101
|
Lee P, Colman RF. Implication by site-directed mutagenesis of Arg314 and Tyr316 in the coenzyme site of pig mitochondrial NADP-dependent isocitrate dehydrogenase. Arch Biochem Biophys 2002; 401:81-90. [PMID: 12054490 DOI: 10.1016/s0003-9861(02)00041-3] [Citation(s) in RCA: 13] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/27/2022]
Abstract
Sequence alignment of pig mitochondrial NADP-dependent isocitrate dehydrogenase with eukaryotic (human, rat, and yeast) and Escherichia coli isocitrate dehydrogenases reveals that Tyr316 is completely conserved and is equivalent to the E. coli Tyr345, which interacts with the 2'-phosphate of NADP in the crystal structure [Hurley et al., Biochemistry 30 (1991) 8671-8678]. Lys321 is also completely conserved in the five isocitrate dehydrogenases. Either an arginine or lysine residue is found among the enzymes from other species at the position corresponding to the pig enzyme Arg314. While Arg323 is not conserved among all species, its proximity to the coenzyme site makes it a good candidate for investigation. The importance of these four amino acids to the function of pig mitochondrial NADP-isocitrate dehydrogenase was studied by site-directed mutagenesis. Mutants (R314Q, Y316F, Y316L, K321Q, and R323Q) were generated by a megaprimer polymerase chain reaction method. Wild-type and mutant enzymes were expressed in E. coli and purified to homogeneity. All mutant and wild-type enzymes exhibited comparable molecular weights indicative of the dimeric enzyme. Mutations do not cause an appreciable change in enzyme secondary structure as revealed by circular dichroism measurements. The kinetic parameters (V(max) and K(M) values) of K321Q and R323Q are similar to those of wild-type, indicating that Lys321 and Arg323 are not involved in enzyme function. R314Q exhibits a 10-fold increase in K(M) for NADP as compared to that of wild-type, while they have comparable V(max) values. These results suggest that Arg314 contributes to the affinity between the enzyme and NADP. The hydroxyl group of Tyr316 is not required for enzyme function since Y316F exhibits similar kinetic parameters to those of wild-type. Y316L shows a 4-fold increase in K(M) for NADP and a decrease in V(max) as compared to wild-type, suggesting that the aromatic ring of the Tyr of isocitrate dehydrogenase contributes to the affinity for coenzyme, as well as to catalysis. The K(i) for NAD of R314Q, Y316F, and Y316L is comparable to that of wild-type, indicating that the Arg314 and Tyr316 may be located near the 2'-phosphate of enzyme-bound NADP.
Collapse
Affiliation(s)
- Peychii Lee
- Department of Chemistry and Biochemistry, University of Delaware, Newark 19716, USA
| | | |
Collapse
|
102
|
Camacho M, Rodríguez-Arnedo A, Bonete MJ. NADP-dependent isocitrate dehydrogenase from the halophilic archaeon Haloferax volcanii: cloning, sequence determination and overexpression in Escherichia coli. FEMS Microbiol Lett 2002; 209:155-60. [PMID: 12007799 DOI: 10.1111/j.1574-6968.2002.tb11125.x] [Citation(s) in RCA: 14] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/30/2022] Open
Abstract
A gene encoding NADP-dependent Ds-threo-isocitrate dehydrogenase was isolated from Haloferax volcanii genomic DNA by using a combination of polymerase chain reaction and screening of a lambda EMBL3 library. Analysis of the nucleotide sequence revealed an open reading frame of 1260 bp encoding a protein of 419 amino acids with 45837 Da molecular mass. This sequence is highly similar to previously sequenced isocitrate dehydrogenases. In the alignment of the amino acid sequences with those from several archaeal and mesophilic NADP-dependent isocitrate dehydrogenases, the residues involved in dinucleotide binding and isocitrate binding are well conserved. We have developed methods for the expression in Escherichia coli and purification of the enzyme from H. volcanii. This expression was carried out in E. coli as inclusion bodies using the cytoplasmic expression vector pET3a. The enzyme was refolded by solubilisation in 8 M urea followed by dilution into a buffer containing EDTA, MgCl(2) and 3 M NaCl. Maximal activity was obtained after several hours incubation at room temperature.
Collapse
Affiliation(s)
- Mónica Camacho
- División de Bioquímica y Biología Molecular, Facultad de Ciencias, Universidad de Alicante, Ap. 99, E-03080, Spain
| | | | | |
Collapse
|
103
|
Singh SK, Miller SP, Dean A, Banaszak LJ, LaPorte DC. Bacillus subtilis isocitrate dehydrogenase. A substrate analogue for Escherichia coli isocitrate dehydrogenase kinase/phosphatase. J Biol Chem 2002; 277:7567-73. [PMID: 11751849 DOI: 10.1074/jbc.m107908200] [Citation(s) in RCA: 29] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
In Escherichia coli, the homodimeric Krebs cycle enzyme isocitrate dehydrogenase (EcIDH) is regulated by reversible phosphorylation of a sequestered active site serine. The phosphorylation cycle is catalyzed by a bifunctional protein, IDH kinase/phosphatase (IDH-K/P). To better understand the nature of the interaction between EcIDH and IDH-K/P, we have examined the ability of an IDH homologue from Bacillus subtilis (BsIDH) to serve as a substrate for the kinase and phosphatase activities. BsIDH exhibits extensive sequence and structural similarities with EcIDH, particularly around the phosphorylated serine. Our previous crystallographic analysis revealed that the active site architecture of these two proteins is almost completely conserved. We now expand the comparison to include a number of biochemical properties. Both IDHs display nearly equivalent steady-state kinetic parameters for the dehydrogenase reaction. Both proteins are also phosphorylated by IDH-K/P in the same ratio (1 mole of phosphate per mole of monomer), and this stoichiometric phosphorylation correlates with an equivalent inhibition of IDH activity. Furthermore, tandem electrospray mass spectrometry demonstrates that BsIDH, like EcIDH, is phosphorylated on the corresponding active site serine residue (Ser-104). Despite the high degree of sequence, functional, and structural congruence between these two proteins, BsIDH is surprisingly a much poorer substrate of IDH-K/P than is EcIDH, with Michaelis constants for the kinase and phosphatase activities elevated by 60- and 3,450-fold, respectively. These drastically disparate values might result from restricted access to the active site cavity and/or from the lack of a potential docking site for IDH-K/P.
Collapse
Affiliation(s)
- Satinder K Singh
- Department of Biochemistry, Molecular Biology, and Biophysics, University of Minnesota, 321 Church Street SE, Minneapolis, MN 55455, USA
| | | | | | | | | |
Collapse
|
104
|
Cloning and expression of Aspergillus niger icdA gene encoding mitochondrial NADP+-specific isocitrate dehydrogenase. J Biosci Bioeng 2002. [DOI: 10.1016/s1389-1723(02)80005-6] [Citation(s) in RCA: 14] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022]
|
105
|
Steen IH, Madern D, Karlström M, Lien T, Ladenstein R, Birkeland NK. Comparison of isocitrate dehydrogenase from three hyperthermophiles reveals differences in thermostability, cofactor specificity, oligomeric state, and phylogenetic affiliation. J Biol Chem 2001; 276:43924-31. [PMID: 11533060 DOI: 10.1074/jbc.m105999200] [Citation(s) in RCA: 54] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
With the aim of gaining insight into the molecular and phylogenetic relationships of isocitrate dehydrogenase (IDH) from hyperthermophiles, we carried out a comparative study of putative IDHs identified in the genomes of the eubacterium Thermotoga maritima and the archaea Aeropyrum pernix and Pyrococcus furiosus. An optimum for activity at 90 degrees C or above was found for each IDH. PfIDH and ApIDH were the most thermostable with a melting temperature of 103.7 and 109.9 degrees C, respectively, compared with 98.3 and 98.5 degrees C for TmIDH and AfIDH, respectively. Analytical ultracentrifugation revealed a tetrameric oligomeric state for TmIDH and a homodimeric state for ApIDH and PfIDH. TmIDH and ApIDH were NADP-dependent (K(m)((NADP)) of 55.2 and 44.4 microm, respectively) whereas PfIDH was NAD-dependent (K(m)((NAD)) of 68.3 microm). These data document that TmIDH represents a novel tetrameric NADP-dependent form of IDH and that PfIDH is a homodimeric NAD-dependent IDH not previously found among the archaea. The homodimeric NADP-IDH present in A. pernix is the most common form of IDH known so far. The evolutionary relationships of ApIDH, PfIDH, and TmIDH with all of the available amino acid sequences of di- and multimeric IDHs are described and discussed.
Collapse
Affiliation(s)
- I H Steen
- Department of Microbiology, University of Bergen, P. O. Box 7800, Jahnebakken 5, N-5020 Bergen, Norway
| | | | | | | | | | | |
Collapse
|
106
|
Singh SK, Matsuno K, LaPorte DC, Banaszak LJ. Crystal structure of Bacillus subtilis isocitrate dehydrogenase at 1.55 A. Insights into the nature of substrate specificity exhibited by Escherichia coli isocitrate dehydrogenase kinase/phosphatase. J Biol Chem 2001; 276:26154-63. [PMID: 11290745 DOI: 10.1074/jbc.m101191200] [Citation(s) in RCA: 51] [Impact Index Per Article: 2.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Isocitrate dehydrogenase from Bacillus subtilis (BsIDH) is a member of a family of metal-dependent decarboxylating dehydrogenases. Its crystal structure was solved to 1.55 A and detailed comparisons with the homologue from Escherichia coli (EcIDH), the founding member of this family, were made. Although the two IDHs are structurally similar, there are three notable differences between them. First, a mostly nonpolar beta-strand and two connecting loops in the small domain of EcIDH are replaced by two polar alpha-helices in BsIDH. Because of a 13-residue insert in this region of BsIDH, these helices protrude over the active site cleft of the opposing monomer. Second, a coil leading into this cleft, the so-called "phosphorylation" loop, is bent inward in the B. subtilis enzyme, narrowing the entrance to the active site from about 12 to 4 A. Third, although BsIDH is a homodimer, the two unique crystallographic subunits of BsIDH are not structurally identical. The two monomers appear to differ by a domain shift of the large domain relative to the small domain/clasp region, reminiscent of what has been observed in the open/closed conformations of EcIDH. In Escherichia coli, IDH is regulated by reversible phosphorylation by the bifunctional enzyme IDH kinase/phosphatase (IDH-K/P). The site of phosphorylation is Ser(113), which lies deep within the active site crevice. Structural differences between EcIDH and BsIDH may explain disparities in their abilities to act as substrates for IDH-K/P.
