fig: _ .. 4...... a .mi 5 13. 4.3.1... 2‘ :1: 3.... 5: :5 ‘ , . . ‘ .2 ‘ x ‘ n‘. 9“ . ‘ . ‘ 3.4.? "um“ trig... 5..th . do an, ‘ .z m. y a v: . S . ’33. .551... K . . .y . . V in: .x . . 2 :53. . . ‘ , 31153. £35.: ‘ , . a . . . . a. tv .5. ‘ . . . , 35:12:»?ith «7hr? , . , . s: I \ —————> 000' 2 Cl / '00C '000 Cl (:1 Scheme 1. Reaction catalyzed by deA. In this introduction, 1 first compare the amino acid sequences of members of the oc-KG-dependent dioxygenase superfamily to indicate relationships and differences between these enzymes. Next, I summarize previous research on deA and other biochemically characterized a-KG-dependent dioxygenases. Finally, I review studies on the occurrence of the tfdA gene, which encodes deA, in various 2,4-D degrading soil bacterial isolates and communities. The introductory chapter will conclude with an overview of the research presented in this dissertation. oc-KG-dependent dioxygenases are common in plants, animals, and the majority of aerobic microorganisms. Furthermore, data from recent genome sequence projects 1‘.) suggest their presence in nematodes and insects (y-butyrobetaine hydroxylase in Caenorhabditis elegans, Genbank # 266523 and Drosophila melanogaster, Genbank # AL109630, respectively). Since 1967, when prolyl hydroxylase was characterized as the first a-KG-dependent dioxygenase (45), this class of enzymes has come to include more than forty members (Table 1). The reactions catalyzed by these enzymes range from steps in the synthesis of hormones (36, 91), antibiotics (58), collagen (54), and compatible solutes (30, 44) to the decomposition of pyrimidines (1 1 l, 115), amino acids (21) and other small molecules (21, 27, 40). In light of the diversity of substrates for a-KG dependent dioxygenases and the different types of oxidative reactions they catalyze, it is not surprising that their sequences exhibit little overall similarity. Amino acid sequence analysis of a-KG—dependent dioxygenases Although there is no overall sequence identity uniting all enzymes in the a-KG- dependent dioxygenase superfamily, comparison of their primary structures finds subgroups of related proteins. More than twenty distinct enzymes with significant levels of sequence identity to isopenicillin-N—synthase (IPNS) (Table 1) have been described (8, 58). The IPNS-like enzymes (Group I) include the majority of characterized oc-KG- dependent dioxygenases. A second cluster of enzymes related by sequence (Group H) is comprised of deA, four previously characterized oc-KG-dependent dioxygenases, and four putative enzymes yet to be studied. The Group II enzymes do not exhibit significant Table 1. Characterized enzymes in the a—KG—dependent dioxygenase superfamily. Enzyme Source for purified References enzyme Group I IPNS-like enzymes Iscmenecillin N synthase (IPNS)a Fungi, Bacteria (93) Deacetoxycephalosporin C synthetase (DAOCS) Fungi, Bacteria (56) Deacetylcephalosporin C synthase Fungi, Bacteria (102) Aminocyclopropane carboxylate (ACC) oxidasea Plant (135) Ethylene-forming enzyme Bacteria (83) Deacetoxyvindoline 4-hydrmglase Plant (14) Hyocyamine-6S—hydroxylase Plant (69) F lavone synthase I Plant (9) Flavanol synthase Plant (43) F lavanone 3-hydroxylase Plant (10) Leucoanthocyanidin hydroxylase Plant (99) Gibberellin-ZB-hydroxylase Plant (37) for review Gibberellin 3B-hydroxylase Plant " Gibberellin-7-oxidase Plant " Gibberellin l3-oxidase Plant " Gibberellin-ZO-oxidase Plant " Procollagen Prolyl-4—Hydroxylase Mammal, Bird, (54) for review, (23) Virus Procollagen Aspartyl Hydroxylase Mammal (55, 129) Procollagen Lysyl Hydroxylase Mammal (74, 81, 89, 125) Group II deA-like enzymes deA Bacteria (27) Taurine/oc-KG dioxygenase (TauD) Bacteria (21) Sulfonate/a—KG dioxygenase Fungi (40) Clavaminate synthase (CS) Bacteria (11) y-Butyrobetaine hydroxylase Bacteria, Mammal (30, 62) Group III Proline-4-Hydroxylase Bacteria (1 O6) Phytanoyl-COA Hydroxylase Mammal (46) Miscellaneous Thymine hydroxylaseb Fun& (1 15) Pyrimidine-deoxynucleoside- 1 dioxygnaseb Fungi (1 l 1 ) Trirnethyl-lysine dioxygenaseb Mammal (44) Proline 3-hydroxylase° Bacteria (77) a Enzyme does not use a—KG as a cosubstrate bNo sequence is available for this enzyme ° Enzyme does not have any sequence identity to other characterized enzymes similarity to IPNS and other Group I enzymes. Group IH contains four enzymes that, despite their very disparate substrates, have obvious sequence relatedness. Future studies on the structure and activity of enzymes in this class may uncover a relationship between Group III and either of the other two groups of a-KG-dependent dioxygenases; at present, however, sequence comparisons predict that they form a distinct cluster. Table 2 summarizes the elements common to the enzymes in each group. Finally, a fourth group contains or-KG-dependent dioxygenases that either exhibit no sequence similarity to enzymes in the other three classes or their amino acid sequence is not yet published. IPNS-like a—KG-dependent dioxygenases (Group I) IPNS is an Fe(II)—dependent enzyme that converts L-oc-amino-B-adipoyl-L- cysteinyl-D-valine (ACV) to isopenicillin N (Scheme 2) with chemistry that is similar to the mechanism of a-KG-dependent dioxygenases (although a—KG is not used as a co- substrate). IPNS has between 25 and 35% amino acid identity to over twenty enzymes involved in antibiotic synthesis (58) and the formation of plant hormones (91). For comparison, sequences for the same enzyme from different plant genera have, on average, more than 60% amino acid identity. Alignment of Group I sequences indicates that a His-X—Asp-Xss-His motif is found in all of these enzymes (8) (Table 2 A). Crystal structures of IPNS (94) and the related enzyme deacetoxycephalosporin C synthase (DAOCS) (122) supported biophysical studies that predicted these residues to be ligands to the iron metallocenter. These same structures, combined with further sequence comparisons (58, 63), reveal a conserved Arg-X-Ser arrangement involved in oc-KG binding in Group I enzymes. P4 Lys B. Group H H D x DN H x R deA WHSDSSFQ 147 WNVGDLVMWDNRCVLH 10 R TauD WHTDVTFI 153 WQPNDIAIWDNRVTQH 10 R YSD WHTDVVSY 147 WEPHSVVIWDNRRVQH 11 R AtsK WHTDVTFV 134 WEAGDVAIWDNRATQH 10 R Pch LHWDGMYL 147 WRSDDLVIADNLTLLH 10 R CarC FHADGGLL 148 WEDGDLLIMDNRRVIH 11 R SyrI YHNESSHL 192 WRKGDVVMLDNMLAAH 8 R y-BBH LHTDLPTR 139 LEAGQLWCFDNRRVLH 11 R cs FHTEMAYH 123 LEPGDLLIVDNFRTTH 11 K x R C. III H N R HGSGTNMSPHPR H TIRR HGSGQNKTQGFR RPR Table 2. Conserved amino acids among subgroups within the a-KG-dependent dioxygenase superfamily. Amino acids in common among the IPNS-like a-KG—dependent dioxygenases (A), the deA-like enzymes (B) and the proline 4-hydroxylase group (C) in the a—KG—dependent dioxygenase superfamily are shown. The numbers in columns labeled “X” represent the number of amino acids intervening between the conserved regions. The unique identifiers for the protein sequences in Group I are IPNS, isopenicillin N synthase (IPNS_STRJU); DAOCS, Deacetoxycephalosporin synthase (pir2:A39204); FL3H, F lavanone 3- hydroxylase (SW:FL3H_HORVU); VanB, vanillate biosynthesis enzyme B (AF 009673); P4H, human prolyl 4-hydroxylase (P13674); LysH, human lysyl hydroxylase (Q02809). The sequences in Group H are deA, 2,4-D/oc-KG dioxygenase (M16730); TauD, taurine/a-KG dioxygenase (P37610); YSD, S. cerevisae sulfonate/oc-KG dioxygenase (Q12358); AtsK, putative oc-KG dioxygenase from sulfate-ester transport gene cluster (AF 126201) Pch, pyoverdine biosynthesis enzyme (AF 002222); CarC, carbapenem biosynthesis enzyme C (AAD38231); SyrI, syringomycin biosynthesis enzyme I (AAB63253); y—BBH, y-butyrobetaine hydroxylase (P80193); CS, clavaminate synthase (A44241). Group III sequences are Pro4H, proline 4—hydroxylase (D78338); HtxA, putative hypophosphite/oc—KG dioxygenase (AF023463); PPHYH, Phytanoyl-CoA hydroxylase (NP_006205); MmcH, mitomycin C biosynthesis protein (AF 127374). Several enzymes involved in the hydroxylation of peptidyl amino acids are grouped together by activity and placed into Group I based on the presence of the above motif. Although these key active site residues are maintained in prolyl-4—hydroxylase (15, 23), lysyl hydroxylase (74, 81, 125), and aspartyl hydroxylase (47), their sequences lack detectable levels of similarity to other Group I enzymes. The importance of the amino acids in the above motif for activity has been confirmed by mutagenesis studies (described below). 4—Hydroxyphenylpyruvate dioxygenase (4-I-IPPD) is an Fe(H)- dependent enzyme that catalyzes ring hydroxylation concomitant with decarboxylation of the a-keto acid side chain. Mechanistic similarities in addition to seeming conservation of the above motif prompted inclusion of 4-I-IPPD with Group I enzymes of or-KG dependent dioxygenase superfamily. Upon introduction of less related 4-I-IPPD sequences into alignments, the H—X—D motif was no longer absolutely conserved, casting doubt on the relationship of this enzyme to the a-KG-deperrdent dioxygenases. A recently released crystal structure of 4-HPPD indicates a protein fold more similar to that of extradiol (ring-cleaving) dioxygenases than H’NS and DAOCS. Enzymes with similarity to T fdA (Group 11) As recently as 1995, no enzymes with significant sequence similarity to deA were known. Over the past five years, however, multiple deA homologs have been identified due to the dramatic increase in genome sequence data and the development of more powerful sequence analysis tools, such as BLASTP (l) and Position-Specific Iterative (PSD-BLAST (2). Now, it is apparent that deA is related to four characterized it. 33%»on 0M8 ozgza .M3< Ammmmomafiwv H ego £35585 onaowfibnm dram AwmeOv 035?»on OMébfisobsm 89.38% a dm> 54$}; 035% 321550 .50 51$qu N 035% 33833 .mmu mammgomac 2:50 assesses 0592/93 “mgm ”AmmmooNv .3633 mtfiecfozmcb Sow 3258623 05830559-.» .mU 2mm.» Amaflownb .Qm acmofionzmwm 80m 033x953 0380905591» .Emm-> 8m 25% wow: moocoscom 2F .oocozcom beam 05 mm pom: mm 0535 £033 no mcmwcoaow bfiwzm await 32on msowofioEo: (Ho fiwcfl 95 $53qu “scuba BC. .ozumafiofi some 5 68mm 05 mm “dogma Gauge as masses 8% item? 2:. imbued Hmaqfimzmm Ho max—gm mHmo_ m3“ 8:253 05 £0.53 H26 8:253 (Ho Ewan: 06. use @385: “$5 3:52 338m 50m .8359»on Eucaoaocrwvfid gage: gaming—Eco 8:253 Eon 2.33. .m 3.38 - Sufi exam - 95 8:2 was: - use 05% figs. - v3.4 3Q: 3Q: 22: 48mm - 5% mega an: :. S: 3qu 892 - a? 2:2 ER 3 :8 ~32. 8% ~23 - Go a :3 53 was: 2&2 BEN ow S: 03; - mmo oREm KEN ENE meson «.32 N3: EE gem - maa 3N5 8:8 ”RE 8%: 8&2 SS: ER 392 8min - mmms ER: 9&2 ~82 SQ: KERN KEN % Sm % SN ow :8 am: m - mo mmms $5 95 use Me< 85 new 50 $0 mgm mmms mo mmms oc-KG-dependent dioxygenases and to sequences of several putative a-KG-dependent dioxygenases. Amino acid sequence identities from pairwise comparisons range from fifteen to forty percent (Table 3). deA is most closely related to taurine/OL-KG dioxygenase (TauD) (21), Saccharomyces cerevisiae sulfonate/oc-KG dioxygenase (YSD) (40) and three putative Ot- KG—dependent dioxygenases: Pch, involved in synthesis of the siderophore pyoverdine (109), CarC, which is necessary for formation of the B—lactamase inhibitor carbapenem (72), and AtsK, a putative dioxygenase in a sulfate-ester degradation gene cluster in Pseudomonas putida (127). Although they have not been biochemically characterized, the roles of Pch, CarC and AtsK were identified through various transposon mutagenesis studies. The relationships between these six sequences are detected by a standard BLASTP analysis. More sensitive analyses with PSI-BLAST find relationships between deA and more distantly related a-KG-dependent dioxygenases. PSI-BLAST detects lower levels of relatedness by creating a position-specific matrix of conserved amino acids to rescore and realign sequences that receive low BLAST scores. As this process is repeated, unrelated sequences with chance high scores are downgraded as the matrix of positionally-conserved amino acids improves. By these methods, relationships between deA and two other well-characterized oc-KG-dependent dioxygenases, y- butyrobetaine hydroxylase and clavaminate synthase (Table 3), become evident. In addition, the sequence of a third enzyme, SyrB, which is required for synthesis of syringomycin, also aligns with the deA-like sequences. Of the twenty nine sequences identified as relatives of deA after six iterations of PSI—BLAST, all were either previously characterized oc-KG dependent dioxygenases or hypothetical proteins with the characteristic conserved amino acids described below. Comparison of all nine Group H a-KG-dependent dioxygenase sequences finds several strictly conserved residues (Table 2B, Appendix A). As in the Group I enzymes, an H-X-D-like motif is present in all sequences, although in both clavaminate synthase and SyrB, the aspartate is replaced by a chemically-equivalent glutamate residue. A stretch of amino acids near the C-terminus with high levels of similarity across these sequences, contains the only other invariant histidine, suggesting that this residue may be analogous to the third metal ligand in the Group I motif described above. Conserved asparagine and arginine residues near this histidine are also apparent and the latter may play a role similar to the a-KG-binding arginine in IPNS and DAOCS. Group III a-KG-dependent dioxygenases A third group within the oc—KG-dependent dioxygenase superfamily contains two characterized enzymes, proline 4-hydroxylase (73) and phytanoyl-CoA hydroxylase (46), and two putative a-KG-dioxygenases, hypophosphite oxidase (HtxA) (73) and MmcH which is involved in mitomycin C biosynthesis (67). Although no homology is detected between these sequences and either Group I or Group II or-KG-dependent dioxygenases by BLAST comparisons, the motif common to this group is clearly similar to those described above (Table 2C). Again, there is a conserved H-X-D motif and two additional invariant histidines in this small set of sequences (Appendix A) and in uncharacterized homologs in other bacteria (120). A His-Xlo-R motif (similar to Group I and Group H enzymes) identifies a good candidate for the third metal ligand; however, biochemical 10 studies or structural information is needed to establish the roles of these residues. In humans, non-functional phytanoyl-CoA hydroxylase results in Refsum disease due to the inability to break down phytanic acid (3,7,11,15-tetramethylhexadecanoic acid), a degradation product of chlorophyll (12). It is interesting to note that two of the disease- causing alleles of phytanoyl-CoA hydroxylase, which encode inactive enzyme, have mutations of either the conserved arginine or an invariant asparagine nearby. Summary of a-KG-dependent dioxygenase sequence analyses Sequence comparisons of members of the or-KG-dependent dioxygenase superfamily find three distinct subgroups of related enzymes; no overall sequence identity was detected between enzymes from different subgroups. Comparison of more closely related sequences allows for the identification of potentially important, conserved residues. The consensus motifs for the three subgroups are similar (an H-X-D motif distant from an H-Xlo-R arrangement) and likely identify residues important for active site structure. As the number of oc-KG-dependent dioxygenase sequences grows, the three groups may merge into a continuum of sequences. Alternatively, as other sequences are found, distinctions between the groups may be maintained. Differences in the arrangement of the conserved residues across the superfamily may be indicative of the mechanisms that drove the divergent evolution of these enzymes. 11 Biochemical and biophysical characterization of a-KG-dependent dioxygenases Biochemistry of T fdA and related enzymes deA was detected in a 2,4—D-degrading Arthrobacter sp. in 1967 (l 16). Preliminary studies on the degradation of 18O-ether—labeled 2,4—D found the products to be 18O-2,4—dichlorophenol and glyoxalate. Thus, it was proposed that the 2,4—D decomposition proceeds through a hemiacetal intermediate that spontaneously breaks down to yield the reaction products (Scheme 1). Almost twenty—five years later, the mechanism of purified deA was examined (27), proving this enzyme to be an or-KG- dependent dioxygenase. Various characteristics of recombinant deA, purified from E. coli cells containing the Ralstonia eutropha JMP134 tfdA gene, have been elucidated over the past five years. deA, with a predicted subunit molecular weight of 31.7 kDa has an apparent molecular weight of 50 :r 2.5 kDa by gel filtration chromatography suggesting that the native protein exists as a homodimer. The high affinity for or-KG is demonstrated by a Km of 3.2 i 0.54 uM. Several different oc-keto—acids, including oc-ketoadipate and pyruvate, are less efficient substitutes for oc-KG as the cosubstrate. The km and Km for 2,4-D are 529 i 16 min'] and 17.5 uM, respectively. 2,4—D can be replaced with other aromatic compounds such as phenoxyacetic acids (27), substituted cinnamic acids, naphthoxyacetic acid and benzofurancarboxylic acid (R. Padmakumar and J. Hotopp, unpublished data). deA enzymes from both R. eutropha JMP134 and Burkholderia cepacia RASC (1 12) selectively abstract the pro-R hydrogen atom from the prochiral side-chain carbon adjacent to the ether bond of 2,4—D (98). Interestingly, there is evidence of a deA-like oc-KG-dependent dioxygenase that has the opposite stereoselectivity from the deAs described above (98, 114), however, this enzyme has not been cloned or characterized (Appendix B). Distinct oc-KG-dependent dioxygenases in Sphingomonas herbicidovorans act on each enantionmer of mecoprop ((R,S)-2-(4—chloro-2- methylphenoxy)—propionic acid), a chiral analog of 2,4—D (84). Three characterized enzymes with similarity to deA include TauD, which releases sulfite from taurine, y-butyrobetaine hydroxylase, which is responsible for the final step in camitine biosynthesis, and clavaminate synthase, which catalyzes three distinct transformations in the biosynthesis of the B-lactamase inhibitor clavulinic acid. The diverse substrates of these enzymes are compared in Figure 1. While all four enzymes have approximately the same apparent KD for iron, other kinetic parameters vary significantly. The Km for oc-KG ranges from 3 uM for deA to 260 uM for clavaminate synthase. deA and TauD are the only enzymes that can substitute other oc-keto acids for OL-KG. The kcm/Km (min'l 11M “1) is 31.1, 2.0, 0.3, 0.013 for deA, TauD, y-butyrobetaine hydroxylase, and clavaminate synthase, respectively. The wide range of catalytic efficiencies likely reflects the diversity of substrates. In each case, the hydroxylated carbon is one or two carbons away from either a carboxylate or sulfonate group indicating that a positively charged amino acid might be involved in substrate binding. The pro-R hydrogens abstracted by y-butyrobetaine hydroxylase (22) and clavaminate synthase (5) are in similar relative positions to the carboxylate moiety, which is two carbons away from the hydroxylation site. The pro-S hydrogen removed by deA (98) is in the opposite position with respect to the carboxylate. The stereochemistry of TauD catalysis has not been examined. 13 2,4-D taurine \\S/O‘ +H3N/>‘,$/ \\O H 2“ Cl ‘ ? CH O H§L\\l 3 ll §b N ’ H3C/ + O- COOH y-butyrobetaine deoxyguanidinoproclavaminic acid Figure 1. Substrates of deA-like enzymes Structures of substrates for deA, TauD, y-butyrobetaine hydroxylase, and clavaminate synthase are compared. The abstracted proton is indicated by an arrow. 14 For maximal activity, many a-KG-dependent dioxygenases require ascorbate to reduce the iron after non-productive decarboxylation of a-KG that is uncoupled from oxidation of substrate. Although deA activity is stimulated by the presence of ascorbate - in the reaction mix (27), deA does not appear to undergo uncoupled turnover of oc-KG in the absence of substrate. Even with the addition of ascorbate, deA is inactivated in a matter of minutes in the presence of iron under oxic conditions. Inactivation of deA appears to occur by both turnover—dependent and a tumover-independent mechanisms (97). TauD activity is also stimulated by 50% in the presence of ascorbate at concentrations over 200 uM. y-Butyrobetaine hydroxylase activity is stimulated six-fold by the addition of 100 M potassium ions; the Km for both substrates is also reduced in the presence of potassium (130). Without potassium present, y—butyrobetaine hydroxylase performs the uncoupled decarboxylation of oc-KG. Higher rates of uncoupling are achieved by the addition of product and its enantiorner (42). Clavaminate synthase also catalyzes the uncoupled decarboxylation of oc-KG, but the rate is greatly reduced (1 in 890 turnovers) in the presence of substrate (60). Beyond the differences in kinetic parameters and activity towards a-KG in the absence of substrate, other characteristics of these enzymes can be compared. All four enzymes have a pH optimum of 6.5-7. TauD (subunit size of 37.4 kDa) and y- butyrobetaine hydroxylase, (43 kDa) both exist as homodimers (21, 62). In contrast, gel filtration chromatography of clavaminate synthase gives an estimated mass of 36 kDa indicating that the enzyme is not rnultimeric. Clavaminate synthase is unique in that it catalyzes hydroxylation, desaturation, and cyclization reactions (Scheme 2), all coupled to oc-KG decarboxylation, at the same active site (11, 60). Kinetic studies suggest that 15 there is no release of the substrate between the desaturation and cyclization reaction. It is interesting to recognize the similarities between the substrates of clavaminate synthase and those of IPNS and DAOCS. Despite a lack of sequence similarity, and all three enzymes participate in the synthesis of B-lactam antibiotics. In addition, a poorly characterized enzyme, CarC, which shows high levels of identity to deA over the entire sequence, is involved in formation of carbapenem, another B-lactam-containing compound (Scheme 2). Biochemical studies of other a-KG-dependent dioxygenases There is significant diversity in the subunit size and quaternary structure within the oc-KG-dependent dioxygenase superfamily. Among the enzymes related to IPNS, most are between 38 and 45 kDa (similar in size to deA-like enzymes) and many (flavone synthase (9), gibberellin 3B-hydroxylase (3 7), gibberellin 20-oxidase (100), hyoscyamine (34)), but not all (flavanone 3-hydroxylase (10), ACC oxidase (49), thymine hydroxylase (115)), of the enzymes in this subgroup exist in a monomeric state. Crystal structure information suggests that apo-DAOCS exists as a trimer but dissociates to a monomer upon metal binding (122). The three mammalian enzymes involved in collagen maturation, aspartyl hydroxylase, lysyl hydroxylase and prolyl hydroxylase, show additional differences from other a-KG-dependent dioxygenases. Aspartyl hydroxylase has a subunit size of 85 kDa, but a 52 kDa catalytic domain still has full catalytic activity (47). Lysyl hydroxylase has been identified as an 87 kDa monomer in chick embyros and bovine skin, and a dimer of 70 and 110 kDa subunits in porcine skin. Prolyl hydroxylase, with a subunit size of 61 kDa, forms an 01202 heterodimer in 16 mammals and an 043 dimer in C. elegans (54). In both cases, the [3 subunit is protein disulfide isomerase. In all three enzymes, the C-terminal catalytic domain appears to be linked to a separate N-terminal region involved in cellular localization (52). Furthermore, a recently discovered prolyl 4-hydroxylase from Paramecium bursaria Chorella virus- 1, an algal virus, that is 20% identical to the C-terminus of mammalian prolyl 4-hydroxylase lacks the N-terminal localization domain; thus, it’s subunit size is closer to that of bacterial and fungal a-KG-dependent dioxygenases (23). Almost all a—KG-dependent dioxygenases have an apparent KD for iron of approximately 2 uM. The Km for oc-KG ranges from 5 uM to between 140 uM for different enzymes. The Km for 02 has been determined to be 30-40 uM for prolyl hydroxylase (82), 60 uM for thymine hydroxylase (41), and 45 uM for desacetoxyvindoline 4-hydroxylase (14). Not surprisingly, the measured Km values for the unique substrates vary from nanomolar to near milimolar concentrations. Substrate interaction kinetics and product inhibition studies in four a-KG-dependent dioxygenases, desacetoxyvindoline 4-hydroxylase (l4), prolyl hydroxylase (82), lysyl hydroxylase (92), and thymine hydroxylase (41), all indicate an ordered ter-ter mechanism for substrate binding. Competitive inhibition studies indicated that in three of the four enzymes, oc-KG binds first, followed by the appropriate substrate, then 02. In prolyl hydroxylase, the order of Oz and substrate binding are reversed (82). The oxidized product, C02, and succinate are then released sequentially in that order. The unproductive decarboxylation of a-KG in the absence of substrates provides additional evidence that oc-KG can bind without the substrate present. 17 lPN S H Fe(ll) HOOC N m NH2 0 O N 02 21120 ' COOH H H HOOC N HOOC N S DAOCS Fe(II) O NH; 0 O N 7:» NH2 0 N / 0; H20 COOH aKG succinate (‘03 NH CS A Fe(ll) HO NH N A N NH O N NHZ COOH O ' H 2 succmate aKG (:02 COOH HO CS ..... ““110 Fe(ll) H N N 0 NH; 7? o ”“2 COOH 02 H20 COOH aKG succinate C02 ..... ““110 ,,_....u\\O H CS N Fe(ll) -—— 0% NH2 7?» N NH? 0 COOH 02 ”20 COOH uKG succinate C02 CarC acetyl-CoA N / ————> ——> y-glutamyl phosphate 0 COOH Scheme 2. Reactions catalyzed by IPN S, DAOCS, clavaminate synthase and CarC. Although in deA and related enzymes Ot-KG decarboxylation is relatively tightly coupled to substrate oxidation, prolyl 4—hydroxylase and lysyl hydroxylase (80) have a near absolute requirement for ascorbate. Early spectroscopic studies with Fe(H)-prolyl hydroxylase demonstrated that ascorbate serves to reduce the iron metallocenter which becomes oxidized to an Fe(HI) form upon brief incubation with a—KG. Ascorbate is consumed stoichiometrically with respect to uncoupled decarboxylation (80). Catalysis is not as dependent on the presence of ascorbate in bacterial proline hydroxylases (78). Studies with analogs of oc-KG and ascorbate suggest that both occupy a similar or overlapping site in prolyl 4—hydroxylase and both compounds chelate the metal (65). In addition to IPNS, aminocyclopropane carboxylate (ACC) oxidase does not require or-KG for activity. ACC oxidase, responsible for the generation of ethylene in plants, requires Fe(II) and C02 (49). The role for CO; has yet to be determined, but it has been suggested that bicarbonate, a carbamylated lysine, or CO; itself may ligate the iron (49). Stoichiometric amounts of ascorbate are consumed during the course of the reaction, perhaps suggesting that ascorbate replaces a-KG. Site-directed mutagenesis and chemical modification studies of a—KG-dependent dioxygenases Mutagenesis studies of several enzymes within Group I of the oc-KG-dependent dioxygenase superfamily have confirmed the importance of the conserved residues (H-X- D-Xss-H-Xlo-R). In IPNS, replacement of His 214, Asp 216 or His 272 with leucine completely inactivates the enzyme (1 13). The crystal structure of IPNS proves these residues to be metallocenter ligands (94, 95). Although a fourth metal ligand, Gln 330, is 19 evident in the substrate-free structure, this residue is neither conserved nor critical for activity (101). The importance of the corresponding residues in flavanone 3B- hydroxylase, which is 23% identical to IPNS over the entire sequence, has also been demonstrated (63). Alteration of His 220 or Asp 222 to glutamine causes a greater than 99% loss of activity but the resultant enzymes only show a modest increase in the apparent KD for iron. Mutation of His 278 to glutamine completely abolished activity, and mutation of a conserved Arg in position 288 to lysine or glutamine corresponded to decreased affinity for a-KG (63). Additional studies have been carried out with the enzymes that hydroxylate peptidyl substrates. Mutation of His 656, Asp 658, and His 708 rendered lysyl hydroxylase inactive (88). Likewise, alteration of the second conserved histidine in aspartate hydroxylase (His 675) completely abolishes activity (70) and appears to cause a decrease in metal and a—KG binding (47). Replacement of His 675 in aspartate hydroxylase with iron-binding amino acids restores partial activity (70). In prolyl hydroxlase, alteration His412, Asp 416, His483 and Lys493 in the human-derived enzyme had significant effects on activity. Prolyl hydroxylase variants in which either of the histidines was mutated were inactive regardless of the substitution (59). Chemically- equivalent substitutions of remaining conserved residues, Glu for Asp 414 and Arg for Lys 493, cause an increase in the Km for iron and oc-KG, respectively. Mutation of a fourth histidine in prolyl hydroxylase, His 501, leads to a 95% loss of activity, a small increase in Km for a-KG, and an increase in the rate of uncoupled oc-KG decarboxylation suggesting that this residue is also present at or near the active site (79). 