Skip to main content
Eukaryotic Cell logoLink to Eukaryotic Cell
. 2010 Jun;9(6):835–846. doi: 10.1128/EC.00039-10

Cryptococcal Interactions with the Host Immune System

Kerstin Voelz 1, Robin C May 1,*
PMCID: PMC2901644  PMID: 20382758

Abstract

Opportunistic pathogens have become of increasing medical importance over the last decade due to the AIDS pandemic. Not only is cryptococcosis the fourth-most-common fatal infectious disease in sub-Saharan Africa, but also Cryptococcus is an emerging pathogen of immunocompetent individuals. The interaction between Cryptococcus and the host's immune system is a major determinant for the outcome of disease. Despite initial infection in early childhood with Cryptococcus neoformans and frequent exposure to C. neoformans within the environment, immunocompetent individuals are generally able to contain the fungus or maintain the yeast in a latent state. However, immune deficiencies lead to disseminating infections that are uniformly fatal without rapid clinical intervention. This review will discuss the innate and adaptive immune responses to Cryptococcus and cryptococcal strategies to evade the host's defense mechanisms. It will also address the importance of these strategies in pathogenesis and the potential of immunotherapy in cryptococcosis treatment.


The basidiomycetous yeast genus Cryptococcus includes the two medically important pathogens C. neoformans and C. gattii. These two species are further divided into C. neoformans serotypes A (C. neoformans var. grubii), D (C. neoformans var. neoformans), and A/D and C. gattii serotypes B and C (formerly C. neoformans var. gattii) based on differential antibody recognition of the polysaccharide capsule (135). The two pathogenic species show different geographical distributions. C. neoformans is globally distributed and has been isolated from various natural sources, with particularly high concentrations occurring in avian guano, rotting vegetables, and soil. In contrast, C. gattii is geographically restricted to tropical and subtropical regions, with the notable exception of British Columbia. In tropical and subtropical regions, it has been found to be associated with the eucalyptus species Eucalyptus camaldulensis, Eucalyptus tereticornis, Eucalyptus rudis, and Eucalyptus gomphocephala (64, 172). C. neoformans causes mainly opportunistic infections in immunocompromised patients with underlying conditions, such as HIV, leukemia, and other cancers, or in those taking corticosteroid medication (135). Serotype A is responsible for the majority of cryptococcosis cases in immunocompromised hosts (135). In contrast, C. gattii affects mainly immunocompetent individuals. The recent and spreading cryptococcosis outbreak in healthy individuals in British Columbia has highlighted the potential of C. gattii to act as an emerging pathogen (84, 85, 121). In addition, other non-C. neoformans/non-C. gattii species, such as Cryptococcus laurentii and Cryptococcus albidus, have recently started to emerge as potential human pathogens (83).

Cryptococcal infection can be asymptomatic, chronic, or acute. Typically, an initial pulmonary infection can spread systemically, with a particular predilection for the central nervous system. Pulmonary infections are in most cases asymptomatic. However, they can involve coughing, pleuritic chest pain, fever, dyspnoea, weight loss, and malaise. Pneumonia and acute respiratory distress syndrome have been reported mainly for immunocompromised patients (17, 141). Cryptococcosis of the central nervous system is life threatening and presents as meningitis or meningoencephalitis, with symptoms such as headache, increased intracranial pressure, fever, lethargy, coma, personality changes, and memory loss. Less common are secondary infections of the skin, lungs, prostate, and eye (135). A recent publication estimated 957,900 cases of cryptococcal meningitis resulting in 624,700 deaths globally each year (150). It is the leading cause of death in HIV-infected individuals, with an incidence of 30% and a mortality of 30 to 60%. The mortality rate in transplant patients is even higher (20 to 100%) (Centers for Disease Control and Prevention) (135).

The dramatic course of Cryptococcus infections in immunocompromised individuals shows the importance of an intact immune response to the pathogen. This review will consider both the host's innate and adaptive immune responses to C. neoformans and C. gattii together with the pathogens' strategy to undermine these defense mechanisms and how current knowledge might be applied to improve anticryptococcal therapy.

INNATE IMMUNE RESPONSE TO CRYPTOCOCCUS

A variety of innate factors interfere with the establishment of cryptococcal infection. Besides physical barriers, such as the skin and the nasal mucosa, the anticryptococcal activity of human serum and saliva has been described repeatedly (810, 65, 74, 143, 181). However, the complement system and phagocytic effector cells are the major players in the nonspecific host immune response to Cryptococcus.

Complement response to Cryptococcus.

The complement system is an antipathogen cascade of serum proteins that can be activated by the classical (antibody-mediated), lectin, or alternative (microbial-surface-mediated) pathway. All three pathways eventually converge in the formation of the C3-convertase and the cleavage of C3 into C3a and C3b. C3b either facilitates pathogen opsonization and thus enhances uptake by phagocytic cells or enables cleavage of C5 into C5a and C5b. C5a functions, together with C3a, as a mediator of inflammatory responses and attracts phagocytic effector cells, whereas C5b initiates the formation of the membrane attack complex (C5b, C6, C7, C8, C9) (76).

Observations from animal model systems and human patients have repeatedly shown the importance of the complement system during cryptococcal infections (Fig. 1). The survival time of C. neoformans-infected guinea pigs and mice treated with cobra venom to deplete late complement components (C3 to C9) is shortened and the ability to clear C. neoformans from extraneural sites reduced (37). Mice deficient in C5 are more susceptible to intravenously injected C. neoformans and succumb three times quicker than C5-positive mice due to acute and fatal pneumonia (158, 159). Furthermore, patients presenting with cryptococcal fungemia show reduced levels of C3 and alternative complement factor B (122). Brain sections from patients with cryptococcal meningitis do not show C3 binding to the yeast (186). In contrast, however, the survival time of C4-deficient guinea pigs is similar to that of normal guinea pigs after infection with C. neoformans, indicating that the alternative activation pathway is the major protective complement pathway during infections with C. neoformans (37).

Fig. 1.

Fig. 1.

Summary of the complement pathways activated upon infection with Cryptococcus neoformans. The yeast can activate the classical (antibody-mediated) and alternative (microbial-surface-mediated) pathways. Both pathways eventually converge in the formation of the C3-convertase and the cleavage of C3 into C3a and C3b. C3b either facilitates pathogen opsonization and enhances uptake by phagocytic cells or enables the cleavage of C5 into C5a and C5b. C5a functions as a mediator of inflammatory responses and attracts phagocytic effector cells, whereas C5b initiates the formation of the membrane attack complex (C5b, C6, C7, C8, C9). However, C. neoformans is resistant to pore formation and cell lysis by the membrane attack complex (165). fB, factor B; fD, factor D.

This finding is supported by results from in vitro analysis of the complement binding dynamics of C. neoformans. Diamond et al. (38) reported the consumption of complement components by C. neoformans when it was incubated with normal or C4-deficient guinea pig serum, while C. neoformans-dependent activation of the complement cascade can be reconstituted from six proteins (factor D, factor B, factor H, factor I, C3, and properdin) belonging to the alternative pathway (96). It was estimated that approximately 107 to 108 C3 fragments bind to an encapsulated cryptococcal cell and, dependent on the source of serum, localize inside and at the outer edge of the capsule (56, 88, 93, 203). The binding of C3 starts characteristically with a lag phase of 4 to 6 min, followed by rapid binding of C3 fragments to encapsulated yeast cells when they are incubated in normal human serum (95). Incubation with Mg-EGTA to chelate calcium required for classical pathway activation did not change the C3 binding dynamics, supporting the idea of a dominant role for the alternative complement pathway during cryptococcosis (95). According to the complement model, the activation of the alternative pathway relies on the spontaneous decomposition of C3 to C3b and Bb upon pathogen interaction. Similarly, closer examination of the binding process of C3 to yeast cells by immunofluorescence revealed an initial slow C3 deposition in small loci with subsequent expansion to larger areas after a lag (95). Investigation into the molecular form of bound C3 by SDS-PAGE of eluted radioactively labeled fragments revealed a rapid decay of the C3 hydrolysis product C3b into iC3b, which in turn is the dominant fragment bound to the cryptococcal capsule (92, 153).

The two major functions of the complement system during cryptococcosis are to stimulate the chemotaxis of phagocytic effector cells and enhance the uptake of cryptococcal cells by these phagocytes. Early evidence for the involvement of complement in the opsonization of C. neoformans was drawn from phagocytosis assays with heat-inactivated serum (39). Assays with phagocytic cells and serum, depleted of specific components of the complement pathways, revealed the requirement of the complement pathway for phagocytosis of cryptococci by neutrophils (38), polymorphonuclear leukocytes, and monocytes in an antibody-free situation (30). However, although the alternative pathway is sufficient for yeast opsonization, the classical pathway is required for optimal opsonization kinetics (38). Laxalt and Kozel (100) and Diamond and Erickson (36) observed the chemotactic potential of serum; both serum-opsonized encapsulated and nonencapsulated C. neoformans cells are able to chemotactically attract neutrophils and monocytes in vitro. As C5-deficient mice are more susceptible to cryptococcosis (158, 159) and closer investigations revealed a lack of neutrophil accumulation in pulmonary vessels, C5a, the C5 cleavage product, seems to be the chemotactic active component (Fig. 1) (113).

The cryptococcal polysaccharide capsule is a well-known factor required for the pathogen's virulence, e.g., by inhibiting phagocytosis (reviewed in reference 202). C. neoformans mutants with a capsule-deficient phenotype are avirulent in mice (19, 98). Several studies with encapsulated and nonencapsulated C. neoformans strains also showed a difference in complement activation dependent on capsulation. The capsule inhibits the binding of mannan-binding lectin and thus the activation of the complement system via the lectin pathway in C. neoformans (149). The total numbers of bound C3 molecules are similar in different cryptococcal strains, independent of the capsule size (93). However, a comparison of the depositions of C3 on C. neoformans strain 145 grown under capsule-inducing and non-capsule-inducing conditions indicates a relationship between capsule diameter and C3 density under high yeast cell concentrations. The small-capsule variant bound more C3 molecules per cubic micrometer of capsule than the population with a large capsule (94). In contrast, the noncapsular strain 602 accumulates significantly less C3 on its cellular surface than capsulated strains (88, 95, 97). The decay rate of C3b to iC3b is lower in nonencapsulated C. neoformans than in capsulated forms (70% versus 100%, respectively, after 20 min of incubation with factors H and I) (92, 153). Furthermore, the acapsular strain 602 activates not only the alternative but also the classical complement pathway (95). C3 binding to the acapsular strain occurs immediately, without any lag phase, and rather than the characteristic small C3 deposition sites in encapsulated strains, sudden and rapid binding of C3 to the entire cell surface can be observed. This accumulation is ongoing and does not stop after 8 min, as seen in encapsulated C. neoformans. However, the dynamics described for capsular strains can be reinduced by treatment with Mg-EGTA (95). Interestingly, there seems to be a species-specific difference in activation of the complement system. Although the binding efficiency of C3 to C. gattii serotype B and C strains is higher than to C. neoformans serotype A and D strains (161), the total accumulation of C3 on the cellular surface of C. gattii serotype B and C strains is only half of that of the latter (194, 198).

The complement system is the first line of defense against Cryptococcus in the bloodstream and, by opsonizing the pathogen and attracting immune effector cells, performs important preparations for the subsequent host defense response. The capsule probably functions as an inhibitor of complement-related host responses, such as uptake by phagocytosis, by inhibiting the classical pathway and constricting C3-convertase activity by the efficient removal of C3b, an essential part of the alternative C3-convertase, and thus restricting activity amplification. In addition, the differences in complement activation between C. neoformans and C. gattii might be important for the increased virulence of certain serotype B strains. Recent data from survival assays of factor B- or C3-deficient mice after C. gattii infection have suggested that complement pathways other than the alternative activation pathway contribute to the host's protection (129).

