1
|
Sun W, Wang S, Tan X, Guo L, Liu W, Tian W, Zhang H, Jiang T, Meng W, Liu Y, Kang Z, Lü C, Gao C, Xu P, Ma C. Production of α-ketoisovalerate with whey powder by systemic metabolic engineering of Klebsiella oxytoca. Microb Cell Fact 2024; 23:264. [PMID: 39367476 PMCID: PMC11452931 DOI: 10.1186/s12934-024-02545-4] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 07/20/2024] [Accepted: 09/29/2024] [Indexed: 10/06/2024] Open
Abstract
BACKGROUND Whey, which has high biochemical oxygen demand and chemical oxygen demand, is mass-produced as a major by-product of the dairying industry. Microbial fermentation using whey as the carbon source may convert this potential pollutant into value-added products. This study investigated the potential of using whey powder to produce α-ketoisovalerate, an important platform chemical. RESULTS Klebsiella oxytoca VKO-9, an efficient L-valine producing strain belonging to Risk Group 1 organism, was selected for the production of α-ketoisovalerate. The leucine dehydrogenase and branched-chain α-keto acid dehydrogenase, which catalyzed the reductive amination and oxidative decarboxylation of α-ketoisovalerate, respectively, were inactivated to enhance the accumulation of α-ketoisovalerate. The production of α-ketoisovalerate was also improved through overexpressing α-acetolactate synthase responsible for pyruvate polymerization and mutant acetohydroxyacid isomeroreductase related to α-acetolactate reduction. The obtained strain K. oxytoca KIV-7 produced 37.3 g/L of α-ketoisovalerate from lactose, the major utilizable carbohydrate in whey. In addition, K. oxytoca KIV-7 also produced α-ketoisovalerate from whey powder with a concentration of 40.7 g/L and a yield of 0.418 g/g. CONCLUSION The process introduced in this study enabled efficient α-ketoisovalerate production from low-cost substrate whey powder. Since the key genes for α-ketoisovalerate generation were integrated in genome of K. oxytoca KIV-7 and constitutively expressed, this strain is promising in stable α-ketoisovalerate fermentation and can be used as a chassis strain for α-ketoisovalerate derivatives production.
Collapse
Affiliation(s)
- Weikang Sun
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Shuo Wang
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Xiaoxu Tan
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Leilei Guo
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Wei Liu
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Wenjia Tian
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Hui Zhang
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Tianyi Jiang
- School of Municipal and Environmental Engineering, Shandong Jianzhu University, Jinan, 250101, China
| | - Wensi Meng
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Yidong Liu
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Zhaoqi Kang
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Chuanjuan Lü
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Chao Gao
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China
| | - Ping Xu
- State Key Laboratory of Microbial Metabolism, Shanghai Jiao Tong University, Shanghai, 200240, China
| | - Cuiqing Ma
- State Key Laboratory of Microbial Technology, Shandong University, NO.72 Binhai Road, Qingdao, 266237, China.
| |
Collapse
|
2
|
dos Santos PF, Moreira DS, Baba EH, Volpe CM, Ruiz JC, Romanha AJ, Murta SM. Molecular characterization of lipoamide dehydrogenase gene in Trypanosoma cruzi populations susceptible and resistant to benznidazole. Exp Parasitol 2016; 170:1-9. [DOI: 10.1016/j.exppara.2016.08.006] [Citation(s) in RCA: 7] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 02/25/2016] [Revised: 07/18/2016] [Accepted: 08/22/2016] [Indexed: 01/15/2023]
|
3
|
Pseudomonas aeruginosa Uses Dihydrolipoamide Dehydrogenase (Lpd) to Bind to the Human Terminal Pathway Regulators Vitronectin and Clusterin to Inhibit Terminal Pathway Complement Attack. PLoS One 2015; 10:e0137630. [PMID: 26368530 PMCID: PMC4569481 DOI: 10.1371/journal.pone.0137630] [Citation(s) in RCA: 26] [Impact Index Per Article: 2.9] [Reference Citation Analysis] [Abstract] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/30/2015] [Accepted: 08/19/2015] [Indexed: 11/19/2022] Open
Abstract
The opportunistic human pathogen Pseudomonas aeruginosa controls host innate immune and complement attack. Here we identify Dihydrolipoamide dehydrogenase (Lpd), a 57 kDa moonlighting protein, as the first P. aeruginosa protein that binds the two human terminal pathway inhibitors vitronectin and clusterin. Both human regulators when bound to the bacterium inhibited effector function of the terminal complement, blocked C5b-9 deposition and protected the bacterium from complement damage. P. aeruginosa when challenged with complement active human serum depleted from vitronectin was severely damaged and bacterial survival was reduced by over 50%. Similarly, when in human serum clusterin was blocked by a mAb, bacterial survival was reduced by 44%. Thus, demonstrating that Pseudomonas benefits from attachment of each human regulator and controls complement attack. The Lpd binding site in vitronectin was localized to the C-terminal region, i.e. to residues 354-363. Thus, Lpd of P. aeruginosa is a surface exposed moonlighting protein that binds two human terminal pathway inhibitors, vitronectin and clusterin and each human inhibitor when attached protected the bacterial pathogen from the action of the terminal complement pathway. Our results showed insights into the important function of Lpd as a complement regulator binding protein that might play an important role in virulence of P. aeruginosa.
Collapse
|
4
|
Jiang W, Jiang Y, Bentley GJ, Liu D, Xiao Y, Zhang F. Enhanced production of branched-chain fatty acids by replacing β-ketoacyl-(acyl-carrier-protein) synthase III (FabH). Biotechnol Bioeng 2015; 112:1613-22. [DOI: 10.1002/bit.25583] [Citation(s) in RCA: 15] [Impact Index Per Article: 1.7] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 12/01/2014] [Revised: 02/09/2015] [Accepted: 02/23/2015] [Indexed: 01/13/2023]
Affiliation(s)
- Wen Jiang
- Department of Energy, Environmental, and Chemical Engineering; Washington University in St. Louis; 1 Brookings Drive, Saint Louis Missouri 63130
| | - Yanfang Jiang
- Department of Energy, Environmental, and Chemical Engineering; Washington University in St. Louis; 1 Brookings Drive, Saint Louis Missouri 63130
| | - Gayle J. Bentley
- Division of Biological & Biomedical Sciences; Washington University in St. Louis; Saint Louis Missouri
| | - Di Liu
- Division of Biological & Biomedical Sciences; Washington University in St. Louis; Saint Louis Missouri
| | - Yi Xiao
- Department of Energy, Environmental, and Chemical Engineering; Washington University in St. Louis; 1 Brookings Drive, Saint Louis Missouri 63130
| | - Fuzhong Zhang
- Department of Energy, Environmental, and Chemical Engineering; Washington University in St. Louis; 1 Brookings Drive, Saint Louis Missouri 63130
- Division of Biological & Biomedical Sciences; Washington University in St. Louis; Saint Louis Missouri
| |
Collapse
|
5
|
Hallström T, Mörgelin M, Barthel D, Raguse M, Kunert A, Hoffmann R, Skerka C, Zipfel PF. Dihydrolipoamide Dehydrogenase of Pseudomonas aeruginosa Is a Surface-Exposed Immune Evasion Protein That Binds Three Members of the Factor H Family and Plasminogen. THE JOURNAL OF IMMUNOLOGY 2012; 189:4939-50. [DOI: 10.4049/jimmunol.1200386] [Citation(s) in RCA: 26] [Impact Index Per Article: 2.2] [Reference Citation Analysis] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/19/2022]
|
6
|
Benamara H, Rihouey C, Jouenne T, Alexandre S. Impact of the biofilm mode of growth on the inner membrane phospholipid composition and lipid domains in Pseudomonas aeruginosa. BIOCHIMICA ET BIOPHYSICA ACTA-BIOMEMBRANES 2010; 1808:98-105. [PMID: 20849811 DOI: 10.1016/j.bbamem.2010.09.004] [Citation(s) in RCA: 46] [Impact Index Per Article: 3.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Received: 04/02/2010] [Revised: 09/02/2010] [Accepted: 09/03/2010] [Indexed: 11/30/2022]
Abstract
Many studies using genetic and proteomic approaches have revealed phenotypic differences between planktonic and sessile bacteria but the mechanisms of biofilm formation and the switch between the two growth modes are not well understood yet. In this study, we focused on inner membrane lipidome modifications when Pseudomonas aeruginosa cells were grown as biofilm. Lipid analyses were performed by Electrospray Ionization Mass Spectrometry. Results showed a drastic decrease of the uneven-numbered chain phospholipids and a slight increase of long chain PEs in sessile organisms as compared with planktonic counterparts, suggesting a better lipid stability in the bilayer and a decrease in membrane fluidity. The impact of sessile growth on lipid domains was then investigated by Brewster Angle Microscopy (BAM) and Atomic Force Microscopy (AFM). Observations showed that inner membrane lipids of P. aeruginosa formed domains when the pressure was close to physiological conditions and that these domains were larger for lipids extracted from biofilm bacteria. This is coherent with the mass spectrometry analyses.
