve collagen-like fibers on the surface of the endospore interacting with mucins on the nematode cuticle (201).

Pasteuria ramosa

P. ramosa infects water fleas of the Daphnia genus. Infected hosts are completely sterilized and have a reduced life span. A collagen-like protein containing GXY repeats (Pcl1a [Pasteuria collagen-like protein 1a]), one of 37 collagen-like proteins encoded in the P. ramosa genome, was identified in a proteomic analysis of spore surface proteins of P. ramosa (202, 203). The specific location of the Pcl1a protein in the spore has not been determined.

THE EXOSPORIUM AS A SPORE DISPLAY PLATFORM

The BclA collagen-like glycoprotein is abundantly expressed on the surfaces of spores of B. anthracis, B. cereus, and B. thuringiensis. The N-terminal 35 amino acids of BclA have been shown to be sufficient for localization of the protein around the spore and its stable incorporation onto the exosporium basal layer surface (57, 65). It is possible to display foreign proteins on the surfaces of B. anthracis, B. cereus, or B. thuringiensis spores by incorporating the 35 BclA N-terminal codons at the N-terminal end of the foreign gene open reading frame and expressing it in sporulating cells by utilizing the bclA promoter and ribosome binding site. The foreign protein is synthesized by the sporulating cells and stably incorporated in large numbers at the spore surface. The recombinant protein-decorated spores are easily purified and can be employed as microparticles for industrial, bioremediation, or vaccine applications (H. Y. Hsieh, G. C. Stewart, B. M. Thompson, and C. H. Lin, unpublished data). Enzymes that retain activity when tethered at the N terminus are ideal for these biocatalytic processes. The spore display system is ideally suited for soil environments, where spores can remain in place following application, whereas free enzymes tend to wash out and are more readily inactivated.

CONCLUSIONS

Our understanding of the composition, assembly process, and function of the spore exosporium layer remains incomplete. The presence of exosporium proteins in large SDS-resistant complexes hampers studies of specific protein-protein interactions. As with other aspects of endospore formation, genetic studies will likely be the best route to a more complete understanding of the composition, assembly pathway, and function of this outermost spore layer. The distribution of sugar residues on the spore surface glycoproteins and the roles that they may play are understudied aspects of spore biology. Atomic force microscopy was recently utilized to begin to define the distribution of sugars on the surfaces of B. cereus spores (205). Regulation of exosporium formation warrants further investigation. Recently, it was discovered that production of spores of a particular strain of B. cereus under conditions of anaerobiosis resulted in defects in the exosporium structure (206). Based on what we know regarding spore formation by B. anthracis, it could be anticipated that reduced oxygen tension might result in reduced sporulation efficiencies. However, structural defects in the exosporia of the formed spores were unexpected. How exosporia evolved in different species by using unique building blocks will be a fascinating field of study as our understanding of non-B. cereus group exosporium composition improves. It is apparent that the studies begun in the 19th century by Robert Koch still have a lot to disclose regarding the biology of endospores.

ACKNOWLEDGMENTS

Work in my laboratory was supported by grant AI101093 from the National Institutes of Health.

I thank Michael Calcutt for his critical reading of the manuscript and for helpful suggestions and Krista Spreng for the electron micrographs shown in Fig. 8.

Biography

• 

George C. Stewart received his B.S. degree from North Texas State University in 1975 and his Ph.D. degree from the University of Texas Health Science Center at Dallas (now UT Southwestern Medical Center) in 1980. Following postdoctoral studies in bacterial molecular genetics at the University of North Carolina at Chapel Hill Medical School, he held faculty positions at the University of Kansas, University of South Carolina School of Medicine, Kansas State University College of Veterinary Medicine, and University of Missouri College of Veterinary Medicine, where he is currently the McKee Professor of Microbial Pathogenesis and Chairman of the Department of Veterinary Pathobiology. He is interested in the genetics and physiology of Bacillus anthracis, with an emphasis on the structure and synthesis of the exosporium layer of the spore.

Notes

This review is dedicated to the memory of Alvin Fox, who started with me down the road of B. anthracis spore research.

