american society for microbiologyeukaryotic cell

59
American Society for Microbiology Eukaryotic Cell Skip to main page content Home Current Issue Archive Alerts About ASM Contact us Tech Support Journals. ASM.org keywords GO Advanced » User Name Password Sign In Expand+ Eukaryotic Cellec.asm.org 1. doi: 10.1128/EC.2.4.655-663.2003 Eukaryotic Cell August 2003 vol. 2 no. 4 655-663

Upload: shannon-levine

Post on 05-May-2017

214 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: American Society for MicrobiologyEukaryotic Cell

American Society for Microbiology Eukaryotic Cell Skip to main page content

Home Current Issue Archive Alerts About ASM Contact us Tech Support Journals.  ASM.org

keywords  GOAdvanced »

User Name   Password   Sign In 

Expand+

Eukaryotic Cellec.asm.org

1. doi: 10.1128/EC.2.4.655-663.2003 Eukaryotic Cell August 2003 vol. 2 no. 4 655-663

A Yeast under Cover: the Capsule of Cryptococcus neoformans

1. Indrani Bose 1 ,

Page 2: American Society for MicrobiologyEukaryotic Cell

2. Amy J. Reese 1 ,3. Jeramia J. Ory 1 ,4. Guilhem Janbon 2 ,* and5. Tamara L. Doering 1 ,*

+ Author Affiliations

1. 1Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, Missouri 63110

2. 2Unité de Mycologie Moléculaire, Institut Pasteur, 75724 Paris Cedex, France

Few fungi are pathogenic to humans. Of these, Cryptococcus neoformans has emerged as an important cause of mortality in immunocompromised patients, especially those with AIDS. As a result, extensive research efforts have addressed the pathogenesis and virulence of this organism.

C. neoformans is a basidiomycetous fungus that is ubiquitous in the environment, where it is found in soil, in association with certain trees, and in bird guano (16). Because of its ubiquity, it has been suggested that most people are exposed to C. neoformans early in life (41). The fungus is heterothallic, with mating types MATa and MATα. Asexual reproduction takes place either by budding or, in the case of MATα cells, by haploid fruiting in response to nutrient deprivation or exposure to the mating pheromone a factor (106). Sexual reproduction occurs when cells of opposite mating types come together to form a heterokaryon that ultimately leads to the production of basidia and basidiospores (64). Desiccated cells and the spores formed by haploid fruiting or sexual reproduction have all been suggested to serve as infective particles, which must be less than 2 μm in diameter to penetrate the lung parenchyma (44, 86). Infection occurs when the fungal particles are inhaled and enter the alveolar space. In most immunocompetent individuals, this infection is either cleared or remains dormant until an immune imbalance leads to further development. In the setting of compromised immune function, however, the fungus disseminates, with particular tropism for the central nervous system. In severe cases, cryptococcal infection progresses to a meningoencephalitis that is fatal if left untreated.

C. neoformans virulence is mediated predominantly by a polysaccharide capsule that surrounds its cell wall and has multiple effects on the host immune system. This structure provides a physical barrier that interferes with normal phagocytosis and clearance by the immune system. Capsule components inhibit the production of proinflammatory cytokines, deplete complement components (by efficiently binding them), and reduce leukocyte migration to sites of inflammation (11). The capsule also constitutes the major diagnostic feature of cryptococcosis, because its components can be detected in the bloodstream and it can be visualized with light microscopy by using

Page 3: American Society for MicrobiologyEukaryotic Cell

India ink staining. The capsule excludes the ink particles and forms characteristic halos (Fig. 1A) whose diameters are often several times that of the cell. The elaborate structure of the capsule may also be appreciated by electron microscopy (Fig. 1B and C).

View larger version:

In this page In a new window

Download as PowerPoint Slide

FIG. 1.

Views of the cryptococcal capsule. (A) Differential interference contrast micrograph of cells that were mixed with India ink after induction of capsule formation by growth in low-iron medium (80). Scale bar, 3 μm. (B) Thin-section micrograph of a budding cell fixed in the presence of ruthenium red dye (80). Scale bar, 1 μm. (C) Quick-freeze, deep-etch image of the edge of a cell, with an arc of cell wall separating the cell interior (lower left) from the abundant capsule fibers emanating upwards. Scale bar, 0.15 μm. All images were obtained as described by Pierini and Doering (80).

Given the importance of the capsule in cryptococcal disease, tremendous effort has been applied in recent years to understanding its biology. This review focuses on the resulting advances in our understanding of the structure and synthesis of the capsular components, the incorporation of these components into the existing capsular network, the association between the capsule and the cell wall, and the regulation of capsule growth.

 Next Section

CAPSULE STRUCTURE

The capsule of C. neoformans is composed primarily of polysaccharides. Studies of its structure have taken advantage of the fact that capsular material is shed copiously into the external milieu. This material, called the exopolysaccharide, can easily be collected, purified, and used for structural studies. Historically, chemical degradation followed by paper, column, and/or gas-liquid chromatography and mass spectrometry was used to determine the composition of the capsular material. These techniques

Page 4: American Society for MicrobiologyEukaryotic Cell

identified two polysaccharide components: an abundant glucuronoxylomannan (GXM) and a minor galactoxylomannan (GalXM). Their precise structures were solved by the application of nuclear magnetic resonance techniques (27) and are shown in Fig. 2.

View larger version:

In this page In a new window

Download as PowerPoint Slide

FIG. 2.

Structures of the cryptococcal capsule components. (A) GXM of serotype B (27); (B) GalXM (96). Linkages between sugars are printed next to the arrows connecting monosaccharides; arrow position is not intended to represent linkage. Blue, mannose; gray, glucuronic acid; red, xylose; green, galactose. 6-O-acetylation of mannose, which modifies the GXM backbone, is not shown.

GXM makes up about 90% of the capsule mass. Its backbone consists of mannose residues that are α-1,3 linked and decorated with xylosyl and glucuronyl side groups (Fig. 2A). Roughly two of every three mannose residues are also 6-O-acetylated (8, 26, 61), with a preference for unbranched mannose but with some acetylation of the mannose which is substituted with glucuronic acid (56). GXM differs in the degree of xylose addition and acetylation in various cryptoccocal strains. These variations, differentiated by specific antibody binding, contribute to the classification of C. neoformans into four serotypes, A through D (35, 108). A fifth serotype, AD, has been proposed, as some strains exhibit characteristics of both the A and D serotypes (50). The antiserum classification system, however, is not equipped to distinguish the subtle heterogeneity in GXM structure that may correlate with the range of virulence seen among C. neoformans isolates. Therefore, it has also been proposed that C. neoformans strains be classified on the basis of the minimum GXM repeating unit, as determined by proton nuclear magnetic resonance (27).

Analysis of GXM structure is complicated by variability on several levels. One instructive study involved isolates from separate episodes of cryptococcosis in individual patients (25). In several cases, the strains remained the same, as determined by analysis of restriction fragment length polymorphisms (29) and electrophoretic karyotyping (78), but alterations in GXM structure resulted in their assignment to

Page 5: American Society for MicrobiologyEukaryotic Cell

different serotypes (25). Similar results have been observed after prolonged cultivation either in vitro or during cryptococcal infection in animal models (36). To further complicate the picture, all cells in a population may not bear identical GXM, as assessed by their binding of anticapsular monoclonal antibodies (F. Moyrand and G. Janbon, unpublished results).

GalXM constitutes about 7% of the capsular mass and has a more elaborate structure than GXM (96). It is built on an α-1,6-linked galactose polymer which bears side chains of different lengths on alternate galactose residues (Fig. 2B). As with GXM, the side chain structures, which consist in this case of galactosyl, mannosyl, and xylosyl residues, vary between serotypes. The cap67 mutant strain has been particularly useful in the elucidation of GalXM structure since it lacks GXM, thus simplifying analysis (55, 96, 97).

A third polysaccharide component identified in culture filtrates of C. neoformans is mannoprotein, which forms a minor fraction of the exopolysaccharide (76, 83). These mannoproteins have been the focus of immunologic studies, as they appear to be involved in the induction of cell-mediated immunity (18, 28, 70, 75, 82) and cytokine production (48,81), both critical functions for clearing the fungus. Several of these proteins have been purified from acapsular strains of Cryptococcus by concanavalin A affinity (49, 66), and one study suggests that they move through the cell wall and are released into the external environment (99). It is not clear whether the mannoproteins are capsule components. For this review, they are not considered part of the capsule structure and will not be discussed further.

Previous Section Next Section

MUTANTS IN CAPSULE SYNTHESIS

The structures of the C. neoformans capsule polysaccharides are complex (Fig. 2), and their biosynthesis requires many gene products (31). Both genetic and biochemical techniques have been applied to elucidate the steps in this pathway. One productive series of studies was based on analysis of acapsular cryptococcal strains (52). By complementing these mutants, Kwon-Chung and colleagues (19-22) identified four genes required for capsule production. These CAP genes were shown to be essential for virulence in mouse models of cryptococcosis, emphasizing the importance of the capsule as a virulence factor. Although proteins encoded by two of them have been localized, Cap60p to the nuclear membrane (20) and Cap10p to the cytoplasm (21), the biochemical functions of the CAP gene products have yet to be determined.

