% imp . . 2 2 .3 22 2.. fiudawfivémw yfianwmw LP . .. 4 i ummfls 66%.. w. . . ,t..£. . . . hut 2 2 2:2 5 , wisp?» .. aw, m... . ,. $2.... .Uwuhqfilw a . .l 5. rhulflW-H r (I. . 22.32.. ,m .‘fl I. I anal} . 15““? 2 . .. c 3.1.73. . flit. a...” fifikflrnfifin 3&2}. . 2 . III: e a . ha 2!. .zu2234fiuflhw . 51v ‘1 .v‘!‘ .ung l.“ , .22 .hu»?§itisfl.u2 :1... iii... :21- .212? 2.... 12.222.223.222! 5 , Lav). our-IVA xvi... .21}... 1...... are . .. . 1.. 322...... {ft .1. (.1 a. 2.! 142: . ”.2122. u... I543. . I. Itiv 1).... Alt full... ..I.Irl.n..’hly....zx!u It...“ . . . .‘.t‘ F1 FIGURE 4.2: Denaturing polyacrylamide gel electrophoretic analysis of Xan variants purified from E. coli. ......................................................... 133 FIGURE 4.3: Fluorescence spectroscopic analysis of XanA interaction with Fe(II). ..................................................................................... 135 FIGURE 4.4: Fluorescence spectroscopic analysis of XanA interaction with OLKG .................................................................................... 137 FIGURE 4.5: UV-visible spectra of 1-methy1xanthine, 9-methy1xanthine, and their products. ............................................................................ 142 FIGURE 4.6: Interaction of XanA with 6,8-DHP. .................................. 144 FIGURE 4.7: General mechanism for slow-binding inhibition. ................... 145 FIGURE 4.8: Interaction of Q101A with 6,8-DHP. ................................. 147 FIGURE 4.9: Interaction of K122A with 6,8-DHP. ................................. 148 FIGURE 4.10: Interaction of E1 37A with 6,8-DHP. ................................ 149 FIGURE 4.11: Interaction of D138A with 6,8-DHP. ................................ 150 FIGURE 4.12: Interaction of Q356A with 6,8-DHP. ................................ 151 FIGURE 4.13: Interaction of C357A with 6,8-DHP. ................................ 152 FIGURE 4.14: Interaction of N358A with 6,8-DHP. ................................ 153 FIGURE 4.15: Inactivation of XanA by DTNB. .................................... 156 FIGURE 4.16: Time-dependent loss of XanA activity for highly diluted enzyme. .................................................................................... 158 FIGURE 4.17: Oxygen consumption studies of XanA and its variants. .......... 160 FIGURE 4.18: Hydrogen peroxide production by Q101A and wild-type XanA.163 2 FIGURE 4.19: Proposed xanthine binding on the active site. ..................... 166 CHAPTER 5 . FIGURE 5.1: TauD active site. ......................................................... I80 xiv FIGURE 5.2: General mechanism of TauD. ......................................... 181 FIGURE 5.3: UV-visible absorption spectra of XanA complexes. ............... 188 FIGURE 5.4: UV-visible absorption spectra of {FeNO}7 complexes of TauD and XanA. ................................................................................ 189 FIGURE 5.5: Thermodynamics and kinetics of NO binding to TauD and XanA. ..................................................................................... 191 FIGURE 5.6: EPR spectra of the {FeNO}7 complexes of TauD and XanA. 193 FIGURE 5.7: One-dimensional ESEEM spectra of different TauD complexes prepared in H20 and 60% 2H20 buffer at 172.0 mT and 340.0 mT. .............. 195 FIGURE 5.8: One-dimensional ESEEM spectra of different XanA complexes prepared in H20 and 60% szO buffer at 172.0 mT and 340.0 mT. .............. 198 FIGURE 5.9: HYSCORE spectra of different TauD complexes prepared in H20 and 60% 21120 buffer at 172.0 mT. ............................................. 203 FIGURE 5.10: HYSCORE spectra of taurine-aKG-Fe(II)-TauD complex prepared in H20 buffer at 172.0 mT (A), 340.0 mT (C) and 60% 2H20 buffer at 172.0 mT (B), 340.0 mT (D). ....................................................... 204 FIGURE 5.11: HYSCORE spectra of different TauD complexes prepared in H20 and 60% szO buffer at 340.0 mT. ............................................. 206 FIGURE 5.12: UV-Vis spectra and HYSCORE of ternary complex taurine-NOG-Fe(II)-TauD. ............................................................ 208 FIGURE 5.13: HYSCORE spectra at 172.0 mT of taurine-zH-labeled-aKG -Fe(II)-TauD prepared in 60% 2&0 buffer (A) and 90% szO buffer. ......... 209 FIGURE 5.14: HYSCORE spectra of 2H-Iabeled taurine-aKG-Fe(II)-TauD complex prepared in H20 buffer at 172.0 mT (A) and 340.0 mT (B). ........... 210 FIGURE 5.15: HYSCORE spectra of taurine-aKG-Fe(II)-TauD complex prepared in 60% 2H20 buffer at 172.0 mT (A), 182.0 mT (B) and 192.0 mT (C). ........................................................................................ 214 FIGURE 5.16 Two depictions of TauD active site and the XanA active site. 215 FIGURE 5.17: HYSCORE spectra at 172.0 mT of taurine—aKG-Fe(I.I)-TauD XV 1V:- HI~ Hf Flt» Hg llt . pk" at I Fe but Cl‘ Flt (ng W248F prepared in H20 buffer (A) and 60% 2H20 buffer (B). .................. 216 FIGURE 5.18: HYSCORE spectra of different XanA complexes prepared in H20 and 60% 2&0 buffer at 172.0 mT. ............................................. 217 FIGURE 5.19: HYSCORE spectra of different XanA complexes prepared in H20 and 60% 21120 buffer at 340.0 mT. ............................................. 218 FIGURE 5.20: HYSCORE spectra of xanthine-aKG-Fe(II)-XanA complex prepared in H20 buffer at 172.0 mT (A), 340.0 mT (C) and 60% 2H20 buffer at 172.0 mT (B), 340.0 mT (D). ....................................................... 219 FIGURE 5.21: HYSCORE spectra at 172.0 mT of Fe(II)-TauD (A), (XKG- Fe(II)-TauD (B), and taurine-aKG-Fe(II)-TauD complex prepared in H20 buffer. ..................................................................................... 226 CHAPTER 6 FIGURE 6.1: HYSCORE spectra of NO treated ternary complex TauD-Fe- aKG-taurine in 50 mM Tris-H20, pH = 8.0 at (A) 3400 G and (B) 1720G 244 xvi aKG ABTS CAS 2,4-D DAOCS DEA/NO 2,8-DHP 6,8-DHP DSBH DTNB EPR ESEEM F3H F IH HIF HPPD HYSCORE IPNS JHDM LMCT ABBREVIATIONS a-Ketoglutarate 2,2’-Azino-bis(3-ethylbenzthiazoline—6-sulfonic acid) Clavaminate synthase 2,4-Dichlor0phenol Deacetoxycephalosporin C synthase Diethylammonium (Z)-1-(N,N-diethylamino)diazen-l-ium -1,2-diolate 2,8-Dihydroxypurine 6,8-Dihydroxypurine Double-stranded B-helix 5,5’-Dithiobis(2-nitrobenzoic acid) Electron paramagnetic resonance Electron spin-echo envelope modulation Flavanone 3fl-hydroxylase HIF-a-specific asparaginyl hydroxylases Hypoxia —inducible factor 4-Hydroxyphenylpyruvate dioxygenase Hyperfine sublevel correlation spectroscopy Isopenicillin N synthase ijC domain-containing histone demethylase Li gand-to-metal charge-transfer xvii .\1l M1 N1? \l 01 n4 ll.|.1 ll: J! P3 MLCT Moco NO NOG NTA OPDA P3H P4H PNGase F SDS-PAGE TauD deA XAS XanA XO Metal-to-ligand charge-transfer Molybdopterin cofactor Nitric acid N-oxalylglycine Nitrilotriacetic acid o-Phenylenediamine Proline 3—hydroxylase Prolyl 4-hydroxylases N-glycosidase F Sodium dodecyl sulfate-polyacrylamide gel electrophoresis Taurine/OLKG dioxygenase 2.4-D dioxygenase X-ray absorption spectroscopy Xanthine/(XKG dioxygenase. Xanthine oxidase xviii CHAPTER 1 INTRODUCTION (if d1 31 Fe(II)/a—KETOGLUTARATE-DEPENDENT HYDROXYLASES Fe(II)/a-ketoglutarate (aKG)—dependent hydroxylases catalyze a wild range of reactions including protein side-chain modification (some of which are involved in transcriptional regulation), repair of alkylated DNA/RNA, biosynthesis of antibiotics and plant products, lipid metabolism, and synthesis or decomposition of a wide variety of small molecules (1-13). Structural studies reveal a conserved double-stranded B-helix (DSBH) core of this superfamily, also called a B-strand “jellyroll” structure. This conserved structure is responsible for binding the iron through three amino acid side chains in a His'-X-Asp/Glu-X,,-His2 motif (where n can range from 44 to over 150). In addition, the metal is bound by aKG using its C-l carboxylate and C-2 keto groups. Binding of co-substrate and substrate triggers the ligation of dioxygen to metal, stimulates the oxidative decomposition of aKG to succinate and C02, and leads to the formation of a highly reactive Fe(IV)-oxo species which is proposed to hydroxylate the primary substrate. The overall reaction is illustrated in Figure 1.1. In this chapter, I briefly introduce several representatives of this superfamily of enzymes, including selected Fe(II)/0tKG-dependent hydroxylases as well as others that are structurally or mechanistically related. I discuss the general structural and mechanistic features of these enzymes based on kinetic and spectroscopic studies. Finally, I provide an outline of my thesis to highlight the purpose of my studies. FIGURE 1.1: General reaction of Fe(Il)/0tKG-dependent hydroxylases. Fe(ID/aKG—dependent R-H hydroxylases ; R-OH 0H ' 0 H0 0 0 + 002 H00 H00 aKG succinate FIGURE 1.2: General reaction of prolyl 4-hydroxylases (P4H). 0 R §~R R\ §—R N “ P4H N O 7’" 02 + “KG CO2 + succinate OH Prolyl side chain 4-Hydroxyprolyl side chain FIGURE 1.3: General reaction of HIF—a-specific asparaginyl hydroxylases (FIH). 02 + aKG C02 + succinattce3 Asparaginyl side chain [S-Hydroxy asparaginyl side chain Protein Modification Prolyl 4-hydr0xylase Prolyl 4-hydroxylase (P4H), the first Fe(II)/aKG-dependent hydroxylase identified (14), catalyzes the hydroxylation of proline residues to yield the trans-4-hydroxyprolyl group as shown in Figure 1.2. In mammals, P4H is the key enzyme in the biosynthesis of collagens (2), a family of extracellular matrix proteins, as 4—hydroxyproline residues are essential for the folding of the newly synthesized collagen polypeptide chains into triple-helical structure. The enzyme recognizes the Gly-X-Pro motif and about 10% of the Pro position is modified. P4H has also been isolated from plants (15, 16), where it hydroxylates proline-rich structural glycoproteins of the cell walls. Genes encoding P4H have been cloned from human, plant, insect, nematode, and other sources, including Paramecium bursaria Chlorella virus-l (15-20). Generally, the protein is composed of two subunits, a catalytic subunit on, which contains separate catalytic and peptide substrate-binding domains, and a protein disulfide isomerase subunit, [3, except for algal and plant enzymes which are identified as monomer (15, 16). His412, Asp414 and His483 were proposed to be the metal ligands for human enzyme via a mutagenesis study (21, 22), and aKG is thought to chelate the Fe(II) center in a bidentate fashion as described above with the C-5 carboxylate forming a salt bridge with Lys493. The peptide substrate-binding domain consisting of residues 144-244 of the human 0; subunit was crystallized and diffracted to at least 3A (23), however, there is no complete structure reported since the catalytic domain was not included in the crystal. Transcriptional Regulation Prolyl hydroxylase domain-containing enzymes & F actor-inhibiting HIF Fe(II)/aKG-dependent hydroxylases were known for decades to catalyze protein post-translational modifications (24-26), and included the prolyl, lysyl, and aspartyl(asparaginyl) hydroxylases. Recently, some representatives of these enzymes were shown to play a novel role involving regulation of the hypoxic response, one of the most important ways in which animals respond to reduced levels of dioxygen (3, 4, 27-29). Hypoxia —inducible factor (HIF) is an (113 heterodimeric transcription factor enabling the transcription of an array of genes that work to compensate for the effects of low oxygen tension. Under low oxygen condition, the HIF-a subunit translocates to the nucleus, dimerizes with HIF-B and together they recruit coactivators, such as p300, to initiate a transcriptional response. Under normal dioxygen concentraions, HIF-(x is targeted for degradation by two independent pathways. One path involves hydroxylation of either Pro402 or Pr0564 by HIF-a-specific prolyl hydroxylase domain-containing enzymes, using an oxygen, Fe(II), and OtKG-dependent reaction (3, 4, 30). The hydroxylated HIF-(x forms a complex with the von Hippel-Lindau tumor suppressor protein, elongin B, and elongin C, resulting in polyubiquitinylation and destruction of the transcription factor subunit. The other pathway involves hydroxylation at the pro-S position of the B-carbon of Asn803 in HIF-ot through factor-inhibiting HIF (FIH) (5, 31-33) as shown in Figure 1.3, The action of this HIF-a-speciflc asparaginyl hydroxylase prevents interaction of HIF-oz with the p300 transcription coactivator, thus repressing HIF transcriptional activity. An X-ray bf 311 3C1 pr: CU] of 8p: an the CI. crystallography study predicted that FIH is comprised of a B-strand jellyroll core with both Fe(II) and the co-substrate OLKG bound in the active site (34). The metal ligands are Hisl99, Asp201, and His279, whereas Tyr145, Thr196, and Ly5214 stabilize binding of the (xKG C-5 carboxylate in a unique type of interaction. The crystal structure of FIH, in complex with Fe(II), OLKG and the C-terminal transactivation domain of HIP-0t, was also obtained. Asn803 of CAD is precisely orientated in the active site to allow hydroxylation to occur at its B carbon; however, oxidation was prevented by the anaerobic conditions used for crystallization. Similarly, anaerobic conditions in the cell prevents HIF hydroxylation allowing it to function as a transcription factor. Jumonji C (ijC)- domain-containing histone demethylases Covalent histone modifications have an important role in regulating a wide range of processes including gene activity, chromatin structure, dosage compensation and epigenetic memory (35). One such modification is methylation, occurring on arginine and lysine residues, the extent of which is controlled by a balance between enzymes that catalyze the addition and removal of this modification (36). It was long thought that histone methylation was irreversible until a novel ijC domain-containing protein, JHDMI (ijC domain-containing histone demethylase 1), that specifically demethylates histone H3 at lysine 36 (H3K36) was discovered in 2005 (3 7). From then on, more JHDM members have been identified and JHDMs have become the third and largest class of demethylase enzymes. PADI4 (petidylarginine deiminase 4) converts methyl-arginine to citrulline as opposed to directly reversing arginine m: ll} 16‘ 111; Ii}; 11] l'C‘. yl methylation, so it cannot strictly be considered a histone demethylase (38, 39). LSDl (lysine specific demethylase 1), a representative of a second class of enzymes, directly reverses histone H3K4 or H3K9 modifications by an oxidative demethylation reaction in which flavin is cofactor (40, 41). It is worth noticing that the state of histone methylation, in addition to the site of lysine modification, is important in determining the functional outcome of this epigenetic modification. Unlike the other two classes, which can only remove mono- and dimethyl lysyl modifications, the JHDMs can remove all three histone lysine-methylation states. In this very active area of investigation, JHDMs already have been shown to reverse H3K36 (JHDMI) (37), H3K9 (JHDM2A) (42) and both H3K9 and H3K36 (JHDM3 and JMJDZA-D) methylation (43-46). A set of 98 ijC-domain-containing proteins from human to yeast has been reported based on the analysis of public protein-domain databases (47). Curiously, FIH, functioning as an asparagine hydroxylase for the HIF-a transcription factor, was included in this enzyme family. Even though FIH is only found in higher eukaryotes such as mice and humans, it is still a useful template for study of other ijC-domain-containing proteins. The reaction catalyzed by JHDMs uses Fe(II) and OLKG as cofactors to hydroxylate the methyl group of the modified lysyl side chain of histones, with the resulting intermediate spontaneously decomposing as shown in Figure 1.4. In agreement with the reaction, formaldehyde and succinate were detected during the demethylation by JHDMlA (a human JHDMl homologue) (37). Based on the recently solved structure (2.28 A) of the catalytic domain of JMJDZA (48), the ijC dc 311 hr. domain folds into eight conserved B-sheets forming the typical jellyroll-like structure. The iron atom is chelated by three absolutely conserved residues: Hisl88, Glu190, and Hi5276. Two water molecules binding to the metal center also were observed in the native structure. In the presence of OLKG, the two water molecules are replaced by two oxygen atoms from aKG which associate with Fe(II) via the C-1 carboxylate and C-2 keto groups. aKG is further stabilized by three hydrogen bonds formed between OLKG and the side chains of Tyrl32, Asn198 and Lys206. The potential substrate binding site and binding mechanism were proposed, but a complex structure of JMJD2A and its cognate substrate peptide is still required to verify this model. Repair of Alkylated DNA/RNA AlkB Cellular DNA can be damaged by various intracellular and environmental alkylating agents to produce alkylation base lesions which may cause genetic changes that lead to diseases such as cancer (49-51). In E. coli, an adaptive responsive pathway mediated by Ada protein is initiated by this treatment (52). Methyl transfer to Cys69 of Ada converts it into an activator that increases expression of alkB and two other genes, alkA and aidB. AlkA is a glycosylase that cleaves specific methylated bases from DNA and performs the first step of the well-known base excision repair pathway of base lesions. The precise function of AidB is still unknown. The activity of AlkB was unknown for almost 20 years after its involvement in alkylation damage was shown; however, based on phenotypic studies, AlkB was known to function as a DNA repair enzyme that demethylates l-methyladenine and 3-methylcytosine of FIGURE 1.4: General reaction of ijC—domain-containing histone demethylases (J HDM). .CHa CH20H +H2N ‘ +H2N Spontaneous 0Ez + u—fffm GCO2 + succinate HCHO R s N R H O H o Monomethylated lysyl side chain Lysyl side Chill“ FIGURE 1.5: General reactions of AlkB. NH2 NH; NH; 4. I” «II: °“’°H ”f“ / / /d transitions belonging to the six-coordinated distorted octahedral center; these investigations also showed that (XKG binding is reduced in the absence of metal ion, confirming that the metal ion binds prior to the co-substrate (138, 145, 146).This octahedral arrangement of the Fe(II) ion was also indicated by the XAS studies with deA (86). Moreover, the Mdssbauer spectrum of 57Fe-TauD shows a signal with an isomer shift 5: 1.27 10.05 mm/s and quadrupole splitting AEQ = 3.06 10.05 mm/s, consistent with high-spin F e(II) in the active site (143). In summary, the holoenzyme species of this superfamily contains a distorted six-coordinated high-spin Fe(II) ion center, surrounded by two His, one Asp/Glu and three water molecules. Two of the solvent molecules were replaced upon binding of aKG which exhibits bidentate interactions with the iron atom via C-1 carboxylate and C-2 keto groups, Figure 1.15B. Crystal structures showing this bidentate association of orKG 29 FIGURE 1.15: Proposed mechanism for the hydroxylation reaction catalyzed by Fe(II)/aKG-dependent dioxygenase. See text for a description of the various intermediates. xi ”.1 o: A .0 o 9:38... s o. o 92>”? ma... . \..mo.... 0 o m :.a\ \ 0\1/ I.“ Inc o 80. 80 ~ 1% >0.E>mu.\..0 ...u ._ ..... l1~o I.u\ ~0/ InO ....u .../h... :0. QE~>GU ....... ... I.m\m,o. .0 til I.» a... c». P: 0w 0. o 1 on 9:32.... ..u ...... o o 9:33. ...... no.2. o o. \ \ o z.u\\ 1.9.0.1 :.m\ :80 80. r 000. 80.. j $10.0 . c at o . O\ 0:553. ...... .102 055$... ..._<./...01110 0 m I-“\ ~lo I—uvm‘al DEV/P/ :5 .. 000. 15 000. X are available for many enzymes, such as FIH, CAS, DAOCS and AtsK (5, 6, 118, 120). The five-membered ring formed by OLKG binding to the metal is associated with the metal-to—ligand charge-transfer (MLCT) absorption transitions with visible maxima around 500 nm that can be observed in the absence of oxygen. These transitions are now accepted as characteristic of bidentate (IKG binding to the metal ion even though the intensities and maxima of these transitions differ slightly in each enzyme, indicating differences in overlap between d orbitals of Fe and 1t. orbitals of OtKG among different enzymes (142). The lilac-colored chromophore exhibits a 3...... at 530 nm for both (xKG-Fe(II)-TauD and orKG-Fe(II)-deA (140, 141), whereas the feature is observed at 500 nm in orKG-Fe(II)-AlkB (I). Resonance Raman (RR) spectroscopy detected two vibrational transitions with the complex orKG-Fe(II)-TauD in H20: 470 cm"I indicative of metal-ligand stretching vibrations, and 1688 cm'1 due to the C=0 vibration of orKG (147). There are two types of coordination chemistry shown in Figure 1.15 for the OtKG bound state, intermediate B (in line) and B’(off line). As described before, the C-1 carboxylate is trans to the proximal His ligand for the in-line binding mode while it is trans to the distal His ligand for the off-line binding mode. Under the off-line binding mode, the ferryl-oxo initially is oriented away from the substrate binding site. In order to accomplish substrate oxidation, migration is required of either the metal ligands or the ferry] intermediate in the later step of the catalytic cycle. The primary substrate is bound nearby, but not in contact with, the metal ion at the active site, Figure 1.15C, however, substrate binding has a direct influence on the 31 metal’s coordination number. Crystal structures of substrate-(XKG-Fe(II)-enzyme complex has been resolved for a few proteins, such as FIH, CAS, TauD and AlkB (6, 34, 78, 79). These structures indicate that substrate binding induces a conversion from six-coordinate octahedral to five-coordinate square-pyramidal geometry by the release of the last H20 molecule. CAS is an exception by only exhibiting an enlonged metal to H20 bond distance (6). The departure of the H20 molecule is critical for enzyme activity since 02 is supposed to bind in this open site to carry out the hydroxylation reaction. It was shown that poor substrate binding sometimes retains six-coordination while inclusion of native substrate almost invariably induces five-coordinate rearrangements (6, 78, 121). One exception is AlkB, where the substrate-bound enzyme retains a six-coordinate metal site (63, 148). Multiple types of spectroscopies have been applied to monitor the effects of primary substrate binding on the active site in the presence of orKG Upon substrate binding, the UV/Vis spectra often exhibit a perturbation of the MLCT features including a slight increase in intensity, greater resolution of the transitions, and a blue shifi (resulting in maxima at 520 nm for TauD and 515 nm for deA) (140, 141). For TauD, RR also exhibited 10-cm’l shift in features to 470 cm" and 1688 cm'I in the presence of taurine (1 4 7); this result was attributed to a switch from six-coordination to five-coordination. Also, the Mbssbauer spectra of taurine-aKG-Fe(II)-TauD complex gave a signal with an isomer shift 5 = 1.16 $0.05 mm/s and quadrupole splitting AEQ= 2.76 $0.05 mm/s, in accord with a reduction in coordination number (143). For CAS, the d—>d transitions in the presence of substrate identified by near-infrared MCD were in agreement with a . 32 five-coordination site (145, 14G. Substrate binding in the off-line mode is shown as Figure 1150. It was reported that at least one third of the available crystal structures of this superfamily indicate that the sixth site on the ferrous ion in the tertiary complex is orthogonal to the substrate (142). Crystal structure studies on CAS and AlkB provide good examples of enzymes that likely require rearrangements around the active site (6, 63, 129). From intermediate C to the end of catalytic cycle is the least understood part of the mechanism. Only one intermediate, a ferryl-oxo species, has been directly identified. This species was first detected in TauD, but has also been observed in prolyl-4-hydroxylase (P4H) and the halogenase Cth3 (143, 149—151). This intermediate from TauD has been investigated most thoroughly. The presumed ferryl-oxo species exhibited an absorption near 318 nm by UV/Vis stopped-flow spectroscopy (143). Using rapid freeze-quench techniques to trap the intermediate and carrying out subsequent analysis by EPR and Mbssbauer spectroscopy, a species with isomer shift 8: 0.31 10.03 mm/s and quadrupole splitting AEQ= 0.88 10.03 mm/s was detected and assigned to an integer spin with S2 (143). Cryoreduction of this intermediate formed a high-spin Fe(III) species; thus, the intermediate was identified as some type of Fe(IV) species (143). Use of C-1 deuterated taurine decreased the rate of decay of this intermediate by 37 fold, indicating that the Fe(IV) species participates in hydrogen abstraction from the substrate (152). Furthermore, RR and cryogenic continuous flow were applied to directly identify this intermediate as Fe(IV)=02’ by detecting its isotope-sensitive vibrations (787 cm'1 for '80 and 821 cm" for 160) (149). 33 The Fe-O distance of 1.62 A was determined by EXAFS, consistent with literature Fe(IV)-oxo models (153-155). The identification of Fe(IV)=02' is a big step for understanding the mechanism of this enzyme superfamily, but there are still many unsolved questions. None of the 02-dependent intermediates leading to or following Fe(IV)=02'has been identified. Even though we knew that Fe(IV)=02' is responsible for the hydrogen atom abstraction from TauD, whether or not this highly reactive species has the capacity to facilitate all of the observed reactions of this superfamily is unknown. Alternatively, different enzymes might use different activated oxygen species to carry out their unique reactions. Nevertheless, this general mechanism involving formation of Fe(IV)=02' is very appealing and could reasonably accommodate the desaturation, ring-expansion, epimerization, halogenation and other reactions of this enzyme superfamily. For instance, CAS catalyzes cyclization and desaturation reactions from proclavaminic acid. The ferryl-oxo species could abstract the C4, (S) hydrogen as illustrated in Figure 1.16, which leads to intermediate B, with Fe(III)-OH and the C4’-centered radical (156). An attack of the C3-bound hydroxyl at C4'coupled with hydrogen atom transfer to Fe(III)-OH leads to the product complex C. Desaturation can be explained in a similar way. The Fe(IV)=02' species could first hydroxylate the substrate, followed by a dehydration reaction to form the double bond. Alternatively, the Fe(IV)=02’ could initiate two hydrogen atom transfers from the substrate directly to the product, Figure 1.17. Regardless of the diversity of overall reactions (that are observed, the initial processes of hydrogen abstraction could be quite similar for each of the enzymes. The most recently discovered subgroup of 34 Fe(II)/orKG-dependent dioxygenase, the halogenases, replace the monodentate carboxylate of the facial triad with a halide ion, Figure 1.18. The Fe(IV)=02' is thought to abstract hydrogen atom from the substrate, then the newly formed radical and chloride atom will combine together to give halogenated product (151, 157). For the “in line” mode, the Fe(IV)=02' points to the intended position of the primary substrate allowing the hydrogen abstraction and subsequent reactions to proceed directly. In contrast, the Fe(IV)=02' points away from the substrate for the “ off line” mode and migration of this ferryl-oxo species or shifting of the metal ligands is necessary. One possibility is that orKG reorients from “off line” (Figure 1.15B,) to “in line” (Figure 1.15B) binding mode, so that the subsequent steps will follow those of Figure 1.15 B to F. Another option is that orKG retains its off-line orientation resulting in the formation of Fe(IV)=02‘ pointing away from the substrate, as shown in Figure 1.15 B, to F’. Release of C02 will provide an open site for one H2O molecule to bind, resulting in the formation of dihydroxylated intermediate, Figure 1.15F", which can lose a molecule of water to complete the ferryl-oxo migration, Figure 1.15 F'——» F" -—>F. Subsequently, Fe(IV)=02' is positioned near the substrate and the following steps are equivalent to those in Figure 1.15 F to A. Currently, the crystal structures of O2 adducts have only been determined for two mononuclear non-heme iron enzymes, naphthalene dioxygenase and homoprotocatechuate 2,3-dioxygenase (I58, 159). Both of these proteins contain 2-His-1-carboxylate facial triad, but their proposed reaction mechanisms are totally different from that of the Fe(II)/01KG dependent dioxygenases and neither use OtKG as 35 co-substrate. For both enzymes, in the presence of primary substrate, 02 binds to the iron center through a side-on rather than end-on mode which is favored by theoretical calculations (160). Even though these enzymes provide valuable information about an 02 binding mode, the Fe(II)/OtKG dioxygenases are unlikely share this side-on 02 binding fashion due to their different protein structure and reaction mechanism. It is well known that reactive oxygen species can cause severe damage to the cell, and effective control of these reactive species is very important in viva (161-163). As mentioned earlier, oxygen generally is the last substrate to bind to the metal ion so that activation of oxygen will not occur until all the necessary cofactors or co-substrates are in the proper position. This strategy guarantees a high coupling efficiency, so the substrate covens into product concomitant with the oxidative decarboxylation of (xKG. This is not always the situation in vitro, where the primary . substrates may be absent or inhibitors or poor substrates may be present. Under these conditions uncoupled reactions may take place and lead to enzyme inactivation and self-modification, such as reported deA, AlkB, TauD, and HPPD (1, 11], 164-168). Aerobic purification of HPPD resulted in a blue chromophore consistent with an Fe(III) tyrosinate and the tyrosinate probably is formed from hydroxylation of one of the phenyl rings near the active site (110). When deA reacted with dioxygen and (xKG in the absence of primary substrate, a weak chromophore with 11m, around 580 nm was observed (164). RR and EPR were applied to identify this species as a Fe(III) hydroxyindolate product arising from the hydroxylation of Trp112 adjacent to the metal ion. A similar result was also observed using TauD, which first forms a 36 FIGURE 1.16: Proposed mechanism for oxidative cyclization catalyzed by CAS. 0 o 0 ,Ef COOH J:Nl COOH 1:“ Fe(1v1=o2- Fe(III)—OH- Fe(Il10H20 COOH HO ' HO _" NH, NH, NH: A B C FIGURE 1.17: Proposed mechanism for a desaturation reaction. R, R R R R1’ 1‘4 R1“? 3*4 R2)::: H ) H ( R1 =8 H6 Glu/Asp ‘0 (2:0 Glu/Asp}: 0.