Imidazole Glycerol Phosphate Synthase from Thermotoga maritima

QUATERNARY STRUCTURE, STEADY-STATE KINETICS, AND REACTION MECHANISM OF THE BIENZYME COMPLEX*

Silke Beismann-Driemeyer and Reinhard SternerDagger

From the Universität zu Köln, Institut für Biochemie, Otto-Fischer-Str. 12-14, D-50674 Köln, Germany

Received for publication, March 6, 2001, and in revised form, March 21, 2001

    ABSTRACT
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Imidazole glycerol phosphate synthase, which links histidine and de novo purine biosynthesis, is a member of the glutamine amidotransferase family. In bacteria, imidazole glycerol phosphate synthase constitutes a bienzyme complex of the glutaminase subunit HisH and the synthase subunit HisF. Nascent ammonia produced by HisH reacts at the active site of HisF with N'-((5'-phosphoribulosyl)formimino)-5-aminoimidazole-4-carboxamide-ribonucleotide to yield the products imidazole glycerol phosphate and 5-aminoimidazole-4-carboxamide ribotide. In order to elucidate the interactions between HisH and HisF and the catalytic mechanism of the HisF reaction, the enzymes tHisH and tHisF from Thermotoga maritima were produced in Escherichia coli, purified, and characterized. Isolated tHisH showed no detectable glutaminase activity but was stimulated by complex formation with tHisF to which either the product imidazole glycerol phosphate or a substrate analogue were bound. Eight conserved amino acids at the putative active site of tHisF were exchanged by site-directed mutagenesis, and the purified variants were investigated by steady-state kinetics. Aspartate 11 appeared to be essential for the synthase activity both in vitro and in vivo, and aspartate 130 could be partially replaced only by glutamate. The carboxylate groups of these residues could provide general acid/base catalysis in the proposed catalytic mechanism of the synthase reaction.

    INTRODUCTION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

During the synthesis of various biomolecules including amino acids, nucleotides, and coenzymes, the amido group of glutamine is transferred to a large variety of acceptor substrates by glutamine amidotransferases (GATases)1 (1, 2). GATases catalyze two separate reactions at two active sites that are either located on a single polypeptide chain or on different subunits. In the glutaminase reaction, glutamine is hydrolyzed to glutamate and ammonia, which in the synthase reaction is added to an acceptor substrate that is specific for each GATase.

There are two classes of GATases that can be discriminated by catalytically essential residues in their glutaminase domains (1, 3). The key feature of class I GATases is the catalytic triad Cys-His-Glu. Recent x-ray structure determinations of three class I GATases, Escherichia coli carbamoyl phosphate synthase (4, 5), E. coli GMP synthase (6), and S. solfataricus anthranilate synthase (7) indicate a common fold of their glutaminase domains, which is similar to the well known alpha /beta hydrolase fold (8). Class II GATases belong to the large family of Ntn hydrolases (9), and their only catalytically essential amino acid is the conserved N-terminal cysteine. The corresponding synthase domains within each class are structurally, evolutionary, and functionally unrelated (2), supporting the hypothesis that glutamine-hydrolyzing enzymes were recruited independently by previously ammonium-dependent enzymes (3). Along these lines, most GATases can use ammonium salts as an alternative source of ammonia (1, 2).

The imidazole glycerol phosphate (ImGP) synthase is a class I GATase, which in bacteria constitutes a bienzyme complex consisting of the glutaminase subunit HisH and the synthase subunit HisF (10, 11). The ammonia produced by HisH reacts with the substrate of HisF, which is N'-((5'-phosphoribulosyl) formimino)-5-aminoimidazole-4-carboxamide-ribonucleotide (PRFAR). The products of this reaction, ImGP and 5- aminoimidazole-4-carboxamide ribotide (AICAR), are further used in histidine and de novo purine biosynthesis, respectively. In yeast, the glutaminase and synthase activities are located on a single polypeptide chain, which is termed HIS7 (12, 13). The mechanism of glutamine hydrolysis by HisH can be deduced from the class I GATase carbamoyl phosphate synthase (5). However, due to the lack of a high resolution structure for the bienzyme complex, the ammonia transfer from HisH to HisF and the mechanism of the HisF reaction are only poorly understood.

In order to address these questions, the thermostable variants tHisF and tHisH from Thermotoga maritima were produced in E. coli, purified and characterized by hydrodynamic and spectroscopic measurements, limited proteolysis, and steady-state enzyme kinetics. Moreover, the high resolution x-ray structure of isolated tHisF (14) was used to identify and probe amino acid residues that are potentially involved in catalysis of the synthase reaction. It was shown that tHisH is activated by complex formation with tHisF containing ImGP or a substrate analogue bound to its active site. A flexible loop region in tHisF appears to be important for these functional interactions with tHisH. Furthermore, two aspartate residues at the active site of tHisF were demonstrated to be essential for catalysis. One of them was partially replaceable by glutamate, as shown by saturation random mutagenesis and complementation in vivo of an E. coli hisF deletion strain. Based on these findings, a chemically plausible mechanism for the HisF synthase reaction was derived, which involves general acid/base catalysis.

    EXPERIMENTAL PROCEDURES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

DNA Manipulation and Sequence Analysis-- Preparation of DNA, digestion with restriction endonucleases, and DNA ligation were performed as described (15). Oligonucleotides were purchased from Metabion. DNA was amplified by PCR using cloned Pfu polymerase (Stratagene). For PCR with mutagenic oligonucleotides, Taq polymerase (Roche Diagnostics) was used. DNA was extracted from agarose gels using the QIAquick gel extraction kit (Qiagen). DNA sequencing was performed by the "Göttingen Genomics Laboratory" and by the "Zentrum für Molekulare Medizin der Universität Köln," using standard methods. N-terminal protein sequencing was performed by Dr. Paul Jenö (Biozentrum, University of Basel) and the "Zentrum für Molekulare Medizin der Universität Köln," again using standard methods.

Subcloning of the hisF and hisH Genes from T. maritima-- The hisF gene variants from Thermotoga maritima (thisF; Ref. 16) encoding the synthase subunit of imidazole glycerol phosphate synthase were cloned into the expression vector pET11c (Novagen) using the restriction enzymes NdeI and BamHI. The hisH gene of T. maritima (thisH; Ref. 16) was amplified by PCR, using purified chromosomal DNA (Qbiogen) as a template. The oligonucleotides 5'-GGT GTG ATA GCA TGC GTA TCG-3' with a SphI-site (in boldface type) and 5'-CTA CCA AGC TTC TGA AGA GAT CTA TCG-3' with a HindIII-site (in boldface type) were used as 5'- and 3'-primers, respectively. Using the two newly introduced restriction sites, the amplified DNA fragment was cloned into the vector pDS56/RBSII/SphI (17) to yield the plasmid pDS56/RBSIISphI-thisH. All inserts were entirely sequenced to exclude inadvertent PCR mutations.

