Accessibility / Report Error

Cryptococcal pathogenic mechanisms: a dangerous trip from the environment to the brain

Abstract

Cryptococcus neoformans is an opportunistic pathogenic yeast that causes serious infections, most commonly of the central nervous system (CNS). C. neoformans is mainly found in the environment and acquired by inhalation. It could be metaphorically imagined that cryptococcal disease is a “journey” for the microorganism that starts in the environment, where this yeast loads its suitcase with virulence traits. C. neoformans first encounters the infected mammalian host in the lungs, a site in which it must choose the right elements from its “virulence suitcase” to survive the pulmonary immune response. However, the lung is often only the first stop in this journey, and in some individuals the fungal trip continues to the brain. To enter the brain, C. neoformans must “open” the main barrier that protects this organ, the blood brain barrier (BBB). Once in the brain, C. neoformans expresses a distinct set of protective attributes that confers a strong neurotropism and the ability to cause brain colonisation. In summary, C. neoformans is a unique fungal pathogen as shown in its ability to survive in the face of multiple stress factors and to express virulence factors that contribute to the development of disease.

Key words
Cryptococcus neoformans; capsule; melanin; intracellular pathogen; metabolic adaptation; dissemination


Invasive fungal infections are estimated to cause over 1.5 million deaths per year worldwide, with the vast majority of these infections occurring in immunocompromised patients (Brown et al. 2012Brown GD, Denning DW, Gow NAR, Levitz SM, Netea MG, White TC. Hidden killers: human fungal infections. Sci Transl Med. 2012; 4(165): 165rv13.). Over the last few decades, the emergence of HIV infection in particular has led to the rise in cases of cryptococcal meningoencephalitis. Caused by the basidiomycete yeast Cryptococcus neoformans, cryptococcosis globally results in approximately 215,000 infections per year, leading to 180,000 deaths patients (Rajasingham et al. 2017Rajasingham R, Smith RM, Park BJ, Jarvis JN, Govender NP, Chiller TM, et al. Global burden of disease of HIV-associated cryptococcal meningitis: an updated analysis. Lancet Infect Dis. 2017; 17(8): 873-81.). C. neoformans can be isolated from the environment in many regions of the world, resulting in nearly universal exposure to this fungus among human populations. However, symptomatic disease after exposure is relatively rare. Defects in cell-mediated immunity, especially as directed by CD4+ lymphocytes, are the most common risk factors for developing invasive cryptococcal disease. Additional predisposing factors include solid organ or bone marrow transplantation-associated immunosuppression, treatment with corticosteroids, treatment with tumor necrosis factor-a inhibitors, various malignancies, sarcoidosis, chronic liver disease, and renal failure (Casadevall and Perfect 1998Casadevall A, Perfect JR. Cryptococcus neoformans. Washington DC: ASM Press; 1998., Baddley et al. 2008Baddley JW, Perfect JR, Oster RA, Larsen RA, Pankey GA, Henderson H, et al. Pulmonary cryptococcosis in patients without HIV infection: factors associated with disseminated disease. Eur J Clin Microbiol Infect Dis. 2008; 27(10): 937-43., Maziarz and Perfect 2016Maziarz EK, Perfect JR. Cryptococcosis. Infect Dis Clin North Am. 2016; 30(1): 179-206.). Cryptococcosis is a common AIDS-defining illness and a leading cause of mortality among adults with HIV in Sub-Saharan Africa (Rajasingham et al. 2017Rajasingham R, Smith RM, Park BJ, Jarvis JN, Govender NP, Chiller TM, et al. Global burden of disease of HIV-associated cryptococcal meningitis: an updated analysis. Lancet Infect Dis. 2017; 17(8): 873-81.). Despite the advent of antiretroviral therapy, which drastically reduced the number of HIV cases in the developed world, C. neoformans remains a major problem in resource-limited regions. Furthermore, while the number of AIDS-associated cases of cryptococcal disease has decreased overall, the incidence of disease in solid organ transplant patients and other non-AIDS-associated cases has increased (Bratton et al. 2012Bratton EW, El Husseini N, Chastain CA, Lee MS, Poole C, Stürmer T, et al. Comparison and temporal trends of three groups with cryptococcosis: HIV-infected, solid organ transplant, and HIVnegative/non-transplant. PLoS One. 2012; 7(8): e43582.).

Although this fungus is found primarily in the environment, it possesses features that allow survival and proliferation within a human host. Moreover, C. neoformans must be able to move from the lungs, the most common initial site of infection, to the central nervous system (CNS), the most common site of symptomatic disease. To accomplish this journey, C. neoformans has developed inducible and highly regulated cellular processes that favor fungal survival despite formidable host defenses.

CRYPTOCOCCUS IN THE ENVIRONMENT AND ACQUISITION OF INFECTION: GETTING THE SUITCASE READY AND FULL OF VIRULENCE ATTRIBUTES

C. neoformans is frequently found in the environment in association with pigeon guano, as well as in association with a variety of trees and soils (Emmons 1955Emmons CW. Saprophytic sources of Cryptococcus neoformans associated with the pigeon (Columbia livia). Am J Epidemiol. 1955; 62(3): 227-32., Litvintseva et al. 2011Litvintseva AP, Carbone I, Rossouw J, Thakur R, Govender NP, Mitchell TG. Evidence that the human pathogenic fungus Cryptococcus neoformans var. grubii may have evolved in Africa. PLoS One. 2011; 6(5): e19688., Chowdhary et al. 2012Chowdhary A, Rhandhawa HS, Prakash A, Meis JF. Environmental prevalence of Cryptococcus neoformans and Cryptococcus gattii in India: an update. Crit Rev Microbiol. 2012; 38(1): 1-16.). While C. neoformans does not generally cause symptomatic disease in pigeons due to their high body temperature, these birds are thought to be a reservoir contributing to the global dispersion of this pathogen (Littman and Borok 1968Littman ML, Borok R. Relation of the pigeon to cryptococcosis: natural carrier state, heat resistance and survival of Cryptococcus neoformans. Mycopathol Mycol Appl. 1968; 35(3): 329-45., Litvintseva et al. 2011Litvintseva AP, Carbone I, Rossouw J, Thakur R, Govender NP, Mitchell TG. Evidence that the human pathogenic fungus Cryptococcus neoformans var. grubii may have evolved in Africa. PLoS One. 2011; 6(5): e19688.). C. neoformans is found throughout the world and can infect a wide variety of hosts, including cats, dogs, koalas, dolphins, and even plants (Lester et al. 2004Lester SJ, Kowalewich NJ, Bartlett KH, Krockenberger MB, Fairfax TM, Malik R. Clinicopathologic features of an unusual outbreak of cryptococcosis in dogs, cats, ferrets, and a bird: 38 cases (January to July 2003). J Am Vet Med Assoc. 2004; 225(11): 1716-22., McGill et al. 2009McGill S, Malik R, Saul N, Beetson S, Secombe C, Robertson I, et al. Cryptococcosis in domestic animals in Western Australia: a retrospective study from 1995-2006. Med Mycol. 2009; 47(6): 625-39., Kido et al. 2012Kido N, Makimura K, Kamegaya C, Shindo I, Shibata E, Omiya T, et al. Long-term surveillance and treatment of subclinical cryptococcosis and nasal colonization by Cryptococcus neoformans and C. gattii species complex in captive koalas (Phascolarctes cinereus). Med Mycol. 2012; 50(3): 291-8., Venn-Watson et al. 2012Venn-Watson S, Daniels R, Smith C. Thirty year retrospective evaluation of pneumonia in a bottlenose dolphin Tursiops truncatus population. Dis Aquat Organ. 2012; 99(3): 237-42., Pennisi et al. 2013Pennisi MG, Hartmann K, Lloret A, Ferrer L, Addie D, Belák S, et al. Cryptococcosis in cats: ABCD guidelines on prevention and management. J Feline Med Surg. 2013; 15(7): 611-8., Warpeha et al. 2013Warpeha KM, Park YD, Williamsonb PR. Susceptibility of intact germinating Arabidopsis thaliana to human fungal pathogens Cryptococcus neoformans and C. gattii. Appl Environ Microbiol. 2013; 79(9): 2979-88.). Studies estimate that 50% of urban children have been exposed to Cryptococcus by age 2 (Goldman et al. 2001Goldman DL, Khine H, Abadi J, Lindenberg DJ, Pirofski La, Niang R, et al. Serologic evidence for Cryptococcus neoformans infection in early childhood. Pediatrics. 2001; 107(5): E66.). A related Cryptococcus species, Cryptococcus gattii is found in more restricted geographic regions. Unlike C. neoformans, C. gattii is found almost exclusively in association with certain plant species, including eucalyptus trees (Ellis and Pfeiffer 1990Ellis DH, Pfeiffer TJ. Natural habitat of Cryptococcus neoformans var. gattii. J Clin Microbiol. 1990; 28(7): 1642-4.). Infections with C. gattii have some overlapping features with those due to C. neoformans; however, C. gattii tends to cause disease more often in people without clear immunodeficiencies. Additionally, C. gattii infections often present with focal brain abscesses, instead of more generalised CNS infections (Mitchell et al. 1995Mitchell DH, Sorrell TC, Allworth AM, Heath CH, McGregor AR, Papanaoum K, et al. Cryptococcal disease of the CNS in immunocompetent hosts: influence of cryptococcal variety on clinical manifestations and outcome. Clin Infect Dis. 1995; 20(3): 611-6., Speed and Dunt 1995Speed B, Dunt D. Clinical and host differences between infections with thetwo varieties of Cryptococcus neoformans. Clin Infect Dis. 1995; 21(1): 28-34., Byrnes et al. 2011Byrnes EJ, Bartlett KH, Perfect JR, Heitman J. Cryptococcus gattii: an emerging fungal pathogen infecting humans and animals. Microbes Infect. 2011; 13(11): 895-907.). Interestingly, murine inhalation models of C. gattii infections demonstrate that this Cryptococcus species tends to more frequently cause focal pulmonary disease rather than dissemination to the CNS. These clinical and experimental data suggest that C. gattii may possess microbial features that favor localised and tissue-specific infections rather than widespread systemic dissemination (Ngamskulrungroj et al. 2012Ngamskulrungroj P, Chang Y, Sionov E, Kwon-Chung KJ. The primary target organ of Cryptococcus gattii is different from that of Cryptococcus neoformans in a murine model. MBio. 2012; 3(3): e00103-12.).

C. neoformans has several well characterised virulence attributes, including the ability to grow at mammalian body temperature and the production of polysaccharide capsule, melanin, urease, and phospholipases. C. neoformans has also developed several strategies to survive and replicate within phagocytic cells. However, despite these successful virulence strategies, C. neoformans does not require the mammalian host to complete its lifecycle, leading to its designation as an “accidental pathogen” (Casadevall and Pirofski 2007Casadevall A, Pirofski LA. Accidental virulence, cryptic pathogenesis, martians, lost hosts, and the pathogenicity of environmental microbes. Eukaryot Cell. 2007; 6(12): 2169-74.). Instead, it has been hypothesised that these virulence traits were acquired for survival in the environment, and then re-purposed in the setting of mammalian infection. For example, the antiphagocytic capsule has been hypothesised to protect cells from environmental desiccation. Similarly, the antioxidant melanin pigment, which is required for survival in vivo, is thought to shield against UV radiation-induced cellular damage (Aksenov et al. 1973Aksenov SI, Babyeva IP, Golubev VI. On the mechanism of adaptation of microorganisms to conditions of extreme low humidity. Life Sci Space Res. 1973; 11: 55-61., Nosanchuk and Casadevall 2006Nosanchuk JD, Casadevall A. Impact of melanin on microbial virulence and clinical resistance to antimicrobial compounds. Antimicrob Agents Chemother. 2006; 50(11): 3519-28.).

It has also been suggested that C. neoformans may have evolved to survive encounters with free-living soil microbes, such as amoebae and nematodes (Steenbergen and Casadevall 2003Steenbergen JN, Casadevall A. The origin and maintenance of virulence for the human pathogenic fungus Cryptococcus neoformans. Microbes Infect. 2003; 5(7): 667-75., Casadevall and Pirofski 2007Casadevall A, Pirofski LA. Accidental virulence, cryptic pathogenesis, martians, lost hosts, and the pathogenicity of environmental microbes. Eukaryot Cell. 2007; 6(12): 2169-74.). Amoebae including Acanthamoeba castellanii and Dictyostelium discoideum can interact and ingest C. neoformans in a manner similar to mammalian macrophages (Steenbergen et al. 2001Steenbergen JN, Shuman HA, Casadevall A. Cryptococcus neoformans interactions with amoebae suggest an explanation for its virulence and intracellular pathogenic strategy in macrophages. Proc Natl Acad Sci USA. 2001; 98(26): 15245-50.). C. neoformans is able to kill these organisms, and many of the virulence attributes described above were also shown to be important for survival within amoebae (Steenbergen et al. 2001Steenbergen JN, Shuman HA, Casadevall A. Cryptococcus neoformans interactions with amoebae suggest an explanation for its virulence and intracellular pathogenic strategy in macrophages. Proc Natl Acad Sci USA. 2001; 98(26): 15245-50.). Similar protozoa, as well as bacteria and insects, have been isolated from pigeon guano and shown to influence C. neoformans growth (Ruiz et al. 1982Ruiz A, Neilson JB, Bulmer GS. Control of Cryptococcus neoformans in nature by biotic factors. Med Mycol. 1982; 20(1): 21-9.). Interestingly C. neoformans, but not related non-pathogenic cryptococcal species, can kill the nematode Caenorhabditis elegans (Mylonakis et al. 2002Mylonakis E, Ausubel FM, Perfect JR, Heitman J, Calderwood SB. Killing of Caenorhabditis elegans by Cryptococcus neoformans as a model of yeast pathogenesis. Proc Natl Acad Sci USA. 2002; 99(24): 15675-80.). Interaction with A. castellanii, as well as the wax moth Galleria mellonella, induced C. neoformans capsule and the formation of giant/titan cells, similar to what has been observed in mammalian models of infection (Chrisman et al. 2011Chrisman CJ, Albuquerque P, Guimaraes AJ, Nieves E, Casadevall A. Phospholipids trigger Cryptococcus neoformans capsular enlargement during interactions with amoebae and macrophages. PLoS Pathog. 2011; 7(5): e1002047., García-Rodas et al. 2011García-Rodas R, Casadevall A, Rodríguez-Tudela JL, Cuenca-Estrella M, Zaragoza O. Cryptococcus neoformans capsular enlargement and cellular gigantism during Galleria mellonella infection. PLoS One. 2011; 6(9): e24485.).

Role of spores - C. neoformans isolated from the environment grows almost exclusively as a haploid budding yeast. This asexual form replicates by mitosis. C. neoformans also has a defined sexual cycle in which it grows in a filamentous form. Mating occurs when partners of opposite mating types (MATa and MATa) fuse and form filaments with distinct nuclei and specialised clamp cells. The dikaryotic filaments eventually produce a basidium, a terminally differentiated structure at the end of a growing hyphae, in which nuclear fusion and meiosis occur to produce chains of haploid a and a basidiospores at the basidial head (Kwon-Chung 1976Kwon-Chung KJ. Morphogenesis of Filobasidiella neoformans, the sexual state of Cryptococcus neoformans. Mycologia. 1976; 68(4): 821-33.). Strains of the a mating type can also undergo monokaryotic haploid fruiting or same sex mating (Wickes et al. 1996Wickes BL, Mayorga ME, Edman U, Edman JC. Dimorphism and haploid fruiting in Cryptococcus neoformans: association with the alpha-mating type. Proc Natl Acad Sci USA. 1996; 93(July): 7327-31., Lin et al. 2005Lin X, Hull CM, Heitman J. Sexual reproduction between partners of the same mating type in Cryptococcus neoformans. Nature. 2005; 434(7036): 1017-21.). Interestingly, most clinical and environmental isolates are exclusively a mating type, leading to questions about the frequency of sexual reproduction in nature (Kwon-Chung and Bennett 1978Kwon-Chung KJ, Bennett JE. Distribution of α and a mating types of Cryptococcus neoformans among natural and clinical isolates. Am J Epidemiol. 1978; 108(4): 337-40.). However, Litvintseva et al. (2003)Litvintseva AP, Marra RE, Nielsen K, Heitman J, Vilgalys R, Mitchell TG. Evidence of sexual recombination among Cryptococcus neoformans serotype A isolates in Sub-Saharan Africa. Eukaryot Cell. 2003; 2(6): 1162-8. identified fungal populations in Botswana in which the proportion of MATa and MATa isolates is relatively even. Further analysis revealed evidence of clonal expansions and recombination among this population (Litvintseva et al. 2003Litvintseva AP, Marra RE, Nielsen K, Heitman J, Vilgalys R, Mitchell TG. Evidence of sexual recombination among Cryptococcus neoformans serotype A isolates in Sub-Saharan Africa. Eukaryot Cell. 2003; 2(6): 1162-8., Chen et al. 2015Chen Y, Litvintseva AP, Frazzitta AE, Haverkamp MR, Wang L, Fang C, et al. Comparative analyses of clinical and environmental populations of Cryptococcus neoformans in Botswana. Mol Ecol. 2015; 24(14): 3559-71.).

It has long been debated whether desiccated yeasts or spores are the predominant infectious particles of C. neoformans. Despite a lack of evidence for frequent sexual reproduction as a mechanism to generate infectious basidiospores, monokaryotic haploid fruiting can potentially result in the production of abundant spores in the absence of a mating partner. While a liquid suspension of fungal cells is generally used in murine models of cryptococcal infection, several studies have demonstrated that spores are capable of producing infection (Sukroongreung et al. 1998Sukroongreung S, Kitiniyom K, Nilakul C, Tantimavanich S. Pathogenicity of basidiospores of Filobasidiella neoformans var. neoformans. Med Mycol. 1998; 36(6): 419-24., Giles et al. 2009Giles SS, Dagenais TRT, Botts MR, Keller NP, Hull CM. Elucidating the pathogenesis of spores from the human fungal pathogen Cryptococcus neoformans. Infect Immun. 2009; 77(8): 3491-3500., Velagapudi et al. 2009Velagapudi R, Hsueh YP, Geunes-Boyer S, Wright JR, Heitman J. Spores as infectious propagules of Cryptococcus neoformans. Infect Immun. 2009; 77(10): 4345-55., Springer et al. 2013Springer DJ, Saini D, Byrnes EJ, Heitman J, Frothingham R. Development of an aerosol model of Cryptococcus reveals humidity as an important factor affecting the viability of Cryptococcus during aerosolization. PLoS One. 2013; 8(7): e69804.). Detailed spore analyses have been delayed due to difficulties in purifying large numbers of spores to homogeneity. However, recent advances in spore isolation have led to the characterisation of spore morphology, stress tolerance, and surface/coat composition (Botts et al. 2009Botts MR, Giles SS, Gates MA, Kozel TR, Hull CM. Isolation and characterization of Cryptococcus neoformans spores reveal a critical role for capsule biosynthesis genes in spore biogenesis. Eukaryot Cell. 2009; 8(4): 595-605.). These studies revealed that the spore surface is composed of specialised polysaccharides, which are thought to aid in persistence in the environment (Botts et al. 2009Botts MR, Giles SS, Gates MA, Kozel TR, Hull CM. Isolation and characterization of Cryptococcus neoformans spores reveal a critical role for capsule biosynthesis genes in spore biogenesis. Eukaryot Cell. 2009; 8(4): 595-605.). Recent comparative proteomic analyses highlighted proteins involved in spore composition as well as proteins important for spore germination and initiation of vegetative growth (Huang et al. 2015Huang M, Hebert AS, Coon JJ, Hull CM. Protein composition of infectious spores reveals novel sexual development and germination factors in Cryptococcus. PLoS Genet. 2015; 11(8): e1005490.).

