mm. . if.» «.1 mm”. H I. .15... .2, 3. .V {’5 .a I y . 2 .. :41! £7.13! 3‘31‘3 III. . V 2&1 memy a 6:2" ‘ ‘ 5 Michigan State University This is to certify that the dissertation entitled Functional Analysis of OmtA and Subcellular Localization of Aflatoxin Enzymes in Aspergillus parasiticus presented by Li-Wei Lee has been accepted towards fulfillment of the requirements for the Doctoral degree in Food Science and Environmental Toxicology "lflw2;¢d;€/ Major Professor’s Sdé‘hature 1/«Z flbB\ FluG 4L Asexual \7‘ flbC ‘3' sporulation , —Flb ~' 3. FadA-GT? + SfaD (Y) ——> Colony growth , Sterigmatocystin T T Biosynthesis ND Figure 1.1. A model for genetic control of Aspergillus growth, asexual sporulation and sterigmatocystin biosynthesis. 161,000 ppb of aflatoxin in wild-type spores) (Palmgren et al., 1986). This result implied that aflatoxin is synthesized mainly in mycelia. The question of whether aflatoxin is synthesized in spores or requires transport to spores remains unanswered. 1.4. Aflatoxin biosynthesis. Aflatoxin gene cluster. Aflatoxins are secondary metabolites produced by filamentous fungi, including A. flavus, A. parasiticus, A. nomius and A. pseudotamarii (Ito et al., 2001). More than 17 enzymes are involved in aflatoxin biosynthesis and the genes encoding these proteins are found clustered in a 75 kb genomic DNA fragment (Figure 1.2; and Bhatangar et al., 2003). Recently, it was demonstrated that location within the aflatoxin cluster plays an essential role in the regulation of aflatoxin gene expression. When nor-1 was relocated to other chromosomal positions outside the cluster (pyrG and niaD), the expression of this gene was highly reduced (Chiou et al., 2002). An enhancer element is proposed to be located inside or surrounding the cluster to positively regulate aflatoxin gene expression. In addition, clustering of secondary metabolite genes may provide an advantage for horizontal gene transfer and, therefore, favor cluster dispersal and survival as a unit (Walton, 2000). Dispersed pathway genes can be transmitted vertically but do not favor horizontal transfer because a single pathway gene has no advantage to the recipient host. The synthesis of secondary metabolites may be subject to the natural selection to benefit the producing fungi. A dothistromin gene cluster, homologous to the aflatoxin and sterigmatocystin clusters, was isolated from the pine needle pathogen Dathistrama pim' (Bradshaw et al., 2002). Dothistromin is a difuranoanthraquinone toxin with structure similar to versicolorin B. However, norB cypA One pathway-specific I regulator aflR aflJ adhA estA norA ver- 1 A verA avnA verB ava omtB omtA ordA vbs fl— ¢— —O —O-D—O IOQ—Q-I-C-CIQ—O d—I-Dfi—I-DQ—Q—I-b cpr moxY 0rdB ‘- Figure 1.2. Schematic presentation of the aflatoxin gene cluster. Arrows indicate the direction of gene transcription. dothistromin can not be converted to aflatoxin by A. parasiticus, suggesting that A. parasiticus has no enzyme to convert this toxin to an aflatoxin intermediate. One DNA fragment containing four biosynthetic genes, datA, B, C and D was cloned from D. pini. The amino acid sequences of dotA and dotC have 80.2 and 31.2% identity to ver-l and aftT (toxin transporter), respectively. Because dothistromin is a phytotoxin targeted to mature pine embryos and leaf callus, DotC was proposed to transport dothistromin out to the target in host plant. This proposed general secretion mechanism may protect fungal cells from this toxin. Aflatoxin genes. aflR. aflR encodes a pathway-specific regulator for many genes in the aflatoxin gene cluster. AflR can bind to a consensus DNA sequence, 5’-TCG N5 CGA-3’ (consensus AF LR binding site) in the promoters of regulated genes. Disruption of aflR eliminated accumulation of aflatoxin and pathway enzymes (Cary et al., 2002) except for the transcript from aflJ (Chang and Yu, 2002). The function of AflR may be necessary but not sufficient to activate certain pathway genes whose promoters do not contain perfect AflR binding site, for example, ava. Other transcription factors may be also involved in the regulation of aflatoxin genes and responsible for basal level of aflatoxin gene expression. :11”. aflJ is located adjacent to aflR in the aflatoxin gene cluster. A gene disruption experiment showed aflJ is involved in Aflatoxin biosynthesis. No aflatoxin was detected in aflJ mutant while other aflatoxin gene transcripts including pksA, nor-1, ver—l and omtA were detected (Meyers et al., 1998). The actual function of this gene is not clear. No aflJ homolog can be found in DNA/protein sequence databases to date. Based on the Peroxisomal Targeting Sequence 2 (PTS2) found in this gene, one proposed function of Afll is the transport of Aflatoxin proteins into microbodies (Meyers et al., 1998). Recently, several reports suggested that AflJ may function to regulate early— pathway gene transcription (Du, ASM Meeting 2000) possibly through interaction with AflR as demonstrated in a yeast two hybrid system (Bhatangar, 2000; Chang, 2003). fasl, fas2 and pks. The structural genes involved in aflatoxin biosynthesis include pksA, fas-IA, fas-2A, nor-1, avnA, ava, ver—I, omtB, omtA, ordA and vbs (Figure 1.2; Bhatangar et al., 2003). The 6-carbon compound hexanoyl-CoA is the starter unit in aflatoxin biosynthesis and is generated by fatty acid synthases Fasl and FasZ using one acetyl-CoA and two malonyl-CoA molecules (Mahanti et al., 1996; Watanabe et al., 1996). Malonyl-CoA and acetyl-CoA are first converted to hexanoyl-CoA by these two enzymes. The hexanoyl-CoA is further extended to noranthrone (NA) by polyketide synthase (PKS) using 7 molecules of malonyl-CoA (Minton and Townsend, 1997). nor-1. Norsolorinic acid (NOR) is the first stable intermediate in aflatoxin biosynthesis. NOR is convert to averantin (AVN) by the enzyme Nor-1 (Zhou and Linz, 1999). Mutation of the nor-1 gene in A. parasiticus greatly reduced aflatoxin production and resulted in accumulation of the red pigment, NOR (Trail et al., 1994). A low level of aflatoxin production remained suggesting that an alternative pathway or enzyme(s) could substitute Nor-I function but with lower enzymatic efficiency. avnA. The reaction to convert AVN to 5’-hydroxyaverantin (HAVN) is catalyzed by the enzyme AvnA. Disruption of avnA in A. parasiticus resulted in a non-aflatoxin producing strain and accumulation of the aflatoxin precursor AVN (Yu et al., 1997). AvnA is a cytochrome P-450 monoxygenase and contains 495 amino acids with a calculated molecular mass of 56.3 kDa. 10 adhA and norA. HAVN can be converted to averufin (AVR) by AdhA and NorA (Bhatangar et al., 2003). An adhA mutant was reported to produce trace amounts of aflatoxins (Chang, ASM Meeting 2000). Based on the observation that adhA and nor-1 mutants still produced aflatoxins, it seems likely that enzymes involved in the early stages of secondary metabolism share some activities with other primary metabolic enzymes. ava. AVR is converted to versiconal hemiacetal acetate (VHA) by Ava (Yu et al., 2000) and may also involve Cpr (Bhatangar et al., 2003). ava is located downstream of omtB and upstream of verB. Ava contains 285 amino acid with a calculated molecular mass of 30.9 kDa (Yu et al., 2000). The function of Ava has not been determined yet but it may function as a dehydrogenase based on its amino acid sequence. estA. The conversion of VHA to versiconal (VAL) is catalyzed by an esterase (Anderson and Green, 1994). The gene thought to encode this esterase is estA, located downstream of adhA and upstream of narA in the gene cluster (Yu, et al., 2002). Based on the cDNA sequence, this enzyme has a putative molecular mass of 34 kDa (Yu, et al., 2002). In the native form, however, this enzyme may form a dimer of 60 kDa as shown by gel filtration (Kusumoto and Hsieh, 1996). The expression pattern of estA may differ from other aflatoxin genes based on its ability to be expressed in non-aflatoxin inducing medium (PMS) (Anderson and Green, 1994; Yu, et al., 2002). However, this conclusion was based on RT-PCR; additional data generated from different methods such as Western blot analysis or Northern blot analysis may be required to confirm this statement. The second copy of this gene, estA2, located outside of this gene cluster, however, was not expressed in either inducing or non-inducing media (Yu, et al., 2002). This provides 11 additional evidence for a positional effect on the aflatoxin gene cluster reported previously (Chiou et al. 2002). vbs and verB. The side-chain of Versiconal (VAL) is cyclized by Versicolorin B synthase (VBS) to generate Versicolorin B (VERB) (Lin and Anderson, 1992; Silva, 1996 and 1997). The bisfuran ring structure of AFB, is formed at this step. VBS may require glycosylation to obtain enzyme activity (Silva, 1996 and 1997). The native VBS purified from A. parasiticus has a molecular mass of 78 kDa. Compared to the deduced molecular mass of 70.3 kDa, a post—translational modification (glycosylation) of the native protein was suggested. A double bond is introduced into the bifuran structure of VERB by a desaturase called VerB (verB) to generate versicolorin A (VERA). The bifuran ring is responsible for the mutagenic nature of aflatoxin AF 8,. ver-l and moxY. The conversion of versicolorin A (VERA) to demethyl sterigmatocystin (DMST) may require several steps (oxidation, keto-reduction, and decarboxylation) and several enzymes (Keller et al., 1995). Ver-l, a keto-reductase, was confirmed to be involved in this conversion (Liang, 1996 and 1997). Based on similarity with a polyhydroxynaphalene reductase (66% identity and 82% similarity) which is involved in melanin biosynthesis in Magnapartha grisea, the proposed function for Ver—l is to catalyze a deoxygenation of VA to 6-deoxy VA. Disruption of the ver-I gene in Aspergillus parasiticus NR-l (NiaD') resulted in a non-aflatoxin producing strain, VAD- 102 (Liang et al.,1996). The moxY gene, located in the aflatoxin gene cluster, also was suggested to be involved in conversion of VERA to DMST (Bhatnagar et al., 2003). However, no specific data were presented to support this idea. In A. nidulans, stcS, which encodes a monooxygenase, was demonstrated to be involved in VERA to DMST conversion (Keller et al., 1995). 12 omtB. The conversion of DMST to sterigmatocystin (ST) is carried out by OmtB (Yaba et al., 1989 and Yu et al., 2000). The calculated molecular mass of OmtB is 43.1 kDa. OmtB as well as OmtA are the only two methyl-transferases utilized in the aflatoxin biosynthetic pathway. omtB is located downstream of omtA and upstream of ava in the aflatoxin gene cluster. The promoter of omtB contains an AflR binding site but the downstream gene, ava, does not. The intergenic region between the omtB translation stop codon and the start codon of ava is only 173 bp. Yu et al. (2000) suggested that ava may be co-transcribed with omtB from the omtB promoter and subsequently the transcript is processed into two mRNAs. omtA. OmtA is a methyltransferase that catalyzes the conversion of ST to O- methyl-sterigmatocystin (OMST) by transferring a methyl group from the cofactor S- adenosyl-methionine (SAM) (Bhatnagar etal., 1987 and Cleveland et al., 1987). Although the molecular structure of OMST is close to ST, several chemical characteristics of OMST are more similar to AF 8,. Under UV illumination, the fluorescence emitted from ST shows a weak brick-red color while emission is pale blue from OMST and blue from AF B,. The migration rate (Rf) in TLC analysis showed that the Rf of ST is 0.97, OMST is 0.44, and AFB, is 0.37 using ether-methanol-water (96:3:1) as a development system. In a chloroform-acetone (95:5) development system, the Rfof ST is 0.74, OMST is 0.24 and AFBI is 0.22 (Bhatnagar et al., 1987). The first native OmtA purified from a cell-free extract of A. paraciticus showed one band on a non-denaturing gel with a molecular mass of 168 kDa. This protein was shown to contain two subunits, 110 and 58 kDa, by SDS-PAGE (Bhatnagar er al., 1988). This large enzyme was extremely unstable and displayed a low reaction rate in vitro. Therefore, Bhatnagar et al. suggested that a (hydrophobic) environment is required to 13 stabilize the hydrophobic ST (substrate), SAM (cofactor), OMST (product) and this enzyme (Bhatnagar et al., 1988). However, using a similar method, a different protein (OmtA) was purified with the same O-methyl-transferase activity. The second purified protein had a smaller molecular mass of 40 kDa (Keller et al., 1993; Liu et al., 1993). Antisera raised against the 40 kDa purified protein was used to screen a cDNA expression library resulting in the cloning of omtA (Yu et al., 1993). The molecular mass derived from the OmtA cDNA is 46 kDa. Comparing the N-terminal sequence of the purified protein (40 kDa) with the deduced amino acid sequence (46 kDa), Yu et al. suggested that OmtA might contain a leader sequence of 41 amino acids. This leader sequence may function to translocate OmtA across a membrane structure (Yu et al., 1993). To explain the relationship between the two purified enzymes, 40 kDa and 168— kDa, Yu et al. suggested that OmtA subunits were posttranslationally modified and polymerized to become the large protein (Yu et al., 1993). However, the antibody generated from the large protein (168- kDa) has very little reactivity with the 40 kDa OmtA protein, suggesting that these two enzymes are different proteins (Keller, 1993). No further study was carried out on this large purified protein. In the current study, we conducted a knockout of the omtA gene in A. paraciticus C810 and conducted an in viva feeding experiment. Our results suggested that OmtA (46 kDa) plays a major role in conversion of ST to OMST in viva (Lee et al., 2002). ordA. OrdA is the final enzyme involved in the aflatoxin biosynthetic pathway. The proposed reactions for the conversion of OMST to AFB, include parahydroxylation, demethylation, reduction, ring cleavage, and conversion of a six-membered ring to a five- membered ring (Bhatnagar et al., 1991). Therefore, more than one enzyme was proposed to be involved in this conversion. In one early study, the partially purified oxidoreductase 14 showed four major bands on a non-denaturing gel with a molecular mass about 180 - 200 kDa and seven major proteins (165, 105, 88, 64, 43, 38, and 26 kDa) on SDS-PAGE (Bhatnagar et al., 1991). However, none of the bands excised from the native gel showed enzymatic activity. Recently, ordA encoding an enzyme activity able to convert OMST to AF B1 was cloned from A. flavus (Prieto and Woloshuk, I997). The putative molecular mass of OrdA is 60.2 kDa. Based on the amino acid sequence, OrdA is predicted to function as a monooxygenase. Yeast transformed with a single gal l-ordA construct (ordA fused with inducible gall promoter) were able to convert exogenous OMST to AFB, (Yu et al., 1998). Because no other intermediates between OMST and AFB, were reported, it is possible that this final reaction step requires only OrdA. However, the possibility can not be ruled out that other enzymes may help OrdA in more general reactions rather than the specific step in both filamentous fungi and yeast. The co-purified proteins with oxidoreductase activity observed in the early study could represent these helper proteins (Bhatnagar et al., 1991). Two additional proteins were identified that are required for production of the G group aflatoxins and these are detected in microsomal and cytosolic fractions (Yabe et al., 1999). 1.5. Organelles involved in secondary metabolism in filamentous fungi- the microbody and vacuole. For most secondary metabolites including aflatoxin, the biological functions are not clear. Specific secondary metabolites (including penicillin and preheliminthosporol) are believed to enhance the survival of the producing organism by inhibition of other micro-organisms in the same grth environment (Calvo etal., 2002; Akesson, et al., 15 1996.). Characterization of the sub-cellular compartments of secondary metabolites can aid in understanding their biological functions; storage and secretion mechanism, and even efficient production via industrial fermentation. Penicillin is perhaps the best-studied secondary metabolite (Van der Kamp et al., 1999). The biosynthesis of penicillin in Penicillium chrysagenum involves three enzyme activities that are distributed in at least in two sub-cellular compartments. The major storage compartment for the substrates, or-aminoadipic acid, L-cysteine and L-valine, is the vacuole (Van der Kamp et al., 1999). The first and second pathway enzymes were localized in the cytosol (Van der Lende, 2002). The final enzyme was localized in the microbody (Muller et al., 1991; Van der Lende, 2002). It was suggested that penicillin is possibly synthesized and stored in this membrane-bound organelle (Muller et al., 1991). To have full enzyme activity, it is important for biosynthetic enzymes to be confined in the proper environment under conditions in which optimal pH, required cofactors and a transport system for substrate are provided. Based on these observations, 1 hypothesized that microbodies and vacuoles represent two compartments for aflatoxin synthesis. The biogenesis, protein import pathway and biological function of these two organelles are briefly reviewed in the following sections. Microbody. Micobodies can be divided into two subgroups based on the enzymes they contain: peroxisomes (contain flavin-linked oxidase) and glyoxysomes (contain enzymes of the glycoxylate cycle). The specific fiinction of microbodies also varies in different cell types. In hepatocytes, peroxisomes can catalyze several metabolic pathways including the biosynthesis of ether lipids, catabolism of long chain fatty acids and production of cholesterol (Faumgart and Baumgart, 1993). Unlike animals where B- 16 oxidation occurs primarily in the mitochondria, B-oxidation in plants and fungi is exclusively a peroxisomal function (Waterham and Cregg, 1997). The biogenesis of peroxisomes can occur either by fission or budding from an existing peroxisome or proliferation independent of a pre-existing peroxisome (Markham et al., 1994; Rapp et al., 1996; Valenciano et al., 1998). The proliferation of peroxisomes involves several maturation steps- from low-density organellar structures (pre- peroxisome) to mature peroxisomes (Hettema et al., 1998). The low-density organellar structures fine-peroxisome) contain several peroxisomal proteins including Pex10p and ScPexl 1p that are responsible for independent peroxisome proliferation in yeast (Waterham et al., 1997). PexlOp (34-48 kDa) is a peroxisomal integral-membrane protein. ScPexl 1p (27 kDa) is a peroxisomal membrane-associated protein. Induction of both genes leads to an over-proliferation of peroxisomes. Proteins targeted to the peroxisome are mainly dependent on the Peroxisomal Targeting Sequence (PTS) on the protein. Three kinds of PTS have been identified: PTS- 1, 2 and internal PTS (Subramani, 1998). PTS-1 and PTS-2 are commonly found in peroxisomal proteins whereas the internal PTS is not yet clearly defined. PTS-l is a tripeptide sequence, Ser-Lys-Leu (SKL) or conserved variants (S/A(/C))(K/R/H)L, located at the carboxyl-terminus (Waterham et al., 1997). PTS-l is not cleaved after import into the peroxisome. PTS-2 is an amino-terminal motif with a consensus sequence Arg-Leu-Xs-His/Gln-Leu (RLX5H/QL). PTS-2 may be cleaved after import to the peroxisome (Waterham et al., 1997). Internal PTS motifs may be contained in some peroxisomal matrix proteins with or without PTS-1 and/or PTS-2. For example, the yeast catalase A with PTS-l deleted is still able to be imported into the peroxisome (Purdue and Lazarow, 1994). The potential position of internal PTS for this protein is located 17 between residues 104 and 126. Since the receptor for internal PTS has not been identified, a proposed function for intemal PTS motifs is to associate and co-translocate with other PTS-l-containing proteins into the peroxisome (Subramani, 1998). Unlike mitochondria and ER proteins that require an unfolded conformation for translocation, peroxisomal proteins can be imported in either extended monomeric conformation or large preassembled structures (Waterham and Cregg, 1997). The monomer without a PTS-2 sequence can be dimerized with a PTS-Z-containing monomer and translocated into peroxisomes (Rachubinski and Subramani, 1995). Formation of large structures via PTS-1 receptor-PTS-2 receptor by protein-protein interaction has been demonstrated in a yeast system (Rachubinski and Subramani, 1995; Tabak et al., 1999). The yeast (Saccharomyces cerevisiae) PTS-1 receptor (PTS-1R5), Pex5p, contains a loosely conserved 34 amino acid repeat called the tetratricopeptide (TPR) domain. The yeast PTS-2 receptor (PTSZR), Pex7p, contains six copies of a motif of 40 amino acid residues (WD40 motif). WD is a Trp (VD-Asp (D) pair at the carboxyl terminus of the motif. The results from two-hybrid experiments indicate that the binding between Pex7p and Pex5p is mediated by interaction between the TPR domain and the WD motif (Rachubinski and Subramani, 1995). Other proteins such as Pexlp and Pex6p with AAA domain (ATPases associated with diverse cellular activities) are also involved in protein import (Tabak et al., 1999). These proteins are associated with vesicles and interact in an ATP-dependent manner in Pichia pastoris (Tabak et al., 1999). A simple model for operation of a PTS—containing protein-import pathway is demonstrated in Figure 1.3 (modified from Tabak et al., 1999). First, a PTS-containing protein forms a complex with its receptor in the cytosol. Second, this complex binds to docking proteins (translocon) in and on the peroxisomal membrane. Third, the PTS 18 protein is released into the peroxisomal matrix and the receptor is recycled. The docking proteins consist of three proteins, Pex13p (an integral peroxisomal membrane protein), Pexl4p (a peroxisomal membrane-associated protein) and Pex17p (membrane-associated protein). Two additional helper proteins, hsp70—class cytoplasmic chaperones (Hsp70) and J-domain—containing proteins (Djplp), are required in the recognition of PTS- containing proteins by PTS receptors and/or the formation of PTS-1 and PTS-2 receptors complex during the translocation process (Hettema et al., 1998). Based on a PSORT search, the 6-aminopenicillanic acid acyltransferase from Penicillium chrysogenum and Aspergillus nidulans contain a carboxyl-tenninal tripeptide ARL (PTS-1). The experiments also confirmed that 6-aminopenicillanic acid acyltransferase is localized in microbodies. A PSORT search suggested that several aflatoxin enzymes contain PTS-l sequences but the locations of PTS-1 sequences in those enzymes are different from the suggested location (carboxyl-terminal). Among them, AflJ even contains both PTS—1 and 2, located at the appropriate sites in the protein. The PSORT report for aflatoxin enzymes is shown in Figure 1.4. So far, no published data have been obtained to confirm the functional significance of these targeting signals. Vacuole. The origin of vacuoles is not clear. It was suggested that the endoplasmic reticulum may be involved directly in vacuole biogenesis and the Golgi complex involved in increasing vacuole size (Klionsky, 1990). In vacuole biogenesis, more than 40 proteins are required. Mutations in vacuole biosynthetic genes usually result in fragmentation of the vacuole into small vacuoles, or vesicles, or unusual membrane inclusions (Ohsumi et al., 2002). Proteins such as newly synthesized hydrolytic enzymes destined for the vacuole are transported through the endoplasmic reticulum and Golgi 19 PTS protein 62 Helpers A Hel ers \/ \:/ P . Receptors Peroxisomal l Translocon membrane L Peroxisomal matrix © Figure 1.3. A schematic presentation of the peroxisomal-targeting sequence-dependent protein-import pathway for proteins destined for the peroxisomal matrix. 