Site-directed Mutagenesis of thisF-- Point mutations were introduced into thisF by PCR-based methods using either mutagenic 5'-primer for construction of thisF_C9A, D11N, K19S, or the megaprimer method (18) for construction of thisF_D51N, N103A, D130N, D176N, and D183N. In both approaches, the plasmid pET11c-thisF (19) was used as the template. The following mutagenic 5'-primers were used (NdeI restriction sites are in boldface type, and base substitutions to introduce an amino acid exchange are underlined): 5'-TGA TGA AGA CAT ATG CTC GCT AAA AGA ATA ATC GCG GCT CTC GAT-3' for construction of C9A, 5'-TGA TGA AGA CAT ATG CTC GCT AAA AGA ATA ATC GCG TGC CTC AAT GTG AAA GAC-3' for construction of D11N, and 5'-TGA TGA AGA CAT ATG CTC GCT AAA AGA ATA ATC GCG TGT CTC GAT GTG AAA GAC GGT CGT GTG GTG AGC GGA ACG AAC TTC-3' for construction of K19S. In all PCRs, the oligonucleotide 5'-CCG GAT CCA GCG TCA TCA CAA-3' containing a BamHI restriction site (in boldface type) was used as the 3'-primer. For the production of megaprimers, the following mutagenic oligonucleotides were used (base substitutions to introduce restriction sites for the control of the reaction are in boldface type, and base substitutions to introduce an amino acid exchange are underlined): 5'-CGC GGT AAT ATT CAG AAA AAC GAG-3' with a new SspI restriction site for construction of D51N, 5'-CAC AGC CGC AGT GGC TAT GCT CAC CTT GTC-3' with a new TspRI site for construction of N103A, 5'-CAC TCT TTT TGC ATT AAT CGC CAC GAC-3' with a new VspI restriction site for construction of D130N, 5'-GAC AGA AAC GGC ACC AAA TCG G-3' with a new BshNI restriction site for construction of D176N, and 5'-GGC ACA AAA TCG GGT TAC AAC ACT GAG ATG ATA AGG-3' with a new TspRI restriction site for construction of D183N. For production of the megaprimers for thisF_D51N, N103A, and D130N, the corresponding mutagenic oligonucleotides listed above were used as 3'-primers, and the oligonucleotide 5'-TGA TGA AGA CAT ATG CTC GCT AAA AG-3' (NdeI restriction site in boldface type) was used as 5'-primer. The megaprimers were purified by agarose gel electrophoresis and used in the second PCR as 5'-primers, whereas the oligonucleotide 5'-CCG GAT CCA GCG TCA TCA CAA-3' (BamHI restriction site in boldface type) was used as 3'-primer. The construction of the variants thisF_D176N and D183N followed the same protocol except that in the first PCR the mutagenic oligonucleotides were used as 5'-primers and the oligonucleotide 5'-CCG GAT CCA GCG TCA TCA CAA-3' (BamHI restriction site in boldface type) was used as 3'-primer. The purified megaprimers were used as 3'-primers, and the oligonucleotide 5'-TGA TGA AGA CAT ATG CTC GCT AAA AG-3' (NdeI restriction site in boldface type) was used as 5'-primer. The resulting full-length products were digested with NdeI and BamHI and ligated into pET11c. In order to confirm the base substitutions and to exclude inadvertent additional ones, all thisF gene variants were entirely sequenced.

Randomization of thisF Codons-- The thisF codons representing amino acids 11 and 130 were randomized in PCR-based approaches using degenerated primers and pET11c-thisF as template. For randomization of position 11, the oligonucleotide 5'-TAT ACG CAT GCT CGC TAA AAG AAT AAT CGC GTG CCT CNN SGT GAA GAC-3' with a SphI restriction site (in boldface type) was used, where N represents equal molar mixtures of all four bases, and S represents an equal molar mixture of G and C. In a PCR, the degenerate oligonucleotide was used as 5'-primer, and the oligonucleotide 5'-GTC GAC GGA TCC ACA ACC CCT CCA G-3' with a BamHI restriction site (in boldface type) was used as 3'-primer to yield an ensemble of thisF_D11NNS gene variants. For randomization of position 130, a megaprimer was produced using the degenerate oligonucleotide 5'-GTC GTG GCG ATT NNS GCA AAA AGA-3' as 3'-primer and the oligonucleotide 5'-TAT ACG CAT GCT CGC TAA AAG AAT AAT CGC-3' with a SphI restriction site (in boldface type) as 5'-primer. In a second PCR, the purified megaprimer was used as 5'-primer, and the oligonucleotide 5'-GTC GAC GGA TCC ACA ACC CCT CCA G-3' with a BamHI restriction site (in boldface type) was used as 3'-primer to yield an ensemble of thisF_D130NNS gene variants.

In Vivo Complementation-- The PCR-amplified ensembles of thisF genes containing randomized codons at amino acid position 11 or 130 were ligated into a modified pDS56/RBSII vector (termed pTNA), which contains a truncated derivative of the tryptophanase operon promotor (20) that permits constitutive gene expression in E. coli. The auxotrophic E. coli strain UTH860 (Delta hisF), which carries a mutant ehisF gene that encodes an inactive HisF protein (21), was transformed separately with the two plasmid libraries and plated onto LB medium containing 150 µg ml-1 ampicillin (15). The resulting lawns of at least 105 clones were rinsed off the plates, and plasmid DNA was prepared from the mixture of colonies. The Delta hisF strain was retransformed with these plasmid libraries, and aliquots were streaked on LB medium plates (nonselective plates) or on minimal medium plates (22) without histidine (selective plates) and incubated at 37 °C. From the nonselective plates, 16 colonies with randomized thisF codon 11 and 18 colonies with randomized thisF codon 130 were picked, and the corresponding thisF genes were entirely sequenced.

The mutational saturation is given by the following,
p≤1−<FENCE><AR><R><C> </C></R><R><C> </C></R></AR>1−<LIM><OP>∑</OP><LL>i=1</LL><UL>m</UL></LIM> f<SUB>i</SUB></FENCE><SUP>n</SUP> (Eq. 1)
where p is the probability that, for m randomized codons appearing with the relative frequency fi, practically all possible amino acid combinations are present in a library containing n independent clones (23). Under the assumption that all 32 permitted codons appear with equal frequency (fi = <FR><NU>1</NU><DE>32</DE></FR>), mutational saturation (p >=  0.99) of one codon (m = 1) is attained with a library of more than 145 independent clones. Although the bases within codons 11 and 130 were not randomly distributed, every allowed base was found at all three codon positions. None of the thisF genes contained additional mutations. Considering the size of the two libraries, each of which contained at least 105 independent clones, it can be concluded that all 20 amino acids were represented at both randomized positions. To select for functional amino acids at position 11 or 130 of tHisF, for each library one selective plate was incubated for various time periods. A number of clones appeared on both plates overnight and, after 48 h, additional colonies appeared on the selective plate with thisF genes that were randomized at position 130. Incubation was continued for 1 week, but no additional colonies appeared. A number of colonies that grew on selective medium after different periods of time were picked, and the thisF_D11NNS or thisF_D130NNS genes that encoded functional tHisF proteins were sequenced.

