Phenotypic Characterization and Gene Expression Analyses of a Penicillium marneffei Septin Mutant by Daniel E. Kennedy, II Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science in the Biological Sciences Program YOUNGSTOWN STATE UNIVERSITY May, 2012 Phenotypic Characterization and Gene Expression Analyses of a Penicillium marneffei Septin Mutant Daniel E. Kennedy, II I hereby release this thesis to the public. I understand that this thesis will be made available from the OhioLINK ETD Center and the Maag Library Circulation Desk for public access. I also authorize the University or other individuals to make copies of this thesis as needed for scholarly research. Signature: Daniel E. Kennedy, II, Student Date Approvals: Dr. Chester Cooper, Thesis Advisor Date Dr. Gary Walker, Committee Member Date Dr. Jonathan Caguiat, Committee Member Date Dr. David Asch, Committee Member Date Peter J. Kasvinsky, Dean of School of Graduate Studies and Research Date ii ABSTRACT Penicillium marneffei is the only known thermally dimorphic species of the Penicillium genus, growing as a filamentous mold at 25˚C and adopting yeast morphology at 37˚C. Endemic to Southeast Asia, there is no known reservoir for P. marneffei. Nonetheless, the evolutionary adaptation of phase transition has allowed P. marneffei to become a highly aggressive, opportunistic pathogen. Since the emergence of AIDS, P. marneffei infections have become increasingly common due to its increased virulence when invading immunocompromised individuals with most cases being reported in Thailand. Phase transition in P. marneffei is tightly regulated at the transcriptional level. Gene expression in P. marneffei is regulated when subjected to higher or lower temperatures. These thermally dependent, phase specific regulatory mechanisms are of current interest, as understanding such mechanisms will potentially provide targets for chemotherapeutic agents in treating P. marneffei infections in the future. Presently, an Agrobacterium tumefaciens-mediated transformation (AMT)-generated mutant in septin function is being studied in order to establish its role in phase transition and normal vegetative growth. Preliminary results indicate that the disruption mutation in the aspC gene caused aberrant morphological characteristics. In addition, aspC expression in the aspC mutant was greater than that in the wild-type strain. iii ACKNOWLEDGEMENTS Dr. Chet Cooper Thank you for allowing me to work in your lab. Your advice has helped me keep life in perspective and will help guide me both in my personal and professional endeavors. The skills I have acquired under your guidance will benefit me the rest of my career. From providing recommendations to tolerating my shenanigans, I appreciate everything you have done for me. Drs. Caguiat, Walker, and Asch Thank you for answering the multitude of questions I have had during my time in the Proteomics and Genomics Research Group. You have all contributed to my development as a scientist. Andrew Holmes I knew only a few basic laboratory techniques upon entering the lab as a graduate student. You taught me the importance of being independent, self-reliant, and the ability to improvise when necessary in order to accomplish goals. Dr. Lorna Gallagher Thank you for providing me with a different perspective on science. Your assistance made it possible for me to complete my project in a timely fashion. iv TABLE OF CONTENTS ABSTRACT………………………………………………………………………………ii ACKNOWLEDGEMENTS ……………………………………………………………...iii TABLE OF CONTENTS………………………………………………………………....iv LIST OF FIGURES…………………………………………………………………..…...v LIST OF TABLES……………………………………………………………………….vi CHAPTERS I. Introduction……………………………………………………………………….1 II. Materials and Methods…………………………………………………………...25 III. Results…………………………………………………………………………....32 IV. Discussion………………………………………………………………………..61 V. References………………………………………………………………………..69 v LIST OF FIGURES Figure Page 1. Comparison of seven day old wild-type and mutant strains cultured on PDA 33 2. CR disruption assay of wild-type strain 34 3. CR disruption assay of aspC mutant strain 35 4. H 2 O 2 disruption assay of wild-type strain 35 5. H 2 O 2 disruption assay of aspC mutant strain 36 6. CW disruption assay of wild-type strain 37 7. CW disruption assay of aspC mutant strain 38 8. CR disruption assay of wild-type and aspC mutant strains 38 9. Wild-type slide cultures 41 10. aspC mutant slide cultures 42 11. Wild-type H 2 O 2 slide cultures 43 12. aspC mutant H 2 O 2 slide cultures 44 13. Wild-type CR slide cultures 45 14. aspC mutant CR slide cultures 46 15. Wild-type CW slide cultures 47 16. aspC mutant CW slide cultures 48 17. Differential expression of septin genes in wild-type strain 51 18. Differential expression of aspC in wild-type 52 19. Differential expression of septin genes in wild-type and aspC mutant 53 20. Differential expression of aspC in wild-type and mutant strains 54 21. Differential expression of septin genes in wild-type and mutant strains in H 2 O 2 55 22. Differential expression of aspC in wild-type and mutant strains in H 2 O 2 56 23. Differential expression of septin genes in wild-type and mutant strains in CR 57 24. Differential expression of aspC in wild-type and mutant strains in CR 58 25. Differential expression of septin genes in wild-type and mutant strains in CW 59 vi 26. Differential expression of aspC in wild-type and mutant strains in CW 60 LIST OF TABLES Table Page 1. Primers used in qPCR 31 1 Chapter I Introduction Fungi are highly active participants in most of the world’s ecosystems. The roles of most fungi are limited to degrading organic matter in their respective environments. Some fungi exist as symbionts with plants and other organisms, while others are harmful, opportunistic pathogens. While most Penicillium species appear to be quite common and are not involved in infection of humans, several species from the Penicillium genus, including Penicillium marneffei, have been documented as causative agents of a condition called penicilliosis (Cooper and Haycocks 2000). However, P. marneffei differs from other Penicillium species in that it is the only thermally dimorphic species of the genus, undergoing phase transition, also termed the dimorphic switch, after entering the host organism’s body. Penicillium marneffei typically grows as a mold at temperatures between 8.0˚C and 31.0˚C (Cao et al. 2007). After internalization of P. marneffei, however, phase transition occurs, with concomitant changes in morphology. In addition, up- or down- regulation of various genes can be observed, as P. marneffei further develops as a yeast. Penicilliosis due to P. marneffei, while extremely rare in those with healthy immune systems, has proven terminal without treatment in those with and without suppressed immune systems. Penicilliosis, which usually manifests itself as a systemic yeast infection in its latter stages, has also become the third most common AIDS-defining illness (Supparatpinyo et al. 1994). With penicilliosis due to P. marneffei on the rise, the P. marneffei genome, including genes coding for proteins involved in cell wall construction and maintenance, phase transition, and those proteins that allow P. marneffei to survive in host organism phagocytic cells, has recently become a subject of 2 interest in a search for potential targets of chemotherapeutic agents. However, in order for such genes or their corresponding proteins to be targeted, P. marneffei’s genome should be further annotated. History and Epidemiology of P. marneffei Penicillium marneffei was initially isolated from a bamboo rat grown in captivity for experimental infections at the Pasteur Institute of Indochina, Dalat, South Vietnam in 1956 (Capponi et al. 1956). Despite experiments designed to reveal the native reservoir of P. marneffei, results have not yielded a definitive origin. Rather, it seems the conditions related to infection are host-specific factors along with environmental risk factors; age and an agricultural occupation are the most common independent factors (Chariyalerstak et al. 1997). Specifically, exposure to soil during the nearly seven months-long rainy season of Southeast Asia, from May through October, appear to have the highest corollary effect of any risk factor (Chariyalerstak et al. 1996). However, similar experiments designed to uncover the correlation of infection in native bamboo rats to infection in native human hosts have also failed to elucidate any known vectors of transmission (Cooper 1998; Chariyalerstak et al. 1996). These findings point to P. marneffei infections existing as a result of sapronosis, rather than zoonosis (Vanittanakom et al. 2006). In other words, an environmental reservoir should exist, rather than direct rat-human interactions being responsible for infection. The trail of infection is further convoluted when taking into account that many experiments attempting to isolate P. marneffei from soil samples near areas home to infected human and rat hosts alike have failed to yield positive results (Joshi et al. 2003; Vanittanakom et al. 1995). Vanittanakom et al. demonstrated P. marneffei could survive several weeks in 3 sterilized soil, while P. marneffei’s survival was reduced to only a few days grown in non-sterilized soil. Such findings indicate that competing fungi decrease the abundance of P. marneffei, thus limiting the potential for a native reservoir in Thailand. The first documented case of the disease was reported by Professor Gabriel Segretain in 1959, who also identified the fungus as a new Penicillium species (Segretain 1959). It was during a laboratory experiment that he accidentally inoculated himself with a fungus-contaminated needle and later developed a nodule at the site of infection, lymphangitis, and lymphadenopathy. The first naturally occurring case was described in 1973 in an American minister with Hodgkin’s disease who had lived in Southeast Asia (DiSalvo et al. 1973). Subsequently, mostly due to the rise of HIV/AIDS infection, increased travel to southeast Asia and non-sterile intravenous drug use engaged in by some of those who do, cases of penicilliosis due to P. marneffei infection have been documented in Cambodia (Bailloud et al. 2002), China (Liao et al. 2002), Hong Kong (Chang et al. 1998; Ko 1994; Tsang et al.1991; Tsui 1992; Wong et al. 1998), India (Ranjana et al. 2002), Malaysia (Rokiah et al. 1995), Taiwan (Chang et al. 1995; Chiang et al. 1998; Hsueh et al. 2000; Hung et al. 1998; Liu et al. 1994), Thailand (Sathapatayavongs et al. 1989), Vietnam (Hien et al. 2001; Huynh et al. 2003), Australia (Heath et al. 1995; Jones and See 1992), Belgium (Depraetere et al. 1998), France (Grise et al. 1997; Hilmarsdottir et al. 1994; Hilmarsodottir et al. 1993; Valeyrie et al. 1999), Germany (Rimek et al. 1999; Sobottka et al. 1996), Japan (Mohri et al. 2000; Tsunemi et al. 2003), Sweden (Julander and Petrini 1997), Switzerland (Borradori et al. 1994; Garbino et al. 2001; Kronauer et al. 1993), The Netherlands (Hulshof et al. 1990; Kok et 4 al. 1994), the United Kingdom (Bateman et al. 2002; McShane et al.1998; Peto et al. 1988; Vilar et al. 2000), and the United States (Nord et al. 1998). Medical Significance Although most P. marneffei related cases of penicilliosis have occurred in Southeast Asia, the source of P. marneffei remains largely unknown. Globally, P. marneffei has proven itself highly pathogenic in AIDS patients, becoming the third most common AIDS-defining illness and yielding the third highest rate of opportunistic pathogenesis among studied AIDS patients (Cooper and Haycocks 2000). Penicilliosis in both HIV-positive and HIV-negative individuals results in death if left untreated. However, penicilliosis in HIV-positive individuals requires a much more lengthy treatment protocol than in their HIV-negative counterparts, due to the commonality of relapse after cessation of antifungal therapy. Penicilliosis in HIV-positive individuals usually occurs in the advanced stages of HIV infection due to the depletion of CD4+ T cells, and those infected typically exhibit symptoms including fever, anemia, leukopenia, weight loss, diarrhea, cough, massive hepatosplenomegaly, and generalized lymphadenopathy (Noritomi et al. 2005; Desakorn et al. 2002). Affected patients also present characteristic, multiple, umbilicated skin lesions on the face and torso. Treatment of penicilliosis due to P. marneffei involves lengthy rounds of oral and intravenous antifungal agents and, in some cases, surgery (Cooper and Haycocks 2000). The pathogenesis of P. marneffei makes treatment of infections highly difficult, increasing the urgency in understanding this opportunistic pathogen. 