20 Together, these studies conclusively show the importance of histidines and carboxylates in the creation of the reactive metallocenter of or-KG-dependent dioxygenases. In most enzymes, no alternative amino acids (Arg, Asp, Glu, Gln, Ala) can substitute for a metal-binding histidine, although in certain instances, low levels of activity can be seen with a His to Gln substitution. Alteration of the conserved aspartate to even the very similar glutamate cannot always retain activity, yet in clavaminate synthase and the sequence of SyrB, glutamate replaces aspartate in the metal binding motif. In the cases presented above, the conserved residues are identified through sequence comparisons to IPNS and related enzymes. To date, no mutagenesis studies on enzymes without similarity to IPNS have been published. Spectroscopic experiments for metallocenter characterization of a—KG-dependent dioxygenases Early insight into the metallocenter structure of enzymes in the OL-KG-dependent dioxygenase superfamily came from spectroscopic studies of IPNS. Electron paramagnetic resonance (EPR) spectra of inactive Cu(11)- and Co(II)—substituted IPNS, used to model the EPR silent Fe(H) form of the enzyme, are characteristic of a five— or six-coordinate tetragonally distorted site (75). Electron spin-echo envelope modulation (ESEEM) spectroscopic data are indicative of two equatorial histidines to the Cu(II) form of the enzyme, one of which is displaced upon substrate binding (48). Extended X-ray absorption fine structure (EXAFS) (105) spectroscopy demonstrates the presence of a thiolate ligand to the metal upon substrate binding. Site-directed mutagenesis of cysteine 21 residues ruled out a potential role for endogenous S-containing ligands thus indicating that the metal-bound sulfur is derived from substrate (85). Spectroscopic studies of three Group H oc-KG-dependent-dioxygenases, clavaminate synthase, deA and TauD, demonstrate common features among these three enzymes as weH as similarity to IPNS. Again, initial EPR characterization was performed on the inactive Cu(H)-substituted form of the enzyme. Structural information on the Cu(H)-metallocenter is likely relevant to the Fe(H) conformation as Cu(H) binds competitively with respect to iron with an approximate K, of 1-3 uM (131). Later studies, using F e(H)— nitric oxide (NO) complexes to derive an EPR signal, were also performed on deA (39). In the absence of substrates, EPR spectra of both Fe(H) and Cu(H)- substituted deA indicate the presence of two histidines bound to the metal. ESEEM studies of Cu(H)-deA agree with the EPR data, and also identify three metal-bound waters in the remaining coordination sites. Addition of 2,4-D to the enzyme does not affect the coordination geometry, but binding of a-KG decreases the rhombicity of the metallocenter and displaces two solvent molecules indicating direct interaction between oc-KG and the metal. ESEEM studies of the Cu(H)-oc-KG—2,4—D deA ternary complex indicated the loss or shift of a histidine ligand from the equatorial plane. Studies with the F eCH) deA are consistent with the rearrangement of ligands and loss of the third water upon substrate binding to give a five-coordinate metal center. Nitric oxide, which also serves as an oxygen analog, binds reversibly to Fe-deA and F e-or-KG-deA complexes, but shows a greatly increased affinity for the metal when both substrates are present. The interaction between oc-KG and the metal was studied in detail using UV- visible, magnetic circular dichroism (MCD) and circular dichroism (CD) spectroscopies 22 with clavaminate synthase. These techniques revealed that upon binding of a—KG to clavaminate synthase, the six-coordinate geometry is distorted. The formation of a pink intermediate arising from ligand to metal charge transfer between oc-KG and the iron indicates that oc-KG binds to the metal through both the C-1 carboxylate and the C-2 keto group based on comparisons to model compounds (87). A similar mode for binding of Ot- KG to deA and TauD is predicted from the results of recent experiments (39, 96). Crystal structures of IPNS and DA OCS IPNS has been crystallized in the presence of Mn(H) and Fe(H) (94, 95). Both structures show that the metallocenter is ligated by His 214, Asp 216, His 270, and Gln 330. The active site is buried within the hydrophobic pocket of an eight-strand “jelly-roll” motif. In the F e(H)-IPNS structure with substrate bound, Gln 330 and one water is displaced upon addition of the tripeptide substrate, ACV. In the Fe-ACV-IPNS complex, the iron is five-coordinate which agrees with the geometry observed in the spectroscopic experiments described above. Furthermore, nitric oxide binds to the open axial coordination site forming an octahedral complex as revealed by the crystallographic studies. The crystal structure of DAOCS is the only published structure of an oc-KG dependent dioxygenase. The anaerobic addition of F e(H) to crystals of apo-DAOCS leads to only small structural changes. In DAOCS, ferrous ion is ligated by residues corresponding to the metal ligands in IPNS, and three water molecules. As predicted by the spectroscopic experiments described above, a-KG displaces two solvent molecules to chelate the iron. The C-l carboxylate binds opposite to His 243 (the distal histidine of the 23 H-X-D-Xss-H motif) and the C-2 keto group binds trans to the metal-bound aspartate. The only accessible site, located in a hydrophobic pocket, is occupied by a solvent molecule. The crystal structures of IPNS and DAOCS lend very strong support for the models of a-KG—dependent dioxygenase metallocenter structure as deduced from sequence alignments, site-directed mutagenesis studies, and spectroscopic experiments. Although the substrate-free configuration of the two rnetallocenters are similar, the differences in the chemistry performed by these two enzymes, particularly the use of or— KG as a co-substrate in DAOCS and not in IPNS, is reflected in the differences in ligands and coordination geometries upon substrate binding. There is no available structure of an a-KG-dependent dioxygenase that performs hydroxylation chemistry. Crystal structures for two additional a-KG-dependent dioxygenases, clavaminate synthase (Group H) and proline 4-hydroxylase (Group IH) will soon accompany the structures of IPNS and DAOCS as described in a seminar by Prof. C. Schofield. Despite the distinct differences between the three groups of sequences, these enzymes appear to fold on a similar scaffold. This exciting finding will allow us to better assimilate the large amount of spectroscopic, stereochemical, and biochemical data into a universal model for enzymes in this superfamily. Within the subgroups, information from one crystal structure will greatly aid in the prediction of substrate binding residues and potential catalytic residues in other related enzymes. A 2-His-1-carboxylate motif is also found in other non-heme, ferrous ion dependent enzymes such as lipoxygenase, extradiol dioxygenase, and pterin-dependent tyrosine hydroxylase (38). These three ligands anchor the iron in the active sites, and, as 24 demonstrated by site-directed mutagenesis studies, are clearly important for establishing the proper geometry and reactivity for interaction with substrates. Thus, the three remaining coordination sites provide the opportunity for substrates to interact directly with the iron to modulate activity. Mechanism of a-KG-dependent dioxygenases a-KG-dependent dioxygenases perform a wide range of oxidative reactions including hydroxylations, cyclizations, ring expansions and desaturations using or—KG as a reductant of diatomic oxygen. Despite the diversity of substrates and reactions, the initial catalytic steps likely proceed in a similar fashion in these disparate enzymes. Kinetic, spectroscopic, and structural data, in addition to mutagenesis data and the studies on the behavior of model compounds, have lent support to the Hanauske-Abel and Gt‘rnzler model (33) that was proposed for prolyl hydroxylase. The expanded and updated catalytic model for this superfamily of enzymes predicts the generation of the ferryl-oxo species (Fe(IV)=O) via the oxidative decarboxylation of Ot-KG. A summary of the steps proposed in the oxyferryl mechanism for oc—KG—dependent dioxygenases are summarized in Figure 2. The iron in the resting enzyme is bound by two histidines and one carboxylate with three waters occupying the open coordination sites. Spectroscopic studies indicate that oc-KG displaces at least one water molecule upon binding to the metal through the C-1 carboxylate and C-2 keto groups. Chelation of the iron by or-KG alters the coordination geometry and redox environment sufficiently to induce binding of oxygen in an axial position. In some enzymes, the presence of substrates may be necessary for oxygen to bind to the metal. During the course of oxidative decarboxylation, one oxygen atom is incorporated into succinate, the other forming an 25 Resting metallocenter /0 ° o2 C|> / Fe(H>\ (ll—KG Fe(III)-"‘O His Asp ‘. - 1 "0 COO 9H His2 R-C-R O hydroxylated x substrate 0 0\ Fe(lll)—OH R-C-R Fe(fi 0 \\ .O - \ ,0 coo H ‘ o Fe(lV)=O R-C-R sztrate O CO SUCClnate /O\O/“\/\COO' 2 Fe(ll) Figure 2. Oxyferryl mechanistic model for oc-KG-dependent dioxygenases. 26 Fe(IV)-oxo species. In hydroxylation reactions (shown here), the reactive oxygen species abstracts a hydrogen atom from the substrate to yield a hydroxide radical that recombines with the substrate to give the final product. During cyclization, ring expansion, or desaturation reactions, the iron-bound oxo species is reduced by two electrons to water. Degradation of 2,4—D and Ecology of deA deA was initially characterized for its ability to degrade 2,4-D (1 10), one of the most commonly used herbicides in agriculture (121). Despite its widespread use, 2,4-D does not accumulate in the environment due to its rapid consumption by soil bacteria. The estimated half-life for 2,4-D in oxic soils is approximately seven days (132). Over the past few decades, researchers have been particularly interested the mechanism by which soil microorganisms come to exploit 2,4-D, a xenobiotic compound, as a source for carbon and energy. The best characterized pathway for 2,4—D degradation is encoded by genes found on plasmid pJP4 isolated from Ralstonia eutropha (formerly Alcaligenes eutrophus) JlVfP134 (Figure 3) (17, 19). In this bacterium, 2,4—D is converted to 2,4-dichlorophenol which is then broken down by series of enzymes encoded by zdeCDEF to form [3- ketoadipate, a Kreb’s cycle intermediate. A thorough review of the enzymes involved in 2,4-D degradation as well as an overview of other 2,4-D degradation pathways can be found in a recent chapter by Hausinger et al. (35) and will be only briefly summarized here. After deA catalyzes the initial transformation of 2,4—D, 2,4-dichlorophenol hydroxylase (deB) hydroxylates the aromatic ring to give dichlorocatechol. deB is a homotetramer with 63 kDa subunits and is similar in sequence and mechanism to 27 2,4-D 2,4-dichlorophenol 3,5-dichlorocatechol 2,4-dichloro-c1’s, crs-muconate oAcoo - OOC Cl OOC Cl deA deB deC 1 Cl C deD 000 'OOC rde/ deE chromosomal O enzymes B-ketoadipate 2-chlorom aleyl acetate cis-Z-chlorodienelactone R A, .5 DC” E11 17!! BH. K T C D; EF. 3.. Figure 3. Pathway for the degradation of 2,4—D by Ralstonia eutropha JMP134(pJP4) The enzymes involved in 2,4-D degradation include deA, 2,4-dichlorophenoxyacetic acid/oc-KG dioxygenase; deB, 2,4—dichlorophenol hydroxylase; deC, 3,5- dichlorocatechol dioxygenase; deD, chloromuconate cycloisomerase; deE, dienelactone hydrolase; and deF, maleylacetate reductase. The region of the pJP4 plasmid encoding these enzymes is also shown. The tfd gene names in italics correspond to the enzymes defined above. tde encodes 2,4—D perrnease and tde/S encodes a LysR- type regulator of the tfd genes. The above figure is adapted from van der Meer et al. (126). 28 enzymes responsible for the hydroxylation of salicylate and other aromatics. Each subunit contains a molecule of FAD and activity is dependent on NAD(P)H as a cosubstrate. deC is an Fe(H)-dependent intradiol dioxygenase that converts 3,5-dich10rocatechol to 2,4-dichloro-cis,cis-muconate. The fourth enzyme involved in 2,4-D degradation, choromuconate isomerase, or deD, converts 2,4—dichloro-cis,cis-muconate to 2- chlorodienelactone which is then hydrolyzed to 2-chloromalylacetate by dienelactone hydrolase or deE. The resultant chloromaleylacetate is then degraded by either a chromosomally—encoded maleylacetate reductase, which has been purified from R. eutropha JMPl34, or by deF, an enzyme which, by sequence, appears to have the same activity. The genes encoding deC, deD, deE, and deF are cotranscribed and are highly similar to operons involved in the degradation of other chlorinated aromatic compounds such as chlorobenzoate and trichlorobenzene (126). As shown in Figure 3, there are several other tfd genes associated with the 2,4—D degradative pathway in addition to the catabolic enzymes. The transcription of tfdA and the tdeDEFB operon are regulated by a LysR-type regulator, deR/ S, which activates transcription in response to the inducer 2,4-dichloro-cis,cis-muconate (24).There are two identical copies of the gene encoding deR/S. Recently, an energy-dependent protein that facilitates the uptake of 2,4—D, deK, was described and found to be similar to transporters for 4-hydroxybenzoate and other aromatic compounds (61). In addition, a second set of genes similar to tdeDEF B and a truncated regulatory gene, tde, are located in between the two sets of tfd genes. Because the second set of tfd genes (zdeHC ”E ”F ”311) is not required for 2,4-D degradation, but is induced in the presence of 29 2,4-D, they may have resulted from a gene duplication or gene transfer event that failed to produce a functional 2,4-D degradation pathway (126). Researchers have used the pJP4-encoded 2,4—D degradation pathway as a model to study the assembly of genes necessary to enable an organism to degrade a xenobiotic compound. Several features of the zfd operon, including the duplication of the regulatory element, a second copy of the tdeDEF operon, and the presence of the ISJP4 insertion element sequences in tde and near tde, suggest recent evolution of the 2,4—D degradation cassette. The data presented below provide additional evidence that the necessary genes are widespread in environmental bacteria and that they have been assembled multiple times to enable single organisms to degrade 2,4—D. Enrichments from soils ((86) and references therein), aquifers and water samples (3) readily degrade 2,4-D, although often only after an acclimation period (29, 103, 107). Soulas (107) modeled the dynamics the 2,4—D degradation as the result of competition between two populations with different 2,4—D degradation rates. In this simulation, the acclimation period reflects the time required for competitive 2,4wD degraders to achieve dominance. In support of this theory, phylogenetically diverse 2,4-D degrading isolates were obtained by directly plating soil bacteria onto 2,4-D-containing medium, but the diversity of 2,4-D degraders in liquid. enrichment cultures, a more competitive enviroMent, was greatly reduced (20). Experiments using pristine soils argue that genetic exchange may also play a role in 2,4-D degradation. Soil enrichments from six pristine locations metabolized 3-chloroben20ate (3-CBA) and 2,4-D (‘29). While 3-CBA was catabolized without a lag period and 3--CBA-degrading cultures were readily isolated from/these enrichments, 2,4-D was generally not consumed until the second ‘week of .30 incubation and only a small number 2,4-D degrading bacteria were obtained (29). These data suggest that 2,4-D is degraded by a consortium in environments without a history of 2,4-D treatment. Diverse bacteria capable of 2,4-D degradation have been isolated from various environmental samples and enrichments (76). Survey studies of isolates from soil, water and sewage sludge have found distinct taxa capable of growth on 2,4-D coexisting in certain environments (3, 20, 117). Over the past thirty years, many 2,4-D degrading bacteria from different genera have been characterized including strains of Pseudomonas (3, 31), Bordetella (31) and Alcaligenes (3, 18, 32), Variovorax (123), Burkholderia (112), Halomonas (66), Flavobacterium (13) Sphingomonas (50), Acinetobacter (6), Arthrobacter (116), Corynebacterium (3), Nocardioides (5 7), Azotobacter (4), Achromobacter (108), and Streptomyces (128). Further, genetic and biochemical studies of many of these strains demonstrate the pervasiveness of deA in these 2,4-D degradation pathways. Sphingomonas strains that degrade 2,4-D are commonly found in Michigan soils, but these isolates do not have a tfdA gene that can be detected by gene probes or a PCR-based screening method for different alleles of the tfdA gene (50). Nevertheless, biochemcial characterization of the Sphingomonas isolates indicate that 2,4-D degradation is initiated by an oc-KG-dependent dioxygenase (104). The tfd genes often reside on mobile genetic elements. Plasmids belonging to at least two different incompatibility groups have been isolated from Alcaligenes species (18) and soils (118). Plasmid capture experiments were used to isolate pEMT8, a 75-kb plasmid that encodes tfdA. This plasmid does not appear to contain tde or tdeDEF and fails to confer activities of the downstream enzymes to a 2,4-D' Ralstom’a eutropha 31 recipient (119). Nearly identical alleles of tfdA have been detected on plasmids isolated from soils (68) and on the chromosome of plasmid-cured Burkholderia sp. suggesting that the gene may travel on transposons as well as plasmids (112). Evidence for a conjugative plasmid capable of integration into the chromosome of different strains has also been published (51). The pJP4 plasmid can be transferred to bacteria of different genera such as Variovorax, Pseudomonas, and Burkholderia in soils (16, 25) and similar plasmids have been isolated from F lavobacterium sp. (13, 50). Different plasmids (7, 64) and putative degradative transposons, such as Tn5 53 0 carried on pIJBl (134) have also been purified from 2,4-D degrading strains. Additional proof of the extensive intergeneric transfer of tfdA is the distribution of at least three different alleles of tfdA across the [3- Proteobacteria (7 1). In addition to studies on the dispersal of 2,4—D degradative ability, much attention has been paid to the creation or assembly of this pathway. Several experiments have been designed to determine if there is one common ancestor to tfd-based gene cluster or if these genes were assembled multiple times from other biodegradative pathways. Xia et al. (133) reported that DNA from soils with no or low exposure to 2,4-D do not hybridize to the pJP4 tfdA gene probe and no 2,4-D degrading isolates could be cultured from these soils. Upon exposure to herbicide, 2,4-D degraders with different tfdA and tde genes become dominant in the different plots. Because there was not a major shift in the total community structure in response to 2,4-D application, based on random amplified polymorphic DNA (RAPD) fingerprint analysis, it seems that either a 2,4-D degrading organism was created from the resident community or a mobile element carrying the necessary genes was acquired and transferred among resident organisms. Phylogenetic 32 analyses of different 2,4-D degrading bacteria also indicate independent recruitment of the {fd genes into different organisms during the assembly of the 2,4-D catabolic pathway (124). Dendrograms plotting the relationships of each of three tfd alleles (g’dA, tde, and tde) from a collection of 2,4-D degraders are significantly different indicating that these three genes have not evolved together (28, 124). The seemingly ubiquitous nature of 2,4-D degrading organisms from both agricultural and pristine soils as well as from aquatic environments argues that this metabolic characteristic is important for competitiveness in the environment. Although some bacteria can degrade 2,4—D by other means (4, 53, 57), deA appears to play a major role in the breakdown of 2,4-D in the environment. Of all the enzymes in the pJP4- encoded pathway, deA is the most unique to 2,4-D degradation. Phenols and catechols are intermediates in numerous catabolic pathways, thus different enzymes exist to catalyze their degradation. No alternative role for deA has been identified. Overview of thesis In this thesis, I present my research on the characterization of the active site of deA and a study of related enzymes. Chapter 2 focuses specifically on studies relating to the structure and activity of deA from Ralstonia eutropha JMP134. I describe the site- directed mutagenesis of ten residues and the characterization of the resultant enzymes to determine the role, if any, of these amino acids in deA structure or activity. These studies are an important contribution to research on the structure and mechanism of deA and the related enzyme TauD. In Chapter 3, I describe the cloning and characterization of an a-KG-dependent dioxygenase from Saccharomyces cerevisiae to determine if it 33 possesses activity similar to deA. Work with a strain of S. cerevisiae in which this gene was disrupted confirms our findings that this enzyme is a sulfonate/oc-KG dioxygenase. Chapter 4 describes my research on putative deA-like oc-KG-dependent dioxygenases in soil bacteria. In the last chapter, I summarize my current thoughts on the role of oc-KG- dependent dioxygenases in the environment and propose future directions for research on deA and deA-like enzymes. 34 10. 11. Literature cited Altschul, S. F., W. Gish, W. Miller, E. W. Myers, and D. J. Lipman. 1990. Basic local alignment search tool. J Mol Biol 215:403-10. Altschul, S. F., T. L. Madden, A. A. Schiiffer, J. Zhang, Z. Zhang, W. Miller, and D. J. Lipman. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 25:3389-3402. Amy, P. S., J. W. Schulke, L. M. Frazier, and R. J. Seidler. 1985. Characterization of aquatic bacteria and cloning of genes specifying partial degradation of 2,4-dichlorophenoxyacetic acid. Appl. Environ. Microbiol. 49:1237-1245. Balajee, S., and A. Mahadevan. 1990. Dissimilation of 2,4- dichlorophenoxyacetic acid by Azotobacter chroococcum. Xenobiotica 20:607— 617. Baldwin, J. E., K. D. Merritt, C. J. Schofield, S. W. Elson, and K. H. Baggaley. 1993. Studies on the stereospecificity of the clavaminic acid synthase catalyzed hydroxylation reaction. J. Chem. Soc., Chem. Comm.:l30l-1302. Bell, G. R. 1957. Some morphological and biochemical characteristics of a soil bacterium which decomposes 2,4-dichlorophenoxyacetic acid. Can. J. Microbiol. 3: 82 1-840. Bhat, M. A., M. Tsuda, K. Horiike, M. N ozaki, C. S. Vaidyanathan, and T. N akazawa. 1994. Identification and characterization of a new plasmid carrying genes for degradation of 2,4-dichlorophenoxyacetate from Pseudomonas cepacia CSV90. Appl. Environ. Microbiol. 60:307-312. Borovok, I., O. Landman, R. Kreisberg-Zakarin, Y. Aharonowitz, and G. Cohen. 1996. Ferrous active site of isopenicillin N synthase: genetic and sequence analysis of the endogenous ligands. Biochemistry 35: 1981-1987. Britsch, L. 1990. Purification and characterization of flavone synthase I, a 2- oxoglutarate-dependent desaturase. Arch. Biochem. Biophys. 282: 1 52-160. Britsch, L., and H. Grisebach. 1986. Purification and characterization of (2S)- flavanone 3-hydroxylase from Petunia hybrida. Eur. J. Biochem. 156:569-577. Busby, R. W., M. D.-T. Chang, R. C. Busby, J. Wimp, and C. A. Townsend. 1995. Expression and purification of two isozymes of clavaminate synthase and initial characterization of the iron binding site. J. Biol. Chem. 270:4262—4269. 35 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. Chahal, A., M. Khan, S. G. Pai, E. Barbosa, and 1. Singh. 1998. Restoration of phytanic acid oxidation in Refsum disease fibroblasts from patients with mutations in the phytanoyl-CoA hydroxylase gene. FEBS Lett. 429: 1 19-122. Chaudry, G. R., and G. H. Huang. 1988. Isolation and characterization of a new plasmid from a F lavobacterium sp. which carries the genes for the degradation of 2,4-dichlorophenoxyacetate. J. Bacteriol. 170:3897-3902. De Carolis, E., and V. De Luca. 1993. Purification, characterization, and kinetic analysis of a 2-oxoglutarate-dependent dioxygenase involved in vindoline biosynthesis of Catharanthus roseus. J. Biol. Chem. 268:5504-5511. De Long, L., S. P. J. Albracht, and A. Kemp. 1982. Prolyl 4-hydroxylase activity in relation to the oxidation state of enzyme-bound iron. Biochim. Biophys. Acta. 704:326-332. DiGiovanni, G. D., J. W. N eilson, I. L. Pepper, and N. A. Sinclair. 1996. Gene transfer of Alcaligenes eutrophus JMP134 plasmid pJP4 to indigenous soil recipients. Appl. Environ. Microbiol. 62 :252 1 —2526. Don, R. H., and J. M. Pemberton. 1985. Genetic and physical map of the 2,4- dichlorophenoxyacetic acid-degrading plasmid pJP4. J. Bacteriol. 161:466-468. Don, R. H., and J. M. Pemberton. 1981. Properties of six pesticide degradation plasmids isolated from Alcaligenes paradoxus and Alcaligenes eutrophus. J. Bacteriol. 145:681-686. Don, R. H., A. J. Weightman, H. J. Knackmuss, and K. N. Timmis. 1985. Transposon mutagenesis and cloning analysis of the pathways for degradation of 2,4-dichlorophenoxyacetic acid and 3-chorobenzoate in Alcaligenes eutrophus JMP134 (pJP4). J. Bacteriol. 161:85-90. Dunbar, J., S. White, and L. Forney. 1997. Genetic diversity through the looking glass: effect of enrichment bias. Appl. Environ. Microbiol. 63: 1326-133 1. Eichhorn, E., J. R. van der Ploeg, M. A. Kertesz, and T. Leisinger. 1997. Characterization of oc—ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J. Biol. Chem. 272:23031-23036. Englard, S., J. S. Blanchard, and C. F. Midelfort. 1985. y-Butyrobetaine hydroxylase: stereochemical course of the hydroxylation reaction. Biochemistry 24:1110-1116. Eriksson, M., J. Myllyharju, H. Tu, M. Hellman, and K. I. Kivirikko. 1999. Evidence for 4-hydroxyproline in viral proteins. Characterization of a viral prolyl 4-hydroxylase and its peptide substrates. J Biol Chem 274:2213 1-22134. 36 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. Filer, K., and A. R. Harker. 1997. Identification of the inducing agent of the 2,4-dichlorophenoxyacetic acid pathway encoded by plasmid pJ P4. Appl. Environ. Microbiol. 63:317-320. Friedrich, B., M. Meyer, and H. G. Schlegel. 1983. Transfer and expression of the herbicide-degrading plasmid pJP4 in aerobic autotrophic bacteria. Arch. Microbiol. 134:92-7. Fukumori, F ., and R. P. Hausinger. 1993. Alcaligenes eutrophus JMP134 "2,4- dichlorophenoxyacetate monooxygenase" is an oc—ketoglutarate-dependent dioxygenase. J. Bacteriol. 175:2083-2086. Fukumori, F., and R. P. Hausinger. 1993. Purification and characterization of 2,4-dichlorophenoxyacetate/OL-ketoglutarate dioxygenase. J. Biol. Chem. 268:24311-24317. Fulthorpe, R. R., C. McGowan, O. V. Maltseva, W. E. Holben, and J. M. Tiedje. 1995. 2,4-Dichlorophenoxyacetic acid-degrading bacteria contain mosaics of catabolic genes. Appl. Environ. Microbiol. 61:3274-3281. Fulthorpe, R. R., A. N. Rhodes, and J. M. T iedje. 1996. Pristine soils mineralize 3-chlorobenzoate and 2,4-dichlorophenoxyacetate via different microbial p0pulations. Appl. Environ. Microbiol. 62:1159-1 166. Galland, S., F. Le Borgne, D. Guyonnet, P. Clouet, and J. Demarquoy. 1998. Purification and characterization of the rat liver gamma-butyrobetaine hydroxylase. Mol. Cell. Biochem. 178:163-8. Greer, L. E., J. A. Robinson, and D. R. Shelton. 1992. Kinetic comparison of seven strains of 2,4-dichlorophenoxyacetic acid-degrading bacteria. Appl. Environ. Microbiol. 58: 1027-1030. Gunalan, and J. C. Fournier. 1993. Effect of microbial competition on the survival and activity of 2,4-D degrading Alcaligenes xylosoxidans subsp. denitrificans added to soil. Lett. Appl. Microbiol. 16: 178-181. Hanauske—Abel, H. M., and V. Giinzler. 1982. A stereochemical concept for the catalytic mechanism of prolyl hydroxylase. Applicability to classification and design of inhibitors. J. Theor. Biol. 94:421-455. Hashimoto, T., and Y. Yamada. 1987. Purification and characterization of hyoscyamine 6B—hydroxylase, from root cultures of Hyoscyamus niger L. Eur. J. Biochem. 164:277-285. 