Phagocytic effector cells in the host's immune response to Cryptococcus.

Research over the last few decades has shown the importance of phagocytic effector cells during a host's immune response to cryptococcal infections. This section will describe how the yeast cells are taken up and discuss the interaction with the different types of phagocytes. A particular focus will be given to the interaction between cryptococcal cells and macrophages.

(i) Phagocytosis.

Uptake of cryptococcal cells has been shown repeatedly by a variety of leukocytes (15), such as rat and mouse peritoneal macrophages (62, 89, 134), guinea pig pulmonary macrophages (16), human neutrophils and macrophages (40), and swine microglia (110). Phagocytosis is triggered by direct recognition of the yeast or by receptor-mediated recognition via complement or antibodies (135). Conserved structures, such as the components of the cryptococcal capsule, can be directly recognized by pattern recognition receptors. C. neoformans glucuronoxylomannan (GXM) can bind to Toll-like receptor 4 (167), and the mannose receptor of dendritic cells (DCs) binds to mannoproteins expressed on the yeast's surface (154) and, in the absence of complement to CD18, to a subunit of the complement receptors (40). Serum-opsonized (i3Cb) C. neoformans is recognized either by the complement receptor CR1 (CD35) or by the heteromeric β2-integrins CR3 (CD11b/CD18) and CR4 (CD11c/CD35) (108, 203). Studies with the receptors heterologously expressed in Chinese hamster ovary cells indicate that binding to any of the receptors occurs independently, with the greatest binding avidity being shown for CR3, followed by CR1 and CR4 (109). Antibody-opsonized yeast cells are recognized by Fcγ receptor molecules expressed on the surfaces of macrophages, neutrophils, and dendritic cells (62, 115, 136, 180).

(ii) Dendritic cells.

Cryptococcus neoformans is phagocytosed by human primary DCs in vitro (82, 180) and also in the mouse model (197). Dendritic cells function as major antigen-presenting cells that constantly monitor the current antigen population and modulate adaptive immune responses accordingly (28). During cryptococcal infections, DCs are thought to be the major initiators of protective cell-mediated immunity (11, 147). Not only are major antigens (e.g., mannoproteins and glycoantigens) for the activation of anticryptococcal T-cell responses dominantly presented by DCs (107, 125), but also the induction of T-cell responses by DCs is much more efficient than by alveolar or peritoneal macrophages (125, 180).

(iii) Neutrophils.

Neutrophils are thought to contribute strongly to the innate immune response to cryptococcosis. Upon cryptococcal challenge, the number of polymorphonuclear cells increases at the site of infection in animal models (54, 152). In vitro, the oxidative burst exerted by neutrophils effectively kills C. neoformans (21, 124, 131). However, despite rapid antimicrobial activity in vivo, cryptococci are only partially cleared from the infected site (54, 152). Interestingly, induced neutropenia in mice increases survival after pulmonary challenge with C. neoformans (128), a counterintuitive finding that may be due to the absence of neutrophils, resulting in more interleukin 4 (IL-4) and IL-10 signaling, which, in turn, modifies the inflammatory status and reduces tissue damage due to the aggressive oxidative burst (128). Together with data showing neutrophils being present in infected tissues only in low numbers and at early stages of infection (46), this might suggest a more immune-regulatory than antimicrobial role for neutrophils.

Besides killing pathogens by means of respiratory bursts, neutrophils also produce antimicrobial peptides and proteins as part of the antimicrobial response (101). One such family of antimicrobial peptides—the defensins—is found in abundance in human, rat, rabbit, and guinea pig neutrophils; however, expression is lacking in mouse neutrophils (43). As mice are routinely used as a system to model cryptococcal disease, this finding needs to be considered when assessing the role of neutrophils in the anticryptococcal immune defense. The lack of defensins might also be a reason for the high susceptibly of mice to challenge with the yeast.

(iv) Macrophages.

The importance of macrophages in cryptococcal infections has become more and more obvious in the last decade. Research has revealed an intriguing interaction between the phagocytic effectors and yeast cells that revealed C. neoformans as an intracellular parasite (48). Cryptococcus has developed a unique method to manipulate host macrophages. After phagocytosis, C. neoformans can survive and proliferate within these infected host cells, eventually leading to host cell lysis (5, 33, 46, 48, 118, 187). Intracellular proliferation occurs despite the harsh environment within macrophages, and the yeast does not seem to utilize strategies that are known from other intracellular pathogens to manipulate the host cell. In contrast to pathogens such as Listeria monocytogenes or Shigella flexneri, C. neoformans can reside in the phagosome and does not have to escape into the cytoplasm to establish the intracellular niche (106, 168, 174). Moreover, C. neoformans does not inhibit phagosome-lysosome fusion (106), as has been shown for Legionella pneumophila (70), nor does the yeast interfere with phagosome maturation or acidification, as occurs during infections with Histoplasma capsulatum or Mycobacterium species (106, 168, 178, 179). Instead, C. neoformans survives and replicates in the acidic phagolysosome, and in fact, any increase in phagosomal pH (e.g., by experimental addition of chloroquine or ammonium chloride) leads to reduced intracellular proliferation (105). Host cell lysis is a common escape route for intracellular pathogens. To date, there is not much information on the mechanisms of cryptococcal lytic escape (201). However, given its documented role as a virulence factor (25), phospholipase B (plb1) is a potential candidate molecule that may mediate the permeabilization of C. neoformans-containing phagosomes (47, 187).

Besides lytic escape, a novel expulsive mechanism (Fig. 2) by which the yeast can exit macrophages without killing the host cell, thus avoiding a local inflammatory response, has recently been described (4, 117). In contrast to other expulsive mechanisms, such as the actin-based protrusion shown by Listera monocytogenes, Rickettsia spp., Shigella flexneria, and Burkholderia pseudomallei (177), cryptococcal expulsion occurs without obvious involvement of the actin cytoskeleton or damage to either the pathogen or the host (4, 117). The process is dependent on live cryptococci and occurs very rapidly, requiring less than 60 s, with events being randomly distributed over time (4, 6, 117). Expulsion events are independent of the route of uptake (117), although antibody-opsonized yeast cells tend to be expelled as a clump of cells that subsequently continue to replicate as a biofilm (6), whereas complement-opsonized yeast cells are released individually (6). In addition to having the ability to extrude from host cells, Cryptococcus can be laterally transferred from one macrophage to another (5, 118), an event which, like expulsion, does not occur with heat-killed cryptococci or latex beads (5) and is independent of the route of uptake or cryptococcal strain (118). In contrast to extrusion, however, lateral transfer is an actin-dependent process that can be inhibited by treatment with the actin depolymerization drug cytochalasin D (118).

Fig. 2.

Fig. 2.

Cryptococcal expulsion from within a macrophage. Cryptococcus can exit macrophages in a novel nonlytic way that does not involve the killing of the host cell or the yeast. (A to C) Time-lapse images of two intracellular yeasts within a macrophage. (D) Four hours into the experiment, the yeast is suddenly expelled. (E to F) Both the macrophage and the yeast remain alive after this process, as shown by continuing proliferation.

Results from restriction fragment length polymorphism analysis suggest that initial infection with C. neoformans often occurs in early childhood and can be followed by a long latent phase (55, 58). However, C. neoformans is generally capable of dissemination to other organs within the human body and shows a predilection for the central nervous system (73). The interaction between Cryptococcus and macrophages might be the key to explaining not only how cryptococcal infection remains latent but also how dissemination within the host is achieved. The intracellular environment is clearly beneficial to the pathogen, as it offers protection from the immune system and thus undisturbed proliferation. However, the purpose of expulsion might be more subtle. Although expulsion from macrophages subjects the yeast cells to a greater immune attack within the extracellular environment, after sufficient intracellular replication it might also lead to fungemia and thus a general breakdown in host immunity. In addition, the so-called “Trojan horse” model suggests that replication within, lateral transfer between, and eventual expulsion from macrophages might offer a potential explanation for how C. neoformans stays latent and spreads within the host without triggering immediate immune responses (20, 22, 162). Expulsion and lateral transfer, in particular, might be involved in allowing the yeast to cross the blood-brain barrier either by using macrophages as a trafficking vehicle or by becoming directly transferred to the endothelial layer of the barrier and then expelled into the central nervous system.

Taken together, these possibilities raise the issue of whether macrophages exert a beneficial or deleterious effect during infection. In this context, two recent studies have demonstrated that the absence of macrophages or monocytes was associated with prolonged survival (20, 81). These finding suggest that in certain situations, macrophages are in fact responsible for the development of the disease. In addition, differences in the activities of macrophages have been correlated with the differential susceptibilities of different hosts to infection, highlighting the importance of macrophages in determining the outcome of the disease (164, 200).

ADAPTIVE IMMUNE RESPONSE TO CRYPTOCOCCUS

The development of an adaptive immune response is essential to overcoming cryptococcal infection. This section will concentrate on the antibody and cell-mediated immune responses to Cryptococcus.

Antibody-mediated immune response to Cryptococcus.

With Cryptococcus being a facultative intracellular pathogen, there is some controversy about the importance of antibody-mediated immune mechanisms for effective microbial clearance. Several cases of cryptococcosis have been reported in patients with primary or acquired B-cell, antibody, or lymphoproliferative deficiencies (18), and antibodies against cryptococcal proteins and capsular polysaccharides are routinely found in individuals without an apparent infection (1, 51, 71), indicating either latent or asymptomatic cleared infections. Passive administration of capsule-binding antibody prolongs survival and/or reduces fungal burdens in experimental cryptococcosis (52, 139). Anticryptococcal antibodies elicit their protective function by opsonizing pathogens for Fc receptor-dependent phagocytosis and by activating the classical complement pathway. In addition, direct antibody opsonization of cryptococci, leading to complement-independent (but complement receptor-dependent via CD18) uptake into macrophages, has been described (145, 183). The specificity of antibodies seems to be of great importance for protective efficacy. Generally, it is thought that the two domains of antibodies fulfill two different functions: the variable region is responsible for antigen binding and the constant region for the effector functions. In studies with C. neoformans, it has been shown that this classical view might have to be rethought, since antibodies identical in their variable regions but differing in their constant regions show diverse binding affinity and specificity to a univalent peptide antigen (29, 184, 185). In addition, antibody subgroups developed from the same B cell can be either protective or nonprotective according to their staining patterns (annular or punctuate). Although derived from the same B cell, the two subtypes recognize spatially distinct areas of the cryptococcal capsule, and only a punctuate opsonization pattern can induce a protective immune response in mice. The preponderance of protective compared to nonprotective antibodies determines the efficacy of the antibody-mediated response to C. neoformans infection. It appears that small differences in antibody epitope specificities may have a large impact on the protective effect afforded by anticryptococcal antibodies. Those that are protective show a punctuate distribution within the capsule, and those with an annular pattern of capsular distribution are nonprotective (138). Furthermore, enhancement of cryptococcal disease has been described due to excess antibodies triggering immune unresponsiveness (90, 163). The concentration of antibody plays an important role for its protective functions. In mouse models, administration of antibody in higher concentrations can be less effective than in lower concentrations when mice are subsequently challenged with C. neoformans. This so-called prozone-like activity suggests that antibodies during cryptococcosis can be protective, nonprotective, or even disease enhancing depending on the antibody isotype and dose (182, 183).