REFERENCES

1. Henriques AO, Moran CP Jr. 2007. Structure, assembly, and function of the spore surface layers. Annu Rev Microbiol 61:555–588. 10.1146/annurev.micro.61.080706.093224. [Abstract] [CrossRef] [Google Scholar]
2. Ghosh S, Korza G, Maciejewski M, Setlow P. 2015. Analysis of metabolism in dormant spores of Bacillus species by 31P-NMR of low molecular weight compounds. J Bacteriol 197:992–1001. 10.1128/JB.02520-14. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
3. McKenney PT, Driks A, Eichenberger P. 2013. The Bacillus subtilis endospore: assembly and functions of the multilayered coat. Nat Rev Microbiol 11:33–44. 10.1038/nrmicro2921. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
4. Setlow P. 2014. Germination of spores of Bacillus species: what we know and do not know. J Bacteriol 196:1297–1305. 10.1128/JB.01455-13. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
5. Setlow P, Johnson EA. 2012. Spores and their significance, p 45–79. In Doyle MP, Buchanan R (ed), Food microbiology, fundamentals and frontiers, 4th ed ASM Press, Washington, DC. [Google Scholar]
6. Logan NA. 2012. Bacillus and relatives in foodborne illness. J Appl Microbiol 112:417–429. 10.1111/j.1365-2672.2011.05204.x. [Abstract] [CrossRef] [Google Scholar]
7. Bottone EJ. 2010. Bacillus cereus, a volatile human pathogen. Clin Microbiol Rev 23:382–398. 10.1128/CMR.00073-09. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
8. Horii T, Notake S, Tamai K, Yanagisawa H. 2011. Bacillus cereus from blood cultures: virulence genes, antimicrobial susceptibility and risk factors for blood stream infection. FEMS Immunol Med Microbiol 63:202–209. 10.1111/j.1574-695X.2011.00842.x. [Abstract] [CrossRef] [Google Scholar]
9. Sasahara T, Hayashi S, Morisawa Y, Sakihama T, Yoshimura A, Hirai Y. 2011. Bacillus cereus bacteremia outbreak due to contaminated hospital linens. Eur J Clin Microbiol Infect Dis 30:219–226. 10.1007/s10096-010-1072-2. [Abstract] [CrossRef] [Google Scholar]
10. Glare TR, O'Callaghan M (ed). 2000. Bacillus thuringiensis: biology, ecology, and safety. John Wiley & Sons, Ltd, Chichester, United Kingdom. [Google Scholar]
11. Palma L, Muñoz D, Berry C, Murillo J, Caballero P. 2014. Bacillus thuringiensis toxins: an overview of their biocidal activity. Toxins 6:3296–3325. 10.3390/toxins6123296. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
12. Dubberke ER, Olsen MA. 2012. Burden of Clostridium difficile on the healthcare system. Clin Infect Dis 55:S88–S92. 10.1093/cid/cis335. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
13. Paredes-Sabja D, Shen A, Sorg JA. 2014. Clostridium difficile spore biology: sporulation, germination, and spore structural proteins. Trends Microbiol 22:406–416. 10.1016/j.tim.2014.04.003. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
14. Pizarro-Guajardo M, Olguín-Araneda V, Barra-Carrasco J, Brito-Silva C, Sarker MR, Paredes-Sabja D. 2014. Characterization of the collagen-like exosporium protein, BclA1, of Clostridium difficile spores. Anaerobe 25:18–30. 10.1016/j.anaerobe.2013.11.003. [Abstract] [CrossRef] [Google Scholar]
15. Miller BA, Chen LF, Sexton DJ, Anderson DJ. 2011. Comparison of the burdens of hospital-onset, healthcare facility-associated Clostridium difficile infection and of healthcare-associated infection due to methicillin-resistant Staphylococcus aureus in community hospitals. Infect Control Hosp Epidemiol 32:387–390. 10.1086/659156. [Abstract] [CrossRef] [Google Scholar]
16. Kutty PK, Woods CW, Sena AC, Benoit SR, Naggie S, Frederick J, Evans S, Engel J, McDonald LC. 2010. Risk factors for and estimated incidence of community-associated Clostridium difficile infection, North Carolina, USA. Emerg Infect Dis 16:197–204. 10.3201/eid1602.090953. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
17. Lindström M, Heikinheimo A, Lahti P, Korkeala H. 2011. Novel insights into the epidemiology of Clostridium perfringens type A food poisoning. Food Microbiol 28:192–198. 10.1016/j.fm.2010.03.020. [Abstract] [CrossRef] [Google Scholar]
18. Scallan E, Hoekstra RM, Angulo FJ, Tauxe RV, Widdowson MA, Roy SL, Jones JL, Griffin PM. 2011. Foodborne illness acquired in the United States—major pathogens. Emerg Infect Dis 17:7–15. 10.3201/eid1701.P11101. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
19. Yang CC, Hsu PC, Chang HJ, Cheng CW, Lee MH. 2013. Clinical significance and outcomes of Clostridium perfringens bacteremia—a 10-year experience at a tertiary care hospital. Int J Infect Dis 17:e955–e960. 10.1016/j.ijid.2013.03.001. [Abstract] [CrossRef] [Google Scholar]
20. Lee HL, Cho SY, Lee DG, Ko Y, Hyun JI, Kim BK, Seo JH, Lee JW, Lee S. 2014. A fatal spontaneous gas gangrene due to Clostridium perfringens during neutropenia of allogeneic stem cell transplantation: case report and literature review. Infect Chemother 46:199–203. 10.3947/ic.2014.46.3.199. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
21. Allaart JG, van Asten AJ, Gröne A. 2013. Predisposing factors and prevention of Clostridium perfringens-associated enteritis. Comp Immunol Microbiol Infect Dis 36:449–464. 10.1016/j.cimid.2013.05.001. [Abstract] [CrossRef] [Google Scholar]
22. Centers for Disease Control and Prevention. 2014. Botulism. CDC, Atlanta, GA: http://www.cdc.gov/nczved/divisions/dfbmd/diseases/botulism/. [Google Scholar]
23. Espelund M, Klaveness D. 2014. Botulism outbreaks in natural environments—an update. Front Microbiol 5:287. 10.3389/fmicb.2014.00287. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
24. Rosow LK, Strober JB. 2015. Infant botulism: review and clinical update. Pediatr Neurol 52:487–492. 10.1016/j.pediatrneurol.2015.01.006. [Abstract] [CrossRef] [Google Scholar]
25. Eijlander RT, de Jong A, Krawczyk AO, Holsappel S, Kuipers OP. 2014. SporeWeb: an interactive journey through the complete sporulation cycle of Bacillus subtilis. Nucleic Acids Res 42:D685–D691. 10.1093/nar/gkt1007. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
26. Koch R. 1876. Untersuchungen über Bakterien. V. Die Äetiologie der Milzbrand-Krankheit, begründet auf die Entwicklungsgeschichte des Bacillis anthracis. Beitr Biol Pflanz 2:277–310. [Google Scholar]
27. Flügge C. 1886. Die Mikroorganismen, 2nd ed Verlag F C W Vogel, Leipzig, Germany. [Google Scholar]
28. Hachisuka Y, Kojima K, Sato T. 1966. Fine filaments on the outside of the exosporium of Bacillus anthracis spores. J Bacteriol 91:2382–2384. [Europe PMC free article] [Abstract] [Google Scholar]
29. Kramer MJ, Roth IL. 1968. Ultrastructural differences in the exosporium of the Sterne and Vollum strains of Bacillus anthracis. Can J Microbiol 14:1297–1299. 10.1139/m68-217. [Abstract] [CrossRef] [Google Scholar]
30. Driks A. 2002. Maximum shields: the assembly and function of the bacterial spore coat. Trends Microbiol 10:251–254. 10.1016/S0966-842X(02)02373-9. [Abstract] [CrossRef] [Google Scholar]
31. Sylvestre P, Couture-Tosi E, Mock M. 2002. A collagen-like surface glycoprotein is a structural component of the Bacillus anthracis exosporium. Mol Microbiol 45:169–178. 10.1046/j.1365-2958.2000.03000.x. [Abstract] [CrossRef] [Google Scholar]
32. Sylvestre P, Couture-Tosi E, Mock M. 2003. Polymorphism in the collagen-like region of the Bacillus anthracis BclA protein leads to variation in exosporium filament length. J Bacteriol 185:1555–1563. 10.1128/JB.185.5.1555-1563.2003. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
33. Steichen C, Chen P, Kearney JF, Turnbough CL Jr. 2003. Identification of the immunodominant protein and other proteins of the Bacillus anthracis exosporium. J Bacteriol 185:1903–1910. 10.1128/JB.185.6.1903-1910.2003. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