37 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. Hausinger, R. P., F. Fukumori, D. A. Hogan, T. M. Sassanella, Y. Kamagata, H. Takami, and R. E. Wallace. 1996. Biochemistry of 2,4-D degradation: evolutionary implications, p. 35-51. In K. Horikoshi, M. Fukuda, and T. Kudo (ed.), Microbial Diversity and Genetics of Biodegradation. Japan Scientific Press, Tokyo. Hedden, P. 1992. 2-Oxog1utarate-dependent dioxygenases in plants: mechanism and function. Biochem. Soc. Trans. 20:373-377. Hedden, P. 1997. The oxidases of giberellin biosynthesis: Their function and mechanism. Physio]. Planta. 101 :709-71 9. Hegg, E. L., and L. Que Jr. 1997. The 2-His-1-carboxylate facial triad. Eur. J. Biochem. 250:625-629. Hegg, E. L., A. K. Whiting, J. McCracken, R. P. Hausinger, and J. Que, L. 1999. The herbicide-degrading a-keto acid enzyme deA: metal coordination environment and mechanistic insights. Biochemistry In press. Hogan, D. A., T. A. Auchtung, and R. P. Hausinger. 1999. Cloning and characterization of a sulfonate/a—ketoglutarate dioxygenase from Saccharomyces cerevisiae. J. Bacteriol. 181:5876—5879. Holme, E. 1975. A kintic study of thymine 7-hydroxylase from Neurosporra crassa. Biochemistry 14:4999-5003. Holme, E., S. Lindstedt, and I. Nordin. 1982. Uncoupling in the y— butyrobetaine hydroxylase reaction by D- and L-carnitine. Biochem. Biophys. Res. Comm. 107:518-524. Holton, T. A., F. Brugliera, and Y. Tanaka. 1993. Cloning and expression of flavonol synthase from Petunia hybrida. Plant J 4:1003-10. Hulse, J. D., S. R. Ellis, and L. M. Henderson. 1978. Camitine biosynthesis. [3- Hydroxylation of trimethyllysine by an a-ketoglutarate-dependent mitochondrial dioxygenase. J. Biol. Chem. 253:1654-9. Hutton Jr., J. J., A. L. Tappel, and S. Udenfriend. 1967. Cofactor and substrate requirements of collagen proline hydroxylase. Arch. Biochem. Biophys. 118:231- 240. Jansen, G. A., S. J. Mihalik, P. A. Watkins, C. Jakobs, H. W. Moser, and R. J. Wanders. 1998. Characterization of phytanoyl-coenzyme A hydroxylase in human liver and activity measurements in patients with peroxisomal disorders. Clin. Chim. Acta 271:203-11. 38 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. Jia, S., K. McGinnis, W. J. VanDusen, C. J. Burke, A. Kuo, P. R. Griffin, M. K. Sardana, K. O. Elliston, A. M. Stern, and P. A. Friedman. 1994. A fully active catalytic domain of bovine aspartyl (asparaginyl)-B-hydroxylase expressed in Escherichia coli: characterization and evidence for the identification of an active-site region in vertebrate-OL-ketoglutarate-dependent dioxygenases. Proc. Natl. Acad. Sci. 91 :7227-723 1. Jiang, F., J. Peisach, L.-J. Ming, L. Que, Jr., and V. J. Chen. 1991. Electron spin-echo envelope modulation studies of the Cu(H)-substituted derivative of isopenicillin N synthase: a structural and spectroscopic model. Biochemistry 30:11437-11445. John, P. 1997. Ethylene biosynthesis: The role of l-aminocyclopropane-l- carboxylate (ACC) oxidase, and its possible evolutionary origin. Physiologia Plantarum 100:583-592. Ka, J. 0., W. E. Holben, and J. M. Tiedje. 1994. Genetic and phenotypic diversity of 2,4-dichlorophenoxyacetic acid (2,4-D)-degrading bacteria isolated from 2,4-D-treated field soils. Appl. Environ. Microbiol. 60:1106-11 15. Ka, J. 0., and J. M. Tiedje. 1994. Integration and excision of a 2,4- dichlorophenoxyacetic acid degradative plasmid in Alcaligenes paradoxus and evidence of its natural intergeneric transfer. J. Bacteriol. 176:5284-5289. Kellokumpu, S., R. Sormunen, J. Heikkinene, and R. Myllyla. 1994. Lysyl hydroxylase, a collage processing enzyme, exemplifies a novel class of luminally- oriented peripheral membrane protein in the endoplasmic reticulum. J. Biol. Chem. 269:30524-30529. Kilbane, J. J., D. K. Chatterjee, J. S. Karns, S. T. Kellogg, and A. M. Chakrabarty. 1982. Biodegradation of 2,4,5-trichlorophenoxyacetic acid by a pure culture of Pseudomonas cepacia. Appl. Environ. Microbiol. 44:72-78. Kivirikko, K. I., and T. Pihlajaniemi. 1998. Collagen hydroxylases and the protein disulfide isomerase subunit of prolyl 4-hydroxylases. Adv. Enzymol. Relat. Areas Mol. Biol. 72:328-398. Korioth, F., C. Gieffers, and J. Frey. 1994. Cloning and characterization of the human gene encoding aspartyl B-hydroxylase. Gene 150:395-399. Kovacevic, S., B. J. Weigel, M. B. Tobin, T. D. Ingolia, and J. R. Miller. 1989. Cloning, characterization, and expression in Escherichia coli of the Streptomyces clavuligerus gene encoding deacetoxycephalosporin C synthetase. J. Bacteriol. 171 :754-60. 39 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. Kozyreva, L. P., Y. V. Shurukhin, Z. I. Finkelstein, B. P. Baskunov, and L. A. Golovleva. 1993. Metabolism of the herbicide 2,4—D by a Nocardioides simplex strain. Mikrobiol. 62:110-119. Kreisberg-Zakarin, R., I. Borovok, M. Yanko, Y. Aharonowitz, and G. Cohen. 1999. Recent advances in the structure and function of isopenicillin N synthase. Antonie Van Leeuwenhoek 7 5:33—39. Lamberg, A., T. Pihlajaniemi, and K. I. Kivirikko. 1995. Site-directed mutagenesis of the or subunit of the human prolyl 4-hydroxylase. J. Biol. Chem. 270:9926-9931. Lawlor, E. J., S. W. Elson, S. Holland, R. Cassels, and J. E. Hodgson. 1994. Expression in Escherichia coli of a clavaminic acid synthase isozyme: 3 trifuctional oxygenases involved in clavulanic acid biosynthesis. Tetrahedron 50:8737-8748. Leveau, J. H., A. J. B. Zehnder, and J. R. Van der Meer. 1998. The tde gene product facilitates the uptake of 2,4-dichlorophenoxyacetate by Ralstonia eutropha JMP134(pJP4). J. Bacteriol. 180:2237-2243. Lindstedt, G., and S. Lindstedt. 1970. Cofactor requirements of y-butyrobetaine hydroxylase from rat liver. J. Biol. Chem. 245:4178-4180. Lukacin, R., and L. Britsch. 1997. Identification of strictly conserved histidine and arginine residues as part of the active site in Petunia hybrida flavanone 3B- hydroxylase. Eur. J. Biochem. 249:748-757. Mae, A. A., R. O. Martis, N. R. Ausmees, V. M. K6iv, and A. L. Heinaru. 1993. Characterization of a new 2,4-dichlorophenoxyacetic acid degrading plasmid pEST4011: physical map and localization of catabolic genes. J. Gen. Microbiol. 139:3165-3170. Majamaa, K., V. Giinzler, H. M. Hanauske—Abel, R. Myllylii, and K. I. Kivirikko. 1986. Partial identity of the 2-oxoglutarate and ascorbate binding sites of prolyl 4-hydroxylase. J. Biol. Chem. 261:7819-7823. Maltseva, O. V., C. McGowan, R. Fulthorpe, and P. J. Oriel. 1996. Degradation of 2,4-dichlorophenoxyacetic acid by haloalkaliphilic bacteria. Microbiol. 142:1115-1122. Mao, Y. Q., M. Vargolu, and D. H. Sherman. 1999. Molecular characterization and analysis of the biosynthetic gene cluster for the antiturnor antibiotic mitomycin C from Streptomyces lavendulae NRRL2564. Chem. Biol. 6:251-263. 40 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. Matheson, V. G., L. F. F orney, Y. Suwa, C. H. Nakatsu, A. J. Sexstone, and W. E. Holben. 1996. Evidence for acquisition in nature of a chromosomal 2,4- dichlorophenoxyacetic acid/oc-ketoglutarate dioxygenase gene by different Burkholderia spp. Appl. Environ. Microbiol. 62:2457-2463. Matsuda, J., S. Okabe, T. Hashimoto, and Y. Yamada. 1991. Molecular cloning of hyoscyamine 6 beta-hydroxylase, a 2-oxoglutarate- dependent dioxygenase, from cultured roots of Hyoscyamus niger. J. Biol. Chem. 266:9460- 9464. McGinnis, K., G. M. Ku, W. J. VanDusen, J. Fu, V. Garsky, A. M. Stern, and P. A. Friedman. 1996. Site-directed mutagenesis of residues in a conserved region of bovine apartyl (asparaginyl) B-hydroxylase: evidence that histidine 67 5 has a role in binding Fe+2. Biochemistry 35:3957-3962. McGowan, C., R. Fulthorpe, A. Wright, and J. M. Tiedje. 1998. Evidence for interspecies gene transfer in the evolution of 2,4-dichlorophenoxyacetic acid degraders. Appl. Environ. Microbiol. 64:4089-4092. McGowan, S. J., M. T. Holden, B. W. Bycroft, and G. P. Salmond. 1999. Molecular genetics of carbapenem antibiotic biosynthesis. Antonie Van Leeuwenhoek 75:135-141. Metcalf, W. W., and R. S. Wolfe. 1998. Molecular genetic analysis of phosphite and hypophosphite oxidation by Pseudomonas stutzeri WM88. J. Bacteriol. 180:5547-5558. Miller, R. L., and H. H. Varner. 1979. Purification and enzymic properties of lysyl hydroxylase from fetal porcine skin. Biochemistry 18:5928-5932. Ming, L.-J., L. Que, Jr., A. Kriauciunas, C. A. Frolik, and V. J. Chen. 1990. Coordination chemistry of the metal binding site of isopenicillin N synthase. Inorg. Chem. 29:1111-1112. Miwa, N ., and S. Kuwatsuka. 1991. Characteristics of 2,4- dichlorophenoxyacetic acid (2,4-D)-degrading microorganisms isolated from different soils. Soil Sci. Plant Nutr. 37 :575-58 1. Mori, H., T. Shiasaki, K. Yano, and A. Ozaki. 1997. Purification and cloning of a proline 3-hydroxylase, a novel enzyme which hydroxylates fiee L-proline to cis— 3-hydroxyl-L-proline. J. Bacteriol. 179:5677-5683. Mori, H., T. Shibasaki, Y. Uozaki, K. Ochiai, and A. Ozaki. 1996. Detection of novel proline 3-hydroxylase activities in Streptomyces and Bacillus spp. by regio- and stereospecific hydroxylation of L-proline. Appl. Environ. Microbiol. 62:1903-1907. 41 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. Myllyharju, J., and K. I. Kivirikko. 1997. Characterization of the iron- and 2- oxoglutarate-binding sites of human prolyl 4-hydroxylase. EIVIBO J. 16:1173- 1 180. Myllylii, R., K. Majamaa, V. Giinzler, H. N. Hanauske-Abel, and K. I. Kivirikko. 1984. Ascorbate is consumed stoichiometrically in the uncoupled reactions catalyzed by prolyl 4-hydroxylase and lysyl hydroxylase. J. Biol. Chem. 259:5403-5405. Myllyla, R., T. Pihlajaniemi, L. Pajunen, T. Turpeenniemi-Hujanen, and K. I. Kivirikko. 1991. Molecular cloning of chick lysyl hydroxylase. Little homology in primary structure to the two types of subunit of prolyl 4- hydroxylase. J. Biol. Chem. 266:2805-2810. Myllylii, R., L. Tuderman, and K. I. Kivirikko. 1977. Mechanism of the prolyl hydroxylase reaction: 2. Kinetic analysis of the reaction sequence. Eur. J. Biochem. 80:349-357. Nagahama, K., T. Ogawa, T. Fujii, M. Tazaki, S. Tanase, Y. Merino, and H. Fukuda. 1991. Purification and properties of an ethylene-forming enzyme from Pseudomonas syringae pv. phaseolicola PK2. J. Gen. Microbiol. 137:2281-2286. Nickel, K., M. J.-F. Suter, and H.-P. E. Kohler. 1997. Involvement of two oc- ketoglutarate-dependent dioxygenases in enantioselective degradation of (R)- and (8)-mecoprop by Sphingomonas herbicidovorans MI-I. J. Bacteriol. 179:6674- 6679. Orville, A. M., V. J. Chen, A. Kriauciunas, M. R. Harpel, B. G. Fox, E. Miinck, and J. D. Lipscomb. 1992. Thiolate ligation of the active site F e+2 of isopenicillin N synthase derived from substrate rather than endogenous cysteine: spectroscopic studies of site-specific Cys to Ser mutated enzymes. Biochemistry 31:4602-4612. Ou, L.-T. 1984. 2,4-D degradation and 2,4-D degrading microorganisms in soils. Soil Sci. 137:100-107. Pavel, E. G., J. Zhou, R. W. Busby, M. Gunsior, C. A. Townsend, and E. 1. Solomon. 1998. Circular dichroism and magnetic circular dichroism spectroscopic studies of the non-heme ferrous active site in clavaminate synthase and its interaction with ct-ketoglutarate cosubstrate. J. Am. Chem. Soc. 120:743- 754. Pirskanen, A., A.-M. Kaimio, R. Myllylii, and K. I. Kivirikko. 1996. Site- directed mutagenesis of human lysyl hydroxylase expressed in insect cells. J. Biol. Chem. 271:9398-9402. 42 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. Pirskanen, A., A.-M. Kaimio, R. Myllylii, and K. I. Kivirikko. 1996. Site— directed mutagenesis of human lysyl hydroxylase expressed in insect cells. Identification of histidine residues and an aspartic acid residue critical for catalytic activity. J. Biol. Chem. 271:9398-9402. Prescott, A. G. 1993. A dilemma of dioxygenases (or where biochemsitry and molecular biology fail to meet). J. Exp. Bot. 44:849-861. Prescott, A. G., and P. John. 1996. Dioxygenases: Molecular structure and role in plant metabolism. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47 :245-271. Puistola, U., T. M. Turpeenniemi-Hujanen, R. Myllylii, and K. I. Kivirikko. 1980. Studies on the lysyl hydoxylase reaction: H. Inhibition kinetics and the reaction mechanism. Biochim. Biophys. Acta 611 :5 1-60. Ramon, D., L. Carramolino, C. Patino, F. Sanchez, and M. A. Penalva. 1987. Cloning and characterization of the isopenicillin N synthetase gene mediating the formation of the beta-lactam ring in Aspergillus nidulans. Gene 57:171-81. Roach, P. L., I. J. Clifton, V. Fiiliip, K. Harlos, G. J. Barton, J. Hajdu, I. Andersson, C. J. Schofield, and J. E. Baldwin. 1995. Crystal structure of isopenicillin N synthase is the first from a new structural family of enzymes. Nature 375:700-704. Roach, P. L., I. J. Clifton, C. M. H. Hensgens, N. Shibata, C. J. Schofield, J. Hajdu, and J. E. Baldwin. 1997. Structure of isopenicillin N synthase complexed with substrate and the mechanism of penicillin formation. Nature 387 :827—830. Ryle, M. J., R. Padmakumar, and R. P. Hausinger. 1999. Stopped-flow kinetic analysis of Escherchia coli taurine/OL-ketoglutarate dioxygenase: interactions with oc-ketoglutarate, taurine, and oxygen. Biochem. In press. Saari, R. E., and R. P. Hausinger. 1998. Ascorbic acid-dependent turnover and reactivation of 2,4- dichlorophenoxyacetic acid/alpha-ketoglutarate dioxygenase using thiophenoxyacetic acid. Biochemistry 37 :3035-3042. Saari, R. E., D. A. Hogan, and R. P. Hausinger. 1999. Stereospecific degradation of the phenoxypropionate herbicide dichlorprop. J. Molec. Catalysis B: Enzymatic 266:421-428. Saito, K., M. Kobayashi, Z. Gong, Y. Tanaka, and M. Yamazaki. 1999. Direct evidence for anthocyanidin synthase as a 2-oxoglutarate-dependent oxygenase: molecular cloning and functional expression of cDNA from a red forma of Perilla frutescens. Plant Journal 17: 181-1 89. 43 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. Saito, T., S. S. Kwak, Y. Kamiya, H. Yamane, A. Sakurai, N. Murofushi, and N. Takahashi. 1991. Plant Cell. Physiol. 32:239-245. Sami, M., T. J. N. Brown, P. L. Roach, C. J. Schofield, and J. E. Baldwin. 1997. Glutamine--330 is not essential for activity in isopenicillin N synthase from Aspergillus nidulans. FEBS Lett. 405: 191-194. Samson, S. M., J. E. Dotzlaf, M. L. Slisz, G. W. Becker, R. M. Van Frank, L. E. Veal, W. K. Yeh, J. R. Miller, S. W. Queener, and T. D. Ingolia. 1987. Cloning and expression of the fungal expandase/hydroxylase gene involved in cephalosporin biosynthesis. Biotechnology 5:1207-1214. Sandmann, E. R. I. C., M. A. Loos, and L. P. van Dyk. 1988. Microbial degradation of 2,4-dichlorophenoxyacetic acid in soil. Rev. Environ. Contamin. Tox. 101:1-53. Sassanella, T. M., F. Fukumori, M. Bagdasarian, and R. P. Hausinger. 1997. Use of 4-nitrophenoxyacetic acid for the detection and quantitation of 2,4- dichlorophenoxyacetic acid (2,4-D)/oc-ketoglutarate dioxygenase activity in 2,4-D degrading microorganisms. Appl. Environ. Microbiol. 63:1 189-1 191. Scott, R. A., S. Wang, M. K. Eidsness, A. Kriauciunas, C. A. Frolik, and V. J. Chen. 1992. X-ray absorption spectroscopic studies of the high-spin iron(II) active site of isopenicillin N synthase: evidence for F e-S interaction in the enzyme-substrate complex. Biochemistry 31 :45 96-601 . Shibasaki, T., H. Mori, S. Chiba, and A. Ozaki. 1999. Microbial proline 4- hydroxylase screening and gene cloning. Appl. Environ. Microbiol. 65:4028- 4031. Soulas, G. 1993. Evidence for the existence of different physiological groups in the microbial community responsible for 2,4-D mineralization in soil. Soil Biol. Biochem. 25:443-449. Steenson, T. I., and N. Walker. 1957. The pathway of breakdown of 2,4- dichloro- and 4-chloro-2-methyl-phenoxyacetic acid by bacteria. J. Gen. Microbiol. 16:146-155. Stintzi, A., Z. Johnson, M. Stonehouse, U. Ochsner, J. Meyer, M. Vasil, and K. Poole. 1999. The pvc gene cluster of Pseudomonas aeruginosa: role in synthesis of the pyoverdine chromophore and regulation by Pth and PvdS. J. Bacteriol. 181:4118-4124. Streber, W. R., K. N. Timmis, and M. H. Zenk. 1987. Analysis, cloning, and high-level expression of 2,4-dichlorophenoxyacetate monooxygenase gene tfdA of Alcaligenes eutrophus JMP134. J. Bacteriol. 169:2950-2955. 44 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. Stubbe, J. 1985. Identification of two alpha—ketoglutarate-dependent dioxygenases in extracts of Rhodotoru/a glutinis catalyzing deoxyuridine hydroxylation. J. Biol. Chem. 260:9972—9975. Suwa, Y., A. D. Wright, F. Fukumori, K. A. Nummy, R. P. Hausinger, W. E. Holben, and L. J. Forney. 1996. Characterization of a chromosomally encoded 2,4-dichlorophenoxyacetic acid/a-ketoglutarate dioxygenase from Burkhola’eria sp. strain RASC. Appl. Environ. Microbiol. 62:2464-2469. Tan, D. S. H., and T.-S. Sim. 1996. Functional analysis of conserved histidine residues in Cephalosporin acremoniumisopenicillin N synthase by site-directed mutagenesis. J. Biol. Chem. 271:889-894. Tett, V. A., A. J. Willetts, and H. M. Lappin-Scott. 1997. Biodegradation ofthe chlorophenoxy herbicide (R)—(+)—mecoprop by Alcaligenes denitrificans. Biodegradation 8:43-52. Thornburg, L. D., M. T. Lai, J. S. Wishnok, and J. Stubbe. 1993. A non-heme protein with heme tendencies: an investigation of the substrate specificity of thymine hydroxylase. Biochemistry 32: 14023-14033. Tiedje, J. M., and M. Alexander. 1969. Enzymatic cleavage of the ether bond of 2,4-dichlorophenoxyacetate. J. Agr. Food Chem. 17: 1080-1084. Tonso, N. L., V. G. Matheson, and W. E. Holben. 1995. Polyphasic characterization of a suite of bacterial isolates capable of degrading 2,4-D. Microb. Ecol. 3:3-24. Top, E. M., W. E. Holben, and L. J. Forney. 1995. Characterization of diverse 2,4-dichlorophenoxyacetic acid-degradative plasmids isolated from soil by complementation. Appl. Environ. Microbiol. 61: 1691-1698. Top, E. M., O. V. Maltseva, and L. J. Forney. 1996. Capture ofa catabolic plasmid that encodes only 2,4-D/oc-ketoglutarate dioxygenase (deA) by genetic complementation. Appl. Environ. Microbiol. 62:2470-2476. Unfinished Genome Database Website. 11 October 1999. http://www.ncbi.nlm.nih.gov/BLAST/ unfmishedgenomehtml. United States Department of Agriculture. 1996. Biologic and economic benefits from the use of phenoxy herbicides in the United States NAPIAP Report No. 1-PA-96. Valegard, K., A. C. T. van Scheltinga, M. D. Lloyd, T. Hara, S. Ramaswamy, A. Perrakis, A. Thompson, H.-J. Lee, J. E. Baldwin, C. J. Schofield, J. Hajdu, 45 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. and I. Andersson. 1998. Structure of a cephalosporin synthase. Nature 394:805- 809. Vallaeys, T., L. Albino, G. Soulas, A. D. Wright, and A. J. Weightman. 1998. Isolation and characterization of a stable 2,4-dichlorophonoxyacetic acid degrading bacterium, Variovorax paradoxus, using chemostat culture. Biotechnol. Lett. 20:1073-6. Vallaeys, T., L. Courde, C. McGowan, A. D. Wright, and R. R. Fulthorpe. 1999. Phylogenetic analyses indicate independent recruitment of diverse gene casettes during assemblage of the 2,4-D catabolic pathway. FEMS Microbiol. Ecology 28:373-382. Valtavaara, M., H. Papponene, A.-M. Pirttilii, K. Hiltunen, H. Helander, and R. Myllylii. 1997. Cloning and characterization of a novel human lysyl hydroxylase isoform highly expressed in pancreas and muscle. J. Biol. Chem. 272:6831-6834. van der Meer, J. R., W. M. de Vos, S. Harayama, and A. J. B. Zehnder. 1992. Molecular mechanisms of genetic adaption to xenobiotic compounds. Microbiol. Rev. 56:677-694. Vermeij, P., J. Hummerjohann, T. Leisinger, and M. A. Kertesz. 1999. Unpublished. Walker, R. L., and A. S. Newman. 1956. Microbial decomposition of 2,4- dichlorophenoxyacetic acid. Appl. Microbiol 4:201-207. Wang, Q., W. J. VanDusen, C. J. Petroski, V. M. Garsky, A. M. Stern, and P. A. Friedman. 1991. Bovine liver aspartyl B—hydroxylase. Purification and characterization. J. Biol. Chem. 266: 14004-14010. Wehbie, R. S., N. S. Punekar, and H. A. Lardy. 1988. Rat liver y-butyrobetaine hydroxylase catalyzed reaction: influence of potassium, substrates and substrate analogues on hydroxylation and decarboxylation. Biochemistry 27 :2222-2228. Whiting, A. K., L. Que J r., R. E. Saari, R. P. Hausinger, M. A. Fredrick, and J. McCracken. 1997. Metal coordination environment of a Cu(H)-substituted Ot- keto acid-dependent dioxygenase that degrades the herbicide 2,4—D. J. Amer. Chem. Soc. 119:3413-3414. Wilson, R. D., J. Geronimo, and J. A. Armbruster. 1997. Dissipation in field soils after application of 2,4-D dimethylamine salt and 2,4-D 2-ethylhexyl ester. Environmental Toxicology and Chemistry 16: 1239-1246. 46 133. 134. 135. Xia, X., J. Bollinger, and A. Ogram. 1995. Molecular genetic analysis of the response of three soil microbial communties to the application of 2,4-D. Mol. Ecol. 4:17-28. Xia, X.-S., S. Aathithan, K. Oswiecimska, A. R. W. Smith, and I. J. Bruce. 1998. A novel plasmid pIJBl possessing a putative 2,4-dichlorophenoxyacetate degradative transposon TN5530 in Burkholderia cepacia strain 2a. Plasmid 39: 1 54-1 59. Zhang, Z., C. J. Schofield, J. E. Baldwin, P. Thomas, and P. John. 1995. Expression, purification and characterization of l-aminocyclopropane-l- carboxylate oxidase from tomato in Escherichia coli. Biochem. J. 307 :77-85. 47 CHAPTER 2 CHARACTERIZATION OF THE ACTIVE SITE OF 2,4- DICHLOROPHENOXYACETIC ACID/a-KETOGLUTARATE DIOXYGENASE Many of the results in this chapter have been submitted for publication to the Journal of Biological Chemistry in a manuscript entitled “Site-directed mutagenesis of 2,4-dichlorophenoxyacetic acid/a-ketoglutarate dioxygenase” by Hogan, D.A., S. R. Smith, E. A. Saari, J. McCracken, and R. P. Hausinger. 48 Introduction 2,4-Dichlorophenoxyacetic acid (2,4-D)/oc—ketoglutarate (cc-KG) dioxygenase (deA) is an Fe(H)— and oc-KG—dependent enzyme that catalyzes the first step in degradation of the herbicide 2,4-D. This enzyme couples the oxidative decarboxylation of a-KG to the hydroxylation of a side—chain carbon atom. The resultant hemiacetal spontaneously decomposes to form 2,4-dichlorophenol and glyoxalate (6). Mechanistically, deA resembles numerous other OL-KG—dependent dioxygenases from plants, animals, fimgi, and bacteria that catalyze similar hydroxylation reactions at unactivated carbon centers (4, 28). The best—studied oc-KG-dependent hydroxylases, including prolyl and lysyl hydroxylase (14) and flavanone hydroxylase (8, 28), show sequence identity to isopenicillin N synthase (IPNS) and thus fall into Group I of the oc- KG-dependent dioxygenase superfamily (defined in Chapter 1). These enzymes have an H-X-D-Xss-H-Xlo-R motif in common (2, l8) and site-directed mutagenesis studies have confirmed the importance of these residues (15, 18, 21, 25, 27). Through various biochemical and biophysical studies of Group I enzymes, these amino acids are implicated in creating the F e(H) and oc-KG binding sites (17, 29, 30, 33). The sequence of deA does not have significant overall identity to the sequences of Group I enzymes. Although deA is not similar in sequence to the Group I oc-KG-dependent dioxygenases, it is clearly homologous to a number of other enzymes (Group II). Sequences of Escherichia coli taurine/(x-KG dioxygenase (TauD) (5) and sulfonate/oc-KG dioxygenase from Saccharomyces cerevisiae (11) have 25-30% overall identity to deA. Alignment of deA, TauD and sulfonate/oc-KG dioxygenase indicates the conservation of 49 an H-X—D motif (His 113 and Asp 115 in deA) and three other histidines (Appendix A). One invariant histidine, His 167, is positioned to create a three amino acid arrangement similar to the H—X—D-Xss-H configuration of metal ligands in Group I enzymes, but there are no good candidates for a residue corresponding to the conserved arginine in the Group I motif. In addition, sequence comparisons between deA and more distantly related Group II enzymes, such as y-butyrobetaine hydroxylase and clavaminate synthase, do not show conservation of a histidine corresponding to His 167. Alignment of Group II dioxygenase sequences indicates that His 262 is absolutely conserved and is in close proximity to an invariant arginine which may be analogous to the OL-KG-binding arginine in deacetoxycephalosporin synthase (DAOCS) and other Group I enzymes. In this study, we used site-directed mutagenesis methods to examine the roles of the all nine histidines and Asp l 15 in deA. Previously published work showed that deA was inactivated by diethylpyrocarbonate, a histidine-selective reagent, and provided evidence consistent with the presence of multiple histidines in the active site (7). Spectroscopic studies of deA showed the presence of two equatorially—bound imidazole nitrogens as ligands to the active site metal, and indicated that one imidazole ligand may be displaced or shifted to an axial position upon substrate binding (3, 34). Based on analyses of different deA variants, I identify several likely metal—binding residues and provide evidence that another one or two histidines may aid in substrate coordination. 50 Experimental Procedures Recombinant Plasmids All plasmids were constructed from pUS311 (7), a pUCl9 derivative that contains the Ralstonia eutropha Jl\/[P134 tfdA gene (Figure 4). The H8A, Dl 15A, H213A, H216A, H235A, H245A, and H262A deA variants were created by direct mutation of tfdA in pUS311 by the Stratagene Quickchange System (Stratagene, La Jolla, CA). All mutagenic primers are listed in Table 4. Two alternative approaches were used to construct the three remaining variants. Plasmids encoding H113A and H167A deA variants were created by Clontech mutagenesis of pXHtfdA, a pUC19 plasmid containing the 5’-XbaI-HindII fragment of the tfdA gene (Figure 4). To create the complete gene containing the indicated mutations, the XbaI—HindII fragment was cloned into pHthdA, which contains the 3’ end of the tfdA gene. pHthdA was constructed in two steps. First, the 1.4—kb XbaI-Sall fragment from pUS311, containing the complete tfdA gene, was cloned into pBC KS' (Stratagene) cut with XbaI and XhoI to create pBthdA. This step had the benefit of eliminating a HindH site that interfered with further cloning steps. The 727-bp HindH-Kpnl fragment of pBthdA was subcloned into pBC KS‘ cut with the same enzymes to give pHthdA. Similarly, the gene encoding H200A deA was made by mutagenesis of pXthdA, a pUC19 plasmid containing the 5’-XbaI—Xhol fragment of tfdA. The altered XbaI-XhoI fragment was then inserted into pHthdA cut with Xba I and XhoI to give the complete H200A tfdA gene. The identity of 51 Figure 4. Restriction maps of the plasmids used in construction of tfdA mutants. Abbreviations for the restriction enzyme sites are as follows: H, HindH; X, Xbal; Xh, XhoI; Sl, SalI; K, Kpnl; So, Sad; N, Nrul. The vector backbone is indicated in the box on the left. The tfdA open reading frame is shown as a wide, solid line. x N H Xh N SIfH so pUC19 I I I I I I pUS311 x N H Xh N SI/Xh ch KS I I I I I I I pBthdA X N H x N H Xh @' I '4 pXHtfdA HthdA ch KS I I I H p Table 4. Sequences of mutagenic primers used to create altered tfdA genes Mutant Primer Sequencea H8A 5’- CGC AAA TCC CCT TGC TCC TCT TTT CGC C —3’ H113A 5’- GTC GCT GGC CCA AGC TGG-3’ D115A 5’—GCA CAG CGC CAG CTC CTT TCA-3’ H167A 5’-GCG TGC CGA GCA GTA CGC ACT G—3’ HZOOA 5’—GGT TCG AAC CGC CGC CGG CTC—3’ H213A 5’-GCT CGC GGC CGC GCC GAT-3’ H216A 5,-CCT TCG ACG GCG CTC GCG—3’ H235A 5,-GCT TCT CGA GGC GAC ACA G—3’ H245A 5’-GTG TAC CGG GCT CGC TGG AAC-37 H262A 5’—CGT TCT TGC ACG CGG ACG CAG—3’ dea—MBPF 5’—TCT CTA GAG TGA GCG TCG TCG CAA ATC C-3’ dea-MBPR 5’-GTC AAG CTT GGT TGC GTA CAT CTT GTG G-3’ 21The reverse primer used for the creation of all mutants was the complement of the forward primer. 