20 & AflJ l. SKL motif (signal for peroxisomal protein): ARL (a.a. 354; n.t. 438) 2. SKL2: 2nd peroxisomal targeting signal: found [KLASIPLHL at a.a. 286] Nor-1 SKL motif (signal for peroxisomal protein): ARL (a.a. 225; n.t. 271) AvnA 1. SKL motif (signal for peroxisomal protein): SHL (a.a. 211; n.t. 396) 2. Peroxisomal proteins: Status: positive VBS SKL motif (signal for peroxisomal protein): No. Ver-l SKL motif (signal for peroxisomal protein): AKL (a.a. 75; n.t. 262) OmtA SKL motif (signal for peroxisomal protein): SHL (a.a. 81; n.t. 418) OrdA 1. GvH: Examining signal sequence (von Heijne) Possible cleavage site: 23 2. Seems to have a cleavable N-term signal sequence. 3. SKL motif (signal for peroxisomal protein): AHL (a.a. 84; n.t. 528) B_._ Acyl coenzyme A: 6-aminopenicillanic acid acyltransferase (from Penicillium chrysogenum and Aspergillus nidulans) 21 1. SKL motif (signal for peroxisomal protein): SKL motif found at the C-terminus 2. Peroxisomal proteins: positive. Figure 1.4. PSORT (Prediction of Protein Localization Sites version 6.4) prediction for the protein targeting sequences (PTS-I and PTS-2) in (A) Aflatoxin enzymes, AflJ, Nor- 1, AvnA, VBS, Ver-l, OmtA and OrdA from A. parasiticus, and (B) 6-aminopenicillanic acid acyltransferase from Penicillium chrysogenum and A. nidulans. 22 secretion pathway (Klionsky, 1990 and 1997). These proteins are carried by vesicles and finally released into the vacuole. The process is completed‘by fusion between vesicles and the vacuole with the help of small GTPases. The small GTPases recycle between donor (vesicle) and receptor (vacuole) membranes (Ohsumi et al., 2002). GTP-bound GTPase directs the fusion of vesicle to the vacuole. Upon fusion, GTPase hydrolyzes GTP to GDP and is released from the membrane with the assistance of GTPase—activating protein. The GDP-bound GTPase then returns to the donor membrane and GDP is replaced by GTP with the assistance of guanine exchange factor. The vacuole is a single-membrane bound organelle with highly dynamic structure (Markham, 1994). Morphological changes in the vacuole occur rapidly and may correspond to the grth stage (Markham, 1994; Paul, et al., 1994). Hyphal vacuolation and fragmentation was only observed after a rapid growth phase during fungal fermentation (Paul, at al., 1994). Recently, it was discovered that the vacuole is not just an empty space as its electron microscopic appearance suggests but an organelle with multiple possible functions (Markham, 1994). Therefore, vacuoles may consist of a series of functionally distinct subclasses of organelles to compartmentalize a variety of biological activities (Markham, 1994). The detailed functions of vacuoles in filamentous fungi are not completely understood but may be similar to those in yeast (Klionsky, 1990). Here, vacuoles are large storage sites for metabolites, ions and amino acids. Ions (Cafi, Mg”, Zn”, Fe”), including potential toxic ions (Srii, Co“, Pb”), and inorganic phosphorus (in the form of polyphosphate) (Markham, 1994). S-adenosyl-methionine (AdoMet) is an important methyl donor involved in a great variety of transmethylation reactions. In S. cerevisiae, S-adenosyl-methionine accumulates to high levels (about 60% of intracellular S-adenosyl-methionine) in vacuoles by a specific transport system 23 (Farooqui et al., 1983; Schwencke et al., 1976; Svihla et al., 1969). S-adenosyl- methionine in vacuoles can be directly detected by microscopy under utraviolet light (at a wavelength of 260 or 265 nm) (Farooqui et al., 1983; Schwencke et al., 1976; Svihla et al., 1969). Results from subcellular fractionation revealed that two metabolically distinct S-adenosyl-methionine pools exist in yeast. The labile pool exists in the cytosol while the stable pool is in the vacuole (Farooqui et al., 1983). The vacuole is also an important nitrogen reserve. Basic amino acids like arginine are concentrated in vacuole. The amino acids a-aminoadipic acid, L-cysteine and L-valine used in penicillin synthesis are also stored in vacuoles. Recently, the vacuole was found to be an important compartment for lipid degradation and glycerol production at a specific stage during appressoria formation in Magnaparthe grisea , suggesting that partial or complete enzymatic pathways may be compartmentalized in vacuoles (Weber et al., 2001). Vacuoles control the intracellular pH homeostasis (Thumm, 2000). The pH in vacuoles is maintained at pH6 in Neurospora crassa. The membrane ATPase can pump H+ from the cytosol into the vacuole, generating a membrane potential of about 25-40 mV in N. crassa. This electron-chemical potential is the driving force to transport ions and amino acids across the vacuolar membrane. Vacuolar hydrolases (e.g. aminopeptidase I) mature in this organelle after cleavage of a signal peptide. Fungal vacuoles are an equivalent of mammalian lysosomes in that they can be involved in degradation and recycling of unneeded proteins and even whole organelles (Amor et al., 2000). In Saccharomyces, cytoplasmic proteins can be transported into the vacuole by macroautophagy or the cytoplasm to vacuole targeting (cvt) pathway without involvement of receptors. These two pathways are regulated by nutrients (e.g. glucose or ethanol) and nutritional conditions (nutrition limitation) (Klionsky, 1997 and Thumm, 2000). 24 1.6. Compartmentalization of penicillin in Penicillium chrysogenum. The biosynthesis of penicillin in Penicillium chrysagenum involves three enzyme activities and three amino acids (or-aminoadipic acid, L-cysteine and L-valine) (Van der Kamp et al., 1999). In the first reaction, or-aminoadipic acid, L-cysteine and L-valine are condensed to a 6-(L-a-aminoadipyl)-L-cysteinyl-D-valine tripepetide (ACV) catalyzed by the enzyme 8—(L-or-aminoadipyl)-L-cysteinyl-D-valine synthetase (ACV synthetase). The substrate amino acids (a-aminoadipic acid, L-cysteine and L-valine) are stored in vacuoles. In early studies, ACV synthetase was localized in two subcellular compartments, a Golgi-like organelle (Kurylowicz et al., 1987) and vacuoles (Lenderfeld et al., 1993). Recently, ACV synthetase was also suggested to localize in the cytosol (Van der Lende et al., 2002). The second reaCtion is conversion of ACV to isopenicillin N (IPN) by the enzyme, isopenicilin N-synthetase. This enzyme has been localized to Golgi-like organelles (Kurylowicz et al., 1987) and the cytosol (Van der Lende et al., 2002). Most evidence suggests that the cytosol is the true compartment for this enzyme (Muller et al., 1991 and Van der Lende et al., 2002). The final reaction entails the conversion of IPN to penicilin G by Acyl-coenzyme A: isopenicillin N acyltransferase (IAT). IAT was localized in peroxisomes (microbodies) (Mfiller et al., 1991 and Van der Lende et al., 2002). In this step, the aminoadipyl side-chain of IPN is substituted by phenylacetyl (PA) or phenoxylacetyl (POA) to yield penicilin G or penicilin V, respectively (van de Kamp et al., 1999). Prior to this reaction, the PA and POA have to be activated to their thioesters (PA-CoA and POA-CoA) by acetyl-CoA synthetase (ACS) or PA-CoA ligase (PCL). ACS is a cytosolic enzyme. PCL contains a PTS-1 sequence, 25 suggesting that this enzyme may target to microbodies where it could function together with IAT (van de Kamp et al., 1999). Based on localization results, Miiller et al. suggested that penicillin is stored in the microbody together with IAT (Mfiller et al., 1991). To summarize these studies, Van der Lende et al. suggested that the penicillin biosynthetic activities are likely confined to two major compartments, the cytosol and microbodies (Van der Lende et al., 2002). Other compartments observed in earlier studies were proposed to be artifacts generated from the experimental procedures. A detailed discussion is presented in two publications (Muller et al., 1991 and Van der Lende etal., 2002). For the localization of ACV synthetase in vacuoles, van der Lende et al. suggested that this artifact occurred during the isolation of protoplasts and organelles (Van der Lende et al., 2002). Because the procedure is time-consuming (one day) and is conducted under nutrition starvation conditions (cell wall digestion), this enzyme potentially could relocate from the cytosol into the vacuole for protein degradation. For the localization of ACV synthetase and isopenicilin N-synthetase in Golgi-like organelles, Muller et al. suggested that organelles would be disrupted during grinding of mycelia and subsequent fractionation procedure (Muller et al., 1991). The broken organelles could be reformed into new organelles with morphology similar to other organelles. To avoid potential artifacts introduced by vigorous cell disruption and time- consuming organelle isolation procedures, another strategy for in situ localization of protein is immuno-electron microscopy. In immuno-electron microscopic analysis, the sample is processed first by fixation to prevent protein relocation and organelle disruption. Therefore, the fixation method plays a crucial role in this analysis. In general, the sample can be fixed by chemical crosslinking (chemical fixation) or low temperature freezing (cryofixation). It is widely accepted that it is very difficult to obtain 26 simultaneously good preservation of both cellular ultrastructure and protein antigenicity. For example, chemical fixation usually requires a longer incubation time because crosslinking will inhibit chemical penetration and slow down the fixation process. Prolonged crosslinking with chemicals like glutaradehyde may provide a good preservation of ultrastructure but result in loss of antigenicity. The cryofixation method is commonly considered superior to chemical fixation in preservation of ultrastructure and protein antigenicity. This method physically fixes cellular components in a very short time (ms) without any modification of protein antigenicity (Akesson et al., 1996 and Muller et al., 1991). However, membrane components are easily extracted during substitution of water with organic solvent. Although a low concentration of chemical fixative can be added in the substitution solvent to maintain the membrane structure, this may cause the loss of antigenicity. 27 REFERENCES Adams T. H., J. K. Wieser, and J-H. Yu. 1998. Asexual sporulation 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. Akesson, H., E. Carlemalm, E. Everitt, T. Gunnarsson, G. Odham, and H-B Jansson. 1996. Immunocytochemical localization of phytotoxins in Bioplaris sorokiniana. Fungal Gene. Biol. 20:205-216. Amor, C., A. I. Dominguez, J. R. D. Lucas, and F. Laborda. 2000. The catabolite inactivation of Aspergillus nidulans isocitrate lyase occurs by specific autophagy of peroxisomes. Arch. Microbiol. 174:59-66. Anderson, J. A., and L. D. Green. 1994. Timing of apperence of versiconal hemiacetate esterase and versiconal cyclase activity in cultures of Aspergillus parasiticus. Mycopathologia. 126: 169-1 72. Asao, T., G. Buchi, M. M. Abdel-Kader, S.B. Chang, E. L. Wick, and G. N. Wongan. 1963. Aflatoxin B and G. J. Am. Chem. Soc. 85:1706-1707. Bannasch, P., N. I. Khoshkhou, and H. J. Hacker. 1995. Synergistic hepatocacinogenic effect of hepadnaviral infection and dietary aflatoxin B1 in woodchucks. Cancer Res. 55:3318-3330. Becroft, D. M. 0., and D. R. Webster. 1972. Aflatoxins and Reye’s disease. Br. Med. J. 4:117. Bhatnagar D., S. P. McCormick , L. S. Lee , and R. A. Hill. 1987. Identification of O- methylsterigmatocystin as an aflatoxin B1 and G1 precursor in Aspergillus parasiticus. Appl. Environ. Microbiol. 53:1028-33. Bhatnagar D., A. H. Ullah, and T. E. Cleveland. 1988. Purification and characterization of a methyltransferase from Aspergillus parasiticus SRRC 163 involved in aflatoxin biosynthetic pathway. Prep. Biochem. 18:321-49. Bhatnagar D., T. E. Cleveland, and D. G. 1. Kingston. 1991. Enzymological evidence for separate pathways for aflatoxin B1 and B2 biosnythesis. Biochimie. 30:434324350. 28 Bhatnagar D., K. C. Ehrlich, and T. E. Cleveland. 1992. Oxidation-reduction reactions in biosynthesis of secondary metabolites, p. 255-286. In D. Bhatnagar, E. B. Lillehoj, and D. K. Arora (ed.), Handbook of applied mycology, vol. 5. Mycotoxins in ecological systems. Marcel Dekker, Inc., New York. Bhatnagar D., K. C. Ehrlich, and T.E. Cleveland. 2003. Molecular genetic analysis and regulation of aflatoxin biosynthesis. Appl. Microbiol. Biotechnol. 61 :83-93. Bradshaw, R. E., D. Bhatnagar, R. J. Ganley, C. J. Gillman, B. J. Monahan, and J. M. Seconi. 2002. Dothistroma pini, a forest pathogen, contains homologs of aflatoxin biosynthetic pathway genes. Appl. Microbiol. Biotechnol. 68:2885-2892. Calvo, A. M., R. A. Wilson, J. W. Bok, and N. P. Keller. 2002. Relationship between secondary metabolism and fungal development. Microbiol. Mol. Biol. Rev. 66:447-459. Cary, J. W., 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 systems. Report 139. Chang P.-K., J. Yu, K. C. Ehrlich, S. M. Boue, B. G. Montalbano, D. Bhatnagar, and T. E. Cleveland. 2000. Dirsruption of the aflatoxin biosynthetic gene adhA in Aspergillus parasiticus leads to accumulation of 5'-Hydroxyaverantin. ASM 100th Meeting. Chang P.-K., and J. Yu. 2002. Characterization of a partial duplication of the aflatoxin gene cluster in Aspergillus parasiticus ATCC56775. Appl. Microbiolo. Biotechnol. 58:632-636. Chang P.-K. 2003. The Aspergillus parasiticus protein AF LJ interacts with the aflatoxin pathway-specific regulator AF LR. Mol. Gen. Genomics. 268:71 1-719. Chiou, C.-H., M. Miller, D. L. Wilson, F. Trail, and J. Linz. 2002. Chromosomal location plays a role in regulation of aflatoxin gene expression in Aspergillus parasiticus. Appl. Environ. Microbiol. 68:306-315. Cleveland T. E., A. R. Lax, L. S. Lee, and D. Bhatnagar. 1987. Appearance of enzyme activities catalyzing conversion of sterigmatocystin to aflatoxin Bl in late- growth-phase Aspergillus parasiticus cultures. Appl. Environ. Microbiol. 53:1711-3. D’Souza, C. A., B. N. Lee, and T. H. Adams. 2001. Characterization of the role of the F luG protein in asexual development in Aspergillus nidulans. Microbiol. Genetic. 158:1027-1036. 29 Du W. L., and G. A. Payne. 2000. Function of aflJ in regulating Aflatoxin Biosynthesis. ASM 100th General Meeting. Eaton D. L. and E. P. Gallagher. 1994. Mechanisms of aflatoxin carcinogenesis. Ann Rev. Pharmacol. Toxicol. 34:135-72. F arooqui, J. 2., H. W. Lee, S. Kim, and W. K. Paik. 1983. Studies on the compartmentation of S-adenosyl-L-methionine in Saccharomyces cerevisiae and isolated rat hepatocytes. Biochimica et Biophysica Acta. 757:342-352. Faumgart E. and E. Baumgart. 1993. Electron Microscopic cytochemistry in biomedicine. Charpter 9. Cell Organelles. p 491-502. by CRC Press, Inc. F eng, G. H., and T. J. Leonard. 1998. Culture conditions control expression of the genes for aflatoxin and sterigmatocystin biosynthesis in Aspergillus parasiticus and Aspergillus nidulans. Appl. Environ. Microbiol. 64:2275-2277. anleni, N., J. E. Smith, J. Lacey, and G. Gettinby. 1997. Effects of temperature, water activity, and incubation time on production of aflatoxins and cyclopiazonic acid by an isolate of Aspergillusflavus in surface agar culture. Appl. Environ. Microbiol. 63:1048-1053. Guzman-de—Pena, D., and J. Ruiz-Herrera. 1997. Relationship between aflatoxin biosynthesis and sporulation in Aspergillus parasiticus. Fungal Genet Biol. 21:198-205. Hettema E. H., C. C. M. Ruigrok, M. G. Koerkamp, M. van den Berg, H. F. Tabak, B. Distel, and I. Braakman. 1998. The cytosolic DnaJ-like protein djplp is involved specifically in peroxisomal protein import. J. Cell Biol. 142:421-34. 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 G alpha protein- dependent signaling pathway. EMBO J. 6:4916-23. Ito, Y., S. W. Peterson, D. T. Wicklow, and T. Goto. 2001. Aspergillus pseudotamarii, a new aflatoxin producing species in Aspergillus section Flavi. Mycol. Res. 105:233-239. Keller N. P., H. C. Dischinger, JR., D. Bhatnagar, T. E. Cleveland, and A. H. J. Ullah. 1993. Purification of a 40-kDa methyltransferase activity in the aflatoxin biosynthetic pathway. Appl. Environ. Microbiol. 36:270-273. Keller N. P., S. Segner, D. Bhatnagar, and T. E. Cleveland. 1995. stcS, a putative P- 450 monooxygenase, is required for the conversion of versicolorin A to sterigmatocystin in Aspergillus nidulans. Appl. Environ. Microbiol. 61:3628-3632. Klionsky, D. J., P. K. Herman, and S. D. EMR. 1990. Fungal vacuole: composition, function and biogenesis. Microbiol. Rev. 54:266-292. 30 Klionsky, D. J. 1997. Protein transport from the cytoplasm into vacuole. J. Membrane Biol. 157: 105-115. 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., and D. P. Hsieh. 1996. Purification and characterization of the esterase involved in the aflatoxin biosynthesis in Aspergillus parasiticus. Can. J. Microbiol. 8:804-810. Lee, L.-W., C.-H., Chiou, and J. E. Linz. 2002. Function of native OmtA in viva and expression and distribution of this protein in colonies of Aspergillus parasiticus. Appl. Environ. Microbiol. 68:5718-5727. 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., C. D. Skory and J. E. Linz. 1996. Characterization of the function of the ver-l and ver-IB genes, involved in aflatoxin biosynthesis in Aspergillus parasiticus. Appl Environ Microbiol. 62:4568-4575. Liang S-H., T -S. Wu, R. Lee, F. S. Chu, and J. E. Linz. 1997. Analysis of mechanism regulating expression of the ver-l gene, involved in aflatoxin biosynthesis. Appl Environ Microbiol. 63: 1058-1065. Lin, B.-K., and J. A. Anderson. 1992. Purification and properties of versiconal cyclase from Aspergillus parasiticus. Arch. Biochem. Biophy. 293:67-70. Lin B. H., N. P. Keller, D. Bhatnagar, T. E. Cleveland and F. S. Chu. 1993. Production and characterization of antibodies against sterigmatocystin O- methyltransferase. Food Agric. Immunol. 5:155-164. Luchese, R. H. and W. F. Harrigan. 1993. Biosynthesis of aflatoxin- the role of nutritional factors. J. Applied Bacteriology. 74: 5-14. Mahanti, N., D. Bhatnagar, J. W. Cary, J. Joubran, and J. E. Linz. 1996. Structure and function of fas-lA, a gene encoding a putative fatty acid synthetase directly involved in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 62:191- 195. Markham, P. 1995. Organelles of filamentous fungi. Pp. 75-98. In N. A. R. Gow, and G. W. Gooday (ed), The growing fungus, lst ed. Chapman & Hall, London, UK. 31 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. Minto, R. E., and C. A. Townson. 1997. Enzymology and molecular biology of aflatoxin biosynthesis. Chem. Rev. 97:2537-2555. Miiller W.H., T.P. van der Krift, A.J.J. Krouwer, H.A.B. Wosten, L.H.M. van der Voort, E.B. Smaal, and A.J. Verkleij. 