34. Redmond C, Baillie LW, Hibbs S, Moir AJ, Moir A. 2004. Identification of proteins in the exosporium of Bacillus anthracis. Microbiology 150:355–363. 10.1099/mic.0.26681-0. [Abstract] [CrossRef] [Google Scholar]
35. Giorno R, Mallozzi M, Bozue J, Moody K-S, Slack A, Qiu D, Wang R, Friedlander A, Welkos S, Driks A. 2009. Localization and assembly of proteins comprising the outer structures of the Bacillus anthracis spore. Microbiology 155:1133–1145. 10.1099/mic.0.023333-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
36. Ball DA, Taylor R, Todd SJ, Redmond C, Couture-Tosi E, Sylvestre P, Moir A, Bullough PA. 2008. Structure of the exosporium and sublayers of spores of the Bacillus cereus family revealed by electron crystallography. Mol Microbiol 68:947–958. 10.1111/j.1365-2958.2008.06206.x. [Abstract] [CrossRef] [Google Scholar]
37. Couture-Tosi E, Ranck J-L, Haustant G, Pehau-Arnaudet G, Sachse M. 2010. CEMOVIS on a pathogen: analysis of Bacillus anthracis spores. Biol Cell 102:609–619. 10.1042/BC20100080. [Abstract] [CrossRef] [Google Scholar]
38. Kailas L, Terry C, Abbott N, Taylor R, Mullin N, Tzokov SB, Todd SJ, Wallace BA, Hobbs JK, Moir A, Bullough PA. 2011. Surface architecture of endospores of the Bacillus cereus/anthracis/thuringiensis family at the subnanometer scale. Proc Natl Acad Sci U S A 108:16014–16019. 10.1073/pnas.1109419108. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
39. Rodenburg CM, McPherson SA, Turnbough CL Jr, Dokland T. 2014. Cryo-EM analysis of the organization of BclA and BxpB in the Bacillus anthracis exosporium. J Struct Biol 186:181–187. 10.1016/j.jsb.2014.02.018. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
40. Scherrer R, Cabrera Beaman T, Gerhardt P. 1971. Macromolecular sieving by the dormant spore of Bacillus cereus. J Bacteriol 108:868–873. [Europe PMC free article] [Abstract] [Google Scholar]
41. Bailey-Smith K, Todd SJ, Southworth TW, Proctor J, Moir A. 2005. The ExsA protein of Bacillus cereus is required for assembly of coat and exosporium onto the spore surface. J Bacteriol 187:3800–3806. 10.1128/JB.187.11.3800-3806.2005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
42. Boydston JA, Yue L, Kearney JF, Turnbough CL Jr. 2006. The ExsY protein is required for complete formation of the exosporium of Bacillus anthracis. J Bacteriol 188:7440–7448. 10.1128/JB.00639-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
43. Giorno R, Bozue J, Cote C, Wenzel T, Moody KS, Mallozzi M, Ryan M, Wang R, Zielke R, Maddock JR, Friedlander A, Welkos S, Driks A. 2007. Morphogenesis of the Bacillus anthracis spore. J Bacteriol 189:691–705. 10.1128/JB.00921-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
44. Johnson MJ, Todd SJ, Ball DA, Shepherd AM, Sylvestre P, Moir A. 2006. ExsY and CotY are required for the correct assembly of the exosporium and spore coat of Bacillus cereus. J Bacteriol 188:7905–7913. 10.1128/JB.00997-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
45. McPherson SA, Li M, Kearney JF, Turnbough CL Jr. 2010. ExsB, an unusually highly phosphorylated protein required for the stable attachment of the exosporium of Bacillus anthracis. Mol Microbiol 76:1527–1538. 10.1111/j.1365-2958.2010.07182.x. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
46. Sousa JC, Silva MT, Balassa G. 1976. An exosporium-like outer layer in Bacillus subtilis spores. Nature 263:53–54. 10.1038/263053a0. [Abstract] [CrossRef] [Google Scholar]
47. Sousa JC, Silva MT, Balassa G. 1978. Ultrastructure and development of an exosporium-like outer spore envelope in Bacillus subtilis. Ann Microbiol (Paris) 129B:339–362. [Abstract] [Google Scholar]
48. Waller LN, Fox N, Fox KF, Fox A, Price RL. 2004. Ruthenium red staining for ultrastructural visualization of a glycoprotein layer surrounding the spore of Bacillus anthracis and Bacillus subtilis. J Microbiol Methods 58:23–30. 10.1016/j.mimet.2004.02.012. [Abstract] [CrossRef] [Google Scholar]
49. Wunschel D, Fox K, Black G, Fox A. 1994. Discrimination among the Bacillus cereus group, in comparison to B. subtilis, by structural carbohydrate profiles and ribosomal RNA spacer region PCR. Syst Appl Microbiol 17:625–635. [Google Scholar]
50. Fox A, Black GE, Fox K, Rostovtseva S. 1993. Determination of carbohydrate profiles of Bacillus anthracis and Bacillus cereus including identification of O-methyl methylpentoses by using gas chromatography-mass spectrometry. J Clin Microbiol 31:887–894. [Europe PMC free article] [Abstract] [Google Scholar]
51. Abe K, Kawano Y, Iwamoto K, Arai K, Maruyama Y, Eichenberger P, Sato T. 2014. Developmentally-regulated excision of the SPβ prophage reconstitutes a gene required for spore envelope maturation in Bacillus subtilis. PLoS Genet 10:e1004636. 10.1371/journal.pgen.1004636. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
52. Roels S, Losick R. 1995. Adjacent and divergently oriented operons under the control of the sporulation regulatory protein GerE in Bacillus subtilis. J Bacteriol 177:6263–6275. [Europe PMC free article] [Abstract] [Google Scholar]
53. McKenney PT, Driks A, Eskandarian HA, Grabowski P, Guberman J, Wang KH, Gitai Z, Eichenberger P. 2010. A distance-weighted interaction map reveals a previously uncharacterized layer of the Bacillus subtilis spore coat. Curr Biol 20:934–938. 10.1016/j.cub.2010.03.060. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
54. Imamura D, Kuwana R, Takamatsu H, Watabe K. 2011. Proteins involved in formation of the outermost layer of Bacillus subtilis spores. J Bacteriol 193:4075–4080. 10.1128/JB.05310-11. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
55. Jiang S, Wan Q, Krajcikova D, Tang J, Tzokov SB, Barak I, Bullough PA. 2015. Diverse supramolecular structures formed by self-assembling proteins of the Bacillus subtilis spore coat. Mol Microbiol 97:347–359. 10.1111/mmi.13030. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
56. Steichen CT, Kearney JF, Turnbough CL Jr. 2007. Non-uniform assembly of the Bacillus anthracis exosporium and a bottle cap model for spore germination and outgrowth. Mol Microbiol 64:359–367. 10.1111/j.1365-2958.2007.05658.x. [Abstract] [CrossRef] [Google Scholar]
57. Thompson BM, Stewart GC. 2008. Targeting of the BclA and BclB proteins to the Bacillus anthracis spore surface. Mol Microbiol 70:421–434. 10.1111/j.1365-2958.2008.06420.x. [Abstract] [CrossRef] [Google Scholar]
58. Thompson BM, Hoelscher BC, Driks A, Stewart GC. 2012. Assembly of the BclB glycoprotein into the exosporium and evidence for its role in the formation of the exosporium ‘cap’ structure in Bacillus anthracis. Mol Microbiol 86:1073–1084. 10.1111/mmi.12042. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
59. Thompson BM, Waller LN, Fox KF, Fox A, Stewart GC. 2007. The BclB glycoprotein of Bacillus anthracis is involved in exosporium integrity. J Bacteriol 189:6704–6713. 10.1128/JB.00762-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
60. Piggot PJ, Coote JG. 1976. Genetic aspects of bacterial endospore formation. Bacteriol Rev 40:908–962. [Europe PMC free article] [Abstract] [Google Scholar]
61. Ohye DF, Murrell WG. 1973. Exosporium and spore coat formation in Bacillus cereus T. J Bacteriol 115:1179–1190. [Europe PMC free article] [Abstract] [Google Scholar]
62. Gerhardt P, Ribi E. 1964. Ultrastructure of the exosporium enveloping spores of Bacillus cereus. J Bacteriol 88:1774–1789. [Europe PMC free article] [Abstract] [Google Scholar]
63. Sylvestre P, Couture-Tosi E, Mock M. 2005. Contribution of ExsFA and ExsFB proteins to the localization of BclA on the spore surface and to the stability of the Bacillus anthracis exosporium. J Bacteriol 187:5122–5128. 10.1128/JB.187.15.5122-5128.2005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