} "(0‘0 Jew/Asp Fe'," _. ‘0 His/ I His’Hi: ‘ ’ ‘ ’ Hi His 00 00' FIGURE 1.18: General reaction of halogenase. X Halogenase 1 R-CH3 7"? R-CH, 0: + “KG C0, + succinate 37 tyrosyl radical then catalyzes the self-hydroxylation of Tyr73 to form a Catechol which binds Fe(III) to develop a chromOphore with with 11m, near 550 nm, as identified by RR as a Fe(lII)-catecholate species (166). These enzymes in their ferric states are inactive, but some reductants, such as ascorbate and DTT, were reported to be able to restore part of the enzyme activity (132). Completing reaction cycles in uncoupled turnover reactions is commonly thought of as the reason for the ascorbate requirement of many Fe(II)/01KG-dependent hydroxylases. Of interest, representatives of these enzymes operate in organisms where ascorbate is not present (131), so they could not be restored to a functional state by this reductant. It’s still unclear if the self-hydroxylation observed in vitro is of physiological relevance. It’s been suggested that the enzyme builds in a number of sacrificial amino acids that are susceptible to oxygenation adjacent to the active site metal ion, so that in the absence of substrates or when substrate is not positioned properly, the enzyme internally quenches the reactive oxygen species rather than releasing it to solvent (13]). Given the facts that the uncoupling reactions rates are much slower than the catalytic reactions involving substrate (132, 166), and the relatively low cellular dioxygen concentration, self-inactivation is unlikely to be physiologically relevant in viva. THESIS OUTLINE The following chapters describe my studies on the purification and characterization of recombinant xanthine hydroxylase (XanA) from Aspergillus nidulans, including the investigation of the substrate-binding mode by site-directed 38 mutagensis and examination of the metallocenter coordination chemistry by various kinds of spectroscopies. In Chapter 2, I describe the purification of recombinant A. nidulans XanA as a His-tag version, and I provide evidence that XanA belongs to the family of Fe(II)/(XKG-dependent hydroxylases. I also examine the effects of pH, different kinds of buffers, and salt concentration on the enzymatic activity to optimize the assay conditions. In addition, I characterize the substrate and co-substrate concentration dependence to characterize the kinetic properties of XanA. Of interest, comparison of XanA expressed in different hosts revealed very different quaternary structures and posttranslational modifications. In Chapter 3, I describe more extensive kinetic characterizations, including effects of different divalent metals, co-substrate analogs, and substrate analogs to understand their binding modes to the active site. Finally, I use a homology model, created by a colleague using TauD as the template, to help understand the active site structure. The studies described in Chapter 1 and 2 were published in Biochemistry 2007, 46, 5293-5304. In Chapter 4, I describe the preparation and characterization of mutant proteins with putative active site residues mutated to alanine based on the homology model constructed from Chapter 3. Xanthine, OLKG alternative substrate analogs, inhibitors and chemical regents were applied to characterize the kinetic properties of the seven active mutants. Kinetic parameters from single mutants were compared with wild-type XanA to gain more insight into the substrate binding mode. The study of oxygen consumption when assayed without primary substrate provided more information on the potential of the relevance of these residues to the active site. Combining the kinetic characterization 39 and oxygen consumption studies, several important substrate binding interactions are proposed. In Chapter 5, I show that binding of orKG and xanthine to anaerobic Fe(II)-XanA generated MLCT transitions typical of this enzyme family. In addition, I show that nitric oxide (N0), 3 good oxygen surrogate, can be used to prepare various {FeNO}7 complexes, which were extensively characterized by various kinds of spectroscopies, including UV/Vis, EPR, and one- and two-dimensional ESEEM. I used these different {FeNO}7 intermediates of XanA to investigate the coordination chemistry at different reaction stages. For comparision, I used TauD, a well characterized member of this superfamily, as a reference to better interpret the results of my spectroscopic studies on XanA. Finally, I provide a concluding chapter that summarizes the remaining questions and places my investigations in broader perspective. 40 REFERENCES (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) Trewick, S. C., Henshaw, T. F., Hausinger, R. P., Lindahl, T., and Sedgwick, B. (2002) Oxidative demethylation by Escherichia coli AlkB directly reverts DNA base damage. Nature 419, 174-178. Kivirikko, K. I., and Pihlajaniemi, T. (1998) Collagen hydroxylases and the protein disulfide isomerase subunit of prolyl 4-hydroxylase. Adv. Enzymol. Rel. Areas Mol. Biol. 72, 325-398. Ivan, M., Kondo, K., Yang, H., Kim, W., Valiando, J., Ohh, M., Salic, A., Asara, J. M., Lane, W. S., and Kaelin, W. G, Jr. (2001) HIFor targeted for VHL-mediated destruction by proline hydroxylation: implications for 02 sensing. Science 292, 464-467. Jaakkola, P., Mole, D. R., Tian, Y.-M., Wilson, M. 1., Gielbert, J., Gaskell, S. J., von Kn'egsheim, A., Hebestreit, H. F., Mukherji, M., Schofield, C. J., Maxwell, P. H., Pugh, C. W., and Ratcliffe, P. J. (2001) Targeting of HIF-or to the von Hippel-Lindau ubiquitylation complex by 02-regulated prolyl hydroxylation. Science 292, 468-472. Dann, C. E., III, Bruick, R. K., and Deisenhofer, J. (2002) Structure of a factor-inhibiting hypoxia-inducible factor 1: an essential asparaginyl hydroxylase involved in the hypoxic response pathway. Proc. Natl. Acad. Sci. USA 99, 15351-15356. Zhang, Z., Ren, J., Stammers, D. K., Baldwin, J. E., Harlos, K., and Schofield, C. J. (2000) Structural origins of the selectivity of the trifunctional oxygenase clavaminic acid synthase. Nature Struct. Biol. 7, 127-133. Clifton, I. J., Hsueh, L.-C., Baldwin, J. E., Harlos, K., and Schofield, C. J. (2001) Structure of proline 3-hydroxylase. Evolution of the family of 2-oxoglutarate dependent dioxygenases. Eur: J. Biochem. 268, 6625-6636. Lukacin, R., Groning, 1., Pieper, U., and Matem, U. (2000) Site-directed mutagenesis of the active site serine290 in flavanone 3B-hydroxylase from Petunia hybrida. Eur: J. Biochem. 267, 853-860. Hedden, P., and Phillips, A. L. (2000) Gibberellin metabolism: new insights revealed by the genes. Trends Plant Sci. 5, 523-528. De Carolis, E., Chan, E, Balsevich, J., and De Luca, V. (1990) Isolation and 41 (11) (12) (13) (14) (15) (16) (17) (18) (19) characterization of a 2-oxoglutarate dependent dioxygenase involved in the second-to-last step in vindoline biosynthesis. Plant Physiol. 94, 1323-1329. Hashimoto, T., and Yamada, Y. (1987) Purification and characterization of hyoscyamine 6B-hydroxylase from root cultures of Hyoscyamus niger L. Hydroxylase and epoxidase activities of enzyme preparations. Eur. J. Biochem. 164, 277-285. ' Jansen, G A., Mihalik, S. J., Watkins, P. A., Jakobs, C., Moser, H. W., and Wanders, R. J. A. (1998) Characterization of phytanoyl-coenzyme A hydroxylase in human liver and activity measurements in patients with peroxisomal disorders. Clin. Chim. Acta 271, 203-211. Fukumori, F., and Hausinger, R. P. (1993) Purification and characterization of 2,4-dichlorophenoxyacetate/or-ketoglutarate dioxygenase. J. Biol. Chem. 268, 24311-24317. Hutton, J. J., Jr., Trappel, A. L., and Udenfriend, S. (1966) Requirement for or-ketoglutarate, ferrous ion and ascorbate for collagen proline hydroxylase. Biochem. Biophys. Res. Commun. 24, 179-184. Hieta, R., and Myllyharju, J. (2002) Cloning and characterization of a low molecular weight prolyl 4-hydroxylase from Arabidopsis thaliana. Effective hydroxylation of proline-rich, collagen-like, and hypoxia-inducible transcription factor a—like peptides. J. Biol. Chem. 277, 23965-23971. Kaska, D. D., Myllyléi, R., Giinzler, V., Gibor, A., and Kivirikko, K. I. (1988) Prolyl 4-hydroxy1ase from Volvox carteri. A low-Mr enzyme antigenically related to the or subunit of the invertebrate enzyme. Biochem. J. 256, 257-263. Helaakoski, T., Vuori, K., R., M., Kivirikko, K. I., and Pihlajaniemi, T. (1989) Molecular cloning of the or-subunit of human prolyl 4-hydroxylase: the complete cDNA-derived amino acid sequence and evidence for alternative splicing of RNA transcripts. Proc. Natl. Acad. Sci. USA 86, 4392-4396. Annunem, P., Koivunen, P., and Kivirikko, K. I. (1999) Cloning of the Ot-subunit of prolyl 4-hydroxylase from Drosophila and expression and characterization of the corresponding enzyme tetramer with some unique properties. J. Biol. Chem. 274, 6790-6796. Veijola, J., Koivunen, P., Annunem, P., Pihlajaniemi, T., and Kivirikko, K. I. (1994) Cloning, Baculovirus expression, and characterization of the or subunit of prolyl 4-hydroxylase from the nematode Caenorhaditis elegans. This or subunit forms an active OLB dimer with the human protein disulfide 42 (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) (30) isomerase/B subunit. J. Biol. Chem. 269, 26746-26753. Ericksson, M., Myllyharju, J., Tu, H., Hellman, M., and Kivirikko, K. I. (1999) . Evidence for 4-hydroxyproline in viral prOteins. Characterization of a viral prolyl 4-hydroxylase and its peptide substrates. J. Biol. Chem. 274, 22131-22134. ‘ Lamberg, A., Pihlajaniemi, T., and Kivirikko, K. I. (1995) Site-directed mutagenesis of the or subunit of human prolyl 4-hydroxylase. Identification of three histidine residues critical for catalytic activity. J. Biol. Chem. 270, 9926-9931. Myllyharju, J., and Kivirikko, K. I. (1997) Characterization of the iron- and 2-oxoglutarate-binding sites of human prolyl 4-hydroxylase. The EMBO J. 16, 1173-1180. Pekkala, M., Hieta, R., Kursula, P., Kivirikko, K. I., Wierenga, R. K., and Myllyharju, J. (2003) Crystallization of the proline-rich-peptide binding domain of human type I collagen prolyl 4-hydroxylase. Acta Crystallogr: D59, 940-942. Myllyharju, J. (2003) Prolyl 4-hydroxylases, the key enzymes of collagen biosynthesis. Matrix Biol. 22, 15-24. Pirskanen, A., Kaimio, A. M., Myllyla, R., and Kivirikko, K. I. (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, 9398-402. Walsh, G, and Jefferis, R. (2006) Post-translational modifications in the context of therapeutic proteins. Nat. Biotechnol. 24, 1241-52. Semenza, G L. (2000) HIF-1 and human disease: one highly involved factor. Genes Dev. 14, 1983-91. Schofield, C. J., and Ratcliffe, P. J. (2004) Oxygen sensing by HIF hydroxylases. Nat. Rev. Mol. Cell Biol. 5, 343-54. Hewitson,iK. S., McNeil], L. A., Elkins, J. M., and Schofield, C. J. (2003) The role of iron and 2-oxoglutarate oxygenases in signalling. Biochem. Soc. Trans. 31,510-515. Bruick, R. K., and McKnight, S. L. (2001) A conserved family of prolyl-4-hydroxylases that modify HIF. Science 294, 1337-1340. 43 (31) (32) (33) (34) (35) (36) (37) (38) (39) (40) Hewitson, K. S., McNeil], L. A., Riordan, M. V., Tian, Y.-M., Bullock, A. N., Welford, R. W., Elkins, J. M., Oldham, N. J., Bhattacharya, S., Gleadle, J. M., Ratcliffe, P. J., Pugh, C. W., and Schofield, C. J. (2002) Hypoxia-inducible factor (HIF) asparagine hydroxylase is identical to factor inhibiting HIF (FIH) and is related to the cupin structural familly. J. Biol. Chem. 2 7 7, 26351-26355. Lando, D., Peet, D. J., Whelan, D. A., Gonnan, J. J., and Whitelaw, M. L. (2002) Asparagine hydroxylation of the HIF transactivation domain: a hypoxic switch. Science 295, 858-861. McNeill, L. A., Hewitson, K. S., Claridge, T. D., Seibel, J. F., Horsfall, L. E., and Schofield, C. J. (2002) Hypoxia-inducible factor asparginyl hydroxylase (FIH-1) catalyses hydroxylation at the B-carbon of asparagine-803. Biochem. J. 367, 571-575. Elkins, J. M., Hewitson, K. S., McNeill, L. A., Seibel, J. F., Schlemminger, 1., Pugh, C. W., Ratcliffe, P. J., and Schofield, C. J. (2003) Structure of factor-inhibiting hypoxia-inducible factor (HIF) reveals mechanism of oxidative modification of HIF-101. J. Biol. Chem. 278, 1802-1806. Martin, C., and Zhang, Y. (2005) The diverse functions of histone lysine methylation. Nat. Rev. Mol. Cell Biol. 6, 838-49. Strahl, B. D., and Allis, C. D. (2000) The language of covalent histone modifications. Nature 403, 41-5. Tsukada, Y.-I., Fang, J., Erdjument-Bromage, H., Warren, M. E., Borchers, C. H., Tempst, P., and Zhang, Y. (2006) Histone demethylation by a family of ijC domain-containing proteins. Nature 439, 811-816. Wang, Y., Wysocka, J., Sayegh, J., Lee, Y.-H., Perlin, J. R., Leonelli, L., Sonbuchner, L. S., McDonald, C. H., Cook, R. G, Dou, Y., Roeder, R. G, Clarke, 8., Stallcup, M. R., Allis, C. D., and Coonrod, S. A. (2004) Human PAD4 regulates histone arginine methylation levels via demethylimination. Science 306, 279-283. Cuthbert, G L., Daujat, S., Snowden, A. W., Erdjument-Bromage, H., Hagiwara, T., Yamada, M., Schneider, R., Gregory, P. D., Tempst, P., Bannister, A. J., and Kouzarides, T. (2004) Histone deimination antagonizes arginine methylation. Cell 118, 545-53. Shi, Y., Lan, F., Matson, C., Mulligan, P., Whetstine, J. R., Cole, P. A., Casero, R. A., and Shi, Y. (2004) Histone demethylation mediated by the nuclear amine oxidase homolog LSDl. Cell 119, 941-53. 44 (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) Metzger, E., Wissmann, M., Yin, N., Muller, J. M., Schneider, R., Peters, A. H., Gunther, T., Buettner, R., and Schule, R. (2005) LSDl demethylates repressive histone marks to promote androgen-receptor-dependent transcription. Nature 43 7, 436-9. Yamane, K., Toumazou, C., Tsukada, Y.-I., Erdjument-Bromage, H., Tempst, P., Wong, J., and Zhang, Y. (2006) JHDM2A, a ijC-containing H3K9 demethylase, facilitates transcription activation by androgen receptor. Cell 125, 1-13. Cloos, P. A. C., Christensen, J., Agger, K., Maiolica, A., Rappsilber, J., Antal, T., Hansen, K. H., and Helin, K. (2006) The putative oncogene GASC] demethylates tri- and dimethylated lysine 9 on histone H3. Nature 442, 307-311. Whetstine, J. R., Nottke, A., Lan, F., Huart, M., Smolikov, S., Chen, Z., Spooner, E., Li, B, Zhang, G, Colaiacovo, M., and Shi, Y. (2006) Reversal of histone lysine trimethylation by the JMJD2 family of histone demethylases. Cell 125, 467-481. Klose, R. J ., Yamane, K., Bae, Y., Zhang, D., Erdjument-Bromage, H., Tempst, P., Wong, J., and Zhang, Y. (2006) The transcriptional repressor JHDM3A demethylates trimethyl histone H3 lysine 9 and lysine 36. Nature 442, 312-316. Fodor, B. D., Kubicek, S., Yonezawa, M., O'Sullivan, R. J., Sengupta, R., Perez-Burgos, L., Opravil, S., Mechtler, K., Schotta, G, and Jenuwein, T. (2006) ijd2b antagonizes H3K9 trimethylation at pericentric heterochromatin in mammalian cells. Genes Dev 20, 1557-62. Klose, R. J., Kallin, E. M., and Zhang, Y. (2006) ijC-domain-containing proteins and histone demethylation. Nat. Rev. Genet. 7, 715-27. Chen, Z., Zang, J., Whetstine, J., Hong, X., Davrazou, F., Kutateladze, T. G, Simpson, M., Mao, 0., Pan, C. H., Dai, S., Hagman, J., Hansen, K., Shi, Y., and Zhang, G (2006) Structural insights into histone demethylation by JMJD2 family members. Cell 125, 691-702. Sedgwick, B., and Lindahl, T. (2002) Recent progress on the Ada response for inducible repair of DNA alkylation damage. Oncogene 21, 8886-8894. Sedgwick, B. (2004) Repairing DNA-methylation damage. Nat. Rev. Molec. Cell Biol. 5, 148-157. 45 (51) (52) (53) (54) (55) (56) (57) (58) (59) (60) (61) Rydberg, B., and Lindahl, T. (1982) Nonenzymatic methylation of DNA by the intracellular methyl group donor S-adenosyl-L-methionine is a potentially mutagenic reaction. EMBO J. 1, 211-6. Teo, 1., Sedgwick, B., Kilpatrick, M. W., McCarthy, T. V., and Lindahl, T. (1986) The intracellular signal for induction of resistance to alkylating agents in E. coli. Cell 45, 315-24. Kataoka, H., Yamamoto, Y., and Sekiguchi, M. (1983) A new gene (alkB) of Escherichia coli that controls sensitivity to methyl methane sulfonate. J. Bacterial. 153, 1301-1307. Chen, B. J., Carroll, P., and Samson, L. (1994) The Escherichia coli AlkB protein protects human cells against alkylation-induced toxicity. J. Bacteriol. 176, 6255-6261. Aravind, L., and Koonin, E. V. (2001) The DNA-repair protein AlkB, EGL-9, and leprecan define new families of 2-oxoglutarate-dependent- and iron-dependent dioxygenases. Genome Biology 2, 7.1-7.8. Falnes, P. 0., Johansen, R. F., and Seeberg, E. (2002) AlkB-mediated oxidative demethylation reverses DNA damage in Escherichia coli. Nature 419, 178-182. Duncan, T., Trewick, S. C., Koivisto, P., Bates, P. A., Lindahl, T., and Sedgwick, B. (2002) Reversal of DNA alkylation damage by two human dioxygenases. Proc. Natl. Acad. Sci. USA 99, 16660-16665. Koivisto, P., Duncan, T., Lindahl, T., and Sedgwick, B. (2003) Minimal methylated substrate and extended substrate range of Escherichia coli AlkB protein, a l-methyladenine-DNA dioxygenase. J. Biol. Chem. 278, 44348-44354. Wei, Y.-F., Carter, K. .C., Wang, R.-P., and Shell, B. K. (1996) Molecular cloning and fitnctional analysis of a human cDNA encoding an Escherichia coli AlkB homolog, a protein involved in DNA alkylation damage repair. Nucl. Acids Res. 24, 931-937. Kurowski, M. A., Bhagwat, A. S., Papaj, G, and Bujnicki, J. M. (2003) Phylogenomic identification of five new human homologues of the DNA repair enzyme AlkB. BMC Genomics 4, 48. Colombi, D., and Gomes, S. L. (1997) An alkB gene homolog is differentially transcribed during the Caulobacter crescentus cell cycle. J. Bacteriol. 179, 46 (62) (63) (64) (65) (66) (67) (68) (69) (70) (71) (72) 3139-45. Aas, P. A., Otterlei, M., Falnes, P. 0., Vagbe, C. B., Skorpen, F., Akbari, M., Sundheim, 0., Bjoras, M., Slupphaug, G, Seeberg, E., and Krokan, H. E. (2003) Human and bacterial oxidative demethylases repair alkylation damage in both RNA and DNA. Nature 421, 859-863. Yu, B., Edstrom, W. C., Hamuro, Y., Weber, P. C., Gibney, B. R., and Hunt, J. F. (2006) Crystal structures of catalytic complexes of the oxidative DNA/RNA repair enzyme AlkB. Nature 439, 879-884. Matagne, A., Dubus, A., Galleni, M., and Frere, J. M. (1999) The beta-lactamase cycle: a tale of selective pressure and bacterial ingenuity. Nat Prod Rep 16, 1-19. Lloyd, M. D., Mem't, K. D., Lee, V., Sewell, T. J., Wha-Son, B., Baldwin, J. E., Schofield, C. J., Elson, S. W., Baggaley, K. H., and Nicholson, N. H. (1999) Product-substrate engineering by bacteria: studies on Clavaminate synthase, a trifunctional dioxygenase. Tetrahedron 55, 10201-10220. Busby, R. W., Chang, M. D.-T., Busby, R. C., Wimp, J., and Townsend, C. A. (1995) Expression and purification of two isozymes of Clavaminate synthase and initial characterization of the iron binding site. General error analysis in polymerase chain reaction amplification. J. Biol. Chem. 270, 4262-4269. Marsh, E. N., Chang, M. D.-T., and Townsend, C. A. (1992) Two isozymes of Clavaminate synthase central to Clavulanic acid formation: cloning and sequencing of both genes from Streptomyces clavuligerus. Biochemistry 31, 12648-12657. Bohm, B. A. (1998) Introduction to Flavonoids, Harwood Academic Publishers, Amsterdam, The Netherlands. Prescott, A. G (1993) A dilemma of dioxygenases (or where biochemistry and molecular biology fail to meet). J. Expert Bat. 44, 849-861. Heller, W., and Forkmann, G. (1993) The flavonaids, advances in research since 1986, 399-425. Britsch, L., and Grisebach, H. (1986) Purification and characterization of (ZS)-flavanone 3-hydroxylase from Petunia hybrida. Eur: J. Biochem. 156, 569-577. Lukacin, R., Groning, 1., Schiltz, E., Britsch, L., and Matem, U. (2000) 47 (73) (74) (75) (76) (77) (78) (79) (80) (81) Purification of recombinant flavanone 3B-hydroxylase from Petunia hybrida and assignment of the primary site of proteolytic degradation. Arch. Biochem. Biophys. 375, 364-370. Lukacin, R., and Britsch, L. (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. Uria-Nickelsen, M. R., Leadbetter, E. R., and Godchaux, W., 111. (1993) Sulfonate utilization by enteric bacteria. J. Gen. Microbial. 139, 203-208. Van Der Ploeg, J. R., Weiss, M. A., Saller, E., Nashimoto, H., Saito, N., Kertesz, M. A., and Leisinger, T. (1996) Identification of sulfate starvation-regulated genes in Escherichia coli: 3 gene cluster involved in the utilization of taurine as a sulfur source. J. Bacterial. I 78, 5438-5446. Fukumori, E, and Hausinger, R. P. (1993) Alcaligenes eutraphus JMP134 "2,4-dichlorophenoxyacetate monooxygenase" is an or-ketoglutarate-dependent dioxygenase. J. Bacteriol. 175, 2083-2086. Eichhom, E., van der Ploeg, J. R., Kertesz, M. A., and Leisinger, T. (1997) Characterization of a-ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J. Biol. Chem. 272, 23031—23036. Elkins, J. M., Ryle, M. J., Clifton, I. J., Dunning Hotopp, J. 0, Lloyd, J. S., Burzlaff, N. 1., Baldwin, J. E., Hausinger, R. P., and Roach, P. L. (2002) X-ray‘ crystal structure of Escherichia coli taurine/a-ketoglutarate dioxygenase complexed to ferrous iron and substrates. Biochemistry 41, 5185-5192. O'Brien, J. R., Schuller, D. J., Yang, V. S., Dillard, B. D., and Lanzilotta, W. N. (2003) Substrate-induced conformational changes in Escherichia coli taurine/a-ketoglutarate dioxygenase and insight into the oligomeric structure. Biochemistry 42, 5547-5554. Hausinger, R. P., Fukumori, F., Hogan, D. A., Sassanella, T. M., Kamagata, Y., Takami, H., and Saari, R. E. (1997) Biochemistry of 2,4-D degradation: evolutionary implications, in Microbial Diversity and Genetics of Biodegradatian (Horikoshi, K., Fukuda, M., and Kudo, T., Eds.) pp 35—51, Japan Scientific Press, Tokyo. Dunning Hotopp, J. C., and Hausinger, R. P. (2001) Alternative substrates for 2,4-dichlorophenoxyacetate/or-ketoglutarate dioxygenase. J. Molec. Catalysis B 15,155-162. 48 (82) (83) (84) (85) (86) (87) (88) (89) (90) (91) Tett, V. A., Willetts, A. J., and Lappin-Scott, H. M. (1997) Biodegradation of . the chlorophenoxy herbicide (R)-(+)-mecoprop by Alcaligenes denitrificans. Biodegradation 8, 43-52. Nickel, K., Suter, M. J.-F., and Kohler, H.-P. E. (1997) Involvement of two or-ketoglutarate-dependent dioxygenases in enantioselective degradation of (R)- and (S)-mecoprop by ‘Sphingomanas herbicidavarans MH. J. Bacteriol. 179, 6674-6679. Westendorf, A., Miiller, A., and Babel, W. (2003) Purification and characterization of the enantiospecific dioxygenases from Delftia acidovorans MCl initiating the degradation of phenoxypropionate and phenoxyacetate herbicides. Acta Biotechnol 23, 3-17. Miiller, T. A., Zavodszky, M. I., Feig, M., Kuhn, L. A., and Hausinger, R. P. (2006) Structural basis for the enantiospecificities of R- and S-phenoxypropionate/or-ketoglutarate dioxygenases. Prat. Science 15, 1356-1368. Cosper, N. J., Stalhandske, C. M. V., Saari, R. E., Hausinger, R. P., and Scott, R. A. (1999) X-ray absorption spectroscopic analysis of Fe(II) and Cu(II) forms of an herbicide-degrading Ot-ketoglutarate dioxygenase. J. Biol. Inorg. Chem. 4, 122-129. Whiting, A. K., Que, L., Jr., Saari, R. E., Hausinger, R. P., Fredrick, M. A., and McCracken, J. (1997) Metal coordination environment of a Cu(II)-substituted ' or-keto acid-dependent dioxygenase that degrades the herbicide 12,4-D. J. Am. Chem. Soc. 119, 3413-3414. Dunning Hotopp, J. C., and Hausinger, R. P. (2002) Probing the 2,4-dichlorophenoxyacetate/or-ketoglutarate dioxygenase substrate binding site by site-directed mutagenesis and mechanism-based inactivation. Biochemistry 41, 9787-9794. Hogan, D. A., Smith, S. R., Saari, E. A., McCracken, J., and Hausinger, R. P. (2000) Site-directed mutagenesis of 2,4-dichlorophenoxyacetic acid/Ot-ketoglutarate dioxygenase. Identification of residues involved in metallocenter formation and substrate binding. .1. Biol. Chem. 275, 12400-12409. Hille, R. (2005) Molybdenum-containing hydroxylases. Arch. Biochem. Biophys. 433, 107-116. Cultrone, A., Scazzocchio, C., Rochet, M., Montero-Moran, G, Drevet, C., 49 (92) (93) (94) (95) (96) (97) (98) (99) (100) and Femandez-Martin, R. (2005) Convergent evolutiOn of hydroxylation mechanisms in the fungal kingdoms: molybdenum cofactor-independent hydroxylation of xanthine via or-ketoglutarate dependent dioxygenase. Molec. Microbiol. 57, 276-290. Kreisberg-Zakarin, R., Borovok, 1., Yanko, M., Aharonowitz, Y., and Cohen, G (1999) Recent advances in the structure and function of isopenicillin N synthase. Antonie van Leeuwenhoek 75, 33-39. Schofield, C. J., Baldwin, J. E., Byford, M. F., Clifton, I. J., Hensgens, C., and Roach, P. L. (1997) Proteins of the penicillin biosynthesis pathway. Curr: Opinion Struct. Biol. 7, 857-864. Pang, C. P., Chakravarti, B., Adlington, R. M., Ting, H. H., White, R. L., Jayatilake, G S., Baldwin, J. E., and Abraham, E. P.’(l984) Purification of isopenicillin N synthetase. Biochem J 222, 789-95. Baldwin, J. E., and Abraham, E. (1988) The biosynthesis of penicillins and cephalosporins. Nat. Prod. Rep. 5, 129-45. Castro, J. M., Liras, P., Laiz, L., Cortes, J., and Martin, J. F. (1988) Purification and characterization of the isopenicillin N synthase of Streptomyces lactamdurans. J. Gen. Microbiol. 134, 133-41. Samson, S. M., Belagaje, R., Blankenship, D. T., Chapman, J. L., Perry, D., Skatrud, P. L., VanFrank, R. M., Abraham, E. P., Baldwin, J. E., Queener, S. W., and et al. (1985) Isolation, sequence determination and expression in Escherichia coli of the isopenicillin N synthetase gene from Cephalasporium acremanium. Nature 318, 191-4. Chen, V. J., Orville, A. M., Harpel, M. R., Frolik, C. A., Surerus, K. K., Munck, E., and Lipscomb, J. D. (1989) Spectroscopic studies of isopenicillin N synthase. A mononuclear nonheme Fez” oxidase with metal coordination sites for small molecules and substrate. .1. Biol. Chem. 264, 21677-21681. Jiang, F., Peisach, J., Ming, L.-J., Que, L., Jr., and Chen, V. J. (1991) Electron spin echo envelope modulation studies of the Cu(I)-substituted derivative of isopenicillin N synthase: a structural and spectroscopic model. Biochemistry 30, 11437-11445. Randall, C. R., Zang, Y., True, A. E., Que, L., Jr., Chamock, J. M., Garner, C. D.; Fujishima, Y., Schofield, C. J ., and Baldwin, J. E. (1993) X-ray absorption studies of the ferrous active site of isopenicillin N synthase and related model complexes. Biochemistry 32, 6664-6673. ‘ 50 (101) (102) (103) (104) (105) (106) (107) (108) (109) (110) (111) Scott, R. A., Wang, S., Eidsness, M. K., Kriauciunas, A., Frolik, C. A., and Chen, V. J. (1992) X-ray absorption spectroscopic studies of the high-spin iron(II) active site of isopenicillin N synthase: evidence for Fe-S interaction in the enzyme-substrate complex. Biochemistry 31, 4596-4601. Roach, P. L., Clifton, I. J., Ffilbp, V., Harlos, K., Barton, G J., Hajdu, J., Andersson, K., Schofield, C. J., and Baldwin, J. E. (1995) Crystal structure of isopenicillin N synthase is the first from a new structural family of enzymes. Nature (London) 375, 700-704. Landman, 0., Borovok, 1., Aharonowitz, Y., and Cohen, G (1997) The glutamine ligand in the ferrous iron active site of isopenicillin N synthase of Streptomyces jumonjinensis is not essential for catalysis. FEBS Lett. 405, 172-174. Roach, P. L., Clifton, I. J., Hensgens, C. M. H., Shibata, N., Schofield, C. J., Hajdu, J., and Baldwin, J. E. (1997) Structure of isopenicillin N synthase complexed with substrate and the mechanism of penicillin formation. Nature (London) 387, 827-830. Moran, G R. (2005) 4-Hydroxyphenylpyruvate dioxygenase. Arch. Biochem. Biophys. 433, 117-128. Gissen, P., Preece, M. A., Willshaw, H. A., and McKiernan, P. J. (2003) Ophthalmic follow-up of patients with tyrosinaemia type I on NTBC. J. Inherit. Metab. Dis. 26, 13-6. Wada, G H., Fellman, J. H., Fujita, T. S., and Roth, E. S. (1975) Purification and properties of avian liver p-hydroxyphenylpyruvate hydroxylase. J. Biol. Chem. 250, 6720-6. Lindblad, B., Lindstedt, G, Lindstedt, S., and Rundgren, M. (1977) Purification and some properties of human 4-hydroxyphenylpyruvate dioxygenase (I). J. Biol. Chem. 252, 5073-84. Buckthal, D. J ., Roche, P. A., Moorehead, T. J ., Forbes, B. J., and Hamilton, G A. (1987) 4-Hydroxyphenylpyruvate dioxygenase from pig liver. Methalds Enzymol. 142, 132-8. Lindstedt, S., and Rundgren, M. (1982) Blue color, metal content, and substrate binding in 4-hydroxyphenylpyruvate dioxygenase from Pseudomonas sp. strain P. J. 874. J. Biol. Chem. 257, 11922-31. Bradley, F. C., Lindstedt, S., Lipscomb, J. D., Que, L., Jr., Roe, A. L., and 51 (112) (113) (114) (115) (116) (117) (118) (119) (120) Rundgren, M. (1986) 4-Hydroxyphenylpyruvate dioxygenase is an iron-tyrosinate protein. J. Biol. Chem. 261, 11693-11696. Serre, L., Sailland, A., Sy, D., Boudec, P., Rolland, A., Pebay-Peyroula, E., and Cohen-Addad, C. (1999) Crystal structure of Pseudomonas fluorescens 4-hydroxyphenylpyruvate dioxygenase: an enzyme involved in the tyrosine degradation pathway. Structure 7, 977-988. Ryle, M. J., and Hausinger, R. P. (2002) Non-heme iron oxygenases. Curr. Opinion Chem. Biol. 6, 193-201. Johnson-Winters, K., Purpero, V. M., Kavana, M., Nelson, T., and Moran, G R. (2003) (4-Hydroxyphenyl)pyruvate dioxygenase from Streptomyces avermitilis: the basis for ordered substrate addition. Biochemistry 42, 2072-2080. Gunsior, M., Ravel, J ., Challis, G L., and Townsend, C. A. (2004) Engineering p-hydroxyphenylpyruvate dioxygenase to a p-hydroxymandelate synthase and evidence for the proposed benzene oxide intermediate in homogentisate formation. Biochemistry 43, 663-674. Brownlee, J. M., Johnson-Winters, K., Harrison, D. H. T., and Moran, G R. (2004) Structure of the ferrous form of (4-hydroxyphenyl)pyruvate dioxygenase from Streptomyces avermitilis in complex with the therapeutic herbicide, NTBC. Biochemistry 43, 6370-6377. Stirk, H. J., Woolfson, D. N., Hutchinson, E. G, and Thornton, J. M. (1992) Depicting topology and handedness in jellyroll structures. FEBS Lett 308, 1-3. Valegard, K., Terwisscha van Scheltinga, A. 0, Lloyd, M. D., Hara, T., Ramaswamy, S., Perrakis, A., Thompson, A., Lee, W.-J., Baldwin, J. E., Schofield, C. J., Hajdu, J., and Andersson, I. (1998) Structure of a cephalosporin synthase. Nature (London) 394, 805-809. Lloyd, M. D., Lee, H.-J., Harlos, K., Zhang, Z.-H., Baldwin, J. E., Schofield, C. J., Chamock, J. M., Gamer, C. D., Hara, T., Terwisscha Van Scheltinga, A. C., Valegard, K., Viklund, J. A. C., Hajdu, J., Andersson, I., Danielsson, A., and Bhikhabhai, R. (1999) Studies on the active site of deacetoxycephalosporin C synthase. J. Mol. Biol. 287, 943-960. Miiller, I., Kahnert, A., Pape, T., Sheldrick, G M., Meyer-Klaucke, W., Dierks, T., Kertesz, M., and Uson, I. (2004) Crystal structure of the alkylsulfatase AtsK: insights into the catalytic mechanism of the Fe(II) tx-ketoglutarate-dependent dioxygenase superfamily. Biochemistry 43, 52 (121) (122) (123) (124) (125) (126) (127) (128) (129) (130) (131) 3075-3088. Clifton, I. J., Doan, L. X., Sleeman, M. C., Topf, M., Suzuki, H., Wilmouth, R. C., and Schofield, C. J. (2003) Crystal structure of carbapenem synthase (CarC). J. Biol. Chem. 278, 20843-20850. Clifton, I. J., McDonough, ‘M. A., Ehrismann, D., Kershaw, N. J., Granatino, N., and Schofield, C. J. (2006) Structural studies on 2-oxoglutarate oxygenases and related double-stranded B-helix fold protein. J. Inorg. Biochem. 100, 644-669. Broderick, J. B. (1999) Catechol dioxygenases. Essays Biochem. 34, 173-189. Costas, M., Mehn, M. P., Jensen, M. P., and Que, L., Jr. (2004) Dioxygen activation at mononuclear nonheme iron active sites: enzymes, models, and intermediates. Chem. Rev. 104, 939-986. Lancaster, D. E., McDonough, M. A., and Schofield, C. J. (2004) Factor inhibiting hypoxia-inducible factor (FIH) and other asparaginyl hydroxylases. Biochem. Soc. Trans. 32, 943-5. Hausinger, R. P. (2004) Fe(II)/or-ketoglutarate-dependent hydroxylases and related enzymes. Crit. Rev. Biochem. Mal. Biol. 39, 21-68. Valegard, K., Terwisscha van Scheltinga, A. C., Dubus, A., Ranghino, G, Oster, L. M., Hajdu, J., and Andersson, I. (2004) The structural basis of cephalosporin formation in a mononuclear ferrous enzyme. Nat. Struct. Mol. Biol. 11, 95-101. Lee, H. J., Lloyde, M. D., Clifton, I. J., Baldwin, J. E., and Schofield, C. J. (2001) Kinetic and crystallographic studies on deacetoxycephalosporin C synthase (DAOCS). J. Mol. Biol. 308, 937-948. Zhang, Z., Ren, J.-s., Harlos, K., McKinnon, C. H., Clifton, I. J., and Schofield, C. J. (2002) Crystal structure of a Clavaminate synthase-Fe(II)-2-oxog1utarate-substrate-NO complex: evidence for metal centred rearrangements. FEBS Lett. 517, 7-12. Price, J. C., Barr, E. W., Hoffart, L. M., Krebs, C., and Bollinger, J. M., Jr. (2005) Kinetic dissection of the catalytic mechanism of taurinezor-ketoglutarate dioxygenase (TauD) from Escherichia coli. Biochemistry 44, 8138-8147. Hewitson, K. S., Granatino, N., Welford, R. W., McDonough, M. A., and 53 (132) (133) (134) (135) (I36) (137) (138) (139) (140) Schofield, C. J. (2005) Oxidation by 2-oxoglutarate oxygenases: non-haem iron systems in catalysis and signalling. Philos. Transact. A Math. Phys. Eng. Sci. 363, 807-28; discussion 1035-40. Ryle, M. J., Koehntop, K. D., Liu, A., Que, L., Jr., and Hausinger, R. P. (2003) Interconversion of two oxidized forms of TauD, a non-heme iron hydroxylase: evidence for bicarbonate binding. Proc. Natl. Acad. Sci. USA 100, 3790-3795. Neidig, M. L., Decker, A., Choroba, O. W., Huang, F., Kavana, M., Moran, G. R., Spencer, J. B., and Solomon, E. I. (2006) Spectroscopic and electronic structure studies of aromatic electrophilic attack and hydrogen-atom abstraction by non-heme iron enzymes. Proc Natl Acad Sci U S A 103, 12966-73. Holme, E. (1975) A kinetic study of thymine 7-hydroxylase from Neurospara crassa. Biochemistry 14, 4999-5003. Tuderman, L., R., M., and Kivirikko, K. I. (1977) Mechanism of prolyl hydroxylase reaction. 1. Role of co-substrates. Eur: J. Biochem. 80, 314-348. Johnson-Winters, K., Purpero, V. M., Kavana, M., and Moran, G R. (2005) Accumulation of multiple intermediates in the catalytic cycle of (4-hydroxyphenyl)pyruvate dioxygenase from Streptomyces avermitilis. Biochemistry 44, 7189-7199. Lee, H.-J., Lloyd, M. D., Clifton, I. J., Harlos, K., Dubus, A., Baldwin, J., E., Frere, J.