A. n ‘ , . a.» .. . 3.7. .- ; .. _ . .u..a,....§%. ‘ , , _ , ‘ ‘ aft giwfifixmflififlfifli m a ..u.. my? \ n=. . ‘ , . i . T u .4 2w... #ww IWKF. . . . M, M... . . I n 5 ii. 1 JR . .. l . ; fig-3.: n. a V2}. . 3.5.3“? Lag amt arm .3 am. .. : .6? .§ W5; .L. .X If :53»... 3w. a (an: . 2.1;? F: . Wizsnrflu. i 1....- “on-5’ M' @336. _ hfifinfiw , . ‘ ‘ 32.». .,L .3». ‘3‘..,...4§3£.$ , , . , :9. .2 ., (310qu LIBRARY Suzi/16914 Michigan State University This is to certify that the dissertation entitled POSITION-DEPENDENT REGULATION OF THE NOR-1 PROMOTER IN THE AFLATOXIN GENE CLUSTER AND lMMUNO—LOCALIZATION OF PROTEINS INVOLVED IN AFLATOXIN BIOSYNTHESIS presented by Ching-Hsun Chiou has been accepted towards fulfillment of the requirements for the Doctoral degree in Food Science and Environmental Toxicology @2117? £46 [Major Professors Signaturé [Z— 3 - (>3 Date MSU is an Affirmative Action/Equal Opportunity institution _.-_.—u-o-.-o-o—c-—o- PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE. 6/01 c:/CIRC/DateDue.p65—p.15 POSITION-DEPENDENT REGULATION OF THE NOR-1 PROMOTER IN THE AFLATOXIN GENE CLUSTER AND IMMUNO-LOCALIZATION OF PROTEINS INVOLVED IN AFLATOXIN BIOSYNTHESIS By Ching-Hsun Chiou A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Food Science and Human Nutrition 2003 ABSTRACT POSITION-DEPENDENT REGULATION OF THE NOR-1 PROMOTER IN THE AF LATOXIN GENE CLUSTER AND IMMUNO-LOCALIZATION OF PROTEINS INVOLVED IN AF LATOXIN BIOSYNTHESIS By Ching-Hsun Chiou Aflatoxins (AF) are highly toxic and carcinogenic fungal secondary metabolites. AF contamination in crops including com, peanuts, cottonseed, and specific tree nuts has caused significant agricultural economic and human health concerns throughout the world. My research goal was to study regulation of AF biosynthesis in Aspergillus parasiticus. As along term goal, these studies should aid efforts to discover strategies to efficiently eliminate AF contamination of agricultural commodities and in turn to ensure product safety and consumer health. Biosynthesis of AF in the filamentous fungi in the genus Aspergillus utilizes a complex pathway consisting of at least eighteen enzymes and at least one positive regulatory factor, AflR; these genes are clustered within a 75 kb chromosomal region. Expression of most of the AF genes in the cluster is co-regulated but the mechanisms are not clear. The localization and interaction of these enzymes to synthesize AF at the cellular level has yet to be determined. Therefore, two hypotheses are included in my study. 1) Expression of genes involved in aflatoxin biosynthesis is positively regulated within the AF gene cluster. 2) Enzymes involved in AF biosynthesis are localized within developmental structures (e.g., conidiophores) in the colony. The objectives of my research were to: 1) study the positional-effects on AF gene (nor-I) expression in the gene cluster; 2) generate highly-specific antibodies against an AF protein (anti-VBS) for immuno-localization studies; 3) develop methods to localize AF proteins within a fungal colony; and 4) localize proteins involved in AF biosynthesis in a time-course dependent fractionated colony. To study positional effects on nor-I promoter function, a plasmid pAPGUSNP containing the nor-1 promoter sequence fused to a reporter gene (uidA encoding B- glucuronidase, GUS) plus a pyrG selectable marker was constructed. Transformation of pAPGUSNP into CSIO (ver-I, wh-I, pyrG) resulted in homologous recombination at 5' nor-I, 3' nor-1, or the pyrG locus. The plasmid integration site was screened by a novel rapid DNA extraction/PCR assay and confirmed by Southern hybridization analysis. The nor-1 promoter was positively regulated in the AF gene cluster while outside the cluster (pyrG locus), nor-1 promoter activity decreased dramatically. The data supported the hypothesis that expression of AF genes is cluster position-dependent. The second hypothesis was tested by studying localization AF proteins utilizing anti-VBS antibodies raised against recombinant Maltose Binding Protein-VBS expressed in Escherichia coli (generated in this study), and other antibodies available in the laboratory (anti- Nor-l , Ver-l , and OmtA). A 72 h-old colony was fractionated based on 24 h intervals to correlate the time of grth with AF protein distribution. Colony fractions were fixed and embedded in paraffin to generate 4-5 pm sections for immuno- fluorescence labeling. Confocal Laser Scanning Microscopy (CLSM) demonstrated that Nor-1, Ver-l , OmtA, and VBS were evenly distributed among different cell types and not concentrated in conidiophores, contrary to the hypothesis. These proteins were not evenly distributed throughout a 72 h—old colony. Among three fractions, colony fraction S2 (24 to 48 h growth) exhibited the highest abundance of Nor-1 , Ver—l , OmtA, and VBS (analyzed by CLSM and quantitative fluorescence intensity analysis), the highest levels of AF Bl (determined by enzyme-linked-immunosorbent-assay), and the highest numbers of newly-developed conidiophores. These data support a spatial and temporal association between AF protein expression/AF production and sporulation at the colony level. ACKNOWLEDGMENTS I would first like to thank my major professor, Dr. John Linz, for his support and guidance throughout this study. He is a good mentor who teaches me science and many other precious things beyond the science. I admire Dr. Linz as a good scientist and a knowledgeable person. Appreciation is also expressed to Dr. Karen Klomparens, Dr. James Pestka, Dr. Gale Strasburg, and Dr. Jay Schroeder for serving on my guidance committee. I am also especially gratefiil to Dr. Shirley Owens and others who assisted me in the Center for Advanced Microscopy at Michigan State University. Thanks to Li-Wei Lee, Mike Miller, Matt Rarick, and other members in the Linz laboratory for all the help they have provided. I also appreciate my parents, family, and friends, who always stand by me and show me their support and encouragement. iv TABLE OF CONTENTS LIST OF TABLES .................................................................................. ix LIST OF FIGURES ........................................................................................................ x CHAPTER 1 Literature review .............................................................................................................. 1 DELETERIOUS EFFECTS OF AFLATOXINS .................................................. 1 Background and history .............................................................................. 1 Economic impact and regulation of AF in foods ......................................... 3 Environmental factors and controlling AF contamination .......................... 4 Toxicological effects of AF ........................................................................ 5 Aflatoxin and liver cancer ............................................................................ 6 Bioactivation of AFB1 .................................................................... 6 Detoxification of AF ........................................................................ 6 AF B1-DNA adducts and AF B1 indcued p53 gene mutations ........ 6 A synergistic effect between AFB1 and hepatitis B virus (HBV) ......................................................................................... 8 AFLATOXIN BIOSYNTHESIS .......................................................................... .9 Aflatoxin gene cluster ...................................................................... 9 Regulation of AF biosynthesis ....................................................... 9 Partially-duplicated AF gene cluster ............................................. 13 Sugar utilization gene cluster ........................................................ 14 SECONDARY METABOLISM AND FUNGAL DEVELOPMENT .............. 14 Association between secondary metabolism and fungal development ......................................................................................... 14 G-Protein signaling pathway ..................................................................... 14 FluG .......................................................................................................... 1 5 AF bioynthesis and conidiation ............................................................... 15 LOCALIZATION OF AF ASSOCIATED PROTEINS ................................... .17 Penicillin biosynthesis as an example ........................................................ 17 Organelles possibly involved in AF biosynthesis18 Microbodies .................................................................................. 18 Woronin bodies ............................................................................. 18 Vesicles ......................................................................................... 19 Vacuoles ........................................................................................ 19 Endoplasmin reticulum (ER) ........................................................ 21 CONFOCAL LASER SCANNING MICROSCOPY (CLSM) IN THE STUDY OF FUNGAL BIOLOGY ..................................................................................... 22 Introducion to CLSM ................................................................................ 22 Image quality enhancement by CLSM ..................................................... 23 CLSM Imaging ...................................................................................... 25 Optical sections ........................................................................... .25 Z-series .......................................................................................... 25 Phi-z sectioning ............................................................................. 26 Time-lapse and live imaging .......................................................... 26 Fluorescence imaging ................................................................... .26 CHAPTER2 A rapid DNA extraction method for screening large numbers of Aspergillus transformants via polymerase chain reaction ............................................................. 31 ABSTRACT ......................................................................................................... 31 INTRODUCTION ............................................................................................... 3] MATERIAL AND METHODS .......................................................................... 32 RESULTS AND DISCUSSION ........................................................................... 33 ACKNOWLEDGEMENTS ................................................................................. 4 1 CHAPTER3 Position-dependent regulation of the nor-1 promoter in the aflatoxin gene cluster ............................................................................................................................... 42 ABSTRACT ......................................................................................................... 42 INTRODUCTION ............................................................................................... 43 MATERIALS AND METHODS ....................................................................... 45 Strains and grth conditions .................................................................. 45 Construction of pAPGUSNP ................................................................... 45 Fungal transformation ............................................................................... 48 Identification of the site of plasmid integration ........................................ 48 Rapid DNA extraction procedure for PCR analysis ..................... 48 PCR analysis ................................................................................. 48 Solid-culture GUS activity assay .............................................................. 49 Southern hybridization analysis ................................................................ 49 5' nor-1 integration ........................................................................ 51 3' nor-1 integration ........................................................................ 51 Integration at pyrG ........................................................................ 51 RESULTS .............................................................................................................. 51 Generation of pAPGUSNP ..................................................................... .51 Transformation of A. parasiticus CSIO with pAPGUSNP ...................... 56 Solid-culture GUS assay ........................................................................... 56 Determination of pAPGUSNP integration site within the chromosome ......................................................................................... 56 Confirmation of pyrG expression in pyrG+ transformants ....................... 59 DISCUSSION ....................................................................................................... 59 ACKNOWLEDGEMENTS ................................................................................. 63 CHAPTER4 Immuno-localization of Nor-l, Ver-l, and OmtA proteins involved in aflatoxin biosynthesis in Aspergillus parasiticus .......................................................................... 64 INTRODUCTION ............................................................................................... 64 MATERIALS AND METHODS ........................................................................ 67 vi Fungal strains ............................................................................................ 67 Time-dependent fractionation of colonies grown on solid medium .......... 67 Preparation of paraffin-embedded fungal sections .................................... 67 Immuno-fluorescence labeling of firngal cells .......................................... .69 Confocal Laser Scanning Microscopy (CLSM) ........................................ 70 Quantitative fluorescence intensity analysis ............................................. 70 RESULTS .............................................................................................................. 71 Immuno-localization of Nor-1 and Ver-l .................................................. 71 Immuno-localization of OmtA .................................................................. 71 Quantitative fluorescence intensity analysis ............................................. 79 DISCUSSION ....................................................................................................... 85 ACKNOWLEDGEMENTS ................................................................................. 89 CHAPTERS Distribution and sub-cellular localization of the aflatoxin enzyme versicolorin B synthase (VBS) in time-fractionated colonies of Aspergillusparasiticus .................. 90 ABSTRACT ......................................................................................................... 90 INTRODUCTION ............................................................................................... 91 MATERIALS AND METHODS ........................................................................ 94 Fungal strains and growth conditions ....................................................... 94 Production of antibodies against native VBS ........................................... 94 Generation of MBP (Maltose Binding Protein)-VBS fusion protein ....... 95 Production and purification of PAb against MBP-VBS ........................... 95 Time-dependent fractionation of fiingal colonies ..................................... 96 Fungal protein extraction .......................................................................... 96 SDS-Polyacrylamide Gel Electrophoresis (SDS-PAGE) and Western blot analysis ........................................................................... 97 Detection of AFB1 in colony fiactions ..................................................... 97 Thin Layer Chromatagraphy (TLC) ............................................. 97 Enzyme-Linked-Immunosorbent-Assay (ELISA) ........................ 97 Preparation of paraflin-embedded fungal sections and immuno- fluorescence labeling ............................................................................ 98 Confocal Laser Scanning Microscopy (CLSM) ........................................ 99 Quantitative fluorescence intensity analysis ............................................. 99 Statistical analysis ................................................................................... 100 Preparation of ultra-thin fungal sections, immuno-gold labeling, and TEM analysis ............................................................................ . 100 RESULTS ............................................................................................................ 100 Generation of PAb against native firngal VBS ......................................... 100 Expression of MBP-VBS fusion protein and production of PAb against VBS ........................................................................................ 101 Expression of VBS in A. parasiticus grown on solid medium ............... . 106 Confocal Laser Scanning Microscopy (CLSM) of time-dependent expression of VBS in a 72 h-old firngal colony .................................. 109 Fluorescence intensity quantification and statistical analysis ................. 109 Production of AFB1 in colony fractions ................................................. 109 vii Sub-cellular localization of VBS ............................................................ . 109 DISCUSSION ..................................................................................................... l 16 ACIG‘IOWLEDGEMENTS ............................................................................... 124 BIBLIOGRAPHY ......................................................................................................... .125 viii LIST OF TABLES Table 2.1. Primer sequences and PCR cycling parameters used for analysis of fungal genomic DNA templates ........................................................................... 34 Table 3.1. Primer sequences and PCR cycling parameters used for analysis of site of integration using DNA templates prepared by the rapid method ............................. 50 Table 3.2. Summary of GUS activities and plasmid integration sites in A. paraciticus CSIO transformants carrying the nor::GUS firsion plasmid pAPGUSNP (Chiou), and NR-l harboring pAPGUSNNB (Wilson, 1998) and pNG3.0 (Miller, 2003) ............... 61 Table 4.1. Quantative fluorescence intensity analysis of A. parasiticus SU-l and AF SIO immuno-fluorescence labeled with Nor-1, Ver-l and OmtA antibodies ..................... 86 Table 5.1. Quantitative fluorescence intensity analysis of A. parasiticus SU—l and AF SlO labeled with polyclonal antibodies to VBS and OmtA ............................ 113 Table 5.2. Quantitative detection (ELISA) of AFB1 in colony fi'actions of SU-I cultured on YES solid media for 72 h ..................................................................... 114 LIST OF FIGURES Figure 1.1. Chemical structures of AFB1, AFB2, AFG], AFG2, AFM], AFM2, and AFB 1-8, 9-epoxide ................................................................................... 2 Figure 1.2. Generally accepted pathway for sterigmatocystin and aflatoxin biosynthesis .......................................................................................... 10 Figure 1.3. The G-protein signaling pathway involved in ST biosynthesis in A. nidulans ............................................................................................... 16 Figure 1.4. Diagram of the light path in a confocal instrument ............................... 24 Figure 1.5. CLSM Z-series of A. parasiticus SU-l conidiophores stained with SYTOX green fluorescence dye .............................................................................. 26 Figure 1.6. CLSM single optical section and phi-z scan of A. parasiticus SU-l microbodies .......................................................................................... 28 Figure 2.1. Homologus recombination of pAPGUSNP into the A. parasiticus chromosome and primers used in integration analysis ......................................... 35 Figure 2.2. Amplification of PCR targets using the rapid DNA extraction technique. . .37 Figure 3.1. Restriction endonuclease maps of plasmids containing the nor-1::GUS construct related to this study ..................................................................... 46 Figure 3.2. Southern hybridization analysis of site of integration in pAPGUSNP transformants ........................................................................................ 52 Figure 3.3. Southern hybridization analysis of pAPGUSNP integrated at pyrG ................................................................................................... 54 Figure 3.4. PCR analysis to detect site of integration in pAPGUSNP transformants. . . 57 Figure 3.5. Solid-culture GUS activity assay on CSIO transformants harboring pAPGUSNP .......................................................................................... 58 Figure 4.1. Time-dependent fractionation of firngal colonies ................................ 68 Figure 4.2. Immuno-fluorescence confocal microscopy of Nor-1 and Ver-l in A. parasiticus SU-l and AF SlO (cyIR knockout mutant) grown on YES agar for 72 h ........ 72 Figure 4.3. CLSM immuno-fluorescence detection of Nor-1 in A. parasiticus SU-1.. ...74 Figure 4.4. CLSM immuno-fluorescence detection of Ver—l in A. parasiticus SU-l . . . ...76 Figure 4.5. Immuno-fluorescence labeling of Ver-l antibodies in the control strain A. parasiticus AF 810 and the wild-type strain SU-l .............................................. 78 Figure 4.6. OmtA protein localization in time-fractionated colonies of A. parasiticus SU- 1, AF S10 ((ng knockout), C 810 and LW1432 (omtA knockout) grown on YES agar for 72 h ........................................................................................................ 80 Figure 4.7. Protein localization using OmtA PAb and anti-SKL, and detection of nuclei using SYTOX Green ................................................................................ 82 Figure 4.8. CLSM immuno-fluorescence detection of OmtA in A. parasiticus SU-l . . . .84 Figure 5.1. Western blot analysis of VBS detected by PAb raised against native firngal VBS ......... . ........................................................................................ 102 Figure 5.2. SDS-PAGE of recombinant MBP-VBS fusion protein ........................ 103 Figure 5.3. MBP-VBS treated with Factor Xa to generate MBP (42 kDa) and VBS (52 kDa) .................................................................................................. 104 Figure 5.4. Western blot analysis of VBS produced in liquid culture using highly specific PAb raised against MBP-VBS ................................................................... 107 Figure 5.5. Western blot analysis of VBS expressed in A. parasiticus SU-I colonies (solid culture) and colony fractions .............................................................. 108 Figure 5.6. CLSM of VBS protein distribution in time-fractionated colonies of A. parasiticus SU-l (S1, S2, S3) and AF 810 (R1, R2, and R3) grown on YES agar for 72h ................................................................................................... 110 Figure 5.7. Thin Layer Chromatagraphy analysis of AFB1 production in colony fractions ............................................................................................. 111 Figure 5.8. CLSM immuno-fluorescence of VBS A. parasiticus SU-l grown in solid culture ............................................................................................... 112 Figure 5.9. CLSM dual detection for localization of VBS and nuclei in A. parasiticus SU-l .................................................................................................. 115 Figure 5.10. Immuno-gold labeling of VBS in A. parasiticus AF $10 and SU-l analyzed by TEM ............................................................................................. 117 CHAPTER 1 Literature review DELETERIOUS EFFECTS OF AFLATOXINS Background and history. Aflatoxins (AF) are secondary metabolites mainly produced by the filamentous fungi Aspergillus flaws and A. parasiticus (Wilson and Payne, 1994; CAST, 2003). They are less frequently produced by A. nomius, A. pseudotamari, A. bombycis, and one isolate of A. ochraceoroseus (Klich et al., 2000; Ito et al., 2001; Bhatnagar et al., 2003). A. parasiticus synthesizes four major aflatoxins including aflatoxin B1(AF B1), AFB2, AFG1, and AFG2 (Asao et al., 1963). Animals ingest AF and produce metabolites AF M1 and AF M2 in milk (Fig. 1.1). Among these AF molecules, AF B1 is the most toxic and carcinogenic (Eaton and Groopman, 1994; Mclean and Dutton, 1995). AP contamination of corn, peanuts, cotton seed, and certain tree nuts has caused significant agricultural economic and human health problems throughout the world (CAST, 2003). This mycotoxin first aroused public attention in the early 19608 with the incidence of "Turkey X" disease (Blount, 1961). Large quantities (approximtely 10, 000) of turkeys died from acute hepatotoxicity caused by contamination of AF in their peanut feed (Siller and Ostler, 1961). Rainbow trout that received AF contaminated diets were discovered to develop liver cancer in the United States (Sinnhuber et al., 1977). In the 19708, severe pre-harvest AF contamination was reported in the Southern and midwestem United States (Anderson, 1975). A. flavus and AF contamination are frequently seen in the Southern US., consistent with the observation that serious outbreaks are related to below-average rainfall and above-average temperature (Payne, 1998). Dietary exposure to AF has been associated with a significant incidence of human liver cancer. Studies on experimental animals indicated that AF B1 is a potent mutagen AF B1-8,9-epoxide I 40A, / ... Figure 1.1. Chemical structures ofAFB], AFBz, AFG1, AFG2, AFM], AFMz, and AF Bl-8, 9-epoxide. and hepato-carcinogen (reviewed in Eaton and Groopman, 1994). Based on this information, the International Agency for Research on Cancer (IARC) identified AF Bl as a group I carcinogen in humans (IARC, 1993). Several studies also indicated that infection of hepatitis B virus (HBV) causes a synergistic effect with AF B1 induced p53 mutation related to hepatocarcinogenesis (HCC) (Hsia et a1, 1992; Harris and Hollstein, 1993; Jackson and Groopman, 1999, CAST, 2003). Hepatitis C viral (HCV) infection is the most important risk factor for HCC in western countries; however, no adequate epidemiological data on the relative roles of HCC and AF were available (Henry et al., 1999; Montalto et al., 2002). Economic impact and regulation of AF in foods. Approximately 25% of food crops are affected by mycotoxin contamination (including AF) every year throughout the world and this results in serious health and economic concerns (Mannon and Johnson, 1985). Since 1977, documents regarding mycotoxin regulation and control have been published by the Food and Agricultural Organization (FAO) (Boutrif, 1995). Over 50 countries developed guidelines to regulate mycotoxins. The United States set action levels for aflatoxin under the Food, Drug and Cosmetic Act (See. 402) (Sharma and Salunkhe, 1991; Bhatanagar et al., 2002; CAST, 2003). The US. Food and Drug Administration (FDA) prohibits interstate commerce of feed grain containing more than 20 parts per billion (ppb, or ug/kg) AF B] and prohibits the sale of milk and eggs containing more than 0.5 ppb AF M1. The action level for a feed crop varies for different animals depending on their sensitivity to AP. Over the world, the maximum allowed levels for aflatoxins in foodstuff range fiom zero tolerance to 50 ppb (Van Egrnond, 1993; Bhatanagar et al., 2002). Contamination of agricultural commodities with AF poses a significant threat to the food supply world-wide. In the United States, the mean economic losses of crops due to mycotoxin contamination (including AF, fumonisins, and deoxynivalenol) are estimated to be $932 million. The loss attributed to AF contamination on food crops (corn and peanuts) was approximately $47 million per year and approximately $225 million per year in feed corn (CAST, 2003). Thus, prevention and control of mycotoxin contamination is an important issue. Environmental factors and control of AF contamination. In the field, A. flaws and A. parasiticus are able to grow over the temperature range from 12 to 48 0C (optimum 35 to 37 0C) and with a water activity as low as 0.80 (optimum 0.95). However, AF production in these fungi requires a smaller range of temperature (28 to 33 0C, optimum 31°C) and water activity (0.85 to 0.97, optimum 0.90) (Hussein and Brasel, 2001; Bhatanagar et al., 2002). Pre-harvest contamination of AF on agriculture commodities is usually related to the environmental conditions that generate fungal growth and toxin synthesis. Several methods have been applied to control AF at this stage, including use of fimgicides and insecticides (insects facilitate Aspergillus infection of plants), modified irrigation practice, and the use of atoxigenic biocompetitive strains (Cotty and Bhatnagar, 1994; Cary et al., 2000; Bhatnagar et al., 2002). However, these procedures have limited effects, particularly in years of