64. Steichen CT, Kearney JF, Turnbough CL Jr. 2005. Characterization of the exosporium basal layer protein BxpB of Bacillus anthracis. J Bacteriol 187:5868–5876. 10.1128/JB.187.17.5868-5876.2005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
65. Tan L, Turnbough CL Jr. 2010. Sequence motifs and proteolytic cleavage of the collagen-like glycoprotein BclA required for its attachment to the exosporium of Bacillus anthracis. J Bacteriol 192:1259–1268. 10.1128/JB.01003-09. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
66. Thompson BM, Hsieh HY, Spreng KA, Stewart GC. 2011. The co-dependence of BxpB/ExsFA and BclA for proper incorporation into the exosporium of Bacillus anthracis. Mol Microbiol 79:799–813. 10.1111/j.1365-2958.2010.07488.x. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
67. Thompson BM, Binkley JM, Stewart GC. 2011. Current physical and SDS extraction methods do not efficiently remove exosporium proteins from Bacillus anthracis spores. J Microbiol Methods 85:143–148. 10.1016/j.mimet.2011.02.009. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
68. Cybulski RJ Jr, Sanz P, McDaniel D, Darnell S, Bull RL, O'Brien AD. 2008. Recombinant Bacillus anthracis spore proteins enhance protection of mice primed with suboptimal amounts of protective antigen. Vaccine 26:4927–4939. 10.1016/j.vaccine.2008.07.015. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
69. Chesnokova ON, McPherson SA, Steichen CT, Turnbough CL Jr. 2009. The spore-specific alanine racemase of Bacillus anthracis and its role in suppressing germination during spore development. J Bacteriol 191:1303–1310. 10.1128/JB.01098-08. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
70. McKevitt MT, Bryant KM, Shakir SM, Larabee JL, Blanke SR, Lovchik J, Lyons CR, Ballard JD. 2007. Effects of endogenous d-alanine synthesis and autoinhibition of Bacillus anthracis germination on in vitro and in vivo infections. Infect Immun 75:5726–5734. 10.1128/IAI.00727-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
71. Raines KW, Kang TJ, Hibbs S, Cao GL, Weaver J, Tsai P, Baillie L, Cross AS, Rosen GM. 2006. Importance of nitric oxide synthase in the control of infection by Bacillus anthracis. Infect Immun 74:2268–2276. 10.1128/IAI.74.4.2268-2276.2006. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
72. Weaver J, Kang TJ, Raines KW, Cao GL, Hibbs S, Tsai P, Baillie L, Rosen GM, Cross AS. 2007. Protective role of Bacillus anthracis exosporium in macrophage-mediated killing by nitric oxide. Infect Immun 75:3894–3901. 10.1128/IAI.00283-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
73. Bergman NH, Anderson EC, Swenson EE, Niemeyer MM, Miyoshi AD, Hanna PC. 2006. Transcriptional profiling of the Bacillus anthracis life cycle in vitro and an implied model for regulation of spore formation. J Bacteriol 188:6092–6100. 10.1128/JB.00723-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
74. Liu H, Bergman NH, Thomason B, Shallom S, Hazen A, Crossno J, Rasko DA, Ravel J, Read TD, Peterson SN, Yates J III, Hanna PC. 2004. Formation and composition of the Bacillus anthracis endospore. J Bacteriol 186:164–178. 10.1128/JB.186.1.164-178.2004. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
75. Boydston JA, Chen P, Steichen CT, Turnbough CL Jr. 2005. Orientation within the exosporium and structural stability of the collagen-like glycoprotein BclA of Bacillus anthracis. J Bacteriol 187:5310–5317. 10.1128/JB.187.15.5310-5317.2005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
76. Daubenspeck JM, Zeng H, Chen P, Dong S, Steichen CT, Krishna NR, Pritchard DG, Turnbough CL Jr. 2004. Novel oligosaccharide side chains of the collagen-like region of BclA, the major glycoprotein of the Bacillus anthracis exosporium. J Biol Chem 279:30945–30953. 10.1074/jbc.M401613200. [Abstract] [CrossRef] [Google Scholar]
77. Réty S, Salamitou S, Garcia-Verdugo I, Hulmes DJ, Le Hégarat F, Chaby R, Lewit-Bentley A. 2005. The crystal structure of the Bacillus anthracis spore surface protein BclA shows remarkable similarity to mammalian proteins. J Biol Chem 280:43073–43078. 10.1074/jbc.M510087200. [Abstract] [CrossRef] [Google Scholar]
78. Liu C-Q, Nuttall SD, Tran H, Wilkins M, Streltsov VA, Alderton MR. 2008. Construction, crystal structure and application of a recombinant protein that lacks the collagen-like region of BclA from Bacillus anthracis spores. Biotechnol Bioeng 99:774–782. 10.1002/bit.21637. [Abstract] [CrossRef] [Google Scholar]
79. Tamborrini M, Oberli MA, Werz DB, Schürch N, Frey J, Seeberger PH, Pluschke G. 2009. Immuno-detection of anthrose containing tetrasaccharide in the exosporium of Bacillus anthracis and Bacillus cereus strains. J Appl Microbiol 106:1618–1628. 10.1111/j.1365-2672.2008.04129.x. [Abstract] [CrossRef] [Google Scholar]
80. Lequette Y, Garénaux E, Combrouse T, Del Lima Dias T, Ronse A, Slomianny C, Trivelli X, Guerardel Y, Faille C. 2011. Domains of BclA, the major surface glycoprotein of the B. cereus exosporium: glycosylation patterns and role in spore surface properties. Biofouling 27:751–761. 10.1080/08927014.2011.599842. [Abstract] [CrossRef] [Google Scholar]
81. Brahmbhatt TN, Darnell SC, Carvalho HM, Sanz P, Kang TJ, Bull RL, Rasmussen SB, Cross AS, O'Brien AD. 2007. Recombinant exosporium protein BclA of Bacillus anthracis is effective as a booster for mice primed with suboptimal amounts of protective antigen. Infect Immun 75:5240–5247. 10.1128/IAI.00884-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
82. Hahn UK, Boehm R, Beyer W. 2006. DNA vaccination against anthrax in mice—combination of anti-spore and anti-toxin components. Vaccine 24:4569–4571. 10.1016/j.vaccine.2005.08.031. [Abstract] [CrossRef] [Google Scholar]
83. Cote CK, Kaatz L, Reinhardt J, Bozue J, Tobery SA, Bassett AD, Sanz P, Darnell SC, Alem F, O'Brien AD, Welkos SL. 2012. Characterization of a multi-component anthrax vaccine designed to target the initial stages of infection as well as toxaemia. J Med Microbiol 61:1380–1392. 10.1099/jmm.0.045393-0. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
84. Brahmbhatt TN, Janes BK, Stibitz ES, Darnell SC, Sanz P, Rasmussen SB, O'Brien AD. 2007. Bacillus anthracis exosporium protein BclA affects spore germination, interaction with extracellular matrix proteins, and hydrophobicity. Infect Immun 75:5233–5239. 10.1128/IAI.00660-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
85. Bozue JA, Parthasarathy N, Phillips LR, Cote CK, Fellows PF, Mendelson I, Shafferman A, Friedlander AM. 2005. Construction of a rhamnose mutation in Bacillus anthracis affects adherence to macrophages but not virulence in guinea pigs. Microb Pathog 38:1–12. 10.1016/j.micpath.2004.10.001. [Abstract] [CrossRef] [Google Scholar]
86. Dong S, Chesnokova ON, Turnbough CL Jr, Pritchard DG. 2009. Identification of the UDP-N-acetylglucosamine 4-epimerase involved in exosporium protein glycosylation in Bacillus anthracis. J Bacteriol 191:7094–7101. 10.1128/JB.01050-09. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
87. Dong S, McPherson SA, Tan L, Chesnokova ON, Turnbough CL Jr, Pritchard DG. 2008. Anthrose biosynthetic operon of Bacillus anthracis. J Bacteriol 190:2350–2359. 10.1128/JB.01899-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
88. Dong S, McPherson SA, Wang Y, Li M, Wang P, Turnbough CL Jr, Pritchard DG. 2010. Characterization of the enzymes encoded by the anthrose biosynthetic operon of Bacillus anthracis. J Bacteriol 192:5053–5062. 10.1128/JB.00568-10. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