Collapse
Affiliation(s)
- S K Singh
- Department of Biochemistry, Molecular Biology, and Biophysics, University of Minnesota, Minneapolis, Minnesota 55455, USA
| | | | | | | |
Collapse
|
107
|
Stoddard BL. Trapping reaction intermediates in macromolecular crystals for structural analyses. Methods 2001; 24:125-38. [PMID: 11384188 DOI: 10.1006/meth.2001.1174] [Citation(s) in RCA: 25] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022] Open
Abstract
The development of "time-resolved" crystallographic methods, including trapping of reaction intermediates and rapid data collection, allows the comparative study of discrete structural species formed during a macromolecular reaction, such as enzymatic catalysis, ribozyme cleavage, or a protein photocycle. The primary technical details that must be addressed in such studies are the reaction initiation, the accumulation of a specific reaction species throughout the crystal, the lifetime of that species and of the crystal under the experimental conditions, and the method used to collect X-ray data. Methods of reaction initiation range from substrate diffusion, which is appropriate for the visualization of very long-lived intermediates, to photolysis, which is appropriate for the accumulation of rate-limited species with half-lives ranging from milliseconds to nanoseconds. This review discusses various methods for initiating turnover in crystals and trapping rate-limiting species for structural studies.
Collapse
Affiliation(s)
- B L Stoddard
- Division of Basic Sciences, Program in Structural Biology, Fred Hutchinson Cancer Research Center, 1100 Fairview Avenue North, A3-023, Seattle, Washington 98109, USA.
| |
Collapse
|
108
|
Panisko EA, McAlister-Henn L. Subunit interactions of yeast NAD+-specific isocitrate dehydrogenase. J Biol Chem 2001; 276:1204-10. [PMID: 11042198 DOI: 10.1074/jbc.m005056200] [Citation(s) in RCA: 27] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Yeast mitochondrial NAD(+)-specific isocitrate dehydrogenase is an octamer composed of four each of two nonidentical but related subunits designated IDH1 and IDH2. IDH2 was previously shown to contain the catalytic site, whereas IDH1 contributes regulatory properties including cooperativity with respect to isocitrate and allosteric activation by AMP. In this study, interactions between IDH1 and IDH2 were detected using the yeast two-hybrid system, but interactions between identical subunit polypeptides were not detected with this or other methods. A model for heterodimeric interactions between the subunits is therefore proposed for this enzyme. A corollary of this model, based on the three-dimensional structure of the homologous enzyme from Escherichia coli, is that some interactions between subunits occur at isocitrate binding sites. Based on this model, two residues (Lys-183 and Asp-217) in the regulatory IDH1 subunit were predicted to be important in the catalytic site of IDH2. We found that individually replacing these residues with alanine results in mutant enzymes that exhibit a drastic reduction in catalysis both in vitro and in vivo. Also based on this model, the two analogous residues (Lys-189 and Asp-222) of the catalytic IDH2 subunit were predicted to contribute to the regulatory site of IDH1. A K189A substitution in IDH2 was found to produce a decrease in activation of the enzyme by AMP and a loss of cooperativity with respect to isocitrate. A D222A substitution in IDH2 produces similar regulatory defects and a substantial reduction in V(max) in the absence of AMP. Collectively, these results suggest that the basic structural/functional unit of yeast isocitrate dehydrogenase is a heterodimer of IDH1 and IDH2 subunits and that each subunit contributes to the isocitrate binding site of the other.
Collapse
Affiliation(s)
- E A Panisko
- Department of Biochemistry, University of Texas Health Science Center, San Antonio, Texas 78229-3900, USA
| | | |
Collapse
|
109
|
Abstract
Orthologs typically retain the same function in the course of evolution. Using beta-decarboxylating dehydrogenase family as a model, we demonstrate that orthologs can be confidently identified. The strategy is based on our recent findings that substitutions of only a few amino acid residues in these enzymes are sufficient to exchange substrate and coenzyme specificities. Hence, the few major specificity determinants can serve as reliable markers for determining orthologous or paralogous relationships. The power of this approach has been demonstrated by correcting similarity-based functional misassignment and discovering new genes and related pathways, and should be broadly applicable to other enzyme families.
Collapse
Affiliation(s)
- R Chen
- Department of Biochemistry, College of Medicine, University of Saskatchewan, Saskatoon, Canada.
| | | |
Collapse
|
110
|
Chen R, Yang H. A highly specific monomeric isocitrate dehydrogenase from Corynebacterium glutamicum. Arch Biochem Biophys 2000; 383:238-45. [PMID: 11185559 DOI: 10.1006/abbi.2000.2082] [Citation(s) in RCA: 34] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
The monomeric isocitrate dehydrogenase (IDH) of Corynebacterium glutamicum is compared to the topologically distinct dimeric IDH of Escherichia coli. Both IDHs have evolved to efficiently catalyze identical reactions with similar pH optimum as well as striking specificity toward NADP and isocitrate. However, the monomeric IDH is 10-fold more active (calculated as kcat/Km.isocitrate/Km.NADP) and 7-fold more NADP-specific than the dimeric enzyme, favoring NADP over NAD by a factor of 50,000. Such an extraordinary coenzyme specificity is not rivaled by any other characterized dehydrogenases. In addition, the monomeric enzyme is 10-fold more specific for isocitrate. The spectacular substrate specificity may be predominantly attributed to the isocitrate-assisted stabilization of catalytic complex during hydride transfer. No significant overall sequence identity is found between the monomeric and dimeric enzymes. However, structure-based alignment leads to the identification of three regions in the monomeric enzyme that match closely the three motifs located in the central region of dimeric IDHs and the homologous isopropylmalate dehydrogenases. The role of Lys253 as catalytic residue has been demonstrated by site-directed mutagenesis. Our results suggest that monomeric and dimeric forms of IDHs are functionally and structurally homologous.
Collapse
Affiliation(s)
- R Chen
- Department of Biochemistry, College of Medicine, University of Saskatchewan, Saskatoon, Canada.
| | | |
Collapse
|
111
|
Deng Z, Wang X, Kemp RG. Site-directed mutagenesis of the fructose 6-phosphate binding site of the pyrophosphate-dependent phosphofructokinase of Entamoeba histolytica. Arch Biochem Biophys 2000; 380:56-62. [PMID: 10900132 DOI: 10.1006/abbi.2000.1881] [Citation(s) in RCA: 4] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
Attempts to define the active site of pyrophosphate-dependent phosphofructokinase (PPi-PFK) using homology modeling based on the three-dimensional structure of the ATP-dependent PFKs from bacteria have been frustrated by low sequence identity between PPi- and ATP-PFKs in their carboxyl terminal halves. In the current study, alanine scanning mutagenesis of residues in the carboxyl terminal half of the PPi-PFK of Entamoeba histolytica coupled with comparative sequence analysis and computational modeling is used to identify residues that contribute to fructose 6-phosphate (fructose 6-P) binding. Of seven alanine mutants that were generated by site-directed mutagenesis, Arg377, Ser392, Arg405, Lys408, His415, His416, and Arg423, only the last mutant, Arg423Ala, was found to have dramatically lower affinity for fructose 6-P. Mutation of Arg 423 decreased k(cat) by 10,000-fold and decreased apparent affinity for fructose 6-P by 126-fold, while the K(m) for PPi increased only 4-fold. The second greatest effect was seen with Arg377Ala, which had a nearly 10-fold decrease in apparent affinity and an approximate 60-fold decrease in maximal activity. Another residue, Tyr420, was chosen for mutagenesis by its complete identity in all other PPi-PFK. This residue and its homologue in Escherichia coli ATP-PFK, His249, were mutated and shown to be very important for substrate binding in both enzymes.
Collapse
Affiliation(s)
- Z Deng
- Department of Microbiology and Immunology, The Chicago Medical School, North Chicago, Illinois 60064, USA
| | | | | |
Collapse
|
112
|
Fujita M, Toyooka Y, Tamegai H, Eguchi T, Kakinuma K. Arg-94 is crucial to the catalysis of 3-isopropylmalate dehydrogenase from Thermus thermophilus HB8. ACTA ACUST UNITED AC 2000. [DOI: 10.1016/s1381-1177(99)00091-0] [Citation(s) in RCA: 3] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/16/2022]
|
113
|
Karr DB, Emerich DW. Bradyrhizobium japonicum isocitrate dehydrogenase exhibits calcium-dependent hysteresis. Arch Biochem Biophys 2000; 376:101-8. [PMID: 10729195 DOI: 10.1006/abbi.1999.1687] [Citation(s) in RCA: 4] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
Bradyrhizobium japonicum NADP(+)-dependent isocitrate dehydrogenase was purified both from cultured cells and from the symbiotic form of the bacteria and was found to be identical in terms of N-terminal amino acid sequence, kinetics, and physicochemical properties. Magnesium and glycerol were absolute requirements for maintaining enzyme activity. The N-terminal amino acid sequence of the enzyme was more similar to the sequences from soybean and yeast than to other bacterial sequences. There was no immunological cross-reaction of antibodies from B. japonicum isocitrate dehydrogenase to extracts of soybean, pea, or Escherichia coli, but there was detectable, although weak, cross-reaction of antibodies from E. coli with the B. japonicum enzyme. B. japonicum isocitrate dehydrogenase displayed strong inhibition by NADH, indicating that during symbiotic nitrogen fixation the enzyme activity would be markedly reduced in planta. The enzyme displayed a calcium-dependent hysteresis, with a pronounced lag lasting as long as 2 min. Hysteresis was evident at concentrations of magnesium less than 0.5 mM and calcium greater than 1 microM. The hysteresis could be alleviated by excess magnesium or by 1,2-bis(2-aminophenoxy)ethane-N,N,N',N'-tetraacetic acid. The results suggest two roles for magnesium during catalysis; one magnesium may be needed to convert the enzyme into the steady-state form and the second needed for chelation of isocitrate for catalysis. The calcium-dependent hysteretic behavior of B. japonicum NADP(+)-isocitrate dehydrogenase suggested that this metal could serve as an intracellular regulator during symbiosis.