Evidence from mouse models suggests that the protective effect of antibodies is at least partly due to an interaction with cell-mediated immunity. Mice defective in CD4, gamma interferon (IFN-γ), and Th1/Th2-associated cytokines cannot be protected by passive administration of IgG1 antibodies, whereas mice deficient in CD8, natural killer (NK) cells, or complement factor C3 can (12, 165, 199). Thus, antibody-mediated immunity is a significant part of the host defense that integrates into a complex network of different elements of protective anticryptococcal immunity.

Cell-mediated immune response to Cryptococcus.

The number of cryptococcosis cases drastically increased with the onset of the AIDS pandemic, and to date the highest incidence is still found in HIV-stricken sub-Saharan Africa (150). Besides HIV patients, individuals with extensive corticosteroid treatment, organ transplantation, leukemia, or lymphoma and sarcoidosis belong to a group at high risk for cryptococcal infections (135). The common feature of all of these predispositions is a defect in cell-mediated immunity (CMI). This part of the host defense contributes to cryptococcal killing either directly by cytotoxic effects or indirectly by regulatory functions of natural killer cells or T lymphocytes. NK, CD4+, and CD8+ cells all exhibit direct antimicrobial activity to C. neoformans (103, 206), and the secreted proteins granulysin and perforin are able to induce both cryptococcal permeabilization and lysis (44, 191). Although NK cells express both proteins, perforin is the main mediator of anticryptococcal killing via a PI3K-dependent ERK1/2 signaling pathway (120, 126, 195). In contrast, CD4+ and CD8+ lymphocytes utilize the anticryptococcal function of granulysin that is expressed upon activation of STAT5 and PI3K in the presence of IL-2/IL-15 and IL-15, respectively (119, 205, 206). In HIV patients, these two pathways are defective, resulting in inefficient killing of C. neoformans (206).

The regulatory arm of CMI seems to be an even more important part of fungal clearance than the ability to directly lyse cryptococci. The outcome of cryptococcosis depends on the immune status of the infected individual and the expression of host cytokines generated in response to the pathogen. Both Th1 and Th2 cytokines are involved in protection against C. neoformans, but whereas Th1-associated cytokines are essential for natural immunity, Th2-associated immunity is not protective in mice (12, 68, 72). Increased expression of Th1 cytokines, such as tumor necrosis factor alpha (TNF-α) and IFN-γ, results in improved fungal control (50, 80, 130, 196), while IFN-γ knockout mice show an increased fungal burden (7). In addition, Müller et al. have recently demonstrated a significant role for IL-17 and the associated Th17 response in modulating the survival of C. neoformans-infected mice (86, 140). In contrast, Th2 cytokines, such as IL-4 and IL-13, reduce the host's ability to deal with C. neoformans in vivo (13, 31, 78, 140). The incidence of cryptococcosis increases throughout the course of HIV infection and correlates with the loss of the Th1 response in HIV-infected patients (3) and with a Th2 cytokine profile in transplant patients (171). Thus, the Th1-Th2-Th17 balance is essential for the survival of cryptococci.

In addition to DCs and neutrophils, primary lymphocytes, NK cells, and γδ T cells are involved in the maintenance of this cytokine balance during infection (128, 142, 188, 197). NK cells produce high concentrations of IFN-γ and IL-4 that trigger Th1-mediated immunity but not Th2-related immune responses (79). This activation is opposed by the function of γδ T cells; depletion of γδ T cells in a mouse model leads to a decreased fungal burden and reduced IFN-γ levels (188). Since Th1 immunity is proinflammatory, exaggerated Th1 activation during infection might have negative consequences for individuals, and thus γδ T cells might function as downregulators of Th1 responses to sustain a healthy Th1-Th2 balance (77). However, C. neoformans is able to actively change the Th1-Th2 balance toward a Th2 profile by expressing eicosanoids (e.g., prostaglandins and leukotrienes), which are potent inhibitors of Th1-type immunity (146). In addition, expression of the virulence factor urease promotes a Th2 immune response within the lungs via an unknown mechanism (26).

Cytokine signaling also leads to downstream activation or inhibition of antimicrobial effects in other immune cells, such as phagocytic effector cells; Th1 cytokines activate macrophages to create an oxidative and nitrosative burst as microbicidal mechanisms (classically activated macrophages) (72, 168), whereas Th2-polarized host responses lead to inhibition of phagocyte activity (alternatively activated macrophages) and enhanced susceptibility to C. neoformans (87). This alternative activation is associated with upregulated expression of genes involved in tissue repair, such as arginase-1 and the mannose receptor (reviewed in reference 59). Arginase-1 competes with inducible nitric oxidase for the substrate l-arginine and so decreases the synthesis of nitric oxide (66). In fact, intracellular cryptococcal proliferation is significantly higher, and the occurrence of expulsion significantly lower, in macrophages activated by the Th2 cytokine IL-4 or IL-13 than in cells stimulated by Th1 cytokines, such as IFN-γ and TNF-α (Fig. 3) (190). Likewise, the number of intracellular cryptococci is increased in alveolar macrophages isolated from IFN-γ knockout mice (7). It is therefore likely that a shift to a Th2 environment results in activation changes in phagocytic effector cells and hence to a change in the composition of phagolysosomes. However, IFN-γ treatment increases the intracellular growth of C. neoformans in human macrophages (104, 157), suggesting that aspects of this response may be host species specific.

Fig. 3.

Fig. 3.

Th1-Th2-Th17 balance during cryptococcosis. The ability of macrophages to inhibit cryptococcal growth is strongly dependent on cytokine balance. A Th1 and/or Th17 cytokine profile leads to less intracellular C. neoformans and C. gattii proliferation, whereas a dominant Th2 cytokine profile increases cryptococcal proliferative potential. The three rows show intracellular yeast proliferation at 0, 9, 18, and 21 h after treatment with the Th1 cytokine TNF-α, the Th17 cytokine IL-17, and the Th2 cytokine IL-13.

CRYPTOCOCCAL IMMUNE EVASION STRATEGIES IN HEALTHY AND IMMUNOCOMPROMISED PATIENTS

Pathogens have evolved a great variety of strategies to gain advantage within the host's environment and to successfully undermine the host's defense mechanisms. This section will focus on strategies applied by Cryptococcus that counteract the host's immune responses and then discuss how this scientific knowledge may contribute to the development of more efficient therapeutic regimes.

Cryptococcal immune evasion strategies.

Infections are associated with a continuous struggle between the pathogen and the host. To achieve an advantage in the host cell environment, Cryptococcus expresses a wide range of virulence factors (e.g., capsule, melanin, secreted enzymes) that can modify the host's immune response to improve pathogen survival.

The cryptococcal polysaccharide capsule is the best-studied virulence factor (reviewed in reference 202). It inhibits phagocytosis by macrophages (14, 60, 102), dendritic cells (189), and neutrophils (39, 40, 91), and consequently, nonencapsulated strains are phagocytosed three times more effectively by human leukocytes (15, 27). This effect seems to result from a masking mechanism of opsonins by which the polysaccharides form a barrier between the opsonins and their receptors (127). Once infection is established, high concentrations of free glucuronoxylomannan, the major component of the cryptococcal capsule, are found in the patient's bodily fluids, making it likely that many antibodies form immune complexes before they can efficiently opsonize yeast cells (155). In addition, C. neoformans expresses the factor antiphagocytic protein 1 (APP1), which inhibits uptake through a complement-mediated mechanism of binding to complement receptors 2 and 3 (114, 176).

Besides having antiphagocytic properties, the cryptococcal capsule provides protection against reactive oxygen species (ROS) and nitrogen species within host cells. Capsule enlargement upon interaction with phagocytic effector cells in vivo and in vitro (45) correlates with cryptococcal susceptibility to ROS; ROS kill cells with larger capsules less efficiently than they kill cells with smaller capsules (201). This effect may be due to the action of glucuronoxylomannan; this major component of the cryptococcal capsule reduces cryptococcal killing and the production of superoxide in primary human neutrophils (137). In fact, C. neoformans shows high phenotypic plasticity; a process called phenotypic switching enables a switch in cryptococcal morphology between smooth and mucoid, with the latter cell type showing increased survival in murine macrophages (53, 63, 75). In addition, several secreted enzymes are also involved in detoxifying oxygen and nitrogen radicals: superoxide dismutase (SOD) (24), an alternative oxidase (AOX1) (2), a flavinhemoglobin denitrosylase (FHB1) (32), urease (URA1) (26), glutathione peroxidase (GLR1) (132), and thiol peroxidase (TSA1) (132). Finally, the enzyme laccase appears to protect C. neoformans in multiple ways from oxidative and nitrosative stresses; it is involved in the production of the antioxidant pigment melanin via the sphingolipid pathway (67, 192, 193) and also protects C. neoformans in a melanin-independent manner by an iron oxidase function that may maintain iron in an oxidized form, thereby inhibiting production of hydroxyl radicals (111).

Upon uptake, phagocytic effector cells decrease the phagolysosomal pH to below 5.5 to improve microbial killing (179). However, C. neoformans can survive in these low-pH conditions; indeed, artificially increasing the pH, by adding agents such as chloroquine, actually inhibits cryptococcal survival (106). The enzyme inositol phosphosphingolipid-phospholipase C (ISC1) is important for adaptation to the acidic environment, and deletion of isc1 renders cryptococcal cells more susceptible to acidic, oxidative, and nitrosative stresses (166). ISC1 generates phytoceramides that are known to play an important role in regulating PMA1 (49, 57), an ATPase that is involved in the regulation of intracellular pH (69, 173, 175), as well as oxidative (170) and nitrosative (204) stresses in Saccharomyces cerevisiae.

Glucosylceramide (GlcCer), a glycosphingolipid found at the surfaces of C. neoformans cells, has been identified as a new regulator of fungal virulence in recent years (160). Knockout of the GlcCer synthase 1 (GCS1) gene results in a very interesting phenotype in mouse models where the mutant is rendered avirulent following nasal inhalation yet causes fatal disease when injected intravenously (160). The Δgcs1 mutant strain also shows a specific growth defect under high CO2 and at neutral pH and grows well within macrophages (160). Within tissues, the CO2 concentration is relatively high, at 5%, compared to 0.04% in the atmosphere, suggesting that the sphingolipid might be involved in adaptation to the conditions within the host environment and that the mutant is impaired in traversing the lung tissue to reach the intracellular niche (133).

The functions of many of the cryptococcal virulence factors mentioned above depend on the availability of metal ions and thus cation homeostasis. A recent study has suggested that anti-inflammatory cytokines might enhance iron uptake and storage by macrophages by suppressing the activation of iron regulatory proteins 1 and 2, leading to translational repression of the iron storage protein ferritin or translational activation of the membrane receptor for iron uptake (148). This would result in increased metal ion availability and thus increased activity of virulence factors.

Immunotherapy in cryptococcal disease.