89. Waller LN, Stump MJ, Fox KF, Harley WH, Fox A, Stewart GC, Shahgholi M. 2005. Identification of a second collagen-like glycoprotein produced by Bacillus anthracis and demonstration of associated spore-specific sugars. J Bacteriol 187:4592–4597. 10.1128/JB.187.13.4592-4597.2005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
90. Leski TA, Caswell CC, Pawlowski M, Klinke DJ, Bujnicki JM, Hart SJ, Lukomski S. 2009. Identification and classification of bcl genes and proteins of Bacillus cereus group organisms and their application in Bacillus anthracis detection and fingerprinting. Appl Environ Microbiol 75:7163–7172. 10.1128/AEM.01069-09. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
91. Thompson BM, Hoelscher BC, Driks A, Stewart GC. 2011. Localization and assembly of the novel exosporium protein BetA of Bacillus anthracis. J Bacteriol 193:5098–5104. 10.1128/JB.05658-11. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
92. Moody KL, Driks A, Rother GL, Cote CK, Brueggemann EE, Hines HB, Friedlander AM, Bozue J. 2010. Processing, assembly and localization of a Bacillus anthracis spore protein. Microbiology 156:174–183. 10.1099/mic.0.033407-0. [Abstract] [CrossRef] [Google Scholar]
93. Fazzini MM, Schuch R, Fischetti VA. 2010. A novel spore protein, ExsM, regulates formation of the exosporium in Bacillus cereus and Bacillus anthracis and affects spore size and shape. J Bacteriol 192:4012–4021. 10.1128/JB.00197-10. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
94. Todd SJ, Moir AJ, Johnson MJ, Moir A. 2003. Genes of Bacillus cereus and Bacillus anthracis encoding proteins of the exosporium. J Bacteriol 185:3373–3378. 10.1128/JB.185.11.3373-3378.2003. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
95. Little S, Driks A. 2001. Functional analysis of the Bacillus subtilis morphogenetic spore coat protein CotE. Mol Microbiol 42:1107–1120. 10.1046/j.1365-2958.2001.02708.x. [Abstract] [CrossRef] [Google Scholar]
96. Aronson A, Goodman B, Smith Z. 2014. The regulated synthesis of a Bacillus anthracis spore coat protein that affects spore surface properties. J Appl Microbiol 116:1241–1249. 10.1111/jam.12452. [Abstract] [CrossRef] [Google Scholar]
97. Qu N, Peng Q, Qiu L, Liu C, Chen Z, Zhang J, Song F, Li J. 2013. Transcriptional regulation of exosporium basal layer structural gene exsB in Bacillus thuringiensis by σK and GerE. Wei Sheng Wu Xue Bao 53:241–248. [Abstract] [Google Scholar]
98. García-Patrone M, Tandecarz JS. 1995. A glycoprotein multimer from Bacillus thuringiensis sporangia: dissociation into subunits and sugar composition. Mol Cell Biochem 145:29–37. 10.1007/BF00925710. [Abstract] [CrossRef] [Google Scholar]
99. Lequette Y, Garénaux E, Tauveron G, Dumez S, Perchat S, Slomianny C, Lereclus D, Guérardel Y, Faille C. 2011. Role played by exosporium glycoproteins in the surface properties of Bacillus cereus spores and in their adhesion to stainless steel. Appl Environ Microbiol 77:4905–4911. 10.1128/AEM.02872-10. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
100. Krogh A, Larsson B, von Heijne G, Sonnhammer EL. 2001. Predicting transmembrane protein topology with a hidden Markov model: application to complete genomes. J Mol Biol 305:567–580. 10.1006/jmbi.2000.4315. [Abstract] [CrossRef] [Google Scholar]
101. Severson KM, Mallozzi M, Bozue J, Welkos SL, Cote CK, Knight KL, Driks A. 2009. Roles of the Bacillus anthracis spore protein ExsK in exosporium maturation and germination. J Bacteriol 191:7587–7596. 10.1128/JB.01110-09. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
102. Gai Y, Liu G, Tan H. 2006. Identification and characterization of a germination operon from Bacillus thuringiensis. Antonie Van Leeuwenhoek 89:251–259. 10.1007/s10482-005-9026-x. [Abstract] [CrossRef] [Google Scholar]
103. Terry C, Shepherd A, Radford DS, Moir A, Bullough PA. 2011. YwdL in Bacillus cereus: its role in germination and exosporium structure. PLoS One 6:e23801. 10.1371/journal.pone.0023801. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
104. Ragkousi K, Eichenberger P, van Ooij C, Setlow P. 2003. Identification of a new gene essential for germination of Bacillus subtilis spores with Ca2+-dipicolinate. J Bacteriol 185:2315–2329. 10.1128/JB.185.7.2315-2329.2003. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
105. Reference deleted. [Google Scholar]
106. Lai EM, Phadke ND, Kachman MT, Giorno RSV, Vazquez JA, Maddock JR, Driks A. 2003. Proteomic analysis of the spore coats of Bacillus subtilis and Bacillus anthracis. J Bacteriol 185:1443–1454. 10.1128/JB.185.4.1443-1454.2003. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
107. Cybulski RJ Jr, Sanz P, Alem F, Stibitz S, Bull RL, O'Brien AD. 2009. Four superoxide dismutases contribute to Bacillus anthracis virulence and provide spores with redundant protection from oxidative stress. Infect Immun 77:274–285. 10.1128/IAI.00515-08. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
108. Passalacqua KD, Bergman NH, Herring-Palmer A, Hanna P. 2006. The superoxide dismutases of Bacillus anthracis do not cooperatively protect against endogenous superoxide stress. J Bacteriol 188:3837–3848. 10.1128/JB.00239-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
109. Spreng KA, Thompson BM, Stewart GC. 2013. A genetic approach for the identification of exosporium assembly determinants of Bacillus anthracis. J Microbiol Methods 93:58–67. 10.1016/j.mimet.2013.01.019. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
110. Saile E, Koehler TM. 2006. Bacillus anthracis multiplication, persistence, and genetic exchange in the rhizosphere of grass plants. Appl Environ Microbiol 72:3168–3174. 10.1128/AEM.72.5.3168-3174.2006. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
111. Hugh-Jones M, Blackburn J. 2009. The ecology of Bacillus anthracis. Mol Aspects Med 30:356–367. 10.1016/j.mam.2009.08.003. [Abstract] [CrossRef] [Google Scholar]
112. Williams G, Linley E, Nicholas R, Baillie L. 2013. The role of the exosporium in the environmental distribution of anthrax. J Appl Microbiol 114:396–403. 10.1111/jam.12034. [Abstract] [CrossRef] [Google Scholar]
113. Chen G, Driks A, Tawfiq K, Mallozzi M, Patil S. 2010. Bacillus anthracis and Bacillus subtilis spore surface properties and transport. Colloids Surf B Biointerfaces 76:512–518. 10.1016/j.colsurfb.2009.12.012. [Abstract] [CrossRef] [Google Scholar]
114. Andersson A, Rönner U. 1998. Adhesion and removal of dormant, heat-inactivated, and germinated spores of three strains of Bacillus cereus. Biofouling 13:51–67. 10.1080/08927019809378370. [CrossRef] [Google Scholar]
115. Faille C, Tauveron G, Le Gentil-Lelièvre Slomianny CC. 2007. Occurrence of Bacillus cereus spores with a damaged exosporium: consequences on the spore adhesion on surfaces of food processing lines. J Food Prot 70:2346–2353. [Abstract] [Google Scholar]
116. Faille C, Lequette Y, Ronse A, Slomianny C, Garénaux E, Guerandel Y. 2010. Morphology and physico-chemical properties of Bacillus spores surrounded or not with an exosporium: consequences on their ability to adhere to stainless steel. Int J Food Microbiol 143:125–135. 10.1016/j.ijfoodmicro.2010.07.038. [Abstract] [CrossRef] [Google Scholar]
117. Tauveron G, Slomianny C, Henry C, Faille C. 2006. Variability among Bacillus cereus strains in spore surface properties and influence on their ability to contaminate food surface equipment. Int J Food Microbiol 110:254–262. 10.1016/j.ijfoodmicro.2006.04.027. [Abstract] [CrossRef] [Google Scholar]
118. Mercier-Bonin M, Dehouche A, Morchain J, Schmitz P. 2011. Orientation and detachment dynamics of Bacillus spores from stainless steel under controlled shear flow: modelling of the adhesion force. Int J Food Microbiol 146:182–191. 10.1016/j.ijfoodmicro.2011.02.025. [Abstract] [CrossRef] [Google Scholar]