Collapse
Affiliation(s)
- Hayette Benamara
- PBS laboratory, UMR 6270, FR 3038, CNRS, Proteomic Platform of the IFRMP23, University of Rouen, 76821 Mont-Saint-Aignan cedex, France
| | | | | | | |
Collapse
|
7
|
Abstract
Lactic acid bacteria (LAB) constitute a diverse group of Gram positive obligately fermentative microorganisms which include both beneficial and pathogenic strains. LAB generally have complex nutritional requirements and therefore they are usually associated with nutrient-rich environments such as animal bodies, plants and foodstuffs. Amino acids represent an important resource for LAB and their utilization serves a number of physiological roles such as intracellular pH control, generation of metabolic energy or redox power, and resistance to stress. As a consequence, the regulation of amino acid catabolism involves a wide set of both general and specific regulators and shows significant differences among LAB. Moreover, due to their fermentative metabolism, LAB amino acid catabolic pathways in some cases differ significantly from those described in best studied prokaryotic model organisms such as Escherichia coli or Bacillus subtilis. Thus, LAB amino acid catabolism constitutes an interesting case for the study of metabolic pathways. Furthermore, LAB are involved in the production of a great variety of fermented products so that the products of amino acid catabolism are also relevant for the safety and the quality of fermented products.
Collapse
Affiliation(s)
- María Fernández
- Instituto de Productos Lácteos de Asturias CSIC, Crta de Infiesto s/n, Villaviciosa, Asturias, Spain
| | | |
Collapse
|
8
|
Insertional inactivation of branched-chain alpha-keto acid dehydrogenase in Staphylococcus aureus leads to decreased branched-chain membrane fatty acid content and increased susceptibility to certain stresses. Appl Environ Microbiol 2008; 74:5882-90. [PMID: 18689519 DOI: 10.1128/aem.00882-08] [Citation(s) in RCA: 89] [Impact Index Per Article: 5.6] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Staphylococcus aureus is a major community and nosocomial pathogen. Its ability to withstand multiple stress conditions and quickly develop resistance to antibiotics complicates the control of staphylococcal infections. Adaptation to lower temperatures is a key for the survival of bacterial species outside the host. Branched-chain alpha-keto acid dehydrogenase (BKD) is an enzyme complex that catalyzes the early stages of branched-chain fatty acid (BCFA) production. In this study, BKD was inactivated, resulting in reduced levels of BCFAs in the membrane of S. aureus. Growth of the BKD-inactivated mutant was progressively more impaired than that of wild-type S. aureus with decreasing temperature, to the point that the mutant could not grow at 12 degrees C. The growth of the mutant was markedly stimulated by the inclusion of 2-methylbutyrate in the growth medium at all temperatures tested. 2-Methylbutyrate is a precursor of odd-numbered anteiso fatty acids and bypasses BKD. Interestingly, growth of wild-type S. aureus was also stimulated by including 2-methylbutyrate in the medium, especially at lower temperatures. The anteiso fatty acid content of the BKD-inactivated mutant was restored by the inclusion of 2-methylbutyrate in the medium. Fluorescence polarization measurements indicated that the membrane of the BKD-inactivated mutant was significantly less fluid than that of wild-type S. aureus. Consistent with this result, the mutant showed decreased toluene tolerance that could be increased by the inclusion of 2-methylbutyrate in the medium. The BKD-inactivated mutant was more susceptible to alkaline pH and oxidative stress conditions. Inactivation of the BKD enzyme complex in S. aureus also led to a reduction in adherence of the mutant to eukaryotic cells and its survival in a mouse host. In addition, the mutant offers a tool to study the role of membrane fluidity in the interaction of S. aureus with antimicrobial substances.
Collapse
|
9
|
Bunik VI, Degtyarev D. Structure-function relationships in the 2-oxo acid dehydrogenase family: substrate-specific signatures and functional predictions for the 2-oxoglutarate dehydrogenase-like proteins. Proteins 2008; 71:874-90. [PMID: 18004749 DOI: 10.1002/prot.21766] [Citation(s) in RCA: 68] [Impact Index Per Article: 4.3] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/05/2022]
Abstract
Structural relationship within the family of the thiamine diphosphate-dependent 2-oxo acid dehydrogenases was analyzed by combining different methods of sequence alignment with crystallographic and enzymological studies of the family members. For the first time, the sequence similarity of the homodimeric 2-oxoglutarate dehydrogenase to heterotetrameric 2-oxo acid dehydrogenases is established. The presented alignment of the catalytic domains of the dehydrogenases of pyruvate, branched-chain 2-oxo acids and 2-oxoglutarate unravels the sequence markers of the substrate specificity and the essential residues of the family members without the 3D structures resolved. Predicted dual substrate specificity of some of the 2-oxo acid dehydrogenases was confirmed experimentally. The results were used to decipher functions of the two hypothetical proteins of animal genomes, OGDHL and DHTKD1, similar to the 2-oxoglutarate dehydrogenase. Conservation of all the essential residues confirmed their catalytic competence. Sequence analysis indicated that OGDHL represents a previously unknown isoform of the 2-oxoglutarate dehydrogenase, whereas DHTKD1 differs from the homologs at the N-terminus and substrate binding pocket. The differences suggest changes in heterologous protein interactions and accommodation of more polar and/or bulkier structural analogs of 2-oxoglutarate, such as 2-oxoadipate, 2-oxo-4-hydroxyglutarate, or products of the carboligase reaction between a 2-oxodicarboxylate and glyoxylate or acetaldehyde. The signatures of the Ca2+-binding sites were found in the Ca2+-activated 2-oxoglutarate dehydrogenase and OGDHL, but not in DHTKD1. Mitochondrial localization was predicted for OGDHL and DHTKD1, with DHTKD1 probably localized also to nuclei. Medical implications of the obtained results are discussed in view of the possible associations of the 2-oxo acid dehydrogenases and DHTKD1 with neurodegeneration and cancer.
Collapse
Affiliation(s)
- Victoria I Bunik
- School of Bioinformatics and Bioengineering, Lomonosov Moscow State University, Moscow 119992, Russian Federation.
| | | |
Collapse
|
10
|
Bunik VI, Denton TT, Xu H, Thompson CM, Cooper AJL, Gibson GE. Phosphonate analogues of alpha-ketoglutarate inhibit the activity of the alpha-ketoglutarate dehydrogenase complex isolated from brain and in cultured cells. Biochemistry 2005; 44:10552-61. [PMID: 16060664 DOI: 10.1021/bi0503100] [Citation(s) in RCA: 71] [Impact Index Per Article: 3.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/28/2022]
Abstract
The alpha-ketoglutarate dehydrogenase complex (KGDHC), a control point of the tricarboxylic acid cycle, is partially inactivated in brain in many neurodegenerative diseases. Potent and specific KGDHC inhibitors are needed to probe how the reduced KGDHC activity alters brain function. Previous studies showed that succinyl phosphonate (SP) effectively inhibits muscle and Escherichia coli KGDHC [Biryukov, A. I., Bunik, V. I., Zhukov, Yu. N., Khurs, E. N., and Khomutov, R. M. (1996) FEBS Lett. 382, 167-170]. To identify the phosphonates with the highest affinity toward brain KGDHC and with the greatest effect in living cells, we investigated the ability of SP and several of its ethyl esters to inhibit brain KGDHC, other alpha-keto acid-dependent enzymes, and KGDHC in intact cells. At a concentration of 0.01 mM, SP and its phosphonoethyl (PESP) and carboxyethyl (CESP) esters completely inhibited isolated brain KGDHC even in the presence of a 200-fold higher concentration of its substrate [alpha-ketoglutarate (KG)], while the diethyl (DESP) and triethyl (TESP) esters were ineffective. In cultured human fibroblasts, 0.01 mM SP, PESP, or CESP produced 70% inhibition of KGDHC. DESP and TESP were also inhibitory in the cell system, but only after preincubation, suggesting the release of their charged groups by cellular esterases. Thus, SP and its monoethyl esters target cellular KGDHC directly, while the di- and triethyl esters are activated in intact cells. When tested on other enzymes that bind KG or related alpha-keto acids, SP had minimal effects and its two esters (CESP and TESP) were ineffective even at a concentration (0.1 mM) 1 order of magnitude higher than that which inhibited cellular KGDHC activity. The high specificity in targeting KGDHC, penetration into cells, and minimal transformation by cellular enzymes indicate that SP and its esters should be useful in studying the effects of reduced KGDHC activity on neuronal and brain function.