5 General Biology, Dimorphism, and Pathogenicity As with other fungi, P. marneffei’s ability to regulate cell wall construction is critical for growth. Because cell wall construction is so important to the survival of P. marneffei, it should come as no surprise that the process is tightly regulated. Cell wall synthesis involves simple molecules arranged into conglomerate complexes and enzymes, the latter being responsible for arranging those molecules in such intricate patterns. Although P. marneffei’s cell wall composition has yet to be completely characterized, a general pattern of cell wall construction for each phylum of fungi has been established. Like many studied models of the Phylum Ascomycota, P. marneffei’s cell wall is presumed to be composed of two polysaccharides, namely chitin and glucans, and mannoproteins (Applegarth and Bozoian 1967; Bacon et al. 1960; Domer et al. 1967; Kanetsuna et al. 1969). These macromolecules can be further divided into structural and matrix components, the latter serving to cross-link the former. Chitin and β-(1→ 3) - and β-(1→ 6)-glucans provide structural support in the form of microfibrils, while mannoproteins and α (1→ 3)-glucan are found in the matrix of the cell wall. Because the mold phase morphology manifests itself as apical growth at the hyphal tip, also termed the extension zone, those polysaccharides that provide the cell wall rigidity, chitin and β- glucans, are abundant in the cell wall behind the extension zone but exist in relatively low levels in the extension zone cell wall itself. More precisely, chitin chains are synthesized at the apex by chitin synthase and subsequently inserted into the cell wall as the apex continues to grow. Similarly, β-(1→ 3)-glucan chains are synthesized at the apex by glucan synthase, inserted into the wall, and further linked to one another by β-(1→ 6)- linked side chains. Structural polysaccharides aggregate laterally into increasingly 6 complex crystalline microfibrils at higher levels farther from the extension zone (Pancaldi et al. 1984). Increasing the structural polysaccharide content in the hyphal cell wall allows the fungus to maintain rigidity (i.e., protection) in the non-expanding portion of its cell wall, while growth continues at the tip. Cell wall synthetic enzymes, chitin synthase and glucan synthase, are most likely transported to the apical cell wall in vesicles (Deacon 2006). Once embedded in the cell wall chitin synthase is likely activated via protease. Cytosolic N-acetylglucosamine (GlcNAc) is bound to a sugar nucleotide substrate, uridine triphosphate (UTP), yielding UDP-GlcNAc, which is then added to an elongating chitin chain, in a reaction cleaving UDP, by chitin synthase. Comparatively, glucan synthase uses another sugar nucleotide substrate, cyotosolic UDP-glucose, to synthesize glucan chains. In contrast with chitin synthase, however, glucan synthase is a dimeric G-protein. Though, in both cases of microfibrillar formation, chitin and glucan chains are extruded through to the outer surface of the membrane and into the cell wall, where they provide structural integrity. Perhaps the most puzzling trait of P. marneffei is its mode of pathogenesis. Penicillium marneffei grows as mold and yeast forms at 25˚C and 37˚C, respectively, and is the only known thermally dimorphic species of the Penicillium genus (Canovas and Andrianopoulos 2007). At 25˚C, P. marneffei displays asexual reproduction, during which a filamentous body, the hypha, exhibits periodic branching and apical growth. From these filamentous bodies arise conidiophores, composed of specialized cells (Andrianopoulos 2002). Secondary stalks called rama may also be produced from basal, vegetative, hyphal cells. Rama then differentiate to produce the sterigmata cell types: metulae and phialides. Distinctive from vegetative, apically-growing, multinucleate 7 hyphal cells, metulae and phialides are produced via budding division and are uninucleate. In turn, the metula then buds one or more phialides. Phialides then generate conidia in a basipetal manner, producing chains of conidia. However, should the nutritional requirements (e.g., amount or type of carbon source) not be met, rama may be absent. Consequently, sterigmata bud directly from the hyphal cell, or shorter, unbranched stalks may produce conidiating stergimata cells. In addition, carbon-limiting conditions also cause conidiophore production earlier than would occur otherwise. Once conidia have been freed from the phialide, they disperse and may germinate once finding suitable growth conditions. These conidia grow isotropically at first, then produce a germ tube. The latter marks the beginning of polarized hyphal growth (Canovas and Andrianopoulos 2007). Asexual development in P. marneffei is controlled by a number of regulatory and signaling genes, some of which have been characterized. Highly similar homologues of these genes have been well characterized in Aspergillus nidulans, including brlA and abaA, with the former activating the latter (Adams et al. 1988). In P. marneffei, brlA is believed to encode a C 2 H 2 zinc finger protein, which consequently activates abaA (Adams et al. 1998). The abaA gene product is an ATTS/TEA protein (Andrianopoulos and Timberlake 1994; Burglin 1991). ATTS/TEA proteins in P. marneffei are DNA- domain transcriptional regulators suspected of playing roles in cell- cycle events, conidiation, and hyphal-yeast development (Borneman et al. 2000). AbaA, in turn, causes vegetative hyphae to begin conidiophore production (Borneman et al. 2000) and removes conidiation from the influence of other inductive signals (Cooper and Vanittanakom 2008). abaA deletion strains of P. marneffei were shown to be deficient in conidiation, producing conidiophore stalks possessing abnormal sterigmata cells 8 incapable of producing conidia (Borneman et al. 2000). abaA, as well as brlA, is further regulated by a member of the APSES family of transcription factors, stuA, which is responsible for sterigmata, metula, and phialide development during conidiation (Borneman et al. 2002). stuA contains a basic helix-loop-helix (bHLH) DNA binding motif. Experimentation has shown that, while the stuA transcript was undetectable during vegetative hyphal growth, it was highly induced once the conidiation program had begun. stuA deletion strains exhibited conidiophores lacking the characteristic sterigmata cells. However, conidia continued to be produced basipetally directly from the tips of conidiophore stalks, showing stuA to be essential for sterigmata cell differentiation but not for conidiation. Adding to the complexity of the gene regulation pathway during asexual development are gasA, cflA, and cflB. The Gα subunit of GasA inhibits asexual development by repressing brlA, maintaining a state of vegetative growth (Zuber et al. 2002). In contrast, P. marneffei Rho GTPases Cdc42 and Rac homologues, cflA and cflB, respectively, carry out overlapping functions affecting growth at both 25˚C and 37˚C (Boyce et al. 2001; Boyce et al. 2003). cflA is required for germination of conidia at 25˚C and 37˚C, polarized growth and division of hyphae at 25˚C, activating the p21 activated kinases (PAKs) pakA and pakB (Richman et al. 1999; Lamson et al. 2002; Benton et al. 1997; Gladfelter et al.2004; Peter et al. 1996), and polarized growth of yeast cells at 37˚C. Like cflA, cflB is also responsible for cellular polarization and division during hyphal growth, but, unlike cflA, cflB also plays a role in conidiophore cell differentiation to the extent that cflB deletion strains result in yeast-like morphology and inappropriate yeast cell production at 25˚C (Boyce et al. 2009; Boyce et al. 2003). Similarly, CflB has experimentally been shown to colocalize with actin in the tips of 9 vegetative hyphal cells and at sites of septation, and, while CflB is not required for localization of cortical spots and actin at nascent septation sites, it is required for actin localization at the hyphal tip (Boyce et al. 2003). However, cflB deletion in similar experiments did not affect yeast cell production, with cellular polarization occurring relatively normally at 37˚C. The two aforementioned P. marneffei PAKs are associated with a multitude of signaling and morphogenic activities, which play a role in the more severe pathological effects caused by P. marneffei. For instance, in pakA deletion strains, conidial germination failed to occur after conidia had been phagocytized by macrophages, limiting the ability of intramacrophage hyphal growth and subsequent arthroconidiation, strongly suggesting that at least systemic penicilliosis due to P. marneffei requires pakA expression. Related experimentation has also shown pakA to be essential for conidial germination and correct yeast cell morphogenesis at 37˚C, and is required for conidial germination in vivo (Boyce and Andrianopoulos 2007). Experimental results have also revealed pakB to be an integral cell polarity gene during infection (in vivo), but not in vitro (Boyce et al. 2009). Macrophages infected with wild- type P. marneffei conidia contained yeast cells, whereas macrophages infected with pakB deletion strains contained highly branched, septate, hyphal cells but no yeast cells. This data suggests that PakB is essential for pathogenicity by P. marneffei. Those afflicted with penicilliosis have most likely inhaled the conidia. Once a conidium reaches an alveolus, its ability to recognize fibronectin (Vanittanakom et al. 2006) and bind laminin in the bronchoalveolar epithelium via a scialic acid-specific lectin is suspected to aid adherence (Rongrungruang and Levitz 1999; Boyce et al. 2009). The conidium is to be phagocytized by monocyte-derived, alveolar macrophages, thus 10 initiating infection and potentially causing systemic penicilliosis in the host. Once the conidia have been subjected to an environmental temperature of 32˚C to 37˚C (Cao et al. 2007), the dimorphic switch occurs. Penicillium marneffei morphology and physiology at 37˚C differ markedly from that at 25˚C. In a process termed arthroconidiation, cellular and nuclear division become coupled (Borneman et al. 2000), and the hyphae form an additional internal septum at already existing septa, creating a double septum and a means to free the arthroconidia (Chan and Chow 1990; Garrison and Boyd 1973). Once the arthroconidia have been freed from the hyphal body they enter the yeast phase, during which cells grow in a less polarized manner than germinating conidia nor grow apically as hyphae do, but divide via fission (Chan and Chow 1990; Garrison and Boyd 1973; Pitt 1979; Segretain 1959). The arthroconidia represent the pathogenic growth phase of P. marneffei. However, subjecting the yeast phase cells to 25˚C reverses the dimorphic switch, causing arthroconidia to polarize and begin apical growth; cellular and nuclear growth will uncouple; and mold phase growth will be resumed (Cooper and Haycocks 2000; Canovas and Andrianopoulos 2007). It is the dimorphic switch that makes P. marneffei unique among the Penicillia. As with asexual development, yeast cell morphogenesis in P. marneffei is tightly controlled. The previously described hyphal-yeast dimorphic switch is most notably characterized by the coupling of cellular and nuclear division, septation of cells leading to cell separation, and division of subsequent yeast cells via fission (Andrianopoulos 2002). Unlike during asexual development, stuA does not play a role dimorphic switching in P. marneffei (Borneman et al. 2001; Borneman et al. 2002). Conversely, the inhibitory effects of TupA, a transcription factor that is also homologous to genes associated with 11 the regulation of asexual development in other filamentous fungi, that normally repress the activation of abaA, and thus the dimorphic switch, at 25˚C, are overcome by the exposure to the increased temperature cue of 37˚C (Todd et al. 2003). Consequently, abaA is expressed, and AbaA, the ATTS protein, causes coupling of cellular and nuclear division. Similarly, the effects of CflA, PakB, and PakA aid in both conidial germination and polarized growth, once conidia have been phagocytized by macrophages. As P. marneffei is the only dimorphic Penicillium species it is important to keep in mind that the dimorphic switch is not only an identifying characteristic of the species, but one of many factors that have allowed P. marneffei to become such a virulent pathogen of mammals. In the context of human- fungal interactions, P. marneffei’s genome is of interest due to the tightly controlled manner in which this fungus grows. Once gene functions have been elucidated, they may serve as potential drug targets. Some of P. marneffei’s genome has already been annotated, including some genes that serve as virulence factors. Possible Virulence Factors of P. marneffei Other Penicillium spp. have been documented as fungal pathogens of humans (Ramirez 1982). However, based on clinical diagnostic testing, P. marneffei is responsible for the overwhelming majority of penicilliosis cases. Penicillium marneffei is only one of about 40 Penicillia species that affect HIV-negative individuals and one of a handful that have been documented as opportunistic pathogens of those who are HIV- positive (Cooper and Haycocks 2000). This raises the question of why P. marneffei accounts for such a large percentage of penicilliosis cases. The answer can possibly be 12 found in P. marneffei’s physiology. The dimorphic switch itself appears to be the primary virulence factor of P. marneffei, as the switch is required for yeast cell morphogenesis and, hence, for systemic pathogenesis in host organisms (Canovas and Andrianopoulos 2007; Andrianopoulos 2002; Boyce et al. 2009). Other physiological characteristics of P. marneffei may also play a role in its virulence. Upon phagocytosis by macrophages conidia are subjected to environmentally unfavorable conditions inside phagosomes (Pongpom et al. 2005). These organelles are typically the site of reactive oxygen species (ROS) production, including superoxide, hydrogen peroxide, and hydroxyl radicals, termed respiratory burst, during infection. As a result, these ROS act on, and ultimately destroy, the conidia and yeast cells contained within the phagosome. While the exact defense mechanism utilized by P. marneffei has not been completely charted, a possible answer lies in the cpeA gene, which encodes a catalase-peroxidase enzyme that exhibits