Page 6: American Society for MicrobiologyEukaryotic Cell

To investigate whether the CAP genes are unique or members of larger gene families, we used their sequences to search the available C. neoformans genome. Because CAP59 andCAP60 are close homologs, they were treated as part of the same gene family and only CAP10, CAP59, and CAP64 were used to search sequence data from The Institute for Genomic Research (www.tigr.org). The protein sequence for each gene was compared to the sequences of the assembled B3501 contig database by using TBLASTN (4), and hits were considered positive if a good match to at least half of the query sequence existed in the same contig (a good match was defined as a score greater than 50 bits, with an expect value below 10−10). There are at least two homologs of CAP10 in the genome, one of which is the recently identified CAP1 (65). Two additional homologs matched the first and last 50 amino acids of Cap10p but had weak homology to the rest of the protein. Searches with the CAP59 sequence revealed two homologous sequences in addition to CAP60. CAP64is also a member of a gene family, with six homologs in C. neoformans.

To assess whether CAP genes were specific for Cryptococcus in the fungal world, we used the CAP10, CAP59, and CAP64 sequences to search all fungi with publicly available genome sequences (Table 1). The only fungus with sequences homologous to all three is another basidiomycete, the white rot fungus (Phanerochaete chrysosporium). Hybridization studies previously suggested that CAP59 is unique to the genus Cryptococcus (79). However, the broader technique of sequence analysis indicates that many fungi, both pathogenic and benign, contain homologs of both CAP59 and CAP10. This suggests that these families of genes may play roles other than capsule synthesis, as these other fungi lack capsules. Unfortunately, no genome sequencing projects are under way for other encapsulated fungi. We also looked for CAP gene homologs beyond fungi (as assessed by BLink analysis of all genes and gene products in the nonredundant National Center for Biotechnology Information database). Interestingly, CAP10 has matches in metazoans with sequences of unknown function in human, mouse, and Drosophila. CAP59 also matches a hypothetical gene in the Leishmania major genome. CAP64, in contrast, has no detectable homologs outside theMycota.

View this table:

In this window In a new window

TABLE 1.

Presence of CAP genes in the fungal worlda

Page 7: American Society for MicrobiologyEukaryotic Cell

A second set of genes that were identified genetically as having a role in capsule synthesis are the CAS genes. These were discovered when one of our laboratories screened mutagenized C. neoformans cells for those that were unable to bind, or were defective in binding, monoclonal antibodies directed against the capsule (74). Analysis of these CASmutants identified six complementation groups, Cas1 through Cas6 (74), of which several members have now been examined. The CAS1 gene (56) is one of a family of genes with orthologs in human, Drosophila, and plants. Analysis of the cas1 mutant capsule showed defects in O acetylation, and interestingly, cas1Δ strains were more virulent than the parental wild type in a murine model of cryptococcosis (56). The CAS2 gene (74) was found to correspond to the previously identified UXS1 gene. UXS1 encodes a UDP-glucuronate decarboxylase which forms UDP-xylose from UDP-glucose (5). This gene is necessary for both capsule xylosylation and fungal virulence (74).

Previous Section Next Section

STEPS REQUIRED FOR CAPSULE SYNTHESIS

Genetic screens have clearly proven valuable in identifying genes whose products are required for capsule synthesis. Biochemical studies have tremendous power to complement this approach and help elucidate the functions of these gene products. These experiments have generally focused on the specific events in capsule biosynthesis that are outlined in Fig. 3. The first to be considered here is the production of the activated precursors that serve as donors for capsule components. GXM synthesis requires GDP-mannose, UDP-xylose, and UDP-glucuronic acid as biosynthetic precursors; GalXM additionally requires UDP-galactose. The acetyl donor to GXM has not been identified. Since GXM is the major component of the capsule and has a mannan backbone, biosynthesis of GDP-mannose is a critical early step. This requires the activity of three enzymes that are highly conserved in the fungal world, phosphomannose isomerase, phosphomannomutase, and GDP-mannose pyrophosphorylase (or GDP-mannose synthase) (Fig. 3, element A). The gene encoding phosphomannose isomerase in C. neoformans (MAN1) has been identified and cloned (107). man1 mutants are defective in capsule production and GXM secretion, have morphological abnormalities, and are avirulent.

View larger version:

In this page In a new window

Page 8: American Society for MicrobiologyEukaryotic Cell

Download as PowerPoint Slide

FIG. 3.

Steps in GXM synthesis. Proteins with a role in biosynthesis are indicated by red letters and are discussed further in the text. (A) Phosphomannose isomerase; (B) UDP-glucose dehydrogenase; (C) UDP-glucuronic acid decarboxylase; (D) nucleotide sugar transporters specific for GDP-mannose (D1), UDP-glucuronic acid (D2), and UDP-xylose (D3); (E) mannosyltransferase; (F) xylosyltransferase; (G) glucuronosyltransferase; and (H) acetyltransferase. The color coding is the same as described in the legend to Fig. 2, with the addition of purple to represent glucose.

The nucleotide sugars UDP-glucuronic acid (which donates a major component of GXM) and UDP-xylose (required for both GXM and GalXM) can be made from UDP-glucose by the sequential actions of two enzymes. UDP-glucose dehydrogenase leads to the formation of UDP-glucuronic acid (Fig. 3, element B), while UDP-glucuronate decarboxylase converts UDP-glucuronic acid to UDP-xylose (Fig. 3, element C). These activities were detected in C. neoformans by Jacobson and Payne (51, 53), and subsequent work in one of our laboratories has identified the genes encoding both of these enzymes. One of these genes (UXS1), encoding UDP-glucuronate decarboxylase, was found through its homology to a bacterial gene that had been hypothesized to have a similar function (5). Although this decarboxylase is important for synthetic processes across biological kingdoms, the sequence for this enzyme had not previously been obtained in any system.

Once biosynthetic precursors are made, the next probable step in capsule synthesis is their localization to the site of polysaccharide construction. This is likely to require several specific transporters (Fig. 3, element D), since the charged precursors are generally formed in the cytosol, while most eukaryotic glycan synthesis occurs in membrane-bound compartments. Several genes encoding putative nucleotide sugar transporters have been identified in C. neoformans by sequence homology performed in our laboratories. Functional assays of these putative transporters should clarify their potential role in capsule formation and cryptococcal biology.

After appropriate precursors are transported into the compartment of capsule biosynthesis, the monosaccharides must be assembled into polymers. As previously noted (31), there are at least 11 linkages present in the various configurations of GXM and GalXM. Each linkage probably reflects the function of a different sugar transferase, because these enzymes are usually specific for the substrate, the moiety transferred, and the bond formed. Constructing GXM and GalXM would therefore require an array of sugar transferases (for examples, see Fig. 3, elements E to G) and an acetyltransferase (Fig. 3, element H).

Page 9: American Society for MicrobiologyEukaryotic Cell

Several distinct mannosyl transferase activities have been identified in C. neoformans by using crude membrane preparations. Two of these, identified in efforts to isolate enzymes involved in capsule synthesis, are in fact α-1,2-mannosyltransferases and therefore lack the requisite activity (32, 104). A third, however, can transfer mannose from GDP-mannose to α-1,3-linked mannobiose and would therefore be capable of building the GXM backbone (32). This enzyme has been purified, and its peptide sequence has been used to obtain the corresponding gene sequence. Inhibition of its expression shows that it is required for normal capsule synthesis (U. Sommer, H. Liu, and T. L. Doering, unpublished data). In addition, xylosyltransferase (Fig. 3, element G) and glucuronyltransferase (Fig. 3, element F) activities have been detected in crude membrane preparations of C. neoformans (103). Study of these activities led to the proposal that the order of addition of side groups to the mannan backbone is first, acetyl; second, glucuronyl; and third, xylosyl (103). However, subsequent investigation of GXM structures in a mutant lacking xylose suggests that acetylation follows the addition of xylose (74).

Previous Section Next Section

CAPSULE ASSEMBLY AND DEGRADATION

Once individual capsule components are generated by the biosynthetic pathways discussed above, they must be assembled into the elaborate meshwork of the mature capsule. How are these new materials incorporated to generate and maintain this crucial virulence factor? Capsule synthesis has been investigated in two physiologically important situations: bud formation, which occurs with concomitant encapsulation of the daughter cell, and capsule synthesis on mature cells, which is induced by altering growth conditions (80). In budding cells, new-capsule synthesis is directed into the bud, with no incorporation of the maternal capsule into the nascent polysaccharide structure overlying the daughter cell (Fig. 4). In mature cells, new capsule material is incorporated just outside the cell wall, at the inner aspect of the existing capsule structure (80). This displaces the existing material outwards, although there is also some mixing of old and new polysaccharide and an overall increase in density of the polysaccharide structure with time. Material displaced by new synthesis could conceivably be shed from the cell into the environment or it could remain associated with the capsule, perhaps fixed in place by cross-linking processes such as those suggested to occur during cell wall maturation (67). To fully understand these processes, the mechanism by which the capsule is initially associated with the cell must first be defined.

View larger version:

Page 10: American Society for MicrobiologyEukaryotic Cell

In this page In a new window

Download as PowerPoint Slide

FIG. 4.