Collapse
Affiliation(s)
- Victoria I Bunik
- School of Bioinformatics and Bioengineering and Belozersky Institute of Physico-Chemical Biology, Moscow State University, Moscow 119992, Russian Federation.
| | | | | | | | | | | |
Collapse
|
11
|
Shingaki R, Kasahara Y, Iwano M, Kuwano M, Takatsuka T, Inoue T, Kokeguchi S, Fukui K. Induction of L-form-like cell shape change of Bacillus subtilis under microculture conditions. MICROBIOLOGY (READING, ENGLAND) 2003; 149:2501-2511. [PMID: 12949175 DOI: 10.1099/mic.0.26259-0] [Citation(s) in RCA: 9] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/18/2022]
Abstract
A remarkable cell shape change was observed in Bacillus subtilis strain 168 under microculture conditions on CI agar medium (Spizizen's minimal medium supplemented with a trace amount of yeast extract and Casamino acids). Cells cultured under a cover glass changed in form from rod-shaped to spherical, large and irregular shapes that closely resembled L-form cells. The cell shape change was observed only with CI medium, not with Spizizen's minimum medium alone or other rich media. The whole-cell protein profile of cells grown under cover glass and cells grown on CI agar plates differed in several respects. Tandem mass analysis of nine gel bands which differed in protein expression between the two conditions showed that proteins related to nitrate respiration and fermentation were expressed in the shape-changed cells grown under cover glass. The cell shape change of CI cultures was repressed when excess KNO3 was added to the medium. Whole-cell protein analysis of the normal rod-shaped cells grown with 0.1% KNO3 and the shape-changed cells grown without KNO3 revealed that the expression of the branched-chain alpha-keto acid dehydrogenase complex (coded by the bfmB gene locus) was elevated in the shape-changed cells. Inactivation of the bfmB locus resulted in the repression of cell shape change, and cells in which bfmB expression was induced by IPTG did show changes in shape. Transmission electron microscopy of ultrathin sections demonstrated that the shape-changed cells had thin walls, and plasmolysis of cells fixed with a solution including 0.1 M sucrose was observed. Clarifying the mechanism of thinning of the cell wall may lead to the development of a new type of cell wall biosynthetic inhibitor.
Collapse
Affiliation(s)
- Ryuji Shingaki
- Department of Oral Microbiology, Okayama University Graduate School of Medicine and Dentistry, Shikata-cho, Okayama 700-8525, Japan
| | - Yasuhiro Kasahara
- Department of Bioresource Science, Ibaraki University, School of Agriculture, Ami, Ibaraki 300-0393, Japan
| | - Megumi Iwano
- Department of Intercellular Communication, Graduate School of Biological Science, Nara Institute of Science and Technology, Ikoma 630-0101, Japan
| | - Masayoshi Kuwano
- Department of Bioinformatics and Genomics, Graduate School of Information Science, Nara Institute of Science and Technology, Ikoma 630-0101, Japan
| | - Tomomasa Takatsuka
- Department of Operative Dentistry, Okayama University Graduate School of Medicine and Dentistry, Shikata-cho, Okayama 700-8525, Japan
| | - Tetsuyoshi Inoue
- Department of Oral Microbiology, Okayama University Graduate School of Medicine and Dentistry, Shikata-cho, Okayama 700-8525, Japan
| | - Susumu Kokeguchi
- Department of Oral Microbiology, Okayama University Graduate School of Medicine and Dentistry, Shikata-cho, Okayama 700-8525, Japan
| | - Kazuhiro Fukui
- Department of Oral Microbiology, Okayama University Graduate School of Medicine and Dentistry, Shikata-cho, Okayama 700-8525, Japan
| |
Collapse
|
12
|
Madhusudhan KT, Sokatch JR. Purification of branched-chain keto acid dehydrogenase regulator from Pseudomonas putida. Methods Enzymol 2001; 324:329-35. [PMID: 10989441 DOI: 10.1016/s0076-6879(00)24242-0] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/17/2023]
Abstract
BkdR can be isolated in nearly pure form as a tetramer by this procedure, which involves hyperexpressing bkdR from a plasmid, purification by chromatography on DEAE-Sepharose CL-6B, heparin-Sepharose CL-6B, and dialysis to precipitate BkdR. BkdR is relatively insoluble in aqueous buffers but can be kept in solution in buffer with 50% (v/v) glycerol and 0.2 M NaCl. Cultures of E. coli DH5 alpha (pJRS119) should be maintained at 30 degrees to promote plasmid stability. Because BkdR is prone to form intermolecular disulfide bonds, buffers for SDS-PAGE should contain fresh 0.5% (v/v) 2-mercaptoethanol.
Collapse
Affiliation(s)
- K T Madhusudhan
- Central Arkansas Veterans Healthcare System, Pathology and Laboratory Medicine, Little Rock 72205, USA
| | | |
Collapse
|
13
|
Takahashi N, Sato T, Yamada T. Metabolic pathways for cytotoxic end product formation from glutamate- and aspartate-containing peptides by Porphyromonas gingivalis. J Bacteriol 2000; 182:4704-10. [PMID: 10940008 PMCID: PMC111344 DOI: 10.1128/jb.182.17.4704-4710.2000] [Citation(s) in RCA: 108] [Impact Index Per Article: 4.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Metabolic pathways involved in the formation of cytotoxic end products by Porphyromonas gingivalis were studied. The washed cells of P. gingivalis ATCC 33277 utilized peptides but not single amino acids. Since glutamate and aspartate moieties in the peptides were consumed most intensively, a dipeptide of glutamate or aspartate was then tested as a metabolic substrate of P. gingivalis. P. gingivalis cells metabolized glutamylglutamate to butyrate, propionate, acetate, and ammonia, and they metabolized aspartylaspartate to butyrate, succinate, acetate, and ammonia. Based on the detection of metabolic enzymes in the cell extracts and stoichiometric calculations (carbon recovery and oxidation/reduction ratio) during dipeptide degradation, the following metabolic pathways were proposed. Incorporated glutamylglutamate and aspartylaspartate are hydrolyzed to glutamate and aspartate, respectively, by dipeptidase. Glutamate is deaminated and oxidized to succinyl-coenzyme A (CoA) by glutamate dehydrogenase and 2-oxoglutarate oxidoreductase. Aspartate is deaminated into fumarate by aspartate ammonia-lyase and then reduced to succinyl-CoA by fumarate reductase and acyl-CoA:acetate CoA-transferase or oxidized to acetyl-CoA by a sequential reaction of fumarase, malate dehydrogenase, oxaloacetate decarboxylase, and pyruvate oxidoreductase. The succinyl-CoA is reduced to butyryl-CoA by a series of enzymes, including succinate-semialdehyde dehydrogenase, 4-hydroxybutyrate dehydrogenase, and butyryl-CoA oxidoreductase. A part of succinyl-CoA could be converted to propionyl-CoA through the reactions initiated by methylmalonyl-CoA mutase. The butyryl- and propionyl-CoAs thus formed could then be converted into acetyl-CoA by acyl-CoA:acetate CoA-transferase with the formation of corresponding cytotoxic end products, butyrate and propionate. The formed acetyl-CoA could then be metabolized further to acetate.
Collapse
Affiliation(s)
- N Takahashi
- Department of Oral Biochemistry, Tohoku University School of Dentistry, 4-1 Seiryo-machi, Aoba-ku, Sendai, 980-8575, Japan.
| | | | | |
Collapse
|
14
|
Takahashi N, Yamada T. Pathways for amino acid metabolism by Prevotella intermedia and Prevotella nigrescens. ORAL MICROBIOLOGY AND IMMUNOLOGY 2000; 15:96-102. [PMID: 11155172 DOI: 10.1034/j.1399-302x.2000.150205.x] [Citation(s) in RCA: 34] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 11/23/2022]
Abstract
Pathways for amino acid metabolism by Prevotella intermedia and Prevotella nigrescens were investigated. Prevotella strains grew anaerobically in tryptone-based medium and their growth increased upon the addition of aspartate to the medium. Washed cells of tryptone-grown strains metabolized aspartate to succinate, acetate, fumarate, malate, formate and ammonia, while from tryptone they produced isobutyrate and isovalerate in addition to the end products from aspartate. Cell extracts obtained from the tryptone-grown cells had aspartate ammonia-lyase for the conversion of aspartate to fumarate. Methylviologen-dependent fumarate reductase was found to reduce fumarate to succinate. A series of enzymatic activities, including fumarase, NAD-dependent malate dehydrogenase, oxaloacetate decarboxylase, methylviologen-dependent pyruvate oxidoreductase, phosphotransacetylase and acetate kinase, was detected for the oxidative conversion of fumarate to acetate. Pyruvate formate-lyase and NAD-dependent formate dehydrogenase were also found for the production and consumption of formate, respectively. Methylviologen: NAD(P) oxidoreductase was found to be responsible for linkage between these reductive and oxidative pathways. Furthermore, the cell extracts had branched-chain amino acid aminotransferase and methylviologen-dependent branched-chain 2-oxoacid oxidoreductase, concomitantly with NAD-dependent glutamate dehydrogenase. Valine and leucine could be converted to isobutyryl CoA and isovaleryl CoA, respectively, through the sequential catalyses of these enzymes, and consequently to isobutyrate and isovalerate, respectively.
