Your activity: 74 p.v.
your limit has been reached. plz Donate us to allow your ip full access, Email: sshnevis@outlook.com

Biology of Candida infections

Biology of Candida infections
Author:
Wiley A Schell, MS
Section Editor:
Carol A Kauffman, MD
Deputy Editor:
Keri K Hall, MD, MS
Literature review current through: Dec 2022. | This topic last updated: Jan 19, 2022.

INTRODUCTION — Candidiasis refers to the range of infections caused by species of the fungal genus Candida; these infections can be acute or chronic, localized or systemic. Disseminated candidiasis is life threatening. The great majority of candidiasis is caused by Candida albicans. C. albicans is a common commensal organism in the oropharyngeal cavity, gastrointestinal tract, and vagina of humans but is capable of causing opportunistic infection following disruption of the normal flora, a breach of the mucocutaneous barrier, or a defect in host cellular immunity. C. albicans can be detected as normal microbiota in about 50 percent of individuals [1].

The basic mycology and pathogenesis of candidiasis will be reviewed here. An overview of Candida infections and the epidemiology, pathogenesis, clinical manifestations, diagnosis, and treatment of candidemia and invasive candidiasis are presented separately; other forms of candidiasis are also discussed elsewhere. (See "Overview of Candida infections" and "Candidemia in adults: Epidemiology, microbiology, and pathogenesis" and "Clinical manifestations and diagnosis of candidemia and invasive candidiasis in adults" and "Management of candidemia and invasive candidiasis in adults".)

MYCOLOGY — The genus Candida encompasses about 200 species [2]. They can be found among humans and other mammals, birds, insects, arthropods, fish, animal waste, plants, mushrooms, naturally occurring high-sugar substrates (eg, honey, nectar, grapes) and fermentation products, dairy products, soil, freshwater, seawater, and on airborne particles [3-5].

Infection in humans was first described as oral thrush by Hippocrates in the fifth century BC. In 1853, Charles Robin microscopically observed budding cells and filaments in epithelial scrapings, and he named the fungus Oidium albicans. Subsequently, more than 160 synonyms, including Monilia albicans, were used before Candida albicans became the accepted name for the species.

At least 30 Candida species have caused infection in humans [2]. The most common of these are C. albicans, C. glabrata, C. krusei, C. parapsilosis, and C. tropicalis. C. parapsilosis has been recognized as a heterogeneous species, and it was proposed that it be split into three morphologically and physiologically indistinguishable species: C. parapsilosis, C. metapsilosis, and C. orthopsilosis [6]. Phylogenetic analyses show that C. glabrata is more closely related to Saccharomyces cerevisiae than to the C. albicans group [7-9], and future taxonomic revision is possible.

In 1995, Candida dubliniensis was described as a new species to accommodate a subset of isolates previously identified as C. albicans [10]. C. dubliniensis is seen primarily among HIV-infected patients [11]. An environmental source for this species may be the digestive tract of marine birds [5]. There is no clear clinical or experimental evidence that significant pathogenicity differences exist between C. albicans and C. dubliniensis [12], and it is not necessary to routinely distinguish C. dubliniensis from C. albicans in the clinical laboratory [13].

Candida auris was described as a new species after it was detected from a patient in Japan in 2009. Since then, cases have been reported from around the world including much of the United States [14]. C. auris is of particular concern because isolates cause serious infections, are often multidrug resistant, can be associated with outbreaks in health care facilities, and can be misidentified. C. auris infection was designated as a nationally notifiable disease in the United States in 2018. (See "Candidemia in adults: Epidemiology, microbiology, and pathogenesis", section on 'C. auris emergence'.)

All of the clinically common Candida species are diploid, except for the haploid C. glabrata and C. krusei. Of the aforementioned species, only C. krusei has been determined to have a classic heterothallic sexual cycle [15]. However, mating and genetic recombination does occur in C. albicans and C. tropicalis through a parasexual cycle [9,16-18]. An apparently similar mating mechanism exists in C. dubliniensis, and interspecific mating occurs between C. albicans and C. dubliniensis [19].

Sequencing of the C. albicans genome has demonstrated that this species possesses 6419 genes contained in 8 chromosomes [20]. These data will facilitate proteomic studies, which in turn could lead to better understanding of key biological characteristics of these species, including pathogen-host interactions, improved diagnostics, antifungal drug resistance, novel molecular targets for antifungal therapy, biofilm formation and dynamics, cell signaling, and responses to stress, nutrients, temperature, and pH [21,22].

Growth, morphology, and pleomorphism — Unlike the so-called systemic dimorphic fungi (eg, Histoplasma capsulatum), the morphology of a given Candida species is fundamentally the same whether observed in vitro or in vivo [23]. C. glabrata grows as a small, elliptical, unicellular budding yeast at all times (picture 1). Rarely, buds of C. glabrata can adhere to one another in rudimentary short chains. In marked contrast, C. albicans, C. krusei, C. parapsilosis, and C. tropicalis form elliptical budding cells that typically are larger than those of C. glabrata; they can also form elaborate and well-developed multicellular filaments (picture 2), particularly when in contact with a solid substrate, such as human tissue or agar culture media.

C. krusei and C. parapsilosis can be considered dimorphic because they exhibit budding and pseudohyphal forms. C. albicans and C. tropicalis can form true hyphae (in addition to buds and pseudohyphae) and thus can be considered pleomorphic. Each of these species is capable of exhibiting budding and/or filamentous morphologies in infected tissue.

Usually, the presence of budding yeast and filaments in infected tissue is indicative of candidiasis. However, similar morphologies occur in infections caused by certain opportunistic molds (eg, Purpureocillium lilacinum [formerly Paecilomyces lilacinus]). These mold infections can be histopathologically misdiagnosed as candidiasis if histology is the sole diagnostic criterion used [24-26]. Correlation of histologic and culture results is recommended.

Pseudohyphae and hyphae — Pseudohyphae are morphologically and ontogenically distinct from hyphae. They are formed when buds remain attached to each other and subsequently elongate via differential rates of wall synthesis at various points along the cell wall. Elongation then stops and each terminal cell produces a new apical bud, which elongates in turn. This repeated process of budding and elongation can result in extensive filamentation [27]. Side branches initiate as buds and develop in the same manner. A constriction typically remains and can be seen at the origin of each bud (picture 3). The branching process is regulated in part by a vacuolar protein [28] and by calcineurin [29].