52 all final constructs was confirmed by sequence analysis. To insert the genes encoding H167A and H262A variants of deA into a plasmid that would allow for IPTG-controlled expression, the XbaI—Sall fragments from the corresponding plasmids described above were cloned into pET23a (N ovagen) prepared with the same enzymes. To create the maltose-binding protein (MBP)-deA fusion proteins, the wild-type tfdA gene was amplified from pUS311 with deA-MBPF and deA-MBPR primers (Table 4) to create an Xbal site directly upstream of the GTG start codon of the tfdA gene and a HindIII restriction site 54 bp downstream of the stop codon. The PCR product was cloned directly into the pGEM-T vector according to the manufacturer’s instructions (Promega, Madison, WI). The XbaI-HindIII fragment was isolated from the resulting plasmid and cloned into the pMAL-cZ vector that had been digested with the same enzymes. The identity of the newly created malE—tfdA gene fusion was confirmed by sequencing. Substitution of the mutation-containing internal NruI fragment for the same fragment of the wild-type malE-tfdA gene created MBP-fusion forms of altered deAs. First, pMAL—tfa’A was digested with Nrul and the vector fragment was purified and religated to create pMAL-t‘fa’AANruI. The resultant plasmid was linearized with NruI and dephosphorylated with calf intestine alkaline phosphatase prior to ligation with the NruI fragments isolated from the previously described mutant genes. Constructs were confirmed by restriction analysis. Protein Purification H8A, H1 13A, D1 15A, H216A, H235A and wild—type deA proteins were purified from E. coli DHSa cells carrying pUS311 and its mutated derivatives according to a previously described protocol (7). In addition, the non-mutated enzyme and the deA 53 variants H167A, H200A, H213A, H245A and H262A were purified as MBP-deA fusion proteins from E. coli DH50t by the protocol described in the pMAL Protein Fusion and Purification System Manual (New England Biolabs, Beverly, MA). Analysis of Kinetic Parameters Specific activities of the wild-type and variant deA proteins were determined by a previously described spectrophotometric assay (7). The typical assay mixture contained 1 mM 2,4-D, 1 mM Ot-KG, 100 uM (NH4)2Fe(SO4)2, and 100 uM ascorbic acid in 10 rnM MOPS buffer (pH 6.75) at 30°C. The reactions were quenched by the addition of EDTA to a concentration of 5 mM. 2,4-Dichlorophenol was quantified by reaction with 4-aminoantipyrene followed by measurement of the absorbance at 510 nm. One unit of activity was defined as the amount of enzyme required to produce one umol of dichlorophenol-min". Protein concentrations were determined using the BioRad Protein Assay with bovine serum albumin as a standard. For calculation of the kcat values, the deA variants were assumed to have M, = 31,600 and the MBP-deA variants were assigned Mr = 74,500. The low Km for or-KG (~2-5 uM for the wild-type enzyme) precluded use of the 4-aminoantipyrene assay for accurate determination of this value. The alternative method used to measure the Km for oc-KG quantified the amount of 14C02 liberated from a—[l— 14C]-KG during the course of the reaction (7). 54 Native Protein Analysis by Gel Filtration Size exclusion chromatography was used to estimate the native molecular weights of deA, MBP—deA, and mutant proteins. The proteins were chromatographed on a Superose 6 gel filtration colurrm (1.0 x 30 cm, Pharmacia-Upjohn) in 20 mM Tris buffer (pH 7.5), 1 mM EDTA, and 200 mM NaCl at a flow rate of 0.2 ml°min‘1. The elution volumes were compared to those for gel filtration standards (Bio-Rad) including thyroglobulin, 670 kDa; bovine gamma globulin, 158 kDa; chicken ovalbumin, 44 kDa; myoglobin, 17 kDa; and vitamin B12, 1350 Da. Spectroscopic Analysis Proteins for electron paramagnetic resonance (EPR) and electron spin-echo envelope modulation (ESEEM) spectroscopic analyses were exchanged into 25 mM MOPS (pH 6.75) by repeated concentration and dilution in Centricon 30 (Amicon) centrifugal concentrators. The final subunit concentration was 0.5 mM for the non—fusion forms of deA and 0.4 mM for the MBP-deA proteins. CuClz was added to a concentration of 450 uM and 350 uM, respectively. Buffered solutions of oc-KG and 2,4- D were added to final concentrations of 5 mM. Glycerol was present at 40% in all samples. X—band EPR spectra were obtained at 77K on a Bruker ESP-300E spectrometer. ESEEM data were collected on a home-built spectrometer; the microwave bridge of this instrument has been previously described in detail (20). Data collection and analyses were controlled by a Power Computing model 200 Power PC using software written with LabView v. 5.01 (National Instruments). Electron spin echoes were digitized, averaged, 55 and integrated by a Tektronix model 620B digital oscilloscope interfaced to the spectrometer computer via an [BEE-488 bus. Two four-channel delay and gate generators (Stanford Research Systems model DG535), a Bruker BH-15 magnetic field controller, and a Hewlett-Packard model 8656B radiofrequency synthesizer were also interfaced using [BEE-488 protocol. Data were collected using a reflection cavity that employed a folded rnicrostrip resonator (16). A three-pulse stimulated echo sequence (90°-t-90°-T- 90°) was used. ESEEM spectra were generated by Fourier transformation of the time— domain data using dead time reconstruction (23). Simulations of the experimental data were performed on a Sun SparcII work station. Simulation programs were written in FORTRAN and based on the density matrix formalism developed by Mims (22). Software for the frequency analysis of the experimental and simulated data was written in Matlab (Mathworks, Natick, MA). Results Production of the mutant T fdAs Initially, all of the mutant genes were expressed from their pUC19-based plasmids except for tfdA genes encoding H113A, H167A and H200A deA, which were in pBC KS‘-derived plasmids. Using the standard protocol to produce soluble, wild-type deA (growth at 30°C to early stationary phase), only the H8A, H113A, D1 15A, H216A, and H23 5A variants existed as soluble proteins. All of the other deA variants were present as inclusion bodies even when grown at lower temperatures (22°C), in M9 minimal medium, or in LB broth containing 660 mM sorbitol and 2.5 mM betaine (1). In addition, 56 IPTG-controlled production of H167A and H262A proteins from mutant genes cloned into pET23a did not yield soluble samples even when the harvested cell pellets were suspended in buffer containing 20% glycerol to limit protein aggregation. To overcome the solubility problems for the five deA variants, MBP-deA fusion proteins were created. Wild-type deA and the MBP—deA fusion protein had essentially identical km and very similar Km values for oc-KG and 2,4-D (Table 5). A slight increase in the apparent KD F e(II) may reflect some metal binding capacity of MBP. Since the presence of the fusion protein did not appear to greatly affect the kinetic parameters of wild-type enzyme, similar fusion proteins were created for the H167A, H200A, H213A, H245A, and H262A deA variants. Kinetic analyses of altered T fdAs Results from kinetic analyses of the four active mutant proteins are summarized in Table 5. H8A deA was soluble and active, but was rapidly proteolyzed to an inactive form. By electrophoretic comparisons, the cleavage site appeared to be the same as in wild-type deA (between Arg77 and Phe78) (7). The rate of proteolysis of H8A deA was enhanced compared to that seen for the wild-type enzyme despite the presence of EDTA and protease inhibitors in the purification buffer. Because purified H8A deA was more than 75% degraded, the catalytic rate constant was calculated on the basis of the estimated amount of intact enzyme. These calculations indicate rates and Km values similar to those for the wild-type enzyme. Similarly, the kinetic parameters for H23 5A deA were comparable to the native enzyme. In contrast, two variants exhibited differences from wild-type enzyme in their kinetic parameters. The H213A MBP-deA 57 variant exhibited a twenty-fold reduction in kcat and a ten-fold increase in Km for 2,4-D. In addition, H216A deA had a modest (2.5-fold) increase in the Km for 2,4-D and no change in catalytic rate. The other kinetic parameters for HZ 1 3A MBP-deA and H216A deA (Km for a—KG and KD for ferrous ion) did not differ significantly from the wild-type values. Six soluble deA variants (H113A, D115A, MBP-H167A, MBP-H200A, MBP- H245A and MBP-H262A) exhibited no activity even when assayed with elevated substrate and cofactor concentrations (10 mM oc-KG, 5 mM 2,4-D, 250 uM Fe(H)). Table 5. Summary of the kinetic parameters for active deA variants deA Sample kcat Km aKG K.,, 2,4-D KD Fe(II) min" 11M M M wild-type 442 i 33 4.9 i 0.73 19.3 i 3.7 2.0 i 0.5 MBP-fusion 411 i 23 9.9 i 1.8 33.7 i 2.5 15.2 i 5.4 H8Aa 284i6.9 6.51L 1.8 17.7:38 1.9:r 0.7 MBP-H213A 22.1 i 2.9 8.3 i 2.9 318 i 44 27.7 i 9.8 H216A 474 i 47 6.05 i 2.6 52 i 5 2.4 i 1.7 H235A 284 i 25 9.6 i 1.6 34 i 8 1.4 :1.7 a More than 75% of the protein was present as the degradation product. The kcat was estimated from the active, full-length fraction. Evaluation of the Structural Consequences of the Mutations To assess whether the inactive mutant proteins assumed conformations similar to the wild-type enzyme, their apparent molecular weights were estimated by gel filtration analysis. The observed size of wild-type deA was found to be 51 kDa by comparison to protein standards, suggesting that deA forms a compact dimer or an elongated monomer. The elution volume for both H113A and D115A corresponded exactly to wild- type deA indicating that these proteins are not significantly altered in their quaternary structure. MBP-deA eluted both in the void volume (approximately 25 % of the protein) and at a position corresponding to 216 kDa (roughly 75 % of the protein), suggesting that 58 MBP-deA forms at least a dimer. Because each MBP-deA subunit is comprised of two domains separated by a thirteen amino acid linker, the resultant protein may migrate with a larger apparent molecular weight. MBP-H167A and MBP-H200A samples {demonstrated the same two-peak profile as MBP-deA, but with larger proportions eluting in the void volume. MBP-H245A and MBP-H262A proteins were soluble, however, gel filtration analysis indicated the presence of only highly aggregated material eluting in the void volume. Because these mutant proteins exhibited aberrations in their folding properties, the catalytic role of His 245 and His 262, if any, could not be assessed. EPR Spectroscopic Characterization of Variants with Altered Metal Sites The deA metallocenter properties were examined by EPR spectroscopy. For the purposes of these experiments, the diamagnetic Fe(II) metallocenter of deA was substituted with spectroscopically active Cu(H) ion. Although the Cu(H) form of deA is inactive, Cu(H) binds competitively with respect to F e(H) (K, = l-3 uM) and Cu(H)- substituted deA has been used previously to study the metal coordination environment of this enzyme in the presence and absence of substrates (3, 34). Spectral parameters of wild-type Cu(H)-deA, Cu(H)-deA + oc-KG, and Cu(H)-deA + oc-KG + 2,4-D (Figure 5A and Table 6), agreed well with those reported previously (34). Earlier studies of Cu(H)-substituted wild-type deA indicated that the metal was bound in a tetragonal environment with a mixture of O and N ligands in the equatorial plane. Upon addition of oc-KG and 2,4—D, the spectral parameters were altered to a more rhombic signal with accompanying resolution of ligand hyperfine coupling (Figure 5A). These results suggest that binding of the cofactor and substrate to the enzyme leads to a better-defined copper 59 Figure 5. X-band CW-EPR spectra of Cu(H)-substituted deA variants. EPR spectra were obtained at 77 K for the following samples (substrate concentrations were 5 leI, when present). A, (Top to bottom) Wild-type Cu-deA in the absence of substrates, with or-KG, and with a—KG + 2,4-D. B, H113A Cu-deA in the absence of substrates, with a-KG, and with a-KG + 2,4-D. C, D115A Cu-deA in the absence of substrates, with oc-KG, and with oc-KG + 2,4-D. D, MBP-H167A Cu—deA in the absence of substrates, with a-KG, and with oc-KG + 2,4-D. E, MBP-H200A Cu-deA in the absence of substrates and in the presence of OL-KG + 2,4-D. Spectral parameters are listed in Table 6. 60 Figure 5. X-band cw-EPR spectra of Cu(II) substituted deA variants A wt deA ’/ gt Cu only Cu+aKG 811 I Cu+aKG+2AJD L__J A11 2500 2750 3000 3250 3500 3750 Gauss B H113A deA :- 2500 2750 3000 3250 3500 3750 Gauss C D115A deA Cu only Cu+aKG Cu+aKG+2A{) 2500 2750 3000 3250 3500 3750 Gauss 61 Figure 5. (con’t) D H167A lVIBP-deA Cu only Cu + OtKG + 2,4-D 3250 3500 3750 E H200A IVIBP-deA Cu + ot—KG + 2,4-D 2500 2750 3000 3250 3500 3750 Gauss 62 Table 6. Summary of the EPR spectral parameters for Cu(H)-substituted deA variants. 811 A11 g1 mT wild-type deA 2.34 16.3 2.07 wild-type deA/oc—KG 2.36 15.1 2.07 wild—type deA/a—KG/2,4—D 2.38 12.0 2.09 H113A 2.30 16.6 2.07 H113A/0L-KG 2.33 15.8 2.07 H113A/0t-KG/2,4-D 2.35 14.0 2.08 D115A 2.30 17.0 2.06 D115A/0t—KG 2.30 16.2 2.06 2.34 16.8 2.06 D115A/0t-KG/2,4-D 2.30 16.2 2.06 2.34 16.8 2.06 MBP—H167A 2.30 17.1 2.07 lVfBP-Hl67A/ot—KG/2,4-D 2.30 17.1 2.07 MBP-H200A 2.30 17.1 2.07 2.34 16.1 2.07 MBP-H200A/0t-KG/2,4-D 2.30 17.1 2.07 2.34 16.1 2.07 63 site in the enzyme (34) and the decreased values for A” upon addition of substrates (AH =12.0 mT) indicated a significant distortion from planarity. The effects of the amino acid substitutions on metallocenter structure in the deA variants w1th quaternary structures s1milar to the corresponding wild— type protein (H113A, D1 15A, MBP-H167A and MBP-H200A deA) were assessed by EPR spectroscopy. Because there were no significant differences between the spectra of MBP- deA and the non—fusion wild-type form of the protein (data not shown), H167A and H200A deA were analyzed as MBP fusion. In all cases, EPR spectra of the Cu(II)— substituted deA variants showed contributions from multiple copper species. The mixture of copper centers may have resulted from either copper interactions in other regions of the protein or the presence of multiple binding modes at the active site. The EPR spectra for Cu—H113A deA (Figure 5B and Table 6) differed significantly from spectra of wild-type enzyme. Irnportantly, these spectra showed clear changes in Cu(II) g-values and hyperfine tensor principal values upon substrate additions consistent with the binding of copper to the active site pocket. The broadening of the EPR signal in the gH region upon addition of or-KG suggested a mixture of copper coordination environments consistent with a less tightly bound metal site. The EPR spectrum of D1 15A deA (Figure 5C) had parameters similar to those observed in the copper-substituted wild-type protein (Table 6), although with less resolution of A“. Addition of oc-KG (in the presence or absence of 2,4-D) had a dramatic effect on the appearance of the D115A data, enhancing resolution of the Cu hyperfine peaks at g“ and the ligand hyperfine structure in the g i region (Figure 5C). The appearance of these superhyperfine interactions may have resulted from a subtle shift in 64 the orientation of the principal g—tensor such that the imidazole ligands occupied an increasingly equatorial position (31). These results were consistent with the formation of a tighter or more regular copper binding site upon addition of the cofactor. Again, the perturbations of the spectrum upon a—KG addition were consistent with copper binding to the active site. MBP-H167A deA (Figure 5D) and lVIBP-H200A deA (Figure 5E) EPR spectra were poorly resolved and showed evidence for multiple modes of copper binding. Significantly, few changes were seen upon addition of either oc—KG or 2,4-D. The MBP— H200A EPR spectrum (Figure SE) was more defined than that of H1 13A or MBP— H167A, and clearly showed a second set of resonances of lower amplitude with parameters identical to those observed for the copper-substituted wild—type enzyme (Table 6). The broadness of the resonances in the MBP-H167A sample suggested that a similar situation may occur in that enzyme as well. ESEEM Spectroscopic Characterization of Variants with Altered Metal Sites ESEEM spectroscopy was used to gain additional insight into the number of nitrogen ligands that create the copper centers in the deA variants. As reported previously, ESEEM analysis of wild-type Cu(H)—deA indicated the presence of two equatorial histidines directly coordinated to the metal (34). The Fourier transformed 3- pulse ESEEM data, collected near g“, for wild-type Cu(II) deA with oc—KG showed sharp peaks at 0.6, 0.9, and 1.5 MHz and a broad feature at 3.5 MHz (Figure 6). This signature is typical of copper bound by an equatorial imidazole. The narrow combination bands at 2.1, 2.5, and 3.1 MHz (Figure 6) indicated the presence of more than one histidine per 65 copper site (19, 20, 24). Simulations of the ESEEM data for the wild-type enzyme plus a-KG, generated with the density matrix approach of Mims (22) and the parameters for copper coordinated by two identical equatorial histidines, are superimposed on the experimental data in Figure 6. ESEEM data for Cu(H)—deA in the presence of or—KG plus 2,4-D were analyzed by a similar method (data not shown). A substantial decrease in the modulation intensity and disappearance of the combination bands (34) upon addition of both substrates is consistent with the loss of one histidine from the equatorial plane. Similar analyses of Cu(H)—substituted deA variants were used to compare their metallocenter structure to that of the wild—type enzyme. Three pulse ESEEM patterns were generated for the H113A, MBP-H167A, and D115A variants of deA (Figure 7) as well as for the MBP-H200A samples (data not shown). To normalize the data, the amplitudes in the spectra for all four mutants were corrected to account for differences in baseline frequencies. Samples for each deA variant gave rise to low frequency modulations indicative of equatorially—bound histidine ligands. The ESEEM spectra from H113A (Figure 7A) and MBP-H200A (not shown) variants were nearly identical and show modulations that were considerably weaker than those measured for MBP—H167A (Figure 7B) and D1 15A (Figure 7C). As for wild-type spectra, the data were simulated with a program based on Mims' density matrix formalism, assuming a system of spherical symmetry. In each case, the ratio of modulated signal intensity to unmodulated signal intensity was suggestive of copper bound in more than one coordination environment. The differences in strength of copper binding to the active site can be visualized in Figure 7 by comparing of the peak heights of a given spectrum (modulated frequency) to the background signal that is observed at the end of data collection upon shift to the direct 66 0.8 8 0.6’ , 3 E- 0 4 < ' — _ 0.2 i 0 — 1 12 l l l I I I I l l l l a) ‘C 3 E _ < /\ 7 '_ - u. .I \ NINA/WW V‘s/Ivar, BWMi/LMW‘KV ‘-'. . . l . i . 1 i . i l . . . I i _ 0 2 4 6 8 10 frequency (MHz) Figure 6. 3-Pulse ESEEM spectra of Cu-substituted wild-type deA with 2-histidine simulation data. ESEEM spectrum of wild-type Cu(H)-deA + or-KG (solid) and computer—simulated data for Cu(H) bound by two identical histidines (dashed) (A) and the corresponding frequency domain spectrum (B). Parameters for data collections were magnetic field = 2800 G, t = 375 ns, T = 50 ns, v = 8.98 GHz; 4.2 K. Cu hyperfine parameters: g“: 2.05, gz=2.34, Axx=Ayy=20 MHz, A2: 487 MHz, Superhyperfine parameters: gN = 0.40347, A150: 1.7 MHz,radius= 2.75 A, [0,¢]=[2.37,2.07] radians, e2qQ = 1.60 MHZ, n = 0.75, NQI [0t,[3,y]=[ .50,1.1,0.1] radians. The simulated data were treated with an exponential ((_rs+ T‘) 3500 background decay fimction was applied to allow the amplitudes of the initial 1.5 usec of the simulations to match the data. decay fiinction to give y'jj = (y — 0. 17) * exp + 0.17 For each simulation, 3 67 current (unmodulated frequency). Although this ratio was less than the wild-type value for all of the deA variants, H113A and MBP-HZOOA had a markedly decreased capacity to constrain copper to one geometry in the absence of substrates. Fourier transformations of the data shown in Figure 7 yielded major features at 0.7, 1.5, and 4.0 MHz indicative of histidines equatorially coordinated to Cu(H) (24). Although Cu—bound histidines were visible in the H113A and MBP-H200A spectra, the ESEEM signals were too weak to be consistent with one dominant Cu(II) conformation, and thus were omitted from further analyses. The frequency and Fourier transformed data, with the corresponding simulations, are shown for D115A (Figure 8) and MBP- H167A deA (Figure 9). In both cases, the large unmodulated component of the data required a severe decay function in order to simulate the data. In addition, it was necessary to use the average of several simulation data sets with varying isotropic hyperfine coupling constants to compensate for the presence of multiple copper binding environments identified by EPR spectroscopy. Although the combination lines were weaker in the data for D115A than for wild-type deA, the combination line at approximately 3.0 MHZ was observed and its frequency was predicted nicely by the 2 His simulation data. Thus, these data were most consistent with two histidine ligands to the Cu(II) site in this protein. The broadness of the Fourier transformed spectrum for D115A deA and the subsequent difficulty in observing the combination lines most likely reflects a distribution of 14N superhyperfine couplings from multiple Cu(H) centers in D115A deA which is consistent with our use of an average of several isotropic hyperfine coupling constants to simulate the data. 68 I I I I I I I I I l I I I I l l l I I I I (”Va .... M.,, A - "MW”r-rM»Mimrwm. _ s A - 3 E \ ,va 4 8‘ IIW/W W., W’W'I‘rtw B __ < WWW I 8 ' IIWflIntrIW/sim WWW “twnmwn ‘ .5 ” “W WWW -‘\ ‘ To 0. 4 -- IF WW WWI - E - c - o _ _ z - __ 0. 2 -— _ O _ l l l i l l l i 1 l l i l l l l l l i L: 0 2 4 6 8 10 r+T (usec) Figure 7. 3-Pulse ESEEM time-domain spectra of deA variants. Spectra for H113A Cu-deA (A) (magnetic field = 3050 G, r = 300 ns, T = 40 ns); MBP- H167A Cu-deA (B) (magnetic field = 3100 G, r = 300 ns, T = 40 ns) and D115A Cu- deA (C) (magnetic field = 3050 G, r = 385 ns, T = 55 ns, v = 8.8 GHz) are compared. All were collected at 4.2 K. 69 _ , I'IT , I I I I l I I l I I I I I I I I _ a) j I I go.8_Il,III A — -*—'—' ” 1 1 I: I.‘ - a i III! I. . — E 0-6 _— III: 1,// \ ”V” j 'C) _1IfIII1 ‘/ ,\ 143% ““WWMN.W W a 04 _— I ., -—--wwmm —_ E I i. 0.2 '— L O _ 4 Z — I 0 — l l l I I l I I I. L I 1 l L I l 1 1 I 0 2 8 1O T+T(IlS€C) 1O LI I I I I I I I I I I 1 +51 - 8 3}, B ‘ EI ‘; E #3 'u III/L ,I, d a 4 i 2,, i E " 1 1'. , - < 2 -_':;\/I I ,11 /~, g E :I/I '1 I Inc? ,/\., 3 o 3' :. a... / Wwww _ \. ‘f’.‘ — -2 _ 1 e 1 I I I ‘1 I J I 1 I i 1 1 I 1 i 1 0 2 4 6 8 10 Frequency (MHz) Figure 8. Spectral simulation of ESEEM data for D115A-Cu-deA. Experimental (solid line), computer-simulated l-His (dashed line), and simulated 2-His (dotted line) (A) and the corresponding time—domain spectra (B) for copper-substituted D115A deA are shown. Magnetic Field = 3050 G, ‘C = 230 ns, T = 50 ns G, v = 8.8 GHZ, 4.2 K, gN = 0.40347, eZqQ = 1.55 MHz, 11 = 0.87, Axx = 1.6 MHZ, Ayy = 1.9 MHz, An = 2.2 MHz. The simulated data were treated with an exponential decay function to give . " (1+ T) 0'8 y” = (y, — 0.05) * exp ( 600 — 0.05. 70 _\ ... I I I I I I I I I I I I I l I I I I _ 3 O 8 ’ ’IIIIII/i 'I A ‘ —Q. HI". , I IN _ ' .' ‘-V..\ i ‘ e ‘ 8 — “I - .E 0.4 _ i _ E : I : §02 _ I _ Z I I 0 — I l I I I I I I I I I I l I I I I I in ~ 0 2 4 6 8 1O T+T (usec) I III I I I I I I I I I I I I I I I I I I I I I I Til Ti r1 I—I T I I I r IE 5:? B 3 a) ; J; ' '- . 1 g I III/I: I ,r/IPI‘“\ _f < . II I r '- 3 I— II Wig .. / \ : LL 1 I I! . 2 7 w. / \\ “ ”ca/‘JA’m/Vi II /( II/i \4/ a wwwwf \/ i l l‘\‘11/- I4 I . I I I . I I I . é . #44} . I . . l I I E 0 1 2 3 4 5 6 7 8 Frequency (MHz) Figure 9. Spectral simulation of ESEEM data for Cu(II) MBP-H167A deA. Experimental (solid line), computer-simulated l-His (dashed line), and 2—His (dotted line) time -domain spectra (A) and the corresponding Fourier transformed data (B) are shown. Magnetic field = 3100 G, I = 300 ns, T = 40 ns, V = 8.8 GHz, 4.2 K, gN = 0.40347, eZqQ = 1.60 MHZ, n = 0.82, AXX = 1.55 MHz, Ayy = 1.9 MHz, An = 2.0 MHz. The simulated data were treated with an exponential decay function to give (I + D)“ y'” = (yfl — 0.03)*eprW —0.03. 71 Simulations of the MBP-H167A deA also suggested the presence of two histidines per copper (Figure 9). The peak broadness in the experimental data resulted in our inability to resolve the features at 0.6 and 0.9 MHz, indicative of a heterogeneous population of copper sites. Simulations achieved by averaging of multiple isotropic hyperfine coupling constants, as in those for D1 15A, predicted the coalescence of these peaks. The number of averaged data sets required to simulate the MBP-H167A spectra indicated that the copper center in this protein was more disordered than that in D1 15A deA. The increased number of unconstrained copper binding environments in MBP- H167A deA may have resulted from non-specific binding to the MBP domain or a small population of misfolded deA protein, as observed in gel filtration experiments. The Cu-substituted forms of H1 13A, D115A, MBP-H167A, and MBP-H200A deA species were also analyzed in the presence of oc-KG and 2,4—D. Although the Cu(H) site in wild-type deA appeared to undergo significant rearrangement upon binding of substrates, only slight variations in the lineshape of the higher frequency peak at ~4.0 MHz were observed for H1 13A, MBP-H167A, and MBP-His 200A. The substrate- dependent changes in the copper site of H1 13A deA upon substrate addition that were observed by EPR spectroscopy were not well defined in the ESEEM spectra, again indicating a less constrained copper site. Addition of a-KG and 2,4-D to D115A deA led to enhanced definition of the D1 15A copper site similar to what was observed by EPR analysis (Figure 5). The ternary complex of Cu(H)-D1 15A deA exhibited coherence in the frequency data at longer times with a slightly increased depth of modulation. These changes were not accompanied by a significant increase in unmodulated intensity (Figure 10) indicating that the copper was predominantly found in 72 - l 3 I I I I I I I I I I l l I I I I l I I .— I z/I‘I/I»m A 2 g :- f\\ ‘F \MmfiiffimMfiWWM-Wowlw'xMWW-fi j a - I I 3 - E ’ In [I I I m If" A\ . l - < ; III \/ \,J \ d1 V k'W’I—xL"kadeMM/MTTVHWVMVAWfl/‘r‘WY’F‘NijvI 8 — /\ I I g I I \d I 3 (U " M l 7 CE) :_ VII A/mu/ WM \MW I. _ Z 3 I : F— l J I I l I J I I l l I l I l I l I L I I-.L O 2 6 8 10 r+T (usec) I I T I I I I I I I l I ' I [A B ‘2 W/ \_\ - AW FT Amplitude lllllll Frequency (MHz) Figure 10. 3-pulse ESEEM of Cu(II) D115A T fdA (4.2K). (Top to bottom) ESEEM spectra for Cu(H) D1 15A deA in the absence of substrate, with 5 mM a—KG, and with 5 mM a-KG + 5 mM 2,4-D (A) and the corresponding Fourier transformed data (B) are shown. Conditions for these measurements were identical to those of Figure 6. 