1991. Localization of the pathway of the penicillin biosynthesis in Penicillium chrysogenum. EMBO Journal. 10:489-495. 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. Palmgren M.S., and L. S. Lee. 1986. Separation of mycotoxin-containg sources in grain dust and determination of threir mycotoxin potential. 66:105-108. 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. A. 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 Dekker, Inc., New York. Prieto R. and GP. Woloshuk. 1997. 0rd 1, an oxidoreductase gene responsible for conversion of O-methylsterigmatocystin to aflatoxin in Aspergillus/laws. Appl Environ. Microbiol. 63:1661-6. Purdue RE. and RB. Lazarow. 1994. Peroxisomal biogenesis: Multiple pathways of protein import. J. Biol. Chem. 268:30065-68. Rachubinski R.A., and S. Subraman. 1995. How proteins penerate peroxisomes. Cell. 83:525-528. Rapp S., R. Saffrich , M. Anton, U. J akle, W. Ansorge, K. Gorgas, and W.W. Just. 1996. Micritubule-based peroxisome movement. J. Cell Sci. 109:837-849. Reiss J. 1982. Development of Aspergillus parasiticus and formation of aflatoxin B1 under the influence of conidiogenesis affecting compounds. Arch. Microbiol. 133:236-8. Sell, 8., J. M. Hunt, H. A. Dunsford, and F. V. Chisari. 1991. Synergy between hepatitis B virus expression and chemical hepatocacinogens in transgenic mice. Cancer Res. 51:1278-1285. Schwencke, J., and H. De Robichon-Szulmajster. 1976. The transport of S-adenosyl- L-methionine in the isolated yeast vacuoles and spheroplasts. Eur. J. Biochem. 65:49-60. 32 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. Smela, M. E., S. S. Currier, E. A. Bailey, and J. M. Essigmann. 2001. The chemistry and biology of aflatoxin B1: from mutational spectrometry to carcinogenesis. Carcinogenesis. 22:535-545. Subramani, S. 1998. Components involved in peroxisome import, biogenesis, proliferation, turnover, and movement. Physiol. Rev. 78:171-88. Svihla, G., J. L. Dainko, and F. Schlenk. 1969. Ultraviolet micrography of penetration of exogenous cytochrome c into the yeast cell. J. Bacteriol. 100:498-504. Tabak, H. F., I. Braakman, and B. Distel. 1999. Peroxisomes: simple in the function but complex in maintenance. Trends in cell biology. 9:447-453. Thumm, M. 2000. Structure and function of the yeast vacuole and its role in autophagy. Microscopy Research and Technique. 51: 563-572. Trail F., P-K. Chang, J. Cary, 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. Valenciano, S., J.R. De Lucas, I. Van der Klei, M. Veenhuis, and F. Laborda. 1998. Charatcerization of Aspergillus nidulans peroxisomes by immunoelectron microscopy. Arch. Microbiol. 170:370-376. Van der Kamp, M., A. J. M. Driessen, and W. N. Konings. 1999. Compartmentalization 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. 5-(L-or- 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. Walton, J. D. 2000. Horizontal gene transfer and the evolution of secondary metabolite gene cluster in fungi: a hypothesis. Fungal Gene. Biol. 30:167-171. Watanabe, C. M. H., D. Wilson, J. E. Linz, and C. A. Townsend. 1996. Demonstration of the catalytic roles and evidence for the physical association of the type 33 I fatty acid synthases and a polyketide synthase in the biosynthesis of aflatoxin B 1. Chem. Biol. 3:463-469. Waterham H. R., and J. M. Cregg. 1997. Peroxisome biogenesis. Bioessays. 19:57-66. Weber, R. W. S., G. E. Wakley, E. Thines, and N. J. Talbot. 2001. The vacuole as element of the lytic system and sink for lipid droplets in maturing appressoria of Magnaparthe grisea. Protoplasma. 216: 101 -1 12. Wicklow, D. T., and O. L. Shotwell. 1983. Intrafungal distribution of aflatoxins among conidia and sclerotia of Aspergillus flavus and Aspergillus parasiticus. Can. J. Microbiol. 29:1-5. Wild, C. P., and P. C. Turner. 2002. The toxicology of aflatoxins as a basis for public health decisions. Mutagenesis. 17:471-481. 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. Yabe, K., Y. Ando, J. Hashimoto, and T. Hamasaki. 1989. Two distinct O- methyltransferases in aflatoxin biosynthesis. Appl Environ Microbiol. 55:2172-7. Yabe, K., M. Nakamura, and T. Hamasaki. 1999. Enzymatic formation of G-group aflatoxins and biosynthetic relationship between G- and B-group aflatoxins. Appl. Environ. Microbiol. 65:3867-3872. 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. 59:3564- 71. Yu J., P.-K. Chang, 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 averrufin in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 63:1349-1356. Yu, J., P. K. Chang, K. C. Ehrlich, .1. 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 monooxygenase encoded by ordA that is involved in the biosynthesis of aflatoxins B1, GI, 82, and G2. Appl. Environ. Microbiol. 64:4834-41. Yu, J., C. P. Woloshuk, D. Bhatnagar, and T. E. Cleveland. 2000. Cloning and Characterization of ava and omtB genes involved in aflatoxin biosynthesis in three ASpergillus species. Gene. 248: 1 57-167. 34 Yu, J., P. K. Chang, D. Bhatnagar, and T. E. Cleveland. 2002. Cloning and functional expression of an esterase gene in Aspergillus parasiticus. Mycopathologia. 156:227-234. 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. 35 CHAPTER 2 Genetic Disruption of omtA in Aspergillus parasiticus C810 INTRODUCTION Aflatoxins, highly toxic and carcinogenic secondary metabolites produced by the filamentous fungi Aspergillus parasiticus, A. flavus, A. nomius, and A. tamarii (6,11), frequently contaminate food and feed crops such as corn, cotton, peanuts and tree nuts resulting in health risks to animals and humans (6). Most genes involved in aflatoxin biosynthesis are clustered in a 75-kb genomic DNA region that carries a positive regulator, AflR, and structural genes including omtA (17). Although much has been learned about the molecular biology and biochemistry of aflatoxin biosynthesis, little is known about the location of aflatoxin enzymes within a fungal colony or within a fungal cell. Early studies identified two enzymes, a 168 kDa protein and a 40 kDa protein (OmtA), in Aspergillus parasiticus capable of converting sterigmatocystin to O- methylsterigmatocystin (3,5) in vitro; about 60% of the enzyme activity in cell extracts was associated with the 168 kDa protein and 40% of the activity was associated with OmtA (4). Both enzymes required the co-factor S-adenosylmethionine. The 168-kDa protein was purified to near homogeneity (3) and consisted of two subunits (58 and 110 kDa). OmtA (40 kDa) was also purified to homogeneity (8,9). The gene encoding OmtA (omtA) was cloned from a cDNA expression library using OmtA-specific polyclonal antibodies raised to OmtA protein expressed in E. coli (9, 16); omtA was later shown to 36 reside in the aflatoxin gene cluster (17) supporting its proposed role in aflatoxin synthesis. Because sterigmatocystin and O-methlysterigmatocystin could be converted to aflatoxin B, in vivo (2), they were generally accepted as aflatoxin pathway intermediates. Two important issues remained unresolved. 1) Although the 168 kDa protein and OmtA converted sterigmatocystin to O-methylsterigmatocystin in vitra, it was not clear which of these activities was necessary and sufficient to catalyze this reaction in vivo (7, 10, 11, 14). 2) Because the methyl group on O-methlysterigmatocystin is removed during the subsequent oxidoreduction reaction and because structurally related molecules are incorporated into AFB, in feeding experiments with similar or higher efficiency than sterigmatocystin (7), it was not clear if this reaction was necessary for aflatoxin synthesis. To address these issues, I generated an omtA gene disruption mutant (LW1432). Feeding studies conducted on LW1432 demonstrated a critical role for OmtA and the reaction catalyzed by this enzyme in aflatoxin synthesis in viva. MATERIALS AND METHODS Fungal strains. A. parasiticus SU1(NRRL5862, ATCC 56775) is a wild—type, aflatoxin-producing strain. A. parasiticus CSIO (ver-I wh-I pyrG) was derived from A. parasiticus ATCC36537 (ver-I wh-I) (l3). LW1418 and LW1432 (ver-I wh-I omtA) were generated in this study by disrupting the omtA gene in CSIO. AF 810 is a non- aflatoxin producing aflR knockout strain derived from A. parasiticus NR1 (niaD) derived fi'om SUI (provided by Dr. Jeff Cary, USDA, New Orleans, LA.). Construction of omtA disruption vector pLW14. Plasmid pLW14 was constructed by inserting pyrG, which encodes orotidine monophosphate decarboxylase 37 (13), into the coding region of omtA at the SphI site (Figure 2.1). A plasmid carrying omtA genomic DNA was kindly provided by Dr. Fun Sun Chu (University of Wisconsin- Madison, WI). The 2.5kb pyrG fragment was generated by polymerase chain reaction (PCR) using plasmid pPG3J (12) as template. The primers used to amplify pyrG carried an SphI (GCATGC) restriction site to facilitate cloning (5'- GTAGAAGTTCAGQATGCTGATGG-3' and 5'-GAGTATCACAGTCAG_GCAT§C ACGTC-3'). The reaction conditions for thermal cycling were: 95° C for 5 min followed by 35 cycles at 95° C for 1 min, 62° C for 1 min, and 72° C for 3 min. The reaction was completed by incubation at 72° C for 10 min. Transformation and screening for omtA gene-disrupted strains. Circular or linear plasmid (8 ug) digested with HindIII was used in transformation experiments using A. parasiticus CSIO as the recipient strain (13). Protoplasts were generated by digestion of mycelium (harvested 17 h after initiation of germination) with Novozyme 234 and transformation was conducted by a PEG method as described previously (13). The selection of omtA-disrupted strains was achieved in four steps. 1) pyrG-positive clones (pyrG+; uridine prototrophs) were selected by grth on CZ agar (Czapek Dox Agar; DIFCO Laboratories, Detroit, MI), a minimal growth medium (13). 2) Spores from 3 to 4 day old transfonnant colonies grown on CZ plates were transferred by sterile toothpick to coconut agar medium (1) supplemented with sterigmatocystin at a concentration of 20 11ng and also to CZ agar medium. CSIO protoplasts regenerated and tolerated sterigmatocystin supplementation of the agar during active growth and then utilized sterigmatocystin to synthesize aflatoxin BI as active growth slowed or ceased. 3) Colonies that did not fluoresce blue under long-wavelength UV light (254 nm) after a 38 _Mzoi 668 Hindlll 689 ”‘1‘?“ 4'-'~-—..\ '1‘ "'. in“ “ :fi'.!,‘_;‘L-L ‘ J.» .4 .~: 53.73; / /m Amp ’1 Akoll778 ,/ cnnhA / i ‘ grifflephl l 8 O 7 Xbal 6208._ . 1 I... Hindlll 6149““?‘1‘7 /' I<3 $31, X%a15880 p~' 1XWOI2747 Sphl4267“ Figure 2.1. Restriction endonuclease map of plasmid pLW14. 39 3-day incubation at 29° C in the dark on coconut agar medium were selected and secondary metabolites extracted and analyzed by thin-layer chromatography by a previously published procedure (15). 4) The genotype of suspected omtA-disrupted strains was confirmed by Southern hybridization analysis. OmtA null mutants were further analyzed for ability to convert sterigmatocystin or O-methylsterigmatocystin to aflatoxin B1 by a feeding experiment, as described below. Southern hybridization analysis. Conidiospores (5 x 106) isolated from transforrnants were inoculated into 100 ml of YES and incubated with shaking at 150 rpm for 48 to 72 h at 29° C in the dark. One gram of freshly harvested mycelium was used for isolation of genomic DNA using a published procedure (13). Southern hybridization analysis was performed using a Non-Radioactive Hybridization Kit (Roche Molecular Biochemicals, Indianapolis, IN) and a procedure provided by the manufacturer. To identify the presence of pyrG sequences integrated within the chromosomal omtA gene, genomic DNA was digested with HindIII and EcoRI, separated by electrophoresis on a 1% agarose gel, transferred to Nytran membrane, and probed with digoxigenin-labeled 2.9 Kb HindIII fiagment contained in the omtA gene. Northern hybridization analysis. Conidiospores (5 x 106) isolated from LW1418, LW1432, CSIO and SUI were inoculated into 100 ml of YES and incubated with shaking at 150 rpm for 48 h at 29° C in the dark. Total RNA was purified from freshly ground mycelium used TRIzol reagent (GibcoBRL, Rockville, MD). Northern hybridization analysis was performed using a Non-Radioactive Hybridization Kit (Roche Molecular Biochemicals, Indianapolis, IN) and a procedure provided by the manufacturer. The total RNA (35 u g) was separated by electrophoresis (60 V, 4 h) on a 1% formaldehyde-agarose gel, transferred to Nytran membrane, and probed with a 40 digoxigenin-labeled 2.9 Kb HindIII fragment contained in the omtA gene. The gel was stained with ethdium bromide (in 0.5 M ammonium acetate, at a final concentration of 0.5 pg per ml) for 40 min before the transferring procedure. Feeding studies of omtA gene disruption strains. One gram of mycelia from the same culture as used in Southern hybridization analysis was inoculated into 10 ml of YES media supplemented with either sterigmatocystin or O-methylsterigrnatocystin at a concentration of 20 pg per ml and incubated at 29° C for 24 h in the dark with shaking at 150 RPM. The fungal mycelium and culture medium were extracted with 10 ml of chloroform at 4° C for 16 h. The chloroform fraction was evaporated under N2 gas and resuspended in 200 pl of acetone. Ten pl samples were analyzed by TLC using ether/methanol/water (96:32]) as the solvent system. RESULTS Generation of omtA disruption mutants. TLC analysis of cell extracts from pyrG+ transforrnants tentatively identified 5 omtA gene disruption isolates. Two isolates (LW1418 and LW1432) were identified from among 41 pyrG+ colonies transformed with circular plasmid LW14. Southern hybridization analysis of genomic DNAs isolated from LW1418 and LW1432 detected a 2.5 kb pyrG DNA fragment within the omtA gene (Figure 2.2.). Northern hybridization analysis of total RNAs isolated from LW1418 and LW1432 did not detect any bands but a 1.6 kb band was detected in CS 10 and SUI, confirming the gene disruption event in these isolates (Figure 2.3.). 41 Figure 2.2. Southern hybridization analysis of omtA gene disruption strains. A. Restriction endonuclease map of omtA locus in omtA disruption strains LW1418 and LW1432. B. Southern hybridization analysis. Genomic DNAs isolated from LW1418 and LW1432, CSIO (parent strain) and SUI (wild-type) were digested with restriction enzymes HindIII and EcoRI. DNAs were resolved and hybridized to a digoxigenin- labeled 2.9 kb HindIII omtA DNA fragment using standard methods. Strains with omtA- disruption were predicted to contain a 5.4 kb HindIII fragment consisting of the 2.5 kb pyrG selectable marker inserted into the omtA locus. A 5.2 kb fragment in the omtA- disrupted strains was also predicted to replace the wild-type EcoRI fragment (2.7 kb). Numbers to the right of the blot represent the approximate size of the detected fragments in Kilobase pairs. 42 omtA E 1H 1; IS H E E: EcoRI, a: Hindu], s: Sphl l ‘—— ] ”'6 Probe: 2.9 Kb HindIII fragment Digestion: Hind III EcoRl Wild type: 2.9 Kb 8 Kb + 2.7 Kb Mutant: 5.4 Kb 8 Kb + 5.2 Kb mum Ecol“ F I l 1 1°. a 9.". 2 ~ 2 1!. 1'. '7 9. E. 2'. #9: 8 5 a. a 8 g 3 a, __ 5.4 Kb Figure 2.2. 43 Figure 2.3. Northern blot analysis of omtA gene disruption strains. (A) omtA transcript accumulation assessed by Northern hybridization analysis. Total RNAs isolated from LW1418 and LW1432, CSIO (parent strain) and SU-l (wild-type) were resolved and hybridized to a digoxigenin-labeled 2.9 kb HindIII omtA DNA fragment using standard methods. Strains with omtA-disruption were predicted to contain no omtA transcript and a 1.6 kb transcript was predicted to be detected in both CSIO and SU-l. Equal loading of RNA is demonstrated by Ethidium bromide staining of RNA as shown in panel (B). Numbers to the left of the gel represent the approximate size of standard molecular fragments in Kilobase. 44 9.49 l 7.46 4.40 2.37 1.35 0.24 Figure 2.3. Northern Blot Analysis no N E E a a E E a _- .. - <— OmtA EtBr Stain ‘— 1.6 kb 45 In vivo feeding experiments. No O-methylsterigmatocystin or aflatoxin B1 could be detected in LW1418 or LW1432 supplied with (fed) exogenous sterigmatocystin; in contrast, each isolate converted exogenously supplied O- methylsterigrnatocystin to aflatoxin B1 (Figure 2.43). The parental strain CS10 converted either sterigmatocystin or O-methylsterigmatocystin to aflatoxin B1 (Figure 2.4.A). These data suggest that LW 1418 and LW1432 carry mutant copies of omtA, but the genes encoding later pathway enzymes and the regulatory genes involved in aflatoxin biosynthesis are still functional. DISCUSSION The initial goal of this study was to determine if OmtA is necessary and sufficient to convert sterigmatocystin to O-methylsterigmatocystin in vivo and if this reaction is necessary for aflatoxin synthesis. When the same concentration of either sterigmatocystin or O-methylsterigmatocystin was fed to the parent strain CSIO (wild type OmtA and OrdA activities) nearly equal concentrations of O-methylsterigmatocystin and aflatoxin B, were generated. This observation could result for one of at least three potential reasons. I) The intermediates were fed in excess of the capacity of the pathway enzyme to carry out complete conversion in the 24 hour period allotted. 2) The entry of the intermediates into the cell or into the enzyme active sites was rate-limiting. 3) The intermediates stimulate a feedback repression system downregulating activities of one or both enzymes. Nevertheless, it was clear that CSIO could convert either sterigmatocystin or O- methylsterigmatocystin to aflatoxin B, while LW1432 and LW1418 could convert 46 Figure 2.4. Thin layer chromatography of extracts of CS 1 0 and omtA-disruption strains LW1418 and LW1432. (A) Thin layer chromatography of extracts of CS10 supplied with (fed) exogenous sterigmatocystin (ST) or O-methylsterigmatocystin (OMST). Lane 1 and 7, AF 81, AF 82, AF G1 and AFGz standard mixture. Lane 2 and 6, OMST standard. Lane 3, CSIO fed no pathway intermediate. Lane 4, CSIO fed sterigmatocystin. Lane 5, CSIO fed 0- methylsterigmatocystin. Ten pl samples were analyzed using chloroform/acetone (95: 5) as the solvent system. (B) Thin layer chromatography of extracts of omtA -disrupted strains LW1418 and LW1432 supplied with (fed) exogenous sterigmatocystin (ST) or 0- methylsterigmatocystin (OMST). Metabolic scheme for aflatoxin biosynthesis in strain LW1418 and LW1432 is shown at the top of the thin layer chromatograph. The ver-I and omtA genes involved in aflatoxin biosynthesis are non-functional in LW1418 and 1432. Lanes 1 and 8, OMST standard. Lanes 2 and 7, AF B1, AF Bz, AF G1 and AFG2 standard mixture. Lane 3, LW1418 fed OMST. Lane 4, LW1418 fed ST. Lane 5, LW1432 fed OMST. Lane 6, LW1432 fed ST. Abbreviations: VA, versicolorin A; DMST, demthylsterigmatocystin. Ten pl samples were analyzed using ether/methanol/water (96: 3:1) as the solvent system. 47 OMST AFB, AFG, B ver-I dmrA omtA ordA VA ’X" omsri>sr WX+ OMST ans, 1 2 3 4 5 6 7 8 OMST AFB, AFG, Figure 2.4. 48 O-methylsterigmatocystin to aflatoxin B, but were unable to convert sterigmatocystin to either O-methylsterigmatocystin or aflatoxin B,. These data clearly demonstrate that OmtA is necessary for efficient conversion of sterigmatocystin to O- methylsterigmatocystin in viva in A. parasiticus CSIO under the specified growth and assay conditions. The data also demonstrate that this reaction is necessary for aflatoxin biosynthesis. It is not yet clear if OmtA alone is sufficient to carry out this reaction in viva. We cannot rule out the possibility that the 168 kDa protein contributes to this reaction, but its contribution appears negligible compared to OmtA. In support of this idea, LW1432 (omtA disruption) could not convert exogenously supplied sterigmatocystin to detectable levels of either O-methylsterigmatocystin or aflatoxin B, but could convert O-methylsterigmatocystin to aflatoxin B1 in viva. In contrast, CSIO (parent of LW1432) converted either sterigmatocystin or O-methylsterigmatocystin to aflatoxin B,. These data demonstrate that the omtA disruption is gene-specific and does not affect the activity of late pathway enzymes or regulators of aflatoxin synthesis. 49 REFERENCES l. Arseculeratne, S. N., L. M. De Silva, S. Wijesundera, and C. H. S. R. Bandunatha. 1969. Coconut as a medium for the experimental production of aflatoxin. Appl. Microbiol. 18:88-94. 2. Bhatnagar, D., S. P. McCormick, L. S. Lee , R. A. Hill. 1987. Identification of O- methylsterigmatocystin as an aflatoxin B1 and GI precursor in Aspergillus parasiticus. Appl. Environ. Microbiol. 53:1028-33. 3. Bhatnagar, D., A. H. Ullah, and T. E. Cleveland. 1988. Purification and characterization of a methyltransferase from Aspergillus parasiticus SRRC 163 involved in aflatoxin biosynthetic pathway. Prep. Biochem. 18:321-49. 4. Bhatnagar, D., T. E. Cleveland, and E. B. Lillehoj. 1989. Enzymes in aflatoxin B1 biosnythesis: Strategies for identifying pertinent genes. Mycopathologia. 107:75-83. 5. Cleveland, T. E., A. R. Lax, L. S. Lee, and D. Bhatnagar. 1987. Appearance of enzyme activities catalyzing conversion of sterigmatocystin to aflatoxin BI in late- growth-phase Aspergillus parasiticus cultures. Appl. Environ. Microbiol. 53:1711-3. 6. Eaton, D. L. and J. D. Groopman (eds.). 1994. The toxicology of aflatoxins: human health, veterinary, and agricultural significance. San Diego, CA: Academic. 