drought when fungal growth and toxin synthesis are elevated. Ongoing research is focused on alternative strategies to control AF contamination in the field. Time of harvest is also critical to restrain the levels of AF contamination. For example, contaminated corn should be harvested early and a subsequent drying procedure may help to prevent an increase in AF contamination that occurs post-harvest. Post- harvest contamination can also be controlled by maintaining proper storage conditions (CAST, 2003). For example, limiting the water activity of the harvest grains below 0.85 and maintaining the storage temperature below 10°C may prevent A. flavus spore germination (Marin et al., 1998; CAST, 2003). Toxins on contaminated products can be removed or detoxified by various physical, chemical and biological methods (Bhatnagar et al., 2002; CAST, 2003). For example, toxins can be removed by filtration or adsorption onto clays or activated charcoal. Oil roasting or dry roasting peanuts (Marth and Doyle, 1979) and microwave treatment (F arag et al., 1996) may partially destroy AF although they are heat stable. AF in contaminated grains may be extracted by a selected solvent mixture (e.g., 95% ethanol, 90% aqueous acetone, and 80% isopropanol), yet this procedure is costly and product safety remains a concern (CAST, 2003). Toxicological effects of AF. The degree of AF-induced toxic effects in animals varies, depending on species, dose, exposure duration, and nutritional status of the host (CAST, 2003). The major organ targeted by AF is the liver. Acute aflatoxicosis often produces hepatitis in animals and humans, and ingestion of a high dose may be lethal. Acute hepatitis has been related to consumption of foodstuff contaminated with high doses of AF, especially corn (Ngindu et al., 1982). Under severe conditions, 25% mortality has been reported in outbreaks in India (Krishnamachari et al., 1975). Sublethal doses generate chronic toxicity and low levels of chronic exposure can result in cancer. A sensitive model to study AF toxicity is trout, and the LD50 was estimated to be 0.5 to 1.0 mg/kg (CAST, 2003). A high incidence of hepatoma was seen in trout fed with 80 ppb AF contaminated diet. Studies in cattle showed chronic clinical signs including reduced weight gain, irnmuno-suppression, lower milk production, liver damage, and lower reproduction (Bodine and Mertens, 1983). In humans, chronic dietary exposure to AF has a high risk of generating hepatocellular carcinoma (HCC). A high level of AF daily intake (1.4 ug)(Wild et al., 1992) and a high incidence of hepatitis B infection were associated with HCC (Jackson and Groopman, 1999). Epidemiological studies indicated that an increase in AF ingestion from 3 to 222 ng/kg body weight per day corresponded to elevated liver cancer incidence (from 2 to 35 cases per 100,000 population per year) (Jackson and Groopman, 1999). Aflatoxin and liver cancer. Bioactivation of AFB1. AF B1 naturally produced in the fungus is not a potent mutagen and carcinogen to humans (and other animals); however, the liver cytochrome P450 enzymes generate highly reactive molecules during the oxidative metabolic processes (Eaton and Groopman, 1994). Hence, the major targets for AF Bl toxicity are hepatocytes. In vitro studies have shown that biotransformation of AF B1 requires molecular oxygen and NADPH. The biotransforrned AFB], the AF B1-8,9-epoxides (Fig. 1.1), are potent bioreactive compounds causing mutation and carcinogenesis. The epoxide molecules covalently bind to proteins, DNA and RNA molecules to generate adducts. Adduct formation in turn impairs the normal molecular and cellular fimctions for the targets (McLean and Dutton, 1995). Detoxification of AF. Detoxification of AF B1-8,9—epoxides mainly depends on conjugation with glutathione (GSH). Glutathione S transferase (GST) activity is therefore an important determinant for the susceptibility of different species to the toxic effects of AFB1 (Eaton and Gallagher, 1994). Different animal species also showed different responses to AFBl toxicity due to variability in bioactivation and detoxification processes. In addition to humans, rat, rainbow trout and turkey are susceptible to AF B1 induced toxicity, whereas mice, ducks and chickens showed lower toxic response (Mclean and Dutton, 1995). It is likely that genetic variability in the expression of cytochrome P450 enzymes may result in substantial differences between individuals in susceptibility to the carcinogenic effects of aflatoxins. For example, glutathione S transferase is the predominant detoxifying activity of AF B1-8,9-epoxides in mice. However, in humans, GST has very little activity toward AF B1-8,9-epoxides, suggesting higher human sensitivity to the genotoxic effect of AF B1 (Hussein and Brasel, 2001). AFB1-DNA adducts and AFB1 induced p53 gene mutations. The N7-guanine residue of DNA is a potential target for AF B1-8,9-epoxides, forming a AF Bl-N7-guanine adduct. Adduct formation or depurination can lead to mutations in DNA during DNA replication, mainly resulting in GC—-)TA transversions (Eaton and Groopman, 1994). Mutation in the p53 tumor suppressor gene by AF B1 epoxide has also proposed to be involved in hepatocarcinogenesis (Massey et al., 1995; Mclean and Dutton, 1995; Jackson and Groopman, 1999). The p53 gene is a tumor suppressor gene that regulates cell growth. The biological functions of P53 are involved in control of cell cycle, DNA replication and repair, cell differentiation and apoptosis (Harris and Hollstein, 1993; Shen et al., 1996). Mutations in the p53 tumor suppressor gene were frequently detected in HCC patients (Greenblatt et al., 1994). For example, codon 249 in exon VII of p53 has been identified as a mutational hot-spot in HCC but not in other tumor types (Yang, et al., 1997; Jackson and Groopman, 1999). A point mutation in the third base of codon 249 of p53 results in a G—)T transversion (AGG—-> AGT) and a change in amino acid sequence from Arg to Ser. This mutation has been fi'equently associated with high levels of AF B1 exposure in HCC patients. The mutation of the p53 tumor suppressor gene at codon 247-250 (nucleotide 14067-14075) induced by AF B1 has been studied (Aguilar et al., 1993). AFB] was activated by rat liver rrricrosomes and mutagenesis at codons 247-250 of p53 gene was analyzed in human hepatocarcinoma cells, HepG2, by restriction fragment length polymorphism/polymerase chain reaction (RFLP/PCR). The results indicated that AF B1 induces the G—-)T transversion at the third position of codon 249 with relatively high frequency. However, cells exhibiting different bioactivation and detoxifying systems may influence the results. Mace et al. (1997) investigated AFB 1 -induced DNA adduct formation and p53 mutations in human liver cells expressing various cytochrome p450 isozymes. The results suggested that CYP1A2 is the most effective isoforrn in the activation of AF B1 to cytotoxic products and CYP3A4 for genotoxic metabolite formation. Cells containing CYP1A2 and CYP3A4 also showed a higher level of AF B1- DNA adduct formation and increased levels of p53 mutations (Mace, et al., 1997). A synergistic effect between AFB1 and hepatitis B virus (HBV). Exposure to AFB] alone has been verified as an important risk factor for development of human HCC. A high prevalence of hepatitis B virus (HBV) infection frequently occurred along with a high level of aflatoxin contamination in many regions of the world. HBV infection and AF B] exposure have been suggested to have synergistic effects on tumor formation as well as p53 gene mutation, (Lunn et al., 1997; Yang, et al., 1997). This was supported by a study using a transgenic mouse model that expresses hepatitis B surface Ag (HBsAg) and genetic inbred p53 null mice to determine the association of AF B] , HBV and p53 expression on liver tumor formation (Ghebranious and Sell, 1998). The results showed that AF B] induced a higher level of tumor formation than HBV infection alone. A synergistic effect between AFB] and HBV infection increased HCC. The highest risk group for liver cancer formation was generated with combinations of AF B] ingestion, HBV expression and p53 mutation. Epidemiological studies also suggest that high levels of AF B] exposure and HBV infection may have synergistic effects on HCC induction. Shen and Ong (1996) summarized the results from more than thirty reports from 1991 to 1995 with more than 1500 human HCC samples that were examined for p5 3 mutation with respect to different AFB] exposure levels. It was found that the frequency of codon 249 mutations in the p53 tumor suppressor gene directly parallels the environmental AFB] exposure level. Lasky and Mahden (1997) performed similar research by retrieving twenty studies that provided analytical data in p53 mutations in HCC patients. The analysis suggested that high levels of AF exposure (such as Southern Africa and China) were associated with the hot-spot mutation (coden 249) in p53 in a high proportion (approximately 50%) of HCC patients. On the contrary, codon 249 mutations were not detected in areas with low AF exposure such as Japan, Europe and the USA (Jackson and Groopman, 1999). AFLATOXIN BIOSYNTHESIS Aflatoxin gene cluster. Biosynthesis of AF in A. flavus and A. parasiticus from the starter unit, acetate, requires at least eighteen enzymatic activities (Woloshuck and Prieto, 1998; Payne and Brown, 1998; Minto and Townsend, 1998; Motomura et al., 1999; Cary et al., 2000; Bhatnagar et al., 2003). A. nidulans possesses a similar biosynthetic pathway to synthesize the secondary metabolite Sterigrnatocystin (ST), a precursor of AF, but this species lacks the final enzymatic activities (OmtA and OrdA) to convert ST to AF (Brown, et al., 1996). Many of the ST/AF genes have been identified and they were found clustered within a 75 kb chromosomal DNA region (Trail et al, 1995a; Woloshuck & Prieto, 1998; Payne & Brown, 1998; Bhatnagar et al., 2003)(Figure 1.2). These genes include aflR, a pathway specific regulatory factor of this gene cluster (Chang et al., 1993; Woloshuck et al, 1994), and many structural genes including pksA (Chang et al., 1995; Feng and Leonard, 1995),fas1,fas2 (Mahanti et al., 1996; Trail et al., 1995), nor-1 (Chang et al, 1992 ), avnA (Yu et al., 1997), avf-I (Prieto et al., 1996), estA (Yu et al., 2002), vbs (Silva et al., 1996), ver-I (Skory et al., 1992), omtA (Yu et al., 1993) and ordA (Yu et al., 1998). Expression of the AF genes appears to be co-regulated within the gene cluster. Most of the structural genes are activated under AF producing conditions and the timing of their expression is similar (Skory et al., 1993; Trail et al., 1995b; Brown et al., 1996). However at least one AF gene, estA (encodes an esterase activity that converts versiconal hemiacetal [VHA] to versiconal [VAL]), was recently reported to have a different expression pattern compared to other AF genes (Yu et al., 2002). Regulation of AF biosynthesis. An important positive regulator of the AF biosynthetic pathway is AflR. This GAL4-type transcription factor contains a Cy86Zn2- type DNA binding domain and is required for biosynthesis of AF (Chang et al., 1995b; Ehrlich et al.,1999a; Ehrlich, et al., 2003). Many of the structural genes involved in AF biosynthesis such as nor-I , ver—I, and omtA, as well as aflR itself, contain the Figure 1.2. Generally accepted pathway for sterigmatocystin and aflatoxin biosynthesis. (Adatpted from Bhatnagar et al., 2003) 10 1 Dehydrogenase "0’3 I 2 Monooxygenase cypA I 3 AF Transporter aflT I 4 Polyketide synthase pksA I I avnA I 5 Reductase —:Inor-1 15 Monooxgenase I A 16 Desaturase It verB _ ' L 6 Eggfhggigz fas-z 17 Dehydrogenase 3"“ 7 Fatty acid 1 f 1 18 Me transferase I omtB S nthase-1 35' y 19 Me transferase II omtA I 8 Transcription A aflFLI — Factor 20 Monooxgenase I ordA 9 Co-activator V aflJ - 21 Ver B synthase I Vbs 10 Dehydrogenase v adhA I I- "' 22 Monooxgenase I cpr I 11 Esterase II estA I _ 12 A I-alcohol " de ydrogenase ‘I norA I 23 Monooxgenase I 13 Dehydrogenase ver-1 24 Monooxgenase 14 Monooxgenase I “AI NADH oxidase Hexose transporter Glucosidase Transcrp. Factor “m Figure 1.2 palindromic consensus AflR binding site (5'-TCGN5CGR-3') in their promoter region and require AflR for expression (Ehrlich et al., 1999b). In addition to aflR, an adjacent gene aflJ possibly plays a role in regulation of AF biosynthesis. aflR and ale are divergently transcribed; both genes share part of the promoter regions (Meyer et al., 1998). Disruption of ale down-regulated AF production but the transcripts of other structural genes including pksA, nor-1, ver-I, and omtA were still present (Meyer et al., 1998). AflJ protein did not show any known sequence identity to other enzymes, yet its peptide sequence contains three putative membrane spanning domains and a peroxisomal targeting sequence, PTSl (Meyer et al., 1998). Thus, AflJ was suggested to be involved in transport of AF proteins to specific targets (Meyer et al., 1998). However, the actual function of AflJ has not been identified. It is not known why AF genes are clustered but the organization and coordination of expression of AF genes appears to be important (Trail et al., 1995b; Brown et al., 1996). In AF producing Aspergillus spp., the transcription factor AflR is an important positive regulator but AflR alone is not sufficient to regulate AF biosynthesis. Some Aspergillus spp., e.g. A. sojae and A. oryzae used in fermentation of foods, contain homologues of AF biosynthesis genes including aflR but they are not aflatoxigenic (Trail, 19953). A 28-amino acid zinc finger binding domain located at the C-terminus of AflR (CTSCASSK-VRCTKEKPACARCIERGLAC) was present in A. flavus, A. parasiticus, A. sojae, and some strains of A. oryzae (Chang et al., 1995b). aflR and other structural genes such as nor-1 and ver-I were present in some A. sojae and A. oryzae strains but they do not produce AF (Whatson et al., 1999). Transcripts of the aflR homologue in the non-AF producing A. oryzae were also not detected under AF inducing conditions by reverse-transcription polymerase chain reaction (RT-PCR) (Kusumoto et al., 1998). Absence of AF biosynthesis in koji mold (A. sojae) was later confirmed to be caused by defects in aflR expression, resulting in lack of expression of AF related genes (Matsushima et al., 2001). In addition, some mutant strains contain an incomplete gene 12 cluster affecting gene expression and ST/AF production, and implicating a loss of functional regulation in these species (Cary, et al., 1999; Butchko, et al., 1999). These data also suggested that in addition to AF LR, other mechanisms appear to regulate biosynthesis of AF. The cluster effects for expression of AF genes play an important role in AF biosynthesis. Significance of AF genes organized in a cluster was investigated as one major objective in my study (using nor-1 gene as an example, Chapter 3). The results confirmed that the AF gene cluster exerts a positive position-dependent regulatory influence on the nor-1 promoter (Chiou et al., 2002). Partially-duplicated AF gene cluster. A second, partially-duplicated AF cluster was discovered in some strains of A. parasiticus (Liang et al., 1996; Chang and Yu, 2002). This partial AF gene cluster contains homologs of aflR-aflJ-adhA—estA -norA-ver1 plus omtB, yet is missing verA-avnA -verB-ava, omtA, ordA, and vbs present in the complete AF cluster in A. parasiticus strain 56775 (F ig. 1.2) (Chang and Yu, 2002). The truncated AF gene cluster contained several mutations and did not seem to be functional. Several structural genes were found to be defective, including pre-termination of verIB (Liang et al., 1996) and norA2, point mutations in adhAZ, and a large deletion in omtB (Chang and Yu, 2002). Mutations were also found in the second copy aflR (aflRZ), and a specific point mutation in nucleotide residue 896 (with a T to A substitution) can be used to discriminate between aflRI and aflRZ (Cary et al., 2002). However, this mutation did not result in a pre-terminated gene product. AflR2 was not able to activate transcription in a yeast two hybrid system (Cary et al., 2002), and it could not complement the loss of functional activity of the primary AflRl (Chang and Yu, 2002). An A. parasiticus aflR knockout mutant AF 810, was disrupted in aflRI but not aflRZ. This strain did not produce AF, any precursors in the pathway, or transcripts of aflRZ and other AF genes such as nor-I , ver-I , and omtA analyzed by Northern hybridization analysis. However, 13 using RT-PCR, the transcripts of aflRZ and these AF genes were detected at low levels (Cary et al., 2002). The aflRI knockout strain AF S10 was used as a negative control in my AF protein localization studies (Chapter 4 and 5). Sugar-utilization gene cluster. Downstream of the AF gene cluster following a 5 kb spacer region, a sugar utilization gene cluster consisting of 4 genes (sugR, hxtA, gch, and nadA) was identified (Yu et al., 2000). Expression of hxtA, which encodes a hexose-transport protein, followed the same timing of expression under AF inducing conditions (glucose minimum salt medium) as AF genes such as omtA (Yu et al., 2002). This sugar utilization cluster, or the hexose-transporter thA specifically, may play a role in regulation of AF biosynthesis, since simple sugars, e.g. glucose, were found to induce AF biosynthesis (Buchanan and Lewis, 1984; Skory et al. 1993; Miller, 2003). SECONDARY METABOLISM AND FUNGAL DEVELOPMENT Association between secondary metabolism and fungal development. In Aspergillus as well as many other microorganisms, the production of secondary metabolites is frequently associated with developmental processes, e.g., sporulation (Bennett, 1983; Horinouchi et al., 1994, Guzman-De-Pena and Ruiz-Herrera, 1997; Adam and Yu, 1998; Adams et al., 1998; Calvo et al., 2002). The hypothesis is best supported by the gram-positive bacterial genus Streptomyces in which morphological differentiation resembles development of filamentous fungi. In Streptomyces griseus, secondary metabolism and formation of aerial mycelium, a structure required for subsequent generation of spores, are both triggered by a 'y-butyrolactone containing molecule, A-factor (Horinouchi and Beppu, 1994; Kato et al., 2002). G-Protein signaling pathway. In A. nidulans, coordinated regulation of asexual sporulation and biosynthesis of the mycotoxin sterigmatocystin (ST) has been demonstrated (Adams and Yu, 1998; Adams et al., 1998). The balance between vegetative growth, sporulation, and ST production is regulated by two antagonistic 14 factors, F adA (Fluffy Autolytic Dominant), and F le (Fluffy, Low _b_rlA; brlA is a conidial vesicle specific gene) involved in a G-protein signaling pathway (Yu et al., 1996; Hicks et al, 1997)(Figure 1.3). The fadA gene encodes the Ga subunit of a heterotrimeric G protein and activation of this gene facilitates cell proliferation and vegetative growth, while flbA encodes a RGS (Regulator of G protein signaling) domain that antagonizes the function of F adA. Inactivation of the F adA G protein signaling pathway is required for both ST biosynthesis and asexual sporulation (Hicks et al, 1997). Conversely, developmental activation requires at least partial inactivation of fadA. Dominant activating mutations of fadA as well as loss of function mutations offlbA resulted in inhibition of both conidiation and ST production. Conversely, overexpression offlbA or a dominant interfereing mutation of fadA leads to precocious ST gene expression and ST accumulation, as well as unscheduled conidiation (Hicks et al, 1997). FluG. Another factor F luG, acting up-stream of this G-protein linked pathway, is associated with synthesis of an extracellular, low molecular weight diffusible factor leading to asexual conidiation and ST production (Lee and Adams, 1994). Loss of function mutations in fluG resulted in a complete block in ST production and sporulation, similar phenotypes to loss of function mutations in brlA and dominant activation mutation in fadA (Hicks, et al., 1997). The FluG-dependent asexual sporulation pathway seems to be in competition with the vegetative growth pathway promoted by FadA. Whereas F adA negatively regulates ST biosynthesis, F luG is believed to promote ST production via inactivation of FadA signaling (D'Souza et al., 2001). The detailed mechanisms of this G-protein signaling pathway observed in Aspergillus are under investigation. AF biosynthesis and conidiation. Secondary metabolites and conidiation are usually observed when cells enter the stationary phase of grth (Betina, 1995). In A. flavus and A. parasiticus, AF was detected in developmental structures including conidia 15 Carbon Stress Nitrogen Stress \ .- .1.“ s . ,1-“7‘ . . ,g \ .. _. _- T fle- IlbD- fIbB\ 2" Receptor? brIA -—> Development / 0+5» .. .L Extracellular T""';’I"" """"""""" ' FluG signal ' ’ . “1:le ; D —. *zzFadi-GTP.” Colony proliferation and Aytolysls .v Growthslgnal GDP ? ———->afIR—> ST production . J k. A , Figure 1.3. The G-protein signaling pathway involved in ST biosynthesis in A. nidulans. (Adapted from Adams et al., 1998) 16 and sclerotia (Wicklow and Schotwell, 1983). A. parasiticus with a defect in fluP, a gene that encodes a large polyfunctional enzyme with sequence similarity to polyketide synthetases, did not develop conidiospores and AF production was reduced 3 to 5 fold compared to the wild type (Zhou et al., 1999). Production of ST/AF in Aspergillus spp. typically accompanies conidiation but the converse is not always true because ST/AF can be synthesized in liquid and submerged shake culture where conidiation does not usually occur (Kale et al. 1994). However, Guzman-de-Pena et al. (1998) indicated a lower level production of AF in liquid culture compared to AF production fi'om the same mass of A. parasiticus grown on solid media with conidiation. Also, chemicals that suppress conidiation such as fluoroacetic acid (ReiB, 1982) and diaminbutanone, an ornithine decarboxylase inhibitor (Guzman-de-Pena et al., 1998) reduced production of AF in A. parasiticus. LOCALIZATION OF PROTEINS INVOLVED IN AF BIOSYNTHESIS One of the goals of my study was to determine where AF proteins are synthesized within a cell and to understand how these enzymatic activities associate with each other in order to make the final AF molecules. Penicillin biosynthesis as an example. In Penicillium chrysogenum, biosynthesis of the secondary metabolite penicillin is similar to ST/AF biosynthesis in Aspergillus because penicillin biosynthetic genes are also organized in a cluster. Three gene products, ACVS (aminoadipyl-cysteinyl-valine tripeptide synthetase), IPNS (isopenicillin N-synthetase), and AT (acyl CoA: 6-iso-penicillin-N-acyltransferase) involved in penicillin biosynthesis have been isolated and localized (Muller et al., 1991). ACVS was thought to be associated with a Golgi-like organelle (Kurylowicz et al., 1987) and vacuoles (Lenderfeld et al., 1993), but later was confirmed to be localized in the cytoplasm (Van der Lende et al., 2002). The second enzyme IPNS was also found to be a cytosolic protein (Van der Kamp et al., 1999; Van der Lende et al., 2002). By irnmuno- 17 transmission electron microscopy, only the final enzyme activity, AT, was found associated with microbodies (aka. peroxisomes) (Muller et al., 1995). Organelles possibly involved in AF biosynthesis. This section reviews several organelles that may be associated with AF biosynthesis in A. parasiticus. l) Microbodies. Microbodies (aka. peroxisomes) are found universally in fungi, but the number, shape and function are highly variable (Markham, 1995). Even within a species, generation of microbodies may be affected by various factors including age, developmental stage, and nutrient source. Microbodies contain a variety of enzyme activities and are potentially involved in a wide range of degradative and biosynthetic activities (Carson and Cooney, 1990). There are two types of microbodies with different enzyme activities. The peroxisome contains catalases and enzymes involved in fatty acid B-oxidation, while the glyoxysome houses the glyoxylate cycle and also enzymes involved in fatty acid B-oxidation (Keller et al., 1991). Microbodies are common in fungal spores, where they are associated with lipid inclusions (Maxewell et al., 1977). Two types of Beroxisomal Targeting Sequence (PTS) have been identified. PTSl is a carboxyl-terminal tri-peptide sequence Ser—Lys-Leu (SKL) and conserved variants (S/A/(CXK/R/HXL/MXGould et al., 1987). The second matrix protein targeting signal, PTSZ, is an N-terminal nonapeptide motif with the consensus sequence (R/K)(LN/I)X5- (H/Q)(L/A)(Subramani, 1996). However, many peroxisomal enzymes have neither a PTSl nor a PTSZ motif (Waterham and Cregg, 1999). None of the AF proteins contains the conserved PTSl or PTSZ located at the C-terminus or N-terminus, respectively. However, the PTSl sequence was detected in AflJ, OmtA and OrdA, but not at the C- terrninus (predicted by PSORT version 6.4). 2) Woronin bodies. Woronin bodies are highly refractile particles seen in over 50 species of filamentous fimgi (Markam and Collinge, 1987; Jedd and Chua, 2000). This organelle is bounded by a single bilayer membrane and the major component is protein (Head, 1989). The Woronin body has a spherical to oval shape and the size is 18 approximately 0.1-0.75 pm, slightly larger than the septa] pore (Markham, 1995; Markham and Collinge, 1987; Jedd and Chua, 2000). Woronin bodies are usually located in the septa and have been suggested to function by plugging the damage within the septa] pore in response to hyphal injury (Colling and Markham, 1985; Tenney et al.,- 2000). This function indicates that the Woronin body is capable of moving within hyphal cells, implying a possible role in transport of AF. The phytotoxin prehelminthosporol has been found localized in Woronin bodies in Bioplaris sorokiniana (Akesson, et al., 1996). The Woronin body was suggested to be derived fiom peroxisomes in A. nidulans (V alenciano et al., 1998). A gene encoding a major Woronin body protein, hex], was recently identified in Neurospora crassa (Tenney et al., 2000; Jedd and Chua, 2000). The Hexl protein contains the C-terminal PTSI sequence, suggesting a relationship between the woronin body and the peroxisome. In Neurospora crassa , Woronin bodies are associated with catalase-negative peroxisomes (J edd and Chua, 2000). 3) Vesicles. Vesicles are small but they are important and very abundant in filamentous fimgi (Markam, 1995). Vesicle sizes vary with diameters ranging from approximately 30 nm to 400 nm (Markam, 1995). A high density of vesicles is usually clustered in the hyphal apex; they are involved in transport of various structural proteins, enzymes, and other cellular components including lipids and polysaccharides (Markam, 1995). The vesicles are able to migrate through the septal pore and the movement is associated with microtubule components (Hoch et al, 1987). There are many types of vesicles such as wall vesicles, microvesicles, and chitosomes, with respect to their particular functions. 4) Vacuoles. Vacuoles are the largest organelles in eukaryotic microorganisms. The large vacuoles of fungi have many features similar to animal lysosomes. Fungal vacuoles, similar to higher plant vacuoles, are acidic compartments containing hydrolases and lysosomal activities (Klinlosky, 1997; Ashford, 1998); the acidic environment is 19 regulated by a vacuolar ATPase (Wada and Anraku, 1994). The functions of vacuoles are associated with intracellular homeostasis, storage of metabolites, and degradation of proteins (Thrumm, 2000). This organelle is also involved in intracellular protein transport such as endocytosis, protein secretion, and intracellular protein localization (Wendland et al., 1998). The structure and dynamic activities of fungal vacuoles indicate that this organelle has the capacity to interact and transport material over long hyphal distances (Ashfor, 1998). Under the microscope, the vacuole appears as a large, hemi-spherical hollow-like, fluid -filled organelle surrounded by a single membrane (Markham, 1995; Ohsumi et al., 2001; Ohneda et al., 2002). Vacuoles are often observed in aged cells and they are not usually present in the active apical tip regions (Markham, 1995). Some large fungal vacuoles may contain small membrane-bound vesicles, which are derived from fusion of multivesicular bodies with large vacuoles (Hoch and Staples, 1983). Over 40 vacuolar protein-sorting (vps) and vacuolar-morphology (vam) associated genes have been identified in Saccharomyces cerevisiae (Raymond et al., 1992). One vacuole resident protein, carboxypeptidase Y (CPY), has been used as a vacuole marker protein (Johnson et al., 1987). In A. nidulans, cypA that encodes CPY (Ohsumi et al., 2001), and vpsA involved in vacuolar biogenesis (Tarutani et al., 2001) have been cloned and characterized. Klinonsky (1998) reviewed vacuole-related functions and suggested that of all the organelles, the vacuole may be the most adaptive compartment as a delivery mechanism. In fed-batch fermentation of P. chrysogenum, during the switch fiom the rapid growth to the production phase as nutrients became limited, a high degree of mycelial fragmentation was seen along with heavy vacuolation (Paul et al., 1994), implicating an association of vacuole biogenesis and secondary metabolism. In addition, immunogold-transmission electron microscopy localized OmtA, an enzyme involved in a late stage AF biosynthesis, in the cytoplasm as well as in the vacuole in cells located near the substrate surface of the 20 colony (Lee et al., manuscript submitted). In this regard, the vacuole is hypothesized to play a role in AF biosynthesis. 