Cell wall and newly synthesized capsule overlying nascent buds. Cell walls were tagged with fluorescein (green), and capsules were labeled with fluorophore-linked antibody to GXM (red). Cells were then returned to culture to allow bud emergence. A comparison of differential interference contrast (left) and fluorescence (right) images shows that none of the original label extends beyond the mother-bud neck (arrow), indicating that there is no incorporation of maternal glycans in the new bud. Reprinted with permission from Blackwell Publishing (80).

Early studies of how capsule material associates with the cell were based on the observation that capsular exopolysaccharide could bind to acapsular cells. Capsule material purified from growth medium of wild-type cryptococcal cells was first shown to bind to acapsular mutants and to protect them from phagocytosis (12, 13, 91). This interaction was specific for cryptococcal cells and cryptococcal capsule material (62) and was shown to be saturable and reversible (63). Chemical treatments of acceptor cells and purified capsule material were subsequently investigated to test their effects on binding (89). These studies led to a model in which polysaccharide capsule material binds to a specific receptor on the surface of acapsular mutants.

To pursue the question of how capsule polysaccharides interact with the cell, an in vivo assay for the binding of capsule to acapsular cells has been employed. In this assay, conditioned medium from wild-type cells serves as a capsule source and indirect immunofluorescence with a fluorophore-tagged anticapsule antibody is used to evaluate the transfer of polysaccharides to acapsular acceptor cells. Studies using this model strongly implicate cell wall glycan in capsule binding (A. J. Reese and T. L. Doering, submitted for publication).

The relationship of the capsule to the cell wall necessitates an examination of C. neoformans cell wall composition. While much is known about the cell wall structures ofSaccharomyces cerevisiae and Candida albicans, considerably less is known about those of other fungal organisms (Table 2). Common cell wall structural building blocks include β-glucans and chitin (14, 15). Other possible cell wall components in fungi include glycoproteins (particularly mannoproteins), α-glucans (14), and melanin (17). In S. cerevisiae (90) andCandida albicans (60), the main cell wall network is formed by branched β-1,3-linked glucans, with chitin and several types of mannoproteins linked directly to it. Other fungi, including C. neoformans, have a high

Page 11: American Society for MicrobiologyEukaryotic Cell

α-1,3-glucan content, and their cell wall structures might be considerably different. Binding assays indicate that α-1,3-glucan at the cryptococcal surface is required for acapsular cells to bind exogenous capsule material. Posttranscriptional silencing of a cryptococcal α-glucan synthase gene yields cells with poor ability to assemble capsules, although the components are still made (Reese and Doering, submitted). It is interesting that in the pathogenic fungi Blastomyces dermatitidis and Histoplasma capsulatum, virulence is associated with α-1,3-glucan levels (47, 59), although these fungi are not encapsulated.

View this table:

In this window In a new window

TABLE 2.

Major cell wall components of selected fungal organismsa

While studies of capsule synthesis have progressed, capsule degradation remains poorly understood. Although enzymes from soil microbes can degrade capsules (37), no cryptococcal or human enzymes with this activity have been found. Nonetheless, capsule material must be degraded or remodeled by the cell to allow bud emergence, since buds are coated with newly synthesized capsule material (Fig. 4) (80). The mammalian host may also have mechanisms for capsule breakdown, but further studies in this area are needed to clarify these events.

Previous Section Next Section

REGULATION OF THE CAPSULE

The C. neoformans capsule varies tremendously in size and morphology under different growth conditions. As early as 1958, Littman studied the effects of various nutrients on capsule formation and formulated a capsule-inducing medium containing thiamine, sodium glutamate, mineral salts, maltose, and sucrose (68). Low-iron medium (98), alterations in pH, and a high ratio of CO2 to HCO3 (43) also induce capsule formation, while other medium additives, such as 1 M sodium chloride, suppress capsule formation (54). Suppression of capsule synthesis by sodium chloride is specific, as other hypertonic solutions do not have this effect. Capsule size also increases dramatically during infection (6, 40). The extent of capsule formation depends on the tissue location, with cells in the brain having larger capsules than those in the lungs (84). This is probably due to differences in the CO2/HCO3 ratio or

Page 12: American Society for MicrobiologyEukaryotic Cell

other capsule-modulating factors in those sites, again stressing the highly regulated nature of this virulence factor.

Efforts to understand capsule regulation have focused attention on second messengers, which mediate aspects of growth and development in eukaryotic cells. In particular, studies have addressed the second messenger cyclic AMP (cAMP), which is produced in cells by the enzyme adenylyl cyclase. The activity of adenylyl cyclase in mammalian cells is regulated by heterotrimeric G proteins (39), which are in turn activated by seven transmembrane receptors of the β-adrenergic receptor family. Binding of an appropriate ligand to the receptor leads the heterotrimeric G protein to exchange GDP for GTP, causing dissociation of the α subunit from the βγ dimer (Fig. 5). Depending on the organism, either one of these products may activate an adenylyl cyclase. The subsequent production of cAMP has diverse effects, one of the best known of which is the activation of cAMP-dependent protein kinase A (PKA). PKA in turn regulates downstream targets, including transcription factors, leading to broad changes in cellular physiology (92).

View larger version:

In this page In a new window

Download as PowerPoint Slide

FIG. 5.

cAMP signaling pathway in the regulation of the capsule. See text for explanation. Red symbols, Pkr1p; black circles, Pka1p; blue-gray circles, cAMP.

cAMP has been shown to regulate capsule production in C. neoformans (2, 3). Implicated in the upstream pathway are the GPA1 and CAC1 gene products, which have been shown to control intracellular cAMP production. The GPA1 product is homologous to the S. cerevisiae Gα subunit Gpa2p (94), while the CAC1 product is the adenylyl cyclase (3). Bothgpa1 and cac1 mutants are unable to induce capsule production and are correspondingly avirulent, but overproduction of CAC1 can suppress the gpa1 mutant defects (3). These defects can also be corrected by the addition of cAMP to the medium, thereby bypassing the requirement for Gpa1p and Cac1p and supporting the crucial role that signaling via cAMP plays in capsule formation.

Page 13: American Society for MicrobiologyEukaryotic Cell

Among the potential targets of cAMP, the C. neoformans PKA1 gene product has a clear role in capsule production (33), sinces a pka1 strain is acapsular and avirulent. In this case, capsule production is not restored by exogenous cAMP, suggesting that PKA acts downstream of cAMP in the signaling pathway (Fig. 5). Further investigations of PKA in C. neoformans are based on its subunit structure. Inactive PKA is a tetramer of two catalytic subunits (encoded by PKA1) and two regulatory subunits (encoded by PKR1). Binding of cAMP to the regulatory subunits causes them to undergo a conformational change and to dissociate from the catalytic subunits, releasing the latter from inhibition (Fig. 5). Loss-of-function mutations of PKR1 should, therefore, lead to constitutive activation of PKA1. This is indeed the case, as pkr1 mutants produce larger capsules than wild-type cells do and are hypervirulent in animal models. pkr1 mutants can also suppress the capsule production defect of gpa1 cells (33). However, the capsules produced by the pkr1 gpa1 double mutant are larger than those produced by the pkr1 single mutant. This suggests that GPA1 may have additional targets, yet to be identified, that modulate capsule production (34). Once active, Pka1p kinase can cause changes in gene expression and physiology by phosphorylation of effector proteins. One candidate target of PKA, which is involved in positive regulation of capsule production, is the transcription factor encoded by STE12, which is found in two mating-type-specific forms (105, 110). This transcription factor is probably only one of several targets (Fig. 5), because loss of function of STE12α or STE12a leads only to a decrease in capsule size (23, 24, 110) and overexpression of STE12α does not overcome the capsule production defect of a pka1 strain (33).

In many other systems, the Ras signaling protein plays a role in the activation of adenylyl cyclase. However, RAS1 does not appear to be involved in capsule production in C. neoformans (1, 2). Other major signaling pathways in C. neoformans that have been investigated include the mitogen-activated protein kinase pathway and the signaling pathway of cyclophilins. Although both of these pathways have been shown to regulate virulence and mating (100, 101), they are not clearly implicated in capsule production.

Previous Section Next Section

OTHER CAPSULES

C. neoformans is unique in being the only fungus that is both pathogenic to humans and in possession of a true polysaccharide capsule outside its cell wall. Tremella mesenterica, a close phylogenetic relative, possesses a similar capsule but does not cause disease (30). Several other fungi have capsule-like structures, but they have not been well characterized. These fungi include Malassezia furfur (73), Rhinosporidium

Page 14: American Society for MicrobiologyEukaryotic Cell

seeberi (93), Trichosporon beigelii (72), Blastocystis hominis (71), and Sporothrix schenckii (38).

Among other microbes, many gram-negative and gram-positive bacteria also bear polysaccharide capsules. The numerous examples among pathogens include Streptococcus pneumoniae, Neisseria meningitidis, and Pseudomonas aeruginosa (109). Capsules protect bacterial cells from desiccation, mediate their adherence to host tissue, and assist in evasion of the host immune response (85). Some bacteria also shed exopolysaccharides into their surroundings as a method of avoiding immune system clearance. There is a tremendous variety of polysaccharide structures present in bacterial capsules (85, 109). All of them, however, are quite distinct from those of the C. neoformans capsule and, therefore, require an independent set of enzymes for their synthesis and assembly.