73 l ... a single environment. ESEEM spectra of the D115A deA ternary complex showed resolution of all three combination lines from the electron spin manifold that gave rise to the intense low-frequency peaks (Figure 10B). These data provided convincing evidence for two equatorial histidine ligands to the copper in D1 15A deA in the presence of substrates and supported the conclusions drawn from substrate—free Cu (H)-Dl 15A deA ESEEM spectra. The enhanced intensity and definition upon substrate binding suggested that OI-KG constrained the geometry of the copper at the active site. Discussion Our results demonstrate the importance of multiple histidines and one aspartate for deA activity. EPR, ESEEM, and extended x—ray absorption fine structure spectroscopies of both the Fe(H)- and Cu(H)-forms of wild-type deA had previously indicated that the resting enzyme has two imidazoles and a mixture of bound nitrogens and oxygens as ligands to the metal (3, 34). Changes in the Cu-EPR spectral parameters upon addition of OL-KG are consistent with an increased number of oxygen ligands to the metal, but ESEEM and X-ray absorption data indicated that both imidazole ligands are retained. Addition of 2,4-D, however, leads to the displacement or g-tensor reorientation of a histidine out of the equatorial plane (3, 34). Mutation of His 1 13 or Asp 1 15 to alanine both eliminates activity and greatly alters the structure of Cu(H) metallocenter site. Significantly, however, substrate binding perturbs the spectra of these sites consistent with the metal binding to the active site pocket. These residues comprise the H-X-D motif which is conserved in enzymes related to deA (Appendix A) and common to all other known oc-KG-dependent dioxygenases 74 (2) (Chapter 1). Site-directed mutagenesis studies analogous to those described here demonstrated the importance of this motif in a number of Group I enzymes including aspartyl, lysyl, prolyl, and flavanone hydroxylases (15, 18, 21, 25, 27). Crystal structures of IPNS and DAOCS established these residues as ligands to the metal. My studies support the proposal that residues in the H—X—D motif are ligands to the metal in deA and, most likely, other Group H enzymes. Cu(H)-D115A deA ESEEM spectra indicate the presence of two imidazole ligands, whereas similar analyses of the H1 13A proteins shows evidence of multiple copper binding environments. Clearly, the absence of the aspartate ligand results in a significantly altered binding site, producing a less structurally-defined copper geometry, and contributes to the increased damping of the ESEEM. Addition of oc-KG to this mutant, however, results in increased resolution in the EPR spectrum and ESEEM pattern, indicating that the binding of this cofactor locks the geometry of the active site into place. Together, these data may suggest that the histidines are primarily responsible for Cu(H) binding, and Asp 115 may contribute more to the orientation of the metal. The properties of copper predict a much stronger interaction with nitrogen rather than oxygen ligands; in fact, there are no mononuclear copper enzymes known to have a carboxylate metal ligands (13). Because of differences in the preferred environments of Fe(H) and Cu(H), it is difficult to draw conclusions about the specific electronic role of Asp 115 in the native Fe(H) deA based on these observations. The identity of the second metal binding histidine, predicted by various spectroscopic studies, is less clear. A previously described motif (H-X-D—Xss-H-XIO-R- X-S) common to IPNS and related enzymes (2, 18) would best match His 167 in deA 75 except that the terminal arginine is displaced. Spectral simulations of H167A, however, are consistent with the presence of two imidazole nitrogens per copper in this protein. Furthermore, His 167 is replaced with an arginine in the deA sequence from pH B in Burkholderia cepacia (Appendix A). Thus, it is unlikely that His 167 is one of the three metal ligands necessary to create the metallocenter. EPR and ESEEM analyses of H200A deA suggest that His 200 is important for creation of the metallocenter. Although His 200 is conserved among the enzymes with high identities to deA, such as TauD from E. coli and sulfonate/a-KG dioxygenase from yeast, a corresponding histidine is not found in more distantly related Group H enzymes such as y-butyrobetaine hydroxylase or clavaminate synthase (Appendix A). Mutational analysis of prolyl 4-hydroxylase (25) provided evidence for a non-ligand histidine at active site that is important for a—KG binding and controlling the decarboxylation of OI-KG in the absence of substrate. Perhaps His 200 serves a similar function in deA providing an explanation for the changes in active site structure upon its alteration. Sequence comparisons of all Group H enzymes reveal that His 262 in deA is the only other invariant histidine and I suggest that this is most likely the third metal—binding residue. Supporting my assessment of His 262 as a metal ligand, this residue lies ten amino acids away from Arg 273 creating an H—XIo-R motif similar to that in DAOCS and related enzymes (18). It is surprising, however, that the H262A variant cannot fold properly and forms inclusion bodies when overproduced in E. coli. For comparison, wild-type deA is routinely purified in its apo—enzyme form and deA activity is immediately detected upon addition of Fe (11). Together, these data suggest a more general motif for a—KG-dependent dioxygenases, H-X-D/E—Xn-H-XIO_13- R (the carboxylate occurs as a glutamate in clavaminate synthase, see Appendix A), 76 where Xn varies with the different subgroups within the superfamily (Group I, n=44-65; Group [I n=123-153; Group III, n=74—104). His 213 and His 216 may also be present at or near the deA active site. MBP— H213A and H216A deA variants had increased Km values for 2,4-D and the former enzyme also demonstrated a decreased catalytic rate. These data may indicate that His 213 (perhaps also assisted by His 216) facilitates 2,4-D binding by electrostatic interactions between the imidazole nitrogen and the substrate carboxylate group; however, additional studies are required to test this hypothesis. The importance of the 2.4-D carboxylate for binding to deA is indicated by the inability of phenoxyethanol to serve as a substrate (7). Neither His 213 nor His 216 are conserved among other related oc-KG—dependent dioxygenases. The divergence of sequences in this region may reflect differences in the residues required to bind substrates unique to each enzyme. Our results provide no evidence to indicate that His 8, His 235, or His 245 are critical for catalysis. The H8A variant exhibits enhanced protease sensitivity, but the full- length protein possesses near wild-type activity. Similarly, the H23 5A protein is nearly unaffected in its activity. Neither of these residues is conserved in homologous sequences. As described for the H262A deA variant, His 245 appears to be structurally important. Because H262A and H245A derivatives of deA exist in non—native, highly aggregated conformations, we were unable to assess whether these residues may also play a catalytic role. Although His 245 is not conserved in these sequences, it is in the C- terminal half of the protein which in other a—KG—dependent dioxygenases show a higher level of similarity than the overall sequence comparisons (12). This region might be more constrained for structural reasons. 77 The H—X—D/E-Xn-H motif, which is found in all known oc-KG dependent dioxygenases, is an example of the more general structural arrangement (2- His-1- carboxylate) found in non-heme Fe (H) enzymes including extradiol dioxygenases and pterin-dependent hydroxylases (9). It is proposed that these three ligands anchor the iron in the active site of the enzyme while leaving three open coordination sites available for the binding of substrate, cofactor and oxygen. Several studies suggest that 02 reacts much more readily with such iron centers after the displacement of solvent molecules by more electronegative substrate molecules (9). In addition, it is the diversity of exogenous ligands that enable non-heme Fe(H)-dependent oxygenases to catalyze such a wide range of reactions. Our proposed model of the active site of deA is shown in Figure 11. His 113, Asp l 15 and a third histidine tentatively identified as His 262 are shown as ligands to the metal. As in the crystal structures of both DAOC S and IPNS, water molecules are thought to occupy the three remaining coordination sites in substrate-free enzyme. The model includes a likely binding mode for oc-KG based on the DAOCS crystal structure (3 3). Recent spectroscopic studies of clavaminate synthase (26), TauD (32) and deA (10) are consistent with a similar mode of oc—KG binding in these enzymes. We have no evidence to suggest direct binding of 2,4-D to the metal center, and indicate the presence of a separate 2,4-D binding site adjacent to the metal in the model. Mutation of His 213 affects both substrate binding and the rate of catalysis so I show this residue at the active site to indicate potential substrate interaction, but its precise role is still to be defined. In the absence of a crystal structure, the site—directed mutagenesis, metallocenter spectroscopy, and kinetic analysis studies described here provide the most complete 78 / NH U + \\\ )0 O \‘ O 2000 2 >7Asp115 -ooc l /‘0 o—Fe(n)——N / \ His113 000- N» n :k NH His 262 Figure 11. Model for the active site of deA. The three residues proposed to ligate the ferrous ion of deA are shown along with a possible coordination mode for the substrates, oc-KG and 2,4-D. His 213, implicated in binding 2,4-D, is also illustrated. 79 working model of the deA active site. This model begins to define the necessary elements for catalysis in Group H oc-KG-dependent dioxygenases and finds similarities between these enzymes and other well-studied representatives of this superfamily. 80 Literature Cited Blackwell, J. R., and R. Horgan. 1991. A novel strategy for the production of a highly expressed recombinant protein in an active form. FEBS 295:1-12. Borovok, I., O. Landman, R. Kreisberg-Zakarin, Y. Aharonowitz, and G. Cohen. 1996. Ferrous active site of isopenicillin N synthase: genetic and sequence analysis of the endogenous ligands. Biochemistry 35:1981—1987. Cosper, N. J., C. M. V. Stfilhandske, R. E. Saari, R. P. Hausinger, and R. A. Scott. 1999. X—Ray absorption analysis of Fe(H) and Cu(II) forms of an herbicide-degrading OL-ketoglutarate dioxygenase. J. Biol. Inorg. Chem. 4:122— De Carolis, E., and V. De Luca. 1994. 2—Oxoglutarate—dependent dioxygenases and related enzymes: biochemical characterization. Phytochem. 36:1093—1 107. Eichhorn, E., J. R. van der Ploeg, M. A. Kertesz, and T. Leisinger. 1997. Characterization of oc—ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J. Biol. Chem. 272:23031-23036. Fukumori, F., and R. P. Hausinger. 1993. Alcaligenes eutrophus JMP134 "2,4- dichlorophenoxyacetate monooxygenase" is an oc-ketoglutarate-dependent dioxygenase. J. Bacteriol. 175:2083-2086. Fukumori, F., and R. P. Hausinger. 1993. Purification and characterization of 2,4-dichlorophenoxyacetate/oc-ketoglutarate dioxygenase. J. Biol. Chem. 268:24311-24317. Hedden, P. 1992. 2—Oxoglutarate-dependent dioxygenases in plants: mechanism and function. Biochem. Soc. Trans. 20:373-377. Hegg, E. L., and L. Que Jr. 1997. The 2—His-l—carboxylate facial triad. Eur. J. Biochem. 250:625-629. Hegg, E. L., A. K. Whiting, J. McCracken, R. P. Hausinger, and J. Que, L. 1999. The herbicide-degrading oc-keto acid enzyme deA: metal coordination environment and mechanistic insights. Biochemistry In press. Hogan, D. A., T. A. Auchtung, and R. P. Hausinger. 1999. Cloning and characterization of a sulfonate/Obketoglutarate dioxygenase from Saccharomyces cerevisiae. J. Bacteriol. 181:5876-5879. Jia, S., K. McGinnis, W. J. VanDusen, C. J. Burke, A. Kuo, P. R. Griffin, M. K. Sardana, K. O. Elliston, A. M. Stern, and P. A. Friedman. 1994. A fully 81 20. 21. 22. active catalytic domain of bovine aspartyl (asparaginyl)—B-hydroxylase expressed in Escherichia coli: characterization and evidence for the identification of an active-site region in vertebrate—OL-ketoglutarate-dependent dioxygenases. Proc. Natl. Acad. Sci. 91:7227—7231. Karlin, 8., Z. Y. Zhu, and K. D. Karlin. 1997. The extended environment of mononuclear metal centers in protein structures. Proc. Natl. Acad. Sci. U S A 94: 14225—14230. Kivirikko, K. I., and T. Pihlajaniemi. 1998. Collagen hydroxylases and the protein disulfide isomerase subunit of prolyl 4—hydroxylases. Adv. Enzymol. Relat. Areas Mol. Biol. 72:328-398. Lamberg, A., T. Pihlajaniemi, and K. I. Kivirikko. 1995. Site-directed mutagenesis of the on subunit of the human prolyl 4-hydroxylase. J. Biol. Chem. 270:9926-9931. Lin, C. P., M. K. Bowman, and J. R. Norris. 1985. J. Mag. Res. 65:369-374. Lloyd, M. D., H. J. Lee, K. Harlos, Z. H. Zhang, J. E. Baldwin, C. J. Schofield, J. M. Charnock, C. D. Garner, T. Hara, A. C. Terwisscha van Scheltinga, K. Valegard, J. A. Viklund, J. Hajdu, I. Andersson, A. Danielsson, and R. Bhikhabhai. 1999. Studies on the active site of deacetoxycephalosporin C synthase. J. Mol. Biol. 287:943—960. Lukacin, R., and L. Britsch. 1997. Identification of strictly conserved histidine and arginine residues as part of the active site in Petunia hybrida flavanone 3B- hydroxylase. Eur. J. Biochem. 249:748-757. McCracken, J., J. B. Cornelius, and J. Peisach. 1989. Quantitative aspects of nitrogen-14 electron spin-echo envelope modulation in Cu(H)-proteins, p. 155- 161. In C. P. Keijzers, E. J. Riejerse, and J. Schmidt (ed.), Pulsed EPR: A New Field of Applications. North Holland Publishers, Amsterdam. McCracken, J., D.-H. Shin, and J. L. Dye. 1992. Pulsed EPR studies of polycrystalline cesium hexamethyl hexacyclen sodide. Appl. Mag. Res. 3:305— 316. McGinnis, K., G. M. Ku, W. J. VanDusen, J. Fu, V. Garsky, A. M. Stern, and P. A. Friedman. 1996. Site-directed mutagenesis of residues in a conserved region of bovine apartyl (asparaginyl) B-hydroxylase: evidence that histidine 675 has a role in binding Fe+2. Biochemistry 35:3957-3962. Mims, W. B. 1972. Phys. Rev. B: Solid State 5:2409-2419. 82 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Mims, W. B. 1984. Elimination of the dead-time altifact in electron spin—echo envelop specta. J. Magn. Res. 59:291-306. Mims, W. B., and J. Peisach. 1978. The nuclear modulation affect in electron spin echoes for complexes of Cu2+ and imidazole with 1‘4N and 15N. J. Chem. Phys. 69:4921-4930. Myllyharju, J ., and K. I. Kivirikko. 1997. Characterization of the iron- and 2- oxoglutarate-binding sites of human prolyl 4—hydroxylase. EMBO J. 16:1 173— 1180. Pavel, E. G., J. Zhou, R. W. Busby, M. Gunsior, C. A. Townsend, and E. 1. Solomon. 1998. Circular dichroism and magnetic circular dichroism spectroscopic studies of the non-heme ferrous active site in clavaminate synthase and its interaction with a—ketoglutarate cosubstrate. J. Am. Chem. Soc. 120:743— 754. Pirskanen, A., A.-M. Kaimio, R. Myllylfi, and K. I. Kivirikko. 1996. Site- directed mutagenesis of human lysyl hydroxylase expressed in insect cells. J. Biol. Chem. 271:9398-9402. Prescott, A. G., and P. John. 1996. Dioxygenases: Molecular structure and role in plant metabolism. Annu. Rev. Plant Physiol. Plant Mol. Biol. 47:245-271. Roach, P. L., I. J. Clifton, V. Fiiltip, K. Harlos, G. J. Barton, J. Hajdu, I. Andersson, C. J. Schofield, and J. E. Baldwin. 1995. Crystal structure of isopenicillin N synthase is the first from a new structural family of enzymes. Nature 375:700-704. Roach, P. L., I. J. Clifton, C. M. H. Hensgens, N. Shibata, C. J. Schofield, J. Hajdu, and J. E. Baldwin. 1997. Structure ofisopenicillin N synthase complexed with substrate and the mechanism of penicillin formation. Nature 387:827—830. Rotilio, G., A. F. Agro, L. Calabrese, F. Bossa, P. Guerrieri, and B. Mondovi. 1971. Studies on the metal sites of copper proteins: ligands of copper in hemocuprein. Biochemistry 10:616-621. Ryle, M. J., R. Padmakumar, and R. P. Hausinger. 1999. Stopped-flow kinetic analysis of Escherchia coli taurine/OL—ketoglutarate dioxygenase: interactions with a—ketoglutarate, taurine, and oxygen. Biochem. In press. Valegard, K., A. C. T. van Scheltinga, M. D. Lloyd, T. Hara, S. Ramaswamy, A. Perrakis, A. Thompson, H.-J. Lee, J. E. Baldwin, C. J. Schofield, J. Hajdu, and I. Andersson. 1998. Structure of a cephalosporin synthase. Nature 394:805- 809. 83 34. Whiting, A. K., L. Que Jr., R. E. Saari, R. P. Hausinger, M. A. Fredrick, and J. McCracken. 1997. Metal coordination environment of a Cu(H)-substituted Ot- keto acid-dependent dioxygenase that degrades the herbicide 2,4-D. J. Amer. Chem. Soc. 119:3413-3414. 84 CHAPTER 3 CLONING AND CHARACTERIZATION OF A SULFONATE/oc- KETOGLUTARATE DIOXYGEN ASE FROM Saccharomyces cerevisiae This chapter is based largely on the published article: Hogan, D.A., T.A. Auchtung, and R. P. Hausinger (1999) “Cloning and characterization of a sulfonate/0L- ketoglutarate dioxygenase from Saccharomyces cerevisiae.” J. Bacteriol. 181:5876-5879. 85 Introduction The Saccharomyces cerevisiae genome sequence project revealed an open reading frame, YLL057c, that encodes a protein with 27 % identity to the herbicide-degrading enzyme deA (5) and 31.5 % identity to taurine/oc—KG dioxygenase, TauD (4). The product of YLL057c has lower levels of amino acid sequence identity to other a—KG- dependent dioxygenases such as y-butyrobetaine hydroxylase and clavaminate synthase. Alignment of the deduced amino acids from YLLOS 7c to these four sequences shows conservation of the H-X-D and H-X—R motifs that are signatures for the Group H subset of the a—KG-dependent dioxygenase superfamily that I defined in the Introduction of this thesis. The presence of these residues, which were proven to be important in the site— directed mutagenesis studies of deA described in Chapter 2, provides strong evidence that the product of YLL057c is an OL-KG-dependent dioxygenase. Given the nearly equivalent levels of identity between deA, TauD, and the putative yeast dioxygenase, it is impossible to predict the natural substrate for the yeast enzyme. Although there are no reports of 2,4-D degradation in S. cerevisiae, yeast may contain an enzyme involved in the decomposition of other naturally occurring aromatic compounds such as aromatic amino acids or lignin degradation products. Alternatively, the YLL05 7c product may resemble TauD in its ability to decompose taurine, a naturally occurring sulfonate (6). A number of yeast can use aliphatic sulfonates, such as taurine, cysteate and isethionate, as alternative sulfur sources (17). Sulfonate utilization by S. cerevisiae only occurs under aerobic conditions, is independent of sulfate utilizing enzymes, and requires sulfite reductase, consistent with the formation of sulfite prior to assimilation. This pattern of sulfonate utilization is similar to that in Escherichia coli and 86 other enteric bacteria (16, 18), suggesting the presence of a common pathway in these diverse organisms. In E. coli, TauD hydroxylates the carbon atom in the C-S bond of taurine to give an unstable intermediate that spontaneously decomposes to sulfite and aminoacetaldehyde (4). There are no reports of an analogous enzyme in fungi. In this chapter, I report both the characterization of the S. cerevisiae YLL05 7c gene product and studies with the YLL05 7c deletion mutant to determine whether this enzyme plays a role that is analogous to deA in 2,4-D degrading microorganisms or to TauD in E. coli. Materials and Methods Cloning and expression of YLL05 7c. Open reading frame YLL057c (Genbank Accession # 273162) was PCR amplified and cloned into pET23a for expression in E. coli. The 1236-bp YLL057c sequence, which contains no introns, was amplified from XPM-5392 (ATCC) containing a 30-kb fragment of S. cerevisiae chromosome XH, with primers 5 ’-CAT ATG TAC AGA GGA CGT CGT CGA G -3’ and 5’-CTA CAA CAC TTT TCG TCT CCG AGG- 3 ’. The forward primer contained a 5’ extension that introduced an NdeI restriction site directly upstream of the start ATG codon. Template DNA was added by transferring a small amount of material from a lPM-5392 plaque, obtained by plating on E. coli Y1090 (13), directly to the PCR reaction mix. The PCR reaction mix included 20 mM Tris-HCl (pH 8.4), 50 mM KCl, 1.5 mM MgC12, 25 uM dNTPs (Gibco), 1 uM of each primer and 0.05 U Taq polymerase (Gibco). The temperature program was comprised of an initial denaturation step at 95°C for 2 min 10 sec, 35 cycles of a denaturation step at 95°C for 1 87 min, an annealing step at 55°C for 1 min, a 2 min 10 sec extension step at 68°C, and a final extension at 68°C for 6 min. The resulting 1734-bp PCR product (which included additional downstream sequence) was cloned into pCR2.l TOPO vector (Invitrogen) according to the manufacturer’s instructions to give plasmid pCR2.1—YDO. The cloned gene was cut from pCR2.1-YDO with NdeI and SacI and inserted into pET23a cut with the same enzymes to create pBlO. To create an N—terminal fusion of the YLL057c gene product to the maltose-binding protein (MBP), YLL057c was amplified from pBlO using the protocol described above with primers 5’-TCT CTA GAA TGT CTC CTG CAG CAG C —3’ and 5’-GTC AAG CTT AAA GAA GTG TTG TCG CCG—3’. The forward primer introduced an Xbal site directly upstream of the start site for in—frame insertion into pMAL-c2 (New England Biolabs). The amplification product was directly cloned into pGEM-T (Promega) to give pGTYDOlé. YLL057c was cut from pGTYDOl6 with XbaI and HindHI, and ligated to pMAL-c2 prepared with the same enzymes to create pMBPYDO, a fusion of the malE gene to YLL057c. Production and purification of YLL05 7c gene product. Various efforts to directly express the yeast gene in E. coli DHSOL from pBlO failed to achieve high level production of the desired protein. In contrast, a MBP fusion of this product was produced in E. coli DH50L(pMBPYDO), consistent with an increased stability and decreased toxicity imparted by the N—terminal extension. Cultures were grown to an A600 of 0.4-0.6 at 30 °C in LB containing 100 ug/ml ampicillin, induced with 0.3 mM IPTG, and incubated for three hours. Harvested cells were suspended 20 mM Tris buffer (pH 7.5) containing 1 mM EDTA, 200 mM NaCl, and 1 mM 88 dithiothreitol, and lysed by two passages through a French pressure cell at 120 MPa. The cell lysates were spun at 100,000 x g to yield the clarified cell extracts. The 89.7 kD fusion protein was readily purified by passing cell extracts over an amylose column according to manufacturer’s instructions followed by elution in the above buffer amended with 10 mM maltose. Characterization of MBP-dioxygenase fusion protein. The MBP-dioxygenase fusion protein was assayed for activity towards 2,4-D using the 4—aminoantipyrene assay described previously (5). Activity towards sulfonate- containing compounds was determined by quantifying the amount of sulfite released using Ellman’s reagent (5,5’-dithiobis(2-nitrobenzoic acid)) (4). Standard assay conditions were defined as 10 mM dimethylglutarate (pH 7.0), 1 mM a—KG, and 50 uM Fe(H)-ascorbic acid salt (Sigma) for 5 min at 30°C in the presence of varied substrate concentrations. Consumption of OL-[ 14C—1]-KG was determined using a previously described method (5). One unit is defined as one umol of product formed per min using the standard assay conditions. Studies with the YLL05 7c deletion mutant The S. cerevisiae YLL057c deletion mutant, BY4742-11545 and the corresponding wild type strain, BY4742, were obtained from Research Genetics (Huntsville, AL). Overnight cultures for inoculation of sulfur limited media were grown in yeast peptone dextrose (YPD) broth. Mutant cultures were amended with 50 pig/ml geneticin sulfate antibiotic. To assess the growth of S. cerevisiae BY4742—11545 and 89 BY4742 with alternative sulfur sources, strains were grown aerobically in defined medium (3) with isethionate (250 uM), taurine (250 uM), sulfate (250 uM) or with no sulfur source added. Since geneticin, is sold as a sulfate salt, the antibiotic was omitted from experiments using low sulfur medium. Results and Discussion The product of YLL057c was found to be an F e(H)— and OL-KG-dependent dioxygenase capable of utilizing a variety of sulfonates (Figure 12, Scheme 3) but was inactive on 2,4-D. Maximum rates (Vmax), affinities (Km), and catalytic efficiencies (kcat/Km) for various substrates are compared in Table 7. Whereas taurine was a relatively poor substrate for the yeast sulfonate/oc—KG dioxygenase (with a moderate reaction rate and high Km), isethionate (2—hydroxy-ethanesulfonate) and taurocholate were tumed over more rapidly and bound with much higher affinity by the enzyme. According to their catalytic efficiencies, the best substrates were the synthetic compounds MOPS and N—phenyltaurine. Several of the better substrates for this enzyme had large adducts on the amino group, indicating a relaxed substrate specificity for this portion of the molecule. Of the substrates listed in Table 7, only taurine, MOPS (4) and taurocholate (1) were also substrates of E. coli TauD. Notably, TauD activity toward taurine exhibits a kcat of 133 min-1, which is the same order of magnitude as found here for the yeast enzyme. While taurine utilization followed simple Michaelis-Menten kinetics 90 Q‘NNEI—OH N-phenyltaurine I/\N/\/\E——-OH MOPS ”0 ~ 0“ lsethionic acid I) o H l? NH/VETOH Taurocholic acid HO OH O H2N\/\L| T _ g———_OH aunne Figure 12. Structures of substrates for S. cerevisiae sulfonate/oc-KG dioxygenase. COO- R\/fi\ coo- R 0 2+ + HSO - + + \/\ + + 02 Fe H 3 C02 ——> C00- C00- Scheme 3. Reaction mechanism of S. cerevisiae sulfonate/a-KG dioxygenase. 91 (Figure 13A), the kinetic properties for several other substrates were complicated by apparent inhibition of the enzyme at elevated substrate concentrations (Figure 13B and C). In those cases, the data were fit to the standard equation for substrate inhibition (Equation 1). Equation 1 v = VmaXISl/([SJ+Km +[512/Kn Inhibition constants (Ki) are listed in Table 7. TauD exhibited similar substrate inhibition profiles for certain substrates other than taurine (1). High concentrations of substrate are also inhibitory to y-butyrobetaine hydroxylase (K, is approximately 1 mM) (19), an enzyme that catalyzes the conversion of y-butyrobetaine to camitine and to prolyl hydroxylase (10). In y-butyrobetaine hydroxylase researchers demonstrated that the uncoupled decarboxylation of oc-KG at high substrate concentrations did not explain the apparent inhibition. The peptide substrate of prolyl hydroxylase is a competitive inhibitor with respect to iron and oc-KG for that enzyme (10). Sulfonate/OL-KG dioxygenase activity was not detected towards the following compounds: cysteate, aniline-2-sulfonic acid, homotaurine, picrylsulfonic acid, sulfosalicylic acid, 3-[(3—cholamidopropyl) dimethylammonio] propanesulfonic acid (CHAPS), 4-(2-hydroxyethyl)piperazine-1- propanesulfonic acid (EPPS), piperazine-N,N'—bis(2-ethanesulfonic acid) (PIPES), and 3- (cyclohexylamino)-1-propanesulfonic acid (CAPS). 92 Table 7. Sulfonate specificity and kinetic parameters for sulfonate/a-KG dioxygenase. Substratea 3-N-Morpholino- propanesulfonic acid (MOPS) [0-20 mM] N-Phenyltaurine [0-3 mM] 2-Hydroxyethanesulfonate (Isethionate) [0-20 mM] 3-Morpholino-2-hydroxy- propanesulfonic acid (MOPSO) [0-20 mM] Taurocholate [0-10 mM] 2-(4-Morpholino)- ethanesulfonic acid (MES) [0-20 mM] 2-(Cyclohexylamino) ethanesulfonic acid (CITES) [0-20 mM] N—(Tris[hydroxymethy1] methyl)-2-amino- ethanesulfonic acid (TES) [0-15 mM] N-[2-Hydroxyethyl] piperazine-N'-[2-hydroxy- propanesulfonic acid] (HEPPSO) [0-20 mM] N-(Tris(hydroxymethyl)— methyl)-3-aminopropane- sulfonic acid (TAPS) [0-20 mM] 4-(2-Hydroxyethyl)-1- piperazineethanesulfonic acid (HEPES) [0-20 mM] 2-Aminoethanesulfonic acid (Taurine) [0-20 mM] Vmax (umol/min/mg) 1.80 i 0.06 0.81 i 0.03 1.26 i- 0.05 1.51 i 0.03 0.99 i 0.05 1.48 i 0.05 0.56 i- 0.02 0.75 i 0.01 1.01 i 0.08 1.61i0.06 0.82 i 0.02 0.46 i 0.04 Km (mM) 0.24 i- 0.02 0.12:0.0’1 0.27 i 0.05 0.33 i- 0.02 0.42 i 0.05 1.94i0.14 2.57 i 0.34 3.60:0.12 5.04 i 1.07 10.38 i- .80 16.62 i 0.80 10.37 i 1.82 kcat/Km (min'l/mM) 670.9 606.4 418.4 411.2 210.7 68.4 19.6 18.7 18.0 13.9 4.4 4.0 Ki (111M) 34.3 i 5.8b 1.5:r0.1b 20.3 i 2.5b 28.4 i 2.4b 8.2:r1.1b 45.2 i 6.6b a The concentrations of substrate examined are shown in brackets b Data for these substrates were fit to Equation 1 c No inhibition was detected. Kinetic parameters were fit to the Michaelis-Menten equation. 93 22:: $56693: $38695 850 2: E8 2.6m: GOSBEE 88593 How cognac Eavcfim 2t 8 2.8 @558 Sm 830:3 :oEoEéfiEBE @3985 8 E 803 Sec 05. 2205239 CV was .Bmaflfiofl RC .0538 A»: :23 693mg 33 ogm .5352“ amazemxxfic UM-5\o§:o.:=m mo 3:06:33. nocubsuuoo oaabmazw .3 2sz :25 @825me ES 855% or mN m mN 0 ON m_. or m 0 ON 9. or m o H“V . Hm H \H . - .8. \fm - no. N H H .1 OVw ..or .1—. . . H H ..q t. .7/ . . . . o/ T - x/fi/ 8- -2 . / . . /// . H \ H n n.r fl! 1 r .. - / -3. /P -8. no .8 .. . l 0 l 1 l . . . o . H \e U .. . ...... .N....._..... ...i .09. 1.. _ _ .mN (Btu/n) Ail/woe amoeds 94 Additional properties of S. cerevisiae sulfonate dioxygenase Several additional properties of the MBP-dioxygenase fusion protein were characterized. The Km for a-KG was determined to be 30 i 16 uM. The approximate KD for F e(H) was found to be 44 i 5 uM which is ten-fold higher than the value observed for other OL-KG-dependent dioxygenases (see Chapter 1). The increase in apparent KD for iron may reflect non-specific metal binding by lVIBP. Temperature stability experiments, conducted by incubating the protein at varying temperatures for 1 hr prior to measuring activity at 30 °C, showed that the enzyme lost 77% of its activity after incubation at 37°C and was completely inactivated at higher temperatures. The fusion protein was shown to be cleaved by Factor Xa (New England Biolabs) to yield active 47 kDa enzyme. There was no significant difference in kcat towards MOPS for the fusion protein and the free enzyme. Phenotype of S. cerevisiae YLL05 7c deletion mutant Studies with a S. cerevisiae YLL057c deletion mutant, BY4742-11545 (Research Genetics, Huntsville, AL), demonstrated the importance of sulfonate/a-KG dioxygenase for growth on alternative sulfur sources. Comparison of the growth profiles for these two strains clearly showed that deletion of the sulfonate/oc-KG dioxygenase gene decreased the extent to which S. cerevisiae could use taurine and isethionate as the sole source of sulfur (Figure 14). The rate at which BY4742 grew in medium with isethionate was also affected (data not shown). The growth rate and final yield was identical for the two strains in medium containing sulfate. 95 0.6 0.5 0.4 3 0.3 ~ 0.2 . Absorbance (600 nm) 0.1 ~ _I:u_l:_ Sulfate lsethionate Taurine No Sulfur added Sulfur Source Figure 14. Role of YLL057c in the utilization of different sulfur sources. Growth of S. cerevisiae BY4742 (solid bars) and the YLL057c deletion strain BY4742-11545 (open bars) in medium containing either sulfate, isethionate, taurine, or no added sulfur source was measured as absorbance at 600 nm of stationary phase cultures. Error bars represent the standard deviation between replicate cultures. 