Collapse
Affiliation(s)
- D B Karr
- Department of Biochemistry, University of Missouri, Columbia, Missouri, 65211, USA
| | | |
Collapse
|
114
|
Soundar S, Danek BL, Colman RF. Identification by mutagenesis of arginines in the substrate binding site of the porcine NADP-dependent isocitrate dehydrogenase. J Biol Chem 2000; 275:5606-12. [PMID: 10681542 DOI: 10.1074/jbc.275.8.5606] [Citation(s) in RCA: 33] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Pig heart mitochondrial NADP-dependent isocitrate dehydrogenase is the most extensively studied among the mammalian isocitrate dehydrogenases. The crystal structure of Escherichia coli isocitrate dehydrogenase and sequence alignment of porcine with E. coli isocitrate dehydrogenase suggests that the porcine Arg(101), Arg(110), Arg(120), and Arg(133) are candidates for roles in substrate binding. The four arginines were separately mutated to glutamine using a polymerase chain reaction method. Wild type and mutant enzymes were each expressed in E. coli, isolated as maltose binding fusion proteins, then cleaved with thrombin, and purified to yield homogeneous porcine isocitrate dehydrogenase. The R120Q mutant has a specific activity, as well as K(m) values for isocitrate, Mn(2+), and NADP(+) similar to wild type enzyme, indicating that Arg(120) is not needed for function. The specific activities of R101Q, R110Q, and R133Q are 1.73, 1.30, and 19.7 micromols/min/mg, respectively, as compared with 39.6 units/mg for wild type enzyme. The R110Q and R133Q enzymes exhibit K(m) values for isocitrate that are increased more than 400- and 165-fold, respectively, as compared with wild type. The K(m) values for Mn(2+), but not for NADP(+), are also elevated indicating that binding of the metal-isocitrate complex is impaired in these mutants. It is proposed that the positive charges of Arg(110) and Arg(133) normally strengthen the binding of the negatively charged isocitrate by electrostatic attraction. The R101Q mutant shows smaller, but significant increases in the K(m) values for isocitrate and Mn(2+); however, the marked decrease in k(cat) suggests a role for Arg(101) in catalysis. The V(max) of wild type enzyme depends on the ionized form of an enzymic group of pK 5.5, and this pK(aes) is similar for the R101Q and R120Q enzymes. In contrast, the pK(aes) for R110Q and R133Q enzymes increases to 6.4 and 7.4, respectively, indicating that the positive charges of Arg(110) and Arg(133) normally lower the pK of the nearby catalytic base to facilitate its ionization. These results may be understood in terms of the structure of the porcine NADP-specific isocitrate dehydrogenase generated by the Insight II Modeler Program, based on the x-ray coordinates of the E. coli enzyme.
Collapse
Affiliation(s)
- S Soundar
- Department of Chemistry, University of Delaware, Newark, Delaware 19716, USA
| | | | | |
Collapse
|
115
|
Weiss C, Zeng Y, Huang J, Sobocka MB, Rushbrook JI. Bovine NAD+-dependent isocitrate dehydrogenase: alternative splicing and tissue-dependent expression of subunit 1. Biochemistry 2000; 39:1807-16. [PMID: 10677231 DOI: 10.1021/bi991691i] [Citation(s) in RCA: 12] [Impact Index Per Article: 0.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/30/2022]
Abstract
NAD+-dependent isocitrate dehydrogenase (IDH), a key regulatory enzyme in the Krebs cycle, is a multi-tetrameric enzyme. At least three of the subunits in the core tetramer of mammals are unique gene products. Subunits 1/beta and 2/gamma are considered to be regulatory, while subunits 3,4/alpha, comprising half the tetramer, are catalytic. The full sequence was obtained for the major subunit 1 cDNA in bovine heart, IDH 1-A. A second cDNA, rare in heart, was also identified (IDH 1-B). Differences in the two were confined to the 3'-region, suggesting alternative splicing. Screening of brain, kidney, and liver RNA showed the presence of IDH 1-A and 1-B and a third major species, IDH 1-C. Amplification of bovine genomic DNA by PCR across the regions of difference produced a single product. Comparison of the genomic and mRNA sequences showed that IDH 1-A resulted from splicing of exon W to exon Y, eliminating intron w, exon X, and intron x. IDH 1-B was formed by splice junctions between exon W, exon X, and exon Y. IDH 1-C resulted from splicing of exon W to exon X and subsequent retention of intron x. The 2 proteins predicted from these 3 mRNAs are identical over their first 357 residues. Protein IDH 1-A, resulting from a termination codon within exon Y, contains an additional 26 residues. Proteins IDH 1-B and 1-C derive from a common termination codon within exon X and contain an additional 28 residues. The two C-terminal regions differ notably in the number and nature of charged residues, resulting in proteins with a charge difference of 3.2 at pH 7.0. Subunit 1 sequences previously reported from other species grouped with one or the other of the bovine proteins. No evidence was found for alternative splicing in subunit 3,4/alpha. The results of the present study, together with recent work on the 2/gamma subunit [Brenner,V., Nyakatura, G., Rosenthal, A., and Platzer, M. (1998) Genomics 44, 8], indicate that the regulatory subunits of the enzyme, but not the catalytic, possess alternatively spliced forms varying in C-terminal properties with tissue-specific expression. The finding is suggestive of a mechanism for modulation of allosteric regulation tailored to the needs of different tissues.
Collapse
Affiliation(s)
- C Weiss
- Department of Biochemistry, State University of New York Downstate Medical Center, 450 Clarkson Avenue, Brooklyn, New York 11203, USA
| | | | | | | | | |
Collapse
|
116
|
Cherbavaz DB, Lee ME, Stroud RM, Koshland DE. Active site water molecules revealed in the 2.1 A resolution structure of a site-directed mutant of isocitrate dehydrogenase. J Mol Biol 2000; 295:377-85. [PMID: 10623532 DOI: 10.1006/jmbi.1999.3195] [Citation(s) in RCA: 26] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
Isocitrate dehydrogenase catalyses the two step, acid base, oxidative decarboxylation of isocitrate to alpha-ketoglutarate. Lysine 230 was suggested to act as proton donor based on geometry and spatial proximity to isocitrate. To clarify further the role of lysine 230, we co-crystallized the lysine-to-methionine mutant (K230M) with isocitrate and with alpha-ketoglutarate. Crystals were flash-frozen and the two structures were determined and refined to 2. 1 A. Several new features were identified relative to the wild-type structure. Seven side-chains previously unplaced in the wild-type structure were identified and included in the model, and the amino acid terminus was extended by an alanine residue. Many additional water molecules were identified. Examination of the K230M active sites (K230M isocitrate and K230M-ketoglutarate) revealed that tyrosine 160 protrudes further into the active site in the presence of either isocitrate or alpha-ketoglutarate in K230 M than it does in the wild-type structure. Also, methionine 230 was not as fully extended, and asparagine 232 rotates approximately 30 degrees toward the ligand permitting polar interactions. Outside the active site cleft a tetragonal volume of density was identified as a sulfate molecule. Its location and interactions suggest it may influence the equilibrium between the tetragonal and the orthorhombic forms of isocitrate dehydrogenase. Differences observed in the active site water structure between the wild-type and K230M structures were due to a single point mutation. A water molecule was located in the position equivalent to that occupied by the wild-type epsilon-amine of lysine 230; a water molecule in that location in K230M suggests it may influence catalysis in the mutant. Comparison of K230M complexed with isocitrate and alpha-ketoglutarate illuminates the influence a ligand has on active site water structure.
Collapse
Affiliation(s)
- D B Cherbavaz
- Department of Biochemistry, University of California, San Francisco, San Francisco, CA, 94143-0448, USA
| | | | | | | |
Collapse
|
117
|
Huang YC, Soundar S, Colman RF. Affinity cleavage at the divalent metal site of porcine NAD-specific isocitrate dehydrogenase. Protein Sci 2000; 9:104-11. [PMID: 10739252 PMCID: PMC2144432 DOI: 10.1110/ps.9.1.104] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/19/2022]
Abstract
A divalent metal ion, such as Mn2+, is required for the catalytic reaction and allosteric regulation of pig heart NAD-dependent isocitrate dehydrogenase. The enzyme is irreversibly inactivated and cleaved by Fe2+ in the presence of O2 and ascorbate at pH 7.0. Mn2+ prevents both inactivation and cleavage. Nucleotide ligands, such as NAD, NADPH, and ADP, neither prevent nor promote inactivation or cleavage of the enzyme by Fe2+. The NAD-specific isocitrate dehydrogenase is composed of three distinct subunits in the ratio 2alpha:1beta:1gamma. The results indicate that the oxidative inactivation and cleavage are specific and involve the 40 kDa alpha subunit of the enzyme. A pair of major peptides is generated during Fe2+ inactivation: 29.5 + 10.5 kDa, as determined by SDS-PAGE. Amino-terminal sequencing reveals that these peptides arise by cleavage of the Val262-His263 bond of the alpha subunit. No fragments are produced when enzyme is incubated with Fe2+ and ascorbate under denaturing conditions in the presence of 6 M urea, indicating that the native structure is required for the specific cleavage. These results suggest that His263 of the alpha subunit may be a ligand of the divalent metal ion needed for the reaction catalyzed by isocitrate dehydrogenase. Isocitrate enhances the inactivation of enzyme caused by Fe2+ in the presence of oxygen, but prevents the cleavage, suggesting that inactivation occurs by a different mechanism when metal ion is bound to the enzyme in the presence of isocitrate: oxidation of cysteine may be responsible for the rapid inactivation in this case. Affinity cleavage caused by Fe2+ implicates alpha as the catalytic subunit of the multisubunit porcine NAD-dependent isocitrate dehydrogenase.