Collapse
Affiliation(s)
- N Takahashi
- Department of Oral Biochemistry, Tohoku University School of Dentistry, 4-1 Seiryo-machi, Aoba-ku, Sendai, 980-8575, Japan
| | | |
Collapse
|
15
|
Ward DE, Ross RP, van der Weijden CC, Snoep JL, Claiborne A. Catabolism of branched-chain alpha-keto acids in Enterococcus faecalis: the bkd gene cluster, enzymes, and metabolic route. J Bacteriol 1999; 181:5433-42. [PMID: 10464218 PMCID: PMC94053 DOI: 10.1128/jb.181.17.5433-5442.1999] [Citation(s) in RCA: 41] [Impact Index Per Article: 1.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Genes encoding a branched-chain alpha-keto acid dehydrogenase from Enterococcus faecalis 10C1, E1alpha (bkdA), E1beta (bkdB), E2 (bkdC), and E3 (bkdD), were found to reside in the gene cluster ptb-buk-bkdDABC. The predicted products of ptb and buk exhibited significant homology to the phosphotransbutyrylase and butyrate kinase, respectively, from Clostridium acetobutylicum. Activity and redox properties of the purified recombinant enzyme encoded by bkdD indicate that E. faecalis has a lipoamide dehydrogenase that is distinct from the lipoamide dehydrogenase associated with the pyruvate dehydrogenase complex. Specific activity of the ptb gene product expressed in Escherichia coli was highest with the substrates valeryl-coenzyme A (CoA), isovaleryl-CoA, and isobutyryl-CoA. In cultures, a stoichiometric conversion of alpha-ketoisocaproate to isovalerate was observed, with a concomitant increase in biomass. We propose that alpha-ketoisocaproate is converted via the BKDH complex to isovaleryl-CoA and subsequently converted into isovalerate via the combined actions of the ptb and buk gene products with the concomitant phosphorylation of ADP. In contrast, an E. faecalis bkd mutant constructed by disruption of the bkdA gene did not benefit from having alpha-ketoisocaproate in the growth medium, and conversion to isovalerate was less than 2% of the wild-type conversion. It is concluded that the bkd gene cluster encodes the enzymes that constitute a catabolic pathway for branched-chain alpha-keto acids that was previously unidentified in E. faecalis.
Collapse
Affiliation(s)
- D E Ward
- Department of Biochemistry, Wake Forest University Medical Center, Winston-Salem, North Carolina 27157, USA
| | | | | | | | | |
Collapse
|
16
|
Debarbouille M, Gardan R, Arnaud M, Rapoport G. Role of bkdR, a transcriptional activator of the sigL-dependent isoleucine and valine degradation pathway in Bacillus subtilis. J Bacteriol 1999; 181:2059-66. [PMID: 10094682 PMCID: PMC93617 DOI: 10.1128/jb.181.7.2059-2066.1999] [Citation(s) in RCA: 110] [Impact Index Per Article: 4.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
A new gene, bkdR (formerly called yqiR), encoding a regulator with a central (catalytic) domain was found in Bacillus subtilis. This gene controls the utilization of isoleucine and valine as sole nitrogen sources. Seven genes, previously called yqiS, yqiT, yqiU, yqiV, bfmBAA, bfmBAB, and bfmBB and now referred to as ptb, bcd, buk, lpd, bkdA1, bkdA2, and bkdB, are located downstream from the bkdR gene in B. subtilis. The products of these genes are similar to phosphate butyryl coenzyme A transferase, leucine dehydrogenase, butyrate kinase, and four components of the branched-chain keto acid dehydrogenase complex: E3 (dihydrolipoamide dehydrogenase), E1alpha (dehydrogenase), E1beta (decarboxylase), and E2 (dihydrolipoamide acyltransferase). Isoleucine and valine utilization was abolished in bcd and bkdR null mutants of B. subtilis. The seven genes appear to be organized as an operon, bkd, transcribed from a -12, -24 promoter. The expression of the bkd operon was induced by the presence of isoleucine or valine in the growth medium and depended upon the presence of the sigma factor SigL, a member of the sigma 54 family. Transcription of this operon was abolished in strains containing a null mutation in the regulatory gene bkdR. Deletion analysis showed that upstream activating sequences are involved in the expression of the bkd operon and are probably the target of bkdR. Transcription of the bkd operon is also negatively controlled by CodY, a global regulator of gene expression in response to nutritional conditions.
Collapse
Affiliation(s)
- M Debarbouille
- Unité de Biochimie Microbienne, Institut Pasteur, URA 1300 du Centre National de la Recherche Scientifique, 75724 Paris Cedex 15, France.
| | | | | | | |
Collapse
|
17
|
Madhusudhan KT, Huang N, Braswell EH, Sokatch JR. Binding of L-branched-chain amino acids causes a conformational change in BkdR. J Bacteriol 1997; 179:276-9. [PMID: 8982009 PMCID: PMC178690 DOI: 10.1128/jb.179.1.276-279.1997] [Citation(s) in RCA: 11] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 02/03/2023] Open
Abstract
BkdR is the positive transcriptional activator of the inducible bkd operon of Pseudomonas putida. Evidence is accumulating that L-branched-chain amino acids are the inducers of the operon, and the data obtained in this study show that they induce a conformational change in BkdR. Addition of L-branched-chain amino acids increased the susceptibility of BkdR to trypsin with the cleavage between Arg-51 and Gln-52 on the C-terminal side of the DNA-binding domain. L-Valine also caused an increased fluorescence emission intensity and produced significant changes in the circular dichroism spectrum of BkdR. Analytical ultracentrifugation confirmed earlier data obtained from gel filtration that BkdR was a tetramer with a Stokes radius of 32 +/- 3 A and an axial ratio of 2:1.
Collapse
Affiliation(s)
- K T Madhusudhan
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190, USA
| | | | | | | |
Collapse
|
18
|
Hester K, Luo J, Burns G, Braswell EH, Sokatch JR. Purification of active E1 alpha 2 beta 2 of Pseudomonas putida branched-chain-oxoacid dehydrogenase. EUROPEAN JOURNAL OF BIOCHEMISTRY 1995; 233:828-36. [PMID: 8521848 DOI: 10.1111/j.1432-1033.1995.828_3.x] [Citation(s) in RCA: 10] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/31/2023]
Abstract
Active E1 component of Pseudomonas putida branched-chain-oxoacid dehydrogenase was purified from P. putida strains carrying pJRS84 which contains bkdR (encoding the transcriptional activator) and bkdA1 and bkdA2 (encoding the alpha and beta subunits). Expression was inducible, however, 45-, 39- and 37-kDa proteins were produced instead of the expected 45-kDa and 37-kDa proteins. The 45-kDa protein was identified as E1 alpha and the 37-kDa and 39-kDa proteins were identified as separate translational products of bkdA2 by their N-terminal sequences. The N-terminal amino acid of the 39-kDa protein was leucine instead of methionine. The 45-, 39- and 37-kDa proteins were also produced in wild-type P.putida. Translation of bkdA1 and bkdA2 from an Escherichia coli expression plasmid produced only 45-kDa and 39-kDa proteins, with N-terminal methionine on the 39-kDa protein. The insertion of guanine residues 5' to the first ATG of bkdA2 did not affect expression of E1 beta in P. putida including the N-terminal leucine which appears to eliminate the possibility of ribosome jumping. The Z-average molecular mass of the E1 component was determined by sedimentation equilibrium to be 172 +/- 9 kDa compared to a calculated value of 166 kDa for the heterotetramer and a Stokes radius of 5.1 nm. E1 alpha Ser313, which is homologous to the phosphorylated residue of rat liver E1 alpha, was converted to alanine resulting in about a twofold increase in Km, but no change in Kcat. S315A and S319A mutations had no effect on Km or Kcat indicating that these residues do not play a major part in catalysis of E1 alpha beta 2.
Collapse
Affiliation(s)
- K Hester
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190, USA
| | | | | | | | | |
Collapse
|
19
|
Madhusudhan KT, Huang N, Sokatch JR. Characterization of BkdR-DNA binding in the expression of the bkd operon of Pseudomonas putida. J Bacteriol 1995; 177:636-41. [PMID: 7836297 PMCID: PMC176638 DOI: 10.1128/jb.177.3.636-641.1995] [Citation(s) in RCA: 35] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/27/2023] Open
Abstract
The bkd operon of Pseudomonas putida consists of the structural genes encoding the components of the inducible branched-chain ketoacid dehydrogenase. BkdR, a positive regulator of the bkd operon and a homolog of Lrp of Escherichia coli is encoded by a structural gene adjacent to, and divergently transcribed from, the bkd operon of P. putida. BkdR was purified from E. coli containing bkdR cloned into pCYTEXP1, an expression vector. The molecular weight of BkdR obtained by gel filtration indicates that BkdR is a tetramer, and the abundance of BkdR in P. putida was estimated to be about 25 to 40 copies of the tetramer per cell. BkdR bound specifically to the region between bkdR and bkdA1, the latter being the first gene of the bkd operon. One BkdR-DNA complex was observed in gel mobility shift patterns. Approximately 100 bp was protected from the action of DNase I by BkdR, and the addition of L-branched-chain amino acids enhanced the appearance of hypersensitive sites in the protected region. There are four potential BkdR-DNA binding sequences in this region based on similarity to Lrp-binding consensus sequences. Like many other transcriptional activators, BkdR regulates expression of its structural gene. DNAs from several gram-negative bacteria hybridized to a probe containing bkdR, indicating the presence of bkdR-like genes in these organisms.