The new generation of antifungals, echinocandins, are not active against C. neoformans and in consequence are not used in clinical practice, thus limiting the range of antifungals available to treat the disease (123). Current antifungal treatment regimes involve combination therapy of amphotericin B, flucytosine, and fluconazole (151). The introduction of highly active antiretroviral therapy (HAART) has reduced the incidence of cryptococcosis in developed countries but not the short-term mortality, and hence HAART has not improved the clinical outcome of cryptococcosis. Despite rapid clinical intervention, the 3-month mortality of HIV patients with acute cryptococcal meningoencephalitis is as high as 20% (42, 112). This poor prognosis has resulted in the need for exploration of alternative treatment regimes, such as immunotherapy, passive immunization, and cytokine-based treatment strategies. One promising approach might be the use of adjunctive passive immunotherapy with monoclonal antibodies. As the administration of anticapsular antibodies is protective in infected mice (41, 61), a phase I trial with the monoclonal antibody 18B7 was conducted with HIV patients who had recovered from cryptococcal meningitis (99). Therapy was well tolerated for antibody concentrations up to 1 mg/kg of body weight, with higher doses showing pharmacological effects (99). A second antibody, 2G8, targeting the cell wall glucan exerts remarkable anticryptococcal activity in vitro and in vivo and might be a good candidate as a new therapeutic agent (156). Thus, there is clear potential for anticryptococcal-antibody-mediated therapy. However, it remains a significant challenge to design treatment strategies considering the complex pharmacodynamics of administrated antibodies and antigens within the host as well as drug-associated toxicity. Vaccination with GXM-tetanus toxoid conjugates in mouse model systems has also shown promise (34, 35) but is not yet approved for clinical use.

Cytokine-based treatments have also been proposed for cryptococcosis treatment. IFN-γ levels at the site of infection correlate with fungal burden (169), and administration of IFN-γ has successfully improved the outcome of systemic cryptococcosis in mouse model systems (23, 116) and in one human patient (144). However, although IFN-γ seems to have a protective role in mice (7, 68) and to increase fungicidal activity of murine macrophages (50), the same treatment reduces anticryptococcal activity in human macrophages (104, 157), suggesting that this line of treatment may be less advantageous for human patients.

CONCLUSIONS

C. neoformans is now recognized to be a facultative intracellular pathogen of host cells. This intracellular location provides a niche to escape host immune mechanisms (e.g., complement and antibodies) and also reduces the exposure to antifungal agents. C. neoformans has developed a wide range of mechanisms to adapt to the intracellular niche and counteract the host's immune response. Future research will need to consider the ability to parasitize host cells in order to advance therapeutic schemes. Targeting intracellular survival and growth and/or cryptococcal virulence factors expressed during intracellular parasitism might offer new strategies to improve anticryptococcal treatment. Finally, since C. neoformans is a major pathogen of immunocompromised patients, new strategies need to consider targeted modification of the patient's immune system, for instance by the selective administration of proinflammatory cytokines, to encourage the expression of a protective immune profile.

Biographies

graphic file with name zek9990935450004.jpg

Kerstin Voelz did her undergraduate studies in biology at Friedrich Schiller University in Jena, Germany. While working on her undergraduate thesis on the transcription of genes involved in carotene metabolism in the zygomycete Rhizopus oryzae, she developed an interest in fungal biology. After finishing her first degree, she entered the University of Birmingham, United Kingdom, on a Darwin Fellowship to combine her passion for fungal research with her interest in the continuous struggle between pathogens and their hosts. Her Ph.D. research, conducted with Robin May, focuses on the interaction between the facultative pathogenic yeast Cryptococcus and macrophages. In particular, she is interested in the influence of immune signaling on this interaction and how Cryptococcus evades the human immune system to cause disease.

graphic file with name zek9990935450005.jpg

Robin C. May is a principal investigator in infectious disease at the University of Birmingham, United Kingdom. He studied biological sciences at the University of Oxford before completing a Ph.D. on the actin cytoskeleton under the supervision of Laura Machesky at University College London and, later, at the University of Birmingham. In 2001, he moved to Utrecht, Netherlands, to take a postdoctoral position working on RNA interference with Ronald Plasterk, funded by the Human Frontier Science Program. He was appointed a principal investigator at the University of Birmingham in 2005. Work in his laboratory focuses on host-pathogen interactions in three major areas: the fungal disease cryptococcosis, the Gram-positive pathogen Streptococcus agalactiae, and the innate immune system of the nematode Caenorhabditis elegans.

Footnotes

Published ahead of print on 9 April 2010.