The recognition of spores compared to yeast-like cells by the immune system has also been investigated, revealing important differences in how these two morphological states are sensed (Giles et al. 2009Giles SS, Dagenais TRT, Botts MR, Keller NP, Hull CM. Elucidating the pathogenesis of spores from the human fungal pathogen Cryptococcus neoformans. Infect Immun. 2009; 77(8): 3491-3500., Walsh et al. 2017Walsh NM, Wuthrich M, Wang H, Klein B, Hull CM. Characterization of C-type lectins reveals an unexpectedly limited interaction between Cryptococcus neoformans spores and Dectin-1. PLoS One. 2017; 12(3): e0173866.). Unlike yeast cells, spores are readily phagocytosed by macrophages, inside which they can germinate and replicate. One caveat, however, is that activated macrophages can rapidly kill ungerminated spores, which are highly susceptible to ROS (Giles et al. 2009Giles SS, Dagenais TRT, Botts MR, Keller NP, Hull CM. Elucidating the pathogenesis of spores from the human fungal pathogen Cryptococcus neoformans. Infect Immun. 2009; 77(8): 3491-3500.). Therefore, the ability for spores to produce an active infection is dependent on their ability to germinate prior to macrophage activation and killing. Once germinated, the budding yeast cells can grow both intracellularly, and extracellularly in the host, prompting investigators to refer to this microbe as a facultative intracellular pathogen.

FIRST STOP OF THE TRIP: ADAPTATION TO THE LUNG ENVIRONMENT (CHOOSING THE RIGHT TRAIT FROM THE VIRULENCE SUITCASE)

A key step in the trip of the cryptococcal disease is its arrival at the lungs. The immune response of this organ is complex and specialised because it is frequently exposed to a large number of exogenous particles suspended in the air, like dust and microorganisms. For this reason, this organ has complex and specialised immune responses to control the continuous challenge from external threats. One of the main mechanisms of defense in the lung depends on tissue-resident macrophages present in the alveoli, which phagocytose and remove exogenous particles and microorganisms. In addition, the lung contains the surfactant system, which is a mixture of phospholipids and glycoproteins whose main function is to maintain the superficial tension during respiration. Some of these surfactant proteins also have antimicrobial properties, as they can bind to microbes and induce phagocytosis.

After inhalation, the infectious particles of C. neoformans have to evade this complex immune response and replicate. Survival in this environment is not an intrinsic property of most microbes. In fact, in animal models, most fungi cannot cause lung infection. For example, the immune response of the lung typically results in complete clearance of most Candida species. Even in the case of Aspergillus fumigatus, a filamentous fungus that can cause pulmonary infection in immunosuppressed individuals, colonisation of the mouse lungs only occurs when the animals are immunosuppressed. For this reason, C. neoformans is a remarkable fungal pathogen due to its effective evasion of the lung immune system, and in fact, it behaves like other primary fungal pathogens, such as Histoplasma or Paracoccidioides species. In the next sections, we will briefly describe the virulence factors and adaptation mechanisms elicited by C. neoformans that produce its adaptation to the lung.

Metabolic adaptation to temperature, nutrients and metals - While there are over 1.5 million fungi, only a handful of these are capable of growing at elevated temperatures, including human body temperature (37°C). Within the basidiomycetes, pathogenic Cryptococcus species are the only organisms known to grow well at high temperatures (Perfect 2006Perfect JR. Cryptococcus neoformans: the yeast that likes it hot. FEMS Yeast Res. 2006; 6(4): 463-8.). It has been demonstrated that growth at 37°C can protect against the accumulation of deleterious mutations, suggesting a role for this trait in genomic stability beyond contributing to pathogenesis in a mammalian host (Xu 2004Xu J. Genotype-environment interactions of spontaneous mutations for vegetative fitness in the human pathogenic fungus Cryptococcus neoformans. Genetics. 2004; 168(3): 1177-88.). Considered one of the main virulence attributes of this organism, many investigators have worked to characterise the proteins important for tolerance to high temperature. Components of the mitochondrial antioxidant response, including manganese superoxide dismutase, have been shown to be important for high temperature growth, as well as virulence (Giles et al. 2005Giles SS, Batinić-Haberle I, Perfect JR, Cox GM. Cryptococcus neo-formans mitochondrial superoxide dismutase: an essential link between antioxidant function and high-temperature growth. Eukaryot Cell. 2005; 4(1): 46-54.). Trehalose, a sugar made by fungi and not by mammals, protects C. neoformans from internal and external stresses, including high temperature. Components of the trehalose biosynthesis pathway are required for high temperature growth as well as virulence in a variety of infection models (Petzold et al. 2006Petzold EW, Himmelreich U, Mylonakis E, Rude T, Toffaletti D, Cox GM, et al. Characterization and regulation of the trehalose synthesis pathway and its importance in the pathogenicity of Cryptococcus neoformans. Infect Immun. 2006; 74(10): 5877-87.). A number of C. neoformans signal transduction pathways also have major roles in sensing and responding to high temperature stress. These include the calcium/calmodulin/calcineurin pathway (Odom et al. 1997Odom A, Muir S, Lim E, Toffaletti DL, Perfect J, Heitman J. Calcineurin is required for virulence of Cryptococcos neoformans. EMBO J. 1997; 16(10): 2576-89., Kraus and Heitman 2003Kraus PR, Heitman J. Coping with stress: calmodulin and calcineurin in model and pathogenic fungi. Biochem Biophys Res Commun. 2003; 311(4): 1151-7.), MAP kinase pathways including the PKC/cell wall integrity pathway (Kraus et al. 2003Kraus PR, Fox DS, Cox GM, Heitman J. The Cryptococcus neoformans MAP kinase Mpk1 regulates cell integrity in response to antifungal drugs and loss of calcineurin function. Mol Microbiol. 2003; 48(5): 1377-87., Gerik et al. 2005Gerik KJ, Donlin MJ, Soto CE, Banks AM, Banks IR, Maligie MA, et al. Cell wall integrity is dependent on the PKC1 signal transduction pathway in Cryptococcus neoformans. Mol Microbiol. 2005; 58(2): 393-408., Gerik et al. 2008Gerik KJ, Bhimireddy SR, Ryerse JS, Specht CA, Lodge JK. PKC1 is essential for protection against both oxidative and nitrosative stresses, cell integrity, and normal manifestation of virulence factors in the pathogenic fungus Cryptococcus neoformans. Eukaryot Cell. 2008; 7(10): 1685-98.) and the high osmolarity glycerol (HOG) response pathway (Bahn and Kojima 2005Bahn Y, Kojima K. Specialization of the HOG pathway and its impact on differentiation and virulence of Cryptococcus neoformans. Mol Biol Cell. 2005; 16(5): 2285-2300., Bahn et al. 2007Bahn YS, Geunes-Boyer S, Heitman J. Ssk2 mitogen-activated protein kinase kinase kinase governs divergent patterns of the stress-activated HOG1 signaling pathway in Cryptococcus neoformans. Eukaryot Cell. 2007; 6(12): 2278-89.), and the Ras signaling pathway (Alspaugh et al. 2000Alspaugh JA, Cavallo LM, Perfect JR, Heitman J. RAS1 regulates filamentation, mating and growth at high temperature of Cryptococcus neoformans. Mol Microbiol. 2000; 36(2): 352-65.). Aside from their roles in thermotolerance, these pathways contribute to the fungal response to other stress responses, and each plays a central role in virulence.

In addition to adapting to the high temperature of the mammalian host, C. neoformans must also adapt to limitations and/or influxes of essential nutrients and metals. Analysis of the C. neoformans genes expressed in the context of murine infected lungs showed the up-regulation of many genes involved in carbon metabolism (Hu et al. 2008Hu G, Cheng P-Y, Sham A, Perfect JR, Kronstad JW. Metabolic adaptation in Cryptococcus neoformans during early murine pulmonary infection. Mol Microbiol. 2008; 69(6): 1456-75.). Hu and colleagues also found that various transporters, including those for monosaccharides, acetate, iron, and copper, were all induced in the murine lung (Hu et al. 2008Hu G, Cheng P-Y, Sham A, Perfect JR, Kronstad JW. Metabolic adaptation in Cryptococcus neoformans during early murine pulmonary infection. Mol Microbiol. 2008; 69(6): 1456-75.). Similarly, a study of the transcriptomes of two clinical isolates from human CSF demonstrated that upregulated genes were enriched for GO terms associated with cellular metabolism in these in vivo clinical samples compared to ex vivo incubated samples (Chen et al. 2014Chen Y, Toffaletti DL, Tenor JL, Litvintseva AP, Fang C, Mitchell TG, et al. The Cryptococcus neoformans transcriptome at the site of human meningitis. MBio. 2014; 5(1): 1-10.).

Iron availability is an important aspect of cryptococcal pathogenesis, and detailed studies have explored the role of this metal in various aspects of its physiology. C. neoformans and other microbes compete with the host for iron, and iron sequestration is a basic component of host “nutritional immunity”. This metal is required for both capsule and melanin synthesis, and excess iron can contribute to exacerbated meningoencephalitis in mouse models of infection (Barluzzi et al. 2002Barluzzi R, Saleppico S, Nocentini A, Boelaert JR, Neglia R, Bistoni F, et al. Iron overload exacerbates experimental meningoencephalitis by Cryptococcus neoformans. J Neuroimmunol. 2002; 132(1-2): 140-6.). C. neoformans has several enzymes and transporters that aid in the acquisition of iron from the host (reviewed in Jung and Kronstad 2008Jung WH, Kronstad JW. Iron and fungal pathogenesis: a case study with Cryptococcus neoformans. Cell Microbiol. 2008; 10(2): 277-84.). Under the transcriptional control of the central iron regulator, Cirl, C. neoformans possesses many cell surface proteins that facilitate iron uptake into the fungal cell. These surface proteins include iron reductases that reduce extracellular iron to allow transport into the cell, iron permeases such as Cft1, and plasma membrane ferroxidases such as Cfo1 to convert iron atoms to biologically optimised oxidation states.

C. neoformans must also be able to sense and respond to the essential metal copper. Copper is both simultaneously required and detrimental for C. neoformans growth in vivo. Copper is an important cofactor for a number of enzymatic reactions, in addition to being required for the enzymatic activity involved in melanin formation. However, there is increasing evidence that it is used by the host as a microbicide; innate immune cells upregulate copper importers to accumulate copper in the phagosome (White et al. 2009White C, Lee J, Kambe T, Fritsche K, Petris MJ. A role for the ATP7A copper-transporting ATPase in macrophage bactericidal activity. J Biol Chem. 2009; 284(49): 33949-56.). Furthermore, alveolar cells isolated from mice challenged with C. neoformans were shown to have increased expression of copper importers and higher levels of intracellular copper (Ding et al. 2013Ding C, Festa RA, Chen YL, Espart A, Palacios Ò, Espín J, et al. Cryptococcus neoformans copper detoxification machinery is critical for fungal virulence. Cell Host Microbe. 2013; 13(3): 265-76.). As with iron, much work has gone into defining the proteins and enzymes required for the response to and regulation of copper uptake and overload. The transcription factor Cuf1 has been shown to regulate the response to both high and low copper conditions (Waterman et al. 2007Waterman SR, Hacham M, Hu G, Zhu X, Park Y-D, Shin S, et al. Role of a CUF1/CTR4 copper regulatory axis in the virulence of Cryptococcus neoformans. J Clin Invest. 2007; 117(3): 794-802., Ding et al. 2011Ding C, Yin J, Tovar EMM, Fitzpatrick DA, Higgins DG, Thiele DJ. The copper regulon of the human fungal pathogen Cryptococcus neoformans H99. Mol Microbiol. 2011; 81(6): 1560-76.). In the lungs and the phagosomes of innate immune cells, C. neoformans experiences high copper conditions, in which Cuf1 directs the upregulation of the copper-detoxifying metallothioneins, CMT1 and CMT2, and downregulates the expression of copper importers (Ding et al. 2013Ding C, Festa RA, Chen YL, Espart A, Palacios Ò, Espín J, et al. Cryptococcus neoformans copper detoxification machinery is critical for fungal virulence. Cell Host Microbe. 2013; 13(3): 265-76., Sun et al. 2014Sun TS, Ju X, Gao HL, Wang T, Thiele DJ, Li JY, et al. Reciprocal functions of Cryptococcus neoformans copper homeostasis machinery during pulmonary infection and meningoencephalitis. Nat Commun. 2014; 5: 5550.). In contrast to its mutational state while in the lung, C. neoformans experiences low copper conditions during brain infection during which Cuf1 directs the transcriptional induction of the CTR1 and CTR4 copper importers, among other proteins to control copper homeostasis (Ding et al. 2013Ding C, Festa RA, Chen YL, Espart A, Palacios Ò, Espín J, et al. Cryptococcus neoformans copper detoxification machinery is critical for fungal virulence. Cell Host Microbe. 2013; 13(3): 265-76., Sun et al. 2014Sun TS, Ju X, Gao HL, Wang T, Thiele DJ, Li JY, et al. Reciprocal functions of Cryptococcus neoformans copper homeostasis machinery during pulmonary infection and meningoencephalitis. Nat Commun. 2014; 5: 5550.).

Upon transitioning to the host environment, C. neoformans must also adapt to the relatively alkaline pH of the mammalian lung. Changes in ambient pH can induce stress on many important cellular processes, including nutrient uptake, protein stability and function, and membrane and cell wall stability and maintenance. The Rim alkaline response pathway is the main signaling pathway responsible for sensing and responding to changes in external pH (reviewed in Selvig and Alspaugh 2011Selvig K, Alspaugh JA. pH response pathways in fungi: adapting to host-derived and environmental signals. Mycobiology. 2011; 39(4): 249-56.). Activation of the pathway occurs when alkaline pH is sensed at the cell surface by the membrane sensing complex composed of the Rra1 membrane protein and members of the ESCRT machinery (Hu et al. 2013Hu G, Caza M, Cadieux B, Chan V, Liu V, Kronstad J. Cryptococcus neoformans requires the ESCRT protein Vps23 for iron acquisition from heme, for capsule formation, and for virulence. Infect Immun. 2013; 81(1): 292-302., Hu et al. 2015Hu G, Caza M, Cadieux B, Bakkeren E, Do E, Jung WH, et al. The endosomal sorting complex required for transport machinery influences haem uptake and capsule elaboration in Cryptococcus neoformans. Mol Microbiol. 2015; 96(5): 973-92., Ost et al. 2015Ost KS, O’Meara TR, Huda N, Esher SK, Alspaugh JA. The Cryptococcus neoformans alkaline response pathway: identification of a novel rim pathway activator. PLoS Genet. 2015; 11(4): e1005159.). The assembled ESCRT complexes serve as a scaffold for the proteolysis complex composed of Rim20, Rim23, and the Rim13 protease, which cleaves the Rim101 transcription factor so that it can transit to the nucleus to regulate gene expression (O’Meara et al. 2010O’Meara TR, Norton D, Price MS, Hay C, Clements MF, Nichols CB, et al. Interaction of Cryptococcus neoformans Rim101 and protein kinase A regulates capsule. PloS Pathog. 2010; 6(2): e1000776., O’Meara et al. 2014O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84., Ost et al. 2015Ost KS, O’Meara TR, Huda N, Esher SK, Alspaugh JA. The Cryptococcus neoformans alkaline response pathway: identification of a novel rim pathway activator. PLoS Genet. 2015; 11(4): e1005159.). Rim101 directly regulates genes required for various stress responses including low iron and elevated salt concentrations (O’Meara et al. 2014O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84.). It is also required for proper formation of the polysaccharide capsule and proper cell wall maintenance in response to host conditions (O’Meara et al. 2010O’Meara TR, Norton D, Price MS, Hay C, Clements MF, Nichols CB, et al. Interaction of Cryptococcus neoformans Rim101 and protein kinase A regulates capsule. PloS Pathog. 2010; 6(2): e1000776., O’Meara et al. 2013O’Meara TR, Holmer SM, Selvig K, Dietrich F, Alspaugh JA. Cryptococcus neoformans Rim101 is associated with cell wall remodeling and evasion of the host immune responses. MBio. 2013; 4(1): 1-13., O’Meara et al. 2014O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84., Ost et al. 2017Ost KS, Esher SK, Wager CML, Walker L, Wagener J, Munro C, et al. Rim pathway-mediate alterations in the fungal cell wall influence immune recognition and inflammation. MBio. 2017; 8(1): e02290-16.).

Adaptation to free radicals: melanin and antioxidant mechanisms - Once inside the phagosome, C. neoformans must also adapt to reactive oxygen species (ROS) in order to survive within this environment. Melanin is perhaps the most well-known factor involved in ROS tolerance. In C. neoformans, melanin synthesis depends on laccase enzymes, which uses dopaminergic precursors, mainly L-DOPA to produce the pigment. In fact, both laccase activity and the accumulation of melanin pigments are required for pathogenesis (Kwon-Chung and Rhodes 1986Kwon-Chung KJ, Rhodes JC. Encapsulation and melanin formation as indicators of virulence in Cryptococccus neoformans. Infect Immun. 1986; 51(1): 218-23., Williamson 1994Williamson PR. Biochemical and molecular characterization of the diphenol oxidase of Cryptococcus neoformans: identification as a laccase. J Bacteriol. 1994; 176(3): 656-64.). Melanin in C. neoformans accumulates at the cell wall and confers resistance to many different types of stresses (Nosanchuk and Casadevall 2003Nosanchuk JD, Casadevall A. The contribution of melanin to microbial pathogenesis. Cell Microbiol. 2003; 5(4): 203-23.). Melanised C. neoformans strains were less susceptible than melanin deficient strains to nitrosative and oxidative stresses (Wang and Casadevall 1994Wang Y, Casadevall A. Susceptibility of melanized and nonmelanized Cryptococcus neoformans to nitrogen- and oxygen-derived oxidants. Infect Immun. 1994; 62(7): 3004-7.). As a free radical scavenger, melanin is capable of neutralising ROS (Jacobson and Hong 1997Jacobson ES, Hong JD. Redox buffering by melanin and Fe(II) in Cryptococcus neoformans. J Bacteriol. 1997; 179(17): 5340-6.). Additionally, laccases, the enzymes responsible for making melanin, interfere with the oxidative burst of phagocytes in part by sequestering and oxidising iron during infection (Jacobson and Hong 1997Jacobson ES, Hong JD. Redox buffering by melanin and Fe(II) in Cryptococcus neoformans. J Bacteriol. 1997; 179(17): 5340-6., Liu et al. 1999Liu L, Tewari RP, Williamson PR. Laccase protects Cryptococcus neoformans from antifungal activity of alveolar macrophages. Infect Immun. 1999; 67(11): 6034-9.). Melanin and the Lac1 laccase enzyme have also been demonstrated to facilitate dissemination of C. neoformans from the lung to the CNS (Noverr et al. 2004Noverr MC, Williamson PR, Fajardo RS, Huffnagle GB. CNLAC1 is required for extrapulmonary dissemination of Cryptococcus neoformans but not pulmonary persistence. Infect Immun. 2004; 72(3): 1693-9.+).