Collapse
Affiliation(s)
- K T Madhusudhan
- Department of Biochemistry and Molecular Biology, University of Oklahoma, Health Sciences Center, Oklahoma City 73190
| | | | | |
Collapse
|
20
|
Skinner DD, Morgenstern MR, Fedechko RW, Denoya CD. Cloning and sequencing of a cluster of genes encoding branched-chain alpha-keto acid dehydrogenase from Streptomyces avermitilis and the production of a functional E1 [alpha beta] component in Escherichia coli. J Bacteriol 1995; 177:183-90. [PMID: 8002616 PMCID: PMC176571 DOI: 10.1128/jb.177.1.183-190.1995] [Citation(s) in RCA: 24] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/28/2023] Open
Abstract
A cluster of genes encoding the E1 alpha, E1 beta, and E2 subunits of branched-chain alpha-keto acid dehydrogenase (BCDH) of Streptomyces avermitilis has been cloned and sequenced. Open reading frame 1 (ORF1) (E1 alpha), 1,146 nucleotides long, would encode a polypeptide of 40,969 Da (381 amino acids). ORF2 (E1 beta), 1,005 nucleotides long, would encode a polypeptide of 35,577 Da (334 amino acids). The intergenic distance between ORF1 and ORF2 is 73 bp. The putative ATG start codon of the incomplete ORF3 (E2) overlaps the stop codon of ORF2. Computer-aided searches showed that the deduced products of ORF1 and ORF2 resembled the corresponding E1 subunit (alpha or beta) of several prokaryotic and eukaryotic BCDH complexes. When these ORFs were overexpressed in Escherichia coli, proteins of about 41 and 34 kDa, which are the approximate masses of the predicted S. avermitilis ORF1 and ORF2 products, respectively, were detected. In addition, specific E1 [alpha beta] BCDH activity was detected in E. coli cells carrying the S. avermitilis ORF1 (E1 alpha) and ORF2 (E1 beta) coexpressed under the control of the T7 promoter.
Collapse
Affiliation(s)
- D D Skinner
- Bioprocess Research, Central Research Division, Pfizer Inc., Groton, Connecticut 06340
| | | | | | | |
Collapse
|
21
|
Madhusudhan KT, Lorenz D, Sokatch JR. The bkdR gene of Pseudomonas putida is required for expression of the bkd operon and encodes a protein related to Lrp of Escherichia coli. J Bacteriol 1993; 175:3934-40. [PMID: 8320210 PMCID: PMC204820 DOI: 10.1128/jb.175.13.3934-3940.1993] [Citation(s) in RCA: 40] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/29/2023] Open
Abstract
Branched-chain keto acid dehydrogenase is a multienzyme complex which is required for the metabolism of the branched-chain amino acids in Pseudomonas putida. The structural genes encoding all four proteins of the bkd operon have been cloned, and their nucleotide sequences have been determined (G. Burns, K. T. Madhusudhan, K. Hatter, and J. R. Sokatch, p. 177-184 in S. Silver, A. M. Chakrabarty, B. Iglewski, and S. Kaplan [ed.], Pseudomonas: Biotransformations, Pathogenesis, and Evolving Biotechnology, American Society for Microbiology, Washington D.C., 1990). An open reading frame which encoded a protein with 36.5% amino acid identity to the leucine-responsive regulatory protein (Lrp) of Escherichia coli was found immediately upstream of the bkd operon. Chromosomal mutations affecting this gene, named bkdR, resulted in a loss of ability to use branched-chain amino acids as carbon and energy sources and failure to produce branched-chain keto acid dehydrogenase. These mutations were complemented in trans by plasmids which contained intact bkdR. Mutations affecting bkdR did not have any effect on transport of branched-chain amino acids or transamination. Therefore, the bkdR gene product must affect expression of the bkd operon and regulation must be positive. Mutations affecting bkdR could also be complemented by plasmids containing lrp of E. coli. This is the first instance of a Lrp-like protein demonstrated to regulate expression of an operon outside of E. coli.
Collapse
Affiliation(s)
- K T Madhusudhan
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | |
Collapse
|
22
|
Palmer JA, Madhusudhan KT, Hatter K, Sokatch JR. Cloning, sequence and transcriptional analysis of the structural gene for LPD-3, the third lipoamide dehydrogenase of Pseudomonas putida. EUROPEAN JOURNAL OF BIOCHEMISTRY 1991; 202:231-40. [PMID: 1722146 DOI: 10.1111/j.1432-1033.1991.tb16367.x] [Citation(s) in RCA: 23] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/28/2022]
Abstract
The third lipoamide dehydrogenase structural gene of Pseudomonas putida, lpd3, was isolated from a library of P. putida PpG2 DNA cloned in Escherichia coli TB1. The nucleotide sequence of lpd3 and its flanking regions indicate that lpd3 is not part of an operon, which is unique for a prokaryotic lipoamide dehydrogenase. An open reading frame was found 207 bases upstream from the start of transcription, but is encoded on the strand opposite lpd3. There is no evidence of an open reading frame immediately downstream from lpd3. The coding region of lpd3 consists of 1401 bp, providing for 466 amino acids plus a stop codon with a G/C content of 62.4%. The transcriptional start site was located 33-bp upstream from the start of translation. The third lipoamide dehydrogenase (LPD-3) shares amino acid identity with the other two lipoamide dehydrogenases of P. putida, 45% with that of the 2-oxoglutarate dehydrogenase and pyruvate multienzyme complexes, and 45.9% with the lipoamide dehydrogenase of the branched-chain oxoacid complex. LPD-3 is more closely related to eukaryotic lipoamide dehydrogenases since it has 53.6% amino acid sequence identity with pig and human lipoamide dehydrogenases and 51.1% identity with yeast lipoamide dehydrogenase. LPD-3 was not produced in wild-type P. putida PpG2 under a variety of growth conditions. However, LPD-3 was produced in P. putida PpG2 carrying pSP14, a pKT240-based clone with the entire lpd3 gene plus 104 bases of the leader. The only demonstrated role of LPD-3 in P. putida is as a substitute for lipoamide dehydrogenase of the 2-oxoglutarate dehydrogenase and pyruvate multienzyme complexes when the latter is inactive or missing.
Collapse
Affiliation(s)
- J A Palmer
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | | | |
Collapse
|
23
|
Schulze E, Benen JA, Westphal AH, de Kok A. Interaction of lipoamide dehydrogenase with the dihydrolipoyl transacetylase component of the pyruvate dehydrogenase complex from Azotobacter vinelandii. EUROPEAN JOURNAL OF BIOCHEMISTRY 1991; 200:29-34. [PMID: 1908777 DOI: 10.1111/j.1432-1033.1991.tb21044.x] [Citation(s) in RCA: 14] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/29/2022]
Abstract
The interaction between lipoamide dehydrogenase (E3) and dihydrolipoyl transacetylase (E2p) from the pyruvate dehydrogenase complex was studied during the reconstitution of monomeric E3 apoenzymes from Azotobacter vinelandii and Pseudomonas fluorescens. The dimeric form of E3 is not only essential for catalysis but also for binding to the E2p core, because the apoenzymes as well as a monomeric holoenzyme from P. fluorescens, which can be stabilized as an intermediate at 0 degree C, do not bind to E2p. Lipoamide dehydrogenase from A. vinelandii contains a C-terminal extension of 15 amino acids with respect to glutathione reductase which is, in contrast to E3, presumably not part of a multienzyme complex. Furthermore, the last 10 amino acid residues of E3 are not visible in the electron density map of the crystal structure and are probably disordered. Therefore, the C-terminal tail of E3 might be an attractive candidate for a binding region. To probe this hypothesis, a set of deletions of this part was prepared by site-directed mutagenesis. Deletion of the last five amino acid residues did not result in significant changes. A further deletion of four amino acid residues resulted in a decrease of lipoamide activity to 5% of wild type, but the binding to E2p was unaffected. Therefore it is concluded that the C-terminus is not directly involved in binding to the E2p core. Deletion of the last 14 amino acids produced an enzyme with a high tendency to dissociate (Kd approximately 2.5 microM). This mutant binds only weakly to E2p. The diaphorase activity was still high. This indicates, together with the decreased Km for NADH, that the structure of the monomer is not appreciably changed by the mutation. Rather the orientation of the monomers with respect to each other is changed. It can be concluded that the binding region of E3 for E2p is constituted from structural parts of both monomers and binding occurs only when dimerization is complete.