REFERENCES

  • 1.Abadi J., Pirofski L. 1999. Antibodies reactive with the cryptococcal capsular polysaccharide glucuronoxylomannan are present in sera from children with and without human immunodeficiency virus infection. J. Infect. Dis. 180:915–919 [DOI] [PubMed] [Google Scholar]
  • 2.Akhter S., McDade H. C., Gorlach J. M., Heinrich G., Cox G. M., Perfect J. R. 2003. Role of alternative oxidase gene in pathogenesis of Cryptococcus neoformans. Infect. Immun. 71:5794–5802 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Altfeld M., Addo M. M., Kreuzer K. A., Rockstroh J. K., Dumoulin F. L., Schliefer K., Leifeld L., Sauerbruch T., Spengler U. 2000. T(H)1 to T(H)2 shift of cytokines in peripheral blood of HIV-infected patients is detectable by reverse transcriptase polymerase chain reaction but not by enzyme-linked immunosorbent assay under nonstimulated conditions. J. Acquir. Immune Defic. Syndr. 23:287–294 [DOI] [PubMed] [Google Scholar]
  • 4.Alvarez M., Casadevall A. 2006. Phagosome extrusion and host-cell survival after Cryptococcus neoformans phagocytosis by macrophages. Curr. Biol. 16:2161–2165 [DOI] [PubMed] [Google Scholar]
  • 5.Alvarez M., Casadevall A. 2007. Cell-to-cell spread and massive vacuole formation after Cryptococcus neoformans infection of murine macrophages. BMC Immunol. 8:16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Alvarez M., Saylor C., Casadevall A. 2008. Antibody action after phagocytosis promotes Cryptococcus neoformans and Cryptococcus gattii macrophage exocytosis with biofilm-like microcolony formation. Cell. Microbiol. 10:1622–1633 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Arora S., Hernandez Y., Erb-Downward J. R., McDonald R. A., Toews G. B., Huffnagle G. B. 2005. Role of IFN-gamma in regulating T2 immunity and the development of alternatively activated macrophages during allergic bronchopulmonary mycosis. J. Immunol. 174:6346–6356 [DOI] [PubMed] [Google Scholar]
  • 8.Baum G. L., Artis D. 1961. Fungistatic effects of cell free human serum. Am. J. Med. Sci. 242:761–770 [DOI] [PubMed] [Google Scholar]
  • 9.Baum G. L., Artis D. 1961. Growth inhibition of Cryptococcus neoformans by cell free human serum. Am. J. Med. Sci. 241:613–616 [DOI] [PubMed] [Google Scholar]
  • 10.Baum G. L., Artis D. 1963. Characterization of the growth inhibition factor for Cryptococcus neoformans (Gifc) in human serum. Am. J. Med. Sci. 246:53–57 [DOI] [PubMed] [Google Scholar]
  • 11.Bauman S. K., Nichols K. L., Murphy J. W. 2000. Dendritic cells in the induction of protective and nonprotective anticryptococcal cell-mediated immune responses. J. Immunol. 165:158–167 [DOI] [PubMed] [Google Scholar]
  • 12.Beenhouwer D. O., Shapiro S., Feldmesser M., Casadevall A., Scharff M. D. 2001. Both Th1 and Th2 cytokines affect the ability of monoclonal antibodies to protect mice against Cryptococcus neoformans. Infect. Immun. 69:6445–6455 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Blackstock R., Murphy J. W. 2004. Role of interleukin-4 in resistance to Cryptococcus neoformans infection. Am. J. Respir. Cell Mol. Biol. 30:109–117 [DOI] [PubMed] [Google Scholar]
  • 14.Bolanos B., Mitchell T. G. 1989. Phagocytosis of Cryptococcus neoformans by rat alveolar macrophages. J. Med. Vet. Mycol. 27:203–217 [PubMed] [Google Scholar]
  • 15.Bulmer G. S., Sans M. D. 1967. Cryptococcus neoformans. II. Phagocytosis by human leukocytes. J. Bacteriol. 94:1480–1483 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bulmer G. S., Tacker J. R. 1975. Phagocytosis of Cryptococcus neoformans by alveolar macrophages. Infect. Immun. 11:73–79 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Campbell G. D. 1966. Primary pulmonary cryptococcosis. Am. Rev. Respir. Dis. 94:236–243 [DOI] [PubMed] [Google Scholar]
  • 18.Casadevall A., Pirofski L. 2005. Insights into mechanisms of antibody-mediated immunity from studies with Cryptococcus neoformans. Curr. Mol. Med. 5:421–433 [DOI] [PubMed] [Google Scholar]
  • 19.Chang Y. C., Kwon-Chung K. J. 1994. Complementation of a capsule-deficient mutation of Cryptococcus neoformans restores its virulence. Mol. Cell. Biol. 14:4912–4919 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Charlier C., Nielsen K., Daou S., Brigitte M., Chretien F., Dromer F. 2009. Evidence of a role for monocytes in dissemination and brain invasion by Cryptococcus neoformans. Infect. Immun. 77:120–127 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Chaturvedi V., Wong B., Newman S. L. 1996. Oxidative killing of Cryptococcus neoformans by human neutrophils. Evidence that fungal mannitol protects by scavenging reactive oxygen intermediates. J. Immunol. 156:3836–3840 [PubMed] [Google Scholar]
  • 22.Chretien F., Lortholary O., Kansau I., Neuville S., Gray F., Dromer F. 2002. Pathogenesis of cerebral Cryptococcus neoformans infection after fungemia. J. Infect. Dis. 186:522–530 [DOI] [PubMed] [Google Scholar]
  • 23.Clemons K. V., Lutz J. E., Stevens D. A. 2001. Efficacy of recombinant gamma interferon for treatment of systemic cryptococcosis in SCID mice. Antimicrob. Agents Chemother. 45:686–689 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Cox G. M., Harrison T. S., McDade H. C., Taborda C. P., Heinrich G., Casadevall A., Perfect J. R. 2003. Superoxide dismutase influences the virulence of Cryptococcus neoformans by affecting growth within macrophages. Infect. Immun. 71:173–180 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Cox G. M., McDade H. C., Chen S. C., Tucker S. C., Gottfredsson M., Wright L. C., Sorrell T. C., Leidich S. D., Casadevall A., Ghannoum M. A., Perfect J. R. 2001. Extracellular phospholipase activity is a virulence factor for Cryptococcus neoformans. Mol. Microbiol. 39:166–175 [DOI] [PubMed] [Google Scholar]
  • 26.Cox G. M., Mukherjee J., Cole G. T., Casadevall A., Perfect J. R. 2000. Urease as a virulence factor in experimental cryptococcosis. Infect. Immun. 68:443–448 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Cross C. E., Collins H. L., Bancroft G. J. 1997. CR3-dependent phagocytosis by murine macrophages: different cytokines regulate ingestion of a defined CR3 ligand and complement-opsonized Cryptococcus neoformans. Immunology 91:289–296 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Crowley M., Inaba K., Steinman R. M. 1990. Dendritic cells are the principal cells in mouse spleen bearing immunogenic fragments of foreign proteins. J. Exp. Med. 172:383–386 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Dam T. K., Torres M., Brewer C. F., Casadevall A. 2008. Isothermal titration calorimetry reveals differential binding thermodynamics of variable region-identical antibodies differing in constant region for a univalent ligand. J. Biol. Chem. 283:31366–31370 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Davies S. F., Clifford D. P., Hoidal J. R., Repine J. E. 1982. Opsonic requirements for the uptake of Cryptococcus neoformans by human polymorphonuclear leukocytes and monocytes. J. Infect. Dis. 145:870–874 [DOI] [PubMed] [Google Scholar]
  • 31.Decken K., Kohler G., Palmer-Lehmann K., Wunderlin A., Mattner F., Magram J., Gately M. K., Alber G. 1998. Interleukin-12 is essential for a protective Th1 response in mice infected with Cryptococcus neoformans. Infect. Immun. 66:4994–5000 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.de Jesus-Berrios M., Liu L., Nussbaum J. C., Cox G. M., Stamler J. S., Heitman J. 2003. Enzymes that counteract nitrosative stress promote fungal virulence. Curr. Biol. 13:1963–1968 [DOI] [PubMed] [Google Scholar]
  • 33.Del Poeta M. 2004. Role of phagocytosis in the virulence of Cryptococcus neoformans. Eukaryot. Cell 3:1067–1075 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Devi S. J. 1996. Preclinical efficacy of a glucuronoxylomannan-tetanus toxoid conjugate vaccine of Cryptococcus neoformans in a murine model. Vaccine 14:841–844 [DOI] [PubMed] [Google Scholar]
  • 35.Devi S. J., Schneerson R., Egan W., Ulrich T. J., Bryla D., Robbins J. B., Bennett J. E. 1991. Cryptococcus neoformans serotype A glucuronoxylomannan-protein conjugate vaccines: synthesis, characterization, and immunogenicity. Infect. Immun. 59:3700–3707 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Diamond R. D., Erickson N. F., III 1982. Chemotaxis of human neutrophils and monocytes induced by Cryptococcus neoformans. Infect. Immun. 38:380–382 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Diamond R. D., May J. E., Kane M., Frank M. M., Bennett J. E. 1973. The role of late complement components and the alternate complement pathway in experimental cryptococcosis. Proc. Soc. Exp. Biol. Med. 144:312–315 [DOI] [PubMed] [Google Scholar]
  • 38.Diamond R. D., May J. E., Kane M. A., Frank M. M., Bennett J. E. 1974. The role of the classical and alternate complement pathways in host defenses against Cryptococcus neoformans infection. J. Immunol. 112:2260–2270 [PubMed] [Google Scholar]
  • 39.Diamond R. D., Root R. K., Bennett J. E. 1972. Factors influencing killing of Cryptococcus neoformans by human leukocytes in vitro. J. Infect. Dis. 125:367–376 [DOI] [PubMed] [Google Scholar]
  • 40.Dong Z. M., Murphy J. W. 1997. Cryptococcal polysaccharides bind to CD18 on human neutrophils. Infect. Immun. 65:557–563 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Dromer F., Charreire J., Contrepois A., Carbon C., Yeni P. 1987. Protection of mice against experimental cryptococcosis by anti-Cryptococcus neoformans monoclonal antibody. Infect. Immun. 55:749–752 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Dromer F., Mathoulin-Pelissier S., Launay O., Lortholary O. 2007. Determinants of disease presentation and outcome during cryptococcosis: the CryptoA/D study. PLoS Med. 4:e21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Eisenhauer P. B., Lehrer R. I. 1992. Mouse neutrophils lack defensins. Infect. Immun. 60:3446–3447 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Ernst W. A., Thoma-Uszynski S., Teitelbaum R., Ko C., Hanson D. A., Clayberger C., Krensky A. M., Leippe M., Bloom B. R., Ganz T., Modlin R. L. 2000. Granulysin, a T cell product, kills bacteria by altering membrane permeability. J. Immunol. 165:7102–7108 [DOI] [PubMed] [Google Scholar]
  • 45.Feldmesser M., Kress Y., Casadevall A. 2001. Dynamic changes in the morphology of Cryptococcus neoformans during murine pulmonary infection. Microbiology 147:2355–2365 [DOI] [PubMed] [Google Scholar]
  • 46.Feldmesser M., Kress Y., Novikoff P., Casadevall A. 2000. Cryptococcus neoformans is a facultative intracellular pathogen in murine pulmonary infection. Infect. Immun. 68:4225–4237 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Feldmesser M., Mednick A., Casadevall A. 2002. Antibody-mediated protection in murine Cryptococcus neoformans infection is associated with pleotrophic effects on cytokine and leukocyte responses. Infect. Immun. 70:1571–1580 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Feldmesser M., Tucker S., Casadevall A. 2001. Intracellular parasitism of macrophages by Cryptococcus neoformans. Trends Microbiol. 9:273–278 [DOI] [PubMed] [Google Scholar]
  • 49.Ferreira T., Mason A. B., Slayman C. W. 2001. The yeast Pma1 proton pump: a model for understanding the biogenesis of plasma membrane proteins. J. Biol. Chem. 276:29613–29616 [DOI] [PubMed] [Google Scholar]
  • 50.Flesch I. E., Schwamberger G., Kaufmann S. H. 1989. Fungicidal activity of IFN-gamma-activated macrophages. Extracellular killing of Cryptococcus neoformans. J. Immunol. 142:3219–3224 [PubMed] [Google Scholar]
  • 51.Fleuridor R., Lyles R. H., Pirofski L. 1999. Quantitative and qualitative differences in the serum antibody profiles of human immunodeficiency virus-infected persons with and without Cryptococcus neoformans meningitis. J. Infect. Dis. 180:1526–1535 [DOI] [PubMed] [Google Scholar]
  • 52.Fleuridor R., Zhong Z., Pirofski L. 1998. A human IgM monoclonal antibody prolongs survival of mice with lethal cryptococcosis. J. Infect. Dis. 178:1213–1216 [DOI] [PubMed] [Google Scholar]
  • 53.Fries B. C., Taborda C. P., Serfass E., Casadevall A. 2001. Phenotypic switching of Cryptococcus neoformans occurs in vivo and influences the outcome of infection. J. Clin. Invest. 108:1639–1648 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Gadebusch H. H., Johnson A. G. 1966. Natural host resistance to infection with Cryptococcus neoformans. IV. The effect of some cationic proteins on the experimental disease. J. Infect. Dis. 116:551–565 [DOI] [PubMed] [Google Scholar]