5) Endoplasmic reticulum (ER). The ER is a continuous membrane system, yet various domains are present in this single organelle that perform different functions (Markhem, 1995; Voeltz et al., 2003). The ER appears to make physical contact with the membranes of many organelles such as nuclei (nuclear envelopes) and vacuoles, and dynamic contact with cytoplasmic vesicles via the continuous traffic system (Beckette et al., 1974; Wedlich-SOlender et al., 2002). This organelle is involved in protein folding and several post-translational modifications in the ER lumen, and translocation of proteins (e.g. secretory proteins) across the ER membrane (V oletz et al., 2003; Ellgaard and Helenius, 2003). N-glycosylation of proteins along the secretory pathway is an important protein quality control mechanism. This proof reading system serves beyond the levels of transcription and translation to ensure native conformation of newly synthesized proteins and the fidelity of cellular functions (Ellgaard and Helenius, 2003; Roth, 2002; Helenius and Aebi, 2001; Parodi, 2000). After proteins reach their native tertiary and quaternary structure, they subsequently translocate to their final cellular destination. In general, the proteins involved in the secretory pathway are targeted to the plasma membrane, vesicles, or extracellar space for secretion (Roth, 2002; Ellgaard and Helenius, 2003). As part of the ER quality control, misfolded proteins are translocated fiom the ER to the cytoplasm and subjected to an ER-associated degradation (ERAD) process and subsequently destroyed by the 26S proteosome (Kostova and Wolf, 2003). According to protein sub-cellular localization prediction programs listed in the Expasy molecular server (http://us.expasy.org), proteins targeted to the ER usually require an N-terminal hydrophobic signal peptide and N-glycosylation site(s) (Asn-X-Ser/Thr) (Signal IP; Target P; NetGlyc 1.0; Ausubel et al., 2003). BR resident proteins firrther require an ER-retention signal, KDEL, within the C-terminal peptide sequence (Teasdale 21 and Jackson, 1996; Harter and Wieland, 1996). Among AF proteins, only VBS was confirmed to be N-glycosylated (Silva et al., 1996) and appeared to be modified in the ER (see Chapter 5). Based on protein sequence prediction, only one other protein, OrdA involved in the final step of AF biosynthesis, contains a putative N-terminal signal peptide and one N-glycosylation site, suggesting OrdA may be associated with the ER secretory pathway like VBS (ExPasy molecular server). Both proteins were not predicted to be ER resident proteins. Association of AF proteins along the ER secretory pathway may be significant to their protein functions, and possibly play a role in regulation of AF biosynthesis. CONFOCAL LASER SCANNING MICROSCOPY (CLSM) IN THE STUDY OF FUNGAL BIOLOGY Introduction to CLSM. Microscopy serves as an useful tool to conduct studies in cell biology including analyses of cell dynamics, morphogenesis, cytochemistry, fungal spore structure and development (Spear et al., 1999). Fungal cellular organelles can also be visualized; however, the resolution of light microscopy has limitations (theoretical maximum resolution is 0.2 pm) in traditional light microscopes (Paddock, 2000). In conjunction with fluorescence-based techniques such as immuno-fluorescence labeling as well as the recent application of the green fluorescent protein (OF P), fluorescence tagged cellular targets can be tracked by fluorescence microscopy. In the past, conventional epi- fluorescence microscopy has been widely used. Confocal laser scanning microscopy (CLSM) provides a wide range of advantages over traditional light microscopy (and epi- fluorescence microscopy) although CLSM is based on the fundamental principles used to light microscopes. The first stage-scanning confocal microscope was invented in mid-1957 by Minsky for imaging of neural networks in unstained brain cells (Minsky, 1961; 1988). Modern confocal microscopes are based on the configuration of this prototype (with the 22 use of a pinhole) and an enhanced model became commercially available in 1987 after air-cooled lasers and PC computers and software became prevalent (Amos et al., 1987). The characteristic feature of CLSM is the ability to generate "confocal images" that eliminate the interference of out-of-focus light captured by the detector. This confocal effect is achieved by use of a confocal pinhole positioned between the specimen and the detector in the light path (Figure 1.4). CLSM utilizes one or more focused beams of light illuminated by the laser (light amplification by stimulated gmission of radiation) which scans across the sample generating "optical sections" at various focal planes. Thus, contrast and resolution are enhanced on each single optical section. This feature is particularly beneficial for viewing thick samples since CSLM is capable of scanning through specimens (up to hundreds of microns in thickness) instead of mechanical sectioning of samples and acquiring serial optical sections (Z-sections) (Whallon, 1998). Image quality enhancement by CLSM. Since CLSM utilizes the basic configuration of the light microscope, the instrument resolution per se, that is influenced by the wavelength of light, the objective lens, and properties of the specimen itself, is not improved drastically (Paddock, 2001). However, the final image resolution of modern CLSM is improved by the following factors. First, stable multiple wavelength lasers instead of conventional tungsten-halogen and/or mercury lamps are used to generate bright point sources of light to scan the samples. Typically air-cooled lasers are used including the helium-neon laser emitting at 543 nm and 633 nm and the argon-ion laser emitting at 488 nm and 514 nm. The laser spot size affects instrument resolution; similar to scanning electron microscopy, the smaller the beam spot, the greater the resolution (Czymmeck et al., 1994). The spot size is also affected by the wavelength of light and nmneric aperture (NA) of the objective lens (Dresolution= 0.61X [Wavelength]/ NAobjective)- For example, the spot size for the 488 nm laser with a high NA (1.4) lens is approximately 250 nm (Czymmeck et al., 1994). In addition, a laser focused to a small 23 Beam Splitter L I aser CollectingLens p To A M Mom... T p | . /. Objective Lens Confocal Unit With Pinhole Focal Plane Outof-Focus Area —Path 01 ln‘Focus Light Figure 1.4. Diagram of the light path in a confocal instrument. (Adapted from Czymmek et al., 1994) 24 point (as equipped in the CLSM models; Zeiss 210 and Zeiss LSM5 PASCAL, Carl-Zeiss Inc., Germany) generates better resolution than a laser focused through a slit (as equipped in Meridian Insight®, Meridian Instrument, Inc., Okemos, Mich.). Theoretical resolution of CLSM under 488 nm excitation with a 1.4 NA oil objective was estimated to be 160 nm (Brelje et al., 1993) and approximately 130 run under ideal conditions (Carlsson and Aslund, 1987). Second, image detection in modern confocal microscopes utilizes sensitive and low noise detectors (e.g., photomutiplier tube (PMT) or a digitized charge-coupled device (CCD) camera, particularly the cooled-charged CCD with reduced noise read out. In addition, advanced CLSM is equipped with fast microcomputers with image processing capability and elegant software to analyze images. High quality output images can be processed directly in a computer imaging system or recorded on hard copy devices (Poddock, 2001). Advanced image analysis programs also allow CLSM digitized images to be processed further, for example, reconstruction of three-dimensional (3-D) images from Z-stacks (Czymmek et al., 1994), or pixel analysis to quantify fluorescence intensity (Spear et al., 1999; Chapter 5 in this dissertation). CLSM imaging. CLSM is capable of generating many types of microscopic images. General light microscopy applications, such as phase contrast, polarization, and differential interference contrast (DIC), can be conducted via CLSM with enhanced image qualities. However, an important feature of CLSM is detection of fluorescence. The characteristic functions are described as follows. 1) Optical sections. The basic unit of confocal microscopy is the single optical section (confocal image). The contrast of an optical section is enhanced more intensely than non-confocal images generated by epi-fluorescence microscopy. 2) Z-series. For one specific field of interest, several optical sections can be collected as a series to generate a-Z-stack. The CLSM is commonly equipped with a stepping motor to control the movement of the stage with a pre-determined distance (step 25 size). Using a macro program, CLSM can acquire images from the top (or from the bottom) of a specimen, scan through the sample with a pre-set step size, and acquire as well as store the whole series of optical sections on the hard drive to generate a Z-series (Z-stack). The Z-series can be further processed into a 3-D presentation, stereo images, or viewed as a movie depending on the program capabilities. A representative Z-series of A. parasiticus SU-l condiophores stained with SYTOX green nuclear staining dye (Molecular Probes, Eugene, OR) is shown in Fig. 1.5. 3) Phi-z sectioning. Optical sectioning is not limited to the horizontal focal plane but also can be collected in the vertical plane to generate a phi-z or x-z section. Phi-z sectioning is accomplished by scanning the laser across a defined single line. It allows signals to be detected fi'om another angle of view. For example, the microbodies of A. parasiticus SU-l immuno-labeled with anti-SKL antibodies followed by secondary antibodies conjugated with Alexa 488 probe are displayed in a single optical section (Fig. 1.6, A). Analysis by phi-z scanning indicated that several rrricrobodies are located inside the hyphae (Fig. 1.6, B). 4) Time-lapse and live imaging. Since CLSM is capable of scanning through the sample without mechanical sectioning, live cells may be studied. This application may be somewhat difficult to accomplish, for example, if the experiment requires fixation of the cell, or if the target signals are not stable under illumination by the laser. In addition, the position of the specimen has to be maintained under the microscope. Nevertheless, time- lapse imaging allows visualization of dynamic events at a rate close to realtime or video- rate imaging (Whallon, 1998). 5) Fluorescence imaging. CLSM is a powerful tool for fluorescence imaging. The argon-ion laser emits at 488 nm and generates optimum excitation spectra for many commonly used fluorescence probes, including FITC (fluorescein isothiocyanate (Abs/Em 490/523 nm), Alexa 488 (Abs/Em 495/519 nm), the red-shift green fluorescent protein 26 Figure 1.5. CLSM Z-series of A. parasiticus SU-l conidiophores stained with SYTOX green fluorescence dye. Fungal thin sections (4 pm) derived from colony fraction 82 (24- 48 h growth) of a 72 h-old SU-l colony grown on YES agar were stained with SYTOX green dye (Molecular Probes). (l) - (10): Ten optical sections with step size (2 section interval) 750 nm. (A) An extended focus image of this CLSM Z-stack. (B) A bright field image in the middle of the Z-stack. Bars, 25 um. 27 Phi-z scan _> Figure 1.6. CLSM single optical section and phi-z scan of A. parasiticus SU-l microbodies. Fungal thin sections (4 pm) derived from colony fraction 32 (24-48 h growth) of a 72 h-old SU-l colony grown on YES agar were immuno-labeled with anti- SKL antibodies (Zymed) followed by secondary antibody conjugated with Alexa 488 probe (Molecular Probes). CLSM was performed using a Zeiss 210 confocal scanning microscope (Carl Zeiss Inc.). (A) Single optical section. (B) Phi-z scan. The arrows indicate the location of microbodies inside the cell. Bar, 5 pm. 28 variant, EGF P (A bs/Em 488/507 nm), and a nuclear staining dye, SYTOX green (Abs/Em 504/523 nm). CLSM can also perform multiple fluorescence detection, for example, dual channel detection of green and red fluorescence on the same image (Chapter 5, Fig. 5.9). This is particularly useful if these fluorophores have minimum overlapping emission spectra to avoid cross-talking signals. Mixed krypton-argon gas lasers are popular for multiple-wavelength confocal microscopy because they emit at three well-separated wavelengths (488 and 568 and 647 nm) allowing detection of green to far-red fluorophores (e.g. F ITC; Lissarnine rhodarnine B, Abs/Em 570/590 nm; Cy5, Abs/Em 649/666 nm)(Ausubel et al., 2003). However, some fluorophores excite at shorter waves (e.g., DAPI [4,'6-diamidino-2-phenylindale, Abs/Em 458/461 nm] nuclear stain, and CMAC [7-amino-4-chloromethylcoumarin, Abs/Em 353/466 nm] vacuolar stain); detection of these probes requires CLSM equipped with expensive UV lasers. Photobleaching may cause significant problems for fluorescence detection, particularly when samples are illuminated with high intensity lasers. It is a quenching phenomenon that causes fading of fluorescence signals, yet the mechanism is not fully understood. One important class of bleaching is photodynarnics and this involves interaction of fluorophores with light and oxygen. The amount of photodamage depends on the concentration of molecular oxygen and the distance between the fluorophore, molecular oxygen and other cellular constituents (Herman, 1998). Protection against photobleaching can be achieved by reducing excitation energy or exposure time; however, this will also reduce the yield of the signal. Other methods to prevent photobleaching include deoxygenation of solutions; use of singlet oxygen quenchers such as histidine, diphenylisobenzofuran, or crocetin (a water soluble acratenoid); or antifade reagents such as n-propyl-gallate. Several antifade mounting media are also commercially available (e.g. Vectashield [Vector] and Prolong antifade mounting media [Molecular Probes]). 29 In summary, CLSM is capable of generating higher contrast and resolution images compared to traditional light microscopy by reducing background interference significantly. Advanced image detection systems and analysis programs render this instrument more capable of analyzing cellular structures or other aspects in detail. In this dissertation, CLSM is an important tool to study localization of proteins involved in AF biosynthesis in the filamentous fungus A. parasiticus (Chapter 4 and 5). 30 CHAPTERZ A rapid DNA extraction method for screening large numbers of Aspergillus transformants via polymerase chain reaction ABSTRACT The efficiency of screening fungal transformants using available methods, particularly in large quantity, has been a problem in fungal molecular biology. A rapid DNA extraction method was developed for screening Aspergillus transformants via polymerase chain reaction (PCR). Fungal cells were collected directly from colonies on a culture plate into microcentrifuge tubes containing TE buffer (10 mM Tris-HCl, 1.0 mM EDTA, pH 8.0). The cell suspension was boiled for five minutes to release nucleic acids. After a brief centrifugation step (12,000 g for 5 min), fungal cell extracts (supernatant) were obtained. PCR of diluted fungal extracts with appropriate primers resulted in amplicons up to 2.5 kb with Taq DNA polymerase and up to 3.5 kb with PfiiTurbo DNA polymerase. The procedure was successfully utilized for amplification of native fungal genes as well as identification of the integration site of plasmid sequences after homologous recombination. INTRODUCTION Fungal transformation provides an important tool for the study of fungal biology. However, limited numbers of transformants can be screened conveniently due to labor- intensive assay procedures. The presence of the fungal cell wall limits the efficiency of genomic DNA isolation. Disruption of the fungal cell wall to release nucleic acids is usually accomplished by the laborious procedure of grinding under liquid nitrogen. DNA purification also commonly includes proteolytic enzyme digestion, phenol/chloroform extraction, and precipitation of DNA (Ausubel et al., 2003; Borgia et al., 1994). The quantity and quality of the genomic DNA isolated by these procedures tends to vary 31 significantly. A high content of polysaccharides and nucleases represent significant obstacles for obtaining high quality fungal genomic DNA. Hence, my goal was to develop a rapid, inexpensive, and efficient DNA extraction method for screening large numbers of fungal transformants. I present a simple procedure to obtain crude DNA templates directly from fungal colonies that was successfully used to amplify endogenous fungal genes as well as to detect the integration site of plasmid sequences afier homologous recombination. Twenty-four fungal colonies can be easily assayed in a three to four hour period including time for PCR and electrophoretic analysis. MATERIAL AND METHODS Aspergillus parasiticus CS10 (wh-l , ver-l pyrG) (Skory et al., 1990) was transformed with plasmid pAPGUSNP (Fig. 2.1, A) and the transformants were selected on Czapek-Dox medium (CZ medium, Difco) for ability to synthesize uridine monophosphate. As soon as the prototrophic transformants sporulated (approximately 2 days), spores were transferred onto duplicate CZ plates with a toothpick. After three days (or as soon as colonies could be observed), colonies were analyzed. A sterile 200 pl plastic pipette tip (Lee and Cooper, 1995) was used to transfer a portion of the fungal colony (including conidiospores and mycelia) into a microcentrifuge tube containing 100 pl TE buffer (10 mM Tris-HCl, 1.0 mM EDTA, pH 8.0). This process was repeated up to three times to ensure that sufficient fungal cells were transferred. Tubes were heated in a boiling water bath for 5 min and placed on ice. The samples were centrifuged at 12,000 g for 5 min at 40C. The supernatant (firngal extract) containing crude DNA templates was transferred into a new microcentrifuge tube. The extract was further diluted with TE buffer (undiluted, 1/5, 1/20, 1/ 100). In most assays, 2.5 pl of the 1:5 dilution resulted in good amplification in a 50 pl PCR reaction. The remainder of the fungal extract was 32 stored at -80°C and could be successfully analyzed by PCR for at least three months. Diluted samples stored at 4°C could be analyzed by PCR for at least one week. PCR analysis was conducted on DNA extracted from A. parasiticus transformants and the recipient strain, CS10. Four primer pairs were used in the assay (Table 2.1). Primer pair J L136 and JL137 produced a 1.4 kb exon H fiagment from the fungal gene encoding versicolorin B synthase (vbs) (Silva et al., 1996). Three primer pairs, JL99 and JL100, JL102 and JL103, JL221 and JL222 were designed to determine the integration site of pAPGUSNPyrG in the 5' region of nor-1, in the 3' region of nor-I or at the pyrG locus repectively. One primer from each pair targeted a sequence located within the fungal genome. The other primer targeted one site within the plasmid sequence (Figure 2.1). A 50 pl PCR reaction mixture consisted of 2.5 pl of the diluted fungal extract, 20 mM Tris-HCl pH 8.8, 10 mM KC], 10 mM (NH4)2SO4, 0.1% Triton X-100, 0.1 mg/ml BSA, 0.5 - 1.0 pM of each primer, 200 pM each dNTP, 2.0 - 3.0 mM MgC12 and either 2.5 units of Taq thermal stable DNA polymerase (5 U/pl, Gibco BRL) or 2.5 units of PfuTurbo DNA polymerase (2.5 U/ pl, Stratagene). PCR cycling parameters were adjusted depending on primer design and the target size (Table 2.1). PCR was performed in a Perken-Elmer GeneAmpli System 9600 thermal cycler. All PCR reactions were initialized by a denaturation step at 95°C for 3 min followed by 35 cycles of denaturation, annealing and extension. Reactions were terminated with an extension reaction at 72°C for 10 min. PCR products (10 p] of reaction) were observed by electrophoresis on a 1% agarose gel using standard procedures (Ausubel et al., 2003). RESULTS AND DISCUSSION DNA was extracted by the boiling procedure from seven representative transformants and the recipient strain, CS10 (Figure 2.2). The DNA was analyzed by PCR with appropriate primer pairs. A 1.4 kb fragment was generated from all eight 33 Table 2.1. Primer sequences and PCR cycling parameters used for analysis of fungal genomic DNA templates. Primer Target Primer Sequence PCR cycling parameters Endogenous vbs ExonII JL136 1.4 kb 5' AGG crc GAA AGG CGC ATA CGA 3' JL137 5' GT0 GTC TAC TGC CCA Gcc ATC 3' Integration of pAPGUSNP, 3' region of nor-l gene JL102 2.1 kb 5' CGC AAG GTG AGG GIT CGA ACC GAG G 3' JL103 5' CCG CAG CAG GGA GGc AAA CAA TGA A 3' Integration of pAPGUSNP, 5' region of nor-1 gene JL 99 3.2 kb 5'TITCACGGGTI‘GGGGTI‘TCTACAGG3' JL100 5'GACGGGGAACCTCTI‘TACAAACATC3' Integration of pAPGUSNP at pyrG JL221 3.5 kb 5' CCA GAA AAr GCG AAG GTA AGT GCT TC3' JL222 5' GCG ACA CGG AAA TGT TGA ATA crc AT3' 95°C 1', 66°C 1', 72°C 2' 95°C 1', 65°C 1', 72°C 3' 95°C 1', 62°C 1', 72°C 5' 95°C 1', 55°C 1', 72°C 5' 34 Figure 2.1. Homologus recombination of pAPGUSNP into the A. parasiticus chromosome and primers used in integration analysis. (A) Plasmid map of pAPGUSNP. (B) pAPGUSNP integrated within the 5' region of the nor-1 gene generates a 3.2 kb PCR fiagment using primers J L99 and J L100. (C) pAPGUSNP integrated within the 3' region of the nor-1 gene generates a 2.1 kb PCR fragment using primers JL102 and JL103. (D) pAPGUSNP integrated at the pyrG locus generates a 3.5 kb PCR fiagment using primers JL221 and JL222. (E) Amplification of endogenous vbs gene fi'agment in A. parasiticus using primers JL136 and JL137. 35 Plasmid Xhol 10600 al 2100 Sac19600 pA PGUSNP Sca|2900 12100 bp Scal8300 3' nor-1 Chl 5000 C|a|5900 Sphl 5200 Genome D B C X 5' X X 3' I: genomic pyrG genomb nor-1 vbs E -1 JL13 5'nor-1 genome no; genogne 6'>‘.JL137 W~~”~— _ _ _ . 1.4 kb JL100> ‘JL99 3.2 kb nor-1 genome C 5' 3' _ _ _ _ s\\\\\‘ vbs 1.4 kb JL103WL102 2.1kb pyrG genome Mx//zxx/xw K\\\\V vbs 1.4 kb JL221, ‘JL222 3.5 kb Figure 2.1 36 1081012345672. ACS101234567¢X174 3.5kb —> 2.1kb —> Figure 2.2. Amplification of PCR targets using the rapid DNA extraction technique. The assay was performed on eight A. parasiticus colonies, including the host strain, CS10, and seven representive transformants. (A) Integration analysis using various primer sets. Cell extracts of C S10, and transformants 100, 149, 164 were PCR amplifed with primers JL221 and JL222. A 3.5 kb target fragment was produced in transformants 100, 149, 164 Canes 1-3) but not the host strain CS10. PCR of cell extracts from 57 and 104 (Lanes 4- 5) with primers JL102 and JL103 produced a 2.1 kb fragment. PCR of cell extracts from transformants 17 and 24 with primers JL99 and JL100 generated a 3.2 kb fragment (Lanes 6-7). All DNA extracts produced the endogenous 1.4 kb vbs fragment with primers JL136 and JL137 (shown in B). 37 strains fi'om the endogenous vbs gene (Figure 2.2, B). Using the same fungal extracts, the three sets of primer pairs for integration analysis successfully determined the integration sites of pAPGUSNP within the genome in the transformants (Figure 2.2, A). With primer pair JL221 and JL222, PCR amplification of cell extracts from transformants 100, 149 and 164 (Lanes 1-3) produced a 3.5 kb PCR product, indicating that the plasmid integrated at the chromosomal pyrG locus while PCR amplification of cell extracts from the host strain CS10 did not generate this fiagment. PCR amplification of cell extracts fiom transfonnants 57 and 104 (Lanes 4 and 5) produced a 2.1 kb fiagment with primer pair JL102 and JL103, indicating integration of pAPGUSNP within the 3' region of the nor-I gene while cell extracts of transformants 17 and 24 (Lanes 6 and 7) produced a 3.2 kb fiagment indicating integration within the 5' region of the nor-1 gene. The integration location of pAPGUSNP in each strain was confirmed by Southern hybridization analysis (data shown in Chapter 3). After demonstrating the utility of this DNA extraction procedure, over 500 transformants were screened. The assay correctly identified the site of integration in over 80% of the transformants. Several genomic DNA extraction methods for plant tissues and fungi were developed to simplify the extraction procedure and to allow large numbers of samples to be analyzed. However, these methods still have some drawbacks. For example, some procedures applied to fungi (Cenis et al. 1992) and plant tissues (Steiner et al. 1995) required mechanical grinding of tissues for isolation of DNAs. Alternatively, cell disruption was achieved by temperature extremes in Podospora anserina (Lecellier and Gael 1994) or by alkaline lysis in Neurospora and Aspergillus to prepare DNA for dot blot analysis (Min et a]. 1995). DNA extracted from whole colonies of yeast cells treated with NaOH was used directly in PCR (Wang et al. 1996), but the method could not reliably amplify targets greater than 600 bp (Steiner et a]. 1995). The fungal cell wall can also be digested by specific enzymes to facilitate release of nucleic acids. For example, yeast cells were treated with zymolase (Ling et al. 1995) 38 and the filamentous firngus Aspergillus niger was treated with Novozyme 234 (Zeijl et al. 1998). Both assays could amplify targets up to 3 kb. However, these methods utilized costly enzymes, which require additional time for preparation and cell wall digestion. A rapid spore PCR assay that used conidiospores ofAspergiIlusfizmigatus directly fiom culture plates for screening gene disruption events has been reported (Aufauvre- Brown et a]. 1993). Low numbers of Magnaporthe. grisea spores (104-10°) were also used directly in PCR without extraction of DNA (Xu and Hamer 1995). Spores from five strains of filamentous fungi were treated with microwave irradiation to facilitate release of DNA into the supernatants (Ferreira and Glass 1996). However, it is important to note that these direct spore PCR methods could only amplify PCR targets smaller than 1 kb. Attempts to produce a 1.4 kb PCR fragment using the direct PCR methods described above were not successful. The fungal DNA extraction procedure presented here is simple and effective. Only a small quantity of the firngal cells must be acquired directly from the culture plates eliminating the need for extended culture. Fungal colonies grown for 48 h to 72 h were easily analyzed. Alternatively, colonies could be replicated onto several assay plates reducing the growth time to less than 36 hours. I observed that too much material could introduce a high level of contaminants for PCR, reducing effectiveness. This problem was overcome in most cases by increasing the dilution of the sample prior to PCR analysis. The initial version of our rapid DNA extraction utilized sonication to release DNA. Sonication has been used successfully with other cell types, e.g. bacterial cells (Ausubel et al. 2003). In our laboratory, conidiospores of Aspergillus parasiticis SU-l (2x107 spores/ml) were disrupted by sonication (output control level 4, 10 pulse x 4 cycles, SONIFIER® cell disruptor W—350, Branson Sonic Power Co.). An increase in optical density at 260 nm (A260= 0.34) suggested the release of nucleic acids into the lysates. High temperature treatment of cell lysates was attempted to inactivate cellular 39 nucleases and other PCR inhibitory factors. However, when the efficiency of sonication, boiling, and a combination of the two to release DNA was tested (A260= 0.701, 0.962, 0.989 respectively), the boiling process by itself resulted in the effective release of nucleic acids. Both procedures provided DNA templates for PCR amplification (data not shown). However, the boiling method was nearly as effective and did not require expensive equipment. Boiled fimgal lysates may still contain PCR inhibitory factors which cannot be heat inactivated. This problem can be minimized by diluting the extracts with TE buffer to the point where sufficient DNA templates remain while minimizing the level of inhibitors. In our hands, a 2.5 pl aliquot from the 1:5 dilution of the fungal extract worked well in the 50 pl PCR reaction. Dilution of the original cell extract of transformants generated fi'om other fungal species may have to be adjusted appropriately. Success of PCR amplification also depended on the reaction conditions (e.g., optimum MgC12 concentration, proper annealing temperature and extension time). I observed that the activities of specific thermal stable DNA polymerases played an important role in determining success of amplifying long target fragments. T aq polymerase (Gibco BRL) inconsistently amplified targets up to 2.5 kb using this DNA extraction procedure. Addition of PCR adjuncts and stablizers such as BSA (10-100 pl/ml) and DMSO (1-10%) (Stratagene manual) had limited effects on enhancement of product yield. However, PfirTurbo polymerase (Stratagene) efficiently amplified target fiagrnents up to 3.5 kb. Longer targets were not tested in this assay. Conditions for sample storage also play a role in determining success of the analysis. In our hands, fungal colonies on plates could be stored at 4°C for at least a week before analysis. Once samples were collected into TE buffer, fungal extracts (supernatants) could be used for up to one week of storage at 4°C and up to three months of storage at -80°C. However, the sooner the extracts were used, the better the results. Preparation of fungal extracts for only 10 to 12 samples at one time is recommended. 