both catalase activity (the ability to break down hydrogen peroxide into water and molecular oxygen) and peroxidase activity (the ability to use hydrogen peroxide to oxidize other compounds). While most other microorganisms are known to possess similar enzymes, although usually as independent catalase and peroxidase enzymes, some bacteria and fungi possess the bifunctional form of the enzyme. Partial experimental characterization of the P. marneffei yeast phase proteome has revealed an increase in the expression of catalase-peroxidase relative to that of the mold phase. In addition, the P. marneffei catalase-peroxidase amino acid sequence has been shown to be 45-69% identical to that found in other fungi and bacteria. Pharmacological studies have also reported similar findings (Taramelli et al. 2001). The antimicrobial chloroquine was shown to raise the intravacuolar pH of studied 13 phagosomes, binding the free protons within the phagosomes and possibly inhibiting the vacuolar proton- ATPase, which maintains a proton gradient across the vacuolar membrane and perpetuates an acidic intravacuolar environment. Both of the aforementioned effects contribute to raising the intravacuolar pH. Because P. marneffei grows optimally at an acidic pH (5.5-6.0) the increased pH most likely inhibits yeast cells’ potential for survival. The observed increase in intravacuolar pH may also inhibit pH-dependent yeast virulence factors. Penicillium marneffei has been documented to express acid phosphatase activity, which has also been reported in other intracellular pathogens and whose optimal pH is 5. Acid phosphatase activity is suspected to aid in the survival of intracellular pathogens by limiting the previously described respiratory burst that occurs in the phagosomes of phagocytic cells. These data suggest that catalase- peroxidase plays a major role in the survival of P. marneffei in the macrophage phagosome, making possible for infection to continue and potentially develop into the more severe systemic penicilliosis. In addition to catalase-peroxidase activity, P. marneffei also activates the copper, zinc-superoxide dismutase enzyme encoding gene, sodA, upon internalization by macrophages (Thirach et al. 2008). Up-regulation of this gene during yeast cell development may also aid in P. marneffei’s ability to survive the harsh conditions encountered during intracellular growth. Yeast cells are also subjected to conditions of inadequate nutrient supplies within macrophage phagosomes. In order to combat such conditions, P. marneffei differentially expresses genes that aid in utilizing non-preferred carbon sources (Canovas and Andrianopoulos 2006; Thirach et al. 2008). Experimental data has proven the differential 14 expression of the glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and the glyoxylate cycle enzyme isocitrate lyase encoding genes, gpdA and acuD, respectively, after induction of dimorphic switching (Thirach et al. 2008). Mold phase cells exhibited significantly higher levels of GAPDH than their yeast phase counterparts. Conversely, yeast phase cells produced more isocitrate lyase than those of the mold phase. Because GAPDH has been associated with glycolysis, a pathway typically implemented under conditions that provide organisms with sufficient sources of carbon, the down-regulation of gpdA in yeast cells suggests that intracellular macrophage conditions are not conducive to the growth of P. marneffei during infection. However, in typical evolutionary fashion of many pathogens, P. marneffei has evolved to overcome the hostile environment of its host’s phagocytic cells by utilizing alternative sources of energy. During yeast cell development acuD is expressed. Consequently, isocitrate lyase is produced, allowing implementation of the glyoxylate cycle. Isocitrate lyase affords P. marneffei the ability to bypass one of two decarboxylation steps in the TCA cycle by hydrolyzing isocitrate, a six carbon molecule, to succinate, a four carbon molecule, and glyoxylate, a two carbon molecule (Canovas and Andrianopoulos 2006; Thirach et al. 2008). Penicillium marneffei, as well as other intracellular pathogens, is suspected to use glyoxylate or other two carbon molecules, perhaps products of fatty-acid degradation, in the glyoxylate cycle in gluconeogenesis. Similarly, the P. marneffei acuD gene has also been shown to be required for growth on gluconeogenic carbon sources, including acetate and fatty-acids, and is independently regulated via dimorphic switching, partially controlled by the transcriptional activator AbaA. Hence, the expression of acuD 15 increases, along with microbial virulence, during the infectious stage of growth in P. marneffei. While many more possible virulence factors are likely to exist, these genes have been experimentally shown to be up- or down-regulated depending upon the dimorphic switch, which is itself a characteristic that increases the virulence of this already pathogenic organism. Further experimentation focusing on P. marneffei’s genome is imperative to understanding the mechanisms and pathways involved in pathogenesis. The Septins In addition to the cell wall synthesizing enzymes and other constitutive, cell wall molecules is the family of highly, evolutionarily conserved family of proteins known as the septins. Septins are found in a wide range of organisms, ranging from yeast, to mice, to humans (Pan et al. 2007). The roles of septins, which include the formation of septa, polarized growth, initiating sites of new growth, and the processes of cellular division, are similar in most eukaryotes, excluding plants (Momany et al. 2001; Longtine et al. 1996). Septins were first characterized in the budding yeast S. cerevisiae in association with the mother- bud neck created during cellular division, and were later recognized as products of several cell-division-cycle genes (CDC3, CDC10, CDC11, and CDC12) in S. cerevisiae (Byers and Goetsch 1976; Haarer and Pringle 1987; Kim et al. 1991; Ford and Pringle 1991). Near the cytoplasmic surface of the mother- bud neck lies a ring comprised of filaments, each approximately 10 nm long and characteristically observed in the mother- bud neck. It was in this ring that the septins were originally documented. Interestingly, experiments on temperature sensitive S. cerevisiae mutants grown at the 16 non-permissive temperature resulted in the suppression of that particular septin gene, revealing that if any one of the four septins that form the mother- bud neck complex were not to be expressed, not only would cells not divide properly, but the septin ring also fails to form, resulting in chains of elongated cells with multiple bud sites as well as multiple nuclei (Byers and Goetsch 1976). Such evidence suggests that septins play a larger role than merely transiently interacting with the eukaryotic cytoskeleton. As may be expected with genes involved in cellular division, S. cerevisiae CDC genes have since been revealed to have homologues in many other species of fungi and animals, including the filamentous fungus Aspergillus nidulans (Momany et al. 2001; Lindsey et al. 2010). The A. nidulans CDC3, CDC10, CDC11, CDC12 homologues are aspB, aspD, aspA, and aspC, respectively. Aspergillus nidulans also contains a fifth septin found only in filamentous fungi, aspE, and its septins share 35% homology in their amino acid sequences, most of which is contained in the central regions, a property of all septins (Momany et al. 2001; Juvvadi et al. 2011). Structurally, these proteins contain three predicted P-loop domains with GTPase activity, which suggests septin complex formation and function may be dependent upon nucleotide binding and subsequent hydrolysis (Longtine et al. 1996; Saraste et al. 1990). Another secondary structure, the coiled- coil domain, is also present in most septins, excluding aspD, near their C- terminus and is thought to play a role in septin- septin or other septin- protein interactions (Lupas et al. 1991). Expression experiments (Momany et al. 2001) have shown that in synchronized, conidiating cultures aspB was expressed the greatest, followed by aspC, aspA, aspE, and finally aspD. 17 While a difference in expression of individual septins is evident, it must be emphasized that the septins do actually interact with one or all of the other septins, including co-localization and supplemental expression in instances when deletion strains are created. Experimentation using green fluorescent protein (GFP)- tagged AspA and AspC, as well as aspA, aspC, and aspAaspC deletion strains, has shown both proteins play similar roles in cellular growth and development in A. nidulans, where they form the characteristic ring-like structure through which newly formed cellular bodies grow, and, in the cases of aspA, aspC, and aspAaspC deletion strains, either provide a partial function to support their partner’s absence or function improperly due to some level of dependence on their septin partner (Lindsey and Momany 2006; Lindsey et al. 2010). GFP- tagged AspA and AspC were observed localizing to sites of cell wall manipulation, including sites of germ tube emergence, newly branching hyphae, septation, at the tips of mitotically active hyphae, and even during development of new conidiophore cell layers. However, while AspA and AspC are both certainly involved in septation, they are not necessarily required for septation to be completed or even present. For instance, aspA deletion strains exhibited septation levels, albeit delayed, similar to the wild-type strain (Lindsey et al. 2010). In contrast, the aspC deletion strain exhibited only about a third of the level of wild-type septation, indicating that, while AspA and AspC do co-localize in many instances, AspC plays a more specialized role in septation than does AspA. Furthermore, deletion of either or both aspA or aspC results in both the disruption of spatial patterning of germ tube emergence and an increase in the total number of germ tubes per germinating conidium and hinders proper conidiation. Subsequently, in an effort to uncover any degree of co-localization between AspA and AspC, A. nidulans 18 aspA deletion strains with a GFP- tagged AspC and aspC deletion strains with a GFP- tagged AspA were created. Photographs taken via fluorescence microscopy revealed that AspC-GFP failed to localize in aspA deletion strains. Similarly, AspA-GFP localized abnormally in aspC deletion strains. Additional studies inserting aspC into S. cerevisiae CDC12 deletion strains have shown that the Cdc12 homologue can actually alter morphology in budding yeast to resemble that of filamentous fungi (Lindsey et al. 2010). Such S. cerevisiae mutant strains resembled hyphal growth, with characteristically large, fuzzy colonies with a high percentage of elongated, filament-like cells, as opposed to the smaller, smoother colonies with ovoid or round shape observed in yeast growth. Further analysis revealed that the native AspA homologue, Cdc11, co-localized with the aspC product at the bases of the filament-like cells, termed atypical pseudohyphae. Like aspA and aspC, the CDC3 homologue, aspB, is also involved in the formation of septa, branches, and asexual reproductive structures, with subtle variations (Westfall and Momany 2001; Momany and Hamer 1997). AspB has been documented localizing to forming septa in a step-wise and regulated manner (Westfall and Momany 2001). In A. nidulans, a single ring of filamentous actin (F-actin) is first formed, a step proven to be crucial, as mycelia treated with F- actin depolymerizing agents were shown to lack the typical localization pattern of AspB to septation sites. Once the F- actin ring is formed, AspB co-localizes in a ring- like structure to the septation site. After the F- actin and AspB rings have co-localized, deposition of cell wall material begins, invagination of the cell wall is initiated, and the forming septum can be visualized. In budding yeast or animal cells this invagination would continue until two daughter cells were produced. However, in filamentous fungi, contraction of the actin ring does not 19 lead to complete cytokinesis. Cytokinesis in filamentous fungi results in septation of hyphae and compartmentalization of cellular organelles, including cytoplasm and the mitotically produced nuclei. Unlike the actin ring, though, AspB does not contract. Instead, the AspB ring separates into two distinct rings, each bordering one side of the septum. As invagination continues, the actin ring contracts until it is eventually no longer associated with the septum. Nearing the completion of septation, the basal ring of AspB disappears, and only the apical AspB ring is left, until it too is lost. Once septation has been completed, neither actin nor AspB are associated with the septum. Related analysis of AspB and chitin rings in hyphal tip cells also showed a pattern of maturation within the process of septation itself, with the most mature septa near the apical end of the cell and the least mature septa near the basal end. In addition to septation, AspB may also play roles in branching and conidiophore development; serves as a molecular marker for branching, as it localizes premitotically to new branch points; is present at sites of secondary germ tube emergence, but not of the primary germ tube, from germinating conidia; and localizes at the aerial hypha/metulae, metulae/phialide, and the phialide/conidium interfaces, but disappears as the descendant cell types mature. The fourth, and final, known S. cerevisiae homologue in many filamentous fungi is