In contrast, true hyphae elongate by a polarized growth process of apical synthesis that does not involve budding. Buds are not present at the hyphal tips; as a result, true hyphae do not exhibit the periodic constrictions that characterize pseudohyphae (picture 4) [30]. C. albicans hyphae have been shown to contain the Spitzenkörper, a vesicle-laden, actin-rich structure, well known in molds, that correlates with tip growth rate and direction [31]. Hyphae develop by a process that is fundamentally different from that of pseudohyphal growth.

Septa — Septa form by centripetal synthesis from the interior of the cell wall inward toward the center of the cell and are present in pseudohyphae, hyphae, and between mature bud cells [30]. The septa provide internal support of the tubular filaments, and a central pore in each septum assures cytoplasmic continuity between cells [30].

White-opaque switching — In addition to yeast, pseudohyphal, and hyphal forms, C. albicans exhibits additional phenotypes of white and opaque colonies. The two phases differ in virulence and gene expression, and reversible switching between phenotypes occurs at a known frequency although the extent to which environmental (including host) stimuli influence switching is not fully known [32,33]. For mating to occur, white cells must first switch to the opaque form. Opaque-phase cells are prevalent as skin colonizers and are less frequently phagocytized by neutrophils in vitro, whereas white-phase cells are more commonly associated with candidemia and are preferentially phagocytized once germ tubes begin to form [34-38].

Genetic control — Signals that trigger each of the growth forms gradually are becoming better understood [39]. Signals include pH, temperature, carbon and nitrogen availability, oxygenation, serum, hormones, and density of the Candida cells within the infected host [40]. For example, one of the key factors, as indicated by studies of C. albicans, is pH of the microenvironment [41-43]. Several pH-regulated genes involved in growth and morphology have been identified: PHR1 is expressed when the ambient pH is 5.5 or higher and correlates with filamentous growth, whereas PHR2 is expressed when the ambient pH is lower than 5.5 (as in the vagina) and correlates with yeast-like growth [41]. Thus, the importance of these genes might vary according to the site of infection. For example, a null mutant of PHR1 was avirulent in a mouse model of systemic infection (the systemic pH being well above 5.5) but maintained its ability to produce vaginal infection (where the pH is approximately 4.5). Conversely, a PHR2 null mutant was able to cause systemic but not vaginal infection [44]. Mutants without Phr1p, a protein product of the PHR1 gene, also lack cross-linking activity associated with the cell wall glucan when grown at pH 8 [45]. Adaptation of C. albicans to the neutral-alkaline environment of the human body by pH sensing is determined mainly by the RIM 101 signal transduction pathway [46].

Several other genes have been identified that regulate the switch from the budding yeast to the filamentous form. Double mutants of EFG1 and CPH1 were nonfilamentous (ie, locked in the yeast form) and unable to form hyphae or pseudohyphae in response to stimuli, such as serum and macrophages [47]. Heat-shock protein 90 (Hsp90), a repressor of hyphal development, can be partially compromised at 30ºC, thus allowing hyphal initiation [48]. Several dozen genes have been shown to contribute to biofilm formation by C. albicans, and a role for quorum-sensing molecules has been postulated as well [43,49-52].

The gene EFG1 encodes a protein Efg1 that contains conserved regions with homology to the human Myc protein; Efg1 appears to activate pseudohyphal formation and downregulate hyphal development [53,54]. In one study of C. albicans, the gene CaRSR1, a RAS-related gene, was necessary for budding and also contributed to germ-tube emergence and cell elongation, which is part of the process of hyphal formation [55]. Alterations in the gene led to decreased virulence. G1 cyclins CLN1, CLN2, and CLN3 have been shown to play a key role in bud formation and regulation of yeast-hyphal transition [56,57]. Various protein kinases in the Cdc2 subfamily also have been shown to regulate filamentous growth [58]. Tor1 is required for contact-dependent induction of C. albicans filaments on artificial surfaces and could be required in vivo as well [59]. A role for quorum sensing as a signaling mechanism for hyphal development has also been suggested [60-63].

Both budding and filamentous forms presumably play a role in the progression of infection in humans [24,25,64]. The ongoing formation and detachment of unicellular buds can facilitate hematogenous dissemination of the yeast following angioinvasion. In contrast, filamentation enhances the ability of Candida species to invade solid tissue via a burrowing process [53]. Nonfilamentous C. albicans can be avirulent [65]. Some investigators have suggested that the hyphae of C. albicans invade epithelial cells using thigmotropism (movement based upon touch) [60,66]. Others argue that chemotropism plays a more important role [67].

Detection in the microbiology laboratory — Candida species can be expected to appear in culture within one to two days in most cases. C. glabrata is an exception; it may take several days to grow.

Definitive identification of yeast isolates traditionally has been based on a combination of morphologic features and physiologic testing. For example, rapid presumptive identification of C. albicans can be accomplished by observing formation of germ tubes (the initiation of true hyphal growth) when incubated in proteinaceous liquid (eg, plasma) at 35ºC for two to three hours (picture 5). Another phenotypic approach employs rapid colorimetric detection of the enzymes L-proline aminopeptidase and beta-galactosaminidase (C. albicans Screen). In addition, a chromogenic agar culture medium (CHROMagar) allows presumptive phenotypic identification of C. albicans, C. tropicalis, and C. krusei (picture 6), and a modified formula can distinguish C. dubliniensis as well [68]. C. auris colonies reportedly are pink to beige on the chromogenic culture medium and grow well at 42°C [69-71]. Traditional methods can further characterize yeast isolates using carbohydrate assimilation tests, which require two to three days for completion. These nutritional assimilation tests use either manual kits or automated methods. However, not all of these tests reliably identify C. auris isolates and careful testing using updated database software or instrumentation would be needed [70,71].