In addition to melanin, combined proteomic and genetic analyses have identified several other cellular processes involved in the nitrosative stress response, from canonical cellular stress response pathways to cell wall maintenance, signal transduction, intracellular transport, transcriptional control, respiration, and metabolism (Missall et al. 2006Missall TA, Pusateri ME, Donlin MJ, Chambers KT, Corbett JA, Lodge JK. Posttranslational, translational, and transcriptional responses to nitric oxide stress in Cryptococcus neoformans: implications for virulence. Eukaryot Cell. 2006; 5(3): 518-29.). Other classical enzymes, including copper- and zinc-containing superoxide dismutase and components of the thioredoxin and glutathione antioxidant systems, have been highlighted in the response to oxidative and nitrosative stresses, as well as in promoting fungal virulence (Cox et al. 2003Cox GM, Harrison TS, McDade HC, Taborda CP, Heinrich G, Casadevall A, et al. Superoxide dismutase influences the virulence of Cryptococcus neoformans by affecting growth within macrophages. Infect Immun. 2003; 71(1): 173-80., Missall and Lodge 2005aMissall TA, Lodge JK. Function of the thioredoxin proteins in Cryptococcus neoformans during stress or virulence and regulation by putative transcriptional modulators. Mol Microbiol. 2005a; 57(3): 847-58., Missall and Lodge 2005bMissall TA, Lodge JK. Thioredoxin reductase is essential for viability in the fungal pathogen Cryptococcus neoformans. Eukaryot Cell. 2005b; 4(2): 487-9., Missall et al. 2005Missall TA, Cherry-Harris JF, Lodge JK. Two glutathione peroxidases in the fungal pathogen Cryptococcus neoformans are expressed in the presence of specific substrates. Microbiology. 2005; 151(8): 2573-81.). Surprisingly, catalases, enzymes that detoxify hydrogen peroxide, were shown not to play a major role in ROS stress tolerance in C. neoformans, perhaps due to functional redundancy with other elements of ROS defense (Giles et al. 2006Giles SS, Stajich JE, Nichols C, Gerrald QD, Alspaugh JA, Dietrich F, et al. The Cryptococcus neoformans catalase gene family and its role in antioxidant defense. Eukaryot Cell. 2006; 5(9): 1447-59.).

The polysaccharide capsule - The most characteristic feature of C. neoformans is its capsule, a complex network of polysaccharides present around the cell wall. This structure has been extensively studied for decades, but there are still key aspects about its biology that remain unknown. The capsule is not required for the replication of the yeast in regular laboratory conditions, as acapsular mutants can divide as well as wild type strains. However, the capsule is very important for virulence (Fromtling et al. 1982Fromtling RA, Shadomy HJ, Jacobson ES. Decreased virulence in stable, acapsular mutants of Cryptococcus neoformans. Mycopathologia. 1982; 79(1): 23-9., Chang and Kwon-Chung 1994Chang YC, Kwon-Chung KJ. Complementation of a capsule-deficient mutation of Cryptococcus neoformans restores its virulence. Mol Cell Biol. 1994; 14(7): 4912-9.). Prior studies have demonstrated that the polysaccharide capsule contributes to disease in two complementary ways. First, it confers a protective shield to the yeast against the multiple challenges produced by the immune system. Additionally, its components exert a large number of deleterious effects on the host (reviewed in Vecchiarelli 2000Vecchiarelli A. Immunoregulation by capsular components of Cryptococcus neoformans. Med Mycol. 2000; 38(6): 407-17., Zaragoza et al. 2009Zaragoza O, Rodrigues ML, de Jesus M, Frases S, Dadachova E, Casadevall A. The capsule of the fungal pathogen Cryptococcus neoformans. Adv Appl Microbiol. 2009; 68: 133-216., O’Meara and Alspaugh 2012O’Meara TR, Alspaugh JA. The Cryptococcus neoformans capsule: a sword and a shield. Clin Microbiol Rev. 2012; 25(3): 387-408., Vecchiarelli and Monari 2012Vecchiarelli A, Monari C. Capsular material of Cryptococcus neoformans: virulence and much more. Mycopathologia. 2012; 173(5-6): 375-86.). For this reason, the capsule is considered the main virulence factor of this yeast.

Capsular composition and capsule organisation - The capsule is mainly composed of two complex polysaccharides: glucuronoxylomannan (GXM) and glucuronoxylomannogalactan (GXMGal) (Bose et al. 2003Bose I, Reese AJ, Ory JJ, Janbon G, Doering TL. A yeast under cover: the capsule of Cryptococcus neoformans. Eukaryot Cell. 2003; 2(4): 655-63., Janbon 2004Janbon G. Cryptococcus neoformans capsule biosynthesis and regulation. FEMS Yeast Res. 2004; 4(8): 765-71., Heiss et al. 2009Heiss C, Klutts JS, Wang Z, Doering TL, Azadi P. The structure of Cryptococcus neoformans galactoxylomannan contains beta-Dglucuronic acid. Carbohydr Res. 2009; 344(7): 915-20.). In turn, GXM is composed of a chain of mannose residues with substitutions of xylose and glucuronic acid. In the case of GXMGal, the main component is a chain of galactose molecules with substitutions of mannose, xylose and glucuronic acid. Many of the proteins and enzymes involved in the synthesis of the capsule have been defined (reviewed in Doering 2009Doering TL. How sweet it is! Cell wall biogenesis and polysaccharide capsule formation in Cryptococcus neoformans. Annu Rev Microbiol. 2009; 63: 223-47.), but there are still important aspects that remain uncharacterised. Although it is known that the polysaccharide capsule is organised as interwoven fibres (Pierini and Doering 2001Pierini LM, Doering TL. Spatial and temporal sequence of capsule construction in Cryptococcus neoformans. Mol Microbiol. 2001; 41(1): 105-15., McFadden et al. 2007McFadden DC, Fries BC, Wang F, Casadevall A. Capsule structural heterogeneity and antigenic variation in Cryptococcus neoformans. Eukaryot Cell. 2007; 6(8): 1464-73., Frases et al. 2009Frases S, Pontes B, Nimrichter L, Viana NB, Rodrigues ML, Casadevall A. Capsule of Cryptococcus neoformans grows by enlargement of polysaccharide molecules. Proc Natl Acad Sci USA. 2009; 106(4): 1228-33.), the mechanisms by which these fibres are assembled remain to be elucidated. Interestingly, there is strong evidence to indicate that the capsule polymers are branched, forming micro-gel like structures (Cordero et al. 2011Cordero RJB, Frases S, Guimarães AJ, Rivera J, Casadevall A. Evidence for branching in cryptococcal capsular polysaccharides and consequences on its biological activity. Mol Microbiol. 2011; 79(4): 1101-17., Araújo et al. 2016Araújo GR, Fontes GN, Leão D, Rocha GM, Pontes B, Sant’Anna C, et al. Cryptococcus neoformans capsular polysaccharides form branched and complex filamentous networks viewed by high-resolution microscopy. J Struct Biol. 2016; 193(1): 75-82.). In addition to being present on the cell surface, capsular polysaccharides can also be found in extracellular vesicles (EVs) (Rodrigues et al. 2007Rodrigues ML, Nimrichter L, Oliveira DL, Frases S, Miranda K, Zaragoza O, et al. Vesicular polysaccharide export in Cryptococcus neoformans is a eukaryotic solution to the problem of fungal trans-cell wall transport. Eukaryot Cell. 2007; 6(1): 48-59.). These structures have therefore been proposed as a mechanism for the extracellular export of capsule components, both for targeting to the cell surface and for release into the surrounding environment. It is still unknown how EVs are formed and trafficked to the outer surface of the cell, allowing release and attachment of capsule components. There are many genes required for capsule polysaccharide synthesis (reviewed in Doering 2000Doering TL. How does Cryptococcus get its coat? Trends Microbiol. 2000; 8(12): 547-53., Bose et al. 2003Bose I, Reese AJ, Ory JJ, Janbon G, Doering TL. A yeast under cover: the capsule of Cryptococcus neoformans. Eukaryot Cell. 2003; 2(4): 655-63., O’Meara and Alspaugh 2012O’Meara TR, Alspaugh JA. The Cryptococcus neoformans capsule: a sword and a shield. Clin Microbiol Rev. 2012; 25(3): 387-408.), however a large proportion of these genes still have uncharacterised functions.

The capsule as a protective structure - Before and during the interaction with the host, the presence of a capsule confers resistance to multiple types of stress. For example, it protects the fungal cell against environmental challenges such as dehydration (Aksenov et al. 1973Aksenov SI, Babyeva IP, Golubev VI. On the mechanism of adaptation of microorganisms to conditions of extreme low humidity. Life Sci Space Res. 1973; 11: 55-61.). Furthermore, some of its roles are required during an actual infection. During infection the capsule contributes to evasion of phagocytosis-mediated killing by alveolar macrophages through several mechanisms. First, it impairs the recognition of cell wall epitopes by macrophage receptors, contributing to phagocyte avoidance (Kozel and Gotschlich 1982Kozel TR, Gotschlich EC. The capsule of Cryptococcus neoformans passively inhibits phagocytosis of the yeast by macrophages. J Immunol. 1982; 129(4): 1675-80.). In addition, the capsular polysaccharides have antioxidant properties, protecting the fungal cell from the toxic effects of reactive oxygen species produced in the phagolysosome (Zaragoza et al. 2008Zaragoza O, Chrisman CJ, Castelli MV, Frases S, Cuenca-Estrella M, Rodríguez-Tudela JL, et al. Capsule enlargement in Cryptococcus neoformans confers resistance to oxidative stress suggesting a mechanism for intracellular survival. Cell Microbiol. 2008; 10(10): 2043-57.).

Changes in capsular size and structure as mechanisms of immune evasion - The capsule is a dynamic structure that changes its composition, structure, and size depending on the environmental conditions. Among these phenomena, one of the best studied is the change in size. The capsule diameter is normally small during growth in rich media, however there is a significant increase in its size after interaction with the host (Feldmesser et al. 2001Feldmesser M, Kress Y, Casadevall A. Dynamic changes in the morphology of Cryptococcus neoformans during murine pulmonary infection. Microbiology. 2001; 147: 2355-65.). This enlargement has been described during infection in animal models and phagocytic cells (Chrisman et al. 2011Chrisman CJ, Albuquerque P, Guimaraes AJ, Nieves E, Casadevall A. Phospholipids trigger Cryptococcus neoformans capsular enlargement during interactions with amoebae and macrophages. PLoS Pathog. 2011; 7(5): e1002047.), the non-vertebrate host G. mellonella (García-Rodas et al. 2011García-Rodas R, Casadevall A, Rodríguez-Tudela JL, Cuenca-Estrella M, Zaragoza O. Cryptococcus neoformans capsular enlargement and cellular gigantism during Galleria mellonella infection. PLoS One. 2011; 6(9): e24485.), and even environmental predators such as free-living amoebas (Chrisman et al. 2011Chrisman CJ, Albuquerque P, Guimaraes AJ, Nieves E, Casadevall A. Phospholipids trigger Cryptococcus neoformans capsular enlargement during interactions with amoebae and macrophages. PLoS Pathog. 2011; 7(5): e1002047.). Furthermore, there are several factors that induce this transition in vitro, such as CO2 (Granger et al. 1985Granger DL, Perfect JR, Durack DT. Virulence of Cryptococcus neoformans: regulation of capsule synthesis by carbon dioxide. J Clin Invest. 1985; 76(2): 508-16.), iron limitation (Vartivarian et al. 1993Vartivarian SE, Anaissie EJ, Cowart RE, Sprigg HA, Tingler MJ, Jacobson ES. Regulation of cryptococcal capsular polysaccharide by iron. J Infect Dis. 1993; 167(1): 186-90.), mammalian serum (Zaragoza et al. 2003aZaragoza O, Fries BC, Casadevall A. Induction of capsule growth in Cryptococcus neoformans by mammalian serum and CO2. Infect Immun. 2003a; 71(11): 6155-64.) and nutrient limitation (Zaragoza and Casadevall 2004Zaragoza O, Casadevall A. Experimental modulation of capsule size in Cryptococcus neoformans. Biol Proced Online. 2004; 6(1): 10-4.). This process seems to be important from a clinical point of view, since there is a correlation between ex vivo capsule size and the intracranial pressure of patients affected by cryptococcal meningoencephalitis (Robertson et al. 2014Robertson EJ, Najjuka G, Rolfes MA, Akampurira A, Jain N, Anantharanjit J, et al. Cryptococcus neoformans ex vivo capsule size is associated with intracranial pressure and host immune response in HIV-associated cryptococcal meningitis. J Infect Dis. 2014; 209(1): 74-82.). Capsule enlargement poses a significant change for the cells and it is believed that it is an energy-costly process that requires protein synthesis and the correct functioning of mitochondria (Trevijano-Contador et al. 2017Trevijano-Contador N, Rossi SA, Alves E, Landín-Ferreiroa S, Zaragoza O. Capsule enlargement in Cryptococcus neoformans is dependent on mitochondrial activity. Front Microbiol. 2017; 8: 1423.). During infection, capsule enlargement confers resistance to complement-mediated phagocytosis (Zaragoza et al. 2003bZaragoza O, Taborda CP, Casadevall A. 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. 2003b; 33(7): 1957-67.) and contributes to killing-avoidance in macrophages (Zaragoza et al. 2008Zaragoza O, Chrisman CJ, Castelli MV, Frases S, Cuenca-Estrella M, Rodríguez-Tudela JL, et al. Capsule enlargement in Cryptococcus neoformans confers resistance to oxidative stress suggesting a mechanism for intracellular survival. Cell Microbiol. 2008; 10(10): 2043-57.). Cells with larger capsules are also more resistant to oxidative stress, antimicrobial peptides and antifungal compounds (i.e., amphotericin B).

The capsule also can undergo other rearrangements that have profound consequences for pathogenesis and immune evasion. For example, the structure and organisation of the polysaccharide fibres can substantially change in the host. There are several monoclonal antibodies (mAbs) to the capsule available, and the binding properties of these mAbs to C. neoformans cells obtained from in vivo samples is variable, even changing during the course of infection (Garcia-Hermoso et al. 2004Garcia-Hermoso D, Dromer F, Janbon G. Cryptococcus neoformans capsule structure evolution in vitro and during murine infection. Infect Immun. 2004; 72(6): 3359-65.). These dynamic capsular changes result in a very heterogeneous population of cryptococcal cells that differ in their epitope composition, which impairs the effectiveness of a proper immune response. Furthermore, changes in capsule structure have been also related dissemination efficiency and to brain invasion (Garcia-Hermoso et al. 2004Garcia-Hermoso D, Dromer F, Janbon G. Cryptococcus neoformans capsule structure evolution in vitro and during murine infection. Infect Immun. 2004; 72(6): 3359-65.). In addition, the structure of the capsule can undergo microevolution in vitro, making the microbial population phenotypically and antigenically variable in laboratory cultures depending on the growth conditions (McFadden et al. 2006McFadden D, Zaragoza O, Casadevall A. The capsular dynamics of Cryptococcus neoformans. Trends Microbiol. 2006; 14(11): 497-505.).

Finally, the density of the polysaccharide fibers also increases in vitro (Maxson et al. 2007Maxson ME, Cook E, Casadevall A, Zaragoza O. The volume and hydration of the Cryptococcus neoformans polysaccharide capsule. Fungal Genet Biol. 2007; 44(3): 180-6.) and during infection (Gates et al. 2004Gates MA, Thorkildson P, Kozel TR. Molecular architecture ofthe Cryptococcus neoformans capsule. Mol Microbiol. 2004; 52(1): 13-24.). Although the consequences of this increase in density are not fully known, it produces a capsular structure that is less permeable to elements of the immune response, such as antibodies, complement or antimicrobial peptides.

Exopolysaccharides as virulence factors - The polysaccharides of the capsule are not only attached to the cell, but they are also released into the medium (exopolysaccharides). During infection, extracellular capsular polysaccharides can be found in tissues, CSF, and blood. Recent work has demonstrated that the release of exopolysaccharides is a regulated process in C. neoformans that depends on environmental cues and distinct genes (Denham et al. 2017Denham ST, Verma S, Reynolds RC, Worne CL, Daugherty JM, Lane TE, et al. Regulated release of cryptococcal polysaccharide drives virulence and suppresses immune infiltration into the central nervous system. Infect Immun. 2017; pii: IAI.00662-17.). These polysaccharides seem to contribute to the development of disease through multiple mechanisms. Among them, both GXM and GXMGal can cause apoptosis of several types of immune cells through activation of FasL/Fas (Chiapello et al. 2003Chiapello LS, Aoki MP, Rubinstein HR, Masih DT. Apoptosis induction by glucuronoxylomannan of Cryptococcus neoformans. Med Mycol. 2003; 41(4): 347-53., Monari et al. 2005bMonari C, Pericolini E, Bistoni G, Casadevall A, Kozel TR, Vecchiarelli A. Cryptococcus neoformans capsular glucuronoxylomannan induces expression of fas ligand in macrophages. J Immunol. 2005b; 174(6): 3461-8., Monari et al. 2006Monari C, Bistoni F, Vecchiarelli A. Glucuronoxylomannan exhibits potent immunosuppressive properties. FEMS Yeast Res. 2006; 6(4): 537-42., Monari et al. 2008Monari C, Paganelli F, Bistoni F, Kozel TR, Vecchiarelli A. Capsular polysaccharide induction of apoptosis by intrinsic and extrinsic mechanisms. Cell Microbiol. 2008; 10(10): 2129-37., Villena et al. 2008Villena SN, Pinheiro RO, Pinheiro CS, Nunes MP, Takiya CM, Dosreis GA, et al. Capsular polysaccharides galactoxylomannan and glucuronoxylomannan from Cryptococcus neoformans induce macrophage apoptosis mediated by Fas ligand. Cell Microbiol. 2008; 10(6): 1274-85.). Secreted polysaccharides can also impair Ab production, induce complement depletion (Macher et al. 1978Macher AM, Bennett JE, Gadek JE, Frank MM. Complement depletion in cryptococcal sepsis. J Immunol. 1978; 120(5): 1686-90.), inhibit leukocyte migration (Dong and Murphy 1995Dong ZM, Murphy JW. Intravascular cryptococcal culture filtrate (CneF) and its major component, glucuronoxylomannan, are potent inhibitors of leukocyte accumulation. Infect Immun. 1995; 63(3): 770-8., Dong et al. 1999Dong ZM, Jackson L, Murphy JW. Mechanisms for induction of Lselectin loss from T lymphocytes by a cryptococcal polysaccharide, glucuronoxylomannan. Infect Immun. 1999; 67(1): 220-9., Ellerbroek et al. 2002Ellerbroek PM, Hoepelman AIM, Wolbers F, Zwaginga JJ, Coenjaerts FEJ. Cryptococcal glucuronoxylomannan inhibits adhesion of neutrophils to stimulated endothelium in vitro by affecting both neutrophils and endothelial cells. Infect Immun. 2002; 70(9): 4762-71.), reduce immune cell infiltration to the brain (Denham et al. 2017Denham ST, Verma S, Reynolds RC, Worne CL, Daugherty JM, Lane TE, et al. Regulated release of cryptococcal polysaccharide drives virulence and suppresses immune infiltration into the central nervous system. Infect Immun. 2017; pii: IAI.00662-17.), and stimulate the production of cytokines and chemokines (Monari et al. 2005aMonari C, Bistoni F, Casadevall A, Pericolini E, Pietrella D, Kozel TR, et al. Glucuronoxylomannan, a microbial compound, regulates expression of costimulatory molecules and production of cytokines in macrophages. J Infect Dis. 2005a; 191(1): 127-37., Vecchiarelli et al. 2011Vecchiarelli A, Pericolini E, Gabrielli E, Chow S-K, Bistoni F, Cenci E, et al. Cryptococcus neoformans galactoxylomannan is a potent negative immunomodulator, inspiring new approaches in anti-inflammatory immunotherapy. Immunotherapy. 2011; 3(8): 997-1005.). Furthermore, these polysaccharides are recognised by several types of immune receptors, such as CD18, CD14 and toll-like receptors (TLRs) (Shoham et al. 2001Shoham S, Huang C, Chen J-M, Golenbock DT, Levitz SM. Tolllike receptor 4 mediates intracellular signaling without TNF-α release in response to Cryptococcus neoformans polysaccharide capsule. J Immunol. 2001; 166(7): 4620-6., Taborda and Casadevall 2002Taborda CP, Casadevall A. CR3 (CD11b/CD18) and CR4 (CD11c/CD18) are involved in complement-independent antibody-mediated phagocytosis of Cryptococcus neoformans. Immunity. 2002; 16(6): 791-802., Yauch et al. 2004Yauch LE, Mansour MK, Shoham S, Rottman JB, Levitz SM. Involvement of CD14, toll-like receptors 2 and 4, and MyD88 in the host response to the fungal pathogen Cryptococcus neoformans in vivo. Infect Immun. 2004; 72(9): 5373-82.).