The value of capsules has also been appreciated beyond the microbial world with the use of microencapsulation techniques in the food industry. In this application, alginate or other polysaccharides are used to microencapsulate bacteria (57). This process makes use of the protective nature of polysaccharide capsules to assist in the survival of probiotic bacteria.

Previous Section Next Section

FUTURE DIRECTIONS

The biogenesis of the capsule in Cryptococcus is a complicated process, and rapid advances in understanding its structure, synthesis, and regulation have been seen in the past few years. However, there are vast areas that are still unexplored and numerous questions that remain unanswered. Filling these gaps will require concerted effort in all fields of biology—biophysical, biochemical, genetic, and cell biological.

The structure of the major capsular polysaccharides has been established and suggests the synthetic precursors required, but how is the synthesis of these polysaccharides initiated? Similarly, the question of whether the addition of new material to the capsule occurs in units of single monosaccharides, short oligosaccharides, or extensive polysaccharide chains remains unanswered.

Our present understanding of the biochemical reactions in capsule synthesis is patchy. The participants in some of the predicted reactions, such as GXM acetylation, have not yet been defined. Other capsular synthetic activities have been biochemically identified only in lysates and have yet to be studied in detail or correlated with a corresponding gene. Only a few have been purified and cloned. Without pure enzyme preparations or strains deficient in specific activities, it becomes extremely difficult to

Page 15: American Society for MicrobiologyEukaryotic Cell

determine the order of the reactions that add side chain subunits to the polysaccharide backbone and, therefore, to delineate the entire biosynthetic pathway.

There are several approaches that can identify additional genes with potential roles in capsule synthesis. These include identifying gene families within Cryptococcus, finding cryptococcal genes with homology to those encoding relevant proteins in other organisms, and examining mutants with altered capsules. Once important enzymes are identified and functionally validated, their localization will be crucial in determining the sites and organelles involved in capsule synthesis. Determination of the transport pathways that carry capsule components to the cell surface is another area that is important but largely unexplored.

In the search for the molecular basis of capsule regulation, a cAMP-signaling pathway that modulates capsule production has been identified. Genetics has been used to define many of the members of this pathway. However, unanswered questions remain. What is the nature of the signal that activates the pathway? What are the identities of the transmembrane receptor and of downstream effectors? A variety of nutrients and environmental conditions also affect capsule production, but what are the regulatory pathways by which they do so? The regulation of capsule structure presents fascinating questions in addition to those about how much capsule is made. Capsule structure depends on environment (95), and studies with anticapsule antibodies suggest that structure may vary among cells in a population (Moyrand and Janbon, unpublished results), but the mechanisms controlling these differences are yet to be defined.

The capsule polysaccharides that are shed into the external milieu enable C. neoformans to resist the host immune response. Little is known, however, about how this occurs. Are the old parts of the capsule sloughed off as new material is added, or is the release of capsular material a regulated process under specific cellular control? Also, how does degradation of the capsule take place? The molecular mechanisms of these processes remain obscure.

Finally, although the structures of the main components of the capsule are now known, how do these components combine to form the three-dimensional structure of the capsule? Electron micrographs show an elaborate network of fibers that extend out from the cell wall. However, these images do not address whether these strands are composed of one or multiple polysaccharide components or how they group and branch. A better understanding of the biochemical and biophysical properties of the capsule is needed to answer these questions.

For many years, the study of the cryptococcal capsule was hindered by the absence of appropriate tools. This picture has been completely redrawn by the recruitment and

Page 16: American Society for MicrobiologyEukaryotic Cell

adaptation of new technologies to the cause of cryptococcal research, from genome sequencing (45, 88) and RNA interference (69) to structural studies. These tools, coupled with the facts that gene disruptions are now routinely performed and animal models are in place to study in vivo effects, have considerably advanced our understanding of the Cryptococcus capsule. The coming years will be an exciting time for studies of cryptococcal biology.

Previous Section Next Section

ACKNOWLEDGMENTS

We thank members of the Doering lab and Tom Kozel for helpful discussions, Wandy Beatty for assistance in the preparation of Fig. 1, Andy Alspaugh and Chad Rappleye for comments on the manuscript, and John Heuser for quick-freeze deep-etch electron microscopy.

Studies of capsule synthesis in the Doering laboratory are supported by NIH grants GM66303 and AI49173, a Burroughs Wellcome Fund New Investigator Award in Molecular Pathogenic Mycology, an award from the Edward J. Mallinckrodt, Jr., Foundation, and a Craig Faculty Fellowship. Studies of capsule synthesis in the Janbon laboratory are supported by Ensemble contre le SIDA (grant AO13-3) and the Pasteur Institute.

Previous Section Next Section

FOOTNOTES

*Corresponding author. Mailing address for Guilhem Janbon: Unité de Mycologie Moléculaire, Institut Pasteur, 25 rue du Dr. Roux, 75724 Paris Cedex, France. Phone: 33 145 688356. Fax: 33 145 688420. E-mail: [email protected]. Mailing address for Tamara L. Doering: Department of Molecular Microbiology, Washington University School of Medicine, St. Louis, MO 63110. Phone: (314) 747-5597. Fax: (314) 362-1232. E-mail: [email protected].

Previous Section 

REFERENCES

1. ↵

Page 17: American Society for MicrobiologyEukaryotic Cell

Alspaugh, J. A., L. M. Cavallo, J. R. Perfect, and J. Heitman. 2000. RAS1 regulates filamentation, mating and growth at high temperature of Cryptococcus neoformans. Mol. Microbiol. 36:352-365.

CrossRef Medline

2. ↵ Alspaugh, J. A., J. R. Perfect, and J. Heitman. 1997. Cryptococcus neoformans mating and virulence are regulated by the G-protein α subunit GPA1 and cAMP.Genes Dev. 11:3206-3217.

Abstract/FREE Full Text

3. ↵ Alspaugh, J. A., R. Pukkila-Worley, T. Harashima, L. M. Cavallo, D. Funnell, G. M. Cox, J. R. Perfect, J. W. Kronstad, and J. Heitman. 2002. Adenylyl cyclase functions downstream of the Gα protein Gpa1 and controls mating and pathogenicity of Cryptococcus neoformans. Eukaryot. Cell 1:75-84.

Abstract/FREE Full Text

4. ↵ Altschul, S. F., W. Gish, W. Miller, E. W. Myers, and D. J. Lipman. 1990. Basic local alignment search tool. J. Mol. Biol. 215:403-410.

CrossRef Medline

5. ↵ Bar-Peled, M., C. L. Griffith, and T. L. Doering. 2001. Functional cloning and characterization of a UDP-glucuronic acid decarboxylase: the pathogenic fungusCryptococcus neoformans elucidates UDP-xylose synthesis. Proc. Natl. Acad. Sci. USA 98:12003-12008.

Abstract/FREE Full Text

6. ↵

Page 18: American Society for MicrobiologyEukaryotic Cell

Bergman, F. 1965. Studies on capsule synthesis of Cryptococcus neoformans. Sabouraudia 4:23-31.

Medline

7. Bernard, M., and J.-P. Latgé. 2001. Aspergillus fumigatus cell wall: composition and biosynthesis. Med. Mycol. 39:9-17.

Medline

8. ↵ Bhattacharjee, A. K., J. E. Bennett, and C. P. Glaudemans. 1984. Capsular polysaccharides of Cryptococcus neoformans. Rev. Infect. Dis. 6:619-624.

Medline

9. Borges-Walmsley, M., D. Chen, X. Shu, and A. Walmsley. 2002. The pathobiology of Paracoccidioides brasiliensis. Trends Microbiol. 10:80-87.

CrossRef Medline

10.Brandhorst, T., and B. Klein. 2000. Cell wall biogenesis of Blastomyces dermatitidis. Evidence for a novel mechanism of cell surface localization of virulence-associated adhesin via extracellular release and reassociation with cell wall chitin. J. Biol. Chem. 275:7925-7934.

Abstract/FREE Full Text

11.↵ Buchanan, K. L., and J. W. Murphy. 1998. What makes Cryptococcus neoformans a pathogen? Emerg. Infect. Dis. 4:71-83.

Medline

12.↵ Bulmer, G. S., and M. D. Sans. 1967. Cryptococcus neoformans. II. Phagocytosis by human leukocytes. J. Bacteriol. 94:1480-1483.

Abstract/FREE Full Text

Page 19: American Society for MicrobiologyEukaryotic Cell

13.↵ Bulmer, G. S., and M. D. Sans. 1968. Cryptococcus neoformans. III. Inhibition of phagocytosis. J. Bacteriol. 95:5-8.

Abstract/FREE Full Text

14.↵ Cabib, E., B. Bowers, A. Sburlati, and S. Silverman. 1988. Fungal cell wall synthesis: the construction of a biological structure. Microbiol. Sci. 5:370-375.

Medline

15.↵ Cabib, E., T. Drgon, J. Drgonová, R. A. Ford, and R. Kollár. 1997. The yeast cell wall, a dynamic structure engaged in growth and morphogenesis. Biochem. Soc. Trans. 25:200-204.

Medline

16.↵ Casadevall, A., and J. R. Perfect. 1998. Cryptococcus neoformans. ASM Press, Washington, D.C.

17.↵ Casadevall, A., A. L. Rosas, and J. D. Nosanchuk. 2000. Melanin and virulence in Cryptococcus neoformans. Curr. Opin. Microbiol. 3:354-358.