7. Gengan R. M., A. A. Chuturgoon, D. A. Mulholland and M. F. Dutton. 1999. Synthesis of sterigmatocystin derivatives and their biotransformation to aflatoxins by a blocked mutant of Aspergillus parasiticus. Mycopathologia 144:115-122. 8. Keller, N. P., H. C. Dischinger, JR., D. Bhatnagar, T. E. Cleveland and A. H. J. Ullah. 1993. Purification of a 40-kDa methyltransferase activity in the aflatoxin biosynthetic pathway. Appl. Environ. Microbiol. 59:479-484. 9. Liu, B. H., N. P. Keller, D. Bhatnagar, T. E. Cleveland and F. S. Chu. 1993. Production and characterization of antibodies against sterigmatocystin O- methyltransferase. Food Agric. Immunol. 5:155-164. 10. Mito, R. E. and C. A. Townsend. 1997. Enzymology and molecular biology of aflatoxin biosynthesis. Chem. Rev. 97:2537-2555. 11. Payne, G. A. and M. P. Brown. 1998. Genetics and physiology of aflatoxin biosynthesis. Annu. Rev. Phytopathol. 36:329-362. 50 12. Reichard, U., G. T. Cole, T. W. Hill, R. Ruchel and M. Monod. 2000. Molecular characterization and influence on fungal development of ALP2, a novel serine proteinase from Aspergillusfumigatus. Int. J. Med. Microbiol. 290:549—558. 13. Skory, C. D., J. S. Horng, J. J. Pestka, and J. E. Linz. 1990. Transformation of Aspergillus parasiticus with the homologous gene (pyrG) involved in pyrimidine biosynthesis. Appl. Environ. Microbiol. 56:3315-3320. 14. Sweeney, M. J. and A. D. W. Dobson. 1999. Molecular biology of mycotoxin biosynthesis. FEMS Micobiology Letters. 175:149-163. 15. Trail, F., P-K. Chang, J. Cary 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. 16. Yu, J., J. W. Cary, D. Bhatnagar, T. E. Cleveland, N. P. Keller, F. S. Chu. 1993. Cloning and characterization of a cDNA from Aspergillus parasiticus encoding an O- methyltransferase involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 59:3564- 71. 17. Yu, J., P-K. Chang, J. W. Cary, M. Wright, D. Bhatnagar, T. E. Cleveland, G. A. Payne, and J. E. Linz. 1995. Comparative mapping of aflatoxin pathway gene clusters in Aspergillus parasiticus and Aspergillusflavus. Appl. Environ. Microbiol. 61 :2365-2371. 51 CHAPTER 3 Expression/Distribution of the OmtA Protein in Colonies of Aspergillus parasiticus INTRODUCTION Previous studies on accumulation of aflatoxin enzymes and expression of aflatoxin genes including omtA were conducted primarily in liquid medium using submerged shake culture (batch fermentation) (6,22,25); these conditions induce aflatoxin synthesis but do not support asexual sporulation (conidiation). Because a close regulatory association has been demonstrated between aflatoxin synthesis and conidiation (5,7,9) we hypothesized a spatial and temporal association between omtA expression and conidiospore development. Our goal was to develop a grth model that would closely mimic regulation of toxin synthesis in soil and on the host plant. We developed a novel time-dependent colony fractionation protocol to study OmtA accumulation in fungal colonies grown on solid medium; these conditions support toxin synthesis and conidiation. This protocol also allowed analysis of OmtA distribution to different cell types in fungal colonies. OmtA-specific polyclonal antibodies were generated against OmtA fusion protein (MBP-OmtA) and purified by affinity chromatography using a LW1432 protein extract. OmtA was not detected in 24 h old colonies but was detected in 48 h old colonies using Western blot analysis; the protein accumulated in all regions of a 72 h old colony including cells (0 to 24 h old near the colony margin) in which little conidiophore development was observed. OmtA in older parts of the colony (24 to 72 h) was partly degraded. F luorescence-based immuno-histochemical analysis conducted on 52 thin sections of paraffin-embedded fungal cells from time-fractionated fungal colonies demonstrated that OmtA is evenly distributed among different cell types and is not concentrated in conidiophores. These data suggest that OmtA accumulates in newly formed fungal tissue and then is proteolytically cleaved as cells in that section of the colony age. The data also suggest that OmtA is localized to specific areas within a fungal cell but it is not yet clear if these areas correspond to 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. MATERIALS AND METHODS Fungal strains. A. parasiticus SU1(NRRL5862, ATCC 56775) is a wild-type, aflatoxin-producing strain. A. parasiticus CSIO (var-1 wh-I pyrG) was derived from A. parasiticus ATCC36537 (ver-I wh-I) (21). LW1418 and LW1432 (ver-I wh-I omtA) were generated in this study by disrupting the omtA gene in CSIO. AF $10 is a non- aflatoxin producing aflR knockout strain derived from A. parasiticus NR-I (niaD) which in turn was derived from SUI. Construction of OmtA expression vector pLW12. An omtA cDNA was generated by reverse transcriptase- PCR (RT-PCR). Template RNA was isolated from A. parasiticus strain SUI cultured in YES media for 48 to 72 h using Trizol reagent and a procedure supplied by the manufacturer (GibcoBRL, Rockville, MD). For first strand cDNA synthesis, 48 pg of total RNA were incubated in the RT-PCR mix at 37° C for 2 h. All chemicals used in the RT-PCR were purchased from GibcoBRL (Rockville, MD). 53 The 20 pl reaction mixture contained: 4 pl of 5 X first strand buffer, 2 pl of 0.1 M DTT, 1 pl of 10 mM dNTP, 2 pl of M-MLV Reverse Transcriptase (200 U per pl), and 1 pl of Oligo (dT) primer (0.5 pg per pl). One primer for omtA amplification contained a HindIII (AAGCTT) restriction site and the other primer an XbaI (TCTAGA) restriction site to facilitate cloning (5'-CCCTCTAGAATG GCACTACCGAGCAAAG-3' and 5'- TGCAAQCTTCTACTTGCGCAAACGCAGT-3'). One pl of the resulting cDNA mixture was used as template for PCR. The cycling conditions were: 95° C for 5 min followed by 35 cycles of 95° C for 1 min; 55° C for 1 min, and 72° C for 2 min. The mixture was incubated at 72° C for 10 min to complete the reaction. The omtA PCR fragment (1260 bp) was digested with restriction enzymes XbaI and HindIII and cloned into plasmid pMAL-c2 (New England Biolabs, Beverly, MA) digested with the same enzymes. The resulting plasmid construct, pLW12, was transformed into E. cali DHSa. The proper construction of pLW12 in clones expressing MBP-OmtA was confirmed by restriction enzyme analysis of purified plasmid DNA isolated by the Qiagen mini prep plasmid kit (Qiagen, Valencia, CA). The size of the fusion protein was determined by small-scale expression studies. E cali DHS or carrying pLW12 was incubated in 5 m1 of LB broth containing Ampicillin (100 pg per ml) for 16 h. One ml of bacterial culture was saved as non-induced control. The remaining 4 ml of culture was induced to express fusion protein by addition of 0.3 mM of IPTG for 3 h. Conversion of sterigmatocystin to O-methylsterigmatocystin by maltose binding protein -OmtA. The pMAL Protein Fusion and Purification System from New England Biolabs (Beverly, MA) was utilized to express and purify Maltose Binding Protein-OmtA. Large-scale preparation of Maltose Binding Protein-OmtA was conducted 54 for the study of enzymatic activity and antigen purification. E cali DH5 or carrying pLW12 was grown in 500 ml of rich media (10 g tryptone, 5 g yeast extract, 5 g NaCl, 2% glucose) containing ampicillin (100 pg per ml). Fusion protein synthesis was induced by addition of IPTG (0.3 mM), cells were sonicated (Sonifier cell disrupter W-350; Fisher Scientific, Pittsburgh, PA) and the fusion protein was purified by amylose affinity column chromatography according to a protocol supplied by the manufacturer (New England Biolabs, Beverly, MA). Amylose resin-purified Maltose Binding Protein and Maltose Binding Protein - OmtA were used for analysis of enzyme activity. In a time-course experiment, three l-ml reactions were prepared. Each reaction mixture contained 100 pg of Maltose Binding Protein-OmtA, 20 pg of sterigmatocystin, and 600 pg of S-adenosylmethionine. Sterigmatocystin, O-methylsterigmatocystin and S-adenosylmethionine were purchased from Sigma (St. Louis, MO). The reaction mixtures were incubated at room temperature for appropriate times (20-min time points for 60 min; one tube per time point). Three control reactions were also incubated for 60 min. Each contained the same reagents as above except the Maltose Binding Protein-OmtA or sterigmatocystin or S- adenosylmethionine were omitted. Reactions were stopped by extraction with 4 volumes of chloroform. Chloroforrn extracts were dried under nitrogen gas and resuspended in 100 pl of acetone. Five pl of extract was analyzed by TLC using chloroform lacetone (95:5) as a development system (23). OmtA antibody production and purification. OmtA polyclonal antibodies were generated in rabbits using purified MBP-OmtA as antigen. Each of 2 rabbits (New Zealand White) was injected subcutaneously with 300 pg of protein in TitreMax adj uvant 55 (Cthx, Norcross, GA) at 1: 1 (V/V) ratio. After 35 days, the animals received one booster injection with 200 pg of protein in TitreMax adjuvant. Serum was obtained 4 weeks after the boost. The IgG fraction was purified from rabbit serum by precipitation with ammonium sulfate using standard procedures (2). The IgG fraction of OmtA antibodies was firrther purified by affinity chromatography using a protein extract prepared from LW1432 grown in YES media for 48 h. This column was prepared by conjugating 11 mg of total protein to 2 ml of Aminolink coupling gel (coupling efficiency approximately 80%) using a procedure supplied by the manufacturer (Pierce Chemical Company, Rockford, IL). For antibody purification, one ml of IgG (8 mg) was loaded onto the gel bed in this column and an additional 2 ml of PBS was loaded to cover the gel bed. The column with antibodies was incubated at room temperature for 1.5 h at room temperature. After incubation, PBS was allowed to flow-through and antibody-containing fractions were identified by absorbance at 280nm. The protein concentration was determined by a BIO-RAD protein assay (Hercules, CA) using Dye Reagent and BSA (Fraction V, Sigma, St. Louis, MO) as standard. This highly purified antibody preparation (3 mg per ml) was used in Western blot analysis and immuno-fluorescence microscopy. Western blot analysis. To determine specificity of anti-OmtA antibodies, fungal proteins were extracted from mycelia for Western blot analysis. Fungal strains including SUI, CSIO, LW1418, LW1432, LW1468 and LW1470 were grown in the dark at 29°C (shaking at 150RPM) in 100 ml of YES liquid media for 24, 48 and 72 h or only 48 h. The mycelia were harvested, pulverized under liquid nitrogen and the fungal proteins were extracted in TSA buffer (2 mM Tris-Cl, pH8.0; 40 mM NaCl and 0.025 % sodium azide) containing complete protease inhibitors (Roche Molecular Biochemicals, 56 Indianapolis, IN). Proteins were resolved by SDS polyacrylamide gel electrophoresis using standard methods (2). For Western blot analysis, each lane on the 12% SDS-PAGE contained 30pg protein. The primary Ab was column-purified anti-OmtA IgG (1 pg per ml) and the secondary Ab consisted of a 10,000-fold dilution of goat anti-rabbit IgG alkaline phosphate conjugate (Schleicher & Schuell, Keene, NH). A BCIP/NBT colorimetric detection system was utilized (Roche Molecular Biochemicals, Indianapolis, IN). The antibody against native OmtA was kindly provided by Dr. Fun Sun Chu (University of Wisconsin-Madison, WI). This antibody was also subtraction-purified with a LW1432 strain (omtA gene disrupted) protein extract before use in Western blot analysis (Figure 3.2.). 1.4 pl of antibody preparation was incubated with 100 pl (800 pg) of LW1432 protein extract for I h at 4° C. After incubation, the preparation was centrifuged at 12,000 x g for 15 min at 4° C to pellet cross-reactive antibodies. The supernatant was mixed with 14 ml of blocking reagent and used in Western blot analysis. Time-dependent fractionation of colonies grown on solid medium. To determine the accumulation and distribution of OmtA in fungal colonies grown on solid culture media, conidiospores (2 x 105) ofA. parasiticus SUI, AFSIO, c310, and LW1432 were inoculated onto the center of YES agar or PDA agar (for SUI only) overlaid with sterile cellophane membranes and incubated at 29° C in the dark. Some colonies were analyzed after 24 or 48 h of growth. 72 h-old colonies were fractionated into three concentric rings based on area covered at three time points (72, 48 and 24 h). For example (Fig. 5D), a SUI colony with a diameter of 4.2 cm was fractionated to SI which contained mycelia from the colony center out to a distance of 0.8 cm (ages 48 to 72 57 h), 82 which contained mycelia from 0.8 to 2.5 cm (24 to 48 h) and S3 which contained mycelia from 2.5 to 4.2 cm (0 to 24 h). The harvested mycelia from appropriate sections of the colony were pulverized under liquid nitrogen and the fungal proteins were extracted in TSA buffer containing complete protease inhibitors. Western blot analysis was performed as described above. Immuno-localization of OmtA protein. Immuno-localization of OmtA protein was conducted on SUI colony fractions. In addition, AF S10, CSIO, and LW1432 colonies were fractionated following the same scheme to generate analogous fractions R1, R2, R3, C1, C2, C3, and L1, L2, L3, respectively. Preparation of paraffin-embedded fungal sections. Samples from fungal colony fractions were embedded in paraplast (Sigma, St. Louis, MO) using a published procedure (2) with the following modifications. Fungal tissues were fixed with Streck tissue fixative (Streck Laboratory Inc., Omaha, NE) at 4° C overnight, 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 min, 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 paraffin/xylene mixture (1:1; v/v) for 15 min, 60 min and overnight at 600C and finally for 8 h in 100% paraffin at 60° C (three changes of paraffin during incubation). The paraffin embedded sample blocks were hardened in a plastic mold (V WR Scientific, Detroit, MI) at RT. Sample blocks were cut into 4 pm thick sections 58 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-labeling. Coverslips with paraffin-embedded fungal 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 immuno-labeled with primary antibody (purified anti-OmtA IgG; 20 pg/ml) or anti-SKL (1:500, Zymed, So. San Francisco, CA) at room temperature for 1.5 h, followed by secondary antibody conjugated with fluorescent probe (goat anti-rabbit IgG-Alexa 488 conjugate [5 pg/ml]; Molecular Probes; Eugene, OR), at room temperature 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 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 with 488 nm laser line. The 40 X oil objective lens (Zeiss Plan-NeoFlura, NA: 1.3) was used to acquire all images. The 59 Alexa 488 fluorescence probe (Abs 495 nm/ Em 519 nm) was detected using LP 520 or BP 520-560 barrier filters. Fluorescence image analysis of SUI, AF S10, CSIO and LW1432 was conducted under the same instrument parameter settings. RESULTS Western blot analysis and protein localization using PAb against native OmtA. The antibodies against purified native OmtA were kindly provided by Dr. Fun Sun Chu (University of Wisconsin-Madison, WI). The specificity of this PAb was tested on SUI (wild type) cultured in YES liquid medium for 24, 48 or 72 h. One primary band (larger than 45 kDa) was detected in Western blot analysis of protein extracts prepared from 24 h culture and three primary bands were detected in 48 or 72 h cultures (Figure 3.1.A). Protein localization using a slide-culture method (12) showed the primary signal was detected in conidiophores and conidiospores (Figure 3.1 .B). This antibody was further purified by subtraction with a protein extract from LW1432 (omtA gene disruption strain) and the specificity was tested by Western blot analysis. Two primary bands were detected in SUI and CSIO and one intense band was detected in all omtA knock-out strains, suggesting that this purified antibody still cross-reacted with one protein (Figure 3.2.). Purification of Maltose Binding Protein-OmtA. Maltose Binding Protein- OmtA in DH50L cell extracts (Figure 3.3.A) was purified by amylose column chromatography. Most of the purified Maltose Binding Protein-OmtA fusion protein was approximately 88 kDa in mass (Figure 3.3.8); this is consistent with a mass calculated 60 from the fusion of 42 kDa MBP and 45 kDa OmtA predicted using nucleotide sequence data. Very little proteolytic cleavage of the protein was detected by this analysis. The fusion protein remained soluble in the bacterial cytoplasm allowing purification of 39 mg of MBP-OmtA per liter of bacterial culture. Enzymatic conversion of sterigmatocystin to O-methylsterigmatocystin by Maltose Binding Protein-OmtA. Purified OmtA fusion protein efficiently converted sterigmatocystin to O- methylsterigrnatocystin in the presence of S-adenosylrnethionine (Figure 3.3.C) within 60 min. The cofactor, S-adenosylmethionine, was required for this conversion. Without exogenous S-adenosylmethionine, no O-methylsterigmatocystin could be detected in the presence of the Maltose Binding Protein-OmtA. As expected, no O-methylsterigmatocystin was detected in the reaction that contained S- adenosylrnethionine without substrate (sterigmatocystin), in a reaction without added Maltose Binding Protein-OmtA, or in a reaction with Maltose Binding Protein but without the Maltose Binding Protein-OmtA (data not shown). Production of PAb against OmtA fusion protein. To obtain highly specific anti-OmtA polyclonal antibodies (PAb), amylose-purified MBP-OmtA was used as antigen to produce polyclonal antibodies in two rabbits. The IgG fraction of the rabbit serum was further purified by affinity chromatography using a protein extract from LW1432 (omtA gene disruption strain). The specificity of anti-OmtA PAb was tested on SUI (wild type) cultured in YES liquid medium for 24, 48 or 72 h. No OmtA could be detected in Western blot analysis of protein extracts prepared from SUI grown on YES liquid medium for 24 h while one primary band (approximately 45 kDa) could be 61 Figure 3.1. Western blot analysis and fluorescence microscopy using polyclonal antibodies raised against partially purified fungal OmtA proteins. The native OmtA antibody was kindly provided by Dr. Fun Sun Chu (University of Wisconsin-Madison, W1). (A) Analysis of SUI crude protein extracts isolated at three time points (24, 48 and 72 h) from fungal cultures grown in YES liquid medium. The approximate mass of the primary signal (45 kDa) is shown to the right of the blots. (B) OmtA protein localization in a colony of A. parasiticus SUI grown on YES agar for 48 h. Fungal tissue from slide-culture was digested with Novezyme and probed with OmtA PAbs (20 pg/ml) followed by FITC-conjugated goat anti-rabbit IgG (1:50). The protein localization method used was reported previously (12). 62 Western Blot Analysis: anti-OmtA 24H 48H 72H lfi F—I l_‘l MW. 45 _ ————-—-—-...,...‘-m--—_- --w- 31 _ — -— —- -— B. Bright Field OmtAFITC IYNK 1111111111,, , willfl ‘11“ :3,“ ““3,“ \11 ll . ‘ 110““ “m“? 1 mill lsllllll Figure 3.1. 63 ab 6'! co N no a v-1 I") F! ("I ‘9 l‘ g c r. z z z z z a t— "‘ 5 3 3 3 B 3 3 m a o .4 ...i ...r .1 .1 .2 Em" - U 66kDa OmtA» 19191:; - g __' ” """'"" """"‘" .2-.. ““'“"' _ 45 kDa . J 31 kDa Figure 3.2. Western blot analysis of firngal protein extracts using subtraction-purified PAb raised against native OmtA. Crude protein extracts were isolated from wild-type aflatoxin-producing strain SUI, parent strain CSIO, and omtA knock-out strains, LW1418-1, LW1432-1, LW1418, LW1432, LW1468 and LW1470, grown in YES liquid medium for 48 h. LW1418-l and LW1432-1 are single spore isolates of LW141 8 and LW1432, respectively. The native OmtA antibodies used were subtraction-purified with LW1418 protein extract two times. The possible OmtA reactive band (45 kDa) is indicated by an arrow at the left of the blots. Lane STD contains molecular mass standards; mass of standards is shown at the right of the blot. 64 Figure 3.3. Conversion of ST to OMST by affinity purified MBP-OmtA fusion protein. (A) SDS-PAGE of IPTG- induced bacterial crude extracts containing MBP-OmtA or MBP. (B) SDS-PAGE of affinity purified MBP-OmtA. Lanes MWI and MW2 represent molecular mass standards; molecular mass is indicated to the left and right of the gel. (C) Thin layer chromatography to catalyze the conversion of ST to OMST by affinity purified MBP-OmtA and controls. Five pl of reaction mixture was analyzed after appropriate time periods by TLC using chloroform /acetone (95:5) as the developing system. Control reactions included all reagents except ST, SAM, or MBP-OmtA, respectively. OMST standards are located in lanes on both ends of the TLC (STD: OMST). 65 E" I” Z a: 'F o z i 96 MWI .3’ - .-.. kDa Purified MBP-OmtA 1"“ kDa MWI l—fi MWI MW2 RD: 201 97 116 - - - - - a so 66 - " - 31 - - g 32 -_ ..._ ,__. .fi 3 1% i C a? 3 ' S a 3 + 3 + w MBP-OmtA % + 3.” m g +ST+SAM z ‘11 2 g 3 20’ 4o: 60’ 60’ 60’ 60’ E U) m -I -1 Figure 3.3. 