aspD. While documentation of the AspD septin is scant, some brief experimental characterization has been conducted in the human pathogenic fungus, A. fumigatus (Juvvadi et al. 2011). Experimentation using GFP- tagged AspD has revealed that AspD localizes at many sites as ring-like structures or associated with discontinuous tubules in the cytoplasm of the hyphal body. GFP- tagged AspD also was present at both forming 20 and mature septa, as a layer in the vesicle of the aerial hypha at the point of phialide emergence, and in phialides at the point of conidiogenesis. AspE, a septin found exclusively in filamentous fungi, has been briefly characterized in A. fumigatus (Juvvadi et al. 2011). GFP- tagged AspE, rather than exhibiting localization patterns seen in septins AspA, -B, or -C, appeared to localize to cytoplasmic networks of long, interconnecting filaments, which resemble microtubule networks. Another contrasting observation of GFP- tagged AspE was that characterized hyphae lacked the characteristic septin ring. However, AspE was observed at newly formed septa. Transformational Mutagenesis While many elements of P. marneffei’s biology have remained elusive, some studies have been successful in creating ways of revealing gene function in P. marneffei. Agrobacterium tumefaciens, the Gram-negative bacterium responsible for producing crown gall tumors in plants, has been put to use in transformation studies (Michielse et al. 2005). Agrobacterium tumefaciens contains a DNA sequence coding for tumor-inducing (Ti) factors. Part of the sequence is transferred to a host plant cell in a single stranded form as a plasmid, where it is integrated into the plant’s genome. The transferred DNA sequence is termed T-DNA. T-DNA sequences contain genes encoding plant growth regulating enzymes and a virulence region, which not only houses the vir genes needed for tumorigenicity, but may also encode genes responsible for the construction, transport, and integration of T-DNA into the host genome. In addition, the Ti plasmid is surrounded by repeated sequences acting as signals for delivery of DNA to the host cell. 21 The presence of A. tumefaciens alone, however, is not enough for DNA transfer to occur. vir genes must be induced prior to co-cultivation of A. tumefaciens and host strains. In order to induce the vir genes phenolic compounds, such as acetosyringone, have been used. Although A. tumefaciens has been thoroughly used in transformations of other fungi, Agrobacterium tumefaciens – mediated transformation (AMT) has recently been applied to P. marneffei in several studies, yielding morphological mutants (Zhang et al. 2008). Zhang et al., (2008) transformed P. marneffei to create at least seven morphological mutants. Sequence analysis of those mutants revealed mutations in genes whose products included transporter proteins, transcription regulators, synthases, kinases, and revealed an unknown mutation that caused an excess of mucus to be produced at the surface of the colony. A further improved AMT system (Kummasook et al. 2010) created many mutants, a number of which remain uncharacterized. One of these mutants is the subject of the present study. Objective Each of the five asp genes in Aspergillus species have been shown to be differentially expressed and their respective products localized to varying locations in the filamentous fungal cell. Based on previous experiments with the related ascomycetous fungi A. nidulans and A. fumigatus, a degree of functional overlap has been revealed between septins, with some physiological processes still occurring in strains with one or another deleted septin. The current body of reported data suggests that in the AMT- generated aspC mutant of P. marneffei, not only will the level of expression of aspC decrease relative to the wild-type, but the remaining four septins will be up-regulated in a compensatory effort to provide proper homeostatic control for the cell. 22 The first of two major goals in this proposal is to phenotypically characterize the aspC mutant of P. marneffei, beginning with its most superficial macroscopic qualities, including color, degree of conidiation, and degree of colony growth, followed by a more detailed, microscopic examination, such as conidiophore structure and any potentially observable phenotypes due to the disruption mutation of the aspC gene. Additionally, because septins are extensively involved in processes requiring cell wall manipulation, both the wild-type and mutant strains will be cultured in an environment that stresses the cell wall using two polysaccharide-binding dyes, Congo red (CR) and Calcofluor white (CW). Chitin biosynthesis, structure, and morphology in the Basidiomycete Mucor rouxii was strongly inhibited or reduced by varying concentrations of the polysaccharide- binding dye Congo red (CR) in vitro (Bartnicki- Garcia et al. 1994). Higher concentrations of CR stopped chitin formation entirely or formed a slimy, brightly colored dye-chitin complex. Subjecting M. rouxii to lower concentrations of CR resulted in a red precipitate and nearly colorless solution, indicating chitin had completely bound the CR. Exposure to CR also produced much thinner chitin fibrils relative to chitin fibrils formed in the absence of CR and a gel-like substance, rather than the crystalline structure of chitin assembled in the absence of CR. Similarly, CR has also been used as a microscopic fluorescence growth indicator of hyphal growth in Aspergillus niger (Matsuoka et al. 1994). Subjecting hyphae to increasing concentrations of CR yielded morphological anomalies including swelling and curling of hyphae. Fluorescence measurements also indicated that as long as hyphae were growing (i.e., cell wall surface area and the number of chitin molecules increased, thus increasing the number of CR binding sites) the fluorescence of CR-bound hyphae increased linearly with time until 23 miconazole, a hyphal growth inhibitor, was added. Examined sections of hyphal cell walls displayed greater increased fluorescence levels behind the apex than in the apex cell walls, suggesting an uncertain binding affinity of CR for chitin and providing some evidence as to the type and location of chitin bound by CR. However, the CR-chitin binding event occurs in a stepwise manner (Herth 1980). Because the crystallization process is disturbed, it appears that CR has a greater binding affinity for polymerized chitin than for the larger, and more elaborate, crystalline structure of side-linked chitin microfibrils. Thus, CR-chitin reactions usually follow this binding pattern. Calcofluor white (CW), another polysaccharide- binding dye, has also been shown to have similar effects on both filamentous fungi and budding yeasts (Roncero and Duran 1985). Subjecting Geotrichum lactis to CW resulted in cellular lysis at mycelial tips, while Saccharomyces cerevisiae exposed to CW resulted in incomplete cytokinesis. Subjecting both the wild-type and mutant strains to CR and CW should lead to an increase in septin expression. The second goal of this proposal is to document the differential expression, if any, of each asp gene between the mold and yeast phases of both the wild-type and mutant strains cultured under normal conditions (e.g., cultured in Sabouraud dextrose agar [BD Biosciences] ((SAB); 20g/L dextrose and 10g/L peptone) or minimal medium ((MM); 20mL 50x salt solution; 1mL trace elements (40mg Na 2 B 4 O 7 ·7H2O; 400mg CuSO 4 ·5H2O; 800mg FeSO 4 ·7H 2 O; 800mg NaMoO·2H 2 O; 8g ZnSO 4 ·7H 2 O); 10g glucose; and 66.3mL of 0.3M L-glutamine; at a pH of 6.5) without any supplements) and under conditions which stress the fungal cell wall. Again, this may potentially reveal the roles some genes play in growth and cellular development, as well as phase transition. 24 The identification of the aspC gene’s exact function brings with it more crucial evidence needed to view the larger scope of how P. marneffei interacts with its environment, particularly in its role as an often aggressive, opportunistic pathogen of immunocompromised individuals. 25 Chapter II Materials and Methods Strains, Maintenance, and Growth Conditions The P. marneffei wild-type strain (strain F4; CBS 119456) was procured from the hemoculture of an AIDS patient from the central laboratory, Maharaj Nakorn Chiang Mai Hospital in January of 1999 (Kummasook 2010). The P. marneffei strains were cultured on potato dextrose agar (PDA [Difco brand] Becton- Dickinson, Franklin Lakes, NJ) for seven to ten days at 25˚C. Conidial suspensions for further experimentation were prepared by scraping the surface of the fungal culture with a cotton swab pre-moistened with sterilized water and suspending the mycelial tissue in sterilized water. The tissue suspension was filtered through sterilized glass wool (Corning, Acton, MA) in a screen cap (BioRad Laboratories, Inc., Hercules, CA) attached to a 50 mL conical tube to yield a suspension of isolated conidia via centrifugation (Gifford and Cooper 2009). The concentration of the resultant conidial suspension was measured by counting conidia with a haemocytometer (http://www.ruf.rice.edu/~biolabs/methods/microscopy/cellcounting.html) under a phase contrast microscope. After the concentration of the conidial suspension was obtained, it was used to make dilutions as necessary. Penicillium marneffei stock cultures were suspended in 15- 30% glycerol and stored at -80˚C. Penicillium marneffei mutant strain I209 was derived using a modified AMT system (Kummasook 2010; Kummasook et al. 2010). Strain I209 was later revealed to be an aspC disruption mutant (Eric Price, personal communication). 26 Morphological Analysis and Characterization Morphological characterization of P. marneffei and mutant strain I209 was conducted on synchronous cultures. Both wild-type and mutant strains were inoculated onto PDA plates by collecting mycelial tissue from strains cultured on PDA and subsequently streaking that tissue with a sterilized, metal inoculating loop, then incubated for seven to ten days at 25˚C and 37˚C in order to obtain isolate colonies for the purposes of single colony observation and photographic documentation. To observe asexual developmental structures, slide cultures (Harris 1986) were prepared and photographed using differential interference contrast (DIC) microscopy. Cell Wall Disruption Assays For cell wall disruption experiments in liquid cultures, both the wild-type and mutant strain were subjected to each condition (e.g., incubation times and temps; chemical supplements; and control cultures with no supplements). Fifty milliliters each of SAB and MM were inoculated with conidia from both fungal strains to a final concentration of at least 1.0 x 10 6 cells/mL and incubated at 25˚C and 37˚C for two days. Those cultures to be supplemented with H 2 O 2 (Fisher) were treated with H 2 O 2 four hours before ending the incubation period (44 h) to a final concentration of 2.0 mM. Congo red (CR) (Amresco) and Calcofluor white (CW) (Sigma- Aldrich) cell wall disruption experiments were performed under the same parameters as those implemented in the H 2 O 2 experiments with the exception of the final concentrations of the respective supplement. CR was added to cultures to a final concentration of 20 μg/mL. CW was added to cultures to a final concentration of 80 μM. 27 Cell wall disruption experiments performed on PDA plates were conducted on both strains, and the following supplements and concentrations were used: Congo red (0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, 1, 2, 3, 4, 5, 6, 7, 8, 9, and 10 μg/mL); Calcofluor white (10, 20, 30, 40, 50, 60, and 70 μM); and hydrogen peroxide (0.75, 0.5, and 0.25 mM). Conidial dilutions (10 μL of 1 x 10 4 cells/mL) were pipetted onto the agar plate. Genomic DNA Extraction Genomic DNA was extracted according to the protocol from the ZR Fungal/Bacterial DNA Miniprep kit (Zymo Research Corp., www.zymoresearch.com). Extracted DNA was further concentrated via ethanol precipitation (Gallagher 2010). The resulting DNA samples were subjected to spectrophotometry to measure their concentrations as double-stranded nucleotides. A dilution factor of 1:10 was used, and corresponding data was recorded. Gel Electrophoresis In order to assess the quality of genomic DNA, samples were subjected to agarose gel electrophoresis. To make a 1.0% agarose gel, 1.0 g agarose I (Amresco) was dissolved in a flask containing 100 mL 1 x TAE buffer (50 x TAE buffer: 242 g Tris base; 57.1 mL glacial acetic acid; 100 mL 0.5 M EDTA); to prepare 1 x TAE buffer, 20 mL of 50 x TAE buffer were added to 1 L of ddH 2 O) by boiling in a microwave. While waiting for the agarose solution to cool enough to handle the flask with bare hands, the electrophoresis rigging was assembled by placing the gel tray in the electrophoresis chamber and inserting the wedges into their respective slots to ensure the agarose solution did not leak out of the gel chamber. Once the flask had cooled enough to touch, 28 the agarose solution was poured in to the gel chamber, the comb was inserted into the gel to create wells for loading samples, and the gel was allowed to solidify. Once solidified, the wedges were gently removed, enough TAE buffer to cover the gel by three or four millimeters was poured into the chamber, and the comb was gently removed, as to not tear or otherwise damage the solidified gel. Five microliters of each DNA sample were then individually mixed with one microliter of E-Z Vision loading dye (BioRad) on parafilm via pipetting. After the DNA and dye were mixed, the sample