Collapse
Affiliation(s)
- Y C Huang
- Department of Chemistry and Biochemistry, University of Delaware, Newark 19716, USA
| | | | | |
Collapse
|
118
|
Xu Y, Bhargava G, Wu H, Loeber G, Tong L. Crystal structure of human mitochondrial NAD(P)+-dependent malic enzyme: a new class of oxidative decarboxylases. Structure 1999. [DOI: 10.1016/s0969-2126(99)80115-4] [Citation(s) in RCA: 41] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/16/2022]
|
119
|
Oudot C, Jaquinod M, Cortay JC, Cozzone AJ, Jault JM. The isocitrate dehydrogenase kinase/phosphatase from Escherichia coli is highly sensitive to in-vitro oxidative conditions role of cysteine67 and cysteine108 in the formation of a disulfide-bonded homodimer. EUROPEAN JOURNAL OF BIOCHEMISTRY 1999; 262:224-9. [PMID: 10231385 DOI: 10.1046/j.1432-1327.1999.00395.x] [Citation(s) in RCA: 11] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/20/2022]
Abstract
Isocitrate dehydrogenase kinase/phosphatase (IDHK/P) is a homodimeric enzyme which controls the oxidative metabolism of Escherichia coli, and exibits a high intrinsic ATPase activity. When subjected to electrophoresis under nonreducing conditions, the purified enzyme migrates partially as a dimer. The proportion of the dimer over the monomer is greatly increased by treatment with cupric 1,10 phenanthrolinate or 5,5'-dithio-bis(2-nitrobenzoic acid), and fully reversed by dithiothreitol, indicating that covalent dimerization is produced by a disulfide bond. To identify the residue(s) involved in this intermolecular disulfide-bond, each of the eight cysteines of the enzyme was individually mutated into a serine. It was found that, under nonreducing conditions, the electrophoretic patterns of all corresponding mutants are identical to that of the wild-type, except for the Cys67-->Ser which migrates exclusively as a monomer and for the Cys108-->Ser which migrates preferentially as a dimer. Furthermore, in contrast to the wild-type enzyme and all the other mutants, the Cys67-->Ser mutant still migrates as a monomer after treatment with cupric 1,10 phenanthrolinate. This result indicates that the intermolecular disulfide bond involves only Cys67 in each IDHK/P wild-type monomer. This was further supported by mass spectrum analysis of the tryptic peptides derived from either the cupric 1,10 phenanthrolinate-treated wild-type enzyme or the native Cys108-->Ser mutant, which show that they both contain a Cys67-Cys67 disulfide bond. Moreover, both the cupric 1,10 phenanthrolinate-treated wild-type enzyme and the native Cys108-->Ser mutant contain another disulfide bond between Cys356 and Cys480. Previous results have shown that this additional Cys356-Cys480 disulfide bond is intramolecular [Oudot, C., Jault, J.-M., Jaquinod, M., Negre, D., Prost, J.-F., Cozzone, A.J. & Cortay, J.-C. (1998) Eur. J. Biochem. 258, 579-585].
Collapse
Affiliation(s)
- C Oudot
- Institut de Biologie et Chimie des Protéines, CNRS, Lyon, France
| | | | | | | | | |
Collapse
|
120
|
Olano J, Soler J, Busto F, De Arriaga D. Chemical modification of NADP-isocitrate dehydrogenase from Cephalosporium acremonium evidence of essential histidine and lysine groups at the active site. EUROPEAN JOURNAL OF BIOCHEMISTRY 1999; 261:640-9. [PMID: 10215879 DOI: 10.1046/j.1432-1327.1999.00297.x] [Citation(s) in RCA: 7] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/20/2022]
Abstract
NADP-isocitrate dehydrogenase from Cephalosporium acremonium CW-19 has been inactivated by diethyl pyrocarbonate following a first-order process giving a second-order rate constant of 3.0 m-1. s-1 at pH 6.5 and 25 degrees C. The pH-inactivation rate data indicated the participation of a group with a pK value of 6.9. Quantifying the increase in absorbance at 240 nm showed that six histidine residues per subunit were modified during total inactivation, only one of which was essential for catalysis, and substrate protection analysis would seem to indicate its location at the substrate binding site. The enzyme was not inactivated by 5, 5'-dithiobis(2-nitrobenzoate), N-ethylmaleimide or iodoacetate, which would point to the absence of an essential reactive cysteine residue at the active site. Pyridoxal 5'-phosphate reversibly inactivated the enzyme at pH 7.7 and 5 degrees C, with enzyme activity declining to an equilibrium value within 15 min. The remaining activity depended on the modifier concentration up to about 2 mm. The kinetic analysis of inactivation and reactivation rate data is consistent with a reversible two-step inactivation mechanism with formation of a noncovalent enzyme-pyridoxal 5'-phosphate complex prior to Schiff base formation with a probable lysyl residue of the enzyme. The analysis of substrate protection shows the essential residue(s) to be at the active site of the enzyme and probably to be involved in catalysis.
Collapse
Affiliation(s)
- J Olano
- Departmento de Bioquímica y Biología Molecular, Universidad de León Spain
| | | | | | | |
Collapse
|
121
|
Rhode DJ, Martin BL. Localized structural effects of electrostatic interactions in a thermostable enzyme. Biochem Biophys Res Commun 1999; 258:179-83. [PMID: 10222256 DOI: 10.1006/bbrc.1999.0571] [Citation(s) in RCA: 2] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/22/2022]
Abstract
The sequence and resolved structure of thermotrophic isopropylmalate dehydrogenase (IPMDH) and a related protein, mesotrophic isocitrate dehydrogenase (IDH), were compared emphasizing clusters of charged residues identified from X-ray crystallographic studies (Wallon, G., Kryger, G., Lovett, S. T., Oshima, T., Ringe, D., and Petsko, G. A. (1997) J. Mol. Biol. 266, 1016-1031). Mesotrophic isocitrate dehydrogenase was used for comparison because crystallographic data for a mesotrophic form of IPMDH was not available in the database. The structural features in the region of these clusters were compared and localized conformational differences were found in the thermotroph compared to the mesotroph. Because the overall topology of the two proteins is similar, it was concluded that these localized structural differences induced by electrostatic interactions between charged residues in the thermotrophic enzyme were responsible for the enhanced thermal stability of proteins from thermotroph.
Collapse
Affiliation(s)
- D J Rhode
- Department of Biochemistry, University of Tennessee, 858 Madison Avenue, Memphis, Tennessee, 38163, USA
| | | |
Collapse
|
122
|
Pitson SM, Mendz GL, Srinivasan S, Hazell SL. The tricarboxylic acid cycle of Helicobacter pylori. EUROPEAN JOURNAL OF BIOCHEMISTRY 1999; 260:258-67. [PMID: 10091606 DOI: 10.1046/j.1432-1327.1999.00153.x] [Citation(s) in RCA: 66] [Impact Index Per Article: 2.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/02/2023]
Abstract
The composition and properties of the tricarboxylic acid cycle of the microaerophilic human pathogen Helicobacter pylori were investigated in situ and in cell extracts using [1H]- and [13C]-NMR spectroscopy and spectrophotometry. NMR spectroscopy assays enabled highly specific measurements of some enzyme activities, previously not possible using spectrophotometry, in in situ studies with H. pylori, thus providing the first accurate picture of the complete tricarboxylic acid cycle of the bacterium. The presence, cellular location and kinetic parameters of citrate synthase, aconitase, isocitrate dehydrogenase, alpha-ketoglutarate oxidase, fumarate reductase, fumarase, malate dehydrogenase, and malate synthase activities in H. pylori are described. The absence of other enzyme activities of the cycle, including alpha-ketoglutarate dehydrogenase, succinyl-CoA synthetase, and succinate dehydrogenase also are shown. The H. pylori tricarboxylic acid cycle appears to be a noncyclic, branched pathway, characteristic of anaerobic metabolism, directed towards the production of succinate in the reductive dicarboxylic acid branch and alpha-ketoglutarate in the oxidative tricarboxylic acid branch. Both branches were metabolically linked by the presence of alpha-ketoglutarate oxidase activity. Under the growth conditions employed, H. pylori did not possess an operational glyoxylate bypass, owing to the absence of isocitrate lyase activity; nor a gamma-aminobutyrate shunt, owing to the absence of both gamma-aminobutyrate transaminase and succinic semialdehyde dehydrogenase activities. The catalytic and regulatory properties of the H. pylori tricarboxylic acid cycle enzymes are discussed by comparing their amino acid sequences with those of other, more extensively studied enzymes.
Collapse
Affiliation(s)
- S M Pitson
- School of Biochemistry and Molecular Genetics, University of New South Wales, Sydney, Australia
| | | | | | | |
Collapse
|
123
|
Lancien M, Gadal P, Hodges M. Molecular characterization of higher plant NAD-dependent isocitrate dehydrogenase: evidence for a heteromeric structure by the complementation of yeast mutants. THE PLANT JOURNAL : FOR CELL AND MOLECULAR BIOLOGY 1998; 16:325-33. [PMID: 9881153 DOI: 10.1046/j.1365-313x.1998.00305.x] [Citation(s) in RCA: 30] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 05/22/2023]
Abstract
NAD-dependent isocitrate dehydrogenase (IDH) is a key enzyme controlling the activity of the citric acid cycle. Despite more than 30 years of work, the plant enzyme remains poorly characterized. In this paper, a molecular characterization of the plant IDH is presented. Starting from probes defined according to sequence comparisons, three full-length cDNAs named Ntidha, Ntidhb and Ntidhc encoding different IDH subunits have been isolated from a Nicotiana tabacum cell suspension library. Sequence comparisons of the tobacco IDH subunits with the E. coli NADP-dependent enzyme, and the yeast IDH1 and IDH2 subunits suggested that only IDHa had the capacity to be catalytic as IDHb and IDHc were lacking certain residues implied in catalysis. The ability of antibodies raised against the recombinant IDHa protein to preferentially cross-react with IDH2 indicated that IDHa was more closely related to IDH2 than to IDH1. Complementation of yeast single IDH mutants showed that IDHb and IDHc could replace the function of the yeast regulatory IDH1 subunit. Although IDHa was unable to complement the IDH2 mutant, its catalytic function was revealed by the ability of two heteromeric enzymes, composed of either IDHa with IDHb or IDHa with IDHc, to replace IDH function in a yeast double mutant lacking both subunits. Expression studies at the protein and mRNA levels show that each subunit is present in both root and leaf tissues and that the three IDH genes respond in the same way to nitrate addition. Taken together, such observations suggest that the physiologically active enzyme is composed of the three different subunits. These results show for the first time that the plant IDH is heteromeric and that IDH subunit composition appears to be conserved between plant and animal kingdoms.