Collapse
Affiliation(s)
- E Schulze
- Department of Biochemistry, Agricultural University, Wageningen, The Netherlands
| | | | | | | |
Collapse
|
24
|
van Berkel WJ, Benen JA, Snoek MC. On the FAD-induced dimerization of apo-lipoamide dehydrogenase from Azotobacter vinelandii and Pseudomonas fluorescens. Kinetics of reconstitution. EUROPEAN JOURNAL OF BIOCHEMISTRY 1991; 197:769-79. [PMID: 2029906 DOI: 10.1111/j.1432-1033.1991.tb15970.x] [Citation(s) in RCA: 25] [Impact Index Per Article: 0.8] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/29/2022]
Abstract
The apoenzymes of lipoamide dehydrogenase from pig heart and from Pseudomonas fluorescens were prepared at pH 2.7 and pH 4.0, respectively, using a hydrophobic interaction chromatography procedure recently developed for lipoamide dehydrogenase from Azotobacter vinelandii and other flavoproteins [Van Berkel et al. (1988) Eur. J. Biochem. 178, 197-207]. The apoenzyme from pig heart, having 5% of residual activity, shows an equilibrium between the monomeric and dimeric species. Both the yield and the degree of reconstitution of dimeric holoenzyme is 75% of starting material under optimal conditions. The kinetics of reconstitution of pig heart apoenzyme differ slightly from that obtained with the apoenzyme prepared by acid ammonium sulfate precipitation at pH 1.5 [Kalse, J. F. and Veeger, C. (1968) Biochim. Biophys. Acta 159, 244-256]. The apoenzyme from P. fluorescens is in the monomeric state and shows negligible residual activity. The yield and degree of reconstitution of the dimeric holoenzyme is more than 90% of starting material. Reconstitution of the apoenzymes from A. vinelandii and P. fluorescens involves minimally a two-step sequential process. Initial flavin-binding results in regaining of full dichloroindophenol activity, quenching of tryptophan fluorescence and strong increase of FAD fluorescence polarization. In the second step, dimerization occurs as reflected by regain of lipoamide activity, strongly increased FAD fluorescence and increased hyperchroism of the visible absorption spectrum. The kinetics of FAD-induced dimerization are strongly dependent on the apoenzyme used. At 0 degrees C, the monomeric apoenzyme-FAD complex is either stabilized (P. fluorescens) or only transiently detectable (A. vinelandii). Dimerization of P. fluorescens enzyme is strongly stimulated in the presence of NADH.
Collapse
Affiliation(s)
- W J van Berkel
- Department of Biochemistry, Agricultural University, Wageningen, The Netherlands
| | | | | |
Collapse
|
25
|
Palmer JA, Hatter K, Sokatch JR. Cloning and sequence analysis of the LPD-glc structural gene of Pseudomonas putida. J Bacteriol 1991; 173:3109-16. [PMID: 1902462 PMCID: PMC207904 DOI: 10.1128/jb.173.10.3109-3116.1991] [Citation(s) in RCA: 19] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/29/2022] Open
Abstract
Pseudomonas putida is able to produce three lipoamide dehydrogenases: (i) LPD-glc, which is the E3 component of the pyruvate and 2-ketoglutarate dehydrogenase complexes and the L-factor for the glycine oxidation system; (ii) LPD-val, which is the specific E3 component of the branched-chain keto acid dehydrogenase complex and is induced by growth on leucine, isoleucine, or valine; and (iii) LPD-3, which was discovered in a lpdG mutant and whose role is unknown. Southern hybridization with an oligonucleotide probe encoding the highly conserved redox-active site produced three bands corresponding to the genes encoding these three lipoamide dehydrogenases. The complete structural gene for LPD-glc, lpdG, was isolated, and its nucleotide sequence was determined. The latter consists of 476 codons plus a stop codon, TAA. The structural gene for LPD-glc is preceded by a partial open reading frame with strong similarity to the E2 component of 2-ketoglutarate dehydrogenase of Escherichia coli. This suggests that lpdG is part of the 2-ketoglutarate dehydrogenase operon. LPD-glc was expressed in Pseudomonas putida JS348 from pHP4 which contains a partial open reading frame corresponding to the E2 component, 94 bases of noncoding DNA, and the structural gene for lpdG. This result indicates that lpdG can be expressed separately from the other genes of the operon.
Collapse
Affiliation(s)
- J A Palmer
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | |
Collapse
|
26
|
Madhusudhan KT, Huang G, Burns G, Sokatch JR. Transcriptional analysis of the promoter region of the Pseudomonas putida branched-chain keto acid dehydrogenase operon. J Bacteriol 1990; 172:5655-63. [PMID: 2211503 PMCID: PMC526879 DOI: 10.1128/jb.172.10.5655-5663.1990] [Citation(s) in RCA: 15] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/30/2022] Open
Abstract
Branched-chain keto acid dehydrogenase is a multienzyme complex produced by Pseudomonas putida when it is grown in a minimal medium containing branched-chain amino acids. A 1.87-kilobase (kb) DNA fragment was cloned and sequenced which contained 0.24 kb of the E1 alpha structural gene and 1.6 kb of upstream DNA. There were 854 base pairs (bp) of noncoding DNA upstream of bkdA1, the first gene of the bkd operon, and 592 bp between the transcriptional and translational starts. The G + C content of the noncoding region was 56.7% compared with 65.2% for all the structural genes of the operon. A partial open reading frame was found on the strand opposite that of the bkd operon beginning at base 774. When the bkd promoter was cloned into the promoter probe vector pKT240, streptomycin resistance was obtained in P. putida but not Escherichia coli with the promoter in both orientations, which indicates either that the bkd promoter is bidirectional or that there are two promoters in this region. A series of ordered deletions on both sides of the proposed site of the start of transcription revealed that almost 700 bp upstream of the start of translation were required for expression. Streptomycin resistance was also obtained in an rpoN mutant of P. putida KT2440 containing constructs with the intact bkd promoter, indicating that the bkd operon does not require the rpoN sigma factor for expression. Another construct containing the bkd promoter, bkdA1, and bkdA2 in pKT240 was used to transform P. putida JS113, a mutant which was unable to produce the E1 subunits of the branched-chain keto acid dehydrogenase. In this case, very high inducible expression of the bkd operon was obtained.
Collapse
Affiliation(s)
- K T Madhusudhan
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | | | |
Collapse
|
27
|
Dietrichs D, Andreesen JR. Purification and comparative studies of dihydrolipoamide dehydrogenases from the anaerobic, glycine-utilizing bacteria Peptostreptococcus glycinophilus, Clostridium cylindrosporum, and Clostridium sporogenes. J Bacteriol 1990; 172:243-51. [PMID: 2294086 PMCID: PMC208424 DOI: 10.1128/jb.172.1.243-251.1990] [Citation(s) in RCA: 29] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/31/2022] Open
Abstract
Three different dihydrolipoamide dehydrogenases were purified to homogenity from the anaerobic glycine-utilizing bacteria Clostridium cylindrosporum, Clostridium sporogenes, and Peptostreptococcus glycinophilus, and their basic properties were determined. The enzyme isolated from P. glycinophilus showed the properties typical of dihydrolipoamide dehydrogenases: it was a dimer with a subunit molecular mass of 53,000 and contained 1 mol of flavin adenine dinucleotide and 2 redox-active sulfhydryl groups per subunit. Only NADH was active as a coenzyme for reduction of lipoamide. Spectra of the oxidized enzyme exhibited maxima at 230, 270, 353, and 453 nm, with shoulders at 370, 425, and 485 nm. The dihydrolipoamide dehydrogenases of C. cylindrosporum and C. sporogenes were very similar in their structural properties to the enzyme of P. glycinophilus except for their coenzyme specificity. The enzyme of C. cylindrosporum used NAD(H) as well as NADP(H), whereas the enzyme of C. sporogenes reacted only with NADP(H), and no reaction could be detected with NAD(H). Antibodies raised against the dihydrolipoamide dehydrogenase of C. cylindrosporum reacted with extracts of Clostridium acidiurici, Clostridium purinolyticum, and Eubacterium angustum, whereas antibodies raised against the enzymes of P. glycinophilus and C. sporogenes showed no cross-reaction with extracts from 42 organisms tested.