was loaded into its respective well. In order to ascertain the relative size of each DNA sample, either a 100 base pair (bp) or 1kbp molecular ladder was loaded into a separate well. The electrodes were then oriented appropriately so that the power supply was connected to the electrophoresis chamber. The power supply was turned on, and the samples were subjected to 120 V for about one hour or as long as it took for the tracking dye to reach about three-quarters of the length of the gel, allowing adequate separation of ladder bands. The power supply was turned off, electrodes detached, and the gel tray was gently removed from the chamber, briefly allowing any excess buffer to drain. Because E-Z Vision, which is visible under UV light, was loaded with each sample, the use of the known human carcinogen ethidium bromide was avoided prior to imaging the DNA- containing gel. Total RNA Extraction and Purification Wild-type and mutant strains were cultured for two days in SAB broth and liquid MM under previously described conditions. RNA extraction was performed using the RNeasy Mini Kit (Qiagen, Valencia, CA). The RNA extraction protocol for plants and fungi was followed with the following exceptions: Steps 2 and 3 in the manufacturer’s 29 protocol were not used. At least two RNA samples were purified from each culture, and two cDNA samples were later synthesized from each RNA sample, yielding a total of at least four cDNA samples. DNase treatment was performed using DNase I (Invitrogen, Carlsbad, CA) according to the manufacturer’s instructions. The concentration of each RNA sample was measured via spectrophotometry and the corresponding data was recorded. The procedure for measuring RNA sample concentrations was the same as for genomic DNA, with the exception of the dilution factor, which was 1:100 for RNA samples. cDNA Synthesis, and Qualtiy Control Experiments Purified total RNA was used as a template for cDNA synthesis using the protocol for oligo(dT) 20 primers in the iScript Select cDNA Synthesis Kit (BioRad). An additional cDNA synthesis was performed in parallel using the protocol for cDNA synthesis with anchored oligo(dT) 18 primers in the Transcriptor First Strand cDNA Synthesis Kit (Roche, Switzerland) in order to compare the efficacy in cDNA synthesis of each kit. In both cases, cDNA synthesis was performed according to the manufacturer’s instructions. Concentrations of cDNA samples were then measured via spectrophotometry. For the measurements of cDNA sample concentrations, the procedure was the same as genomic DNA with two exceptions: a 1:100 dilution factor, rather than a 1:10 dilution factor, was used in the measurement of cDNA sample concentrations; and cDNA samples were measured as double-stranded nucleotides. After cDNA sample concentrations were determined, cDNA dilutions were prepared to a final concentration of 50 ng/μL. 30 As a control experiment to determine whether or not cDNA samples were contaminated, cDNA samples (50 ng/μL), as well as genomic DNA (50 ng/μL), were subjected to a polymerase chain reaction (PCR) employing primers to amplify a portion of the calmodulin gene. To amplify the calmodulin genes, cDNA and DNA samples were subjected the following reaction: 2 μL cDNA or DNA template (50 ng/μL); 4 μL nuclease- free water; 2 μL calmodulin forward primer (10 μM); 2 μL calmodulin reverse primer (10 μM); 10 μL Go- Taq Green Master Mix, 2x (Promega, Madison, WI, USA) in a final volume of 20 μL. PCR amplification conditions were 5 min for denaturation at 95 ° C, followed by 35 cycles of 30 sec at 95 ° C, 30 sec at an annealing temperature of 62 ° C, and 1 min at 72 ° C, with a final extension step of 7 min at 72 ° C. Because the genomic DNA calmodulin product is larger than its cDNA counterpart, the DNA PCR product would migrate through the gel more slowly than the cDNA PCR product. Any evident DNA contamination in cDNA samples ran in the 2% agarose gel resulted in the exclusion of those DNA-contaminated cDNA samples. Two- Step quantitative (q) PCR Once cDNA samples had been verified as non-contaminated via calmodulin control PCR, multiple cDNA samples from cultures grown under the same conditions (e.g., all wild-type mold cDNA samples from SAB cultures treated with Congo red; all mutant yeast cDNA samples from MM cultures treated with Calcofluor white; etc.) were combined with one another to yield a master stock of cDNA from both phases of each fungal strain under the various conditions to which they were subjected. A 1:100 dilution of each sample was prepared for use in the qPCR. 31 All samples were subjected to a qPCR of each individual septin gene, along with calmodulin gene as the control, using the same parameters. Each qPCR reaction included 2.5 μL cDNA template and 18 μL master mix, which contained 3.5 μL nuclease-free water, 2 μL each of forward and reverse primers at 10 μM each, and 10 μL iQ SYBR Green Supermix (BioRad). qPCR amplification conditions were 3 min for denaturation at 95 ° C, followed by 40 cycles of 15 sec at 95 ° C, 30 sec at an annealing temperature of 60 ° C, and 91 cycles of 1 min for extension at 50 ° C. Table 1- Potentially differentially expressed genes and their corresponding primers Gene Subjected to qPCR Primers’ Sequences calmodulin F 5’- TCA CAA CCT CCC TTC GAT TC -3’ R 5’- TCT CTT CCT CGG AGT CGG TA -3’ aspB F 5’- TCC GCA AAG GAT TCA ACT TC -3’ R 5’- GGT CTC CAT GAG TCG TCG TT -3’ aspC F 5’- GAC AAC CCC AAG TTC AAG GA -3’ R 5’- ACG AAA ATG TAC ACC GCA CA -3’ aspD F 5’- CGG TGA CCA AGT CAA CAA TG -3’ R 5’- GTC GCT CTT CCA AGG TCA AG -3’ aspE F 5’- GCA AGG ACC CTA CCC CTT AC -3’ R 5’- CCT CGA ATT TGC TCT CCA AG -3’ 32 Chapter III Results Macroscopic Observations Observational analysis of isolated colonies and slide cultures revealed stark macroscopic contrasts between P. marneffei wild-type and the aspC mutant strains. Isolated colonies obtained from strains cultured on PDA (Figure 1) showed differences in the gross amount of conidiation, patterns of conidiation, and pigmentation of conidia. Wild-type P. marneffei cultures exhibited thicker mycelial tissue growth and conidia appeared to germinate faster than the I209 strain. Conidiation also revealed a difference in pigmentation between wild-type and mutant strains, with wild-type conidia presenting a drab, olive- green pigment and the I209 strain displaying a vibrant, yellow pigment. Incidentally, conidiation patterns were also noticeably different from one another. While isolated wild-type colonies conidiated in a seemingly uniform pattern, isolated colonies of the I209 strain appeared to conidiate distally at first and progressively conidiate in a proximal, concentric pattern. In stark contrast to aspC mutant cultures incubated at 25 ° C, no growth was observed on PDA plates incubated at 37 ° C. However, wild-type P. marneffei growth appeared normal (no photographs were taken as evidence). In addition to unsupplemented PDA plates, experiments using PDA plates supplemented with H 2 O 2 , CR, and CW exhibited no growth of the aspC mutant. Thus, further experimentation on PDA and alternative solid media at 37 ° C were abandoned. 33 Figure 1. A), Wild-type P. marneffei cultured on PDA and incubated at 25˚C for seven days; B), Penicillium marneffei mutant I209 cultured on PDA and incubated at 25˚C for seven days; A and B), in the corner of each photo is a close- up of an isolated colony of the respective strain. 34 Sensitivity Plate Assays Cell wall disruption assays were monitored from the third day of incubation to the seventh, and the radial growth was measured in millimeters. All plate assays included a negative control, that being PDA with no supplement. Initially, a lag in growth was absent at 10, 9, 8, 7, 6, and 5 μg/mL of CR and also at 0.75 mM H 2 O 2 (Figures 2-6). This was confirmed by the increase in radial growth recorded as the concentrations of each supplement decreased, relative to the PDA- only negative control. However, the CW plate assay did not provide the expected results (Figures 7 and 8). Rather than the radial growth of each colony being slowed at higher CW concentrations and gradually increasing as the CW concentration decreased, as was seen in the H 2 O 2 and CR plate assays, radial growth of both the wild-type and aspC mutant grew steadily, without any obvious or evident CW- related hindrance relative to the negative control. Figure 2. Radial growth of the wild-type strain over a 5 day period during CR disruption plate assay. 35 Figure 3. Radial growth of the aspC mutant over a 5 day period during CR plate disruption assay. Figure 4. Radial growth of wild-type over a 5 day period during H 2 O 2 disruption plate assay. 36 Figure 5. Radial growth of aspC mutant over a 5 day period during H 2 O 2 disruption plate assay. 37 Figure 6. Cell wall disruption plate assay using CR supplemented PDA. Wild-type (left) and I209 (right) after incubating at 25 ° C for 4 days. The left column begins at the top with the negative control (PDA only) plate continuing with 10, 9, 8, 7, 6, 5, 4, 3, 2 μg/mL. The right column begins at the top with 1 μg/mL continuing with 0.9, 0.8, 0.7, 0.6, 0.5, 0.4, 0.3, 0.2, 0.1 μg/mL. 38 Figure 7. Radial growth of wild-type over a 5 day period during CW disruption plate assay. Figure 8. Radial growth of aspC mutant over a 5 day period during CW disruption plate assay. 39 Microscopic Observations The most revealing morphological differences were observed via imaging mounted slide cultures of wild-type and aspC strains (Figures 9 and 10, respectively). Wild-type P. marneffei exhibited growth and development already well- characterized in primary and secondary literature. Hyphal cells developed into foot cells, which continued to grow apically until the asexual developmental program was triggered, consequently leading to conidiation (Canovas and Andrianopoulos 2007). Extensive conidiation was routinely observed in the wild-type, both as loose, single and long chains and those still attached to the conidiophore itself. Several forms of conidiophores were observed in the wild-type cultured on PDA, with some conidiophores producing several phialides, each with a chain of conidia, while others developed only one or two phialides from which a chain of conidia budded. Few of these characteristics were observed in the aspC mutant. The aspC mutant displayed several traits divergent from those of the wild-type and unique to its mutation. Primarily, conidiophores were markedly different. aspC mutant conidiophores did not characteristically exhibit the developmentally organized conidiophores seen in the wild- type. Rather, aspC mutant asexual development appeared to be disorganized, with a spatial arrangement highly varied from the wild-type. Nor was there an abundance of conidia, as free, single conidia or chains, attached or otherwise observed in aspC mutant slide cultures. In contrast to the wild-type, mutant conidiophores produce phialides that eventually come to bear a single conidium. All observed aspC mutant conidiophores displayed this pattern of conidiation. Observational data of the aspC disruption mutant 40 are in parallel with those data of an aspC deletion strain previously characterized (Lindsey 2010). Slide culture microanalysis of wild-type cultures revealed very little difference between the negative control (PDA only) culture and those supplemented with H 2 O 2 or CW (Figures 11 and 12, respectively). However, CR wild-type cultures did show some signs of decreased asexual development (Figure 13). Wild-type cultures subjected to 25 μg/mL of CR exhibited a high degree of aberrant conidiophore types compared to the negative control and also to the lower end of CR concentrations (15 μg/mL). Wild-type cultures subjected to 20 μg/mL of CR also showed some inhibition of conidiation and aberrant conidiophore types. Similarly, the aspC mutant’s asexual development did not seem to be affected by either CR or CW compared to the mutant control culture (Figures 14 and 15, respectively). Although, the aspC mutant did appear to be affected at the upper range of the H 2 O 2 assay (3 mM H 2 O 2 ; Figure 16). At 3 mM H 2 O 2 , mutant conidiophores generally displayed structures finer than those observed at lower concentrations. 