40 This appears to reduce degradation of nucleic acids which occurs during sample preparation. This rapid DNA extraction method for PCR analysis provides an easy, inexpensive, convenient and efficient assay procedure for screening a large number of Aspergillus transformants and may be applicable to a wide variety of other fungal genera. This rapid procedure to prepare DNA templates fi'om fungal colonies provides sufficient purity for PCR analysis of the site of plasmid integration into the A. parasiticus genome (Chiou et al., 2002). This development was critical for current and future studies on the regulation of the AF promoters (such as nor-1 and ver-I) because detection of the correct site of integration in the genome is essential to accurately measure environmental influences on the aflatoxin gene promoters. In this regard, the speed and effectiveness of the rapid assay makes screening of large numbers of fungal transformants practical. ACKNOWLEDGEMENTS Special thanks to David Wilson, a former student working in the laboratory, for providing the primers to analyze the plasmid integration sites at nor-1 by PCR. This rapid DNA isolation/PCR method is published as part of the data in a manuscript "Chiou, C. H., M. M. Miller, D. L. Wilson, F. Trail and J. E. Linz. 2002. Chromosomal location plays a role in regulation of aflatoxin gene expression in Aspergillus parasiticus. Appl. Environ. Microbiol. 68:306-315". 41 CHAPTER3 Position-dependent regulation of the nor-I promoter in the aflatoxin gene cluster ABSTRACT Expression of most of the AF genes within the 75 kb gene cluster in Aspergillus spp. appears to be co-regulated but the mechanism is not yet clear. I hypothesized that regulation of AF genes is cluster position-dependent; i.e., AF genes are positively regulated within the AF gene cluster. In this study, the nor-1 gene that encodes norsolorinic acid reductase, an early AF pathway enzymatic activity, was used as a study model. Promoter activity of nor-1 was compared within the AF gene cluster (nor-1 locus) and outside the cluster (pyrG locus). A plasmid pAPGUSNP containing the nor-1 promoter sequence fused to a reporter gene (uidA, B-glucuronidase, GUS) plus a selectable marker (pyrG) was constructed to study positional effects on nor-1 promoter function. Transformation of pAPGUSNP into C810 (ver-I, wh-I, pyrG) resulted in homologous recombination at 5' nor-1, 3' nor-1, or the pyrG locus. The plasrrrid integration site was screened by a rapid DNA extraction/PCR assay and confirmed by Southern hybridization analysis. GUS reporter activity was analyzed by a solid-culture GUS assay. Among 178 transformants analyzed, 42 transformants showed positive GUS activity (24.0%). Based on a rapid PCR assay, random selection of 14 GUS+ transformants indicated that plasmid integration occurred within the nor-1 gene (11 within 5' nor-1 and 3 within 3' nor-1). Southern hybridization analysis confirmed that of 10 GUST isolates analyzed, they all harbored pAPGUSNP within the nor-1 locus (7 at 5' nor-1; 2 at 3' nor-1; 1 showed multiple integration at both sites). None of the 11 transformants that harbored the norzzGUS fusion at the pyrG locus expressed detectable GUS activity. The results agree with previous observations that reduced ver-I and nor-1 42 promoter activity is expressed at niaD (encodes nitrate reductase). These data support our hypothesis that expression of AF genes is cluster position-dependent. INTRODUCTION Many of the genes involved in aflatoxin (AF) biosynthesis in Aspergillus spp. have been identified and they were found clustered within a 75 kb chromosomal DNA region (Trail et al., 1995; Payne and Brown. 1998; Woloshuck and Prieto, 1998.). Expression of these genes appeared to be coordinately regulated although details of the regulatory mechanism are not yet clear. The nor-1 gene in the filamentous fungus Aspergillus parasiticus encodes a ketoreductase activity that converts norsolorinic acid (NOR), the first stable intermediate in the AF biosynthetic pathway, to averantin (AVN) (Trail et al., 1994; Zhou and Linz, 1999). The nor-1 gene is expressed under AF inducing conditions, concurrent with many other AF genes (Skory et al., 1993; Zhou and Linz, 1999). The timing of nor-I expression and protein accumulation was consistent with the accumulation of AF (Skory et al., 1993). Previously, Liang et al. (1997) reported preliminary evidence that the organization of AF genes in a cluster may play a role in regulation of AF biosynthesis. The ver-I promoter, which drives expression of a gene encoding a ketoreductase activity involved in a middle step of AF biosynthesis, was fused to uidA (encodes B-glucuronidase; GUS) to generate a reporter plasmid, pHD 6.6, containing niaD (encodes nitrate reductase) as a selectable marker. pHD6.6 was transformed into A. parasiticus NR—l (niaD) resulting in integration predominantly at the ver-I or niaD locus via homologous recombination. Single copy integration of pHD6.6 at niaD resulted in a 500-fold reduction in ver-I promoter activity when compared with single copy integration at the ver-I locus. Similar results have been shown for two norzzGUS constructs located at niaD and the nor-1 locus (Wilson, 1998; Chiou et al., 2002; Miller, 2003). GUS reporter activity was detected at 43 high levels when the nor-1 promoter was positioned at the nor-1 region within the AF gene cluster but was not detected at niaD. However, in A. parasiticus, niaD expression is under the control of nitrogen catabolite repression mediated by areA and the pathway specific induction of nirA. In the presence of nitrate, niaD is positively regulated. On the other hand, nitrate has been shown to down regulate AF biosynthesis (Payne and Brown. 1998). Lack of ver-I and nor-1 expression at niaD could have resulted from niaD- dependent regulation. In this study, nor-1 promoter activity expressed at another chromosomal locus, pyrG, was investigated. The pyrG gene encodes orotidine monophosphate (OMP) decarboxylase that converts OMP to uridine monophophate (UMP), a compound essential for cell growth (Skory et al., 1993). Unlike niaD, expression of pyrG has not been reported to be associated with regulation of AF biosynthesis. In this study, a plasmid (pAPGUSNP) containing four major DNA fiagrnents was constructed. This plasmid includes the 5' nor-I fragment (upstream promoter sequences of the nor-1 gene), a GUS reporter (obtained fiom the Escherichia coli uidA gene which encodes B-glucuronidase), the 3' nor-I fragment (the nor-I terminator and down stream sequences), and a selectable marker pyrG. Transformation of pAPGUSNP into A. parasiticus CS 10 (ver-I , wh-I , pyrG), an uridine auxotroph, resulted in homologous recombination of pAPGUSNP into 5 ' nor-1, 3' nor-1 or pyrG locus. The plasmid integration site was screened by a rapid DNA extraction/PCR assay and confirmed by Southern hybridization analysis. GUS activity was analyzed by a solid culture GUS assay. The results showed that the nor-1 gene was positively regulated only when the promoter was positioned at its native localization (nor-1). When the nor-1 promoter was placed outside the cluster (at pyrG), expression of nor-I was reduced dramatically. This result was in agreement with our previous studies on positional effects at nor-1 and niaD (Wilson, 1998; Miller, 2003). These data provide strong support for the hypothesis that the AF gene cluster exerts a positive position-dependent regulatory influence on the nor-1 promoter. 44 MATERIALS AND METHODS Strains and growth conditions. Escherichia coli DHSa F'e [F'/endAI hstI 7 (r1; mk')supE44thi-1 recAI gyrA (Nalr)Are1A1 (lacZYA argF)u169: (m80 AlacZ M15) (Gibco BRL, Life Technologies Inc., Rockville, MD) was used to amplify plasmid DNA using standard procedures (Ausubel et al., 2003). A. parasiticus CSlO (wh-I, ver-I, pyrG) (Skory et al., 1990) was derived from A. parasiticus ATCC 36537 (wh-I, ver-I) by nitrosoguanidine mutagenesis and was used as a recipient for the reporter construct carrying the pyrG selectable marker (pAPGUSNP). To grow the recipient strain CSlO, YES medium or Czapek Dox medium (CZ, Difco, Detroit, MI) was supplemented with uridine (100 pg/ml). Construction of pAPGUSNP. Plasmid pAPGUSN (Fig. 3.1, A) contains the nor-1::GUS backbone and was generated by Dr. Frances Trail, Department of Botany and Plant Pathology, Michigan State University. The _5'n_0r-L fragment (3.0 kb) contains flanking sequences, promoter, and nucleotides encoding the first 21 amino acid residues of Nor-1. This fiagment was fused in fiarne to a 1.8 kb GUS reporter (B-glucuronidase) derived from E. coli uidA. The 3'_n_0£1_ fragment (2.0 kb) contains nucleotide sequences encoding the last 6 amino acids of Nor-1, the transcription terminator, and flanking sequences. To generate pAPGUSNNP, pAPGUSN was digested with Nhel and the ends filled-in with Klenow fiagment of DNA Poll (Roche Molecular Biochemicals, Indianapolis, Ind.). The 2.5 kb pyrG gene was amplified by PCR with Taq DNA Polymerase (Gibco BRL, Rockville, MD), pPG3J (Skory et al., 1990) as template, and appropriate primers (5'-GTAGAAGTTGTTCAGTAGCTGATGG-3' and 5'- CGAGTATCACAGTCAGGACTCCACG-3'). The resulting PCR fiagment was end- 45 Figure 3.1. Restriction endonuclease maps of plasmids containing the nor-1 ::GUS construct related to this study. Only restriction endonuclease sites relevant to this study are shown. (A) pAPGUSN contains the nor-I ::GUS backbone (generated by Dr. Trail, Department of Botany and Plant Pathology, Michigan State University). (B) pAPGUSNP contains the nor—1::GUS fusion plus the pyrG selectable marker (Chiou, this study). (C) pAPGUSNNB contains the nor-1 ::GUS fusion plus the niaD selectable marker (generated by Wilson, 1998). (D) pBGN3.0 contains the nor-1 ::GUS fusion plus the niaD selectable marker (generated by Miller, 2003). 46 PSI] 800 P511 1000 A Sea] 8300 Sea] 2 100 pAPGUSN 9600 bp Sca12900 Hindlll 3000 Xhol ] 3' nor-l Clal 5930 ‘3‘: L Xhol 10600 I 'i "H C Ia] 5000 EcoRl 4800 pyrG Sea] 2100 Sac] 9600 I Sea] 2900 PAPGUSNP iinmd1113000 12100 bp 0 Xhol | P51] 800 Amp Pstl 1000 GUS . A Sea] 8300 \ Sea] 2100 5' “0“ - Hindlll 3000 - Pstl 13800 5;: 5‘0“ 480° ‘ Cla15000 Cla15900 Sph15200 ma], pAPGUSNNB Gus 17000 bp / 1' EcoRl 4300 D / Clal 5000 3' nor-l / / / / Sea] “5400 Ascl 418 / Clal 5030 Clal 1218 Clal2118 Sea] 8300 \\ 1 i I'fI'\Noti 4226 i' Sea] 4326 7 Sea] 5126 P511 6226 P511 6426 Not] 7226 Pstl I0400 Figure 3.1 47 polished by Pfir polymerase (Stratagene, LaJolla, CA) and blunt-end ligated into pAPGUSN to generate pAPGUSNP (Fig. 3.1, B). Fungal transformation. Plasmid pAPGUSNP (5 -10 pg) was transformed into protoplasts of CS10 (an uridine auxotroph) according to a method described previously (Horng et al., 1990; Skory, et al., 1993). Uridine prototrophic transformants were selected on CZ agar medium with 20% sucrose as osmotic stabilizer. Agar plates were incubated in the dark at 290C for 3 to 5 days. Identification of the site of plasmid integration. (i) Rapid DNA extraction procedure for PCR analysis. Plasmid site of integration was screened by a rapid DNA extraction and PCR method discussed in Chapter 2 (Chiou et al., 2002). Transforrnants were grown until sporulation (usually 3 to 5 days) and conidiospores were transferred to duplicate Czapek Dox agar plates using a sterile toothpick or sterile pipette tip. After 3 to 4 days (or as soon as colonies could be observed), colonies were analyzed. A sterile 200 pl plastic pipette tip was used to transfer a portion of the fungal colony (including conidiospores and mycelia) into a microcentrifuge tube containing 100 pl TE buffer (10 mM Tris-HCI, pH 8.0, 1 mM EDTA). The process was repeated up to 3 times to ensure that sufficient fungal cells were transferred. Tubes were heated in a boiling water bath for 5 min and placed on ice. The samples were centrifuged at 12,000 x g for 5 min at 40C. The supernatant (fungal extract), containing DNA, was transferred to a new microcentrifuge tube. The extract was diluted with TE buffer (undiluted, 1:5, 1:20, and 1:100). In most assays, 2.5 pl of the 1:5 dilution resulted in good amplification in a 50 pl PCR reaction. The remainder of the fungal extract was stored at -800C and could be analyzed by PCR for at least 3 months. Diluted samples stored at 40C could be analyzed for at least 1 week. (ii) PCR analysis. PCR analysis was conducted on DNA extracted from A. parasiticus transformants and’the recipient strain A. parasiticus CS10. The primer pairs utilized depended upon the plasmid vector used and the site of integration analyzed 48 (Table 3.1). A 50 p1 reaction consisted of 2.5 pl of the diluted fungal extract, 20 mM Tris-HCl (pH 8.8), 10 mM KC], 10 mM (NH4)ZSO4, 0.15% Triton X-100, 0.1 mg/ml BSA, 0.5 to 1.0 pM of each primer, 200 pM of each dNTP, 2.0 to 3.0 mM MgC12, and either 2.5 units of Pfu Turbo polymerase (2.5 U/pl, Stratagene, LaJolla, CA) or 2.5 units of Taq polymerase (5 U/pl) (Gibco BRL). PCR cycling depended on primer design and the target size (Table 3.1). PCR was performed in a Gene Ampli System 9600 (Perkin- Elmer) thermal cycler. All PCR reactions were initiated by a denaturation step at 950C for 3 min followed by 35 cycles of denaturation, annealing, and extension. Reactions were terminated with an extension reaction at 720C for 10 min. PCR products were observed by electrophoresis on a 1% agarose gel using standard procedures (Ausubel et al., 2003). Solid-culture GUS activity assay. A solid-culture assay to analyze GUS activity was developed based on published procedures (Kolar et al., 1991; Verdoes et al., 1994). Transforrnants were transferred from selective medium to agar plates containing an aflatoxin-inducing solid growth medium (YES) overlaid with sterilized circular 82 mm Nytran Plus membranes (Schleicher and Schuell Inc., Keene, NH). Plates were incubated in the dark at 290C for 36 to 55 h. Membranes with attached fungal colonies were removed from the surface of the agar plates and immersed in liquid nitrogen for 30 s. Membranes were allowed to thaw and immersed again in liquid nitrogen for 30 s. Afier thawing, membranes were incubated in GUS substrate solution (60 mM NazHPO4, 40 mM NaH2P04, 10 mM KC], 1 mM MgSO4, 0.27% B-mercaptoethanol. 0.04% X GLU [Gold Biotechnology, St. Louis, MO]) for up to 24 h. Positive GUS activity resulted in a blue color in the colony Southern hybridization analysis. The plasmid integration site was confirmed by Southern hybridization analysis. Fungal genomic DNA was purified from A. parasiticus transformants grown in YES liquid shake culture (the recipient strain CSlO required 49 Table 3.1. Primer sequences and PCR cycling parameters used for analysis of site of integration using DNA templates prepared by the rapid method. Primer Target Primer Sequence PCR cycling size (kb) parameters Integration in 5' region of nor-1 gene PCR 1 IL 99 3.1 kb S'rrr CAC GGG TTG GGG TTT CTA CAG G 3' 950C 1', 620C 1', 720C 5' JLIOO 5'GAc GGG GAA ccr crr TAC AAA CAT c3' Integration in 3' region of nor-1 gene PCR 2 JL102 2.1 kb 5'CGC AAG GTG AGG GTT CGA ACC GAG G3' 950C 1', 680C 1', 720C 3' JL103 5'CCG CAG CAG GGA GGc AAA CAA TGA A3' Integration in pyrG PCR 3 JL221 3.5 kb 5'CCA GAA AAT GCG AAG GTA AGT GCT Tc3' 950C 1', 550C 1', 720C 5' JL222 5'GCG ACA CGG AAA TGT TGA ATA crc AT3' Endogenous vbs ExonII (Control) JL136 1.4 kb S'AGG CTC GAA AGG CGC ATA CGA3' 950C 1', 680C 1', 720C 2' JL137 5'GTG GTC TAC TGC CCA GCC ATC3' 50 supplementation with 100 pg/ ml Uridine) for 48 h at 29°C with five, 6 mm glass beads (Skory et al., 1993). Southern hybridization analysis was performed using digoxigenin labeled probes and a non-radioactive hybridization detection procedure provided by the manufacturer (Roche Molecular Biochemicals, Indianapolis, Ind). The restriction enzymes and probes used to analyze the integration site are summarized in Table 3.1. The integration scheme and resulting blots are shown in Figures 3.2 and 3.3. (i) 5' nor-1 integration. Genomic DNA of transformants was digested with Xhol and probed with Probe 1. DNA fi'om A. parasiticus CS10 generates a 5.8-kb band while DNA from a 5' integrant of pAPGUSNP generates 13.0 and 3.4-kb bands (Fig. 3.2). (ii) 3' nor-1 integration. Genomic DNA of transformants was digested with ScaI and probed with Probe 2. DNA from A. parasiticus CSlO generates a 3.0-kb band while DNA from a 3' integrant of pAPGUSNP generates 4.0 and 4.4-kb bands (Fig. 3.2). (iii) Integration at pyrG. Genomic DNA of transformants was digested with SphI and SacI, respectively, and probed with a 2.5-kb pyrG fragment generated by PCR (Probe 3). DNA fi'om A. parasiticus CSlO digested with SphI generates a 12.6-kb band while DNA from a pyrG integrant generates 16.0 and 8.8-kb bands. CSlO genomic DNA digested with SacI generates a 8.1-kb band while DNA from a pyrG integrant generates 5.9 and 14.3-kb bands. (Fig. 3.3). RESULTS Generation of pAPGUSNP. Plasmid pAPGUSNP was constructed by blunt-end ligation of a 2.5 kb pyrG fragment into a 9.6 kb nor-1 ::GUS backbone derived from pPAPGUSN (Fig. 3.1, A). The rate of successful blunt-end ligation was low. The ligated plasmids were transformed into E. coli DHSOL. Among 1200 transformants generated, only three transformants contained the 2.5 kb pyrG gene; these were identified by DNA dot blot analysis using digoxigenin-labeled pyrG fragment (2.5 kb) as a probe. Restriction digestion analysis indicated only one transformant harbored the correct 51 Figure 3.2. Southern hybridization analysis of site of integration in pAPGUSNP transformants. Restriction endonuclease map of pAPGUSNP is shown on the top. (A) Schematic for Southern hybridization and PCR analyses of pAPGUSNP integration into 5' region of nor-1 and Southern hybridization data. A. parasiticus genomic DNA was digested with XhoI and probed with Probe 1. Single crossover integration of pAPGUSNP into the 5’ region of nor-1 results in two bands, 19 kb and 3.4 kb, and the disappearance of the 5.8-kb wild-type band. (B) Schematic for Southern hybridization and PCR analysis of pAPGUSNP integration into the 3' region of nor-1 and Southern hybridization data. Genomic DNA was digested with Sea] and probed with Probe 2. Single crossover integration of pAPGUSNNB into the 3’ region of nor-1 results in two bands, 4.0 kb and 4.4 kb, and the disappearance of the 3.4-kb wild-type band. Abbreviations: X = Xhol; Sc = ScaI; C = Clal. 52 Xhol 10600 Sac] 9600 Scal 8300 A 5’nor—1 X ScSc X CCSc — nor-l Probe 1 '2 .. .._,.-. -..__ «..I X 5.8 kb X X . PCR] I "5 "' .. - GUS pyrG - nor-l I I l l 19 kb I 3.4 kbl 3.1 kb C810 24 37 4O 60 84 17 32 57 -231 kb '- Vl‘;"""- 9.41m — 6.5 kb 5.8 kit—>3! — 4.31m -— 2.3 kb Figure 3.2 pAPGUSNP 12100 bp 11 Seal 2100 Sea] 2900 HindIII 3000 EcoRI 4800 Cla15000 B 3’nor-1 X X ScSc X CCSc ENE. r-tl'r . ., a], Or- 4* '1‘” _H .. nor- 3.4 kb— CS‘IO 24 37 40 60 84 17 32 57 —23.1kb U VI.- . _ 9.4 kb «7 — 6.5 kb 4.41m.> '- .. V “' .. "‘ “a. 4.0 kb—> - we - 4.3 kb 3.4kb—> “U 'w-”~~ — 2.3 kb 53 Figure 3.3. Southern hybridization analysis of pAPGUSNP integrated at pyrG. Restriction endonuclease map of pAPGUSNP is shown on the top. Genomic DNA was digested with Sphl (A) or SacI (B) and hybridized with the digoxigenin-labeled pyrG probe (2.5 kb, Probe 3). With Sphl restriction endonuclease digestion (panel A), single crossover integration of pAPGUSNP into pyrG results in two bands, 16 kb and 8.8 kb, and the disappearance of the 12.6-kb wild-type band. With Sacl restriction endonuclease digestion (panel B), single crossover integration of pAPGUSNP into pyrG results in two bands, 14.3 kb and 5.9 kb, and the disappearance of the 8.1-kb wild-type band. Abbreviations: Sp = SphI; Sa = Sad. 54 ........ Xhol l0600 Sea] 2100 Sacl9600 Sea] 2900 HindIIl 3000 pAPGUSNP 12100 bp Sea] 8300 EcoRI 4800 Clal 5000 1——1 A SphI 3.5 kb B $5161 (308- GUS- l I I l csm1oo 21121 133 135 142145154 14 95 105 57 cs101oo 21121 133 135 142149154 14 95 105 57 l l i 15.0 115...: , 12.5 lib->5» ~ -' w l 8'8 kb- 14.3 115->- . I 5.1 kb- 5.9115. 1 i i 1 Figure 3.3 55 construct. This plasmid (pAPGUSNP) carries nor-1:: GUS and pyrG oriented in opposite directions, resulting in divergent transcription the two genes (Fig. 3.1, B). Transformation of A. parasiticus C810 with pAPGUSNP. Transformation of the A. parasiticus uridine-auxotrophic strain CS 10 with plasmid pAPGUSNP was performed in three separate experiments. The pyrG+ uridine-prototrophic transformants were selected on CZ media with 20% sucrose. In total, 178 transformants were generated and subsequently analyzed using a solid-culture GUS activity assay. Solid-culture GUS assay. All of the pyrG+ transformants selected on CZ medium were analyzed for expression of GUS on AF inducing medium (YES agar plate). Solid-culture GUS assay of these transformants (178 in total) resulted in identification of 42 GUS+ isolates (24%). Randomly selected GUS+ or GUS' isolates were subsequently analyzed by PCR or Southern hybridization to determine the plasmid integration sites. Determination of pAPGUSNP integration site within the chromosome. Fourteen GUS + isolates were randomly selected and screened by a rapid PCR assay for site of integration (Fig. 3.4). The rapid PCR assay utilized a quick and easy template preparation procedure (see Materials and Methods) and the results were comparable to PCR using purified genomic DNA (Chiou et al., 2002). All 14 GUS + isolates had pAPGUSNP integrated at 5' nor-1 (11 isolates) or 3' nor-1 (3 isolates). The site of integration of pAPGUSNP was confirmed in 4 of these isolates and 6 additional randomly selected GUS + isolates by Southern hybridization analysis (Fig. 3.2). Of these 10 isolates, 7 had pAPGUSNP integrated at 5' nor-1 and 2 carried pAPGUSNP at 3' nor- 1. Isolate 57 was likely a multiple integrant of pAPGUSNP; these data suggested this isolate carried one copy of pAPGUSNP at both 5' and 3' nor-1 (Fig. 3.2). In contrast, of 11 isolates with pAPGUSNP integrated at pyrG, none expressed GUS activity (Southern hybridization analysis shown in Fig. 3.3). 56 A 1051012345673. 3.1 kb 3.5 kb —> 2.1 kb —> B 108101234 567¢x174 1.4 kb Figure 3.4. PCR analysis to detect site of integration in pAPGUSNP transformants. Template DNA for 5' and 3' nor-1 and pyrG integration assays was prepared with a rapid boiling procedure. (A) GUS+ and GUS' transformants (lanes 1 through 7) and CS10 (negative control) were analyzed for 5' nor-1, 3' nor-I , or pyrG integration. Primers PCR3 were used to detect pyrG integration (lanes 1, 2, and 3; GUS' transformants) and generated a 3.5-kb PCR fragment. Primers PCR2 were used to detect 3' nor-1 integration (lanes 4 and 5; GUS+ transformants) and generated a 2.1-kb PCR fragment. Primers PCR] were used to detect 5' nor-1 integration (lanes 6 and 7; GUS+ transformants) and generated a 3.1-kb PCR fiagrnent. (B) DNA from all 7 transformants and the control strain CS10 was amplified with a primer pair specific for exon III of the versicolorin B synthase gene (positive control). A 1.4-kb PCR fragment was expected in all fungal isolates. Lanes containing molecular size markers in panels A and B are represented by AHindlIl digest and ¢x174 HaeIII digest. Nucleotide sequences of all primers are found in Table 3.1. 57 YES Figure 3.5. Solid-culture GUS activity assay on CSIO transformants harboring pAPGUSNP. The recipient strain CS10 (control) and representative pyrG+ transformants were inoculated onto Nytran membranes overlaid onto YES agar medium (panels C and D) or YES supplemented with 5 mg uridine per plate (panels A and B). A GUS assay was conducted on filters in panels B, C and D. Transformants carried a single copy of pAPGUSNP integrated at 5' nor-1 (transformants 17 and 32), at 3' nor-I (transformant 37), or at pyrG (transformants 100 and 121). Transformants 24 and 60 carried multiple copies of pAPGUSNP with at least one copy integrated at 5' nor-I . Transformant 57 carried multiple copies of pAPGUSP with at least one copy integrated at 3' nor-1. Transformants l4 and 133 carried multiple copies of pAPGUSNP with at least one copy at pyrG and no copies at nor-I . 58 Confirmation of pyrG expression in pyrG+ transformants. Expression of pyrG in A. parasiticus CS 1 0 (used as the recipient) was analyzed under these aflatoxin- inducing growth conditions (Fig. 3.5). Growth of A. parasiticus CS10 (wh-I, ver-I, pyrG) on YES required supplementation with uridine (100 pg/ml) while transformation of CSIO with a functional pyrG gene overcame this requirement demonstrating that a functional copy of pyrG must be present and expressed for growth of CS10 on YES medium. DISCUSSION Based on reduced ver-I promoter activity expressed at the niaD locus observed previously (Liang et al., 1997), I hypothesized that location within the AF cluster provided a positive regulatory influence on AF gene promoter function under aflatoxin- inducing growth conditions. Suppression of promoter activity of another AF gene, nor-1, at niaD was also demonstrated prior to the current study (Wilson, 1999; Miller, 2003). However, alteration of promoter activity may be specifically affected by the niaD gene since niaD is only expressed under non-AF inducing conditions using nitrate as the sole nitrogen source; otherwise the gene is repressed. Reduction of ver-I and nor-1 promoter activity at the niaD locus may be a consequence of regional control. Thus, investigation of positional effects of AF promoter activity at an additional locus provides supporting data for our previous observations. In this study, promoter activity of nor-1 located in the pyrG locus and its native chromosomal location was investigated. The pyrG locus was chosen for two reasons. 1) The selectable marker (pyrG) that encodes uridine monophosphate decarboxylase was cloned previously (Skory et al., 1990). 2) Expression of pyrG is not associated with regulation of AF biosynthesis, unlike niaD. The results generated in this study were consistent with results generated in analysis of nor-1 promoter activity within the AF gene cluster and at niaD (Chiou et al., 2002). First, the percentage of GUS+ transformant is similar (approximately 24%). This 59 study utilized a plasmid pAPGUSNNP (Fig. 3.1, B) containing the nor-1::GUS fusion (derived from pAPGUSN, Fig. 3.1, A) plus pyrG as a selectable marker; while pAPGUSNNB (Fig. 3.1, C) and pBNG3.0 (Fig. 3.1, D) contained a similar nor-1::GUS construct and niaD as a selectable marker. Transformation of pAPGUSNNB and pBNG3.0 into A. parasiticus NR1 (niaD) resulted in 23% (25 out of 149 transformants analyzed) and 25% (29 out of 117 transformants analyzed) GUS+ transformants, respectively (data summarized in Table 3.2). Second, these data confirmed the preliminary observation by Liang et al. (1997) that chromosomal location plays a role in regulation of AF gene expression. Expression of nor-1::GUS at the nor-1 locus was regulated similarly to the native nor-I gene, while placement at two other positions (niaD and pyrG) outside of the aflatoxin gene cluster resulted in no detectable GUS activity. The targeting of reporter constructs to specific chromosomal locations for promoter analysis has been advised by other investigators studying expression of fungal genes (Hammer, and Timberlake, 1987; Timberlake and Marshall, 1989; van Gorcom et al., 1986). Although the mechanisms of position- dependent gene expression have not been fully elucidated in filamentous fungi, it has been hypothesized that enhancer elements may be responsible for the position-dependent effect (Kinsey and Rambosek, 1984). The hypothesis that positive cis-acting factors have regional control over the transcription of AF genes is reasonable, since removal of AF reporter fusions from the AF gene cluster results in reduced GUS expression. An alternative explanation for the lack of GUS activity at the niaD and pyrG loci is that a transcriptionally inactive chromatin structure exists at these two sites. We think that such inactivation is unlikely for two reasons. First, NR-l transformed with pGAPN2B (B—tubulin::GUS) expressed GUS activity at similar levels at 24, 48, and 72 h when integrated at the niaD locus, although the level of GUS activity in the pGAPN2B transformants was approximately ten-fold lower than GUS activity in D8D3 60 Table 3.2. Summary of GUS activities and plasmid integration sites exhibited in A. paraciticus transformants carrying nor: :GUS fusion in CS 1 0 transformants carrying pAPGUSNP (Chiou, in this study). and NR-l harboring pAPGUSNNB (Wilson, 1998) and pBNG3.0 (Miller, 2003). 