96 Role of sulfonate/a-K G dioxygenase in S. cerevisiae The yeast sulfonate/a-KG dioxygenase may allow cells to exploit one or more of the many naturally occurring sulfonates. Some studies suggest that sulfonate sulfur is the major form of organic sulfur in soils (2). Specific examples of environmental sources of sulfonates include the sulfonolipids of gliding bacteria, the capsules of some Gram positive bacteria (1 l), and the taurine produced in all vertebrates and invertebrates (6). In addition, common oxidation products of cysteine and other S-containing amino acids include sulfoacetaldehyde (9, 14), sulfopyruvate, sulfolactate, cysteate (20), and taurine (20). In mice, the gut microflora have been shown to convert taurine and taurocholate to isethionate (20). Perhaps significantly, a taurocholate transporter (YBTl) (12) resides on the same segment of chromosome XH as YLL057c. The bile acid transporter has been localized to the vacuolar membrane and efficiently imports taurocholate into the vacuole. Sulfonate/oc-KG dioxygenase may play an additional role in the degradation of taurine- conjugated compounds in viva. Candida albicans, another fungus, contains a hypothetical protein sequence with more than 50% amino acid identity to the sulfonate/or-KG dioxygenase suggesting that other fungi have similar pathways to take advantage of these alternative sources of sulfur (15). The fact that S. cerevisiae BY4742-11545 could grow to some degree on isethionate suggests that an additional pathway exists for sulfonate utilization. Examples of enzymes capable of the liberation of sulfite from sulfonates include a flavin mononucleotide- dependent monooxygenase from Pseudomonas aeruginosa (7) and a sulfo-lyase from an Acinetobacter isolate (8). Redundancy of sulfonate catabolic pathways accentuates the importance of these compounds as sources of cell sulfur. 97 10. 11. 12. Literature cited Auchtung, T. A. 1999. Unpublished observations. Autry, A. R., and J. W. Fitzgerald. 1990. Sulfonate S: A major form of forest soil organic sulfur. Biol. Fertil. Soils 10:50-56. Cherest, H., and Y. Surdin-Kerjan. 1992. Genetic analysis of a new mutation conferring cysteine auxotrophy in Saccharomyces cerevisiae: updating of the sulfur metabolism pathway. Genetics 130:51-58. Eichhorn, E., J. R. van der Ploeg, M. A. Kertesz, and T. Leisinger. 1997. Characterization of OL—ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J. Biol. Chem. 272:23031-23036. Fukumori, F., and R. P. Hausinger. 1993. Purification and characterization of 2,4-dichlorophenoxyacetate/oc-ketog1utarate dioxygenase. J. Biol. Chem. 268:24311-24317. Huxtable, R. J. 1992. Physiological actions of taurine. Physiol. Rev. 72:101-143. Kertesz, M., K. Sehmidt—Larbig, and T. Wiiest. 1999. A novel reduced flavin mononucleotide-dependent methanesulfonate sulfonatase encoded by the sulfur regulated msu operon of Pseudomonas aeruginosa. J. Bacteriol. 181:1464-1473. King, J. E., R. Jaouhari, and J. P. Quinn. 1997. The role of sulfoacetalaldehyde sulfo-lyase in the mineralization of isethionate by an environmenal Acinetobacter isolate. Microbiology 143:2339-2343. Kondo, H., H. Anada, K. Ohsawa, and M. Ishimoto. 1971. Formation of sulfoacetaldehyde from taurine in bacterial extracts. J. Biochem. 69:621-623. Myllyla, R., L. Tuderman, and K. I. Kivirikko. 1977. Mechanism of the prolyl hydroxylase reaction: 2. Kinetic analysis of the reaction sequence. Eur. J. Biochem. 80:349-357. Ohtomo, T., K. Yoshida, and C. L. San Clemente. 1981. Effect of bile acid derivatives on taurine biosynthesis and extracellular slime production in encapsulated Staphylococcus aureus S-7. Infec. Irnmun. 31:798-807. Ortiz, D. F., M. V. St. Pierre, A. Abdulmessih, and I. M. Arias. 1997. A yeast ATP-binding cassette-type protein mediating ATP-dependent bile acid transport. J. Biol. Chem. 272:15358-15365. 98 13. 14. 15. 16. 17. 18. 19. 20. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Shimamoto, G., and R. S. Berk. 1979. Catabolism of taurine in Pseudomonas aeruginosa. Biochim. Biophys. Acta. 569:287-292. Unfinished Genome Database Website. 11 October 1999. http://www.ncbi.n1m.nih.gov/BLAST/ unfinishedgenomehtml. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux IH. 1994. Comparative aspects of utilization of sulfonate and other sulfur sources by Escherichia coli K12. Arch. Microbiol. 161:434-438. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux III. 1993. Sulfonate-sulfur assimilation by yeasts resembles that of bacteria. FEMS Microbiol. Letts 114:73-78. Uria-Nickelsen, M. R., E. R. Leadbetter, and W. Godchaux III. 1993. Sulphonate utilization by enteric bacteria. J. Gen. Microbiol. 139:203-208. Wehbie, R. S., N. S. Punekar, and H. A. Lardy. 1988. Rat liver y-butyrobetaine hydroxylase catalyzed reaction: influence of potassium, substrates and substrate analogues on hydroxylation and decarboxylation. Biochemistry 27 :2222-2228. Weinstein, C. L., and O. W. Griffith. 1988. Cysteinesulfonate and 13- sulfopyruvate metabolism. J. Biol. Chem. 263:3735-3743. 99 CHAPTER 4 DISTRIBUTION OF tfdA IN THE ENVIRONIVIENT Many of these results have been published in the article: Hogan, D. A., D. H. Buckley, C. H. Nakatsu, T. M. Schmidt, and RP. Hausinger (1997). “Distribution of the tfdA gene in soil bacteria that do not degrade 2,4-dichlorophenoxyacetic acid (2,4-D).” Microb. Ecol. 34:90-96. 100 Introduction The tfa’A gene (30) encodes 2,4-D/a-KG dioxygenase (deA) (9), an enzyme that catalyzes removal of the acetate side chain of 2,4-D yielding 2,4-dichlorophenol. Thus far, tfdA is uniquely associated with 2,4-D catabolism, and is most often found with the other tfd genes necessary to complete the degradation of 2,4-D. While tfdA is unlike any previously characterized genes involved in aromatic compound decomposition, the other tfd genes are similar to components of a number of different catabolic pathways. For instance, genes similar in sequence to lde encode hydroxylases that act on various substituted or unsubstituted phenols in a variety of microorganisms (15, 39). Likewise, the tdeDEF operon is homologous to other catechol degradation genes involved in many different aromatic degradative pathways (3 9). Because the downstream tfd genes are closely related to common bacterial genes, it is likely that homologs to tfdA can also be identified. Several pieces of evidence suggest that the assembly of the tfd genes to form 2,4- D degradation plasmids has occurred numerous times. First, the 2,4-D degradative genes are found in different arrangements on a variety of different plasmid backbones (35, 37). Second, plasmids within the same incompatibility group can have different alleles of the genes that encode the first three enzymes of the 2,4-D degradative pathway (10, 36). Third, on some plasmids, tfdA is physically distant from the zdeDEFB gene cluster (7). For there to be a “source” of tfd genes in the environment prior to their incorporation in 2,4-D degradative pathways, these genes must either be maintained without selection or serve an alternative purpose. Consistent with a widespread availability of tfdA homologs, 101 evidence for deA or deA-like activity has been found in diverse phylogenetic groups including the y-(21) and B-Proteobacteria (2, 6, 10, 32), a member of the Flexibacter- Cytophaga— Bacteroides phylum (3), and Arthrobacter sp. in the High G+C Subdivision of the Gram-positive phylum (34). To assess the natural distribution of tfdA, I used PCR primers internal to the tfdA gene to screen a collection of soil microorganisms, isolated on a non-selective medium, for the presence of tfdA or tfdA—like genes. In this chapter, I provide evidence that tfdA- like sequences are present in phylogenetically diverse soil microbes that do not degrade 2,4-D. These data suggest that tfdA is maintained in natural bacterial populations, possibly for a purpose other than 2,4-D degradation. Materials and Methods Bacterial strains Bacterial strains were isolated from soils obtained from a Long Term Ecological Research (LTER) plot (24) at the Kellogg Biological Research Station in Hickory Comers, IVH, USA which had been managed as a single cropping system to produce grain. Soil extracts were serially diluted in 50 mM phosphate buffer (pH 7), then plated on R2A medium (Difco, Detroit, MI) and incubated at 28°C for 10 days. Plates with 30 to 150 colonies were divided into quadrants. All colonies from one quadrant were purified on R2A agar and incubated at 28°C for three to four days. Additional colonies were chosen on the basis of colony morphology to include several groups of similar isolates as well as a representation of distinctive colony morphologies. The origins of Alcaligenes eutrophus 102 JMP134 (pJP4) (6), JMP228 (30), and Burkhola'eria sp. RASC (32) have been described previously. I6S rDNA partial sequence analyses Total DNA was obtained from these isolates by detergent lysis (1) and 168 rDNA was amplified by PCR using the eubacterial primers complementary to Escherichia coli nucleotide positions 8-27 and 1510-1492 (8). Amplified rDNA was purified with Ultra- free MC (30,000 MWCO) centrifugal filter units (Millipore, Bedford, MA), then sequenced with primers complementary to E. coli positions 536-519 or 8-27 (27). The rDNA sequence was determined by automated fluorescent sequencing performed by the MSU-DOE-PRL Plant Biochemistry Facility using an ABI Catalyst 800 for Taq cycle sequencing and the ABI 373A Sequencer for the analysis of products. Partial 16S rDNA sequences were aligned manually against a collection of 168 rRNA sequences (18) on the basis of conserved regions of sequence and secondary structure of 16S rRNA. Regions of ambiguous alignment were omitted from the subsequent analysis. Phylogenetic trees were generated using the Neighbor Joining method and the topology was confirmed by Parsimony and Maximum Likelihood analyses (3 3). PCR Amplification of the tfdA gene Total DNA samples were isolated from each of the LTER strains (1). Primers were derived from conserved regions of sequence in the tfdA genes of pJ P4 and Burkholderia strain RASC (79% identical to tfdA—pJP4). The sequence of the forward 103 primer, TVU, was 5’-ACG GAG TTC TG(C/T) GA(C/T) ATG-3’; the sequence of the reverse primer, TVL, was 5’-AAC GCA GCG (G/A)TT (G/A)TC CCA-3’(3 8). The PCR mix contained 5 ul 10X PCR buffer (Gibco, Gaithersberg, MD), 1.5 mM MgC12, 1 uM primer TVU, 1 uM primer TVL, 1 mM dNTPs (Boehringer Mannheim, Indianapolis, W), 1.5 units of Taq polymerase (Gibco), 10-100 ng template DNA, and distilled water to bring the final volume to 50 al. The PCR amplification protocol consisted of thirty five cycles with a one min denaturation at 94°C, one min amiealing at 55°C and one min extension at 72°C. The predicted product size was 360 bp. Amplified products were separated on a 1% agarose gel and compared to PCR products generated from JMP134 and RASC genomic DNA as template. Generation of PCR products was confirmed by at least one additional PCR amplification reaction. Products were purified with Promega PCR mini—prep columns (Promega, Madison, WI). Selected PCR products were sequenced with either primer TVL or TVU by the method described for 168 rDNA above. Dot Blot of PCR products PCR products (~50 ng) were applied to a Hybond N nylon membrane (Amersharn, Arlington Heights, H.) in 1 ul aliquots and UV cross-linked to the membrane with a Stratalinker (Stratagene, La J olla, CA). To create probes, PCR products obtained using DNA from JMP134 and RASC were labeled with digoxigenin using the DIG High Prime labeling mix (Boehringer Mannheim) according to the manufacturer’s instructions. Hybridization, detection, and probe removal were performed according to protocols in the Genius Detection System manual (Boehringer Mannheim). To assure high stringency, the 104 hybridization solution contained 50% forrnamide and hybridizations and stringency washes were at 68°C. Hybridization signals were detected by the colorimetric detection system (Boehringer Mannheim). Analysis 0f2,4-D degradation and deA activity Isolates were streaked on R2A agar and incubated at 30°C for three days to allow for growth of all cultures. Isolated colonies were transferred to tubes containing 3 ml of A+N medium (41) amended with 0.1% yeast extract and 1 mM 2,4-D. Cultures were incubated for two weeks at 30°C on a shaker. After incubation, cells were removed by centrifugation and the supernatant was filtered (0.45 urn) into an autosampler vial. Duplicate samples of each culture were either analyzed immediately or frozen for later analysis. 2,4-D and 2,4-dichlorophenol concentrations were determined by FH’LC with 70/30 methanol-0.1% phosphoric acid as eluent, with a Lichosorb RP-18 column (EM Separations, Gibbstown, NJ) and monitored with a UV detector at 230 nm. For induction experiments, cells were grown on MMO medium (29) with 5 mM succinate, harvested and washed in phosphate-buffered saline, then resuspended in medium with phenoxyacetic acid, 3-chlorobenzoate, 3-chlorocatechol, catechol, succinate (all 1 mM final concentration), or with no carbon source. deA activity was measured by detection of 2,4-dichlorophenol with 4-aminoantipyrene (9). Incorporation of radiolabel from 14C-2,4-D Single colonies from cultures grown on R2A plates were transferred to sterile nylon membranes overlaid on R2A agar. After overnight incubation at 30°C, all cultures had visible growth. Each filter was replicated (25), then both master and replica filters 105 were transferred to minimal agar containing either l4C-ring-labeled or chain—labeled 2,4- D and incubated for three days at room temperature. The filters were exposed to X-ray film (Kodak, Rochester, NY) for 18 hours to detect incorporation of radiolabeled carbon. Results Composition of the LTER collection The analysis of 16S rDNA partial sequence from strains in the LTER collection showed that these isolates represent diverse phylogenetic divisions (Figure 15,Table 8). The taxonomic divisions used to describe the isolates are defined in the Ribosomal Database Project (18). Among the seventy-six isolates analyzed, twenty seven were found to cluster within the Bacillus-Lactobacillus-Streptococcus subdivision and thirteen strains fall within the High G+C Subdivision of the Gram-positive phylum. The remaining isolates included five a-Proteobacteria, two B-Proteobacteria, eleven y-Proteobacteria, and eighteen members of the Cytophaga group. In a polyphasic taxonomy study, more than 95% of these isolates were found to be distinct taxa based on BIOLOG analyses (Biolog, Hayward, CA), fatty acid methyl ester (FAME) profiles (13), and genomic fingerprints from rep-PCR analyses (40). The polyphasic taxonomy study of the LTER isolates, including refined phylogenetic analyses based on complete 16S rDNA sequence is underway and will be presented elsewhere. Because the LTER collection includes both closely related isolates as well as organisms from several different phyla, this collection provides a good opportunity to test for the dispersion of tfdA between distantly related bacteria as well as for the presence of tfdA in similar strains within a bacterial species. 106 Identification of [@211 in LT ER collection PCR amplification with internal tfdA primers generated products using genomic DNA from twenty eight of the seventy six isolates. The products were similar in size to the tfdA product amplified from M134 and RASC (360 bp), and hybridized at high stringency to labeled products from PCR amplification of tfdA from JMP134 (Figure 16) and Burkholderia RASC. Amplification product from each isolate was assigned to either the tfdApm or tfdARAsc hybrization group. When primers TVU and TVL were applied directly to the membrane, hybridization to the probe was not detected. Amplified PCR products from four isolates were sequenced to assess the similarity of the tfdA-like gene in organisms from different phylogenetic divisions to each other and to previously characterized tfdA genes. Three of the isolates, LTER 1 and LTER 11 from the Bacillus-Lactobacillus-Streptococcus subdivision as well as LTER 8 from the Rhizobium-Agrobacterium group were identical to the tfdA partial sequence (Genbank U22499) from Halomonas sp. I-l 8, a moderate halophile (21) and to the tfdA sequence identified on plasmid pEMT8 (36) (Dr. Terrence Marsh, unpublished data). Although pEMT8 does not confer the ability to degrade 2,4-D to its host, it can complement R. eutropha(pBH501aE) (36), a tfdA" derivative of pJ P4, to restore the 2,4-D degradation phenotype (37) suggesting that it does encode a functional deA. The sequence of the PCR product from LTER 40, a member the B-Proteobacteria division, was identical to the sequence of tfdA found on pJP4. 107 Figure 15. 16S rDNA phylogeny of the LTER isolates Phylogeny is based on partial sequences corresponding to E. coli rDNA sequence between positions 27 and 519. Names of phylogenetic divisions are derived from the RDP 5.0 release (18). The scale bar indicates the number of nucleotide changes per base position analyzed. Isolates that produce a PCR amplification product with primers internal to the tfdA gene are enclosed in boxes. The abbreviation Rub. indicates Rubrivivax, th. indicate Rhizobium, and Ag. signifies Agrobacterium. LTER isolates in dark boxes yielded a PCR product belonging to tfdApm hybridization group; grey boxes indicate tfdA-like genes more similar to ddARAsc- 108 LTER28 i———— LTER97 LTER73 LTER63 LTER89 I dnorfiqns seuowopnasd L. {trenssl _j LTER64 __ , 'treaav ”"3 .J 38’- LTERBB ;;}E; LTER79 LTER54 ‘ LTER47 {l-Lrensr LTER41 £3 . : z 3 IIEHEHI § LTER37 _ 3% LTER36 g LTERSO £3 "-———-LTERSQ g LTER86 ——_j:fLTER74 LTER84 LTER21 f——"LTER52 1— LTER57 J F—————— tTERBt —— LTER70 LTER94 LTER69 IIIEHIII LTER1 LTER62 LTERIOO LTER13 IIIEEHEII LTER42 LTER15 IIIEHIHI LTERZ LTEFl32 ' LTERSS LTER44 LTER43 LTERS IIIEHEEI LTERl 7 LTE R34 LTERZQ LTERG LTERl O LTEH71 0.10 109 I euaroeg eldrnd I wnMud seploueroeg-BfiaudoMQ-Jeiaeqrxem I wnMud anmsod mars (D S- 9.0: 3115; Kim ._ =0 3 D) a? 9 co c 0’0 330. 9m) .10 _s.3 033 m3 :0 5' $3 3” mm U) 2’ I! 0’ 93' 92 -s?-ss U3 ‘g-m - a _l _ O ‘5. o 'o 3’ 9) no I!) O H o C ‘0 _l __l 22f“ "— l ._ F9 3 0o -3co «:9; cm _. :1 — c 3 03 W 9. E' ‘9 r- m m 9. o ' U) m cr 9:: fig —2g 92 u: <: m III—v a? 00) 3:“ (D P. o o o O o C (n (n 0b .09 Ca: ‘0: U) T _ 2’ "‘ (D a)? “E «0 ED ‘83 :3 02"§9 ‘2 80 U) :95an +3< .—Jg'_ _ Q) (0 a: 2,4-D degradation by LTER isolates HPLC analyses showed that 2,4-D was not degraded to a significant extent by any of the LTER isolates after a two week incubation. In contrast, JMP134 (pJP4) or Burkholderia sp. RASC completely consumed the 2,4-D during this time period. Uptake experiments provided additional evidence that none of the LTER isolates utilized 2,4-D for cellular carbon. Neither ring- nor chain-labeled 14C-2,4-D was incorporated into the LTER cultures after a three day incubation period while JMP134(pJP4) produced a strong positive signal after incubation on medium with either radiolabeled substrate. Experiments to determine if the [fl-like gene encodes a functional deA To determine if the tfdA-like genes encode functional enzymes, selected isolates from different phylogenetic groups were tested for deA activity and the presence of a deA—like protein in cell extracts. Ten isolates from different phylogenetic groups were selected for fiirther study including: LTER 28 (B-Proteobacterium); LTER 38 and LTER 82 (or-Proteobacteria); LTER 40 (y-Proteobacterium); LTER 20 (Cytophaga Group); LTER 1, LTER 11 and LTER 24 (Bacillus group); and LTER 3 and LTER 5 (Arthrobacter group). While JMP134 and RASC had high levels of deA activity after induction with 2,4-D, none of the cell lysates from the above LTER strains exhibited any activity after induction with 2,4-D, phenoxyacetic acid, 3-chlorobenzoate (3-CBA), catechol, 3-chlorocatechol, succinate or with no added carbon source. JMP134 was the only strain that degraded 3-CBA, as determined by HPLC analysis. Western blot analysis with anti-deA(pJP4) antibodies did not indicate the presence of a cross-reacting deA in 110 1 2 3 4 8 9 10 11 I2 14 4Q o :T '3 # Q G . C 18 20 22 24 25 27 28 31 38 40 O . c O 48 65 67 68 69 75 77 82 90 NC «... . {3 . P R 69~ Figure 16. Hybridization of amplified products from LTER isolates to a digoxigenin-labeled fragment internal to the tfdA gene. The numbers correspond to the LTER isolates that yielded the template DNA for the PCR reaction; 69* is an amplified product that is 40 bp longer than the predicted product size; as a negative control, NC, PCR primers were applied directly to the membrane; P is the amplified product from JMP134 (pJ P4); R, amplified product from Burkholderia strain RASC. Results from hybrization to RASC probe not shown. 111 Table 8. Distribution of isolates with tfdA among the different phylogenetic groups within the LTER collection. Phylogenetic subdivison Total no. No. of isolates % of isolates with of isolates with tfd/la {fdA pfll RASC a—Proteobacteria 5 2 2 60 B-Proteobacteria 2 O 2 1 OO y-Proteobacteria 1 1 1 2 27 Cytophaga 18 0 1 6 High G+C Gram-positives 13 1 5 46 Bacillus-Lactobacillus- 27 5 7 40 Streptococcus Total 76 9 19 37 a Amplified PCR products were categorized into hybridization groups based on the strength of the hybridization signal to the pJP4 and RASC probe. 112 lysates from any of the LTER isolates. In a separate experiment, cell extracts of Halomonas sp. LI 8 grown on 2,4-D possessed a band of approximately 32 kDa (similar in size to pJP4 and the predicted MW of pEMT8 deA) that cross reacted with the anti- deA (pJP4) antibody, however, RASC, grown with 2,4-D as the sole source of carbon, did not. Mating experiments similar to those published by Top et al. (3 6) were also performed to determine if any of the LTER isolates could transfer a functional tfdA to a recipient strain. No 2,4-D+ transconjugants were generated after incubation of the LTER isolates with R. eutropha JMP228 carrying pBHSOlaE (a tfdA‘ pJP4-derived plasmid) or plasmid-free R. eutropha JMP228 (N alR) (36). Discussion A significant percentage (37%) of organisms in this collection of soil isolates contains tfdA or tfdA-like genes. Two aspects of these data are particularly interesting. First, the tfdA-like genes are found in very phylogenetically diverse taxa (Figure 15, Table 8) including both Gram positive and Gram negative bacteria. Second, isolates within the same phylogenetic division contain tfdA-like genes from different hybridization groups (Table 8). Previous research provides evidence for tfdA in only a small number of Gram- positive 2,4-D-degraders, such as Arthrobacter sp. (34), compared to a vast collection of tfdA-containing Gram-negative bacteria (2, 6, 21 , 32, 35). This is the first report of a tfdA- like gene in Bacillus or C ytophaga. Despite the fact that 2,4-D is readily degraded by many different microbial communities (26), the isolation of axenic cultures that degrade 2,4-D has proven to be 113 more difficult (11, 17). These data suggest that, in many instances, 2,4-D is not degraded by individual organisms, but by the concerted activities of several different bacteria. The presence of tfdA in isolates that do not degrade 2,4-D may indicate the source of the 2,4- D-degradation potential of microbial communities. Perhaps prolonged exposure to 2,4-D is necessary to select for organisms or plasmids that contain both tfdA and the tdeCDEF genes. If tfdA is present in bacteria lacking the additional genes needed for 2,4—D mineralization, deA activity may be difficult to detect in pure culture due to the accumulation of toxic products, such as 2,4-dichlorophenol, or lack of the proper inducer molecule (19). The high identity of the partial tfdA—like sequences, amplified from the DNA of phylogenetically diverse bacteria, to previously characterized tfdAs suggests extensive horizontal transfer of these genes. The tfdA gene is most often found on self- transmissible, broad host-range plasmids that confer the ability to degrade 2,4-D (2, 6, 36). Several studies (4-6) have shown that plasmid pJP4, which encodes all of the enzymes necessary for 2,4-D degradation, can be maintained in a number of different bacteria. It is worth noting, however, that not all organisms containing pJP4 express the 2,4-D-degrading phenotype (6). Genetic characterization of plasmid-encoded 2,4-D degradation pathways indicate that tfdA most often resides on transposable elements. Plasmids that confer the 2,4-D+ phenotype are often unstable. Multiple researchers have reported that growth on non- selective media induces the loss of a 20- to 40-kb fragment from a plasmid concurrent with the inability to degrade 2,4-D (2, 7, 20, 42). A plasmid isolated from Michigan soils, pEMT8, (3 7) contains tfdA but does not have the other genes required for 2,4-D 114 degradation. The pEMT8 tfdA, which is identical to the partial sequences amplified from three of the LTER isolates, is adjacent to a transposase of the ISI 1 1 1/ISl328/IS15 33 family of elements (Dr. Terrence Marsh, unpublished data). Similar transposable elements have been found in diverse organisms including strains of Streptomyces (16), Mycobacterium (l6), Pseudomonas (Genbank # AF052750) and Rhizobium (Genbank# P55615). With this, it is reasonable to hypothesize that the tfdA—like genes in the LTER isolates are associated with a similar transposable element. Consistent with this hypothesis is the observation that the tfdA-like gene can be lost, as evidenced by the inability to amplify the 360 bp product (data not shown), upon multiple transfers of the isolates on rich medium. The wide spread presence of tfdA, even on a transposon, in organisms that do not degrade 2,4-D may indicate that these genes serve an important metabolic purpose apart from herbicide degradation. There are several examples of genes that reside primarily on mobile elements, including those encoding virulence factors (14), plant-microbe symbiosis functions (31), and antibiotic resistance (23). The strategy of maintaining important genes in only a fraction of the population by use of mobile genetic elements is poorly understood, but may prove to be important to microbial ecology. Naturally- occurring compounds with structural similarities to 2,4-D that are potential substrates of native deA include the numerous aryl-ether compounds that are released during fungal degradation of lignin (22) or the different halogenated aromatic metabolites produced by fungi and bacteria (12, 28). Alternatively, it is possible that the natural substrate for deA shows little resemblance to 2,4-D and the ability to degrade this compound is fortuitous. 115 Acknowledgements. I am grateful to Helen Corlew-Newman and Frank Lowes for providing the LTER strains and the past and present members of the Research on Microbial Evolution (ROME) Laboratory for helpful suggestions and discussions. Particular thanks go to Terry Marsh for providing the unpublished pEMT8 tfdA and transposase sequences and to Olga Maltseva for providing DNA and cell material from Halomonas strain I—18. 116 10. Literature cited Ausubel, F. M., R. Brent, R. E. Kingston, D. D. Moore, J. G. Seidman, J. A. Smith, and K. Struhl. 1987. Current protocols in molecular biology. John Wiley and Sons, New York. Bhat, M. A., M. Tsuda, K. Horiike, M. Nozaki, C. S. Vaidyanathan, and T. Nakazawa. 1994. Identification and characterization of a new plasmid carrying genes for degradation of 2,4-dichlorophenoxyacetate from Pseudomonas cepacia CSV90. Appl. Environ. Microbiol. 60:307-312. Chaudry, G. R., and G. H. Huang. 1988. Isolation and characterization of a new plasmid from a F lavobacterium sp. which carries the genes for the degradation of 2,4-dichlorophenoxyacetate. J. Bacteriol. 170:3897-3902. Daane, L. L., J. A. E. Molina, E. C. Berry, and M. J. Sadowsky. 1996. Influence of earthworm activity on gene transfer from Pseudomonas fluorescens to indigenous soil bacteria. Appl. Environ. Microbiol. 62:515-521. DiGiovanni, G. D., J. W. N eilson, I. L. Pepper, and N. A. Sinclair. 1996. Gene transfer of Alcaligenes eutrophus JMP134 plasmid pJP4 to indigenous soil recipients. Appl. Environ. Microbiol. 62:2521-2526. Don, R. H., and J. M. Pemberton. 1981. Properties of six pesticide degradation plasmids isolated from Alcaligenes paradoxus and Alcaligenes eutrophus. J. Bacteriol. 145:681-686. Don, R. H., A. J. Weightman, H. J. Knackmuss, and K. N. Timmis. 1985. Transposon mutagenesis and cloning analysis of the pathways for degradation of 2,4-dichlorophenoxyacetic acid and 3-chorobenzoate in Alcaligenes eutrophus JMP134 (pJP4). J. Bacteriol. 