Purification of tHisF, tHisF Variants, and tHisH-- Wild-type tHisF and its variants containing individual amino acid exchanges were purified as described (19). The yield was between 7 and 17 mg of purified enzyme per g of wet cell mass. Heterologous expression of thisH was conducted in E. coli W3110 Delta trpEA2 cells containing pDS56/ RBSIISphI-thisH and the repressor plasmid pDMI, 1, as described for hisA from T. maritima (19). The cells were grown in 1 liter of LB medium supplemented with 0.15 mg/ml ampicillin and 0.075 mg/ml kanamycin. Overexpression of thisH was induced by adding 1 mM isopropyl-1-thio-beta -D-galactopyranoside at an optical density at 600 nm of about 0.6, and incubation was continued overnight. The cell suspension was washed with 100 mM potassium phosphate buffer at pH 7.5, containing 1 mM EDTA and 1 mM dithiothreitol, resuspended (5 ml of buffer per g wet cell mass), and lysed by sonification (Branson Sonifier W-250, 2 × 2 min, 50% pulse, 0 °C). According to SDS-PAGE, about 60% of tHisH were found in the soluble fraction of the cell extract. Benzonase (Merck) (50 units) was added to this fraction, which was then incubated for 1 h at 37 °C and subsequently for 20 min at 75 °C. The resulting suspension was centrifuged (Sorvall SS34, 12,000 rpm, 30 min, 4 °C), and the pellet, which contained heat-labile host proteins, was discarded. The supernatant was dialyzed against 10 mM Tris/HCl buffer at pH 8.0, containing 1 mM EDTA and 1 mM dithiothreitol, and loaded on an anion exchange column (POROS HQ20; 1 × 10 cm, PE Biosystems) that was equilibrated with the same buffer at room temperature. The column was washed with four volumes of equilibration buffer, and bound proteins were eluted with 1.5 liters of a linear gradient of 0-1 M sodium chloride at pH 8.0. tHisH eluted between 130 and 150 mM sodium chloride, as judged from SDS-PAGE and conductivity measurements. Fractions containing tHisH were pooled, dialyzed against 10 mM potassium phosphate buffer at pH 7.5, containing 1 mM dithiothreitol, loaded on a hydroxylapatite column (3.6 × 20 cm; Novartis) that was equilibrated with the same buffer, and eluted with 2 liters of a linear gradient of 10-500 mM potassium phosphate. tHisH eluted between 100 and 150 mM potassium phosphate with a purity above 95%, as judged from SDS-PAGE. The purification yielded ~5 mg of tHisH per g of wet cells. Fractions containing pure tHisH were pooled, dialyzed against 50 mM potassium phosphate buffer at pH 7.5, containing 1 mM EDTA and 1 mM dithiothreitol, concentrated to 1.8 mg/ml using Centricon-10 concentration devices (Millipore), and shock-frozen in liquid nitrogen.

Analytical Methods-- Purification of proteins was followed by electrophoresis on 12.5 or 15% SDS-polyacrylamide gels using the system of Laemmli (24) and staining with Coomassie Blue. During purification, the protein concentration was determined according to Bradford (25). The concentration of purified proteins was determined with molar extinction coefficients at 280 nm that were calculated from the amino acid sequence (26). Analytical gel filtration was performed at a flow rate of 0.5 ml/min on a Superdex 75 column (1 × 30 cm; Amersham Pharmacia Biotech) that was equilibrated with 50 mM potassium phosphate at pH 7.5, containing 300 mM sodium chloride. Apparent molecular masses were determined from the corresponding elution volumes, using a calibration curve that was obtained with standard proteins. Sedimentation equilibrium runs were performed in a Beckman analytical ultracentrifuge (model Optima XLA), following the absorption at 278 nm. Runs with tHisF were performed as described (27). Runs with tHisH were performed at 24,000 rpm and protein concentrations of 12 and 23 µM in 50 mM potassium phosphate, pH 7.5, at 20 °C, containing 25 mM potassium chloride. Runs with tHisH-tHisF were performed at 18,000 and 24,000 rpm with 20 µM protein in 50 mM potassium phosphate at pH 7.5 at 20 °C, containing 300 mM sodium chloride. To determine apparent molecular masses, the runs were analyzed as described (27). Fluorescence spectra were recorded with a F-4500 spectrofluorimeter (Hitachi) or a Cary Eclipse spectrofluorimeter (Varian). Proteolytic stability was tested at room temperature by incubating 10 nmol of substrate protein with 64 pmol of trypsin in 1 ml of 50 mM potassium phosphate, pH 7.5. The reaction was stopped after different time intervals by adding one volume of 2× SDS-PAGE sample buffer and heating for 5 min at 95 °C. The time course of proteolysis was followed on Tris-Tricine gels containing 20% acrylamide (28).

Steady-state Enzyme Kinetics-- The ammonia-dependent activity of isolated tHisF was measured by recording entire progress curves in 50 mM Tris acetate buffer, pH 8.5, at 25 °C. In order to determine KmPRFAR, the enzyme was saturated with ammonia by adding 100 mM ammonium acetate corresponding to 14.4 mM NH3 at pH 8.5. PRFAR, 20 µM, were synthesized in situ from ProFAR, using a molar excess of HisA from T. maritima (19) and completely converted by tHisF to ImGP and AICAR. The reaction was quantified by the decrease in absorption at 300 nm, using Delta epsilon 300(PRFAR-AICAR) = 5.64 mM-1 cm-1 (11). In order to determine KmNH3, the reaction was performed in the presence of 50 µM PRFAR at various concentrations of ammonium acetate between 0 and 200 mM, corresponding to 0 and 35 mM NH3 at pH 8.5. The glutamine-dependent activity of the tHisH-tHisF complex was measured in an analogous way as the ammonia-dependent reaction of tHisF, but the pH value was set to 8.0. In order to determine KmPRFAR, the reaction was performed with 20 µM ProFAR and 5 mM L-glutamine. To determine KmGln, the reaction was performed in the presence of 50 µM PRFAR at various concentrations of glutamine between 0 and 7 mM. The progress curves were analyzed with the integrated form of the Michaelis-Menten equation (29), yielding values for Km and Vmax. The glutaminase activity of tHisH in complex with liganded tHisF was measured in a coupled enzymatic assay with bovine liver glutamate dehydrogenase (Sigma). Glutamate that was produced by tHisH was oxidized by a molar excess of glutamate dehydrogenase in the presence of NAD+, yielding 2-oxoglutarate and NADH + H+ + NH4+. The reaction was quantified by the increase in absorption at 340 nm, using Delta epsilon 340(NADH-NAD+) = 6300 M-1 cm-1. The values for KmGln and Vmax were deduced from initial velocity measurement at various glutamine concentrations.

    RESULTS
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Production and Purification of tHisF, tHisH, and the tHisH-tHisF Complex-- tHisF was produced and purified as described previously (19). The thisH gene (16) was cloned into the expression vector pDS56/RBSII/SphI. Proteins were expressed in E. coli from this plasmid under the control of a lac operator system (17). tHisH was purified from the soluble fraction of the cell homogenate by first heat-precipitating E. coli host proteins, followed by anion exchange and hydroxylapatite chromatography. The final preparation was more than 98% pure as judged by SDS-PAGE and gel filtration chromatography on Superdex 75 (data not shown). The presence of the N-terminal methionine residues of both purified tHisF and tHisH were confirmed by N-terminal protein sequencing. The complex of tHisF and tHisH (tHisH-tHisF) was prepared by mixing equal molar amounts of both proteins, followed by gel filtration chromatography in order to remove any unintentional surplus of either of the components. The concentration of the purified proteins was determined by absorption spectroscopy, using calculated molar extinction coefficients at 280 nm (26) of 11,500 M-1 cm-1 for tHisF, 17,400 M-1 cm-1 for tHisH, and 28,900 M-1 cm-1 for the tHisH-tHisF complex.