-M., and Schofield, C. J. (2001) Alteration of the co—substrate selectivity of deacetoxycephalosporin C synthase. The role of arginine 258. J. Biol. Chem. 276, 18290-18295. Pavel, E. G, Zhou, J., Busby, R. W., Gunsior, M., Townsend, C. A., and Solomon, E. I. (1998) Circular dichroism and magnetic circular dichroism spectroscopic studies of the non-heme ferrous active site in Clavaminate synthase and its interaction with or-ketoglutarate cosubstrate. J. Am. Chem. Soc. 120, 743-753. Neidig, M. L., Kavana, M., Moran, G R., and Solomon, E. I. (2004) CD and MCD studies of the non-heme ferrous active site in (4-hydroxyphenyl)pyruvate dioxygenase: correlation between oxygen activation in the extradiol and alpha-KG-dependent dioxygenases. J. Am. Chem. Soc. 126, 4486-7. Ryle, M. J., Padmakumar, R., and Hausinger, R. P. (1999) Stopped-flow kinetic analysis of Escherichia coli taurine/a—ketoglutarate dioxygenase: (141) (142) (143) (144) (145) (146) (147) interactions with a-ketoglutarate, taurine, and oxygen. Biochemistry 38, 15278-15286. Hegg, E. L., Whiting, A. K., Saari, R. E., McCracken, J ., Hausinger, R. P., and Que, L., Jr. (1999) Herbicide-degrading or-keto acid-dependent enzyme deA: metal coordination environment and mechanistic insights. Biochemistry 38, 16714-16726. Purpero, V., and Moran, G R. (2007) The diverse and pervasive chemistries of the Ot-keto acid dependent enzymes. J. Biol. Inorg. Chem. 12, 587-601. Price, J. C., Barr, E. W., Tirupati, B., Bollinger, J. M., Jr., and Krebs, C. (2003) The first direct characterization of a high-valent iron intermediate in the reaction of an or-ketoglutarate-dependent dioxygenase: a high-spin Fe(IV) complex in taurine/(x-ketoglutarate dioxygenase (TauD) from Escherichia coli. J. Biol. Chem. 42, 7497-7508. Borowski, T., Bassan, A., and Siegbahn, P. E. M. (2004) Mechanism of dioxygen activation in 2-oxoglutarate-dependent enzymes: 3 hybrid DFT study. Chem. Eur: J. 10,1031-1041. Zhou, J., Gunsior, M., Bachmann, B. 0., Townsend, C. A., and Solomon, E. I. (1998) Substrate binding to the a—ketoglutarate-dependent non-heme iron enzyme Clavaminate synthase 2: coupling mechanism of oxidative decarboxylation and hydroxylation. J. Am. Chem. Soc. 120, 13539-13540. Zhou, J., Kelly, W. L., Bachmann, B. 0., Gunsior, M., Townsend, C. A., and Solomon, E. I. (2001) Spectroscopic studies of substrate interactions with Clavaminate synthase 2, a multifunctional or-KG-dependent non-heme iron enzyme: correlation with mechanisms and reactivities. J. Am. Chem. Soc. 123, 7388-7398. Ho, R. Y. N., Mehn, M. P., Hegg, E. L., Liu, A., Ryle, M. A., Hausinger, R. P., and Que, L., Jr. (2001) Resonance Raman studies of the iron(II)-or-keto acid chromophore in model and enzyme complexes. J. Am. Chem. Soc. 123, 5022-5029. (148) Bleijlevens, B., Shivarattan, T., Sedgwick, B., Rigby, S. E., and Matthews, S. J. (149) (2007) Replacement of non-heme Fe(II) with Cu(II) in the a-ketoglutarate dependent DNA repair enzyme AlkB: spectroscopic characterization of the active site. J. Inorg. Biochem., ASAP. Proshlyakov, D. A., Henshaw, T. F., Monterosso, G R., Ryle, M. J., and Hausinger, R. P. (2004) Direct detection of oxygen intermediates in the 55 (150) (151) (152) (153) (154) (155) (156) (157) non-heme Fe enzyme taurine/a-ketoglutarate dioxygenase. J. Am. Chem. Soc. 126, 1022-1023. Hoffart, L. M., Barr, E. W., Guyer, R. B., Bollinger, J. M., Jr., and Krebs, C. (2006) Direct spectroscopic detection of a C-H-cleaving high-spin Fe(IV) complex in a prolyl-4-hydroxylase. Proc. Natl. Acad. Sci. USA 103, 14738-14743. ‘ Galonic, D. P., Barr, E. W., Walsh, C. T., Bollinger, J. M., Jr., and Krebs, C. (2007) Two interconverting Fe(IV) intermediates in aliphatic chlorination by the halogenase Cth3. Nat Chem Biol 3, 113-6. Price, J. C., Barr, E. W., Glass, T. E., Krebs, C., and Bollinger, J. M., Jr. (2003) Evidence for hydrogen abstraction from C1 of taurine by the high-spin Fe(IV) intermediate detected during oxygen activation by taurineza-ketOglutarate dioxygenase (TauD). J. Am. Chem. Soc. 125, 13008-13009. Wolter, T., Meyer-Klaucke, W., Muther, M., Mandon, D., Winkler, H., Trautwein, A. X., and Weiss, R. (2000) Generation of oxoiron (IV) tetramesitylporphyrin pi-cation radical complexes by m—CPBA oxidation of ferric tetramesitylporphyrin derivatives in butyronitrile at - 78 degrees C. Evidence for the formation of six-coordinate oxoiron (IV) tetramesitylporphyrin pi-cation radical complexes FeIV = 0(tmp*)X (X = Cl-, Br-), by Mossbauer and X-ray absorption spectroscopy. J. Inorg. Biochem. 78, 117-22. Penner—Hahn, J. E., McMurry, T. 1., Renner, M., Latos-Grazynsky, L., Eble, K. S., Davis, I. M., Balch, A. L., Groves, J. T., Dawson, J. H., and Hodgson, K. 0. (1983) X-ray absorption spectroscopic studies of high valent iron porphyrins. Horseradish peroxidase compounds I and II and synthetic models. J. Biol. Chem. 258, 12761-4. Riggs-Gelasco, P. J., Price, J. C., Guyer, R. B., Brehm, J. H., Barr, E. W., Bollinger, J. M., Jr., and Krebs, C. (2004) EXAFS spectroscopic evidence for an Fe=0 unit in the Fe(IV) intermediate observed during oxygen activation by taurinezalpha-ketoglutarate dioxygenase. J. Am. Chem. Soc. 126, 8108-9. Borowski, T., de Marothy, S., Broclawik, E., Schofield, C. J., and Siegbahn, P. E. (2007) Mechanism for cyclization reaction by clavaminic acid synthase. Insights from modeling studies. Biochemistry 46, 3682-91. Blasiak, L. C., Vaillancourt, F. H., Walsh, C. T., and Drennan, C. L. (2006) Crystal structure of the non-haem iron halogenase SyrB2 in syringomycin biosynthesis. Nature 440, 368-71. 56 (158) (159) (160) (161) (162) (163) (164) (165) (166) (167) (168) Karlsson, A., Parales, J. V., Pareles, R. 15., Gibson, D. T., Eklund, H., and Ramaswamy, S. (2003) Crystal structure of naphthalene dioxygenase: side- -on binding of droxygen to 1ron Science 299, 1039- 1042. Kovaleva, E. G, and Lipscomb, J. D. (2007) Crystal structures of Fe2+ dioxygenase superoxo, alkylperoxo, and bound product intermediates. Science 316, 453-7. Goddard, W. A., 3rd, and Olafson, B. D. (1975) Ozone model for bonding of an O2 to heme in oxyhemoglobin. Proc Natl Acad Sci U S A 72, 2335-9. Valko, M., Rhodes, C. J., Moncol, J., Izakovic, M., and Mazur, M. (2006) Free radicals, metals and antioxidants in oxidative stress-induced cancer. Chem. Biol. Interact. 160, 1-40. Weiss, S. J. (1982) Neutrophil-mediated methemoglobin formation in the erythrocyte. The role of superoxide and hydrogen peroxide. J. Biol. Chem. 25 7, 2947-53. Simon, R. H., Scoggin, C. H., and Patterson, D. (1981) Hydrogen peroxide causes the fatal injury to human fibroblasts exposed to oxygen radicals. J. Biol. Chem. 256, 7181-6. Saari, R. E., and Hausinger, R. P. (1998) Ascorbic acid-dependent turnover and reactivation of 2,4-dichlorophenoxyacetic acid/a-ketoglutarate dioxygenase using thiophenoxyacetic acid. Biochemistry 3 7, 3035-3042. Welford, R. W. D., Schlemminger, I., McNeill, L. A., Hewitson, K. S., and Schofield, C. J. (2003) The selectivity and inhibition of AlkB. J. Biol. Chem. 278, 10157-10161. Ryle, M. J., Liu, A., Muthukumaran, R. B., Ho, R. Y. N., Koehntop, K. D., McCracken, J., Que, L., Jr., and Hausinger, R. P. (2003) 02- and or-ketoglutarate-dependent tyrosyl radical formation in TauD, an Ot-keto acid-dependent non-heme iron dioxygenase. Biochemistry 42, 1854-1862. Lindstedt, S., and Rundgren, M. (1982) Blue color, metal content, and substrate binding in 4-hydroxyphenylpyruvate dioxygenase from Pseudomonas Sp. strain P. J. 874. J. Biol. Chem. 25 7, 11922-11931. Liu, A., Ho, R. Y. N., Que, L., Jr., Ryle, M. J., Phinney, B. S., and Hausinger, R. P. (2001) Alternative reactivity of an a—ketoglutarate-dependent iron(II) oxygenase: enzyme self-hydroxylation. J. Am. Chem. Soc. 123, 5126-5127. 57 CHAPTER 2 PURIFICATION AND PROPERTIES OF ASPERGILL US N UDULANS XANTHINE HYDROXYLASE The work described in this chapter was combined with additional studies and published: Montero-Moran*, G. M.; Li*, M.; Rendon-Huerita, E.; Jourdan, F.; Lowe, D. J.; Stumpf‘f-Kane, A. W; Feig, M.; Scazzocchio, C.; and Hausinger, R. P. “Purification and characterization of the Fe(II)- and oi-Ketoglutarate-Dependent Aspergillus nidulans Xanthine Hydroxylase from Aspergillus nidulans” Biochemistry, 2007, 46 (18), 5293-5304 (‘ Co-first author). The studies briefly mentioned here regarding purification, Mr estimation, kinetic characterzaion, glycosylation analysis and phosphorylation testing of protein from the A. nudulans host were carried out by other authors. One figure included here (the subunit size comparison from both hosts examined by SDS-PAGE) was provided by Dr. Montero-Moran. 58 ABSTRACT Hisfi-tagged xanthine/a-ketoglutarate (OLKG) dioxygenase (XanA) of Aspergillus nidulans was purified from both the fungal mycelium and recombinant Escherichia coli cells, and the properties of the two forms of the protein were compared. The kinetic parameters were similar for XanA from the two sources (ka 30-70 3", Km of OLKG 31-50 “M, Km of xanthine ~45 11M, and pH optima at 7.0-7.4); however, the protein properties were markedly distinct. Evidence was obtained for both N- and O-linked glycosylation on the fungus-derived XanA, which aggregated into an apparent dodecamer, while bacterial-derived XanA was free of glycosylation and behaved as a monomer. Furthermore, the phosphorylation status differed for the two enzyme forms and the fungus-derived sample was shown to undergo extensive truncation at its amino terminus. The sites of posttranslational modification on the two forms of the enzyme are discussed in terms of a homology model of XanA. These studies represent the first biochemical characterization of purified xanthine/aKG dioxygenase. 59 INTRODUCTION Most organisms that metabolize xanthine possess a molybdopterin cofactor (Moco)-containing enzyme that hydroxylates the substrate to form uric acid while transferring electrons to NAD (xanthine dehydrogenase) or oxygen (xanthine oxidase) (I). These enzymes, referred to here as xanthine hydroxylases, are conserved throughout living organisms, including archaea, bacteria, fungi, plants, and metazoans. In 2005, a novel mechanism for xanthine metabolism was discovered in certain fungi (2). This finding arose out of the observation that all mutants of Aspergillus nidulans defective in xanthine dehydrogenase (i.e., with mutations affecting the structural gene th, the cnx genes for Moco synthesis, or th for sulfuration of Moco) retained the ability to grow on xanthine as sole nitrogen source (2, 3). A mutation affecting this alternative process was identified, and the cognate gene xanA was localized to chromosome VIII (4). Subsequently, the xanA gene and its homologues from Schizosaccharomyces pombe and Neurospora crassa were cloned and further homologues were identified in several other fungi (but not outside the fungal kingdom). The XanA sequence shows some similarity with the TauD group of Fe(II)- and a—ketoglutarate (orKG)-dependent dioxygenases (2), including a clear conservation of the Fe(II)- and aKG-binding sites. This homology suggested that the alternative xanthine oxidation mechanism present in some fungi might utilize an Fe(II)-dependent xanthine/OLKG dioxygenase. Such activity, depicted in Figure 2.1, was demonstrated in both crude and partially purified extracts of fungal mycelia of strains that expressed the 60 FIGURE 2.]: General mechanism of F e(11)/0tKG-dependent xanthine hydroxylases. H J; 6 5| :8\> Fe(II), XanA ; 1 I N>=o N 0 H O N H H OH o HO O 02 + O + 002 H00 "°° aKG succinate 61 xanA gene (2). The wide range of Fe(II)/0tKG hydroxylases utilize a diverse array of primary substrates (reviewed in (5)); however, XanA is the first described enzyme of this group to hydroxylate a free purine base. In the fungal kingdom, this enzyme coexists with the classical xanthine hydroxylase; i.e., some fungi possess both xanthine hydroxylase and xanthine/OLKG dioxygenase, while others possess only one or the other. Notably, yeasts as evolutionarily distant as S. pombe and Kluveromyces lactis are able to metabolize xanthine through the activity of a XanA homologue (2). They lack a classical xanthine dehydrogenase, and they are incapable of synthesizing Moco, which is universally present in the classical xanthine hydroxylases. The discovery of the novel Fe(II)/0LKG-dependent XanA enzyme poses both evolutionary and mechanistic problems. Is the xanthine-binding site of the newly identified enzyme at all similar to that of the classical xanthine hydroxylases (6, 7) or to the recently described xanthine transporters (8, 9) Is the mechanism of hydroxylation similar to that described for TauD (10, 11) What are the evolutionary advantages and disadvantages of possessing the Moco-containing and Fe(II)/0LKG-dependent enzymes. As a first step towards answering the above questions, I purified the Hisb-tagged versions of XanA from Escherichia coli and compared it to the corresponding protein purified from fungal mycelium. I confirm that the enzyme is an Fe(II)/0tKG dioxygenase and determine the pH dependence and kinetic parameters associated with the Fe(II), ocKG, and xanthine concentration dependencies. Comparison studies with mycelium-derived 62 enzyme show similar kinetic properties but the proteins differ in quaternary structure and identity of posttranslational modifications. In addition, fungus-derived protein is truncated at its amino terminus. Finally, 1 use a homology model of XanA to provide insights into the sites of posttranslational modification. These studies present the first detailed biochemical analysis of a purified Fe(II)-dependent xanthine/OLKG dioxygenase. More detailed kinetic inhibition and homology modeling studies are described in Chapter 3. 63 EXPERIMENTAL PROCEDURES The plasmid pxanA-Hiso was provided by Gabriela M. Montero-Moran, Institut de Ge'ne'tique et de Microbiologie, Universite’ Paris-Sud, Batiment 409, UMR 8621 CNRS, 91405 Orsay Cedex, France. Growth of E. coli Cells Overproducing A. nidulans XanA. Plasmid pxanA-Hi56 was transformed into XLlBlue E. coli cells (Stratagene) as described by Hanahan (12). A single colony of E. coli XLlBlue (pxanA-Hisé) was used to inoculate 50 mL of Luria Base Broth (Difco) containing 100 11g mL'l ampicillin. The cells were grown to saturation at 37 °C for 16 h and used to inoculate l L of LB media. The culture was grown at 37 ° C with vigorous shaking until reaching OD600 0.5-0.7, adjusted to contain 0.4 mM isopropyl-B-D-thiogalactopyranoside, transferred to 25 °C, and vigorously shaken for 16 h. Alternatively, E. coli XLlBlue (pxanA-Hisé) was grown aerobically in 200 mL LB-ampicillin medium at 37 °C to an OD600 of 0.6, induced with 0.75 mM IPTG, and incubated at 25 °C for 14 h with constant shaking (140 rpm). In either case, cultures were harvested by centrifugation at 8,000-9,000 rpm for 10 min at 4 ° C. Purification of His6-tagged XanA from E. coli. Approximately 5 g of E. coli XLBlue (pxanA-Hisé) cell paste was suspended in 10-15 mL of lysis buffer containing 100 mM Tris, pH 8.0, 300 mM NaCl, 25 mM imidazole, and trace amounts of lysozyme, leupeptin, DNase I and RNase A. This suspension was incubated at room temperature for 30 min, transferred to an ice bath for 30 min, disrupted by using a French pressure cell at ~500 psi at 4 ° C, and spun for 45 min at 100,000 g. The soluble cell extracts (30 mL) were loaded 64 onto a 10 mL Ni-nitrilotriacetic acid (NTA) column using 100 mM Tris buffer, pH 8.0, containing 300 mM NaCl and 25 mM imidazole, followed by elution with 100 mM Tris buffer, pH 8.0, with 300 mM NaCl, 250 mM imidazole and 15 % glycerol. The fractions containing XanA were collected and incubated with 1 mM EDTA at 4 °C for 5 h, then concentrated to 5-10 mg mL'I by using a Centriprep (Amicon Corp.) with a YM-lO membrane. Enzyme Assays. Xanthine/aKG dioxygenase activity was measured at 25 °C by using the following typical assay conditions (total volume of 1 mL): 1 mM orKG, 40 11M Fe(NH4)2(SO4)2, and 200 uM xanthine in 50 mM MOPS buffer, pH 7.4. Variations of these conditions included use of alternate buffers, different pH values, and varied concentrations of substrates. The absorbance at 294 nm was monitored to determine the uric acid production (8294 12,200 M’I cm") with a correction for loss of the xanthine absorbance at this wavelength (measured 8294 2,000 M’l cm") for an overall change in 3294 of 10,200 M" cm". Units of activity (U) were defined as umol min’1 of uric acid produced and the specific activity (U mg") was measured as umol min‘l (mg of purified XanA)". In addition to the above spectroscopic assay, xanthine/aKG dioxygenase activity was measured by two alternative methods. Oxygen consumption measurements were carried out in air-saturated MOPS medium (pH 7) at 25 °C by using a Clark-type oxygen electrode. Quantification of aKG consumed during the reaction was assessed by HPLC. Aliquots (300 uL) of the reaction mixtures were quenched by addition of 5 uL 6 M H2SO4, the 65 samples were centrifuged for 5 min at 20,000 g, and the supernatant was chromatographed on an Aminex HPX-87H column (Bio-Rad Laboratories) in 0.013 M H2804 with detection by using a differential refractometer (Waters, Model R401). Metal Analyses. The iron concentration was measured by utilizing the KMnO4 oxidation, ascorbate reduction, and ferrozine chelation protocol of Bienert (13). Protein Analytical Methods. Routine determinations of protein concentration were carried out by the method of Bradford (14) with bovine serum albumin used as the standard. Qualitative measurement of protein overexpression and assessment of protein purity made use of sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) (15), with stacking and running gels containing 5 % and 12 % acrylamide. Standard proteins used for comparison included phosphorylase b, M, 97,400; bovine serum albumin, M, 66,200; ovalbumin, M, 45,000; carbonic anhydrase, M, 31,000; trypsin inhibitor, M, 21,500; and lysozyme, M, 14,400 (Bio-Rad Laboratories). The native size of XanA isolated from E. coli was estimated by gel filtration chromatography using a Protein-pak Diol(OH) 10 um column (Waters, 0.5 or 1 min mL", room temperature in 100 mM Tris buffer, pH 7.5, containing 300 mM NaCl), a SuperoseTM 6 HR 10/30 GL column (Pharrnacia, 1 mL min'1 in 50 mM MOPS, pH 6.8, containing 0.15 M NaCl). The calibration proteins were thyroglobulin, M, 670,000; y-globulin, M, 158,000; ovalbumin, M, 44,000; myoglobin, M, 17,000; and vitamin B12, M, 1,350 (Bio-Rad). Mass Spectrometry. (Assisted by Dan Jones, Department of Biochemistry & Chemistry, MSU). Mass spectrometry analyses were performed by using a Waters 66 (Milford, MA) LCT Premier mass spectrometer coupled to a Shimadzu (Columbia, MD) LC-20AD HPLC and SIL-5000 autosampler. Samples were analyzed using electrospray ionization in positive ion mode. On-line desalting and separation from detergents was performed using a Thermo Hypersil-Keystone BetaBasic cyano column (1.0 x 10 mm) coupled to the electrospray ionization probe.Aliquots were injected onto the column using a flow rate of 0.1 mL/min of 95% solvent A (0.15% aqueous formic acid)/5% solvent B (acetonitrile). Gradient elution was performed by using the following parameters: (0-1 min: 95%A/5%B; linear gradient to 30%A/70%B at 6 min; hold at 30%A/70%B until 9 min). Electrospray spectra were processed using MassLynx software (Waters, Milford, MA), and zero-charge state mass spectra were obtained by deconvolution using the MaxEntl algorithm. MALDI mass spectra were generated on a Voyager-DE STR mass spectrometer (Applied Biosystems, Foster City, CA) in positive ion linear mode using sinapinic acid as matrix. Samples were processed using strong cation exchange ZipTips (ZipTipSCX, Millipore, Billerica, MA) to remove detergent and reversed phase C18 ZipTips for desalting, following the manufacturer's recommended protocols, before spotting the MALDI target. Structural Homology Modeling. (Assisted by Michael Feig and Andrew W. Stumpff-Kane, Department of Biochemistry and Molecular Biology, MSU.) Details related to the creation of the homology model using TauD (PDB code IOS7, chain A) as a template structure (16, 1 7) are described in Chapter 3 (18). 67 RESULTS Properties of Recombinant XanA in E. coli Cell Extracts. When assayed by using standard conditions (1 mM aKG, 40 11M Fe(NH4)2(SO4)2, and 200 uM xanthine in 50 mM MOPS buffer, pH 7.4, 25‘ °C), fresh extracts of E. coli cells overproducing Hisé-tagged XanA exhibited xanthine/orKG dioxygenase specific activity of approximately 19 U (mg protein)’1 (Figure 2.2A). As expected, the activity was strictly dependent on the presence of aKG in the assay as previously described for the crude enzyme extracted from A. nidulans (Figure 2.28) (2). Unlike the earlier findings, however, significant activity was detected in the absence of added Fe(II) (16.9 U mg") (Figure 2.2C) and some activity was retained when 5 mM EDTA was included in the assay buffer (2.01 U mg"), Figure 2.2D. These data indicate that endogenous Fe(II) is tightly bound to the bacterial-produced enzyme sample. The activity of these extracts was irreversibly lost over 3 weeks when stored at 4 °C using a protein concentration of 15-17 mg mL", or after 1 week in the added presence of 1 mM EDTA. While long-term incubation of cell extracts with EDTA was undesirable, studies with enriched enzyme samples showed that EDTA treatment provided stable apoprotein (when maintained at high protein concentrations) that could be activated by addition of ferrous ions. Purification of XanA from the E. coli. XanA produced in E. coli XLlBlue (pxan-His6) was purified to homogeneity from cell extracts (Figure 2.3A) by Ni-NTA chromatography. Inclusion of 15 % glycerol in the chromatography buffers, which were maintained on ice, helped to minimize protein precipitation during purification. About 5 % of the soluble 68 FIGURE 2.2: Activity assay of XanA cell extracts. (A) Complete assay: 15 pg XanA crude extract, 40 uM F e(ll), 1 mM aKG, 100 uM xanthine in 50 mM MOPS, pH=7.4, 25 °C. (B) without aKG. (C) without Fe(II). (D) with 5 mM EDTA. UV-Vis spectra were recorded from 240 to 320 nm every 15 seconds after initiating the enzymatic reaction by adding primary substrate. °2io ' 230 Y 260 ' 300 320 Wavelengthmm) Wavelengtlflnm) Abs 0.0 . a - . - v 240 200 280 300 320 Wavelengthmm) Wavelengthmm) 69 activity was located in the flow-though fractions, perhaps indicating that some XanA interacts tightly with other proteins that fail to bind the resin. As measured by the standard assay protocol, the Ni-NTA column fractions containing purified XanA accounted for 33 % of the activity that had been observed in cell extracts. When this pool was treated with 1 mM EDTA at 4 ° C for 5 h and then concentrated to 5-12 mg mL", the activity increased such that 60 % of the activity of cell extracts was recovered and yielded a final specific activity for the purified enzyme of 70-80 U (mg protein)" at 25 °C which corresponds to a km, of 49-56 s'l (assuming M, = 42 kDa per subunit). The enzyme recovered from the Ni-NTA column contained 0.26 to 0.5 moles of Fe per mole of subunit according to the colorimetric assay, while that incubated with EDTA lacked detectable Fe, and a sample incubated with exogenous Fe(II) and then chromatographed on a Sephadex G-25 gel filtration column contained 1.3 moles of Fe per mole of subunit. Concentrated XanA derived from E. coli was stable for at least one month at 4 °C when stored in 100 mM Tris buffer, pH 8.0, containing 300 mM NaCl, 250 mM imidazole, 1 mM EDTA and 15 % glycerol, or at least two months if frozen at -80 ° C. For comparison, the specific activity of the isolated protein from the firngal host was measured as 22-40 U (mg protein)" at 30 °C equivalent to a km, of 15.4-30 5" (again assuming M, = 42 kDa per subunit), comparable with those from bacterial host. (The activity of XanA isolated from fungal host was measured by Gabriela M. Montero-Moran, Institut de Génétique et de Microbiologie, Université Paris-Sud, Batiment.) 70 FIGURE 2.3: SDS-PAGE analysis of the purified XanA from E. coli and A. nidulans. (A) Analysis of fractions during purification of XanA from E. coli. Lane 1, molecular weight standards; lanes 2 and 3, suspension of cultures before and after induction by IPTG; lane 4, cell extracts afier lysis; lane 5, cell pellet after lysis; lane 6, flow-through fraction from Ni-NTA chromatography; lane 7-13, fractions of purified XanA obtained after Ni-NTA chromatography. (B) Comparison of the purified XanA protein derived from A. nidulans and E. coli. Lane M, markers; lane 1, sample purified from the fungus; lane 2, protein isolated from the bacterium (7 pg each). Stacking and running gels contain 5 % and 12 % acrylamide. 12345070 910111213 kDa 97 . as 97.4 .. 45 86.2 45.0 1 31 .0 . 11.: 21.5 ‘ 14.4 J ‘ 7I Effects of pH on Stability and Activity of XanA from E. coli. The effect of pH on the stability of XanA was examined. After incubating the XanA samples in various pH buffers at 4 °C for 3 h, the activity remaining was examined by using the standard assay procedure. The results (data not shown) indicate that each purified enzyme is stable over a wide pH range (7.0-11.0). The effect of pH on the activity of the enzyme was also examined. XanA isolated from E. coli was assayed by using a series of different buffers as depicted in Figure 2.4. The enzyme activity exhibited a narrow pH optimum of 7.0-8.0 with pH 7.4 yielding optimal activity for most buffers (Tris, MOPS, MES, imidazole, and HEPES). Kinetic Analyses of XanA from E. coli. For the E. coli-derived enzyme, the results of studies using varied orKG concentrations provided a Km of 3 1.1 i 1.6 uM and km of 66.5 s" at 25 °C (Figure 2.5A), while those for varied xanthine concentrations provided a Km of 45.2 :t 3.6 uM and kca, of 71.4 s" (Figure 2.5B). Fe(lI) is required for xanthine/aKG dioxygenase activity, with half-maximal activity at ~7 11M when using the apoprotein isolated from the bacterium (Figure 2.6). The kinetic parameters were very similar for XanA isolated from the two sources. When the protein isolated from A. nidulans enzyme was assayed at 30 °C with varied aKG or xanthine concentrations the measured Km values were 50 11M 2t 6 and 46 d: 4 pM, respectively, but with a smaller Item that ranged from 15 s" to 30 5" depending on the preparation provided by Gabriela M. Montero-Moran. 72 FIGURE 2.4: pH dependence of XanA activity. The activity of XanA (0.51 pg ml") derived from E. coli was assayed in the following buffers (50 mM) and pH values at 25 °C: imidazole, pH 6.0-8.2 (O);MES, pH 666.8 (+); MOPS, pH 6.6-7.8 (X); HEPES, pH 7.0-8.2 (V); Tris, pH 7.4-9.0 (o);CHES, pH 9.4-9.8 (A); CAPS, pH 9.6-10.0 (0). Assay solutions also contained 1 mM aKG, 40 uM Fe(NH4)2(SO4)2, and 200 uM xanthine. 73 FIGURE 2.5: Substrate and co-substrate concentration dependencies of XanA. The effects of varying the concentrations of (A) aKG and (B) xanthine on xanthine/aKG dioxygenase activity were examined for the E. coli-derived protein at 25 °C. Except for the compound being varied, the assay solutions contained 40 pM Fe(II), 1 mM aKG, and 200 uM xanthine in 50 mM MOPS buffer, pH 7.4. The data were fit to the Michaelis-Menten equation. 1M' Utmg 3-$-§-$_ 0 0 100200300400500000 alpha-KG (fl) 4 oo- - 1 oo- 4 EN' 8 D zo- 07 V I ' U ' U V V f1 j o 40 so 120 100 zoo "WIN (W) 74 The stoichiometry of the enzymatic reaction was examined for XanA by UV-Vis spectroscopy and 02 electrodes. The degradation of 100 pM xanthine was accompanied by the production of 92 uM uric acid as measured by UV-Vis spectrosocopy, and this coincided with the consumption Of 110 uM oxygen determined by O; electorde (data not shown). Differential Protein Properties of XanA Purified from the Two Host Cells. Gel filtration chromatography provided an estimated M, of 39-42 kDa for the native enzyme isolated from the bacterial host (consistent with a monomeric structure) (Figure 2.7); however, similar analysis of the protein from the fungal host showed that it was oligomeric, with an approximate M, of 500 kDa that was consistent with about twelve subunits per native molecule (data from Gabriela M. Montero-Moran.) (18).The SDS-PAGE results highlight a key difference between the XanA proteins isolated from the two sources; i.e., the apparent M, of the E. coli-derived protein is larger than that of the protein derived from the fungus (Figure 2.3B) (Provided by Montero-Morén). This finding led us to investigate the possibility of unique posttranslational modifications in the proteins produced in the bacterial and eukaryotic hosts. G M. Montero-Moran and co-workers have shown the presence of N-glycosylation, O-glycosylation and Thr phosphorylation in the protein purified from A. nidulans, but both Thr and Ser phosphorylation in the one isolated from bacterial host (18). Mass spectrometric methods were used to further characterize the two enzyme forms. Electrospray ionization mass spectrometry of bacteria-derived XanA indicated a single 75 FIGURE 2.6: Fe(ll) concentration dependence of XanA. The effects of varying the concentration of Fe(II) on xanthine/aKG dioxygenase activity were examined by using the E. coli-derived protein at 25 °C in solutions containing 1 mM aKG and 200 1.1M xanthine in 50 mM MOPS buffer, pH 7.4. 76 FIGURE 2.7: Measurement of native size of XanA derived from E. coli. (I) represents the standards and (1:1) represents the XanA isolated form E. coli. 6 .- I 15 ' E 0’ I .04 . 3 I I l . 9 12 15 18 elution volume (mL) 77 species with a molecular mass of 41,992 Da (data not shown), which matches very well to the theoretical mass (41,996.50 Da; using the ExPASY Compute pI/Mw tool at ca.expasy.org) for the His6-tagged protein missing its amino-terminal Met residue. This sample provided no evidence Of phosphorylation (differing from the immunological detection of phosphoserine and phosphothreonine by our collaborators) or glycosylation. In contrast to this single species, the fungus-derived protein sample exhibited a complex electrospray ionization mass spectrum centered near 36,000 Da (Figure 2.8). The spectrum of Figure 2.8 includes features separated by 162 mass units, consistent with glycosylation involving hexose sugars, as well as features separated by 80 mass units, indicating phosphorylation. The smallest component of the spectrum exhibits a mass of 35,171 Da, indicating that the non-glycosylated and non-phosphorylated fungal protein is severely truncated compared to the theoretical mass of full-length protein of 42,127.69. This truncation must occur at the N-terminus, since the C-terrninal Hisé-tag was used for enzyme purification, and consists of approximately 60 residues. 78 FIGURE 2.8: Mass spectrometric analysis of fungus-derived XanA. The XanA protein isolated from A. nidulans was analyzed by electrospray ionization mass spectrometry. The figure depicts a series of peaks separated by 162 mass units and 80 mass units, consistent with glycosylation and phosphorylation, respectively. Emomiucaw em a $3 3.. _m,...§ 5.52 ”9.0 $928 88b graham: as 3568 856 so- 882. 895.0 853 o magma swarm Swab ammo Bum» a 88mm 86mm 883 «Om Em mm+ am Bach 83mm 3086 gm mflwub 33mm 3 wwmumm WEE .m wunoc makes . wan: .m Smwu wmmoo muooo wumoo 79 DISCUSSION In this study I describe the isolation and general properties of xanthine/aKG dioxygenase, a novel enzyme found exclusively in the fungal kingdom (2). Purification. Immobilized metal ion chromatography is very effective for purifying recombinant XanA from the bacterial host cells. His-tagged versions of several other representative Fe(ll)/0LKG dioxygenases have been studied, including AlkB, deA, SdpA, phytanoyl-CoA hydroxylase, and the oxygen-sensing prolyl 4- and asparaginyl hydroxylases (19-23). The His-rich sequences generally have modest if any effect on activity. XanA Enzyme Stability. Purified XanA is unstable at room temperature (denaturing within an hour even in buffers containing 15 % glycerol), when agitated (e.g., during stirring in an Amicon concentrator), or when incubated at pH values below 6.5 (the pI estimated for XanA is 5.82 according to the ExPASy ProtParam tool at ca.expasy.org). EDTA treatment enhances the activity of the highly concentrated bacterial-derived enzyme and this compound is maintained in the storage buffer to ensure maximal lifetime of the activity. I attribute the enhancement effect of EDTA to its ability to remove inhibitory Ni(II) (co-eluted with the enzyme from the NTA resin) and Fe(III) (the oxidized, inactive state of the metal) from the enzyme so that the apoprotein can bind Fe(lI) in the assay buffer. Several other Fe(II)/0LKG dioxygenases are purified as their apoprotein forms by inclusion of chelators (24, 25) to prevent Fe(II) oxidation and to eliminate Ni(II), if purified using an NTA column. In contrast, some representatives of 80 these enzymes have been purified anaerobically to assure that the metal remains in its reduced form (26, 27). XanA Enzyme Activity. The kinetic properties of XanA as purified from E. coli (~70 U mg ", Km of 31 pM for orKG, and Km of 45 uM for xanthine at pH 7.4) compare well with those of XanA isolated from A. nidulans (30 U mg", and Km values of 50 uM and 46 uM at pH 7.0), and these results are compatible with those reported earlier for enriched sample from the fungus (40 U mg", 50 uM, and 23 uM) (2). The reaction requires Fe(II) (half-maximal activity at 7 uM for standard conditions), consistent with the results of . related family members. Posttranslational Modifications. My comparison of recombinant XanA purified from E. coli with that of A. nidulans provided by my collaborators reveals very different quaternary structures and posttranslational modifications. Whereas the protein derived from the bacterial source chromatographs on a size exclusion column as a monomer, that isolated from the fungus is much larger, with apparent M, of 500 kDa. The sample from the fungus was also glycosylated and phosphorylated while the one from bacterial was suggested to contain both phosphoserine and phosphothreonine by our collaborators. Treatment of the fungal protein with PNGase F was found to result in a dramatic shift in electrophoretic mobility, but not all glycosylation was removed by this process indicating the presence of both N— and O-glycoconjugates. Comparing the size of the subunit from both sources along with the mass spectrometry results suggested an extensive truncation takes place at the N-terminus of the fungal enzyme. Despite the extensive differences in 81 post-translational modifications between the two forms of the enzyme the kinetic parameters are nearly identical, consistent with the modifications not affecting catalytic activity. Analogous biochemical comparisons between the enzyme form isolated from its native eukaryotic host and the form isolated from E. coli have not been reported for other F e(II)/aKG dioxygenase family members. An extant question is the role of the extensive N-terminal processing of enzyme isolated from its natural fungal host. I speculate that this feature may relate to cellular targeting of the XanA enzyme. It is not unreasonable to propose that an oxidoreductase such as XanA could be peroxisomal. In fact, the immediate downstream enzyme, urate oxidase, has been shown to be peroxisomal in every organism where its localization has been investigated, including amoebas, mammals and plants (28-31). A urate oxidase-green fluorescent protein fusion shows a particulate intracellular distribution in Aspergillus nidulans, fully consistent with a peroxisomal localization (G. Langousis and G. Diallinas, unpublished data). XanA does not possess a C-terminal peroxisomal targeting signal (variations of an SKL tripeptide, denoted PTSI); however, it contains a RSALYTHL sequence (residues 40-47) that resembles the PTSZ N-terminal import sequence (variations of RLXsHL) found in a minority of peroxisomal proteins throughout the eukaryotes (32-34). In some instances, including mammalian (35), yeast (36), and plant (37) peroxisomal proteins, it has been shown that the PST2 is contained in a pre-sequence that is cleaved upon peroxisomal entry. The function of XanA glycosylation could be understood in this context as a mechanism to protect the protein from further proteolysis. 82 FIGURE 2.9: Location of potential sites of glycosylation and phosphorylation in XanA. The positions of selected residues are depicted in a homology model, described in the chapter 3, with the potential N-glycosylation site (Asn118) in blue, Thr and Ser residues that could be either glycosylated or phosphorylated (T hr5 and Ser37) in orange and magenta, potential 0-glycosylation sites (Thr10, Thr195, and Ser316) in red, and other possible phosphothreonines (Thr38, Thr58, Thr66, Thr120, and Thr274) or phosphoserines (Ser15, Serl42, Serl99, Ser208, Ser217, Ser2l8, and Ser34l) illustrated in yellow and cyan. The amino terminus of the fungus-derived protein is truncated through the first approximately 60 residues. The brown sphere shows the postulated position of Fe(II). Additional glycosylation (Ser231) and phosphorylation (Thrl77 and Ser184) are not shown because they are located on non-modeled protein loops that represent insertions compared to the template structure (TauD, PDB code 1087, chain A). stems Ser142 56'2" 83 Insights from a Homology Model. A XanA homology model (Figure 2.9), whose generation is described in detail in Chapter 3 (18), provides critical insights into the likely sites of posttranslational modification in the protein. This model is based on the full-length sequence of the XanA protein, but we assume that the overall fold is maintained even in the truncated version derived from the fungal host because this region of the protein is external to the DSBH core of the protein. Consistent with the PNGase F results that demonstrate the presence of N-glycosylation sites in the fungus-derived protein, the NetNGlyc 1.0 server (R. Gupta, E. Jung, and S. Brunak, unpublished) predicts glycosylation of Asnl.18, and the homology model places this residue on the protein surface. The PNGase F-treated fungal-derived protein still stains as a glycoprotein, consistent with O-Iinked glycoconjugates in the protein. Thr5 and ThrIO are predicted to be glycosylated by the NetOGlyc 3.1 server (38) while the YingOYang 1.2 server (R. Gupta, J. Hansen, and S. Brunak, unpublished) predicts Thr5, Ser37, Thr195, Ser23 1 , and Ser3l6 as possible sites of glycosylation. The XanA homology model reveals that all of these positions except for Thr195 are surface exposed or, for Ser23], on surface loops that were not modeled. On the basis of my mass spectrometry results showing truncation of approximately 60 residues from the N-terminus, I suggest Ser23] or Ser3l6 as the most likely sites of O-glycosylation. A wide range of potential Ser, Thr, or Tyr phosphorylation sites are identified in XanA by the NetPhos 2.0 server (39) and by Prosite (www.cxpasvorg). Phosphoproteins are common in eukaryotes and are well known in E. coli (40). Given the immunological 84 evidence for phosphothreonine in XanA from both sources, the Thr phosphorylation sites are of special interest and are predicted by at least one program to include Thr5, Thr38, Thr58, Thr66, Thr120, Thrl77, and Thr274. The first three residues are likely to be missing in the fungus-derived protein based on the mass spectrometry results. The homology model predicts that Thr66 and Thr120 are buried (although their disposition is less clear for the truncated protein), whereas the other residues are exposed to the surface or, for Thrl77, on a loop that was not modeled. The bacterial-derived protein also reacts with antibodies directed against phosphoserine, whereas XanA derived, from the fungus reacted only weakly with these antibodies. We attribute this difference in reactivity between the two proteins to the phosphorylation site being inaccessible in the fungal-derived sample, either due to nearby glycosylation or due to being located at a subunit interface in the multimeric protein. Sites of Ser phosphorylation predicted by at least one program include Ser15, Ser37, Serl42, Ser184, Serl99, Ser208, Ser217, Ser218, and Ser34l. The first two residues are likely to be absent in the fungus-derived protein. Residues Ser199 and Ser341 are predicted to be somewhat buried in the model, Ser184 occurs on a loop that could not be modeled, and the other residues are predicted to be surface exposed. Of note, Ser341 would be made more inaccessible by glycosylation of Thr195; thus, potentially explaining the source specificity behavior for Ser phosphorylation. 85 REFERENCES (1) (2) (3) (4) (5) (6) (7) (8) (9) Hille, R. (2005) Molybdenum-containing hydroxylases. Arch. Biochem. Biophys. 433, 107-116. Cultrone, A., Scazzocchio, C., Rochet, M., Montero-Moran, G, Drevet, C., and F emandez-Martin, R. (2005) Convergent evolution of hydroxylation mechanisms in the fungal kingdoms: molybdenum cofactor-independent hydroxylation of xanthine via or-ketoglutarate dependent dioxygenase. Molec. Microbiol. 5 7, 276-290. Darlington, A. J ., and Scazzocchio, C. (1968) Evidence for an alternative pathway of xanthine oxidation in Aspergillus nidulans. Biochim. Biophys. Acta 166, 557-568. SeaIy-Lewis, H. M., Scazzocchio, C., and Lee, S. (1978) A mutation defective in xanthine alternative pathway of Aspergillus nidulans: its use to investigate the specificity of uaY mediated induction. Molec. Gen. Genet. 164, 303-308. Hausinger, R. P. (2004) Fe(lI)/or-ketoglutarate-dependent hydroxylases and related enzymes. Crit. Rev. Biochem. Mol. Biol. 39, 21-68. Glatigny, A., Hof, P., Romao, M. J., Huber, R., and Scazzocchio, C. (1998) Altered specificity mutations define residues essential for substrate positioning in xanthine dehydrogenase. J. Mol. Biol. 278, 431-438. Truglio, J. J ., Theis, K., Leimkt'rhler, S., Rappa, R., Rajagopalan, K. V., and Kisker, C. (2002) Crystal structure of the active and alloxanthine-inhibited forms of xanthine dehydrogenase from Rhodobacter capsulatus. Structure 10, 115-125. Goudela, S., Karatza, P., Koukaki, M., Frilingos, S., and Diallinas, G (2005) Comparative substrate recognition by bacterial and fungal purine transporters of the NAT/NCS2 family. Molec. Membrane Biol. 22, 263-275. Koukaki, M., Vlanti, A., Goudela, S., Pantazopoulou, A., Gioule, H., Toumaviti, S., and Diallinas, G (2005) The nucleobase-ascorbate transporter (NAT) signature motif in UapA defines the function of the purine translocation pathway. J. Mol. Biol. 350, 499-513. 86 (10) (11) (12) (13) (14) (15) (16) (17) (18) (19) Eichhom, E., van der Ploeg, J. R., Kertesz, M. A., and Leisinger, T. (1997) Characterization of or-ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J. Biol. Chem. 272, 23031-23036. Grzyska, P. K., Ryle, M. J., Monterosso, G R., Liu, J., Ballou, D. P., and Hausinger, R. P. (2005) Steady-state and transient kinetic analyses of taurine/a-ketoglutarate dioxygenase: effects of oxygen concentration, alternative sulfonates, and active site variants on the Fe(IV) intermediate. Biochemistry 44, 3845-3855. Hanahan, D. (1983) Studies on transformation of Escherichia coli with plasmids. J. Mol. Biol. 166, 557-580. Beinert, H. (1978) Micro methods for the quantitative determination of iron and copper in biological material. Meth. Enzymol. 54, 435-445. Bradford, M. M. (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. 72, 248-254. Laemmli, U. K. (1970) Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature (London) 22 7, 680-685. Elkins, J. M., Ryle, M. J., Clifton, I. J., Dunning Hotopp, J. C., Lloyd, J. S., Burzlaff, N. 1., Baldwin, J. E., Hausinger, R. P., and Roach, P. L. (2002) X-ray crystal structure of Escherichia coli taurine/a-ketoglutarate dioxygenase complexed to ferrous iron and substrates. Biochemistry 41, 5185-5192. O'Brien, J. R., Schuller, D. J., Yang, V. S., Dillard, B. D., and Lanzilotta, W. N. (2003) Substrate-induced conformational changes in Escherichia coli taurine/a-ketoglutarate dioxygenase and insight into the oligomeric structure. Biochemistry 42, 5547-5554. Montero-Moran, G M., Li, M., Rendon-Huerta, E., Jourdan, F., Lowe, D. J., Stumpff-Kane, A. W., Feig, M., Scazzocchio, C., and Hausinger, R. P. (2007) Purification and characterization of the Fe(II)- and or-ketoglutarate-dependent xanthine hydroxylase from Aspergillus nidulans. Biochemistry 46, 5293-5304. Trewick, S. C., Henshaw, T. F., Hausinger, R. P., Lindahl, T., and Sedgwick, B. (2002) Oxidative demethylation by Escherichia coli AlkB directly reverts DNA base damage. Nature 419, 174-178. 87 (20) (21) (22) (23) (24) (25) (26) (27) (28) (29) Muller, T. A., Zavodszky, M. I., F eig, M., Kuhn, L. A., and Hausinger, R. P. (2006) Structural basis for the enantiospecificities of R- and S-phenoxypropionate/or-ketoglutarate dioxygenases. Prot. Science 15, 1356-1368. Searls, T., Butler, D., Chien, W., Mukherji, M., Lloyd, M. D., and Schofield, C. J. (2005) Studies on the specificity of pro- and mature-forms of phytanoyl-CoA 2-hydroxylase and mutation of the iron binding ligands. J. Lipid Res. 46, 1660-1670. McDonough, M. A., Li, V., Flashman, E., Chowdhury, R., Mohr, C., Liénard, B. M. R., Zondlo, J., Oldham, N. J., Clifton, I. J., Lewis, J., McNeil], L. A., Kurzeja, R. J. M., Hewitson, K. S., Yang, E., Jordan, S., Syed, R. S., and Schofield, C. J. (2006) Cellular oxygen sensing: crystal structure of hypoxia-inducible factor prolyl hydroxylase (PHD2). Proc. Natl. Acad. Sci. USA 103, 9814-9819. Dann, C. E., III, Bruick, R. K., and Deisenhofer, J. (2002) Structure of a factor-inhibiting hypoxia-inducible factor 1: an essential asparaginyl hydroxylase involved in the hypoxic response pathway. Proc. Natl. Acad. Sci. USA 99, 15351-15356. Ryle, M. J., Padmakumar, R., and Hausinger, R. P. (1999) Stopped-flow kinetic- analysis of Escherichia coli taurine/or-ketoglutarate dioxygenase: interactions with or-ketoglutarate, taurine, and oxygen. Biochemistry 38, 15278-15286. Fukumori, F., and Hausinger, R. P. (1993) Purification and characterization of 2,4-dichlorophenoxyacetate/a-ketoglutarate dioxygenase. J. Biol. Chem. 268, 24311-24317. Mishina, Y., Chen, L. X., and He, C. (2004) Preparation and characterization of the native iron(III)-containing DNA repair AlkB protein directly from Escherichia coli. J. Am. Chem. Soc. 126, 16930-16936. Vaillancourt, F. H., Yin, J., and Walsh, C. T. (2005) SyrBZ in syringomycin E biosynthesis is a nonheme Fell a-ketoglutarate- and Oz-dependent halogenase. Proc. Natl. Acad. Sci. USA 102, 10111-10116. Muller, M., and Moller, K. M. (1969) Urate oxidase and its association with peroxisomes in Acanthamoeba sp. Eur: J. Biochem. 9, 424-430. Leighton, F., Poole, B., Lazarow, P. B., and de Duve, C. (1969) The synthesis and 88 (30) (31) (32) (33) (34) (35) (36) (37) (38) (39) turnover of rat liver peroxisomes. I. Fractionation of peroxisome proteins. J. Cell. Biochem. 41, 521-535. Nguyen, T., Zelechowska, M., Foster, V., Bergmann, H., and Verma, D. P. S. (1985) Primary structure of the soybean nodulin-35 gene encoding uricase II localized in the peroxisomes of uninfected cells of nodules. Proc. Natl. Acad. Sci. USA 82, 5040-5044. Breidenbach, R. W., Kahn, A., and Beevers, H. (1968) Characterization of glyoxysomes from castor bean endosperm. Plant Physiol. 43, 705-713. de Hoop, M. J., and Ab, G (1992) Import of proteins into peroxisomes and other microbodies. Biochem. J. 286, 657-669. Petriv, O. 1., Tang, L., Titorenko, V. 1., and Rachubinski, R. A. (2004) A new definition for the consensus sequence of the peroxisome targeting signal type 2. J. Mol. Biol. 341, 119-134. Reumann, S. (2004) Specification of the peroxisome targeting signals type 1 and type 2 of plant peroxisomes by bioinformatics analyses. Plant Physiol. 135, 783-800. Swinkels, B. W., Gould, S. J., Bodnar, A. G, Rachubinski, R. A., and Subramaniam, S. (1991) A novel, cleavable peroxisomal targeting signal at the amino-terminus of rat 3-ketoacyl-C0A thiolase. EMBO J. 10, 3255-3262. Lee, J. G, Lee, Y. J., Lee, C. H., and Maeng, P. J. (2006) Mutational and functional analysis of the cryptic N-terminal targeting signal for both mitochondria and peroxisomes in yeast peroxisomal citrate synthase. Biochem (Tokyo) 140, 121-133. Kato, A., Hayashi, M., Kondo, M., and Nishimura, M. (2000) Transport of peroxisomal proteins synthesized as large precursors in plants. Cell Biochem. Biophys. 32, 269-275. Julenius, K., Molgaard, A., Gupta, R., and Brunak, S. (2005) Prediction, conservation analysis and structural characterization of mammalian mucin-type O-glycosylation sites. Glycobiology 15, 153-164. Blom, N., Gammeltoft, S., and Brunak, S. (1999) Sequence- and structure-based prediction of eukaryotic protein phosphorylation sites. J. Mol. Biol. 295, 89 1351-1362. (40) Cortay, J.-C., Rieul, C., Duclos, B., and Cozzone, A. J. (1986) Characterization of the phosphoproteins of Escherichia coli cells by electrophoretic analysis. Eur. J. Biochem. 159, 227-237. 90 CHAPTER 3 KINETIC CHARACTERIZATION, ISOTOPE EFFECTS, AND EFFECTS OF ALTERNATE METAL IONS, SALTS, AND SUBSTRATE ANALOGUES ON ASPERGILL US NIDULANS XANTHINE HYDROXYLASE The studies described in this chapter were combined with additional investigations and published: Montero-Moran*, G. M.; Li*, M.; Rendon-Huerita, E.; Jourdan, F.; Lowe, D. J .; Stumpff-Kane, A. W; Feig, M.; Scazzocchio, C.; and Hausinger, R. P. “Purification and characterization of the Fe(II)- and a-Ketoglutarate-Dependent Aspergillus nidulans Xanthine Hydroxylase from Aspergillus nidulans” Biochemistry, 2007, 46 (18), 5293-5304 (. Co-first author). Michael Feig and Andrew W. Stumpff-Kane constructed the homology model, David J. Lowe and Fabrice Jourdan synthesized the 8-hydroxypurine, 2,8-dihydroxypurine and 6,8-dihydroxypurine, and Christine Drevet provided the alignment figure of XanA with putative orthologous sequences from several fungi. 91 ABSTRACT Xanthine/or-ketoglutarate (OtKG) dioxygenase (XanA) of Aspergillus nidulans was previously purified from fungal mycelium and recombinant Escherichia coli cells, and the two forms of the enzyme were shown to exhibit distinct posttranslational modifications while retaining similar kinetic parameters as described in Chpater 2. Here, I characterize several additional kinetic properties of the XanA isolated form E. coli. The enzyme exhibits no significant isotope effect when using 8-2H-xanthine; however, it demonstrates a two-fold solvent deuterium isotope effect. Cu(II) and Zn(II) potently inhibit the Fe(II)-specific enzyme, presumably by binding to the Fe(II) site, whereas Co(Il), Mn(ll), and Ni(II) are weaker inhibitors. NaCI inhibits the enzyme, resulting in decreases in km, and increases in Km of both ocKG and xanthine. The 01KG cosubstrate can be substituted by a—ketoadipate (resulting in a 9-fold decrease in kca, and a 5-fold increase in the Km compared to the normal (x-keto acid), while the OtKG analogue N-oxalyl glycine is an effective competitive inhibitor (K, 0.12 uM). No alternative purines are able to effectively substitute for xanthine as a substrate, and only one purine analogue (6,8-dihydroxypurine, DHP) results in significant inhibition. A homology model of XanA was generated on the basis of the structure of the related enzyme TauD (PDB code lOS7) and used to provide insights into the mode of substrate and inhibitor binding. These studies present the first analysis of xanthine/OLKG dioxygenase isotope effects, interactions With alternative metal ions and NaCl, and behavior with substrate analogues. 92 INTRODUCTION In Chapter 2, I described the purification and basic properties of the Fe(II)-dependent xanthine/a-ketoglutarate (aKG) dioxygenase (XanA) of Aspergillus nidulans. This enzyme, exclusively associated with the fungal kingdom, catalyzes the oxidative decarboxylation of OLKG, forming succinate and carbon dioxide, concomitant with xanthine hydroxylation to uric acid, as shown'in Figure 2.1. I showed that fungus-derived XanA differed from the recombinant form isolated from Escherichia coli both in quaternary structure and identity of posttranslational modifications, but the two enzyme forms exhibited very similar kinetic parameters. In contrast to the phylogenetically restricted XanA enzyme, a molybdopterin cofactor (Moco)-containing enzyme xanthine oxidoreductase or xanthine hydroxylase (1) is present in all kingdoms of life, including fungi, and carries out an analogous reaction, the key step of which is depicted in Figure 3.1. A glutamic acid general base activates the hydroxyl group coordinated to Mo(Vl) (chelated by an enedithiolate moiety of the pterin) and the resulting nucleophile attacks the C—8 position of xanthine with concomitant hydride transfer to Mo=S. The product subsequently dissociates and the resulting Mo(IV)-SH intermediate reoxidizes by sequential electron transfer steps through two [2Fe-2S] clusters and an enzyme bound flavin adenine dinucleotide. The reduced flavin passes on the electrons to NAD (xanthine dehydrogenase) or in some cases to oxygen (xanthine oxidase). Whereas the Moco-containing xanthine hydroxylases have been extensively characterized, little is known of the Fe(II)/0tKG-dependent XanA. 93 FIGURE 3.1: General mechanism of Moco-containing xanthine oxidases // r) AIL/Em. f“ /:/\:V\H o\o u/KO /M\°§o \O_¢o Glu Glu 94 IZ NH The Fe(II)/0tKG hydroxylases encompass a wide range of enzymes with diverse primary substrates (reviewed in (2)). The prolyl, lysyl, and aspartyl(asparaginyl) hydroxylases catalyze posttranslational modifications of proteins, in some cases associated with hypoxic signaling (3). J ij domain-containing proteins have been shown to catalyze methylated-histone demethylation reactions (4). AlkB repairs l-methyl-A or 3-methyl-C lesions in DNA (or RNA) by using an analogous oxidative dealkylation reaction (5, 6). The lipid-metabolizing enzyme phytanoyl-CoA hydroxylase participates in the metabolism of phytanic acid, and deficiency of this enzyme leads to Refsum disease (7). Finally, a wide variety of small molecules are synthesized or decomposed by members of this enzyme family. For example, plants synthesize gibberellins, flavonoids, and some alkaloids by action of these enzymes (8), thymine 7-hydroxylase sequentially hydroxylates the methyl group of free thymine (9), and deA, deA, and SdpA decompose specific phenoxyalkanoic acid herbicides by hydroxylation of the side chain (10, 11). The best-characterized representative of this protein family is taurine/OtKG dioxygenase (TauD), responsible for the decomposition of taurine (2-aminoethanesulfonate) to aminoacetaldehyde and sulfite (12). In particular, a series of transient kinetic studies have demonstrated the interrnediacy of an Fe(IV)-0x0 species in the TauD catalytic mechanism (13-19). In addition, the TauD crystal structure has been solved (20, 21) and aberrant chemistry leading to side chain hydroxylation reactions has been extensiVely defined (22-24). The XanA sequence shows some similarity with the TauD group of dioxygenases (25), including a clear conservation of the Fe(II)- and 95 aKG-binding sites. The studies reported here continue our characterization of the bacteria-derived forms of A. nidulans XanA. I describe its solvent and substrate kinetic isotope effects, define the effects of different divalent metal ions and several OLKG and xanthine analogues, and I identify alternate substrates and inhibitors. Finally, a homology model of XanA was generated and I use it to provide insights into the mode of substrate/inhibitor binding. These studies present the first analysis of isotope effects, interactions with alternative metal ions and NaCl, and behavior of substrate analogues for xanthine/OLKG dioxygenase. 96 EXPERIMENTAL PROCEDURES Purification of Bacterial-Derived XanA. XanA was purified from recombinant E. coli cells as previously described in Chapter 2 (26). Enzyme Assays. Xanthine/OtKG dioxygenase activity was measured at 25 °C by using the following typical assay conditions (total volume of 1 mL): 1 mM OtKG, 40 uM Fe(NH4)2(SO4)2, and 200 uM xanthine in 50 mM MOPS buffer, pH 7.4. Variations of these conditions included use of varied concentrations of these and other additives. The absorbance at 294 nm was monitored to determine the uric acid production (8294 12,200 M" cm") with a correction for loss of the xanthine absorbance at this wavelength (measured 8294 2,000 M" cm") for an overall change in 8294 of 10,200 M" cm". Initial rates were calculated by using linear data collected over a time period where 10% or less of the substrate was consumed. Units of activity (U) were defined as umol min" of uric acid produced and the specific activity (U mg") was measured as umol min" (mg of purified XanA)". Sources and Synthesis of Chemical Analogues of aKG and Xanthine. OLKG, a—ketoadipate, a—ketobutyric acid, pyruvate, phenylpyruvate, 4-hydroxyphenylpyruvate, purine, 6-methylpurine, 2-hydroxypurine, 2-hydroxy-6-methylpurine, hypoxanthine, xanthine, l-methylxanthine, 3-methylxanthine, 7-methylxanthine, 9-methylxanthine, allopurinol, and allantoin were from Sigma-Aldrich. N-oxalylglycine (NOG) was a gift from Dr. Nicolai Burzlaff. 8-Hydroxypurine and 6,8-dihydroxypurine (6,8-DHP) were prepared from the 97 commercially available (Sigma Aldrich) 4,5-diaminopyrimidine and 4,5-diamino-6-hydroxypyrimidine, respectively, following a modified method described previously (27). 2,8-Dihydroxypurine (2,8-DHP) was prepared following a published method (28). Each of these compounds was provided by David J. Lowe and Fabrice Jourdan. 8-2H-Xanthine Preparation. 8-2H-xanthine was prepared by incubating a 1 % (w/v) xanthine solution in szO (99.9 %, Sigma & Aldrich), with 0.3 M NaOzH (> 99 % 2H, Sigma-Aldrich) in a serum vial sealed with a butyl rubber stopper for 20 h in a 100 °C oven. The proton-deuterium exchange was monitored by using NMR spectroscopy to integrate the resonance at 6.9 ppm due to the proton on C-8. The final solution was diluted and neutralized to pH 7.0. The 8-2H-xanthine precipitated from the solution and was dried under vacuum for at least 3 h. Yields were approximately 90 %. The purity of 8-2H-xanthine was monitored by electrospray ionization mass spectrometry, and the conversion was shown to be 98.4 % as a molar ratio. 21120 Solvent Isotope Eflect. In 1 mL 50 mM MOPS, p211 8.0 (p2H values were determined by adding 0.4 to the pH 7.6 electrode reading), 0.5] pg of purified XanA was assayed in 40 uM Fe(NH4)2(SOo)2, 1 mM aKG and 0-200 uM xanthine. All buffers were prepared in szO. These data were compared to assays carried out in the same conditions at pH 7.6. Of note, the E. coli-derived enzyme exhibits optimal activity over a range from pH 7.0 to 8.0. Structural Modeling. (Assisted by Andrew W. Stumpff-Kane and Michael Feig, 98 Department of Biochemistry and Molecular Biology, MSU.) A homology model of XanA was generated on the basis of the structure of the related enzyme TauD (PDB code lOTJ, chain A, and PDB code IOS7, chain A) (21). XanA is 18 % identical to TauD over 280 residues, and has three significant additional regions of 15, 22, and 18 amino acids. Multiple templates for the XanA structure were obtained with PSI-BLAST (29, 30) starting both from the amino acid sequence of XanA and from that of several closely related proteins for which no structure has yet been reported, and with 3D-Jury (31) using the Bioinfo Meta Server (http://bioinfo.pl/Meta). Additional suboptimal alignments for. each template were generated using probA (32) to produce a large pool of possible models. A structural model was constructed from each alignment, with side chains reconstructed using the MMTSB Tool Set (33, 34), and the best model from the entire set of models was selected according to combined energy scores from DFIRE (35), MMGB/SA (36), and RAPDF (3 7), using a correlation-based approach (38) in combination with clustering. The orKG and iron were placed into the active site according to the TauD structure. 99 RESULTS Isotope Eflects. To test for substrate or solvent isotope effects on the overall reaction rate, activity assays were carried out by using 8—2H-xanthine or in 2H20 and compared to control studies with unlabeled xanthine in H20. No significant isotope effect was observed when the reaction was carried out with 8—2H-xanthine (98.4 % enriched with 2H) and compared to non-labeled substrate (data not shown). In contrast, a significant solvent deuterium isotope effect was observed (Figure 3.2), with Vmax reduced by 50 % (72.1 U mg" dropping to 34.0 U mg") while Km was nearly unaffected (45.2 and 48.9 mM, respectively). Eflect of Other Metal Ions and Salt. Metal ions other than Fe(II) were tested and found to be unable to stimulate activity when added to the apoprotein. When various metal ions were added to the assay buffer in concentrations equivalent to that of Fe(II), both Cu(II) and Zn(ll) completely inhibited the xanthine/aKG dioxygenase activity, with partial inhibition observed with Co(Il), Mn(ll), and (much less pronounced) Ni(II) (Figure 3.3). The inactive metal ions are presumed to compete for the Fe(II)-binding site. The activity of XanA was shown to decrease in buffers containing NaCl (Figure 3.4A). Kinetic analyses revealed that 0.5 M NaCl salt increased the Km of ocKG to 0.74 i 0.08 mM (Figure 3.4B), increased the Km of xanthine to 105 i 5.8 uM (Figure 3.4C), and decreased the kca, to 35.4 s'1 and 43.2 s", respectively in the two studies. These findings indicate that the Km of OLKG is significantly affected by ionic strength, while the Km of xanthine and kca, are less affected, and demonstrate that the salt content of fractions 100 FIGURE 3.2: Solvent deuterium isotope effect on XanA activity. The effects of varying the concentration of xanthine on xanthine/orKG dioxygenase activity were examined at 25 °C in 50 mM MOPS buffer, pZH 8.0 (I) (p2H values were determined by adding 0.4 to the pH electrode reading) or pH 7.6 (A) containing 40 uM Fe(ll), 1 mM OLKG, and 0-200 11M xanthine. o ' 5'0 160 130 260 Xanthine (pM) 101 FIGURE 3.3: Divalent cation inhibition of XanA. The effects of several divalent cations (M", at 40 uM concentration) on xanthine/(XKG dioxygenase activity were examined by using the E. coli-derived protein at 25 °C in solutions containing 40 uM Fe(II), 1 mM aKG and 200 uM xanthine in 50 mM MOPS buffer, pH 7.4. Samples included (I) Fe(II) only and Fe(Il) plus (2) Mg(ll), (3) Mn(ll), (4) Co(II), (5) Ni(II), (6) Zn(II), (7) Cu(II). 80 60 1 2 3 4 5 6 7 Fe(II)+M(II) 102 FIGURE 3.4: NaCl inhibition of XanA. (A) The effects of varying NaCl concentrations on xanthine/aKG dioxygenase activity were examined by monitoring the absorbance at 294 nm over time for the E. coli-derived protein at 25 °C in solutions containing 40 11M Fe(II), 1 mM aKG and 200 11M xanthine in 50 mM. MOPS buffer, pH 7.4. The concentrations of NaCl examined were: 0 mM (6); 200 mM (I); 400 mM (A); 600 mM (V); 800 mM (C). Using 0.5 M NaCl, the effects of varying the concentrations of (B) (XKG and (C) xanthine were examined. Data in panels B and C represent initial rates and were fit to the Michaelis-Menten equation. 0.71 45‘ I 0.3! E 0.5: g 30" a 0.4‘ 5 .3 ‘ 15 B < 0.3: 0.2 ‘ v v v v I *7 T V ‘l 0 jjjjjjjjjj o 40 30 120 150 o 1 2 3 4 5 Time (3) alpha-KG (W) 45" 3o, Ulrng .51 c 0 V v f 1 v I V T T u . o 40 00 120 160 200 xanthine (9") 103 recovered during protein isolation must be considered when assaying the enzyme. aK G Analogues. In addition to (xKG, a-ketoadipic acid is a cosubstrate of XanA and results in an activity of 9.2 U (mg protein)" or about 1/10 of that observed with (XKG Kinetic analyses revealed a kca, of 7.6 s" and 3 Km of 0.16 mM for this alternative co-substrate compared to a kca, of 61 s" and Km of 31 11M for OLKG (26). In contrast, pyruvate, (x-ketobutyric acid, phenylpyruvate, and 4-hydroxyphenylpyruvate were not used as co-substrates. NOG, a known inhibitor of several Fe(II)/01KG-dependent dioxygenase family members (39-43), was shown to compete with OLKG and provided a K, of 0.12 11M for inhibition of the enzyme (Figure 3.5). Xanthine Analogues. XanA was shown to be highly specific for xanthine. On the basis of the spectroscopic assay using standard conditions with 12 nM enzyme, no activity was detected when the enzyme was assayed with 80, 100, or 200 11M hypoxanthine, l-methylxanthine, 3-methylxanthine, 7-methylxanthine, 9-methylxanthine, purine, 6-methy1purine, 2-hydroxypurine, 8-hydroxypurine, 2,8-dihydroxypurine, 2-hydroxy-6-methylpurine, allopurinol, allantoin, or adenosine diphosphate. Similarly, significant inhibitory effects were not observed with 100 or 200 11M of any of these compounds (although very modest inhibition was noted with 2,8-DHP). The one xanthine-like compound that does inhibit the enzyme, but does not serve as a substrate, is 6,8-DHP. The kinetic inhibition mechanism of 6,8-DHP will be discussed in more detail in Chapter 4. Structural Model of XanA . XanA was aligned with TauD and a homology model was 104 FIGURE 3.5: NOG inhibition of XanA. (A) The effects of varying concentrations of NOG (0, 0.2, 0.6, 1.2, and 2 uM) on xanthine/ocKG dioxygenase activity were examined by using the E. coli-derived protein at 25 °C in solutions containing 40 1.1M Fe(II), 1 mM aKG and 200 11M xanthine in 50 mM MOPS buffer, pH 7.4. Each set of initial rate data was fit to the Michaelis-Menten equation. (B) A replot of the values of apparent Km divided by apparent Vmax as a function of inhibitor concentration. 1001 no.1 20 / A ’ , 0 vvvvv . vi ~fi - . . o 100 200 300 400 500 600 alpha-Kama) 10 ' 8 l- K 36 ‘ E4 - B 2 I- oI + L l 0 1 2 3 ~061le 105 created by my collaborators using the TauD structure (20, 21) as the template. The overall sequence identity is 18%, however the sequence identity of the most relevant structural regions near the active site is 33%. Moreover, given recent advances in structure prediction, sequence identities as low as 18% identity are commonly sufficient to support comparative modeling with a good match of predicted secondary structure elements (44, 45) as in the case of XanA. Consequently, it is reasonable to assume that our model of the XanA structure captures the overall features correctly, while structural details near the active site are likely represented more accurately. Nevertheless, any predicted model without further experimental validation remains speculative and is subject to some level of uncertainty. As illustrated in Figure 3.6A, the XanA protein is predicted to contain the DSBH - fold comprised of eight B-strands with connecting loops, as is typical of this enzyme family (2, 46). Three loops in the sequence, comprising residues 72-88, 173-190, and 219-231, had no cotmterparts in TauD and were not modeled (indicated by boxes at the appropriate positions in the figure), but these are all distant from the putative active site region. The homology model contains the Fe(II)-binding site (Hisl49, Asp151, and His340) expected from prior sequence alignments (25). The co-substrate (shown in yellow) was positioned into the model so as to chelate F e(II) in a similar fashion as OLKG ' occurs in TauD. The aKG C-5 carboxylate is predicted to form a salt bridge with Arg352 (depicted in red), while Ly5122 (in green) is well positioned to stabilize the C-1 carboxylate of the co-substrate. 106 FIGURE 3.6: Homology model of XanA. A, Ribbon diagram depicting the XanA homology model that was predicted using TauD (PDB code 1087) as a template. The polypeptide chain is depicted in blue to red coloration (N-terminus to C-terminus) with gaps indicated by boxes for the non-modeled loops involving residues 72-88, 173-190, and 219-231. The postulated metal ligands (Hisl49, Asp151, and His340) are illustrated in green, yellow, and red, respectively, to the right of the magenta Fe(II) sphere. Bound aKG (yellow) chelates the metals and is suggested to be stabilized by a salt bridge to the C-5 carboxylate involving Arg352 (red) and a hydrogen bond to the C-1 carboxylate via Lysl22 (green). B, Putative active site pocket derived from the XanA homology model. A closeup view (20 A slab) of the XanA homology model depicting possible positions of residues at the active site pocket, shown in the same orientation as illustrated in Figure 9. OtKG in stick form is shown chelating the metal with its carbon atoms colored yellow. The three side chains that bind the metal are shown in stick form using green (Hisl49), yellow (Asp151), and red (His340) coloration, as in Figure 9. Residues predicted to line the active site pocket are shown in stick form with their carbon atoms in white, while other residues are shown as lines with green carbon atoms. 107 FIGURE 3.6: 108 DISCUSSION In this study I describe several properties of xanthine/OLKG dioxygenase, a novel xanthine-metabolizing enzyme found exclusively in the fungal kingdom (25). Below, I place my findings on isotope effects, interactions with various metal ions, salts, and substrate analogues, into the larger context of other Fe(II)/01KG dioxygenases, and I relate key findings to our homology model of the protein. Isotope Eflects. Although the xanthine C-H bond is broken at 08 during turnover, substitution of the proton at this position by 2H did not lead to a substrate isotope effect. This result demonstrates that C-H cleavage is not the rate-detennining step in the reaction. The deuterated substrate might be useful in future experiments to examine individual steps in the reaction by using transient kinetic approaches, as was elegantly demonstrated with deuterated substrate and stOpped-flow techniques for TauD (15) and, more recently, prolyl 4-hydroxylase (4 7). In contrast to the situation with labeled xanthine, a solvent isotope effect was observed upon substituting H20 with D20 (Figure 3.2). This substitution had. little effect on the Km of xanthine while decreasing Vmax by 40 % compared to the assay in H20. This result suggests that a chemical group possessing an exchangeable proton is important in the rate-detennining step of the overall reaction. Options for the protonatable group include a general base or general acid protein side chain or a metallocenter species such as Fe(III)-OOH or Fe(III)-OH. The finding of a solvent deuterium isotope effect contrasts with the case of TauD, where product release is the slow step in catalysis and no solvent isotope effect is observed (13, 19). 109 Metal Ion and Salt Eflects. Zn(II) and Cu(II) are potent inhibitors of XanA, and several other metals also inhibit the enzyme (Figure 3.3). This situation resembles that known for related enzymes such as TauD where Co(ll) and Ni(II) inhibition has been studied (39), Clavaminate synthase where Co(lI) inhibition was characterized (48), or deA where the Cu(II)-inhibited enzyme was analyzed (49). The inhibitory metal ions are likely to substitute for Fe(II) and utilize the same set of amino acid side chain ligands. The inhibitory effects of NaCl on XanA (Figure 3.4) represent, to my knowledge, the only systematic characterization of any salt effect on an Fe(II)/01KG dioxygenase. The presence of salt leads to a large increase in Km of (XKG, a small increase in Km of xanthine, and a reduction in km. I attribute the Km effect to the ability of salt to interfere with salt bridge formation and other stabilizing interactions between OLKG or xanthine and the protein. Similar salt effects are likely to apply to a wide range of other family members; thus, one must exercise caution in the choice of ionic strength when doing enzyme assays. Co-Substrate and Substrate Specificity. The co-substrate specificity of XanA is somewhat more relaxed than that for the primary substrate, with a—ketoadipic acid (with one extra carbon compared to aKG) also yielding activity. The increase in Km and decrease in km, for the incorrectly-sized analogue is easily rationalized in terms of the XanA homology model where both the C-1 and C—5 carboxylates of aKG are predicted to interact with the protein (with Ly5122 and Arg352, respectively) while also chelating the active site metal ion. Alternative OL-ketoacids are known to support (xKG-dependent 110 activities of several other enzyme family members including TauD, deA, deA, SdpA, and an alkyl sulfatase (10, 12, 50, 51). The aKG homologue NOG is a competitive inhibitor of XanA (Figure 3.5), consistent with its known inhibition of several other representatives of the Fe(Il)/(xKG-dependent hydroxylases. Of interest, the measured K, of NOG for XanA (0.12 11M) is well below the Km of OLKG in this enzyme and much below the reported 290 11M K, for inhibition of TauD (39). More generally, the K, of NOG can vary widely among family members (e.g., it is reported to be l.9-7.0 11M for collagen prolyl 4-hydroxylase (40) and 1.2 mM for an oxygen-sensing asparaginyl hydroxylase (43)), presumably due to distinct interactions with the active site protein side chains in the target enzymes. XanA proved to be exquisitely specific to its primary substrate. For example, allopurinol (a known substrate and inhibitor of xanthine oxidase (52)), l-methylxanthine and 2-hydroxy-6-methylpurine (alternative substrates of the Moco-containing enzyme (53, 54), and several other purine-type compounds were neither substrates nor inhibitors of XanA. Of the compounds tested, only 6,8-DHP bound tightly to the enzyme (more date . and discussion of this inhibitor are provided in Chapter 4). For comparison, other members of the Fe(II)/01KG-dependent dioxygenases range widely in their substrate specificities. The prolyl hydroxylases involved in the hypoxic response appear to be highly specific for recognizing a single prolyl residue in the HlFlor protein (55, 56). By contrast, deA utilizes a wide range of phenoxyacetic acids (10) and a yeast orKG/sulfonate dioxygenase metabolizes a diverse array of sulfonates (5 7). ll] Postulated Substrate-Binding Mode. Additional structural or mutagenesis studies are required to characterize the mode of substrate binding to XanA; however, our homology model (Figure 3.6) allows me to identify potential key active site residues (Figure 3.6B). the model depicts a pocket adjacent to the metallocenter and lined by a series of putative active site residues (Gln99, ProlOO, GlnlOl, IlellO, Thr120, Lysl22, Glul37, Ala152, leu154, Gln356, and Asn358) that are generally well conserved in sequences of XanA orthologues (Figure 3.7). Pr0100, GlnlOl, Gln356, and Asn358 are universally conserved in the XanA sequences. Gln99 counterparts also are present often, but Val occupies this position in some representatives. Lysl22, which is suggested to stabilize the OLKG C-l carboxylate and bind substrate, either is retained or conservatively replaced by Arg or Thr. In some fungi, various residues (Lys, Gln, Arg, Asn, and Glu) replace Thr120, but all of these are ble to function in hydrogen bonding. Similarly, Asp, Gln, and Ser, all capable of similar hydrogen bond interactions, replace Glul37 in other XanA homologues. Among the hydrophobic residues predicted to surround the active site, Ile110 is replaced by Val or Phe, Ala152 is retained or replaced by Ser, Leu154 is strictly conserved, and Leu251 (not shown) is retained or replaced by other large side chains in other homologues. Aromatic groups, often involved in tt-tt stacking interactions with nucleic acids and known to occur in xanthine hydroxylase (58, 59) and uric acid oxidase (60), do not appear to be important for binding xanthine in xanthine/orKG dioxygenase. These predictions set the stage for future chemical modification, mutagenesis, and structural efforts to test these interactions. 112 FIGURE 3.7: Comparison of sequences from XanA orthologues in selected fungi. Black shading represents identical sequences and gray shading indicates conservative replacements. chryaosporfum cinereual neotorman graminearum crassa nidulans maydie po 9 alblcans cinereuaz a n m c v z m a are chryaosporium Cinereusl pombe albicans c1nareus2 onnererone U a 2 H B Q chryaoapor1um KEL KKSILHPHLFTIP PQVOLIGN' cinereusl 7 ' ' KKSILHPDLKTI PO' VQ neo oto gramlnearum crassa nidulans maydls pombe albicans Y' 1 . KSO cinereusz .2 S A' Yb H'-----EK13RSS'LSfll§l b 00md>2100v chrysosporlum graminearum crassa nidulans maydls pombe albicans cinereusz nOmCDZ’HOn'U chrysosporium s 'K I-hlnnGTGDh= cinereuel 17 OTYDHCTGDFIFI neotorman R" gramlnearum asea nidulans maydls albtcans ctnereusZ O O m d > z m n 0 v 0 chrysoeporiun clnereual neotorman graminearum ‘g a u m albicans cinereusz 9999>=wonv D E E 113 FIGURE 3.7': OOMQVZ’UOO'U OOMC’Z‘OO'U OOMC’Z'UOO"! .chrysosporium .clnereusl .neotorman .graminearum .crassa .nldulans .maydls .pombe .albtcans .ctnereusz .chrysosporlum .ctnereusl neotorman graminearum . crassa .ntdulans mayors albtcans c1nereus2 .chryaosporlum .clnoreuel .neotorman .gramrnearum .craaaa .nrdulans .maydls .albrcans .clnereusz 432 380 387 413 371 -—---—--------------—-—----------- LQVAVRRVVV ........................ ll4 REFERENCES (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) Hille, R. (2005) Molybdenum-containing hydroxylases. Arch. Biochem. Biophys. 433, 107-116. Hausinger, R. P. (2004) Fe(II)/01-ketoglutarate-dependent hydroxylases and related enzymes. Crit. Rev. Biochem. Mol. Biol. 39, 21-68. Schofield, C. J., and Ratcliffe, P. J. (2004) Oxygen sensing by hydroxylases. Nat. Rev. Molec. Cell Biol. 5, 343-354. Tsukada, Y.-l., Fang, J., Erdjument-Bromage, H., Warren, M. E., Borchers, C. H., Tempst, P., and Zhang, Y. (2006) Histone demethylation by a family of ijC domain-containing proteins. Nature 439, 811-816. Trewick, S. C., Henshaw, T. F., Hausinger, R. P., Lindahl, T., and Sedgwick, B. (2002) Oxidative demethylation by Escherichia coli AlkB directly reverts DNA base damage. Nature 419, 174-178. Aas, P. A., Otterlei, M., Falnes, P. O., Vagbe, C. B., Skorpen, F., Akbari, M., Sundheim, O., Bjoras, M., Slupphaug, G, Seeberg, E., and Krokan, H. E. (2003) Human and bacterial oxidative demethylases repair alkylation damage in both RNA and DNA. Nature 421, 859-863. Mukherji, M., Chien, W., Kershaw, N. J., Clifton, I. J., Schofield, C. J., Wierzbicki, A. S., and Lloyd, M. D. (2001) Structure-function analysis of phytanoyl-CoA 2-hydroxylase mutations causing Refsum's disease. Human Molec. Genet. 10, 1971-1982. Prescott, A. G, and John, P. (1996) Dioxygenases: molecular structure and role in plant metabolism. Annu. Rev. Plant Physiol. Plant Molec. Bio]. 47, 245-271. Thomburg, L. D., Lai, M.-T., Wishnok, J. S., and Stubbe, J. (1993) A non-heme iron protein with heme tendencies: an investigation of the substrate specificity of thymine hydroxylase. Biochemistry 32, 14023-14033. Fukumori, F., and Hausinger, R. P. (1993) Purification and characterization of 2,4-dichlorophenoxyacetate/or-ketoglutarate dioxygenase. J. Biol. Chem. 268, 24311-24317. 115 (11) (12) (13) (14) (15) (16) (17) (18) (19) Miiller, T. A., Zavodszky, M. 1., Feig, M., Kuhn, L. A., and Hausinger, R. P. (2006) Structural basis for the enantiospecificities of R- and S-phenoxypropionate/(x-ketoglutarate dioxygenases. Prot. Science 15, 1356-1368. Eichhom, E., van der Ploeg, J. R., Kertesz, M. A., and Leisinger, T. (1997) Characterization of a—ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J. Biol. Chem. 2 72, 23031-23036. Ryle, M. J., Padmakumar, R., and Hausinger, R. P. (1999) Stopped-flow kinetic analysis of Escherichia coli taurine/Ot-ketoglutarate dioxygenase: interactions with a—ketoglutarate, taurine, and oxygen. Biochemistry 38, 15278-15286. Price, J. C., Barr, E. W., Tirupati, B., Bollinger, J. M., Jr., and Krebs, C. (2003) The first direct characterization of a high-valent iron intermediate in the reaction of an a—ketoglutarate-dependent dioxygenase: a high-spin Fe(IV) complex in taurine/Ot-ketoglutarate dioxygenase (TauD) from Escherichia coli. J. Biol. Chem. 42, 7497-7508. Price, J. C., Barr, E. W., Glass, T. E., Krebs, C., and Bollinger, J. M., Jr. (2003) Evidence for hydrogen abstraction from C1 of taurine by the high-spin Fe(IV) intermediate detected during oxygen activation by taurineza-ketoglutarate dioxygenase (TauD). J. Am. Chem. Soc. 125, 13008-13009. Proshlyakov, D. A., Henshaw, T. F., Monterosso, G R., Ryle, M. J., and Hausinger, R. P. (2004) Direct detection of oxygen intermediates in the non-heme Fe enzyme taurine/Ot-ketoglutarate dioxygenase. J. Am. Chem. Soc. 126, 1022-1023. Riggs-Gelasco, P. J., Price, J. C., Guyer, R. B., Brehm, J. H., Barr, E. W., Bollinger, J. M., Jr., and Krebs, C. (2004) EXAF S spectroscopic evidence for an Fe=O unit in the Fe(IV) intermediate observed during oxygen activation by taurineza-ketoglutarate dioxygenase. J. Am. Chem. Soc. 126, 8108-8109. Price, J. C., Barr, E. W., Hoffart, L. M., Krebs, C., and Bollinger, J. M., Jr. (2005) Kinetic dissection of the catalytic mechanism of taurineza-ketoglutarate dioxygenase (TauD) from Escherichia coli. Biochemistry 44, 8138-8147. Grzyska, P. K., Ryle, M. J., Monterosso, G R., Liu, J., Ballou, D. P., and Hausinger, R. P. (2005) Steady-state and transient kinetic analyses of taurine/or-ketoglutarate dioxygenase: effects of oxygen concentration, alternative sulfonates, and active site variants on the Fe(IV) intermediate. Biochemistry 44, 116 (20) (21) (22) (23) (24) (25) (26) (27) 3845-3855. Elkins, J. M., Ryle, M. J., Clifton, I. J., Dunning Hotopp, J. C., Lloyd, J. S., Burzlaff, N. 1., Baldwin, J. E., Hausinger, R. P., and Roach, P. L. (2002) X-ray crystal structure of Escherichia coli taurine/Ot-ketoglutarate dioxygenase complexed to ferrous iron and substrates. Biochemistry 41, 5185-5192. O'Brien, J. R., Schuller, D. J., Yang, V. S., Dillard, B. D., and Lanzilotta, W. N. (2003) Substrate-induced conformational changes in Escherichia coli taurine/Ot-ketoglutarate dioxygenase and insight into the oligomeric structure. Biochemistry 42, 5547-5554. Ryle, M. J., Liu, A., Muthukumaran, R. B., Ho, R. Y. N., Koehntop, K. D., McCracken, J., Que, L., Jr., and Hausinger, R; P. (2003) 02- and or-ketoglutarate-dependent tyrosyl radical formation in TauD, an a—keto acid-dependent non-heme iron dioxygenase. Biochemistry 42, 1854-1862. Ryle, M. J., Koehntop, K. D., Liu, A., Que, L., Jr., and Hausinger, R. P. (2003) Interconversion of two oxidized forms of TauD, a non-heme iron hydroxylase: evidence for bicarbonate binding. Proc. Natl. Acad. Sci. USA 100, 3790-3795. Koehntop, K. D., Marimanikkuppam, S., Ryle, M. J., Hausinger, R. P., and Que, L., Jr. (2006) Self-hydroxylation of taurine/a—ketoglutarate dioxygenase. Evidence for more than one oxygen activation mechanism. J. Biol. Inorg. Chem. 11, 63-72. Cultrone, A., Scazzocchio, C., Rochet, M., Montero-Moran, G, Drevet, C., and Fernandez-Martin, R. (2005) Convergent evolution of hydroxylation mechanisms in the fungal kingdoms: molybdenum cofactor-independent hydroxylation of xanthine via or-ketoglutarate dependent dioxygenase. Molec. Microbiol. 5 7, 276-290. Montero-Moran, G M., Li, M., Rendon-Huerta, E., Jourdan, F., Lowe, D. J., Stumpff—Kane, A. W., Feig, M., Scazzocchio, C., and Hausinger, R. P. (2007) Purification and characterization of the Fe(II)- and on-ketoglutarate-dependent xanthine hydroxylase from Aspergillus nidulans. Biochemistry 46, 5293-5 304. Albert, A., and Brown, D. J. (1954) Pruine studies. Part 1. Stability to acid and alkali. Solubility. Ionization. Comparison with pteridines. J. Chem. Soc, 2060-2071. 117 (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) (38) Dornow, a., and Hinz, E. (1958) Syntheses of nitrogen-containing heterocycles. XVIII. ortho-condensations of heterocyclic o-aminocarboxylic acid derivatives. Chem. Ber. 91, 1834-1840. Altschul, S. F., Madden, T. L., Schaffer, A. A., Zhang, J. H., Zhang, Z., Miller, W., and Lipman, D. J. (1997) Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucl. Acids Res. 25, 3389-3402. Schaffer, A. A., Aravind, L., Madden, T. L., Shavirin, S., Sponge, J. L., Wolf, Y. I., Koonin, E. V., and Altschul, S. F. (2001) Improving the accuracy of PSI-BLAST protein database searches with composition-based statistics and other refinements. Nucleic Acids Research 29, 2994-3005. Ginalski, K., Elofsson, A., Fisher, D., and Rychlewski, L. (2003) 3D-Jury: a simple approach to improve protein structure predictions. Bioinformatics 19, 1015-1018. Muckstein, U., Holfacker, I. L., and Stadler, P. F. (2002) Stochastic pairwise alignments. Bioinformatics 18 (Suppl. 2), 8153-8160. Feig, M., Rotkiewicz, P., Kolinski, A., Skolnick, J., and Brooks, C. L., 111. (2000) Accurate reconstruction of all-atom protein representations from side-chain-based low-resolution models. Proteins 41, 86-97. Feig, M., Karanicolas, J., and Brooks, C. L., 111. (2004) MMTSB tool set: enhanced sampling and multiscale modeling methods for applications in structural biology. J. Molec. Graphics Modeling 22, 377-395. Zhang, C., Liu, S., Zhu, Q. Q., and Zhou, Y. Q. (2005) A knowledge-based energy function for protein-ligand, protein-protein, and protein-DNA complexes. J. Med. Chem. 48, 2325-2335. Feig, M., and Brooks, C. L., 111. (2002) Evaluating CASP4 predictions with physical energy functions. Proteins: Struct. Funct. Genet. 49, 232-245. Samudrala, R., and Moult, J. (1998) An all-atom distance-dependent conditional probability discriminatory function for protein structure prediction. J. Mo]. Bio]. 275, 895—916. Stumpff-Kane, A. W., and Feig, M. (2006) A correlation-based method for the enhancement of scoring functions on funnel-shaped energy landscapes. Proteins: ll8 (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) Struct., F unct. Bioinform. 63, 155-164. Kalliri, E., Grzyska, P. K., and Hausinger, R. P. (2005) Kinetic and spectroscopic investigation of Co”, Ni”, and N-oxalylglycine inhibition of the Fe(II)/(x-ketoglutarate dioxygenase, TauD. Biochem. Biophys. Res. Commun. 338, 191-197. Baader, E., Tschank, G, Baringhaus, K. H., Burghard, H., and Gunzler, V. (1994) Inhibition of prolyl 4-hydroxylase by oxalyl amino acid derivatives in vitro, in isolated microsomes and in embryonic chicken tissues. Biochem. J. 300, 525-530. Cunliffe, C. J., Franklin, T. J., Hales, N. J ., and Hill, G B. (1992) Novel inhibitors of prolyl 4—hydroxylase. 3. Inhibition by the substrate analogue N-oxaloglycine and its derivatives. J. Med. Chem. 35, 2652-2658. Chan, D. A., Sutphin, P. D., Denko, N. C., and Giaccia, A. J. (2002) Role of prolyl hydroxylation in oncogenically stabilized hypoxia-inducible factor-101. J. Biol. Chem. 277, 40112-40117. McDonough, M. A., McNeill, L. A., Tilliet, M., Papmichael, C. A., Chen, Q.-Y., Banerji, B., Hewitson, K. S., and Schofield, C. J. (2005) Selective inhibition of factor inhibiting hypoxia-inducible factor. J. Am. Chem. Soc. 127, 7680-7681. Tramontano, A., and Morea, V. (2003) Assessment of homology-based predictions in CASP5. Proteins: Struct. Funct. Genet. 53, 352-368. Kryshtafovych, A., Venclovas, C., Fidelis, K., and Moult, J. (2005) Progress over the first decade of CASP experiments. Proteins: Struct. Funct. Genet. S7, 225-236. Clifion, I. J., McDonough, M. A., Ehrismann, D., Kershaw, N. J., Granatino, N., and Schofield, C. J. (2006) Structural studies on 2-oxoglutarate oxygenases and related double-stranded B-helix fold protein. .1. Inorg. Biochem. 100, 644-669. Hoffart, L. M., Barr, E. W., Guyer, R. B., Bollinger, J. M., Jr., and Krebs, C. (2006) Direct spectroscopic detection of a C-H-cleaving high-spin Fe(IV) complex in a prolyl-4-hydroxylase. Proc. Nat]. Acad. Sci. USA 103, 14738-14743. Busby, R. W., Chang, M. D.-T., Busby, R. C., Wimp, J., and Townsend, C. A. (1995) Expression and purification of two isozymes of Clavaminate synthase and initial characterization of the iron binding site. General error analysis in 119 (49) (50) (51) (52) (53) (54) (55) (56) (57) polymerase chain reaction amplification. J. Biol. Chem. 270, 4262-4269. Hegg, E. L., Whiting, A. K., Saari, R. E., McCracken, J., Hausinger, R. P., and Que, L., Jr. (1999) Herbicide-degrading or-keto acid-dependent enzyme deA: metal coordination environment and mechanistic insights. Biochemistry 38, 16714-16726. Miiller, T. A., Fleischmann, T., van der Meer, J. R., and Kohler, H.-P. E. (2006) Purification and characterization of two enantioselective (x-ketoglutarate-dependent dioxygenases, deA and SdpA, from Sphingomonas herbicidovorans MH. Applied and Environmental Microbiology 72, 4853-4861. Kahnert, A., and Kertesz, M. A. (2000) Characterization of a sulfur-regulated oxygenative alkylsulfatase from Pseudomonas putida S-3l3. J. Biol. Chem. 275, 31661-31667. Elion, G B., Kovensky, A., and Hitchings, G. H. (1966) Metabolic studies of allopurinol, an inhibitor of xanthine oxidase. Biochem. Pharmacol. 15, 863-880. Edmondson, D., Ballou, D. P., Vn Heuvelen, A., Palmer, G, and Massey, V. (1973) Kinetic studies on the substrate reduction of xanthine oxidase. J. Biol. Chem. 248, 6135-6144. McWhirter, R. B., and Hille, R. (1991) The reductive half-reaction of xanthine oxidase. Identification of spectral intermediates in the hydroxylation of 2-hydroxy-6-methylpurine. J. Biol. Chem. 266, 23724-23731. Ivan, M., Kondo, K., Yang, H., Kim, W., Valiando, J., Ohh, M., Salic, A., Asara, J. M., Lane, W. S., and Kaelin, W. G, Jr. (2001) HIFa targeted for VHL-mediated destruction by proline hydroxylation: implications for 02 sensing. Science 292, 464-467. Jaakkola, P., Mole, D. R., Tian, Y.-M., Wilson, M. I., Gielbert, J., Gaskell, S. J., von Kriegsheim, A., Hebestreit, H. F., Mukherji, M., Schofield, C. J ., Maxwell, P. H., Pugh, C. W., and Ratcliffe, P. J. (2001) Targeting of HIF-01 to the von Hippel-Lindau ubiquitylation complex by Oz—regulated prolyl hydroxylation. Science 292, 468-472. Hogan, D. A., Auchtung, T. A., and Hausinger, R. P. (1999) Cloning and characterization of a sulfonate/(x-ketoglutarate dioxygenase from Saccharomyces cerevisiae. J. Bacteriol. 181, 5876-5879. . 120 (58) Okamoto, K., Matsumoto, K., Hille, R., Eger, B. T., Pai, E. F., and Nishino, T. (2004) The crystal structure of xanthine oxidoreductase during catalysis: implications for reaction mechanism and enzyme inhibition. Proc. Natl. Acad. Sci. USA 101, 7931-7936. (59) Truglio, J. J ., Theis, K., Leimkiihler, S., Rappa, R., Rajagopalan, K. V., and Kisker, (60) C. (2002) Crystal structure of the active and alloxanthine-inhibited forms of xanthine dehydrogenase from Rhodobacter capsulatus. Structure 10, 115-125. Retailleau, P., Colloc'h, N., Vivarés, D., Bonneté, F., Castro, E., El Hajji, M., Momon, J.-P., Monard, G, and Prangé, T. (2004) Complexed and ligand-free high-resolution structures of urate oxidase (on) from Aspergillus flavus: a reassignment of the active-site binding mode. Acta Crystallogr. D60, 453-462. 121 CHAPTER 4 Characterization of Active Site Variants of Xanthine Hydroxylase from Aspergillus nidulans Tina Muller constructed the XanA variants and carried out oxygen consumption studies. 122 ABSTRACT Xanthine/a-ketoglutarate (OtKG) dioxygenase (XanA) is a non-heme mono- nuclear Fe(ll) enzyme that decarboxylates orKG to succinate and C02 while catalyzing the hydroxylation of Xanthine to generate uric acid. In the absence of a XanA crystal structure, a homology model was used to target several putative active site residues for mutagenesis. Wild-type XanA and ten enzyme variants were purified from recombinant Escherichia coli cells and extensively characterized. From analysis of the quenching of the endogenous fluorescence of XanA, the H149A, D151A, and H340A mutants displayed significant increases in K, of Fe(II) and the K122A variant exhibited a large increase in Kd of OLKG, consistent with the proposed roles of the corresponding residues in Fe(II)- and OtKG-binding. The H149A and DlSlA variants were inactive whereas the H340A variant exhibited 0.13 U mg" (0.17 % of the wild-type enzyme). The N358A variant exhibited the next largest change in xanthine-related kinetics with a 12-fold larger Km and 2-fold decrease in km, compared to wild-type XanA, pointing to a key role of Asn358 in catalysis. The E137A and D138A variants demonstrated enhanced activity with 9-methy1xanthine, a poor substrate of the enzyme, consistent with Glul37 and Aspl 38 being proximal to N-9 of substrate. The Q356A and N358A variants had significantly increased K?” over control protein for 6,8-dihydroxypurine, identified as a slow-binding competitive inhibitor of XanA, suggesting that Gln356 and Asn358 hydrogen bond with the C-6 hydroxyl group of substrate. In contrast, the K?” decreased for the E137A and D138A proteins, consistent with repulsion between these carboxylates and the 123 deprotonated C-8 hydroxyl group suspected to bind Fe(II). Support for Cys357 residing at the active site was obtained by using thiol-specific reagents that inactivated wild-type enzyme with partial protection by substrate, whereas the C357A variant was resistant to these reagents. In the absence of substrates, the Q101A, Q356A, and C357A variants showed elevated ferroxidase activity, indicating increased oxygen reactivity of their metallocenters and/or enhanced Fe(II) access to those sites. These results were combined into a model depicting Fe(II) and substrate interaction with the XanA active site and provide insight into the specificity of the enzyme and selected aspects of its reactivity. 124 INTRODUCTION Xanthine dehydrogenases and xanthine oxidases are molybdopterin cofactor (Moco)-containing enzymes that transform xanthine into uric acid (1). These enzymes are conserved throughout living Organisms, including archaea, bacteria, fungi, plants, and metazoans. Surprisingly, mutants of Aspergillus nidulans with known defects in xanthine dehydrogenase activity (i.e., with mutations affecting the structural gene th, the cnx genes for Moco synthesis, or hrB for sulfuration of Moco) were found to retain the ability to grow on xanthine as sole nitrogen source (2). This capacity arises from an alternative xanthine-degrading activity encoded by the xanA gene found only in selected fungi (3). Recombinant Hisé-tagged XanA protein was purified from both A. nidulans mycelia and Escherichia coli cells, extensively characterized, and shown to be an Fe(II)- and a-ketoglutarate (OLKG)-dependent hydroxylase that catalyzes the reaction shown in Figure 2.1 (4). On the basis of its sequence similarity to taurine/aKG dioxygenase or TauD (5), a well-studied and crystallographically characterized member of the Fe(II)/0tKG dioxygenase family (6, 7, 8), a homology model was constructed for XanA (4). The overall sequence identity of XanA and TauD is only 18 % over 280 residues and three loops in XanA could not be modeled; however, the relevant structural regions near the active sites are 33 % identical and it is reasonable to assume that the model captures the overall features correctly. The XanA model (Figure 4.1) predicts that (i) Fe(II) binds to the Hisl49, Asp151, and His340‘ residues of the protein, (ii) (xKG chelates the metal by using its C-l carboxylate and C-2 keto group, with its C-l carboxylate further stabilized by 125 FIGURE 4.1: Depiction of the putative active site pocket of XanA revealed from a homology model. Hisl49, Asp151, and His340 side chains (blue) and aKG (yellow) coordinate the Fe(H)(orange sphere). Seven residues lining the putative active site and capable of participating in hydrogen-bonding interactions are colored green (Gln99, Ly5122, Glul37, Asp138, Gln356, Cys357, Asn358), with another shown in purple (GlnlOl) at the active site entrance. :1 (3513357 .. Asn358 ,“ \ C ‘\ . , Asp151 l’ ‘ ~ --- Gln99 \ Hrs34o Gln356 Lys122 A W "b , .9 H1949 :. p *1 4’ 1 :3 F Gln101 -. arKG . l . \\~_ 1 Glu137 Asp138 126 interaction with Ly3122 and its C-5 carboxylate forming a salt bridge with Arg352, and (iii) xanthine binds in an active site pocket lined with potential hydrogen bond donors or acceptors (Gln99, GlnlOl, Glul37, Asp138, Ly3122, Gln356, Cys357, and Asn358) whereas aromatic residues are not important for binding substrate. In this study, the three residues thought to bind Fe(II) and each of the potential hydrogen-bondin g residues at the XanA active site was replaced by Ala, a residue that is incapable of hydrogen bonding. Fluorescence quenching methods were used to assess the K, values of Fe and OtKG for the mutant enzymes. The kinetic properties of the variants were analyzed with xanthine, the altemate substrate 9-methylxanthine, and 6,8-dihydroxypurine (6,8-DHP), which was shown to be a slow-binding competitive inhibitor of XanA. Chemical modification studies were carried out with thiol-specific reagents for the wild-type enzyme and the C357A variant. Finally, the reactivity of each metallocenter with oxygen was examined in the absence of substrate. On the basis of these results, we identify critical residues that participate in binding of Fe(II), orKG, and the primary substrate, and we obtain insight into the enzyme reactivity. 127 EXPERIMENTAL PROCEDURES Site-Directed Mutagenesis, Protein Overproduction, and Enzyme Purification. The Q99A, Q101A, K122A, E137A, D138A, H149A, D151A, H340A, Q356A, C357A, and N358A variants of XanA were created by mutagenesis of xanA (encoding a Hisb-tagged version of XanA, termed wild-type XanA for convenience) in a modified version of vector pThioHisC (4) using the Quickchange II site-directed mutagenesis kit (Stratagene). Each mutation was confirmed by sequence analysis (Davis Sequencing, Davis, CA). Plasmids were transformed into XLlBlue E. coli cells (Stratagene) and the variant proteins were overproduced during cell growth in Luria Base Broth (Difco) and induced by using isopropyl-B-D-thiogalactopyranoside as described previously (4). Cultures were harvested by centrifugation at 8,000 g for 10 min at 4 °C, the cells were disrupted by use of a French pressure cell, membranes and insoluble materials were removed by ultracentrifugation (45 min at 100,000 g), and the wild-type and variant forms of XanA were purified by using Ni-loaded nitrilotriacetic acid (NTA)-bound resin, as previously reported (4). Fluorescence Quenching Analyses of Fe(II) and 6K0 Binding. Fluorescence measurements were made on a luminescence spectrometer, model LS-50B (Perkin Elmer Limited, UK). The temperature of the cells was maintained at 25 °C. Fluorescence measurements were carried out at an excitation wavelength of 280 nm (10 nm band-width) with emission monitored from 300 to 400 nm (5 nm bandwidth). Samples were prepared in 50 mM Tris buffer, pH 8.0, and Fe(II) or aKG was added to 35 11M or 350 11M, respectively. The data were fit to equation 1, where [L] is the 128 ligand (Fe(II) or aKG) concentration. Afluorescence = Afluorescencemax [L]/(Kd + [L]) (1) Enzyme Assays. Enzymatic assays of XanA and its variants were carried out at 25 °C by using the following typical assay conditions (total volume of 1 mL): 1 mM aKG, 40 11M Fe(NH4)2(SO4)2, and 200 11M xanthine in 50 mM MOPS buffer, pH 7.4 (4). For kinetic analyses, the concentrations of orKG or xanthine were varied while holding the concentrations of the other components constant. The absorbance at 294 nm was monitored to determine the uric acid production (overall change in 8294 of 10,200 M" cm"). Units of activity (U) were defined as umol min'l of uric acid produced and the specific activity (U mg") was measured as umol min" (mg of purified XanA)" as described before (4). Analogous studies were carried out with 9-methylxanthine (Sigma-Aldrich), but in this case the absorbance at 291 nm was used to calculate the production of 9-methyluric acid (overall change in 829, of 6,800 M'l cm"). Trace activity was detected with l-methylxanthine (Sigma-Aldrich), but complete kinetic studies could not be performed. In addition to the spectrophotometric measure of xanthine conversion to uric acid (or the use of substituted xanthines to form substituted uric acids), the various forms of the enzyme were analyzed for ferroxidase activity (reduction of oxygen by using excess ferrous ions as reductant) by use of a Clark-type oxygen electrode. These assays were carried out in air-saturated MOPS medium (pH 7.4) at 25 °C. Less than 12% of the initial levels of 02 were consumed in these assays, and the data were fit to equation 2. [02], is the 02 concentration at time t, [02],, is the initial 02 concentration, vb 129 is the background rate of 02 reduction under the given conditions, v0 is the ferroxidase initial velocity, and kapp is the apparent first-order rate constant for the transition from V0 to vb. [02]: = [0216 — vb! —(vo - vb)[1-6XP(-kappt)l/kapp (2) As a complement to the electrode assay for 02 reduction, the production of hydrogen peroxide was quantified by using a spectrophotometric assay. Timed aliquots of reaction mixtures were added to assay solutions containing 100 mM potassium phosphate (pH 5.0), 8.7 mM 2,2’-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) (ABTS), and approximately 0.003 units of horseradish peroxidase (Sigma-Aldrich). The oxidation of ABTS was monitored at 420 nm and compared to a standard curve generated with hydrogen peroxide. A final method to assay the activity of the enzyme focused on quantification of the OtKG consumed during the reaction. Aliquots (250 11L) of reaction mixtures were . incubated for selected time periods, quenched by addition of 1 mL 0.5 mg/ml o-phenylenediamine (OPDA, Sigma-Aldrich) stock solution (dissolved in 1 M phosphoric acid, pH 2, containing 0.25% (v/v) B-mercaptoethanol), and the samples were heated for 3 min at 100 °C. The absorbance at 333 nm was monitored to determine aKG consumption by comparison to standard curves. Kinetic Analysis of 6,8-DHP Inhibition. Spectrophotometric progress curves (containing 90 or more data points, typically at 15 5 intervals) were initiated by adding XanA to solutions containing several fixed concentrations of substrate and selected concentrations of 6,8-DHP (graciously provided by D. J. Lowe and F. 130 Jourdan). The data were analyzed according to equation 3 by using a nonlinear regression program to give the individual parameters for each progress curve: A (absorbance at 294 nm), A0 (for baseline correction), v, (initial velocity), v, (steady-state velocity), and kobs (apparent first-order rate constant for the transition from v, to v,). A = A0 + Vst + (Vi' Vs)[]' exp('kObst)]/ kobs (3) 131 RESULTS Site-directed variants of recombinant A. nidulans XanA, an Fe(II)- and OtKG-dependent hydroxylase of xanthine, were investigated to further characterize the protein features important to catalysis. In the absence of a crystal structure for XanA, we relied on apreviously reported homology model of XanA (4) that identifies the likely triad of metal-binding residues and predicts the active site pocket is lined by eight residues capable of hydrogen bonding interactions with substrate. To test this hypothetical‘model,‘ Ala was used to replace each of these residues and the enzymatic properties of the resulting variant proteins were characterized. In particular, the binding of Fe(lI) and OtKG to the XanA variants was assessed by fluorescence quenching methods and the kinetic properties were studied using xanthine, the alternative substrate '9-methylxanthine, and 6,8-DHP, a known XanA inhibitor (4). In addition, the behavior of wild-type enzyme and C357A variant were compared using thiol-specific chemical modification reagents. Finally, the ferroxidase activities of the active mutant proteins Were examined in the absence of primary substrate. Fe(II) and otKG Binding to XanA and its Variants. Eleven XanA variants (Q99A, Q101A, K122A, E137A, D138A, H149A, D151A, H340A, Q356A, C357A, and N358A) were produced in E. coli XLlBlue cells as Hisé-tagged fusion proteins, along with the wild-type protein. Seven variant proteins were purified to homogeneity from cell extracts by Ni-NTA chromatography whereas three (H149A, D151A, and H340A) were enriched by this chromatographic step (Figure 4.2), and Q99A XanA failed to bind to Ni-loaded NTA resin. We interpret the Q99A XanA results to indicate protein 132 FIGURE 4.2: Denaturing polyacrylamide gel electrophoretic analysis of XanA variants purified from E. coli. Lane I, markers; lane 2, Q101A (1.5 pg) ; lane 3, K122A (4.2 pg); lane 4, E137A (1.0 pg); lane 5, D138A (1.2 pg); lane 6, Q356A (0.5 pg); lane 7, C357A (0.3 pg); lane 8, N358A (6.4 pg); lane 9, H149A (0.17 pg); lane 10, D151 (0.8 pg); lane 11, H340A (0.13 pg) variant protein. Stacking and running gels contain 5 % and 12 % acrylamide, respectively. kDa l 2 3 4 5 6 7 8 97.4 66.2 45.0 31.0 21.5 l 4.4 133 misfolding during overexpression, and this variant was not further investigated. The binding of OLKG and xanthine to XanA was previously shown to result in quenching of the endogenous fluorescence of the protein (4), thus allowing estimation of the K., of the substrates. The apparent quenching of fluorescence by xanthine is complicated by the significant absorption of the excitation wavelength by the substrate, precluding the confident estimation of the K, of xanthine. In contrast, the fluorescence quenching approach was extended to study the K, of both OLKG and Fe(II) (neither of which absorbs at 280 nm) for the mutant proteins (Figure 4.3 and 4.4, and Table 4.1). The three variants involving the putative ligands to the metallocenter displayed significantly increases in the Kd of Fe(II), consistent with their metal binding assignments. Similarly, the K122A variant exhibited an increased Kd of (xKG, consistent with its suggested function in stabilizing the binding of the cosubstrate. Of interest, the fluorescence of the H149A variant was not quenched by addition of OLKG perhaps suggesting that it may an important role in protein folding. Kinetic Comparison of XanA and its Variants with Xanthine. Although all variants appeared to be equally overproduced, the Q99A, H149A, D151A, and H340A variants displayed no activity in cell extracts. When purified, H340A XanA exhibited trace activity (0.13 U mg") corresponding to 0.17 % of that associated with wild-type enzyme. These four variants were not further analyzed kinetically. The kinetic parameters for XanA and the seven active mutant proteins are listed in Table 4.2. All of these variants exhibit perturbations in their Km (xanthine), K.,, (OtKG), and kca, compared to the control enzyme, consistent with the mutant proteins having 134 FIGURE 4.3: Fluorescence spectroscopic analysis of XanA interaction with Fe(II). The changes in the endogenous fluorescence (excitation at 280 nm and emission at 335 nm) due to quenching by Fe(II) were examined in 50 mM Tris buffer, pH. 8.0, solutions containing (A) 0.12 pM XanA, (B) 0.0] pM H149A, (C) 0.03 pM D151A, (D) 0.04 pM H340A, (E) 0.05 pM Q101A, (F) 0.12 pM K122A, (G) 0.05 pM E137A, (H) 0.07 pM D138A, (I) 0.02 pM Q356A, (J) 0.02 pM C357A, and (K) 0.07 pM N358A at 25 °C. Fe(II) concentrations ranged from 0-35 pM. soo, , A 400- is... 02001 2 l ‘2 100- 0 .............. 0 5 10 15 20 25 30 35 Fe(pM) ' Fe(ull) 1401 1604 . C D 120. . 1 l 0 120+ £1”: E 3 304 a ‘ Q ‘ 08° 8 30‘ :I s . c . IT. 40“ < ‘0‘ 4 20- . 1 0 V I jfi'fiij V I V I V I V I o yyyyyyy y—v 1 v v v ' 0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35 F001") Fe(pM) ' E 100. 500- 1 d F 00. 8 ‘ .4004 504 g i g ( gm- 2 I g‘o‘ 0200- E l E , <20j <1oo. o l V I V I V I '1fi' I V I o vvvvvvvvvvvvvv 0 5 10 15 20 25 30 35 0 5 1° 15 2° 25 30 35 Fe (1m) Fe M.) 135 IHGURE43: 250- 1 O A ‘ «II 33 ..I O O A A Fluorescence A 13 A v'VUVVVVVj’fI" 5 101520253035 Fowl) ' 5'10'115121072‘5'31113'5 F001") K f ' v ' v ' w r f vvvvv 5 10 15 20 F901") 136 AFluorIseenee 3. .3. 3- .8. J .3 9 A Fluomcenge 8 8 - l A """ 1'5'20'2'5'30' Fe(v") '5110'1'5f20'2'5'50'35 F301”) FIGURE 4.4: Fluorescence spectroscopic analysis of XanA interaction with aKG The changes in the endogenous fluorescence due to quenching by OLKG in 50 mM Tris, pH. 8.0, buffer were examined in solutions containing (A) 0.12 pM XanA, (B) 0.03 pM D151A, (C) 0.04 pM H340A, (D) 0.50 pM Q101A, (E) 0.12 pM K122A, (F) 0.13 pM E137A, (G) 0.18 pM D138A, (H) 0.05 pM Q356A, (I) 0.10 pM C357A and (J) 0.11 pM N358A, at 25 °C. . so. 150 A ‘ B 50- 0 . 0 §1Wd 540. u I l 30- s E . 250. |€20. u. < ‘ <1 10- 1 o r I v u v y w I o i l V I V I V I 0 100 200 300 400 0 100 200 300 400 aKetuM) aKG(pM) . 00 . D 6°. C 70 a“? °° - §4o. £50 . O 530. 340 2 20* $30 Ill- It 3 < . 1:20 10- <10 0‘-.-.-,#. am 0 100 200 300 400 0 100 200 300 400 200. E 1204 F 1 l 3150- C 8 §’°: g1“ goo: 17. E401 <501 <1 1 , 20- 0 V I ' I fi'i I 'i l o if I ‘ W ' U ' I 0 100 200 300 400 0 100 200 300 400 1.1400111) aKthM) 137 FIGURE 4.4: 301 A fluorescence O 4“ 104 160 ' 260 ' 360 ' 400 aKG(pM) 160'260'2410w 460 aKG (MI) 138 A Fluorescence S D D j l A A 160 ' 260 a 330 aKG (M) 160 ' zoo ' 360 aKG(pM) 400 TABLE 4.]: Determination of Kd of Fe(II) and OLKG for XanA and Selected Variantsa Mutant Kd of Fe(II) (uM) Kd of aKG (11M) XanA 85:03 115 :5 Q101A 21.5:12 108:12 K122A 9.3 : 0.6 236 : 29 E137A 11.7: 0.6 146: 13 D138A 12.3 :1.0 136: 14 H149A 36.8 :1.3 NDb D151A 28.2: 1.4 169: 16 H340A 30.7 :14 175 :16 Q356A 15.9 : 0.6 108 : 6 C357A 13.5:09 124: 14 N358A 10.1 : 1.0 168 :9 a Estimated on the basis of quenching of the endogenous fluorescence with excitation at 280 nm and emission at 335 nm at 25 °C. The protein concentrations: 0.12 M for XanA, 0.01 M for H149A, 0.03 M for D151A, 0.04 uM for H340A, 0.05 uM for Q101A in the Fe(II) assay and 0.5 uM in the aKG assay, 0.12 11M for K122A, 0.05 uM for E137A with Fe(II) and 0.13 uM in the (xKG assay, 0.07 11M for D138A in the Fe(II) assay and 0.18 11M in the (XKG assay, 0.02 M for Q356A in the Fe(II) assay and 0.05 11M in the OtKG assay, 0.02 M for C357A in the Fe(II) assay and 0.10 M in the aKG. assay, and 0.07 uM for N358A in the Fe(II) assay and 0.1] uM in the aKG assay. b ND, fluorescence quenching was not detected. 139 TABLE 4.2: Substrate, Cosubstrate, and Substrate Analog Kinetic Parameters of XanA and Selected Variantsa XanA Xanthineb (JtKGc 9-Methylxanthined sample Km kcat kcat/ Km Km kcat kcat/ Km Km kcat kcat/ Km 01M) 6") (.844 s") (11M)(s")01M" s") (mM) (s") (mM' s") Wild-Type 45 :4 72:2 1.6 31 :2 67:1 2.16 0.40: 3.8: 9.5 0.07 0.3 0101A 122 : 9 36 :1 0.30 59: 6 21 :1 0.36 0.78: 0.66: 0.85 0.07 0.03 K122A 88:5 60:] 0.68 32:2 42:1 1.31 0.88: 1.9: 2.16 0.07 0.1 E137A 71:5 40:1 0.56 41:8 31:1 0.76 1.09: 5.9: 5.41 0.17 0.5 D138A 69 : 6 34 : 2 0.49 53 : 7 25 :1 0.47 0.58 : 4.25 : 7.33 0.09 0.24 Q356A 57 : 5 35 : 1 0.61 42 : 3 28 : 1 0.67 0.68 : 1.4 : 2.05 0.07 0.1 C357A 58 : 5 32 :1 0.55 47 : 4 19 :1 0.40 0.55 : 4.5 : 8.18 0.05 0.2 N358A 554:24 38:2 0.07 48:8 17:] 0.34 1.0: 1.26: 1.26 0.4 0.35 a Except for the compound being analyzed, the assay solutions contained 40 11M Fe(II), 1 mM aKG, and 200 11M xanthine in 50 mM MOPS buffer, pH 7.4, and were maintained at 25 °C. b Xanthine concentrations ranged from 0-250 11M, except for the N358A variant, which ranged from 0-400 pM. c aKG concentrations ranged from 0-500 pM. d 9-methy1xanthine concentrations ranged from 0-1.0 mM. 140 modifications involving active site residues. Alternative Substrates for XanA and its Variants. In agreement with prior results (4), XanA was shown to be incapable of using 3-methylxanthine, 6-methylpurine, or 7-methylxanthine as the primary substrate. In contrast, trace levels of activity were detected for l-methylxanthine (about 10'4 of that observed for xanthine) when using highly concentrated enzyme solutions, and low levels of activity were observed with 9-methylxanthine. The absorbance spectra of the products arising from l-methylxanthine and 9-methylxanthine exhibited features at 289 nm and 291 nm (Figure 4.5), consistent with the formation of l-methyluric acid and 9-methy1uric acid (9), respectively. Generation of the product from 9-methylxanthine was coupled to 02 consumption, as shown by use of an oxygen electrode, and to the decomposition of 0tKG, as analyzed by the OPDA assay (data not shown). The kinetic parameters associated with XanA utilization of 9-methy1xanthine for wild-type enzyme and the seven variants are shown in Table 2. For wild-type XanA, the Km of 9-methy1xanthine is 8.9-fold larger than that for xanthine while the kw is approximately 5 % of that for the true substrate. Comparison of the data obtained for the control enzyme with those for the mutant proteins provides insight into potential protein interactions involving the 9-methyl group. For example the E137A variant exhibited a 1.6-fold increase in km for 9—methylxanthine versus the wild-type enzyme. Although not statistically significant, 3 small increase (1.1-fold) in km also was noted for the D138A variant. These results suggest that substituting Ala for the carboxylate at position 137, and perhaps also at position 138, partially compensates for the 141 FIGURE 4.5: UV-visible spectra of l-methylxanthine, 9-methylxanthine, and their products. (A) Spectrum of 9-methylxanthine in assay solution (solid) and the sample after 30 min of reaction (dotted) at 25 0C. The assay solution included 40 11M Fe(II), 1 mM aKG, 3.6 uM XanA, and 200 uM 9-methylxanthine in 50 mM MOPS, pH 7.4. (B) Spectrum of l-methylxanthine in assay solution (solid) and the sample after 100 min of reaction at 25 °C (dotted), using conditions as above. (C) The spectrum of the product arising from the enzymatic transformation of 1-methy1xanthine was calculated from B (dotted line) by measuring the absorption at 290 nm (0.747 absorbance units), determining £290 for l-methylxanthine (1,497 M'I cm") and l-methylun'c acid (10,273 M'l cm'l), using these results to estimate that the concentrations of l-methylxanthine (149 0M) and l-methylun'c acid (51 11M) which totaled to 200 11M, and subtracting the spectrum of 149 1.1M l-methylxanthine from the spectrum of the mixture. This was compared to the spectrum of authentic 1-methyluric acid (solid line) .3 om ‘ 142 presence of a methyl group at N-9 of xanthine by allowing for more productive active site binding. 6,8-DHP Inhibition. 6,8-DHP was previously shown to inhibit wild-type XanA (4), but its mechanism of inhibition Was not examined. By contrast only very modest inhibition was noted with 2,8-DHP, and 8-hydroxypurine did not inhibit the enzyme (4). These results indicated that XanA interactions with the C-6 oxygen were most important for inhibitor binding. To examine the inhibitory effect of 6,8-DHP on XanA, progress curves of uric acid production were obtained under different inhibitor and substrate concentrations. The curves obtained for a constant substrate concentration with varied levels of inhibitor reveal a time-dependent inhibition of XanA by 6,8-DHP (Figure 4.6A) whereby the initial rate constant, vi, decreases to a steady-state rate constant, vs, with an apparent. first-order rate constant, kobs. The values of V., vs, and [robs were characterized according to equation 2 by using a variety of assay conditions. For varied inhibitor concentrations, kub, exhibited a hyperbolic dependence on the concentration of 6,8-DHP (Figure 4.6B), consistent with a slow-binding inhibitory mechanism involving the reversible binding of inhibitor followed by a conformational change to create a tightly bound state (10). To distinguish the type of slow-binding inhibition mechanism, progress curves were obtained for a constant concentration of 6,8-DHP in the presence of varied substrate concentration (Figure 4.6C). A replot of kobs versus substrate concentration reveals a negative slope (Figure 4.6‘D) and confirms the mechanism as being competitive (11), as depicted in Figure 4.7. Thus, 143 FIGURE 4.6: Interaction of XanA with 6,8-DHP. (A) Progress curves for XanA inhibition by varied concentrations of 6,8-DHP. 2.4 nM XanA was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 11M Fe(II), 1 mM aKG, 200 11M xanthine, and 0 mM (n), 0.02 mM (.), 0.04 mM (A), 0.06 mM (V), 0.1 mM (0) or 0.12 mM (4) 6,8-DHP. The production of uric acid was monitored at 294 nm and the progress curves were fit to equation 1. (B) Dependence of kobs on the concentration of 6,8-DHP. The curve was fit to equation 2. (C) XanA progress curves were obtained as above, but in the presence of 0.1 mM 6,8-DHP and 50 11M (4), 100 1.1M (V), 150 11M (A), 200 11M (0) or 250 11M (I) xanthine. The progress curves were fit to equation I. (D) Dependence of kobs on the concentration of xanthine. -" 0.00251 0 O b 1. x 1 '1 1 ‘1 I 1 I I A A 1 :1" ,x z ‘7 0.00201 ‘ t 0 v § 0.0015{ B .3 0.0010' 400 600 30° 100° 120° o'msno002004000000 030 03? Time (sec) 5 B-DHP (mM) 0 e e e d 4 - 1 A n - 1 - Abs 294 nm 2.01110"- 2.71110“- ..2‘ 2.01110": . 3 2.5x10"« 0 .31 1: 2.41110 1 D 0 4 2 2.31110 . 2.2x1o"- ' “Twang,“ “°° 5'0 ' 160 ' 150 ' 200 ' 250 ' Xanthine (0M) IOD‘A 144 FIGURE 4.7: General mechanism for slow-binding inhibition. 145 6,8-DHP competes with substrate to form an initial enzyme inhibitor complex (E-I) that slowly transforms to a more stable complex (E-I‘). For such a mechanism, kobs varies with the inhibitor concentration [I] according to equation 4 (10) , where K, = k4/k3. This equation simplifies to equation 5 when one defines K?” = Ki(1 + [S]/Km). kobs = k6 + k5[I]K./(1 + [SI/Km + [ll/K1) (4) kobs = k6+ kill/(1’9“pp + [1]) (5) Using the data from Figure 4.68 for wild-type XanA, k5 = 2.0 X 10'3 8", k6 = 8.1 x 104's", and Kiapp = 12.6 11M. k5 is larger than k6 suggesting the rate of formation of B]. complex is much faster than its dissociation, which is in agreement with slow-binding inhibition mechanism. The slow-binding inhibition kinetic properties of the active variant forms of XanA with 6,8-DHP were analyzed in a similar manner (see Figures 4.8-4.14) and the k5, k6 and K-f'pp are compared to the control enzyme in Table 4.3. Compared with the K?” of 12.6 11M calculated for wild-type XanA, the values for the E137A and D138A proteins were smaller, ranging from 3.0-3.6 11M. This decrease in K?” is consistent with elimination of negative interactions between 6,8-DHP (possibly involving the deprotonated C-8 hydroxyl group that likely binds to Fe(II)) and the Glul37 or Aspl38 carboxylates; i.e., the charge repulsion is eliminated in the two mutant proteins containing the small and hydrophobic Ala side chains. Kf’pp increased significantly for both the Q356A and N358A proteins, suggesting that stabilizing interactions provided by the Gln356 and Asn358 side chains are abolished in these mutants. In. particular, this result is consistent with hydrogen bonding between the C-6 hydroxyl group (or the corresponding keto tautomer) of 6,8-DHP and Gln356 and 146 FIGURE 4.8: Interaction of Q101A with 6,8-DHP. (A) Progress curves for Q101A XanA inhibition by varied concentrations of 6,8-DHP. 1.9 uM Q101A was added to 50 mM MOPS buffer, pH 7.4, at 25 0C containing 40 11M Fe(II), 1 mM aKG 200 11M xanthine, and 0 mM (I), 0.01 mM (0), 0.02 mM (A), 0.04 mM (V), 0.06 mM (4) or 0.08 mM (b) 6,8-DHP. The production of uric acid was monitored at 294 nm and the progress curves were fit to equation 1. (B) Dependence of kobs on the concentration of 6,8-DHP. The curve was fit to equation 2. (C) XanA progress curves were obtained as above, but in the presence of 0.1 mM 6,8-DHP and 50 1.1M (I), 100 11M (0), 150 0M (A), 200 11M (V) or 250 11M (4) xanthine. The progress curves were fit to equation 1. (D) Dependence of kobs on the concentration of xanthine. 10« 0.004- . ‘ ‘ - 0.04 0.0031 5 ~ . v 0.0. ‘7 a ‘ 3 0002‘ g 0.40 g 1 ‘ -K 0.W1< < 0.2. A 1 B 00.-.-.-r. 0.000..-.-.r-- o 400 000 1200 0.00 0.02 0.04 0.00 0.00 “”9 (39°) 0,11-an (mM) 0.0« . . 0.0000- 05 . 0.0000- 5 .- . 3 . 4‘ 0.0000. 3 . I a 1 N 30.“: n 1 2 3 0.0040. . 0.0030. D ‘12'00' 00'100‘100'200'200' Time (sec) Xanthine (pM) 147 FIGURE 4.9: Interaction of K122A with 6,8-DHP. (A) 6.1 uM K122A was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 11M Fe(II), 1 mM OLKG, 200 11M xanthine, and 0 mM (I), 0.01 mM (0), 0.02 mM (A), 0.04 mM (V), 0.06 mM (4), 0.08 mM (D) or 0.1 mM (0) 6,8-DHP. Panels B-D are the same as for Figure 83. 1.0- 0.31 1 0,51 0,41 Abs 294nm 0.2‘ 7 A '0'400‘060'12100v '0.00'0.04'0.00'0.'12 W" 189°) 6,8-DHP (mM) Abs 294 nm C 1 D o '400'000'12'00' 50*150'150'2fifi250' Time (sec) Xanthine (pM) 148 FIGURE 4.10: Interaction of E137A with 6,8-DHP. (A) 0.27 uM E137A was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 1.1M Fe(II), 1 mM aKG, 200 11M xanthine, and 0 mM (I), 0.005 mM (0), 0.01 mM (A), 0.02 mM (V), 0.04 mM (4), 0.06 mM (b) or 0.08 mM (0) 6,8-DHP. 0.54 11M E137A was used for plot (C) and (D). Panels B-D are the same as for Figure S3. 6.31 0.51 E 0.4. 3 0.3: N . 3 0.2. < 0.1: 0.0 fi.vA.-.r 0.000.:.-.-r~ffi 0 400 000 1200 0.00 0.02 004 0.00 0.00 Time (sec) 6,8-DHP (mM) . 0.020- 00. J ' 0.010- 2 ~ 1 F I ' '10 0.0124 g V a 30000 .n .3 . . < 0.0044 D ' - 0.0.-.-.fi.- .-.-.f.fi.. 0 400 000 1200 00 100 150 200 200 Time (sec) Xanthine (an) 149 FIGURE 4.11: Interaction of D138A with 6,8-DHP. (A) 0.83 11M D138A was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 11M Fe(II), 1 mM OtKG, 200 11M xanthine, and 0 mM (I), 0.005 mM (0), 0.01 mM (A), 0.02 mM (V), 0.04 mM (4), 0.06 mM (D) or 0.08 mM (0') 6,8-DHP. 1.6 11M D138A was used for plot (C) and (D). Panels B-D are the same as for Figure S3. 0.0- :_. 0.003- E 0.0< 1: . ,7: . 3 0,4. g 0.002: N 10 1 .3 02 '3 < . .1: 0.001 0.0 . - A - - f ' ' OlomI ' I r U V V v V ‘ ° ‘°° '°° 12°° 0.00 0.02 0.04 0.00 0.00 71"“ ““1 0.0-011p (mM) 0.00301 0.0. ' S E 0.00324 0.4- c 3 3 1 N ‘ N 00 0.0020 .n 0.21 3 < , C < 1 D ' I 0.0: va v - w - 0.0024 1a.r.-.-+. 0 400 000 1200 so 100 150 200 250 Time (sec) Xanthine (till) 150 FIGURE 4.12: Interaction of Q356A with 6,8-DHP. (A) 53.5 nM Q356A was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 1.1M Fe(II), 1 mM aKG, 200 1.1M xanthine, and 0 mM (I), 0.1 mM (0), 0.2 mM (A), 0.3 mM (V), 0.4 mM (4) or 0.5 mM (b) 6,8-DHP. Panels B-D are the same as for Figure S3. 0.0- A g ‘7 3’, 0.41 3 N ‘ 3 3 0.21 g “ A 0.0 v v v 1 . I - l T I V I j I V I fl I v 0 400 000 1200 0.0 0.1 0.2 0.3 0.4 0.5 Time (sec) 6,8-DHP (mM) 0-3‘ 0.0070. 0.00054 ' 0.0 . E p 0.00004 . C In 1 v 0.4« v 0.0055- G a . " ‘5 0.0050- - _ 3 0.2 4‘ . < C 0.0045: D . 0.0..fifi...- 0.0040.-.-.-.-,- 0 400 000 1200 50 100 150 200 250 Xanthine (all) Xanthine (0'4) 151 FIGURE 4.13: Interaction of C357A with 6,8—DHP. (A) 2.3 11M C357A was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 11M Fe(II), 1 mM OtKG, 200 11M xanthine, and 0 mM (I), 0.02 mM (0), 0.06 mM (A), 0.10 mM (V), 0.14 mM (4) or 0.18 mM (b) 6,8-DHP. Panels B-D are the same as for Figure S3. E c v a: N ( B _ .-.-.. °-°°°‘.-.-.-.+.-. 0 400 000 1200 0.00 0.04 0.00 0.12 0.10 0.20 Time (sec) 3.8-0"? ("1") 0.0040 E0.0 0.0030: c ,1" v 0.44 'g, 0.0032 ' 0 v N a ' £03 30.0020 ' .1: . D . 0.0fifi—f.rfiv 0.0024 W. ‘2‘” 50 100 150 200 250 Time (39°) Xanthine (01111) 152 FIGURE 4.14: Interaction of N358A with 6,8-DHP. (A) 1.4 uM N358A was added to 50 mM MOPS buffer, pH 7.4, at 25 °C containing 40 uM Fe(II), 1 mM (XKG, 200 uM xanthine, and 0 mM (I), 0.02 mM (0), 0.04 mM (A), 0.10 mM (V), 0.15 mM (4), 0.20 mM (D) or 0.30 mM (0) 6,8-DHP. 2.3 uM N358A was used for plot (C) and (D). Panels B-D are the same as for Figure S3. 0.51 0.00324 0.4- 0.0030. E 1 A J c '7 00020 V0.31 . u ‘ OD “—- v 4 N “03 3 0.0020‘ n o 4 <01 4‘ 0.0024. °'° ' ' - T - . - 0.0022 - . - i f 1 ° ‘°° °°° 12°° 0.0 0.1 0.2 0.3 Time (sec) 6,8-DHP (mM) 0.000 1.2- i I ‘ 0.m7- . 1,0. ‘ . A ‘ E 0.0. ,7 0.006 ‘ ‘ am; 3 0.0: a . 4 °‘ .0 0.4. 00.0042 , 1 * I D ‘M‘ C 0.0034 ' . 0.0-,-,+1_ 1-.er-.... 0 400 000 1200 50 100 150 200 250 Time (sec) Xanthine 0,1”) 153 TABLE 4.3: Inhibition of Wild-Type XanA and its Variants by 6,8-DHPa XanA sample 1‘5 (5") hi§0 Kiapp (HM) “manme QHHA K122A E137A D138A Q3 56A C357A N358A (ZOiQDx103 (l7iODx103 (17:0nxnfi (18i00x103 (sciomx103 (58ionx103 (lliQDx103 (94irmxio‘ (&li0&x104 (54irnxlo‘ (20:0nxio‘ (25irnx104 (lliLDx104 (L2i0nx103 (Isionxiot (23ionx103 12.6 i 2.4 9.8:t 1.7 lO.4:tl.l 3.] i0.4 3.6 :t 0.4 lOZiIZ 16.6 d: 2.0 81i39 a Kinetic analyses were carried out as described in the text, with the kinetic parameters defined according to equation 5. 154 Asn358. For most of the mutants, k5 is larger than k6; however, N358A exhibited the opposite result (k5 = 9.4 x 10'4 3", k6 = 2.3 x 10'3 s"). This finding is consistent with the N358A enzyme lacking an important stabilizing hydrogen bond(s) to the C-6 hydroxyl group of 6,8-DHP, perhaps involved in the k5 transition and formation of the E-I.complex (Figure 4.5). Identification of a Reactive T hi0! at the XanA Active Site. The homology model of XanA depicts Cys357 at the active site (Figure 4.1), a prediction that was directly tested by chemical modification studies. As illustrated in Figure 4.15A, the incubation of XanA with various concentrations of 5,5’-dithiobis (2-nitrobenzoic acid) (DTNB or Ellman’s reagent, specific for reacting with thiol groups) led to concentration- dependent, first-order losses of activity. Significantly, the added presence of xanthine provided some protection to the enzyme (A symbols in Figure 4.15A), consistent with a reaction between DTNB and a Cys residue at the active site. In contrast, OLKG had much less of a protective effect when examined alone ([3 symbols) or with xanthine (data not shown). Inactivation studies also were carried out with iodoacetamide (Figure 4.15B), a less specific reagent to thiol groups than DTNB. Higher concentrations of iodoacetamide were required to obtain XanA inactivation rates compared to DTNB. Either xanthine or (xKG provided some protection against enzyme inactivation by this reagent, and the combined presence of xanthine and aKG provided increased protection against iodoacetamide. These results obtained with iodoacetamide agree well with those using DTNB and are consistent with the presence of a reactive Cys residue near the xanthine-binding site. When the DTNB 155 FIGURE 4.15: Inactivation of XanA by DTNB. (A) XanA (51 pg ml") was incubated on ice with O (O), 2.5 uM (I), 125 uM (A), 750 11M (0) or 1.25 mM (0) DTNB in 50 mM MOPS buffer, pH 7.4, for the indicated times, then diluted lOO-fold into buffer containing 40 11M Fe(II)and 1 mM aKG After blanking the spectrophotometer, xanthine was added to 200 pM. Additional samples containing 750 uM DTNB were examined with added 0.2 mM xanthine (--A--) or 1 mM aKG (--D--). The combination of added xanthine plus OLKG was equivalent to added xanthine alone. (B) XanA inactivation by iodoacetamide. The enzyme (51 pg ml") was incubated on ice with O (O), 20 mM (A), 50 mM (I), 60 mM (0) and 100 mM (x) iodoacetamide in 50 mM MOPS buffer, pH 7.4, for the indicated times, then diluted lOO-fold into buffer containing 40 uM Fe(II)and 1 mM aKG After blanking the spectrophotometer, xanthine was added to 200 pM. Additional samples containing 60 mM iodoacetamide were examined with added 0.2 mM xanthine (-—+--) or 0.2 mM xanthine plus 1 mM OLKG (--D--). The inclusion of OtKG alone was equivalent to the case of added xanthine. (C) C357A XanA (51 ug ml") was incubated on ice with 0 (O), 15 uM (I), 0.75 mM (A) and 1.25 uM (0) DTNB in 50 mM MOPS buffer, pH 7.4, for the indicated times, and diluted and assayed as above. (D) Time—dependent inactivation of wild-type XanA (I) and the C357A variant (O). Time (min) Time (min) 0 5 10 15 20 25 0 5 10 15 20 25 . M {R . ' I E 43.02 ,, A O E 0.0121 g’ o I A " ‘ ' 5 .0“ _ . . g 0.000- _ 32 I: . 3: '0-03 f C 0 12 0.00M . - .000 - g ‘ 2 . o 5 0.000. .01 - 03‘032'03'03003113274 156 inactivation studies were repeated with the C357A variant enzyme, the reagent yielded very little inactivation compared to the control enzyme (Figure 4.15C and D). This result confirms the positioning of Cys357 at the XanA active site. During the course of the above studies, XanA was found to be labile when examined at dilute concentrations (e.g., 12 nM or 0.51 pg mL"), such as those used in routine enzyme assays, whereas the concentrated XanA samples were quite stable during storage. For example when highly diluted XanA was incubated at 25 °C in 50 mM MOPS buffer (pH 7.4), the activity decreased by ~70 % in about 6 min (Figure 4.16). Inclusion of 40 pM Fe(NH4)3(SO4)2 during this incubation stabilized the protein such that ~55 % activity was retained over this time period. Addition of both 40 pM Fe(NH4)2(SO4)2 and 1 mM aKG further stabilized the protein, with ~80 % activity remaining after 6 min. HPLC measurements showed no detectable change of OLKG concentrations during these incubations. The approximately first-order loss of activity in the presence of 40 pM Fe(NH4)2(SO4)2 and 1 mM aKG was not affected by inclusion of 80 pM ascorbate (data not shown); however, the protein was found to be more stable in the absence of oxygen. The oxygen-dependent instability of XanA apoprotein hinted at the reactivity of a cysteine thiol in the protein. Consistent with this hypothesis, the activity of the Cys357 protein was quite stable when examined at dilute (12 nM) concentrations (Figure 4.16). This result confirms that metal-independent oxidation of Cys357 is associated with XanA inactivation; although the mechanism of this inactivation process remains unclear. 157 FIGURE 4.16: Time-dependent loss of XanA activity for highly diluted enzyme. Wild-type XanA was diluted to 0.5] pg ml'1 in aerobic buffer, 50 mM MOPS (pH 7.4), without additives (V), with 40 pM Fe(II) (A), and with 40 pM Fe(II) plus 1 mM aKG (0), or was diluted into anaerobic buffer (I). For comparison, the C357A variant (0.51 pg ml") was examined in aerobic buffer (0). At the indicated time points, the missing components were added and assays (25 °C) were initiated by adding 200 pM xanthine. i 4 6 ii i 1‘? Time (min) 158 Oxygen Reactivity ananA in the Absence of Primary Substrate. The addition of XanA to a buffered solution containing 40 pM Fe(II) led to a slow rate of 0; consumption, similar to the background rate of consumption that occurred in the absence of enzyme (data not shoWn). For the Q101A, Q356A, and C357A mutant proteins, and to a lesser extent the E137A variant, a distinct pattern was observed with 15-20 pM 02 being consumed with a much greater initial rate constant, denoted v0 in Table 4.4. This rapid phase of oxygen consumption also was observed for enzyme samples added to solutions containing Fe(II) plus (XKG (Figure 4.17). The presence of cosubstrate often led to an increase in V.,, which is especially noticeable for the K122A and N358A variants (Table 4.4). Control experiments carried out with Q101A and E137A forms of XanA confirmed that the amount of oxygen consumed in this rapid process correlated with the concentration of added Fe(II), so that approximately twice the amount of O; was consumed when using 100 pM Fe(II). The use of higher concentrations of the metal ion in the assay allowed for a more reliable calculation of v0 for the wild-type enzyme (Table 4.4) even though the background rate of oxygen consumption also was increased. We attribute the rapid phase of oxygen consumption in the above studies to enzyme-catalyzed ferroxidase activity in which the exogenous metal ions are used to provide electrons for the reduction of O; to hydrogen peroxide. Support for this activity was obtained by using the ABTS assay for quantifying peroxide production. To illustrate, the Q101A variant stimulated the production of hydrogen peroxide significantly faster than the wild-type protein, which generated H203 more rapidly 159 FIGURE 4.17: Oxygen consumption studies of XanA and its variants. An oxygen electrode was used to monitor 0; consumption at 25 °C in 50 mM MOPS buffer, pH 7.4, containing 1 mM aKG and 40 pM Fe(II) (circles) or 100 pM Fe(II) (triangle), and 1 pM enzyme (wild-type, dark blue; Q101A, cyan; D138A, purple; C357A, green). Enzyme- free control samples (black symbols, no lines) were analyzed for comparison. Oz-concentratlon (nM) 220 time (sec) 160 TABLE 4.4: Oxygen Reactivity of XanA and Selected Variants in the Absence of Primary Substratea Mutant Fe(II) only Fe(II) + OLKG v0 (11M s") vo (11M 8“) . 0.030 d: 0.002 W - . i . ild type 0 023 0 0001 (0.136 i 0.002) 0.46 i 0.01 .3 :1: . 0101A 0 8 0 008 (0.240 d: 0.002) K122A 0.05] i 0.001 0.143 i 0.001 0.192 :1: 0.018 3 . i . El 7A 0 17 O 03 (0.264 i 0.004) D138A 0.061 i 0.003 0.097 3: 0.002 Q356A 0.39 at 0.01 0.48 :t 0.03 C357A 0.70 i 0.03 0.73 :t 0.01 N358A 0.064 i 0.001 0.159 i 0.002 8 These representative data were obtained by using an oxygen electrode to measure the effects of adding enzyme samples (final concentrations of 1 pM) to solutions containing 50 mM MOPS buffer, pH 7.4, and 40 pM Fe (or 100 pM Fe for the values shown in parentheses), or the same plus 1 mM OLKG. The data shown for each sample are derived from replicate measurements obtained on the same day. The immediate response (V.,) is attributed to ferroxidase activity with electrons from excess Fe(II) used to reduce oxygen. 16] than the non-protein controls (Figure 4.18) when assayed in the presence of 120 pM Fe(II) and 1 mM aKG,. The rates of H202 production correlated with the rates of oxygen consumption for the mutant examined. The H2O2 produced in these reactions accounted for 48-63 % of the expected levels of this product (with the low stoichiometry likely due to the known Fe(III) inhibition of peroxidase). The increased levels of ferroxidase activity observed in several of the mutants compared to the wild-type enzyme could reasonably relate to increased accessibility of exogenous Fe(II) to the enzyme metallocenter or to altered binding of (XKG resulting in enhanced metal center reactivity. 162 FIGURE 4.18: Hydrogen peroxide production by Q101A and wild—type XanA. The wild-type (O) and mutant (I) XanA samples (1 pM) were incubated with 120 pM Fe(II) and 1 mM (xKG in 50 mM MOPS buffer, pH 7.4, at room temperature. At selected time intervals, aliquots Were mixed with ABTS assay solution (pH 5.0) containing horseradish peroxidase and the absorbance was monitored at 420 nm. A non-protein control sample (A) was examined for comparison. 2.5- 1 . I I I 2.0- . ° . O A ‘ O A 2 1.5. ‘ :3. A U ‘ . N 10. O N ' A I 0.5- 0.0 - - 1 - . - 1 - . O OH} ..I O ..5 0| N O N 0' Time (min) 163 DISCUSSION To identify active site residues and better define the reactivity of XanA, a recently identified Fe(II)/aKG-dependent xanthine hydroxylase (4), we extensively characterized a suite of enzyme Variants (Q99A, Q101A, K122A, E137A, D138A, H 149A, D151A, H340A, Q3 56A, C357A, and N358A variants). The residues targeted for mutagenesis, suspected metal ligands or potentially capable of hydrogen-bonding interactions with substrate, were selected on the basis of a homology model of the enzyme (4) that depicts them binding Fe(II) lining the active site pocket (Figure 4.1). Significantly, these residues are well conserved in sequences of XanA orthologues (4). For example, Gln101, Hisl49, Asp151, His340, Gln356, Cys357, and Asn358 are universally conserved in the available XanA sequences. Gln99 often is present, but Val occupies this position in some representatives. Lysl22, which was suggested to stabilize interactions with the C-1 carboxylate of OLKG, either is- retained or conservatively replaced by Arg or Thr. Although Asp, Gln, and Ser replace Glul37 in other XanA orthologues, these replacement residues are capable of similar hydrogen bond interactions. More flexibility exists for Asp138 which is replaced by Glu, Gln, His, and Lys in other XanA sequences. Aromatic residues are not predicted to lie near the active site, reducing the possibility of 1H: stacking interactions with the base, though other hydrophobic residues are suggested to occur in this region. In addition to these various side chains, the substrate might also interact with unidentified backbone amide and carbonyl groups. Model ofXanthine Binding Based on Kinetic Analyses of XanA and its Variants. I64 The results from our extensive kinetic analyses of wild type and variant XanA enzymes with substrates and inhibitors were combined along with determination of the Kd of Fe(II) and OLKG along with the ferroxidase activities into a model for xanthine binding to XanA, as depicted in Figure 4.19 (with substrate entering the active site from the right and some hydrogen bonds likely to arise from interactions with backbone amide groups). In this figure, the Fe(II) is shown bound by the Hisl49, Asp151, and His340 side chains, as is almost certain from sequence comparisons to other related enzymes (4). Mutations involving these residues abolished or greatly diminished activity and led to an increase in the Kd of Fe(II). Furthermore, OLKG is illustrated as chelating the metal, in agreement with other members of this enzyme family, with stabilization provided by the appropriately positioned Arg352 (supported by sequence comparisons) and Ly5122. Consistent with this role for Ly5122, the K122A variant displayed an increase in the Kd of OLKG The primary substrate is shown binding to the enzyme active site via a constellation of hydrogen bonding interactions (some of which cannot be identified from our analyses), so that the disruption of any one hydrogen bond will lead to only modest effects on the kinetics. The Km and kw parameters measured by steady-state kinetics for xanthine utilization by the wild type and variant enzymes cannot be used to directly infer which residues bind the substrate; however, this baseline information allows for such inferences to be made when combined with the data obtained from studies using 9-methylxanthine and 6,8-DHP. In addition, the ferroxidase measurements of the variants provide added insight about putative active site residues. The suspected positioning and role of each 165 FIGURE 4.19: Proposed xanthine binding on the active site. CYS357 , . . H ASH358‘ ' \ O\~ 9 ASP151 i. Gln356 \ 0 7 . |‘ ~ ‘ ‘ ‘ N ' a ’ ' H6340: Ffi LYS122 . 3‘00- :0: 2 .5 m 1: c 0: 1- 3 0.4‘ H ,5 .E 0.0- . u 4 0 12 10 20 24 20 ll 4 8 12 16 20 24 28 MHz 197 MHz FIGURE 5.8: One-dimensional ESEEM spectra of different XanA complexes prepared in H20 and 60% 2H2O buffer at 172.0 mT and 340.0 mT. Fe(II)-XanA complex examined at 172.0 mT in H2O buffer (A) and 60% 2H2O buffer (B), aKG-Fe(II)-XanA complex examined at 172.0 mT in H2O buffer (C) and 60% 2H2O buffer (D), xanthine-aKG-Fe(II)-XanA complex examined at 172.0 mT in H2O buffer (E) and 60% 2H2O buffer (F). Fe(II)-XanA complex examined at 340.0 mT in H2O buffer (G) and 60% 2H2O buffer (H), aKG-Fe(II)-XanA complex examined at 340.0 mT in H2O buffer (1) and 60% 2H2O buffer (J), xanthine-aKG-Fe(II)-XanA complex examined at 340.0 mT in H2O buffer (K) and 60% 2H2O buffer (L). Protein concentration, 0.45 mM; ferrous ammonium sulfate, slightly less than 0.45 mM; aKG and taurine, 1 mM. ESEEM data were collected under the conditions as described in experimental procedures. I98 FIGURE 5.8: A 2.5 «PO A .— 2.0 5 1.5 a? In c 1.0 2 .E 0.5 0.9. I l I I I I 0 4 01210202420 MHz A20 C q o 2.0 .. 51.5 2" 3 1.0 C 3 0.5 .E 0.0 0 4 01210202420 MHz A10 E q o I- 1.2 x v £00 2 o 0.4 H .E 0.0. . . . . . . . 0 4 01210202420 MHz 6. B {‘5‘ o 0-4. :5 3. .é‘ 102- 5 E“ t 4 0 1210202420 MHz 8- D A 9°0- ._ 5 4 .é‘ 2 2 .9. .E c' I l I l T I I‘ 0 4 01210202420 MHz 5 F A «9 4 o ‘- x3 V .22 In t: 31 IE 9| l I jfffi'I'fi‘ o 4 01210202420 MHz FIGURE 5.8: 12 16 20 24 28 4 0 76543210 A93 .0 3.3.35 v—Ij 1'2'1'0 2'0 2'4 20 Y 0 4 - 7 an-.. 4321 _. 5 33:35 MHz MHz 12 10 20 24 20 4 . 2 1 0 am 2. 5 3.3.35 0. 0'1'21'0 2'0 54'2'0 4 1 v #0 v a I 4 2 n. AMo —. ...».c hmcmufi 0 MHz 0 12 10 20 24 23 MHz 4 No.3 My 3_m.m_3:_ MHz 8 6 4 2 A02. 5 33:35 0 12 10 2'0 24 20 1 MHz 200 the proton Larmor frequency at g = 2. The low frequency region that contains 14N and 1H couplings are more complicated than was observed at g = 4 since bound NO might also contribute to this feature. A striking difference is that the broad peaks contributed from the strong proton couplings observed at g = 4 disappeared at g = 2 spectra. The corresponding spectra obtained with XanA at the two fields (Figure 5.8) were very similar to those observed for TauD. For both enzymes, when probed at g = 4, the contributions from the coupled nuclei were overlapped making it necessary to use the two dimensional, 4-pulse HYSCORE method to resolve contributions from bound histidine nitrogens, coordinated H20, and ambient H20. Two-Dimensional ESEEM of {FeN0}7 adducts of T auD. TauD was used as the model system to develop HYSCORE spectroscopy as a tool for analysis of Fe(II)/(1K6 dioxygenases. As described in the following paragraphs, the HYSCORE analysis of TauD included the use of H20 and 2H-labeled buffers, studies at several magnetic field strengths, examination of 2H-labeled substrates, investigation of the effect of an inhibitor, and the study of W248F variant enzyme. Comparison of HYSCORE spectra collected at 172.0 mT (perpendicular to the Fe-NO axis) for TauD samples prepared in aqueous buffer versus buffer containing 60% 2H20 reveal features associated with exchangeable protons and perturbations induced by the binding of co-substrates (Figure 5.9). In particular, the HYSCORE spectrum of the enzyme with Fe(II) alone shows two major cross-correlations in the (+,+) quadrant, Figure 5.9A. The stronger correlations lying in low frequencies (0.94, 1.86 MHz) and (1.77, 0.88 MHZ), which are shown more clearly in Figure 5.10A, are 20] likely due to 14N couplings. The HYSCORE spectrum also shows at least three pairs of high frequency correlations, (5.32, 9.14 MHZ) (9.01, 5.76 MHZ), (5.79, 9.57 MHz) (9.50, 5.89 MHZ) and (4.1, 11.2 MHZ) (11.3, 3.9 MHZ). The correlations with higher frequencies exhibit a weak, disordered system of overlapping IH arches which most likely reflects a high degree of disorder at the site prior to co-substrate addition. When aKG was added, Figure 5.9C , the low frequency feature remains the same, whereas, the higher frequency correlations, (3.94, 11.58 MHZ) (11.58, 3.78 MHZ), (4.85, 11.64 MHz) (11.86, 4.33 MHZ) and (5.37, 9.70 MHZ) (9.70, 5.55 MHZ), become better defined and take on the appearance of a wedge-shape that could be indicative of a 1H hyperfine coupling of rhombic symmetry. Subsequent addition of substrate taurine to yield the NO-bound ternary taurine-aKG-Fe(II)-TauD complex at the active site yielded a new IH hyperfine coupling, at frequency correlations (4.48, 11.80 MHZ) and (11.61, 4.73 MHZ), with smoother contours and better resolution as shown in Figure 5.9E. In addition, there is a less resolved coupling at (2.1, 12.0 MHZ) and (12.3, 2.2 MHZ) lying right beside the major lH interactions. Figure 5.9B, D and F show the HYSCORE spectra at g = 4 of Fe(II)-TauD, OLKG-Fe(II)-TauD and taurine-aKG-Feal) -TauD complexes prepared in 60% 2H20 buffer. Compared to signals in H20 buffer, the low frequency couplings slightly changed in shape, Figure 5.10B, and in their correlation peaks (1.03, 2.04 MHZ) and (2.07, 1.00 MHZ). For complexes Fe(II)-TauD and aKG—Fe(II)-TauD, the cross peaks at high frequency disappeared from the spectrum when the samples are exchanged against 2H20 buffer, however, in the presence of taurine, this proton correlation is not exchangeable. 202 FIGURE 5.9: HYSCORE spectra of different TauD complexes prepared in H20 and 60% 2H20 buffer at 172.0 mT. Fe(II)-TauD complex examined in H20 buffer (A) and 60% 2H20 buffer (B), aKG-Fe(H)-TauD complex examined in H20 buffer (C) and 60% 2H20 buffer (D), taurine-aKG-Fe(II)-TauD complex examined in H20 buffer (E) and 60% 2H20 buffer (F). buffer. Protein concentration, 2 mM; ferrous ammonium sulfate, slightly < 2 mM; aKG and taurine, 4 mM. 1H A 15J E 10< E 10. E " a I: s. l: 5, r. .i'!’ j ('1‘ 3? x '1 5 1'0 1'5 '0 12 (MHz) 15‘ C 15- ‘1'" 1o. . E g 7: 1o- \. :I: a -.. a C 5‘ c 5. ml! 2 e . r u 5 1a 15 9" 1'2 (”112) 15. E 1" 1: 104 ‘7 1o- : I .2, s a t s- I: c- r’. r“; -. 't 5 1'0 15 50 12 (MHz) 5 1'0 12 (M Hz) D 203 FIGURE 5.10: HYSCORE spectra of taurine-aKG-FeaD-TauD complex prepared in H20 buffer at 172.0 mT (A), 340.0 mT (C) and 60% 2H20 buffer at 172.0 mT (B), 340.0 mT (D). All the spectra were examined at a threshold high enough to show l4N couplings. Protein concentration, 2 mM; ferrous ammonium sulfate, slightly < 2 mM; aKG and taurine, 4 mM. 15 1s. A B 101 .. 1o- 1; i“ z a a E s. I: 5. o . . . o . r . . . 0 1o 15 o 5 1o 15 f2 (MHZ) f2 (MI-ll) 2° 0 C 15d 1s~ ,3: 1.7 ‘5»? 1o. 1 10‘ I = a I: I: s- 5‘ o T— ' I f 0 V I ' I I o 10 1s 20 o 5 1o 15 1'2 ("H11 72 (MHz) 204 Figure 5.11 exhibits the HYSCORE spectra that result from repeating the measurement and processing procedure described for Figure 5.9 at 340.0 mT. An intact contour centered at the diagonal of (+,+) quadrant (14.5, 14.5 MHZ) was observed for all three Complexes, Fe(II)-TauD, (xKG-Fe(II)-TauD and taurine-aKG-Fe(II)—TauD, whether they were prepared in H20 or 60% 2H20 bufi'er. However, the feature at low frequency changed its appearance as substrate was added. Two weak I4N couplings were shown for complexes Fe(II)-TauD and (xKG-Fe(II)-TauD at correlations (2.26, 2.94 MHz) (2.94, 2.26 MHZ) and (1.34, 3.34 MHZ) (3.33, 1.35 MHZ), Figure 5.11A and C. In the presence of taurine, only one coupling remained at correlations (2.15, 3.04 MHz) and (3.04, 2.16 MHZ), Figure 5.11E. The low frequency couplings are very similar for all three complexes when they were prepared in 2H20 buffer, as illustrated in Figure 5.11B, D, F and 5.11D. This result is probably due to the dominate feature of 2H signals. One significant change for the spectra taken at 340.0 mT compared to those taken at 172.0 mM is that the 1H cross peaks at high frequency are no longer exist. Three possibilities were considered to account for the origin of the non-exchangeable protons in NO-treated taurine-aKG-Fe(II)-TauD. First, this lH could come from C-3 of aKG. To test this option, NO-treated taurine-NOG-Fe(II)-TauD in H20 was examined by HYSCORE spectroscopy. NOG is a known inhibitor of several Fe(II)/OLKG-dependent dioxygenase family members (42-45), including XanA and TauD (31, 46), and is known to compete with OLKG for binding to the metallocenter. The addition of NOG to Fe(II)-TauD yields a broad 205 FIGURE 5.11: HYSCORE spectra of different TauD complexes prepared in H20 and 60% 2P120 buffer at 340.0 mT. Fe(II)-TauD complex examined at in H20 buffer (A) and 60% szO buffer (B), OLKG-Fe(II)-TauD complex examined in H20 buffer (C) and 60% 2H20 buffer (D), taurine-aKG-FeaD-TauD complex examined in H20 buffer (B) and 60% 2H20 buffer (F). Protein concentration, 2 mM; ferrous ammonium sulfate, slightly < 2 mM; OtKG and taurine, 4 mM. 20 A , 3 i “‘ W; ‘37 1o- 1;; I I 10. a . a E a: ‘ sq 5. _ ‘ o .2 1“ .1 : ‘~x~._ ,9 o f a o a o 5 1o 15 o 5 1o 15 20 12 (MHz) :2 (MHz) 20 20 1s. "71 x,“ 16 4 {Ix 2“. \ f; a 2: 1o- 51' 1o. 2 2 V v E I s L . ‘ rt 0.3 1151:: \ ‘1 a: .‘ Kit”; {5 o ' r —' ‘r fii r v 0 —v 1 fl 1 v I f o 5 1o 15 20 o 5 1o 15 an 12 (MHz) f2 (MHZ) 20 m 15- ,g. ‘5‘ Q A i A i N i 10~ z: 101 = a I: I: 5 5. a gig. 1 °- *- r“ 0 v f T r v 1 r o ' r ' I ' T T o 5 10 1s 20 o 5 1o 15 20 1'2 ("H21 1'! (MHz) 206 absorbance near 400 nm rather than the 530 nm MLCT feature observed for (XKG (46). After adding taurine and NO to this sample, the protein complex exhibited an electronic spectrum with a well-defined peak at 408 nm and a feature between 550 nm and 800 nm, Figure 5.12A. Very similar 111 cross peaks were observed in the HYSCORE spectrum of this sample in H30 at 172.0 mT compared to that generated by adding NO to the taurine-aKG-Fe(II)-TauD complex (Figure 5.128). This result suggested that the non-exchangeable proton likely does not arise from the C-3 position of OLKG Further confirmation of this point was obtained by using aKG that was incubated for 24 hr in szO at 25 °C to exchange the protons at C-3. This sample was confirmed to be predominantly intact by the demonstration that its use generated 70% of the activity (km) for freshly prepared aKG When the 172.0 mT HYSCORE spectra of NO-treated tauine-zH-labeled-(xKG-Fe(II)-TauD (2 mM protein in 60% 2H20 or 0.45 mM TauD in 90% 2H20) were compared to the unlabeled samples, no significant differences were noted (Figure 5.13), again emphasizing that C-3 OLKG protons are unlikely to be the source of non-exchangeable protons observed in the sample. A second possible source of the non-exchangeable protons is from the primary substrate. To examine this option, HYSCORE spectra were obtained in H20 for NO-treated taurine-aKG-Fe(II)-TauD where the taurine was deuterated at C1. The non-exchangeable IH was still present in the HYSCORE spectra at 172.0 mT, although a slight change in the low frequency region was observed at 340.0 mT (Figure 5.14D), a new correlation appeared at (2.21, 2.21 MHZ) compared to 207 FIGURE 5.12: UV-Vis spectra and HYSCORE of ternary complex taurine-NOG-Fe-(ID- TauD. (A) Left panel: UV-Vis Spectra of anaerobic TauD (black), Fe(II)-TauD (red), NOG-Fe(m-TauD (blue) and taurine-NOG-Fe-(II)-TauD (cyan) obtained in 50 mM Tris buffer, pH 8.0. The TauD subunit (2 mM) was mixed with ferrous ammonium sulfate (slightly < 2 mM) and adjusted to contain 4 mM orKG and 4 mM taurine. Right panel, difference spectra of Fe(II)-TauD in the presence of NOG (solid) and NOG plus taurine (dashed). HYSCORE spectra of taurine-NOG-Fe(II)-TauD complex prepared in H20 buffer at 172.0 mT (B) and 340.0 mT (C). 400 A ‘7 E 300 0 F . I E 2001 v ‘0 10° 1 ' 1 fi I ' I ' j ' ' ' ' ' ' 400 500 000 700 .00 M m m m m "m "I“ 20 15" B C 1 T " _|~ [‘11,].— 15‘ k 1 A 3'? ~ 1 , :1? 1o 2 1‘ s z E 54 3: 5.1 1 . ‘1 ‘6: “r‘ 1 a. o ', Z, . ' L F , I o _ 1 , . I fl 1 . 0 5 1o 15 0 5 10 15 20 f2 (MHZ) f2 (MHZ) 208 FIGURE 5.13: HYSCORE spectra at 172.0 mT of taurine-ZH—labeled-0tKG-Fe(II)-TauD prepared in 60% 2H20 buffer (A) and 90% 2H20 buffer. Protein concentration for (A) is 2 mM and (B) is 0.45 mM; ferrous ammonium sulfate for (A) is slightly < 2 mM and (B) is < 0.45 mM; OLKG and taurine, 4 mM for (A) and 1 mM for (B). q, , 4 ~._ , ‘1 :- , ' ~j _. ' ' 10- 7" A > ’ . ‘— N ( 19 .‘1‘3 ‘ D 1 I 11 .‘\ 2 :1 v 1) c <7 _ F X) V I0- 5 - 1 o | A; I ' 0 5 10 15 12 (M Hz) 1 5 - B ..., 1o - A N C?) I .1,- . E 31.13)- v r A.» a: 5‘ i x v: V. V:‘\ o . 9 . . 0 5 10 15 r2 (M Hz) 209 FIGURE 5.14: HYSCORE spectra of 2H-labeled taurine-orKG-FeflD-TauD complex prepared in H20 buffer at 172.0 mT (A) and 340.0 mT (B). Protein concentration, 2 mM; ferrous ammonium sulfate, < 2 mM; (xKG and taurine, 4 mM. 1 5 - A . 3'— , ~ g 9- ,. ‘ it 34;. ‘12-. . ..7‘!‘ ”j .. . -g-‘g' _ . 4., §~ ‘1 :1 \ ‘—\ _' ,‘ .11- M “\m _ f. ._ 7‘ 971-1 r", . . I . 1 1 0 - .. n. A C‘ 1' 4 r;‘),‘ N r / 1 li (7:4): Hi? 2 \ { .r/ V" 1 can; . . x F ‘11.": ,i ‘ '1 \‘I ‘ . ‘ E x ‘1. c‘ ..X '1 h 5 q ' g . m \. .1 ‘~ . .1.» 1 x; ‘ j ' 2X: .‘ 2,, j. ",1. ‘.\ -' Its 11., L .‘i'i‘ 4 In ‘. I 0 a 20 15~ 1’21: 46}- J 12 (MHz) 210 non-deuterated spectrum, Figure 5.10C. The third possibility to account for the observed non-exchangeable 1H is that it derives from a side chain of TauD. To attempt to identify this proton, the distance between the lH and Fe was estimated by analyzing the cross-peak lineshape of the HYSCORE spectrum for NO-treated taurine-aKG-Fe(II)-TauD in H20 at 172.0 mT. These cross peaks are indicative of an axial hyperfine coupling and can be analyzed for their dipole-dipole and isotropic hyperfine contributions, |T| and lAisol, respectively, using a graphical analysis developed by Dikanov and Bowman (4 7). In this analysis, an arc is constructed down the long-axis of the contour and correlated frequency pairs (vmvB) are read from points along the arc. One then constructs a plot of V.,? vs. v52 and extracts the slope, Qa, and y-intercept, Ga, from a linear least squares fit to the points. A150 and |T| are then determined from knowing the Larmor frequency of the nucleus, v), and the system of equations given below. v. , the 1H Larmor frequency at g = 4, as 7.323 MHz, [3 (Bohr constant) is 9.27408 x 10'24 JT", 6,, (Bohr magneton constant) is 5.05082x 10'27 JT", g. (electron g factor) is 4.00, gn (nuclear g factor) is 5.586, and h (Planck’s constant) is 6.62618x 10'34 Is. Fa = [i4vi(Qa + 1)]/(Qa - l) (2) m = (2/3)-{a[oa(r., a 4v.)]/2v. — 4v.2 + ref/4} ”1’ (3) A.,, = (Fa - T)/2 (4) T = gcgnflcfln/hr3 (5) In practice, the assignment of the labels or and B to the frequency pairs is arbitrary and one just has to pick a convention and stick to it. The signs for the first 211 term in the square root argument for calculating [Tl are chosen to yield a real root. The result of using this system of equations is two unique pairs of values for Also and T. One set features Am and T parameters of the same sign, while for the other, Aiso and T have opposite signs. For the higher are of Figure 5913, the v0.2 vs. V32 plot yielded Q, = -0.6364 and GCl = 108.47 MHZZ. Plugging these numbers into equation 2-5 gave T = 5.0 MHZ, Aiso = 1.1 MHZ and r, the distance between the transition metal ion and nuclei of interest, approximate 3.2 A. A field dependent experiment was carried out for the taurine-aKG-Fe(II)TauD sample in 60% 2HZO at 172.0 mT, 182-.0 mT, and 192.0 mT to further study the gradual hyperfine coupling changes and determine the Aim and T values, Figure 5.15. After simulation, the non-exchangeable lH cross-peaks are characterized by Aiso = 0.75 i 0.3 MHZ, T = 5.0 MHZ, and fire = 90° ('H is perpendicular to the Fe-NO bond axis). This result is consistent with the graphical analysis described above. Examination of the TauD crystal structure reveals six residues that could position one or more protons at approximately this distance: the three metal ligands (His99, Asp101, and His255) and Asn95, Trp248, and Arg270, Figure 5.16A. Since the crystal structure was obtained under anaerobic conditions, the NO ligand was modeled into the crystal structure using bond lengths and angles from model compounds studies and placing the Fe-NO bond opposite to the axial histidine ligand (41). In particular, proton of C-7 in the benzene ring of Trp248 provided the shortest distance of 3.7 A to the iron center, Figure 5.16B. To test the importance of the Trp248 C-7 proton relative to the HYSCORE spectrum, we made use of the W248F 212 variant that had previously been characterized and shown to retain 37% of the activity (km) of wild-type TauD (26). In addition, this mutant protein exhibits UV—visible spectra similar to the control enzyme when incubated anaerobically with aKG and taurine. The taurine—aKG-Fe(II)-TauD W248F complex was treated with NO and examined by HYSCORE at 172.0 mT (Figure 5.17). When using 2 mM W248F, the lH hyperfine coupling was still clearly observed and remained non-exchangeable in 60% 2&0 buffer. Furthermore, the simulation result from taurine-aKG-Fe(II)-TauD resembled spectrum of taurine-(xKG-Fe(II)-TauD W248F complex as illustrated in Figure 5.17C. Analogous experiments were not carried out with Arg270 mutants because this residue is well conserved in TauD sequences and we have found that the R270K variant is inactive. In the case of Asn95, the available variants (N95A and N95D) exhibit very large increases in Km (taurine) and would not be suitable for these studies. Two-Dimensional ESEEM of {FeN0}7 adducts of XanA. The HYSCORE spectra of XanA at 172.0 mT (Figure 5.18) and 340.0 mT (Figure 5.19) exhibited patterns very similar to TauD. A weak, exchangeable IH hyperfine coupling at (5.73, 9.23 MHZ) and (9.37, 5.48 MHZ) was observed before addition of (xKG and xanthine; stronger, exchangeable IH couplings appeared when orKG was added; and further interaction of xanthine caused a new, non-exchangeable lH coupling at (4.25, 11.71 MHZ) and (11.69, 3.78 MHz). The low frequency l4N couplings are similar to TauD spectra in 2H20 prepared buffer, comparing Figure 5.20A to Figure 5.11A. When the 172.0 mT spectra were examined at a less sensitive contour or the spectra were 213 FIGURE 5.15: HYSCORE spectra of taurine-aKG-Fe(II)-TauD complex prepared in 60% 2H20 buffer at 172.0 mT (A), 182.0 mT (B) and 192.0 mT (C). Red dots are simulations. Protein concentration, 2 mM; ferrous ammonium sulfate, slightly < 2 mM; aKG and taurine, 4 mM. 14. i 7 . 3172023. 12... WEN A. .f 10 . g E . a 5 E “‘r 2. - c 6 - , ..‘ c~ : f2 (MHz) f1 (MHz) . . :41}; .1 . :43! ‘ . '6 is 1'0 r2 (MHz) 214 FIGURE 5.16: Two depictions of the TauD active site (A, B) and the XanA active site (C). aKG in stick form with its carbon atoms colored yellow is shown chelating the metal (red sphere). The three side chains that bind the metal are shown in stick form with carbon in yellow, oxygen in red and hydrogen in white. Residues located close to the metal center and potentially capable of accounting for the non-exchangeable hyperfine coupling with the paramagnetic center are shown in stick form with their carbon atoms in blue and nitrogen in dark blue. Panel B highlights the distance and geometry of Trp248 versus the NO-bound metal center of TauD. Panel C highlights the distance and geometry of Asp138 and ILe150 to metal center of XanA. rm. Taurlm 270 Arg Asn95 His” HI8255 215 FIGURE 5.17: HYSCORE spectra at 172.0 mT of taurine-aKG-Fe(H)—TauD W248F prepared in H20 buffer (A) and 60% 2H20 buffer (B). Protein concentration is 2 mM; ferrous ammonium sulfate, slightyly < 2 mM; aKG and taurine, 4 mM. Spectrum (C) shows the spectrum (A) overlapping with simulations in red. is. A 1.4.. , A' 34 ) :7 1o. 6 ‘ 1 I 1. 5 415:3 [in I: s_ I" e r I I I I I 5 10 15 f2 (M Hz) 15- B k Na ’1; 10- ; .1}: I: 5_ ’ N i, :2 e rs” I l 5 10 15 f2 (M Hz) 1 14.- 12:- 31;. C 10?» E i 1 a “i c 6) 4: 21 » 1 o-.. . ......L,....,._.__,.....z..n.. ,4. ._.L_.., . . ..l._.._ o s 3 1o 12 14 f2 (MHz) 216 FIGURE 5.18: HYSCORE spectra of different XanA complexes prepared in H20 and 60% 21120 buffer at 172.0 mT. Fe(II)-XanA complex examined at in H20 buffer (A) and 60% 2H20 buffer (B), aKG-Fe(II)-XanA complex examined in H20 buffer (C) and 60% 2H20 buffer (D), xanthine-aKG—FeaD-XanA complex examined in H20 buffer (B) and 60% 2H20 buffer (F). Protein concentration, 0.45 mM; ferrous ammonium sulfate, slightly < 0.45 mM; ocKG and xanthine, 1 mM. 15. A 154 : .t'.’ B . . '_ o C ' v5. “ ‘ " o §10- L’Q-S : .. £10. 3 o ‘ 1 ‘ o a .3 ‘ a 9 . Q h a , ~(f? . a . a -» o . - . .1 ‘5 _ (1.. J., t 3‘ ° Q I: 5. ~.. , ., . 3 . . t i .. . ‘_ .. o . a 1 o o. o -A . Q Swear» b - - e 0 AZ . f 1 j I 0+ 7 .1 {533%. . 1 0 I 10 15 o 5 1o 1; ”(mill amuz) 15- C 15" . D 4 ‘ J ‘7' 1o "3' 9 4. 10* fl: 1 ' ct'g C- J: E 1 “$35 kart: a {rd (3 3, I: . v‘ " I: 5 e 0 I'Ir o v u v 1 v I o 1* r V I Y I 0 5 10 1‘ 0 5 10 15 f2(Ml-lz) WNW) 151 E 15 F .~ g 1 )‘c a . ' my . is». A101 5' “1°. . N N :I: I 1 ' 3, es 5 % I: .4 ° : s- . o. 8 2i? 3:3 ,. - o T I o fir I 0 5 10 16 0 I 10 15 f2 (M Hz) f2 (MHz) 217 FIGURE 5.19: HYSCORE spectra of different XanA complexes prepared in H20 and 60% 2HgO buffer at 340.0 mT. Fe(ID-XanA complex examined at in H20 buffer (A) and 60% 2H20 buffer (B), aKG-Fe(II)—XanA complex examined in H20 buffer (C) and 60% 2H20 buffer (D), xanthine-aKG-FeflD-XanA complex examined in H20 buffer (E) and 60% 2H20 buffer (F). Protein concentration, 0.45 mM; ferrous ammonium sulfate, slightly < 0.45 mM; OLKG and xanthine, 1 mM. an an A B . {5* q I~ 1‘ \§ 1‘ k f 10. 5:. 10: a a c ‘ a: s- s‘ . J 1 av, o r ‘— r —'— I o ' V ‘7 Y 5 1o 15 an o 5 1o 15 12 (MHz) f2(MHz) 20 C D as. ,3 p 15‘ ‘\+ ‘37 10‘ A 1 a: 3'? . 3.. .5. 10 I: s. t h 1 ft, New 0 I I 7 o ' I f 1 i I 5 1o 15 o 6 1o 16 nwuz) f2(MH2) 20 . 154 3 12‘ 10+ 1.: 1 a: . .2. g 10 I: I: sq 5_ 9 o ' f V f ' o ' I I ‘— f o s 10 15 o 5 10 1s f2(MHz) tullflz) 218 FIGURE 5.20: HYSCORE spectra of xanthine-aKG-Fe(H)—XanA complex prepared in H20 buffer at 172.0 mT (A), 340.0 mT (C) and 60% 2H20 buffer at 172.0 mT (B), 340.0 mT (D). All the spectra were cut threshold high enough to show l4N couplings. Protein concentration, 0.45 mM; ferrous ammonium sulfate, slightly < 0.45 mM; ocKG and xanthine, 1 mM. 16¢ A 16‘ B 4;: 10~ 7: 10< I I a a , I: 5‘ c 54 ‘05 ‘ 41"}"1 o e . 1 . . . o . . . f . o 5 1o 15 o 5 1o 16 1'2 (MHZ) f2(MHz) , - a 20 154 C Q-% 1 D . , 15‘ a. 10. ... 1 N N I :r: 104 a a I: ‘1 a: ‘4 Q a 1 a o ' I f r V I o 1 I f I r , . o 5 1o 15 o 5 1o 15 20 ”(MM 12 (MHz) 219 studied using 340.0 mT field, non-exchangeable protons were not convincingly observed and the features associated with nitrogen atoms were much less obvious than in the case of TauD. The potential donor of this non-exchangeable proton coule be Asp138, as shown in Figure 5.16C, or IlelSO, which occupies the similar position of Trp248 in TauD. 220 DISCUSSION Characterization of the coordination environment of most Fe(II) enzymes is hampered by the difficulty of obtaining high—quality crystal structures of these proteins under anaerobic conditions. One approach to overcome this hurdle is to make use of various spectroscopic methods to examine the active site environments. Whereas Fe(II) sites generally are uninformative when examined by electronic and EPR spectroscopies, prior studies with TauD (20, 28) and other Fe(II)/aKG dioxygenase family members (40, 48) have shown that the (xKG-bound Fe(II) species exhibit a visible chromophore attributed to a .MLCT transition (36) and that their NO-bound centers yield visible species that are paramagnetic and capable of study by EPR methods. Here, we use the highly soluble, easily obtained, and well-characterized TauD enzyme as a model system to extend these methods to include one- and two-dimensional ESEEM, and then we apply these methods to the study of XanA, a recently identified member of this group of enzymes. As expected from TauD and other family members, (xKG-Fe(II)-XanA exhibited characteristic MLCT transitions that were more defined in the presence of xanthine. Three partially resolved bands, 58] nm, 506 nm and 465 nm, could be assigned to electron transfer from Fe(II) dyz, d xiy’, d 2’ to the 1!. orbital of aKG respectively (36). In the presence of NO, XanA developed a distinct UV-visible spectrum as seen in many Fe(II) proteins. The two main features, were previously assigned by Solomon and coworkers to NO' to Fe(III) charge transfer transitions, in which the most intense transition at 443 nm for TauD and 435 nm for XanA are due 22] to the interaction of the out-of-plane (the plane created by the bend Fe-N=O bond) NO" 21:. to Fe(III) dyz orbitals, and the broad transitions around 650 nm for both enzymes could be due to the overlap of two transitions, the in-plane NO' 21:. to Fe(III) d,z and Fe(IIl) d,y to Fe(III) dxif orbitals (41). NO-treated TauD and XanA complexes in the absence of substrate exhibited a nearly axial signals of S = 3/2 resulting from the strong antiferromagnetic coupling of the high-spin Fe(III) (S = 5/2) with NO' (S = 1) (41). In the presence of primary substrate, both enzymes exhibited a mixed signal at g = 4 composed of a major axial species (~70%) and a more rhombic one (~30%). Increasing the co-substrate or substrate ratio to protein from 2.0 to 5.0 did not alter the overall EPR lineshape and the percentages of each component. Moreover, the diagnostic MLCT transition related to aKG bidentate binding to Fe(II) did not increase when the ratios of added (XKG or taurine to protein increased. These results indicated that the two different species were not caused by slow equilibrium between the protein and substrates, but more likely caused by slightly different ligand environments of the ternary complexes taurine-aKG-Fe(II)-TauD or xanthine-aKG-Fe(II)-XanA. Different environments could be associated with distinct Fe-NO angles arising from different ligand geometries of the equatorial histidine and aspartate. Both of the observed species could be catalytic relevant, or only one might lead to turnover. Neither UV-Vis nor HYSCORE spectroscopies are sensitive enough to discriminatethe basis of the two species. One-dimensional ESEEM spectra of {FeNO}7 adducts of TauD gave mixed 222 signals attributed to l4N (bound histidine and NO) and 'H (coordinated H20, ambient H20, and protons from the protein). Even though 1 could not clearly identify the source of each feature by l-D ESEEM, the individual signals were useful for later studies. The general features of the spectra include the following, (i) Two distinct regions were present in the spectra including a low frequency region, from 0-5 MHz, and a high frequencydomain, usually above proton Larmor frequency. The high frequency signals are more resolved, especially, for the case of a non-exchangeable proton/protons that appears in the presence of substrate. (ii) The low frequency features are more complicated, probably due to a mixture of different 14N and 1H couplings. (iii) ESEEM spectra were dominated by 2H modulations when protein samples were prepared in 60% 2H20 buffer. The important advantage of HYSCORE techniques lies in the creation of cross peaks whose coordinates are nuclear frequencies from opposite-electron-spin manifolds. By using HY SCORE, our ability to identity couplings from different nuclei is greatly increased, as well as the capability to discriminate noise from signals. When TauD is incubated with Fe(II) alone, two water molecules should be coordinated to the Fe center, consistent with HY SCORE spectra at g = 4 which showed a substantial distribution of exchangeable lH hyperfine couplings attributed to accessible water molecules. There should be no bound water molecules when (xKG is added, however, better defined proton couplings appeared in the spectrum. My explanation is that even though there is no metal bound water, some ambient water which binds near the {FeNO}7 site are still close enough to modulate the electronic 223 center. Such protons may be attributed to the C-1 protonated carboxylate of (XKG, or those interacting with the oxygen of the Fe-NO adduct. These proton cross-peaks disappear from the spectrum when the samples are exchanged against 2&0 buffer since they come from ambient water. The most interesting result is the appearance of the non-exchangeable proton in the ternary complex. This contribution is not derived from (xKG or taurine, as shown from comparison studies of perdeuterated taurine and the aKG surrogate, NOG. Both graphical analysis and simulation of field dependent spectra for the taurine-aKG-Fe(II)-TauD sample gave consistent results: Aiso z 1.0 MHz, T z 5.0 MHz, r z 3.2 A and this proton is perpendicular to Fe-NO bond axis, in agreement with the disappearance of cross-peaks at g = 2. Interestingly, xanthine-aKG-Fe(II)-XanA exhibited a very similar non-exchangeable lH signals with much lower intensity, while only exchangeable 1H couplings were observed for complexes without primary substrate. From these observations, it appears that whatever residue gives rise to this strong signal, it is coupled to the {FeNO}7 center in this unique fashion only when primary substrate is added. Perhaps this residue plays a role in positioning substrate or in guiding the chemistry of the center once the high valent iron-0x0 intermediate is prepared. Based on TauD the crystal structure, this proton could be contributed from any of six residues, including the three metal ligands or Trp248, ,Asn95, and Arg270. W248F TauD has been used to test the importance of Trp248 and its HYSCORE spectrum showed no substantial difference with wild-type TauD. These results do not completely rule out Trp248 as a candidate for this coupling since the phenylalanine residue partially resembles tryptOphan structurally, and it is 224 known that W248F retains 37% of the activity of wild-type TauD. The three metal ligands are less likely as a source of this proton since no non-exchangeable proton cross-peaks were observed for Fe(II)-TauD and aKG—Fe(II)-TauD. However, the nitrogen couplings were greatly affected by adding primary substrate, thus implying that the positions of metal ligands may shift upon addition of taurine to create an optimized site for 02 coordination, compatible with the finding that NO has the greatest affinity to the enzyme in the presence of both co-substrate and substrate. Further mutagenesis studies will help to locate this proton; however, the mutants need to be active to confirm proper folding of the protein, so the residue could not simply be mutated to alanine. The lack of crystal structure of XanA makes it even more difficult to locate this proton donor. The HYSCORE spectra of XanA exhibited patterns very similar to TauD, however, the poor signal-to-noise ratio forces one to set the contour plot threshold high. As a result, only the most intense spots instead of complete contour were observed for XanA complexes and an accurate comparison with TauD results is difficult. To overcome this problem, I attempted to set the threshold of TauD HYSCORE spectra high enough so that only the most intense couplings can be seen, as shown in Figure 5.2]. The comparison of XanA with this plot reveals that the correlations of proton hyperfine couplings are not exactly the same for XanA and TauD in the Fe(II) bound only and OLKG-Fe(II) bound complexes, as shown in Figure 5.21 and Figure 5.18A, C and E. All the XanA samples exhibited a less sensitive contour and much less obvious signals than in the case of TauD, however, this loss of useful information is not just from the decreased sample 225 FIGURE 5.21: HYSCORE spectra at 172.0 mT of Fe(II)-TauD (A), aKG-Fe(II)-TauD (B), and taurine-aKG-Fe(II)-TauD complex prepared in H20 buffer. The threshold of the spectra was adjusted to show ("‘1’ the most intense spots. Protein concentration, 2 mM; ferrous ammonium sulfate, slightly < 2 mM; aKG and taurine, 4 mM. 15- A . 104 '”"""‘~‘~‘r‘2- ,, £ ‘17.: 1.,} E. Q c ‘. 1 "3. o s 10 1s rum-I2) 151 B 1 "f‘x;\ ..-“ -“1 :E a B s 3 o .‘ r ' 1 ' I o 5 1o 15 rzwuz) 15- C 1 \ :10‘ I E e t ‘1 \ 1 J3 o - . o 5 1o 15 f2 (MHZ) 226 concentration of XanA, as compared with TauD at lower concentration (data not shown). These facts imply that structural differences exist between the two proteins. The differences could be associated with changes in the coordination geometry from metal ligands or aKG, or the active site may be affected by surrounding residues to different extents. Nevertheless, it is not surprising to see these differences based on discrepencies observed before (32), including selectivity for primary substrate, spacings between the conserved Asp to the distal His, solvent isotope effect, MLCT transitions of Fe(II)-orKG-protein complex, electronic rhombicity for the ternary complex, also the reactivity of the replaced proton (H at C-1 of taurine, H at C-8 of xanthine). In summary, I applied different spectroscopies, including UV-Vis, EPR, and one- and two- dimensional ESEEM, to gain insights into XanA, a recently described Fe(II)/(1K6 hydroxylase family member that is the first representative to catalyze the oxidation of purine base. A well-characterized member, TauD, was used as a model system. (xKG-Fe(II)-XanA exhibited characteristic MLCT transitions that were altered by the presence of xanthine verifying the (XKG bidentate binding to Fe(II). NO treated XanA complexes developed similar spectra of UV-Vis, EPR, one- and two- dimensional ESEEM compared with those of TauD. This study provides the first spectroscopic information for XanA and yields insights into the coordination properties of Fe(II) in this protein. An non-exchangeable proton coupling signal was observed for the ternary complex, this new feature only exists in the presence of the primary substrate and may play an important role of positioning the substrate in the 227 binding pocket. 228 REFERENCES (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) Hausinger, R. P. (2004) Fe(II)/or-ketoglutarate-dependent hydroxylases and related enzymes. Critical Reviews in Biochemistry and Molecular Biology 39, 21-68. Clifton, I. J., McDonough, M. A., Ehrismann, D., Kershaw, N. J., Granatino, N., and Schofield, C. J. (2006) Structural studies on 2-oxoglutarate oxygenases and related double-stranded B-helix fold protein. Journal of Inorganic Biochemistry 100, 644-669. Purpero, V., and Moran, G. R. (2007) The diverse and pervasive chemistries of the a-keto acid dependent enzymes. J. Biol. Inorg. Chem. 12, 587-601. Kivirikko, K. I., and Pihlajaniemi, T. (1998) Collagen hydroxylases and the protein disulfide isomerase subunit of prolyl 4-hydroxylase. Advances in Enzymology and Related Areas of Molecular Biology 72, 325-398. Schofield, C. J., and Ratcliffe, P. J. (2004) Oxygen sensing by hydroxylases. Nature Reviews in Molecular Cell Biology 5, 343-354. Tsukada, Y.-I., Fang, J., Erdjument-Bromage, H., Warren, M. E., Borchers, C. H., Tempst, P., and Zhang, Y. (2006) Histone demethylation by a family of ijC domain-containing proteins. Nature 439, 811-816. Trewick, S. C., Henshaw, T. F ., Hausinger, R. P., Lindahl, T., and Sedgwick, B. (2002) Oxidative demethylation by Escherichia coli AlkB directly reverts DNA base damage. Nature 419, I74-178. Aas, P. A., Otterlei, M., Falnes, P. O., Vagbe, C. B., Skorpen, F., Akbari, M., Sundheim, 0., Bjoras, M., Slupphaug, G., Seeberg, E., and Krokan, H. E. (2003) Human and bacterial oxidative demethylases repair alkylation damage in both RNA and DNA. Nature 421, 859-863. Mukherji, M., Chien, W., Kershaw, N. J., Clifton, I. J., Schofield, C. J., Wierzbicki, A. S., and Lloyd, M. D. (2001) Structure-function analysis of phytanoyl-CoA 2-hydroxylase mutations causing Refsum's disease. Human Molecular Genetics 10, 1971-1982. Prescott, A. G., and John, P. (1996) Dioxygenases: molecular structure and role in plant metabolism. Annual Review of Plant Physiology and Plant Molecular Biology 4 7, 245-271. 229 (11) (12) (13) (14) (15) (16) (17) (18) (19) (20) Thomburg, L. D., Lai, M.-T., Wishnok, J. S., and Stubbe, J. (1993) A non-heme iron protein with heme tendencies: an investigation of the substrate specificity of thymine hydroxylase. Biochemistry 32, 14023-14033. Fukumori, F., and Hausinger, R. P. (1993) Purification and characterization of 2,4-dichlorophenoxyacetate/a-ketoglutarate dioxygenase. Journal of Biological Chemistry 268, 24311-24317. Miiller, T. A., Zavodszky, M. I., Feig, M., Kuhn, L. A., and Hausinger, R. P. (2006) Structural basis for the enantiospecificities of R- and S—phenoxypropionate/a—ketoglutarate dioxygenases. Protein Science 15, 1356-1368. Eichhom, E., van der Ploeg, J. R., Kertesz, M. A., and Leisinger, T. (1997) Characterization of or-ketoglutarate-dependent taurine dioxygenase from Escherichia coli. Journal of Biological Chemistry 272, 23031-23036. Kertesz, M. A. (1999) Riding the sulfur cycle - metabolism of sulfonates and sulfate esters in Gram-negative bacteria. FEMS Microbiology Reviews 24, 135-175. Elkins, J. M., Ryle, M. J., Clifton, l. J., Dunning Hotopp, J. C., Lloyd, J. S., Burzlaff, N. 1., Baldwin, J. E., Hausinger, R. P., and Roach, P. L. (2002) X-ray crystal structure of Escherichia coli taurine/a-ketoglutarate dioxygenase complexed to ferrous iron and substrates. Biochemistry 41, 5185-5192. O'Brien, J. R., Schuller, D. J., Yang, V. S., Dillard, B. D., and Lanzilotta, W. N. (2003) Substrate-induced conformational changes in Escherichia coli taurine/a-ketoglutarate dioxygenase and insight into the oligomeric structure. Biochemistry 42, 5547-5554. Koehntop, K. D., Emerson, J. P., and Que, L., Jr. (2005) The 2-His-l-carboxylate facial triad: a versatile platform for dioxygen activation by mononuclear non-heme iron(II) enzymes. Journal of Biological Inorganic Chemistry 10, 87-93. Hanauske-Abel, H. M., and Gunzler, V. (1982) A stereochemical concept for the catalytic mechanism of prolylhydroxylase. Applicability to classification and design of inhibitors. Journal of Theoretical Biology 94, 421-455. Ryle, M. J., Padmakumar, R., and Hausinger, R. P. (1999) Stopped-flow kinetic analysis of Escherichia coli taurine/a-ketoglutarate dioxygenase: interactions with a-ketoglutarate, taurine, and oxygen. Biochemistry 38, 230 15278-15286. (21) Price, J. C., Barr, E. W., Tirupati, B., Bollinger, J. M., Jr., and Krebs, C. (2003) The first direct characterization of a high-valent iron intermediate in the reaction of an a-ketoglutarate-dependent dioxygenase: a high-spin Fe(IV) complex in taurine/a-ketoglutarate dioxygenase (TauD) from Escherichia coli. Journal of Biological Chemistry 42, 7497-7508. (22) Price, J. C., Barr, E. W., Glass, T. E., Krebs, C., and Bollinger, J. M., Jr. (2003) Evidence for hydrogen abstraction from C1 of taurine by the high-spin Fe(IV) intermediate detected during oxygen activation by taurine:0t-ketoglutarate dioxygenase (TauD). Journal of the American Chemical Society 125, 13008-13009. (23) Proshlyakov, D. A., Henshaw, T. F., Monterosso, G. R., Ryle, M. J., and Hausinger, R. P. (2004) Direct detection of oxygen intermediates in the non-heme Fe enzyme taurine/a-ketoglutarate dioxygenase. Journal of the American Chemical Society 126, 1022-1023. (24) Riggs-Gelasco, P. J., Price, J. C., Guyer, R. B., Brehm, J. H., Barr, E. W., Bollinger, J. M., Jr., and Krebs, C. (2004) EXAFS spectroscopic evidence for an Fe=O unit in the Fe(IV) intermediate observed during oxygen activation by taurineza-ketoglutarate dioxygenase. Journal of the American Chemical Society 126, 8108-8109. (25) Price, J. C., Barr, E. W., Hoffart, L. M., Krebs, C., and Bollinger, J. M., Jr. (2005) Kinetic dissection of the catalytic mechanism of taurineza-ketoglutarate dioxygenase (TauD) from Escherichia coli. Biochemistry 44, 8138-8147. (26) Grzyska, P. K., Ryle, M. J., Monterosso, G. R., Liu, J., Ballou, D. P., and Hausinger, R. P. (2005) Steady-state and transient kinetic analyses of taurine/a-ketoglutarate dioxygenase: effects of oxygen concentration, alternative sulfonates, and active site variants on the Fe(IV) intermediate. Biochemistry 44, 3845-3855. (27) Sinnecker, S., Svensen, N., Barr, E. W., Ye, S., Bollinger, J. M., Jr., Neese, F., and Krebs, C. (2007) Spectroscopic and computational evaluation of the structure of the high-spin Fe(IV)-0x0 intermediates in taurine: alpha-ketoglutarate dioxygenase from Escherichia coli and its His99Ala ligand variant. J Am Chem Soc 129, 6168-79. (28) Muthakumaran, R. B., Grzyska, P. K., Hausinger, R. P., and McCracken, J. (2007) Probing the Fe-substrate orientation for taurine/a-ketoglutarate 231 (29) (30) (31) (32) (33) (34) (35) (36) (37) dioxygenase using deuterium ESEEM spectroscopy. Biochemistry in press. Cultrone, A., Scazzocchio, C., Rochet, M., Montero-Moran, G., Drevet, C., and Feméndez-Martin, R. (2005) Convergent evolution of hydroxylation mechanisms in the fungal kingdoms: molybdenum cofactor-independent hydroxylation of xanthine via a-ketoglutarate dependent dioxygenase. Molecular Microbiology 5 7, 276-290. Hille, R. (2005) Molybdenum—containing hydroxylases. Archives of Biochemistry and Biophysics 433,107-116 Montero-Moran, G. M., Li, M., Rendém-Huerta, E., Jourdan, F ., Lowe, D. J., Stumpff-Kane, A. W., Feig, M., Scazzocchio, C., and Hausinger, R. P. (2007) Purification and characterization of the FeIl- and a-ketoglutarate-dependent xanthine hydroxylase from Aspergillus nidulans. Biochemistry 46, 5293-5304. Montero-Moran, G. M., Li, M., Rendon-Huerta, E., Jourdan, F., Lowe, D. J., Stumpff-Kane, A. W., Feig, M., Scazzocchio, C., and Hausinger, R. P. (2007) Purification and characterization of the Fe(II)- and a-ketoglutarate—dependent xanthine hydroxylase from Aspergillus nidulans. Biochemistry 46, 5293-5304. Gemperle, C., Aebli, G., Schweiger, A., and Ernst, R. R. (1990) PHase cycling in pulse EPR. J. Magn. Reson. 88, 241-256. Hoffman, B. M., Venters, R. A., and Martinsen, J. (1985) General theory of polycrystalline ENDOR patterns. Effects of finite EPR and ENDOR component linewidths. J. Magn. Reson. 62, 537-542. Grzyska, P. K., Miiller, T. A., Campbell, M. G., and Hausinger, R. P. (2007) Metal ligand substitution and evidence for quinone formation in taurine/a-ketoglutarate dioxygenase. Journal of Inorganic Biochemistry 10], 797-808. Pavel, E. 6., Zhou, J., Busby, R. W., Gunsior, M., Townsend, C. A., and Solomon, E. I. (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. Journal of the American Chemical Society 120, 743-753. Zhou, J., Kelly, W. L., Bachmann, B. 0., Gunsior, M., Townsend, C. A., and Solomon, E. I. (2001) Spectroscopic studies of substrate interactions with Clavaminate synthase 2, a multifunctional or-KG-dependent non-heme iron enzyme: correlation with mechanisms and reactivities. Journal of the American Chemical Society 123, 7388-7398. 232 (38) (39) (40) (41) (42) (43) (44) (45) (46) Ho, R. Y. N., Mehn, M. P., Hegg, E. L., Liu, A., Ryle, M. A., Hausinger, R. P., and Que, L., Jr. (200]) Resonance Raman studies of the iron(lI)-0t-keto acid chromophore in model and enzyme complexes. Journal of the American Chemical Society 123, 5022-5029. Enemark, J. H., and Fleltham, R. D. (1974) Principles of structure, bonding, and reactivity for metal nitrosyl complexes. Coordination Chemistry Reviews 13, 339. Hegg, E. L., Whiting, A. K., Saari, R. E., McCracken, J., Hausinger, R. P., and Que, L., Jr. (1999) Herbicide-degrading (Jr-keto acid-dependent enzyme deA: metal coordination environment and mechanistic insights. Biochemistry 38, 16714-16726. Brown, C. A., Pavlovsky, M. A., Westre, T. E., Zhang, Y., Hedman, B., Hodgson, K. 0., and Solomon, E. I. (1995) Spectroscopic and theoretical description of the electronic structure of S = 3/2 iron-nitrosyl complexes and their relation to Oz activation by non-heme iron enzyme active sites. Journal of the American Chemical Society 11 7, 715-732. Baader, E., Tschank, G., Baringhaus, K. H., Burghard, H., and Gunzler, V. (1994) Inhibition of prolyl 4-hydroxylase by oxalyl amino acid derivatives in vitro, in isolated microsomes and in embryonic chicken tissues. Biochemical Journal 300, 525-530. Cunliffe, C. J., Franklin, T. J., Hales, N. J., and Hill, G. B. (1992) Novel inhibitors of prolyl 4-hydroxylase. 3. Inhibition by the substrate analogue N-oxaloglycine and its derivatives. Journal of Medicinal Chemistry 35, 2652-2658. Chan, D. A., Sutphin, P. D., Denko, N. C., and Giaccia, A. J. (2002) Role of prolyl hydroxylation in oncogenically stabilized hypoxia-inducible factor-10L. Journal of Biological Chemistry 277, 40112-40117. McDonough, M. A., McNeill, L. A., Tilliet, M., Papmichael, C. A., Chen, Q.-Y., Banerji, B., Hewitson, K. S., and Schofield, C. J. (2005) Selective inhibition of factor inhibiting hypoxia-inducible factor. Journal of the American Chemical Society 12 7, 7680-7681. Kalliri, E., Grzyska, P. K., and Hausinger, R. P. (2005) Kinetic and spectroscopic investigation of Co", Ni", and N-oxalylglycine inhibition of the Fen/a-ketoglutarate dioxygenase, TauD. Biochemical and Biophysical Research Communications 338, 191-197. 233 (47) (48) Dikanov, S. A., and Bowman, M., K. (1995) Cross-peak lineshape of two-dimensional ESEEM spectra in disordered S=1/2, l=1/2 spin systems. Journal of Magnetic Resonance A 116, 125-128. Hogan, D. A., Smith, S. R., Saari, E. A., McCracken, J., and Hausinger, R. P. (2000) Site-directed mutagenesis of 2,4-dichlorophenoxyacetic acid/a-ketoglutarate dioxygenase. Identification of residues involved in metallocenter formation and substrate binding. Journal of Biological Chemistry 275, 12400-12409. 234 CHAPTER 6 CONCLUSIONS AND FUTURE RESEARCH 235 CONCLUSIONS Moco-independent Aspergillus nidulans xanthine hydroxylase (XanA) was purified as the His6-tagged recombinant protein from both the fungus and a heterologous bacterial host ’ and characterized as a novel member of Fe(II)/orKG-dependent dioxygenase superfamily (l). Ferrous XanA catalyzes xanthine hydroxylation to uric acid, concomitant with the oxidative decarboxylation of aKG, forming succinate and carbon dioxide. Comparison of XanA isolated from A. nidulans and E. coli revealed very different quaternary structures and posttranslational modifications; however, the kinetic properties of XanA purified from both hosts are very similar (E. coli : ~70 U mg 4, Km of 3] uM for OLKG, and Km of 45 uM for xanthine at pH 7.4; A. nidulans : 30 U mg '1, and K.,, values of 50 uM and 46 uM at pH 7.0). Fe(II) is irreplaceable for enzyme activity and some divalent metals, such as Ni(II), Zn(II) and Cu(II), are competitive inhibitors of Fe(II). A solvent isotope effect was observed upon substituting H20 with 2H20. This result suggests that a chemical group possessing an exchangeable proton such as Fe(III)-OOH or Fe(III)-OH is important in the rate-determining step of the overall reaction. Although the xanthine C-H bond is broken at C-8 during turnover, substitution of the proton at this position by 2H did not lead to a substrate isotope effect. This result demonstrates that C-H cleavage is not the rate-determining step in the reaction. The co-substrate can be substituted by a-ketoadipate, but not by other (it-keto acids tested. NOG, a known inhibitor of several other representatives of this superfamily, is a competitive (with aKG) inhibitor of XanA with a K, of 0.12 pM. XanA displays high specificity 236 towards its primary substrate. 9-methylxanthine and l-methylxanthine act as alternative substrates with significant activity loss, respectively around 20 and 104 times less than xanthine. 6,8-DHP is a competitive, slow-binding inhibitor with Kiapp of 12.6 11M, while 2,8-DHP and 8-HP have almost no inhibitory effect, indicating the potentially important role of the C-6 carbonyl or enol for primary substrate binding. The above studies represent the first biochemical characterization of purified xanthine/aKG dioxygenase, and provide valuable baseline information to carry out further mechanistic and spectroscopic studies. Fe(II)/aKG-dependent dioxygenases catalyze a wild range of reactions and utilize substrates ranging from small molecules, such as taurine, to large polymers, such as proteins or DNA/RNA (2). It is well known that the highly conserved DSBH structure acts as a stable platform to anchor the Fe(II) and aKG (2). The metal ion is ligated by the three residues forming a conserved HXD/EXnH motif, while aKG binds Fe(II) in a bidentate fashion through the C-1 carboxylate and C-2 keto groups. The most interesting question is how to bind different substrates within this scaffold to carry out the unique reactions for each enzyme. A homology model of XanA was generated on the basis of the structure of the related enzyme TauD (I). The XanA protein is predicted to contain the DSBH comprised of eight B-strands with the Fe(II) binding ligands (Hisl49, Asp151, and His340), and the co-substrate positioned to chelate Fe(II) in a bidentate fashion. The aKG C-5 carboxylate is predicted to form a salt bridge with Arg352, while Ly5122 is well positioned to stabilize the C-1 carboxylate of the co-substrate. Xanthine binds in an active site pocket lined with 237 potential hydrogen bond donors or acceptors (Gln99,-Gln101, Glul37, Asp138, Ly3122, Gln356, Cys357, and Asn358). Ala was chosen to replace each of the potential hydrogen-bonding residues at the XanA active site and the eight single mutants were constructed and purified. Afier analyzing the kinetic properties of the variants with xanthine, the alternate substrate 9-methylxanthine, 6,8-DHP and thiol-specific reagents, critical residues that participate in binding of the primary substrate were identified and a model for xanthine binding to XanA was proposed, Figure 4.17. Xanthine is shown binding to the enzyme active site via a constellation of hydrogen bonding interactions which explains the high selectivity towards the primary substrate. The xanthine binding model derived from my extensive kinetic comparisons will need to be verified by crystal structure determination, but crystals of XanA are not available currently. After carrying of biochemical characterization and mutational/kinetic studies (focusing on the primary substrate binding mode) of XanA, my ultimate goal and the most challenging part of this project was to further uncover the mechanism of this enzymatic reaction by using of spectroscopic methods. Knowing the structures of each intermediate is the key to understand the mechanism. I started with investigating the coordination chemistry of different enzyme complexes: XanA-Fe, XanA-Fe(II)-aKG and XanA-Fe(II)-0tKG-xanthine. I confirmed the bidentate binding of orKG to Fe(II) by observing the diagnostic MLCT features around 500 nm by UV-Vis spectroscopy of the XanA-Fe(II)-OLKG and XanA-Fe(II)-0tKG-xanthine complexes. Consistent with a shift from 6-coordinate to 5-coordinate geometry, I 238 observed enhanced definition of the spectroscopic features after adding substrate. Since high spin Fe(II) is EPR silent, I used NO, a good 02 surrogate, to bind Fe(II) to form an {FeNO}7 complex which is EPR active. I found that adding primary substrate disturbs the electronic environment of Fe(II) and this is supported by the more rhombic EPR line shape for XanA-Fe(II)-aKG-xanthine compared with XanA-Fe(II) or XanA-Fe(II)-aKG. Electron nuclear hyperfine coupling between Fe(II) and surrounding ligands was investigated by using one- and two-dimensional ESEEM spectroscopies. In this part, I used TauD as a reference and examined even more extensively than XanA. One-dimensional ESEEM spectra for both XanA and TauD show modulations from ”N and 1H. At g = 4, the contributions from these coupled nuclei are overlapped making it necessary to use the two dimensional, 4-pulse HYSCORE method to resolve contributions from bound histidine nitrogens, coordinated H20, and ambient H20. For TauD-Fe(II), HYSCORE spectra show a substantial distribution of exchangeable, lH hyperfine couplings. When co-substrate orKG is added, the 1H HYSCORE is considerably altered with the dominant hyperfine coupling arising from an exchangeable, strongly-coupled proton of rhombic symmetry. Subsequent addition of substrate taurine, to yield the ternary complex at the active site, showed a new, IH hyperfine interaction that was not exchangeable in 2H20. The HYSCORE cross-peaks from this lH show a hyperfine tensor of axial symmetry characterized by a dipole-dipole distance of 3.2 A and an isotropic contribution of 1.1 MHz. Comparison of these data with the X-ray crystal structure of TauD and the results of parallel studies of TauD variants suggests that this 1H is likely from W248. 239 However, the W248F variant did not show any substantial difference in its spectral features compared with those of wild-type TauD. Interestingly, XanA—Fe(II)-0tKG-xanthine exhibited very similar non-exchangeable lH signals with much lower intensity. Exchangeable IH hyperfine couplings were also observed for XanA-Fe(II) and the XanA-Fe(II)-aKG complex. This comparison study suggested that the coordination chemistry of TauD and XanA are very similar, even though the origin of this strong, non-exchangeable 1H is still unclear, its appearance in both enzyme complexes means this structural change induced by adding primary substrate could be quite common for this superfamily and important for understanding the mechanism. FUTURE RESEARCH Constructing Double or Triple Mutants. Eight single mutants (Q99A, Q101A, E137A, D138A, K122A, Q356A, C357A, and N358A) were constructed and, except Q99A, all of them exhibited substantial activities. This finding suggested that modifying only one residue will not cause significant change of primary substrate binding. Double or even triple mutants might affect xanthine binding more efficiently. N358A has shown the most significant Km change, so it could be considered first; Q99A can be excluded; C357A did not show any direct interaction with substrate based on the homology model, so it could be considered the last possibility. So (Q101A, N358A), (K122A, N358A), (E137A, N358A), (D138A, N358A) and (Q3 56A, N358A) could be the first set of double mutants to try out. 240 Determination of Dissociation Constant (K.,). Dissociation constant (Kd) is a more accurate parameter than Km to examine substrate binding. Fluorescence spectroscopy was applied to try to determine the Kd. Fluorescence measurements were carried out at an excitation wavelength of 280 nm with emission monitored from 300 to 400 nm. The binding of Fe, orKG and xanthine each quenched the endogenous fluorescence of the proteins. However, the quenching by xanthine is artificial since xanthine absorbs at 268 nm which is too close to the excitation wavelength 280 nm. So the quenching observed upon adding xanthine is actually caused by fewer photons being absorbed by the protein. So finding a better to determine the dissociation constant (Kd) of xanthine would be very helpful to understand primary substrate binding mode for XanA. Steady-State and Transient Kinetic Studies by Stopped-Flow UV-Vis spectroscopy. Stopped-flow UV-Vis spectroscopy is a powerful technique to detect intermediates forming within ms to s. The chromophores associated with XanA-Fe(II)-0tKG and XanA-Fe(II)-0tKG-xanthine could be examined to see if their formation is fast enough to be catalytically relevant. Next, the oxygen reactivity of XanA-Fe(II)-0tKG and XanA-Fe(II)-aKG-xanthine could be examined. The Fe(IV)=Oz' intermediate has been observed by stopped-flow methods applied to three family members, and in each case exhibited an absorption near 318 nm by UV/Vis. We could assess whether a similar signal is formed in XanA. To enhance the chances of observing this species, one can attempt to slow down the decay reaction. Using 2H20 instead of H20 may be a useful approach since XanA is known to exhibit a significant solvent isotope effect. 241 Alternatively, the intermediate may be longer-lived in the presence of a poor substrates, such as using 9-methylxanthine instead of xanthine. Probably the best approach is to use 8-2H-xanthine instead of xanthine. Even though I did not observed any isotope effect during steady-state kinetic studies, the deuterated substrate is likely to reduce the rate of the hydrogen abstraction in the reaction and should facilitate the transient kinetic approaches. If transient absorption changes are observed with wild-type XanA, single or double mutants which are potentially important for substrate binding or enzymatic reactions also could be examined by this technique. Resonance Raman and Mossbauer Spectroscopy. The UV-Vis spectrum provides information on electron transitions, but in order to further define the properties of the lntennediate species, more sophisticated methods need to be applied, such as resonance Roman and Mossbauer spectroscopy. For RR, typical vibrational modes of each intermediate, combining the information from modeling studies and other members of this superfamily, could help us assign each species. Mossbauer spectroscopy is very sensitive to Fe redox state and coordination, and it detects the Fe center no matter whether it is paramagnetic or diamagnetic, so it is a very powerful technique to investigate Fe-containing proteins. So, freeze-quench Mossbauer analyses may be a useful tool for further studies. More HYSCORE Spectroscopy studies. In order to assign the non-exchangeable 1H in the TauD-Fe(II)—aKG-taurine complex to certain residue, more single mutants 242 could be prepared, such as substitutions involving Asn95 and Arg270, which also lie within 5 A around Fe(II) of TauD. The metal binding ligands are also possible targets to examine if incorporation of deuterated histidine and aspartate could be accomplished. The same strategy is applicable for wild-type XanA and mutants such as D138A (this residue lies within 5A around Fe(II) of XanA), and might be the first mutant to examine. Another interesting study is to assign the strong couplings we observed at g“ = 2.00 (not observed at g f. 4.00) for both TauD and XanA. It could come from NO or axial histidine ligand as shown in Figure 6.1. Problems Application of many techniques mentioned above, such as, HYSCORE, RR, Mossbauer, even UV—Vis, require substantial amounts of concentrated (at least 0.5 mM), active and stable protein. XanA isolated from E. coli barely meets this requirement and only freshly prepared protein can be used. So, the further studies of the XanA enzyme mechanism will benefit from additional investigation designed to improve the protein yield and stability. From my HYSCORE study, I found the non-exchangeable IH only appearing when the primary substrate is present. This suggests that a protein structural change is induced by adding substrate. Mutation of targeted residues could help to locate this proton; however, the mutants need to be active to confirm proper folding of the protein. That is why we could not simply mutate the metal ligands to alanine. Further mutagenesis studies will need to take this problem into account. 243 FIGURE 6.1: HYSCORE spectra of NO treated ternary complex TauD-Fe-OLKG -taurine in 50 mM Tris-H20, pH = 8.0 at (A) 3400 G and (B) 17200 20 15 \Q A \It. N I 101 a E v t _ E 54 I o. o ‘—l ' I ' I 1 I 1 I 1—I fiT -2o -15 -1o -6 o I 10 15 20 f2 (MHz) A 151 7.710- A \1 1 I ., 1 , E 1‘5 ». V a: 51 :§\ 1 1‘ o I ' I ' I ' I t I ‘7I ‘ I -15 -1o -5 o 5 1o 15 12 (MHz) B 244 REFERENCES (1) (2) Montero-Moran, G. M., Li, M., Rendon-Huerta, E., Jourdan, F., Lowe, D. J., Stumpff-Kane, A. W., Feig, M., Scazzocchio, C., and Hausinger, R. P. (2007) Purification and characterization of the Fe(II)- and a-ketoglutarate-dependent xanthine hydroxylase from Aspergillus nidulans. Biochemistry 46, 5293-5304. Hausinger, R. P. (2004) Fe(II)/0t-ketoglutarate-dependent hydroxylases and related enzymes. Crit. Rev. Biochem. Mol. Biol. 39, 21-68. 245 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII WWWWI