66 detected in SUI grown for 48 or 72 h in the same medium (Figure 3.4.A). Western blot analysis of OmtA in colonies grown on solid medium. The accumulation and distribution of OmtA was analyzed in time-fractionated colonies grown on solid YES or PDA agar medium for 72 h (Figure 3.4.B, C and D). OmtA protein was detected in fractions SI, S2, and S3 of a 72 h old SUI colony but not in corresponding colony fractions isolated from AF S10 (aflR knockout) (Figure 3.4.D) or LW1432 (omtA knockout) (Figure 3.4.C). OmtA in colony fractions S1 (48 to 72 h) and 82 (24 to 48 h) showed increased levels of smaller peptides and correspondingly less full-length protein. In contrast, in fraction S3 (0 to 24 h) of a 72 h old colony, full length OmtA was found at higher levels. A similar result was observed using SUI grown on PDA medium and CSIO (parent strain of LW1432) grown on YES agar medium; very little OmtA protein could be detected in fraction SI and C1 (48 to 72 h) while more full-length protein was detected in the youngest colony fraction (S3 and C3; 0 to 24 h) (Figure 3.4.C). No OmtA protein could be detected in any colony fraction of either LW1432 or AF SlO. Confocal Laser Scanning Microscopy. Immuno-fluorescence microscopy was conducted on samples prepared from fractionated colonies grown on YES solid media. To compare the fluorescence labeling intensity between SUI and CSIO and their non- aflatoxin producing counterparts, AF S10 and LW1432, respectively, the samples were viewed under the lowest zoom level during the Confocal Laser Scanning Microscopy (zoom = 20) in order to acquire maximum cell number within a field. Under the same 67 Figure 3.4. Western blot analysis of fungal protein extracts using affinity purified OmtA PAb. (A) Analysis of SUI crude protein extracts isolated at three time points (24, 48 and 72 h) from fungal cultures grown in YES liquid medium. The approximate mass of the primary signal (45 kDa) is shown to the right of the blots in panels A and B, and to the left of blots in panels C and D. (B) Analysis of protein extracts from time—fractionated colonies of SUI grown on PDA agar medium. 72 h-old colonies were fractionated into three concentric rings based on area covered at three time points. SI, 48 to 72 h; S2, 24 to 48 h; S3, 0 to 24 h. (C) Analysis of protein extracts from time-fractionated colonies of CSIO and LW1432 grown on YES agar medium. (D) Analysis of protein extracts from time- fractionated colonies of SUI and AF SIO grown on YES agar medium. Lane STD contains molecular mass standards; mass of standards is shown at the right of panel C. 68 A Medium Liquid YES Strain SUI 24h 48h 72h -—---—1-45kDa YES Agar CSIO LW1432 0.8 1.6 2.7 0.8 2.1 3.5 C1 C2 C3 STD L1 L2 L3 kDa 250 I48 60 ! 42 30 22 Figure 3.4. 69 Medium Strain PDA Agar SUI Diameter (cm) 1.0 2.2 3.3 Fractions 818283 ‘-- + 45 kDa YES Agar SUI AFSIO l j l N 0.8 2.5 4.2 0.8 2.5 4.0 SI 82 S3 R1 R2 R3 45 kDa -> Mb contrast and brightness settings, the samples prepared from the control strain AF SIO (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 SUI (fractions S1, 82, and S3) or CSIO (fractions C1, C2 and C3) (Figure 3.5.). These data are consistent with Western blot analysis of protein extracts isolated from the same colony fractions (Figure 3.4.C and D; i.e., OmtA was only detected in SUI and CSIO but not in AF S10 or LW1432). OmtA was observed in the substrate level mycelium throughout a 72 h old colony grown on YES media; the protein was also detected in the conidiospore-bearing structures (particularly in vesicles; Figure 3.6.B) located in fractions S1 and S2 at nearly equal intensity as in substrate level mycelium. The fluorescent signal was confined to discreet areas (patches) within cells (Figure 3.6.A). However, due to limitations in resolution of confocal laser scanning microscopy, 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 (Figure 3.6.C) (10,23). The observed fluorescent pattern was consistent with localization to these organelles. In addition, we used SYTOX Green to stain nuclei in fungal fractions (Figure 3.6.D); again the fluorescent pattern was consistent with the expected results. Neither anti-SKL nor SYTOX generated the same “patchy” pattern as anti-OmtA. In summary, the data suggest that OmtA protein 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. 70 Figure 3.5. OmtA protein localization in time-fractionated colonies of A. parasiticus SUI, AF S10 (AfR knockout), CSIO and LW1432 (omtA knockout) grown on YES agar for 72 h. Paraffin-embedded fungal sections were immuno-labeled with affinity purified OmtA PAb (20 pg/ml) followed by Alexa 488 conjugated goat anti-rabbit IgG. (A) Fluorescence images of SU-I (SI, S2, and S3) and AF 510 (R1, R2, and R3). (B) Fluorescence images of CSIO (C1, C2 and C3) and LW1432 (L1, L2 and L3). All colony fractions were analyzed under the same instrument settings. Bar = 50 pm. 71 Figure 3.5. 72 Figure 3.6. Protein localization using OmtA PAb and anti-SKL, and detection of nuclei using SYTOX Green. (A) Z-series overlay image of SUI (fraction S2) 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 81) immunolabled with OmtA PAb. (C) Immuno- labeling of SUI (fraction 82) with anti-SKL. The magnification for the left panel is 400X and for right panel is 2,000 X. (D) Fluorescence image of nuclei in SUI (fraction S2) 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. 73 DISCUSSION Because a close regulatory association between aflatoxin synthesis and conidiation has been demonstrated in several studies (7,9), we hypothesized a close temporal and spatial association between OmtA expression and conidiospore development. This study was designed to address this hypothesis. First, it was necessary to develop highly specific OmtA antibodies and a method for analysis of protein accumulation and distribution in cells and fungal colonies grown on solid growth medium. We initially experienced specificity problems with PAb (15) raised to native OmtA that resulted in severe cross-reactivity in Western blot analysis and artifacts in protein localization (Figure 3.1 .). Because aflatoxin enzymes are present at low concentration in the fungus (I I), purification of OmtA from the fungus likely resulted in co-purification of at least trace amounts of other proteins that, although undetectable by SDS-PAGE, still possessed strong antigenicity. To increase specificity, we generated PAb against affinity-purified MBP-OmtA and further purified them by affinity chromatography with a column carrying fungal proteins isolated from LW1432 (omtA knockout). This procedure helped eliminate cross- reactive antibodies. Specificity was demonstrated via Western blot analysis. OmtA could not be detected in either AF $10 or LW1432, strains that do not synthesize this protein. OmtA PAbs did not detect OmtA after 24 h in YES liquid or solid media but did detect a protein of appropriate size (45kDa) at 48 and 72 h. This pattern of accumulation is similar to that observed for other aflatoxin proteins including Nor-1 (26), Ver-l (13,14) 74 and VBS (unpublished data) suggesting that accumulation of these proteins is coordinately regulated. In addition, analysis of fungal extracts from SU-l, CSIO, AF 810 and LW1432 did not show significant cross-reaction of OmtA PAb to other cellular proteins including DmtA; sequence identity between OmtA and DmtA was reported previously (18). Time-dependent colony fractionation provided a practical method for monitoring protein accumulation and distribution in firngal tissues of different ages; similar information could not be obtained in liquid shake culture (batch fermentation). Knowledge of OmtA distribution was essential to identify fungal tissues that were rich in the target protein and allowed successful immuno-histochemistry in this study and immuno-electron microscopy in a study now underway. Since these protocols require very small quantities of fungal tissue (1mm3 for immunoelectron microscopy for example) one could easily miss the relevant protein if it was not uniformly distributed in a colony. Using this colony fractionation protocol, OmtA PAb, and Western blot analysis, we observed OmtA in all fractions of a 72 h old colony (fractions S1, S2, and S3) grown on YES agar. The youngest colony fraction (S3; 0 to 24 h) contained mostly full-length OmtA, while fraction S1 (48 to 72 h) contained less full-length OmtA and more OmtA- derived peptides. In fraction C1 (CSIO, 48 to 72 h) and the oldest fraction of SU-I (48 to 72 h) grown on PDA medium, OmtA was almost undetectable. Similarly, OmtA was not observed in the oldest fraction of a 90 h old colony grown on YES medium. No OmtA protein was detected in 24 h old colonies grown on YES; OmtA was detected in 48 h old colonies on the same medium. Together the data suggest that OmtA biosynthesis in A. parasiticus SUI grown on solid YES media initiates after 24 h. OmtA then accumulates 75 to relatively high levels in the youngest fraction, for example, in C3 and S3 (0 to 24 h) of a 72 h old colony. As cells in that fraction age (at 48 or 72 h), OmtA synthesis appears to decline, the protein is proteolytically cleaved, and disappears completely by 90 h. The rate of OmtA proteolysis appears to be age- and grth medium-dependent. OmtA proteolysis was not observed to the same extent in cells grown in liquid culture. OmtA proteolysis may occur after “accidental” contact with fungal proteases released during mechanical disruption of fungal tissues. However, because the tissue is quickly frozen before disruption and because protease inhibitors are added to the extraction buffer, we hypothesize that OmtA proteolysis occurs as part of a natural process during fungal grth and development. To generate developmental structures in solid culture, fungi may require proteases to digest unneeded structures or metabolic proteins (20). For example, a mutation in one protease, subtilisin-related serine proteinase (ALP2), resulted in smaller conidiophore vesicles (50% reduction) and a lower number of conidia (80% reduction) in A. nidulans (20). We observed that SU-l fungal colonies on PDA, in which colonies grew more slowly than on YES (smaller diameter), also had more severe OmtA proteolysis. Conidiophore structures also could be found in fraction C3 on YES whereas they were absent in S3 on YES. Septal pores form “channels” between adjacent cells in the mycelium and appear to play an important role in cell-to-cell communication in filamentous fungi (3). Cytoplasm, mitochondria and nuclei migrate throughout the mycelium via septal pores (3,16). Because we do not see accumulation of OmtA in 24 h old colonies on YES but do see accumulation of OmtA in cells in the youngest fraction S3 (0 to 24 h) of a 72-h old fungal colony, we hypothesize that migration of OmtA together with cytoplasm and organelles occurs from older cells (fraction S2, 24 to 48 h) to younger cells (fraction S3, 0 76 to 24 h) on solid media. Alternatively, it is also possible that specific regulatory factor(s) involved in aflatoxin biosynthesis (eg. AflR) move fiom the aflatoxin producing cells in fraction S2 to fraction S3 inducing omtA expression in these younger tissues. OmtA-specific PAb were also used in immuno-fluorescence microscopy. The data provided strong evidence that our purification strategy resulted in PAb that were sensitive and specific. Although the paraffin-embeddment procedure and sectioning technique have been used for microscopic analysis of a variety of organisms (4), this is the first reported use in the study of a fungus grown on solid media and the first application in a study of Aspergillus parasiticus. We previously utilized a fungal cell preparation technique for immuno-labeling that required digestion of the cell wall using Novozyme (8). However, in these early localization studies, variation in cell wall digestion resulted in significant artifacts. The paraffin-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, our data suggest that OmtA is not produced exclusively in conidiophores as hypothesized. On the contrary, CLSM micrographs showed that OmtA is distributed throughout the colony and in both conidiophores and vegetative hyphae. Western blot analysis indicated that abundant full-length OmtA was observed in fraction S3, even though conidiophores were nearly absent in this area. In fraction 82, the highest intensity of OmtA was detected under CLSM, and OmtA accumulated to similar levels in 77 vegetative hyphae and conidiophores. 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 altemative 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 paraffin embedded sections to identify specific organelles; for example anti-SKL antibodies (10) 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. The anti-SKL labeled organelles were recently demonstrated to be Woronin bodies by immuno-electron microscopy (data not shown). There are some indirect data available that suggest that at least some aflatoxin proteins are localized in organelles. For example, in A. parasiticus, the aflatoxin enzyme OrdA was found to be membrane associated during protein purification (6). Based on amino acid sequence, another aflatoxin protein, AflJ, was predicted to contain a C- terminal microbody targeting signal, NRY, and three membrane-spanning regions ( 1 7). By comparing the protein sequence of purified OmtA with the predicted amino acid sequence derived from the OmtA cDNA, this protein was proposed to contain a leader sequence that apparently is processed to generate a 42-kDa protein (25). This leader sequence may be required for this enzyme to interact with membranes (25). Based on computer-assisted analysis (PSORT; prediction of protein localization sites Version 6.4: 78 http://psort.nibb.ac.jp/psort), we also speculate that putative peroxisomal targeting signals are present in aflatoxin proteins including Nor-1, AvnA, OmtA and OrdA (unpublished data). In contrast, OmtA (6) and Nor-1 (26) were observed in the post-microsomal cytoplasmic fraction in cell fractionation studies. Therefore, the distribution of aflatoxin enzymes within a fungal cell still is not clearly understood. Localization to specific organelles has been demonstrated for enzymes involved in secondary metabolism. Penicillin is a secondary metabolite produced by several filamentous fungi including A. nidulans. The enzyme, 6-aminopenicillanic acid acyltransferase, which catalyzes the final step of penicillin biosynthesis in Penicillium was localized to a membrane bound organelle, the microbody (19); the authors suggested that penicillin is synthesized in this organelle. In a separate study, prehelminthosporol, a phytotoxin produced by Bipalaris sarokinianawas, was localized in the Woronin body (1) which is thought to be derived from the peroxisome (24). Prehelminthosporol has been shown to disrupt plant plasma membranes and inhibit growth of gram-positive bacteria. Detailed sub-cellular localization of OmtA and several other aflatoxin enzymes using immunoelectron microscopy should help clarify the cellular site of aflatoxin synthesis in A. parasiticus. 79 ACKNOWLEDGEMENTS I. This work was supported by the Michigan Agricultural Experiment Station (MSU), the National Food Safety and Toxicology Center (MSU), the Center for Environmental Toxicology (MSU), and the National Institutes of Health (ROI CA52003-11). 2. Data in chapter 2 and 3 were published in Applied and Environmental Microbiology entitled “Function of native OmtA in viva and expression/distribution of this protein in colonies of Aspergillus Parasiticus”. 3. The LSM imagines in Figure 5 and 6 were generated by Ching-Hsun Chiou. 80 REFERENCES 1. Akesson, H., E. Carlemalm, E. Everitt, T. Gunnarsson, G. Odham, and H-B Jansson. 1996. Immunocytochemical localization of phytotoxins in Biaplaris sarakiniana. Fungal Gene. Biol. 20:205-216. 2. Ausubel, F. M., R. Brent, E. Kingston, D. D. Moore, J. G. Seidman, J. A. Smith, and K. Struhl. 2001. Current protocols in molecular biology. John Wiley and Sons, New York, NY. 3. Belozerskaya, T. A. 1998. Cell-to-cell communication in differentiation of mycelia] fungi. Membr. Cell Biol. 11: 831-840. 4. Bratthauer, GR. 1999. In L.C. Javois (ed.), Methods in Molecular Biology. Human Press, Totowa, New Jersey. 5. Cary, J., J. Linz, and D. Bhatnagar. 2000 Aflatoxins: Biological significance and regulation of biosynthesis. In J. Cary, D. Bhatnagar. and J. Linz (eds.), Microbial foodbome disease: mechanisms of pathogenesis and toxin synthesis. Technomic Publishing, Lancaster, PA. 6. Cleveland, T. E., A. R. Lax, L. S. Lee, and D. Bhatnagar. 1987. Appearance of enzyme activities catalyzing conversion of sterigmatocystin to aflatoxin B1 in late- growth-phase Aspergillus parasiticus cultures. Appl. Environ. Microbiol. 53:1711-3. 7. Guzman-de-Pena, D. and J. Ruiz-Herrera. 1997. Relationship between aflatoxin biosynthesis and sporulation in Aspergillus parasiticus. Fungal Genet. Biol. 21:198-205. 8. Harris, S. D., J. L. Morrell and J. E. Hamer. 1994. Identification and characterization of Aspergillus nidulans mutants defective in cytokinesis. Genetics. 136:517-532. 9. 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 G protein-dependent signaling pathway. EMBO J. 16: 4916-4923. 10. Keller, G-A., S. Krisans, S. J. Gould, J. M. Sommer, C. C. Wang, W. Schliebs, W. Kunau, S. Brody and S. Subramani. 1991. Evolutionary conservation of a microbody targeting signal that targets proteins to peroxisomes, glyoxysomes, and glycosomes. J. Cell Biol. 114:893-904. 11. Keller, N. P., H. C. Dischinger, JR., D. Bhatnagar, T. E. Cleveland and A. H. J. Ullah. 1993. Purification of a 40-kDa methyltransferase activity in the aflatoxin biosynthetic pathway. Appl. Environ. Microbiol. 59:479-484. 81 12. Liang, S.-H. 1996. The function and expression of the ver-I gene and localization of the Ver-l protein involved in aflatoxin B, biosynthesis in Aspergillus parasiticus. PhD Dissertation, Michigan State University, East Lansing. 13. Liang, S-H., C. D. Skory and J. E. Linz. 1996. Characterization of the function of the ver-l and var-1B genes, involved in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 62:4568-4575. 14. Liang, S-H., T-S. Wu, R. Lee, F. S. Chu and J. E. Linz. 1997. Analysis of mechanism regulating expression of the ver-I gene, involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 63:1058-1065. 15. Lin, B. H., N. P. Keller, D. Bhatnagar, T. E. Cleveland and F. S. Chu. 1993. Production and characterization of antibodies against sterigmatocystin O- methyltransferase. Food Agric. Immunol. 5:155-164. 16. Markham, P. 1995. Organelles of filamentous fungi. p. 75-98. In N. A. R. Gow, and G. W. Gooday (ed), The growing fungus, lst ed. Chapman & Hall, London, UK. 17. Meyers, D. M., G. Obrian, W. L. Du, D. Bhatnagar and G. A. Payne. 1998. Characterization of afl J, a gene required for conversion of pathway intermediates to aflatoxin. Appl. Environ. Microbiol. 64:3713-7. 18. Motomura, M., N. Chihaya, T. Shinozawa, T. Hamasaki and K. Yabe. 1999. Cloning and characterization of the O- Methyltransferase I gene (dmtA) from Aspergillus parasiticus associated with the conversion of demethylsterigrnatocystin to sterigmatocystin and dihydrodemethylsterigmatocystin to dihydrosterigmatocystin in aflatoxin biosynthesis. Appl. Environ. Microbiol. 65:4987-4994. I9. Muller, W. H., T. P. van der Krift, A. J. J. Krouwer, H. A. B. Wosten, L. H. M. van der Voort, E. B. Smaal and A. J. Verkleij. 1991. Localization of the pathway of the penicillin biosynthesis in Penicillium chrysogenum. EMBO journal. 10:489-495. 20. Reichard, U., G. T. Cole, T. W. Hill, R. Ruchel and M. Monod. 2000. Molecular characterization and influence on fungal development of ALP2, a novel serine proteinase from Aspergillusfumigatus. Int. J. Med. Microbiol. 290:549-558. 21. Skory, C. D., J. S. Horng, J. J. Pestka, and .I. E. Linz. 1990. Transformation of Aspergillus parasiticus with the homologous gene (pyrG) involved in pyrimidine biosynthesis. Appl. Environ. Microbiol. 56:3315-3320. 22. Skory, C. D., P-K. Chang, J. Cary, and J. E. Linz. 1992. Isolation and characterization of a gene from Aspergillus parasiticus associated with the conversion of versicolorin A to sterigmatocystin in aflatoxin biosynthesis. Appl. Environ. Microbiol. 58:3527-3537. 82 23. Trail, F., P-K. Chang, J. Cary and J. E. Linz. 1994. Structural and firnctional analysis of the nor-l gene involved in the biosynthesis of aflatoxin by Aspergillus parasiticus. Appl. Environ. Microbiol. 60:3315-3320. 24. Valenciano, S., J.R. De Lucas, I. Van der Klei, M. Veenhuis, and F. Laborda. 1998. Charatcerization of Aspergillus nidulans peroxisomes by immunoelectron microscopy. Arch. Microbiol. 170:370-376. 25. Yu, J., J. W. Cary, D. Bhatnagar, T. E. Cleveland, N. P. Keller, F. S. Chu. 1993. Cloning and characterization of a cDNA from Aspergillus parasiticus encoding an O- methyltransferase involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 59:3564- 71. 26. Zhou, R. 1997. The firnction, accumulation, and localization of the Nor-1 protein involved in aflatoxin biosynthesis; the function of the fluP gene associated with sporulation in Aspergillus parasiticus. PhD Dissertation, Michigan State University, East lansing. 