CrossRef Medline

18.↵ Chaka, W., A. F. Verheul, V. V. Vaishnav, R. Cherniak, J. Scharringa, J. Verhoef, H. Snippe, and I. M. Hoepelman. 1997. Cryptococcus neoformans and cryptococcal glucuronoxylomannan, galactoxylomannan, and mannoprotein induce different levels of tumor necrosis factor alpha in human peripheral blood mononuclear cells. Infect. Immun. 65:272-278.

Page 20: American Society for MicrobiologyEukaryotic Cell

Abstract/FREE Full Text

19.↵ Chang, Y. C., and K. J. Kwon-Chung. 1994. Complementation of a capsule-deficient mutation of Cryptococcus neoformans restores its virulence. Mol. Cell. Biol.14:4912-4919.

Abstract/FREE Full Text

20.↵ Chang, Y. C., and K. J. Kwon-Chung. 1998. Isolation of the third capsule-associated gene, CAP60, required for virulence in Cryptococcus neoformans. Infect. Immun. 66:2230-2236.

Abstract/FREE Full Text

21.↵ Chang, Y. C., and K. J. Kwon-Chung. 1999. Isolation, characterization, and localization of a capsule-associated gene, CAP10, of Cryptococcus neoformans. J. Bacteriol. 181:5636-5643.

Abstract/FREE Full Text

22.↵ Chang, Y. C., L. A. Penoyer, and K. J. Kwon-Chung. 1996. The second capsule gene of Cryptococcus neoformans, CAP64, is essential for virulence. Infect. Immun.64:1977-1983.

Abstract/FREE Full Text

23.↵ Chang, Y. C., L. A. Penoyer, and K. J. Kwon-Chung. 2001. The second STE12 homologue of Cryptococcus neoformans is MATa-specific and plays an important role in virulence. Proc. Natl. Acad. Sci. USA 98:3258-3263.

Abstract/FREE Full Text

Page 21: American Society for MicrobiologyEukaryotic Cell

24.↵ Chang, Y. C., B. L. Wickes, G. F. Miller, L. A. Penoyer, and K. J. Kwon-Chung. 2000. Cryptococcus neoformans STE12α regulates virulence but is not essential for mating. J. Exp. Med. 191:871-882.

Abstract/FREE Full Text

25.↵ Cherniak, R., L. C. Morris, T. Belay, E. D. Spitzer, and A. Casadevall. 1995. Variation in the structure of glucuronoxylomannan in isolates from patients with recurrent cryptococcal meningitis. Infect. Immun. 63:1899-1905.

Abstract/FREE Full Text

26.↵ Cherniak, R., and J. B. Sundstrom. 1994. Polysaccharide antigens of the capsule of Cryptococcus neoformans. Infect. Immun. 62:1507-1512.

Abstract/FREE Full Text

27.↵ Cherniak, R., H. Valafar, L. C. Morris, and F. Valafar. 1998. Cryptococcus neoformans chemotyping by quantitative analysis of 1H nuclear magnetic resonance spectra of glucuronoxylomannans with a computer-simulated artificial neural network. Clin. Diagn. Lab. Immunol. 5:146-159.

Abstract/FREE Full Text

28.↵ Coenjaerts, F. E., A. M. Walenkamp, P. N. Mwinzi, J. Scharringa, H. A. Dekker, J. A. van Strijp, R. Cherniak, and A. I. Hoepelman. 2001. Potent inhibition of neutrophil migration by cryptococcal mannoprotein-4-induced desensitization. J. Immunol. 167:3988-3995.

Abstract/FREE Full Text

29.↵

Page 22: American Society for MicrobiologyEukaryotic Cell

Currie, B. P., L. F. Freundlich, and A. Casadevall. 1994. Restriction fragment length polymorphism analysis of Cryptococcus neoformans isolates from environmental (pigeon excreta) and clinical sources in New York City. J. Clin. Microbiol. 32:1188-1192.

Abstract/FREE Full Text

30.↵ De Baets, S., S. Du Laing, C. Francois, and E. J. Vandamme. 2002. Optimization of exopolysaccharide production by Tremella mesenterica NRRL Y-6158 through implementation of fed-batch fermentation. J. Ind. Microbiol. Biotechnol. 29:181-184.

CrossRef Medline

31.↵ Doering, T. L. 2000. How does Cryptococcus get its coat? Trends Microbiol. 8:547-553.

CrossRef Medline

32.↵ Doering, T. L. 1999. A unique α-1,3 mannosyltransferase of the pathogenic fungus Cryptococcus neoformans. J. Bacteriol. 181:5482-5488.

Abstract/FREE Full Text

33.↵ D'Souza, C. A., J. A. Alspaugh, C. Yue, T. Harashima, G. M. Cox, J. R. Perfect, and J. Heitman. 2001. Cyclic AMP-dependent protein kinase controls virulence of the fungal pathogen Cryptococcus neoformans. Mol. Cell. Biol. 21:3179-3191.

Abstract/FREE Full Text

34.↵

Page 23: American Society for MicrobiologyEukaryotic Cell

D'Souza, C. A., and J. Heitman. 2001. Conserved cAMP signaling cascades regulate fungal development and virulence. FEMS Microbiol. Rev. 25:349-364.

CrossRef Medline

35.↵ Evans, E. E. 1950. The antigenic composition of Cryptococcus neoformans. I. A serologic classification by means of the capsular and agglutination reactions. J. Immunol. 64:423-430.

Abstract/FREE Full Text

36.↵ Franzot, S. P., J. Mukherjee, R. Cherniak, L. C. Chen, J. S. Hamdan, and A. Casadevall. 1998. Microevolution of a standard strain of Cryptococcus neoformansresulting in differences in virulence and other phenotypes. Infect. Immun. 66:89-97.

Abstract/FREE Full Text

37.↵ Gadebusch, H. H., and J. D. Johnson. 1961. Specific degradation of Cryptococcus neoformans 3723 capsular polysaccharide by a microbial enzyme. Can. J. Microbiol. 7:53-60.

Medline

38.↵ Garrison, R. G., and F. K. Mirikitani. 1983. Electron cytochemical demonstration of the capsule of yeast-like Sporothrix schenckii. Sabouraudia 21:167-170.

Medline

39.↵ Gilman, A. G. 1984. G proteins and dual control of adenylate cyclase. Cell 36:577-579.

Page 24: American Society for MicrobiologyEukaryotic Cell

CrossRef Medline

40.↵ Goldman, D., S. C. Lee, and A. Casadevall. 1994. Pathogenesis of pulmonary Cryptococcus neoformans infection in the rat. Infect. Immun. 62:4755-4761.

Abstract/FREE Full Text

41.↵ Goldman, D. L., H. Khine, J. Abadi, D. J. Lindenberg, L. Pirofski, R. Niang, and A. Casadevall. 2001. Serologic evidence for Cryptococcus neoformans infection in early childhood. Pediatrics 107:E66. [Online.] http://www.pediatrics.org.

CrossRef Medline

42.Gomez, B. L., J. D. Nosanchuk, S. Diez, S. Youngchim, P. Aisen, L. E. Cano, A. Restrepo, A. Casadevall, and A. J. Hamilton. 2001. Detection of melanin-like pigments in the dimorphic fungal pathogen Paracoccidioides brasiliensis in vitro and during infection. Infect. Immun. 69:5760-5767.

Abstract/FREE Full Text

43.↵ Granger, D. L., J. R. Perfect, and D. T. Durack. 1985. Virulence of Cryptococcus neoformans. Regulation of capsule synthesis by carbon dioxide. J. Clin. Investig.76:508-516.

Medline

44.↵ Hatch, T. F. 1961. Distribution and deposition of inhaled particles in respiratory tract. Bacteriol. Rev. 25:237-240.

FREE Full Text

45.↵

Page 25: American Society for MicrobiologyEukaryotic Cell

Heitman, J., A. Casadevall, J. K. Lodge, and J. R. Perfect. 1999. The Cryptococcus neoformans genome sequencing project. Mycopathologia 148:1-7.

CrossRef Medline

46.Hochstenbach, F., F. M. Klis, H. van den Ende, E. van Donselaar, P. J. Peters, and R. D. Klausner. 1998. Identification of a putative α-glucan synthase essential for cell wall construction and morphogenesis in fission yeast. Proc. Natl. Acad. Sci. USA 95:9161-9166.

Abstract/FREE Full Text

47.↵ Hogan, L. H., and B. S. Klein. 1994. Altered expression of surface α-1,3-glucan in genetically related strains of Blastomyces dermatitidis that differ in virulence.Infect. Immun. 62:3543-3546.

Abstract/FREE Full Text

48.↵ Huang, C., and S. M. Levitz. 2000. Stimulation of macrophage inflammatory protein-1α, macrophage inflammatory protein-1β, and RANTES by Candida albicansand Cryptococcus neoformans in peripheral blood mononuclear cells from persons with and without human immunodeficiency virus infection. J. Infect. Dis. 181:791-794.

Abstract/FREE Full Text

49.↵ Huang, C., S. H. Nong, M. K. Mansour, C. A. Specht, and S. M. Levitz. 2002. Purification and characterization of a second immunoreactive mannoprotein fromCryptococcus neoformans that stimulates T-cell responses. Infect. Immun. 70:5485-5493.