Collapse
Affiliation(s)
- M Lancien
- Institut de Biotechnologie des Plantes (CNRS ERS569), Université Paris XI, Orsay, France
| | | | | |
Collapse
|
124
|
Abstract
Several time-resolved crystallographic structures were determined over the past year, using a variety of trapping protocols and several data collection methods. A significant theme of recent time-resolved work is the importance of parallel comparative studies on the same protein, using different experimental protocols, in order to fully characterize the structural variation of the intermediates formed in the reaction pathway.
Collapse
Affiliation(s)
- B L Stoddard
- Fred Hutchinson Cancer Research Center, Seattle, WA 98109, USA.
| |
Collapse
|
125
|
Stoddard BL, Cohen BE, Brubaker M, Mesecar AD, Koshland DE. Millisecond Laue structures of an enzyme-product complex using photocaged substrate analogs. NATURE STRUCTURAL BIOLOGY 1998; 5:891-7. [PMID: 9783749 DOI: 10.1038/2331] [Citation(s) in RCA: 46] [Impact Index Per Article: 1.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/08/2022]
Abstract
The structure of a rate-limited product complex formed during a single initial round of turnover by isocitrate dehydrogenase has been determined. Photolytic liberation of either caged substrate or caged cofactor and Laue X-ray data collection were used to visualize the complex, which has a minimum half-life of approximately 10 milliseconds. The experiment was conducted with three different photoreactive compounds, each possessing a unique mechanism leading to the formation of the enzyme-substrate (ES) complex. Photoreaction efficiency and subsequent substrate affinities and binding rates in the crystal are critical parameters for these experiments. The structure suggests that CO2 dissociation is a rapid event that may help drive product formation, and that small conformational changes may contribute to slow product release.
Collapse
Affiliation(s)
- B L Stoddard
- Division of Basic Sciences, Program in Structural Biology, Fred Hutchinson Cancer Research Center, Seattle, Washington 98109, USA.
| | | | | | | | | |
Collapse
|
126
|
Imada K, Inagaki K, Matsunami H, Kawaguchi H, Tanaka H, Tanaka N, Namba K. Structure of 3-isopropylmalate dehydrogenase in complex with 3-isopropylmalate at 2.0 A resolution: the role of Glu88 in the unique substrate-recognition mechanism. Structure 1998; 6:971-82. [PMID: 9739088 DOI: 10.1016/s0969-2126(98)00099-9] [Citation(s) in RCA: 53] [Impact Index Per Article: 2.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/08/2023]
Abstract
BACKGROUND 3-Isopropylmalate dehydrogenase (IPMDH) and isocitrate dehydrogenase (ICDH) belong to a unique family of bifunctional decarboxylating dehydrogenases. Although the ICDH dimer catalyzes its reaction under a closed conformation, known structures of the IPMDH dimer (without substrate) adopt a fully open or a partially closed form. Considering the similarity in the catalytic mechanism, the IPMDH dimer must be in a fully closed conformation during the reaction. A large conformational change should therefore occur upon substrate binding. RESULTS We have determined the crystal structure of IPMDH from Thiobacillus ferrooxidans (Tf) complexed with 3-isopropylmalate (IPM) at 2.0 A resolution by the molecular replacement method. The structure shows a fully closed conformation and the substrate-binding site is quite similar to that of ICDH except for a region around the gamma-isopropyl group. The gamma group is recognized by a unique hydrophobic pocket, which includes Glu88, Leu91 and Leu92 from subunit 1 and Val193' from subunit 2. CONCLUSIONS A large movement of domain 1 is induced by substrate binding, which results in the formation of the hydrophobic pocket for the gamma-isopropyl moiety of IPM. A glutamic acid in domain 1, Glu88, participates in the formation of the hydrophobic pocket. The C beta and C gamma atoms of Glu88 interact with the gamma-isopropyl moiety of IPM and are central to the recognition of substrate. The acidic tip of Glu88 is likely to interact with the nicotinamide mononucleotide (NMN) ribose of NAD+ in the ternary complex. This structure clearly explains the substrate specificity of IPMDH.
Collapse
Affiliation(s)
- K Imada
- International Institute for Advanced Research, Matsushita Electric Industrial Co., Ltd., Seika, Japan.
| | | | | | | | | | | | | |
Collapse
|
127
|
Eppink MH, Schreuder HA, van Berkel WJ. Interdomain binding of NADPH in p-hydroxybenzoate hydroxylase as suggested by kinetic, crystallographic and modeling studies of histidine 162 and arginine 269 variants. J Biol Chem 1998; 273:21031-9. [PMID: 9694855 DOI: 10.1074/jbc.273.33.21031] [Citation(s) in RCA: 45] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
The conserved residues His-162 and Arg-269 of the flavoprotein p-hydroxybenzoate hydroxylase (EC 1.14.13.2) are located at the entrance of the interdomain cleft that leads toward the active site. To study their putative role in NADPH binding, His-162 and Arg-269 were selectively changed by site-specific mutagenesis. The catalytic properties of H162R, H162Y, and R269K were similar to the wild-type enzyme. However, less conservative His-162 and Arg-269 replacements strongly impaired NADPH binding without affecting the conformation of the flavin ring and the efficiency of substrate hydroxylation. The crystal structures of H162R and R269T in complex with 4-hydroxybenzoate were solved at 3.0 and 2.0 A resolution, respectively. Both structures are virtually indistinguishable from the wild-type enzyme-substrate complex except for the substituted side chains. In contrast to wild-type p-hydroxybenzoate hydroxylase, H162R is not inactivated by diethyl pyrocarbonate. NADPH protects wild-type p-hydroxybenzoate hydroxylase from diethylpyrocarbonate inactivation, suggesting that His-162 is involved in NADPH binding. Based on these results and GRID calculations we propose that the side chains of His-162 and Arg-269 interact with the pyrophosphate moiety of NADPH. An interdomain binding mode for NADPH is proposed which takes a novel sequence motif (Eppink, M. H. M., Schreuder, H. A., and van Berkel, W. J. H. (1997) Protein Sci. 6, 2454-2458) into account.
Collapse
Affiliation(s)
- M H Eppink
- Department of Biomolecular Sciences, Laboratory of Biochemistry, Wageningen Agricultural University, Dreijenlaan 3, 6703 HA Wageningen, The Netherlands
| | | | | |
Collapse
|
128
|
Chindamporn A, Nakagawa Y, Mizuguchi I, Chibana H, Doi M, Tanaka K. Repetitive sequences (RPSs) in the chromosomes of Candida albicans are sandwiched between two novel stretches, HOK and RB2, common to each chromosome. MICROBIOLOGY (READING, ENGLAND) 1998; 144 ( Pt 4):849-857. [PMID: 9579060 DOI: 10.1099/00221287-144-4-849] [Citation(s) in RCA: 37] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/18/2022]
Abstract
A novel sequence designated HOK, which is next to the RPS, a repetitive sequence specific to Candida albicans, was cloned and sequenced. HOK hybridized with all of the chromosomes on which the RPSs were located, but did not hybridize with chromosome 3, which does not harbour any RPSs. Sequence determination revealed that a portion of HOK has significant homology with the B and C1 fragments of Ca3, which is used as a molecular epidemiological probe. A homology search of the deduced amino acids of HOK against the protein database showed partial homology with an isocitrate dehydrogenase of Saccharomyces cerevisiae, although an ORF large enough to encode the enzyme was not detected. To verify the existence of other sequences homologous with HOK, a portion of the HOK sequence was amplified using PCR. Sequence determination of the 41 clones from the PCR products resulted in at least six HOK-homologous clones. Another RPS-containing clone, RB2, was isolated from the Pstl-digested chromosome R or 1. It was determined that RB2a, one of the subclones from RB2, hybridized with all of the chromosomes, including chromosome 3, with which neither HOK nor RPS hybridized. The hybridization profile also showed that RPS is located between HOK and RB2a on chromosomes other than chromosome 3.
Collapse
Affiliation(s)
- Ariya Chindamporn
- Laboratory of Medical Mycology, Research Institute for Disease Mechanism and Control, Nagoya University School of Medicine, Showa-ku, Nagoya 466, Japan
| | - Yoshiyuki Nakagawa
- Laboratory of Medical Mycology, Research Institute for Disease Mechanism and Control, Nagoya University School of Medicine, Showa-ku, Nagoya 466, Japan
| | - Ikuyo Mizuguchi
- Laboratory of Medical Mycology, Research Institute for Disease Mechanism and Control, Nagoya University School of Medicine, Showa-ku, Nagoya 466, Japan
| | - Hiroji Chibana
- Laboratory of Medical Mycology, Research Institute for Disease Mechanism and Control, Nagoya University School of Medicine, Showa-ku, Nagoya 466, Japan
| | - Matsuko Doi
- Laboratory of Medical Mycology, Research Institute for Disease Mechanism and Control, Nagoya University School of Medicine, Showa-ku, Nagoya 466, Japan
| | - Kenji Tanaka
- Laboratory of Medical Mycology, Research Institute for Disease Mechanism and Control, Nagoya University School of Medicine, Showa-ku, Nagoya 466, Japan
| |
Collapse
|
129
|
Chen R, Greer AF, Dean AM. Structural constraints in protein engineering--the coenzyme specificity of Escherichia coli isocitrate dehydrogenase. EUROPEAN JOURNAL OF BIOCHEMISTRY 1997; 250:578-82. [PMID: 9428712 DOI: 10.1111/j.1432-1033.1997.0578a.x] [Citation(s) in RCA: 15] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
In a previous study we reported on the successful inversion of coenzyme specificity in isocitrate dehydrogenase (IDH) from NADP to NAD [Chen, R., Greer, A. & Dean, A. M. (1995) A highly active decarboxylating dehydrogenase with rationally inverted coenzyme specificity, Proc. Natl Acad. Sci. USA 92, 11666-11670]. Here, we explore alternative means to generate NAD dependence in the NADP-dependent scaffold of Escherichia coli IDH. The results reveal that engineering a preference for NAD is constrained by the architecture of the IDH coenzyme binding pocket and confirms that the substituted Asp344 in the engineered enzyme is the major determinant of coenzyme specificity. Mutations in the 316-325 loop, which forms part of the coenzyme binding site, reduce activity through transmission of long-range conformational changes into the active site some 14 A distant. Conformational changes seen upon substituting Cys332-->Tyr are not directly involved with improving activity. Replacements at Cys201 reveal that subtle changes in the packing of hydrophobic residues (Met and Ile versus Leu) can elicit markedly different responses. We caution against using sequence alignments as the sole guide for mutagenesis and show how a combination of rational design of active-site residues based on X-ray structures and random substitutions at surrounding residues provides an efficient means to improve enzyme preference and catalytic efficiency towards novel substrates.