161:85-90. Eden, P. A., T. M. Schmidt, R. P. Blakemore, and N. R. Pace. 1991. Phylogenetic analysis of Aquaspirillum magnetotacticum using polymerase chain reaction-amplified 16S rRNA-specific DNA. Int. J. Syst. Bacteriol. 42:166-170. Fukumori, F., and R. P. Hausinger. 1993. Purification and characterization of 2,4-dichlorophenoxyacetate/OL-ketoglutarate dioxygenase. J. Biol. Chem. 268:24311-24317. Fulthorpe, R. R., C. McGowan, O. V. Maltseva, W. E. Holben, and J. M. Tiedje. 1995. 2,4-Dichlorophenoxyacetic acid-degrading bacteria contain mosaics of catabolic genes. Appl. Environ. Microbiol. 61:3274-3281. 117 .11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. Fulthorpe, R. R., A. N. Rhodes, and J. M. Tiedje. 1996. Pristine soils mineralize 3-chlorobenzoate and 2,4-dichlorophenoxyacetate via different microbial populations. Appl. Environ. Microbiol. 62:1 159-1 166. Gribble, G. W. 1992. Naturally occurring organohalogen compounds-a survey. J. Nat. Prod. 55:1353-1395. Haack, S. K., H. Garchow, D. A. Odelson, L. J. Forney, and M. J. Klug. 1994. Accuracy, reproducibility, and interpretation of fatty acid methyl ester profiles of model bacterial communities. Appl. Environ. Microbiol. 60:2483-2493. Hacker, J., G. Blum-Oehler, I. Muhldorfer, and H. Tschape. 1997. Pathogenicity islands of virulent bacteria: structure, function and impact on microbial evolution. M01. Microbiol. 23: 1089-1097. Hiiggblom, M. M. 1992. Microbial breakdown of halogenated aromatic pesticides and related compounds. FEMS Microbiol. Rev. 103:29-72. Hoover, T. A., M. H. Vodkin, and J. C. Williams. 1992. A Coxiella burnetti repeated DNA element resembling a bacterial insertion sequence. J. Bacteriol. 174:5540-5548. Ka, J. 0., W. E. Holben, and J. M. Tiedje. 1994. Use of gene probes to aid in recovery and identification of functionally dominant 2,4-dichlorophenoxyacetic acid-degrading populations in soil. Appl. Environ. Microbiol. 60:1116-1120. Larsen, N., G. J. Olsen, B. L. Maidak, M. J. McCaughey, R. Overbeek, T. J. Macke, T. L. Marsh, and C. R. Woese. 1993. The ribosomal database project. Nucl. Acids Res. 21:3021-3023. Leveau, J. H., F. Konig, H. F uchslin, C. Werlen, and J. R. Van Der Meer. 1999. Dynamics of multigene expression during catabolic adaptation of Ralstonia eutropha JMP134 (pJP4) to the herbicide 2, 4- dichlorophenoxyacetate. Mol. Microbiol. 33:396—406. Mae, A. A., R. O. Martis, N. R. Ausmees, V. M. K6iv, and A. L. Heinaru. 1993. Characterization of a new 2,4-dichlorophenoxyacetic acid degrading plasmid pEST4011: physical map and localization of catabolic genes. J. Gen. Microbiol. 139:3165-3170. Maltseva, O. V., C. McGowan, R. Fulthorpe, and P. J. Oriel. 1996. Degradation of 2,4-dichlorophenoxyacetic acid by haloalkaliphilic bacteria. Microbiol. 142:1115-1122. 118 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. Ribbons, D. W. 1987. Presented at the Technology in the 19903: Utilization of lignocellulosic wastes, London, UK. Roberts, M. C. 1996. Tetracycline resistance determinants: mechanisms of action, regulation of expression, genetic mobility, and distribution. FEMS Microbiol. Rev. 19:1-24. Robertson, G. P., and D. W. Freckman. 1995. The spatial distribution of nematode trophic groups across a cultivated ecosystem. Ecology 76: 1425-1432. Sambrook, J., E. F. F ritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Sandmann, E. R. I. C., M. A. Loos, and L. P. van Dyk. 1988. Microbial degradation of 2,4-dichlorophenoxyacetic acid in soil. Rev. Environ. Contamin. Tox. 101:1-53. Schmidt, T. M., E. F. DeLong, and N. R. Pace. 1991. Analysis of a marine picoplankton community by 16S rRNA gene cloning and sequencing. J. Bacteriol. 173:4371-4378. Siuda, J. F., and J. F. deBernardis. 1973. Naturally occuring halogenated compounds. Lloydia 36: 107-143. Stanier, R. Y., N. J. Palleroni, and M. Doudoroff. 1966. The aerobic pseudomonads: a taxonomic study. J. Gen. Microbiol. 43: 159-271. Streber, W. R., K. N. Timmis, and M. H. Zenk. 1987. Analysis, cloning, and high-level expression of 2,4-dichlorophenoxyacetate monooxygenase gene tfdA of Alcaligenes eutrophus JMP134. J. Bacteriol. 169:2950-2955. Sullivan, J. T., and C. W. Ronson. 1998. Evolution of rhizobia by acquisition of a 500-kb symbiosis island that integrates into a phe-tRNA gene [published erratum appears in Proc Natl Acad Sci U S A 1998 Jul 21 ;95(15):9059]. Proc. Natl. Acad. Sci. U S A 95:5145-9. Suwa, Y., A. D. Wright, F. Fukumori, K. A. N ummy, R. P. Hausinger, W. E. Holben, and L. J. Forney. 1996. Characterization of a chromosomally encoded 2,4-dichlorophenoxyacetic acid/a-ketoglutarate dioxygenase from Burkholderia sp. strain RASC. Appl. Environ. Microbiol. 62:2464-2469. Swofford, D. L., and G. J. Olsen. 1990. Phylogeny reconstruction. In D. Hillis and C. Moritz (ed.), Molecular Systematics. Sinaur Associates, Sunderland, MA. 119 34. 35. 36. 37. 38. 39. 40. 41. 42. Tiedje, J. M., and M. Alexander. 1969. Enzymatic cleavage of the ether bond of 2,4-dichlorophenoxyacetate. J. Agr. Food Chem. 17:1080-1084. Tonso, N. L., V. G. Matheson, and W. E. Holben. 1995. Polyphasic characterization of a suite of bacterial isolates capable of degrading 2,4-D. Microb. Ecol. 3:3-24. Top, E. M., W. E. Holben, and L. J. Forney. 1995. Characterization of diverse 2,4-dichlorophenoxyacetic acid-degradative plasmids isolated from soil by complementation. Appl. Environ. Microbiol. 6 1: 1691-1698. Top, E. M., O. V. Maltseva, and L. J. Forney. 1996. Capture of a catabolic plasmid that encodes only 2,4-D/0L-ketoglutarate dioxygenase (deA) by genetic complementation. Appl. Environ. Microbiol. 62:2470-2476. Vallaeys, T ., R. R. Fulthorpe, A. M. Wright, and G. Soulas. 1996. The metabolic pathway of 2,4-dichlorophenoxyacetic acid degradation involves different families of tfdA and tde genes according to PCR-RFLP analysis. FEMS Microbiol. Lett. 20: 163-172. van der Meer, J. R., W. M. de Vos, S. Harayama, and A. J. B. Zehnder. 1992. Molecular mechanisms of genetic adaption to xenobiotic compounds. Microbiol. Rev. 56:677-694. Versalovic, J., M. Schneider, F. J. de Bruijn, and J. R. Lupski. 1994. Genomic fingerprinting of bacteria using repetitive sequence-based polymerase chain reaction. Methods Mol. Cell Biol. 5:25-40. Wyndham, R. C. 1986. Evolved aniline catabolism in Acinetobacter calcoaceticus during continuous culture of river water. Appl. Environ. Microbiol. 51:781-789. Xia, X.-S., S. Aathithan, K. Oswiecimska, A. R. W. Smith, and I. J. Bruce. 1998. A novel plasmid pIJBl possessing a putative 2,4—dichlor0phenoxyacetate degradative transposon TN5530 in Burkholderia cepacia strain 2a. Plasmid 39:154-159. 120 CHAPTER 5 CONCLUSIONS AND FUTURE RESEARCH 121 In this dissertation, I presented my research on deA as both a model enzyme for the a-KG—dependent dioxygenase superfamily and as an enzyme with a unique catalytic activity in nature. The concluding chapter reviews the model for activity in or-KG- dependent dioxygenases and highlights three major questions that need to be addressed to understand the catalytic cycle of these enzymes. Although it is hardly necessary to justify the study of enzymes so critical to different areas of biology, some potential applications of the research on different a—KG-dependent dioxygenases in medicine, industry and agriculture are discussed. Lastly, I describe the ongoing efforts in our laboratory to find the role of deA or a similar enzyme in catabolic pathways other than those for the decomposition of 2,4-D. Understanding the mechanism of oc-KG-dependent dioxygenases Although the a-KG-dependent dioxygenases are not unified by similarities in primary structure, data from spectroscopic and biochemical experiments suggest that these enzymes share many common elements. Thus, the three groups of enzymes, defined by their amino acid sequences (Chapter 1), are likely derived from a common enzyme ancestor that underwent rearrangements, insertions or deletions during the evolution of novel catabolic activities. The similarities in tertiary structure between seemingly unrelated members of the superfamily is revealed by recent X-ray crystallographic characterizations (Prof. C. Schofield, unpublished data). Information taken from the examination of multiple a—KG-dependent dioxygenases has been incorporated into a general mechanistic model for enzymes in this superfamily. 122 The overall reaction catalyzed by oc—KG-dependent dioxygenases can be distilled down to three major steps. In this discussion, the proposed OL-KG-dependent dioxygenase mechanism is compared to that for IPNS, a related enzyme that does not use an or-keto acid as a co—substrate, to illustrate the common elements in the mechanisms of these disparate enzymes. The three stages of the reactions (A-C) are described below and depicted in Figure 17. (A) Substrate binding to the active site alters the coordination environment of the iron center to enable oxygen binding and formation of the first reactive oxygen species. The coordination geometry and electronic nature of ligand donors act in concert to control the reaction between iron and oxygen (19). In enzymes that use oc-KG as a co- substrate, the oc-keto acid chelates the metal through the C-1 carboxylate and the C-2 keto group (11, 25, 28, 36). The replacement of neutral solvent ligands with anionic species is proposed to cause a shift in the iron’s redox potential that may be necessary for subsequent steps in the reaction. In deA (4, 42), clavaminate synthase (25), and DAOCS (36), the unique substrate is not thought to interact directly with the metal, but rather induces a change in the metallocenter geometry from six— to five-coordinate. Together, the altered geometry and changes in redox potential promote 02 binding to the iron and generation of an Fe(HI)-superoxide intermediate. Although IPNS does not bind Ot-KG, a similar shift to five-coordinate geometry is brought about by direct ligation of a substrate thiol to the iron center (26, 27), thus enabling the binding of oxygen and formation of the proposed F e(III)-superoxide species. (B) The first oxidation reaction generates an F e(I V)-oxo intermediate. Iron- containing model compounds with ligands resembling the 2 His-l-carboxylate facial triad 123 and a bound oc-keto acid moiety have been used to study the initial steps in catalysis (10). Upon exposure to diatomic oxygen, the synthesized compound undergoes oxidative decarboxylation concomitant with the hydroxylation of an adjacent aromatic ring. Comparison of the rate for dioxygenase-like activity for this compound to the decreased rates with other model compounds suggests that the decarboxylation of a-KG provides a large driving force for cleavage of the oxygen-oxygen bond (10). This portion of the reaction catalyzed by oc-KG-dependent enzymes is analogous to the first two-electron oxidation reaction in IPNS where a new nitrogen-carbon bond is created upon ring closure and two electrons are transferred from the substrate to oxygen to form a molecule of water. (C) The reactive F e(I V)-oxygen species participates in a hydroxylation reaction or alternative oxidation (desaturation, ring expansion, cyclization) regenerating the F e(II) center. In hydroxylating enzymes, a specific hydrogen atom is abstracted from the substrate to form an F e(111)-OH species. The iron-oxygen bond breaks, yielding a hydroxyl radical which recombines with the substrate to give the hydroxylated product (32). In IPNS, a series of bonds are cleaved homolytically to generate a ring via a new carbon-sulfur bond. Again two electrons are transferred from substrate to reduce the second atom of oxygen to water, thus restoring the iron to its Fe(H) oxidation state. Additional research is needed in three particular areas to provide experimental evidence and additional information for several of elements of the above model. First, the ligands to the metal during different stages of catalysis must be identified and other 124 Figure 17. Comparison of the proposed reaction mechanisms for an OL-KG- dependent hydroxylase and IPN S. Both resting enzymes possess an Fe(II) that is bound by a facial triad of residues in an H- X-D-Xn-H motif. (A) In the oc-KG-dependent hydroxylases (left branch), two water molecules are displaced upon binding of anionic oc-KG to the metal and a third is lost upon binding of a second substrate (S-H) nearby. The activated metal reacts with 02 to form F e(III)—superoxide. A similar intermediate is generated in IPNS by stepwise binding of the substrate and oxygen (right branch). (B) An Fe(IV)-oxo species is generated by a reaction driven by oc-KG decarboxylation (left) or by oxidative cyclization of the lactam ring (right). (C) The highly reactive Fe(IV)-oxo species catalyzes fLuther oxidative chemistry and the metal returns to its resting Fe(H) state as products dissociate. 125 0H2 O H. M H20/ln,_ ('eF +2 ‘\\ ‘81 -00C 000- \Asp 8"" His: -000 O2 02 /o. O/ —OO/lll.le -i>‘3\\\His113 A O— ‘9 ASP115 HIS -OOC O--0\ 3\\\\H|5113 O o-e\/| ASP115 His 8 H C//0 L 0’ ®Fe [+4\\\Hi3113 -OOC /\/coo-'F l e\Asp115 His i //O S- C // HO 9 . Q.,3.\\H'$113 -ooc , e /\/coo-’ | ‘Aspns His 0H2 - His 8 OH HZO/llli..Fe+2‘\\\ 1 Hp" \Asp H152 N H-AA ”Ill/HS O NH-AA N\/H\"’//H H0 is -OOC \\\ \O \ei +3 \\\H15214 H3O CH3 (I\ H20 Hisz7 A59216 H NH-AA N ”III/H . HKJS -ooc\‘\\ VON} I) H3O H e+4-I\\HISZ14 H20 _ ASPzie Hl5270 NH-AA N "’//H H 5‘ 3 -00C H3O CH3 Figure 17 . Comparison of the proposed mechanisms for an a-KG—dependent hydroxylase and IPNS 126 essential residues defined. Multiple studies of several enzymes have thoroughly characterized the iron ligands in the resting enzyme (2 His, 1 Asp, 3 waters). The identity and geometry of the ligands to the metal during the course of the reaction, however, have not been characterized. The order of dissociation of succinate and C02 (or bicarbonate) suggests that these products remain in the active site until after the second oxidation reaction occurs, however, no role for either product has been identified. The bicarbonate dependence of ACC oxidase, an enzyme related to IPNS that does not use OL-KG as a co— substrate, may indicate the involvement of bicarbonate in the catalytic cycle of other enzymes. In addition, mutagenesis studies suggest that other amino acids present at the active site participate in the reaction; these residues need to be identified and characterized. The second set of critical experiments needed to support the proposed model is detection of the oxygen intermediates to confirm their identity. Researchers in our lab and others have proposed experiments that may detect Fe(HI)-containing intermediates (e. g. with bound superoxide or peroxide species) that may be generated during catalysis. F e(HI) is paramagnetic and therefore can be detected by EPR spectroscopy of freeze— quenched samples. Alternatively, stopped-flow UV-visible spectroscopy can also be used to detect reactive intermediates by their characteristic chromophores. Although detection of an Fe(IV)—oxo species may be difficult due to its high reactivity, model compound studies or the use of non-reactive substrate analogs might trap the reactive oxygen intermediate and aid in its identification. The third area of study that is important for the creation of a robust model involves the interactions between the unique substrate and the F e(lV)-oxo species. Apart 127 from IPNS, which has been structurally characterized with different substrate analogs, little is known about the positioning of the substrate or its interaction with the metal in other enzymes. a-KG-dependent dioxygenases demonstrate an amazing flexibility in the number and types of substrates that can be oxidized in a single catalytic center. For example, thymine 7-hydroxylase, which catalyzes the hydroxylation of methyl group of thymine, can also epoxidize olefins, oxidize thiol ethers to sulfones and sulfoxides, and demethylate the l-methyl group of 1-methyl thymine (3 5). Similarly, clavaminate synthase catalyzes a hydroxylation, a desaturation and a ring expansion on three different, though related substrates (Figure 2) (20). In this enzyme, the regio- and stereospecificity of the three reactions suggest significant differences in the binding modes for the structurally similar substrates (20). Alternatively, it has been proposed that clavaminate synthase may modulate its reactivity by altering the position of the F e(IV ) oxo- intermediate in response to different compounds. These experiments clearly show that the substrates themselves play a significant role in influencing the aspects of the reaction; however, the rules governing these interactions are not known. Information about the importance of different substrate characteristics can be obtained by studying the products of alternative substrates, spectroscopic characterization of the enzyme in the presence of different compounds, and ultimately by crystal structure studies of enzyme-substrate complexes. The site-directed mutagenesis studies described in Chapter 2 lay the foundation for examination of some of the questions outlined above. Knowledge of the metallocenter ligands are not only important in the design of model compounds, but are necessary for the interpretation of spectroscopic data and ultimately for our understanding of the 128 catalytic chemistry. For example, with both metal-bound histidines identified, we can now learn more about the rearrangements that occur upon substrate binding and how the substrate is positioned with respect to the metal. Furthermore, the ligands also play a crucial role in controlling the reactivity of the metal. For example, enzymes with heme- bound iron, such as cytochrome P450s, stabilize their iron(IV)—oxo intermediate with the porphyrin ring. In the absence of such a cofactor, oc-KG—dependent dioxygenases may rely on histidines to delocalize the charge on the metal. This potential role for histidines may explain why glutamine or carboxylates are poor substitutes, and may suggest that other amino acid substitutions (His to Tyr or Cys) may be capable of sustaining activity. Continuing efforts to spectroscopically characterize deA (and the related enzyme TauD) with different amino acid substitutions or different substrates and substrate analogs will address some of these important questions. In addition, for the past four years, I have supplied our collaborator Peter Roach at Oxford University with purified deA for use in crystallization studies. Although crystals were obtained, crystallization conditions are still being optimized. Structural characterization has not yet been successful. If a crystal structure of deA or a closely related enzyme becomes available, more specific experiments can be designed to probe the catalytic mechanism. Biological significance and potential applications of oc-KG-dependent dioxygenases Several or-KG-dependent dioxygenases are already being studied for their potential applications to the fields of medicine, industry, and agriculture. Prolyl and lysyl hydroxylases have long been studied for their role in collagen formation. Hydroxylated proline and lysine residues, which comprise between 0.5—5% of the amino acids in collagen or collagen-like proteins, are necessary to stabilize its triple helical structure 129 under physiological conditions (18). People with defects in lysyl hydroxylase suffer from Ehlers-Danlos syndrome, which is characterized by hypermobility of the joints, hyperextensibility of the skin, and other traits indicative of a connective tissue disorder (18). The excessive accumulation of collagen, termed fibrosis, can also impair function in a variety of tissues including the liver, lungs, and kidneys. Although several different inhibitors and inactivators of collagen hydroxylases prevent collagen formation in cultured human skin fibroblasts, none are currently in clinical use (18). In industry, the enzymes involved in the synthesis of B-lactam antibiotics are being scrutinized for opportunities to improve the large-scale synthesis of these complex molecules. Due to their intricate structure which defies chemical synthesis, all penicillin and cephalosporin antibiotics are still produced as natural fermentation products (31). Research on IPNS, DAOCS and related enzymes has focused on improving the specificity, rate and range of precursor substrates in hopes of optimizing steps in the synthesis and purification of these chemicals in addition to creating additional derivatives with different biological properties. Finally, in agriculture, deA has already been used to create transgenic plants that are resistant to the broad—leaf herbicide 2,4—D and related phenoxyacetic acids (1). It has also been proposed that deA could be used in the preparation and processing of phenoxy herbicides (29). Since only one isomer of chiral 2,4-D derivatives, such as mecoprop, is biologically active, deA-containing microorganisms or immobilized enzyme could be used to remove the inactive isomer from a chemically-synthesized racemic mixture, thus reducing the load of xenobiotics to soils. 130 Role of deA in the environment Enzymes involved in the degradation of xenobiotic compounds are generally derived from proteins that act on chemically-similar, natural compounds. For example, benzoate and chlorobenzoate dioxygenases are related in both sequence and mechanism but have differing reactivities towards chlorinated and unsubstituted substrates (24). Similar examples include phenol hydroxylases (9), catechol dioxygenases (8), and benzene dioxygenases (2). deA acts on a variety of natural and synthetic substrates (Figure 18), but none with kinetic parameters approaching those for 2.4-D (R. Padmakumar, J. Hotopp, R. P. Hausinger, unpublished data). One abundant source of natural compounds that resemble known substrates of deA is lignin (9). This compound, the second most abundant polymer on earth, is an assemblage of aromatic units joined primarily through ether linkages. The aerobic degradation of lignin begins with “combustion” of the polymer to its constituent subunits by the action of fungal peroxidases to release a variety of monomeric and dimeric compounds. The types and relative percentages of the different intermonomer linkages found in native lignin are shown in Figure 19, and indicate the range of monomeric and dimeric compounds that are liberated during lignin decomposition. Three types of compounds in particular, B-aryl ethers, derivatives of cinnamic acids (including the corresponding alcohol (coumaryl alcohol) and cinnamic acid esters through the side-chain carboxylate), and phenoxyacetic acids bear particular resemblance to known substrates of deA (17). One approach to determining if deA-like enzymes are involved in alternative degradative pathways, such as lignin degradation, is to examine the roles of proteins phenoxyacetic acids naphthoxyacetic acid thiophenoxyacetic acid 0 I \ \ 0' W / / O O- X cinnamic acids benzofurancarboxylic acid (couman'llic acid) Figure 18. Structures of various substrates of deA OCH3 fHZOH H fHSOH fl H2 /O\CH H CH0 H OOH OOH T | F r—r 2 0 2 “r“ O 0 w fH H3CO OCH, 0— H OH OH OCHg H3CO OH OH B-aiyl ethers (51%) phenylcoumarane (12%) biphenyls (l 1%) pinoresinol (2%) Figure 19. Representative structures of the major intermonomeric linkages in lignin. The relative percentages of these linkages are shown in parentheses (17, 22). 132 encoded by nearby genes. Efforts are currently underway in our laboratory to clone a tfdA-like gene, described in Chapter 4, from isolates that do not appear to degrade 2,4-D. The extragenic sequence of these elements may provide clues about the in vivo role of this gene. Alternatively, the surrounding regions of sequences encoding proteins with high identities to deA may indicate their potential functions in nature. In Bordetella pertussis, the causative agent of whooping cough, there is a hypothetical protein with 42% identity to deA over the entire amino acid sequence (Figure 20). Interestingly, the adjacent open reading frame, which is divergent from the tfdA-like gene, is 37% identical in amino acid sequence to a glutathione-S-transferase (GST) from Sphingomonas (formerly Pseudomonas) paucimobilis SYK-6 strain cleaves the B-aryl ether linkage of dimeric lignin components (Figure 21) (22, 23). Other open reading frames adjacent to the putative dioxygenase gene in B. pertussis are also homologous to enzymes involved in aromatic degradation (Figure 20), including a LysR-type transcriptional regulator (37) and a sequence homologous to clcE, which encodes maleyl acetate reductase (16). Although B. pertussis has not been found in the environment, Bordetella spp. have been isolated from soils (7, 33) and paper effluents (6), and at least one of these strains degrades 2,4-D (7). Apart from lignin degradation, GSTs are associated with a number of other pathways involved in degradation of aromatics. Genes encoding GSTs are found in degradative pathways for biphenyl (Bth) in Burkholderia sp. LB400 (l2), 2,4,5— trichlorophenoxyacetic acid (the product of orf3) in Burkholderia cepacia AC1 100 (5), and polycyclic aromatic hydrocarbons (Xle) in Cycloclasticus oligotrophicus (41 ). Although gene disruption studies demonstrate that these enzymes are not required for 133 _ ORFl-putative OL-KG-dependent dioxygenase (287 aa) with 42% identity to deA (M16730) ORF2-putative glutathione S-tranferase (203 aa) with 37% identity B-etherase, LigE ORF3-putative LysR-type transcriptional regulator with 29% identity to N-terminus of SalR(3l6 aa). ORF4-putative maleyl acetate reductase with 37% identity to ClcE (329 aa) (Q47100) Figure 20. Open reading frames adjacent to a putative oc-KG-dependent dioxygenase in Bordetella pertussis. Genbank accession numbers are shown in parentheses. activity, the multiple occurrences of GST—encoding genes in aromatic degradation operons suggest that they do serve a purpose. Based on the activities observed for other GSTs, proposed functions for the above enzymes include double bond isomerization, dehalogenation, and epoxide cleavage (40), but none of these activities have been demonstrated. Despite the prevalence of lignin and its degradation products in soils, bacterial decomposition of these compounds has not been well studied due to the difficulties in working with such a large, heterogeneous substrate. The scheme for the degradation of lignin-derived aromatics in S. paucimobilis SYK—6 demonstrates the number and diversity of pathways and intermediates involved in this process (Figure 21) (22), but certainly does not represent the only means for lignin mineralization (17). Of the few previously characterized lignin-degrading isolates, most have generally belonged to the Actinomycete group (3, 21, 34, 39). Although not enriched in cultures with polymeric lignin as a substrate, a vast number of isolates including oc-, B-, and y—Proteobacteria and numerous Gram positive bacteria degrade a variety of aromatic compounds, including 134 .QNNEQ 6 632 80% BESS ong EBQEEEQ mgeSeMNxEsw. E .30:on cosmwfiwow flaw: 288% 2588:: we $95:qu cosmcfiwofl AN ohswmm 298 <09 10 M18 omwsowzxomw 0325038805 10 E Ea .:. > 10" £00 10 1o _ All NIO Allin 10 £00 _ 1oN1o - oom1 T 3.5050 n Juno i 010 O|n_VI 1000 010 10N10 emacowoboxnow mgaonamm ogsasooicoam 2:309:00 cofio road 10 -60 foo 10 1o 10 o ...18 oof 1w _ m 1o|o~1 0011010 1000 108 1w 010 o|n_.1 1w 10N1o 135 biphenyls and aromatic acids (38, 39). Thus, it is not unlikely that more of these organisms possess the ability to degrade at least a fraction of the available lignin degradation products. As an aside, different S. paucimobilis strains are commonly isolated from 2,4-D enrichment cultures (13-15), have deA activity (30), but lack a tfdA gene with enough sequence identity to be picked up by gene hybridization methods (13). When I began this research, there were no known enzymes with obvious sequence similarity to deA. Thus, little could be said about important or conserved residues or active site structure. Five years later, deA is related in sequence to a growing number of characterized and putative dioxygenases (Chapter 1). Sequence alignments of deA to two related sulfonate/OL-KG dioxygenases indicated a number of conserved residues that proved to be important for activity (Chapter 2). Comparison of these findings to studies of other enzymes in the superfamily shows both similarities in metallocenter ligands and differences in overall protein structure among the a-KG-dependent dioxygenases. In Chapter 4, I reported the intriguing presence of a tfdA-like gene in diverse soil isolates that do not degrade 2,4-D suggesting that this gene is maintained for a function other than in herbicide degradation. Although no alternative substrate for deA in soil microorganisms has been identified, there is circumstantial evidence for its involvement in the degradation of plant-synthesized ether- and cinnamic acid-containing products derived possibly from lignin. Further biochemical characterization of pJP4-derived deA and dioxygenases from other environmental bacteria are ongoing in the continued search for other oc-KG-dependent dioxygenases with activity similar to that of deA. 136 Literature cited Bayley, C., N. Trolinder, C. Ray, M. Morgan, J. E. Quisenberry, and D. W. Ow. 1992. Engineering 2,4-D resistance into cotton. Theor. Appl. Gen. 83:645- 649. Beil, S., K. N. Timmis, and D. H. Pieper. 