While many clinical laboratories continue to use traditional phenotypic methods for yeast identification, newer tools are gaining acceptance. These include matrix-assisted laser desorption/ionization time of flight mass spectrometry, which offers the ability to accurately identify yeast isolates, including C. auris, in minutes without the need for mycologic expertise [72-75]. In addition, the T2Candida system is capable of detecting Candida species in blood in as little as four hours. Nucleic acid sequence analysis hold promise for identification of fungal isolates and is a highly effective approach in the research laboratory setting, but as yet, there is no regulatory approval for its wider application. (See "Clinical manifestations and diagnosis of candidemia and invasive candidiasis in adults", section on 'Blood cultures'.)

Antigenic structure — The multilayered cell wall of C. albicans consists of glucans, mannans, mannoproteins, proteins, and chitin [76-79]. With the exception of glucans and chitin, which have little, if any, antigenicity, these products are capable of eliciting an immune response [80]. Glucan and chitin appear to lend mechanical strength and stability to the cell wall [45]. Strains deficient in one or more of these components are less virulent [77,78,81,82].

VIRULENCE FACTORS — Candida species have little inherent virulence as illustrated by their ability to exist in a commensal relationship with humans and other animals. Despite this fact, the prevalence of Candida infections has increased in recent decades, and the relative numbers of non-albicans Candida infections are increasing [83]. C. albicans, however, still causes the majority of infections. (See "Candidemia in adults: Epidemiology, microbiology, and pathogenesis", section on 'Prevalence of Candida species'.)

Molecular genetic advances, such as creation of gene knockout mutants, genome sequencing, and proteomic studies permit specific analysis of putative virulence factors [21,84]. It is likely that there are numerous virulence factors, which may play different roles at various sites and stages of a given infection [1,85,86]. As noted above, one such factor appears to be the ambient pH of the host [41-43]. Other examples include calcineurin (a calcium-regulated signaling enzyme) and Hsp90, which have been shown to be important virulence factors [29,87-89]. Calcineurin mutant strains showed no deficits in known virulence factors but were highly attenuated in a murine model of disseminated candidiasis. In addition, inhibition of calcineurin promotes susceptibility to azole antifungal compounds in vivo and in vitro [90,91]. More recently, the GWT1 gene, which encodes an inositol acyltransferase needed for anchoring of fungal proteins to the fungal cell membrane, has been shown to be a potential gene target that provides the basis for a new class of antifungal compounds [22,92].

There also appear to be multiple adherence mechanisms that allow initial attachment to host tissue (or plastic foreign bodies) and subsequent proliferation. For example:

A hyphal-specific surface protein, Hwp1, is needed to form stable attachments to epithelial cells and in the formation of biofilms [59,93,94]. Hwp1 deficiency resulted in decreased virulence in a murine model of candidiasis [95].

Mutations in the genes that regulate the switch from budding yeast to a filamentous form result in avirulent strains [52,65,84,96].

C. albicans deficient in mannosyl transferase demonstrated decreased adherence and virulence [77,97].

In vitro testing has suggested that the ability of C. albicans to synthesize trehalose provides protection against stress factors (such as thermal stress) [98]. In vivo testing of a trehalose synthase mutant strain showed attenuated virulence in a murine model of infection, supporting the further investigation of trehalose phosphate synthase as a potential antifungal compound target [98].

C. albicans also produces extracellular proteinases, phospholipases, lipases, hydrolytic enzymes, and adhesins in vivo [1,43,64,96,99-104]. Among these possible virulence factors, protease-lacking mutants were less virulent than their wild-type parent [1,30,105]. Reduced virulence also was induced by deletion of the gene encoding for the production of phospholipase [106]. This defect did not appear to affect in vitro adhesion but was associated with diminished cell wall penetration. Phospholipase formation reportedly is higher among C. albicans isolates causing infection [43]. In a mouse model, the phospholipase-competent strain had a significantly greater ability to cross the gastric mucosal barrier and invade the kidneys and liver. Mutant strains that lack adhesion Hwp1 showed decreased virulence in an esophageal murine model of candidiasis [107]. Ability to adapt to rapidly changing hypoxic conditions also may be a virulence factor [108-110].

Comparative analysis of the genome of C. albicans with that of the nonpathogenic yeast Saccharomyces cerevisiae has shown that about 7 percent of genes in C. albicans have no homolog in S. cerevisiae. These genes appear to be of the kind that may be responsible for synthesis of adhesions and hydrolytic enzymes, which are believed to be virulence factors [1].

GENETIC FACTORS — Chronic mucocutaneous candidiasis is associated with the genetic conditions polyglandular autoimmune syndrome type I (the autoimmune polyendocrinopathy-candidiasis-ectodermal dystrophy) syndrome and Job's (hyperimmunoglobulin E) syndrome [111-113]. (See "Autosomal dominant hyperimmunoglobulin E syndrome".)

Certain genetic mutations have also been associated with mucocutaneous Candida infections in familial clusters. These are discussed in detail separately. (See "Chronic mucocutaneous candidiasis".)

IMMUNE RESPONSE — Cell-wall constituents in Candida species can evoke both humoral and cellular host immune responses during the course of infection [114,115]. In addition, extracellular proteinases of C. albicans trigger immune responses locally and systemically [1,116]. The cell-mediated immune response is clearly important in host defense against Candida, as reflected by the higher prevalence of infections among individuals whose cellular immune systems are impaired, but the exact mechanisms of induction of this response have not been elucidated. Interleukin (IL-)12 appears to play an important role in the development of a Th1 response to C. albicans in the mouse; IL-12-deficient mice had increased susceptibility to infection introduced via the gastrointestinal tract and also to reinfection [117,118]. The role of interferon (IFN-)gamma seems to be more complex [112,119-121]:

IFN-gamma knockout mice were more susceptible to Candida infection.

Experimental infection or IL-12 administration induced a Th1 response in mice associated with IFN-gamma production.

IL-12 administration to systemically infected mice resulted in uniformly lethal disease, whereas coadministration of IFN-gamma was protective in 70 percent of animals.

In contrast, IL-10 knockout mice displayed resistance to systemic but not gastrointestinal Candida infection [122].

