首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The regulation of phosphoribulokinase (PRK) and glyceraldehyde 3‐phosphate dehydrogenase (GAPDH) was investigated in a freshwater pennate diatom, Asterionella formosa Hassall, and compared to the well‐studied chlorophyte Chlamydomonas reinhardtii P. A. Dang. As has been reported for a marine centric diatom, in A. formosa, PRK was not regulated by reduction with dithiothreitol (DTT) apart from a weak induction in the presence of NADPH and DTT. However, NADPH‐GAPDH was strongly activated when reduced, in contrast to a previous report on a diatom. Surprisingly, it was inhibited by NADPH, unlike in C. reinhardtii, while NADH‐GAPDH was not affected. NADH‐GAPDH was also strongly activated by DTT in contrast to most other photosynthetic cells. In A. formosa, unlike C. reinhardtii, 1,3‐bisphosphoglycerate, the substrate of GAPDH, activated this enzyme, even in the absence of DTT, when using both NADH and NADPH as cofactors. Some of these kinetic behaviors are consistent with regulation by protein–protein interactions involving CP12, a small protein that links PRK and GAPDH in cyanobacteria, green algae, and higher plants. This conclusion was supported by immunodetection of CP12 in crude extracts of A. formosa, using antibodies raised against CP12 from C. reinhardtii. This is the first report of the existence of CP12 in a diatom, but CP12 may be a common feature of diatoms since a bioinformatic search suggested that it was also present in the Thalassiosira pseudonana Hasle et Heimdal genome v3.0. Despite the presence of CP12, this work provides further support for the differential regulation of Calvin cycle enzymes in diatoms compared to green algae.  相似文献   

2.
A4 glyceraldehyde 3-phosphate dehydrogenase (GAPDH) was purified from the green alga Chlamydomonas reinhardtii and was also overexpressed in Escherichia coli. Both purified A4 tetramers of recombinant and native GAPDH were characterized for the first time. The pH optimum for both native and recombinant enzymes was close to 7.8. The pKs of the residues involved in catalysis indicate that a cysteine and a histidine may take part in catalysis by chloroplast GAPDH, as is the case for glycolytic GAPDH. Native and recombinant GAPDH show Michaelis-Menten kinetics with respect to their cofactors, NADH and NADPH, with greater specificity for NADPH. The kinetic parameters are similar to those of the heterotetrameric A2B2 spinach chloroplast GAPDH. Native C. reinhardtii and recombinant GAPDHs exhibit a cooperative behavior towards the substrate 1,3-bisphosphoglycerate (BPGA). This positive cooperativity is specific to the C. reinhardtii enzyme, as higher plant A2B2 GAPDHs show Michaelis-Menten kinetics. Native GAPDH has twofold lower catalytic constant and K0.5 for BPGA than recombinant GAPDH. Mass spectrometry analysis of native GAPDH shows that it is a complex of GAPDH and the small protein CP12. In vitro reconstitution assays indicate that the kinetic differences are the result conformation changes of GAPDH upon association with CP12.  相似文献   

3.

Background

Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is a key enzyme of the glycolytic pathway, reversibly catalyzing the sixth step of glycolysis and concurrently reducing the coenzyme NAD+ to NADH. In photosynthetic organisms a GAPDH paralog (Gap2 in Cyanobacteria, GapA in most photosynthetic eukaryotes) functions in the Calvin cycle, performing the reverse of the glycolytic reaction and using the coenzyme NADPH preferentially. In a number of photosynthetic eukaryotes that acquired their plastid by the secondary endosymbiosis of a eukaryotic red alga (Alveolates, haptophytes, cryptomonads and stramenopiles) GapA has been apparently replaced with a paralog of the host’s own cytosolic GAPDH (GapC1). Plastid GapC1 and GapA therefore represent two independent cases of functional divergence and adaptations to the Calvin cycle entailing a shift in subcellular targeting and a shift in binding preference from NAD+ to NADPH.

Methods

We used the programs FunDi, GroupSim, and Difference Evolutionary-Trace to detect sites involved in the functional divergence of these two groups of GAPDH sequences and to identify potential cases of convergent evolution in the Calvin-cycle adapted GapA and GapC1 families. Sites identified as being functionally divergent by all or some of these programs were then investigated with respect to their possible roles in the structure and function of both glycolytic and plastid-targeted GAPDH isoforms.