Abstract/FREE Full Text

50.↵

Page 26: American Society for MicrobiologyEukaryotic Cell

Ikeda, R., A. Nishikawa, T. Shinoda, and Y. Fukazawa. 1985. Chemical characterization of capsular polysaccharide from Cryptococcus neoformans serotype A-D.Microbiol. Immunol. 29:981-991.

Medline

51.↵ Jacobson, E. S. 1987. Cryptococcal UDP-glucose dehydrogenase: enzymic control of capsular biosynthesis. J. Med. Vet. Mycol. 25:131-135.

Medline

52.↵ Jacobson, E. S., D. J. Ayers, A. C. Harrell, and C. C. Nicholas. 1982. Genetic and phenotypic characterization of capsule mutants of Cryptococcus neoformans. J. Bacteriol. 150:1292-1296.

Abstract/FREE Full Text

53.↵ Jacobson, E. S., and W. R. Payne. 1982. UDP glucuronate decarboxylase and synthesis of capsular polysaccharide in Cryptococcus neoformans. J. Bacteriol.152:932-934.

Abstract/FREE Full Text

54.↵ Jacobson, E. S., M. J. Tingler, and P. L. Quynn. 1989. Effect of hypertonic solutes upon the polysaccharide capsule in Cryptococcus neoformans. Mycoses 32:14-23.

55.↵ James, P. G., R. Cherniak, R. G. Jones, C. A. Stortz, and E. Reiss. 1990. Cell-wall glucans of Cryptococcus neoformans Cap 67. Carbohydr. Res. 198:23-38.

CrossRef Medline

Page 27: American Society for MicrobiologyEukaryotic Cell

56.↵ Janbon, G., U. Himmelreich, F. Moyrand, L. Improvisi, and F. Dromer. 2001. Cas1p is a membrane protein necessary for the O-acetylation of the Cryptococcus neoformans capsular polysaccharide. Mol. Microbiol. 42:453-467.

CrossRef Medline

57.↵ Kailasapathy, K. 2002. Microencapsulation of probiotic bacteria: technology and potential applications. Curr. Issues Intest. Microbiol. 3:39-48.

Medline

58.Kanetsuma, F., and L. M. Carbonell. 1971. Cell wall composition of the yeastlike and mycelial forms of Blastomyces dermatitidis. J. Bacteriol. 106:946-948.

Abstract/FREE Full Text

59.↵ Klimpel, K. R., and W. E. Goldman. 1988. Cell walls from avirulent variants of Histoplasma capsulatum lack α-(1,3)-glucan. Infect. Immun. 56:2997-3000.

Abstract/FREE Full Text

60.↵ Klis, F. M., P. de Groot, and K. Hellingwerf. 2001. Molecular organization of the cell wall of Candida albicans. Med. Mycol. 39(Suppl. 1):1-8.

Medline

61.↵ Kozel, T. R. 1989. Antigenic structure of Cryptococcus neoformans capsular polysaccharides. Immunol. Ser. 47:63-86.

Medline

Page 28: American Society for MicrobiologyEukaryotic Cell

62.↵ Kozel, T. R. 1977. Non-encapsulated variant of Cryptococcus neoformans. II. Surface receptors for cryptococcal polysaccharide and their role in inhibition of phagocytosis by polysaccharide. Infect. Immun. 16:99-106.

Abstract/FREE Full Text

63.↵ Kozel, T. R., and C. A. Hermerath. 1984. Binding of cryptococcal polysaccharide to Cryptococcus neoformans. Infect. Immun. 43:879-886.

Abstract/FREE Full Text

64.↵ Kwon-Chung, K. J. 1975. A new genus, Filobasidiella, the perfect state of Cryptococcus neoformans. Mycologia 67:1197-1200.

CrossRef Medline

65.↵ Lengeler, K. B., D. S. Fox, J. A. Fraser, A. Allen, K. Forrester, F. S. Dietrich, and J. Heitman. 2002. Mating-type locus of Cryptococcus neoformans: a step in the evolution of sex chromosomes. Eukaryot. Cell 1:704-718.

Abstract/FREE Full Text

66.↵ Levitz, S. M., S. Nong, M. K. Mansour, C. Huang, and C. A. Specht. 2001. Molecular characterization of a mannoprotein with homology to chitin deacetylases that stimulates T cell responses to Cryptococcus neoformans. Proc. Natl. Acad. Sci. USA 98:10422-10427.

Abstract/FREE Full Text

67.↵

Page 29: American Society for MicrobiologyEukaryotic Cell

Lipke, P. N., and R. Ovalle. 1998. Cell wall architecture in yeast: new structure and new challenges. J. Bacteriol. 180:3735-3740.

FREE Full Text

68.↵ Littman, M. L. 1958. Capsule synthesis by Cryptococcus neoformans. Trans. N. Y. Acad. Sci. 20:623-648.

Medline

69.↵ Liu, H., T. R. Cottrell, L. M. Pierini, W. E. Goldman, and T. L. Doering. 2002. RNA interference in the pathogenic fungus Cryptococcus neoformans. Genetics160:463-470.

Abstract/FREE Full Text

70.↵ Mansour, M. K., L. S. Schlesinger, and S. M. Levitz. 2002. Optimal T cell responses to Cryptococcus neoformans mannoprotein are dependent on recognition of conjugated carbohydrates by mannose receptors. J. Immunol. 168:2872-2879.

Abstract/FREE Full Text

71.↵ Matsumoto, Y., M. Yamada, and Y. Yoshida. 1987. Light-microscopical appearance and ultrastructure of Blastocystis hominis, an intestinal parasite of man. Zentbl. Bakteriol. Mikrobiol. Hyg. Abt. 1 Orig. A 264:379-385.

72.↵ Melcher, G. P., K. D. Reed, M. G. Rinaldi, J. W. Lee, P. A. Pizzo, and T. J. Walsh. 1991. Demonstration of a cell wall antigen cross-reacting with cryptococcal polysaccharide in experimental disseminated trichosporonosis. J. Clin. Microbiol. 29:192-196.

Page 30: American Society for MicrobiologyEukaryotic Cell

Abstract/FREE Full Text

73.↵ Mittag, H. 1995. Fine structural investigation of Malassezia furfur. II. The envelope of the yeast cells. Mycoses 38:13-21.

Medline

74.↵ Moyrand, F., B. Klaproth, U. Himmelreich, F. Dromer, and G. Janbon. 2002. Isolation and characterization of capsule structure mutant strains of Cryptococcus neoformans. Mol. Microbiol. 45:837-849.

CrossRef Medline

75.↵ Murphy, J. W. 1988. Influence of cryptococcal antigens on cell-mediated immunity. Rev. Infect. Dis. 10(Suppl. 2):S432-S435.

Medline

76.↵ Murphy, J. W., R. L. Mosley, R. Cherniak, G. H. Reyes, T. R. Kozel, and E. Reiss. 1988. Serological, electrophoretic, and biological properties of Cryptococcus neoformans antigens. Infect. Immun. 56:424-431.

Abstract/FREE Full Text

77.Nosanchuk, J. D., B. L. Gómez, S. Youngchim, S. Díez, P. Aisen, R. M. Zancopé-Oliveira, A. Restrepo, A. Casadevall, and A. J. Hamilton. 2002. Histoplasma capsulatum synthesizes melanin-like pigments in vitro and during mammalian infection. Infect. Immun. 70:5124-5131.

Abstract/FREE Full Text

78.↵

Page 31: American Society for MicrobiologyEukaryotic Cell

Perfect, J. R., N. Ketabchi, G. M. Cox, C. W. Ingram, and C. L. Beiser. 1993. Karyotyping of Cryptococcus neoformans as an epidemiological tool. J. Clin. Microbiol. 31:3305-3309.

Abstract/FREE Full Text

79.↵ Petter, R., B. S. Kang, T. Boekhout, B. J. Davis, and K. J. Kwon-Chung. 2001. A survey of heterobasidiomycetous yeasts for the presence of the genes homologous to virulence factors of Filobasidiella neoformans, CNLAC1 and CAP59. Microbiology 147:2029-2036.

Abstract/FREE Full Text

80.↵ Pierini, L. M., and T. L. Doering. 2001. Spatial and temporal sequence of capsule construction in Cryptococcus neoformans. Mol. Microbiol. 41:105-115.

CrossRef Medline

81.↵ Pietrella, D., R. Cherniak, C. Strappini, S. Perito, P. Mosci, F. Bistoni, and A. Vecchiarelli. 2001. Role of mannoprotein in induction and regulation of immunity to Cryptococcus neoformans. Infect. Immun. 69:2808-2814.

Abstract/FREE Full Text

82.↵ Pietrella, D., R. Mazzolla, P. Lupo, L. Pitzurra, M. J. Gomez, R. Cherniak, and A. Vecchiarelli. 2002. Mannoprotein from Cryptococcus neoformans promotes T-helper type 1 anticandidal responses in mice. Infect. Immun. 70:6621-6627.

Abstract/FREE Full Text

83.↵

Page 32: American Society for MicrobiologyEukaryotic Cell

Reiss, E., M. Huppert, and R. Cherniak. 1985. Characterization of protein and mannan polysaccharide antigens of yeasts, moulds, and actinomycetes. Curr. Top. Med. Mycol. 1:172-207.