Neutrophils also contribute to host defense against Candida infection [96,115]. Several studies have shown that granulocyte colony-stimulating factor (G-CSF) increased neutrophil-mediated damage to pseudohyphae of C. albicans, perhaps in association with IFN-gamma [123,124]. One study found decreased salivary flow, diminished levels of secretory immunoglobulin (Ig)A, and reduced activity of salivary polymorphonuclear leukocytes in the patients with oral candidiasis compared with healthy controls [125].

The role of the humoral response in preventing disease progression during Candida infection has been unclear and controversial [126-128]. However, polyclonal and monoclonal antibodies have shown a protective effect in a mouse model of candidiasis [129,130]. Mannan-specific IgG antibodies have been shown to trigger both the classical and alternative complement pathways [131]. These findings suggest that development of a vaccine or short-term protective antibodies may be feasible. In both immunocompetent and immunocompromised mice, a vaccine against C. albicans conferred reduced fungal burden and improved survival [118,132]. Toll-like receptors (TLRs), mainly TLR2 and TLR4, play a key role in recognition of C. albicans and the activation of the host immune response [133]. In addition, the non-toll-like receptor Dectin-1 has been shown to induce host cellular responses following recognition of beta-glucan of the fungal cell wall [134,135].

SUMMARY

Candidiasis refers to the range of infections caused by species of the fungal genus Candida; these infections can be acute or chronic, localized or systemic. Disseminated candidiasis is life threatening. The great majority of candidiasis is caused by Candida albicans although non-albicans infections are increasing in incidence. C. albicans is a common commensal organism in the oropharyngeal cavity, gastrointestinal tract, and vagina of humans but is capable of causing opportunistic infection following disruption of the normal flora, a breach of the mucocutaneous barrier, or a defect in host cellular immunity. (See 'Introduction' above.)

At least 30 Candida species have caused infection in humans. The most common of these are C. albicans, C. glabrata, C. krusei, C. parapsilosis, and C. tropicalis. C. auris was first identified as a cause of infections in 2011, is often multidrug resistant, is associated with outbreaks in health care facilities, and is regarded as a global health problem. (See 'Mycology' above.)

Unlike the so-called systemic dimorphic fungi (eg, Histoplasma capsulatum), the morphology of a given Candida species is fundamentally the same whether observed in vitro or in vivo. C. glabrata grows as a small, elliptical, unicellular budding yeast at all times (picture 1). In marked contrast, C. albicans, C. krusei, C. parapsilosis, and C. tropicalis form elliptical budding cells that typically are larger than those of C. glabrata; they can also form elaborate and well-developed multicellular filaments (picture 2), particularly when in contact with a solid substrate, such as human tissue or agar culture media. (See 'Growth, morphology, and pleomorphism' above.)

The presence of budding yeast and filaments in infected tissue is usually indicative of candidiasis. (See 'Growth, morphology, and pleomorphism' above.)

Cultures can be expected to grow overnight in most cases. Rapid presumptive identification of C. albicans can be made by observing the formation of germ tubes (the initiation of true hyphal growth) when incubated in proteinaceous liquid (eg, plasma) at 35ºC for two to three hours (picture 5) or by rapid colorimetric detection of the enzymes L-proline aminopeptidase and beta-galactosaminidase. Traditionally, definitive identification of yeast species is based on a combination of morphologic features and physiologic testing, but that approach is being eclipsed by alternative methods such as matrix-assisted laser desorption/ionization time of flight mass spectrometry. (See 'Detection in the microbiology laboratory' above.)

Candida species have little inherent virulence as illustrated by their ability to exist in a commensal relationship with humans and other animals. Nevertheless, several virulence factors have been identified, including the ambient pH of the host, calcineurin, and various adherence mechanisms. Progress in the identification of virulence factors and important genes holds promise for the development of new classes of antifungal compounds. (See 'Virulence factors' above.)

Chronic mucocutaneous candidiasis is associated with the genetic conditions polyglandular autoimmune syndrome type I (the autoimmune polyendocrinopathy-candidiasis-ectodermal dystrophy) syndrome and Job's (hyperimmunoglobulin E) syndrome. Other genetic mutations have also been associated with mucocutaneous Candida infections in familial clusters. (See 'Genetic factors' above.)

Cell-wall constituents in Candida species can evoke both humoral and cellular host immune responses during the course of infection. In addition, extracellular proteinases of C. albicans trigger immune responses locally and systemically. The cell-mediated immune response is clearly important in the host defense against Candida, as reflected by the higher prevalence of infections among individuals whose cellular immune systems are impaired. (See 'Immune response' above.)