Collapse
Affiliation(s)
- D Dietrichs
- Institut für Mikrobiologie Universität, Göttingen, Federal Republic of Germany
| | | |
Collapse
|
28
|
Perham RN, Packman LC. 2-Oxo acid dehydrogenase multienzyme complexes: domains, dynamics, and design. Ann N Y Acad Sci 1989; 573:1-20. [PMID: 2699393 DOI: 10.1111/j.1749-6632.1989.tb14983.x] [Citation(s) in RCA: 67] [Impact Index Per Article: 1.9] [Reference Citation Analysis] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/02/2023]
Affiliation(s)
- R N Perham
- Department of Biochemistry, University of Cambridge, England
| | | |
Collapse
|
29
|
Freudenberg W, Dietrichs D, Lebertz H, Andreesen JR. Isolation of an atypically small lipoamide dehydrogenase involved in the glycine decarboxylase complex from Eubacterium acidaminophilum. J Bacteriol 1989; 171:1346-54. [PMID: 2537814 PMCID: PMC209752 DOI: 10.1128/jb.171.3.1346-1354.1989] [Citation(s) in RCA: 30] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/01/2023] Open
Abstract
The lipoamide dehydrogenase of the glycine decarboxylase complex was purified to homogeneity (8 U/mg) from cells of the anaerobe Eubacterium acidaminophilum that were grown on glycine. In cell extracts four radioactive protein fractions labeled with D-[2-14C]riboflavin could be detected after gel filtration, one of which coeluted with lipoamide dehydrogenase activity. The molecular mass of the native enzyme could be determined by several methods to be 68 kilodaltons, and an enzyme with a molecular mass of 34.5 kilodaltons was obtained by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Immunoblot analysis of cell extracts separated by sodium dodecyl sulfate-polyacrylamide or linear polyacrylamide gel electrophoresis resulted in a single fluorescent band. NADPH instead of NADH was the preferred electron donor of this lipoamide dehydrogenase. This was also indicated by Michaelis constants of 0.085 mM for NADPH and 1.1 mM for NADH at constant lipoamide and enzyme concentrations. The enzyme exhibited no thioredoxin reductase, glutathione reductase, or mercuric reductase activity. Immunological cross-reactions were obtained with cell extracts of Clostridium cylindrosporum, Clostridium sporogenes, Clostridium sticklandii, and bacterium W6, but not with extracts of other glycine- or purine-utilizing anaerobic or aerobic bacteria, for which the lipoamide dehydrogenase has already been characterized.
Collapse
Affiliation(s)
- W Freudenberg
- Institut für Mikrobiologie der Universität, Göttingen, Federal Republic of Germany
| | | | | | | |
Collapse
|
30
|
Burns G, Sykes PJ, Hatter K, Sokatch JR. Isolation of a third lipoamide dehydrogenase from Pseudomonas putida. J Bacteriol 1989; 171:665-8. [PMID: 2914869 PMCID: PMC209648 DOI: 10.1128/jb.171.2.665-668.1989] [Citation(s) in RCA: 25] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/03/2023] Open
Abstract
Pseudomonads are the only organisms so far known to produce two lipoamide dehydrogenases (LPDs), LPD-Val and LPD-Glc. LPD-Val is the specific E3 component of branched-chain oxoacid dehydrogenase, and LPD-Glc is the E3 component of 2-ketoglutarate and possibly pyruvate dehydrogenases and the L-factor of the glycine oxidation system. Three mutants of Pseudomonas putida, JS348, JS350, and JS351, affected in lpdG, the gene encoding LPD-Glc, have been isolated; all lacked 2-ketoglutarate dehydrogenase, but two, JS348 and JS351, had normal pyruvate dehydrogenase activity. The pyruvate and 2-ketoglutarate dehydrogenases of the wild-type strain of P. putida were both inhibited by anti-LPD-Glc, but the pyruvate dehydrogenase of the lpdG mutants was not inhibited, suggesting that the mutant pyruvate dehydrogenase E3 component was different from that of the wild type. The lipoamide dehydrogenase present in one of the lpdG mutants, JS348, was isolated and characterized. This lipoamide dehydrogenase, provisionally named LPD-3, differed in molecular weight, amino acid composition, and N-terminal amino acid sequence from LPD-Glc and LPD-Val. LPD-3 was clearly a lipoamide dehydrogenase as opposed to a mercuric reductase or glutathione reductase. LPD-3 was about 60% as effective as LPD-Glc in restoring 2-ketoglutarate dehydrogenase activity and completely restored pyruvate dehydrogenase activity in JS350. These results suggest that LPD-3 is a lipoamide dehydrogenase associated with an unknown multienzyme complex which can replace LPD-Glc as the E3 component of pyruvate and 2-ketoglutarate dehydrogenases in lpdG mutants.
Collapse
Affiliation(s)
- G Burns
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | | | |
Collapse
|
31
|
Carothers DJ, Pons G, Patel MS. Dihydrolipoamide dehydrogenase: functional similarities and divergent evolution of the pyridine nucleotide-disulfide oxidoreductases. Arch Biochem Biophys 1989; 268:409-25. [PMID: 2643922 DOI: 10.1016/0003-9861(89)90309-3] [Citation(s) in RCA: 108] [Impact Index Per Article: 3.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/01/2023]
Abstract
Dihydrolipoamide dehydrogenase (E3) is the common component of the three alpha-ketoacid dehydrogenase complexes oxidizing pyruvate, alpha-ketoglutarate, and the branched-chain alpha-ketoacids. E3 also participates in the glycine cleavage system. E3 belongs to the enzyme family called pyridine nucleotide-disulfide oxidoreductases, catalyzing the electron transfer between pyridine nucleotides and disulfide compounds. This review summarizes the information available for E3 from a variety of species, from a halophilic archaebacterium which has E3 but no alpha-ketoacid dehydrogenase complexes, to mammalian species. Evidence is reviewed for the existence of two E3 isozymes (one for pyruvate dehydrogenase complex and alpha-ketoglutarate dehydrogenase complex and the other for branched-chain alpha-ketoacid dehydrogenase complex) in Pseudomonas species and for possible mammalian isozymes of E3, one associated with the three alpha-ketoacid dehydrogenase complexes and one for the glycine cleavage system. The comparison of the complete amino acid sequences of E3 from Escherichia coli, yeast, pig, and human shows considerable homologies of certain amino acid residues or short stretches of sequences, especially in the specific catalytic and structural domains. Similar homology is found with the limited available amino acid sequence information on E3 from several other species. Sequence comparison is also presented for other member flavoproteins [e.g., glutathione reductase and mercury(II) reductase] of the pyridine nucleotide-disulfide oxidoreductase family. Based on the known tertiary structure of human glutathione reductase it may be possible to predict the domain structures of E3. Additionally, the sequence information may help to better understand a divergent evolutionary relationship among these flavoproteins in different species.
Collapse
Affiliation(s)
- D J Carothers
- Department of Biochemistry, Case Western Reserve University School of Medicine, Cleveland, Ohio 44106
| | | | | |
Collapse
|
32
|
Burns G, Brown T, Hatter K, Sokatch JR. Sequence analysis of the lpdV gene for lipoamide dehydrogenase of branched-chain-oxoacid dehydrogenase of Pseudomonas putida. EUROPEAN JOURNAL OF BIOCHEMISTRY 1989; 179:61-9. [PMID: 2917566 DOI: 10.1111/j.1432-1033.1989.tb14521.x] [Citation(s) in RCA: 45] [Impact Index Per Article: 1.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/03/2023]
Abstract
The production of two lipoamide dehydrogenases by Pseudomonas is so far unique. One, LPD-val, is the specific E3 component of the branched-chain-oxoacid dehydrogenase and the second, LPD-glc, is the E3 component of 2-oxoglutarate dehydrogenase and the L-factor of the glycine oxidation system. The objective of the present research was to determine the nucleotide sequence of the structural gene for LPD-val in order to compare its deduced amino acid structure with that of other redox-active disulfide flavoproteins. Northern blots using mRNA isolated from P. putida grown in media with branched-chain amino acids identified a transcript of 6.2 kb which is long enough to encode all the structural genes for the complex. The nucleotide sequence of the structural gene for LPD-val, lpdV, was determined and consists of 459 codons plus the stop codon. The open reading frame begins two bases after the stop codon for the E2 subunit and is composed of 66.3% G + C. Codon usage is characteristic of moderately strongly expressed genes. There is a ribosome-binding site preceding the ATG start codon and a strong candidate for a rho-independent terminator at the 3' end of the reading frame. The Mr of the protein encoded is 48,164 and when the Mr of FAD is added, the total Mr is 48,949, which is very close to the value of 49,000 obtained by SDS-polyacrylamide gel electrophoresis. Similarity comparisons of LPD-val with sequences of three other lipoamide dehydrogenases showed that LPD-val was somewhat more distantly related. It is probable that the lipoamide dehydrogenases and the glutathione and mercuric reductases evolved from a common ancestral flavoprotein.
Collapse
Affiliation(s)
- G Burns
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | | | |
Collapse
|
33
|
Allison N, Williams CH, Guest JR. Overexpression and mutagenesis of the lipoamide dehydrogenase of Escherichia coli. Biochem J 1988; 256:741-9. [PMID: 3066354 PMCID: PMC1135478 DOI: 10.1042/bj2560741] [Citation(s) in RCA: 19] [Impact Index Per Article: 0.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/04/2023]
Abstract
A 'split-gene' technique for the overexpression and mutagenesis of the gene encoding the lipoamide dehydrogenase of Escherichia coli was developed in order to overcome the instability problems encountered when attempting to mutate the intact gene. The lipoamide dehydrogenase gene, lpd, was dissected into two fragments which were separately subcloned into M13 vectors for mutagenesis in vitro followed by reconstitution in the pJLA504 expression vector under the transcriptional control of the lambda PR and lambda PL promoters and a temperature-sensitive lambda repressor. After thermo-induction, E. coli cells transformed with the plasmid carrying the reconstituted lpd gene contained 4-5 times more lipoamide dehydrogenase activity than is normally found in the wild-type organism. The strategy was used to engineer a Glu-188----Asp replacement in lipoamide dehydrogenase, and this generated an enzyme with markedly different kinetic properties.