41 Figure 9. (A- F), 100x magnification and scale bar represents 0.1 mm; A and B), P. marneffei wild-type asexual developmental structures with conidiopores and dissociated conidia; C), new branching from hyphae with visible septa; D- F), aberrant conidiation observed in P. marneffei wild-type strain. A B C D E F 42 Figure 10. A- B), 100x magnification and scale bar represents 0.1 mm; A- C), I209 conidiophores; D), hyphae of I209; E), close- up of hypha from boxed area in Figure 8. D showing unidentified spots, possibly an effect of the dye- cell wall binding. A B C D E 43 Figure 11. A- F), 100x magnification and scale bar represents 0.1 mm; P. marneffei wild-type H 2 O 2 slide culture assay; A) PDA only; B) 1 mM H 2 O 2 / PDA; C) 1.5 mM H 2 O 2 / PDA; D) 2 mM H 2 O 2 / PDA; E) 2.5 mM H 2 O 2 / PDA; F) 3 mM H 2 O 2 / PDA. A B C D E F 44 Figure 12. A- H) 100x magnification and scale bar represents 0.1 mm; P. marneffei wild-type CW slide culture assay; A) PDA only; B) 10 μM CW/ PDA; C) 20 μM CW/ PDA; D) 30 μM CW/ PDA; E) 40 μM CW/ PDA; F) 50 μM CW/ PDA; G) 60 μM CW/ PDA; H) 70 μM CW/ PDA. A B C D E F G H 45 Figure 13. A- D) 100x magnification and scale bar represents 0.1 mm; P. marneffei wild-type CR slide culture assay; A) PDA only; B) 15 μg/mL CR/ PDA; C) 20 μg/mL CR/ PDA; D) 25 μg/mL CR/ PDA. A B C D 46 Figure 14. A- D) 100x magnification and scale bar represents 0.1 mm; P. marneffei aspC mutant CR slide culture assay; A) PDA only; B) 15 μg/mL CR/ PDA; C) 20 μg/mL CR/ PDA; D) 25 μg/mL CR/ PDA. A B C D 47 Figure 15. A- H) 100x magnification and scale bar represents 0.1 mm; P. marneffei aspC mutant CW slide culture assay; A) PDA only; B) 10 μM CW/ PDA; C) 20 μM CW/ PDA; D) 30 μM CW/ PDA; E) 40 μM CW/ PDA; F) 50 μM CW/ PDA; G) 60 μM CW/ PDA; H) 70 μM CW/ PDA. A B C D E F G H 48 Figure 16. A- D) 100x magnification and scale bar represents 0.1 mm; P. marneffei aspC mutant H 2 O 2 slide culture assay ; A) PDA only; B) 1 mM H 2 O 2 / PDA; C) 1.5 mM H 2 O 2 / PDA; D) 2 mM H 2 O 2 / PDA; E) 2.5 mM H 2 O 2 / PDA; F) 3 mM H 2 O 2 / PDA. A B C D E F 49 Gene Expression Analyses In order to quantitate expression levels of the asp genes, cDNA samples were subjected to qPCR. Expression levels were also to be considered significant only if expression was two- fold or higher than the expression levels observed in the mold phase of the wild-type strain cultured in SAB. If expression levels were less than two-fold greater, then those data were considered to be a trend but not significant. After several qPCR experiments focusing on the expression of aspA in the wild- type and mutant strains, the corresponding results of each experiment revealed DNA contamination. Due to repetitive contamination and time constraints, expression of aspA was not characterized. However, septin gene aspC was universally upregulated in the aspC mutant compared to the wild-type under every condition (Figures 19-26). Under the described culture conditions of P. marneffei wild-type, aspE was frequently expressed at the highest levels, followed by similar expression levels between aspD and aspB, with aspC typically expressed the least (Figures 17, 19, 21, 23, 25). Similarly, the aspC mutant also generally expressed aspE the highest (Figures 17, 19, 21, 23, 25). Unlike the wild-type strain, however, the aspC mutant frequently expressed aspD in greater levels than aspB. As observed in wild-type expression analysis, aspC was usually expressed the least. Unexpectedly, higher expression levels of aspC were recorded in the aspC disruption mutant than were recorded in the wild-type strain in SAB- and MM- only 50 liquid cultures (Figure 19-20). The most extreme difference in aspC expression levels between mutant and wild-type was seen in SAB cultures incubated at 37 ° C, with about a 100-fold increase in expression in the mutant relative to the wild-type. In SAB cultures at both 25 ° C and 37 ° C, an increase in aspD expression levels was observed in the aspC mutant, while much higher expression levels of aspD were present in the mutant in MM cultures at both 25 ° C and 37 ° C. The greatest-fold difference in asp gene family expression was between the wild-type and mutant MM cultures incubated at 25 ° C. Expression levels of aspC, aspD, aspE, and aspB in the mutant were about 10-fold, 70- fold, nine-fold, and four-fold, respectively, greater than those recorded in the wild-type. Treatment of cultures with H 2 O 2 lead to an increase in expression levels of each asp gene under every condition (Figure 21). aspD was expressed the greatest in SAB and MM cultures incubated at 37 ° C, followed by aspC in the SAB culture and aspE in the MM culture. Again, aspE continued to be highly expressed. Expression levels of the asp gene family in CR cultures were higher in the aspC mutant than in the wild-type under nearly every condition, with the exception of aspB in SAB and MM cultures incubated at 37 ° C (Figure 23). aspC was expressed the least under every condition except for in the aspC mutant cultured in SAB at 37 ° C, under which aspC was expressed about 50-fold greater than in its wild-type counterpart. Again, aspE was expressed the most. With the exception of aspE in SAB cultures incubated at 37 ° C, all asp genes were more highly expressed in the aspC mutant than in the wild-type in CW cultures (Figure 25). 51 aspB aspC aspD aspE Figure 17. Differential expression of aspB, aspC, aspD, and aspE in P. marneffei wild-type cultured in SAB and MM incubated for two days at 25 ° C and 37 ° C. 52 Figure 18. Differential expression of aspC in P. marneffei wild-type cultured in SAB and MM and incubated for two days at 25 ° C and 37 ° C. 53 aspB aspC aspD aspE Figure 19. Comparison of differential expression of aspB, aspC, aspD, and aspE in P. marneffei wild-type and aspC mutant cultured in SAB and MM incubated for two days at 25 ° C and 37 ° C. 54 Figure 20. Differential expression of aspC in P. marneffei wild-type and mutant strains cultured in SAB and MM incubated for two days at 25 ° C and 37 ° C. 55 aspB aspC aspD aspE Figure 21. Differential expression of asp genes (B, C, D, and E) in P. marneffei wild-type and aspC mutant cultured in SAB and MM supplemented with H 2 O 2 incubated for two days at 25 ° C and 37 ° C. 56 Figure 22. Differential expression of aspC in P. marneffei in wild-type and mutant strains cultured in SAB and MM supplemented with H 2 O 2 incubated for two days at 25 ° C and 37 ° C. 57 aspB aspC aspD aspE Figure 23. Comparison of differentially expressed asp genes (B, C, D, and E) between P. marneffei wild- type and aspC mutant cultured in SAB and MM supplemented with CR and incubated for two days at 25 ° C and 37 ° C. 58 Figure 24. Differential expression of aspC in P. marneffei wild-type and mutant strains cultured in SAB and MM supplemented with CR incubated for two days at 25 ° C and 37 ° C. 59 aspB aspC aspD aspE Figure 25. Differential expression of aspB, aspC, aspD, and aspE in P. marneffei wild-type and aspC mutant cultured in SAB and MM supplemented with CW incubated for two days at 25 ° C and 37 ° C. 60 Figure 26. Differential expression of aspC in P. marneffei in wild-type and mutant strains cultured in SAB and MM supplemented with CW incubated for two days at 25 ° C and 37 ° C. 61 Chapter IV Discussion Microscopic Analysis Microscopic analysis of wild-type and aspC mutant P. marneffei strains revealed profound differences in the gross anatomy of asexual structures. The lack of completely developed conidiophores and extensive chains of conidia observed in aspC cultures suggests an inability to complete the budding process that produces the chains of conidia found in wild-type cultures. Because the aspC mutant produces phialides that appear to stop budding after one conidium is produced, it may stand to reason that aspC is involved in conidiation and the remaining functional septins are compensating for the mutated gene to the extent of producing a single conidium but are not entirely adequate for proper conidiation. Alternatively, although, perhaps less likely, the presence of conidia in aspC mutant cultures may mean that aspC is not directly involved in conidiation in P. marneffei, but either plays a limited role in conidiation or is more directly involved in some process upstream of conidiation in the asexual developmental program, and the other septins are compensating for the mutated aspC or do not function properly without the effects the aspC product, respectively. Additionally, because conidia bud basipetally, new cell wall material needs to be produced at the conidium- phialide interface. If the P. marneffei aspC is involved in conidiation, then a mutation in aspC may lead to incomplete mitosis, which, in large part requires expansion and subsequent contraction of the cell wall and cell wall associated constituents, potentially yielding the abnormal conidiation pattern observed in the aspC mutant. 62 Sensitivity Assays Sensitivity assays revealed differences both in the general growth rate and the sensitivity levels of each strain to various cell wall disrupting agents. The growth rate of the aspC mutant was markedly slower than that of the wild-type in negative controls. After comparing H 2 O 2 affected wild-type and mutant strain colonies to those cultured on the negative control plates, it was obvious that growth of both wild-type and mutant strains was inhibited at the highest concentration of H 2 O 2 . However, H 2 O 2 presumably does not inhibit fungal growth via interference with the cell wall. H 2 O 2 has been shown to cause an increase in ROS inside fungal mitochondria, leading to the degradation of the mitochondrial membrane potential. In turn, mitochondrial proteins involved in respiration are dysfunctional and the cell consequently dies. Since both strains were noticeably affected by H 2 O 2 , aspC may not be involved in the protective response to H 2 O 2 . As previously described, CR binds chitin in such a way that chitin cannot properly form the larger crystalline networks observed in untreated hyphae. When considering the proposed mechanism and documented effects of CR-chitin binding, it should also be noted that septins serve in a variety of capacities, including the recruitment of proteins involved in chitin biosynthesis and localization (DeMarini et al. 1997; Kozubowski et al. 2003; Sanz et al. 2004). In S. cerevisiae, CDC10, the AspD homologue, has been shown to recruit Bni4, a scaffolding protein involved in cell wall biosynthesis, to the cell wall on the mother’s side of the bud site (DeMarini et al. 1997; Kozubowski et al. 2003). Then, possibly in a concerted mechanism, Bni4 interacts with the activating sub-unit of chitin 63 synthase III (CSIII), Chs4, which interacts with Chs3, the catalytic sub-unit of CSIII (Sanz et al. 2004). Additionally, analyses of Bni4 deletion strains have shown that Bni4 is essential in localizing Chs4 but not necessary for CSIII activity, although Bni4 is involved in anchoring Chs3 to the neck region during bud emergence. These events contribute to process of chitin synthesis and deposition as a whole. In the case of the aspC disruption mutant, a dysfunctional AspC may be present in the septin complex-cell wall interface. A dysfunctional AspC may, in turn, lead to a partial failure in recruiting Bni4 to the cell wall. Hence, chitin synthesis and deposition in the cell wall would be compromised, and the aspC mutant’s growth would be inhibited at higher concentrations of CR. Similar to CR, CW binds chitin and interferes with proper crystallization of chitin in the fungal cell wall. However, no difference in growth was detected between P. marneffei wild-type and mutant strains in the CW assay, which suggests that the range of concentrations used in the assay was too low and that aspC is possibly not involved in cellular response to CW. asp Gene Expression aspC Expression. The aspC mutant strain cultured in SAB at 37 ° C expressed aspC nearly 100-fold higher than the wild-type counterpart. This observation may have two explanations, possibly occurring in tandem. In the case of the difference in aspC expression between the mutant and wild-type strains cultured in SAB at 37 ° C, it is possible that the aspC mutant is recognizing that a dysfunctional AspC is being translated from a transcript of the mutated aspC and is continually signaling that the transcription 64 process for aspC needs to be repeated in order to replace the dysfunctional AspC with a functional version, causing a dramatic increase in mutant aspC transcripts. However, due to the disrupted aspC gene, only aberrant aspC transcripts are produced. This process may be inadvertently resulting in a feed-back loop. Because qPCR quantitates the amount of DNA being transcribed (i.e., transcripts) from cDNA in real time, the software associated with the thermal cycler may be recording the cells’ homeostatic response to correcting an error in its translation of the aspC transcript. This analysis of aspC also applies to the general pattern of expression of aspC in the aspC mutant. The explanation of differential aspC expression between the mutant strain cultured in SAB at 25 ° C and 37 ° C may be less convoluted. P. marneffei’s dimorphic nature causes increased amounts of cell wall reconstruction. Because septins are so extensively involved in cellular processes that affect cell wall shape and construction, the phase transition program may lead to the increase in aspC transcription. However, because the dimorphic switch was not inhibited, it is not likely that aspC is required for P. marneffei’s dimorphic switch. Similarly, in aspC mutant SAB cultures supplemented with H 2 O 2 incubated at 37 ° C, aspC levels were at least 37-fold greater. This discrepancy in aspC levels may again be explained by the presence of a dysfunctional AspC leading to repeated transcription of acpC. However, since these cultures were treated with a chemical that is known to effect the overall growth and development of cells (Qin et al. 2011), it may also stand to reason that aspC is involved in the reaction to protect the cell against the effects of H 2 O 2 exposure. 