  1. Naglik JR, Challacombe SJ, Hube B. Candida albicans secreted aspartyl proteinases in virulence and pathogenesis. Microbiol Mol Biol Rev 2003; 67:400.
  2. Brandt ME, Lockhart SR. Recent Taxonomic Developments with Candida and Other Opportunistic Yeasts. Curr Fungal Infect Rep 2012; 6:170.
  3. Barnett JA, Payne RW, Yarrow D. Yeasts: Characteristics and Identification, 2nd ed, Cambridge University Press, Cambridge 1990.
  4. Meyer AS, Payne RW, Yarrow D. Candida Berkhout. In: The Yeasts: A Taxonomic Study, 4th ed, Kurtzman CP, Fell JW (Eds), Elsevier, New York 1998. p.454.
  5. Nunn MA, Schäefer SM, Petrou MA, Brown JR. Environmental source of Candida dubliniensis. Emerg Infect Dis 2007; 13:747.
  6. Tavanti A, Davidson AD, Gow NA, et al. Candida orthopsilosis and Candida metapsilosis spp. nov. to replace Candida parapsilosis groups II and III. J Clin Microbiol 2005; 43:284.
  7. Wong S, Fares MA, Zimmermann W, et al. Evidence from comparative genomics for a complete sexual cycle in the 'asexual' pathogenic yeast Candida glabrata. Genome Biol 2003; 4:R10.
  8. Diezmann S, Cox CJ, Schönian G, et al. Phylogeny and evolution of medical species of Candida and related taxa: a multigenic analysis. J Clin Microbiol 2004; 42:5624.
  9. Turner SA, Butler G. The Candida pathogenic species complex. Cold Spring Harb Perspect Med 2014; 4:a019778.
  10. Sullivan DJ, Westerneng TJ, Haynes KA, et al. Candida dubliniensis sp. nov.: phenotypic and molecular characterization of a novel species associated with oral candidosis in HIV-infected individuals. Microbiology 1995; 141 ( Pt 7):1507.
  11. Sullivan DJ, Moran GP, Coleman DC. Candida dubliniensis: ten years on. FEMS Microbiol Lett 2005; 253:9.
  12. Asmundsdóttir LR, Erlendsdóttir H, Agnarsson BA, Gottfredsson M. The importance of strain variation in virulence of Candida dubliniensis and Candida albicans: results of a blinded histopathological study of invasive candidiasis. Clin Microbiol Infect 2009; 15:576.
  13. Lockhart SR. Do hospital microbiology laboratories still need to distinguish Candida albicans from Candida dubliniensis? J Clin Microbiol 2011; 49:4415.
  14. Centers for Disease Control and Prevention. Tracking Candida auris. http://www.cdc.gov/fungal/candida-auris/tracking-c-auris.html (Accessed on June 07, 2019).
  15. Segal E, Elad D. Candida species and Blastoschizomyces capitus. In: Topley & Wilson's Microbiology and Microbial Infections, 9th ed, Ajello L, Hay RJ (Eds), Arnold, London 1998. Vol 4, p.423.
  16. Bennett RJ, Johnson AD. Completion of a parasexual cycle in Candida albicans by induced chromosome loss in tetraploid strains. EMBO J 2003; 22:2505.
  17. Bennett RJ, Uhl MA, Miller MG, Johnson AD. Identification and characterization of a Candida albicans mating pheromone. Mol Cell Biol 2003; 23:8189.
  18. Thomson GJ, Kakade P, Hirakawa MP, et al. Adaptation to the dietary sugar D-tagatose via genome instability in polyploid Candida albicans cells. G3 (Bethesda) 2021; 11.
  19. Pujol C, Daniels KJ, Lockhart SR, et al. The closely related species Candida albicans and Candida dubliniensis can mate. Eukaryot Cell 2004; 3:1015.
  20. Jones T, Federspiel NA, Chibana H, et al. The diploid genome sequence of Candida albicans. Proc Natl Acad Sci U S A 2004; 101:7329.
  21. Kniemeyer O, Schmidt AD, Vödisch M, et al. Identification of virulence determinants of the human pathogenic fungi Aspergillus fumigatus and Candida albicans by proteomics. Int J Med Microbiol 2011; 301:368.
  22. Shaw KJ, Schell WA, Covel J, et al. In Vitro and In Vivo Evaluation of APX001A/APX001 and Other Gwt1 Inhibitors against Cryptococcus. Antimicrob Agents Chemother 2018; 62.
  23. Palková Z, Váchová L. Yeast cell differentiation: Lessons from pathogenic and non-pathogenic yeasts. Semin Cell Dev Biol 2016; 57:110.
  24. Schell WA. New aspects of emerging fungal pathogens. A multifaceted challenge. Clin Lab Med 1995; 15:365.
  25. Liu K, Howell DN, Perfect JR, Schell WA. Morphologic criteria for the preliminary identification of Fusarium, Paecilomyces, and Acremonium species by histopathology. Am J Clin Pathol 1998; 109:45.
  26. Schell WA. Histopathology of fungal rhinosinusitis. Otolaryngol Clin North Am 2000; 33:251.
  27. Sudbery P, Gow N, Berman J. The distinct morphogenic states of Candida albicans. Trends Microbiol 2004; 12:317.
  28. Veses V, Casanova M, Murgui A, et al. ABG1, a novel and essential Candida albicans gene encoding a vacuolar protein involved in cytokinesis and hyphal branching. Eukaryot Cell 2005; 4:1088.
  29. Chen YL, Yu SJ, Huang HY, et al. Calcineurin controls hyphal growth, virulence, and drug tolerance of Candida tropicalis. Eukaryot Cell 2014; 13:844.
  30. Odds FC. Candida and candidosis: A Review and Bibliography, 2nd ed, Bailliere Tindall, London 1988.
  31. Crampin H, Finley K, Gerami-Nejad M, et al. Candida albicans hyphae have a Spitzenkörper that is distinct from the polarisome found in yeast and pseudohyphae. J Cell Sci 2005; 118:2935.
  32. Stevenson JS, Liu H. Regulation of white and opaque cell-type formation in Candida albicans by Rtt109 and Hst3. Mol Microbiol 2011; 81:1078.
  33. Brenes LR, Lohse MB, Hartooni N, Johnson AD. A Set of Diverse Genes Influence the Frequency of White-Opaque Switching in Candida albicans. G3 (Bethesda) 2020; 10:2593.
  34. Hnisz D, Schwarzmüller T, Kuchler K. Transcriptional loops meet chromatin: a dual-layer network controls white-opaque switching in Candida albicans. Mol Microbiol 2009; 74:1.