9'38"“ pAPGUSNP“ pBNG3.0b pAPGUSNNBc Target locus pyrG niaD niaD Recipient strain CS10 NR-1 NR-1 it of Total transformants 178 117 149 # of GUS+ 42 29 35 % GUS + 24% 25% 23% “WNW "WW“ “Chiou leller cwrison (1995) (this study) (2003) PCR on GUS+ transformants 14 Integration at 5' nor-1 11 Integration at 3' nor-1 3 Southern analysis on GUS+ 10 21 11 transformants Integration at 5' nor-1 7 12 4 Integration at 3' nor-1 2 9 6 Multiple intgegration at 5' and 3' nor-1 1 Unverified 1 Southern analysis for 608- 11 22 6 transformants Integration at 5' nor-1 0 0 0 Integration at 3' nor-1 0 0 0 Integration at pyrG 1 1 Integration at niaD 22 2 Single crossover at niaD 4 61 (pAPGUSNNB integrated at 3' nor-1; Wilson, 1998) at 48 and 72 h. In contrast, transformant D8D4 (pAPGUSNNB integrated at niaD; Wilson, 1998), when grown under the same conditions, produced no detectable GUS activity at any of these time points (data not shown). Second, in transformants carrying pAPGUSNP, the nor-1 promoter carried on this plasmid was not active at the pyrG locus even though growth on YES medium required expression of the pyrG gene. In addition, several pAPGUSNP transformants had no detectable GUS activity even when they carried multiple copies of the plasmid (none integrated at nor-1). In contrast, other multiple integrrants with at least one copy of pAPGUSNP integrated at nor-l were GUS": Based on these data and preliminary data reported by Liang et al. (1997) I propose that aflatoxin gene expression is influenced by an enhancer element. The clustering of genes involved in the same secondary metabolic pathway is a common theme in the filamentous fungi (Keller and Hohn, 1997). However, cis-acting elements that regulate many genes simultaneously in fungal gene clusters have not been identified. If an enhancer element does influence the regulation of nor-1 transcription, it is located at least 3-kb upstream from the transcription initiation site in the 5' nor-1 region, and at least 1.8-kb downstream the transcription termination site in the 3' region of nor-1 or within the nor-I coding region. Because similar position-dependent expression was observed with the ver-1::GUS reporter construct (Liang et al., 1997), it also is possible that the same cis-acting element is influencing the regulation of both nor- 1 and ver-I . Studies performed on the SpoCl gene cluster in Aspergillus nidulans showed that clustered genes can be coordinately expressed during development, and that placement of cluster genes at ectopic chromosomal locations results in the loss of that coordination (Miller et al., 1987; Timberlake and Barnard, 1981). The physical linkage of aflatoxin biosynthesis genes also may have a regulatory role. Studies designed to establish such a phenomenon and to determine the nature of the regulatory mechanism, 62 may eventually lead to a broader understanding of the expression of secondary metabolism genes and the ability to manipulate their expression in microorganisms. ACKNOWLEDGEMENTS We thank Dr. Frances Trail, Department of Botany and Plant Pathology at Michgan State University, for providing plasmid pAPGUSN. Data in Chapter 3 is published in Applied and Environmental Microbiology in a manuscript entitled "Chromosomal location plays a role in regulation of aflatoxin gene expression in Aspergillus parasiticus" (Chiou et al., 2002). 63 CHAPTER 4 Immuno-localization of Nor-1, Ver-l, and OmtA proteins involved in aflatoxin biosynthesis in Aspergillus parasiticus INTRODUCTION The biosynthesis of aflatoxin is a complex process that involves at least 18 enzyme activities (Minto and Towsend, 1997; Payne and Brown, 1998, Cary et al., 2000). The mechanisms that coordinate these enzyme activities to synthesize AF within a cell are still not clear. In addition, the distribution of these AF enzymes in a fungal colony and their sub-cellular localization within a fungal cell have yet to be determined. The objective of this study was to investigate the distribution and sub-cellular location of representative enzymatic activities in the aflatoxin biosynthetic pathway. This study focused attention on Nor-1 (Trail et al., 1994), Ver-l (Skory et al., 1992) and OmtA (Yu et al., 1993) that catalyze early, middle and late enzymatic steps in the aflatoxin biosynthetic pathway, respectively. Nor-1 is a NADPH-dependent keto-reductase involved in the conversion of norsolorinic acid (NA) to averantin (AVN) (Zhou and Linz, 1999). Ver-l is another keto-reductase involved in conversion of versicolorin A (VERA) to demethyl-sterigmatocystin (DMST)(Skory et al., 1992; Liang et al., 1996). OmtA is a methyltransferase that specifically converts sterigmatocystin (ST) to O- methylsterigrnatocystin (OMST) by transfer of a methyl group fi'om S- adenosylmetlrionine (SAM) (Yu et al., 1993; Lee et al., 2002). Our laboratory produced polyclonal antibodies (PAb) against Nor-l (Zhou, 1998; Lee et al., manuscript submitted), Ver-l (Liang, et al., 1997), and OmtA (Lee et al., 2002) to localize these enzymes in A. parasiticus SU-l and CSIO. An omtA disruptant strain, LW 1432 (th, ver-I , omtA), was also generated to serve as a negative control (Lee et al., 2002). A model for coordinate regulation of sterigmatocystin (ST) production (an AF biosynthetic pathway intermediate in A. parasiticus) and asexual sporulation has been 64 described in A. nidulans (Hicks, et al., 1997; Adams et al., 1998). Based on this model, I hypothesized that the enzymes involved in AF biosynthesis are localized within specific developmental structures; this study focused on an asexual spore-bearing structure, the conidiophore. At the sub-cellular level, according to the complexity of this AF biosynthetic pathway, I hypothesized that enzymes (or some of the enzymes) involved in AF biosynthesis are compartmentalized or closely associated within a cell. Physical interaction of these enzymes may facilitate efficient and accurate catalysis. Preliminary cellular localization of Nor-1 and Ver-l in the wild type AF producing strain A. parasiticus SU—l was performed by Liang (1996) and Zhou (1998) using a slide-culture cel] preparation procedure followed by indirect irnmuno- fluorescence microscopy. However, artifacts were generated by this sample preparation method for several reasons. First, growth conditions and colony morphology were different from colonies generated by center-inoculation of conidiospores on solid media. In the slide-culture method, conidiospores were inoculated on the surface of agar blocks (approximately 8 m2) placed between 2 sterile coverslips. Mycelia as well as the conidiophores attached onto one surface of the coverslips as growth continued. Acquiring nutrients from the agar block likely became less efficient as the colony extended outwards, unlike regular fungal growth on an agar plate in which the nutrients were obtained from the agar surface directly underneath the mycelia. Slide-culture may induce stress responses and cause alteration of fungal development. Second, the fungal cell wall represents a barrier that obstructs cell permeablization and antibody accessibility; these conditions are extremely critical to achieve success in subsequent irnmuno-labeing. Thus, lysing the cell wall is an important step to immuno-localize proteins in the fungal cells. However, digestion of A. parasitius cell wall utilizing Novozyme 234, a partially-purified cell wall lysing enzyme isolated from T richoderma harzianum, did not give a constant degree of cell wall digestion, thus causing artifacts in immuno-labeling. Preliminary indirect fluorescence microscopy detected Nor-l in the 65 cytoplasm in the conidiophores and the substrate level mycelia (Zhou 1998). Ver-l was detected mainly in the conidiophores and the signals were confined to particle-like structures (Liang, 1996). However, these results were not consistently reproducible, and no reliable negative control was available. The goal of my study was to develop a growth model that would closely mimic regulation of toxin synthesis in soil and on the host plant. An improved colony fractionation scheme as well as cell preparation methods were developed. A 72 h-old colony was fractionated based on 24-h intervals to correlate the time of growth with AF protein distribution. Colony fiactions were then fixed and embedded in paraffin to generate sections (4-5 pm) for immuno-labeling without the need to lyse cell walls. In addition, instead of using a conventional epi-fluorescence microscope, fluorescence microscopy was performed using a Confocal Laser Scanning Microscope (CLSM) which provides better resolution of confocal images (optical sections) and broader image analysis capabilities (described in Chapter 1). The results of my study demonstrated that Nor-1, Ver-l and OmtA were evenly distributed among different cell types and were not concentrated in conidiophores, contrary to the hypothesis. In a 72 h-old colony, these proteins were most abundant in colony fraction 2 where the cells had grown for 24 to 48 h. Within a cell, Nor-l and Ver- 1 were primarily localized to the cytoplasm, suggesting that they are cytosolic enzymes. OmtA was also detected in the cytoplasm and the fluorescent signal was confined to discrete areas (patches). Yet due to limitations in resolution of CLSM, it was not clear if those patches represent specific sub-cellular organelles. The pattern of labeling using anti-OmtA was not consistent with localization of OmtA only to nuclei, peroxisomes, or Woronin bodies. However, the data do not rule out the possibility that OmtA was localized to other cell locations as well as these specific organelles. Although Nor-1, Ver-l and OmtA appeared to localize in the cytoplasm, compartrnentalization of these proteins needs to be further explored. 66 MATERIALS AND METHODS Fungal strains. A. parasiticus SU-l (NRRL5862, ATCC 56775) is a wild-type, aflatoxin-producing strain. A non-aflatoxin producing aflR knockout strain, AF S10 (kindly provided by Dr. Jeff Cary, USDA, New Orleans, LA.), was used as a negative control. AF $10 was generated fiom A. parasiticus NR-l (niaD) which was derived from parental strain SU-l (Cary et al., 2002). Two additional strains, A. parasiticus CSIO (ver- 1 wh-I pyrG) (Skory et al., 1993) derived from A. parasiticus ATCC36537 (ver-I wh-I), and an omtA disrupted mutant strain LW1432 (ver-I wh-I omtA) generated by disrupting the omtA gene in CS10 (Lee et al., 2002) were also used to study localization of OmtA. Time-dependent fractionation of colonies grown on solid medium. To determine the accumulation and distribution of Nor-1 , Ver-l, and OmtA in fungal colonies grown on solid culture media, conidiospores (2 x 105) of A. parasiticus SU-l and AF S10, as well as CSIO and LW1432 (to specifically localize OmtA), were inoculated onto the center of YES (2% yeast extract, 6% sucrose, pH 5.8) agar overlaid with sterile cellophane membranes and incubated at 29° C in the dark. Three-day-old colonies of SU-l (S) and AF S10 (R) were fractionated into three concentric rings based on area covered at 24, 48, or 72 h of growth (fiaction 1, 48-72 h old; fraction 2, 24-48 h, and fraction 3, 0-24 h) to generate fractions 81, S2, S3 and R], R2, R3 (Lee et al., 2002). A. parasiticus CSIO, and LW1432 colonies were fractionated following the same scheme to generate analogous fractions C1, C2, C3, and L1 , L2, L3, respectively (Figure 4.1). Preparation of paraffin-embedded fungal sections. Samples from fungal colony fiactions were embedded in paraplast (Sigma, St. Louis, MO) using a published procedure (Ausubel et al., 2003) with the following modifications. Fungal tissues were fixed with Streck tissue fixative (Streck Laboratory Inc., Omaha, NE) at 4° C overnight, 67 SU1:wild type AFS1O (afIR) CS10 (wh 1. var-1.10m?) LW1432 (wh1,ver-1, omtA) 0% ca L2 L3 81, R1, 01, L1: 48-72 h 82, R2, C2, L2 : 24-48 11 SS, R3, CB, L3 : 0-24 h Figure 4.1. Time-dependent fractionation of firngal colonies. A 72 h culture of SU-l (wild type AF producing) grown on YES medium (2% yeast extract, 6% sucrose) was divided into 3 fractions based on 24 h grth intervals (S1: 48- 72 h growth; S2: 24- 72 h growth; S3: 0- 24 h growth). The control strain AF 810, an qflR knockout mutant, was prepared similarly to generate R1, R2 and R3. For detection of OmtA, additional strains, LW1432(ver-1 wh-I omtA) and its parent strain CSIO(ver-1 wh-I pyrG), were used to generate L1, L2, L3 and C1, C2, C3, respectively. 68 and then dehydrated in a graded series of ethanol: 30% (30 min, room temperature); 50% (30 min, room temperature); 70% (overnight, 4° C); 85% (30 min, room temperature, 2 times); 95% (30 nrin, room temperature, 2 times); 100% (30 min, 4° C, 2 times), followed by incubation in 100% xylene (10 min, room temperature, 3 times). Fungal tissues were then incubated in a paraffin/xylene mixture (1:1; v/v) for 15 min, 60 rrrin and overnight at 60°C and finally for 8 h in 100% paraffin at 60° C (three changes of paraffin during incubation). The paraffm embedded sample blocks were hardened in a plastic mold (V WR Scientific, Detroit, M1) at RT. Sample blocks were cut into 4 pm thick sections using a tissue section microtome (AO Spencer 820 microtome, Fisher Scientific). The sections were attached to poly-L-Lysine (Sigma) coated coverslips (22 mm square). Immuno-fluorescence labeling of fungal cells. Coverslips with paraffin- embedded ftmgal sections were placed in a coverslip holder (EMS, Fort Washington, PA) for de-paraffinization and antigen retrieval. Sections were de-paraffinized twice in 100% xylene for 10 min, then rehydrated with a decreasing concentration of ethanol: twice in 100% for 10 min, once in 95%, 70%, 50% for 5 min each, and finally in deioned distilled H20. Antigen retrieval was performed by heating the sections in 10 mM citrate buffer (pH 6.0) at 95° C for 5 min followed by cooling at room temperature for 20 min. Coverslips were rinsed with TBS (Tris buffer saline, pH 7.5) and incubated in blocking solution (1% BSA with 0.1% saponin in TBS) at 4°C overnight. The samples were probed with primary antibodies against Nor-1 (1 to 500 dilution)(Zhou, 1998; Lee et al., submitted), Ver-l (20 pg/ml) (Liang et al., 1997), OmtA (20 pg/ml) (Lee et al., 2002), or anti-SKL (1:500, Zyrned, So. San Francisco, CA) as a positive control, followed by secondary antibody conjugated to a fluorescent probe (goat anti-rabbit IgG- Alexa 488 conjugate [5 pg/ml]; Abs 495 nm/ Em 519 nm) at RT for l h. Coverslips were washed after each antibody treatment with TBS containing 1% BSA and 0.1% saponin followed by two washes with TBS for 10 min. Fungal nuclei were detected using SYTOX Green 69 fluorescence dye (Molecular Probes). The samples were mounted onto microscopic slides with Prolong anti-fade mounting media (Molecular Probes, Eugene, OR). Confocal Laser Scanning Microscopy (CLSM). Fluorescence image detection was performed on a Zeiss 210 Laser scanning microscope or Zeiss LSM5 Pascal Laser Scanning Microscope with a 488 nm laser line. The 40 X oil objective lens (Zeiss Plan- NeoFlura, NA= 1.3) was used to obtain most images. The Alexa 488 fluorescence probe (Abs 495 nnr/Em 519 nm) was detected using LP 520 or BP 520-560 barrier filters. Fluorescence image analysis of SU-l, AF SlO, CS10 and LW1432 was conducted under the same instrument parameter settings. High resolution images were generated using a Zeiss Plan-APOCHROMAT (63 X/ 1.4 Oil Dic, N.A.=1.4) objective along with higher zoom levels in the Zeiss LSM5 Pascal CLSM specifically to demonstrate the localization of Nor-1, Ver-l , and OmtA within a SU-l cell. Single optical sections or extended focus images from CLSM Z- stacks were obtained. Quantitative fluorescence intensity analysis. To quantitate fluorescence intensity, a Meridian IN SIGHT® plus laser-scanning system (Meridian Instruments, Inc., Okemos, MI) connected to a Zeiss Axioskop microscope equipped with IQ Master Program (V2.31) image analysis software was used. The Alexa 488 green fluorescence was detected using a BP 530/30 barrier filter under the 488 nm laser line. All images were captured using a 40X Zeiss Plan-NEOFLUAR oil objective lens (N .A.=1 .3) (Carl Zeiss Inc., Germany) with a 1X cooled charged-couple detector (CCD) allowing analysis of a large number of cells in one 512 x 480 pixel resolution image. Fluorescence image analysis of strains SU-l and AF 810, as well as CSIO and LW1432 was conducted using the same instrument parameter settings. For each colony fiaction, twenty images were acquired for intensity analysis and the average pixel number was reported. 70 RESULTS Immuno-Iocalization of Nor-1 and Ver—l. CLSM detected Nor-1 and Ver-l in all colony fiactions of SU-l (Fig. 4.2a for Nor-1 and Fig. 4.2d for Ver-l). Under the same instrument settings, very little signal was detected in AF 810 (Fig. 4.2b for Nor-l and Fig. 4.2e for Ver-l). Therefore, bright field images of the AF SlO colony fi‘actions are shown to illustrate the typical size and numbers of cells analyzed (Fig. 4.2e and 4.21). Highest fluorescence intensity was detected in fraction 82 for both Nor-1 and Ver-l. Both proteins were equally distributed in substrate level mycelia and conidiophores (Fig 4.3 for Nor-l; Fig. 4.5, C for Ver-l). Within a cell, high resolution optical sections taken using Zeiss LSM5 PASCAL microscope with a Zeiss Plan-APOCHROMAT (63 X Oil Dic, NA=1.4) objective showed that both Nor-l and Ver-l were mainly localized in the cytoplasm. Nor-1 did not associate with specific organelles in either hyphae (Fig 4.3, A to D) or conidiophores (Fig. 4.3, E). Ver-l was mainly present in the cytoplasm (Fig. 4.4, A, and B), frequently associated with the cell wall as well as spore coat, and occasionally detected in particle- like structures (Fig. 4.4, C, and E). However, the presence of Ver-l in locations other than cytoplasm appeared to be non-specific since these labeling patterns were also observed in the control strain AF 810 (Fig. 4.5, A). Thus, both Nor-1 and Ver-l are concluded to be cytosolic proteins. Although the Ver-l antibodies bound to AF S10 non- specifically, the overall fluorescence intensities detected in AF S10 were significantly lower than in SU-l (Fig. 4.2, B and Fig. 4.5). Immuno—localization of OmtA. To compare the fluorescence labeling intensity between SU-l and CS10 and their non-aflatoxin producing counterparts, AF $10 and LW1432, respectively, the samples were viewed under the lowest zoom level using the Zeiss 210 CLSM (zoom = 20) in order to acquire maximum cell number within a field. Under the same contrast and brightness settings, the samples prepared from the control 71 Figure 4.2. Immune-fluorescence confocal microscopy of Nor-1 and Ver—l in A. parasiticus SU-l and AF SIO (aflR knockout mutant) grown on YES agar for 72 h. Colonies of SU-l and AF 810 were divided into 3 fiactions (see Material and Methods). The paraffm embedded fungal sections were immuno—labeled with anti-Nor-l antiserum (panel A) (1:500) or anti-Ver-l IgG (panel B) (20 pg/ml) followed by Alexa 488 conjugated goat anti-rabbit IgG. Fluorescence intensities of SU-l sections irnmuno- labeled with anti-Nor-l and anti-Ver-l are shown in Rows a and d, respectively; and of AF 810 sections in Rows b and e. The related bright field images of the AF SlO colony fractions are shown in Rows c and f (legends labeled in black). Bars, 100 pm. 72 A. Nor-l " B. Ver—1 5' (I Figure 4.2 73 Figure 4.3. CLSM irnmuno-fluorescence detection of Nor-1 in A. parasiticus SU-l. Fungal cells derived fi'om colony fraction 82 were irnmuno-labeled with the Nor-1 antibody followed by goat-anti-rabbit IgG conjugated with Alexa 488 probe. Fluorescence images were acquired in a Zeiss LSM5 Pascal CLSM using a Plan- APOCHROMAT (63 X Oil Dic, NA=1.4) objective and zoom level 3. Immuno- fluorescence of Nor-1 was detected in the cytoplasm of SU-l hyphae shown in single optical sections (panels A, B, C, and D). Immuno-fluorescence of Nor-1 was also detected in conidiophores (panel E). The DIG image related to panel B is shown in panel F. Bars=2pm. 74 Figure 4.3 75 Figure 4.4. CLSM immuno-fluorescence detection of Ver-l in A. parasiticus SU-l . Fungal cells derived from colony fiaction 82 were immuno-labeled with the Ver-l antibody followed by goat-anti-rabbit IgG conjugated with Alexa 488 probe. Fluorescence images were acquired in a Zeiss LSM5 Pascal CLSM using a Plan- APOCHROMAT (63 X Oil Dic, NA=1.4) objective and zoom level 3. Immuno- fluorescence of Ver-l was detected in the cytoplasm of SU-l hyphae shown in single optical sections (panels A and B). Immuno-fluorescence of Ver—l was also detected in conidiophores (panels C and E). The DIG images related to C and E are shown in D and F, receptively. Bars = 2 pm. The non-specific signals of Ver-l associated with cell wall, spore coat, or particle structures are indicated with arrows. 76 ‘W 1.33,. ‘3' r U. A Figure 4.4 77 Figure 4.5. Immuno-fluorescence labeling by Ver—l antibodies in the control snain A. parasiticus AF S 10 and the wild-type strain SU-l. Fungal cells derived from colony fraction R2 and S2 (see Material and Methods) were immuno-labeled with the Ver-l antibody followed by goat-anti-rabbit IgG conjugated with an Alexa 488 probe. Fluorescence images were acquired in a Zeiss 210 CLSM using a Plan-NeoFluar (40 X Oil Dic, NA=1.3) objective. (A) Fluorescence signals detected in the control strain AF 810. (C) Fluorescence signals detected in the wild-type strain SU-l. The bright-field images related to A and C are shown in B and D, respectively. Representative non- specific signals are indicated with arrows. Bar, 50 pm. 78 strain AFSIO (fractions R1, R2, and R3) and LW1432 (fractions L1, L2 and L3) did not show significant fluorescent signals compared to samples prepared from the same colony fractions from wild type SU-l (fractions S1, 82, and S3) or CSIO (fractions C1, C2 and C3) (Fig. 4.6). These data are consistent with Western blot analysis of protein extracts isolated fi'om the same colony fiactions (Lee et al., 2002). OmtA was detected in the substrate level mycelium (Fig. 4.7, A; Fig 4.8, B) as well as conidiospore-bearing structures (Fig. 4.7, B; Fig. 4.8, A) located in fiactions S1 and $2 at nearly equal intensity as in substrate level mycelium. The fluorescent signal was confined to discrete areas (patches) within cells (Fig. 4.7, A). However, due to limitations in resolution of CLSM, it was not clear if these areas are associated with particular organelles. Similar samples were also probed with anti-SKL antibodies that are expected to detect proteins targeted to peroxisomes and Woronin bodies (Fig. 4.7, C) (Keller et al., 1991; Valenciano et al., 1998). The observed fluorescent pattern was consistent with localization to these organelles. In addition, we used SYTOX Green to stain nuclei in fungal fractions (Fig. 4.7, D); again the fluorescent pattern was consistent with the expected results. Under the same magnification, immuno-fluorescence labeling pattern of OmtA (Fig. 4.7, A and B) was different from SKL or SYTOX staining (Fig. 4.7, C and D). Neither anti-SKL not SYTOX generated the same “patchy” pattern as anti-OmtA. In summary, the data suggest that, like Nor-l and Ver-l, OmtA is evenly distributed in all cell types in a fungal colony and does not accumulate to highest levels in conidiophores. The fluorescent signal is not consistent with localization of OmtA to only peroxisomes, Woronin bodies, or nuclei. However, the data do not rule out the possibility that OmtA is localized to other cell locations as well as these specific organelles. Quantitative fluorescence intensity analysis. The CLSM fluorescence intensities represent the abundance of Nor-l , Ver-l , and OmtA shown in three colony 79 Figure 4.6. OmtA protein localization in time-fiactionated colonies of A. parasiticus SU-l, AF SlO (afR knockout), CSIO and LW1432 (omtA knockout) grown on YES agar for 72 h. Paraffin-embedded fimgal sections were irnmuno-labeled with affinity purified OmtA PAb (20 pg/ml) followed by Alexa 488 conjugated goat anti-rabbit IgG. (A) Fluorescence images of SU-l (81, S2, and S3) and AF S10 (R1, R2, and R3). (B) Fluorescence images of CSIO (C1, C2 and C3) and LW1432 (L1, L2 and L3). All colony fiactions were analyzed under the same instrument settings. Bars, 50 pm. 80 Figure 4.6 81 Figure 4.7. Protein localization using OmtA PAb and anti-SKL, and detection of nuclei using SYTOX Green. (A) Z-series overlay image of SU-l (fiaction $2) immuno-labeled with OmtA PAb. The image consists of an overlay of ten consecutive optical sections taken at Z—interval of 800 nm (step size). (B) Fluorescence images of vesicles of conidiophores of SU-l (fraction S1) immuno-labeled with OmtA PAb. (C) Irnmuno- labeling of SU-l (fiaction S2) with anti-SKL. The magnification for the left panel is 400X and for right pane] is 2,000 X. (D) Fluorescence image of nuclei in SU-l (fraction 82) stained with SYTOX. The magnification for the left panel is 400X and for the right panel is 2,000 X. The bar in panels A, B, C, and D represents 10 pm. 82 Figure 4.7 Figure 4.8. CLSM immuno-fluorescence detection of OmtA in A. parasiticus SU-l. Fungal cells derived from colony fraction S2 were immuno-labeled with the OmtA antibody followed by goat-anti-rabbit IgG conjugated with an Alexa 488 probe. Single optical sections were acquired using a Zeiss LSM5 Pascal CLSM with a Plan- APOCHROMAT objective (63 X Oil Dic, NA=1.4). (A) Immuno-fluorescence of OmtA was detected in conidiophores. (B) Immuno-fluorescence of OmtA was detected in hyphae. Bars = 2 pm. 84 fiactions of SU-l. Fluorescence intensities were quantified by measuring the pixel numbers on CLSM micrographs using the IQ Master Program (V2.31) image analysis software. The quantitative pixel analysis to analyze distribution of AF enzymes in colony fractions was consistent with qualitative detection by Western blot analysis and fluorescence microscopy (Lee et al., 2002; Lee et al. submitted). The control strain AF S10 did not produce significant levels of Nor-l, Ver-l and, OmtA compared to SU-l. These data confirm that the highest quantity of these three aflatoxin enzymes occurs in fiaction 2 (Table 4.1). DISCUSSION Because a close regulatory association between aflatoxin synthesis and conidiation has been demonstrated in several studies (Guzman-de-Pena and Ruiz-Herrera, 1997; Hicks et al., 1997, Adams et al., 1998), a close temporal and spatial association between AF protein expression and conidiospore development was hypothesized. To test this hypothesis, antibodies against several AF proteins were generated to immuno- localize these proteins. Previous studies (Liang, 1996; Zhou, 1998) utilized a fungal cell preparation technique for immuno-labeling that required digestion of the cell wall using Novozyme (Harris et al., 1994). However, in these early localization studies, variation in cell wall digestion resulted in significant artifacts. The paraffm-embedment and sectioning procedures developed and described in this study successfully preserved fungal mycelium, developmental structures, organelle structure, and protein antigenicity; they also eliminated the need for cell wall digestion, in turn generating consistent and reproducible immuno-labeling results. This technique provides a useful tool to localize proteins and possibly other compounds in fungal cells and colonies. Despite the close regulatory relationship between sporulation and aflatoxin production, my data suggest that Nor-1, Ver-l, and OmtA are not produced exclusively in conidiophores as hypothesized. On the contrary, CLSM micrographs showed that these 85 Table 4.1. Quantative fluorescence intensity analysis of A. parasiticus SU-l and AF S 1 0 immuno-fluorescence labeled with Nor-l , Ver-l and OmtA antibodies. (Fungal colony) (SU-l) (AF S10) Protein / Colony inaction 81 S2 S3 R1 R2 R3 Nor-l 229.5 858.4 260.9 15.2 18.7 62.3 Ver-l 144.4 297.9 166.5 68.7 89.2 42.3 OmtA 120.0 940.1 146.7 4.3 3.9 21.6 * The data represent the average pixel number in 20 images from each colony fraction. 86 proteins are distributed throughout the colony and in both conidiophores and vegetative hyphae. In fraction S2, the highest intensity of AF proteins was detected by CLSM and the signal was observed at similar levels in vegetative hyphae and conidiophores. CLSM detected Nor-1 and Ver-l mainly in the cytoplasm. These data are consistent with data from previous subcellular fiactionation studies along with Western blot analysis. Nor-1 was mainly detected in the cytosolic fiaction (Zhou, 1998), while Ver-l was present in the cytosol (80%) and the rrricrobody fraction (14%) (Liang, 1996). Based on CLSM data generated in the current study, detection of Ver-l in particle-like structures by Liang ( 1996) (presumably microbodies), was an artifact. Cytosolic immuno-fluorescence labeling of Ver—l was specific in SU-l and not observed in the negative control strain AF S10. However, immuno-fluorescence of Ver—l on the cell wall, spore coat, as well as particle-like structures was observed in both SU-l and AF S10. Ver-l displays 56% protein sequence similarity to scytalone reductase (V idal-cros et al., 1994), and 66% identity and 82% similarity to polyhydroxynaphthalene reductase; both enzymes are involved in melanin biosynthesis. Melanins are dark pigments that accumulate in distinct layers of cell walls and spores of many fungi (Wheeler and Bell., 1988). Thus, the PAb against Ver-l used in this study possibly exilribited cross-reactivity to the cell wall and the spores. In addition, immuno-gold labeling of Ver-l observed by transmission electron microscopy (TEM) revealed that the labeling pattern of Ver-l in particle-like structures in SU-l and AF S10 was not specific to Ver-l. The organelle irnmuno-labeled by Ver-l antibody in both SU-l and AF S10 was identified to be the Woronin bodies. The appearance of this organelle was recently demonstrated in A. parasiticus using PAb against a C-terminal tri-peptide Peroxisomal Targeting Sequence 1, SKL (Ser-Lys-Leu) (Lee, 2003). The fourteen percent of the Ver-l associated with the nricrobody fraction detected by Liang (1996) was possibly due to this non-specific interaction. 