  • 55.Garcia-Hermoso D., Janbon G., Dromer F. 1999. Epidemiological evidence for dormant Cryptococcus neoformans infection. J. Clin. Microbiol. 37:3204–3209 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Gates M. A., Kozel T. R. 2006. Differential localization of complement component 3 within the capsular matrix of Cryptococcus neoformans. Infect. Immun. 74:3096–3106 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Ghannoum M. A. 2000. Potential role of phospholipases in virulence and fungal pathogenesis. Clin. Microbiol. Rev. 13:122–143 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Goldman D. L., Khine H., Abadi J., Lindenberg D. J., Pirofski L., Niang R., Casadevall A. 2001. Serologic evidence for Cryptococcus neoformans infection in early childhood. Pediatrics 107:E66. [DOI] [PubMed] [Google Scholar]
  • 59.Gordon S. 2003. Alternative activation of macrophages. Nat. Rev. Immunol. 3:23–35 [DOI] [PubMed] [Google Scholar]
  • 60.Granger D. L., Perfect J. R., Durack D. T. 1985. Virulence of Cryptococcus neoformans. Regulation of capsule synthesis by carbon dioxide. J. Clin. Invest. 76:508–516 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Graybill J. R., Hague M., Drutz D. J. 1981. Passive immunization in murine cryptococcosis. Sabouraudia 19:237–244 [DOI] [PubMed] [Google Scholar]
  • 62.Griffin F. M., Jr 1981. Roles of macrophage Fc and C3b receptors in phagocytosis of immunologically coated Cryptococcus neoformans. Proc. Natl. Acad. Sci. U. S. A. 78:3853–3857 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Guerrero A., Jain N., Wang X., Fries B. C. 2010. Cryptococcus neoformans variants generated by phenotypic switching differ in virulence through effects on macrophage activation. Infect. Immun. 78:1049–1057 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Harrison T. S. 2000. Cryptococcus neoformans and cryptococcosis. J. Infect. 41:12–17 [DOI] [PubMed] [Google Scholar]
  • 65.Hendry A. T., Bakerspigel A. 1969. Factors affecting serum inhibited growth of Candida albicans and Cryptococcus neoformans. Sabouraudia 7:219–229 [PubMed] [Google Scholar]
  • 66.Hesse M., Modolell M., La Flamme A. C., Schito M., Fuentes J. M., Cheever A. W., Pearce E. J., Wynn T. A. 2001. Differential regulation of nitric oxide synthase-2 and arginase-1 by type 1/type 2 cytokines in vivo: granulomatous pathology is shaped by the pattern of l-arginine metabolism. J. Immunol. 167:6533–6544 [DOI] [PubMed] [Google Scholar]
  • 67.Heung L. J., Luberto C., Plowden A., Hannun Y. A., Del Poeta M. 2004. The sphingolipid pathway regulates Pkc1 through the formation of diacylglycerol in Cryptococcus neoformans. J. Biol. Chem. 279:21144–21153 [DOI] [PubMed] [Google Scholar]
  • 68.Hoag K. A., Lipscomb M. F., Izzo A. A., Street N. E. 1997. IL-12 and IFN-gamma are required for initiating the protective Th1 response to pulmonary cryptococcosis in resistant C.B-17 mice. Am. J. Respir. Cell Mol. Biol. 17:733–739 [DOI] [PubMed] [Google Scholar]
  • 69.Holyoak C. D., Stratford M., McMullin Z., Cole M. B., Crimmins K., Brown A. J., Coote P. J. 1996. Activity of the plasma membrane H(+)-ATPase and optimal glycolytic flux are required for rapid adaptation and growth of Saccharomyces cerevisiae in the presence of the weak-acid preservative sorbic acid. Appl. Environ. Microbiol. 62:3158–3164 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Horwitz M. A. 1983. The Legionnaires' disease bacterium (Legionella pneumophila) inhibits phagosome-lysosome fusion in human monocytes. J. Exp. Med. 158:2108–2126 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Houpt D. C., Pfrommer G. S., Young B. J., Larson T. A., Kozel T. R. 1994. Occurrences, immunoglobulin classes, and biological activities of antibodies in normal human serum that are reactive with Cryptococcus neoformans glucuronoxylomannan. Infect. Immun. 62:2857–2864 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Huffnagle G. B. 1996. Role of cytokines in T cell immunity to a pulmonary Cryptococcus neoformans infection. Biol. Signals 5:215–222 [DOI] [PubMed] [Google Scholar]
  • 73.Idnurm A., Bahn Y. S., Nielsen K., Lin X., Fraser J. A., Heitman J. 2005. Deciphering the model pathogenic fungus Cryptococcus neoformans. Nat. Rev. Microbiol. 3:753–764 [DOI] [PubMed] [Google Scholar]
  • 74.Igel H. J., Bolande R. P. 1966. Humoral defense mechanisms in cryptococcosis: substances in normal human serum, saliva, and cerebrospinal fluid affecting the growth of Cryptococcus neoformans. J. Infect. Dis. 116:75–83 [DOI] [PubMed] [Google Scholar]
  • 75.Jain N., Li L., McFadden D. C., Banarjee U., Wang X., Cook E., Fries B. C. 2006. Phenotypic switching in a Cryptococcus neoformans variety gattii strain is associated with changes in virulence and promotes dissemination to the central nervous system. Infect. Immun. 74:896–903 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Janeway C. A., Travers P., Walport M., Shlomchik M. 2001. Immunobiology: the immune system in health and disease, 5th ed.Garland Science, London, England [Google Scholar]
  • 77.Kawakami K. 2004. Regulation by innate immune T lymphocytes in the host defense against pulmonary infection with Cryptococcus neoformans. Jpn. J. Infect. Dis. 57:137–145 [PubMed] [Google Scholar]
  • 78.Kawakami K., Hossain Qureshi M., Zhang T., Koguchi Y., Xie Q., Kurimoto M., Saito A. 1999. Interleukin-4 weakens host resistance to pulmonary and disseminated cryptococcal infection caused by combined treatment with interferon-gamma-inducing cytokines. Cell. Immunol. 197:55–61 [DOI] [PubMed] [Google Scholar]
  • 79.Kawakami K., Kinjo Y., Yara S., Uezu K., Koguchi Y., Tohyama M., Azuma M., Takeda K., Akira S., Saito A. 2001. Enhanced gamma interferon production through activation of Vα14+ natural killer T cells by α-galactosylceramide in interleukin-18-deficient mice with systemic cryptococcosis. Infect. Immun. 69:6643–6650 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Kawakami K., Kohno S., Kadota J., Tohyama M., Teruya K., Kudeken N., Saito A., Hara K. 1995. T cell-dependent activation of macrophages and enhancement of their phagocytic activity in the lungs of mice inoculated with heat-killed Cryptococcus neoformans: involvement of IFN-gamma and its protective effect against cryptococcal infection. Microbiol. Immunol. 39:135–143 [DOI] [PubMed] [Google Scholar]
  • 81.Kechichian T. B., Shea J., Del Poeta M. 2007. Depletion of alveolar macrophages decreases the dissemination of a glucosylceramide-deficient mutant of Cryptococcus neoformans in immunodeficient mice. Infect. Immun. 75:4792–4798 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Kelly R. M., Chen J., Yauch L. E., Levitz S. M. 2005. Opsonic requirements for dendritic cell-mediated responses to Cryptococcus neoformans. Infect. Immun. 73:592–598 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Khawcharoenporn T., Apisarnthanarak A., Mundy L. M. 2007. Non-neoformans cryptococcal infections: a systematic review. Infection 35:51–58 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Kidd S. E., Bach P. J., Hingston A. O., Mak S., Chow Y., MacDougall L., Kronstad J. W., Bartlett K. H. 2007. Cryptococcus gattii dispersal mechanisms, British Columbia, Canada. Emerg. Infect. Dis. 13:51–57 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kidd S. E., Hagen F., Tscharke R. L., Huynh M., Bartlett K. H., Fyfe M., Macdougall L., Boekhout T., Kwon-Chung K. J., Meyer W. 2004. A rare genotype of Cryptococcus gattii caused the cryptococcosis outbreak on Vancouver Island (British Columbia, Canada). Proc. Natl. Acad. Sci. U. S. A. 101:17258–17263 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Kleinschek M. A., Muller U., Brodie S. J., Stenzel W., Kohler G., Blumenschein W. M., Straubinger R. K., McClanahan T., Kastelein R. A., Alber G. 2006. IL-23 enhances the inflammatory cell response in Cryptococcus neoformans infection and induces a cytokine pattern distinct from IL-12. J. Immunol. 176:1098–1106 [DOI] [PubMed] [Google Scholar]
  • 87.Koguchi Y., Kawakami K. 2002. Cryptococcal infection and Th1-Th2 cytokine balance. Int. Rev. Immunol. 21:423–438 [DOI] [PubMed] [Google Scholar]
  • 88.Kozel T. R. 1993. Activation of the complement system by the capsule of Cryptococcus neoformans. Curr. Top. Med. Mycol. 5:1–26 [PubMed] [Google Scholar]
  • 89.Kozel T. R., Follette J. L. 1981. Opsonization of encapsulated Cryptococcus neoformans by specific anticapsular antibody. Infect. Immun. 31:978–984 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Kozel T. R., Gulley W. F., Cazin J., Jr 1977. Immune response to Cryptococcus neoformans soluble polysaccharide: immunological unresponsiveness. Infect. Immun. 18:701–707 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Kozel T. R., Highison B., Stratton C. J. 1984. Localization on encapsulated Cryptococcus neoformans of serum components opsonic for phagocytosis by macrophages and neutrophils. Infect. Immun. 43:574–579 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Kozel T. R., Pfrommer G. S. T. 1986. Activation of the complement system by Cryptococcus neoformans leads to binding of iC3b to the yeast. Infect. Immun. 52:1–5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Kozel T. R., Pfrommer G. S. T., Guerlain A. S., Highison B. A., Highison G. J. 1988. Strain variation in phagocytosis of Cryptococcus neoformans: dissociation of susceptibility to phagocytosis from activation and binding of opsonic fragments of C3. Infect. Immun. 56:2794–2800 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Kozel T. R., Tabuni A., Young B. J., Levitz S. M. 1996. Influence of opsonization conditions on C3 deposition and phagocyte binding of large- and small-capsule Cryptococcus neoformans cells. Infect. Immun. 64:2336–2338 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Kozel T. R., Wilson M. A., Murphy J. W. 1991. Early events in initiation of alternative complement pathway activation by the capsule of Cryptococcus neoformans. Infect. Immun. 59:3101–3110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Kozel T. R., Wilson M. A., Pfrommer G. S. T., Schlageter A. M. 1989. Activation and binding of opsonic fragments of C3 on encapsulated Cryptococcus neoformans by using an alternative complement pathway reconstituted from six isolated proteins. Infect. Immun. 57:1922–1927 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Kozel T. R., Wilson M. A., Welch W. H. 1992. Kinetic analysis of the amplification phase for activation and binding of C3 to encapsulated and nonencapsulated Cryptococcus neoformans. Infect. Immun. 60:3122–3127 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Kwon-Chung K. J., Rhodes J. C. 1986. Encapsulation and melanin formation as indicators of virulence in Cryptococcus neoformans. Infect. Immun. 51:218–223 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Larsen R. A., Pappas P. G., Perfect J., Aberg J. A., Casadevall A., Cloud G. A., James R., Filler S., Dismukes W. E. 2005. Phase I evaluation of the safety and pharmacokinetics of murine-derived anticryptococcal antibody 18B7 in subjects with treated cryptococcal meningitis. Antimicrob. Agents Chemother. 49:952–958 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Laxalt K. A., Kozel T. R. 1979. Chemotaxigenesis and activation of the alternative complement pathway by encapsulated and non-encapsulated Cryptococcus neoformans. Infect. Immun. 26:435–440 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Lehrer R. I., Ganz T. 1990. Antimicrobial polypeptides of human neutrophils. Blood 76:2169–2181 [PubMed] [Google Scholar]
  • 102.Levitz S. M., DiBenedetto D. J. 1989. Paradoxical role of capsule in murine bronchoalveolar macrophage-mediated killing of Cryptococcus neoformans. J. Immunol. 142:659–665 [PubMed] [Google Scholar]
  • 103.Levitz S. M., Dupont M. P., Smail E. H. 1994. Direct activity of human T lymphocytes and natural killer cells against Cryptococcus neoformans. Infect. Immun. 62:194–202 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Levitz S. M., Farrell T. P. 1990. Growth inhibition of Cryptococcus neoformans by cultured human monocytes: role of the capsule, opsonins, the culture surface, and cytokines. Infect. Immun. 58:1201–1209 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Levitz S. M., Harrison T. S., Tabuni A., Liu X. 1997. Chloroquine induces human mononuclear phagocytes to inhibit and kill Cryptococcus neoformans by a mechanism independent of iron deprivation. J. Clin. Invest. 100:1640–1646 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Levitz S. M., Nong S. H., Seetoo K. F., Harrison T. S., Speizer R. A., Simons E. R. 1999. Cryptococcus neoformans resides in an acidic phagolysosome of human macrophages. Infect. Immun. 67:885–890 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Levitz S. M., Specht C. A. 2006. The molecular basis for the immunogenicity of Cryptococcus neoformans mannoproteins. FEMS Yeast Res. 6:513–524 [DOI] [PubMed] [Google Scholar]