  35. Mallick EM, Bergeron AC, Jones SK Jr, et al. Phenotypic Plasticity Regulates Candida albicans Interactions and Virulence in the Vertebrate Host. Front Microbiol 2016; 7:780.
  36. Soll DR. The role of phenotypic switching in the basic biology and pathogenesis of Candida albicans. J Oral Microbiol 2014; 6.
  37. Sasse C, Hasenberg M, Weyler M, et al. White-opaque switching of Candida albicans allows immune evasion in an environment-dependent fashion. Eukaryot Cell 2013; 12:50.
  38. Lohse MB, Johnson AD. White-opaque switching in Candida albicans. Curr Opin Microbiol 2009; 12:650.
  39. Basso V, d'Enfert C, Znaidi S, Bachellier-Bassi S. From Genes to Networks: The Regulatory Circuitry Controlling Candida albicans Morphogenesis. Curr Top Microbiol Immunol 2019; 422:61.
  40. Kornitzer D. Regulation of Candida albicans Hyphal Morphogenesis by Endogenous Signals. J Fungi (Basel) 2019; 5.
  41. De Bernardis F, Mühlschlegel FA, Cassone A, Fonzi WA. The pH of the host niche controls gene expression in and virulence of Candida albicans. Infect Immun 1998; 66:3317.
  42. Fonzi WA. Downstream functions in Candida morphogenesis (abstract). ASM Conference on Candida and Candidiasis, American Society for Microbiology, Charleston, South Carolina 1999. p. 16.
  43. Boral H, Metin B, Döğen A, et al. Overview of selected virulence attributes in Aspergillus fumigatus, Candida albicans, Cryptococcus neoformans, Trichophyton rubrum, and Exophiala dermatitidis. Fungal Genet Biol 2018; 111:92.
  44. Sun Y, Cao C, Jia W, et al. pH Regulates White-Opaque Switching and Sexual Mating in Candida albicans. Eukaryot Cell 2015; 14:1127.
  45. Popolo L, Vai M. Defects in assembly of the extracellular matrix are responsible for altered morphogenesis of a Candida albicans phr1 mutant. J Bacteriol 1998; 180:163.
  46. Davis DA. How human pathogenic fungi sense and adapt to pH: the link to virulence. Curr Opin Microbiol 2009; 12:365.
  47. Si H, Hernday AD, Hirakawa MP, et al. Candida albicans white and opaque cells undergo distinct programs of filamentous growth. PLoS Pathog 2013; 9:e1003210.
  48. Shapiro RS, Uppuluri P, Zaas AK, et al. Hsp90 orchestrates temperature-dependent Candida albicans morphogenesis via Ras1-PKA signaling. Curr Biol 2009; 19:621.
  49. Finkel JS, Mitchell AP. Genetic control of Candida albicans biofilm development. Nat Rev Microbiol 2011; 9:109.
  50. Albuquerque P, Casadevall A. Quorum sensing in fungi--a review. Med Mycol 2012; 50:337.
  51. Alfatah M, Bari VK, Nahar AS, et al. Critical role for CaFEN1 and CaFEN12 of Candida albicans in cell wall integrity and biofilm formation. Sci Rep 2017; 7:40281.
  52. Shapiro RS, Zaas AK, Betancourt-Quiroz M, et al. The Hsp90 co-chaperone Sgt1 governs Candida albicans morphogenesis and drug resistance. PLoS One 2012; 7:e44734.
  53. Wang A, Raniga PP, Lane S, et al. Hyphal chain formation in Candida albicans: Cdc28-Hgc1 phosphorylation of Efg1 represses cell separation genes. Mol Cell Biol 2009; 29:4406.
  54. Lu Y, Su C, Mao X, et al. Efg1-mediated recruitment of NuA4 to promoters is required for hypha-specific Swi/Snf binding and activation in Candida albicans. Mol Biol Cell 2008; 19:4260.
  55. Yaar L, Mevarech M, Koltin Y. A Candida albicans RAS-related gene (CaRSR1) is involved in budding, cell morphogenesis and hypha development. Microbiology 1997; 143 ( Pt 9):3033.
  56. Zheng X, Wang Y, Wang Y. Hgc1, a novel hypha-specific G1 cyclin-related protein regulates Candida albicans hyphal morphogenesis. EMBO J 2004; 23:1845.
  57. Chapa y Lazo B, Bates S, Sudbery P. The G1 cyclin Cln3 regulates morphogenesis in Candida albicans. Eukaryot Cell 2005; 4:90.
  58. Chen J, Zhou S, Wang Q, et al. Crk1, a novel Cdc2-related protein kinase, is required for hyphal development and virulence in Candida albicans. Mol Cell Biol 2000; 20:8696.
  59. Bastidas RJ, Heitman J, Cardenas ME. The protein kinase Tor1 regulates adhesin gene expression in Candida albicans. PLoS Pathog 2009; 5:e1000294.
  60. Cottier F, Mühlschlegel FA. Sensing the environment: response of Candida albicans to the X factor. FEMS Microbiol Lett 2009; 295:1.
  61. Enjalbert B, Whiteway M. Release from quorum-sensing molecules triggers hyphal formation during Candida albicans resumption of growth. Eukaryot Cell 2005; 4:1203.
  62. Hall RA, Cottier F, Mühlschlegel FA. Molecular networks in the fungal pathogen Candida albicans. Adv Appl Microbiol 2009; 67:191.
  63. Hall RA, Turner KJ, Chaloupka J, et al. The quorum-sensing molecules farnesol/homoserine lactone and dodecanol operate via distinct modes of action in Candida albicans. Eukaryot Cell 2011; 10:1034.
  64. Calderone RA, Gow NA. Host recognition by Candida species. In: Candida and candidiasis, Calderone RA (Ed), ASM Press, Washington, DC 2002. p.67.
  65. Lo HJ, Köhler JR, DiDomenico B, et al. Nonfilamentous C. albicans mutants are avirulent. Cell 1997; 90:939.
  66. Gow NA. Germ tube growth of Candida albicans. Curr Top Med Mycol 1997; 8:43.
  67. Davies JM, Stacey AJ, Gilligan CA. Candida albicans hyphal invasion: thigmotropism or chemotropism? FEMS Microbiol Lett 1999; 171:245.
  68. Sahand IH, Moragues MD, Eraso E, et al. Supplementation of CHROMagar Candida medium with Pal's medium for rapid identification of Candida dubliniensis. J Clin Microbiol 2005; 43:5768.
  69. Kumar A, Sachu A, Mohan K, et al. Simple low cost differentiation of Candida auris from Candida haemulonii complex using CHROMagar Candida medium supplemented with Pal's medium. Rev Iberoam Micol 2017; 34:109.
  70. Spivak ES, Hanson KE. Candida auris: an Emerging Fungal Pathogen. J Clin Microbiol 2018; 56.