Association States-- The molecular masses and the association states of recombinant tHisF, tHisH, and tHisH-tHisF were determined by analytical gel filtration chromatography and equilibrium analytical ultracentrifugation. tHisF eluted from a calibrated Superdex 75 column with an apparent molecular mass of 26.4 kDa (27), which is similar to the molecular mass for the monomer as calculated from the amino acid sequence (27.7 kDa). tHisH eluted with a significantly lower apparent molecular mass (17.4 kDa) than the calculated molecular mass for the monomer (23.1 kDa). A similarly retarded elution has been observed previously for eHisH (11), for unknown reasons. Mixed equal molar amounts of tHisH and tHisF eluted as a single peak with an apparent molecular mass of 41.8 kDa, which is smaller than the calculated molecular mass for the 1:1 complex (50.8 kDa). In order to further clarify their association states, the molecular masses of tHisH, tHisF, and tHisH-tHisF were determined by analytical ultracentrifugation. The data from several independent sedimentation equilibrium runs yielded average molecular masses of 24.6 kDa for tHisH, 49.8 kDa for tHisH-tHisF, and, dependent on protein concentration, between 26.0 and 38.4 kDa for tHisF (27). These results confirm that isolated tHisH and tHisF are essentially monomeric proteins that assemble to a stable 1:1 heterodimeric bienzyme complex. These association states of the T. maritima proteins are identical to those of the E. coli homologs (11), which excludes the idea that a higher association state is responsible for the increased thermostability of the ImGP synthase from the hyperthermophilic compared with that of the mesophilic bacterium (30).

Fluorescence Spectroscopic Analysis-- Fluorescence emission spectra were monitored to determine the relative solvent accessibility of the single tryptophan residues in tHisF (Trp156) and tHisH (Trp123). As judged from the emission maxima at 323 and 339 nm, Trp156 in tHisF appears to be shielded from solvent, while Trp123 in tHisH appears to be partly exposed to solvent (Fig. 1a). A solution containing equal molar amounts of tHisF and tHisH showed an emission maximum at 326 nm, indicating that complex formation leads to the burial of Trp123 of tHisH but does not change the solvent accessibility of Trp156 in tHisF. Along these lines, the titration of tHisH with tHisF leads to a linear shift of the emission maximum from 339 nm to lower wavelengths, until approximately equal molar amounts of tHisH and tHisF are present. An excess of tHisF does not lead to a further shift of the emission maximum (Fig. 1b). These data suggest that, at the applied protein concentrations, the binding of tHisF and tHisH is stoichiometric, indicating a high affinity of the two proteins. Although the thermodynamic dissociation constant Kd cannot be determined by this method, it has to be much smaller than the equivalence concentration, which is 10 µM.


View larger version (15K):
[in this window]
[in a new window]
 
Fig. 1.   Fluorimetric analysis of tHisF, tHisH, and the tHisH-tHisF complex. Conditions were as follows: excitation wavelength, 295 nm; 10 µM enzyme in 10 mM potassium phosphate buffer, pH 7.5, at 25 °C. a, different solvent accessibility of tryptophan residues. The single tryptophan residue of tHisF has its emission maximum at 323 nm, indicating that it is shielded from solvent. The single tryptophan residue of tHisH has its emission maximum at 339 nm, indicating that it is partly accessible to solvent. In the tHisH-tHisF complex, the emission maximum lies at 326 nm, indicating that both tryptophan residues are shielded from solvent. b, formation of the stoichiometric tHisH-tHisF complex. 10 µM tHisH was titrated with tHisF, and complex formation was followed by a shift of the fluorescence emission maximum (Femmax) from 339 nm to lower wavelengths. The titration curve is linear up to about equal molar concentrations of tHisF and tHisH and remains constant at higher ratios, indicating that the thermodynamic dissociation constant for complex formation, Kd, is 10 µM.

Steady-state Enzyme Kinetics-- The catalytic activities of isolated tHisF and tHisH and of the tHisH-tHisF complex were measured under steady-state conditions with coupled enzymatic assays, using absorption spectroscopy (Fig. 2). The synthase activities of the tHisH-tHisF complex and of isolated tHisF were measured, using glutamine and ammonium acetate as ammonia donors, respectively (Fig. 2, b and c). The glutaminase activity of tHisH was measured in the presence and absence of tHisF and the ligands ImGP or ProFAR (Fig. 2d), which were shown to activate the glutaminase activity of eHisH when bound to the active site of eHisF (11). Entire progress curves or initial velocities were measured and fitted to the integrated or the simple Michaelis-Menten equation, respectively. The resulting kcat and Km values were compared with those of the E. coli homologs eHisH, eHisF, and eHisH-eHisF (11) and with those of HIS7 from Saccharomyces cerevisiae, in which the glutaminase and the synthase domains are fused (13).


View larger version (20K):
[in this window]
[in a new window]
 
Fig. 2.   The reactions catalyzed by the tHisH-tHisF complex and the individual enzymes were monitored spectroscopically using coupled enzymatic assays. a, physiological reaction of the imidazole glycerol phosphate synthase (tHisH-tHisF). b, tHisH-tHisF assay. PRFAR was synthesized in situ from ProFAR with a molar excess of HisA from T. maritima (tHisA); the glutamine-dependent conversion of PRFAR into ImGP and AICAR was quantified using Delta epsilon 300(PRFAR-AICAR) = 5640 M-1 cm-1 (11). c, tHisF assay. The ammonia-dependent reaction was measured by replacing glutamine with ammonium acetate. d, tHisH assay. The glutamate produced by the glutaminase activity of tHisH was oxidized by a molar excess of glutamate dehydrogenase (GDH) to 2-oxoglutarate and ammonium; the reaction was quantified by the concomitant reduction of NAD+ to NADH + H+, using Delta epsilon 340(NADH-NAD+) = 6300 M-1 cm-1. The ligand L bound to tHisF is either ImGP or ProFAR. The conditions for the different assays are listed in Table I (for the synthase reactions) and Table II (for the glutaminase reaction).

Table I summarizes the results of the synthase activity measurements. In the ammonia-dependent reaction, at comparable temperatures and pH values, isolated tHisF exhibited 2-5-fold higher catalytic efficiencies (kcat/KmPRFAR and kcat/KmNH3) than isolated eHisF. The higher catalytic efficiency was caused by lower Km values of tHisF for both PRFAR and ammonia, while the turnover rate kcat was slightly lower for tHisF than for eHisF. In the glutamine-dependent reaction, the tHisH-tHisF complex exhibited catalytic activities comparable to tHisF in the ammonia-dependent reaction. In contrast, the catalytic efficiency kcat/KmPRFAR of the eHisH-eHisF complex in the glutamine-dependent reaction was about 25-fold higher than that of eHisF in the ammonia-dependent reaction, due to a much higher KmPRFAR value of eHisF. The decreased activity of eHisF in the ammonia-dependent reaction might be due to inhibition by high concentrations of ammonium chloride, the ammonia donor (11). tHisF was also inhibited by ammonium chloride, which was therefore replaced by ammonium acetate that showed no detectable inhibitory effect at the applied concentrations (data not shown). Similar catalytic efficiencies were observed for the glutamine-dependent synthase reactions of the bifunctional HIS7 enzyme and the tHisH-tHisF complex (Table I).