Conclusions

In this work we found substantial evidence for convergent evolution in GapA/B and GapC1. In many cases sites in GAPDHs of these groups converged on identical amino acid residues in specific positions of the protein known to play a role in the function and regulation of plastid-functioning enzymes relative to their cytosolic counterparts. In addition, we demonstrate that bioinformatic software like FunDi are important tools for the generation of meaningful biological hypotheses that can then be tested with direct experimental techniques.  相似文献   

4.
A gene encoding an sn‐glycerol‐1‐phosphate dehydrogenase (G1PDH) was identified in the hyperthermophilic archaeon Pyrobaculum calidifontis. The gene was overexpressed in Escherichia coli, and its product was purified and characterized. In contrast to conventional G1PDHs, the expressed enzyme showed strong preference for NADH: the reaction rate (Vmax) with NADPH was only 2.4% of that with NADH. The crystal structure of the enzyme was determined at a resolution of 2.45 Å. The asymmetric unit consisted of one homohexamer. Refinement of the structure and HPLC analysis showed the presence of the bound cofactor NADPH in subunits D, E, and F, even though it was not added in the crystallization procedure. The phosphate group at C2’ of the adenine ribose of NADPH is tightly held through the five biased hydrogen bonds with Ser40 and Thr42. In comparison with the known G1PDH structure, the NADPH molecule was observed to be pushed away from the normal coenzyme binding site. Interestingly, the S40A/T42A double mutant enzyme acquired much higher reactivity than the wild‐type enzyme with NADPH, which suggests that the biased interactions around the C2’‐phosphate group make NADPH binding insufficient for catalysis. Our results provide a unique structural basis for coenzyme preference in NAD(P)‐dependent dehydrogenases. Proteins 2016; 84:1786–1796. © 2016 Wiley Periodicals, Inc.  相似文献   

5.
Spinach (Spinacia oleracea L.) chloroplast NAD(P)-dependent glyceraldehyde 3-phosphate dehydrogenase (NAD(P)-GAPDH; EC 1.2.1.13) was purified. The association state of the protein was monitored by fast protein liquid chromatography-Superose 12 gel filtration. Protein chromatographed in the presence of NADP+ and dithiothreitol consisted of highly NADPH-active protomers of 160 kDa; otherwise, it always consisted of a 600-kDa oligomer (regulatory form) favoured by the addition of NAD+ in buffers and with low NADPH-dependent activity (ratio of activities with NADPH versus NADH of 0.2–0.4). Glycerate 1,3-bisphosphate (BPGA) was prepared enzymatically using rabbit-muscle NAD-GAPDH, and purified. Among known modulators of spinach NAD(P)-GAPDH, BPGA is the most effective on a molar basis in stimulating NADPH-activity of dark chloroplast extracts and purified NAD(P)-GAPDH (activation constant, K a= 12 M). It also causes the enzyme to dissociate into 160-kDa protomers. The K m of BPGA both with NADPH or NADH as coenzyme is 4–7 M. NAD+ and NADH are inhibitory to the activation process induced by BPGA. This compound, together with NADP(H) and ATP belongs to a group of substrate-modifiers of the NADPH-activity and conformational state of spinach NAD(P)-GAPDH, all characterized by K a values three- to tenfold higher than the K m. Since NADP(H) is largely converted to NAD(H) in darkened chloroplasts Heineke et al. 1991, Plant Physiol. 95, 1131–1137, it is proposed that NAD+ promotes NAD(P)-GAPDH association into a regulatory conformer with low NADPH-activity during dark deactivation. The process is reversed in the light by BPGA and other substrate-modifiers whose concentration increases during photosynthesis, in addition to reduced thioredoxin.Abbreviations BPGA glycerate 1,3-bisphosphate - Chl chlorophyll - DTT dithiothreitol - FPLC fast protein liquid chromatography - NAD(P)-GAPDH glyceraldehyde 3-phosphate dehydrogenase, NAD(P)-dependent - 3-PGA glyerate 3-phosphate - PGK phosphoglycerate kinase - Prt protein - Tricine N-tris (hydroxymethyl) methyl-glycine This work was supported by grants from the Ministero dell'Università e della Ricerca Scientifica e Technologica in years 1990–1991. We are grateful to Dr. G. Branlant (Laboratoire d'Enzymologie et de Génie Génétique, Vandoeuvre les Nancy, France) for introducing us to the BPGA purification procedure.  相似文献   

6.
A cytochrome P450BM3‐catalyzed reaction system linked by a two‐step cofactor regeneration was investigated in a cell‐free system. The two‐step cofactor regeneration of redox cofactors, NADH and NADPH, was constructed by NAD+‐dependent bacterial glycerol dehydrogenase (GLD) and bacterial soluble transhydrogenase (STH) both from Escherichia coli. In the present system, the reduced cofactor (NADH) was regenerated by GLD from the oxidized cofactor (NAD+) using glycerol as a sacrificial cosubstrate. The reducing equivalents were subsequently transferred to NADP+ by STH as a cycling catalyst. The resultant regenerated NADPH was used for the substrate oxidation catalyzed by cytochrome P450BM3. The initial rate of the P450BM3‐catalyzed reaction linked by the two‐step cofactor regeneration showed a slight increase (approximately twice) when increasing the GLD units 10‐fold under initial reaction conditions. In contrast, a 10‐fold increase in STH units resulted in about a 9‐fold increase in the initial reaction rate, implying that transhydrogenation catalyzed by STH was the rate‐determining step. In the system lacking the two‐step cofactor regeneration, 34% conversion of 50 μM of a model substrate (p‐nitrophenoxydecanoic acid) was attained using 50 μM NADPH. In contrast, with the two‐step cofactor regeneration, the same amount of substrate was completely converted using 5 μM of oxidized cofactors (NAD+ and NADP+) within 1 h. Furthermore, a 10‐fold dilution of the oxidized cofactors still led to approximately 20% conversion in 1 h. These results indicate the potential of the combination of GLD and STH for use in redox cofactor recycling with catalytic quantities of NAD+ and NADP+. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2009  相似文献   