  71. Keighley C, Garnham K, Harch SAJ, et al. Candida auris: Diagnostic Challenges and Emerging Opportunities for the Clinical Microbiology Laboratory. Curr Fungal Infect Rep 2021; 15:116.
  72. Hamprecht A, Christ S, Oestreicher T, et al. Performance of two MALDI-TOF MS systems for the identification of yeasts isolated from bloodstream infections and cerebrospinal fluids using a time-saving direct transfer protocol. Med Microbiol Immunol 2014; 203:93.
  73. Spanu T, Posteraro B, Fiori B, et al. Direct maldi-tof mass spectrometry assay of blood culture broths for rapid identification of Candida species causing bloodstream infections: an observational study in two large microbiology laboratories. J Clin Microbiol 2012; 50:176.
  74. Vlek A, Kolecka A, Khayhan K, et al. Interlaboratory comparison of sample preparation methods, database expansions, and cutoff values for identification of yeasts by matrix-assisted laser desorption ionization-time of flight mass spectrometry using a yeast test panel. J Clin Microbiol 2014; 52:3023.
  75. Wickes BL, Romanelli AM. Diagnostic Mycology: Xtreme Challenges. J Clin Microbiol 2020; 58.
  76. Lussier M, Sdicu AM, Shahinian S, Bussey H. The Candida albicans KRE9 gene is required for cell wall beta-1, 6-glucan synthesis and is essential for growth on glucose. Proc Natl Acad Sci U S A 1998; 95:9825.
  77. Buurman ET, Westwater C, Hube B, et al. Molecular analysis of CaMnt1p, a mannosyl transferase important for adhesion and virulence of Candida albicans. Proc Natl Acad Sci U S A 1998; 95:7670.
  78. Bulawa CE, Miller DW, Henry LK, Becker JM. Attenuated virulence of chitin-deficient mutants of Candida albicans. Proc Natl Acad Sci U S A 1995; 92:10570.
  79. Chauhan N, Li D, Singh P, et al. The Wall of Candida spp. In: Candida and candidiasis, Calderone RA (Ed), ASM Press, Washington, DC 2002. p.159.
  80. Masuoka J. Surface glycans of Candida albicans and other pathogenic fungi: physiological roles, clinical uses, and experimental challenges. Clin Microbiol Rev 2004; 17:281.
  81. Timpel C, Strahl-Bolsinger S, Ziegelbauer K, Ernst JF. Multiple functions of Pmt1p-mediated protein O-mannosylation in the fungal pathogen Candida albicans. J Biol Chem 1998; 273:20837.
  82. Csank C, Schröppel K, Leberer E, et al. Roles of the Candida albicans mitogen-activated protein kinase homolog, Cek1p, in hyphal development and systemic candidiasis. Infect Immun 1998; 66:2713.
  83. Pfaller MA, Diekema DJ, Turnidge JD, et al. Twenty Years of the SENTRY Antifungal Surveillance Program: Results for Candida Species From 1997-2016. Open Forum Infect Dis 2019; 6:S79.
  84. Kadosh D. Regulatory mechanisms controlling morphology and pathogenesis in Candida albicans. Curr Opin Microbiol 2019; 52:27.
  85. Calderone RA, Fonzi WA. Virulence factors of Candida albicans. Trends Microbiol 2001; 9:327.
  86. Staniszewska M. Virulence Factors in Candida species. Curr Protein Pept Sci 2020; 21:313.
  87. Blankenship JR, Wormley FL, Boyce MK, et al. Calcineurin is essential for Candida albicans survival in serum and virulence. Eukaryot Cell 2003; 2:422.
  88. Blankenship JR, Heitman J. Calcineurin is required for Candida albicans to survive calcium stress in serum. Infect Immun 2005; 73:5767.
  89. O'Meara TR, Robbins N, Cowen LE. The Hsp90 Chaperone Network Modulates Candida Virulence Traits. Trends Microbiol 2017; 25:809.
  90. Chen YL, Brand A, Morrison EL, et al. Calcineurin controls drug tolerance, hyphal growth, and virulence in Candida dubliniensis. Eukaryot Cell 2011; 10:803.
  91. Robbins N, Uppuluri P, Nett J, et al. Hsp90 governs dispersion and drug resistance of fungal biofilms. PLoS Pathog 2011; 7:e1002257.
  92. McLellan CA, Whitesell L, King OD, et al. Inhibiting GPI anchor biosynthesis in fungi stresses the endoplasmic reticulum and enhances immunogenicity. ACS Chem Biol 2012; 7:1520.
  93. Kim S, Nguyen QB, Wolyniak MJ, et al. Release of transcriptional repression through the HCR promoter region confers uniform expression of HWP1 on surfaces of Candida albicans germ tubes. PLoS One 2018; 13:e0192260.
  94. Rodríguez-Cerdeira C, Martínez-Herrera E, Carnero-Gregorio M, et al. Pathogenesis and Clinical Relevance of Candida Biofilms in Vulvovaginal Candidiasis. Front Microbiol 2020; 11:544480.
  95. Staab JF, Bradway SD, Fidel PL, Sundstrom P. Adhesive and mammalian transglutaminase substrate properties of Candida albicans Hwp1. Science 1999; 283:1535.
  96. Pappas PG, Lionakis MS, Arendrup MC, et al. Invasive candidiasis. Nat Rev Dis Primers 2018; 4:18026.
  97. Dutton LC, Nobbs AH, Jepson K, et al. O-mannosylation in Candida albicans enables development of interkingdom biofilm communities. mBio 2014; 5:e00911.
  98. Thammahong A, Puttikamonkul S, Perfect JR, et al. Central Role of the Trehalose Biosynthesis Pathway in the Pathogenesis of Human Fungal Infections: Opportunities and Challenges for Therapeutic Development. Microbiol Mol Biol Rev 2017; 81.
  99. Hube B, Naglik J. Extracellular hydrolases. In: Candida and candidiasis, Calderone RA (Ed), ASM Press, Washington, DC 2002. p.107.
  100. Taylor BN, Staib P, Binder A, et al. Profile of Candida albicans-secreted aspartic proteinase elicited during vaginal infection. Infect Immun 2005; 73:1828.
  101. Schaller M, Korting HC, Borelli C, et al. Candida albicans-secreted aspartic proteinases modify the epithelial cytokine response in an in vitro model of vaginal candidiasis. Infect Immun 2005; 73:2758.
  102. Chen YC, Wu CC, Chung WL, Lee FJ. Differential secretion of Sap4-6 proteins in Candida albicans during hyphae formation. Microbiology 2002; 148:3743.