Intracellular survival inside macrophages/recognition by macrophages - Upon entering the lung, one of the first cell types that C. neoformans engages are innate immune phagocytes, in particular alveolar macrophages. C. neoformans has a dynamic relationship with macrophages, and there is data to support their role in both clearance and persistence of this fungus. For example, depletion of macrophages reduces survival in murine models of infection (Monga 1981Monga DP. Role of macrophages in resistance of mice to experimental cryptococcosis. Infect Immun. 1981; 32(3): 975-8.). On the other hand, while classical virulence attributes such as capsule and melanin assist the fungus to minimise phagocytosis and killing by macrophages, C. neoformans requires macrophages for efficient dissemination to the CNS (Charlier et al. 2009Charlier C, Nielsen K, Daou S, Brigitte M, Chretien F, Dromer F. Evidence of a role for monocytes in dissemination and brain invasion by Cryptococcus neoformans. Infect Immun. 2009; 77(1): 120-7.).

Recognition - The interaction between fungi and host begins when fungal factors are recognised by innate immune cell surface receptors, triggering immune cell activation and inducing phagocytosis of the fungus. A number of pattern recognition receptors (PRRs) recognise C. neoformans, including receptors in the Toll-like (TLR), C-type lectin (CLR), and NOD like families (NLR), as well as scavenger receptors. Acapsular strains are readily ingested by phagocytosis through interactions with the mannose receptor (MR) and Dectin-1 (Cross and Bancroft 1995Cross CE, Bancroft GJ. Ingestion of acapsular Cryptococcus neoformans occurs via mannose and β- glucan receptors, resulting in cytokine production and increased phagocytosis of the encapsulated form. Infect Immun. 1995; 63(7): 2604-11., Casadevall and Perfect 1998Casadevall A, Perfect JR. Cryptococcus neoformans. Washington DC: ASM Press; 1998., Heitman et al. 2010Heitman J, Kozel TR, Kwon-Chung KJ, Perfect JR, Casadevall A. Cryptococcus: from human pathogen to model yeast. Washington DC: ASM Press; 2010.). Capsule components can also be recognised by several receptors, including TLR2, TLR4, and the co-receptor CD14 (Shoham et al. 2001Shoham S, Huang C, Chen J-M, Golenbock DT, Levitz SM. Tolllike receptor 4 mediates intracellular signaling without TNF-α release in response to Cryptococcus neoformans polysaccharide capsule. J Immunol. 2001; 166(7): 4620-6., Yauch et al. 2004Yauch LE, Mansour MK, Shoham S, Rottman JB, Levitz SM. Involvement of CD14, toll-like receptors 2 and 4, and MyD88 in the host response to the fungal pathogen Cryptococcus neoformans in vivo. Infect Immun. 2004; 72(9): 5373-82., Yauch et al. 2005Yauch LE, Mansour MK, Levitz SM. Receptor-mediated clearance of Cryptococcus neoformans capsular polysaccharide in vivo. Infect Immun. 2005; 73(12): 8429-32.). While there is opposing evidence as to the importance of TLR2 in the immune response to C. neoformans, it is clear that TLR4 is not required for protection against C. neoformans in mouse models of infection (Yauch et al. 2004Yauch LE, Mansour MK, Shoham S, Rottman JB, Levitz SM. Involvement of CD14, toll-like receptors 2 and 4, and MyD88 in the host response to the fungal pathogen Cryptococcus neoformans in vivo. Infect Immun. 2004; 72(9): 5373-82., Biondo et al. 2005Biondo C, Midiri A, Messina L, Tomasello F, Garufi G, Catania MR, et al. MyD88 and TLR2, but not TLR4, are required for host defense against Cryptococcus neoformans. Eur J Immunol. 2005; 35(3): 870-8., Nakamura et al. 2006Nakamura K, Miyagi K, Koguchi Y, Kinjo Y, Uezu K, Kinjo T, et al. Limited contribution of Toll-like receptor 2 and 4 to the host response to a fungal infectious pathogen, Cryptococcus neoformans. FEMS Immunol Med Microbiol. 2006; 47(1): 148-54.). A major role for MyD88 (the adaptor protein that directs downstream immune signalling from many of the TLRs) has been demonstrated by multiple groups; mice that are deficient in MyD88 succumb to C. neoformans infection at rates significantly faster than WT mice (Yauch et al. 2004Yauch LE, Mansour MK, Shoham S, Rottman JB, Levitz SM. Involvement of CD14, toll-like receptors 2 and 4, and MyD88 in the host response to the fungal pathogen Cryptococcus neoformans in vivo. Infect Immun. 2004; 72(9): 5373-82., Biondo et al. 2005Biondo C, Midiri A, Messina L, Tomasello F, Garufi G, Catania MR, et al. MyD88 and TLR2, but not TLR4, are required for host defense against Cryptococcus neoformans. Eur J Immunol. 2005; 35(3): 870-8.).

The CLR Dectin-2, which recognises mannan in the fungal cell wall, is associated with higher levels of non-protective Th2 cytokines during C. neoformans infection (Nakamura et al. 2015Nakamura Y, Sato K, Yamamoto H, Matsumura K, Matsumoto I, Nomura T, et al. Dectin-2 deficiency promotes Th2 response and mucin production in the lungs after pulmonary infection with Cryptococcus neoformans. Infect Immun. 2015; 83(2): 671-81.). While Dectin-1 can bind to b-glucan on C. neoformans spores (Giles et al. 2009Giles SS, Dagenais TRT, Botts MR, Keller NP, Hull CM. Elucidating the pathogenesis of spores from the human fungal pathogen Cryptococcus neoformans. Infect Immun. 2009; 77(8): 3491-3500.), its role in phagocytosis of spores as well as overall protection against C. neoformans infection appears to be minimal (Nakamura et al. 2007Nakamura K, Kinjo T, Saijo S, Miyazato A, Adachi Y, Ohno N, et al. Dectin-1 is not required for the host defense to Cryptococcus neoformans. Microbiol Immunol. 2007; 51(11): 1115-9., Walsh et al. 2017Walsh NM, Wuthrich M, Wang H, Klein B, Hull CM. Characterization of C-type lectins reveals an unexpectedly limited interaction between Cryptococcus neoformans spores and Dectin-1. PLoS One. 2017; 12(3): e0173866.). Dectin-3 deficiency was also shown not to be a major factor in immunity towards C. neoformans (Campuzano et al. 2017Campuzano A, Castro-Lopez N, Wozniak KL, Wager CML, Wormley FL. Dectin-3 Is not required for protection against Cryptococcus neoformans infection. PLoS One. 2017; 12(1): e0169347.). However, mice deficient in the CLR adaptor protein Card9, were highly susceptible to C. neoformans infection due to decreased influx of INF-g producing cells, suggesting a role for CLR-mediated signalling pathways in protection from cryptococcal infection (Yamamoto et al. 2014Yamamoto H, Nakamura Y, Sato K, Takahashi Y, Nomura T, Miyasaka T, et al. Defect of CARD9 leads to impaired accumulation of gamma interferon-producing memory phenotype T cells in lungs and increased susceptibility to pulmonary infection with Cryptococcus neoformans. Infect Immun. 2014; 82(4): 1606-15.). Finally, both the mannose receptor (MR) and DC-SIGN recognise mannosylated proteins on the C. neoformans cell surface (Mansour et al. 2006Mansour MK, Latz E, Levitz SM. Cryptococcus neoformans glycoantigens are captured by multiple lectin receptors and presented by dendritic cells. J Immunol. 2006; 176(5): 3053-61.), and MR-deficient mice are highly susceptible to infection with C. neoformans (Dan et al. 2008Dan JM, Kelly RM, Lee CK, Levitz SM. Role of the mannose receptor in a murine model of Cryptococcus neoformans infection. Infect Immun. 2008; 76(6): 2362-7.). Together these data suggest that a combination of immune receptors might act in hetero-complexes to recognise the dynamic surface of C. neoformans, leading to complex downstream immune signalling, similar to what has been described for recognition of other fungal species (reviewed in Inoue and Shinohara 2014Inoue M, Shinohara ML. Clustering of pattern recognition receptors for fungal detection. PLoS Pathog. 2014; 10(2): e1003873.).

While capsule and cell wall components can be recognised by several PRRs, encapsulated strains require opsonisation with antibodies or complement for efficient phagocytosis. Anti-capsular antibodies can be recognised by CD19 and Fcg receptors (Netski and Kozel 2002Netski D, Kozel TR. Fc-dependent and Fc-independent opsonization of Cryptococcus neoformans by anticapsular monoclonal antibodies: Importance of epitope specificity. Infect Immun. 2002; 70(6): 2812-9.). The localisation of the antibody binding, as well as antibody isotype, impact the efficiency of phagocytosis (Nussbaum et al. 1997Nussbaum G, Cleare W, Casadevall A, Scharff MD, Valadon P. Epitope location in the Cryptococcus neoformans capsule is a determinant of antibody efficacy. J Exp Med. 1997; 185(4): 685-94., Cleare and Casadevall 1998Cleare W, Casadevall A. The different binding patterns of two immunoglobulin M monoclonal antibodies to Cryptococcus neoformans serotype A and D strains correlate with serotype classification and differences in functional assays. Clin Diagn Lab Immunol. 1998; 5(2): 125-9.). The cryptococcal capsule is also capable of inducing complement activation through the alternative pathway. This activation results in the deposition of complement proteins within the capsule structure (Kozel 1996Kozel TR. Activation of the complement system by pathogenic fungi. Clin. Microbiol. Rev. 1996; 9(1): 34-46.), which can be recognised by CD11b/CD18 and CD11c/CD18 (Taborda and Casadevall 2002Taborda CP, Casadevall A. CR3 (CD11b/CD18) and CR4 (CD11c/CD18) are involved in complement-independent antibody-mediated phagocytosis of Cryptococcus neoformans. Immunity. 2002; 16(6): 791-802.). Similar to antibody-mediated phagocytosis, the efficiency of complement-mediated phagocytosis depends on capsule size and location of complement protein binding (Kozel 1996Kozel TR. Activation of the complement system by pathogenic fungi. Clin. Microbiol. Rev. 1996; 9(1): 34-46., Zaragoza et al. 2003bZaragoza O, Taborda CP, Casadevall A. 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. 2003b; 33(7): 1957-67., Zaragoza et al. 2009Zaragoza O, Rodrigues ML, de Jesus M, Frases S, Dadachova E, Casadevall A. The capsule of the fungal pathogen Cryptococcus neoformans. Adv Appl Microbiol. 2009; 68: 133-216.). Importantly, complement-deficient animals were more susceptible to C. neoformans infection (Rhodes 1985Rhodes JC. Contribution of complement component C5 to the pathogenesis of experimental murine cryptococcosis. Sabouraudia. 1985; 23(3): 225-34.).

Phagocytosis - As a facultative intracellular pathogen, C. neoformans has many strategies to regulate its phagocytosis by immune cells. Perhaps the best studied is the polysaccharide capsule, which itself inhibits phagocytosis by macrophages (Bolanos and Mitchell 1989Bolanos B, Mitchell TG. Phagocytosis of Cryptococcus neoformans by rat alveolar macrophages. J Med Vet Mycol. 1989; 27(4): 203-17., Levitz and DiBenedetto 1989Levitz SM, DiBenedetto DJ. Paradoxical role of capsule in murine bronchoalveolar macrophage-mediated killing of Cryptococcus neoformans. J Immunol. 1989; 142(2): 659-65.). The specific capsule components can also influence its interaction with host cells through differential binding of opsonins as described above (Kozel et al. 1988Kozel TR, Pfrommer GST, Guerlain AS, Highison BA, Highison GJ. Role of the capsule in phagocytosis of Cryptococcus neoformans. Rev Infect Dis. 1988; 10(Suppl. 2): S436-9., Zaragoza et al. 2003bZaragoza O, Taborda CP, Casadevall A. 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. 2003b; 33(7): 1957-67.).

In addition to capsule, Luberto and colleagues identified an antiphagocytic protein, App1, that has an important role in phagocytosis and virulence in C. neoformans. Importantly, this protein was identified in the serum of AIDS patients with disseminated C. neoformans, highlighting its physiological importance (Salgado et al. 1994Salgado DC, de Jesús JM, Bolaños B. Purification and characterization of a cytoplasmic immunosuppressive component from Cryptococcus neoformans by preparative electrophoretic techniques. In: Abstracts of the 94th general meeting of the American Society for Microbiology. 1994 March 23-27; Las Vegas: 1994; p. 77., Luberto et al. 2003Luberto C, Martinez-Mariño B, Taraskiewicz D, Bolaños B, Chitano P, Toffaletti DL, et al. Identification of App1 as a regulator of phagocytosis and virulence of Cryptococcus neoformans. J Clin Invest. 2003; 112(7): 1080-94.). In vitro, treatment of cells with App1 inhibited engulfment in a complement-dependent manner (Stano et al. 2009Stano P, Williams V, Villani M, Cymbalyuk ES, Qureshi A, Huang Y, et al. App1: an antiphagocytic protein that binds to complement receptors 3 and 2. J Immunol. 2009; 182(1): 84-91.). Conversely, applΔ cells were more readily phagocytosed and displayed attenuated virulence in multiple mouse backgrounds (Luberto et al. 2003Luberto C, Martinez-Mariño B, Taraskiewicz D, Bolaños B, Chitano P, Toffaletti DL, et al. Identification of App1 as a regulator of phagocytosis and virulence of Cryptococcus neoformans. J Clin Invest. 2003; 112(7): 1080-94., Del Poeta 2004Del Poeta M. Role of phagocytosis in the virulence of Cryptococcus neoformans. Eukaryot Cell. 2004; 3(5): 1067-75.). Similarly, there is another regulator, Gat201, which mediates phagocytosis avoidance through a capsule-independent mechanism (Liu et al. 2008Liu OW, Chun CD, Chow ED, Chen C, Madhani HD, Noble SM. Systematic genetic analysis of virulence in the human fungal pathogen Cryptococcus neoformans. Cell. 2008; 135(1): 174-88.).

Survival and proliferation inside macrophages Despite actively avoiding phagocytosis, C. neoformans is quite capable of surviving and proliferating inside of phagocytic immune cells. In fact, C. neoformans is viable and replicates within the acidic environment of the phagolysosome (Levitz et al. 1999Levitz SM, Nong S-H, Seetoo KF, Harrison TS, Speizer RA, Simons ER. Cryptococcus neoformans resides in an acidic phagolysosome of human macrophages. Infect Immun. 1999; 67(2): 885-90., Qin et al. 2011Qin QM, Luo J, Lin X, Pei J, Li L, Ficht TA, et al. Functional analysis of host factors that mediate the intracellular lifestyle of Cryptococcus neoformans. PLoS Pathog. 2011; 7(6): e1002078.). Additionally, phagosomes containing C. neoformans experience lysosomal fusion and acquire phagosomal markers, indicating that phagosomal maturation is not inhibited by this pathogen (Coelho et al. 2014Coelho C, Bocca AL, Casadevall A. The intracellular life of Cryptococcus neoformans. Annu Rev Pathol Mech Dis. 2014; 9: 219-38.). More recent work has demonstrated that several of these early markers are prematurely removed and that C. neoformans can subtly alter the phagosome maturation process in order to create a more permissive environment for its survival (Smith et al. 2015Smith LM, Dixon EF, May RC. The fungal pathogen Cryptococcus neoformans manipulates macrophage phagosome maturation. Cell Microbiol. 2015; 17(5): 702-13.). Through a screen to identify host factors that influence intracellular survival, it was shown that C. neoformans hijacks many aspects of macrophage biology, including cytoskeletal elements, cell surface signaling molecules, and vesicle mediated transport systems, to favor its own survival (Qin et al. 2011Qin QM, Luo J, Lin X, Pei J, Li L, Ficht TA, et al. Functional analysis of host factors that mediate the intracellular lifestyle of Cryptococcus neoformans. PLoS Pathog. 2011; 7(6): e1002078.). This study also showed that autophagy proteins are recruited to pathogen-containing vacuoles, supporting C. neoformans infection (Qin et al. 2011Qin QM, Luo J, Lin X, Pei J, Li L, Ficht TA, et al. Functional analysis of host factors that mediate the intracellular lifestyle of Cryptococcus neoformans. PLoS Pathog. 2011; 7(6): e1002078.). Once ingested, C. neoformans induces phagolysosomal damage (Feldmesser et al. 2000Feldmesser M, Kress Y, Novikoff P, Casadevall A. Cryptococcus neo-formans is a facultative intracellular pathogen in murine pulmonary infection. Infect Immun. 2000; 68(7): 4225-37., Tucker and Casadevall 2002Tucker SC, Casadevall A. Replication of Cryptococcus neoformans in macrophages is accompanied by phagosomal permeabilization and accumulation of vesicles containing polysaccharide in the cytoplasm. Proc Natl Acad Sci. 2002; 99(5): 3165-70., Davis et al. 2015Davis MJ, Eastman AJ, Qiu Y, Gregorka B, Kozel TR, Osterholzer JJ, et al. Cryptococcus neoformans - induced macrophage lysosome damage crucially contributes to fungal virulence. J Immunol. 2015; 194(5): 2219-31.), perhaps as a result of the combination of increased cell/capsule growth and secreted C. neoformans proteins, such as phospholipase B (Feldmesser et al. 2000Feldmesser M, Kress Y, Novikoff P, Casadevall A. Cryptococcus neo-formans is a facultative intracellular pathogen in murine pulmonary infection. Infect Immun. 2000; 68(7): 4225-37., Cox et al. 2001Cox GM, McDade HC, Chen SCA, Tucker SC, Gottfredsson M, Wright LC, et al. Extracellular phospholipase activity is a virulence factor for Cryptococcus neoformans. Mol Microbiol. 2001; 39(1): 166-75.). The damaged phagolysosomes display increased membrane permeability which enhances C. neoformans growth by allowing nutrient influx, pH homeostasis, and eventually escape from the macrophage (Davis et al. 2015Davis MJ, Eastman AJ, Qiu Y, Gregorka B, Kozel TR, Osterholzer JJ, et al. Cryptococcus neoformans - induced macrophage lysosome damage crucially contributes to fungal virulence. J Immunol. 2015; 194(5): 2219-31.).

Macrophage exit - Once inside of a macrophage, C. neoformans has several potential fates. The first is being inhibited or killed by the phagocyte. Other options, all of which ultimately lead to fungal escape, include lysis of the macrophage, cell-to-cell transfer to a neighbouring macrophage, and non-lytic exocytosis or “vomocytosis” in which both fungal cell and macrophage survive the interaction (Johnston and May 2013Johnston SA, May RC. Cryptococcus interactions with macrophages: evasion and manipulation of the phagosome by a fungal pathogen. Cell Microbiol. 2013; 15(3): 403-11., Coelho et al. 2014Coelho C, Bocca AL, Casadevall A. The intracellular life of Cryptococcus neoformans. Annu Rev Pathol Mech Dis. 2014; 9: 219-38., De LeónRodríguez and Casadevall 2016De León-Rodríguez CM, Casadevall A. Cryptococcus neoformans: tripping on acid in the phagolysosome. Front Microbiol. 2016; 7(2): 1-9.). Lateral transfer of C. neoformans from one macrophage to another, while a rare event, allows for fungal cells to disseminate while avoiding immune detection. Alvarez and Casadevall (2007)Alvarez M, Casadevall A. Cell-to-cell spread and massive vacuole formation after Cryptococcus neoformans infection of murine macrophages. BMC Immunol. 2007; 8(1): 16. demonstrated that this occurs in an actin-dependent manner, leaving lasting effects on the inhabited macrophage in the form of a large residual vacuole. This process can occur regardless of serotype or opsonisation type, and in multiple mammalian cell lines (Ma et al. 2007Ma H, Croudace JE, Lammas DA, May RC. Direct cell-to-cell spread of a pathogenic yeast. BMC Immunol. 2007; 8(1): 15.).