Medline

84.↵ Rivera, J., M. Feldmesser, M. Cammer, and A. Casadevall. 1998. Organ-dependent variation of capsule thickness in Cryptococcus neoformans during experimental murine infection. Infect. Immun. 66:5027-5030.

Abstract/FREE Full Text

85.↵ Roberts, I. S. 1996. The biochemistry and genetics of capsular polysaccharide production in bacteria. Annu. Rev. Microbiol. 50:285-315.

CrossRef Medline

86.↵ Rodrigues, M. L., C. S. Alviano, and L. R. Travassos. 1999. Pathogenicity of Cryptococcus neoformans: virulence factors and immunological mechanisms.Microbes Infect. 1:293-301.

CrossRef Medline

87.San-Blas, G., and F. San-Blas. 1977. Paracoccidioides brasiliensis: cell wall structure and virulence. Mycopathologia 62(2):77-86.

CrossRef Medline

88.↵ Schein, J. E., K. L. Tangen, R. Chiu, H. Shin, K. B. Lengeler, W. K. MacDonald, I. Bosdet, J. Heitman, S. J. Jones, M. A. Marra, and J. W. Kronstad. 2002. Physical maps for genome analysis of serotype A and D strains of the fungal pathogen Cryptococcus neoformans. Genome Res. 12:1445-1453.

Page 33: American Society for MicrobiologyEukaryotic Cell

Abstract/FREE Full Text

89.↵ Small, J. A., and T. G. Mitchell. 1986. Binding of purified and radioiodinated capsular polysaccharides from Cryptococcus neoformans serotype A strains to capsule-free mutants. Infect. Immun. 54:742-750.

Abstract/FREE Full Text

90.↵ Smits, G. J., J. C. Kapteyn, H. V. D. Ende, and F. M. Klis. 1999. Cell wall dynamics in yeast. Curr. Opin. Microbiol. 2:348-352.

CrossRef Medline

91.↵ Tacker, J. R., F. Farhi, and G. S. Bulmer. 1972. Intracellular fate of Cryptococcus neoformans. Infect. Immun. 6:162-167.

Abstract/FREE Full Text

92.↵ Taylor, S. S., J. A. Buechler, and W. Yonemoto. 1990. cAMP-dependent protein kinase: framework for a diverse family of regulatory enzymes. Annu. Rev. Biochem.59:971-1005.

CrossRef Medline

93.↵ Thianprasit, M., and K. Thangerngpol. 1989. Rhinosporidiosis. Curr. Top. Med. Mycol. 3:64-85.

CrossRef Medline

94.↵

Page 34: American Society for MicrobiologyEukaryotic Cell

Tolkacheva, T., P. McNamara, E. Piekarz, and W. Courchesne. 1994. Cloning of a Cryptococcus neoformans gene, GPA1, encoding a G-protein α-subunit homolog. Infect. Immun. 62:2849-2856.

Abstract/FREE Full Text

95.↵ Tucker, S. C., and A. Casadevall. 2002. Replication of Cryptococcus neoformans in macrophages is accompanied by phagosomal permeabilization and accumulation of vesicles containing polysaccharide in the cytoplasm. Proc. Natl. Acad. Sci. USA 99:3165-3170.

Abstract/FREE Full Text

96.↵ Vaishnav, V. V., B. E. Bacon, M. O'Neill, and R. Cherniak. 1998. Structural characterization of the galactoxylomannan of Cryptococcus neoformans Cap67.Carbohydr. Res. 306:315-330.

CrossRef Medline

97.↵ van de Moer, A., S. L. Salhi, R. Cherniak, B. Pau, M. L. Garrigues, and J. M. Bastide. 1990. An anti-Cryptococcus neoformans monoclonal antibody directed against galactoxylomannan. Res. Immunol. 141:33-42.

CrossRef Medline

98.↵ Vartivarian, S. E., E. J. Anaissie, R. E. Cowart, H. A. Sprigg, M. J. Tingler, and E. S. Jacobson. 1993. Regulation of cryptococcal capsular polysaccharide by iron. J. Infect. Dis. 167:186-190.

Abstract/FREE Full Text

99.↵

Page 35: American Society for MicrobiologyEukaryotic Cell

Vartivarian, S. E., G. H. Reyes, E. S. Jacobson, P. G. James, R. Cherniak, V. R. Mumaw, and M. J. Tingler. 1989. Localization of mannoprotein in Cryptococcus neoformans. J. Bacteriol. 171:6850-6852.

Abstract/FREE Full Text

100. ↵ Wang, P., M. E. Cardenas, G. M. Cox, J. R. Perfect, and J. Heitman. 2001. Two cyclophilin A homologs with shared and distinct functions important for growth and virulence of Cryptococcus neoformans. EMBO Rep. 2:511-518.

CrossRef Medline

101. ↵ Wang, P., C. B. Nichols, K. B. Lengeler, M. E. Cardenas, G. M. Cox, J. R. Perfect, and J. Heitman. 2002. Mating-type-specific and nonspecific PAK kinases play shared and divergent roles in Cryptococcus neoformans. Eukaryot. Cell 1:257-272.

Abstract/FREE Full Text

102. Wang, Y., P. Aisen, and A. Casadevall. 1996. Melanin, melanin “ghosts,” and melanin composition in Cryptococcus neoformans. Infect. Immun. 64:2420-2424.

Abstract/FREE Full Text

103. ↵ White, C. W., R. Cherniak, and E. S. Jacobson. 1990. Side group addition by xylosyltransferase and glucuronyltransferase in biosynthesis of capsular polysaccharide in Cryptococcus neoformans. J. Med. Vet. Mycol. 28:289-301.

Medline

104. ↵ White, C. W., and E. S. Jacobson. 1993. Mannosyl transfer in Cryptococcus neoformans. Can. J. Microbiol. 39:129-133.

Page 36: American Society for MicrobiologyEukaryotic Cell

Medline

105. ↵ Wickes, B. L., U. Edman, and J. C. Edman. 1997. The Cryptococcus neoformans STE12α gene: a putative Saccharomyces cerevisiae STE12 homologue that is mating type specific. Mol. Microbiol. 26:951-960.

CrossRef Medline

106. ↵ Wickes, B. L., M. E. Mayorga, U. Edman, and J. C. Edman. 1996. Dimorphism and haploid fruiting in Cryptococcus neoformans: association with the α-mating type. Proc. Natl. Acad. Sci. USA 93:7327-7331.

Abstract/FREE Full Text

107. ↵ Wills, E. A., I. S. Roberts, M. Del Poeta, J. Rivera, A. Casadevall, G. M. Cox, and J. R. Perfect. 2001. Identification and characterization of the Cryptococcus neoformans phosphomannose isomerase-encoding gene, MAN1, and its impact on pathogenicity. Mol. Microbiol. 40:610-620.

CrossRef Medline

108. ↵ Wilson, D. E., J. E. Bennett, and J. W. Bailey. 1968. Serologic grouping of Cryptococcus neoformans. Proc. Soc. Exp. Biol. Med. 127:820-823.

Abstract/FREE Full Text

109. ↵ Wilson, J., M. Schurr, C. LeBlanc, R. Ramamurthy, K. Buchanan, and C. Nickerson. 2002. Mechanisms of bacterial pathogenicity. Postgrad. Med. J. 78:216-224.

Abstract/FREE Full Text

Page 37: American Society for MicrobiologyEukaryotic Cell

110. ↵ Yue, C., L. M. Cavallo, J. A. Alspaugh, P. Wang, G. M. Cox, J. R. Perfect, and J. Heitman. 1999. The STE12α homolog is required for haploid filamentation but largely dispensable for mating and virulence in Cryptococcus neoformans. Genetics 153:1601-1615.

Abstract/FREE Full Text

American Society for Microbiology

CiteULike

Delicious

Digg

Facebook Google+

Mendeley

Reddit

StumbleUpon

Twitter

What's this?