Collapse
Affiliation(s)
- R Chen
- Department of Biochemistry, College of Medicine, University of Saskatchewan, Saskatoon, Canada
| | | | | |
Collapse
|
130
|
Huang YC, Kumar A, Colman RF. Identification of the subunits and target peptides of pig heart NAD-specific isocitrate dehydrogenase modified by the affinity label 8-(4-bromo-2,3-dioxobutylthio)NAD. Arch Biochem Biophys 1997; 348:207-18. [PMID: 9390193 DOI: 10.1006/abbi.1997.0392] [Citation(s) in RCA: 8] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023]
Abstract
Pig heart NAD-dependent isocitrate dehydrogenase reacts with 8-(4-bromo-2,3-dioxobutylthio)-NAD (8-BDB-TNAD) with incorporation of 1.21 mol of reagent/mol of average subunit when the enzyme reaches the limit of 25% residual activity (Kumar, A., and Colman, R. F., Arch. Biochem. Biophys. 308, 357-366, 1994). Inclusion of NADPH decreases both the extent of inactivation and the reagent incorporation to 0.55 mol/mol of average subunit. We have now isolated the peptides labeled by radioactive 8-(4-bromo-2,3-dioxobutylthio)-[2-3H]NAD and have located them within the sequence of pig heart NAD-dependent isocitrate dehydrogenase. The enzyme is composed of three types of subunits, present as alpha 2 beta gamma. We have separated the subunits from unmodified and 8-BDBT[2-3H]NAD-modified enzymes by HPLC on a C4 reverse-phase column, after pretreatment of the enzymes with sodium dodecyl sulfate or urea, and compared the subunit sequences of the porcine enzyme with those of the corresponding subunits from other mammalian NAD-dependent isocitrate dehydrogenases. The predominant radioactivity of 8-BDBT[2-3H]NAD is observed in the alpha and gamma peaks, and the NADPH-protected enzyme exhibits marked reduction in incorporation into these peaks. However, evidence based on recombination of subunits from modified and unmodified enzymes indicates that only labeling of the alpha-subunit is responsible for inactivation by 8-BDB-TNAD. Cyanogen bromide was used to cleave the modified enzyme, and we purified one labeled peptide from the alpha-subunit (amino acids 84-177) as well as one from the gamma-subunit (amino acids 67-186). In the alpha-subunit, decreased modification by [7-14C]-phenylglyoxal of Arg88 and Arg98 after prior labeling of the enzyme by 8-BDB-TNAD indicates that these residues are the critical target sites of the reactive nucleotide analogue. We conclude that alpha subunit's Arg88 and Arg98 are both at or near the allosteric NADPH sites of the pig heart isocitrate dehydrogenase.
Collapse
Affiliation(s)
- Y C Huang
- Department of Chemistry and Biochemistry, University of Delaware, Newark 19716, USA
| | | | | |
Collapse
|
131
|
Wang FS, Whittam TS, Selander RK. Evolutionary genetics of the isocitrate dehydrogenase gene (icd) in Escherichia coli and Salmonella enterica. J Bacteriol 1997; 179:6551-9. [PMID: 9352899 PMCID: PMC179578 DOI: 10.1128/jb.179.21.6551-6559.1997] [Citation(s) in RCA: 45] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023] Open
Abstract
Sequences of the icd gene, encoding isocitrate dehydrogenase (IDH), were obtained for 33 strains representing the major phylogenetic lineages of Escherichia coli and Salmonella enterica. Evolutionary relationships of the strains based on variation in icd are generally similar to those previously obtained for several other housekeeping and for invasion genes, but the sequences of S. enterica subspecies V strains are unusual in being almost intermediate between those of the other S. enterica subspecies and E. coli. For S. enterica, the ratio of synonymous (silent) to nonsynonymous (replacement) nucleotide substitutions between pairs of strains was larger than comparable values for 12 other housekeeping and invasion genes, reflecting unusually strong purifying selection against amino acid replacement in the IDH enzyme. All amino acids involved in the catalytic activity and conformational changes of IDH are strictly conserved within and between species. In E. coli, the level of variation at the 3' end of the gene is elevated by the presence in some strains of a 165-bp replacement sequence supplied by the integration of either lambdoid phage 21 or defective prophage element e14. The 72 members of the E. coli Reference Collection (ECOR) and five additional E. coli strains were surveyed for the presence of phage 21 (as prophage) by PCR amplification of a phage 21-specific fragment in and adjacent to the host icd, and the sequence of the phage 21 segment extending from the 3' end of icd through the integrase gene (int) was determined in nine strains of E. coli. Phage 21 was found in 39% of E. coli strains, and its distribution among the ECOR strains is nonrandom. In two ECOR strains, the phage 21 int gene is interrupted by a 1,313-bp insertion element that has 99.3% nucleotide sequence identity with IS3411 of E. coli. The phylogenetic relationships of phage 21 strains derived from sequences of two different genomic regions were strongly incongruent, providing evidence of frequent recombination.
Collapse
Affiliation(s)
- F S Wang
- Institute of Molecular Evolutionary Genetics, Mueller Laboratory, Pennsylvania State University, University Park 16802, USA.
| | | | | |
Collapse
|
132
|
Zhao WN, McAlister-Henn L. Affinity purification and kinetic analysis of mutant forms of yeast NAD+-specific isocitrate dehydrogenase. J Biol Chem 1997; 272:21811-7. [PMID: 9268311 DOI: 10.1074/jbc.272.35.21811] [Citation(s) in RCA: 29] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/05/2023] Open
Abstract
Polyhistidine tags were added to the carboxyl termini of the two homologous subunits of yeast NAD+-specific isocitrate dehydrogenase (IDH). The tag in either the IDH1 or IDH2 subunit permits one-step affinity purification from yeast cellular extracts of catalytically active and allosterically responsive holoenzyme. This expression system was used to investigate subunit-specific contributions of residues with putative functions in adenine nucleotide binding. The primary effect of simultaneous replacement of the adjacent Asp-279 and Ile-280 residues in IDH1 with alanines is a dramatic loss of activation by AMP. In contrast, alanine replacement of the homologous Asp-286 and Ile-287 residues in IDH2 does not alter the allosteric response to AMP, but produces a 160-fold reduction in Vmax due to a 70-fold increase in the S0.5 value for NAD+. These results suggest that the targeted aspartate/isoleucine residues may contribute to regulator binding in IDH1 and to cofactor binding in IDH2, i.e. that these homologous residues are located in regions that have evolved for binding the adenine nucleotide components of different ligands. In other mutant enzymes, an alanine replacement of Asp-191 in IDH1 eliminates measurable catalytic activity, and a similar substitution of the homologous Asp-197 in IDH2 produces pleiotropic catalytic effects. A model is presented for the primary function of IDH2 in catalysis and of IDH1 in regulation, with crucial roles for these single aspartate residues in the communication and functional interdependence of the two subunits.
Collapse
Affiliation(s)
- W N Zhao
- Department of Biochemistry, University of Texas Health Science Center, San Antonio, Texas 78284-7760, USA
| | | |
Collapse
|
133
|
Mesecar AD, Stoddard BL, Koshland DE. Orbital steering in the catalytic power of enzymes: small structural changes with large catalytic consequences. Science 1997; 277:202-6. [PMID: 9211842 DOI: 10.1126/science.277.5323.202] [Citation(s) in RCA: 178] [Impact Index Per Article: 6.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
Small structural perturbations in the enzyme isocitrate dehydrogenase (IDH) were made in order to evaluate the contribution of precise substrate alignment to the catalytic power of an enzyme. The reaction trajectory of IDH was modified (i) after the adenine moiety of nicotinamide adenine dinucleotide phosphate was changed to hypoxanthine (the 6-amino was changed to 6-hydroxyl), and (ii) by replacing Mg2+, which has six coordinating ligands, with Ca2+, which has eight coordinating ligands. Both changes make large (10(-3) to 10(-5)) changes in the reaction velocity but only small changes in the orientation of the substrates (both distance and angle) as revealed by cryocrystallographic trapping of active IDH complexes. The results provide evidence that orbital overlap produced by optimal orientation of reacting orbitals plays a major quantitative role in the catalytic power of enzymes.