87 Cells immuno-labeled with OmtA PAb and analyzed by CLSM showed patches of fluorescence within fungal cells suggesting that OmtA is confined to sub-cellular compartments. An alternative explanation is that OmtA is present in the cytoplasm and is thus excluded from cell organelles. To gain more information about sub-cellular localization, we labeled paraffm embedded sections to identify specific organelles; for example anti-SKL antibodies (Keller et al., 1991) were used to label peroxisomes and SYTOX Green was used to detect nuclei. Both probes generated small, regularly shaped signals consistent with the expected organelles, indicating that the double-membrane bound nuclei as well as single-membrane bound microbodies were well-preserved. These observations strongly suggest that we have minimized artifacts due to poor sample preparation. Images using both probes were different than with OmtA PAb. Quantitative fluorescence intensity analysis performed in this study supported qualitative CLSM data for the distribution of AF enzymes in colony fractions. CLSM also generated important imformation for subsequent determination of sub-cellular localization of AF proteins using high resolution TEM. The specimen size required for TEM cell preparation is extremely small (approximately 1 mm2 dimension to be sectioned into 100 nm thickness). Therefore, the finding that the S2 fi'action exhibited the highest density of AF proteins allowed sample acquisition efficiently (from fraction S2) in order to detect these proteins at the sub-cellular level (Lee, 2003). 88 ACKNOWLEDGEMENTS Special thanks to Dr. Shirley Owens (Center for Advanced Microscopy at Michigan State University) for help in quantification of fluorescence intensity and Dr. Jeff Cary (USDA, New Orleans, LA.) for providing the aflR knockout strain, AF S10. Immuno-localization of OmtA is published in Applied and Environmental Microbiology (Lee, L.W., C.H. Chiou, and J .E. Linz. 2002. Function of native OmtA in vivo and expression and distribution of this protein in colonies ofAspergillus parasiticus. Appl. Environ. Microbiol. 68:5718-5727). Immuno-fluorescence analysis of Nor-1 and Ver-l is submitted for publication (Lee, L. W., C. H. Chiou, K. L. Klomparens, and J. E. Linz. 2003. Sub-cellular localization of aflatoxin biosynthetic enzymes in time-dependent fiactionated colonies of Aspergillus parasiticus. Manuscript submitted to Arch. Microbiol.). 89 CHAPTER 5 Distribution and sub-cellular localization of the aflatoxin enzyme versicolorin B synthase (VBS) in time-fractionated colonies of Aspergillus parasiticus ABSTRACT Aflatoxins (AF) are highly toxic and carcinogenic fungal secondary metabolites. Aflatoxin biosynthesis in the filamentous fungi Aspergillus parasiticus and A. flavus requires a complex pathway consisting of at least eighteen enzymatic activities; one of these enzymes, versicolorin B synthase (V BS), catalyzes closure of the bisfuran ring in the racemic versiconal hemiacetal substrate to form versicolorin B. Bisfuran ring formation is important for AF toxicity as the structure is required for the subsequent creation of AFB1-8,9 epoxide, a potent alkylating agent. I analyzed the localization of VBS in the AF producing strain A. parasiticus SU-l grown on solid media using a colony fiactionation technique developed previously. A highly specific polyclonal antibody, raised against a Maltose Binding Protein-VBS firsion protein synthesized in Escherichia coli, was utilized to detect VBS in SU-l grown on a rich solid medium via irnmuno- fluorescence confocal laser scanning microscopy (CLSM) and irnmuno-gold transmission electron microscopy (TEM). VBS was found not only in vegetative hyphae but also in asexual developmental structures, called conidiophores. Western blot and CLSM analyses demonstrated the highest abundance of VBS in colony fraction S2 where cells had grown for 24 to 48h; this fraction also contained the highest levels of newly developed conidiophores. The highest abundance of AF Bl also was detected in fiaction S2, consistent with VBS abundance. At the sub-cellular level, CLSM and TEM detected VBS at similar levels in the cytoplasm and in ring-like structures surrounding nuclei; it is uncertain if enzymatically active VBS is present in either or both locations. 90 INTRODUCTION Aflatoxins (AF), fungal secondary metabolites produced by filamentous fungi including Aspergillus parasiticus, A. flavus, A. nomius, and A. tamarii, are highly toxic and carcinogenic compounds (Cary et al., 2000; Eaton & Groopman, 1994; Cotty and Cardwell,l999). AF contamination in commercial crops including com, peanuts, tree nuts, and cottonseed has generated significant economic and human health concerns throughout the world (CAST, 2003). Among the members of the AF family, AFB] is considered the most dangerous as it is efficiently bioactivated by cytochrome P450 enzymes (particularly in the liver) to form the AF 31- 8,9 epoxide, a potent alkylating agent. The epoxide interacts with the N—7 position of guanine residues in double stranded DNA to generate AFB1-N7-Gua DNA adducts (Bailey, et al., 1996; Mace, et al., 1997; Aguilar et al., 1993; Eaton and Gallagher,1994). The tumor suppressor gene p53 appears to be a frequent target of AF B] since mutation of p5 3 is often seen in hepatic-cancer patients with high levels of AF B] exposure (Yang, et al., 1997). Our goal is to eliminate AF contamination in agricultural commodities and in turn to ensure product safety and consumer health. AF biosynthesis utilizes a complex pathway requiring at leastl 8 enzymatic activities (Minto & Townsend, 1997; Payne, 1998; Bhatnagar et al., 2003). The structural - genes that encode these aflatoxin enzymes, including nor-1 (Trail et al., 1994), ver-l (Skory et al., 1992), omtA (Yu, et al., 1993), vbs (Silva et al., 1996), as well as one regulatory gene aflR (Chang et al., 1993; Woloshuk et al., 1994), are clustered in a 75 kb chromosomal region (Trail et al., 1995; Minto & Townsend, 1997; Payne and Brown, 1998; Cary et al., 2000; Bhatnagar et al., 2003). Although most AF biosynthetic genes have been cloned and characterized, the mechanisms that regulate AF biosynthesis are not completely understood. The positive-acting GAL4-type transcription factor AflR is required for AF biosynthesis (Chang et al., 1993; Ehrlich et al., 1999a; Ehrlich, et al., 2003). A recent report also described cluster-dependent positional effects on AF 91 biosynthesis. The nor-1 promoter is positively regulated within the AF gene cluster whereas outside the cluster (niaD or pyrG), nor-1 promoter activity decreases dramatically (Chiou et al., 2002). To date, little is known about how or where aflatoxin enzymes interact at the cellular level to synthesize AFB]. To address this issue, our laboratory developed a colony fiactionation protocol and irnmuno-fluorescence and immuno-gold based microscopic methods to localize proteins involved in AF biosynthesis. Previously, we analyzed the distribution of OmtA, an enzyme involved in the later stages of AF biosynthesis (Lee et a]. 2002). The highest abundance of OmtA was detected in firngal cells grown for 24 to 48 h (fiaction S2) in A. parasiticus strain SU-l grown on a solid medium (YES). Immuno-fluorescence confocal laser scanning microscopy (CLSM) detected OmtA in the cytoplasm and the protein frequently was confined to discrete areas in cells in this colony fiaction. Transmission electron microscopy (TEM) together with immuno-gold labeling suggested that OmtA and the aflatoxin enzymes Nor-1 and Ver-l are found in the cytoplasm; of particular interest, OmtA was also detected at high levels in vacuole-like organelles in cells residing near the substrate surface of the colony (Lee et al., manuscript submitted). These observations caused us to propose that early aflatoxin enzymes function in the cytoplasm while later pathway enzymes are localized to organelles either to protect cells from the toxic effects of late pathway intermediates or to reduce the level of aflatoxin synthesis via protein turnover. In this report, we investigate the localization of another key enzyme involved in AF biosynthesis, versicolorin B synthase (V BS), in A. parasiticus to determine if the data support or refute the proposed model. VBS catalyzes the cyclization of the racemic versiconal hemiacetal substrate and closure of the bisfuran ring to form versicolorin B (Lin and Anderson, 1992; Anderson and Chung, 1994; Silva et al., 1996). The bisfuran ring structure is required for the subsequent creation of the AF B1-8,9 epoxide and toxicity. vbs encodes a protein of 634 92 11111110 301 116. 33 1. 11111116] and does The pied VBS 210] by $08 atVI mass '1‘. VBS 1 lltltll main' trans IECO loca in r leni 811g amino acids (V BS) that contains 3 putative N—glycosylation sites at amino acid residues 116, 221, and 507 (Silva et al., 1996). VBS displays sequence similarity to several flavin—dependent oxidases and dehydrogenases but it does not catalyze a redox reaction and does not bind FAD or F MN to its truncated FAD binding domain (Silva et al., 1996). The predicted molecular mass of VBS based on the cDNA sequence is 70 kDa, yet native VBS appears as a homo-dimer with a monomeric molecular mass of 78 kDa (estimated by SDS-polyacrylamide gel electrophoresis). N-glycosidase F treatment of native VBS and VBS heterologously expressed in yeast revealed that the discrepancy in molecular mass is attributed to N-glycosylation (Silva et al., 1997). The N-tenninus of the native VBS was also reported to be post-translationally modified (Silva et al., 1996). VBS heterologously expressed in E. coli was not active, whereas VBS expressed in yeast maintained its activity, suggesting that N—glycosylation and perhaps other post- translational modifications are essential for VBS activity (Silva et al., 1997). In the current study, a highly specific PAb was produced in rabbits using a recombinant maltose binding protein (MBP)-VBS fusion expressed in E. coli. VBS localization was conducted by fluorescence-based CLSM and irnmuno-gold TEM. VBS was observed throughout a 72 h-old colony and was detected not only in vegetative hyphae but also at similar levels in conidiophores. The highest levels of VBS, AF B], and newly developed conidiophores were present in colony fiaction S2 where the cells had grown for 24 to 48 h. These observations suggest a strong spatial association between secondary metabolism and asexual sporulation; a regulatory and temporal association between these processes was reported previously (Adams et al., 1998; Guzrnan-de Pann and Ruiz-Herrera, 1997; and Calvo et al, 2002). At the sub-cellular level, VBS was detected in the cytoplasm and at similar levels in ring-like structures surrounding nuclei; based on morphology, these structures were tentatively identified as endoplasmic reticulum (ER). Computer-based sequence analysis suggested that VBS contains a putative signal peptide at the N-terminus. Because VBS is 93 N-link glycosylated, we propose that VBS is associated with the ER/golgi secretary pathway for transient co- and post-translational modification and localizes to the cytoplasm where it participates in aflatoxin synthesis. However, based on the data, it is currently uncertain if the ER, the cytoplasm, or both locations represent the final destination for VBS activity. MATERIALS AND METHODS Fungal strains and growth conditions. The wild type aflatoxin producing strain A. parasiticus SU-l (NRRL5862, ATCC 56775) and a non-aflatoxin producing strain AF S10 (aflR knockout mutant, kindly provided by Dr. Jeffrey Cary, USDA, New Orleans, LA) (Cary et al., 2002) were used to study VBS and other AF enzymes. For liquid culture, a conidiospore suspension (2 x 107 conidiospores) was inoculated into a 250 ml Erlenmeyer flask containing 100 ml YES media (2% yeast extract, 6% sucrose, pH5.5) and incubated at 29°C with constant shaking (150 rpm) for 24 h to 72 h. For solid culture, conidiospores (2 x 105/pl) were center inoculated onto YES agar (1.5%) overlaid with sterile cellophane membranes and cultured in the dark at 29°C for 24 h, 48 h, or 72 h. Production of antibodies against native VBS. Native VBS purified fi'om A. parasiticus SU-l was kindly provided by Dr. Craig Townsend, Johns Hopkins University (McGuire et al., 1996). Approximately 30 pg of protein was mixed in a l: 1 (V N ) ratio with TiterMax adjuvant (Cthx, Norcross, GA) and used to immunize two New Zealand White rabbits. After 30 d, a booster injection with the same dose of VBS/TiterMax emulsion was utilized. Twenty-one (I later, the anti-serum was collected and the serum IgG was precipitated in 33% ammonium sulfate according to standard protocols (Ausubel et al., 2003). 94 Generation of MBP (Maltose Binding Protein)-VBS fusion protein. Three possible reading frames of the vbs exon H fragment were amplified by PCR based on the published sequence (Silva et al., 1996) using A. parasiticus SU-l genomic DNA as a template. To generate VBS clones with each of the three possible reading fiames, PCR amplification was conducted using three forward primers containing 23 nucleotides including an EcoRI restriction endonuclease site (underlined) [(1) 5' agaGAATTC- agAGGCTCGAAAGG-3', (2) 5' tagGAATTCcag AGGCTCGAAAG-3', and (3) 5'- ctaGAATTC-tcagAGGCTCGAAA—3'] and one reverse primer (5'-ATCAAGCTT- GGTCTACTGCCCAG-3') containing a HindIII restriction endonuclease site (underlined). Each 1.4 kb vbs fiagment was individually cloned into the pMAL-c2 expression vector (New England Biolabs, Beverly, Mass.) to generate expression plasmids, pMAL-c2-vbsl , pMAL-c2-vbs2, and pMAL-c2-vbs3. These plasmids were transformed into E. coil to express the 94 kDa MBP-VBS fusion protein. The fusion protein was induced in large scale culture by addition of 0.5 mM IPTG at optimum cell density (O.D.600 approximately 0.6). Cells were disrupted by a brief freeze and thaw cycle followed by sonication on ice (Sonifier cell disruptor W-350; Fisher Scientific, Pittsburgh, PA). The induced soluble MBP-VBS was purified with amylose resin according to the manufacturer's instructions (New England Biolabs, Beverly, Mass). Production and purification of PAb against MBP-VBS. Purified recombinant MBP-VBS was injected into two New Zealand White rabbits. Each rabbit was immunized subcutaneously with approximate 50 pg purified MBP-VBS mixed at a l: 1 (V N ) ratio with TiterMax adjuvant (Cthx, Norcross, Ga). After 32 d, a booster injection with the same amount of protein/TiterMax emulsion was utilized. Anti-sera collection was performed after an additional 30-d and approximately 60 ml of anti-sera were obtained from each rabbit. Serum IgG was precipitated in 33% ammonium sulfate as described previously (Ausubel et al., 2003). Because immuno-gold TEM studies require low background signals, anti-VBS IgG was further purified by affinity 95 chromatography. The anti-VBS IgG was passed through a column containing proteins extracted from AF S10 (non-AF producing, does not synthesize VBS) grown in liquid YES medium for 48 h as described by Lee et al., (2002); the extracted proteins were conjugated to a resin rrrixture of Affi-Gel 10 and Affr-Gel 15 (BioRad, Hercules, Calif). The flow-through fraction (containing the cleaned-up IgG) was dialyzed against phosphate buffered saline (PBS)(Ausubel et al., 2003) and used for immunogold-labeling on ultra-thin fungal sections (described below). Time-dependent fractionation of fungal colonies. A 72 h-old colony grown on YES agar was fiactionated according to Lee et al. (2002). Briefly, at 24 h intervals during growth, the culture plate was marked on the bottom to indicate the location of the outer margin of the colony. After 72 h of culture, the colony was dissected into 3 fractions (concentric rings) using these marks. Fractionation of 72 h-old SU-l colonies (AF producing strain) generated fractions S1 (48 to 72 h), S2 (24 to 48 h), and S3 (0 to 24 h). Fractionation of AF S 1 0 (aflR knockout mutant) resulted in analogous fractions R1, R2, and R3 (using the same scheme). Fungal protein extraction. Mycelia from liquid culture were collected by filtration through miracloth (Calbiochem, La Jolla, Calif.) by aspiration. Mycelia fi'om solid culture fractions were generated as described above and peeled directly from the cellophane membrane. The fungal mycelia were then pulverized under liquid nitrogen with a mortar and pestle and fungal proteins were extracted using TSA buffer (2 mM Tris-Cl, pH 8.0; 40 mM NaCl; and 0.025% sodium azide) containing complete protease inhibitors (2 tablets/10 ml buffer, Roche Molecular Biochemicals, Indianapolis, Ind.). The resulting extracts were centrifuged at 12,000 g for 20 min at 4°C to remove mycelia. The supernatant (fungal protein extract) was transferred to microcentrifuge tubes in small aliquots and stored at -80°C. Protein concentration was determined using Bio-Rad protein assay reagent (Bio-Rad, Hercules, Calif.) and bovine serum albumin (BSA, Sigma-Aldrich, St. Louis, Mo.) as a standard. 96 SDS—Polyacrylamide Gel Electrophoresis (SDS-PAGE) and Western blot analysis. SDS-PAGE was conducted using a standard procedure (Ausubel et al., 2003) to analyze expression of recombinant MBP-VBS and other fungal proteins. The molecular mass was determined by comparison to Low-Range Prestained SDS-PAGE standards or SDS-PAGE Molecular Mass Standards-low range (Bio-Rad, Hercules, Calif.). For immunoblotting, 30 to 60 pg of fungal proteins were separated by SDS-PAGE (electrophoresis on 10% or 12% polyacrylamide gels) and transferred onto PolyScreen PVDF membrane (Dupont Biotechnology Systems, NEN Research Products, Boston, Mass). The membranes were blocked with 1% casein in TBST (10 mM Tris-Cl, pH 8.0; 150 mM NaCl; and 0.05% Tween 20) at RT for 1 h. The blots were probed with anti- VBS IgG (2 pg/ml) followed by goat anti-rabbit IgG alkaline phosphate conjugate (Sigma-Aldrich, St. Louis, Mo.) as the secondary Ab probe (1: 5,000 dilution) at RT for 30 min. Reactive proteins were visualized using the BCIP/NBT colorimetric detection system (Roche Molecular Biochemicals, Indianapolis, Ind.). Detection of AFB] in colony fractions. (1) Thin Layer Chromatagraphy (TLC). Fungal colony fiactions (from replica plates) were placed into glass scintillation vials for measurement of dry weight, followed by extraction of AF B] using 5 ml chloroform. The chloroform extract (5 pl) was dotted onto silica plates (LHPKD Silica Ge] 60 A, 10 cm x 10 cm; Whatrnan International Ltd., Maidstone, England) for TLC with a solvent system consisting of 5% acetone/ 95% chloroform. AF B] was detected under long-wave UV light. (ii) Enzyme-Linked-Immunosorbent—Assay (ELISA). AFB1 in each colony fraction was quantified by a competitive-sandwich ELISA method described previously (Pestka et al., 1980). One ml of AF extract (in chloroform) was evaporated under N2 and re-suspended in 0.5 ml methanol. The assay plates were coated with anti-AFB] antibodies (Sigma-Aldrich), incubated at 42°C over night and blocked with 1% BSA-PBS at 37°C for 30 to 60 min. Serially diluted samples as well as AF B] standard (Sigma-Aldrich) were mixed with the AF B1-horse radish peroxidase (HRP) 97 conjugates (kindly provided by Dr. Pestka, Michigan State University) as a competitor, followed by incubation with the substrate ABTS (2,2'-Azino-bis(3)-ethylbenzothiazoline- 6-sulfonic acid) to generate a detectable signal at an absorption wavelength at 405 nm. The concentration of AFB] was determined according to the standard curve. The levels of AFB1 measured in a colony fiaction were normalized to dry weight (AF B1 [pg]/ fraction weight [g]). Preparation of paraffin-embedded fungal sections and immuno-fluorescence labeling. Fungal samples were derived from a whole 48 h colony or colony fractions generated from a 72 h-old colony. The detailed procedure to prepare the paraffin- embedded fungal sections for immuno-flluorescence microscopy was described previously (Lee, et al., 2002). Specimens (5 mm) were cut from each colony fraction, fixed with Streck tissue fixative (Streck Laboratory Inc., Omaha, Nebr.) at 4° C overnight, followed by dehydration in a graded series of ethanol, and finally embedded in Paraplast (Sigma-Aldrich). Fungal sections (45 pm thickness) were generated using a tissue section nricrotome (AO Spencer 820 nricrotome, Fisher Scientific) and attached to poly- L-Lysine (Sigma-Aldrich) coated coverslips (22 mm square) (Corning Glass Works, Corning, NY). Prior to immuno-labeling, sections required removal of paraffin by 100% xylenes, and re-hydration with decreasing concentrations of ethanol, followed by an antigen retrieval procedure (heating the sections in 10 mM citrate buffer, pH 6.0 at 95° C for 5 min and cooling at room temperature for 20 min). After blocking the samples with TBS containing 1% BSA and 0.1% saponin at 4°C overnight, the samples were immuno- labeled with one of the following primary antibodies: 20 pg/ml of anti-VBS IgG (generated in this paper); 20 pg/ml purified anti-OmtA IgG (Lee et al., 2002): 20 pg/ml anti-Ver—l IgG (Liang et al., 1997); or a 1:500 dilution of Nor-l antiserum (Zhou, 1997; Lee et al., manuscript submitted) at room temperature for 1.5 h, followed by incubation with the secondary antibody conjugated with a fluorescent probe (Goat anti-rabbit IgG- 98 Alexa 488 conjugate [5 pg/ml]; Molecular Probes; Eugene, OR), at RT for l h. Selected coverslips were subsequently stained with SYTOX orange dye (Molecular Probes; Eugene, OR) to localize the nucleus. Confocal Laser Scanning Microscopy (CLSM). To determine the distribution of VBS in colony fiactions, fluorescence image acquisition and quantitative intensity analysis of one colony of SU-l and AF S10 was conducted using a Zeiss Axioskop microscope equipped with a Meridian Insight confocal laser scanning system (Meridian Inc. Okemos, Mich.) as described by Lee, et al. (manuscript submitted). Due to damage to the detector in this microscope, the remaining samples were viewed using a Zeiss LSM5 Pascal Laser Scanning Microscope with the same 40X objective lens (Plan NEOFLUAR 40 X/ 1.3 Oil Dic, NA=1.3). High resolution single confocal images, extended focus image made from Z-stacks, or DIC (Differential Interference Contrast) images were taken using a Zeiss Plan-APOCHROMAT (63 X/ 1.4 Oil Dic, N.A.=l .4) objective. Detection of the Alexa 488 fluorescence probe (A bs/Em 495/519 nm) was conducted using a BP 505-530 emission filter set under excitation with the 488 nm argon- ion laser line and detection of SYTOX orange (Abs/Em 547/570 nm) was performed using the BP 560-615 nm emission filter set under the 543 nm helium-neon laser. Quantitative fluorescence intensity analysis. Samples derived fiom three SU-l replicate colonies were immuno-labeled with anti-VBS or OmtA and subjected to fluorescence intensity analysis by quantifying the pixel number in 20 randomly selected images. Pixels were analyzed using IQ Master Program (V2.31) image analysis software that accompanied the Meridian Insight laser-scanning microscope (Meridian Instrument, Inc., Okemos, Mich.) or on a Macintosh using a NIH image program (developed at the US. National Institutes of Health and available on the web: httpz/lrsb. info. nih.gov/nih- irnagel). For the samples derived from the same immuno-labeling experiment (both SU-l and the negative control AF S 1 0), the CLSM images were acquired under identical 99 instrument settings. For each colony fraction, average pixel number in the 20 images was reported. Statistical analysis. Fluorescence intensity data were statistically analyzed by a complete random design with a split-plot model using SAS software (SAS Institute Inc. Cary, NC) or two-way ANOVA using SigrnaStat (SPSS Inc. Chicago, Ill.). The amount of AF B1 in three colony fractions was analyzed by one-way AN OVA using SigrnaStat. Statistical significance among samples was defined by a P-value not greater than 0.05. Preparation of ultra-thin fungal sections, immuno-gold labeling, and TEM analysis. Ultra-thin sections were generated from colony fiactions 82 and R2. Sample preparation and immuno-gold labeling followed the procedures described by Lee et al. (manuscript submitted). Briefly, samples were prepared by a freeze-substitution method (F legler et al., 1995) and embedded in LR White resin (Ted Pella Inc., Redding, CA). Ultra-thin sections (90-100 nm) were generated using an MTX ultra-microtome (RMC, Tucson, AZ) and mounted onto formvar-coated nickel grids. Sections were irnmuno- labeled with column-purified anti-VBS IgG (50 pg/ml), followed by goat anti-rabbit IgG conjugated with 10 nm gold (Ted Pella. Inc., Redding, CA); sections were subsequently post-stained with 3% uranyl acetate for 20 nrin. Samples were observed using a JEOL IOOCX II transmission electron microscope (Tokyo, Japan) at 100 kV acceleration voltage . RESULTS Generation of PAb against native fungal VBS. An anti-VBS PAb preparation was generated using native VBS purified from A. parasiticus SU-l as an imrnunogen. Only 60 pg of VBS protein was available for immunization split between a primary and booster injection into two rabbits. The resulting PAb recognized VBS heterologously expressed in yeast but also appeared to cross-react with fungal proteins other than VBS. To remove cross-reactive antibodies, we pre-absorbed the IgG fraction with proteins in a 100 fungal extract isolated from SU-l grown for 24 h in liquid YES culture (conditions which generate little detectable VBS expression). The resulting IgG recognized a single sharp band in heterologous VBS expressed in yeast (78 kDa) but several proteins in SU-l. Among these SU-l proteins, two closely associated protein bands (one band equal to 78 kDa and a smaller band of approximately 72 kDa) appeared to be VBS based on pattern of accumulation; they were only detected at 48 h and 72 h but not at 24 h in the SU-l extracts (Fig. 5.1). Although this clean-up procedure reduced the cross-reactivity, the remaining background signals could generate considerable artifacts during an immuno- cytochenristry study. Thus, anti-VBS antibodies with higher specificity were produced using an alternate procedure. Expression of MBP-VBS fusion protein and production of PAb against VBS. To assure production of full-length protein, we amplified (PCR) three reading frames of VBS exon II based on the published sequence (Silva et al., 1996) and cloned these PCR fiagrnents into the pMAL-c2 expression vector to generate the MBP-VBS fusion proteins. E. coli harboring pMAL-c2-vbs1 expressed the expected 94 kDa MBP-VBS firsion protein (Fig. 5.2) whereas E. coli harboring pMAL-c2-vbs2 and pMAL-c2-vbs3 generated pre-maturely terminated protein products (molecular masses of 45 kDa and 55 kDa, respectively). Control E. coli carrying the backbone vector, pMAL-c2 (malE::lacZ) produced a 52 kDa MBP-fl-galactosidase fusion protein (Fig. 5.2). The MBP-VBS fusion protein was purified and injected into two rabbits to generate PAb. We initially intended to isolate VBS by removing MBP from the fusion protein using factor Xa (IEER, Ile-Glu-Glu-Arg recognition). However, treatment fiagrnented VBS as well (Fig. 5.3) even though there is no factor Xa cleavage site found in the VBS exon 11 sequence. Thus cleavage within VBS appeared to be non-specific. We consequently used the entire MBP-VBS fusion protein to generate PAb. ’ Based on Western blot analysis, PAb raised against MBP-VBS displayed a higher specificity (Fig. 5.4) than PAb raised against the native VBS (Fig. 5.1). The serum 101 SU1 l 1 Yeast 24h 48h 72h vas m 4.4. 75 kDa—p «—--- 2:1,; Figure 5.1. Western blot analysis of VBS detected by PAb raised against native fungal VBS. Fungal proteins were extracted from A. parasiticus SU-l grown in liquid YES medium for 24 h, 48 h, and 72 h. Proteins were separated by SDS-PAGE (12% acrylamide gel), transferred onto PVDF membrane and probed with rabbit serum IgG raised against native VBS (2 pg/ml) that had been preabsorbed with protein isolated fi'om SU-l cultured in liquid YES medium for 24 h. Heterologous VBS expressed in yeast is shown as a control. The apparent molecular mass for the target protein is 78 kDa. 102 pMAL-cz pMAL-c2 pMAL-c2 pMAL-c2 -vbs3 MW1 -vbs1 -vbs2 uwz IPTG: '- +II. +11- +31- +1 “may; 135,1 , ~ Wm”; .. --.97 MBP-VBS ~ . “108 ~..—v I 31 21‘ .023 .414 Figure 5.2. SDS-PAGE of recombinant MBP-VBS fusion proteins. Three reading flames of vbs exon 11 sequences were amplified from genomic DNA of SU-l and PCR fiagments cloned into pMAL-cZ expression vector generating three flame-shift plasmids, pMAL-c2-vbsl, pMAL-cZ-vsz, and pMAL-cZ-vbs3 (see Materials and Methods) that were subsequently transformed into E. coli DH501. The in-fiame MBP-VBS had the predicted molecular mass of approximately 94 kDa (a**). As a control, E. coli carrying the backbone vector, pMAL-cZ containing malE::lacZ, expressed only the MBP-fl- galactosidase fusion protein (52 kDa). The induced fusion proteins (IPTG+) are indicated with symbols * or **. 103 Figure 5.3. MBP-VBS treated with Factor Xa to generate MBP (42 kDa) and VBS (52 kDa). The purified MBP-VBS (94 kDa) produced in E. coli was incubated with Factor Xa at RT for 0 h, 2 h, 4 h, 8 h, and 24 h. As a control, an equal amount of MBP-VBS without incubation with Factor Xa was set at RT for 8 h and 24 h. (A) SDS-PAGE analysis. Samples were separated by electrophoresis on a 10% polyacrylamide gel and stained with coomassie blue. Molecular mass markers are shown at both sides of the gel. (B) Western blot analysis. Samples separated by SDS-PAGE were transferred onto PVDF membrane. The blot was incubated with anti-VBS IgG (2 pg/ml, raised against native VBS) and Goat anti- rabbit IgG conjugated with alkaline phosphatase, followed by colorimetric detection by BCIP/NBT. 104 Xa factor treated no Xe I l I I MW 0h 2h 4h 8h 24h 24h 8h MW Him—:25“ ~—-- --~-- --=-.1 :2: 4— MBP-VBS(94kDa) 1".E‘fi ' 1 -... .. r '1':- ""1 43...- - *r 7 “-7- 7-“- --~—- ,, e—*VBS(521toa) "' "" ”*5 "— MBP(42 kDa) 34-1. .. . 25. .. 20-... s on 211 an 511 24hcontrol 2:13-23— —-_-3 2'994—uep-vas ”‘— .'