  • 108.Levitz S. M., Tabuni A. 1991. Binding of Cryptococcus neoformans by human cultured macrophages. Requirements for multiple complement receptors and actin. J. Clin. Invest. 87:528–535 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Levitz S. M., Tabuni A., Kozel T. R., MacGill R. S., Ingalls R. R., Golenbock D. T. 1997. Binding of Cryptococcus neoformans to heterologously expressed human complement receptors. Infect. Immun. 65:931–935 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Lipovsky M. M., Gekker G., Anderson W. R., Molitor T. W., Peterson P. K., Hoepelman A. I. 1997. Phagocytosis of nonopsonized Cryptococcus neoformans by swine microglia involves CD14 receptors. Clin. Immunol. Immunopathol. 84:208–211 [DOI] [PubMed] [Google Scholar]
  • 111.Liu L., Tewari R. P., Williamson P. R. 1999. Laccase protects Cryptococcus neoformans from antifungal activity of alveolar macrophages. Infect. Immun. 67:6034–6039 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Lortholary O., Poizat G., Zeller V., Neuville S., Boibieux A., Alvarez M., Dellamonica P., Botterel F., Dromer F., Chene G. 2006. Long-term outcome of AIDS-associated cryptococcosis in the era of combination antiretroviral therapy. AIDS 20:2183–2191 [DOI] [PubMed] [Google Scholar]
  • 113.Lovchik J. A., Lipscomb M. F. 1993. Role for C5 and neutrophils in the pulmonary intravascular clearance of circulating Cryptococcus neoformans. Am. J. Respir. Cell Mol. Biol. 9:617–627 [DOI] [PubMed] [Google Scholar]
  • 114.Luberto C., Martinez-Marino B., Taraskiewicz D., Bolanos B., Chitano P., Toffaletti D. L., Cox G. M., Perfect J. R., Hannun Y. A., Balish E., Del Poeta M. 2003. Identification of App1 as a regulator of phagocytosis and virulence of Cryptococcus neoformans. J. Clin. Invest. 112:1080–1094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Luo Y., Cook E., Fries B. C., Casadevall A. 2006. Phagocytic efficacy of macrophage-like cells as a function of cell cycle and Fcgamma receptors (FcgammaR) and complement receptor (CR)3 expression. Clin. Exp. Immunol. 145:380–387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Lutz J. E., Clemons K. V., Stevens D. A. 2000. Enhancement of antifungal chemotherapy by interferon-gamma in experimental systemic cryptococcosis. J. Antimicrob. Chemother. 46:437–442 [DOI] [PubMed] [Google Scholar]
  • 117.Ma H., Croudace J. E., Lammas D. A., May R. C. 2006. Expulsion of live pathogenic yeast by macrophages. Curr. Biol. 16:2156–2160 [DOI] [PubMed] [Google Scholar]
  • 118.Ma H., Croudace J. E., Lammas D. A., May R. C. 2007. Direct cell-to-cell spread of a pathogenic yeast. BMC Immunol. 8:15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Ma L. L., Spurrell J. C., Wang J. F., Neely G. G., Epelman S., Krensky A. M., Mody C. H. 2002. CD8 T cell-mediated killing of Cryptococcus neoformans requires granulysin and is dependent on CD4 T cells and IL-15. J. Immunol. 169:5787–5795 [DOI] [PubMed] [Google Scholar]
  • 120.Ma L. L., Wang C. L., Neely G. G., Epelman S., Krensky A. M., Mody C. H. 2004. NK cells use perforin rather than granulysin for anticryptococcal activity. J. Immunol. 173:3357–3365 [DOI] [PubMed] [Google Scholar]
  • 121.MacDougall L., Kidd S. E., Galanis E., Mak S., Leslie M. J., Cieslak P. R., Kronstad J. W., Morshed M. G., Bartlett K. H. 2007. Spread of Cryptococcus gattii in British Columbia, Canada, and detection in the Pacific Northwest, USA. Emerg. Infect. Dis. 13:42–50 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Macher A. M., Bennett J. E., Gadek J. E., Frank M. M. 1978. Complement depletion in cryptococcal sepsis. J. Immunol. 120:1686–1690 [PubMed] [Google Scholar]
  • 123.Maligie M. A., Selitrennikoff C. P. 2005. Cryptococcus neoformans resistance to echinocandins: (1,3)beta-glucan synthase activity is sensitive to echinocandins. Antimicrob. Agents Chemother. 49:2851–2856 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Mambula S. S., Simons E. R., Hastey R., Selsted M. E., Levitz S. M. 2000. Human neutrophil-mediated nonoxidative antifungal activity against Cryptococcus neoformans. Infect. Immun. 68:6257–6264 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Mansour M. K., Latz E., Levitz S. M. 2006. Cryptococcus neoformans glycoantigens are captured by multiple lectin receptors and presented by dendritic cells. J. Immunol. 176:3053–3061 [DOI] [PubMed] [Google Scholar]
  • 126.Marr K. J., Jones G. J., Zheng C., Huston S. M., Timm-McCann M., Islam A., Berenger B. M., Ma L. L., Wiseman J. C., Mody C. H. 2009. Cryptococcus neoformans directly stimulates perforin production and rearms NK cells for enhanced anticryptococcal microbicidal activity. Infect. Immun. 77:2436–2446 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.McGaw T. G., Kozel T. R. 1979. Opsonization of Cryptococcus neoformans by human immunoglobulin G: masking of immunoglobulin G by cryptococcal polysaccharide. Infect. Immun. 25:262–267 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Mednick A. J., Feldmesser M., Rivera J., Casadevall A. 2003. Neutropenia alters lung cytokine production in mice and reduces their susceptibility to pulmonary cryptococcosis. Eur. J. Immunol. 33:1744–1753 [DOI] [PubMed] [Google Scholar]
  • 129.Mershon K. L., Vasuthasawat A., Lawson G. W., Morrison S. L., Beenhouwer D. O. 2009. Role of complement in protection against Cryptococcus gattii infection. Infect. Immun. 77:1061–1070 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Milam J. E., Herring-Palmer A. C., Pandrangi R., McDonald R. A., Huffnagle G. B., Toews G. B. 2007. Modulation of the pulmonary type 2 T-cell response to Cryptococcus neoformans by intratracheal delivery of a tumor necrosis factor alpha-expressing adenoviral vector. Infect. Immun. 75:4951–4958 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Miller M. F., Mitchell T. G. 1991. Killing of Cryptococcus neoformans strains by human neutrophils and monocytes. Infect. Immun. 59:24–28 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Missall T. A., Pusateri M. E., Lodge J. K. 2004. Thiol peroxidase is critical for virulence and resistance to nitric oxide and peroxide in the fungal pathogen, Cryptococcus neoformans. Mol. Microbiol. 51:1447–1458 [DOI] [PubMed] [Google Scholar]
  • 133.Mitchell A. P. 2006. Cryptococcal virulence: beyond the usual suspects. J. Clin. Invest. 116:1481–1483 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Mitchell T. G., Friedman L. 1972. In vitro phagocytosis and intracellular fate of variously encapsulated strains of Cryptococcus neoformans. Infect. Immun. 5:491–498 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Mitchell T. G., Perfect J. R. 1995. Cryptococcosis in the era of AIDS—100 years after the discovery of Cryptococcus neoformans. Clin. Microbiol. Rev. 8:515–548 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Monari C., Casadevall A., Pietrella D., Bistoni F., Vecchiarelli A. 1999. Neutrophils from patients with advanced human immunodeficiency virus infection have impaired complement receptor function and preserved Fcgamma receptor function. J. Infect. Dis. 180:1542–1549 [DOI] [PubMed] [Google Scholar]
  • 137.Monari C., Retini C., Casadevall A., Netski D., Bistoni F., Kozel T. R., Vecchiarelli A. 2003. Differences in outcome of the interaction between Cryptococcus neoformans glucuronoxylomannan and human monocytes and neutrophils. Eur. J. Immunol. 33:1041–1051 [DOI] [PubMed] [Google Scholar]
  • 138.Mukherjee J., Nussbaum G., Scharff M. D., Casadevall A. 1995. Protective and nonprotective monoclonal antibodies to Cryptococcus neoformans originating from one B cell. J. Exp. Med. 181:405–409 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Mukherjee J., Scharff M. D., Casadevall A. 1992. Protective murine monoclonal antibodies to Cryptococcus neoformans. Infect. Immun. 60:4534–4541 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Müller U., Stenzel W., Kohler G., Werner C., Polte T., Hansen G., Schutze N., Straubinger R. K., Blessing M., McKenzie A. N., Brombacher F., Alber G. 2007. IL-13 induces disease-promoting type 2 cytokines, alternatively activated macrophages and allergic inflammation during pulmonary infection of mice with Cryptococcus neoformans. J. Immunol. 179:5367–5377 [DOI] [PubMed] [Google Scholar]
  • 141.Nadrous H. F., Antonios V. S., Terrell C. L., Ryu J. H. 2003. Pulmonary cryptococcosis in nonimmunocompromised patients. Chest 124:2143–2147 [DOI] [PubMed] [Google Scholar]
  • 142.Nanno M., Shiohara T., Yamamoto H., Kawakami K., Ishikawa H. 2007. Gammadelta T cells: firefighters or fire boosters in the front lines of inflammatory responses. Immunol. Rev. 215:103–113 [DOI] [PubMed] [Google Scholar]
  • 143.Nassar F., Brummer E., Stevens D. A. 1995. Different components in human serum inhibit multiplication of Cryptococcus neoformans and enhance fluconazole activity. Antimicrob. Agents Chemother. 39:2490–2493 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Netea M. G., Brouwer A. E., Hoogendoorn E. H., Van der Meer J. W., Koolen M., Verweij P. E., Kullberg B. J. 2004. Two patients with cryptococcal meningitis and idiopathic CD4 lymphopenia: defective cytokine production and reversal by recombinant interferon-gamma therapy. Clin. Infect. Dis. 39:e83–e87 [DOI] [PubMed] [Google Scholar]
  • 145.Netski D., Kozel T. R. 2002. Fc-dependent and Fc-independent opsonization of Cryptococcus neoformans by anticapsular monoclonal antibodies: importance of epitope specificity. Infect. Immun. 70:2812–2819 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Noverr M. C., Cox G. M., Perfect J. R., Huffnagle G. B. 2003. Role of PLB1 in pulmonary inflammation and cryptococcal eicosanoid production. Infect. Immun. 71:1538–1547 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Osterholzer J. J., Milam J. E., Chen G. H., Toews G. B., Huffnagle G. B., Olszewski M. A. 2009. Role of dendritic cells and alveolar macrophages in regulating early host defense against pulmonary infection with Cryptococcus neoformans. Infect. Immun. 77:3749–3758 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Osterholzer J. J., Surana R., Milam J. E., Montano G. T., Chen G. H., Sonstein J., Curtis J. L., Huffnagle G. B., Toews G. B., Olszewski M. A. 2009. Cryptococcal urease promotes the accumulation of immature dendritic cells and a non-protective T2 immune response within the lung. Am. J. Pathol. 174:932–943 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Panepinto J. C., Komperda K. W., Hacham M., Shin S., Liu X., Williamson P. R. 2007. Binding of serum mannan binding lectin to a cell integrity-defective Cryptococcus neoformans ccr4Δ mutant. Infect. Immun. 75:4769–4779 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Park B. J., Wannemuehler K. A., Marston B. J., Govender N., Pappas P. G., Chiller T. M. 2009. Estimation of the current global burden of cryptococcal meningitis among persons living with HIV/AIDS. AIDS 23:525–530 [DOI] [PubMed] [Google Scholar]
  • 151.Perfect J. R., Dismukes W. E., Dromer F., Goldman D. L., Graybill J. R., Hamill R. J., Harrison T. S., Larsen R. A., Lortholary O., Nguyen M. H., Pappas P. G., Powderly W. G., Singh N., Sobel J. D., Sorrell T. C. 2010. Clinical practice guidelines for the management of cryptococcal disease: 2010 update by the Infectious Diseases Society of America. Clin. Infect. Dis. 50:291–322 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Perfect J. R., Lang S. D., Durack D. T. 1980. Chronic cryptococcal meningitis: a new experimental model in rabbits. Am. J. Pathol. 101:177–194 [PMC free article] [PubMed] [Google Scholar]
  • 153.Pfrommer G. S. T., Dickens S. M., Wilson M. A., Young B. J., Kozel T. R. 1993. Accelerated decay of C3b to iC3b when C3b is bound to the Cryptococcus neoformans capsule. Infect. Immun. 61:4360–4366 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Pietrella D., Corbucci C., Perito S., Bistoni G., Vecchiarelli A. 2005. Mannoproteins from Cryptococcus neoformans promote dendritic cell maturation and activation. Infect. Immun. 73:820–827 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Powderly W. G., Cloud G. A., Dismukes W. E., Saag M. S. 1994. Measurement of cryptococcal antigen in serum and cerebrospinal fluid: value in the management of AIDS-associated cryptococcal meningitis. Clin. Infect. Dis. 18:789–792 [DOI] [PubMed] [Google Scholar]
  • 156.Rachini A., Pietrella D., Lupo P., Torosantucci A., Chiani P., Bromuro C., Proietti C., Bistoni F., Cassone A., Vecchiarelli A. 2007. An anti-β-glucan monoclonal antibody inhibits growth and capsule formation of Cryptococcus neoformans in vitro and exerts therapeutic, anticryptococcal activity in vivo. Infect. Immun. 75:5085–5094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Reardon C. C., Kim S. J., Wagner R. P., Kornfeld H. 1996. Interferon-gamma reduces the capacity of human alveolar macrophages to inhibit growth of Cryptococcus neoformans in vitro. Am. J. Respir. Cell Mol. Biol. 15:711–715 [DOI] [PubMed] [Google Scholar]