119. Glomski IJ, Piris-Gimenez A, Huerre M, Mock M, Goossens PL. 2007. Primary involvement of pharynx and Peyer's patch in inhalational and intestinal anthrax. PLoS Pathog 3:e76. 10.1371/journal.ppat.0030076. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
120. Fiole D, Deman P, Trescos Y, Mayol JF, Mathieu J, Vial JC, Douady J, Tournier JN. 2014. Two-photon intravital imaging of lungs during anthrax infection reveals long-lasting macrophage-dendritic cell contacts. Infect Immun 82:864–872. 10.1128/IAI.01184-13. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
121. Fiole D, Douady J, Vial J-C, Quesnel-Hellmann A, Tournier J-N. 2012. Dynamics of rapid spore capture by dendritic cells in the lung alveolus. Am J Respir Crit Care Med 186:e2–e3. 10.1164/rccm.201110-1782IM. [Abstract] [CrossRef] [Google Scholar]
122. Guidi-Rontani C, Weber-Levy M, Labruyère E, Mock M. 1999. Germination of Bacillus anthracis spores within alveolar macrophages. Mol Microbiol 31:9–17. 10.1046/j.1365-2958.1999.01137.x. [Abstract] [CrossRef] [Google Scholar]
123. Cleret A, Quesnel-Hellmann A, Vallon-Eberhard A, Verrier B, Jung S, Vidal D, Mathieu J, Tournier JN. 2007. Lung dendritic cells rapidly mediate anthrax spore entry through the pulmonary route. J Immunol 178:7994–8001. 10.4049/jimmunol.178.12.7994. [Abstract] [CrossRef] [Google Scholar]
124. Russell BH, Liu Q, Jenkins SA, Tuvim MJ, Dickey BF, Xu Y. 2008. In vivo demonstration and quantification of intracellular Bacillus anthracis in lung epithelial cells. Infect Immun 76:3975–3983. 10.1128/IAI.00282-08. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
125. Russell BH, Vasan R, Keene DR, Koehler TM, Xu Y. 2008. Potential dissemination of Bacillus anthracis utilizing human lung epithelial cells. Cell Microbiol 10:945–957. 10.1111/j.1462-5822.2007.01098.x. [Abstract] [CrossRef] [Google Scholar]
126. Oliva CR, Swiecki MK, Griguer CE, Lisanby MW, Bullard DC, Turnbough CL Jr, Kearney JF. 2008. The integrin Mac-1 (CR3) mediates internalization and directs Bacillus anthracis spores into professional phagocytes. Proc Natl Acad Sci U S A 105:1261–1266. 10.1073/pnas.0709321105. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
127. Bozue J, Moody KL, Cote CK, Stiles BG, Friedlander AM, Welkos SL, Hale ML. 2007. Bacillus anthracis spores of the bclA mutant exhibit increased adherence to epithelial cells, fibroblasts, and endothelial cells but not to macrophages. Infect Immun 75:4498–4505. 10.1128/IAI.00434-07. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
128. Oliva C, Turnbough CL Jr, Kearney JF. 2009. CD14-Mac-1 interactions in Bacillus anthracis spore internalization by macrophages. Proc Natl Acad Sci U S A 106:13957–13962. 10.1073/pnas.0902392106. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
129. Gu C, Jenkins SA, Xue Q, Xu Y. 2012. Activation of the classical complement pathway by Bacillus anthracis is the primary mechanism for spore phagocytosis and involves the spore surface protein BclA. J Immunol 188:4421–4431. 10.4049/jimmunol.1102092. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
130. Xue Q, Gu C, Rivera J, Höök M, Chen X, Pozzi A, Xu Y. 2011. Entry of Bacillus anthracis spores into epithelial cells is mediated by the spore surface protein BclA, integrin α2β1 and complement component C1q. Cell Microbiol 13:620–634. 10.1111/j.1462-5822.2010.01558.x. [Abstract] [CrossRef] [Google Scholar]
131. Ghebrehiwet B, Tantral L, Titmus MA, Panessa-Warren BJ, Tortora GT, Wong SS, Warren JB. 2007. The exosporium of B. cereus contains a binding site for gC1qR/p33: implication in spore attachment and/or entry. Adv Exp Med Biol 598:181–197. 10.1007/978-0-387-71767-8_13. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
132. Bozue J, Cote CK, Moody KL, Welkos SL. 2007. Fully virulent Bacillus anthracis does not require the immunodominant protein BclA for pathogenesis. Infect Immun 75:508–511. 10.1128/IAI.01202-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
133. Ross JM. 1957. The pathogenesis of anthrax following the administration of spores by the respiratory route. J Pathol Bacteriol 73:485–494. 10.1002/path.1700730219. [CrossRef] [Google Scholar]
134. Henderson DW, Peacock S, Belton FC. 1956. Observations on the prophylaxis of experimental pulmonary anthrax in the monkey. J Hyg 54:28–36. 10.1017/S0022172400044272. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
135. Friedlander AM, Welkos SL, Pitt MLM, Ezzell JW, Worsham PL, Rose KJ, Ivins BE, Lowe JR, Howe GB, Mikesell P, Lawrence WB. 1993. Postexposure prophylaxis against experimental inhalation anthrax. J Infect Dis 167:1239–1243. 10.1093/infdis/167.5.1239. [Abstract] [CrossRef] [Google Scholar]
136. Heine HS, Bassett J, Miller L, Hartings JM, Ivins BE, Pitt ML, Fritz D, Norris SL, Byrne WR. 2007. Determination of antibiotic efficacy against Bacillus anthracis in a mouse aerosol challenge model. Antimicrob Agents Chemother 51:1373–1379. 10.1128/AAC.01050-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
137. Loving CL, Kennett M, Lee GM, Grippe VK, Merkel TJ. 2007. Murine aerosol challenge model of anthrax. Infect Immun 75:2689–2698. 10.1128/IAI.01875-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
138. Jenkins SA, Xu Y. 2013. Characterization of Bacillus anthracis persistence in vivo. PLoS One 8:e66177. 10.1371/journal.pone.0066177. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
139. Xue Q, Jenkins SA, Gu C, Smeds E, Liu Q, Vasan R, Russell BH, Xu Y. 2010. Bacillus anthracis spore entry into epithelial cells is an actin-dependent process requiring c-Src and PI3K. PLoS One 5:e11665. 10.1371/journal.pone.0011665. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
140. Chung MC, Tonry JH, Narayanan A, Manes NP, Mackie RS, Gutting B, Mukherjee DV, Popova TG, Kashanchi F, Bailey CL, Popov SG. 2011. Bacillus anthracis interacts with plasmin(ogen) to evade C3b-dependent innate immunity. PLoS One 6:e18119. 10.1371/journal.pone.0018119. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
141. Basu S, Kang TJ, Chen WH, Fenton MJ, Baillie L, Hibbs S, Cross AS. 2007. Role of Bacillus anthracis spore structures in macrophage cytokine responses. Infect Immun 75:2351–2358. 10.1128/IAI.01982-06. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
142. Venir E, Del Torre M, Cunsolo V, Saletti R, Musetti R, Stecchini ML. 2014. Involvement of alanine racemase in germination of Bacillus cereus spores lacking an intact exosporium. Arch Microbiol 196:79–85. 10.1007/s00203-013-0946-y. [Abstract] [CrossRef] [Google Scholar]
143. Shatalin K, Gusarov I, Avetissova E, Shatalina Y, McQuade LE, Lippard SJ, Nudler E. 2008. Bacillus anthracis-derived nitric oxide is essential for pathogen virulence and survival in macrophages. Proc Natl Acad Sci U S A 105:1009–1013. 10.1073/pnas.0710950105. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
144. Auerbach S, Wright GG. 1955. Studies on immunity in anthrax. VI. Immunizing activity of protective antigen against various strains of Bacillus anthracis. J Immunol 75:129–133. [Abstract] [Google Scholar]
145. Ward MK, McGann VG, Hogge AL Jr, Huff ML, Kanode RG Jr, Roberts EO. 1965. Studies on anthrax infections in immunized guinea pigs. J Infect Dis 115:59–67. 10.1093/infdis/115.1.59. [Abstract] [CrossRef] [Google Scholar]
146. Little SF, Knudson GB. 1986. Comparative efficacy of Bacillus anthracis live spore vaccine and protective antigen vaccine against anthrax in the guinea pig. Infect Immun 52:509–512. [Europe PMC free article] [Abstract] [Google Scholar]
147. Brossier F, Levy M, Mock M. 2002. Anthrax spores make an essential contribution to vaccine efficacy. Infect Immun 70:661–664. [Europe PMC free article] [Abstract] [Google Scholar]