Articles citing this article

Influence of capsule size on the in vitro activity of antifungal agents against clinical Cryptococcus neoformans var. grubii strains J Med Microbiol March 2012 61:3 384-388

Page 38: American Society for MicrobiologyEukaryotic Cell

o Abstract o Full Text o PDF

Galactoxylomannans from Cryptococcus neoformans Varieties neoformans and grubii Are Structurally and Antigenically Variable Eukaryot Cell July 2010 9:7 1018-1028

o Abstract o Full Text o PDF

A Putative P-Type ATPase, Apt1, Is Involved in Stress Tolerance and Virulence in Cryptococcus neoformans Eukaryot Cell January 2010 9:1 74-83

o Abstract o Full Text o PDF

Lipophilic Dye Staining of Cryptococcus neoformans Extracellular Vesicles and Capsule Eukaryot Cell September 2009 8:9 1373-1380

o Abstract o Full Text o PDF

Production of Extracellular Polysaccharides by CAP Mutants of Cryptococcus neoformans Eukaryot Cell August 2009 8:8 1165-1173

o Abstract o Full Text o PDF

Elucidating the Pathogenesis of Spores from the Human Fungal Pathogen Cryptococcus neoformans Infect. Immun. August 2009 77:8 3491-3500

o Abstract o Full Text o PDF

Capsular Localization of the Cryptococcus neoformans Polysaccharide Component Galactoxylomannan Eukaryot Cell January 2009 8:1 96-103

o Abstract o Full Text o PDF

UGE1 and UGE2 Regulate the UDP-Glucose/UDP-Galactose Equilibrium in Cryptococcus neoformans Eukaryot Cell December 2008 7:12 2069-2077

o Abstract o Full Text o PDF

Page 39: American Society for MicrobiologyEukaryotic Cell

Use of Gene Dosage Effects for a Whole-Genome Screen To Identify Mycobacterium marinum Macrophage Infection Loci Infect. Immun. July 2008 76:7 3100-3115

o Abstract o Full Text o PDF

Cryptococcal Xylosyltransferase 1 (Cxt1p) from Cryptococcus neoformans Plays a Direct Role in the Synthesis of Capsule Polysaccharides J Biol Chem May 2008 283:2114327-14334

o Abstract o Full Text o PDF

Binding of the Wheat Germ Lectin to Cryptococcus neoformans Suggests an Association of Chitinlike Structures with Yeast Budding and Capsular GlucuronoxylomannanEukaryot Cell April 2008 7:4 602-609

o Abstract o Full Text o PDF

Self-Aggregation of Cryptococcus neoformans Capsular Glucuronoxylomannan Is Dependent on Divalent Cations Eukaryot Cell August 2007 6:8 1400-1410

o Abstract o Full Text o PDF

A beta-1,2-Xylosyltransferase from Cryptococcus neoformans Defines a New Family of Glycosyltransferases J Biol Chem June 2007 282:24 17890-17899

o Abstract o Full Text o PDF

Chitosan, the Deacetylated Form of Chitin, Is Necessary for Cell Wall Integrity in Cryptococcus neoformans Eukaryot Cell May 2007 6:5 855-867

o Abstract o Full Text o PDF

The Pathogenic Fungus Cryptococcus neoformans Expresses Two Functional GDP-Mannose Transporters with Distinct Expression Patterns and Roles in Capsule SynthesisEukaryot Cell May 2007 6:5 776-785

o Abstract o Full Text o PDF

Page 40: American Society for MicrobiologyEukaryotic Cell

Galactoxylomannan Does Not Exhibit Cross-Reactivity in the Platelia Aspergillus Enzyme Immunoassay CVI May 2007 14:5 624-627

o Abstract o Full Text o PDF

Parallel {beta}-Helix Proteins Required for Accurate Capsule Polysaccharide Synthesis and Virulence in the Yeast Cryptococcus neoformans Eukaryot Cell April 2007 6:4630-640

o Abstract o Full Text o PDF

The iron- and cAMP-regulated gene SIT1 influences ferrioxamine B utilization, melanization and cell wall structure in Cryptococcus neoformans Microbiology January 2007153:1 29-41

o Abstract o Full Text o PDF

Radial Mass Density, Charge, and Epitope Distribution in the Cryptococcus neoformans Capsule Eukaryot Cell January 2007 6:1 95-109

o Abstract o Full Text o PDF

A Eukaryotic Capsular Polysaccharide Is Synthesized Intracellularly and Secreted via Exocytosis Mol. Biol. Cell December 2006 17:12 5131-5140

o Abstract o Full Text o PDF

CPS1, a Homolog of the Streptococcus pneumoniae Type 3 Polysaccharide Synthase Gene, Is Important for the Pathobiology of Cryptococcus neoformans Infect. Immun.July 2006 74:7 3930-3938

o Abstract o Full Text o PDF

Cryptococcus neoformans Senses CO2 through the Carbonic Anhydrase Can2 and the Adenylyl Cyclase Cac1 Eukaryot Cell January 2006 5:1 103-111

o Abstract o Full Text o PDF

Baseline Correlation and Comparative Kinetics of Cerebrospinal Fluid Colony-Forming Unit Counts and Antigen Titers in Cryptococcal Meningitis The Journal of Infectious Disease August 2005 192:4 681-684

Page 41: American Society for MicrobiologyEukaryotic Cell

o Abstract o Full Text o PDF

Galleria mellonella as a Model System To Study Cryptococcus neoformans Pathogenesis Infect. Immun. July 2005 73:7 3842-3850

o Abstract o Full Text o PDF

Cryptococcus neoformans Galactoxylomannan Contains an Epitope(s) That Is Cross-Reactive with Aspergillus Galactomannan J. Clin. Microbiol. June 2005 43:6 2929-2931

o Abstract o Full Text o PDF

The Genome of the Basidiomycetous Yeast and Human Pathogen Cryptococcus neoformans Science February 2005 307:5713 1321-1324

o Abstract o Full Text o PDF

Cas3p Belongs to a Seven-Member Family of Capsule Structure Designer Proteins Eukaryot Cell December 2004 3:6 1513-1524

o Abstract o Full Text o PDF

UGD1, Encoding the Cryptococcus neoformans UDP-Glucose Dehydrogenase, Is Essential for Growth at 37{degrees}C and for Capsule Biosynthesis Eukaryot CellDecember 2004 3:6 1601-1608

o Abstract o Full Text o PDF

Role of Phagocytosis in the Virulence of Cryptococcus neoformans Eukaryot Cell October 2004 3:5 1067-1075

o Full Text o PDF

The Glycan-Rich Outer Layer of the Cell Wall of Mycobacterium tuberculosis Acts as an Antiphagocytic Capsule Limiting the Association of the Bacterium with MacrophagesInfect. Immun. October 2004 72:10 5676-5686

o Abstract o Full Text o PDF

Page 42: American Society for MicrobiologyEukaryotic Cell

A Genetic Linkage Map of Cryptococcus neoformans variety neoformans Serotype D (Filobasidiella neoformans) Genetics June 2004 167:2 619-631

o Abstract o Full Text o PDF

Cryptococcus neoformans Capsule Structure Evolution In Vitro and during Murine Infection Infect. Immun. June 2004 72:6 3359-3365

o Abstract o Full Text o PDF

Cryptococcus neoformans CAP59 (or Cap59p) Is Involved in the Extracellular Trafficking of Capsular Glucuronoxylomannan Eukaryot Cell April 2004 3:2 385-392

o Abstract o Full Text o PDF

Surface Glycans of Candida albicans and Other Pathogenic Fungi: Physiological Roles, Clinical Uses, and Experimental Challenges Clin. Microbiol. Rev. April 2004 17:2 281-310

o Abstract o Full Text o PDF

Lessons from the Genome Sequence of Neurospora crassa: Tracing the Path from Genomic Blueprint to Multicellular Organism Microbiol. Mol. Biol. Rev. March 2004 68:11-108

o Abstract o Full Text o PDF

Cryptococcus neoformans Gene Expression during Experimental Cryptococcal Meningitis Eukaryot Cell December 2003 2:6 1336-1349

o Abstract o Full Text o PDF

An {alpha}-1,3-Mannosyltransferase of Cryptococcus neoformans J Biol Chem November 2003 278:48 47724-47730

o Abstract o Full Text o PDF

« Previous | Next Article » Table of Contents

Page 43: American Society for MicrobiologyEukaryotic Cell

This Article

1. doi: 10.1128/EC.2.4.655-663.2003 Eukaryotic Cell August 2003 vol. 2 no. 4 655-663

1. Figures 2. » Full Text3. PDF

- Classifications

1.o MINIREVIEW

- Services

1. Email this article to a colleague 2. Similar articles in ASM journals 3. Alert me when this article is cited 4. Alert me if a correction is posted 5. Similar articles in this journal 6. No Web of Science related articles7. Similar articles in PubMed 8. Alert me to new issues of  EC9. Download to citation manager 10.Reprints and Permissions 11.Copyright Information 12.Books from ASM Press 13.MicrobeWorld

+ Citing Articles

+ Google Scholar

+ PubMed

+ Related Content

+ Social Bookmarking

Navigate This Article

1. Top

Page 44: American Society for MicrobiologyEukaryotic Cell

2. CAPSULE STRUCTURE 3. MUTANTS IN CAPSULE SYNTHESIS 4. STEPS REQUIRED FOR CAPSULE SYNTHESIS 5. CAPSULE ASSEMBLY AND DEGRADATION 6. REGULATION OF THE CAPSULE 7. OTHER CAPSULES 8. FUTURE DIRECTIONS 9. ACKNOWLEDGMENTS 10.FOOTNOTES 11.REFERENCES

current issue

September 2013, volume 12, issue 9

1.

- Spotlights in the Current Issue

Sequencing and Characterization of the Transcriptome of Penicillium marneffei Mediator Stabilizes the Active and Silent States of an Epigenetic Switch in a

Fungal Pathogen CMF22 Is a Broadly Conserved Core Component of Motile Flagella

1. Alert me to new issues of  EC

About  EC Subscribers Authors Reviewers Advertisers Inquiries from the Press Permissions & Commercial Reprints ASM Journals Public Access Policy

EC  RSS Feeds 

Page 45: American Society for MicrobiologyEukaryotic Cell

 

1752 N Street N.W. • Washington DC 20036

202.737.3600 • 202.942.9355 fax • [email protected]

Print ISSN: 1535-9778 Online ISSN: 1535-9786

Copyright © 2013 by the American Society for Microbiology.      For an alternate route to EC.asm.org, visit: http://intl-EC.asm.org  |    More Info»