  • 158.Rhodes J. C. 1985. Contribution of complement component C5 to the pathogenesis of experimental murine cryptococcosis. Sabouraudia 23:225–234 [DOI] [PubMed] [Google Scholar]
  • 159.Rhodes J. C., Wicker L. S., Urba W. J. 1980. Genetic control of susceptibility to Cryptococcus neoformans in mice. Infect. Immun. 29:494–499 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Rittershaus P. C., Kechichian T. B., Allegood J. C., Merrill A. H., Jr., Hennig M., Luberto C., Del Poeta M. 2006. Glucosylceramide synthase is an essential regulator of pathogenicity of Cryptococcus neoformans. J. Clin. Invest. 116:1651–1659 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Sahu A., Kozel T. R., Pangburn M. K. 1994. Specificity of the thioester-containing reactive site of human C3 and its significance to complement activation. Biochem. J. 302(Pt. 2):429–436 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Santangelo R., Zoellner H., Sorrell T., Wilson C., Donald C., Djordjevic J., Shounan Y., Wright L. 2004. Role of extracellular phospholipases and mononuclear phagocytes in dissemination of cryptococcosis in a murine model. Infect. Immun. 72:2229–2239 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Savoy A. C., Lupan D. M., Manalo P. B., Roberts J. S., Schlageter A. M., Weinhold L. C., Kozel T. R. 1997. Acute lethal toxicity following passive immunization for treatment of murine cryptococcosis. Infect. Immun. 65:1800–1807 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Shao X., Mednick A., Alvarez M., van Rooijen N., Casadevall A., Goldman D. L. 2005. An innate immune system cell is a major determinant of species-related susceptibility differences to fungal pneumonia. J. Immunol. 175:3244–3251 [DOI] [PubMed] [Google Scholar]
  • 165.Shapiro S., Beenhouwer D. O., Feldmesser M., Taborda C., Carroll M. C., Casadevall A., Scharff M. D. 2002. Immunoglobulin G monoclonal antibodies to Cryptococcus neoformans protect mice deficient in complement component C3. Infect. Immun. 70:2598–2604 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Shea J. M., Kechichian T. B., Luberto C., Del Poeta M. 2006. The cryptococcal enzyme inositol phosphosphingolipid-phospholipase C confers resistance to the antifungal effects of macrophages and promotes fungal dissemination to the central nervous system. Infect. Immun. 74:5977–5988 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Shoham S., Huang C., Chen J. M., Golenbock D. T., Levitz S. M. 2001. Toll-like receptor 4 mediates intracellular signaling without TNF-alpha release in response to Cryptococcus neoformans polysaccharide capsule. J. Immunol. 166:4620–4626 [DOI] [PubMed] [Google Scholar]
  • 168.Shoham S., Levitz S. M. 2005. The immune response to fungal infections. Br. J. Haematol. 129:569–582 [DOI] [PubMed] [Google Scholar]
  • 169.Siddiqui A. A., Brouwer A. E., Wuthiekanun V., Jaffar S., Shattock R., Irving D., Sheldon J., Chierakul W., Peacock S., Day N., White N. J., Harrison T. S. 2005. IFN-gamma at the site of infection determines rate of clearance of infection in cryptococcal meningitis. J. Immunol. 174:1746–1750 [DOI] [PubMed] [Google Scholar]
  • 170.Sigler K., Hofer M. 1991. Activation of the plasma membrane H(+)-ATPase of Saccharomyces cerevisiae by addition of hydrogen peroxide. Biochem. Int. 23:861–873 [PubMed] [Google Scholar]
  • 171.Singh N., Alexander B. D., Lortholary O., Dromer F., Gupta K. L., John G. T., del Busto R., Klintmalm G. B., Somani J., Lyon G. M., Pursell K., Stosor V., Munoz P., Limaye A. P., Kalil A. C., Pruett T. L., Garcia-Diaz J., Humar A., Houston S., House A. A., Wray D., Orloff S., Dowdy L. A., Fisher R. A., Heitman J., Wagener M. M., Husain S. 2008. Pulmonary cryptococcosis in solid organ transplant recipients: clinical relevance of serum cryptococcal antigen. Clin. Infect. Dis. 46:e12–e18 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Sorrell T. C., Ellis D. H. 1997. Ecology of Cryptococcus neoformans. Rev. Iberoam Micol. 14:42–43 [PubMed] [Google Scholar]
  • 173.Soteropoulos P., Vaz T., Santangelo R., Paderu P., Huang D. Y., Tamas M. J., Perlin D. S. 2000. Molecular characterization of the plasma membrane H+-ATPase, an antifungal target in Cryptococcus neoformans. Antimicrob. Agents Chemother. 44:2349–2355 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Southwick F. S., Purich D. L. 1996. Intracellular pathogenesis of listeriosis. N. Engl. J. Med. 334:770–776 [DOI] [PubMed] [Google Scholar]
  • 175.Stadler N., Hofer M., Sigler K. 2001. Mechanisms of Saccharomyces cerevisiae PMA1 H+-ATPase inactivation by Fe2+, H2O2 and Fenton reagents. Free Radic. Res. 35:643–653 [DOI] [PubMed] [Google Scholar]
  • 176.Stano P., Williams V., Villani M., Cymbalyuk E. S., Qureshi A., Huang Y., Morace G., Luberto C., Tomlinson S., Del Poeta M. 2009. App1: an antiphagocytic protein that binds to complement receptors 3 and 2. J. Immunol. 182:84–91 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Stevens J. M., Galyov E. E., Stevens M. P. 2006. Actin-dependent movement of bacterial pathogens. Nat. Rev. Microbiol. 4:91–101 [DOI] [PubMed] [Google Scholar]
  • 178.Strasser J. E., Newman S. L., Ciraolo G. M., Morris R. E., Howell M. L., Dean G. E. 1999. Regulation of the macrophage vacuolar ATPase and phagosome-lysosome fusion by Histoplasma capsulatum. J. Immunol. 162:6148–6154 [PubMed] [Google Scholar]
  • 179.Sturgill-Koszycki S., Schlesinger P. H., Chakraborty P., Haddix P. L., Collins H. L., Fok A. K., Allen R. D., Gluck S. L., Heuser J., Russell D. G. 1994. Lack of acidification in Mycobacterium phagosomes produced by exclusion of the vesicular proton-ATPase. Science 263:678–681 [DOI] [PubMed] [Google Scholar]
  • 180.Syme R. M., Spurrell J. C., Amankwah E. K., Green F. H., Mody C. H. 2002. Primary dendritic cells phagocytose Cryptococcus neoformans via mannose receptors and Fcγ receptor II for presentation to T lymphocytes. Infect. Immun. 70:5972–5981 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Szilagyi G., Reiss F., Smith J. C. 1966. The anticryptococcal factor of blood serum. A preliminary report. J. Investig. Dermatol. 46:306–308 [DOI] [PubMed] [Google Scholar]
  • 182.Taborda C. P., Casadevall A. 2001. Immunoglobulin M efficacy against Cryptococcus neoformans: mechanism, dose dependence, and prozone-like effects in passive protection experiments. J. Immunol. 166:2100–2107 [DOI] [PubMed] [Google Scholar]
  • 183.Taborda C. P., Casadevall A. 2002. CR3 (CD11b/CD18) and CR4 (CD11c/CD18) are involved in complement-independent antibody-mediated phagocytosis of Cryptococcus neoformans. Immunity 16:791–802 [DOI] [PubMed] [Google Scholar]
  • 184.Torres M., Casadevall A. 2008. The immunoglobulin constant region contributes to affinity and specificity. Trends Immunol. 29:91–97 [DOI] [PubMed] [Google Scholar]
  • 185.Torres M., Fernandez-Fuentes N., Fiser A., Casadevall A. 2007. Exchanging murine and human immunoglobulin constant chains affects the kinetics and thermodynamics of antigen binding and chimeric antibody autoreactivity. PLoS One 2:e1310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Truelsen K., Young T., Kozel T. R. 1992. In vivo complement activation and binding of C3 to encapsulated Cryptococcus neoformans. Infect. Immun. 60:3937–3939 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Tucker S. C., Casadevall A. 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. U. S. A. 99:3165–3170 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Uezu K., Kawakami K., Miyagi K., Kinjo Y., Kinjo T., Ishikawa H., Saito A. 2004. Accumulation of gammadelta T cells in the lungs and their regulatory roles in Th1 response and host defense against pulmonary infection with Cryptococcus neoformans. J. Immunol. 172:7629–7634 [DOI] [PubMed] [Google Scholar]
  • 189.Vecchiarelli A., Pietrella D., Lupo P., Bistoni F., McFadden D. C., Casadevall A. 2003. The polysaccharide capsule of Cryptococcus neoformans interferes with human dendritic cell maturation and activation. J. Leukoc. Biol. 74:370–378 [DOI] [PubMed] [Google Scholar]
  • 190.Voelz K., Lammas D. A., May R. C. 2009. Cytokine signaling regulates the outcome of intracellular macrophage parasitism by Cryptococcus neoformans. Infect. Immun. 77:3450–3457 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Voskoboinik I., Smyth M. J., Trapani J. A. 2006. Perforin-mediated target-cell death and immune homeostasis. Nat. Rev. Immunol. 6:940–952 [DOI] [PubMed] [Google Scholar]
  • 192.Wang Y., Aisen P., Casadevall A. 1995. Cryptococcus neoformans melanin and virulence: mechanism of action. Infect. Immun. 63:3131–3136 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Wang Y., Casadevall A. 1994. Susceptibility of melanized and nonmelanized Cryptococcus neoformans to nitrogen- and oxygen-derived oxidants. Infect. Immun. 62:3004–3007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Washburn R. G., Bryant-Varela B. J., Julian N. C., Bennett J. E. 1991. Differences in Cryptococcus neoformans capsular polysaccharide structure influence assembly of alternative complement pathway C3 convertase on fungal surfaces. Mol. Immunol. 28:465–470 [DOI] [PubMed] [Google Scholar]
  • 195.Wiseman J. C., Ma L. L., Marr K. J., Jones G. J., Mody C. H. 2007. Perforin-dependent cryptococcal microbicidal activity in NK cells requires PI3K-dependent ERK1/2 signaling. J. Immunol. 178:6456–6464 [DOI] [PubMed] [Google Scholar]
  • 196.Wormley F. L., Jr., Perfect J. R., Steele C., Cox G. M. 2007. Protection against cryptococcosis by using a murine gamma interferon-producing Cryptococcus neoformans strain. Infect. Immun. 75:1453–1462 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Wozniak K. L., Vyas J. M., Levitz S. M. 2006. In vivo role of dendritic cells in a murine model of pulmonary cryptococcosis. Infect. Immun. 74:3817–3824 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Young B. J., Kozel T. R. 1993. Effects of strain variation, serotype, and structural modification on kinetics for activation and binding of C3 to Cryptococcus neoformans. Infect. Immun. 61:2966–2972 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Yuan R. R., Casadevall A., Oh J., Scharff M. D. 1997. T cells cooperate with passive antibody to modify Cryptococcus neoformans infection in mice. Proc. Natl. Acad. Sci. U. S. A. 94:2483–2488 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Zaragoza O., Alvarez M., Telzak A., Rivera J., Casadevall A. 2007. The relative susceptibility of mouse strains to pulmonary Cryptococcus neoformans infection is associated with pleiotropic differences in the immune response. Infect. Immun. 75:2729–2739 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Zaragoza O., Chrisman C. J., Castelli M. V., Frases S., Cuenca-Estrella M., Rodriguez-Tudela J. L., Casadevall A. 2008. Capsule enlargement in Cryptococcus neoformans confers resistance to oxidative stress suggesting a mechanism for intracellular survival. Cell. Microbiol. 10:2043–2057 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Zaragoza O., Rodrigues M. L., De Jesus M., Frases S., Dadachova E., Casadevall A. 2009. The capsule of the fungal pathogen Cryptococcus neoformans. Adv. Appl. Microbiol. 68:133–216 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Zaragoza O., Taborda C. P., Casadevall A. 2003. The efficacy of complement-mediated phagocytosis of Cryptococcus neoformans is dependent on the location of C3 in the polysaccharide capsule and involves both direct and indirect C3-mediated interactions. Eur. J. Immunol. 33:1957–1967 [DOI] [PubMed] [Google Scholar]
  • 204.Zhao L., Zhang F., Guo J., Yang Y., Li B., Zhang L. 2004. Nitric oxide functions as a signal in salt resistance in the calluses from two ecotypes of reed. Plant Physiol. 134:849–857 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Zheng C. F., Jones G. J., Shi M., Wiseman J. C., Marr K. J., Berenger B. M., Huston S. M., Gill M. J., Krensky A. M., Kubes P., Mody C. H. 2008. Late expression of granulysin by microbicidal CD4+ T cells requires PI3K- and STAT5-dependent expression of IL-2Rbeta that is defective in HIV-infected patients. J. Immunol. 180:7221–7229 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Zheng C. F., Ma L. L., Jones G. J., Gill M. J., Krensky A. M., Kubes P., Mody C. H. 2007. Cytotoxic CD4+ T cells use granulysin to kill Cryptococcus neoformans, and activation of this pathway is defective in HIV patients. Blood 109:2049–2057 [DOI] [PubMed] [Google Scholar]

Articles from Eukaryotic Cell are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES