BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials...

13
Biochemical and Structural Studies of 6-Carboxy-5,6,7,8- tetrahydropterin Synthase Reveal the Molecular Basis of Catalytic Promiscuity within the Tunnel-fold Superfamily * Received for publication, February 7, 2014, and in revised form, June 20, 2014 Published, JBC Papers in Press, July 2, 2014, DOI 10.1074/jbc.M114.555680 Zachary D. Miles 1 , Sue A. Roberts, Reid M. McCarty 2 , and Vahe Bandarian 3 From the Department of Chemistry and Biochemistry, University of Arizona, Tucson, Arizona 85721 Background: The bacterial homolog, 6-carboxy-5,6,7,8-tetrahydropterin synthase, of the eukaryotic 6-pyruvoyltetrahy- dropterin synthase enzyme acts on the same substrate but produces different products. Results: Structural and biochemical studies trace the differential reactivity to four residues. Conclusion: Differential reactivity between the enzyme homologs is a result of small changes in the enzyme active site. Significance: This work furthers our understanding of how novel activities may arise from common protein-folds. 6-Pyruvoyltetrahydropterin synthase (PTPS) homologs in both mammals and bacteria catalyze distinct reactions using the same 7,8-dihydroneopterin triphosphate substrate. The mam- malian enzyme converts 7,8-dihydroneopterin triphosphate to 6-pyruvoyltetrahydropterin, whereas the bacterial enzyme cat- alyzes the formation of 6-carboxy-5,6,7,8-tetrahydropterin. To understand the basis for the differential activities we deter- mined the crystal structure of a bacterial PTPS homolog in the presence and absence of various ligands. Comparison to mam- malian structures revealed that although the active sites are nearly structurally identical, the bacterial enzyme houses a His/ Asp dyad that is absent from the mammalian protein. Steady state and time-resolved kinetic analysis of the reaction catalyzed by the bacterial homolog revealed that these residues are responsible for the catalytic divergence. This study demon- strates how small variations in the active site can lead to the emergence of new functions in existing protein folds. The rapid sequencing of bacterial genomes has led to an explosion of sequence information for proteins whose func- tions are not known. It is rarely possible to predict enzymatic function purely on the basis of sequence because proteins with nearly identical three-dimensional structures can catalyze vastly different reactions. To bridge the gap from sequence to function, a molecular level understanding of the amino acid changes that permit evolution of novel enzymatic activity in existing folds is highly desirable. The tunnel-fold (T-fold) superfamily is comprised of a widely distributed group of enzymes that catalyze transformations leading to the production of purines and pterins (1, 2). The mammalian tunnel-fold enzyme 6-pyruvoyltetrahydropterin (PPH 4 ) 4 synthase (mPTPS) is required for the biosynthesis of 6R-L-erythro-5,6,7,8-tetrahydrobiopterin (BH 4 ) in eukaryotes (Fig. 1) (3). BH 4 is an essential cofactor for many enzymes including nitric-oxide synthase and phenylalanine hydroxylase where it can act as either an electron donor or in oxygen inser- tion (4). In addition, this cofactor has been implicated in a vast array of physiological roles including hyperphenylalaninaemia, cellular proliferation, vascular dysfunction, and various neuro- logical disorders (4, 5). mPTPS catalyzes the conversion of 7,8- dihydroneopterin triphosphate (H 2 NTP) to PPH 4 , which is subsequently converted to BH 4 by the NADPH-dependent sepiapterin reductase (6, 7). More recently, an alternative path- way has been identified in higher organisms, including humans, which allows for bypassing the enzyme sepiapterin reductase through the combined actions of a carbonyl reductase and an aldose reductase or by members of the aldo-keto reductase family of enzymes (8 –11). BH 4 is typically not produced in bac- teria, but is present in glycosylated forms in certain cyanobac- teria and Chlorobium tepidum (12–16). Although most bacteria do not posses BH 4 , a PTPS homolog is found in virtually all bacterial genomes for production of pterins and deazapurines. Previous studies from our laboratory have established that the Escherichia coli PTPS homolog, 6-car- boxy-5,6,7,8-tetrahydropterin synthase, catalyzes the second step in the biosynthesis of pyrrolopyrimidine nucleosides involving the conversion of H 2 NTP to 6-carboxy-5,6,7,8-tetrahydropterin (CPH 4 ) (17). CPH 4 is a precursor to 30 natural products that range from antibiotic and anticancer agents, produced by various strains of Actinomyces, to the modified tRNA base queuosine, which is found in nearly all kingdoms of life (Fig. 1) (18). * This work was supported, in whole or in part, by National Institutes of Health Grant GM72623 (to V. B.) and the Burroughs Welcome Fund. The atomic coordinates and structure factors (codes 4NTN, 4NTM, and 4NTK) have been deposited in the Protein Data Bank (http://wwpdb.org/). 1 Supported by Biological Chemistry Training Grant T32 GM008804 from the National Institutes of Health and Integrative Graduate Education and Research Traineeship Program in Comparative Genomics National Science Foundation Grant 0654435. 2 Supported by National Institutes of Health Grant T32 GM008804. Present address: Division of Medicinal Chemistry, College of Pharmacy, and Dept. of Chemistry and Biochemistry, University of Texas, Austin, TX 78712. 3 To whom correspondence should be addressed: 1041 E. Lowell St., Tucson, AZ 85721. Tel.: 520-621-0389; Fax: 520-621-9288; E-mail: [email protected]. edu. 4 The abbreviations used are: PPH 4 , 6-pyruvoyl-5,6,7,8-tetrahydropterin; H 2 NTP, 7,8-dihydroneopterin triphosphate; CPH 4 , 6-carboxy-5,6,7,8-tetrahydrop- terin; QueD, CPH 4 synthase; mPTPS, mammalian 6-pyruvoyltetrahydropterin synthase; BH 4 , 6R-L-erythro-5,6,7,8-tetrahydrobiopterin; PDB, Protein Data Bank. THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 289, NO. 34, pp. 23641–23652, August 22, 2014 © 2014 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A. AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23641 by guest on November 30, 2020 http://www.jbc.org/ Downloaded from

Transcript of BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials...

Page 1: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

Biochemical and Structural Studies of 6-Carboxy-5,6,7,8-tetrahydropterin Synthase Reveal the Molecular Basis ofCatalytic Promiscuity within the Tunnel-fold Superfamily*

Received for publication, February 7, 2014, and in revised form, June 20, 2014 Published, JBC Papers in Press, July 2, 2014, DOI 10.1074/jbc.M114.555680

Zachary D. Miles1, Sue A. Roberts, Reid M. McCarty2, and Vahe Bandarian3

From the Department of Chemistry and Biochemistry, University of Arizona, Tucson, Arizona 85721

Background: The bacterial homolog, 6-carboxy-5,6,7,8-tetrahydropterin synthase, of the eukaryotic 6-pyruvoyltetrahy-dropterin synthase enzyme acts on the same substrate but produces different products.Results: Structural and biochemical studies trace the differential reactivity to four residues.Conclusion: Differential reactivity between the enzyme homologs is a result of small changes in the enzyme active site.Significance: This work furthers our understanding of how novel activities may arise from common protein-folds.

6-Pyruvoyltetrahydropterin synthase (PTPS) homologs inboth mammals and bacteria catalyze distinct reactions using thesame 7,8-dihydroneopterin triphosphate substrate. The mam-malian enzyme converts 7,8-dihydroneopterin triphosphate to6-pyruvoyltetrahydropterin, whereas the bacterial enzyme cat-alyzes the formation of 6-carboxy-5,6,7,8-tetrahydropterin. Tounderstand the basis for the differential activities we deter-mined the crystal structure of a bacterial PTPS homolog in thepresence and absence of various ligands. Comparison to mam-malian structures revealed that although the active sites arenearly structurally identical, the bacterial enzyme houses a His/Asp dyad that is absent from the mammalian protein. Steadystate and time-resolved kinetic analysis of the reaction catalyzedby the bacterial homolog revealed that these residues areresponsible for the catalytic divergence. This study demon-strates how small variations in the active site can lead to theemergence of new functions in existing protein folds.

The rapid sequencing of bacterial genomes has led to anexplosion of sequence information for proteins whose func-tions are not known. It is rarely possible to predict enzymaticfunction purely on the basis of sequence because proteins withnearly identical three-dimensional structures can catalyzevastly different reactions. To bridge the gap from sequence tofunction, a molecular level understanding of the amino acidchanges that permit evolution of novel enzymatic activity inexisting folds is highly desirable.

The tunnel-fold (T-fold) superfamily is comprised of a widelydistributed group of enzymes that catalyze transformationsleading to the production of purines and pterins (1, 2). Themammalian tunnel-fold enzyme 6-pyruvoyltetrahydropterin(PPH4)4 synthase (mPTPS) is required for the biosynthesis of6R-L-erythro-5,6,7,8-tetrahydrobiopterin (BH4) in eukaryotes(Fig. 1) (3). BH4 is an essential cofactor for many enzymesincluding nitric-oxide synthase and phenylalanine hydroxylasewhere it can act as either an electron donor or in oxygen inser-tion (4). In addition, this cofactor has been implicated in a vastarray of physiological roles including hyperphenylalaninaemia,cellular proliferation, vascular dysfunction, and various neuro-logical disorders (4, 5). mPTPS catalyzes the conversion of 7,8-dihydroneopterin triphosphate (H2NTP) to PPH4, which issubsequently converted to BH4 by the NADPH-dependentsepiapterin reductase (6, 7). More recently, an alternative path-way has been identified in higher organisms, including humans,which allows for bypassing the enzyme sepiapterin reductasethrough the combined actions of a carbonyl reductase and analdose reductase or by members of the aldo-keto reductasefamily of enzymes (8 –11). BH4 is typically not produced in bac-teria, but is present in glycosylated forms in certain cyanobac-teria and Chlorobium tepidum (12–16).

Although most bacteria do not posses BH4, a PTPS homologis found in virtually all bacterial genomes for production ofpterins and deazapurines. Previous studies from our laboratoryhave established that the Escherichia coli PTPS homolog, 6-car-boxy-5,6,7,8-tetrahydropterin synthase, catalyzes the second stepin the biosynthesis of pyrrolopyrimidine nucleosides involving theconversion of H2NTP to 6-carboxy-5,6,7,8-tetrahydropterin(CPH4) (17). CPH4 is a precursor to �30 natural products thatrange from antibiotic and anticancer agents, produced by variousstrains of Actinomyces, to the modified tRNA base queuosine,which is found in nearly all kingdoms of life (Fig. 1) (18).

* This work was supported, in whole or in part, by National Institutes of HealthGrant GM72623 (to V. B.) and the Burroughs Welcome Fund.

The atomic coordinates and structure factors (codes 4NTN, 4NTM, and 4NTK)have been deposited in the Protein Data Bank (http://wwpdb.org/).

1 Supported by Biological Chemistry Training Grant T32 GM008804 from theNational Institutes of Health and Integrative Graduate Education andResearch Traineeship Program in Comparative Genomics National ScienceFoundation Grant 0654435.

2 Supported by National Institutes of Health Grant T32 GM008804. Presentaddress: Division of Medicinal Chemistry, College of Pharmacy, and Dept.of Chemistry and Biochemistry, University of Texas, Austin, TX 78712.

3 To whom correspondence should be addressed: 1041 E. Lowell St., Tucson, AZ85721. Tel.: 520-621-0389; Fax: 520-621-9288; E-mail: [email protected].

4 The abbreviations used are: PPH4, 6-pyruvoyl-5,6,7,8-tetrahydropterin; H2NTP,7,8-dihydroneopterin triphosphate; CPH4, 6-carboxy-5,6,7,8-tetrahydrop-terin; QueD, CPH4 synthase; mPTPS, mammalian 6-pyruvoyltetrahydropterinsynthase; BH4, 6R-L-erythro-5,6,7,8-tetrahydrobiopterin; PDB, Protein DataBank.

THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 289, NO. 34, pp. 23641–23652, August 22, 2014© 2014 by The American Society for Biochemistry and Molecular Biology, Inc. Published in the U.S.A.

AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23641

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 2: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

CPH4 synthase is promiscuous and in addition to H2NTPconverts sepiapterin and PPH4 to CPH4 (19). It is not clear howthis bacterial protein, which is closely related at the sequencelevel to its mammalian homolog involved in the biosynthesis ofBH4, catalyzes such distinct transformations. Herein we report astructural and functional investigation of CPH4 synthase from E.coli (QueD, so named for its role in the biosynthesis of queuosine).Comparison of the structure of QueD to its mammalian homologrevealed amino acid substitutions in the active site that mayaccount for the distinct catalytic outcome of the bacterial enzyme.Steady state and rapid kinetic analysis of wild-type CPH4 synthaseand its site-directed variants have led to insights regarding the roleof these residues in catalysis. This study highlights how minorchanges in this highly conserved active site have led to the emer-gence of catalytic promiscuity and the evolution of catalytic func-tion to support novel biosynthetic pathways.

EXPERIMENTAL PROCEDURES

Materials—All materials were purchased commercially(unless otherwise noted) and were of the highest purity. Allassays were carried out in a Coy anaerobic chamber in an atmo-sphere of 95–97% N2, 3–5% H2. All buffers and materials weredeoxygenated in the chamber several days prior to use.

Expression and Purification of E. coli QueD and Variants,mPTPS, and Sepiapterin Reductase—All proteins were expressedand purified as described previously (19). Selenomethionine-substituted QueD was obtained by growth in minimal mediausing a method to suppress methionine biosynthesis (20). Pro-tein concentration was determined by the Bradford methodusing bovine serum albumin as a standard.

Site-directed Mutagenesis of QueD—Variants of E. coli QueDwere prepared using the Stratagene QuikChange Site-directedMutagenesis Kit using the following primers and their comple-mentary primer (mutated bases underlined): C27A variant,5�-GAAGGGCATAAAGCTGGTCGCCTGCACGGG-3�; D70Nvariant, 5�-CCTACGAGCGCCTCAATCACCATTATCTC-3�; D70N/H71A variant (from the D70N variant sequence),5�-GAGCGCCTCAATGCCCATTATCTCAATG-3�; D54Nvariant, 5�-CGGGCTGGATTATCAATTTCGCTGAACT-AAAAGCGGCG-3�; H25A/D54N variant (from the D54Nvariant sequence), 5�-GTCCCGGAAGGGGCTAAATGTGG-TCGC-3�. The D70N/H71A mutations were introduced intothe H25A/D54N background by successive rounds of mutagen-esis with the primers above. The presence of the expectedmutations was confirmed by sequencing at the University ofMichigan DNA Sequencing Core Facility.

Crystallization and Structure Solution—QueD was crystal-lized by the hanging drop method. For each drop, 2 �l of proteinwas mixed with an equal volume of the precipitant and sus-pended over 0.5 ml of the precipitant solution. Crystals grewunder these conditions overnight. Crystals of the selenome-thionine-substituted protein (20 mg/ml in 20 mM HEPES (pH7.5), 10 mM dithiothreitol)(PDB codes 4NTN and 4NTM) wereobtained with 5% (v/v) 2-methyl-1,3-propanediol, 0.04 M

CaCl2, 0.1 M sodium acetate (pH 4.5) as precipitant. For theC27A variant (16 mg/ml, 20 mM HEPES (pH 7.5), 10 mM dithio-threitol), the precipitant solution contained 4% (v/v) polyethyl-ene glycol (PEG) 4000 and 0.1 M sodium acetate (pH 4.5). Sepi-apterin- or CPH4-bound structures were obtained by soakingcrystals for 1 h in a solution containing 7 mM sepiapterin and17% (v/v) of appropriate precipitant (either PEG4000 or2-methyl-1,3-propanediol). Crystals were transferred into asolution containing 35% of PEG 4000 or 2-methyl-1,3-propane-diol and frozen in liquid nitrogen.

Diffraction data were collected at SSRL on beam lines 9-2 and11-1, and the data were reduced and scaled using Crystal Clear(unliganded selenomethionine-substituted protein) or XDS(21, 22). The structure of the unliganded, selenomethionine-substituted protein was solved by molecular replacement withPHASER (23) using PDB structure 2OBA as the search model.Liganded structures were solved in the same manner using theunliganded protein structure (PDB code 4NTN) as the searchmodel. The structures were refined using REFMAC5 (24) andrebuilt using COOT (25). TLS parameters were refined (oneTLS group per protein chain) for the liganded and unligandedselenomethionine protein structures. Only protein residueswere included in the TLS groups (26). No � cutoff was used inthe refinement. All other calculations were performed using pro-grams from the CCP4 package (27). Coordinates for unligandedselenomethionine-substituted QueD (PDB code 4NTN), sel-enomethionine-substituted protein soaked with sepiapterin (PDBcode 4NTM), and the C27A variant of QueD soaked with sepiap-terin (PDB code 4NTK) were deposited in the Protein Data Bank.Structure figures were prepared using PYMOL (28).

Enzymatic Preparation and Purification of H2NTP—H2NTPwas produced as reported previously (29). The lyophilizedmaterial was dissolved in deoxygenated water in the anaerobicchamber and aliquots were frozen at �80 °C.

HPLC Assays to Detect Turnover of QueD and mPTPS withH2NTP and Sepiapterin—The assays (0.1 ml) contained 20 mM

PIPES/NaOH (pH 7.4), 10 mM dithiothreitol, 10 mM MgCl2,

O NN

N NHPPPO

N

NHN

NHOHHO

O

NH2

O

NH2OH

PPPO

OH N

NH

HN

NH

O

NH2

O

-O

O

H

PPP+

N

NH

HN

NH

O

NH2O

O

N

NH

O

NH2NH

-OO

N

NH

HN

NH

O

NH2OH

OH

OOH N

NH

HN

N

O

NH2

O

-O

O

H

PPP+

NNH

NH2OH

N

NH

HN

NH

O

NH2O

O

preQ0

tetrahydrobiopterin

O

O-

H2NTP

PPH4

CPH4

GCH I

GTP

QueD

mPTPS

H

FIGURE 1. QueD homologs catalyze distinct reactions in mammals and bacteria. Bacterial QueD catalyzes the second step in the biosynthesis of 7-deaza-purine containing compounds, including toyocamycin and the modified tRNA base queuosine (18). mPTPS catalyzes the second reaction in the biosynthesisof tetrahydrobiopterin.

Catalytic Promiscuity in the Tunnel-fold

23642 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289 • NUMBER 34 • AUGUST 22, 2014

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 3: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

and 50 �M substrate and were initiated by the addition of 10 �M

enzyme. The reactions were quenched after 1 h by addition of30% (w/v) trichloroacetic acid to a final concentration of 10%(v/v). Precipitated protein was removed by centrifugation at13,000 � g. The supernatant (75 �l) was placed into a glass vialthat was sealed inside the anaerobic chamber and analyzed byan ion-pairing HPLC method described previously (19). Theeluents were monitored using a diode array detector and thechromatograms were analyzed for appearance of CPH4 (298nm) or disappearance of sepiapterin (420 nm) and H2NTP (330nm), as appropriate.

Steady State Kinetics Experiments with QueD—HPLC-basedassays to determine steady state kinetic parameters for the con-version of H2NTP to CPH4 were carried out as described above.Velocities were obtained from the linear portion of the reactionat each concentration of H2NTP. The enzyme concentrationfor the assays was in the 0.1–1 �M range. Steady state analysiswith sepiapterin as substrate was carried out in the same man-ner as the steady state H2NTP assays except that conversion ofsepiapterin to CPH4 was monitored (in a total assay volume of 1ml) by the change in absorbance at 420 nm (�420 � 10.4 mM�1

cm�1) (30).Single Turnover Stopped-flow UV-Visible Spectroscopy—All

single turnover experiments were carried out using a BioLogicSFM 400 fitted with a stopped-flow head; a stream of nitrogengas maintained anaerobicity in the syringe compartment. Allsolutions were prepared in the anaerobic chamber and loadedinto syringes before being transferred to the instrument. Thesolutions were the same composition as described for the aboveHPLC assays except one syringe contained substrate, whereasthe other contained enzyme, and the two were mixed in a 1:1ratio. To maintain a single turnover regime, the substrate con-centration (50 �M after mixing) was kept below the enzymeconcentration (250 �M after mixing) in all experiments. Dataobtained with the diode array detector in the 259 –701 nmrange were analyzed using KinTek Explorer software (version3.0) (31, 32). In each case, the data were fit to one or moreexponentials (as described under “Results”) to analyze the spec-tral changes corresponding to turnover.

Single Turnover Quenched-flow Experiments—These experi-ments were carried out essentially as described above for thestopped-flow experiments with the exception that the reactionswere quenched with 30% (w/v) trichloroacetic acid to a final con-centration of 10% (v/v). Samples recovered after the quencheswere analyzed by HPLC as described above.

Assays with mPTPS—Reactions were conducted in the samemanner as the HPLC-monitored experiments described forturnover of E. coli QueD with H2NTP or sepiapterin at concen-trations of mPTPS and substrate indicated in the figure legends.

RESULTS

Overall Structure of QueD—A ribbon representation of wild-type QueD is shown in Fig. 2, and refinement statistics are sum-marized in Table 1. QueD is a homohexamer with each mono-mer containing 121 amino acids (33). QueD crystallizes withthe biological assembly, a homohexamer in the asymmetricunit. As with the mPTPS homolog (1, 34), QueD adopts a tun-nel-fold common to proteins that bind purine and pterin sub-

strates. The structure consists of four sequential antiparallel�-strands with two antiparallel �-helices on the concave face ofthe �-sheet between the 2nd and 3rd strands (������). Thesubunits assemble into a hollow barrel with an approximate3-fold axis down the center of the cavity and approximate2-fold axes in the equatorial plane (Fig. 2). In the structureswe are reporting, the packing of hexamers in the unit cell differsbetween crystal forms. The C27A variant crystallized such thatthe 3-fold axis of the barrel is parallel to the crystallographic aaxis for all hexamers. For selenomethionine-substituted pro-tein, which crystallized under different conditions, the hexamer3-fold axis is not parallel to a crystallographic axis and adjacenthexamers are oriented with the 3-fold axes rotated by 90degrees.

The Active Site Architecture of QueD—The six active sites ofQueD are located at the interface of the monomers that com-prise the homohexamer (Fig. 2). As expected from the sequencesimilarity to the mPTPS homolog, and from biochemical stud-ies showing the presence of �1.1 eq of zinc divalent cation permonomer (19), we observe electron density consistent with thecation in each active site. The catalytic zinc ions (shown asgreen spheres) are positioned near the equator and toward theoutside of the assembly, near the confluence of three proteinchains. As with mPTPS, the zinc divalent cation is coordinatedby the imidazole side chains of three histidine residues (His16,His31, and His33) from a single subunit (35). A water moleculeoccupies the fourth coordination position of the zinc cation. Asin mPTPS, the essential Cys27 residue (Cys42 in mPTPS) in theactive site of QueD is in a position to interact with an His71-Asp70 dyad (His89-Asp88 in mPTPS) activating it to catalyzeproton abstraction from the substrate to initiate catalysis (29,35). Interestingly, a second unique dyad, His25-Asp54, is alsopresent in QueD, which we hypothesize is responsible for pro-moting the novel retroaldol cleavage to form CPH4 by addi-tional interactions with Cys27.

In addition to the structure of the wild-type enzyme, we alsodetermined the structure of QueD with the product CPH4 inthe active site (PDB code 4NTM), and the QueD variant C27Awith the substrate sepiapterin in the active site (PDB code4NTK) (Table 1). Illustrations of the active site residues andcorresponding ligand electron density are shown in Fig. 3, Aand B. The electron density that is observed when sepiapterin issoaked into selenomethionine-substituted QueD is consistentwith formation of the product CPH4 that clearly has a shorter sidechain than sepiapterin (Fig. 3A), demonstrating that the crystalsare catalytically active and convert sepiapterin to product.

By contrast, electron density after soaking sepiapterin intocrystals of the catalytically inactive C27A variant shows the sidechain to be long enough to be uncleaved (Fig. 3B). Therefore, weinitially built sepiapterin into the electron density in each of theactive sites. However, after refinement, the terminal methylgroup of the side chain was out of the electron density in all sixactive sites. Moreover, a negative difference density peakappeared centered on the methyl carbon in several, but not allof the active sites. The lack of electron density for the methylgroup and presence of difference peaks persisted through sev-eral refinement cycles. Although kinetic results (discussedbelow) led us to believe that no reaction should occur in crystals

Catalytic Promiscuity in the Tunnel-fold

AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23643

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 4: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

of the C27A variant, we replaced sepiapterin with CPH4 in themodel. Two positive difference peaks appeared, one 2.3 Å fromthe zinc atom, corresponding to the 2�-oxygen of the sepiap-terin side chain, and the second in the plane of the O-C-C-O ofthe side chain corresponding to the methyl group. This place-ment of the methyl carbon requires the second carbon to be sp2

hybridized. Subsequent refinement modeling the ligand as theenol form of sepiapterin resulted in the loss of the differencedensity peaks and led to electron density that covered themethyl group, which is planar with the OCCO of the side chain.The electron density for the methyl carbon is slightly weakerthan that of the surrounding atoms, which may imply somestructural heterogeneity in the bound ligand. Therefore, itappears that whereas the C27A variant is catalytically inac-tive, the predominant form of bound sepiapterin is thedeprotonated enolate form with both oxygen atoms coordi-nated, asymmetrically, to the zinc cation. Stopped-flow evi-dence supporting this tautomerization of sepiapterin will bepresented below.

The bound ligands have an extensive array of hydrogenbonds that satisfy each of the allowed hydrogen bond donor/acceptor sites of the pyrimidine ring (see Fig. 4). All six activesites appear to be fully occupied with ligand in both soaked

structures. No large protein conformational changes occurupon sepiapterin binding the protein. The core root meansquare deviation is 0.42 Å when unliganded and liganded sel-enomethionine structures are overlaid using the SSM algo-rithm as implemented in COOT (36).

Although ligand binding induces no overall changes, thereare active site adjustments when ligands bind. The largestmovements are of the side chains of Glu54 and Phe55, whichmove �2 Å into the active site. The side chain of Glu54 forms ahydrogen bond with N1 of the pterin ring and Phe55 stacks withthe pterin ring of the ligand. Crystal contacts for the selenome-thionine protein�CPH4 complex are different from those of theC27A protein�sepiapterin complex, so these small conforma-tional changes that are the same in both structures are presum-ably not an artifact of crystal contacts. Structures of unligandedand sepiapterin-soaked native QueD have been deposited in theProtein Data Bank (PDB codes 3QN9, 3QN0, and 3QNA).These structures show the same small adjustments in active siteconformation upon ligand binding. It is important to note thatsepiapterin is not the natural substrate of the enzyme and isused as a model for the second step in the reaction catalyzed byQueD, as described in this article. Therefore, we cannot excludethe possibility that there may be more significant conforma-

90

H16

H33

H31 H71

D70

E110

C27H25

D54

E86

FIGURE 2. The ribbon representation of QueD. The six active sites in the homohexamer lie at trimer interfaces and house a catalytic zinc divalent cation (greensphere). A close-up of the active site is shown to illustrate the contributions from adjacent subunits.

Catalytic Promiscuity in the Tunnel-fold

23644 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289 • NUMBER 34 • AUGUST 22, 2014

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 5: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

tional changes that occur on native substrate binding duringthe first half-reaction.

The other significant difference in the active site is rotation ofthe Cys27 sulfur toward the position of His25 causing its rotationaway from the active site. Both Cys27 and His25 occupy alternateconformations in the CPH4-containing crystal (Fig. 3A); in con-formation A, the sulfur of Cys27 and the nitrogen of His25 areseparated by 3.4 Å. In contrast, the distance between theseatoms is 4.1 Å in the unliganded structure.

The structural data highlight the substantial similarities ofmPTPS and QueD, as well as the conspicuous His25-Asp54 dyad

in the vicinity of the substrate, which may be driving the func-tional differences. In the remainder of this article we focus onbiochemical studies of residues Cys27, His25, Asp54, Asp70, andHis71 to better understand their role in catalysis.

Steady State Kinetic Characterization of QueD and Site-di-rected Variants—To understand how the difference in reactiv-ity between QueD and mPTPS may relate to the structural dif-ferences, we examined the effect of mutations to either theconserved or unique dyad on the activity of QueD. In initial endpoint assays to assess activity, QueD was incubated with sepi-apterin or H2NTP and the reaction mixtures were examined by

TABLE 1Data collection and refinement statistics

Data collection statisticsPDB ID 4NTN 4NTM 4NTKModification, ligand SeMet, no ligand SeMet, CPH4 C27A, sepiapterinBeamline SSRL 9–2 SSRL 9–2 SSRL 11–1Wavelength (Å) 0.97915 1.0000 0.9795Crystal class, space group Orthorhombic, I222 Tetragonal, P41212 Orthorhombic, P212121Z 6 6 6Unit cell parameters

a (Å) 107.08 111.30 58.07b (Å) 111.76 111.30 112.60c (Å) 161.44 126.05 115.09

Resolution limits (Å) 56.0–2.00 (2.10–2.00)a 39.0–2.05 (2.10–2.05) 38.4–1.60 (1.69–1.60)No. independent reflections 66,155 (9,592) 50,332 (3,709) 98,203 (13,940)Completeness (%) 99.9 (100.0) 99.6 (95.6) 98.4 (97.0)Redundancy 4.9 (4.7) 7.2 (6.9) 3.3 (3.4)Avg I/Avg �(I) 9.2 (2.0) 33.2 (4.2) 18.1 (2.1)Rsym

b 0.087 (0.635) 0.031 (0.577) 0.033 (0.546)Rpim

c 0.015 (0.349) 0.022 (0.511)Refinement statistics

Resolution limits (Å) 55.90–1.99 33.92–2.05 36.34–1.60R-factord (overall)/No. reflections 0.204/66142 0.226/50256 0.187/98125R-factor (working)/No. reflections 0.203/62782 0.225/47712 0.185/93200R-factor (free)/No. reflections 0.222/3360 0.255/2544 0.217/4925R.m.s. deviation

Bond lengths (Å) 0.006 0.006 0.014Bond angles (°) 1.145 1.13 1.725General planes (Å) 0.004 0.010 0.011

Ramachandran plote

Favored region (%) 98.9 97.8 98.3Allowed region (%) 1.1 2.1 1.7Outliers (No. of residues) 0 1 0

Average B valuesProtein atoms (Å2) 50.5 61.1 27.9Ligand (Å2) 52.7 47.5 27Solvent (Å2) 50.75 48.3 37.8Wilson B factor (Å2) 27.3 49.2 24

a Numbers in parentheses refer to the respective highest resolution data shell in the data set.b Rsym � (��I � I��/�I).c Rpim � ((1/(n-1))1/2 ��I � I��/�I).d R-factor � (��Fo � Fc�/��Fo�), where Fo is the observed structure-factor amplitude and Fc is the calculated structure-factor amplitude.e As calculated using Molprobity (46).

E110

E86

D54

H25

C27

H31H33

H16

H71

D70

E110

E86

D54 H25

C27A

H31

H33H16

H71

D70

B

FIGURE 3. Structures of QueD with bound ligands. The electron density maps used to prepare the figures are those obtained after molecular replacementand refinement but before any ligand was built into the active sites. A, the active site of selenomethionine-substituted QueD with the product CPH4 bound. B,the active site of the C27A inactive variant of QueD. In B, sepiapterin soaked into the crystal has undergone deprotonation and tautomerization to the enolateform. For both figures, the Fo � Fc electron density maps are contoured at 3.0 �.

Catalytic Promiscuity in the Tunnel-fold

AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23645

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 6: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

HPLC for CPH4 production. The data (Fig. 5) show that, asexpected, the C27A variant of QueD cannot convert either sub-strate to CPH4 confirming the essential role of this residue incatalysis and the lack of turnover in the crystal (see above).Mutation of the conserved dyad (D70N/H71A) greatly dimin-ished CPH4 production from H2NTP, whereas mutation of theunique dyad (H25A/D54N) completely abolished this activity.Interestingly, both dyad mutants were able to produce CPH4from sepiapterin. The quadruple variant (H25A/D54N/D70N/H71A), where both dyads are deleted, is completely inactiveunder all conditions tested.

Steady state kinetic analyses of QueD and variants were car-ried out with H2NTP and sepiapterin to quantify the roles of theconserved residues. Unfortunately, we were not able to obtainreliable rates at very low concentrations of substrate, hamper-ing determination of accurate Km values. Therefore, for theforegoing discussion of the stopped-flow data, we only use thedata in Fig. 6 to estimate the lower limit of the turnover numberof wild-type enzyme with H2NTP, and of wild-type H25A/D54N and D70N/H71A variants with sepiapterin. The C27Avariant is catalytically inactive under all conditions examined.

Pre-steady State Kinetic Characterization of Wild-type QueDand Site-directed Variants—To gain additional insights into therole(s) of the conserved residues, single turnover stopped-flowexperiments were undertaken with H2NTP and sepiapterin assubstrates. In experiments where QueD is mixed with H2NTP,an intermediate with a �max at �440 nm builds in an interval of4.5 s after mixing and disappears in the next �50 s (Fig. 7A).The data are consistent with a model that included steps forformation and disappearance of an intermediate. The data arebest fit with three exponentials, the first two corresponding tothe formation (1.5 and 0.6 s�1) and the third to the disappear-ance (0.1 s�1) of the 440 nm transient (Fig. 7B). Because theturnover number of the enzyme with H2NTP is 0.013 s�1 thetwo phases are fast enough to correspond to formation anddisappearance of a kinetically competent intermediate.

Quenched-flow experiments were carried out to furtherprobe the nature of the 440-nm intermediate. In these experi-ments, QueD was mixed with H2NTP and samples were

quenched at various times (0.5 to 60 s) after mixing and ana-lyzed by HPLC. The samples reveal clear accumulation anddisappearance of a peak at �8.0 min whose UV-visible spec-trum and retention time are identical to those of commerciallyavailable sepiapterin. The intermediate reaches maximal con-centration in nearly the same time frame (14 s) as the stopped-flow detected 440-nm intermediate (4.3 s), strongly supportingthe notion that this intermediate is sepiapterin or a close struc-tural analog.

The C27A variant is clearly inactive in overall turnover (Fig.5). Nevertheless, we carried out a stopped-flow analysis of thisvariant with H2NTP and sepiapterin. We observe no detectablechange in the UV-visible spectrum with H2NTP. However, theUV-visible spectrum of sepiapterin undergoes a red shift uponmixing with the enzyme leading to a 440-nm species, which isindistinguishable from that formed during turnover withH2NTP (Fig. 8A). The spectral change at 440 nm is fit to a singleexponential yielding a rate constant of �0.4 s�1 (Fig. 8B).Therefore, whereas the variant is inactive with respect to over-all turnover, it is capable of binding sepiapterin and catalyzingformation of the transient that we observed with H2NTP andthe catalytically active QueD variants. The slower rate constantfor formation of sepiapterin observed in the rapid quenchassays with H2NTP as substrate may reflect the rate of conver-sion of this transient enolate species to sepiapterin in solutionupon being released from the active site after the quench.

The two dyad variants are differentially affected in turnoverwith H2NTP. The unique dyad H25A/D54N variant clearlyforms the same 440-nm intermediate, but this intermediate isnot turned over further (Fig. 7B). Analysis of the data at 440 nmreveals two phases with rate constants of 0.23 and 0.05 s�1.Therefore, the absence of production of CPH4 with the uniquedyad (Fig. 5) indicates loss of ability to convert the 440-nmtransient to product and not a loss of catalytic activity. We wereunable to identify distinct intermediates by HPLC with thisvariant, presumably because of enzymatic or non-enzymaticconversion of the unstable sepiapterin-like intermediate toother compounds. By contrast, we did not observe any transientspecies accumulate when the conserved dyad D70N/H71A var-

HN

N NH

NO

H2N

O

O

O

O

O O

O

HN

NH

HS

HNN

HN

N

O

OZn2+

E110H16N

H31N

H33N

D70’

H71’

C27

H25

D54”I53”

F55”

E86

T84S85

3.1 Å

2.9 Å

2.7 Å 3.1 Å

3.2 Å

2.6 Å

2.9 Å

2.7 Å3.4 Å

2.7 Å

3.4 Å

3.9 Å

T105T106 E133

H23H50

H48 D88’

H89’

C42

N71”

L72”

M70”

E107

OO

2.0 Å

-

-

-

-

3.0 Å-3.3 Å

HN

N NH

NO

H2N

O OHO

O

O O

O

HN

NH

HNN

HN

N

O

OZn2+

E110H16N

H31N

H33N

D70’

H71’

C27A

H25

D54”I53”

F55”

E86

T84S85

2.8 Å

2.8 Å

2.6 Å 3.0 Å2.8 Å

2.9 Å

2.8 Å

3.1 Å

3.2 Å2.7 Å

T105T106 E133

H23H50

H48 D88’

H89’

C42

N71”

L72”

M70”

E107

-

OO

-

2.0 Å2.3 Å

CH3

-

-

-

H

FIGURE 4. Schematic representation of ligand interactions in QueD. A, the active site of QueD with bound CPH4. B, the active site of the C27A variant of QueDwith bound sepiapterin analog. For both figures, the E. coli QueD residues are in black numbering, whereas the corresponding residues in mPTPS are in gray.The �, , and unlabeled residue numbers show that the active site residues are contributed by three adjacent monomers.

Catalytic Promiscuity in the Tunnel-fold

23646 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289 • NUMBER 34 • AUGUST 22, 2014

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 7: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

iant was assayed with H2NTP as substrate in our stopped-flowanalysis (Fig. 7B).

The stopped-flow data with H2NTP as substrate support thehypothesis that a sepiapterin-like molecule is an intermediateof the reaction catalyzed by QueD. When the wild-type enzymeis mixed with sepiapterin, the 420-nm peak corresponding tothe substrate is lost with concomitant build-up of a species with

a �max at �340 nm (Fig. 9A). This 340-nm intermediate thendisappears along with the spectral features near 420 nm over�60 s. The fits of the data at 420 and 340 nm yield rate con-stants of 0.6 and 0.004 s�1 corresponding to conversion of sepi-apterin to the 340-nm intermediate and its subsequent conver-sion to product (Fig. 9B). As with turnover of the wild-typeenzyme with H2NTP, these rates are of the same order of mag-nitude as the turnover number of the enzyme with sepiapterinas substrate (0.01 s�1). Interestingly, the conserved dyad D70N/H71A variant exhibits nearly overlapping kinetic profiles thatare also fit by similar rate constants (0.4 and 0.1 s�1, respec-tively) (Fig. 9B). However, whereas a disappearance of the sub-strate (at 420 nm) is observed with the H25A/D54N variant,consistent with the fact that this variant turns over with sepi-apterin, no intermediates build up at 340 nm during the process(Fig. 9B).

Biochemical Characterization of the Promiscuity of mPTPS—Additional experiments were undertaken to confirm that theonly product of the mPTPS reaction with H2NTP was PPH4.Within 1 min of mixing 0.1 mM mPTPS with 0.1 mM H2NTP, weobserve a peak by HPLC for PPH4 at �5.8 min. PPH4 is also asubstrate for QueD and we do not observe evidence for a similarpeak in any of our studies with QueD. Surprisingly, however, inreactions with mPTPS we observe that in addition to PPH4,CPH4 also forms within the first 5 min and continues to builduntil all the PPH4 is exhausted (Fig. 10A). To confirm that PPH4is the source of the resulting CPH4, the reaction was coupled tosepiapterin reductase. Sepiapterin reductase catalyzes the con-version of PPH4 to tetrahydrobiopterin. Under these condi-tions, CPH4 formation is inhibited and only tetrahydrobiop-terin forms. Moreover, to confirm that the CPH4-formingactivity is enzymatic and does not result from degradativebreakdown of PPH4, we carried out the same reaction in thepresence of 10-fold less mPTPS. Although PPH4 formation wascomplete within the same time frame in this experiment, CPH4formed at a 10-fold reduced rate (data not shown). Finally,mPTPS assayed with sepiapterin also exhibited the capacity toform CPH4 (Fig. 10B). The mPTPS used here is His6-tagged andwas purified by affinity chromatography, whereas QueD used inall the experiments in this article was not and was purified byanion exchange and hydrophobic interaction chromatographysteps. One can envision that mPTPS expressed in E. coli may becontaminated by a small amount of endogenous QueD. How-ever, the unique subunit arrangement that produces the activesite, which is contributed by residues from 3 adjacent subunits,require that the well expressed recombinant protein incorporateat minimum two subunits of endogenous QueD in the proper ori-entation. Finally, mPTPS has been shown to complement E. coliQueD (37). This observation could not be explained previouslybut it is compatible with the biochemical data shown here. Ourobservations support the notion that the capacity for synthesisof CPH4 is a promiscuous activity that is present in mPTPS andthat the unique dyad simply accelerates the production of thistransformation in QueD.

DISCUSSION

mPTPS and QueD both utilize H2NTP as substrate, butdespite a strikingly similar overall-fold and active site architec-

0 5 10 150 5 10 15

H2NTP Sepiapterin

Control

WT

C27A

D70N/H71A

H25A/D54N

Time (min) Time (min)

CPH4 CPH4H2NTP Sepiapterin

D70N/H71AH25A/D54N

FIGURE 5. QueD and variants have differential activity toward sepiap-terin and H2NTP. Under the HPLC conditions employed, CPH4, sepiapterin,and H2NTP elute at 3.8, 8.0, and 15.5 min, respectively. Note that in the chro-matograms H2NTP is detected at 330 nm, whereas CPH4 is detected at 298nm. Traces of the reaction, except for the H2NTP control, are shown at this�max of CPH4 for clarity. The trace for H2NTP as a control was generated with-out the quench due to the instability of H2NTP in acidic conditions.

Catalytic Promiscuity in the Tunnel-fold

AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23647

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 8: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

ture they produce distinctly different major products. Theactive sites of both proteins are composed of a constellation ofresidues that are contributed by adjacent subunits in the bio-logical assembly (34). The substrate in each active site is boundto a conserved zinc divalent cation via the C1� and C2� hydroxylgroups and the proteins retain similar binding interactions with

the substrate (35). Each active site houses an essential cysteineresidue whose activation with a conserved Asp/His dyad is pro-posed to initiate catalysis (29). The most conspicuous structuraldifference between the two active sites is the presence of anadditional His/Asp dyad in the bacterial enzyme on the oppo-site face of the Cys27 residue relative to the conserved dyad.

D70N/H71A

Sepiapterin (µM)

Wild Type

H25A/D54N

Wild Type

k cat (

sec-1

)

0.003

0.007

0.010

0.013

0 20 40 60 80 100 120

0.003

0.007

0.010

0.013

0.017

0 20 40 60 80 100 120

Sepiapterin (µM)Sepiapterin (µM)

H2NTP (µM)

k cat (

sec-1

)k ca

t (se

c-1)

k cat (

sec-1

)

0 20 40 60 80 100 1200 20 40 60 80 100 120

0.007

0.005

0.003

0.001

2.5X10-4

1.7X10-4

8.3X10-5

DC

FIGURE 6. Steady state kinetic analysis of: A, QueD with H2NTP as substrate; B, QueD with sepiapterin as a substrate; C, D70N/H71A variant of QueDwith sepiapterin as a substrate; D, H25A/D54N variant of QueD with sepiapterin as substrate.

0

0.10

0

0.10

300 350 400 450 500 550

Abs

orba

nce

0.035 sec

4.25 sec

60 sec

4.25 sec

0.43

0.44

0.45

0.46

0.47

0.48

0.49

0.50

0.001 0.01 0.1 1 10 100

Time (sec)

A4

40

nm

0.05

0.05

FIGURE 7. Pre-steady state analysis of the reaction catalyzed by QueD and variants with H2NTP as substrate. A, UV-visible spectra from a single turnoverstopped-flow reaction of QueD with H2NTP. There is a clear buildup and disappearance of an intermediate at �440 nm that reaches a maximum absorbanceat �4.25 s. The initial trace is in black, whereas the final trace is in red within the corresponding time frame. B, the UV-visible spectra from a single turnoverstopped-flow experiment of wild-type and variants of QueD reacted with H2NTP and monitored at 440 nm over time. The trace for the wild-type enzyme is inblue, whereas the traces for the H25A/D54N variant and D70N/H71A variants are in orange and green, respectively. There is formation and disappearance of the440-nm intermediate described in A for wild-type QueD. Fits to the data obtained from kinetic analysis are overlaid in black.

Catalytic Promiscuity in the Tunnel-fold

23648 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289 • NUMBER 34 • AUGUST 22, 2014

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 9: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

Therefore, we investigated site-directed variants of QueD todetermine the contribution(s) of each dyad to the functionaldifferences between QueD and mPTPS.

Although the biological role of QueD is to convert H2NTP toCPH4, we have shown previously that the enzyme also utilizessepiapterin as substrate converting it to CPH4 (19). Our work-ing hypothesis was that the reaction likely involves intermedi-ates that resemble this alternate substrate. Indeed time-re-solved studies of QueD reveal the build up and disappearance ofa kinetically competent intermediate with a �max of 440 nm.The identity of the intermediate is difficult to establish unam-biguously; however, rapid quench studies reveal transient for-

mation of sepiapterin in the same time frame as that of the440-nm transient. The differences between the 440-nm tran-sient observed by stopped-flow and the solution spectrum ofsepiapterin may represent tautomerization to the enolate form,which is also observed in the crystal structure of the C27A var-iant, despite the inability of this variant to catalyze the overallreaction.

We hypothesize that the role of the conserved dyad is toinitiate catalysis with H2NTP as substrate, as mutation of theseresidues leads to �80-fold decrease in overall turnover withH2NTP without effecting turnover parameters regarding catal-ysis with sepiapterin. By contrast, mutation of the unique dyad

0

0.10

0.20

300 350 400 450 500

0.035 sec

12.0 sec

Abs

orba

nce

0.48

0.49

0.50

0.51

0.52

0.53

0.54

0.55

0.56

0.001 0.01 0.1 1 10 100Time (sec)

A4

40

nm

FIGURE 8. Pre-steady state analysis of sepiapterin binding to the C27A variant of QueD. A, UV-visible spectra from a single turnover stopped-flow reactionshow binding of sepiapterin to the C27A variant of QueD. Upon binding to the C27A variant, the spectrum of sepiapterin undergoes a red shift as well as achange in extinction coefficient. The initial trace is in black, whereas the final trace is in red within the corresponding time frame. B, the UV-visible spectra fromthe same single turnover stopped-flow experiment monitored at 440 nm over time. The spectral change is best fit by a single exponential with a rate constantof 0.4 s�1. Fits to the data obtained from the kinetic analysis are overlaid in black.

0

0.14

0

0.12

300 350 400 450 500

Wavelength (nm)

Abs

orba

nce

0.035 sec

6 sec

60 sec

6 sec

0.10

A42

0 nm

0

0.05

0.001 0.01 0.1 1 10 100

Time (sec)

A34

0 nm

0

0.06

0.06

FIGURE 9. Pre-steady state analysis of the reaction catalyzed by QueD and variants with sepiapterin as substrate. A, UV-visible spectra from a singleturnover stopped-flow reaction of QueD with sepiapterin. There is a loss of absorbance at 420 nm corresponding to the loss of sepiapterin substrate with aconcomitant buildup of an intermediate at �340 nm during the first 6 s. Then, both disappear over the remainder of the experiment. The initial trace is in black,whereas the final trace is in red within the corresponding time frame. B, the UV-visible absorbance from a single turnover stopped-flow experiment of QueD andsite-directed variants reacted with sepiapterin and monitored at either 420 or 340 nm over time. The trace for wild-type QueD is in blue, whereas the data forthe H25A/D54N and D70N/H71A variants are shown in orange and green, respectively. Both QueD and the D70N/H71A variant turnover sepiapterin and buildand resolve the 340-nm intermediate in a similar fashion. Conversely, the H25A/D54N variant slowly depletes the initial sepiapterin substrate and does notbuild an appreciable 340-nm intermediate. Fits to the data obtained from the global kinetic analysis are overlaid in black.

Catalytic Promiscuity in the Tunnel-fold

AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23649

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 10: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

leads to a �160-fold drop in activity of the enzyme with sepi-apterin and buildup of the 440-nm intermediate when assayedwith H2NTP, which is not carried through further to CPH4. Asdescribed above, the time-resolved data clearly show that asepiapterin analog is an intermediate in the reaction. Thisobservation, in the context of the fact that sepiapterin is a sub-

strate, suggests that the conversion of H2NTP to CPH4 is amultistep process. Although we cannot assign specific roles tothese residues, the dyads clearly have differential contributionsat distinct stages in the catalytic cycle.

A mechanistic paradigm for QueD is shown in Fig. 11. Theproposed mechanism has many features in common with those

0 min

1 min

2 min

5 min

10 min

15 min

20 min

30 min

40 min

60 min

90 min

120 min

CPH4

PPH4

0 5 10 15

CPH4Sepiapterin

Time (min)3 4 5 6 7

Time (min)

A B

FIGURE 10. mPTPS can utilize multiple substrates. A, in addition to catalyzing the reaction of H2NTP to PPH4, mPTPS can also convert PPH4 to CPH4. H2NTP combinedwith mPTPS is completely converted to PPH4 (5.8 min on chromatogram) within 1 min after initiation of the reaction, and after �5 min CPH4 (3.8 min on chromato-gram) begins to appear and builds over the course of the next 2 h. H2NTP and mPTPS were assayed at equimolar concentrations (0.1 mM). B, mPTPS can also convertsepiapterin to CPH4 in the same HPLC assay regime as detailed in the legend to Fig. 5. The top trace is the control assay without enzyme, and the bottom trace is thecomplete activity assay.

NH

N O

O OH

NH

N

O O

NH

N

O O

S

H

H

NH

N

O OH

NH

N

O OH

NH

N

O O

NH

HN

O

NH

HN

O

OH2

H

O

O

Zn Zn Zn

Zn Zn Zn

Zn Zn

- - -

--

Cys27

H

His71

S

Cys27

HHis25

OH

-

2+ 2+ 2+

2+2+2+

2+ 2+

P P P

O P P P

H

Asp70

Asp54

H

OH

-

FIGURE 11. A working model for the mechanism of conversion of H2NTP to CPH4 by QueD.

Catalytic Promiscuity in the Tunnel-fold

23650 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289 • NUMBER 34 • AUGUST 22, 2014

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 11: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

proposed for mPTPS. Specifically, both enzymes have an essen-tial active site cysteine residue, a divalent zinc ion, and a con-served dyad within hydrogen bonding distance to the con-served Cys. The first step in the reaction is the binding of thesubstrate to the active site via hydrogen bonding and coordina-tion of the 1�- and 2�-hydroxyls to the active site zinc divalentcation. In the structure of the C27A variant of QueD complexedwith sepiapterin, the hydroxyl oxygen atoms are bound asym-metrically and at distances of �2 and � 2.3 Å from the zinc ion.In the catalytic cycle the required Cys residue, which is acti-vated by interaction with the conserved Asp70-His71 dyad,abstracts a proton from C2� eliminating the triphosphate and issubsequently tautomerized to a sepiapterin-like intermediate.The unique dyad, either directly or in concert with the catalyt-ically essential Cys27, activates a water molecule for the nexthalf-reaction, which entails the elimination of the C2�-C3� asacetaldehyde to form the CPH4 product. We favor a role forCys27 in this half-reaction as well because the C27A variant iscompletely inactive in conversion of sepiapterin to CPH4.Again, the C27A variant appears to catalyze a change in theconjugation of sepiapterin upon binding leading to a 440-nmspecies, corroborating the planar geometry for sepiapterinobserved in the x-ray crystal structure of the C27A variant. Thismechanism accounts for all our observations on QueD as wellas the literature on the mechanism of the mPTPS enzyme(38 – 42).

Recent studies have shown that the mPTPS enzyme cancomplement production of queuosine in a queD strain ofE. coli (37). This has been somewhat puzzling, as mPTPS hadnever been shown to catalyze the conversion of H2NTP toCPH4. Our biochemical data with purified mPTPS provide anexperimental basis of the complementation. As shown in Fig.10, in addition to converting H2NTP to PPH4, mPTPS also cat-alyzes conversion of PPH4 to CPH4, albeit at vastly slower rates.The extent of modification of tRNA to queuosine under normalgrowth conditions is not known, but the overexpression of themammalian protein is likely to produce the necessary pool ofCPH4 to support the tRNA modification. As is clear in the pro-posed mechanism, the active site zinc cation should be able topromote a number of tautomerizations and one can readilypropose mechanisms by which zinc-bound PPH4 would be con-verted to CPH4. We note that whereas the unique dyad of QueDappears to be necessary for conversion of sepiapterin to CPH4,it is not essential. The active site of mPTPS appears to have beenset up to permit zinc-mediated tautomerizations and smallvariations in active site environment, such as the introductionof the unique dyad, are sufficient to amplify rates of the reac-tions that produce CPH4.

The differences in reactivity between the highly similarmPTPS and bacterial QueD homologs provide an interestingcase of evolution of new catalytic activities in existing enzymefolds. The conversion of H2NTP to CPH4, albeit at very lowlevels, is clearly a promiscuous activity as the biosynthesis ofqueuosine is only carried out in prokaryotes. The concept ofenzyme promiscuity leading to novel metabolic pathways wasfirst proposed by Jensen (43) (for a review, see Ref. 44). Follow-ing a gene duplication event, enzymes that possess some low-level secondary activity will have a selective advantage toward

evolution of a new activity as it will theoretically require feweradvantageous mutations to elevate this ability to a prominentcomponent of a metabolic pathway. An enzyme possessing thislow-level catalytic promiscuity will therefore have an evolu-tionary “head start.” Recently, a model regarding the evolutionof new metabolic pathways, the innovation-amplification-di-vergence model, describes quite exquisitely how these newactivities could be borne of previous weak activities in ancestralenzymes (45). QueD and mPTPS may represent naturallyoccurring examples of the evolutionary models that posit pro-miscuous activity is required for emergence of new biologicalfunction(s).

Acknowledgments—Portions of this research were carried out at theStanford Synchrotron Radiation Lightsource, a Directorate of SLACNational Accelerator Laboratory and an Office of Science User Facil-ity operated for the United States Department of Energy Office ofScience by Stanford University. The SSRL Structural Molecular Biol-ogy Program is supported by the DOE Office of Biological and Envi-ronmental Research, and National Institutes of Health, NIGMSGrant P41GM103393.

REFERENCES1. Colloc’h, N., Poupon, A., and Mornon, J. P. (2000) Sequence and structural

features of the T-fold, an original tunnelling building unit. Proteins 39,142–154

2. Thöny, B., Auerbach, G., and Blau, N. (2000) Tetrahydrobiopterin biosyn-thesis, regeneration and functions. Biochem. J. 347, 1–16

3. Takikawa, S., Curtius, H. C., Redweik, U., Leimbacher, W., and Ghisla, S.(1986) Biosynthesis of tetrahydrobiopterin: Purification and characteriza-tion of 6-pyruvoyl-tetrahydropterin synthase from human liver. Eur.J. Biochem. 161, 295–302

4. Werner, E. R., Blau, N., and Thöny, B. (2011) Tetrahydrobiopterin: bio-chemistry and pathophysiology. Biochem. J. 438, 397– 414

5. Kaufman, S. (1993) New tetrahydrobiopterin-dependent systems. Annu.Rev. Nutr. 13, 261–286

6. Smith, G. K. (1987) On the role of sepiapterin reductase in the biosynthesisof tetrahydrobiopterin. Arch. Biochem. Biophys. 255, 254 –266

7. Switchenko, A. C., Primus, J. P., and Brown, G. M. (1984) Intermediates inthe enzymic synthesis of tetrahydrobiopterin in Drosophila melanogaster.Biochem. Biophys. Res. Commun. 120, 754 –760

8. Milstien, S., and Kaufman, S. (1989) Immunological studies on the partic-ipation of 6-pyruvoyl tetrahydropterin (2�-oxo) reductase, an aldose re-ductase, in tetrahydrobiopterin biosynthesis. Biochem. Biophys. Res. Com-mun. 165, 845– 850

9. Park, Y. S., Heizmann, C. W., Wermuth, B., Levine, R. A., Steinerstauch, P.,Guzman, J., and Blau, N. (1991) Human carbonyl and aldose reductases:new catalytic functions in tetrahydrobiopterin biosynthesis. Biochem. Bio-phys. Res. Commun. 175, 738 –744

10. Iino, T., Takikawa, S. I., Yamamoto, T., and Sawada, H. (2000) The enzymethat synthesizes tetrahydrobiopterin from 6-pyruvoyl-tetrahydropterin inthe lemon mutant silkworm consists of two carbonyl reductases. Arch.Biochem. Biophys. 373, 442– 446

11. Iino, T., Tabata, M., Takikawa, S., Sawada, H., Shintaku, H., Ishikura, S.,and Hara, A. (2003) Tetrahydrobiopterin is synthesized from 6-pyruvoyl-tetrahydropterin by the human aldo-keto reductase AKR1 family mem-bers. Arch. Biochem. Biophys. 416, 180 –187

12. Kong, J. S., Kang, J. Y., Kim, H. L., Kwon, O. S., Lee, K. H., and Park, Y. S.(2006) 6-Pyruvoyltetrahydropterin synthase orthologs of either a single ordual domain structure are responsible for tetrahydrobiopterin synthesis inbacteria. FEBS Lett. 580, 4900 – 4904

13. Choi, Y. K., Hwang, Y. K., and Park, Y. S. (2001) Molecular cloning anddisruption of a novel gene encoding UDP-glucose: tetrahydrobiopterin�-glucosyltransferase in the cyanobacterium Synechococcus sp. PCC 7942.

Catalytic Promiscuity in the Tunnel-fold

AUGUST 22, 2014 • VOLUME 289 • NUMBER 34 JOURNAL OF BIOLOGICAL CHEMISTRY 23651

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 12: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

FEBS Lett. 502, 73–7814. Lee, Y. G., Kim, A. H., Park, M. B., Kim, H. L., Lee, K. H., and Park, Y. S.

(2010) Molecular cloning of cyanobacterial pteridine glycosyltransferasesthat catalyze the transfer of either glucose or xylose to tetrahydrobiop-terin. Appl. Environ. Microbiol. 76, 7658 –7661

15. Chung, H. J., Kim, Y. A., Kim, Y. J., Choi, Y. K., Hwang, Y. K., and Park, Y. S.(2000) Purification and characterization of UDP-glucose:tetrahydrobiop-terin glucosyltransferase from Synechococcus sp. PCC 7942. Biochim. Bio-phys. Acta 1524, 183–188

16. Cho, S. H., Na, J. U., Youn, H., Hwang, C. S., Lee, C. H., and Kang, S. O.(1998) Tepidopterin, 1-O-(L-threo-biopterin-2�-yl)-�-N-acetylglucos-amine from Chlorobium tepidum. Biochim. Biophys. Acta 1379, 53– 60

17. McCarty, R. M., Somogyi, A., Lin, G., Jacobsen, N. E., and Bandarian, V.(2009) The deazapurine biosynthetic pathway revealed: in vitro enzymaticsynthesis of PreQ(0) from guanosine 5�-triphosphate in four steps. Bio-chemistry 48, 3847–3852

18. McCarty, R. M., and Bandarian, V. (2012) Biosynthesis of pyrrolopyrimi-dines. Bioorg. Chem. 43, 15–25

19. McCarty, R. M., Somogyi, A., and Bandarian, V. (2009) Escherichia coliQueD is a 6-carboxy-5,6,7,8-tetrahydropterin synthase. Biochemistry 48,2301–2303

20. Doublié, S. (1997) Preparation of selenomethionyl proteins for phase de-termination. Methods Enzymol. 276, 523–530

21. Pflugrath, J. W. (1999) The finer things in x-ray diffraction data collection.Acta Crystallogr. D Biol. Crystallogr. 55, 1718 –1725

22. Kabsch, W. (2010) Xds. Acta Crystallogr. D Biol. Crystallogr. 66, 125–13223. McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Sto-

roni, L. C., and Read, R. J. (2007) Phaser crystallographic software. J. Appl.Crystallogr. 40, 658 – 674

24. Murshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A.,Nicholls, R. A., Winn, M. D., Long, F., and Vagin, A. A. (2011) REFMAC5for the refinement of macromolecular crystal structures. Acta Crystallogr.D Biol. Crystallogr. 67, 355–367

25. Emsley, P., Lohkamp, B., Scott, W. G., and Cowtan, K. (2010) Features anddevelopment of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486 –501

26. Schomaker, V., and Trueblood, K. N. (1998) Correlation of internal tor-sional motion with overall molecular motion in crystals. Acta Crystallogr.B Struct. Sci. 54, 507–514

27. Winn, M. D., Ballard, C. C., Cowtan, K. D., Dodson, E. J., Emsley, P., Evans,P. R., Keegan, R. M., Krissinel, E. B., Leslie, A. G., McCoy, A., McNicholas,S. J., Murshudov, G. N., Pannu, N. S., Potterton, E. A., Powell, H. R., Read,R. J., Vagin, A., and Wilson, K. S. (2011) Overview of the CCP4 suite andcurrent developments. Acta Crystallogr. D Biol. Crystallogr. 67, 235–242

28. Delano, W. L. (2002) The PyMol Molecular Graphics System. Delano Sci-entific, San Carlos, CA

29. Ploom, T., Thöny, B., Yim, J., Lee, S., Nar, H., Leimbacher, W., Richardson,J., Huber, R., and Auerbach, G. (1999) Crystallographic and kinetic inves-tigations on the mechanism of 6-pyruvoyl tetrahydropterin synthase. J.Mol. Biol. 286, 851– 860

30. Matsubara, M., Katoh, S., Akino, M., and Kaufman, S. (1966) Sepiapterinreductase. Biochim. Biophys. Acta 122, 202–212

31. Johnson, K. A., Simpson, Z. B., and Blom, T. (2009) Global kinetic ex-plorer: a new computer program for dynamic simulation and fitting ofkinetic data. Anal. Biochem. 387, 20 –29

32. Johnson, K. A., Simpson, Z. B., and Blom, T. (2009) FitSpace explorer: an

algorithm to evaluate multidimensional parameter space in fitting kineticdata. Anal. Biochem. 387, 30 – 41

33. Reader, J. S., Metzgar, D., Schimmel, P., and de Crécy-Lagard, V. (2004)Identification of four genes necessary for biosynthesis of the modifiednucleoside queuosine. J. Biol. Chem. 279, 6280 – 6285

34. Nar, H., Huber, R., Heizmann, C. W., Thöny, B., and Bürgisser, D. (1994)Three-dimensional structure of 6-pyruvoyl tetrahydropterin synthase, anenzyme involved in tetrahydrobiopterin biosynthesis. EMBO J. 13,1255–1262

35. Bürgisser, D. M., Thöny, B., Redweik, U., Hess, D., Heizmann, C. W.,Huber, R., and Nar, H. (1995) 6-Pyruvoyl tetrahydropterin synthase, anenzyme with a novel type of active site involving both zinc binding and anintersubunit catalytic triad motif: site-directed mutagenesis of the pro-posed active center, characterization of the metal binding site and mod-elling of substrate binding. J. Mol. Biol. 253, 358 –369

36. Krissinel, E., and Henrick, K. (2004) Secondary-structure matching (SSM),a new tool for fast protein structure alignment in three dimensions. ActaCrystallogr. D Biol. Crystallogr. 60, 2256 –2268

37. Phillips, G., Grochowski, L. L., Bonnett, S., Xu, H., Bailly, M., Blaby-Haas,C., El Yacoubi, B., Iwata-Reuyl, D., White, R. H., and de Crécy-Lagard, V.(2012) Functional promiscuity of the COG0720 family. ACS Chem. Biol. 7,197–209

38. Deng, H., Callender, R., and Dale, G. E. (2000) A vibrational structure of7,8-dihydrobiopterin bound to dihydroneopterin aldolase. J. Biol. Chem.275, 30139 –30143

39. Ghisla, S., Kuster, T., Steinerstauch, P., Leimbacher, W., Richter, W. J.,Raschdorf, F., Dahinden, R., and Curtius, H. C. (1990) 1H NMR and massspectrometric studies of tetrahydropterins: evidence for the structure of6-pyruvoyl tetrahydropterin, an intermediate in the biosynthesis of tetra-hydrobiopterin. Eur. J. Biochem. 187, 651– 656

40. Le Van, Q., Katzenmeier, G., Schwarzkopf, B., Schmid, C., and Bacher, A.(1988) Biosynthesis of biopterin: studies on the mechanism of 6-pyru-voyltetrahydropteridine synthase. Biochem. Biophys. Res. Commun. 151,512–517

41. Maharaj, G., Selinsky, B. S., Appleman, J. R., Perlman, M., London, R. E.,and Blakley, R. L. (1990) Dissociation constants for dihydrofolic acid anddihydrobiopterin and implications for mechanistic models for dihydrofo-late reductase. Biochemistry 29, 4554 – 4560

42. Bracher, A., Eisenreich, W., Schramek, N., Ritz, H., Götze, E., Herrmann,A., Gütlich, M., and Bacher, A. (1998) Biosynthesis of pteridines: NMRstudies on the reaction mechanisms of GTP cyclohydrolase I, pyruvoyltet-rahydropterin synthase, and sepiapterin reductase. J. Biol. Chem. 273,28132–28141

43. Jensen, R. A. (1976) Enzyme recruitment in evolution of new function.Annu. Rev. Microbiol. 30, 409 – 425

44. O’Brien, P. J., and Herschlag, D. (1999) Catalytic promiscuity and theevolution of new enzymatic activities. Chem. Biol. 6, R91-R105

45. Näsvall, J., Sun, L., Roth, J. R., and Andersson, D. I. (2012) Real-time evo-lution of new genes by innovation, amplification, and divergence. Science338, 384 –387

46. Chen, V. B., Arendall, W. B., 3rd, Headd, J. J., Keedy, D. A., Immormino,R. M., Kapral, G. J., Murray, L. W., Richardson, J. S., and Richardson, D. C.(2010) MolProbity: all-atom structure validation for macromolecularcrystallography. Acta Crystallogr. D Biol. Crystallogr. 66, 12–21

Catalytic Promiscuity in the Tunnel-fold

23652 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 289 • NUMBER 34 • AUGUST 22, 2014

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from

Page 13: BiochemicalandStructuralStudiesof6-Carboxy-5,6,7,8 ... · 2, 3–5% H 2. All buffers and materials were deoxygenated in the chamber several days prior to use. Expression and Purification

Zachary D. Miles, Sue A. Roberts, Reid M. McCarty and Vahe BandarianTunnel-fold Superfamily

Synthase Reveal the Molecular Basis of Catalytic Promiscuity within the Biochemical and Structural Studies of 6-Carboxy-5,6,7,8-tetrahydropterin

doi: 10.1074/jbc.M114.555680 originally published online July 2, 20142014, 289:23641-23652.J. Biol. Chem. 

  10.1074/jbc.M114.555680Access the most updated version of this article at doi:

 Alerts:

  When a correction for this article is posted• 

When this article is cited• 

to choose from all of JBC's e-mail alertsClick here

  http://www.jbc.org/content/289/34/23641.full.html#ref-list-1

This article cites 45 references, 7 of which can be accessed free at

by guest on Novem

ber 30, 2020http://w

ww

.jbc.org/D

ownloaded from