Non-lytic exocytosis, or vomocytosis, is similar to cell-to-cell spread in that both the fungal and macrophage cells are viable after fungal escape. This process has been shown to occur in vivo and is dependent on several host factors (Nicola et al. 2011Nicola AM, Robertson EJ, Albuquerque P, Derengowski LS, Casadevall A. Nonlytic exocytosis of Cryptococcus neoformans from macrophages occurs in vivo and is influenced by phagosomal pH. MBio. 2011; 2(4): pi: e00167-11.). It appears to occur after phagosome maturation and is influenced by phagosomal pH. For example, when the pH of the phagosome was raised artificially with weak bases, rates of vomocytosis increased (Ma et al. 2006Ma H, Croudace JE, Lammas DA, May RC. Expulsion of live pathogenic yeast by macrophages. Curr Biol. 2006; 16(21): 2156-60., Nicola et al. 2011Nicola AM, Robertson EJ, Albuquerque P, Derengowski LS, Casadevall A. Nonlytic exocytosis of Cryptococcus neoformans from macrophages occurs in vivo and is influenced by phagosomal pH. MBio. 2011; 2(4): pi: e00167-11.). Concordantly, when acidification of the phagosome was blocked altogether using vacuolar ATPase inhibitors, the rate of vomocytosis decreased (Ma et al. 2006Ma H, Croudace JE, Lammas DA, May RC. Expulsion of live pathogenic yeast by macrophages. Curr Biol. 2006; 16(21): 2156-60., Nicola et al. 2011Nicola AM, Robertson EJ, Albuquerque P, Derengowski LS, Casadevall A. Nonlytic exocytosis of Cryptococcus neoformans from macrophages occurs in vivo and is influenced by phagosomal pH. MBio. 2011; 2(4): pi: e00167-11.). Phagosomal membrane permeabilisation occurs rapidly after uptake of C. neoformans cells and is thought to be another contributing factor to rates of non-lytic exocytosis (Tucker and Casadevall 2002Tucker SC, Casadevall A. Replication of Cryptococcus neoformans in macrophages is accompanied by phagosomal permeabilization and accumulation of vesicles containing polysaccharide in the cytoplasm. Proc Natl Acad Sci. 2002; 99(5): 3165-70., Coelho et al. 2014Coelho C, Bocca AL, Casadevall A. The intracellular life of Cryptococcus neoformans. Annu Rev Pathol Mech Dis. 2014; 9: 219-38., Davis et al. 2015Davis MJ, Eastman AJ, Qiu Y, Gregorka B, Kozel TR, Osterholzer JJ, et al. Cryptococcus neoformans - induced macrophage lysosome damage crucially contributes to fungal virulence. J Immunol. 2015; 194(5): 2219-31.). Actin flashes around the phagosome occur soon after membrane permeabilisation and contribute to blocking non-lytic exocytosis (Johnston and May 2010Johnston SA, May RC. The human fungal pathogen Cryptococcus neoformans escapes macrophages by a phagosome emptying mechanism that is inhibited by Arp2/3 complex-mediated actin polymyerisation. PLoS Pathog. 2010; 6(8): 27-8.). It has also been demonstrated that cytokine signalling has an impact on this process, with Th2-stimulated macrophages having lower rates of non-lytic exocytosis (Voelz et al. 2009Voelz K, Lammas DA, May RC. Cytokine signaling regulates the outcome of intracellular macrophage parasitism by Cryptococcus neoformans. Infect Immun. 2009; 77(8): 3450-7.). In addition to host factors, C. neoformans proteins are also required for non-lytic exocytosis, including phospholipase B1 (Plb1) and the Sec14 protein required for phospholipase secretion (Chayakulkeeree et al. 2011Chayakulkeeree M, Johnston SA, Oei JB, Lev S, Williamson PR, Wilson CF, et al. SEC14 is a specific requirement for secretion of phospholipase B1 and pathogenicity of Cryptococcus neoformans. Mol Microbiol. 2011; 80(4): 1088-1101.).

Morphological changes in C. neoformans and their role during adaptation to the host

Hyphal formation - Many fungi undergo morphological changes during various stages of an infection, such as the transition among Candida species from a yeast-like form to hyphae and pseudohyphae. These filamentous structures are more adherent than blastoconidia, so they are involved in attachment, invasion and dissemination (reviewed in Trevijano-Contador et al. 2016Trevijano-Contador N, Rueda C, Zaragoza O. Fungal morphogenetic changes inside the mammalian host. Semin Cell Dev Biol. 2016; 57: 100-9.). In the case of C. neoformans, this yeast can only form hyphae during sexual reproduction (Casadevall and Perfect 1998Casadevall A, Perfect JR. Cryptococcus neoformans. Washington DC: ASM Press; 1998.), and true hyphae are not believed to significantly contribute to the development of the disease. In contrast, there are other types of morphological changes that can occur in the host that are more relevant to our understanding of the pathogenesis of this microorganism. For example, C. neoformans can form pseudohyphae and they can be occasionally observed in vivo (Lee et al. 2012Lee SC, Phadke S, Sun S, Heitman J. Pseudohyphal growth of Cryptococcus neoformans is a reversible dimorphic transition in response to ammonium that requires Amt1 and Amt2 ammonium permeases. Eukaryot Cell. 2012; 11(11): 1391-8., Magditch et al. 2012Magditch DA, Liu TB, Xue C, Idnurm A. DNA mutations mediate microevolution between host-adapted forms of the pathogenic fungus Cryptococcus neoformans. PLoS Pathog. 2012; 8(10): e1002936.), although their exact function in the adaptation of this yeast to the host remains unknown.

Titan cells - Although filamentous forms can be found in the host, the most well characterised mechanism developed by C. neoformans to adapt to the lung environment is its ability to increase its cells size. In fact, a significant feature of the cryptococcal population in vivo is its size heterogeneity, finding cells in vivo of very different diameters. Cellular enlargement can be achieved not only by capsule growth (which has been described above), but also by a significant increase in the size of the cell body, which leads to the appearance of rounded yeast cells of an abnormal size that can reach up to 100 microns (Okagaki et al. 2010Okagaki LH, Strain AK, Nielsen JN, Charlier C, Baltes NJ, Chrétien F, et al. Cryptococcal cell morphology affects host cell interactions and pathogenicity. PLoS Pathog. 2010; 6(6): e1000953., Zaragoza et al. 2010Zaragoza O, García-Rodas R, Nosanchuk JD, Cuenca-Estrella M, Rodríguez-Tudela JL, Casadevall A. Fungal cell gigantism during mammalian infection. PLoS Pathog. 2010; 6(6): e1000945.). These forms have been termed titan cells due to their huge size (Zaragoza and Nielsen 2013Zaragoza O, Nielsen K. Titan cells in Cryptococcus neoformans: cells with a giant impact. Curr Opin Microbiol. 2013; 16(4): 409-13.). The signals that induce the massive cellular enlargement are unknown. The main intracellular pathway involved in this process depends on cAMP and PKA signaling (Zaragoza et al. 2010Zaragoza O, García-Rodas R, Nosanchuk JD, Cuenca-Estrella M, Rodríguez-Tudela JL, Casadevall A. Fungal cell gigantism during mammalian infection. PLoS Pathog. 2010; 6(6): e1000945.), and several effectors upstream (such as pheromone receptors and Gpr5) and downstream (Rim101) are required for cell growth (Okagaki et al. 2011Okagaki LH, Wang Y, Ballou ER, O’Meara TR, Bahn Y-S, Alspaugh JA, et al. Cryptococcal titan cell formation is regulated by Gprotein signaling in response to multiple stimuli. Eukaryot Cell. 2011; 10(10): 1306-16.). As a consequence, there are alterations in cell cycle regulation that result in genome endoduplication and polyploidy (Okagaki et al. 2010Okagaki LH, Strain AK, Nielsen JN, Charlier C, Baltes NJ, Chrétien F, et al. Cryptococcal cell morphology affects host cell interactions and pathogenicity. PLoS Pathog. 2010; 6(6): e1000953., Zaragoza et al. 2010Zaragoza O, García-Rodas R, Nosanchuk JD, Cuenca-Estrella M, Rodríguez-Tudela JL, Casadevall A. Fungal cell gigantism during mammalian infection. PLoS Pathog. 2010; 6(6): e1000945.). In addition, the capsule of these cells is also very large and composed of a net of polysaccharide fibres that form a structure that is denser then that observed with cells of normal size (Zaragoza et al. 2010Zaragoza O, García-Rodas R, Nosanchuk JD, Cuenca-Estrella M, Rodríguez-Tudela JL, Casadevall A. Fungal cell gigantism during mammalian infection. PLoS Pathog. 2010; 6(6): e1000945.).

The role of titan cells in cryptococcal disease remains to be fully elucidated, however, their involvement in several processes that contribute to immune evasion and long-term persistence has been demonstrated. Titan cells cannot be phagocytosed presumably due to their size, as it was demonstrated that similarly sized synthetic particles could not be readily engulfed by lung phagocytes (Okagaki and Nielsen 2012Okagaki LH, Nielsen K. Titan cells confer protection from phagocytosis in Cryptococcus neoformans infections. Eukaryot Cell. 2012; 11(6): 820-6.). Interestingly, titan cells are also able to confer this phagocytosis resistance to neighbouring, smaller yeast cells (Okagaki and Nielsen 2012Okagaki LH, Nielsen K. Titan cells confer protection from phagocytosis in Cryptococcus neoformans infections. Eukaryot Cell. 2012; 11(6): 820-6.). The exact mechanism by which titan cells are able to provide collateral protection to neighbouring cells has not been precisely defined. However, it has been demonstrated that polyploid cells can produce a variety of haploid and aneuploid daughter cells, promoting rapid adaptations to stress conditions (Gerstein et al. 2015Gerstein AC, Fu MS, Mukaremera L, Li Z, Ormerod KL, Fraser JA, et al. Polyploid titan cells produce haploid and aneuploid progeny to promote stress adaptation. MBio. 2015; 6(5): e01340-15.). Given the extensive surface capsule of titan cells, it is also plausible that secreted exopolysaccharide may influence the surrounding environment.

The signals that trigger titan cell production are unknown. It was first described that co-infection of mice with opposing mating type cells resulted in a significant increase in the proportion of titan cells (Okagaki et al. 2010Okagaki LH, Strain AK, Nielsen JN, Charlier C, Baltes NJ, Chrétien F, et al. Cryptococcal cell morphology affects host cell interactions and pathogenicity. PLoS Pathog. 2010; 6(6): e1000953.), suggesting that the pheromone signalling pathway is required for this transition. Furthermore, titan cell production is strongly dependent on the host environment, and the percentage of these cells observed in vivo varies in different mouse strains. In particular, in mice that develop a Th1 type response (dependent on interferon-y and TNF-α), the proportion of titan cells is low (around 15%). In contrast, in mice that induce Th2 type responses, the proportion of titan cells is very high, even above 50% of the total population of cryptococcal cells (García-Barbazán et al. 2016García-Barbazán I, Trevijano-Contador N, Rueda C, de Andrés B, Pérez-Tavárez R, Herrero-Fernández I, et al. The formation of titan cells in Cryptococcus neoformans depends on the mouse strain and correlates with induction of Th2-type responses. Cell Microbiol. 2016; 18(1): 111-24.). At the moment, the exact correlation between the host immune response and cryptococcal morphology is unknown, but it is hypothesised that Th2 type responses result in a less aggressive environment that facilitates cellular enlargement.

Cell wall rearrangements during infection - In addition to these well-characterised morphological changes, there is increasing evidence that C. neoformans cell wall maintenance plays an important role in its interaction with the host immune system. C. neoformans dramatically alters its cell wall, both in size and composition in response to the host environment (Feldmesser et al. 2001Feldmesser M, Kress Y, Casadevall A. Dynamic changes in the morphology of Cryptococcus neoformans during murine pulmonary infection. Microbiology. 2001; 147: 2355-65., O’Meara et al. 2013O’Meara TR, Holmer SM, Selvig K, Dietrich F, Alspaugh JA. Cryptococcus neoformans Rim101 is associated with cell wall remodeling and evasion of the host immune responses. MBio. 2013; 4(1): 1-13., O’Meara et al. 2014O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84., Ost et al. 2017Ost KS, Esher SK, Wager CML, Walker L, Wagener J, Munro C, et al. Rim pathway-mediate alterations in the fungal cell wall influence immune recognition and inflammation. MBio. 2017; 8(1): e02290-16.). Feldmesser et al (2001)Feldmesser M, Kress Y, Casadevall A. Dynamic changes in the morphology of Cryptococcus neoformans during murine pulmonary infection. Microbiology. 2001; 147: 2355-65. demonstrated that the cell wall thickens over time in the setting of murine pulmonary infection. Additionally, both capsule and titan cell formation involve significant cell wall remodelling; polysaccharide capsule attaches to the cell surface through an interaction with a-(1,3)-glucan (Reese and Doering 2003Reese AJ, Doering TL. Cell wall a-1,3-glucan is required to anchor the Cryptococcus neoformans capsule. Mol Microbiol. 2003; 50(4): 1401-9.), and titan cells have thicker, more chitin-rich cell walls (Wiesner et al. 2015Wiesner DL, Specht CA, Lee CK, Smith KD, Mukaremera L, Lee ST, et al. Chitin recognition via chitotriosidase promotes pathologic type-2 helper T cell responses to cryptococcal infection. PLoS Pathog. 2015; 11(3): e1004701.). Studies investigating the Rim101 transcriptome during murine infection indicated that this pH responsive transcription factor directly regulates cell wall biosynthesis genes in this context (O’Meara et al. 2013O’Meara TR, Holmer SM, Selvig K, Dietrich F, Alspaugh JA. Cryptococcus neoformans Rim101 is associated with cell wall remodeling and evasion of the host immune responses. MBio. 2013; 4(1): 1-13., O’Meara et al. 2014O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84.). Coordinately, in the absence of Rim101, C. neoformans cells expose immunogenic epitopes that ultimately lead to detrimental immune responses (Feldmesser et al. 2001Feldmesser M, Kress Y, Casadevall A. Dynamic changes in the morphology of Cryptococcus neoformans during murine pulmonary infection. Microbiology. 2001; 147: 2355-65., O’Meara et al. 2013O’Meara TR, Holmer SM, Selvig K, Dietrich F, Alspaugh JA. Cryptococcus neoformans Rim101 is associated with cell wall remodeling and evasion of the host immune responses. MBio. 2013; 4(1): 1-13., O’Meara et al. 2014O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84.). These studies highlight how C. neoformans actively remodels its cell surface in response to the host environment in order to avoid immune detection.

THE TRIP CONTINUES: DISSEMINATION THROUGH THE ORGANISM AND ARRIVAL TO THE BRAIN. DIVING OR SAILING?

Although cryptococcal cells are mainly acquired by inhalation, a key step in disease is the dissemination from the lung to the brain, where it causes the most characteristic clinical manifestation of cryptococcal disease, meningoencephalitis. For this reason, the mechanisms of migration of C. neoformans to the brain have been extensively studied (see seminal review in Griffiths et al. 2012Griffiths EJ, Kretschmer M, Kronstad JW. Aimless mutants of Cryptococcus neoformans: failure to disseminate. Fungal Biol Rev. 2012; 26(2-3): 61-72.). This dissemination occurs through the blood vessels, so cryptococcal cells must cross both epithelial and endothelial barriers to transit from the lung alveoli to the bloodstream, and ultimately the CNS.

The first barrier that C. neoformans faces during dissemination is composed of the epithelial cells from the lung, although this interaction has not been characterised in detail. It has been described that both encapsulated and acapsular cells can interact with human lung epithelial cells, with acapsular mutants being able to recognise and attach to this epithelial layer with greater affinity (Merkel and Scofield 1997Merkel GJ, Scofield BA. The in vitro interaction of Cryptococcus neoformans with human lung epithelial cells. FEMS Immunol Med Microbiol. 1997; 19(3): 203-13.). As a result, C. neoformans cells can be internalised by epithelial cells, leading to the death of the host cell. In the case of regular encapsulated cells, the capsular polysaccharide, GXM, plays a major role in the recognition by epithelial cells, and this binding seems to depend on the CD14 receptor. In addition, other cryptococcal proteins, such as the mannoprotein MP84 or phospholipase B, also seem to contribute to epithelial cell binding (Ganendren et al. 2006Ganendren R, Carter E, Sorrell T, Widmer F, Wright L. Phospholipase B activity enhances adhesion of Cryptococcus neoformans to a human lung epithelial cell line. Microbes Infect. 2006; 8(4): 1006-15., Teixeira et al. 2014Teixeira PA, Penha LL, Mendonça-Previato L, Previato JO. Mannoprotein MP84 mediates the adhesion of Cryptococcus neoformans to epithelial lung cells. Front Cell Infect Microbiol. 2014; 4: 106.).

Of particular interest is the interaction of C. neoformans with endothelial cells, particularly those comprising the blood-brain barrier (BBB). Although the BBB selectively protects the brain from extracellular particles, C. neoformans has developed ways to cross this restrictive barrier, both as free-living fungal cells and intracellularly inside macrophages. Elegant real-time in vivo imaging experiments have demonstrated that isolated fungal cells can directly attach to the endothelial surface of the brain microvasculature as the initial step in breaching the BBB (Shi et al. 2010Shi M, Li SS, Zheng C, Jones GJ, Kim KS, Zhou H, et al. Real-time imaging of trapping and urease-dependent transmigration of Cryptococcus neoformans in mouse brain. J Clin Invest. 2010; 120(5): 1683-93.). Cryptococcal cells can subsequently be internalised by the endothelial cells at the apical side and then released at the basolateral side (Chen et al. 2003Chen SH, Stins MF, Huang S-H, Chen YH, Kwon-Chung KJ, Chang Y, et al. Cryptococcus neoformans induces alterations in the cytoskeleton of human brain microvascular endothelial cells. J Med Microbiol. 2003; 52: 961-70.). In this process, it has been shown that hyaluronic acid (HA) present in C. neoformans can be recognised by the CD44 receptor from endothelial cells, suggesting a process of endothelial cell interaction that his conserved among several microbial neuropathogens (Jong et al. 2008Jong A, Wu C, Shackleford GM, Kwon-Chung KJ, Chang YC, Chen H, et al. Involvement of human CD44 during Cryptococcus neo-formans infection of brain microvascular endothelial cells. Cell Microbiol. 2008; 10(6): 1313-26.). Interestingly, inositol produced by the host cells is recognised by C. neoformans, a process that results in an increased production of HA by the fungus (Liu et al. 2013Liu T, Kim J, Wang Y, Toffaletti D. Brain inositol is a novel stimulator for promoting Cryptococcus penetration of the blood-brain barrier. PLoS Pathog. 2013; 9(4): e1003247.). Other cryptococcal elements, such as urease (Olszewski et al. 2004Olszewski MA, Noverr MC, Chen GH, Toews GB, Cox GM, Perfect JR, et al. Urease Expression by Cryptococcus neoformans promotes microvascular sequestration, thereby enhancing central nervous system invasion. Am J Pathol. 2004; 164(5): 1761-71.), phospholipase B (Santangelo et al. 2004Santangelo R, Zoellner H, Sorrell T, Wilson C, Donald C, Djordjevic J, et al. Role of extracellular phospholipases and mononuclear phagocytes in dissemination of cryptococcosis in a murine model. Infect Immun. 2004; 72(4): 2229-39.), and the extracellular protease Mpr1 (Na Pombejra et al. 2017Na Pombejra S, Salemi M, Phinney BS, Gelli A. The metalloprotease, Mpr1, engages AnnexinA2 to promote the transcytosis of fungal cells across the blood-brain barrier. Front Cell Infect Microbiol. 2017; 7: 296.) have been shown to be involved in the process of binding to endothelial cells. In this last case, the Mpr1 protease induces cytoskeleton rearrangements in the endothelial cells and promotes recognition of C. neoformans by Annexin A2. Internalisation of cryptococcal cells by the BBB is also associated with multiple changes in the endothelial cells including rearrangements of the cytoskeleton and changes in the morphology of nuclei, endoplasmic reticulum and mitochondria (Vu et al. 2013Vu K, Eigenheer RA, Phinney BS, Gelli A. Cryptococcus neoformans promotes its transmigration into the central nervous system by inducing molecular and cellular changes in brain endothelial cells. Infect Immun. 2013; 81(9): 3139-47.).