Collapse
Affiliation(s)
- A D Mesecar
- Department of Molecular and Cell Biology, Stanley Hall, University of California, Berkeley, CA 94720, USA
| | | | | |
Collapse
|
134
|
Didierjean C, Rahuel-Clermont S, Vitoux B, Dideberg O, Branlant G, Aubry A. A crystallographic comparison between mutated glyceraldehyde-3-phosphate dehydrogenases from Bacillus stearothermophilus complexed with either NAD+ or NADP+. J Mol Biol 1997; 268:739-59. [PMID: 9175858 DOI: 10.1006/jmbi.1997.0998] [Citation(s) in RCA: 34] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
Mutations have been introduced in the cytosolic glyceraldehyde-3-phosphate dehydrogenase (GAPDH) from Bacillus stearothermophilus in order to convert its cofactor selectivity from a specificity towards NAD into a preference for NADP. In the B-S mutant, five mutations (L33T, T34G, D35G, L187A, P188S) were selected on the basis of a sequence alignment with NADP-dependent chloroplastic GAPDHs. In the D32G-S mutant, two of the five mutations mentioned above (L187A, P188S) have been used in combination with another one designed from electrostatic considerations (D32G). Both mutants exhibit a dual-cofactor selectivity at the advantage of either NAD (B-S) or NADP (D32G-S). In order to analyse the cofactor-binding site plasticity at the molecular level, crystal structures of these mutants have been solved, when complexed with either NAD+ (D32G-Sn, resolution 2.5 A, R = 13.9%; B-Sn, 2.45 A, 19.3%) or NADP+ (D32G-Sp, 2.2 A, 19.2%; B-Sp, 2.5 A, 14.4%). The four refined models are very similar to that of the wild-type GAPDH and as expected resemble more closely the holo form than the apo form. In the B-S mutant, the wild-type low affinity for NADP+ seems to be essentially retained because of repulsive electrostatic contacts between the extra 2'-phosphate and the unchanged carboxylate group of residue D32. Such an antideterminant effect is not well compensated by putative attractive interactions which had been expected to arise from the newly-introduced side-chains. In this mutant, recognition of NAD+ is slightly affected with respect to that known on the wild-type, because mutations only weakly destabilize hydrogen bonds and van der Waals contacts originally present in the natural enzyme. Thus, the B-S mutant does not mimic efficiently the chloroplastic GAPDHs, and long-range and/or second-layer effects, not easily predictable from visual inspection of three-dimensional structures, need to be taken into account for designing a true "chloroplastic-like" mutant of cytosolic GAPDH. In the case of the D32G-S mutant, the dissociation constants for NAD+ and NADP+ are practically reversed with respect to those of the wild-type. The strong alteration of the affinity for NAD+ obviously proceeds from the suppression of the two wild-type hydrogen bonds between the adenosine 2'- and 3'-hydroxyl positions and the D32 carboxylate group. As expected, the efficient recognition of NADP+ is partly promoted by the removal of intra-subunit electrostatic repulsion (D32G) and inter-subunit steric hindrance (L187A, P188S). Another interesting feature of the reshaped NADP+-binding site is provided by the local stabilization of the extra 2'-phosphate which forms a hydrogen bond with the side-chain hydroxyl group of the newly-introduced S188. When compared to the presently known natural NADP-binding clefts, this result clearly demonstrates that an absolute need for a salt-bridge involving the 2'-phosphate is not required to switch the cofactor selectivity from NAD to NADP. In fact, as it is the case in this mutant, only a moderately polar hydrogen bond can be sufficient to make the extra 2'-phosphate of NADP+ well recognized by a protein environment.
Collapse
Affiliation(s)
- C Didierjean
- Laboratoire de Cristallographie et Modélisation des Matériaux Minéraux et Biologiques, IFR Protéines, Université Henri Poincaré Nancy I (URA CNRS 809), Vandoeuvre-lès-Nancy, France
| | | | | | | | | | | |
Collapse
|
135
|
Dean AM, Golding GB. Protein engineering reveals ancient adaptive replacements in isocitrate dehydrogenase. Proc Natl Acad Sci U S A 1997; 94:3104-9. [PMID: 9096353 PMCID: PMC20329 DOI: 10.1073/pnas.94.7.3104] [Citation(s) in RCA: 76] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023] Open
Abstract
Evolutionary analysis indicates that eubacterial NADP-dependent isocitrate dehydrogenases (EC 1.1.1.42) first evolved from an NAD-dependent precursor about 3.5 billion years ago. Selection in favor of utilizing NADP was probably a result of niche expansion during growth on acetate, where isocitrate dehydrogenase provides 90% of the NADPH necessary for biosynthesis. Amino acids responsible for differing coenzyme specificities were identified from x-ray crystallographic structures of Escherichia coli isocitrate dehydrogenase and the distantly related Thermus thermophilus NAD-dependent isopropylmalate dehydrogenase. Site-directed mutagenesis at sites lining the coenzyme binding pockets has been used to invert the coenzyme specificities of both enzymes. Reconstructed ancestral sequences indicate that these replacements are ancestral. Hence the adaptive history of molecular evolution is amenable to experimental investigation.
Collapse
Affiliation(s)
- A M Dean
- Department of Biological Chemistry, The Chicago Medical School, North Chicago, IL 60064-3095, USA
| | | |
Collapse
|
136
|
Wallon G, Kryger G, Lovett ST, Oshima T, Ringe D, Petsko GA. Crystal structures of Escherichia coli and Salmonella typhimurium 3-isopropylmalate dehydrogenase and comparison with their thermophilic counterpart from Thermus thermophilus. J Mol Biol 1997; 266:1016-31. [PMID: 9086278 DOI: 10.1006/jmbi.1996.0797] [Citation(s) in RCA: 121] [Impact Index Per Article: 4.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/04/2023]
Abstract
The basis of protein stability has been investigated by the structural comparison of themophilic enzymes with their mesophilic counterparts. A number of characteristics have been found that can contribute to the stabilization of thermophilic proteins, but no one is uniquely capable of imparting thermostability. The crystal structure of 3-isopropylmalate dehydrogenase (IPMDH) from the mesophiles Escherichia coli and Salmonella typhimurium have been determined by the method of molecular replacement using the known structure of the homologous Thermus thermophilus enzyme. The structure of the E. coli enzyme was refined at a resolution of 2.1 A to an R-factor of 17.3%, that of the S. typhimurium enzyme at 1.7 A resolution to an R-factor of 19.8%. The three structures were compared to elucidate the basis of the higher thermostability of the T. thermophilus enzyme. A mutant that created a cavity in the hydrophobic core of the thermophilic enzyme was designed to investigate the importance of packing density for thermostability. The structure of this mutant was analyzed. The main stabilizing features in the thermophilic enzyme are an increased number of salt bridges, additional hydrogen bonds, a proportionately larger and more hydrophobic subunit interface, shortened N and C termini and a larger number of proline residues. The mutation in the hydrophobic core of T. thermophilus IPMDH resulted in a cavity of 32 A3, but no significant effect on the activity and thermostability of the mutant was observed.
Collapse
Affiliation(s)
- G Wallon
- Rosenstiel Basic Medical Sciences Research Center, Brandeis University, Waltham, MA 02254-9910, USA
| | | | | | | | | | | |
Collapse
|
137
|
Synthesis of conformationally restricted substrate analogs and their interaction with 3-isopropylmalate dehydrogenase derived from Thermus thermophilus. Tetrahedron 1997. [DOI: 10.1016/s0040-4020(97)00104-x] [Citation(s) in RCA: 9] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/17/2022]
|
138
|
Suzuki T, Inoki Y, Yamagishi A, Iwasaki T, Wakagi T, Oshima T. Molecular and phylogenetic characterization of isopropylmalate dehydrogenase of a thermoacidophilic archaeon, Sulfolobus sp. strain 7. J Bacteriol 1997; 179:1174-9. [PMID: 9023199 PMCID: PMC178813 DOI: 10.1128/jb.179.4.1174-1179.1997] [Citation(s) in RCA: 18] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/03/2023] Open
Abstract
The archaeal leuB gene encoding isopropylmalate dehydrogenase of Sulfolobus sp. strain 7 was cloned, sequenced, and expressed in Escherichia coli. The recombinant Sulfolobus sp. enzyme was extremely stable to heat. The substrate and coenzyme specificities of the archaeal enzyme resembled those of the bacterial counterparts. Sedimentation equilibrium analysis supported an earlier proposal that the archaeal enzyme is homotetrameric, although the corresponding enzymes studied so far have been reported to be dimeric. Phylogenetic analyses suggested that the archaeal enzyme is homologous to mitochondrial NAD-dependent isocitrate dehydrogenases (which are tetrameric or octameric) as well as to isopropylmalate dehydrogenases from other sources. These results suggested that the present enzyme is the most primitive among isopropylmalate dehydrogenases belonging in the decarboxylating dehydrogenase family.
Collapse
Affiliation(s)
- T Suzuki
- Department of Life Science, Tokyo Institute of Technology, Yokohama, Japan
| | | | | | | | | | | |
Collapse
|
139
|
Wallon G, Yamamoto K, Kirino H, Yamagishi A, Lovett ST, Petsko GA, Oshima T. Purification, catalytic properties and thermostability of 3-isopropylmalate dehydrogenase from Escherichia coli. BIOCHIMICA ET BIOPHYSICA ACTA 1997; 1337:105-12. [PMID: 9003442 DOI: 10.1016/s0167-4838(96)00157-4] [Citation(s) in RCA: 24] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/03/2023]
Abstract
3-isopropylmalate dehydrogenase (IPMDH) from Escherichia coli was overexpressed, purified and crystallized. The enzyme was characterized and compared to its thermophilic counterpart from Thermus thermophilus strain HB8. As in the thermophile enzyme, the activity of E. coli IPMDH was dependent on the divalent cations, Mg2+ or Mn2+, with Mn2+ being the preferred cation. Activity was also strongly influenced by KCl: 0.3 M were necessary for the optimal activity. At 40 degrees C the K(m) of E. coli IPMDH was 105 microM for IPM and 321 microM for NAD, the kcat was 69 s-1. The half denaturation temperature was 64 degrees C, which was 20 degrees C lower than that of the thermophile enzyme.
Collapse
Affiliation(s)
- G Wallon
- Rosenstiel Medical Sciences Research Center, Brandeis University, Waltham, MA 02254-9110, USA
| | | | | | | | | | | | | |
Collapse
|
140
|
Engineering secondary structure to invert coenzyme specificity in isopropylmalate dehydrogenase. ACTA ACUST UNITED AC 1997. [DOI: 10.1016/s1080-8914(97)80078-6] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register]
|
141
|
Affiliation(s)
- R F Colman
- Department of Biochemistry, University of Delaware, Newark 19716, USA
| |
Collapse
|
142
|
Chen R, Greer A, Dean AM. Redesigning secondary structure to invert coenzyme specificity in isopropylmalate dehydrogenase. Proc Natl Acad Sci U S A 1996; 93:12171-6. [PMID: 8901552 PMCID: PMC37962 DOI: 10.1073/pnas.93.22.12171] [Citation(s) in RCA: 63] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023] Open
Abstract
Rational engineering of enzymes involves introducing key amino acids guided by a knowledge of protein structure to effect a desirable change in function. To date, all successful attempts to change specificity have been limited to substituting individual amino acids within a protein fold. However, the infant field of protein engineering will only reach maturity when changes in function can be generated by rationally engineering secondary structures. Guided by x-ray crystal structures and molecular modeling, site-directed mutagenesis has been used to systematically invert the coenzyme specificity of Thermus thermophilus isopropylmalate dehydrogenase from a 100-fold preference for NAD to a 1000-fold preference for NADP. The engineered mutant, which is twice as active as wild type, contains four amino acid substitutions and an alpha-helix and loop that replaces the original beta-turn. These results demonstrate that rational engineering of secondary structures to produce enzymes with novel properties is feasible.