  103. Felk A, Kretschmar M, Albrecht A, et al. Candida albicans hyphal formation and the expression of the Efg1-regulated proteinases Sap4 to Sap6 are required for the invasion of parenchymal organs. Infect Immun 2002; 70:3689.
  104. Sundstrom P. Adhesins in Candida albicans. Curr Opin Microbiol 1999; 2:353.
  105. Sanglard D, Hube B, Monod M, et al. A triple deletion of the secreted aspartyl proteinase genes SAP4, SAP5, and SAP6 of Candida albicans causes attenuated virulence. Infect Immun 1997; 65:3539.
  106. Ghannoum MA. Potential role of phospholipases in virulence and fungal pathogenesis. Clin Microbiol Rev 2000; 13:122.
  107. Sundstrom P, Balish E, Allen CM. Essential role of the Candida albicans transglutaminase substrate, hyphal wall protein 1, in lethal oroesophageal candidiasis in immunodeficient mice. J Infect Dis 2002; 185:521.
  108. Stichternoth C, Ernst JF. Hypoxic adaptation by Efg1 regulates biofilm formation by Candida albicans. Appl Environ Microbiol 2009; 75:3663.
  109. Grahl N, Cramer RA Jr. Regulation of hypoxia adaptation: an overlooked virulence attribute of pathogenic fungi? Med Mycol 2010; 48:1.
  110. Pradhan A, Avelar GM, Bain JM, et al. Hypoxia Promotes Immune Evasion by Triggering β-Glucan Masking on the Candida albicans Cell Surface via Mitochondrial and cAMP-Protein Kinase A Signaling. mBio 2018; 9.
  111. Holland SM, Vinh DC. Yeast infections--human genetics on the rise. N Engl J Med 2009; 361:1798.
  112. Davidson L, Netea MG, Kullberg BJ. Patient Susceptibility to Candidiasis-A Potential for Adjunctive Immunotherapy. J Fungi (Basel) 2018; 4.
  113. Wang X, van de Veerdonk FL, Netea MG. Basic Genetics and Immunology of Candida Infections. Infect Dis Clin North Am 2016; 30:85.
  114. Thompson A, Orr SJ. Emerging IL-12 family cytokines in the fight against fungal infections. Cytokine 2018; 111:398.
  115. Bojang E, Ghuman H, Kumwenda P, Hall RA. Immune Sensing of Candida albicans. J Fungi (Basel) 2021; 7.
  116. Gabrielli E, Sabbatini S, Roselletti E, et al. In vivo induction of neutrophil chemotaxis by secretory aspartyl proteinases of Candida albicans. Virulence 2016; 7:819.
  117. Mencacci A, Cenci E, Del Sero G, et al. IL-10 is required for development of protective Th1 responses in IL-12-deficient mice upon Candida albicans infection. J Immunol 1998; 161:6228.
  118. Netea MG, Joosten LA, van der Meer JW, et al. Immune defence against Candida fungal infections. Nat Rev Immunol 2015; 15:630.
  119. Lavigne LM, Schopf LR, Chung CL, et al. The role of recombinant murine IL-12 and IFN-gamma in the pathogenesis of a murine systemic Candida albicans infection. J Immunol 1998; 160:284.
  120. Gozalbo D, Maneu V, Gil ML. Role of IFN-gamma in immune responses to Candida albicans infections. Front Biosci (Landmark Ed) 2014; 19:1279.
  121. Richardson JP, Moyes DL. Adaptive immune responses to Candida albicans infection. Virulence 2015; 6:327.
  122. Vazquez-Torres A, Jones-Carson J, Wagner RD, et al. Early resistance of interleukin-10 knockout mice to acute systemic candidiasis. Infect Immun 1999; 67:670.
  123. Roilides E, Holmes A, Blake C, et al. Effects of granulocyte colony-stimulating factor and interferon-gamma on antifungal activity of human polymorphonuclear neutrophils against pseudohyphae of different medically important Candida species. J Leukoc Biol 1995; 57:651.
  124. Gaviria JM, van Burik JA, Dale DC, et al. Modulation of neutrophil-mediated activity against the pseudohyphal form of Candida albicans by granulocyte colony-stimulating factor (G-CSF) administered in vivo. J Infect Dis 1999; 179:1301.
  125. Ueta E, Tanida T, Doi S, Osaki T. Regulation of Candida albicans growth and adhesion by saliva. J Lab Clin Med 2000; 136:66.
  126. Martínez JP, Gil ML, López-Ribot JL, Chaffin WL. Serologic response to cell wall mannoproteins and proteins of Candida albicans. Clin Microbiol Rev 1998; 11:121.
  127. Casadevall A, Cassone A, Bistoni F, et al. Antibody and/or cell-mediated immunity, protective mechanisms in fungal disease: an ongoing dilemma or an unnecessary dispute? Med Mycol 1998; 36 Suppl 1:95.
  128. Netea MG, Brown GD, Kullberg BJ, Gow NA. An integrated model of the recognition of Candida albicans by the innate immune system. Nat Rev Microbiol 2008; 6:67.
  129. Han Y, Cutler JE. Assessment of a mouse model of neutropenia and the effect of an anti-candidiasis monoclonal antibody in these animals. J Infect Dis 1997; 175:1169.
  130. Cutler JE. Candida albicans antibodies and vaccine development (abstract). ASM Conference on Candida and Candidiasis, American Society for Microbiology, Charleston, South Carolina 1999, p. 20.
  131. Zhang MX, Kozel TR. Mannan-specific immunoglobulin G antibodies in normal human serum accelerate binding of C3 to Candida albicans via the alternative complement pathway. Infect Immun 1998; 66:4845.
  132. Spellberg BJ, Ibrahim AS, Avenissian V, et al. The anti-Candida albicans vaccine composed of the recombinant N terminus of Als1p reduces fungal burden and improves survival in both immunocompetent and immunocompromised mice. Infect Immun 2005; 73:6191.
  133. Gantner BN, Simmons RM, Canavera SJ, et al. Collaborative induction of inflammatory responses by dectin-1 and Toll-like receptor 2. J Exp Med 2003; 197:1107.
  134. Kimberg M, Brown GD. Dectin-1 and its role in antifungal immunity. Med Mycol 2008; 46:631.
  135. Gil ML, Gozalbo D. Role of Toll-like receptors in systemic Candida albicans infections. Front Biosci (Landmark Ed) 2009; 14:570.
Topic 2464 Version 22.0

References