C. neoformans can also alter the structure of the tight junctions that attach the cells of the BBB (Olszewski et al. 2004Olszewski MA, Noverr MC, Chen GH, Toews GB, Cox GM, Perfect JR, et al. Urease Expression by Cryptococcus neoformans promotes microvascular sequestration, thereby enhancing central nervous system invasion. Am J Pathol. 2004; 164(5): 1761-71., Charlier et al. 2005Charlier C, Chretien F, Baudrimont M, Mordelet E, Lortholary O, Dromer F. Capsule structure changes associated with Cryptococcus neoformans crossing of the blood-brain barrier. Am J Pathol. 2005; 166(2): 421-32., Vu et al. 2013Vu K, Eigenheer RA, Phinney BS, Gelli A. Cryptococcus neoformans promotes its transmigration into the central nervous system by inducing molecular and cellular changes in brain endothelial cells. Infect Immun. 2013; 81(9): 3139-47.), so it has been suggested that C. neoformans can also transverse the BBB through a paracellular mechanism. In this sense, some addictive drugs (such as methamphetamine) that alter the structure of the BBB tight junctions increase the dissemination of cryptococcal cells to the brain (Eugenin et al. 2013Eugenin EA, Greco JM, Frases S, Nosanchuk JD, Martinez LR. Methamphetamine alters blood brain barrier protein expression in mice, facilitating central nervous system infection by neurotropic Cryptococcus neoformans. J Infect Dis. 2013; 208(4): 699-704.).

There is also strong evidence that C. neoformans can cross the BBB inside phagocytic cells, through a process that is widely known as the “Trojan-horse” dissemination mechanism. This idea was first suggested when it was found that C. neoformans can survive inside phagocytic cells. In the last few years there has been increasing evidence that this dissemination mechanism occurs in vivo. Several elegant studies have demonstrated that macrophages have a paradoxical role during infection because their depletion has a protective role during cryptococcosis and results in reduced fungal burden in brain, lung and spleen (Kechichian et al. 2007Kechichian TB, Shea J, Del Poeta M. Depletion of alveolar macrophages decreases the dissemination of a glucosylceramide-deficient mutant of Cryptococcus neoformans in immunodeficient mice. Infect Immun. 2007; 75(10): 4792-8., Charlier et al. 2009Charlier C, Nielsen K, Daou S, Brigitte M, Chretien F, Dromer F. Evidence of a role for monocytes in dissemination and brain invasion by Cryptococcus neoformans. Infect Immun. 2009; 77(1): 120-7.), suggesting that in fact these phagocytic cells offer a “safe” niche for this fungus and contribute to dissemination. In agreement, when mice are injected with bone marrow-derived monocytes infected with C. neoformans, the fungal burden in target organs is higher compared to infection with the equivalent dose of free living yeasts (Santangelo et al. 2004Santangelo R, Zoellner H, Sorrell T, Wilson C, Donald C, Djordjevic J, et al. Role of extracellular phospholipases and mononuclear phagocytes in dissemination of cryptococcosis in a murine model. Infect Immun. 2004; 72(4): 2229-39., Charlier et al. 2009Charlier C, Nielsen K, Daou S, Brigitte M, Chretien F, Dromer F. Evidence of a role for monocytes in dissemination and brain invasion by Cryptococcus neoformans. Infect Immun. 2009; 77(1): 120-7.). Further evidence has been provided in vitro using models of reconstituted BBB. In these experiments, C. neoformans can transmigrate across an in vitro-generated BBB via transcellular pores when transported inside macrophages (Sorrell et al. 2016Sorrell TC, Juillard P-G, Djordjevic JT, Kaufman-Francis K, Dietmann A, Milonig A, et al. Cryptococcal transmigration across a model brain blood-barrier: evidence of the Trojan horse mechanism and differences between Cryptococcus neoformans var. grubii strain H99 and Cryptococcus gattii strain R265. Microbes Infect. 2016; 18(1): 57-67., Santiago-Tirado et al. 2017Santiago-Tirado FH, Onken MD, Cooper JA, Klein RS, Doering TL. Trojan horse transit contributes to blood-brain barrier crossing of a eukaryotic pathogen. mBio. 2017; 8(1): e02183-16.). Santiago-Tirado et al. (2017)Santiago-Tirado FH, Onken MD, Cooper JA, Klein RS, Doering TL. Trojan horse transit contributes to blood-brain barrier crossing of a eukaryotic pathogen. mBio. 2017; 8(1): e02183-16. further demonstrated that, during this process, several outcomes of the interaction of C. neoformans and macrophages occur, such as fungal replication, non-lytic exocytosis and cell-to-cell transmission of fungal cells. In addition, these authors also observed direct transmission of cryptococcal cells from macrophages to endothelial cells, which suggests that the same fungal cell can transmigrate the BBB through several mechanisms (“Trojan horse” approach for dissemination through the blood stream, and paracellularly through endothelial cells as free cells).

In summary, there is strong evidence that C. neoformans can disseminate and colonise the brain through different mechanisms, although at the moment it is not known the relative contribution of each mechanism (as free yeasts or inside phagocytic cells). Due to the importance of this process for cryptococcal disease, further work is required to characterise this process and envision therapeutic strategies to control brain invasion by this fungus.

FINAL STOP OF THE TRIP: SURVIVAL WITHIN THE CNS

Although survival in the lung and dissemination are key aspects to understand cryptococcal disease, the mechanisms that allow survival in the brain are also very important to define since the most common clinical manifestation of cryptococcal disease is brain infection (Colombo and Rodrigues 2015Colombo AC, Rodrigues ML. Fungal colonization of the brain: anatomopathological aspects of neurological cryptococcosis. An Acad Bras Cienc. 2015; 87(2): 1293-1309.). Once C. neoformans has invaded the CNS, the clinical manifestations of the real tissue (meningitis) as well as from involvement of the brain tissue itself (encephalitis). Therefore, the symptoms of cryptococcal meningoencephalitis can range from a progressive headache to serious neurological symptoms, including coma and death. Moreover, the viscous capsular polysaccharide of this microorganism can trigger increased intracranial pressure, a major source of morbidity in this infection that must be treated aggressively.

Several investigators have explored how this organism is able to survive in the nutrient-poor environment of the cerebrospinal fluid, as well as in the specialised neural tissue. C. neoformans is able to grow in vitro on a very minimal medium composed primarily of cerebrospinal fluid (Chen et al. 2014Chen Y, Toffaletti DL, Tenor JL, Litvintseva AP, Fang C, Mitchell TG, et al. The Cryptococcus neoformans transcriptome at the site of human meningitis. MBio. 2014; 5(1): 1-10.). This fluid has a low carbohydrate and nitrogen content, suggesting that this fungus effectively scavenges essential nutrients and their precursors from nutrient-poor environments.

Transcription patterns of C. neoformans isolated directly from the CNS of infected patients were compared with samples incubated ex vivo on CSF media (Chen et al. 2014Chen Y, Toffaletti DL, Tenor JL, Litvintseva AP, Fang C, Mitchell TG, et al. The Cryptococcus neoformans transcriptome at the site of human meningitis. MBio. 2014; 5(1): 1-10.). Carbohydrate importers and the sodium transporter Ena1 are highly induced in both conditions. Interestingly, the inositol transporter gene family is specifically required for C. neoformans penetration of the blood brain barrier (Liu et al. 2013Liu T, Kim J, Wang Y, Toffaletti D. Brain inositol is a novel stimulator for promoting Cryptococcus penetration of the blood-brain barrier. PLoS Pathog. 2013; 9(4): e1003247.). Inositol is present in high concentration in the brain, suggesting an interesting potential targeting mechanisms for this neuropathogen to the CNS (Liu et al. 2013Liu T, Kim J, Wang Y, Toffaletti D. Brain inositol is a novel stimulator for promoting Cryptococcus penetration of the blood-brain barrier. PLoS Pathog. 2013; 9(4): e1003247.). Additionally, the alkaline-responsive Rim101 transcription factor was also highly induced during CNS infection. Together, these results suggest that C. neoformans must actively adapt to host-specific signals while growing in the CNS, including nutrient deprivation and host pH.

Several lines of evidence implicate a role for laccase activity during CNS infections. First, laccase mutants are highly attenuated for virulence in animal models of cryptococcal infection (Noverr et al. 2004Noverr MC, Williamson PR, Fajardo RS, Huffnagle GB. CNLAC1 is required for extrapulmonary dissemination of Cryptococcus neoformans but not pulmonary persistence. Infect Immun. 2004; 72(3): 1693-9.+). Also, melanised forms of C. neoformans can be isolated from CNS tissue during infection (Nosanchuk et al. 2000Nosanchuk JD, Rosas AL, Lee SC, Casadevall A. Melanisation of Cryptococcus neoformans in human brain tissue. Lancet. 2000; 355(9220): 2049-50.). The transcript levels of cryptococcal laccases are specifically induced by glucose deprivation, a condition known to be present in the CNS (Williamson 1994Williamson PR. Biochemical and molecular characterization of the diphenol oxidase of Cryptococcus neoformans: identification as a laccase. J Bacteriol. 1994; 176(3): 656-64.). Additionally, the substrates for cryptococcal melanin formation include diphenolic compounds such as epinephrine, DOPA, and norepinephrine (Williamson et al. 1998Williamson PR, Wakamatsu K, Ito S. Melanin biosynthesis in Cryptococcus neoformans. J Bacteriol. 1998; 180(6): 1570-2.). The enhanced availability of these diphenolic neurotransmitter molecules in neural tissue has been postulated to be one reason for the neurotropism of this microorganism.

SUMMARY

Cryptococcus neoformans continues to be a significant pathogen among immunocompromised individuals, especially those with advanced HIV infection. As an environmental fungus, this organism has adapted many strategies to survive its trip to disease in the mammalian host. Suggested to have acquired many of its virulence traits from environmental encounters, C. neoformans has been referred to as an “accidental pathogen”.

Beginning its journey in the environment, this fungus can interact and infect many soil microbes, and during its interaction with these microbes, C. neoformans utilises many of its classical virulence attributes. This fungus is introduced into the mammalian host through the inhalation of spores or desiccated yeast. With advanced methods of spore isolation recently described, the role of spores at this initial step has been able to be elucidated more clearly. However, future work will be required to characterise the innate immune responses to these infectious propagules, and how these responses direct the development of disease.

As C. neoformans is inhaled into the mammalian lung, it must adapt to a number of additional stresses including high temperature, increased pH, and changes in essential nutrients and metal concentrations. Ongoing work continues to identify novel upstream and downstream components of the conserved signalling pathways controlling responses to these stresses. This fungus has a dynamic relationship with host phagocytes, in which it actively avoids detection and killing by these cells, but it also requires them for effective CNS dissemination. Inside the host, C. neoformans has also developed ways to alter its morphology in order to facilitate survival, including the production of polysaccharide capsule, titan cell formation, and cell wall rearrangement. Continued efforts to understand this delicate host-pathogen interface will be needed to drive the development of novel methods to direct this response in favour of the host.

Finally, in order to effectively finish its journey to the CNS, this fungus has the ability to traverse the BBB through various means. These include direct traversal through endothelial cells, manipulation of the tight junctions of the BBB, and the “Trojan-horse” mechanism. A greater understanding of how C. neoformans utilises these different means in vivo will provide a path forward for developing new therapeutic targets to control brain invasion.

  • SKE and JAA were supported by NIH (grants R01 AI074677, R01 GM121458-01 and P01 AI104533); JAA was a Burroughs Wellcome Fund New Investigator in Molecular Pathogenic Mycology; OZ is funded by grant SAF2014-5366R from the Spanish Ministry for Economics, Industry and Competitivity and MPY1356/15 from the Instituto de Salud Carlos III.