                              
View this table:
[in this window]
[in a new window]
 
Table I
Steady-state kinetic constants of the ammonia-dependent ImGP synthase reaction of isolated HisF subunits and the glutamine-dependent synthase reaction of HisH-HisF complexes

In the absence of tHisF, tHisH did not show any measurable glutaminase activity. However, as observed with the E. coli enzyme (11), the binding of either the product ImGP or the substrate analogue ProFAR to the active site of tHisF led to a strong stimulation of tHisH. In contrast, neither the T. maritima, nor the E. coli HisH glutaminase activities are stimulated by AICAR, the second product of the ImGP synthase reaction (Ref. 11 and data not shown). The steady-state parameters of the glutaminase activity of tHisH were therefore measured in the presence of tHisF and saturating concentrations of either ImGP or ProFAR (Table II). The kcat of the glutaminase reaction is 0.1 s-1 and lower by a factor of about 4-8 than the kcat value of the glutamine-dependent synthase reaction (Table I), where PRFAR (instead of ImGP or ProFAR) is bound to the active site of tHisF in the tHisH-tHisF complex. Similarly, in the E. coli enzyme, the kcat of the glutaminase reaction is lower by a factor of 3-4 than the kcat of the glutamine-dependent synthase reaction. In addition, the Km values for glutamine are increased about 10-fold (Tables I and II). Obviously, both in the T. maritima and the E. coli ImGP synthase, the activation of the glutaminase reaction by the synthase subunit is more efficient when the native substrate PRFAR is bound rather than ImGP or ProFAR. In contrast, in HIS7, where glutaminase and synthase activities are located on the same bifunctional polypeptide chain, ImGP and ProFAR are equally efficient glutaminase activators as PRFAR (Tables I and II). The dependence of the glutaminase activity of tHisH on the presence of a ligand bound to the active site of tHisF can be used to estimate the thermodynamic dissociation constants Kd for the binding of these ligands. As shown in Table II, ImGP binds with similarly low affinity to tHisF and eHisF, whereas ProFAR binds almost 20-fold more strongly to tHisF than to eHisF.

                              
View this table:
[in this window]
[in a new window]
 
Table II
Steady-state kinetic constants of the glutaminase activities of tHisH-tHisF, eHisH-eHisF, and HIS7

Limited Proteolysis-- The sensitivity of a protein toward proteolytic degradation can provide valuable information on its stability as well as on the flexibility of potential cleavage sites (31). For this reason, tHisF, tHisH and tHisH-tHisF were digested with trypsin, and the results were analyzed by SDS-PAGE (Fig. 3). Isolated tHisF (27.7 kDa) was cleaved once to yield fragments with apparent molecular masses of about 24 kDa (Fig. 3a) and about 4 kDa (not shown). These fragments, which were detectable already after 5 min of incubation, were enriched at the cost of intact tHisF in a time-dependent manner. After 120 min, about one-quarter of tHisF was degraded (Fig. 3a). N-terminal sequencing of the large fragment showed that cleavage occurred C-terminal of Arg27 (Fig. 4). Arg27 is located in a long and presumably flexible region at the C-terminal (active site) face of the beta -barrel (cf. arrow in Fig. 4a). In contrast to tHisF, at the given conditions isolated tHisH was completely resistant to trypsin (data not shown). Incubation of the tHisH-tHisF complex with trypsin again resulted in the degradation of tHisF, albeit with significantly higher rate; after 120 min, almost all tHisF was degraded into fragments of about 24 kDa (Fig. 3b) and about 4 kDa (not shown). These results show that the single trypsin cleavage site at Arg27 is more susceptible in complexed tHisF compared with isolated tHisF.


View larger version (18K):
[in this window]
[in a new window]
 
Fig. 3.   Isolated tHisF is less susceptible to proteolytic attack than complexed tHisF. 10 µM isolated tHisF (a) or a 10 µM concentration of the tHisH-tHisF complex (b) was digested with 64 nM trypsin for the indicated time intervals in 50 mM potassium phosphate, pH 7.5, at 25 °C. tHisF (27.7 kDa) is degraded to one large fragment with Mapp of ~24 kDa and one small fragment with Mapp of ~4 kDa (not shown). This cleavage, which takes place C-terminal of Arg27 (cf. arrow in Fig. 4), occurs more slowly in isolated compared with complexed tHisF, indicating conformational changes in the flexible loop region between strand beta 1 and helix alpha 1. tHisH (23.1 kDa) is resistant to trypsin, both in the tHisH-tHisF complex (b) and in the isolated form (not shown). Band 1, tHisF; band 2, large tryptic fragment of tHisF (a) or a mixture of tHisH and the large tryptic fragment of tHisF (b). M, standard proteins.


View larger version (27K):
[in this window]
[in a new window]
 
Fig. 4.   Conserved and catalytically important amino acid residues in the (beta alpha )8-barrel structure of tHisF (14). a, structure-based sequence alignment. Secondary structural elements are identified by blue arrows (beta  strands) and red cylinders (alpha  helices). Conserved, amino acids that are invariant in all known 25 HisF sequences and identical in at least 22 sequences are shown in uppercase and lowercase, respectively. Exchanged, eight conserved amino acid residues close to the proposed active site of tHisF were replaced individually with structurally similar residues (in boldface type). Amino acids involved in phosphate binding are underlined. The single trypsin cleavage site within tHisF, which is located C-terminal to Arg27, is marked by an arrow (cf. Fig. 3). b, ribbon diagrams showing a side view on the central beta -barrel. P(N) and P(C) are the phosphate ions bound to the N- and C-terminal halves of tHisF. Asp11 and Asp130 (D11 and D130) are essential for catalysis of the tHisF reaction, Asp176 (D176) appears to be important for catalysis, and the flexible region containing Lys19 (K19) and Arg27 (R27) is involved in interactions between tHisF and tHisH that take place upon the reaction of nascent ammonia with PRFAR (side chains in green). Side chains of conserved amino acids at the N-terminal face of the beta -barrel, which might be involved in binding of tHisH, are in black (see "Results" and "Discussion" for details.)

Mutational Analysis of the Active Site of tHisF-- Glutamine hydrolysis reactions catalyzed by other members of the class I GATase family have been characterized in detail, for example in the case of carbamoyl phosphate synthase (5, 32, 33). In contrast, the complicated reaction mechanism that leads to the formation of the imidazole ring catalyzed by HisF (Fig. 2a) has not been investigated so far. The first step toward this goal is to identify the catalytic amino acid residues. These should (a) be invariant in the 25 currently available HisF sequences; (b) lie at the C-terminal face of the central beta -barrel of tHisF, which is the location of the active sites of all known (beta alpha )8-barrel enzymes (34); and (c) be able to reversibly provide and abstract a proton, since the HisF reaction probably involves general acid/base catalysis. Eight residues that fulfill these criteria were replaced by site-directed mutagenesis with structurally similar amino acids lacking the putative functional group (Fig. 4a). The mutant thisF genes were expressed in E. coli, and the respective gene products were purified in the same way as described for the wild-type protein (19). In order to exclude structural perturbations (or denaturation) caused by the introduced amino acid replacements, all purified tHisF variants were analyzed by fluorescence spectroscopy and analytical gel filtration chromatography. No significant difference from wild-type tHisF was detected in any of the variants, confirming the assumption that residues at the C-terminal face of (beta alpha )8-barrels do not contribute significantly to protein stability (35). Also, the association with tHisH was not impaired (data not shown).