148. Enkhtuya J, Kawamoto K, Kobayashi Y, Uchida I, Rana N, Makino S. 2006. Significant passive protective effect against anthrax by antibody to Bacillus anthracis inactivated spores that lack two virulence plasmids. Microbiology 152:3103–3110. 10.1099/mic.0.28788-0. [Abstract] [CrossRef] [Google Scholar]
149. Gauthier YP, Tournier JM, Paucod JC, Corre JP, Mock M, Goossens PL, Vidal DR. 2009. Efficacy of a vaccine based on protective antigen and killed spores against experimental inhalational anthrax. Infect Immun 77:1197–1207. 10.1128/IAI.01217-08. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
150. Vergis JM, Cote CK, Bozou J, Alem F, Ventura CL, Welkos SL, O'Brien AD. 2013. Immunization of mice with formalin-inactivated spores from avirulent Bacillus cereus strains provides significant protection from challenge with Bacillus anthracis Ames. Clin Vaccine Immunol 20:56–65. 10.1128/CVI.00550-12. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
151. Welkos SL, Friedlander AM. 1988. Comparative safety and efficacy against Bacillus anthracis of protective antigen and live vaccines in mice. Microb Pathog 5:127–139. 10.1016/0882-4010(88)90015-0. [Abstract] [CrossRef] [Google Scholar]
152. Cohen S, Mendelson I, Altboum Z, Kobiler D, Elhanany E, Bino T, Leitner M, Inbar I, Rosenberg H, Gozes Y, Barak R, Fisher M, Kronman C, Velan B, Shafferman A. 2000. Attenuated nontoxinogenic and nonencapsulated recombinant Bacillus anthracis spore vaccines protect against anthrax. Infect Immun 68:4549–4558. 10.1128/IAI.68.8.4549-4558.2000. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
153. Aloni-Grinstein R, Gat O, Altboum Z, Velan B, Cohen S, Shafferman A. 2005. Oral spore vaccine based on live attenuated nontoxinogenic Bacillus anthracis expressing recombinant mutant protective antigen. Infect Immun 73:4043–4053. 10.1128/IAI.73.7.4043-4053.2005. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
154. Mendelson I, Gat O, Aloni-Grinstein R, Altboum Z, Inbar I, Kronman C, Bar-Haim E, Cohen S, Velan B, Shafferman A. 2005. Efficacious, nontoxigenic Bacillus anthracis spore vaccines based on strains expressing mutant variants of lethal toxin components. Vaccine 23:5688–5697. 10.1016/j.vaccine.2004.11.077. [Abstract] [CrossRef] [Google Scholar]
155. Köhler SM, Baillie LW, Beyer W. 2015. BclA and toxin antigens augment each other to protect NMRI mice from lethal Bacillus anthracis challenge. Vaccine 33:2771–2777. 10.1016/j.vaccine.2015.04.049. [Abstract] [CrossRef] [Google Scholar]
156. Welkos SL, Cote CK, Rea KM, Gibbs PH. 2004. A microtiter fluorometric assay to detect the germination of Bacillus anthracis spores and germination inhibitory effects of antibodies. J Microbiol Methods 56:253–265. 10.1016/j.mimet.2003.10.019. [Abstract] [CrossRef] [Google Scholar]
157. Cote CK, Bozue J, Moody KL, DiMezzo TL, Chapman CE, Welkos SL. 2008. Analysis of a novel spore antigen in Bacillus anthracis that contributes to spore opsonization. Microbiology 154:619–632. 10.1099/mic.0.2007/008292-0. [Abstract] [CrossRef] [Google Scholar]
158. Goossens PL, Sylvestre P, Mock M. 2007. Of spore opsonization and passive protection against anthrax. Microbiology 153:301–302. 10.1099/mic.0.2006/003210-0. [Abstract] [CrossRef] [Google Scholar]
159. Glomski IJ, Corre JP, Mock M, Goossens PL. 2007. Cutting edge: IFN-gamma-producing CD4 T lymphocytes mediate spore-induced immunity to capsulated Bacillus anthracis. J Immunol 178:2646–2650. 10.4049/jimmunol.178.5.2646. [Abstract] [CrossRef] [Google Scholar]
160. Moberly BJ, Shafa F, Gerhardt P. 1966. Structural details of anthrax spores during stages of transformation into vegetative cells. J Bacteriol 92:220–228. [Europe PMC free article] [Abstract] [Google Scholar]
161. Hachisuka Y, Kozuka S, Tsujikawa M. 1984. Exosporia and appendages of spores of Bacillus species. Microbiol Immunol 28:619–624. 10.1111/j.1348-0421.1984.tb00714.x. [Abstract] [CrossRef] [Google Scholar]
162. Wan Q. 2013. Structure and assembly of Bacillus spore proteins. Ph.D. dissertation. Department of Molecular Biology and Biotechnology, The University of Sheffield, United Kingdom. [Google Scholar]
163. Tomcsik J, Baumann-Grace J. 1959. Specific exosporium reaction of Bacillus megaterium. J Gen Microbiol 21:666–675. 10.1099/00221287-21-3-666. [CrossRef] [Google Scholar]
164. Kim HU, Goepfert JM. 1972. Efficacy of a fluorescent-antibody procedure for identifying Bacillus cereus in foods. Appl Microbiol 24:708–713. [Europe PMC free article] [Abstract] [Google Scholar]
165. Bowen WR, Fenton AS, Lovitt RW, Wright CJ. 2002. The measurement of Bacillus mycoides spore adhesion using atomic force microscopy, simple counting methods, and a spinning disk technique. Biotechnol Bioeng 79:170–179. 10.1002/bit.10321. [Abstract] [CrossRef] [Google Scholar]
166. Gerhardt P, Pankratz HS, Scherrer R. 1976. Fine structure of the Bacillus thuringiensis spore. Appl Environ Microbiol 32:438–440. [Europe PMC free article] [Abstract] [Google Scholar]
167. Smirnova TA, Minenkova IB, Orlova MC, Lecadet MM, Azizbekyan RR. 1996. The crystal-forming strains of Bacillus laterosporus. Res Microbiol 147:343–350. 10.1016/0923-2508(96)84709-7. [Abstract] [CrossRef] [Google Scholar]
168. Hodgkiss W, Ordal ZJ, Cann DC. 1967. The morphology and ultrastructure of the spore and exosporium of some Clostridium species. J Gen Microbiol 47:213–225. 10.1099/00221287-47-2-213. [Abstract] [CrossRef] [Google Scholar]
169. Hodgkiss W, Ordal ZJ. 1966. Morphology of the spore of some strains of Clostridium botulinum type E. J Bacteriol 91:2031–2036. [Europe PMC free article] [Abstract] [Google Scholar]
170. Stevenson KE, Vaughn RHV. 1972. Exosporium formation in sporulating cells of Clostridium botulinum 78A. J Bacteriol 112:618–621. [Europe PMC free article] [Abstract] [Google Scholar]
171. Takumi K, Kinouchi T, Kawata T. 1979. Isolation and partial characterization of exosporium from spores of a highly sporogenic mutant of Clostridium botulinum type A. Microbiol Immunol 23:443–454. 10.1111/j.1348-0421.1979.tb00484.x. [Abstract] [CrossRef] [Google Scholar]
172. Masuda K, Kawata T, Takumi K, Kinouchi T. 1980. Ultrastructure of a hexagonal array in exosporium of a highly sporogenic mutant of Clostridium botulinum type A revealed by electron microscopy using optical diffraction and filtration. Microbiol Immunol 24:507–513. 10.1111/j.1348-0421.1980.tb02854.x. [Abstract] [CrossRef] [Google Scholar]
173. Joshi LT, Phillips DS, Williams CF, Alvousef A, Baille L. 2012. Contribution of spores to the ability of Clostridium difficile to adhere to surfaces. Appl Environ Microbiol 78:7671–7679. 10.1128/AEM.01862-12. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
174. Escobar-Cortés K, Barra-Carrasco J, Paredes-Sabja D. 2013. Proteases and sonication specifically remove the exosporium layer of spores of Clostridium difficile strain 630. J Microbiol Methods 93:25–31. 10.1016/j.mimet.2013.01.016. [Abstract] [CrossRef] [Google Scholar]
175. Mackey BM, Morris JG. 1972. The exosporium of Clostridium pasteurianum. J Gen Microbiol 73:325–338. 10.1099/00221287-73-2-325. [Abstract] [CrossRef] [Google Scholar]
176. Shakhbanov AA. 1980. Ultrastructure of Clostridium septicum. Zh Mikrobiol Epidemiol Immunobiol 7:43–47. [Abstract] [Google Scholar]
177. La Duc MT, Satomi M, Venkateswaran K. 2004. Bacillus odysseyi sp. nov., a round-spore-forming bacillus isolated from the Mars Odyssey spacecraft. Int J Syst Evol Microbiol 54:195–201. 10.1099/ijs.0.02747-0. [Abstract] [CrossRef] [Google Scholar]