REFERENCES

  • Aksenov SI, Babyeva IP, Golubev VI. On the mechanism of adaptation of microorganisms to conditions of extreme low humidity. Life Sci Space Res. 1973; 11: 55-61.
  • Alspaugh JA, Cavallo LM, Perfect JR, Heitman J. RAS1 regulates filamentation, mating and growth at high temperature of Cryptococcus neoformans. Mol Microbiol. 2000; 36(2): 352-65.
  • Alvarez M, Casadevall A. Cell-to-cell spread and massive vacuole formation after Cryptococcus neoformans infection of murine macrophages. BMC Immunol. 2007; 8(1): 16.
  • Araújo GR, Fontes GN, Leão D, Rocha GM, Pontes B, Sant’Anna C, et al. Cryptococcus neoformans capsular polysaccharides form branched and complex filamentous networks viewed by high-resolution microscopy. J Struct Biol. 2016; 193(1): 75-82.
  • Baddley JW, Perfect JR, Oster RA, Larsen RA, Pankey GA, Henderson H, et al. Pulmonary cryptococcosis in patients without HIV infection: factors associated with disseminated disease. Eur J Clin Microbiol Infect Dis. 2008; 27(10): 937-43.
  • Bahn Y, Kojima K. Specialization of the HOG pathway and its impact on differentiation and virulence of Cryptococcus neoformans Mol Biol Cell. 2005; 16(5): 2285-2300.
  • Bahn YS, Geunes-Boyer S, Heitman J. Ssk2 mitogen-activated protein kinase kinase kinase governs divergent patterns of the stress-activated HOG1 signaling pathway in Cryptococcus neoformans Eukaryot Cell. 2007; 6(12): 2278-89.
  • Barluzzi R, Saleppico S, Nocentini A, Boelaert JR, Neglia R, Bistoni F, et al. Iron overload exacerbates experimental meningoencephalitis by Cryptococcus neoformans J Neuroimmunol. 2002; 132(1-2): 140-6.
  • Biondo C, Midiri A, Messina L, Tomasello F, Garufi G, Catania MR, et al. MyD88 and TLR2, but not TLR4, are required for host defense against Cryptococcus neoformans Eur J Immunol. 2005; 35(3): 870-8.
  • Bolanos B, Mitchell TG. Phagocytosis of Cryptococcus neoformans by rat alveolar macrophages. J Med Vet Mycol. 1989; 27(4): 203-17.
  • Bose I, Reese AJ, Ory JJ, Janbon G, Doering TL. A yeast under cover: the capsule of Cryptococcus neoformans Eukaryot Cell. 2003; 2(4): 655-63.
  • Botts MR, Giles SS, Gates MA, Kozel TR, Hull CM. Isolation and characterization of Cryptococcus neoformans spores reveal a critical role for capsule biosynthesis genes in spore biogenesis. Eukaryot Cell. 2009; 8(4): 595-605.
  • Bratton EW, El Husseini N, Chastain CA, Lee MS, Poole C, Stürmer T, et al. Comparison and temporal trends of three groups with cryptococcosis: HIV-infected, solid organ transplant, and HIVnegative/non-transplant. PLoS One. 2012; 7(8): e43582.
  • Brown GD, Denning DW, Gow NAR, Levitz SM, Netea MG, White TC. Hidden killers: human fungal infections. Sci Transl Med. 2012; 4(165): 165rv13.
  • Byrnes EJ, Bartlett KH, Perfect JR, Heitman J. Cryptococcus gattii: an emerging fungal pathogen infecting humans and animals. Microbes Infect. 2011; 13(11): 895-907.
  • Campuzano A, Castro-Lopez N, Wozniak KL, Wager CML, Wormley FL. Dectin-3 Is not required for protection against Cryptococcus neoformans infection. PLoS One. 2017; 12(1): e0169347.
  • Casadevall A, Perfect JR. Cryptococcus neoformans Washington DC: ASM Press; 1998.
  • Casadevall A, Pirofski LA. Accidental virulence, cryptic pathogenesis, martians, lost hosts, and the pathogenicity of environmental microbes. Eukaryot Cell. 2007; 6(12): 2169-74.
  • Chang YC, Kwon-Chung KJ. Complementation of a capsule-deficient mutation of Cryptococcus neoformans restores its virulence. Mol Cell Biol. 1994; 14(7): 4912-9.
  • Charlier C, Chretien F, Baudrimont M, Mordelet E, Lortholary O, Dromer F. Capsule structure changes associated with Cryptococcus neoformans crossing of the blood-brain barrier. Am J Pathol. 2005; 166(2): 421-32.
  • Charlier C, Nielsen K, Daou S, Brigitte M, Chretien F, Dromer F. Evidence of a role for monocytes in dissemination and brain invasion by Cryptococcus neoformans. Infect Immun. 2009; 77(1): 120-7.
  • Chayakulkeeree M, Johnston SA, Oei JB, Lev S, Williamson PR, Wilson CF, et al. SEC14 is a specific requirement for secretion of phospholipase B1 and pathogenicity of Cryptococcus neoformans Mol Microbiol. 2011; 80(4): 1088-1101.
  • Chen SH, Stins MF, Huang S-H, Chen YH, Kwon-Chung KJ, Chang Y, et al. Cryptococcus neoformans induces alterations in the cytoskeleton of human brain microvascular endothelial cells. J Med Microbiol. 2003; 52: 961-70.
  • Chen Y, Litvintseva AP, Frazzitta AE, Haverkamp MR, Wang L, Fang C, et al. Comparative analyses of clinical and environmental populations of Cryptococcus neoformans in Botswana. Mol Ecol. 2015; 24(14): 3559-71.
  • Chen Y, Toffaletti DL, Tenor JL, Litvintseva AP, Fang C, Mitchell TG, et al. The Cryptococcus neoformans transcriptome at the site of human meningitis. MBio. 2014; 5(1): 1-10.
  • Chiapello LS, Aoki MP, Rubinstein HR, Masih DT. Apoptosis induction by glucuronoxylomannan of Cryptococcus neoformans. Med Mycol. 2003; 41(4): 347-53.
  • Chowdhary A, Rhandhawa HS, Prakash A, Meis JF. Environmental prevalence of Cryptococcus neoformans and Cryptococcus gattii in India: an update. Crit Rev Microbiol. 2012; 38(1): 1-16.
  • Chrisman CJ, Albuquerque P, Guimaraes AJ, Nieves E, Casadevall A. Phospholipids trigger Cryptococcus neoformans capsular enlargement during interactions with amoebae and macrophages. PLoS Pathog. 2011; 7(5): e1002047.
  • Cleare W, Casadevall A. The different binding patterns of two immunoglobulin M monoclonal antibodies to Cryptococcus neoformans serotype A and D strains correlate with serotype classification and differences in functional assays. Clin Diagn Lab Immunol. 1998; 5(2): 125-9.
  • Coelho C, Bocca AL, Casadevall A. The intracellular life of Cryptococcus neoformans Annu Rev Pathol Mech Dis. 2014; 9: 219-38.
  • Colombo AC, Rodrigues ML. Fungal colonization of the brain: anatomopathological aspects of neurological cryptococcosis. An Acad Bras Cienc. 2015; 87(2): 1293-1309.
  • Cordero RJB, Frases S, Guimarães AJ, Rivera J, Casadevall A. Evidence for branching in cryptococcal capsular polysaccharides and consequences on its biological activity. Mol Microbiol. 2011; 79(4): 1101-17.
  • Cox GM, Harrison TS, McDade HC, Taborda CP, Heinrich G, Casadevall A, et al. Superoxide dismutase influences the virulence of Cryptococcus neoformans by affecting growth within macrophages. Infect Immun. 2003; 71(1): 173-80.
  • Cox GM, McDade HC, Chen SCA, Tucker SC, Gottfredsson M, Wright LC, et al. Extracellular phospholipase activity is a virulence factor for Cryptococcus neoformans Mol Microbiol. 2001; 39(1): 166-75.
  • Cross CE, Bancroft GJ. Ingestion of acapsular Cryptococcus neoformans occurs via mannose and β- glucan receptors, resulting in cytokine production and increased phagocytosis of the encapsulated form. Infect Immun. 1995; 63(7): 2604-11.
  • Dan JM, Kelly RM, Lee CK, Levitz SM. Role of the mannose receptor in a murine model of Cryptococcus neoformans infection. Infect Immun. 2008; 76(6): 2362-7.
  • Davis MJ, Eastman AJ, Qiu Y, Gregorka B, Kozel TR, Osterholzer JJ, et al. Cryptococcus neoformans - induced macrophage lysosome damage crucially contributes to fungal virulence. J Immunol. 2015; 194(5): 2219-31.
  • De León-Rodríguez CM, Casadevall A. Cryptococcus neoformans: tripping on acid in the phagolysosome. Front Microbiol. 2016; 7(2): 1-9.
  • Del Poeta M. Role of phagocytosis in the virulence of Cryptococcus neoformans. Eukaryot Cell. 2004; 3(5): 1067-75.
  • Denham ST, Verma S, Reynolds RC, Worne CL, Daugherty JM, Lane TE, et al. Regulated release of cryptococcal polysaccharide drives virulence and suppresses immune infiltration into the central nervous system. Infect Immun. 2017; pii: IAI.00662-17.
  • Ding C, Festa RA, Chen YL, Espart A, Palacios Ò, Espín J, et al. Cryptococcus neoformans copper detoxification machinery is critical for fungal virulence. Cell Host Microbe. 2013; 13(3): 265-76.
  • Ding C, Yin J, Tovar EMM, Fitzpatrick DA, Higgins DG, Thiele DJ. The copper regulon of the human fungal pathogen Cryptococcus neoformans H99. Mol Microbiol. 2011; 81(6): 1560-76.
  • Doering TL. How does Cryptococcus get its coat? Trends Microbiol. 2000; 8(12): 547-53.
  • Doering TL. How sweet it is! Cell wall biogenesis and polysaccharide capsule formation in Cryptococcus neoformans Annu Rev Microbiol. 2009; 63: 223-47.
  • Dong ZM, Jackson L, Murphy JW. Mechanisms for induction of Lselectin loss from T lymphocytes by a cryptococcal polysaccharide, glucuronoxylomannan. Infect Immun. 1999; 67(1): 220-9.
  • Dong ZM, Murphy JW. Intravascular cryptococcal culture filtrate (CneF) and its major component, glucuronoxylomannan, are potent inhibitors of leukocyte accumulation. Infect Immun. 1995; 63(3): 770-8.
  • Ellerbroek PM, Hoepelman AIM, Wolbers F, Zwaginga JJ, Coenjaerts FEJ. Cryptococcal glucuronoxylomannan inhibits adhesion of neutrophils to stimulated endothelium in vitro by affecting both neutrophils and endothelial cells. Infect Immun. 2002; 70(9): 4762-71.
  • Ellis DH, Pfeiffer TJ. Natural habitat of Cryptococcus neoformans var. gattii J Clin Microbiol. 1990; 28(7): 1642-4.
  • Emmons CW. Saprophytic sources of Cryptococcus neoformans associated with the pigeon (Columbia livia). Am J Epidemiol. 1955; 62(3): 227-32.
  • Eugenin EA, Greco JM, Frases S, Nosanchuk JD, Martinez LR. Methamphetamine alters blood brain barrier protein expression in mice, facilitating central nervous system infection by neurotropic Cryptococcus neoformans J Infect Dis. 2013; 208(4): 699-704.
  • Feldmesser M, Kress Y, Casadevall A. Dynamic changes in the morphology of Cryptococcus neoformans during murine pulmonary infection. Microbiology. 2001; 147: 2355-65.
  • Feldmesser M, Kress Y, Novikoff P, Casadevall A. Cryptococcus neo-formans is a facultative intracellular pathogen in murine pulmonary infection. Infect Immun. 2000; 68(7): 4225-37.
  • Frases S, Pontes B, Nimrichter L, Viana NB, Rodrigues ML, Casadevall A. Capsule of Cryptococcus neoformans grows by enlargement of polysaccharide molecules. Proc Natl Acad Sci USA. 2009; 106(4): 1228-33.
  • Fromtling RA, Shadomy HJ, Jacobson ES. Decreased virulence in stable, acapsular mutants of Cryptococcus neoformans Mycopathologia. 1982; 79(1): 23-9.
  • Ganendren R, Carter E, Sorrell T, Widmer F, Wright L. Phospholipase B activity enhances adhesion of Cryptococcus neoformans to a human lung epithelial cell line. Microbes Infect. 2006; 8(4): 1006-15.
  • García-Barbazán I, Trevijano-Contador N, Rueda C, de Andrés B, Pérez-Tavárez R, Herrero-Fernández I, et al. The formation of titan cells in Cryptococcus neoformans depends on the mouse strain and correlates with induction of Th2-type responses. Cell Microbiol. 2016; 18(1): 111-24.
  • Garcia-Hermoso D, Dromer F, Janbon G. Cryptococcus neoformans capsule structure evolution in vitro and during murine infection. Infect Immun. 2004; 72(6): 3359-65.
  • García-Rodas R, Casadevall A, Rodríguez-Tudela JL, Cuenca-Estrella M, Zaragoza O. Cryptococcus neoformans capsular enlargement and cellular gigantism during Galleria mellonella infection. PLoS One. 2011; 6(9): e24485.
  • Gates MA, Thorkildson P, Kozel TR. Molecular architecture ofthe Cryptococcus neoformans capsule. Mol Microbiol. 2004; 52(1): 13-24.
  • Gerik KJ, Bhimireddy SR, Ryerse JS, Specht CA, Lodge JK. PKC1 is essential for protection against both oxidative and nitrosative stresses, cell integrity, and normal manifestation of virulence factors in the pathogenic fungus Cryptococcus neoformans Eukaryot Cell. 2008; 7(10): 1685-98.
  • Gerik KJ, Donlin MJ, Soto CE, Banks AM, Banks IR, Maligie MA, et al. Cell wall integrity is dependent on the PKC1 signal transduction pathway in Cryptococcus neoformans. Mol Microbiol. 2005; 58(2): 393-408.
  • Gerstein AC, Fu MS, Mukaremera L, Li Z, Ormerod KL, Fraser JA, et al. Polyploid titan cells produce haploid and aneuploid progeny to promote stress adaptation. MBio. 2015; 6(5): e01340-15.
  • Giles SS, Batinić-Haberle I, Perfect JR, Cox GM. Cryptococcus neo-formans mitochondrial superoxide dismutase: an essential link between antioxidant function and high-temperature growth. Eukaryot Cell. 2005; 4(1): 46-54.
  • Giles SS, Dagenais TRT, Botts MR, Keller NP, Hull CM. Elucidating the pathogenesis of spores from the human fungal pathogen Cryptococcus neoformans. Infect Immun. 2009; 77(8): 3491-3500.
  • Giles SS, Stajich JE, Nichols C, Gerrald QD, Alspaugh JA, Dietrich F, et al. The Cryptococcus neoformans catalase gene family and its role in antioxidant defense. Eukaryot Cell. 2006; 5(9): 1447-59.
  • Goldman DL, Khine H, Abadi J, Lindenberg DJ, Pirofski La, Niang R, et al. Serologic evidence for Cryptococcus neoformans infection in early childhood. Pediatrics. 2001; 107(5): E66.
  • Granger DL, Perfect JR, Durack DT. Virulence of Cryptococcus neoformans: regulation of capsule synthesis by carbon dioxide. J Clin Invest. 1985; 76(2): 508-16.
  • Griffiths EJ, Kretschmer M, Kronstad JW. Aimless mutants of Cryptococcus neoformans: failure to disseminate. Fungal Biol Rev. 2012; 26(2-3): 61-72.
  • Heiss C, Klutts JS, Wang Z, Doering TL, Azadi P. The structure of Cryptococcus neoformans galactoxylomannan contains beta-Dglucuronic acid. Carbohydr Res. 2009; 344(7): 915-20.
  • Heitman J, Kozel TR, Kwon-Chung KJ, Perfect JR, Casadevall A. Cryptococcus: from human pathogen to model yeast. Washington DC: ASM Press; 2010.
  • Hu G, Caza M, Cadieux B, Bakkeren E, Do E, Jung WH, et al. The endosomal sorting complex required for transport machinery influences haem uptake and capsule elaboration in Cryptococcus neoformans Mol Microbiol. 2015; 96(5): 973-92.
  • Hu G, Caza M, Cadieux B, Chan V, Liu V, Kronstad J. Cryptococcus neoformans requires the ESCRT protein Vps23 for iron acquisition from heme, for capsule formation, and for virulence. Infect Immun. 2013; 81(1): 292-302.
  • Hu G, Cheng P-Y, Sham A, Perfect JR, Kronstad JW. Metabolic adaptation in Cryptococcus neoformans during early murine pulmonary infection. Mol Microbiol. 2008; 69(6): 1456-75.
  • Huang M, Hebert AS, Coon JJ, Hull CM. Protein composition of infectious spores reveals novel sexual development and germination factors in Cryptococcus PLoS Genet. 2015; 11(8): e1005490.
  • Inoue M, Shinohara ML. Clustering of pattern recognition receptors for fungal detection. PLoS Pathog. 2014; 10(2): e1003873.
  • Jacobson ES, Hong JD. Redox buffering by melanin and Fe(II) in Cryptococcus neoformans J Bacteriol. 1997; 179(17): 5340-6.
  • Janbon G. Cryptococcus neoformans capsule biosynthesis and regulation. FEMS Yeast Res. 2004; 4(8): 765-71.
  • Johnston SA, May RC. Cryptococcus interactions with macrophages: evasion and manipulation of the phagosome by a fungal pathogen. Cell Microbiol. 2013; 15(3): 403-11.
  • Johnston SA, May RC. The human fungal pathogen Cryptococcus neoformans escapes macrophages by a phagosome emptying mechanism that is inhibited by Arp2/3 complex-mediated actin polymyerisation. PLoS Pathog. 2010; 6(8): 27-8.
  • Jong A, Wu C, Shackleford GM, Kwon-Chung KJ, Chang YC, Chen H, et al. Involvement of human CD44 during Cryptococcus neo-formans infection of brain microvascular endothelial cells. Cell Microbiol. 2008; 10(6): 1313-26.
  • Jung WH, Kronstad JW. Iron and fungal pathogenesis: a case study with Cryptococcus neoformans Cell Microbiol. 2008; 10(2): 277-84.
  • Kechichian TB, Shea J, Del Poeta M. Depletion of alveolar macrophages decreases the dissemination of a glucosylceramide-deficient mutant of Cryptococcus neoformans in immunodeficient mice. Infect Immun. 2007; 75(10): 4792-8.
  • Kido N, Makimura K, Kamegaya C, Shindo I, Shibata E, Omiya T, et al. Long-term surveillance and treatment of subclinical cryptococcosis and nasal colonization by Cryptococcus neoformans and C. gattii species complex in captive koalas (Phascolarctes cinereus). Med Mycol. 2012; 50(3): 291-8.
  • Kozel TR, Gotschlich EC. The capsule of Cryptococcus neoformans passively inhibits phagocytosis of the yeast by macrophages. J Immunol. 1982; 129(4): 1675-80.
  • Kozel TR, Pfrommer GST, Guerlain AS, Highison BA, Highison GJ. Role of the capsule in phagocytosis of Cryptococcus neoformans Rev Infect Dis. 1988; 10(Suppl. 2): S436-9.
  • Kozel TR. Activation of the complement system by pathogenic fungi. Clin. Microbiol. Rev. 1996; 9(1): 34-46.
  • Kraus PR, Fox DS, Cox GM, Heitman J. The Cryptococcus neoformans MAP kinase Mpk1 regulates cell integrity in response to antifungal drugs and loss of calcineurin function. Mol Microbiol. 2003; 48(5): 1377-87.
  • Kraus PR, Heitman J. Coping with stress: calmodulin and calcineurin in model and pathogenic fungi. Biochem Biophys Res Commun. 2003; 311(4): 1151-7.
  • Kwon-Chung KJ, Bennett JE. Distribution of α and a mating types of Cryptococcus neoformans among natural and clinical isolates. Am J Epidemiol. 1978; 108(4): 337-40.
  • Kwon-Chung KJ, Rhodes JC. Encapsulation and melanin formation as indicators of virulence in Cryptococccus neoformans Infect Immun. 1986; 51(1): 218-23.
  • Kwon-Chung KJ. Morphogenesis of Filobasidiella neoformans, the sexual state of Cryptococcus neoformans Mycologia. 1976; 68(4): 821-33.
  • Lee SC, Phadke S, Sun S, Heitman J. Pseudohyphal growth of Cryptococcus neoformans is a reversible dimorphic transition in response to ammonium that requires Amt1 and Amt2 ammonium permeases. Eukaryot Cell. 2012; 11(11): 1391-8.
  • Lester SJ, Kowalewich NJ, Bartlett KH, Krockenberger MB, Fairfax TM, Malik R. Clinicopathologic features of an unusual outbreak of cryptococcosis in dogs, cats, ferrets, and a bird: 38 cases (January to July 2003). J Am Vet Med Assoc. 2004; 225(11): 1716-22.
  • Levitz SM, DiBenedetto DJ. Paradoxical role of capsule in murine bronchoalveolar macrophage-mediated killing of Cryptococcus neoformans J Immunol. 1989; 142(2): 659-65.
  • Levitz SM, Nong S-H, Seetoo KF, Harrison TS, Speizer RA, Simons ER. Cryptococcus neoformans resides in an acidic phagolysosome of human macrophages. Infect Immun. 1999; 67(2): 885-90.
  • Lin X, Hull CM, Heitman J. Sexual reproduction between partners of the same mating type in Cryptococcus neoformans Nature. 2005; 434(7036): 1017-21.
  • Littman ML, Borok R. Relation of the pigeon to cryptococcosis: natural carrier state, heat resistance and survival of Cryptococcus neoformans Mycopathol Mycol Appl. 1968; 35(3): 329-45.
  • Litvintseva AP, Carbone I, Rossouw J, Thakur R, Govender NP, Mitchell TG. Evidence that the human pathogenic fungus Cryptococcus neoformans var. grubii may have evolved in Africa. PLoS One. 2011; 6(5): e19688.
  • Litvintseva AP, Marra RE, Nielsen K, Heitman J, Vilgalys R, Mitchell TG. Evidence of sexual recombination among Cryptococcus neoformans serotype A isolates in Sub-Saharan Africa. Eukaryot Cell. 2003; 2(6): 1162-8.
  • Liu L, Tewari RP, Williamson PR. Laccase protects Cryptococcus neoformans from antifungal activity of alveolar macrophages. Infect Immun. 1999; 67(11): 6034-9.
  • Liu OW, Chun CD, Chow ED, Chen C, Madhani HD, Noble SM. Systematic genetic analysis of virulence in the human fungal pathogen Cryptococcus neoformans. Cell. 2008; 135(1): 174-88.
  • Liu T, Kim J, Wang Y, Toffaletti D. Brain inositol is a novel stimulator for promoting Cryptococcus penetration of the blood-brain barrier. PLoS Pathog. 2013; 9(4): e1003247.
  • Luberto C, Martinez-Mariño B, Taraskiewicz D, Bolaños B, Chitano P, Toffaletti DL, et al. Identification of App1 as a regulator of phagocytosis and virulence of Cryptococcus neoformans J Clin Invest. 2003; 112(7): 1080-94.
  • Ma H, Croudace JE, Lammas DA, May RC. Direct cell-to-cell spread of a pathogenic yeast. BMC Immunol. 2007; 8(1): 15.
  • Ma H, Croudace JE, Lammas DA, May RC. Expulsion of live pathogenic yeast by macrophages. Curr Biol. 2006; 16(21): 2156-60.
  • Macher AM, Bennett JE, Gadek JE, Frank MM. Complement depletion in cryptococcal sepsis. J Immunol. 1978; 120(5): 1686-90.