The steady-state enzyme kinetic constants of the isolated tHisF variants and of the corresponding tHisH-tHisF complexes were determined and compared with the constants of the wild-type enzymes (Tables III and IV). The catalytic efficiencies kcat/Km of both isolated and complexed tHisF_C9A, tHisF_D51N, tHisF_N103A, and tHisF_D183N were not significantly different from wild-type tHisF, ruling out any central catalytic role for the replaced residues. Also, the ammonia-dependent reactions of isolated tHisF_K19S were similarly efficient as those of wild-type tHisF (Table III). In contrast, the efficiencies of the glutamine-dependent reactions of the tHisH-tHisF_K19S complex were significantly impaired (Table IV). The variant tHisF_D176N showed a 40-50 fold decrease in kcat, both in isolated form and in complex with tHisH. The strongest effects, however, were found for tHisF_D11N and tHisF_D130N, the catalytic efficiencies kcat/KmPRFAR of which were decreased by approximately 5 orders of magnitude. While for tHisF_D11N only the kcat was affected, both kcat and Km were drastically impaired in tHisF_D130N. These results suggest that Asp11 and Asp130 play essential roles in the catalysis of the tHisF reaction.

                              
View this table:
[in this window]
[in a new window]
 
Table III
Effect of individual amino acid exchanges on the steady-state enzyme kinetic constants of the ammonia-dependent ImGP synthase reaction catalyzed by isolated tHisF

                              
View this table:
[in this window]
[in a new window]
 
Table IV
Effect of individual amino acid exchanges on the steady-state enzyme kinetic parameters of the glutamine-dependent ImGP synthase reaction catalyzed by the tHisH-tHisF complex

Saturation Random Mutagenesis and Complementation in Vivo-- To further test the crucial role of Asp11 and Asp130 for catalysis, both residues were replaced by all 19 alternative amino acids, and their function in catalysis was probed by selection in vivo. To this end, the corresponding codons at amino acid positions 11 or 130 were subjected to saturation random mutagenesis using degenerate oligonucleotides. The randomization of the third codon position was limited to G and C. This restriction eliminates 32 out of 64 codons, but the remainder codons still represent all 20 amino acids.

Histidine auxotrophic E. coli cells lacking a functional hisF gene (Delta hisF cells) were transformed with a plasmid that allows constitutive expression in E. coli of the cloned thisF_D11NNS or thisF_D130NNS ensembles (20). Transformants were streaked onto selective medium without histidine and incubated at 37 °C; a small aliquot was streaked onto nonselective LB medium. From the nonselective plates, colonies containing thisF_D11NNS or thisF_D130NNS were randomly picked, and the thisF genes were sequenced in order to confirm the random distribution of bases at each position of the codons. Sequencing of thisF from colonies grown on selective medium showed that at position 11 only the wild-type amino acid aspartate allowed functional complementation of the Delta hisF cells. Similarly, all overnight grown colonies on the selective plate with the randomized codon at position 130 coded for the wild-type amino acid aspartate. However, all colonies that appeared on this plate after 48 h contained a codon for glutamate at position 130 (Table V).

                              
View this table:
[in this window]
[in a new window]
 
Table V
Asp11 is essential for tHisF function, but Asp130 can be functionally replaced by Glu
Results of saturation mutagenesis and functional complementation in vivo are shown.

In order to test its catalytic activity in vitro, tHisF_D130E was produced in E. coli, purified and characterized as described above for the other tHisF variants. tHisF_D130E was identical to the wild-type enzyme with respect to fluorescence properties and complex formation with tHisH (data not shown). Its steady-state enzyme kinetic parameters, however, were between those of wild-type tHisF and tHisF_D130N; the kcat value was reduced by a factor of about 400-500, and the KmPRFAR was increased almost 20-fold (Table III).

    DISCUSSION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Catalytic Properties and Allosteric Interactions of T. maritima ImGP Synthase-- Most investigated enzymes from the hyperthermophilic bacterium T. maritima are only marginally active at room temperature, probably due to conformational rigidity at their active sites (36). In contrast, at comparable temperatures of 25 and 30 °C, the isolated synthase subunit tHisF shows a higher catalytic efficiency (kcat/KmPRFAR and kcat/KmNH3) than eHisF, due to much lower Km values (Table I). Similarly, both phosphoribosyl anthranilate isomerase and indoleglycerol phosphate synthase from T. maritima, which are involved in tryptophan biosynthesis, have lower Km values and higher catalytic efficiencies than their homologues from E. coli (37, 38). It has been speculated that the high catalytic efficiency of T. maritima phosphoribosyl anthranilate isomerase is important for rapid processing of its thermolabile substrate, and a similar argument may hold for the procession by tHisF of the extremely labile PRFAR (39). However, the catalytic efficiencies of the overall tHisH-tHisF reaction (kcat/KmPRFAR and kcat/KmGln; Table I) and of the glutaminase reaction (kcat/KmGln; Table II) were significantly lower than those of the corresponding enzymes from E. coli and S. cerevisiae. Thus, the catalytic activity of tHisH seems to limit, at least at room temperature, the overall catalytic efficiency of T. maritima ImGP synthase. This result indicates conformational rigidity of tHisH, which is supported by its resistance to digestion by trypsin (Fig. 3).

The quaternary structure of the tHisH-tHisF complex has to provide the basis for (a) the activation of the glutaminase reaction at the active site of tHisH by the binding of PRFAR (or ImGP or ProFAR) to the active site of tHisF and (b) the transfer of nascent ammonia between tHisH and tHisF in a way that excludes contact with water, which inevitably would protonate ammonia to the nonreactive ammonium ion. The active site of tHisF is located at the C-terminal face of the central beta -barrel (Fig. 4b), but several lines of evidence suggest that tHisH docks to the N-terminal face of the barrel. About half of the highly conserved and invariant residues are located close to the N-terminal face of the barrel of tHisF, indicating an important functional role of this region (Fig. 4). Two of these conserved residues, Arg5 and Glu46, were shown to be important for the glutamine-dependent but not for the ammonia-dependent reaction of the eHisH-eHisF complex (40). Furthermore, the region between strand beta 1 and helix alpha 1, which is located at the C-terminal face of tHisF (Fig. 4), is not protected by tHisH against trypsinolysis (Fig. 3). Moreover, docking of tHisH to the C-terminal face of tHisF would impede access of the large substrate PRFAR and should therefore impair catalytic activity. However, both the KmPRFAR values and the catalytic efficiencies kcat/KmPRFAR are practically identical for isolated tHisF and tHisF in complex with tHisH (Table I).