178. Holt SC, Gauthier JJ, Tipper DJ. 1975. Ultrastructural studies of sporulation in Bacillus sphaericus. J Bacteriol 122:1322–1338. [Europe PMC free article] [Abstract] [Google Scholar]
179. Nicolas L, Regis LN, Rios EM. 1994. Role of the exosporium in the stability of the Bacillus sphaericus binary toxin. FEMS Microbiol Lett 124:271–275. 10.1111/j.1574-6968.1994.tb07296.x. [Abstract] [CrossRef] [Google Scholar]
180. Chen ZX, Dickson DW, Freitas LG, Preston JF. 1997. Ultrastructure, morphology, and sporogenesis of Pasteuria penetrans. Phytopathology 87:273–283. 10.1094/PHYTO.1997.87.3.273. [Abstract] [CrossRef] [Google Scholar]
181. Beaman TC, Pankratz HS, Gerhardt P. 1972. Ultrastructure of the exosporium and underlying inclusions in spores of Bacillus megaterium strains. J Bacteriol 109:1198–1209. [Europe PMC free article] [Abstract] [Google Scholar]
182. Manetsberger J, Hall EA, Christie G. 2014. BMQ_0737 encodes a novel protein crucial to the integrity of the outermost layers of Bacillus megaterium QM B1551 spores. FEMS Microbiol Lett 358:162–169. 10.1111/1574-6968.12520. [Abstract] [CrossRef] [Google Scholar]
183. Manetsberger J, Hall EA, Christie G. 2015. Plasmid-encoded genes influence exosporium assembly and morphology in Bacillus megaterium QM B1551 spores. FEMS Microbiol Lett 362:fnv147. 10.1093/femsle/fnv147. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
184. Yutin N, Galperin MY. 2013. A genomic update on clostridial phylogeny: Gram-negative spore formers and other misplaced clostridia. Environ Microbiol 15:2631–2641. 10.1111/1462-2920.12173. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
185. Permpoonpattana P, Phetcharaburanin J, Mikelsone A, Dembek M, Tan S, Brisson M-C, La Ragione R, Brisson AR, Fairweather N, Hong HA, Cutting SM. 2013. Functional characterization of Clostridium difficile spore coat proteins. J Bacteriol 195:1492–1503. 10.1128/JB.02104-12. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
186. Permpoonpattana P, Tolls EH, Nadem R, Tan S, Brisson A, Cutting SM. 2011. Surface layers of Clostridium difficile endospores. J Bacteriol 193:6461–6470. 10.1128/JB.05182-11. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
187. Abhyankar W, Hossain AH, Djajasaputra A, Permpoonpattana P, Ter Beek A, Dekker HL, Cutting SM, Brul S, de Koning LJ, de Koster CG. 2013. In pursuit of protein targets: proteomic characterization of bacterial spore outer layers. J Proteome Res 12:4507–4521. 10.1021/pr4005629. [Abstract] [CrossRef] [Google Scholar]
188. Phetcharaburanin J, Hong HA, Colenutt C, Bianconi I, Sempere L, Permpoonpattana P, Smith K, Dembek M, Tan S, Brisson MC, Brisson AR, Fairweather NF, Cutting SM. 2014. The spore-associated protein BclA1 affects the susceptibility of animals to colonization and infection by Clostridium difficile. Mol Microbiol 92:1025–1038. 10.1111/mmi.12611. [Abstract] [CrossRef] [Google Scholar]
189. Strong PC, Fulton KM, Aubry A, Foote S, Twine SM, Logan SM. 2014. Identification and characterization of glycoproteins on the spore surface of Clostridium difficile. J Bacteriol 196:2627–2637. 10.1128/JB.01469-14. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
190. Barra-Carrasco J, Olguín-Araneda V, Plaza-Garrido Á, Miranda-Cárdenas C, Cofré-Araneda G, Pizarro-Guajardo M, Sarker MR, Paredes-Sabja D. 2013. The Clostridium difficile exosporium cysteine (CdeC)-rich protein is required for exosporium morphogenesis and coat assembly. J Bacteriol 195:3863–3875. 10.1128/JB.00369-13. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
191. Fimlaid KA, Bond JP, Schutz KC, Putnam EE, Leung JM, Lawley TD, Shen A. 2013. Global analysis of the sporulation pathway of Clostridium difficile. PLoS Genet 9:e1003660. 10.1371/journal.pgen.1003660. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
192. Díaz-González F, Milano M, Olguin-Araneda V, Pizarro-Cerda J, Castro-Córdova P, Tzeng SC, Maier CS, Sarker MR, Paredes-Sabja D. 2015. Protein composition of the outermost exosporium-like layer of Clostridium difficile 630 spores. J Proteomics 123:1–13. 10.1016/j.jprot.2015.03.035. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
193. Paredes-Sabja D, Sarker MR. 2012. Adherence of Clostridium difficile spores to Caco-2 cells in culture. J Med Microbiol 61:1208–1218. 10.1099/jmm.0.043687-0. [Abstract] [CrossRef] [Google Scholar]
194. Paredes-Sabja D, Cofre-Araneda G, Brito-Silva C, Pizarro-Guajardo M, Sarker MR. 2012. Clostridium difficile spore-macrophage interactions: spore survival. PLoS One 7:e43635. 10.1371/journal.pone.0043635. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
195. Panessa-Warren BJ, Tortora GT, Warren JB. 1997. Exosporial membrane plasticity of Clostridium sporogenes and Clostridium difficile. Tissue Cell 29:449–461. 10.1016/S0040-8166(97)80031-6. [Abstract] [CrossRef] [Google Scholar]
196. Brunt J, Cross KL, Peck MW. 2015. Apertures in the Clostridium sporogenes spore coat and exosporium align to facilitate emergence of the vegetative cell. Food Microbiol 51:45–50. 10.1016/j.fm.2015.04.013. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
197. Panessa-Warren BJ, Tortora GT, Warren JB. 1994. Electron-microscopy of C. sporogenes endospore attachment and germination. Scanning 16:227–240. [Google Scholar]
198. Yousten AA, Davidson EW. 1982. Ultrastructural analysis of spores and parasporal crystals formed by Bacillus sphaericus 2297. Appl Environ Microbiol 44:1449–1455. [Europe PMC free article] [Abstract] [Google Scholar]
199. Zhao N, Ge Y, Shi T, Hu X, Yuan Z. 2014. Collagen-like glycoprotein BclS is involved in the formation of filamentous structures of the Lysinibacillus sphaericus exosporium. Appl Environ Microbiol 80:6656–6663. 10.1128/AEM.02238-14. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
200. Davies KG, De Leij FAAM, Kerry BR. 1991. Microbial agents for the biological control plant-parasitic nematodes in tropical agriculture. Trop Pest Manage 37:303–320. 10.1080/09670879109371606. [CrossRef] [Google Scholar]
201. Davies KG. 2009. Understanding the interaction between an obligate hyperparasitic bacterium, Pasteuria penetrans and its obligate plant-parasitic nematode host, Meloidogyne spp. Adv Parasitol 68:211–245. 10.1016/S0065-308X(08)00609-X. [Abstract] [CrossRef] [Google Scholar]
202. Mouton L, Traunecker E, McElroy K, DuPasquier L, Ebert D. 2009. Identification of a polymorphic collagen-like protein in the crustacean bacteria Pasteuria ramose. Res Microbiol 160:792–799. 10.1016/j.resmic.2009.08.016. [Abstract] [CrossRef] [Google Scholar]
203. McElroy K, Mouton L, DuPasquier L, Qi W, Ebert D. 2011. Characterisation of a large family of polymorphic collagen-like proteins in the endospore-forming bacterium Pasteuria ramosa. Res Microbiol 162:701–714. 10.1016/j.resmic.2011.06.009. [Abstract] [CrossRef] [Google Scholar]
204. Reference deleted. [Google Scholar]
205. Wang C, Ehrhardt CJ, Yadavalli VK. 2015. Single cell profiling of surface carbohydrates on Bacillus cereus. J R Soc Interface 12:20141109. 10.1098/rsif.2014.1109. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
206. Abbas AA, Planchon S, Jobin M, Schmitt P. 2014. Absence of oxygen affects the capacity to sporulate and the spore properties of Bacillus cereus. Food Microbiol 42:122–131. 10.1016/j.fm.2014.03.004. [Abstract] [CrossRef] [Google Scholar]

Articles from Microbiology and Molecular Biology Reviews : MMBR are provided here courtesy of American Society for Microbiology (ASM)

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/16084802
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/16084802

Article citations


Go to all (59) article citations

Other citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Funding 


Funders who supported this work.

NIAID NIH HHS (2)