83 CHAPTER 4 Sub-cellular localization of aflatoxin biosynthetic enzymes in time-dependent fractionated colonies of Aspergillus parasiticus INTRODUCTION Aflatoxins are highly toxic and carcinogenic secondary metabolites produced by several filamentous fungi including Aspergillus parasiticus, A. flavus, A. namius, and A. tamarii (12,23). Aflatoxins pose significant health and economic problems in the US and many other locations throughout the world because they frequently contaminate food and feed crops including com, peanuts, tree nuts and cottonseed (9,12). The biosynthesis of aflatoxin is a complex process that involves at least 17 enzyme activities (23). The distribution of these enzymes within fungal colonies and location within fungal cells remain unknown. The objective of this study was to investigate the distribution and sub-cellular location of representative enzymatic activities in the aflatoxin biosynthetic pathway. Toward this end, I focused attention on Nor-1 (29), Ver-l (19) and OmtA (33) that catalyze early, middle and late enzymatic steps in the aflatoxin biosynthetic pathway, respectively. Nor-1 and Ver-I are NADPH-dependent keto-reductases involved in the conversion of norsolorinic acid (NA) to averantin (AVN) and versicolorin A (VERA) to demethylsterigmatocystin (DMST), respectively; OmtA is a methyltransferase that converts sterigmatocystin (ST) to O-methylsterigmatocystin (OMST). 84 Specific secondary metabolites such as penicillin and preheliminthosporol are known to enhance the survival of the producing organism in the growth environment (5,1). However, the biological function of most secondary metabolites, including aflatoxin, is not clear. Characterization of the sub-cellular location of aflatoxin and/or its biosynthetic enzymes should provide important clues about the biological function and mechanisms that mediate intracellular transport, storage, and secretion of this toxic secondary metabolite by the producing firngus. As a potential mechanism to protect fungal cells from the potential deleterious effects of aflatoxin accumulation in the mycelium, I hypothesized that aflatoxin enzymes are compartmentalized in a sub-cellular organelle(s). Further, I hypothesized that the end-product, aflatoxin, is synthesized within this organelle and then secreted directly into the extracellular environment. To address these hypotheses, I analyzed enzyme distribution in fungal colonies by Western blot analysis and immuno-fluorescence microscopy, and canied out in situ localization of aflatoxin enzymes within fungal cells using immuno-microscopic methods following cryofixation. For immuno-transmission electron microscopy (TEM), it had been shown previously that a sample preparation procedure that combines rapid freezing and freeze substitution is superior to chemical fixation in terms of ultrastructure and antigenicity (1). Therefore, I used a modification of published methods (32) to localize Nor-1, Ver-l and OmtA in time-fractionated colonies. Nor-1 and Ver—l enzymes were primarily localized to the cytoplasm in cells 24 to 48 h old (fraction 2) suggesting that they are cytosolic enzymes. OmtA was also detected in the cytoplasm. However, in cells located near the basal (substrate) surface of the colony, large quantities of OmtA were detected in organelles identified as vacuoles. Based on 85 current data, we hypothesize that the relative distribution of OmtA to cytoplasm or vacuoles depends on the age and/or physiological condition of the fungal cells. MATERIALS AND METHODS Fungal strains and time-dependent fractionation of fungal colonies. A. parasiticus SUI (NRRLS 862, ATCC 56775) is a wild-type, aflatoxin-producing strain. AF S10 (aflR via gene disruption) is a non-aflatoxin producing strain derived from A. parasiticus NR-l (niaD) that in turn was derived from SUI (7). Asexual conidiospores (5 x105) of A. parasiticus SUI and AF S10 were inoculated onto the center of YES agar medium (2% yeast extract, 6% sucrose, pH 5.8) overlaid with sterile cellophane membranes and incubated at 29° C in the dark. 72 h-old colonies of SUI (S) and AF S10 (R) were fractionated into three concentric rings based on area covered at 24, 48, or 72 h of growth (fraction 1, 48-72 h old; fraction 2, 24-48 h, and fraction 3, 0-24 h) to generate fractions S1, 82, S3 and R1, R2, R3 respectively as described by Lee et al. (17). The harvested mycelia from appropriate sections of the colony were used in sample preparation for Western blot analysis and for fluorescence and electron microscopy. Western blot analysis of proteins isolated from colony fractions. Sample preparation and Western blot analysis of A. parasiticus proteins isolated from colony fractions were conducted using methods described in Lee et al. (1 7). Each lane on the 12% SDS-PAGE gel contained 60pg protein. Nor-1, Ver-I and OmtA polyclonal antibodies were raised against maltose binding protein fusions generated for each 86 aflatoxin protein (35,18,19,l7). Anti-Nor-l serum (rabbit No. 126) was generated in this study and used at a 1 to 5000 dilution. Anti-Ver-l IgG was used at 2 pg per ml. Immuno-fluorescence labeling and confocal laser scanning microscopy (CLSM). Preparation of paraffin-embedded fungal sections, immuno-labeling, and microscopic analysis via CLSM utilized methods described by Lee et al. (17). The samples were probed with primary antibodies against Nor-1 (1 to 500 dilution), Ver-l (20 pg/ml) (19) or OmtA (20 pg/ml) (17), 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). Fluorescence image detection was performed on a Meridian INSIGHT plus laser-scanning microscope (Meridian Instruments, Inc., Okemos, MI) with a 488 nm laser line. The Alexa 488 green fluorescence was detected using a BP 530/30 banier filter. All images were captured using a 40X Zeiss Plan-NEOFLUAR oil objective lens (N .A.=1 .3) (Carl Zeiss Inc., Germany) with a 1X CCD cool-charged detector allowing analysis of a large number of cells in one 512 x 480 image. Fluorescence image analysis of strains SUI and AF S10 was conducted using the same instrument parameter settings. For all AF 810 samples, bright field images were also generated to demonstrate the size and number of cells analyzed. Quantitative fluorescence intensity analysis. Alexa 488 fluorescence was quantified using IQ Master Program (V2.31) image analysis software that accompanied the Meridian Insight laser-scanning microscope (Meridian Instruments, Inc., Okemos, MI). The CLSM images were acquired under the identical instrument settings for samples 87 analyzed in the same immuno-labeling experiments. For each colony fraction, twenty images were acquired for intensity analysis and the average pixel number was reported. Sample preparation for immuno—electron microscopy. Three to five mm diameter circles of firngal tissue were obtained from fraction 2 of colonies of SUI and AF S10 grown on YES agar plates. These circles were cut from the colony using sterile 200 pl pipette tips with the end removed with a sterile razor blade to generate a cutting edge with the correct diameter. Samples were immediately cryofixed at ——190° C in a commercial Jet-Freezer (RMC MF 7200) and then transferred to a tube containing acetone that had been frozen by immersion in liquid nitrogen. The following substitution procedure was modified from a published method (32). Samples in frozen acetone were then stored at -80° C for 6 (1. Samples were washed twice with fresh, ice-cold acetone (- 20° C) and then immersed in flesh, ice-cold acetone (-20° C) containing 0.2% glutaradehyde and stored at -20° C for 24 h. After washing with fresh,“ ice-cold acetone (- 20° C) three times, acetone in samples was replaced with 100 % ethanol by a graded series of ethanol/acetone (35%, 50%, 75%, 90% [30 min each] and 100 % [two times, 30 min each]). Samples were then infiltrated with a graded series of LR White resin (Ted Pella, Inc., Redding, CA) /ethanol: 5, 10, 20, 40, 60% (3 h each), 90% for 24h and 100% (3 changes) for 2 d at -20° C. Polymerization was carried out under UV light (366 nm) at 4° C for 48 h. Immuno-gold labeling and electron microscopy. Fungal sections for immuno- gold labeling were cut to 90 ~ 100 nm thickness using an MTX ultrarnicrotome (RMC, Tucson, Arizona). Sections were collected onto formvar-coated grids. Each grid 88 contained approximately 10-20 thin sections. For each labeling experiment, two grids each from SUI and AF SIO tissues were used. For each antibody, at least three independent labeling experiments were performed. First, grids with sections were incubated with blocking solution (1% BSA plus 0.1% saponin in TBS) at 4° C overnight. Primary antibody treatment (purified OmtA IgG, 50 pg per ml; anti-Verl IgG, 140 pg per ml; anti-serum: anti-SKL serum, 1: 200; anti-Nor-I, 1: 500) was performed at room temperature for 4 h. Secondary antibody treatment (goat anti-rabbit conjugated with 10 nm gold; 1 to 30-fold dilution) (Ted Pella, Inc., Redding, CA) was performed at room temperature for 2 h. After each antibody incubation, grids were washed once with Tris- buffered saline (pH 7 .5) containing 1% BSA and 0.1% saponin for 5 min followed by five washes with TBS for 25 min total. Finally, grids were washed with dd H20 for 30 s and post stained with 3% uranyl acetate for 20 min. Sections were observed using a JEOL 100CX II transmission electron microscope (Tokyo, Japan) at 100 kV. RESULTS Western blot analysis of Nor-land Ver-l isolated from colony fractions 1, 2, and 3 of A. parasiticus SUI and AF SlO demonstrated the specificity of the antibody preparations used in microscopic analysis (Figure 4.1). Similar analysis was conducted previously on OmtA only, in which increasing quantities of proteolytically cleaved enzyme were detected in fractions 2 and 1 (17); the highest proportion of full-length protein was observed in fraction 2. In contrast, Nor-1 and Ver-I in the current study appeared predominantly in full-length form in all colony fractions. The proteolytically cleaved forms of these two enzymes were not observed in cells grown on solid YES 89 SI 52 83 R1 R2 R3 MWIkDa) 1‘. mg 8 $3388 V'cr-I zs-o - I l B 81 82 SJ R1 R2 R3 MW(kDa) E120 100 80 W 60 fl 5" Nor-l ‘ n 40 31-0“ “mu—.m- .- a 30 - ’ 20 Figure 4.1. Western blot analysis of fungal protein extracts using Ver-I and Nor-1 polyclonal antibodies. Proteins were extracted from time-fractionated colonies of SUI and AF S10 grown on YES agar medium and were subjected to Western blot analysis with Ver-l (Panel A) and Nor-1 (Panel B) polyclonal antibodies. 72 h-old colonies were fractionated into three concentric rings based on area covered at three time points: for S1 and R1, 48 to 72 h; 82 and R2, 24 to 48 h; S3 and R3, 0 to 24 h. The lanes to the left and right contain molecular mass standards. The molecular mass of standards are marked at the right of Panels A and B. The molecular masses of Ver-l and Nor-1 (indicated by arrows at the left) are 28 kDa and 31 kDa, respectively. 90 medium. However, like OmtA, the highest quantity of Nor-l and Ver-l was observed in fraction 2 of SUI. Immuno-fluorescence microscopy via CLSM confirmed the specificity of the Nor-1 and Ver-I polyclonal antibodies for microscopic analysis. The specificity of the OmtA polyclonal antibody was confirmed in a previous study (17). Nor-1 and Ver-I were detected in all colony fractions of SUI (Figure 4.23 for Nor-I and Figure 4.2d for Ver-l). Under the same instrument settings, very little signal was detected in AF 810 (Figure 4.2b for Nor-1 and Figure 4.2e for Ver-l). Therefore, bright field images of the AF S10 colony fractions are shown to illustrate the typical size and numbers of cells analyzed (Figure 4.2e and 20. Quantitative analysis of the fluorescence intensity in the CLSM digital images using IQ Master Program (V2.31) image analysis software confirmed that the highest quantity of these three aflatoxin enzymes occurred in fraction 2 (Table 4.1). Therefore, we focused our immuno-TEM analysis on the proteins in this colony fraction. Using our sample preparation, labeling, and imaging procedures, very few gold particles were observed on control sections of wild-type strain SUI treated with secondary antibodies only (Figure 4.3A and B). Therefore, we interpreted an intense signal (black dots representing 10 nm gold particles) on sections treated with both primary and secondary antibodies to result from the specific interaction between primary antibodies and their target proteins immobilized within the cell. However, we occasionally detected labeling of the cell wall, especially for anti-Ver-l, in certain sections of both SUI and AF S10 (no aflatoxin proteins produced), leading us to conclude that this was non- specific labeling (data not shown). It was reported previously that non- specific labeling may result either from non-specific binding of primary antibodies to the cell wall or from cell wall-reactive antibodies that arise in rabbits exposed to or 91 Figure 4.2. Immuno-fluorescence confocal microscopy of Nor-l and Ver-l in A. parasiticus SUI and AF SIO (aflR knockout mutant) grown on YES agar for 72 h. Colonies of SUI and AF S10 were divided into 3 fractions (see Material and Methods). The paraffin embedded fungal sections were immuno-labeled with anti-Nor-I 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 SUI sections immuno- labeled with anti-Nor-l and anti-Ver-l are shown in Rows a and d, respectively; and of AF SIO sections in Rows b and e. The related bright field images of the AF SIO colony fractions are shown in Rows c and f (legends labeled in black). Bar, 100 pm. 92 A. Nor-l B. Ver-l 5' d Figure 4.2. 93 Table 4.1. Quantative fluorescence intensity analysis of A. parasiticus SUI and AF S10 immuno-fluorescence labeled with Nor-1, Ver-I and OmtA antibodies. (Fungal colony) (SU— 1) (AF S I 0) Protein / Colony fraction S1 82 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. 94 Figure 4.3. Immuno-gold labeling of A. parasiticus SUI and AF 810. In Panels A and B, ultra-thin sections of fungal tissues of the aflatoxin producing strain SUI were prepared for TEM and labeled with secondary antibodies only. Ultra-thin sections of fungal tissues of AF SIO (Panel C and E) and SUI (Panel D and F) were labeled with primary antibodies against Nor-1 (Panel C and D) and Ver-l (Panels E and F), followed by 10 nm gold beads conjugated to goat anti-rabbit IgG secondary antibodies (1 to 30 dilution) as described in Methods. Bars represent 500 nm in A and 250 nm in B, C, D, E and F. Abbreviations: cw, cell wall; mb, microbodies; M, mitochondria; N, nuclei; V, vacuole. 95 .111, ‘L'l'il‘l‘ip‘ “'iifl" . 4"”ijin . l =~' .. . -. -i muwmlhulm‘m .3- mm lllillilnmi.‘ I! ' .1. , ll‘ ., , , . . 1 WW I . . “if j 1‘,” .151 “3‘1““ .‘1 .11 _'i‘ 11,1 , Figure 4.3. infected by yeast or molds before or during antibody production (3, 4). Widely scattered gold particles were also found occasionally associated with the nucleus and mitochondria in both AF S10 and SUI labeled with Nor-I and Ver-l antibodies, suggesting that this was non-specific labeling or resulted from cross-reactive proteins located in these organelles. This background “noise” could be largely eliminated by a further step of antibody purification using affinity subtraction as described in Lee et al. (17). As shown in this study, very little labeling was noted in the cell wall, mitochondria, and nucleus using affinity-purified antibodies to OmtA (Figure 4.4 and Table. 4.1). When highly specific antibodies to Nor-1, Ver-l, and OmtA were used to label sections from fraction 2 of SUI , high signal intensity was observed in the cytoplasm in many but not all cells (Figure 4.3D, F and 4.4C, D). The gold particles frequently aggregated to form clusters (Figure 4.3D and 4.4D); similar clusters were not seen in the cytoplasm of control sections obtained from AF $10 or when only gold-labeled secondary antibodies were used. The immuno-gold labeling results are in good agreement with immuno-fluorescence data (Figure 4.2) in this and our previous study and may help explain the “patchy” appearance of the fluorescent signal observed (17). Of particular interest, OmtA antibodies specifically and intensely labeled small and large organelles primarily in cells located near the basal surface (substrate surface) of the colony in fraction 2 of SUI (Figure 4.4E and 4.5). Based on ultrastructure /morphology, these organelles have been identified as vacuoles (2). The labeling intensity suggested a high concentration of OmtA in these organelles. In contrast, antibodies against Nor-l or Ver-l did not label these organelles in sections from the same location (basal portion of fraction 97 Figure 4.4. Immuno-gold labeling of OmtA in Aspergillus parasiticus AF 810 and SUI. Ultra thin sections of fungal tissues of AF SIO (Panels A and B) and SUI (Panels C and D) were labeled with primary antibodies against OmtA followed by gold-labeled secondary antibodies as described in Methods. Panel B is a schematic representation of the colony fractionation procedure and shows the cell distribution in a typical thin section. The specific location of OmtA in cells near the basal surface of the colony (shaded gray) is also shown in Panel E. Bars present 500 nm in panels A, C and E, and 250 nm in panels B and D. Abbreviations: cw, cell wall; M, mitochondria; N, nuclei; V, vacuole. 98 Figure 4.4. 99 Figure 4.5. Immuno-gold labeling of OmtA in Aspergillus parasiticus SUI. Sections cut from the basal surface of the colony were labeled with primary antibodies against OmtA. The gold-labeled vacuoles inside of cells were numbered (1 to 8) and higher magnification images of these organelles are shown in Panels B, C, D and E. Bars present 2 pm in panels A and 1 pm in B, C, D and E. Abbreviations: cw, cell wall; M, mitochondria; N, nuclei; V, vacuole. 100 \ . 04' rt 1, mini. h '1 ‘il F'I Its ‘,,,li11 . . " , .11. 31111141 ' . Figure 4.5. 101 2) (data not shown). Cells located above the basal region contained smaller organelles which were sporadically labeled by OmtA and not at all by antibodies to Nor-l and Ver- 1. Similar organelles in AF S10 did not label with polyclonal antibodies to any aflatoxin enzymes. The gross ultrastructure of cells observed in thin sections suggested that our fixation, sectioning and labeling procedures successfully preserved the integrity of organelles bounded by double (nuclei and mitochondria) and single (vacuoles, microbodies, Woronin bodies) membranes. In order to confirm that fragile organelles bounded by single membranes could retain both ultrastructure and the antigenicity of internal proteins, we labeled some sections from fraction 2 of SUI and AF S10 with antibodies against peroxisomal targeting signal 1 (PTS-l; anti SKL). Anti-SKL can label proteins exported to peroxisomes (microbodies; mb) and Woronin bodies (derived from microbodies; 30). Microbodies and Woronin bodies were specifically and intensely labeled with anti-SKL (Figure 4.6A, B and C). However, not all microbodies within one cell were labeled to the same signal intensity possibly because the specific protein concentration or protein composition varies between microbodies in the same cell. It has been demonstrated previously that certain proteins lacking PTS-1 can be localized in peroxisomes (21,27); these proteins should not be recognized by the anti-SKL preparation. In our studies, labeling of microbodies by Nor-land OmtA antibodies was rarely observed. However, we did observe that Woronin bodies but not other microbodies were labeled with anti-Ver-l (Figure 4.6D and E). Woronin bodies in AF S10 were also labeled to a similar intensity, suggesting that this is non-specific labeling possibly due to a high antibody concentration used or that “contaminating” antibodies bound to cross- reactive proteins located in this organelle (Figure 4.6F and G). 102 Figure 4.6. Immuno-gold labeling using anti-SKL and anti-Ver-I in A. parasiticus AF S10 and SUI. Ultra-thin sections of firngal tissues obtained from fraction 2 were labeled with primary antibodies to PTS-1 (anti-SKL) (Panels A, B, and C) or primary antibodies to Ver-l (Panels, D, E, F and G) followed by gold-labeled secondary antibodies as described in Methods. Panels A, B, D, F, and G, represent ultra-thin sections prepared from fungal tissues of AF 810 and Panels C and E represent ultra-thin sections prepared from fungal tissues of SUI. Bars present 250 nm. Abbreviations: cw, cell wall; mb, microbodies; wb, woronin body; st, septum; M, mitochondria; V, vacuole. 103 ' i ll 1“ Illlllill . .nll '. \I 1‘ ‘ .ul'lil .1,,IIIII I III. “if.“ A II'IIIIIIIII 1" l "I 1" I‘Maimlr‘riliriilmnllllmi ‘i‘r'rllllilli l'."!'"‘"“ Figure 4.6. 104 DISCUSSION My goal was to analyze the distribution and sub-cellular localization of the aflatoxin enzymes Nor-1, Ver-I, and OmtA in Aspergillus parasiticus grown on YES agar growth medium. Western blot analysis of protein extracts from time-fractionated colonies demonstrated that fraction 2 (cells 24 to 48 h old) contained the highest concentration of each target protein. In addition, proteolytic cleavage of OmtA occurred to a greater extent in “older” fungal tissues (fractions 1 and 2) but this was not observed for Nor-1 and Ver-l under these culture conditions. The specificity and timing of expression of the proteolytic enzymes and the potential role of cleavage in activation/ inactivation of aflatoxin enzymes is clearly of interest for follow up studies. Immuno-fluorescence microscopy confirmed that Nor-I, Ver-l and OmtA are present at highest concentration in fraction 2; the “patchy” appearance of the fluorescent signal suggested that the proteins aggregate in the cytoplasm or are compartmentalized consistent with our previous studies on Nor-1, Ver-l, and OmtA (17, 18, 35). At even higher resolution, TEM analyses confirmed that Nor-1, Ver-I and OmtA are present in the cytoplasm of many, but not all cells in colony fraction 2. In addition, most cells located near the basal surface of fraction 2 were closely packed and contained one to several large organelles heavily labeled with gold-labeled antibodies to OmtA. In some cells, the large and small organelles appeared to fuse together consistent with a model for vacuolar development in Aspergillus reported previously (22). To our knowledge, this is the first report of aflatoxin enzymes localizing to a specific cellular organelle. 