  • Magditch DA, Liu TB, Xue C, Idnurm A. DNA mutations mediate microevolution between host-adapted forms of the pathogenic fungus Cryptococcus neoformans. PLoS Pathog. 2012; 8(10): e1002936.
  • Mansour MK, Latz E, Levitz SM. Cryptococcus neoformans glycoantigens are captured by multiple lectin receptors and presented by dendritic cells. J Immunol. 2006; 176(5): 3053-61.
  • Maxson ME, Cook E, Casadevall A, Zaragoza O. The volume and hydration of the Cryptococcus neoformans polysaccharide capsule. Fungal Genet Biol. 2007; 44(3): 180-6.
  • Maziarz EK, Perfect JR. Cryptococcosis. Infect Dis Clin North Am. 2016; 30(1): 179-206.
  • McFadden D, Zaragoza O, Casadevall A. The capsular dynamics of Cryptococcus neoformans Trends Microbiol. 2006; 14(11): 497-505.
  • McFadden DC, Fries BC, Wang F, Casadevall A. Capsule structural heterogeneity and antigenic variation in Cryptococcus neoformans Eukaryot Cell. 2007; 6(8): 1464-73.
  • McGill S, Malik R, Saul N, Beetson S, Secombe C, Robertson I, et al. Cryptococcosis in domestic animals in Western Australia: a retrospective study from 1995-2006. Med Mycol. 2009; 47(6): 625-39.
  • Merkel GJ, Scofield BA. The in vitro interaction of Cryptococcus neoformans with human lung epithelial cells. FEMS Immunol Med Microbiol. 1997; 19(3): 203-13.
  • Missall TA, Cherry-Harris JF, Lodge JK. Two glutathione peroxidases in the fungal pathogen Cryptococcus neoformans are expressed in the presence of specific substrates. Microbiology. 2005; 151(8): 2573-81.
  • Missall TA, Lodge JK. Function of the thioredoxin proteins in Cryptococcus neoformans during stress or virulence and regulation by putative transcriptional modulators. Mol Microbiol. 2005a; 57(3): 847-58.
  • Missall TA, Lodge JK. Thioredoxin reductase is essential for viability in the fungal pathogen Cryptococcus neoformans Eukaryot Cell. 2005b; 4(2): 487-9.
  • Missall TA, Pusateri ME, Donlin MJ, Chambers KT, Corbett JA, Lodge JK. Posttranslational, translational, and transcriptional responses to nitric oxide stress in Cryptococcus neoformans: implications for virulence. Eukaryot Cell. 2006; 5(3): 518-29.
  • Mitchell DH, Sorrell TC, Allworth AM, Heath CH, McGregor AR, Papanaoum K, et al. Cryptococcal disease of the CNS in immunocompetent hosts: influence of cryptococcal variety on clinical manifestations and outcome. Clin Infect Dis. 1995; 20(3): 611-6.
  • Monari C, Bistoni F, Casadevall A, Pericolini E, Pietrella D, Kozel TR, et al. Glucuronoxylomannan, a microbial compound, regulates expression of costimulatory molecules and production of cytokines in macrophages. J Infect Dis. 2005a; 191(1): 127-37.
  • Monari C, Bistoni F, Vecchiarelli A. Glucuronoxylomannan exhibits potent immunosuppressive properties. FEMS Yeast Res. 2006; 6(4): 537-42.
  • Monari C, Paganelli F, Bistoni F, Kozel TR, Vecchiarelli A. Capsular polysaccharide induction of apoptosis by intrinsic and extrinsic mechanisms. Cell Microbiol. 2008; 10(10): 2129-37.
  • Monari C, Pericolini E, Bistoni G, Casadevall A, Kozel TR, Vecchiarelli A. Cryptococcus neoformans capsular glucuronoxylomannan induces expression of fas ligand in macrophages. J Immunol. 2005b; 174(6): 3461-8.
  • Monga DP. Role of macrophages in resistance of mice to experimental cryptococcosis. Infect Immun. 1981; 32(3): 975-8.
  • Mylonakis E, Ausubel FM, Perfect JR, Heitman J, Calderwood SB. Killing of Caenorhabditis elegans by Cryptococcus neoformans as a model of yeast pathogenesis. Proc Natl Acad Sci USA. 2002; 99(24): 15675-80.
  • Na Pombejra S, Salemi M, Phinney BS, Gelli A. The metalloprotease, Mpr1, engages AnnexinA2 to promote the transcytosis of fungal cells across the blood-brain barrier. Front Cell Infect Microbiol. 2017; 7: 296.
  • Nakamura K, Kinjo T, Saijo S, Miyazato A, Adachi Y, Ohno N, et al. Dectin-1 is not required for the host defense to Cryptococcus neoformans Microbiol Immunol. 2007; 51(11): 1115-9.
  • Nakamura K, Miyagi K, Koguchi Y, Kinjo Y, Uezu K, Kinjo T, et al. Limited contribution of Toll-like receptor 2 and 4 to the host response to a fungal infectious pathogen, Cryptococcus neoformans FEMS Immunol Med Microbiol. 2006; 47(1): 148-54.
  • Nakamura Y, Sato K, Yamamoto H, Matsumura K, Matsumoto I, Nomura T, et al. Dectin-2 deficiency promotes Th2 response and mucin production in the lungs after pulmonary infection with Cryptococcus neoformans Infect Immun. 2015; 83(2): 671-81.
  • Netski D, Kozel TR. Fc-dependent and Fc-independent opsonization of Cryptococcus neoformans by anticapsular monoclonal antibodies: Importance of epitope specificity. Infect Immun. 2002; 70(6): 2812-9.
  • Ngamskulrungroj P, Chang Y, Sionov E, Kwon-Chung KJ. The primary target organ of Cryptococcus gattii is different from that of Cryptococcus neoformans in a murine model. MBio. 2012; 3(3): e00103-12.
  • Nicola AM, Robertson EJ, Albuquerque P, Derengowski LS, Casadevall A. Nonlytic exocytosis of Cryptococcus neoformans from macrophages occurs in vivo and is influenced by phagosomal pH. MBio. 2011; 2(4): pi: e00167-11.
  • Nosanchuk JD, Casadevall A. Impact of melanin on microbial virulence and clinical resistance to antimicrobial compounds. Antimicrob Agents Chemother. 2006; 50(11): 3519-28.
  • Nosanchuk JD, Casadevall A. The contribution of melanin to microbial pathogenesis. Cell Microbiol. 2003; 5(4): 203-23.
  • Nosanchuk JD, Rosas AL, Lee SC, Casadevall A. Melanisation of Cryptococcus neoformans in human brain tissue. Lancet. 2000; 355(9220): 2049-50.
  • Noverr MC, Williamson PR, Fajardo RS, Huffnagle GB. CNLAC1 is required for extrapulmonary dissemination of Cryptococcus neoformans but not pulmonary persistence. Infect Immun. 2004; 72(3): 1693-9.+
  • Nussbaum G, Cleare W, Casadevall A, Scharff MD, Valadon P. Epitope location in the Cryptococcus neoformans capsule is a determinant of antibody efficacy. J Exp Med. 1997; 185(4): 685-94.
  • O’Meara TR, Alspaugh JA. The Cryptococcus neoformans capsule: a sword and a shield. Clin Microbiol Rev. 2012; 25(3): 387-408.
  • O’Meara TR, Holmer SM, Selvig K, Dietrich F, Alspaugh JA. Cryptococcus neoformans Rim101 is associated with cell wall remodeling and evasion of the host immune responses. MBio. 2013; 4(1): 1-13.
  • O’Meara TR, Norton D, Price MS, Hay C, Clements MF, Nichols CB, et al. Interaction of Cryptococcus neoformans Rim101 and protein kinase A regulates capsule. PloS Pathog. 2010; 6(2): e1000776.
  • O’Meara TR, Xu W, Selvig KM, O’Meara MJ, Mitchell AP, Alspaugh JA. The Cryptococcus neoformans Rim101 transcription factor directly regulates genes required for adaptation to the host. Mol Cell Biol. 2014; 34(4): 673-84.
  • Odom A, Muir S, Lim E, Toffaletti DL, Perfect J, Heitman J. Calcineurin is required for virulence of Cryptococcos neoformans EMBO J. 1997; 16(10): 2576-89.
  • Okagaki LH, Nielsen K. Titan cells confer protection from phagocytosis in Cryptococcus neoformans infections. Eukaryot Cell. 2012; 11(6): 820-6.
  • Okagaki LH, Strain AK, Nielsen JN, Charlier C, Baltes NJ, Chrétien F, et al. Cryptococcal cell morphology affects host cell interactions and pathogenicity. PLoS Pathog. 2010; 6(6): e1000953.
  • Okagaki LH, Wang Y, Ballou ER, O’Meara TR, Bahn Y-S, Alspaugh JA, et al. Cryptococcal titan cell formation is regulated by Gprotein signaling in response to multiple stimuli. Eukaryot Cell. 2011; 10(10): 1306-16.
  • Olszewski MA, Noverr MC, Chen GH, Toews GB, Cox GM, Perfect JR, et al. Urease Expression by Cryptococcus neoformans promotes microvascular sequestration, thereby enhancing central nervous system invasion. Am J Pathol. 2004; 164(5): 1761-71.
  • Ost KS, Esher SK, Wager CML, Walker L, Wagener J, Munro C, et al. Rim pathway-mediate alterations in the fungal cell wall influence immune recognition and inflammation. MBio. 2017; 8(1): e02290-16.
  • Ost KS, O’Meara TR, Huda N, Esher SK, Alspaugh JA. The Cryptococcus neoformans alkaline response pathway: identification of a novel rim pathway activator. PLoS Genet. 2015; 11(4): e1005159.
  • Pennisi MG, Hartmann K, Lloret A, Ferrer L, Addie D, Belák S, et al. Cryptococcosis in cats: ABCD guidelines on prevention and management. J Feline Med Surg. 2013; 15(7): 611-8.
  • Perfect JR. Cryptococcus neoformans: the yeast that likes it hot. FEMS Yeast Res. 2006; 6(4): 463-8.
  • Petzold EW, Himmelreich U, Mylonakis E, Rude T, Toffaletti D, Cox GM, et al. Characterization and regulation of the trehalose synthesis pathway and its importance in the pathogenicity of Cryptococcus neoformans. Infect Immun. 2006; 74(10): 5877-87.
  • Pierini LM, Doering TL. Spatial and temporal sequence of capsule construction in Cryptococcus neoformans Mol Microbiol. 2001; 41(1): 105-15.
  • Qin QM, Luo J, Lin X, Pei J, Li L, Ficht TA, et al. Functional analysis of host factors that mediate the intracellular lifestyle of Cryptococcus neoformans. PLoS Pathog. 2011; 7(6): e1002078.
  • Rajasingham R, Smith RM, Park BJ, Jarvis JN, Govender NP, Chiller TM, et al. Global burden of disease of HIV-associated cryptococcal meningitis: an updated analysis. Lancet Infect Dis. 2017; 17(8): 873-81.
  • Reese AJ, Doering TL. Cell wall a-1,3-glucan is required to anchor the Cryptococcus neoformans capsule. Mol Microbiol. 2003; 50(4): 1401-9.
  • Rhodes JC. Contribution of complement component C5 to the pathogenesis of experimental murine cryptococcosis. Sabouraudia. 1985; 23(3): 225-34.
  • Robertson EJ, Najjuka G, Rolfes MA, Akampurira A, Jain N, Anantharanjit J, et al. Cryptococcus neoformans ex vivo capsule size is associated with intracranial pressure and host immune response in HIV-associated cryptococcal meningitis. J Infect Dis. 2014; 209(1): 74-82.
  • Rodrigues ML, Nimrichter L, Oliveira DL, Frases S, Miranda K, Zaragoza O, et al. Vesicular polysaccharide export in Cryptococcus neoformans is a eukaryotic solution to the problem of fungal trans-cell wall transport. Eukaryot Cell. 2007; 6(1): 48-59.
  • Ruiz A, Neilson JB, Bulmer GS. Control of Cryptococcus neoformans in nature by biotic factors. Med Mycol. 1982; 20(1): 21-9.
  • Salgado DC, de Jesús JM, Bolaños B. Purification and characterization of a cytoplasmic immunosuppressive component from Cryptococcus neoformans by preparative electrophoretic techniques. In: Abstracts of the 94th general meeting of the American Society for Microbiology. 1994 March 23-27; Las Vegas: 1994; p. 77.
  • Santangelo R, Zoellner H, Sorrell T, Wilson C, Donald C, Djordjevic J, et al. Role of extracellular phospholipases and mononuclear phagocytes in dissemination of cryptococcosis in a murine model. Infect Immun. 2004; 72(4): 2229-39.
  • Santiago-Tirado FH, Onken MD, Cooper JA, Klein RS, Doering TL. Trojan horse transit contributes to blood-brain barrier crossing of a eukaryotic pathogen. mBio. 2017; 8(1): e02183-16.
  • Selvig K, Alspaugh JA. pH response pathways in fungi: adapting to host-derived and environmental signals. Mycobiology. 2011; 39(4): 249-56.
  • Shi M, Li SS, Zheng C, Jones GJ, Kim KS, Zhou H, et al. Real-time imaging of trapping and urease-dependent transmigration of Cryptococcus neoformans in mouse brain. J Clin Invest. 2010; 120(5): 1683-93.
  • Shoham S, Huang C, Chen J-M, Golenbock DT, Levitz SM. Tolllike receptor 4 mediates intracellular signaling without TNF-α release in response to Cryptococcus neoformans polysaccharide capsule. J Immunol. 2001; 166(7): 4620-6.
  • Smith LM, Dixon EF, May RC. The fungal pathogen Cryptococcus neoformans manipulates macrophage phagosome maturation. Cell Microbiol. 2015; 17(5): 702-13.
  • Sorrell TC, Juillard P-G, Djordjevic JT, Kaufman-Francis K, Dietmann A, Milonig A, et al. Cryptococcal transmigration across a model brain blood-barrier: evidence of the Trojan horse mechanism and differences between Cryptococcus neoformans var. grubii strain H99 and Cryptococcus gattii strain R265. Microbes Infect. 2016; 18(1): 57-67.
  • Speed B, Dunt D. Clinical and host differences between infections with thetwo varieties of Cryptococcus neoformans Clin Infect Dis. 1995; 21(1): 28-34.
  • Springer DJ, Saini D, Byrnes EJ, Heitman J, Frothingham R. Development of an aerosol model of Cryptococcus reveals humidity as an important factor affecting the viability of Cryptococcus during aerosolization. PLoS One. 2013; 8(7): e69804.
  • Stano P, Williams V, Villani M, Cymbalyuk ES, Qureshi A, Huang Y, et al. App1: an antiphagocytic protein that binds to complement receptors 3 and 2. J Immunol. 2009; 182(1): 84-91.
  • Steenbergen JN, Casadevall A. The origin and maintenance of virulence for the human pathogenic fungus Cryptococcus neoformans. Microbes Infect. 2003; 5(7): 667-75.
  • Steenbergen JN, Shuman HA, Casadevall A. Cryptococcus neoformans interactions with amoebae suggest an explanation for its virulence and intracellular pathogenic strategy in macrophages. Proc Natl Acad Sci USA. 2001; 98(26): 15245-50.
  • Sukroongreung S, Kitiniyom K, Nilakul C, Tantimavanich S. Pathogenicity of basidiospores of Filobasidiella neoformans var. neoformans. Med Mycol. 1998; 36(6): 419-24.
  • Sun TS, Ju X, Gao HL, Wang T, Thiele DJ, Li JY, et al. Reciprocal functions of Cryptococcus neoformans copper homeostasis machinery during pulmonary infection and meningoencephalitis. Nat Commun. 2014; 5: 5550.
  • Taborda CP, Casadevall A. CR3 (CD11b/CD18) and CR4 (CD11c/CD18) are involved in complement-independent antibody-mediated phagocytosis of Cryptococcus neoformans Immunity. 2002; 16(6): 791-802.
  • Teixeira PA, Penha LL, Mendonça-Previato L, Previato JO. Mannoprotein MP84 mediates the adhesion of Cryptococcus neoformans to epithelial lung cells. Front Cell Infect Microbiol. 2014; 4: 106.
  • Trevijano-Contador N, Rossi SA, Alves E, Landín-Ferreiroa S, Zaragoza O. Capsule enlargement in Cryptococcus neoformans is dependent on mitochondrial activity. Front Microbiol. 2017; 8: 1423.
  • Trevijano-Contador N, Rueda C, Zaragoza O. Fungal morphogenetic changes inside the mammalian host. Semin Cell Dev Biol. 2016; 57: 100-9.
  • Tucker SC, Casadevall A. Replication of Cryptococcus neoformans in macrophages is accompanied by phagosomal permeabilization and accumulation of vesicles containing polysaccharide in the cytoplasm. Proc Natl Acad Sci. 2002; 99(5): 3165-70.
  • Vartivarian SE, Anaissie EJ, Cowart RE, Sprigg HA, Tingler MJ, Jacobson ES. Regulation of cryptococcal capsular polysaccharide by iron. J Infect Dis. 1993; 167(1): 186-90.
  • Vecchiarelli A, Monari C. Capsular material of Cryptococcus neoformans: virulence and much more. Mycopathologia. 2012; 173(5-6): 375-86.
  • Vecchiarelli A, Pericolini E, Gabrielli E, Chow S-K, Bistoni F, Cenci E, et al. Cryptococcus neoformans galactoxylomannan is a potent negative immunomodulator, inspiring new approaches in anti-inflammatory immunotherapy. Immunotherapy. 2011; 3(8): 997-1005.
  • Vecchiarelli A. Immunoregulation by capsular components of Cryptococcus neoformans Med Mycol. 2000; 38(6): 407-17.
  • Velagapudi R, Hsueh YP, Geunes-Boyer S, Wright JR, Heitman J. Spores as infectious propagules of Cryptococcus neoformans Infect Immun. 2009; 77(10): 4345-55.
  • Venn-Watson S, Daniels R, Smith C. Thirty year retrospective evaluation of pneumonia in a bottlenose dolphin Tursiops truncatus population. Dis Aquat Organ. 2012; 99(3): 237-42.
  • Villena SN, Pinheiro RO, Pinheiro CS, Nunes MP, Takiya CM, Dosreis GA, et al. Capsular polysaccharides galactoxylomannan and glucuronoxylomannan from Cryptococcus neoformans induce macrophage apoptosis mediated by Fas ligand. Cell Microbiol. 2008; 10(6): 1274-85.
  • Voelz K, Lammas DA, May RC. Cytokine signaling regulates the outcome of intracellular macrophage parasitism by Cryptococcus neoformans Infect Immun. 2009; 77(8): 3450-7.
  • Vu K, Eigenheer RA, Phinney BS, Gelli A. Cryptococcus neoformans promotes its transmigration into the central nervous system by inducing molecular and cellular changes in brain endothelial cells. Infect Immun. 2013; 81(9): 3139-47.
  • Walsh NM, Wuthrich M, Wang H, Klein B, Hull CM. Characterization of C-type lectins reveals an unexpectedly limited interaction between Cryptococcus neoformans spores and Dectin-1. PLoS One. 2017; 12(3): e0173866.
  • Wang Y, Casadevall A. Susceptibility of melanized and nonmelanized Cryptococcus neoformans to nitrogen- and oxygen-derived oxidants. Infect Immun. 1994; 62(7): 3004-7.
  • Warpeha KM, Park YD, Williamsonb PR. Susceptibility of intact germinating Arabidopsis thaliana to human fungal pathogens Cryptococcus neoformans and C. gattii Appl Environ Microbiol. 2013; 79(9): 2979-88.
  • Waterman SR, Hacham M, Hu G, Zhu X, Park Y-D, Shin S, et al. Role of a CUF1/CTR4 copper regulatory axis in the virulence of Cryptococcus neoformans J Clin Invest. 2007; 117(3): 794-802.
  • White C, Lee J, Kambe T, Fritsche K, Petris MJ. A role for the ATP7A copper-transporting ATPase in macrophage bactericidal activity. J Biol Chem. 2009; 284(49): 33949-56.
  • Wickes BL, Mayorga ME, Edman U, Edman JC. Dimorphism and haploid fruiting in Cryptococcus neoformans: association with the alpha-mating type. Proc Natl Acad Sci USA. 1996; 93(July): 7327-31.
  • Wiesner DL, Specht CA, Lee CK, Smith KD, Mukaremera L, Lee ST, et al. Chitin recognition via chitotriosidase promotes pathologic type-2 helper T cell responses to cryptococcal infection. PLoS Pathog. 2015; 11(3): e1004701.
  • Williamson PR, Wakamatsu K, Ito S. Melanin biosynthesis in Cryptococcus neoformans J Bacteriol. 1998; 180(6): 1570-2.
  • Williamson PR. Biochemical and molecular characterization of the diphenol oxidase of Cryptococcus neoformans: identification as a laccase. J Bacteriol. 1994; 176(3): 656-64.
  • Xu J. Genotype-environment interactions of spontaneous mutations for vegetative fitness in the human pathogenic fungus Cryptococcus neoformans. Genetics. 2004; 168(3): 1177-88.
  • Yamamoto H, Nakamura Y, Sato K, Takahashi Y, Nomura T, Miyasaka T, et al. Defect of CARD9 leads to impaired accumulation of gamma interferon-producing memory phenotype T cells in lungs and increased susceptibility to pulmonary infection with Cryptococcus neoformans. Infect Immun. 2014; 82(4): 1606-15.
  • Yauch LE, Mansour MK, Levitz SM. Receptor-mediated clearance of Cryptococcus neoformans capsular polysaccharide in vivo Infect Immun. 2005; 73(12): 8429-32.
  • Yauch LE, Mansour MK, Shoham S, Rottman JB, Levitz SM. Involvement of CD14, toll-like receptors 2 and 4, and MyD88 in the host response to the fungal pathogen Cryptococcus neoformans in vivo Infect Immun. 2004; 72(9): 5373-82.
  • Zaragoza O, Casadevall A. Experimental modulation of capsule size in Cryptococcus neoformans Biol Proced Online. 2004; 6(1): 10-4.
  • Zaragoza O, Chrisman CJ, Castelli MV, Frases S, Cuenca-Estrella M, Rodríguez-Tudela JL, et al. Capsule enlargement in Cryptococcus neoformans confers resistance to oxidative stress suggesting a mechanism for intracellular survival. Cell Microbiol. 2008; 10(10): 2043-57.
  • Zaragoza O, Fries BC, Casadevall A. Induction of capsule growth in Cryptococcus neoformans by mammalian serum and CO2. Infect Immun. 2003a; 71(11): 6155-64.
  • Zaragoza O, García-Rodas R, Nosanchuk JD, Cuenca-Estrella M, Rodríguez-Tudela JL, Casadevall A. Fungal cell gigantism during mammalian infection. PLoS Pathog. 2010; 6(6): e1000945.
  • Zaragoza O, Nielsen K. Titan cells in Cryptococcus neoformans: cells with a giant impact. Curr Opin Microbiol. 2013; 16(4): 409-13.
  • Zaragoza O, Rodrigues ML, de Jesus M, Frases S, Dadachova E, Casadevall A. The capsule of the fungal pathogen Cryptococcus neoformans Adv Appl Microbiol. 2009; 68: 133-216.
  • Zaragoza O, Taborda CP, Casadevall A. 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. 2003b; 33(7): 1957-67.

Publication Dates

  • Publication in this collection
    2018

History

  • Received
    01 Feb 2018
  • Accepted
    08 Mar 2018
Instituto Oswaldo Cruz, Ministério da Saúde Av. Brasil, 4365 - Pavilhão Mourisco, Manguinhos, 21040-900 Rio de Janeiro RJ Brazil, Tel.: (55 21) 2562-1222, Fax: (55 21) 2562 1220 - Rio de Janeiro - RJ - Brazil
E-mail: memorias@fiocruz.br