Collapse
Affiliation(s)
- R Chen
- Department of Biological Chemistry, Chicago Medical School, IL 60064-3095, USA
| | | | | |
Collapse
|
143
|
Tipton PA, Quinn TP, Peisach J, Cook PF. Role of the divalent metal ion in the NAD:malic enzyme reaction: an ESEEM determination of the ground state conformation of malate in the E:Mn:malate complex. Protein Sci 1996; 5:1648-54. [PMID: 8844853 PMCID: PMC2143473 DOI: 10.1002/pro.5560050818] [Citation(s) in RCA: 5] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023]
Abstract
The conformation of L-malate bound at the active site of Ascaris suum malic enzyme has been investigated by electron spin echo envelope modulation spectroscopy. Dipolar interactions between Mn2+ bound to the enzyme active site and deuterium specifically placed at the 2-position, the 3R-position, and the 3S-position of L-malate were observed. The intensities of these interactions are related to the distance between each deuterium and Mn2+. Several models of possible Mn-malate complexes were constructed using molecular graphics techniques, and conformational searches were conducted to identify conformers of malate that meet the distance criteria defined by the spectroscopic measurements. These searches suggest that L-malate binds to the enzyme active site in the trans conformation, which would be expected to be the most stable conformer in solution, not in the gauche conformer, which would be more similar to the conformation required for oxidative decarboxylation of oxalacetate formed from L-malate at the active site of the enzyme.
Collapse
Affiliation(s)
- P A Tipton
- Department of Biochemistry, University of Missouri-Columbia 65211, USA.
| | | | | | | |
Collapse
|
144
|
Stoddard BL, Dean A, Bash PA. Combining Laue diffraction and molecular dynamics to study enzyme intermediates. NATURE STRUCTURAL BIOLOGY 1996; 3:590-5. [PMID: 8673602 DOI: 10.1038/nsb0796-590] [Citation(s) in RCA: 16] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 02/01/2023]
Abstract
Two separate techniques, Laue diffraction and computational molecular dynamics (MD) simulations, have been independently developed to allow the visualization and assessment of transient structural states. Recent studies on isocitrate dehydrogenase show that computational MD simulations of an enzymatic Michaelis complex are consistent with difference Fourier electron density maps of the same structure from a Laue experiment. The use of independent MD studies during crystallographic refinement has allowed us to assign with confidence a number of additional contacts and features important for hydride transfer. We find that unrestrained independent MD simulations provides a very useful method of cross-validation for highly mobile atoms in regions of experimental density that are poorly defined. Likewise, information from Laue difference maps provides information about substrate conformation and interactions that greatly facilitate MD simulations.
Collapse
Affiliation(s)
- B L Stoddard
- Fred Hutchinson Cancer Research Center, Seattle, Washington 98104, USA
| | | | | |
Collapse
|
145
|
Zhao WN, McAlister-Henn L. Assembly and Function of a Cytosolic Form of NADH-specific Isocitrate Dehydrogenase in Yeast. J Biol Chem 1996. [DOI: 10.1074/jbc.271.17.10347] [Citation(s) in RCA: 17] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
|
146
|
Abstract
In the process of identifying genes involved in optimization of lymphocyte activation, we have cloned the human mitochondrial NADP-dependent isocitrate dehydrogenase (mNADP-IDH) cDNA. The cDNA and its deduced amino acid (AA) sequence had a high degree of homology with those of the porcine and bovine. The heart and muscle had the highest constitutive expression of the gene. The expression of steady-state mRNA in the resting T and B lymphocytes was low but was induced after mitogen stimulation. The mRNA levels peaked around 48 h and remained elevated at 72 h. At the protein level, the mitochondrial but not cytosolic NADP-IDH activity was augmented after the mitogen stimulation. There was no cell cycle-dependent fluctuation of mNADP-IDH expression in synchronized Jurkat cells. In T and B cells, rapamycin (RAPA) could repress the mitogen-stimulated mNADP-IDH expression, although most of the early or late phase activation-related genes including a G-protein beta subunit-related gene H12.3 were not affected by the drug. The restricted expression of the gene in certain tissues and the activation-related expression in lymphocytes suggest that this gene might be necessary for optimal functions in heart, muscle, and the activated lymphocytes.
Collapse
Affiliation(s)
- H Luo
- Laboratory of Nephrology and Transplantation Immunology, Notre-Dame Hospital Research Center, Montreal, Quebec, Canada
| | | | | |
Collapse
|
147
|
Yang L, Luo H, Vinay P, Wu J. Molecular cloning of the cDNA of mouse mitochondrial NADP-dependent isocitrate dehydrogenase and the expression of the gene during lymphocyte activation. J Cell Biochem 1996; 60:400-10. [PMID: 8867815 DOI: 10.1002/(sici)1097-4644(19960301)60:3%3c400::aid-jcb11%3e3.0.co;2-o] [Citation(s) in RCA: 2] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/02/2023]
Abstract
The current report documents the molecular cloning of the mouse mitochondrial NADP-dependent isocitrate dehydrogenase (mNADP-IDH) cDNA. The cDNA was 1,863 bp in length and contained one open reading frame encoding a 523-residue polypeptide with a predicted molecular weight of 58 kDa. The cDNA and the deduced amino acid (AA) sequence of the mouse mNADP-IDH had a high degree of homology with those of porcine, bovine, alfalfa, and yeast. The recombinant mNADP-IDH expressed in Escherichia coli had active enzymatic function, as well as an expected molecular weight. The heart had the highest constitutive expression of the steady-state mNADP-IDH mRNA, followed by the kidney, while the expression of the gene in other tissues was low. The enzymatic activity of different tissues was in agreement with their mNADP-IDH mRNA levels. The resting lymphocytes had low constitutive expression of the gene, but the steady-state mRNA could be induced 48 h after mitogen stimulation. At the protein level, the resting lymphocytes had low enzymatic activity of mNADP-IDH, but the activity was augmented fivefold after mitogen stimulation. The cytosolic NADP-IDH, on the contrary, remained low or undetectable before and after the mitogen stimulation. Based on our current findings as well as the known roles of the mNADP-IDH in anabolism and in the isocitrate shuttle, it is conceivable that the mNADP-IDH is necessary for optimizing proliferation in lymphocytes.
Collapse
Affiliation(s)
- L Yang
- Department of Medicine, Notre-Dame Hospital Research Center and Nephrology Service, Montreal, Québec, Canada
| | | | | | | |
Collapse
|
148
|
Yang L, Luo H, Vinay P, Wu J. Molecular cloning of the cDNA of mouse mitochondrial NADP-dependent isocitrate dehydrogenase and the expression of the gene during lymphocyte activation. J Cell Biochem 1996. [DOI: 10.1002/(sici)1097-4644(19960301)60:3<400::aid-jcb11>3.0.co;2-o] [Citation(s) in RCA: 11] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/10/2022]
|
149
|
Dean AM, Shiau AK, Koshland DE. Determinants of performance in the isocitrate dehydrogenase of Escherichia coli. Protein Sci 1996; 5:341-7. [PMID: 8745412 PMCID: PMC2143340 DOI: 10.1002/pro.5560050218] [Citation(s) in RCA: 19] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023]
Abstract
The substrate specificity of the NADP-dependent isocitrate dehydrogenase of Escherichia coli was investigated by combining site-directed mutagenesis and utilization of alternative substrates. A comparison of the kinetics of the wild-type enzyme with 2R-malate reveals that the gamma-carboxylate of 2R,3S-isocitrate contributes a factor of 12,000,000 to enzyme performance. Analysis of kinetic data compiled for 10 enzymes and nine different substrates reveals that a factor of 1,650 can be ascribed to the hydrogen bond formed between S113 and the gamma-carboxylate of bound isocitrate, a factor of 150 to the negative charge of the gamma-carboxylate, and a factor of 50 for the gamma-methyl. These results are entirely consistent with X-ray structures of Michaelis complexes that show a hydrogen bond positions the gamma-carboxylate of isocitrate so that a salt bridge can form to the nicotinamide ring of NADP.
Collapse
Affiliation(s)
- A M Dean
- Department of Biological Chemistry, Chicago Medical School, Illinois 60064-3095, USA
| | | | | |
Collapse
|
150
|
Chen R, Grobler JA, Hurley JH, Dean AM. Second-site suppression of regulatory phosphorylation in Escherichia coli isocitrate dehydrogenase. Protein Sci 1996; 5:287-95. [PMID: 8745407 PMCID: PMC2143338 DOI: 10.1002/pro.5560050213] [Citation(s) in RCA: 19] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/01/2023]
Abstract
Inactivation of Escherichia coli isocitrate dehydrogenase upon phosphorylation at S113 depends upon the direct electrostatic repulsion of the negatively charged gamma-carboxylate of isocitrate by the negatively charged phosphoserine. The effect is mimicked by replacing S113 with aspartate or glutamate, which reduce performance (kcat/K(i).isocitrat/ Km.NADP) by a factor of 10(7). Here, we demonstrate that the inactivating effects of the electrostatic repulsion are completely eliminated by a second-site mutation, and provide the structural basis for this striking example of intragenic suppression. N115 is adjacent to S113 on one face of the D-helix, interacts with isocitrate and NADP+, and has been postulated to serve in both substrate binding and in catalysis. The single N115L substitution reduces affinity for isocitrate by a factor of 50 and performance by a factor of 500. However, the N115L substitution completely suppresses the inactivating electrostatic effects of S113D or S113E: the performance of the double mutants is 10(5) higher than the S113D and S113E single mutants. These mutations have little effect on the kinetics of alternative substrates, which lack the charged gamma-carboxylate of isocitrate. Both glutamate and aspartate at site 113 remain fully ionized in the presence of leucine. In the crystal structure of the N115L mutant, the leucine adopts a different conformer from the wild-type asparagine. Repacking around the leucine forces the amino-terminus of the D-helix away from the rest of the active site. The hydrogen bond between E113 and N115 in the S113E single mutant is broken in the S113E/N115L mutant, allowing the glutamate side chain to move away from the gamma-carboxylate of isocitrate. These movements increase the distance between the carboxylates, diminish the electrostatic repulsion, and lead to the remarkably high activity of the S113E/N115L mutant.
Collapse
Affiliation(s)
- R Chen
- Department of Biological Chemistry, Chicago Medical School, Illinois 60064-3095, USA
| | | | | | | |
Collapse
|