65 Expression levels of aspC were also about 50-fold higher in mutant SAB cultures supplemented with CR at 37 ° C. Again, the previously proposed hypothesis for aspC expression in the aspC mutant may explain the recorded aspC expression levels in the mutant strain. In addition, as previously described, CR causes direct cell wall stress. The combined effects of the dysfunctional AspC and CR could be a likely cause of the increase in aspC expression. aspE Expression. The aspC mutant expressed aspE at a higher level under nearly every tested condition. In A. fumigates, AspE is thought to play roles in both septation and intracellular transport. The up-regulation of aspE of more than four-fold in wild-type SAB cultures at 37 ° C can possibly be attributed to the beginning stages of phase transition, during which the cell wall and associated cytoskeleton undergoes tremendous changes, including polymerization of cell wall constituents, transport of enzymes to the cell wall, transport of cellular organelles to new hyphal branches, etc. If aspE were actually involved in intracellular transport, then it would be reasonable to suspect that aspE should be up-regulated in the wild-type at 37 ° C. In the aspC mutant, an increase in aspE expression levels may also be explained by the cell up-regulating aspE to compensate for a dysfunctional AspC, in addition to the possibly increased role of aspE during phase transition. After supplementation with H 2 O 2 , minor decreases were also recorded in the wild- type strain cultured in SAB at 25 ° C and 37 ° C, respectively, as well as in MM at 37 ° C. Conversely, there was about a two-fold increase of aspE observed in the mutant strain cultured in SAB at 37 ° C. Because H 2 O 2 is known to oxidatively stress cells via protein 66 denaturation (Qin et al. 2011), it is possible that some proteins involved in transcription, either at the nuclear or cytosolic signaling levels, were in some way inhibited. Differential expression analysis of aspE after CR exposure revealed an increase in expression of aspE in every condition. An approximate 12-fold increase in aspE expression in the mutant strain cultured in SAB at 37 ° C was observed. Additionally, more than a five-fold increase in aspE expression was recorded in the wild-type strain cultured in SAB at 37 ° C. These fluctuations in aspE suggest that the gene may play a role in the cell’s response to the effects of CR. The expression levels of aspE in both strains cultured with CW were inconclusive. Interestingly, though, a significant increase in aspE was evident in the wild-type strain cultured in SAB at 37 ° C. aspD Expression. aspD was expressed in a greater amount in the aspC mutant than the wild-type in every instance. In addition, up-regulation of aspD was recorded in the wild- type strain at 37 ° C in every instance and nearly every condition in the mutant strain. These results suggest that aspD plays a role in phase transition. After exposure to H 2 O 2 , expression of aspD was approximately 50-fold higher in the aspC mutant cultured in SAB at 37 ° C. While aspD appears to be involved in phase transition, it is inconclusive as to whether aspD is involved in the response to H 2 O 2 . Results also suggest that the expression of aspD is also inhibited by the presence of H 2 O 2 . With the exception of the aspC mutant cultured in MM at 25 ° C, there was essentially little difference in the expression of aspD in the wild-type and mutant strains 67 cultured with and without CR. These results suggest that aspD is not necessarily involved in the response to CR. Treatment with CW promptly lead to an approximate 110-fold increase of aspD in the aspC mutant cultured in SAB at 37C. aspD expression also increased in the mutant cultured in MM at 37 ° C, with nearly a 40-fold increase. These results potentially provide further evidence that aspD is not necessarily involved in the response to CW, but possibly phase transition. aspB Expression. While expression levels of aspB did fluctuate, the largest-fold difference was a seven-fold increase in the mutant strain cultured in MM at 37 ° C. Because aspB has been documented as being primarily involved in septation, the expression levels of aspB were lower than expected. In addition, while not necessarily significant in every case, aspB expression did increase in the yeast phases of both the wild-type and mutant strains, suggesting a role for aspB in dimorphism. While H 2 O 2 and CW did have a small but noticeable effect on aspB expression, CR did not appear to have an effect. These results suggest that aspB may not be involved in the responses to cell wall stress via polysaccharide binding dyes or the response to oxidative stress. In conclusion, these preliminary results, taken together, suggest that aspC plays little to no role in phase transition in P. marneffei. However, the increased levels of aspE suggest the AspE septin may be involved in phase transition. Further asp gene family analyses will be required to state the significance of each septin in P. marneffei. In 68 addition to differential gene expression analyses, experimentation focusing on possible colocalization of septins with cytoskeletal elements, such as microfilaments, microtubules, and thin filaments, should be conducted concomitantly to elucidate the relationship between septin function and morphology. 69 Chapter V References Adams, T. H., Boylan, M. T., & Timberlake, W. E. (1988). brlA is necessary and sufficient to direct conidiophores development in Aspergillus nidulans. Cell, 54: 353- 362. Adams, T. H., Wieser, J. K., Yu, J. H. (1998). Asexual sporulation in Aspergillus nidulans. Microbiol. Mol. Biol. Rev. 62: 35- 54. Andrianopoulos, A. (2002). Control of morphogenesis in the human fungal pathogen Penicillium marneffei. Int. J. Med. Mibcrobiol. 292: 331- 347. Andrianopoulos, A., and Timberlake, W. E. (1994). The Aspergillus nidulans abaA gene encodes a transcriptional activator that acts as a genetic switch to control development. Mol. Cell Biol.14: 2503- 2515. Applegarth, D. A., and Bozoian, G. (1967). Presence of chitin in the cell wall of a griseofulvin-producing species of Penicillium. J. Bacteriol. 94:1787-1788. Bacon, J. S. D., Davidson, E. D., Jones, D., and Taylor, I. F. (1960). The location of chitin in the yeast cell wall. Biochem. J. 101:33c-38c. Bailloud, R., M. Sumanak, B. Sar, C. Y. Sreng, T. Ellman, D. Laureillard, and D. Monchy. 2002. Premiers cas d’infection Penicillium marneffei identifis chez l’immunodprim au Cambodge. J. Mycol. Me´d. (Paris) 12:138–142. Bartnicki- Garcia, S., Persson, J., Chanzy, H. (1994). An electron microscope and electron diffraction study of the effect of Calcofluor and Congo red on the biosynthesis of chitin in vitro. Archives of Biochemistry and Biophysics. 310(1): 6- 15. 70 Bateman, A. C., G. R. Jones, S. O’Connell, F. J. Clark, and M. Plummeridge. (2002). Massive hepatosplenomegaly caused by Penicillium marneffei associated with human immunodeficiency virus infection in a Thai patient. J. Clin. Pathol. 55:143–144. Benton BK, Tinkelenberg A, Gonzalez I, Cross FR (1997) Cla4p, a Saccharomyces cerevisiae Cdc42p-activated kinase involved in cytokinesis, is activated at mitosis. Mol Cell Biol. 17: 5067–5076. Borneman, A. R., Hynes, M. J., Andrianopoulos, A. (2000). The abaA homologue of Penicillium marneffei participates in two developmental programmes: conidiation and dimorphic growth. Molec. Microbiol. 38(5): 1034- 47. Borneman, A. R., Hynes, M. J., Andrianopoulos, A. (2001). An STE12 homolog from the asexual, dimorphic fungus Penicillium marneffei complements the defect in sexual development of an Aspergillus nidulans steA mutant. Genetics. 157(3): 1003- 14. Borneman, A. R., Hynes, M. J., Andrianopoulos, A. (2002). A basic helix-loop-helix protein with similarity to the fungal morphological regulators, Phd1p, Efg1p and StuA, controls conidiation but not dimorphic growth in Penicillium marneffei. Mol. Microbiol. 44(3):621- 31. Borradori, L., J. C. Schmit, M. Stetzkowski, P. Dussoix, J. H. Saurat, and I. Filthuth. (1994). Penicilliosis marneffei infection in AIDS. J. Am. Acad. Dermatol. 31:843–846. Boyce, K. J., and Andrianopoulos, A. (2007). A p21- activated kinase is required for 71 conidial germination in Penicillium marneffei. Plos Patho 3(11): e162. Boyce, K. J., Hynes, M. J., Andrianopoulos, A. (2001). The CDC42 homologue of the dimorphic fungus Penicillium marneffei is required for correct cell polarization during growth but not development. J. Bacteriol. 183(11): 3447- 57. Boyce, K. J., Hynes, M. J., Andrianopoulos, A. (2003). Control of morphogenesis and actin localization by the Penicillium marneffei RAC homologue. Journal of Cell Science. 116: 1249- 60. Boyce, K. J., Schreider, L., Andrianopoulos, A. (2009). In vivo yeast cell morphogenesis is regulated by a p21- activated kinase in the human pathogen Penicillium marneffei. Plus Pathog. 5(11): e1000678. Bürglin,T.R. (1991). The TEA domain: a novel, highly conserved DNA-binding motif. Cell. 66: 11-12. Byers, B., and Goetsch, L. (1976). A highly ordered ring of membrane- associated filaments in budding yeast. J. Cell Biol. 69: 717- 21. Canovas, D., Andrianopoulos, A. (2006). Developmental regulation of the glyoxylate cycle in the human pathogen Penicillium marneffei. Molec. Microbiol. 62(6): 1725- 38. Cao, C., Li, R., Wan, Z., Liu, W., Wang, X., Qiao, J., Wang, D., Bulmer, G., Calderone, R. (2007). The effects of temperature, pH, and salinity on the growth and dimorphism of Penicillium marneffei. Medical Mycology. 45: 401- 407. 72 Capponi, M., P. Sureau, and G. Segretain. (1956). Pe´nicillose de Rhizomys sinensis. Bull. Soc. Pathol. Exot. 49:418–421. Cardoza, R. E., Vizcaino, J. A., Hermosa, M. R., Sousa, S., Gonzalez, F. J., Llobell, A., Monte, E., Gutierrez, S. (2006). Cloning and characterization of the erg1 gene of Trichoderma harzianum: effect of the erg1 silencing on ergosterol biosynthesis and resistance to terbinafine. Fungal Genetic Biol. 43(3): 164- 78. Chan, Y. F., and Chow, T. C. (1990). Ultrastructural observations on Penicillium marneffei in natural human infection. Ultrastruct. Pathol. 14(5): 439- 52. Chang, C. C., S. T. Liao, W. S. Huang, J. D. Liu, and L. S. Shih. (1995). Disseminated Penicillium marneffei infection in a patient with acquired immunodeficiency syndrome. J. Formos. Med. Assoc. 94:572–575. Chang, K. C.,Chan, C. K., Chow, K. C., and Lam, C. W. (1998). Penicillium marneffei infection and solitary pulmonary nodule. Hong Kong Med. J. 4:59–62. Chariyalertsak, S., Sirisanthana, T., Supparatpinyo, K., and Nelson, K. E. (1996). Seasonal variation of disseminated Penicillium marneffei infections in northern Thailand: a clue to the reservoir? J. Infect. Dis. 173:1490–1493. Chariyalertsak, S., Sirisanthana, T., Supparatpinyo, K., Praparattanapan, J., and Nelson, K. E. 1997. Case-control study of risk factors for Penicillium marneffei infection in human immunodeficiency virus-infected patients in northern Thailand. Clin. Infect. Dis. 24:1080–1086. Cooper, C. R., Jr. (1998). From bamboo rats to humans: the odyssey of Penicillium 73 marneffei. ASM News. 64:390–397. Cooper, Jr., C. R., and Haycocks, N. G. (2000). Penicillium marneffei: An insurgent species among the Penicillia. J. Eukaryot. Microbiol. 47(1): 24- 28. Cooper, Jr., C. R., and Vanittanakom, N. (2008). Insights into the pathogenicity of Penicillium marneffei. Future Microbiol. 3(1): 43- 55. Deacon, J. W. (2006). Fungal Biology. 4 th ed. Malden, MA: Blackwell Publishing. 371 p. Depraetere, K., Colebunders, R., Ieven, M., de Droogh, E., Pelgrom, Y., Hauben, E., van Mark, E., and Devroey, C. (1998). Two imported cases of Penicillium marneffei infection in Belgium. Acta Clin. Belg. 53:255–258. Desakorn, V., Simpson, A. J. H, Wuthiekanun, V., Sahassananda, D., Rajanuwong, A., Pitisuttithum, P., Howe, P. A., Smith, M. D., and White, N. J. (2002). Development and evaluation of rapid urinary antigen detection tests for diagnosis of penicilliosis marneffei. J. Clin. Microbiol. 40:3179–3183. Di Salvo, A. F., Fickling, A. M., and Ajello, L.. (1973). Infection caused by Penicillium marneffei: description of first natural infection in man. Am. J. Clin. Pathol. 59:259–263. Domer, J. E., Hamilton, J. G., Harkin, J. C. (1967). Comparative study of the cell walls of the yeastlike and mycelia phases of Histoplasma capsulatum. J. Bacteriol. 94:466-474. Ford, S.K., Pringle, J.R. (1991). Cellular morphogenesis in the Saccharomyces cerevisiae cell cycle: localization of the CDC11 gene product and the timing of events at the budding site. Devel. Genet.12:281–292. 74 Gallagher, L. (2010). Functional genomic analysis of natural product biosynthesis and secretion in Aspergillus fumigatus. Thesis. Garbino, J., Kolarova, L., Lew. D., Hirschel, B., and Rohner, P. (2001). Fungemia in HIV-infected patients: a 12-year study in a tertiary care hospital. AIDS Patients Care Sex. Transm. Dis. 15:407–410. Garrison, R. G. and Boyd, K. S. (1973). Dimorphism of Penicillium marneffei as observed by electron microscopy. Can. J. Microbiol., 19:1305–1309. Gifford, T. and Cooper, C. (2009). Karyotype determination and gene mapping in two clinical isolates of Penicillium marneffei. Medical Mycology. 47(3): 286-295. Gladfelter, A. S., Zyla, T. R., Lew D. J. (2004). Genetic interactions among regulators of septin organization. Eukaryot Cell 3: 847–854. Grise, G., Aouar, M., Brasseur, P., and Humbert, G. (1997). Infection a` Penicillium marneffei: une pathologie a` connaıˆtre. Ann. Biol. Clin. 55:241– 242. Haarer, B. K., and Pringle, J. R. (1987). Immunofluorescence localization of the Saccharomyces cerevisiae CDC12 gene product to the vicinity of the 10- nm filaments in the mother- b ud neck. Mol. Cell Biol. 7: 3678- 3687. Harris, J. K. (1986). Modified method for fungal slide culture. J. Clin. Microbiol. 24(3): 460- 61. Herth, W. (1980). Calcofluor white and Congo red inhibit chitin microfibril assembly of Poterioochromonas : avidence for a gap between polymerization and microfibril formation. J. Cell Biol.87: 442- 450. Heath, T. C., Patel, A., Fisher, D., Bowden, F. J., and Currie, B. (1995). Disseminated Penicillium marneffei presenting illness of advanced HIV infection: 75 a clinicopathological review, illustrated by a case report. Pathology. 27:101–105. Hien, T. V., Loc, P. P., Hoa, N. T. T., Duong, N. M., Quang, V. M., McNeil, M. M., Dung, N. T., and Ashford, D. A.