How is the binding of PRFAR to tHisF at the C-terminal face of the beta -barrel coupled with the glutaminase activity of tHisH, which is presumably bound to the N-terminal face of the barrel? Arg27 is located in a flexible region between strand beta 1 and helix alpha 1 at the C-terminal face of tHisF (Fig. 4). The cleavage of trypsin at Arg27 is accelerated in the tHisH-tHisF complex compared with isolated tHisF (Fig. 3). This result indicates that the binding of tHisH to the N-terminal face of tHisF induces a long range conformational transition that makes Arg27 more available to proteolytic cleavage. The amino acid exchange K19S, moreover, which lies in the same flexible region as Arg27 (Fig. 4), causes larger effects of kcat and KmPRFAR in the glutamine-dependent reaction of the tHisH-tHisF complex (Table IV) than in the ammonia-dependent reaction of isolated tHisF (Table III). Thus, the flexible region between strand beta 1 and helix alpha 1 is involved in interactions between tHisF and tHisH that take place upon the reaction of nascent ammonia with PRFAR. Since uncharged ammonia is the nucleophile to attack PRFAR (Fig. 2), ammonia has to be transferred from tHisH to tHisF without contact with solvent. As a consequence, the transfer will probably occur through a hydrophobic channel connecting the two active sites, but a high resolution x-ray structure of the tHisH-tHisF complex is necessary to verify this hypothesis. Ammonia channels of this kind were recently identified in the GATases carbamoyl phosphate synthase (4, 41), glutamine phosphoribosyl pyrophosphate amidotransferase (42), and asparagine synthase B (43).

Plausible Mechanism of the HisF Reaction-- The mechanism of the cycloligase/lyase HisF reaction (Fig. 2a) is complicated and unique in primary metabolism (13). Two aspartate residues, Asp11 and Asp130, are catalytically essential, since their replacement by asparagines causes an almost complete loss of tHisF activity (Table III). Saturation random mutagenesis of the corresponding codons and complementation studies showed that the function of Asp11 could not be substituted by any other residue. In contrast, the variant tHisF_D130E complemented a Delta hisF strain in vivo, albeit more slowly than wild-type tHisF (Table V). Thus, the restriction for aspartate at position 130 is not as exclusive as that at position 11, but a carboxylate side chain appears to be essential at both positions. These results support the hypothesis that the tHisF reaction depends on general acid/base catalysis, and a chemically plausible reaction mechanism was developed on the basis of this finding (Fig. 5). The proposed sequence of reactions starts with the substitution of the phosphoribulosyl carbonyl oxygen of PRFAR by ammonia, resulting in the release of water and the formation of Imine I. In the next step, the addition of water results in the generation of the first reaction product AICAR and of Imine II. In the subsequent step, the imidazole ring of ImGP is closed, and water is released in a reaction that is catalyzed by the general acid H-A1 and the general base A<UP><SUB>2</SUB><SUP>−</SUP></UP>. In support of this mechanism, ProFAR (and probably also PRFAR) are slowly hydrolyzed in the absence of enzyme to yield AICAR and other unidentified products, and the rate of this reaction increases with increasing proton concentration (44). In order to act as a general acid at physiological pH, the pKa value of an aspartate within an enzyme must be shifted by several units compared with free aspartate. Such a shift should be very sensitive to the immediate environment of the carboxylate function and therefore to the length of the amino acid side chain. As shown by complementation in vivo, Asp11 is absolutely required for catalysis and cannot be functionally replaced by any other amino acid (Table V); it might therefore be identical with the general acid H-A1. Asp130 can be functionally replaced by Glu (Table V) and therefore probably has a more "normal" pKa value; it might therefore be identical with the general base A<UP><SUB>2</SUB><SUP>−</SUP></UP>. The reactions from PRFAR to Imine I and of Imine I to Imine II might also be accelerated by general acid/base catalysis, and Asp11 and Asp130 could play an essential role here as well. Asp176 could also be involved in either of the reaction steps, since the kcat values of the tHisF_D176N variant are decreased about 50-fold compared with wild-type tHisF (Tables III and IV).


View larger version (18K):
[in this window]
[in a new window]
 
Fig. 5.   Plausible mechanism of the tHisF reaction based on general acid/base catalysis. R1, ribosephosphate, R2, 3-phosphoglycerol. The general acid H-A1 and the general base A<UP><SUB>2</SUB><SUP>−</SUP></UP> are probably identical with Asp11 and Asp130, respectively (see "Discussion" for details.)


    ACKNOWLEDGEMENTS

We thank A. Lustig for running the analytical ultracentrifuge; Dr. Paul Jenö, the "Göttingen Genomics Laboratory," and the "Zentrum für Molekulare Medizin der Universität Köln" for protein and DNA sequencing; and Drs. Ron Bauerle, Helmut W. Klein, Kasper Kirschner, and Matthias Wilmanns for helpful suggestions on the HisF reaction mechanism and critical comments on the manuscript. We are grateful to Dr. Andreas Ivens for experimental help in the first stages of this project.

    FOOTNOTES

* This work was supported in part by Deutsche Forschungsgemeinschaft Grant STE 891/3-1 and a Heisenberg Fellowship (to R. S.).The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

Dagger To whom correspondence should be addressed. Tel.: 49-221-470-6432; Fax: 49-221-470-6731; E-mail: Reinhard.Sterner@Uni-Koeln.de.

Published, JBC Papers in Press, March 22, 2001, DOI 10.1074/jbc.M102012200

    ABBREVIATIONS

The abbreviations used are: GATase, glutamine amidotransferase; AICAR, 5-aminoimidazole-4-carboxamide ribotide; ImGP, imidazole glycerol phosphate; HisH, glutaminase subunit of ImGP synthase; eHisH and tHisH, HisH from E. coli and T. maritima, respectively; HisF, synthase subunit of ImGP synthase; eHisF and tHisF, HisF from E. coli and T. maritima, respectively; PRFAR, N'-((5'-phosphoribulosyl) formimino)-5-aminoimidazole-4-carboxamide-ribonucleotide; ProFAR, N'-((5'-phosphoribosyl)formimino)-5-aminoimidazole-4-carboxamide-ribonucleotide; PCR, polymerase chain reaction; PAGE, polyacrylamide gel electrophoresis; Tricine, N-[2-hydroxy-1,1-bis(hydroxymethyl)ethyl]glycine; Pipes, 1,4-piperazinediethanesulfonic acid.