105 The morphology and apparent developmental origin of the labeled organelles in fraction 2 prompted us to identify them as vacuoles. Vacuoles have been associated with several biological functions in fungi including degradation or recycling of proteins and whole organelles (2), storage of metabolites, ions and amino acids, enzyme maturation (e.g. aminopeptidase I) and pH homeostasis (15,28). Vacuoles are also the site for lipid degradation and glycerol production at specific stages of appressoria formation in Magnaparthe grisea (31). In yeast, the biological sulfonium compound, S-adenosyl- methionine (AdoMet), accumulates to high levels in vacuoles and cytoplasm (25,26,13). S-adenosyl-methionine is an important cofactor that provides the methyl group in the reaction catalyzed by OmtA. Based on the available data, it is difficult to determine if the vacuolar localization of OmtA occurs for protein recycling, for enzyme activation, or for some other purpose in A. parasiticus. Although the data appeared quite convincing, it was possible that disruption of organelles occurred during sample preparation and was in part or fully responsible for the observed cytoplasmic location of Nor-1 , Ver-l , and OmtA. To minimize this possibility, we demonstrated that the ultra-structure of Woronin bodies and microbodies (bounded by single membranes) and the antigenicity of their content proteins were preserved using our methods for sample preparation and microscopic analysis. These structures were intensely labeled by anti-SKL but not by antibodies to Nor-1 or OmtA. Some labeling of Woronin bodies did occur with anti-Ver-l, however, the observed labeling occurred in both AF 810 (control) and SUI strains allowing us to interpret this as non-specific labeling. This interpretation appears to be likely because disruption of aflR in A. parasiticus (such as in AF S-IO) results in loss of nor-1, ver-I and omtA gene transcripts 106 (6,7) and proteins (this study). Therefore, it is unlikely that Ver-l exists at any location in AF S I 0. Similarly, the ultrastructure of double-membrane bound organelles (nuclei and mitochondria) as well as other single-membrane bound organelles (vacuoles) were also maintained in this study. With the exception of vacuoles in the basal region of fraction 2 of SUI , these organelles did not label with antibodies against the aflatoxin enzymes. These results suggest that Nor-1, Ver-I, and OmtA do not localize to nuclei, mitochondria, Woronin bodies, or microbodies. Because the absence of signal in these organelles does not appear to be an artifact arising from organelle breakage, these data strongly support the specificity of labeling of the vacuoles observed in the basal region of fraction 2. Two observations from this study are particularly noteworthy. First, intensely labeled, weakly labeled and unlabeled cells could be found adjacent to each other in the same thin section analyzed by TEM. It is known that cell components located under the surface of a section are unable to be labeled. However, the cytoplasm proteins such as Ver-l and Nor-1 in these unlabeled cells were remainly unlabeled even in serial sections. These data suggest that the extent and timing of aflatoxin gene expression varies from cell to cell, even in an area of the colony that is presumably the same age. We propose that this variation may be related to the local concentration of available nutrients or the relative age of fungal cells (5,10,20). Second, OmtA appeared in the cytoplasm in certain cells and in the vacuole in other cells in the same colony fraction. In Saccharomyces, cytoplasmic proteins can be transported into vacuoles by at least two alternative pathways: 1) macroautophagy; and 2) the cytoplasm to vacuole targeting (cvt) pathway. The cvt pathway is regulated by the availability or limitation of specific nutrients (e. g. 107 glucose or ethanol) (15,16,28); to date, no pathway similar to the yeast cvt has been reported in filamentous fungi. However, we propose that transport of OmtA is similar to the cvt pathway in that a specific set of enviromnental changes (cell age, physiological conditions or nutrient availability) may trigger the transport of OmtA into the vacuole. Based on data from this and previous studies, it is reasonable to assign two potential alternative roles for the vacuole in aflatoxin synthesis: 1) as a means to reduce or limit aflatoxin synthesis, OmtA is transported to and inactivated in vacuoles via proteolytic cleavage; 2) as a means to shield the cell from the potential toxic effects of aflatoxin in the mycelium, OmtA is transported to the vacuole together with the late ) aflatoxin pathway intermediate ST. In the vacuole, OmtA is activated and ST is converted to OMST and further converted to AP B, by OrdA (24,34) (catalyzes the final step of aflatoxin biosynthesis) localized in or on the same vacuole. These hypotheses form the basis for our future studies on aflatoxin synthesis. 108 ACKNOWLEDGEMENTS Chapter 4 was submitted for publication in Applied and Environmental Microbiology (2003). Figure 4.2 and Table 4.1 were generated by Ching-Hsun Chiou. We also thank Dr. Shirley Owens for help in quantification of fluorescence intensity and Dr. Xudong Fan in operating the jet-freezer in Center of Advanced Microscopy at Michigan State University. 109 REFERENCES I. Akesson, H., E. Carlemalm, E. Everitt, T. Gunnarsson, G. Odham, and H-B J ansson. 1996. Immunocytochemical localization of phytotoxins in Biaplaris sorakiniana. Fungal Gene. Biol. 20:205-216. 2. Amor, C., A. I. Dominguez, J. R. D. Lucas, and F. Laborda. 2000. The catabolite inactivation of Aspergillus nidulans isocitrate lyase occurs by specific autophagy of peroxisomes. Arch. Microbiol. 174:59-66. 3. Atkin, A. L. 1999. Preparation of yeast cells for confocal microscopy. Methods in Molecular Biology. 122:131-139. 4. Binder, M., A. Hartig, and T. Sata. I996. Immunogold labeling of yeast cells: an efficient tool for the study of protein targeting and morphological alterations due to overexpression and inactivation of genes. Histochem. Cell Biol. 106:115-130. 5. Calvo, A. M., R. A. Wilson, J. W. Bok, and N. P. Keller. 2002. Relationship between secondary metabolism and fungal development. Microbiol. Mol. Biol. Rev. 66:447-459. 6. Cary, J. W., K. C. Ehrlich, M. S. Wright, P.-K. Chang and D. Bhatnagar. 2000. Generation of aflR disruption mutants of Aspergillus parasiticus. Appl. Microbiol. Biotechnol. 53: 680-684. 7. Cary, J. W., 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. 8. CAST. 2003. Council for Agricultural Science and Technology. Mycotoxins: Risks in plant, animal, and human systems. Report 139. 9. Chang, P.-K., J. Yu, D. Bhatnagar, and T. E. Cleveland. 1998. Abstract. The USDA-ARS Aflatoxin Elimination Workshop. 10. Chiou, C.-H., M. Miller, D. L. Wilson, F. Trail, and J. Linz. 2002. Chromosomal location plays a role in regulation of aflatoxin gene expression in Aspergillus parasiticus. Appl. Environ. Microbiol. 68:306-315. 11. Cleveland, T. E., A. R. Lax, L. S. Lee, and D. Bhatnagar. 1987. Appearance of enzyme activities catalyzing conversion of sterigmatocystin to aflatoxin B, in late- growth-phase Aspergillus parasiticus cultures. Appl. Environ. Microbiol. 53:1711-3. 110 12. Eaton, D. L. and J. D. Groopman (eds.). 1994. The toxicology of aflatoxins: human health, veterinary, and agricultural significance. San Diego, CA: Academic. 13. Farooqui, J. L, H. W. Lee, S. Kim, and W. K. Paik. 1983. Studies on the compartmentation of S-adenosyl-L-methionine in Saccharamyces cerevisiae and isolated rat hepatocytes. Biochimica et Biophysica Acta. 757:342-352. 14. Keller, N. P., H. C. Dischinger, JR., D. Bhatnagar, T. E. Cleveland and A. H. J. Ullah. 1993. Purification of a 40-kDa methyltransferase activity in the aflatoxin biosynthetic pathway. Appl. Environ. Microbiol. 36:270-3. 15. Klionsky, D. J., P. K. Herman, and S. D. EMR. 1990. The fungal vacuole: composition, function, and biogenesis. Microbiol. Rev. 54: 266-292. 16. Klionsky, D. J. 1997. Protein transport from the cytoplasm into vacuole. J. Membrane Biol. 157: 105-115. 17. 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. 18. Liang, S.-H. 1996. The function and expression of the ver-l gene and localization of the Ver-l protein involved in aflatoxin B, biosynthesis in Aspergillus parasiticus. PhD Dissertation, Michigan State University, East Lansing. 19. Liang, S.-H., T.-S. Wu, R. Lee, F. S. Chu and J. E. Linz. 1997. Analysis of mechanism regulating expression of the ver-I gene, involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 63:1058-1065. 20. Luchese, R.H., and W. F. Harrigan. I993. Biosynthesis of aflatoxin- the role of nutritional factors. J. Applied Bacteriology. 74: 5-14. 21. Maeting, 1., G. Schmidt, H. Sahm, J. L. Revuelta, Y.-D. Stierhof, and K.-P. Stahmann. 1999. Isocitrate lyase of Ashbya gossypii- transscriptional regulation and peroxisomal localization. FEBS letters. 444: 15-21. 22. 0hsumi, 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. 23. Payne, G. A. and M. P. Brown. 1998. Genetics and physiology of aflatoxin biosynthesis. Annu. Rev. Phytopathol. 36:329-362. 24. Prieto, R., and C. P. Woloshuk. 1997. ardl, an oxidoreductase gene responsible for the conversion of O-methylsterigmatocystin to aflatoxin in Aspergillusflavus. Appl. Environ. Microbiol. 63:1661-1666. 111 25. Schwencke, J., and H. De Robichon-Szulmajster. 1976. The transport of S- adenosyl-L-methionine in the isolated yeast vacuoles and spheroplasts. Eur. J. Biochem. 65:49-60. 26. Svihla, G., J. L. Dainko, and F. Schlenk. 1969. Ultraviolet micrography of penetration of exogenous cytochrome c into the yeast cell. J. Bacteriol. 100:498-504. 27. Taylor, K. M., C. P. Kaplan, X. Gao, and A. Baker. 1996. Localization and targeting of isocitrate lyases in Saccharamyces cerevisiae. Biochem. J. 319:255-262. 28. T humm, M. 2000. Structure and function of the yeast vacuole and its role in autophagy. Microscopy Research and Technique. 51: 563-572. 29. Trail, F., P-K. Chang, J. Cary and J. E. Linz. 1994. Structural and functional analysis of the nar-l gene involved in the biosynthesis of aflatoxin by Aspergillus parasiticus. Appl. Environ. Microbiol. 60:3315-3320. 30. Valenciano, S., J.R. De Lucas, 1. Van der Klei, M. Veenhuis, and F. Laborda. 1998. Characterization of Aspergillus nidulans peroxisomes by immunoelectron microscopy. Arch. Microbiol. 170:370-376. 31. Weber, R. W. S., G. E. Wakley, E. Thines, and N. J. Talbot. 2001 .The vacuole as element of the lytic system and sink for lipid droplets in maturing appressoria of Magnaporthe grisea. Protoplasma. 216: 101-1 12. 32. Xu, H and K. Mendgen. 1994. Endocytosis of 1,3-B-glucans by broad bean cells at the penetration site of the cowpea rust fungus (haploid stage). Planta. 195:282—290. 33. Yu, J., J. W. Cary, D. Bhatnagar, T. E. Cleveland, N. P. Keller, F. S. Chu. 1993. Cloning and characterization of a cDNA from Aspergillus parasiticus encoding an O- methyltransferase involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 59:3564- 71. 34. Yu, J., P.-K. Chang, K. C. Ehrlich, J. W. Cary, D. Bhatnagar, and T. E. Cleveland. 1998. Characterization of the critical amino acids of Aspergillus parasiticus cytochrome P-450 monooxygenease encoded by ordA that is involved in the biosynthesis of aflatoxin B1, G1, B2 and G2. Appl. Environ. Microbiol. 64:4834-4841. 35. 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. PhD Dissertation, Michigan State University, East Lansing. 112 CHAPTER 5 Detection of aflatoxin B, in fungal spores and colonies INTRODUCTION The biological function of aflatoxin in the producing fungus is still not clear. Since aflatoxins are secondary metabolites and synthesized primarily during stationary phase, they are not necessary for primary growth of the cell. It has been shown previously that the production of sterigmatocystein (ST) in A. nidulus and aflatoxin in A. parasiticus is closely associated with asexual sporulation (2), and positively regulated by F le (1). Although the details of the connection between production of aflatoxin and fungal sporulation are entirely not clear, the presence of aflatoxins in asexual conidiospores has been reported (4,6). Since spores are important structures involved in dispersal and colonization, one of the proposed function of aflatoxins is to provide the fungus with a survival advantage by overcoming environmental threats (e.g. resistance to UV light). The aflatoxin concentrations in spores harvested from different media were estimated. Wicklow and Shotwell (1982) reported that conidiospores harvested from Czapek agar cultures (grown in the dark at 28° C for 21 days) contained 1,106 to 54,300 ppb (ng/g) of AFB, in four A. parasiticus strains and 36 to 97,400 ppb in five A. flavus strains. Palmgren and Lee (1986) reported that conidiospores harvested from A. parasiticus (SRRC-2004) cultured on autoclaved rice at 25°C for 7 days contained 166,000 ppb of AFB,. In contrast to aflatoxin distribution in wild-type strains, the distribution of the 113 aflatoxin intermediate, Nosolorinic acid (NA), in a nor-l mutant is mainly in vegetative mycelia. In the present study, I wanted to compare: 1) the distribution of aflatoxin between colony and substrate (agar); and 2) the aflatoxin concentration in spores harvested under wet and dry conditions. Although the aflatoxin gene cluster contains an aflT gene, which may function as a toxin transporter, no data in the literature have demonstrated that a specific secretion system including AflT is used to transport aflatoxin into the environmental media. In this study, I measured the aflatoxin concentration in time fractionated colonies grown for different total times and compared the aflatoxin distribution between fungal colonies and the substrate media. My results showed that, in YES cultures ranging from 2 days to 5 days of age, the ratio of aflatoxin in the fungal mycelium and total aflatoxin produced decreased with culture time. However, for an unknown reason, the ratio increased at days 6 and 7. In addition, culture media affected the distribution of aflatoxin in the mycelium. Compared with 60% in YES culture, only 20% of the total aflatoxin was detected in fungal mycelium cultured on PDA agar. Based on the current data, the nature of the specific transport system for aflatoxin can not be determined yet. My data also showed that there is a 10-fold difference in aflatoxin concentration between conidiospores isolated under wet and dry conditions. The higher concentration obtained from spores harvested under wet conditions is due to absorption of aflatoxin from water used for harvest onto the surface of the spores. 114 Mutants or Treatments Norsolorinic acid ATCC 24690 Averantin ATCC 56774 Averufanin Averufin ATCC 24551 1? Hydroxyversicolorine HVN-I Versiconal Hemiacetal Acetate Dichlorvos 2%: versicolorin A ATCC 36537 AFBl «H.— O-methylsterigmatocystin 4—_ Sterigmatocystin SRRC 2043 Figure 5.1. Mutants in the aflatoxin biosynthetic pathway. 115 MATERIAL AND METHODS Fungal strains. A. parasiticus SU1(NRRL5862, ATCC 56775) is a wild-type, aflatoxin-producing strain. A. parasiticus RHNI (SRRC2043; ordA) accumulates the aflatoxin intermediate OMST and ATCC36537 (var-I) accumulates VA (Figure 5.1). Aflatoxin distribution in time-dependent fractionated colonies of SUI grown on solid medium. To determine the accumulation and distribution of aflatoxin in fungal colonies grown on solid culture media, conidiospores (2 x 105) of A. parasiticus SUI were inoculated onto the center of YES or PDA agar overlaid with steriled cellophane membranes and incubated at 29° C in the dark. Colonies were analyzed after 2, 3, 4, 5, 6 and 7 d of growth. The 2 d-old colonies were harvested directly without fractionation. Colonies older than 2 d were marked at the edge every 24-h as an indicator for future cutting. For example, the 3 d-old colony was fractionated into two concentric rings (ring- 1 and 2) based on area covered at the two time points: Ring-l (48 h) contained mycelia from the colony center out to a distance of 2.5 cm (24 to 72 h) and Ring-2 (0 t024 h) contained mycelia from 2.5 to 4.2 cm (0 to 24 h) (Figure 5.2 and Table 5.1). The harvested mycelia from appropriate fractions were weighed and extracted with chloroform. The chloroform extracts were evaporated and resuspended in methanol. In the same culture, the substrate (agar) and membrane were also analyzed for aflatoxin levels. The aflatoxin concentrations were determined by ELISA (5) Spores harvested by water (wet conditions). 1) In early experiments, the fungus was grown on a nylon membrane on PDA agar plates at 29° C in the dark for 7 d. 116 Figure 5.2. Diagram of colony fractionation. 117 Table 5.1. Fractionation of fungal colonies at different ages. Age\Diamter (cm) 2.5 2.5-4.2 4.2-5.0 5.0-6.5 6.5-8.0 2d Ring-1 3d Ring-I Ring-2 4d Ring-1 Ring-2 Ring-3 5d Ring-I Ring-2 Ring-3 Ring-4 6d Ring-l Ring-2 Ring-3 Ring-4 Ring-5 118 Spores were released from conidiophores into water (harvest water) by mechanical force. Colonies with membranes were immersed into the water and the surface of the colony was scraped with a spatula. The water containing spores was passed through glass wool to remove any hyphal fragments. Spores were pelleted from the water by centrifugation at 3,000 x RPM for 20 min. Both water and spores were measured for aflatoxin concentrations. 2) In later experiments, firngal strains, including SUI, RHNI (SRRC2043) and ATCC36537, were grown on cellophane membranes laid onto PDA agar plates and incubated at 29°C in the dark for 7 d. For spore isolation, the whole colony was removed from the cellophane into a 50-ml conical tube and 35 ml of water was dispensed into the tube. Spores were released into the harvest water by rocking the tube back and forth for 1 min. The spore-containing water was then filtered through glass wool. Spores were pelleted by centrifuged at 20,000 x g for 15 min and resuspend in 1 ml of H20. The resulting spore pellet and the harvest water were used in binding assays as described in the following section to test their ability to absorb aflatoxin (AFB, and G,) and aflatoxin intermediates (OMST and VA) contained in the harvest water. The AFB, and G, or OMST-containing water used in the binding experiment were firrther filtered through 0.22 pm or 0.45 pm filters to eliminate any contaminating spores. In addition, to make sure that the aflatoxin and pathway intermediates would not block the filter membrane, the filtered water and only centrifuge-treated water (without passing through filter) were examined under a microscope and analyzed by TLC analysis in parallel. No spores were observed in the filtered water, but a small quantity of spores were observed in the centrifuge-treated water. In both treatments, AF 8,, AF G, and OMST were all detected in harvest water by TLC analysis (Figure 5.3). 119 Binding of AFB,, AFG,, OMST, and VA in harvest water by spores. Spores isolated from the non-aflatoxin producing strain RHNI (accumulates only OMST) were used to study the binding of AFB, and AFG, in harvest water. Spores isolated from aflatoxin-producing strain SU-l (accumulates only AFB, and G,) were used to study the binding of OMST and VA in harvest water. The same numbers of spores (2 x 107) were used in the binding assay and for extraction of aflatoxin and OMST for comparison. The RHNI spores were mixed with harvest water (1 ml) containing AFB, and G, (saved from a SU-l spore preparation) for 10 min and then span down at 20,000 x g for 15 min. The spores were washed once with 1 ml of pure water and span down again (same conditions). Spores were finally resuspend in 1 ml of water and extracted with 0.5 m1 of chloroform. 10 pl of spore extracts was analyzed in parallel with other extracts, including the harvest water, spores before binding, and toxin extracted from a whole colony, by TLC. The same procedure was applied to the binding of OMST and VA from harvest water by SU-I spores. Spores harvested by cotton-tipped applicators. A method to harvest spores under dry conditions was developed. The colony surface was gently swabbed with cotton- tipped applicators and the spores attached to the applicator were then released into water by stirring the applicator in the water. A new applicator was used each time to touch the colony surface. In this dry-harvested method, water was never directly in contact with the colony during spore harvest. The spore-containing water (total 8 ml) was then filtered through glass wool. Spores were pelleted by centrifugation as described above. The water was saved for future aflatoxin measurement. The spore pellet was dried under vacuum with heat (60 °C) for 4 h. The AFB, concentrations were determined by ELISA. 