1999. Genetic and biochemical analyses of the tee operon suggest a route for evolution of chlorobenzene degradation genes. J. Bacteriol. 181:341-346. Cespedes, R., L. Salas, I. Calderon, B. Gonzales, and R. Vienna. 1992. Microbial and biochemical-characterization of a bacterial consortium isolated from decaying wood by grth on a B-O-4 lignin-related dimeric compound. Arch. Microbiol. 158: 162-170. Cosper, N. J., C. M. V. Stalhandske, R. E. Saari, R. P. Hausinger, and R. A. Scott. 1999. X-Ray absorption analysis of F e(II) and Cu(H) forms of an herbicide-degrading a—ketoglutarate dioxygenase. J. Biol. Inorg. Chem. 4: 122- 129. Daubaras, D. L., C. D. Hershberger, K. Kitano, and A. M. Chakrabarty. 1995. Sequence-analysis of a gene-cluster involved in metabolism of 2,4,5- trichlorophenoxyacetic acid by Burkholderia cepacia AC1100. Appl. Environ. Microbiol. 61:1279-1289. Foss, S., U. Heyen, and J. Harder. 1998. Alcaligenes defragrans sp. nov., description of four strains isolated on alkenoic monoterpenes ((+)-menthene, alpha-pinene, 2-carene, and alpha-phellandrene) and nitrate. Syst. Appl. Microbiol. 21 :23 7-244. Greer, L. E., J. A. Robinson, and D. R. Shelton. 1992. Kinetic comparison of seven strains of 2,4-dichlorophenoxyacetic acid-degrading bacteria. Appl. Environ. Microbiol. 58: 1027-1030. Haggblom, M. M. 1992. Microbial breakdown of halogenated aromatic pesticides and related compounds. FEMS Microbiol. Rev. 103:29-72. Hausinger, R. P., F. Fukumori, D. A. Hogan, T. M. Sassanella, Y. Kamagata, H. Takami, and R. E. Wallace. 1996. Biochemistry of 2,4-D degradation: evolutionary implications, p. 35-51. In K. Horikoshi, M. F ukuda, and T. Kudo (ed.), Microbial Diversity and Genetics of Biodegradation. Japan Scientific Press, Tokyo. 137 20. Hegg, E. L., R. Y. N. Ho, and L. Que Jr. 1999. Oxygen activation and arene hydroxylation by functional mimics of oc-keto acid-dependent iron(II) dioxygenases. J. Am. Chem. Soc. 121:1972-1973. Hegg, E. L., A. K. Whiting, J. McCracken, R. P. Hausinger, and J. Que, L. 1999. The herbicide-degrading oc-keto acid enzyme deA: metal coordination environment and mechanistic insights. Biochemistry In press. Hofer, B., S. Backhaus, and K. N. Timmis. 1994. The biphenyl polychlorinated biphenyl-degradation locus (Bph) of Pseudomonas Sp. LB400 encodes four additional metabolic enzymes. Gene 144:9-16. Ka, J. 0., W. E. Holben, and J. M. Tiedje. 1994. Genetic and phenotypic diversity of 2,4-dichlorophenoxyacetic acid (2,4-D)-degrading bacteria isolated from 2,4-D-treated field soils. Appl. Environ. Microbiol. 60:1 106—1 1 l5. Ka, J. 0., W. E. Holben, and J. M. Tiedje. 1994. Use of gene probes to aid in recovery and identification of functionally dominant 2,4-dichlor0phenoxyacetic acid-degrading populations in soil. Appl. Environ. Microbiol. 60:1116-1120. Ka, J. 0., and J. M. Tiedje. 1994. Integration and excision of a 2,4- dichlorophenoxyacetic acid degradative plasmid in Alcaligenes paradoxus and evidence of its natural intergeneric transfer. J. Bacteriol. 176:5284—5289. Kasberg, T., D. L. Daubaras, A. M. Chakrabarty, D. Kinzelt, and W. Reineke. 1995. Evidence that Operon-ch, operon-de, and operon-Clc encode maleylacetate reductase, the fourth enzyme of the modified ortho pathway. J. Bacteriol. 177:3885-3889. Kirk, T. K., and R. L. Farrell. 1987. Enzymatic combustion - the microbial- degradation of lignin. Ann. Rev. Microbiol. 41:465-505. Kivirikko, K. I., and T. Pihlajaniemi. 1998. Collagen hydroxylases and the protein disulfide isomerase subunit of prolyl 4-hydroxy1ases. Adv. Enzymol. Relat. Areas Mol. Biol. 72:328-398. Lippard, S. J., and J. M. Berg. 1994. Principles of bioinorganic chemistry. University Science Books, Mill Valley, CA. Lloyd, M. D., K. D. Merritt, V. Lee, T. J. Sewell, B. Wha-Son, J. E. Baldwin, C. J. Schofield, S. W. Elson, K. H. Baggaley, and N. H. Nicholson. 1999. Product-substrate engineering by bacteria: Studies on clavaminate synthase, a trifunctional dioxygenase. Tetrahedron 55: 10201-10220. 138 '21. 22. 23. 24. 25. 26. 27. 28. 29. 30. Lodha, S. J., R. A. Korus, and D. L. Crawford. 1991. Synthesis and properties of lignin peroxidase from Streptomyces viridosporus T7a. Appl. Biochem. Biotechnol. 28-9:41 1-420. Masai, E., Y. Katayama, S. Kawai, S. Nishikawa, M. Yamasaki, and N. Morohoshi. 1991. Cloning and sequencing of the gene for a Pseudomonas paucimobilis enzyme that cleaves B-Aryl ether. J. Bacteriol. 173:7950-7955. Masai, E., Y. Katayama, S. Kubota, S. Kawai, M. Yamasaki, and N. Morohoshi. 1993. A bacterial enzyme degrading the model lignin compound [3- etherase is a member of the glutathione-S-transferase superfamily. FEBS Lett. 323: 135-140. N akatsu, C. H., N. A. Straus, and R. C. Wyndham. 1995 . The nucleotide- sequence of the Tn5271 3-chlorobenzoate 3,4—dioxygenase genes (cbaAB) unites the class Ia oxygenases in a single lineage. Microbiol. UK 141:485-495. Pavel, E. G., J. Zhou, R. W. Busby, M. Gunsior, C. A. Townsend, and E. 1. Solomon. 1998. Circular dichroism and magnetic circular dichroism spectroscopic studies of the non-heme ferrous active site in clavaminate synthase and its interaction with a-ketoglutarate cosubstrate. J. Am. Chem. Soc. 120:743- 754. Roach, P. L., I. J. Clifton, V. Fiiliip, K. Harlos, G. J. Barton, J. Hajdu, I. Andersson, C. J. Schofield, and J. E. Baldwin. 1995. Crystal structure of isopenicillin N synthase is the first from a new structural family of enzymes. Nature 375:700-704. Roach, P. L., I. J. Clifton, C. M. H. Hensgens, N. Shibata, C. J. Schofield, J. Hajdu, and J. E. Baldwin. 1997. Structure of isopenicillin N synthase complexed with substrate and the mechanism of penicillin formation. Nature 387:827-830. Ryle, M. J., R. Padmakumar, and R. P. Hausinger. 1999. Stopped-flow kinetic analysis of Escherchia coli taurine/a—ketoglutarate dioxygenase: interactions with oc-ketoglutarate, taurine, and oxygen. Biochem. In press. Saari, R. E., D. A. Hogan, and R. P. Hausinger. 1999. Stereospecific degradation of the phenoxypropionate herbicide dichlorprop. J. Molec. Catalysis B: Enzymatic 266:421-428. Sassanella, T. M., F. Fukumori, M. Bagdasarian, and R. P. Hausinger. 1997. Use of 4-nitrophenoxyacetic acid for the detection and quantitation of 2,4- dichlorophenoxyacetic acid (2,4-D)/0L-ketoglutarate dioxygenase activity in 2,4-D degrading microorganisms. Appl. Environ. Microbiol. 63: 1 1 89-1 191 . 139 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. Schofield, C. J., J. E. Baldwin, M. F. Byford, 1. Clifton, J. Hajdu, C. Hensgens, and P. Roach. 1997. Proteins of the penicillin biosynthesis pathway. Ctur Opin Struct Biol 7 :857—864. Shilov, A. E., and A. A. Shteinman. 1999. Oxygen atom transfer into C—H bond in biological and model chemical systems. Mechanistic aspects. Accounts of Chemical Research 32:763-771. Sturz, A. V., B. R. Christie, B. G. Matheson, and J. Nowak. 1997. Biodiversity of endophytic bacteria which colonize red clover nodules, roots, stems and foliage and their influence on host growth. Biol. Fert. Soil 25:13-19. Thomas, L., and D. L. Crawford. 1998. Cloning of clustered Streptomyces viridosporus T7A lignocellulose catabolism genes encoding peroxidase and endoglucanase and their extracellular expression in Pichia pastoris. Can. J. Microbiol. 44:364-372. Thornburg, L. D., M. T. Lai, J. S. Wishnok, and J. Stubbe. 1993. A non—heme protein with heme tendencies: an investigation of the substrate specificity of thymine hydroxylase. Biochemistry 32:14023-14033. Valegz‘ird, K., A. C. T. van Scheltinga, M. D. Lloyd, T. Hara, S. Ramaswamy, A. Perrakis, A. Thompson, H.—J. Lee, J. E. Baldwin, C. J. Schofield, J. Hajdu, and I. Andersson. 1998. Structure of a cephalosporin synthase. Nature 394:805- 809. van der Meer, J. R., A. C. J. Frijters, J. J. Leveau, R. I. L. Eggen, A. J. B. Zhender, and W. M. De Vos. 1991. Characterization of the Pseudomonas sp. strain P51 tch, a LysR-type transcriptional activator of the tchDEF chlorocatechol oxidative operon, and analysis of the regulatory region. J. Bacteriol. 173:3700-3708. Vicufia. 1988. Bacterial degradation of lignin. Enzyme Microb. Techno]. 10:646— 655. Vicuna, R., B. Gonzalez, D. Seelenfreund, C. Ruttimann, and L. Salas. 1993. Ability of natural bacterial isolates to metabolize high and low—molecular—weight lignin-derived molecules. J. Biotechnol. 30:9-13. Vuilleumier, S. 1997. Bacterial glutathione S-transferases: What are they good for? J. Bacteriol. 179:1431—1441. Wang, Y., P. C. K. Lau, and D. K. Button. 1996. A marine oligobacterium harboring genes known to be part of aromatic hydrocarbon degradation pathways of soil pseudomonads. Appl. Environ. Microbiol. 62:2169—2173. 140 42. Whiting, A. K., L. Que Jr., R. E. Saari, R. P. Hausinger, M. A. Fredrick, and J. McCracken. 1997. Metal coordination environment of a Cu(H)—substituted Ot- keto acid-dependent dioxygenase that degrades the herbicide 2,4-D. J. Amer. Chem. Soc. 119:3413-3414. 141 APPENDIX A AMINO ACID SEQUENCE ALIGNMENTS Amino acid sequence alignments for Group 11 (Figure 22) and Group III (Figure 23) enzyme and different deAs (Figure 24) are shown. Alignments were performed using either the PIMA (3) or CLUSTAL (5) algorithms, which are available through the Baylor College of Medicine (BCM) Search Launcher (4). Figures were generated using GeneDoc software (1, 2). 142 QQIImHZ HDflMm>mmAmmHmm¢ZIIII mZUBfldfiQO>mmD¢Illm>H¢I|UDQUHH>¢O>OHDMW>D¢>Hh>IIIIOMMB> MZQZW>ZMM>MUM IIIIII MMHZUHMEMIIWOIQQH3>HOQWZH¢MQHH>OZW$MHIMWQWHHOUUZWMMOZHQQ mmOZUOmDMHmm llllll Qm>mIIIIOQIlmQN¢ZE HMHKQMHWIIIIDMIIIIEDHW¢ZMM>HEWE HDM¢4QMQQIEHOHIIIIUZHEQUIQMI|lmHHMK>KHOO>ZM¥¢MHzmhzwafimgmmflmwmmmMQWUWMMHWQB 9<¢QHQmWH¥>IIIIUmHM¢OIHMIIlwm>mm>DHmH¢ImMHWUH¢MMQDHIIIIQM UflflleIIIWEHmZ¢>>WE « ovfi * > * OOH * 0w MN>OQOUmUZDmHBQZEmmm>MUQm>OmIImmZHHMmmzmMQODDWmEBDmZQWUOMWHNZHZMQQZBHHM>mQQHmmHmH¢BO¢¢¢mmE * ow * OV * ON r fimo Nmo mnzm qmo 1mmw 1mmw mo>m onmo omw xmum mama «owe amo Nmo muxm qmo 1mmo 1mmw mo>m oumo oww xmua game acme .85on 8m @30me m2me mmmocowfisfi 8826-36 E cowowcfi 8:162 <13. can .Amoxon bum x83 82533 2: mo $3 :3“ 208 E 33828 83362 .Amoxon x87: 3362 33038 bami .2 852an m 2an mo gown: 2: E wastomov ma 2m whomucoE 8:258 was £02.83: :ommmoooa 03H moms—Subs? «nausea—ocAvv—S 3 959.0 mo Eva—EMF. .NN Baum...— 143 VFH th OOH mMN mNN Nwfi Hma ovN wma PNH NvH |>HE..BmmmoWZmOdmr < QHEWEQQHMWBmmmdmm<0mmmW>QEN>BOmeMmWDEIQDHMMU I>HHWNZmOAHE dZmBIEMHQBmmHmmmqwmflwmmmW>QEM>BDmeMOWOEIQDHmm0 lama..ow>mmmoiimommsommm3mmg1mmm-1quzmMmmme iiiiii mmw>zmwomxxmqowwewmqwiIii >me.zuce¢oz>iiqozqmoqomoqmmamnos1qquiiiimae iiiiiii w<2moom2miiii >q aowoemameliq01m>oHomeooum Hooa1monom wwmeo iiiii mw3moommaiiii mmaamwwoaowmomm>ommm>emam zoo 1qm> iiiii wzz iiiiii mmHmomaomom mq>¢wmmziiii lieu.wmmm>zmHmaawqm>oo>omqqqooo 1umqoomwe>Hmo iiiiii memq iiiii x anmwowo iiiii liqeoowoomom>>miimmewm moam m 0H1.w iiiiiiiii w iiiiii mmemo iiiii Qm>m>mmiqum¢m w e>oa1nmz< iiiiiiiiiiiiii moooaiiinwo oqumaoimiii liqeowwammqumaiquaw memH e 051.2oz iiiiiiiiiiiiii ommzaziizao HHmo>o mini lim:o ommm>>>iu> iiiiii mm<0m0¢>xwo m>zmHo b owm . ovm v >Q0q MZHQHWMQmmmHCmHHQIUDMmUmflzwflodldfi wmdw<¢¢BI¢¢D_mflmxmflflm¢mfimfimEmmflmwaID>mQCmQQQWODmmD>mmflo IIIIII mUMMQ>¢GImUQHBmHfimmmmIQQOQflBmMH >mHI Ilmm hmmefldedflmeBm>UIAOH IIIIII QC>QM>¢HQBMHQBQQQOEIIII mHmQOOFBMEI IIQE m>HmOOQMMB¢mmBOm>OD>HH IIIIII INK>AZKWMWOQWHBWKWIIIIHMm>I DKIIHWMWZBHMWH liq MOWQIUNMBHmmflflflhflfi IIIIIIIIII mDmAAQHZEmmlmzlwmm0mm.dqwm II>O Mm IIIIIII Dflwm>HB IIIIIIIIIIIII JflflmNHQMbIIIBQmQHD IIIII Allzm>>>m00m QZMH ImmthHEH IIIII mmZZOmEOmBOSHSQIIIIKEM$MQMMH>WI memodmz IIIIII OZMM>>>OMO Afiflq I>>m>B IIIIIIIII mm¢>mw IIIIIIIIIIIII :QQMfim IIIIII QMOWOQQIIIQEHOOMMMH>MEM <¢OH Idmmw> IIIIIIIII mEHEQm IIIIIIIIIIIII Ib¢0<fi IIIIII ¢MOOOmBIIIHI fiOQQO>>OEMQ QEMQ lfiWMmm IIIIIIII mmMOZ>MHm00mg IIIIII mimzmflm IIIIII flHOODOmIIIQI mmem>Q>mMm01AmmH ONN * OON owH « OOH Hmo Nmo muwm 1mmo 1mmw mo>m oumo omw mmua Dame anwa Acossflacovumogswwm 144 mom OwN me 0mm mvm mmN HmN hmm vom mmN MON DHN mHN me mwN HON me OOH mom NON NmH HON IHHthQ>MHmm IImmHM>¢m9>mQH IBEMEZQH Ullm>HM>w>B>DQH¢MMQIIIflmflmmflHHMMQNOHIImeZHQOM lllll HmM> IIHO>OzmmOMMDBmmIHH>mmmmBUm IIIII mmmz EH>mmZDAJSECQdkameUmmmmmBZOHI ImmmwdHNMD IIIIIIIII mEMOHDEI OhmflfiHhZflllleHmm> :HHMmZQHhH m-mMHmHMHmZDmZIUHMmI mKMWHMWmMHBMmmZQ CMCRBMIMNNWMmOXHbIIHMI mm3mmm2; mrmm3¥HZWHmZHHmmmm0 IIIIIII mMHm>IIH¥I mmmMHUMMI ¥H>m IOBEZMHMB>IUOI H>hl EOHOHflmMI mem IOH>mZQ QZmOBmBm>OmmmMBH34mI HhMHHflmmmx IIIIIII mm>Q>IIHmI BBmDMZ>bHm :H>UmZ; >I3m>Z3xIMM>mmmOHmHUMII> IIW >Ovv > i > ONV OOV > > owm OdeDmQQ UOMmIIII¢> IIIIIIIIII Q>UUmEmeHHm< IIIIIIIII mBMMDHHmIIIH¢Mm>m IIIII d1 Z IIIIIIIIII Q>UOm>MmDmHm¢ IIIIIIIII mHMMDHHmIIIHHMM>m IIIII 0| mm IIIII m OfiHbIIIIBm IIIIIIIIII mMHHUIIMthmm IIIIIIIII mMmefiIIIH>OmHQ IIIII m¢BIIIIUMWQ IIIIII mIIMDXZMImm>mHm IIIIIIIII O Hmw<m¥mIIIIHm>mEHmmwmmlmOMHWEIImHKHZ IIIIIIIII U HmmflmmlllmmzmH¥m¢UWEmQ mmMmImlm thdllllWMme IIIIIII EflflmmOIMmm>Mm30QmIW>Q>NWOMMH>M>IIE®EMMIIM >mHHIIII IIIII m HmIHIBIII>m¢mmmwaomOmZmfldemmIH>IllIIIIHBm hQOHmMIIImHMDm4MHBIQ¢m > IIIII m >mImlIIBMW>BWBM>Mm2mMHM4mQHQmMHmfiwflwm2mE>< KOdemHmmmHmOW¢¢¢BIZmIHHmmZM<>HHMSMOEMMMBMMNMWEmmmmmmDEmIIImHHMWHHHOImHB B IIIII m HmBZ>mIIIhEflZMOfimmMIQHDOHHmMWZH * me * ovm >i ONm x u u .. Hmo Nmo muxm amo 1mmw 1mmw mobm oumo omw Mmum Dame «owe Hwo Nmo mu>m qmo :mmw ammo mo>m oumu omw mmua once «owe Acosusneovaaoa:wwm 145 va mNm ONm mmm mmm HmN MPN NHw Hom mmN \IwN I u " Hmo I u ” Nmo mmm " Ioom>qummmwmommmm¢qz " muwm I u IIII “ qmolmmmw I u U mmmw | n n mUNVm I n . onmo I u . omw I " IIII " mmu¢ | u n QDQH I u u mvwe I ovm I wmmem4>>oomqumoxuozmmmmmemmI " Hmo memmmHeawdmmwwmIqmw IIIII zmoBme>mzqzmoxwo3mmmmmem¢I u Nmo zoqz>>owmm3mmm>mmmHqmm IIIII qmomwszquqommqmmwzowmIIII “ Imuxm mOIq>qHImmqqmoma>Iwowommmoommmqmmmm Ho¢eooquml>HooazmmmmMI " oumo Hoqeqmmmwwzmozqmmqom>mmm4mI " omw IIIwmmHBmmqooo>w>m>mwmm>Hamoewwoo>IHAIommmmmmIImmgmcmgmmmmemxmmo mszqzemzmzHZmono " mwmmm nmfi HmafimHOQquIIoozammw 3Mq3flLE23>mm m mmmxmzHMmzmqumeqmm " mme mmH dma _A>¢>Z>>$mmmzwomm¢ HHEJEZLZSDIII E .dXflZFmeINVNxVmeHHO u momm hmfi boo.ZEZmzm>gmwmmzmmmflImecfldfigzzmmIIIme mezmmemmqqmomqum u woe: oom I owfi I omfi vva " m>wmqHmmqaowmmqmmomomoo>xaHzmmmmmwmmMmHB>oImz>quqmm>mx IIIII mo mmmzmmmoHom u :wmmm om " mqqopmeIIImm>IIqmoHaaquzmzm>mmquIXHamwhmmmmowIHmzomm>wHoquM¢>mm " mxem am " z¢mwgqmmIIImm>IIqomo>mommzoom>HmwaummqemIommoo IIIIII ¢w mqoo>mm ” momm pm . mzqqzwHmIIIoo>IIIzommmmwxzqqmm>m IIIII ozwaqummememo IIIIII m mqqmwmqmmm>¢ u $052 I, OVH I. ONH I OOH I ow m» " om>qzmH>q zmmwmxmomqeq>zzoqewomoommmmdmmHBumHmm<>>mw¢mmqumoq>Hoqmm¢¢mqomz n mwmmm mm " ¢mmmmmHMH moIqwmgoom¢mz u flme «N u quooquq flmIIIImwoxqmaqu: u momm om " anamoquo mI>qommmMm>QZqudoz " :05: «I 00 II 0v i ON * .CSEEUV N 03me E mm 08% 2: 2m mooaoswom “Sosa 2: Sm $385: 86302». 53on .moxon x83 53> BEEoE 8m H SEED we “x3 05 E wommsommv mosEmom .8th88 8a AEVTELV own—>553 qm¢m zzHoxqz>mzmmwmmxm¢Hw>>mmm ZMOmem>mem IIIIIIIIII oomm mumH¢xm IwwxoeqdmOmm Hmowamqmwm>HququI>zmmm Hmmwm>2qmmwmwwzwowqqmz¢o quqmmm IIIIII >momq IIII0¢m4m>qIIIImH¢IIIQm mmmmmmmomquIIIIIIIIIIonmz H>q>>wm lllmQUDHZUAm IIIIWZMWUQQfiMflMQHfiIIIlmH FDWMQDmthOOO IIIIIIIII m>WUmU MZQHUmfl I me + Ovm I ONm * h00.¥ O DIHQQmmmh>HQ MMZ>QE>M¢MZIMWM IIIIIIIII QOHU$bZMZ> IIIIIII UUMZMQNDEQMQ MHHMI m UIHHm>EIQH>H >mOHAHMUMHmDQ¢m lllllllll DBZMMWHflmUhm IIIII UHM>ZM$QIIW> mmmm H UI>HMWQMQQHW mUHQWMHUmwfifidmdfififlfiflHQWDQDfiIWQOmAZO¢DUDU¢m¢mmm>QlIHQ mMO.H OEBQmUQh>Z>m ¢m¢0>¢>9mfizwm0m IIIIIIIII DWD IIIIIIIIIIIIIII mom IIIIIIII 00m ¥ OmN ¥ O©N * OWN * Ewmmm dXHm momm E082 Iwmmm fiXHE momm SUEZ 63.558 .3 2:5 148 ova m¢mzo ” mUIumHmm I IIIIIIIIIIIIIII " meO>OHHm> IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII " mmome IIII “ mHIH ¢Hmqumwm<¢moommmommggoIzmmzw>>mm¢9mmm>moo>m>2mHomqmmwmmmmm “ mHZMQ mmmzoummaoowmmmq>Hquzmwmmmmoommmommzquz¢mzw>>mmmoxwo>m>zmHomqm¢wmmMmm ” ommm 9mmHmamwhoowqmmq>m¢mg2mwm<¢¢oommmommzqozzmmzw>mmmmmmo¢>mwo>m>2mHaflgmmwmmmmm u mme mmmzo0mmHowwmmm>>>4qumwm>mmmomom>mwoqm>2mHamgm > I OOH I om mqumHmm XMHO>OHHM> mmomh H mHIH moz>mHmwomqowmzmmdeoooqumowmm>q>o 2 axo>ow>mqmmqflmmHHm& " mHme moz>mHmowmqoomzmmmmHOOQOquoomm>q>ofi «mm» ¢q20>ow>mqmmq>mmszz " ommm moz>MHmowmqoomzmmmmHOOQooqmoomm>q>o 00%ommmwmm>mqmewmmzH¢I ” mme moz>xHmwwQOwmzmmhfiwOOQOmAmowmm>q>mmmozqmmHwy I V;mmmoHom>o<4mqmmqmzm>>% “ wmhm I om 04 I om IF .moxon 3% E 850% 8a 352%: Bataan £383 5;» 850% 0.8 AN “8%:an mEoEEomxo mmmoqowfisfi E @233 $52me .8338 2d 95% F3 8 523 32.53 éRSQ m8”: .2 $38 52925» “gauge ml .% §§§m @3235? “as: as: ”Ema 63m A859 09E .% 38mg atfiofism x$§m<$ 82338ka 3332355 ROMSSZV Avamvvwaz SEES» £292ch 8on 828563 «6.: 3035a 63:93.3 <5; :3th was 329:8 me “EBay—«I .VN ouswmm 149 hm wOH vOH mOH me th wwN th I I I I I I I E 2 H Q m > Z 3 DC I m E m m 0 E m Q H [.11 fl I41 q |||||||||| >>4DQQBH4MNQWMM4BHQWMMOMZH>UMZQ32>HQU>M3W$MM>®KMOE¢$MQQM4AQ u O>3m00mmmHmUMMOQHB4dmQmmm4H>memwme>UMZD32>QQO>MBWEMW>MMQOB4mMHHM4HQ " IIIIIIIIII >>woo>yaammqmmmmeHQwomwmmg>omzasz>qow>mzmmmw>mxmoe¢mmqgm4qq u llllllllll >>4QDHHB4mmHmmm4wHQMmMOM$Q>UMZDBE>HQG>ZBMImw>hMMOB4meHm4HQZEUm * owN * > OwN I > OWN> * mwml IOmwEU4E4thQhMMOmw4IBm>HmEn)LNS4ZMOWMWNDHOURHEMWZH4MIW4MHOOHMKMMDmmQDDN <>mm0mthflEflth.whkaL LndEm>AmBm2mm24zm0mHmNQHQUQHmmmZQ4WIm4mqumemmommHDQW <>mqwm>m04m¢OHhflthwmw42«KDJMJZ>hh2¢2m3mm0mHIX4$4®H®QEMM®WD4E4$>JL§0 mmE§ZMOmmm%0BOOdHhmmZQ4MEm4mQOOQMMMMDmmHDQI <>mm©aHmcflmflthAmmmwmwmmBm>Am3m>mm24zm0mmmwoBowfiHmmmZQmmwmHmmémfimeqmmmwm0¢mBmHQm3m>mm2mm0mHEO4E4UHWHWMMUWU4ZHm>AmBm>mmE4ZMOmmmWQHQUAHhmmZOAMmm4mAOOAWmmQDQmOOQZ 4>m90m>mm4m4 th4; YI4r>Jh§Z>hmE4ZZO4HWMQBQOQHmmmZQ4MIm4m H0mAmmOQQmmfiflflw > > I. >OON I. omH II> OQH Acosnfincov.vaoh=wwm mqumHmm XMHO>OHHM> mmoma mHIH mesmm 0mOHH©> mmoma mHIH mezmm om g 10;- o‘- _l 1 L L In J 1 1 l 1 I I O 5 10 15 20 time(min) Fig. 4. Stereospecificity of dichlorprop-degrading activity of A. denitrificans cell extracts as measured by chiral HPLC. Concen- trations of (5)-dichlorprop (A). (R)-dichlorprop (O), and 2,4- DCP (0) were measured at selected time points after incubation of A. denitrificans cell extracts (50 ul) with racemic dichlorprop (0.1 mM) and the stande reaction components. Dichlorprop was quantitated by using chiral HPLC and 2.4-DCP was quantitated colorimetrically. Data for the (5)-dichlorprop were fit to a straight line of slope 0.05 uM min" ' , for (R)-dichlorprop to exponential decay with an exponent of —0.067 min‘ '; and for 2.4-DCP to Eq. (1) with an apparent inactivation rate of 0.16 min". trificans isolated for its ability to degrade (R)- mecoprop [14]. Cell extracts of IMP134 and RASC exhibited highest activity towards 2,4«D, but showed ~ 20% of this activity towards dichlorprop. We attribute the cell extract activ— ity towards phenoxy herbicides in both these strains to deA, since TnS disruption of the tfdA gene in either strain will prevent the conversion of 2,4-D to 2,4-DCP by intact cells [5,18]. By contrast to the specificity of the abovemen- tioned strains, A. denitnficans extracts were most active with phenoxypropionates including mecoprop, dichlorprop, 2-(2,4-dibromophen~ oxy)propionate, and 2-(4-chlorophenoxy)pro- pionate, with only minor activity towards the phenoxyacetates 2,4-D and 2—chloro-4-methyl- phenoxyacetic acid (~13 fold less than the activity towards dichlorprop). Cell extracts of A. denitrificans demonstrated a slightly higher me towards the dichlorprop than towards mecoprop, the substrate on which it was iso- lated, but slightly less affinity (higher K m) (Fig. 3). The A. denimficans activity was dependent on a-KG and ferrous ion, and was rapidly lost (in 2 min) in the absence of ascorbic acid. These cofactor requirements are characteristic of or~KG-dependent dioxygenases. 3.3. Stereospecificity of activity in cell extracts The stereospeeificity of herbicide decomposi- tion by cell extracts was measured using an HPLC column with a chiral stationary phase, which permitted direct separation without prior derivatization of the enantiomers of dichlorprop. In agreement with the results using purified enzyme, cell extracts of JMP134 degraded the S enantiomer. Similar results were obtained with 150 160 170 180 190 200 RAAYDALPRDLQSELEGLRAEHYALNSRFLLGDTDYSEAQRNAMPPVNWPLVRTHAGSGRK IIIII II:I:::II:IllllIlllIIlI:IlIIIIII:IIIII|II:IIIIIIIIIIIII A . deni tri fi cans RAAYDDLPEDFKKELQGLRAEHYALNSRFI LGDTDYSESQRNAMP PVSWPLVRTHAGSGRK llllIIIIIIIIllIllttlIllllzlllllllI:tll|l|lllll|lIII:IIIIIIIII JMP134 RASC RAAYDDLPEDFKKELQGLRAEHYALH S RF I LGDTEYS ESQRNAMPPVSWPLI RTHAGS GRK 170 180 190 200 210 220 2 l O 2 2 0 2 3 0 2 4 O 2 5 0 JMP13 4 FL FI GAHASHVEGLPVAEGRMLIAELLEHATQREFVYRHRWNVGDLVM IIIIIIII:I=II lltlllllllllIlIlllI:IlIIl Izllllll A. deni tri fi cans FLFI GAHAGHIEGRPVAEGRMLLAELLEHATQRKFWRHSWKVGDLVM IIIIIlllzlllllllllllltllllllllll Illllllllllllll RASC FLFIGAHASHIEGRPVAEGRMLLAELLEHATQPKFVYRHSWKVGDLVM 230 240 250 260 270 Fig. 5. Comparison of an A. denitrificans deA-like sequence with the enzymes from RASC and JMP134. The deduced sequence encoded by the PCR amplification product arising from A. denitrificans DNA was compared to the indicated portions of the RASC and JMP134 deA sequences. Identical amino acids are indicated by a solid vertical bar and similar amino acids by a dashed bar. 158 R.E. Saari et al. / Journal of Molecular Catalysis B: Enzymatic 6 (1999) 421 —428 427 RASC extracts (data not shown). By contrast, A. denitrificans extracts degraded (R)-dichlor- prop (Fig. 4) with concurrent 2,4-DCP produc- tion. 3.4. Molecular biological comparison of strains The similarity in cofactor requirements among the three strains raised the question of whether the A. denitrificans activity was due to a gene related in sequence to those which encode the two deA enzymes in the other species. A. denitrificans genomic DNA hybridized at low stringency to a probe for tfdAJMPm and at high stringency to a segment of tfdARASC (data not shown). Amplification of A. denitrificans DNA with primers specific for tfdA yielded a 360-bp DNA fragment with a sequence that was 94% identical to that from RASC and 78% to that from JMP134, in a 327-bp overlap. The trans- lated sequence (109 amino acids long) had 95% identity with deA from RASC and 86% iden- tity to that from JMP134 (Fig. 5). 4. Discussion 4.1. R. eutrophus JMP134 deA oxidizes the S isomer of phenoxypropionates a—KG dependent dioxygenases typically ex— hibit stereoselectivity in the hydroxylations they catalyze (e.g., Refs. [28—31]); thus, it was not surprising to find that R. eutrophus JMP134 deA oxidizes a single enantiomer (the S iso— mer) of dichlorprop. Based on this result with phenoxypropionate, the enzyme likely catalyzes stereochemically selective oxidation at the pro- R position when acting on the prochiral 2,4—D. 4.2. Related gene products can degrade oppo— site stereoisomers The above studies demonstrate that the herbi- cide-degrading activity of A. denitrificans ex- hibits opposite stereochemistry than that found in strains JMP134 or RASC; yet, each of] these activities are a-KG-dependent and the enzymes are likely to be genetically related. The present work extends the conclusions derived from stud- ies with a strain of Sphingomonas herbicidouo- rans that is capable of degrading both enan— tiomers of meCOprop [32]. In that strain, two a—KG-dependent dioxygenases appear to be pre- sent, each specific for a separate enantiomer of the substrate [33]. Based on our findings, it is probable that the two S. herbicidouorans en— zymes are closely related in sequence. The abil- ity for closely related enzymes to recognize opposite stereoisomers has precedent. For ex- ample, recent crystallographic characterization of two tropinone reductases demonstrate the presence of almost identical overall protein fold- ing, but the closely related enzymes exhibit opposite substrate stereoselectivities due to dif- ferences in a few charged residues at the active site [34]. 4.3. Biotechnological implications The work described here may have biotech- nological applications. An S isomer selective dichlorprop-degrading enzyme, such as deA, may be used to facilitate the preparation of the desired R isomer, while a gene encoding an R isomer selective enzyme, such as that in A. denitrificans, may be useful for development of herbicide-resistant transgenic plants. Also, in understanding the role of microorganisms in the environmental fate of phenoxypropionates, it is useful to have analytical techniques such as those used in this research capable of distin- guishing the enantiomers. Acknowledgements We thank Peter Chapman for suggesting that dichlorprop could be used to probe the stere— ospecificity of deA, Paul Loconto of the Engi- neering Research Labs of MSU for assistance in setting up the GC method and suggesting the 159 428 RE. Saari et al. /Journul ofMo/ecular Catalysis 8: Enzymatic 6 (1999) 421—428 use of electron capture detection, and Hilary Lappin-Scott for providing us with the meco- prop-degrading strain of A. denitrificans. This work was supported by the National Science Foundation (MCB9603520), the National Sci— ence Foundation—Science and Technology Cen- ter Grant DEB9120006, and the Michigan State University Agricultural Experiment Station. References [1] MA. Loos, in: RC. Kearney, D.D. Kaufman. (Eds) Herbi- cides: Chemistry, Degradation and Mode of Action, V01. 1, Marcel Dekker, New York, 1969, p. 1. [2] IUPAC, Pure Appl. Chem. 69 (1997) 1335. [3] E. Yabuuchi, Y. Kosako, I. Yano, H. Hotta, Y. Nishiuchi, Microbiol. Immunol. 39 (1995) 897. [4] F. Fukumori, R.P. Hausinger, J. Bacteriol. 175 (1993) 2083. [5] W.R. Streber, KN. Timmis, M.H. Zenk. J. Bacteriol. 169 (1987) 2950. [6] RP. Hausinger, F. Fukumori, D.A. Hogan, T.M. Sassanella, Y. Kamagata, H. Takami, RE. Wallace, in: K. Horikoshi, M. Fukuda, T. Kudo (Eds). Microbial Diversity and Genetics of Biodegradation, Japan Scientific Press, Tokyo, 1996, p. 35. [7] P. Ludwig, W. Gunkel, H. Hiinerfuss, Chemosphere 24 (1992) 1423. [8] R. Martens, Pestic. Sci. 9 (1978) 127. [9] M. Négre, M. Gennari, V. Andreoni, R. Ambrosoli, L. Celi, J. Environ. Sci. Health B 25 (1993) 545. [10] A.E. Smith, A.J. Aubin, Can. J. Soil Sci. 70 (1990) 343. [11] A.E. Smith, Bull. Environ. Contam. Toxicol. 34 (1985) 656. [12] A.E. Smith, J. Agric. Food Chem. 33 (1985) 483. [13] A.E. Smith, J. Agric. Food Chem. 25 (1977) 893. [14] V.A. Tett, A.J. Willetts, H.M. Lappin-Scott, FEMS Micro- biol. Lctt. 14(1994) 191. [15] BR. Lyon, Y.L. Cousins, DJ. Llewellyn, ES. Dennis, Transgen. Res. 2 (1993) 162. [16] L. Feng, l.R. Kennedy, Biotech. Bioeng. 54 (1997) 513. [17] RH. Don, J.M. Pemberton. J. Bacteriol. 145 (1981) 681. [18] Y. Suwa, A.D. Wright, F. Fukumori, K.A. Nummy, R.P. Hausinger, W.E. Holben, L.J. Forney, Appl. Environ. Micro- biol. 62 (1996) 2464. [19] R.J. King, K.A. Short, R.J. Seidler. Appl. Environ. Micro- biOl. 57 (1991) 1790. [20] F. Fukumori, RP. Hausinger, J. Biol. Chem. 268 (1993) 2431 1. [21] R. Stanier, N. Pelleroni, M. Douderoff, J. Gen. Microbiol. 43 (1966) 159. [22] B. Blessington, N. Crabb, S. Karkee, A. Northage, J. Chro- matogr. 469 (1989) 183. [23] J. Sambrook, E.F. Fritsch, T. Maniatis. Molecular Cloning: A Laboratory Manual, 2nd edn., Cold Spring Harbor Labora- tory Press, Cold Spring Harbor, NY, 1989. [24] WE. Holben, B.M. Schroeter, V.G.M. Calabrese, R.H. Olsen, J.R. Kukor, V.O. Biederbeck, A.E. Smith, J.M. Tiedje. Appl. Environ. Microbiol. 58 (1992) 3941. [25] DA. Hogan, D.H. Buckley, CH. Nakatsu, T.M. Schmidt, R.P. Hausinger, Microb. Ecol. 34 (1997) 90. [26] Wisconsin Package Version 9.0, Genetics Computer Group (GCG). Madison, WI. [27] RE. Saari, RP. Hausinger, Biochemistry 37 (1998) 3035. [28] J.E. Baldwin, R.A. Field, C.C. Lawrence, K.D. Merritt, C.J. Schofield, Tetrahedron Lett. 34 (1993) 7489. [29] J.E. Baldwin, K.D. Merritt, C.J. Schofield, S.W. Elson, K.H. Baggaley, J. Chem. Soc., Chem. Commun. (1993). 1301. [30] S. Englard, J.S. Blanchard, CF. Midelfort. Biochemistry 24 (1985) 1110. [31] E. Leete, D.I-I. Lucast, Tetrahedron Lett. 38 (1976) 3401. [32] C. Zipper, K. Nickel, W. Angst, H.-P. Kohler. Appl. Envi- ron. Microbiol. 62 (1996) 4318. [33] K. Nickel, M.J.-F. Suter, H.-P.E. Kohler. J. Bacteriol. 179 (1997) 6674. [34] K. Nakajima, A. Yamashita, H. Akama, T. Nakatsu, H. Kato, T. Hashimoto, J. Oda, Y. Yamada, Proc. Natl. Acad. Sci. 95 (1998) 4876. 160 R "‘tiiitittiiiit e 5