- _ u. 1.. . “#VBS - 34 - 25 - 20 Figure 5.3 105 recognized VBS expressed in yeast and native VBS present in a 72 h culture of SU-l grown in YES (Fig. 5.4, A). SU-l cultured in YEP medium (contains peptone instead of sucrose and does not support AF biosynthesis) did not produce detectable VBS (Fig. 5.4, B). VBS was not observed in a 24 h SU—l in YES liquid-shake culture (Fig. 5.4, A and C) while VBS was abundant at 48 h (Fig. 5.4, C). Expression of VBS in A. parasiticus grown on solid medium. In liquid-shake culture, fungal cells are maintained as vegetative hyphae and also produce secondary metabolites at highest levels during a transition between active growth and stationary phase. However, normal developmental processes such as asexual sporulation are not observed under such growth conditions. Thus, I cultured A. parasiticus on a solid medium that supports native morphology and asexual development including the genesis of conidiophores and asexual conidiospores. Western blot analysis using the highly specific VBS PAb (MBP-VBS) was conducted on fungal total proteins isolated from SU-l and AF SlO colonies grown on solid YES medium for 24 h, 48 h, and 72 h (Fig. 5.5). No VBS was detected in the control strain AF 810 (Fig. 5.5, A and B). A 24 h-old SU-l colony did not contain detectable VBS whereas colonies cultured for 48 h and 72 h did contain detectable levels of VBS (Fig. 5.5, A), a result similar to liquid-shake culture (Fig. 5.4, C). We then analyzed time-dependent expression of VBS; a 72 h—old SU—l colony grown on solid YES was fractionated into S] (48 to 72 h), S2 (24 to 48 h), and S3 (0 to 24 h); AF S10 was similarly fractionated to generate R1, R2, and R3. Western blot analysis detected VBS in all three SU-l fractions but not in any AF S1 0 fraction (Fig. 5.5, B). Fraction 82 appeared to contain the highest abundance of VBS. Similar results for OmtA (Lee et al., 2002), Ver-l , and Nor-1 were reported previously (Lee et al., manuscript submitted). The intact form of VBS (78 kDa) appeared in all SU-l samples and little degradation was apparent. 106 A B C Yeast SU1 Yeast SU1 SU1 VBS YES VBS l—‘—'I YES l—'l YEP YES r—-. _ I 24h 72h 72h 72h 24h 48h 72h 103 as . as If 28 a1 22' d ”j Figure 5.4. Western blot analysis of VBS produced in liquid shake culture using highly specific PAb raised against MBP-VBS. (A) Analysis of protein isolated from A. parasiticus SU-l grown in liquid-shake YES (yeast extract sucrose) culture for 24 h or 72 h. (B) Analysis of proteins isolated from SU-l grown in liquid-shake YES or YEP (yeast extract peptone) culture for 72 h. In panels A and B, samples were electrophoresed on 10% polyacrylamide gels and subsequently probed with a 1:5000 dilution of anti-serum. VBS expressed in yeast was used as a control. (C) Time-course expression of VBS in SU-l grown in liquid-shake YES culture for 24 h, 48 h, and 72 h. Proteins were separated by electrophoresis on a 12% polyacrylamide gel and subsequently probed with 2 ug/ml of serum IgG. The primary signal is shown with the molecular mass of 78 kDa. 107 A su1 AFS10 3 5m AFS10 24h 48h 72h 48h 72h MW ;1 S2 :3 IF" R2 R2 MW 73., '-*:2::7skoa WW7 .22--—.—.-m _66 —>-—-—-m “liaise - 45 “3:13? 45 -30 £Q31 '20 Figure 5.5. Western blot analysis of VBS expressed in A. parasiticus SU-l colonies (solid culture) and colony fiactions. (A) VBS expressed in A. parasiticus colonies of different ages. Proteins were isolated from colonies of SU-l and AF SlO (aflR knockout strain) grown on solid YES medium for 24 h, 48 h and 72 h. (B) Distribution of VBS in time-dependent colony fractions. The 72 h-old colonies of SU1 and AFSIO grown on YES solid media overlaid with a cellophane membrane were divided into 3 fractions based on time of growth (81: 48-72 h; S2: 24-48 h; S3: 0-24 h) according to Materials and Methods. Approximately 60 pg of fimgal proteins were loaded in each lane of a 12% polyacrylamide gel. Both blots were probed with specific anti-VBS IgG (2 ug/ml). The molecular mass markers (MW) are shown on the right of each blot. Native VBS has a mass of 78 kDa. 108 Confocal Laser Scanning Microscopy (CLSM) of time-dependent expression of VBS in a 72 h-old fungal colony. Fungal tissues from three colony fractions (72 h-old SU-l or AF SlO colonies grown on YES solid medium) were fixed and embedded in parafi'm to generate sections for immuno-labeling. CLSM detected significant fluorescence in all SU-l fi'actions (Fig. 5.6, A), while little fluorescence was observed in AF S10 (F ig. 5.6, B). Thus, the corresponding AFSIO bright field images are shown to indicate that similar numbers of the AF S10 cells were analyzed (Fig. 5.6, B, insets). Fluorescence intensity analysis via CLSM (Fig. 5.6, A) appeared to parallel Western blot analysis (Fig. 5.5, B) in that highest intensities were observed in fraction S2; here, VBS was present in both substrate level mycelia (Fig. 5.8, A, B, and C) and in conidiophores (Fig. 5.8, D). To confirm this observation, we immuno-localized VBS and OmtA in 2 additional colonies and analyzed the fluorescence intensity quantitatively. Fluorescence intensity quantification and statistical analysis. Pixel analysis was performed on images generated from three independent colonies (designated A, B, and C)(Table 1). Pairwise comparison confirmed that the levels of OmtA in fraction S2 (cultured for 24 to 48 h) were significantly higher than the other two fractions of SU-l (S2 vs. S1, P= 0.0003; S2 vs. S3, P < 0.0001). The levels of VBS detected by fluorescent antibody were consistently higher in fraction S2 than in fi'actions SI and S3 although the difference was not statistically significant (82 vs. 81, P= 0.3529; 82 vs. SB, P= 0.1230). Production of AFB] in colony fractions. TLC analysis did not detect AFB1 in the control strain AF 810 while significant quantities of AF B1 were detected in all 3 SU-l fiactions (Fig. 5.7). The quantity of AF B] extracted from fraction S2 was higher than from fractions 81 and S3 based on TLC (Fig. 5.7) and ELISA (Table 5.2), yet the difference was not statistically significant (P: 0.0589). Sub-cellular localization of VBS. Using CLSM and anti-VBS probes, fluorescent signals appeared in the cytoplasm of hyphae of SU-l in fraction S2 and these 109 A SU-1 Figure 5.6. CLSM of VBS protein distribution in time-fractionated colonies of A. parasiticus SU-l (SI, 82, S3) and AFSIO (R1, R2, and R3) grown on YES agar for 72 h. Paraffin-embedded fimgal sections derived from various colony fractions (see Material and Methods) were immuno-labeled with anti-VBS IgG (20 ug/ml) followed by Alexa 488 conjugated goat-anti rabbit IgG. The associated bright field images (insets in B) demonstrate the numbers of AF 810 cells being analyzed. Bars, 10 pm. 110 AFS10 SU1 Std R1 R2 R3 Std S1 82 S3 Std AFB1 —> AFG1 —’ Figure 5.7. Thin Layer Chromatagraphy analysis of AFB1 accumulation in colony fractions. Colony fractions of A. parasiticus SU-l (Sl, S2, and S3) and AF 810 (R1, R2, and R3) were extracted in 5 ml chloroform and 5 ul were resolved on silica plates. Aflatoxin standard (Std) is shown on both sides and in the middle lane. Fluorescence of AF was detected under long-wave UV light (2F 360 mn). 111 Figure 5.8. CLSM immuno-fluorescence of VBS of A. parasiticus SU-l grown in solid culture. Immune-fluorescence of VBS was detected in the hyphae of SU-l shown as extended focus images of CLSM Z-stacks (A and B) and a single optical section (C). (D) Presence of VBS in a conidiophore shown in an extended focus image. Bars, 5 um. Table 5.1. Quantitative fluorescence intensity analysis“ of A. parasiticus SU-l and AF SlO labeled with polyclonal antibodies to VBS and OmtA. (Fungal colony) (SU-l) (AF S10) Protein / Colony frafction 81 S2 S3 R1 R2 R3 VBS Colony .49 18.46 42.37 15.02 9.48 9.77 8.06 ab 1 1.60 29.00 6.22 5.75 5.73 7.25 Cb 6.48 1 1.92 5.05 3.00 3.26 3.89 OmtA Colony .19 7.53 58.79 9.20 0.29 0.28 1.38 ab 15.22 37.08 2.55 1.04 1.65 2.85 Cb 9.27 28.28 2.29 0.86 1.84 1.69 * The data represent the average pixel number of 20 images fiom each colony fraction. a. Images of colony A were generated by the Meridian Insight confocal laser scanning system and pixel analysis by NIH image program. b- Images of colony B and C were generated using the Zeiss LSM5 Pascal Laser Scanning Microscope and pixel analysis by NIH image program. 113 Table 5.2. Quantitative detection (ELISA) of AF B1 in colony fractions of SU-l cultured on YES solid media for 72 h. (Fungal strain) (SU-l) Colonv fraction 81 S2 S3 ColonyA *AFBI(118/g) 295.59 490.41 245.83 ColonyB 204.82 451.24 252.82 * The levels of AF B1 measured in the colony fraction were normalized to AF B1 (ug)/ fraction weight (g). 114 Figure 5.9. CLSM dual detection for localization of VBS and nuclei in A. parasiticus SU-l. Fungal sections generated from colony fraction 82 were immuno-labeled with anti-VBS IgG (20 ugml) followed by Alexa 488 conjugated goat-anti rabbit IgG (represented in green) and further stained with SYTOX orange dye to contrast the localization of nuclei (represented in red). Irnmuno-fluorescence of VBS (single green channel) is shown in panels A, C, E, and G. The ring-like fluorescence signals are indicated with arrows. The merged dual channel images indicating localization of VBS (represented in green) and nuclei (represented in red) are shown at the right of each of the image pairs (panels B, D, F, and H). Immuno-fluorescence pattern of OmtA (panel I, immuno-labeled with 20 ug/ml anti-OmtA IgG) and nuclear staining are shown in the lower right corner. Bars, 2 pm. Images in this dissertation are presented in color. 115 were fiequently associated with ring-like structures (Fig. 5.9, A, C, E, and G). We stained nuclei using SYTOX-orange fluorescent dye (Abs/Em 543/570) and found that the fluorescent ring-like structures were associated with the outside surface of the nuclei (Fig. 5.9, B, D, F, and H). The cytoplasmic labeling pattern of VBS was clearly different from OmtA (Fig. 5.9, 1, lower right panel). Ultra-thin sections were immuno-labeled with column purified anti-VBS IgG followed by immuno-gold labeling; TEM analysis revealed that VBS was distributed within the cytoplasm and also concentrated in ring-like structures surrounding the nuclei in SU-l (Fig. 5.10, C, D, E, and F), while no labeling was observed in those structures in AF S10 (Fig. 5.10, A and B). Within a single cell, several ring-like structures immuno-labeled with VBS antibodies were observed in both CLSM (Fig. 5.9, A, C, E, and G) and TEM (Fig. 5.10, panel B). It is not clear if the signals were associated with a membrane structure. The data suggested that VBS is localized in the cytoplasm and also likely in the ER (see below). DISCUSSION I initially used native VBS to generate PAb but encountered significant cross reactivity. This could possibly be due to the limited amount of purified VBS available (60 pg for two rabbits) or that purified VBS still contained contaminating fimgal proteins. Similar cross-reactivity was encountered using the glycosylated form of A. niger glucose oxidase to generate PAb to the native enzyme (Witteveen et al., 1992). Native VBS has been confirmed to be N-link glycosylated (Silva et al., 1997). If the glycosylated form of the antigen affects antibody specificity, the native VBS should have been treated with deglycosylating enzyme to remove the N-glycan moiety prior to immunization. Nevertheless, I chose an alternative way to generate highly specific antibodies using methods reported previously, e.g., anti-Ver-l (Liang et al., 1997), anti-Nor—l (Zhou and Linz, 1999), and anti-OmtA (Lee et al., 2002), and this approach proved fruitful. 116 Figure 5.10. Immuno-gold labeling of VBS in A. parasiticus AF S10 and SU-l analyzed by TEM. Ultra-thin fungal sections derived from colony fractions R2 (AF S10 shown in Panels A and B) and S2 (SU-l shown in Panels C, D, E and F) were immuno-labeled with column-purified anti-VBS IgG (50 ug/ml) followed by goat-anti rabbit IgG conjugated with gold particles (10 nm diameter). The large organelles shown in panels B, D, and F were generated at higher magnifications from the cognate hyphae shown in panels A, C, and B, respectively. Abbreviations: CW, cell wall; M, mitochondria; N, nuclei; V, vacuole. Bars represent 1 pm in panels A and E, 0.5 um in panels B and D, and 250 nm in panels D and F. (This figure was generated by Lee, 2003). 117 Figure 5.10 118 The E. coli- derived recombinant MBP-VBS fusion protein was generated and used as an irnmunogen. Compared to other MBP fusion proteins previously generated in our laboratory (Liang et al., 1996) (Zhou, 1997) (Lee, et al., 2001), the yield of MBP-VBS fusion was relatively low. Silva et al. (1997) reported that E. coli expresses VBS in inclusion bodies with low solubility. Although part of my recombinant MBP-VBS was associated with cell particulates, part of the fusion protein was present in the soluble fraction. This may be due to the nature (solubility) of VBS or just simply the large size of the fusion protein (94 kDa). A temporal and regulatory relationship between secondary metabolism and microbial development has been postulated in various organisms (Bennett, 1983; Horinouchi et al., 1994, Guzman-De-Pena and Ruiz-Herrera, 1997; Adams et al., 1998; Calvo et al., 2002). In A. nidulans for example (contains the AF gene cluster but lacks the final enzyme, OrdA), a regulatory association between asexual sporulation and sterigmatocystin (ST) biosynthesis was demonstrated (Adams, et al., 1998). The balance between vegetative growth, sporulation, and ST production is regulated by two antagonistic factors, F le (Fluffy, I_.ow erA; brlA is a conidial vesicle specific gene) and F adA (Fluffy Autolytic Dominant) involved in a G-protein signaling pathway (Yu et al., 1996; Hicks et al, 1997). A dominant activation mutation in fadA completely blocked ST production and sporulation (Hicks et al., 1997). Based on these data, I hypothesized that AF biosynthesis is spatially associated with fimgal development at the cellular level and that AF enzymes are localized specifically in conidiophores (Lee et al., 2002?). However, recent data now clearly demonstrate that this hypothesis is not fully correct (Chapter 4); similar levels of VBS (this study), Nor-1, Ver-l , and OmtA (Chapter 4; Lee et al., submitted) were detected in vegetative hyphae and conidiophores. However, immuno-localization of VBS (this study) and OmtA, Nor-1, and Ver-l (Chapter 4) detected the highest levels of these proteins and AFB1 in fraction 82; interestingly, fraction 82 also contained the highest 119 numbers of newly developed conidiophores. According to Adams et al. (1998), conidiospores are generated approximately 24 h after germination of spores in the original inoculum. Consistent with these findings, conidiophores and conidiospores were observed in fractions S1 (48 to 72 h) and S2 (24 to 48 h) but not in fi'action S3 (0 to 24 h). In addition, most conidiophores in fraction SZ were immuno-labeled with antibodies to VBS and OmtA but significantly fewer were labeled in fraction 81 suggesting that these proteins were present in newly developed conidiophores. Visual microscopic inspection of the conidiophores in fi'action 2 confirms that many of these are immature and exist in various stages of development. In conclusion, our data support a spatial and temporal association between AF protein expression/AF production and sporulation at the colony level. I cannot rule out the possibility that protein stability or proteolytic degradation may contribute to our observations. The factors that directly participate in mediating this close association in A. parasiticus require further investigation. In terms of sub-cellular localization of VBS, both CLSM and TEM indicated the presence of VBS in the cytoplasm as well as in ring-like structures adjacent to the nucleus. I attempted to identify a subcellular targeting signal in VBS using protein localization prediction programs found on the ExPASy Molecular Biology Server (http://us.expasy.org). Since no database is available for filamentous fungi, in most cases we used the sequence analysis database for yeast; occasionally data from animal or plant cells were used. These analyses predicted that VBS has a signal peptide (analyzed by iPOSRT, PSORT version 6.4, Signal IP V2.0, and Target P V1.0), suggesting VBS may be associated with the ER/Golgi secretary pathway. Using PSORTH, the possible subcellular localization for VBS was predicted as follows: Mitochondria: 43.5%; Cytoplasm: 21.7%; Golgi: 17.4%; Endoplastic reticulum: 13.0%; and Vacuole: 4.3%. The observed perinuclear localization of VBS is likely due to its association with the ER and possibly the Golgi. First, the sequence prediction analysis suggests that the close association of VBS with ring like structures is consistent with passage through an 120 ER/Golgi secretory pathway. VBS is a N-link glycosylated protein and the ER is the site for core-glycan (Glc3-Man9-GlcNAcz) addition to the Asn residue. Second, the perinuclear immuno-fluorescence labeling pattern of VBS is reminiscent of several proteins known to target the ER including calreticulin (Roderick et al., 1997; Dyer and Mullen, 2001; Wedlich-Soldner et al., 2002), BR mannosidase (Zuber et al., 2000), nephrin (Yan et al., 2002), and the vacuolar system-associated protein VASAP-60 (Brule et al., 2003). Third, N-glycosylation of VBS is consistent with passage through an ER/Golgi pathway. Native VBS appeared as a doublet (78 kDa and 72 kDa) upon SDS-PAGE. Similar doublet bands have been detected by immuno-blotting of nephrin (Yan et al., 2002) and VASAP-60 (Brule et al., 2003). These proteins localize to ER and are thought to be post-translationally modified. In addition, it is well known that several co- translational and post-translational modifications take place specifically in the ER but not in the cytosol, including cleavage of the signal-peptide, N-link glycosylation, formation of disulfide bonds, and glyco-phosphatidylinositol (GPI)-anchoring (Ellgaard and Helenius, 2003 review). These observations allow us to predict that VBS is present in the ER for N-glycosylation and signal peptide cleavage; other VBS modifications might occur at that location as well. It is not clear if VBS is associated with the Golgi for complex glycan modification. In filamentous fungi, the typical appearance of the Golgi apparatus with the classic dictyosome organization of stacked disc shape cistemae was not observed previously (Markham, 1995); however, a recent study detected the morphology of Golgi bodies in A. niger (Khalaj et al. 2001). Our CLSM and TEM micrographs did not show clear evidence that the Golgi bodies were being immuno-labeled. Also, the enzyme PNGase F, which was used to confirm that VBS is N-Glycosylated (Silva et al., 1997), cleaves intact oligosaccharide chains from the protein; therefore this method cannot determine if VBS is modified with complex glycans (Ausubel et al., 2003). In order to 121 determine the degree of complexity of the glycan attached onto VBS, other types of endoglycosidase and glycoamidase enzymes should be used (Ausubel et al., 2003). N-glycosylation of VBS appeared to be significant for VBS activity and may play a role in stability. In general, protein N—glycosylation along the secretary pathway plays an important role in control of proper folding to the native tertiary and quaternary structure, and for subsequent translocation to the final cellular destination (Ellgaard and Helenius, 2003; Roth, 2002; Helenius and Aebi, 2001; Parodi, 2000). The N- glycosylation and ER protein control system provide proof reading mechanisms to ensure that the native conformation of newly synthesized proteins is maintained; in this regard, they are important for the fidelity of cellular fimctions (Ellgaard and Helenius, 2003 review). My data support the idea that N-glycosylation contributes to protein stability; VBS was significantly more stable than OmtA analyzed using similar methods (colony fiactionation/Westem blot data). A high proportion of intact VBS (78 kDa) remained in colony fraction S1 (48 to 72 h) compared to OmtA in this same fraction (Fig. 5.5 in this study compared to Fig. 5, Lee et al., 2002). It is still uncertain if enzymatically active VBS is present in the ER, cytoplasm, or at both locations. VBS does not contain a typical C-terminal KDEL ER-retention signal (Teasdale and Jackson, 1996; Harter and Wieland, 1996), so presumably it is not an ER resident protein but is only transiently present for modification. However, in most cells studied to date, glycoproteins are transported along the secretary pathway and localized in the extracellular space, the plasma membrane, or in compartments that are involved in secretion and endocytosis (Ellgaard and Helenius, 2003). Therefore, an important issue to resolve is why VBS appears in the cytosol. One possible explanation is that incorrectly folded, immature proteins can be retro-translocated from the ER into the cytosol for degradation through an ER-associated degradation (BRAD) process (Kostova and Wolf, 2003). However, the relative proportion of VBS found in the cytosol and the ER (approximately 50% in each location) 122 strongly suggests that the cytosolic localization of VBS is not likely due to this type of mechanism. Another, more likely explanation is that, since VBS is involved in biosysthesis of a secondary metabolite (i.e. AF ), targeting of this protein to its final cellular destination may be different from typical glycoprotein targeting in primary metabolism in higher eukaryotes. Perhaps the regular glycoproein transport machinery is altered during secondary metabolism. In addition, we cannot rule out the possibility that although VBS appeared to be a cytosolic potein, it may be associated with a particular sub-cellular compartment yet the membrane structure was not preserved during cell preparation. Alternatively, VBS may be anchored in membrane structures while the bulk of the protein remains in the cytoplasm. In support of this, VBS is predicted to contain a putitive transmembrane domain (PSORT version 6.4). Furthermore, mature VBS is a good candidate for a cytosolic protein because addition of a polar glycan generally increases solubility (Helenius and Aebi, 2001). If VBS is functional in the cytosol, it will be interesting to determine if this enzyme interacts with other AF synthesizing enzymes that appeared in the cytoplasm (e. g. Nor-l , Ver-l, and OmtA; Lee et al., manuscript submitted). Based on our preliminary co- immuno-gold labeling of VBS and OmtA using TEM, these two proteins are in close proximity in the cytosol. Future experiments involved in co-localization of various AF synthesizing proteins either by GFP-linked or co-immuno-labeling techniques, subcellular fractionation to identify the organelles associated with AF synthesizing enzymes, and immuno-precipitation studies to determine interactions between AF proteins may provide insights about the coordination of these protein involved in AF biosynthesis. 123 ACKNOWLEDGEMENTS Special thanks to Dr. Xudong Fan (Center for Advanced Microscopy at Michigan State University) for help in Operating the jet-freezer. Data in Chapter 5 were submitted for publication (Chiou, C. H., L. W. Lee, S. A. Owens, J. Whallon, K.L. Klomparens, C. A. Townsend, and J. E. Linz. Distribution and sub-cellular localization of the aflatoxin enzyme versicolorin B synthase (V BS) in time fiactionated colonies of Aspergillus parasiticus.Fungal Genet. Biol.). 124 BIBLIOGRAPHY Adams, T. H., J. K. Wieser, and J.-H. Yu. 1998. Asexual soprulation in Aspergillus nidulans. Microbiol. Mol. Biol. Rev. 62:35-54. Adams T. H. and J-H. Yu. 1998. Coordinate control of secondary metabolite production and asexual sporulation in Aspergillus nidulans. Curr Opinion Microbiol. 1:674-677. Aguilar, F., S. P. Hussain, and P. Cerutti. 1993. Aflatoxin B1 induces the transversion of G—>T in codon 249 of the p53 tumor suppressor gene in human hepatocytes. Proc. Natl. Acad. Sci. USA. 90:8586-8590. Akesson, H., E. Carlemalm, E. Everitt, T. Gunnarsson, G. Odham, and H-B Jansson. 1996. Irnmunocytochemical localization of phytotoxins in Bioplaris sorokiniana. Fungal Gene. Biol. 20:205-216. Amos, W. B., J. G. White, and M. Fordham. 1987. Use of confocal imaging in the study of biological structure. Appl. Optics 26:3239-3234. Anderson, H.W., E. W. Nehring, aand W. R. Wichser. 1975. Aflatoxin contamination of corn in the field. J. Agric Food Chem, 23:775-782. Anderson, J.A., and CH. Chung. 1994. Conversion of versiconal acetate to versicolorin C in extracts from Aspergillus parasiticus. Mycopathologia 110:31-35. Ashfold, A.E. Dynamic pleiomorphic vacuole system: are they endosomes and transport compartments in fungal hyphae? In Advances in Botanical Research incorporating Advacnes in Plant Pathology. Callow, J. A. ed. Academic Press. San Diego, CA. Asao, T., G. Buchi, M. M. Abdel-Kader, S. B. Chang, E. L. Wick, and G. N. Wogan. 1963. Aflatoxins B and G. J. Am. Chem. Soc. 85:1706-1707. Aufauvre-Brown,A., C. M. Tang, and D. W. Holden. 1993. Detection of gene- disruption events in Aspergillus transformants by polymerase chain reaction direct from conidiospores. Curr Genet. 24: 177-178. Ausubel F. M., R. Brent, E. Kingston, D. D. Moore, J. G. Seidman, J. A. Smith, and K. Struhl. 2003. Current Protocol in Molecular Biology. John Wiley & Sons, New York, N. Y. Bailey, E. A., Iyer, R. S., Stine, M. P. Harris, T. M., and Essigmann, J. M. 1996. Mutational properties of the primary aflatoxin B1-DNA adduct. Proc. Natl. Acad. Sci. USA. 93:1535-1539. 125 Beckette, A., I. B. Health, and D. J. McLaughlin. 1974. An Atlas of Fungal Ultrastructure, Longman, London. Bennett, W. J. 1983. Secondary metabolism and differentiation in fungi, p. 1-32. In Bennett, J .W., and A. Ciegler. (ed.), Secondary metabolism and Differentation in Fungi. Maecel Dekker, New York, N. Y. Betina, V. 1995. Differentiation and secondary metabolism in some prokayotes and fungi. Folia. Microbiol. 40:51-67. Bhatanagar, D., J. Yu, and K. C. Ehrlich. 2002. Toxins of filamentous fimgi. In Fungal Allergy and Pathogenecity. Breitenbach, M., R. Crameri, and S. B. Lehrer. eds.Chem Irnmunol. Basel, Karger 81: 167-206. Bhatnagar, D., K. C. Ehrlich, and T. E. Cleveland. 2003. Molecular genetic analysis and regulation of aflatoxin biosynthesis. 61:83-89. Blount, W. P. 1961. Turkey "X" disease. J. Brit. Turkey Fed. 9:52-54. Bodine, A. B., and D. R. Mertens. 1983. Toxicology, metabolism, and physiological effects of aflatoxin in the bovine. pp. 46-50. In U. L. Diener, R. L. Asquith, and J. W. Dickens, Eds. Aflatoxin and Aspergillus flavus in corn. Souther Cooperative Series Bulletin 279. Auburn University, Auburn, Alabama. Borgia, P.T., L.E. Eagleton, and Y. Miao. 1994. DNA preparations from Aspergillus and other filamentous fungi. Biotechniques 17:431-432. Boutrif E. 1995. FAO programmers for prevention, regulation, and control of mycotoxin in food. Nat. Toxins. 3:322-326. Brown, D.W., J.H. Yu, H.S. Kellar, M. Fernandes, T.C., Nesbit, N.P. Keller, T.H. Adams, and T.J. Leonard. 1996. Twenty-five coregulated transcripts defibe a sterigarnatocystin gene cluster in Aspergillus nidulans. Proc. Natl. Acad. Sci. 93:14] 8- 1422. Brelje, T.C., M. W. Wessendorf, and R. L. Sorenson. 1993. Multicolor laser scanning confocal immunofluorescence microscopy: Practical applications and limitations. In. Cell Biological Applications of Confocal Microscopy (B. Matsumoto, ed.) pp. 98-182. Academic Press, San Diego. Brule, S., R. Faure, M. Dore, D. W. Silversides, and J. G. Lussier. 2003. Irnmunolocalization of vacuolar system-associated protein-60 (V ASAP-60). Histochem. Cell Biol. 119:371-381. 126 Buchanan, R. L., and D. F. Lewis. 1984. Regulation of aflatoxin biosynthesis: effects of glucose on activities of various glycolytic enzymes. Appl. Environ. Microbiol. 48:306- 310. Butchko, R.A.E., T.H. Thomas, N.P. Keller. 1999. Aspergillus nidulan mutants defective in stc gene cluster regulation. Genetics 153:715-720. Calvo, A. M., R. A. Wilson, J. W. Bok., and N. P. Keller. 2002. Relationship between secondary metabolism and fungal development. Micro. Mol. Biol. Rev. 66:447-459. Carlsson, K., and N. Aslund. 1987. Confocal imaging for 3-D digital microscopy. Appl. Optics 26:3232-3238. Cary, J.W., N. Barnaby, N.,K.C. Ehrlich, and D. Bhatnagar. 1999. Isolation and characterization of experimentally induced, aflatoxin biosynthetic pathway deletion mutants of Aspergillus parasiticus. Appl. Microbiol. Bitechnol. 51:808-812. Cary, J. J., J. E. Linz., and D. Bhatnagar. 2000. Aflatoxins: biological significance and regulation of biosynthesis. p. 317-362. In J. Cary, D. Bhatnagar, and J. Linz (ed.), Microbial foodborne diseases: mechanisms of pathogenesis and toxin synthesis. Technomic Publishing, Lancaster, Pa. Cary, J. J., M. D. John, K. C. Ehrlich, M. S. Wright, S-H. Liang and J. E. Linz. 2002. Molecular and functional characterization of a second copy of the aflatoxin regulatory gene, AflR, from Aspergillus parasiticus. Biochim Biophys Acta. 1576:316- 323. CAST. 2003. Council for Agricultural Science and Technology. Mycotoxins: Risks in plant, animal, and human system. Report 139. Cenis,J.L. 1992. Rapid extraction of fimgal DNA for PCR amplification. Nucleic Acids Res. 20:2380. Chang, P. K., C. D. Skory, and J. E. Linz. 1992. Cloning of a gene associated with aflatoxin B1 biosynthesis in Aspergillus parasiticus. Curr. Genet. 21:231-233. Chang, P. K., J. W. Cary, D. Bhatnagar, T. E. Cleveland, J. W. Bennett, J. E. Linz, C. P. Woloshuk, and G. A. Payne. 1993. Cloning of the Aspergillus parasiticus apa-2 gene associated with the regulation of aflatoxin biosynthesis. Appl. Environ. Microbiol. 59:3273-3279. Chang, P. K., J. W. Cary, J. Yu, D. Bhatnagar, and T. E. Cleveland. 1995a. The Aspelto et al., 2001; ynthase gene , pksA, a homologs of Aspergillus nidulans wA, is required for a aflatoxin B1 biosynthesis. Mol. Gen. Genet. 