120 RESULTS Aflatoxin distribution in time-dependent fractionated colonies of SUI grown on solid medium. The total aflatoxin produced by a colony increased with time during the first 4 days, declined at d 5, and then increased at d 6 and 7 (Table 5.2). The total aflatoxin in the mycelium also reached a peak at d 4 and then decreased at d 5 (Table 5.2). A similar pattern was observed in the agar (Table 5.2). These results suggest that, for unknown reasons, the detectable aflatoxin decreased between d 4 and 5. The ratio of aflatoxin in the mycelium to total AF B1 tended to decline with time until (1 5 (Table 5.2) and then increased at d 6 (40 and 46 %) and d 7 (61%). By monitoring dry weight in the same fraction isolated from different age of colonies, the results suggest that the firngus is not only growing outward (two-dimensions) but also growing upward (three dimensions) with increased time (Table 5.3). As shown by dry weights measured in Ring-2, the regular rate of growth is 0.12 g per (1 (Table 5.3). Approximately 10 pg of aflatoxin were detected in the Ring 2 (0-24 h-old) of a 72 h-old colony. Over the next 24 h (24-48 h-old), aflatoxin in Ring 2 rapidly accumulated (6-fold increase in concentration) and the accumulation then declined (Table 5.4). The aflatoxin concentrations in each fraction isolated from two 6-d colonies varied even they were cultured at the same time (Table 5.4 and 5.5). This variation may be explained by the observation that a part of the edge of colony-No. 2 made contact with the Petri dish. That contact may affect a normal fungal growth and the normal physiology of the fungal cells. Similarly for 7-d old colony, the colony margin reached the edge of plate, which appeared to result in an increase in aflatoxin content. 121 Table 5.2. Total aflatoxin detected in a colony, agar and membrane from different aged colonies. Day 0‘ sample Membrane Agar Mycelium Total Mycelium ‘"‘° (Its) (Iug) (118) (118) / Agar /T0ta| Colony No. C/o) YES 2 d-l 0.1 8 8 16 100 50 2 0.1 7 10 17 140 59 3 1 10 16 27 160 59 3 d-l 2 63 94 159 150 59 2 6 98 87 191 90 46 4 d-l 2 155 143 300 90 48 2 9 191 133 333 70 40 5 d-l 2 138 77 217 60 35 2 7 153 80 240 60 33 6 d 11 286 218 515 80 42 7 d 14 257 420 691 160 61 PDA 7 d-l 19 301 86 406 30 21 2 25 460 67 552 10 12 Table 5.3. Dry weight (g) of fractionated mycelium from different aged colonies. Age/Fraction Ring 1 Ring 2 Ring 3 Ring 4 Ring 5 Total L.__E__8__fl__fi___£_, YES 2d ,,,,, 0.128 __ 0.128 3 d , 0.200 0.135 0.335 4 d 0.238 0.250 0.216 0.704 W , ~6de 0.407 0.502 0.438 9548 03272222 -2w0306 0.369 0.420 0.55 1 l .273 2.919 7 d 3.107 PDA 7 d-l 1.263 -2 1.328 122 Table 5.4. Total aflatoxin (pg) in fractionated mycelium from different aged colonies. Age/Fraction Ring 1 Ring 2 Ring 3 Ring 4 Ring 5 118 pg pg $8 118 2 d 16 ‘_ 3 d 1 75 12 5 d . 35 29 13 3 6 d-l 47 60 37 19 4 -2 54 87 57 59 12 Table 5.5. Normalized aflatoxin levels (pg of aflatoxin/g dry weight mycelium) in fractionated mycelium from different aged colonies. Age/Fraction Ring 1 Ring 2 Ring 3 Ring 4 Ring 5 Total/w.t Its/g rig/g rig/g Its/g Its/g JIB/g YES 2 d 125 125 3 d 375 .89.. __.____ - 260 4 d 193 296 60 189 5 d 147- 78 ._ 43 14 ..._...70..__. 6d-1 115 120 84 35 12 75 -2 176 236 136 107 .9..- .. 92 _ 7 days 135 PDA 7 d-l 68 -2 50 123 Binding of AFB, and pathway intermediates in water by spores. Spores harvested under wet conditions (by water) contained 10-fold higher AFB, concentration (1350 rig/mg) than under dry conditions (by cotton-tipped applicator) (average 114 rig/mg) (Table 5.6). Based on TLC analysis, the explanation for the higher concentration in wet-harvested spores is that they absorb aflatoxin from water used in spore harvest- this presumably arises from the mycelium. This is supported by the binding assay. In this experiment, SUI spores were demonstrated to bind OMST from water used to harvest RHN-1 spores and RHN-I spores bound AFB, and AF G, from water used to harvest SUI spores. Before the binding assay, TLC analysis showed that only AFB, and G, were detected in extracts of SUI spores and water used to harvest SU-l spores, and only OMST was detected in extracts of RHNI spores and water used to harvest RHN-1 spores (Figure 5.3 and Fig. 5 .4, lane 3 and 4). After binding, AF B,, G,, and OMST were detected in an extract of RHNI spores (Figure 5.4, lane 5 and 6); and were also detected in extracts of SUI spores (Figure 5.6, lane 4). The AFB, and AF G, concentrations detected on RHNI spores were very similar to the amount detected on SU-l spores suggesting efficient binding (Figure 5 .4). VA was not detected in extracts of ATCC36537 spores and water used to harvest spores but was detected in extracts of the whole colony (Figure 5.5). Because VA was not detected in the water used to harvest spore, it is not surprising that it was also not detected in extracts of SU-l and RHN 1 spores after the binding assay (Figure 5.5 and Figure 5.6, lane 2, 3, 6 and 7). Possible explanations for the lack of detectable VA in water used to harvest spores are the solubility of VA in water is very low or the quantity of VA secreted on the colony surface is very low. 124 Table 5.6. Aflatoxin detected in conidiospores isolated under wet and dry conditions from A. parasiticus grown on YES and PDA agar. Spore Water Wt". (mg) Total (ng) Conc. ° Vol. ° (ml) Total (ng) Conc. (ng/ mg) (118/ m1) Dry' YES 3 .0 l 56 52 1 .3 226 l 74 8 194 24 PDA 1 .0 1 3 5 1 3 5 8 72 9 Wet PDA 10.0 16215 1622 1 0.0 1 4 l 77 141 8 9.9 11621 1174 50 63000 1260 8.5 10140 1193 50 42400 848 a: Water used in dry spore-harvest did not directly contact the colony. b: Dry weight. c: Concentration. d: Volume. 125 1234567 U "I " 8 9 . 4 AFB] + OMST AFG1 + Figure 5.3. Thin layer chromatography of aflatoxin and OMST in water used for spore harvest. Lane 1. AFB,, B2, G1, G2 standards. Lane 2. Water used to harvest SUI spores (centrifuged to remove spores). Lane 3. Water used to harvest SUI spores (0.22 um filtrate). Lane 4. Water used to harvest SUI spores (0.45 um filtrate). Lane 5. Water used to harvest RHNI spores (centrifuged to remove spores). Lane 6. Water used to harvest RHNl spores (0.22 um filtrate). Lane 7. Water used to harvest RHNI spores (0.45 um filtrate). Lane 8. Fluid (brown color) collected from the surface of RHNI colony. Lane 9. OMST standard. 126 123456789 a OMST AFB] , AFG1 + Figure 5.4. Thin layer chromatography of RHNI spore extracts. Lane 1. SUI whole colony extract. Lane 2. AFB1,B2,GI and G2 standards. Lane 3. SUI spore extract. Lane 4. RHN 1 spore extract. Lane 5. RHN] spores mixed with water used to harvest SUI spores (0.22 urn filtrate). Lane 6. RHN 1 spores mixed with water used to harvest SUI spores (0.45 um filtrate). Lane 7. Fluid (brown color) collected fi'om the surface of RHNI colony. Lane 8. OMST standard. Lane 9. RHN] whole colony extract. 127 STD 1 2 3 4 5 VA AFB, -> AFG, -> Figure 5.5. Thin layer chromatography of SUI and ATCC36537 spore extracts. Lane 1. SUI spore extract. Lane 2. ATCC 36537 spore extract. Lane 3. ATCC 36537 spore extract. Lane 4. ATCC 36537 whole colony extract. Lane 5. VA standard. 128 123456789 7 VA OMST AFB, AFG1-P " Figure 5.6. Thin layer chromatography of SUI and RHN] spore extracts after binding with water used to harvest ATCC36537 spores. Lane 1. AFB1, B2, G1 and G2 standards. Lane 2. Extract of SUI spores mixed with water used to harvest ATCC36537 spores (0.22 pm filtrate). Lane 3. Extract of ATCC36537 spores. Lane 4. Extract of SUI spores mixed with water used to harvest RHNI spores (0.22 pm filtrate). Lane 5. OMST standard. Lane 6. Extract of RHNI spores mixed with water used to harvest ATCC36537 spores (0.22 pm filtrate). Lane 7. Water used to harvest ATCC36537 spores (0. 22 pm filtrate). Lane 8. VA standard. Lane 9. SUI whole colony extract. 129 AFB, concentrations in spores harvested under dry condition. The AFB, concentrations in spores were also measured in spores harvested under dry conditions (Table 5.6). Spores harvested from two cultures (PDA and YES) contained concentrations from 50 to 174 ng of aflatoxin per mg of spores, which is about 10 fold lower than the results using a wet harvest method. Only 73 to 194 ng of aflatoxin was detected in water used to release spores from cotton tips of applicators. DISCUSSION In preliminary experiments, Aspergillus was cultured on nylon membranes on which the mycelia was firmly attached; the bottom portion of the mycelia could not be completely recovered for analysis and, therefore, this caused difficulty in estimating the aflatoxin concentration in the colony. This problem was solved by replacing the nylon membrane with a cellophane membrane; the mycelia could be fractionated and completely removed from the cellophane membrane The colony fractionation results showed that, from day 2 to day 5, the ratio of aflatoxin in the fungal mycelium to total aflatoxin produced by the colony decreased. However, the ratio then increased at day 6 and 7. One possible explanation is that the colony edge made contact with the Petri dish at day 6. This possibly could be a signal to the fungus that the space and nutrients are limited, resulting in a change in the normal physiological conditions to promote aflatoxin production. Another possible explanation is a pH change in the medium. As reported by Mellon et al., the pH in the medium dropped to be its lowest value at day 5 (under their culture conditions), possibly stimulating higher levels of aflatoxin synthesis in the mycelium (3). 130 The media used was observed to affect the distribution of aflatoxin within and outside of the fiingal colony. Whereas 80% of the aflatoxin was released into the medium when cultured on PDA agar for 7 days, more than 50% of the aflatoxin was found in the mycelium cultured on YES agar. The reason for this is not clear but may be related to the level of sporulation. I can observe two major differences in colony morphology on PDA: more green spores and less mycelial mass. Although it has been previously hypothesized that aflatoxin synthesis and asexual sporulation are tightly associated (2,4,6), no estimation was conducted on the relative distribution of aflatoxin in the fungal mass and the culture medium. Researchers previously measured aflatoxin in the medium only or in whole culture extracts (mycelium and medium). I hypothesize that, since spores are important structures for dispersal and colonization, if aflatoxin is harmful (carcinogen) to the fungus, then the ability to reduce aflatoxin concentration inside of fimgal mycelium prior to and during spore production could be necessary to prevent mistakes in DNA replication caused by DNA adduct formation. Therefore, a related hypothesis is that the substrate mycelium is still highly active in aflatoxin production but the toxin is secreted outside of the mycelium to prevent toxin accumulation. This mechanism could enhance the precise production of spores by conidiophores. It is obvious that spores harvested under dry conditions have a much lower aflatoxin concentration than those collected under wet conditions (with water). The water used to harvest spores contains high aflatoxin concentrations and is responsible for the higher aflatoxin detected in spores as compared to mycelium. Under dry harvest conditions, the water used to release the spores from the cotton-tipped applicator contained very low but detectable amounts of aflatoxin. The traces of aflatoxin in this water may come fi'om the fluid on the colony surface, absorbed by applicator and released 131 into water. In support of this notion the brown-colored liquid collected from the colony surface of RHNI contained significant levels of OMST (Figure 4, lane 7). We believe this fluid on the colony surface was secreted by fungal cells. It has been demonstrated previously that the variation in spore aflatoxin concentration between strains was high and was affected by the culture conditions and spore-harvest methods (4,6). Wicklow and Shotwell (1982) reported that condiospores harvested from Czapek agar cultures (in the dark at 28° C for 21 days) using water contained a range of 1,106 to 54,300 ppb (ng per g) of AFB, in four A. parasiticus strains and a range of 36 to 97,400 ppb in five A. flavus strains. Palmgren and Lee (1986) reported that the condia harvested from A. parasiticus (SRRC-2004) cultured on autoclaved rice and incubated at 25°C for 7 days contained 161,000 ppb (161 ng per mg) of AF B,. In this report, the spores were harvested by vacuum (dry conditions). Based on our data, we suspect that the real aflatoxin concentration in spores in Wicklow (1982) was likely 10-fold lower than that reported because Czapek agar media is not a good aflatoxin-inducing media. In addition, the method used to harvest the spores (similar to the wet condition reported in our study), caused an artificially high concentration of aflatoxin detected in spores; the aflatoxin extracted from the colony bound to the spore surface. The results reported by Palmgren (4) were similar to ours (proximately 126 ng per mg); we believe these data generated using a dry-harvest method more closely represent the actual concentration in spore. My current data show that both fungal hyphae and spores isolated from aflatoxin- producing strains contain aflatoxin. However, for a nar-I mutant, spores (harvested by vacuum) contained very low amount of NA (280 ng per g of spores) as compared with mycelia (760 ug per g) (4). This observation prompts us to propose that aflatoxin is 132 synthesized in the mycelium and then needs to be transported to spores. In this regard, it will be interesting to determine if all aflatoxin intermediates or only certain pathway- intermediates can be accumulated in spores. 133 REFERENCES 1. 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. 2. Guzman-de-Pena D and J. Ruiz-Herrera. 1997. Relationship between aflatoxin biosynthesis and sporulation in Aspergillus parasiticus. Fungal Genet Biol. 21:198-205. 3. Mellon, J. E., P. J. Cotty, and M. K. Dowd. 2000. Influence of lipids with and without other cottonseed reserve materials on aflatoxin B, production by Aspergillus flavus. J. Agric. Food Chem. 48:3611-3615. 4. Palmgren MS. and LS. Lee. 1986. Separation of mycotoxin-containg sources in grain dust and determination of threir mycotoxin potential. 66:105-108. 5. Pestka, J. J., P. K. Gaur, and F. S. Chu. 1980. Quantitation of aflatoxin BI and aflatoxin Bl antibody by an enzyme-linked immunosorbent microassay. Appl. Environ. Microbiol. 40: 1027-1031 . 6. Wicklow, D. T., and 0. L. Shotwell. I983. Intrafungal distribution of aflatoxins among conidia and sclerotia of Aspergillus flavus and Aspergillus parasiticus. Can. J. Microbiol. 29:1-5. 134 CHAPTER 6 Future Studies Based on the current data, it is difficult to determine if the vacuolar localization of OmtA occurs for protein recycling, for enzyme activation, or for some other purpose in A. parasiticus. The compartment for storage of aflatoxin B, in the fungus was also not determined in this study. We propose three alternative models for synthesis and export of aflatoxin from fungal cells that are consistent with data from this and previous studies. We propose these models as a primary means to design future research directions (Figure 6.1). Model I. Aflatoxin enzymes aggregate in large complexes in the cytoplasm to carry out sequential reactions efficiently. Aflatoxins are synthesized in the cytoplasm and are transported to the extra-cellular environment via AflT (with identity to a transproter protein), via another transporter, or via direct release into the extra-cellular environment when cells die. To fine tune the level of (reduce or limit) aflatoxin synthesis, aflatoxin enzymes (OmtA in particular) are transported to and inactivated in vacuoles by proteolytic cleavage. Model 2. Aflatoxin enzymes form complexes in the cytoplasm to generate the late pathway intermediate sterigmatocystin (ST); the intermediates in the pathway upstream from ST are passed directly from one active site to the next in the enzyme complex limiting the quantity of free intermediates in the cytoplasm. OmtA is synthesized and participates as one enzyme in the complex but it is present in inactive form. As ST bound to OmtA accumulates in the complexes, OmtA carries ST into 135 vacuoles in cells near the basal surface of the colony. Proteolytic cleavage activates OmtA that then converts ST to OMST. OrdA, also localized to the vacuole, converts OMST to AFB, that accumulates in the vacuole. As it fills, the vacuole fuses with the cell membrane and releases AFB, into the external environment at the basal surface (substrate surface) of the fungal colony. Purified native OmtA has a molecular weight about 40 kDa. Based on the nucleotide sequence of the cDNA, OmtA has a predicted molecular mass of 45 kDa suggesting that a 5 kDa peptide is cleaved from its N-terminus. This cleavage could inactivate OmtA as proposed in Model 1 or activate OmtA as proposed in Model 2. In support of this idea, Western blot analysis demonstrated that OmtA is cleaved in fungal cells and the first cleavage releases a peptide of the predicted size. Model 3. This model is similar to Model 2 except OmtA is firnctional in the cytoplasm. OmtA and its product, OMST, are transported together to the vacuole where OMST is converted to AFB, by OrdA. Three experiments will be conducted in the future to allow discrimination between these models: 1) co-immunoprecipitation, 2) subcellular fractionation, 3) in situ localization of OrdA by immunoelectron microscopy (IEM). Results from Experiment 1 will answer the question whether aflatoxin proteins can form protein aggregates to carry out several reactions together. Results from Experiment 2 and 3 will answer the question regarding where OrdA (60 kDa) and aflatoxins are compartmentalized. The purified organelles will be used to test their ability to convert OMST to AFB, in feeding experiments; they will also be freeze-substituted and immunolabeled to confirm protein localization. The protein content extracted from isolated organelles will be assayed to determine their aflatoxin protein profile using Western blot analysis. To assist in these experiments, the required OrdA antibody was generated using similar procedure as 136 described in Chapter 3. The specificity of this antibody was tested by Western blot analysis, and the results are shown in Figure 6.2. In addition, a MBP-AflJ fusion protein and AflJ antibody were also produced (Figure 6.3). These reagents will be used in Experiment 1 to test the possibility that the protein-protein interaction between AM (47 kDa) and AflR (and other transcription factors) or proteins involved in microbody targeting pathway. The primers used in PCR to clone ordA cDNA were: 5’- ATG ATT TCTAGA ATA ATT ATT TGT GCG G-3’ and 5’- GTA AAQQTT TCA AAT CAT CTG ATT-3’. The primers used to clone aflJ cDNA were: 5’- TCC TCTAGA ATG ACC TTG ACT GAC CTA G-3’ and 5’- CAT AAGQTT TTA ATA TCG GTT GTC ATC G -3’. (XbaI restriction sequence: TCT AGA; HindIII: AAG CTT). 137 Model 1 Model 2 Cytoplasm fl Medium Vacuole ’ Medium Figure 6.1. Proposed models for synthesis and export of aflatoxin from firngal cells. 138 123456 7 89STDkDa 120 100 so 60 50 40 30 OrdA ——> “*‘ "" ‘ “’ 20 Figure 6.2. Western blot analysis of fungal protein extracts using affinity-purified OrdA PAb. Proteins were extracted from time-fractionated colonies of SU-l and AF SIO grown on YES agar medium. 60 pg of protein was loaded in each lane on SDS-PAGE. OrdA antibody (#137) was purified by a column packed with AF S10 (aflR knockout) protein. OrdA antibody was used at a concentration of 6 pg per ml in the experiment. Lane 1, SI fraction; 2, SZ fraction; 3, S3 fraction; 4, R1 fraction; 5, R2 fraction; 6, R3 fraction; 7, 32cm fraction; 8, S fraction; 9, S6,m fraction. Lane STD contains molecular mass 4cm standards; mass of standards is shown at the right side of blot. SW, SW, and S6cm are fractions isolated from SU-l colonies grown for 90-h and were fractionated according to their diameters. S covers from center to 2cm; Sm“, 2 to 4 cm; SW, 4 to 6 cm. 2cm 139 Figure 6.3. Production of MBP-AflJ fusion protein and AflJ antibody. (A) SDS-PAGE of IPTG- induced bacterial crude extracts containing MBP-AflJ. Lane I, Non-induced bacterial crude extract; 2, IPTG- induced bacterial crude extract; 3, Molecular mass standards; molecular mass is indicated to the right of the gel. (B) SDS-PAGE of affinity purified MBP-AflJ. Lanes 1, Affinity-purified MBP-AflJ; 2, Molecular mass standards; molecular mass is indicated to the right of the gel. (C) Western blot analysis of fungal protein extracts using AflJ PAb (#124). Proteins were extracted from time-fractionated colonies of SU-l and AF SIO grown on YES agar medium. 60 pg of protein was loaded in each lane on SDS-PAGE. AflJ antibody was used at l to 1000 dilution in the experiment. Lane 1, Molecular mass standards; mass of standards is shown at the left side of blot; 2, SI fraction; 3, 82 fraction; 4, S3 fraction; 5, R1 fraction; 6, R2 fraction; 7, R3 fraction. 140 kDal \llllll \ Figure 6.2 1 2 kDa +8." 97 MBP-ADJ H 66 .1 45 d 31 5 6 7 3......“ <— AflJ 141