    REFERENCES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

1. Zalkin, H., and Smith, J. L. (1998) Adv. Enzymol. Relat. Areas Mol. Biol. 72, 87-144[Medline] [Order article via Infotrieve]
2. Massière, F., and Badet-Denisot, M. A. (1998) Cell. Mol. Life Sci. 54, 205-222[CrossRef][Medline] [Order article via Infotrieve]
3. Zalkin, H. (1993) Adv. Enzymol. Relat. Areas Mol. Biol. 66, 203-309[Medline] [Order article via Infotrieve]
4. Thoden, J. B., Holden, H. M., Wesenberg, G., Raushel, F. M., and Rayment, I. (1997) Biochemistry 36, 6305-6316[CrossRef][Medline] [Order article via Infotrieve]
5. Thoden, J. B., Miran, S. G., Phillips, J. C., Howard, A. J., Raushel, F. M., and Holden, H. M. (1998) Biochemistry 37, 8825-8831[CrossRef][Medline] [Order article via Infotrieve]
6. Tesmer, J. J., Klem, T. J., Deras, M. L., Davisson, V. J., and Smith, J. L. (1996) Nat. Struct. Biol. 3, 74-86[Medline] [Order article via Infotrieve]
7. Knöchel, T., Ivens, A., Hester, G., Gonzalez, A., Bauerle, R., Wilmanns, M., Kirschner, K., and Jansonius, J. N. (1999) Proc. Natl. Acad. Sci. U. S. A. 96, 9479-9484[Abstract/Free Full Text]
8. Ollis, D. L., Cheah, E., Cygler, M., Dijkstra, B., Frolow, F., Franken, S. M., Harel, M., Remington, S. J., Silman, I., and Schrag, J. (1992) Protein Eng. 5, 197-211[Abstract]
9. Brannigan, J. A., Dodson, G., Duggleby, H. J., Moody, P. C., Smith, J. L., Tomchick, D. R., and Murzin, A. G. (1995) Nature 378, 416-419[CrossRef][Medline] [Order article via Infotrieve]
10. Alifano, P., Fani, R., Liò, P., Lazcano, A., Bazzicalupo, M., Carlomagno, M. S., and Bruni, C. B. (1996) Microbiol. Rev. 60, 44-69[Free Full Text]
11. Klem, T. J., and Davisson, V. J. (1993) Biochemistry 32, 5177-5186[Medline] [Order article via Infotrieve]
12. Kuenzler, M., Balmelli, T., Egli, C. M., Paravicini, G., and Braus, G. H. (1993) J. Bacteriol. 175, 5548-5558[Abstract]
13. Chittur, S. V., Chen, Y., and Davisson, V. J. (2000) Protein Expression Purif. 18, 366-377[CrossRef][Medline] [Order article via Infotrieve]
14. Lang, D., Thoma, R., Henn-Sax, M., Sterner, R., and Wilmanns, M. (2000) Science 289, 1546-1550[Abstract/Free Full Text]
15. Sambrook, J., Fritsch, E. E., and Maniatis, T. (1989) Molecular Cloning: A Laboratory Manual , 2nd Ed. , Cold Spring Harbor Laboratory, Cold Spring Harbor, NY
16. Thoma, R., Schwander, M., Liebl, W., Kirschner, K., and Sterner, R. (1998) Extremophiles 2, 379-389[CrossRef][Medline] [Order article via Infotrieve]
17. Stüber, D., Matile, H., and Garotta, G. (1990) in Immunological Methods (Lefkovits, I. , and Pernis, B., eds), Vol. 4 , pp. 121-152, Academic Press, Inc., Orlando, FL
18. Sarkar, G., and Sommer, S. S. (1990) BioTechniques 8, 404-407[Medline] [Order article via Infotrieve]
19. Thoma, R., Obmolova, G., Lang, D. A., Schwander, M., Jeno, P., Sterner, R., and Wilmanns, M. (1999) FEBS Lett. 454, 1-6[CrossRef][Medline] [Order article via Infotrieve]
20. Merz, A., Yee, M. C., Szadkowski, H., Pappenberger, G., Crameri, A., Stemmer, W. P., Yanofsky, C., and Kirschner, K. (2000) Biochemistry 39, 880-889[CrossRef][Medline] [Order article via Infotrieve]
21. Goldschmidt, E. P., Cater, M. S., Matney, T. S., Butler, M. A., and Greene, A. (1970) Genetics 66, 219-229[Free Full Text]
22. Vogel, H. J., and Bonner, D. M. (1956) J. Biol. Chem. 218, 97-106[Free Full Text]
23. Darimont, B., Stehlin, C., Szadkowski, H., and Kirschner, K. (1998) Protein Sci. 7, 1221-1232[Abstract/Free Full Text]
24. Laemmli, U. K. (1970) Nature 227, 680-685[Medline] [Order article via Infotrieve]
25. Bradford, M. M. (1976) Anal. Biochem. 72, 248-254[CrossRef][Medline] [Order article via Infotrieve]
26. Pace, C. N., Vajdos, F., Fee, L., Grimsley, G., and Gray, T. (1995) Protein Sci. 4, 2411-2423[Abstract/Free Full Text]
27. Höcker, B., Beismann-Driemeyer, S., Hettwer, S., Lustig, A., and Sterner, R. (2001) Nat. Struct. Biol. 8, 32-36[CrossRef][Medline] [Order article via Infotrieve]
28. Schägger, H., and von Jagow, G. (1987) Anal. Biochem. 166, 368-379[Medline] [Order article via Infotrieve]
29. Hommel, U., Eberhard, M., and Kirschner, K. (1995) Biochemistry 34, 5429-5439[Medline] [Order article via Infotrieve]
30. Jaenicke, R., and Böhm, G. (1998) Curr. Opin. Struct. Biol. 8, 738-748[CrossRef][Medline] [Order article via Infotrieve]
31. Hubbard, S. J. (1998) Biochim. Biophys. Acta 1382, 191-206[Medline] [Order article via Infotrieve]
32. Thoden, J. B., Huang, X., Raushel, F. M., and Holden, H. M. (1999) Biochemistry 38, 16158-16166[CrossRef][Medline] [Order article via Infotrieve]
33. Rishavy, M. A., Cleland, W. W., and Lusty, C. J. (2000) Biochemistry 39, 7309-7315[CrossRef][Medline] [Order article via Infotrieve]
34. Pujadas, G., and Palau, J. (1999) Biologia 54, 231-254
35. Thoma, R., Hennig, M., Sterner, R., and Kirschner, K. (2000) Struct. Fold. Des. 8, 265-276[CrossRef][Medline] [Order article via Infotrieve]
36. Jaenicke, R. (2000) Proc. Natl. Acad. Sci. U. S. A. 97, 2962-2964[Free Full Text]
37. Sterner, R., Kleemann, G. R., Szadkowski, H., Lustig, A., Hennig, M., and Kirschner, K. (1996) Protein Sci. 5, 2000-2008[Abstract/Free Full Text]
38. Merz, A., Knöchel, T., Jansonius, J. N., and Kirschner, K. (1999) J. Mol. Biol. 288, 753-763[CrossRef][Medline] [Order article via Infotrieve]
39. Martin, R. G., Berberich, M. A., Ames, B. N., Davis, W. W., Goldberger, R. F., and Yourno, J. D. (1971) Methods Enzymol. 17, 3-44[CrossRef]
40. Klem, T. J., Chen, Y., and Davisson, V. J. (2001) J. Bacteriol. 182, 989-996[CrossRef]
41. Holden, H. M., Thoden, J. B., and Raushel, F. M. (1998) Curr. Opin. Struct. Biol. 8, 679-685[CrossRef][Medline] [Order article via Infotrieve]
42. Krahn, J. M., Kim, J. H., Burns, M. R., Parry, R. J., Zalkin, H., and Smith, J. L. (1997) Biochemistry 36, 11061-11068[CrossRef][Medline] [Order article via Infotrieve]
43. Larsen, T. M., Boehlein, S. K., Schuster, S. M., Richards, N. G., Thoden, J. B., Holden, H. M., and Rayment, I. (1999) Biochemistry 38, 16146-16157[CrossRef][Medline] [Order article via Infotrieve]
44. Davisson, V. J., Deras, I. L., Hamilton, S. E., and Moore, L. L. (1994) J. Org. Chem. 59, 137-143


Copyright © 2001 by The American Society for Biochemistry and Molecular Biology, Inc.