( 2001). First case of disseminated Penicilliosis marneffei infection among patients with acquired immunodeficiency syndrome in Vietnam. Clin. Infect. Dis. 32:e78–80. Hilmarsdottir, I., Meynard, J. L., Rogeaux, O., Guermonprez, G., Datry, A., Katlama, C., Bruecker, G., Coutellier, A., Danis, M., and Gentilini, M. (1993). Disseminated Penicillium marneffei infection associated with human immuno- deficiency virus: a report of two cases and a review of 35 published cases. J. Acquir. Immune Defic. Synd. 6:466–471. Hilmarsdottir, I., Coutellier, A., Elbaz, J., Klein, J. M., Datry, A., Gue´ho, E., and Herson, S. (1994). A French case of laboratory-acquired disseminated Penicillium marneffei infection in a patient with AIDS. Clin. Infect. Dis. 19:357– 358. (Letter.) Hsueh, P. R., Teng, L. J.,Hung, C. C., Hsu, J. H., Yang, P. C., Ho, S. W., and Luh, K. T. (2000). Molecular evidence for strain dissemination of Penicillium marneffei: An emerging pathogen in Taiwan. J. Infect. Dis. 181:1706–1712. Hulshof, C. M. J., van Zanten, R. A. A., Sluiters, J. F., van der Ende, M. E., Samson, R. S., Zondervan, P. E., and Wagenvoort, J. H. (1990). Penicillium marneffei infection in an AIDS patient. Eur. J. Clin. Microbiol. Infect. Dis. 9:370. Hung, C. C., Hsueh,P. R., Hsieh, S. M., Liu, C. J., Chen, M. Y., and Luh, K. T. (1998). Bacteremia and fungemia in patients with advanced human immunodeficiency virus (HIV) infection in Taiwan. J. Formos. Med. Assoc. 97:690–697. 76 Huynh, T. X., Nguyen, H. C., Dinh Nguyen, H. M., Do, M. T., Odermatt- Biays, S., Degremont, A., and Malvy, D. (2003). Penicillium marneffei infection and AIDS. A review of 12 cases reported in the Tropical Diseases Centre, Ho Chi Minh City (Vietnam). Sante 13:149–153. Juvvadi, P. R., Fortwendel, J. R., Rogg, L. E., Steinbach, W. J. (2011). Differential localization patterns of septins during growth of the human fungal pathogen Aspergillus fumigatus reveal novel functions. Biochem. Biophys. Res.Commun. Jones, P. D., and See,J. (1992). Penicillium marneffei infection in patients infected with human immunodeficiency virus: late presentation in an area of nonendemicity. Clin. Infect. Dis. 15:744. (Letter.) Joshi, A., Gugnani, H. C., and Vijayan, V. K. (2003). Survival of Penicillium marneffei in sterile and unsterile soil. J. Mycol. Med. 13:211–212. Julander, I., and Petrini, B.. 1997. Penicillium marneffei infection in a Swedish HIV -infected immunodeficient narcotic addict. Scand. J. Infect. Dis. 29:320–322. Kanetsuna, F., Carbonell, L. M., Moreno, R. E., and Rodreiguez, J. (1969). Cell wall composition of the yeast and mycelial forms of Paracoccidioides brasiliensis. J. Bacteriol. 97:1036-1041. Kim, H. B., Haarer, B. K., Pringle, J. R. (1991). Cellular morphogenesis in the Saccharomyces cerevisiae cell cycle: localization of the CDC3 gene product and the timing of events at the budding site. J. Cell Biol. 112: 535- 544. Ko, K. F. (1994). Retropharyngeal abscess caused by Penicillium marneffei: an unusual cause of upper airway obstruction. Otolaryngol. Head Neck Surg. 110: 445–446. 77 Kok, I., Veenstra, J., Rietra, P. J., Dirks-Go, S., Blaauwgeers, J. L., and Weigel, H. M. (1994). Disseminated Penicillium marneffei infection as an imported disease in HIV-1 infected patients. Description of two cases and a review of the literature. Netherlands J. Med. 44: 18–22. Kronauer, C. M., Scha¨r, G., Barben, M., and Bu¨hler. H. (1993). Die HIVassoziierte Penicillium-marneffei-infektion. Schweiz Med. Wschr. 123:385– 390. Kummasook, A. Identification of genes associated with fungal morphogenesis in Penicillium marneffei by random insertional mutagenesis. (2010). Ph. D. Dissertation. Kummasook, A., Cooper, Jr., C. R., Vanittanakom, N. (2010). An improved Agrobacterium-mediated transformation system for the functional genetic analysis of Penicillium marneffei. Medical Mycology. 48: 1066- 74. Kavanagh, K. (ed.) (2007). New Insights In Medical Mycology. Springer. Lamson. R. E., Winters, M. J., Pryciak, P. M. (2002). Cdc42 regulation of kinase activity and signaling by the yeast p21-activated kinase Ste20. Mol. Cell Biol. 22: 2939– 2951. Lazo, G. R., Stein, P. A., Ludwig, R. A. (1991). A DNA transformation-competent Arabidopsis genomic library in Agrobacterium. Biotechnology. 9: 963- 7. Liao, X., Y. Ran, Chen, H., Meng, W., Xiang, B., Kang, M., Xiong, Z., Zhuang, J., Peng, X., Deng, C., Li, G., and Liu, W. (2002). Disseminated Penicillium marneffei infection associated with AIDS, report of a case. Zhonghua. Yi. Xue. Za. Zhi. 82:325–329. Lindsey, R., and Momany, M. (2006). Septin localization across kingdoms: 78 three themes with variations. Curr. Opin. Microbiol. 9:559–565. Lindsey, R., Ha, Y., Momany, M. (2010). A septin from the filamentous fungus A. nidulans induces atypical pseudohyphae in the budding yeast S. cerevisiae. PLoS ONE 5(3): e9858. Lindsey, R., Cowden, S., Hernandez-Rodriguez, Y., and Momany, M. (2010). Septins AspA and AspC are important for normal development and limit the emergence of new growth foci in the mutlicellular fungus Aspergillus nidulans. Eukaryotic Cell. 9(1): 155- 163. Liu, M. T., Wong, C. K., and Fung, C. P. (1994). Disseminated Penicillium marneffei infection with cutaneous lesions in an HIV-positive patient. Br. J. Dermatol. 131:280–283. Longtine, M. S., DeMarini, D. J., Valencik, M. L., Al- Awar, O. S., Fares, H., De Virgilio, C., Pringle, J. R. (1996). The septins: roles in cytokinesis and other processes. Current Opinion in Cell Biol. 8: 106- 119. Lupas,A., Dyke, M. V., and Stock, J. (1991). Predicting coiled coils from protein sequences. Science. 252: 1162–1164. Matsuoka, H., Yang, H. C., Homma, T., Nemoto, Y., Yamada, S., Sumita, O., Takatori, K.,Kurata, H. (1995). Use of Congo red as a microscopic fluorescence indicator of hyphal growth. Appl. Microbiol. Biotechnol. 43: 102- 108. McShane, H., Tang, C. M., and Conlon, C. P. (1998). Disseminated Penicillium marneffei infection presenting as a right upper lobe mass in an HIV positive patient. Thorax. 53:905–906. 79 Michielse, C., Hooykaas, P. J. J., van den Hondel, C. A. M. J. J., Ram, A. F. J. (2005). Agrobacterium- mediated transformation as a tool for functional fenomics in fungi. Curr. Genet. 48: 1- 17. Mohri, S., Yoshikawa, K., Sagara, H., and Nakajima, H. (2000). A case of Penicillium marneffei infection in an AIDS patient: the first case in Japan. Nippon Ishinkin Gakkai Zasshi 41:23–26. Momany, M., and Hamer, J. E. (1997). The Aspergillus nidulans septin encoding gene, aspB, is essential for growth. Fungal Genetics and Biology. 21: 92- 100. Momany, M., Zhao, J., Lindsey, R., Westfall, P. J. (2001). Characterization of the Aspergillus nidulans septin (asp) gene family. Genetics. 157: 969- 977. Nord, J., Karter, D., and LaBombardi, V. (1998). An AIDS patient with fever and pancytopenia. Int. J. Infect. Dis. 2:173–175. Noritomi, D. T., Guilherme, L. B., Beer, I., da Silva, A. F., de Cleva, R., Gama, R. J. J. (2005). Multiple brain abscesses due to Penicillium spp. infection. Rev. Inst. Med. Trop. 47: 167- 170. Pan, F., Malmberg, R. L., Momany, M. (2007). Analysis of septins across kingdoms reveals orthology and new motifs. BMC Evolutionary Biology. 7: 103. Pancaldi, S., Poli, F., Dall’Olio, G., Vannial, G. L. (1984). Morphological anomalies induced by Congo red in Aspergillus niger. Arch. Microbiol. 137: 185- 187. Peter, M., Neiman, A. M., Park, H. O., van Lohuizen, M., Herskowitz, I. (1996). Functional analysis of the interaction between the small GTP binding protein Cdc42 and the Ste20 protein kinase in yeast. Embo J. 15: 7046–7059. 80 Peto, T. E. A., Bull. R, Millard, P. R., Mackenzie, D. W., Campbell, C. K., Haines, M. E., and Mitchell, R. G. (1988). Systemic mycosis due to Penicillium marneffei in a patient with antibody to human immunodeficiency virus. J. Infect. 16: 285– 290. Pitt, J. I. (1979). Genus Penicillium and its teleomorphic states Eupenicillium and Talaromyces. Academic Press, New York, N.Y. Pongpom, P., Cooper, Jr., C. R., Vanittanakom, N. (2005). Isolation and characterization of a catalase- peroxidase gene from the pathogenic fungus, Penicillium marneffei. Medical Mycology. 43: 403- 411. Qin, G., Liu, J., Cao, B., Tian, S. (2011). Hydrogen peroxide acts on sensitive mitochondrial proteins to induce death of a fungal pathogen revealed by proteomic analysis. PLoS ONE 6(7): e21945. Ramirez, C. (1982). Manual and Atlas of the Penicillia. Elsevier, Amsterdam. 874 p. Ranjana, K. H., Priyokumar. K., Singh, T. J., Gupta, Ch. C., Sharmila, L., Singh, P. N., and Chakrabarti, A. 2002. Disseminated Penicillium marneffei infection among HIV-infected patients in Manipur state, India. J. Infect. 45: 268– 271. Richman, T. J., Sawyer, M. M., Johnson, D. I. (1999). The Cdc42p GTPase is involved in a G2/M morphogenetic checkpoint regulating the apical-isotropic switch and nuclear division in yeast. J Biol Chem 274: 16861– 16870. Rimek, D., Zimmermann, T., Hartmann, M., M., Prariyachatigul, C., and Kappe, R. (1999). Disseminated Penicillium marneffei infection in an HIVpositive female from Thailand in Germany. Mycoses 42: 25– 28. Rokiah, I., Ng, K. P., and Soo Hoo, T. S. (1995). Penicillium marneffei infection 81 in an AIDS patient—a first case report from Malaysia. Med. J. Malaysia 50: 101– 104. Roncero, C., and Duran, A. (1985). Effect of Calcofluor white and Congo red on fungal cell wall morphogenesis: in vivo activation of chitin polymerization. J. Bacteriol. 163(3): 1180- 1185. Rongrungruang, Y., and Levitz, S. M. (1999). Interactions of Penicillium marneffei with human leukocytes in vitro. Infection and Immunity. 67(9): 4732- 4736. Saraste, M., Sibbald, P. R., Wittinghofer, A. (1990). The P- loop: a common motif in ATP- and GTP- binding proteins. Elsevier Science Publishers Ltd. (UK). Sathapatayavongs, B., Damrongkitchaiporn, S., P. Saengditha, P., Kiatboonsri, P., and Jayanetra, P. (1989). Disseminated penicilliosis associated with HIV infection. J. Infect. 19: 84– 85. Segretain, G. (1959). Penicillium marneffei n. sp., agent d’une mycose du systeme reticuloendothelial. Mycopathol. Mycol. Appl. 11: 327– 353. Sobottka, I., Albrecht, H., Mack, D., Stellbrink, H. J., van Lunzen, J., Tintelnot, K., and Laufs, R. (1996). Systemic Penicillium marneffei infection in a German AIDS patient. Eur. J. Clin. Microbiol. Infect. Dis. 15: 256– 259. Supparatpinyo, K., Khamwan, C., Baosoung, V., Nelson, K. E., and Sirisanthana, T. (1994). Disseminated Penicillium marneffei infection in Southeast Asia. Lancet. 344: 110-113. Taramelli, D., Tognazioli, C., Ravagnani, F., Leopardi, O., Giannulis, G., Boelaert, J. R. (2001). Inhibition of intramacrophage growth of Penicillium marneffei by 4- aminoquinolines. Antimicrobial Agents and Chemotherapy. 45(5): 1450- 1455. 82 Thirach, S., Cooper, Jr., C. R., Vanittanakom, N. (2008). Molecular analysis of the Penicillium marneffei glyceraldehyde- 3- phosphate dehydrogenase-encoding gene (gpdA) and differential expression of gpdA and the isocitrate lyase- encoding gene (acuD) upon internalization by murine macrophages. Journal of Medical Microbiology. 57: 1322- 1328. Todd, R. B., Greenhalgh, J. R., Hynes, M. J., and Andrianopoulos, A. (2003). TupA, the Penicillium marneffei Tup1p homologue, represses both yeast and spore development. Mol. Microbiol. 48: 85– 94. Tsang, D. N. C., Li, P. K. C., Tsui, M. S., Lau, Y. T., Ma, K. F., and Yeoh, E. K. (1991). Penicillium marneffei: another pathogen to consider in patients infected with human immunodeficiency virus. Rev. Infect. Dis. 13: 766– 767. (Letter.) Tsui, W. N., Ma, K. F., and Tsang, D. N. C. (1992). Disseminated Penicillium marneffei infection in HIV-infected subject. Histopathology 20: 287– 293. Tsunemi, Y., Takahashi, T., and Tamaki, T. 2003. Penicillium marneffei infection diagnosed by polymerase chain reaction from the skin specimen. J. Am. Acad. Dermatol. 49: 344– 346. Valeyrie, L., Botterel, F., Minozzi, C., Roger, P., Bourree, P., and Vittecoq, D. (1999). Prolonged fever revealing disseminated infection due to Penicillium marneffei in a French HIV-seropositive patient. AIDS 13: 731– 732. (Letter.) Vanittanakom, N., Cooper, Jr., C. R., Fisher, M. C., Sirisanthana, T. (2006). Penicillium Marneffei infection and recent advances in the epidemiology and molecular biology aspects. Microbiology Reviews. 19(1): 95 – 110. Vanittanakom, N., Mekaprateep, M., Sriburee, P., Vanittanakom, P., and Khanjanasthiti, 83 P. (1995). Efficiency of flotation method in the isolation of Penicillium marneffei from seeded soil. J. Med. Vet. Mycol. 33: 271– 273. Vilar, F. J., Hunt, R., Wilkins, E. G., Wilson, E. G. G., and Jones, N. P. (2000). Disseminated Penicillium marneffei in a patient infected with human immunodeficiency virus. Int. J. Sex. Transm. Dis. AIDS 11: 126– 128. Westfall, P. J., and Momany, M. (2002). Aspergillus nidulans septin AspB plays pre- and post-mitotic roles in septum, branch, and conidiophores development. Molecular Biology of the Cell. 13: 110- 118. Wong, K. H., Lee, S. S., Chan, K. C., and Choi, T. (1998). Redefining AIDS: case exemplified by Penicillium marneffei infection in HIV-infected people in Hong Kong. Int. J. Sex. Transm. Dis. AIDS 9: 555– 556. Zhang, P., Xu, B., Wang, Y., Li, Y., Qian, Z., Tang, S., Huan, S., Ren, S. (2008). Agrobacterium tumefaciens- mediated transformation as a tool for insertional mutagenesis in the fungus Penicillium marneffei. Mycological Research. 112: 943- 949. Zuber, S., Hynes, M. J., and Andrianopoulos, A. (2003). The G-protein α- subunit GasC plays a major role in germination in the dimorphic fungus Penicillium marneffei. Genetics. 164: 487– 499.