Collapse
Affiliation(s)
- N Allison
- Department of Microbiology, University of Sheffield, U.K
| | | | | |
Collapse
|
34
|
Burns G, Brown T, Hatter K, Idriss JM, Sokatch JR. Similarity of the E1 subunits of branched-chain-oxoacid dehydrogenase from Pseudomonas putida to the corresponding subunits of mammalian branched-chain-oxoacid and pyruvate dehydrogenases. EUROPEAN JOURNAL OF BIOCHEMISTRY 1988; 176:311-7. [PMID: 3416875 DOI: 10.1111/j.1432-1033.1988.tb14283.x] [Citation(s) in RCA: 35] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/05/2023]
Abstract
The genes encoding proteins responsible for activity of the E1 component of branched-chain-oxoacid dehydrogenase of Pseudomonas putida have been subcloned and the nucleotide sequence of this region determined. Open reading frames encoding E1 alpha (bkdA1, 1233 bp) and E1 beta (bkdA2, 1020 bp) were identified with the aid of the N-terminal sequence of the purified subunits. The Mr of E1 alpha was 45,158 and of E1 beta was 37,007, both calculated without N-terminal methionine. The deduced amino acid sequences of E1 alpha and E1 beta had no similarity to the published sequences of the E1 subunits of pyruvate and 2-oxoglutarate dehydrogenases of Escherichia coli. However, there was substantial similarity between the E1 alpha subunits of Pseudomonas and rat liver branched-chain-oxoacid dehydrogenases. In particular, the region of the E1 alpha subunit of the mammalian branched-chain-oxoacid dehydrogenase which is phosphorylated, was found to be highly conserved in the Pseudomonas E1 alpha subunit. There was also considerable similarity between the E1 beta subunits of Pseudomonas branched-chain-oxoacid dehydrogenase and human pyruvate dehydrogenase.
Collapse
Affiliation(s)
- G Burns
- Department of Biochemistry and Molecular Biology, University of Oklahoma Health Sciences Center, Oklahoma City 73190
| | | | | | | | | |
Collapse
|
35
|
Burns G, Brown T, Hatter K, Sokatch JR. Comparison of the amino acid sequences of the transacylase components of branched chain oxoacid dehydrogenase of Pseudomonas putida, and the pyruvate and 2-oxoglutarate dehydrogenases of Escherichia coli. EUROPEAN JOURNAL OF BIOCHEMISTRY 1988; 176:165-9. [PMID: 3046941 DOI: 10.1111/j.1432-1033.1988.tb14264.x] [Citation(s) in RCA: 37] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 01/03/2023]
Abstract
The nucleotide sequence of bkdB, the structural gene for E2b, the transacylase component of branched-chain-oxoacid dehydrogenase of Pseudomonas putida has been determined and translated into its amino acid sequence. The start of bkdB was identified from the N-terminal sequence of E2b isolated from branched-chain-oxoacid dehydrogenase of the closely related species, P. aeruginosa. The reading frame was composed of 65.5% G + C with 82.3% of the codons ending in G or C. There was no intergenic space between bkdA2 and bkdB. No codons requiring minor tRNAs were utilized and the codon bias index indicated a preferential codon usage. The bkdB gene encoded 423 amino acids although the N-terminal methionine was absent from E2b prepared from P. aeruginosa. The relative molecular mass of the encoded protein was 45,134 (45,003 minus methionine) vs 47,000 obtained by SDS/polyacrylamide gel electrophoresis. There was a single lipoyl domain in E2b compared to three lipoyl domains in E2p, and one domain in E2o, the transacylases of pyruvate and 2-oxoglutarate dehydrogenases of Escherichia coli respectively. There was significant similarity between the lipoyl domain of E2b and of E2p and E2o as well as between the E1-E2 binding domains of E2b, E2p and E2o. There was no similarity between the E3 binding domain of E2b to E2p and E2o which may reflect the uniqueness of the E3 component of branched-chain-oxoacid dehydrogenase of P. putida. The conclusions drawn from these comparisons are that the transacylases of prokaryotic pyruvate, 2-oxoglutarate and branched-chain-oxoacid dehydrogenases descended from a common ancestral protein probably at about the same time.
Collapse
Affiliation(s)
- G Burns
- Department of Biochemistry and Molecular Biology, University of Oklahoma, Oklahoma City 73190
| | | | | | | |
Collapse
|
36
|
Hummel KB, Litwer S, Bradford AP, Aitken A, Danner DJ, Yeaman SJ. Nucleotide sequence of a cDNA for branched chain acyltransferase with analysis of the deduced protein structure. J Biol Chem 1988. [DOI: 10.1016/s0021-9258(18)68766-6] [Citation(s) in RCA: 31] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/22/2022] Open
|
37
|
Sokatch JR. Purification of branched-chain keto acid dehydrogenase and lipoamide dehydrogenase-valine from Pseudomonas. Methods Enzymol 1988; 166:342-50. [PMID: 3241566 DOI: 10.1016/s0076-6879(88)66046-0] [Citation(s) in RCA: 5] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/04/2023]
|
38
|
Sykes PJ, Sokatch JR. Cloning of genes for branched-chain keto acid dehydrogenase in Pseudomonas putida. Methods Enzymol 1988; 166:350-9. [PMID: 3071713 DOI: 10.1016/s0076-6879(88)66047-2] [Citation(s) in RCA: 2] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/04/2023]
|
39
|
Perham RN, Lowe PN. Isolation and properties of the branched-chain 2-keto acid and pyruvate dehydrogenase multienzyme complex from Bacillus subtilis. Methods Enzymol 1988; 166:330-42. [PMID: 3149394 DOI: 10.1016/s0076-6879(88)66045-9] [Citation(s) in RCA: 13] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/04/2023]
|
40
|
Carothers DJ, Raefsky-Estrin C, Pons G, Patel MS. Rat liver mitochondria contain two immunologically distinct dihydrolipoamide dehydrogenases. Arch Biochem Biophys 1987; 256:597-605. [PMID: 3619448 DOI: 10.1016/0003-9861(87)90617-5] [Citation(s) in RCA: 18] [Impact Index Per Article: 0.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/06/2023]
Abstract
We have raised antisera against dihydrolipoamide dehydrogenase. One antigen was isolated from purified bovine kidney pyruvate dehydrogenase complex (PDC). The other antigen was a commercial preparation of porcine heart dihydrolipoamide dehydrogenase (E3) which did not first involve purification of the alpha-keto acid dehydrogenase complex(es). Both antibody preparations cross-reacted with the E3 components of PDC, alpha-ketoglutarate dehydrogenase complex, and branched-chain keto acid dehydrogenase complex. This demonstrates the immunological identity of the E3 components. These sera totally precipitated E3 activity from the purified complexes, from purified preparations of E3, and from extracts of rat heart and kidney mitochondria. The two sera vary in their reaction with rat liver mitochondrial extracts: the anti PDC-E3 serum left residual E3 activity (approximately 50% of the original) that was precipitable by the anti-E3 anti-serum. This indicates that liver contains two immunologically distinct forms of E3. Metabolic assays measuring the differential effects of the two sera on the glycine decarboxylation reaction suggest that the form which is immunologically nonreactive with the anti-PDC-E3 serum could represent the E3 involved in the glycine cleavage system.
Collapse
|
41
|
Sykes PJ, Burns G, Menard J, Hatter K, Sokatch JR. Molecular cloning of genes encoding branched-chain keto acid dehydrogenase of Pseudomonas putida. J Bacteriol 1987; 169:1619-25. [PMID: 3549697 PMCID: PMC211990 DOI: 10.1128/jb.169.4.1619-1625.1987] [Citation(s) in RCA: 41] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 01/06/2023] Open
Abstract
We cloned the structural genes for the individual subunits of the branched-chain keto acid dehydrogenase multienzyme complex on a 7.8-kilobase EcoRI-SstI restriction fragment of Pseudomonas putida chromosomal DNA by cloning into the broad-host-range vector pKT230. A direct selection system for growth on valine-isoleucine agar was achieved by complementation of P. putida branched-chain keto acid dehydrogenase mutants. The recombinant plasmid, pSS1-1, increased expression of branched-chain keto acid dehydrogenase up to five times in wild-type P. putida. The complex was expressed constitutively in P. putida(pSS1-1) but was inducible in Escherichia coli HB101(pSS1-1) by high valine. E. coli minicells transformed with pSS1-1 produced three polypeptides which did not match the four polypeptides of the purified complex. To resolve this problem, we inserted P. putida DNA from pSS1-1 into pUC18 and pUC19. The pUC-derived plasmids were used as DNA templates in an E. coli transcription-translation system. Four polypeptides were produced from the pUC18-derived plasmid which had the correct molecular weights, showing that the structural genes had been cloned. Since only weak bands were produced with the pUC19-derived plasmid, the direction of transcription was established. The locations and order of all the structural genes of branched-chain keto acid dehydrogenase were located by restriction enzyme mapping.
Collapse
|