248:270-277. 127 Chang, P. K., D. Bhatnagar, D., K. C. Ehrlich, and T. E. Cleveland, and J. W. Bennett. 1995b. Sequence variability in homologs of the aflatoxin pathway gene aflR distinguished species in the Aspergillus Section F lavi. Appl. Environ. Microbiol. 61:40- 43. Chang, P.-K. and J. Yu. 2002. Characterization of a partial duplication of the aflatoxin gene as ATCC 56775. Appl. Microbiol. Biotechnol. 58:632-636. Chiou, C. H., M. M. Miller, D. L. Wilson, F. Trail and J. E. Linz. 2002. Chromosomal location plays a role in regulation of aflatoxin gene expression in Aspergillus parasiticus. Appl. Environ. Microbiol. 68:306-315. Colling,A.J., and P. Markham. 1985. Woronin bodies rapidly plug septal pores of severed Penicillium Chrysogenumhyphae. Experimental. Mycol. 9:80-85. Cotty, P. J., and D. Bhatnagar. 1994. Variability among atoxigenic Aspergillus flavus strains in ability to prevent aflatoxin contamination and production of aflatoxin biosynthetic pathway enzymes. Appl. Environ. Microbiol. 60:2248-2251. Cotty PJ. and KR Cardwell. 1999. Divergence of West African and North American communities of Aspergillus section Flavi. Appl. Environ, Microbiol. 65:2264-2266. Czymmek, K., J. H. Whallon, and Klomparens, K. L. 1994. Confocal microscopy in mycological research. Exp. Mycol. 18:275-293. Gould, J. R., G.-A. Keller, and S. Subramani. 1987. Identifictaion of a peroxisomal targeting signal at the carboxylterrninus of firefly luciferase. J. Cell. Biol. 105:2923-2931. Dyer, J. M. and R. T. Mullen. 2001. Irnmunocytological localization of two plant fatty acids desaturases in the endoplasmic reticulum. F EBS Lett. 494:44-47. D'Souza, C.A., B.N. Lee, and T. Adams. 2001. Characterization of the rold of the FluG protein in asexual development of Aspergillus nidulans. Genetics 158: 1027-1036. Eaton, D. L., and E. P. Gallagher. 1994. Mechanisms of aflatoxin biosynthesis. Annu. Rev. Pharrnacol. Toxicol. 34: 135-172. Eaton, D. L., and J. D. Groopman. (ed.) 1994. The toxicology of aflatoxins: human health, veterinary, and agriculture Significance. Academic Press, San Diego, Calif. Ehrlich, K., J.W. Carry, and B.G. Montalbano. 1999a. Characterization of the promoter for the gene encoding the aflatoxin biosynthetic pathway regulatory protein AFLR. Biochim. Biophys. Acta 1444:412-417. 128 Ehrlich, K., B.G. Montalbano, and J.W. Carry. 1999b. Binding of the C6-zinc cluster protein, AF LR, to the promoters of aflatoxin pathway biosynthesis genes in Aspergillus parasiticus. Gene 230:249-257. Ehrlich, K., B.G. Montalbano, and P. J. Cotty. 2003. Sequence comparision of aflR from different Aspergillus species provide evidence for variability in regulation of aflatoxin production. Fungal Genet. Biol. 38:63-74. Ellis, W. O., J. P. Smith, B. K. Simpson, and J. H. Oldham. 1991. Aflatoxins in food: occurrence, biosynthesis, effects on organisms, detection, and methods of control. Crit Rev. Food Sci. Nutr. 30:403-439. Ellgaard, Land A. Helenius. 2003. Quality control in the endoplasmic reticulum. Nature Rev. 4:181- 191. Farag, R. S., M. M. Rashed, and A. A. Abo-Hagger. 1996. Aflatoxin destruction by microwave heating. Intl. J. Food Sci Nutri. 47: 197-208. Feng, G. H., and T. J. Leonard. 1995. Characterization of the polyketide synthase gene (pksLI) required for aflatoxin biosynthesis in Aspergillus parasiticus. J. Bacteriol. 177:6246-6254. F erreira, A. V. B., and N. Glass. 1996. PCR from fungal spores afier microwave treatment. Fungal Genet News]. 43:25-26. Flegler, S.L., J.W.Heckman, and KL. Klopmarens. 1995. Scanning and transmission electron microscopy an introduction. Chapter 6. p.97- 150. Ghebranious, N.,and S. Sell. 1998. Hepatitis B injury, male gender, aflatoxin, and p53 expression each contribute to hepatocarcinogenesis in trasngenic mice. Hepatology. 27: 383-391. Greenblatt, M. S., W. P. Bennett, M. Hollstein., and C. C. Harris. 1994. Mutations in the p53 tumor suppressor gene: clues to cancer etiology and molecular pathogenesis. Cancer Res. 54:4855-4878. Guzman-De-Pena, D., and J. Ruiz-Herrera. 1997. Relationship between aflatoxin biosynthesis and sporulation in . Fungal Genet. Biol. 21:198-205. Hamer, J. E. and W. E. Timberlake. 1987. Functional organization of the Aspergillus nidulans trpC promoter. Mol. Cell Biol. 7 :2352-2359. Harris, C. C., M. and Hollstein. 1993. Clinical implications of the p53 tumor suppressor gene. N. Engl. J. Med. 329:1318-1327. 129 Harris, 8. D., J. L. Morrell and J. E. Hamer. 1994. Identification and characterization of Aspergillus nidulans mutants defective in cytokinesis. Genetics. 136:517-532. Hartcr, C., and F. Wieland. 1996. The secretary pathway: mechanisms of protein sorting and transport. Biochim. Biophys. Acta. 1286:75-93. Helenius, A., and M. Aebi. 2001. Intracellular fimctions of N-linked glycans. Science 291:2364-2369. Henry, S. H., F. X. Bosch, T. C. Troxell, and P. M. Bolger. 1999. Reducing liver cancer- Global control of aflatoxin. Science 286:2453-2454. Herman, B. 1998. Fluorescence microscopy. Bios Scientific Publisher, Springer-Verlag, New York, Inc. NY. Hicks, J. K., J.-H. Yu, N. P. Keller, and T. H. Adams. 1997. Aspergillus sporulation and mycotoxin production both require inactivation of the FadA Ga protein-dependent signaling pathway. EMBO. J. 16:4916-4923. Hoch. H. C., and R. C. Staples. 1983. Ultrastructural organization of the non- differentiated uredosopre germling of Uromyces phaseoli variety typica. Mycologia 75:795-824. Horinouchi. S., and T. Beppu. 1994. A-factor as a microbiol hormone that controls cellular differentiation and secondary metabolism in Streptomyces griseus. Mol. Microbiol. 12:859-864. Homg, J. S., P. K. Chang, J. J. Pestka, and J. E. Linz. 1990. Development of a homologous transformation system for Aspergillus parasiticus with the gene encoding nitrate reductase. Mol. Gen. Genet. 224:294-296. Hsia, C. C., D. E., Klenier, C. A. Axiotis, A. DiBidceglie, A. M. Nomura, D. E. Stemmermann, and E. Tabor. 1992. Mutations of p53 gene in hepatocellular carcinoma: roles of hepatitis B virus and aflatoxin contamination in the diet. J. Natl. Cancer. Inst. 84:1638-1641. Hussein, H. S., and J. M. Brasel. 2001. Toxicity, metabolism, and impact of mycotoxins on humans and animals. Toxicology 167:101-134. IARC. 1993. IARC monographs on the evaluation of carcinogenic risks to humans. Some naturally occurring substances: food items and constituents, heterocyclic aromatic amines and mycotoxins. Lyon, F rancezlARC. 56, 245-345. Ito, Y., S. W. Peterson, D. T., Wicklow, and T. Goto. 2001. Aspergillus pseudotamarii, a new aflatoxin producing species in Aspergillus section F lavi. Mycol. Res. 105:233-239. 130 Jackson, P. E. and J. D. Groopman. 1999. Aflatoxin and liver cancer. Bailliere's Clin. Gastro. 13:545-555. Jedd, G. and N.-H. Chua. 2000. A new self-assembled peroxisomal vesicle required for efficient resealing of the plasma membrane. Nat. Cell Biol. 2:226-231. Johnson, L. M., V. A. Bankaitis, and S. D. Emr. 1987. Distinct sequence determinants direct intracellular sorting and modification of a yeast vacuolar protease. Cell 48:875- 885. Kale, S. P., D. Bhatnagar, and J.W. Bennett. 1994. Isolation and characterization of morphological varients of deficient in secondary metabolite production. Mycol. Res. 98: 645-652. Kato, J. A. Suzuki., H. Yamazaki, Y. Ohinishi, and S. Horinouchi. 2002. Control by A-factor of a metalloendopeptidase gene involved in areial mycelium formation in Streptomyces griseus. J. Bact. 184:6016-6025. Keller, G.A. 1991. Evolutionary conservation of a microbody targeting signal thatbtagets protein to peroxisomes, glyoxisomes and glycosome. J. Cell. Biol. 114:893-904. Keller, N. P. and T. M. Hohn. 1997. Metabolic pathway gene clusters in filamentous fungi. Fungal. Genet. Biol. 21:17-29. Khalaj, V., J. L. Brookman, and G. D. Robson. 2001. A study of the protein secretary pathway of Aspergillus niger using a glucoamylase-GFP fusion protein. F ung. Genet. Biol. 32:55-65. Kinsey, J. A. and J. A. Rambosek. 1984. Transformation of Neurospora crassa with the cloned 62de (glutamate dehydrogenase) gene. Mol. Cell Biol. 4:117-122. Klinlosky, D. J. 1997. Protein transport from the cytoplasm into the vacuole. J. Membrane Biol. 157:105-115. Klich, M.A., E. J. Mullaney, C. B. Daly, and J. W. Cary. 2000. Molecular and physiological aspects of aflatoxin and sterigmactocystin biosynthesis by Aspergillus tamarii and A. ochraceoroseus. Appl. Microbiol. Biotechnol. 53:605-609. Kostova, Z., and D. H. Wolf. 2003. For whom the bell tolls: protein quality control of the endoplasmic reticulum and the ubiquitin-proteasome connection. EMBO J. 22:2309- 2317. Krishnamachari, K. A. V. R., R. V. Bhat., V. Nagarajan, and T. B.. G. Tilac. 1975, Hepatitis due to aflatoxicosis. Lancet 1:1061-1063. 131 Kurylowicz, W., W. Kurzatkowski, and J. Kurzatkowski. 1987. Biosynthesis of benzylpenicillin by in Penicillium chrysogenum and its golgi apparatus. Arch. Immunol. Ther. Exp. 35:699-724. Kusumoto, K. 1., K. Yabe, Y. Nogata, and H. Ohta. 1998. Transcript of a homolog of aflR, a regulator gene for aflatoxin synthesis in Aspergillus paraasiticus, was not detected in Aspergillus oryzae strains. FEMS Micrbiol. Letts. 169:303-307. Kwon, Y.H., K.S. Wells., and H. C. Koch. 1993. Fluorescence confocal microscopy: Applications in fungal cytology. Mycologia 85:721-733. Lasky, T., and L. Magder. 1997. Hepatocellular carcinoma p53 G>T trasnversion at codon 249: the fingerprint of aflatoxin exposure? Environ. Health. Perspect. 1115:392- 397. Lecellier, G., and P. Silar. 1994. Rapid methods for nucleic acids extraction from Petri dish-grown mycelia. Curr Genet. 25: 122-123. Lee, A. B., and T. A. Cooper. 1995. Improved direct PCR screen for bacteria colonies: Wooden toothpicks inhibit PCR amplification. BioTechniques 18:225-226. Lee, B. N., and T. Adams. 1994. The Aspergillus nidulansfluG gene is required for production of an extracellular developmental signal. Genes Dev. 8:641-651. Lee, L.W., C.H. Chiou, and J.E. Linz. 2002. Function of native OmtA in vivo and expression and distribution of this protein in colonies of Aspergillus parasiticus. Appl. Environ. Microbiol. 68:5718-5727. Lee, L.W. Functional analysis of omtA and subcellular localization of aflatoxin enzymes in Aspergillus parasiticus. .PhD Dissertation, Michigan State University, East Lansing. Lee, L. W., C. H. Chiou, K. L. Klomparens, and J. E. Linz. 2003. Sub-cellular localization of aflatoxin biosynthetic enzymes in time-dependent fractionated colonies of Aspergillus parasiticus. Manuscript submitted to Arch. Microbiol. Lenderfeld, T., D. Ghali, M. Wolschek, E. M. Kubicek-Pranz, and C. P. Kubicek. 1993. Subcellular compartmentation of penicillin biosynthesis in Penicillium chrysogenum. J. Biol. Chem. 268:665-671. Liang, S.-H. 1996. The function and expression of the ver-l gene and localization of the Ver-l protein involved in aflatoxin B1 biosynthesis in Aspergillus parasiticus. PhD Dissertation, Michigan State University, East Lansing. Liang S.-H., C. D. Skory and J. E. Linz. 1996. Characterization of the function of the ver-l and ver-lB genes, involved in aflatoxin biosynthesis in Aspergillus parasiticus. Appl Environ Microbiol. 62:4568-4575. 132 Liang, S.H., T-S Wu, R. Lee, F.S. Chu, and J.E. Linz. 1997. Analysis of mechanisms regulating expression of the ver-l gene, involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 63: 1058-1 065. Lin B. K., and J. A. Anderson. 1992. Purification and properties of versiconal cyclase from Aspergillus parasiticus. Arch. Biochem. Biophys. 293:67-70. Ling, M., F. Merante, and B. H. Robinson. 1995. A rapid and reliable DNA preparation method for screening a large number of yeast clones by polymerase chain reaction. Nucleic Acids Res. 23:4924—4925. Lunn, R. M., Y. J. Zhang, L. W. Wang, C. J. Chan, P. H. Lee, C. H. Lee, W. Y. Tsai, and R. M. Santella. 1997. p53 mutations, chronic hepatitis B virus infection, and aflatoxin exposure in hepatocellular carcinoma in Taiwan. Cancer Res. 57:3471-3477. Mace, K., F. Aguilar, J. S. Wang, P. Vautravers, M. Gomez-Lechon, F. J. Gonzalez, J. Groopman, C. Harris, and M. A. Pfeifer. 1997. Aflatoxin Bl-induced DNA adduct formation and p53 mutations in CYP450-expressing human liver cell lines. Carconigenesis 18: 1291-1297. Mahanti, N., D. Bhatnagar, J. W. cary, J. Joubrab, and J. E. Linz. 1996. Structure and function of fas-lA, a gene encoding a pututive fatty acid synthase directly involved in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 62: 191-195. Mannon, J., and E. Johnson. 1985. Fungi down on the farm. New Scientist 105:12-16. Markham, P. 1995. Organelles of filamentous fungi. Pp. 75-98. In N. A. R. Gow, and G. W. Gooday (ed), The growing fungus, 1st ed. Chapman & Hall, London, UK. Marin, S., V. Sanchis, R. Saenz, A. J. Ramos, I. Vinas, and N. Magan. 1998. Ecological determinants for germination and growth of some Aspergillus and Penicillium spp. from grain. J. Appl. Microbiol 84:25-36. Markham. P., and A. J. Collinge. 1987. Woronin bodies of filamentous fungi. FEMS Microbiol. Rev. 46: 1-1 1. Marth, E. H., and M. P. Doyle. 1979. Update on molds: Degradation aflatoxin. Food Technol 33:81-87. Massey, T. E., R. K. Stewart, J. M. Daniels, and L. Liu. 1995. Biochemical and molecular aspects of mammalian susceptibility to aflatoxin B1 carcinogenicity. Proc. Soc. Exp. Biol. Med. 208:213-227. 133 Matushima K., K. Yashiro, Y. Hanya, K. Abe, K. Yabe, and T. Hamasaki. 2001. Absence of aflatoxin biosynthesis in koji mold (Aspergillus sojae). Appl. Microbiol. Biotechnol. 55:771-776. Maxewell, D.P., V.N. Armcntrout, and LB. Graves. 1977. Microbodies in plant pathogenic fimgi. Annal. Rew. Phyto. 15:119-134. McGuire, S. M., J. C. Silva, E. G. Casillas, and C. A. Townsend. 1996. Purification and characterization of Versicolorin B Synthase from Aspergillus parasiticus. Catalysis of the stereodifferentiating cyclization in aflatoxin biosynthesis essential to DNA interaction. Biochemistry 35:11740-1 1486. McLean, M. and M. F. Dutton. 1995. Cellular interactions and metabolism of aflatoxin: an update. Pharrnac. Ther. 65:163-192. Meyers D. M., G. Obrian, W. L. Du, D. Bhatnagar, and G. A. Payne. 1998. Characterization of aflJ, a gene required for conversion of pathway intermediates to aflatoxin. Appl. Environ. Microbiol. 64:3713-7. Min, J., M. T. Arganoza, J. Ohrnberger, C. Xu, and RA. Akins. 1995. Alternative methods of preparing whole-cell DNA from fungi for dot-blot, restriction analysis, and colony filter hybridization. Anal Biochem. 225:94-100. Miller, B. L., K. Y. Miller, K. A. Roberti, and W. E. Timberlake. 1987. Position- dependent and -independent mechanisms regulate cell-specific expression of the SpoCI gene cluster of Aspergillus nidulans. Mol. Cell Biol. 7 :427-434. Miller, M. J. 2003. Transcriptional regulation of the Aspergillus parasiticus aflatoxin biosynthesis pathway gene nor-I . Ph.D. Dissertation, Michigan State University, East Lansing. Minto, R. E. and C. A. Townsend. 1997. Enzymology and molecular biology of aflatoxin biosynthesis. Chem. Rev. 97:2537-2555. Minsky, M. 1961. Microscopy apparatus. US. Patent 3013467. Minsky, M. 1988. Memoir on inventing the confocal scanning microscope. Scanning l0: 1 28-1 3 8. Montalto, G. M. Cervello, L. Giannitrapani, F. Dantona, A. Terranova, and L. A. M. Castagnetta. 2002. Epidemiology, risk factors, and natural history of hepatocellular carcinoma. Ann. NY. Acad. Sci. 963:13-20. Motomura, M., N. Chihaya, T. Shinozawa, T. Hamasaki, and K. Yabe. 1999. Cloning and characterization of the O-Methyltransferase I gene (dmtA) from associated with the conversion of demethylstrigmatocystin to sterigmatocystin of dihydro- 134 demethylsterigrnatocystin to dihydrosterigrnatocystin in aflatoxin biosynthesis. Appl. Environ. Microbiol. 65:4987-4994. Muchen, K. F. , R. P. Misra, and M. Z. Humayun. 1983. Sequence specificity in aflatoxin Bl-DNA interactions. Proc. Natl. Acad. Sci. USA 80:6-10. Milfler W.H., T.P. van der Krift, AJ.J. Krouwer, H.A.B. Wosten, L.H.M. van der Voort, E.B. Smaal, and AJ. Verkleij. 1991. Localization of the pathway of the penicillin biosynthesis in Penicillium chrysogenum. EMBO Journal. 10:489-495. Ngindu, A., P. R. Kenya, D. M. Ocheng, T. N. Omonde, W. Ngarc, D. Gatei, B. K. Johnson, J. A. Ngira, H. Nandwa, J. Jansen, J. N. Kaviti, and T. A. Siongok. 1982. Outbreak of acute hepatitis caused by aflatoxin poisioning in Kenya. Lancet 1:1346-1348. Ohneda, M., M. Arioka., H. Nakajima, and K. Kitamoto. 2002. Visualization of vacuoles in Aspergillus oryzae by expression of CYP-EGF P. Fungal Genet. Biol. 37 :29- 38. Ohsumi, K., M. Arioka, H. Nakajima, and K. Kitamoto. 2002. Cloning and characterization of a gene (avaA) from Aspergillus nidulans encoding a small GTPase involved in vacuole biogenesis. Gene. 291:77-84. Paddock, S. W. An introduction to confocal imaging. Confocal Microscopy Methods and Protocols. Ed. S. Poddock Human Press Inc. Totowa, NJ. Meth. Mol. Biol. 122:1-34. Parodi, A. J. 2000. Rold of N-oligosaccharide endoplasmic reticulum processing reactions in glycoprotein folding and degradation. Biochem. J. 348: 1-13. Paul, G. C., C. A. Kent, and C. R. Thomas. 1994. Hyphal vacuolation and fragmentation in Penicillin chrysogenum. Biotechnol. Bioeng. 44:655-660. Payne, G. 1998. Process of contamination by aflatoxin-producing fungi and their impact on crops. pp 279-306. In K. K. S. Sinha and D. Bhatnagar (Eds) Mycotoxins in Agriculture and Food Safety. Marcel Dekkeer, Inc. New York. Payne, G. A., and M. P. Brown. 1998. Genetics and physiology of aflatoxin biosynthesis. Annu. Rev. Phytopathol. 36:329-362. 135 Pestka, J.J., P. K. Gaur, and F. S. Chu. 1980. Quantification of aflatoxin BI and aflatoxin B1 antibody by an enzyme-linked immunosorbent microassay. Appl. Environ. Microbiol. 40: 1027 - 1 03 1 . Prieto R., G. L. Yousibova, and C. P. Woloshuk. 1996. Identifictaion of aflatoxin biosynthetic genes by genetic complementation in a mutant of Aspergillus parasiticus lacking the aflatoxin gene cluster. Appl. Environ. Microbiol. 62:3567-3571. Raymond, C. K., C. J. Roberts, K. E. Moore, 1. Howald, and T. H. Stevens. 1992. Biogenesis of the vacuole in Saccharomyces cerevisiae. Int. Rev. Cytol. 139:59-120. ReiB, J. 1982. Development of Aspergillus parasiticus and formation of aflatoxin B1 under the influence of conidiogenesis affecting compounds. Arch. Microbiol. 133:236-8. Roderick, H. L., A. K. Campbell, and D. H. Llewellyn. 1997. Nuclear localisation of calreticulin in vivo is essential by its interaction with glucocorid receptors. FEBS Lett. 405:181-185. Roth, J. 2002. Protein N-Glycosylation along the secretary pathway: relationship to organelle topography and function protein quality control and cell intercations. Chem. Rev. 102:285-303. Sharma, P. R., and D. K. Salunkhe. 1991. Mycotoxin and phytotoxins in human and animal health. Boca Raton, CRC Press. Shen, H. M., and C. H. Ong. 1996. Mutations of the p53 tumor suppressor gene and ras oncogenes in aflatoxin hepatocarcinogenesis. Mut. Res. 366:23-44. Siller , W. G., and D.C. Ostler. 1961. The histopathology of an enterohepatic syndrome of turkey poults. Vet. Rec. 73:134-148. Silva, J. C., R. E. Minto, C. E. Barry, K. A. Holland, and C. A. Townsend. 1996. Isolation and characterization of the versicolorin B synthase gene from Aspergillus parasiticus. J. Biol. Chem. 271:13600-13608. Silva, J. C. and C. A. Townsend. 1997. Heterologous expression, isolation, and characterization of Versicolonin B Synthase from Aspergillus parasiticus. J. Biol. Chem. 272:804-813. Sinnhuber, R. O., J. D. Hendriks, J. H. Wales, and G. B. Putnam. 1977. Neoplasms in rainbow trout, a senstive animal model for environmental carcinogenesis. Ann. N. Y. Acad. Sci. 298:389-408. Skory, C. D., J. S. Horng, J. J. Pestka, and J. E. Linz. 1990. Transformation of Aspergillus parasiticus with a homologus gene (pyrG) involved in pyrimidine biosynthesis. Appl. Environ. Microbiol. 56:3315-3320. 136 Skory, C. D., J. S. Horng, J. J. Pestka, and J. E. Linz. 1992. Isolation and characterization of a gene from Aspergillus parasiticus associated with the conversion of versicolorin A to sterigmatocyctin in aflatoxin biosynthesis. Appl. Environ. Microbiol. 58:3527-3537. Skory, C. D., P.-K. Chang, and J. E. Linz. 1993. Regulated expression of the nor-1 and ver-I genes associated with aflatoxin biosynthesis. Appl. Environ. Microbiol. 59:1642- 1646. Spear, R. N., D. Cullen., and J. H. Andrews. 1999. Fluorescent labels, confocal microscopy, and quantitative image analysis in study fungal biology. Methods In Enzymol. 307:607-623. Steiner, J.J., C.J. Poklemba, R.G. Fjellstrom, and L.F. Elliott. 1995. A rapid one-tube genomic DNA extraction process for PCR and RAPD analysis. Nucleic Acids Res. 13:2569-2570. Subramani, S. 1996. Protein translocation into peroxisomes. J. Biol. Chem. 271:32483- 32486. Tarutani, Y., K. Ohsumi, M. Arioka, H. Nakajima, and K. Kitamoto. 2001. Cloning and characterization ofAspergillus nidulans vpsA gene which is involved in vacuolar biogenesis. Gene 268:23-30. Teasdale, R. D., and M. R. Jackson. 1996. Signal mediated sorting of membrane proteins between the endoplasmic reticulum and the Golgi apparatus. Annu. Rev. Cell Dev. Biol. 12:27-54. Tenney, K. 1. Hunt, J. Sweigard, J. I. Pounder, C. McClain, E. J. Bowman, and B. J. Bowman. 2000. hex], a gene unique to filamentous fungi, encodes the major protein of the Woronin body and functions as a plug for septal pores. Fungal Genet. Biol. 31:205- 217. Thurmm, M. 2000. Structure and function of the yeast vacuole and its role in autophagy. Microsc, Res. Tech. 51:563-572. Timberlake, W. E. and E. C. Barnard. 1981. Organization of a gene cluster expressed specifically in the asexual spores of A. nidulans. Cell 26:29-37. Timberlake, W. E. and M. A. Marshall. 1989. Genetic engineering of filamentous fungi. Science 244:1313-1317. Trail, F., P-K Chang, and J. E. Linz. 1994. Structural and functional analysis of the nor-1 gene involved in the biosynthesis of aflatoxin by Aspergillus parasiticus. Appl. Environ. Microbiol. 60:3315-3320. 137 Trail, F., N. Mahanti, and J. E. Linz. 1995a. Molecular biology of aflatoxin biosynthesis. Microbiology 141:755-765. Trail, F., N. Mahanti, M. Rarick, R. Mehigh, S.-H. Liang, R. Zhou, and J.E. Linz. 1995b. Physical and transcripitional map of an aflatoxin gene cluster in Aspergillus parasiticus and functional disruption of gene involved early in the aflatoxin pathway. Appl. Environ. Microbiol. 61: 1665-1673. Van der Kamp, M., A. J. M. Driessen, and W. N. Konings. 1999. Compartrnentalization and transport in B-Lactam antibiotic biosynthesis by filamentous fungi. Antonie van Leeuwenhoek. 75:41-78. Van der Lende, T. R., M. van de Kamp, M. van den Berg, K. Sjollema, R. A. L. Bovenberg, M. Veenhuis, W. N. Konings and A. J. M. Driessen. 2002. 8-(L-a- aminoadipyl)-L-cysteinyl-D-valine synthetase, that mediates the first committed step in penicillin biosynthesis, is a cytosolic enzyme. Fungal Gene. Biol. 37:49-55. Valenciano, 8., JR. De Lucas, 1. Van der Klei, M. Veenhuis, and F. Laborda. 1998. Charatcerization of Aspergillus nidulans peroxisomes by irnmunoelectron microsc0py. Arch. Microbiol. 170:370-376. Van Egmond, H. P. 1993. Rationale for regulatory programmes for mycotoxins in human foods and animal feeds. Food Addit. Contam. 10:29-36. van Gorcom, R. F ., P. J. Punt, P. H. Pouwels, and C. A. van den Hondel. 1986. A system for the analysis of expression signals in Aspergillus. Gene 48:211-217. van Zeijl,C. M. J., E. H. M. van de Kamp, P.J. Punt, G. C. M. Selten, B. Hauer, R.F.M. van Gorcom, and C. A. M. J. J. van der Hondel. 1998. An improved colony- PCR method for filamentous fungi for amplification of PCR-fragments of several kilobases. J. Biotech. 59:221-224. Verdoes, J. C., P. J. Punt, A. H. Stouthamer, and C. A. M. J. J. van den Hondel. 1994. The effect of multiple copies of the upstream region on expression of the Aspergillus niger glucoamylase encoding gene. Gene 145: 179-187. Vidal-cros, A., F. Viviani, G. Labesse, M. Boccara, and M. Gaudry. 1994. Polyhydroxy-naphthalene reductase involved in melanin biosynthesis in Magnaporthe grisea-purification, cDNA cloning and sequencing. Eur. J. Biolchem. 219:985-992. Voeltz, G. K., M. M. Rolls, and T. A. Rapoport. 2003. Structureal organization of the endoplasmic reticulum. EMBO reports 3:944-950. Wada, Y. and Y. Anraku. 1994. Chemiosomic coupling of ion transport in the yeast vacuole: its rile in acidification inside organelles. J. Bioenerg. Biomembr. 26:631-637. 138 Wang, H., S. E. Kohalmi, and AJ. Culter. 1996. An improved method for polymerase chain reaction using whold yeast cells. Anal Biochem. 237: 145-146. Waterham, H. R., and J. M. Cregg. 1999. Peroxisome biogenesis. BioAssays 19:57-66. Watson, A. J., L. J. Fuller, D. J. Jeemes, and D. B. Archer. 1999. Homologs of aflatoxin biosynthesis genes and sequence of aflR in Aspergillus oryzae and Aspergillus sojae. Appl. Environ. Microbiol. 65:307-310. Wedlich-Sfilender, R. I. Schulz, A. Straube, and G. Steinberg. 2002. Dynein supports motility of endoplasmic reticulum in the funus Ustilago maydis. Mol. Cell. Biol. 13:965- 977. Wendland, B., S. D. Emr, and H. Riezmanet. 1998. Protein traflic in the yeast endocytic and vacuolar protein sorting pathways Curr. Opin. Cell. Biol. 10:512-522. Whallon, J. H. 1998. Introduction to laser scanning confocal microscopy. Laser Scanning Microscope Laboratory, Michigan State University. East Lansing, Mich. Wheeler, M. H. and A. A. Bell. 1988. Melanins and their importance in pathogenic fimgi. Current Topics in medical Mycology. 2:338-387. Wicklow, D. T., and O. L. Schotwell. 1983. Intrafungal distribution of aflatoxins among conidia and sclerotia of Aspergillus flaws and Aspergillus parasiticus. Can. J. Microbiol. 29:1-5. Wild, C. P., G. J. Hudson, and G. Sbbioni. 1992. Dietary intake of aflatoxins and the level of albumin-bound aflatoxin in peripheral blood in The Gambia, West Afiica. Cancer Epidemiol. Biomarkers Prev. 1:229—234. Wilson, D. L. 1998. Expression analysis of an aflatoxin nor-1 luidA reporter fusion in Aspergillus parasiticus. Master Thesis, Michigan State University, East Lansing. Wilson, D. M., and G. A. Payne. 1994. Factors Affecting Aspergillus flavus Group Infection and Aflatoxin Contamination of Crops. In Eaton, D.L. and Groopman, J .D. (Eds), The Toxicology of Aflatoxins. Academic Press, SanDiego. pp309-346. Witteveen, C. F.B., M. Veenhuis, and J. Visser. 1992. Localization of glucose oxidase and catalase activities in Aspergillus niger. Appl. Environ. Microbiol. 58:1190-1194. Woloshuk, C. P., K. R. Foutz, J. F. Brewer, D. Bhatnagar, T. E. Cleveland, and G. A. Payne. 1994. Molecular characterization of aflR, a regulatory locus for aflatoxin biosynthesis. Appl. Environ. Microbiol. 60:2408-2414. 139 Woloshuck, C. P. and R. Prieto. 1998. Genetic organization and function of the afaltoxin Bl biosynthetic genes. FEMS Micro. Lett. 160:169-176. Xu, J.R., and J. E. Hamer. 1995. Spore PCR. Fungal Genet Newsl. 42:80. Yan, K., J. Khoshnoodi, V. Ruotsalainen., and K. Tryggvason. 2002. N-link glycosylation is critical for the plasma membrane localization of Nephrine. J. Am. Soc. Nephrol. 13:1385-1389. Yang, M., H. Zhou, R.Y. C. Kong, W. F. Fong, L. Ren, , X. Liao, Y. Wang, W. Zhuang, and S. Yang. 1997. Mutations at codon 249 of p53 gene in human hepatocellular carcinomas from Tongan, China. Mutat. Res. 381:25-29. Yu, J., J. W. Cary, D. Bhatnagar, T. E. Cleveland, N. P. Keller, and F. S. Chu. 1993. Cloning and characterization of a cDNA from Aspergillus parasiticus encoding an O- methyltransferase involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 59:3546- 3571 . Yu, J., Chang, P. K., J. W. Cary,D. Bhatnagar, and T. E. Cleveland. 1997. avnA, a gene encoding a cytochrome P-450 monooxygenase, is involved in the conversion of averantin to averufin in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 63: 1349-1356. Yu, J., P. K. Chang, E. C. Ehrlich., J. W. Cary, B. Montalbano,. J. M. Dyer, D. Bhatnagar, and T. E. Cleveland. 1998. Characterization of the critical amino acids of an Aspergillus parasiticus cytochrome P-450 mono-oxygenase encoded by ordA involved in biosynthesis of aflatoxin B1, G1, B2 and 02. Appl. Environ. Microbiol. 64:4834-4841. Yu, J., P. K. Chang, D. Bhatnagar, and T. E. Cleveland. 2000. Cloning of a sugar utilization gene cluster in Aspergillus parasiticu. Biochem. Biophy. Acta 1493:211-214. Yu, J., P. K. Chang, D. Bhatnagar, and T. E. Cleveland. 2002. Cloning and fimctional expression of an esterase gene in Aspergillus parasiticus. Mycopathologia. 156:227-234. Yu, J. H., J. Weisser, and J. H. Adams. 1996. The Aspergillus F le RGS domain protein antagonizes G-protein signaling to block proliferation and allow development. EMBO J. 15:5184-5190. Zhou, R. 1997. The function, accumulation, and localization of the Nor-1 protein involved in aflatoxin biosynthesis; the function of the fluP gene associated with sporulation in Aspergillus parasiticus. Ph.D. Dissertation, Michigan State University, East Lansing. Zhou, R., and J. E. Linz. 1999. Enzymatic function of the Nor-l protein involved in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 65:5639- 5641. 140 Zhou, R., R. Rasooly, and J. E. Linz. 2000. Isolation and analysis of fluP, a gene associated with hyphal growth and sporulation in Aspergillus parasiticus. Mol. Gen. Genet. 264:514-520. Zuber, C., M. J. Spiro., B. Guhl, R. G. Spiro., and J. Roth. 2000. Golgi apparatus irnmunolocalization of endomannosidase suggests post-endoplasmic reticulum glocose trimming: implications for quality control. Mol. Biol. Cell. 11:4227-4240. 1' 141 11111111111111111111111