7.
The growing importance of biocatalysis in the syntheses of enantiopure molecules results from the benefits of enzymes regarding selectivity and specificity of the reaction and ecological issues of the process. Ene‐reductases (ERs) from the old yellow enzyme family have received much attention in the last years. These flavo‐enzymes catalyze the trans‐specific reduction of activated C?C bonds, which is an important reaction in asymmetric synthesis, because up to two stereogenic centers can be created in one reaction. However, limitations of ERs described in the literature such as their moderate catalytic activity and their strong preference for NADPH promote the search for novel ERs with improved properties. In this study, we characterized nine novel ERs from cyanobacterial strains belonging to different taxonomic orders and habitats. ERs were identified with activities towards a broad spectrum of alkenes. The reduction of maleimide was catalyzed with activities of up to 35.5 U mg?1 using NADPH. Ketoisophorone and (R)‐carvone, which were converted to the highly valuable compounds (R)‐levodione and (2R,5R)‐dihydrocarvone, were reduced with reaction rates of up to 2.2 U mg?1 with NADPH. In contrast to other homologous ERs from the literature, NADH was accepted at moderate to high rates as well: Enzyme activities of up to 16.7 U mg?1 were obtained for maleimide and up to 1.3 U mg?1 for ketoisophorone and (R)‐carvone. Additionally, excellent stereoselectivities were achieved in the reduction of (R)‐carvone (97–99% de). In particular, AnabaenaER3 from Anabaena variabilis ATCC 29413 and AcaryoER1 from Acaryochloris marina MBIC 11017 were identified as useful biocatalysts. Therefore, novel ERs from cyanobacteria with high catalytic efficiency were added to the toolbox for the asymmetric reduction of alkenes. Biotechnol. Bioeng. 2013; 110: 1293–1301. © 2012 Wiley Periodicals, Inc.  相似文献   

8.
9.
10.
In this study we introduce a computationally‐driven enzyme redesign workflow for altering cofactor specificity from NADPH to NADH. By compiling and comparing data from previous studies involving cofactor switching mutations, we show that their effect cannot be explained as straightforward changes in volume, hydrophobicity, charge, or BLOSUM62 scores of the residues populating the cofactor binding site. Instead, we find that the use of a detailed cofactor binding energy approximation is needed to adequately capture the relative affinity towards different cofactors. The implicit solvation models Generalized Born with molecular volume integration and Generalized Born with simple switching were integrated in the iterative protein redesign and optimization (IPRO) framework to drive the redesign of Candida boidinii xylose reductase (CbXR) to function using the non‐native cofactor NADH. We identified 10 variants, out of the 8,000 possible combinations of mutations, that improve the computationally assessed binding affinity for NADH by introducing mutations in the CbXR binding pocket. Experimental testing revealed that seven out of ten possessed significant xylose reductase activity utilizing NADH, with the best experimental design (CbXR‐GGD) being 27‐fold more active on NADH. The NADPH‐dependent activity for eight out of ten predicted designs was either completely abolished or significantly diminished by at least 90%, yielding a greater than 104‐fold change in specificity to NADH (CbXR‐REG). The remaining two variants (CbXR‐RTT and CBXR‐EQR) had dual cofactor specificity for both nicotinamide cofactors.  相似文献   

11.
Robust microorganisms are necessary for economical bioethanol production. However, such organisms must be able to effectively ferment both hexose and pentose sugars present in lignocellulosic hydrolysate to ethanol. Wild type Saccharomyces cerevisiae can rapidly ferment hexose, but cannot ferment pentose sugars. Considerable efforts were made to genetically engineer S. cerevisiae to ferment xylose. Our genetically engineered S cerevisiae yeast, 424A(LNH-ST), expresses NADPH/NADH xylose reductase (XR) that prefer NADPH and NAD+-dependent xylitol dehydrogenase (XD) from Pichia stipitis, and overexpresses endogenous xylulokinase (XK). This strain is able to ferment glucose and xylose, as well as other hexose sugars, to ethanol. However, the preference for different cofactors by XR and XD might lead to redox imbalance, xylitol excretion, and thus might reduce ethanol yield and productivity. In the present study, genes responsible for the conversion of xylose to xylulose with different cofactor specificity (1) XR from N. crassa (NADPH-dependent) and C. parapsilosis (NADH-dependent), and (2) mutant XD from P. stipitis (containing three mutations D207A/I208R/F209S) were overexpressed in wild type yeast. To increase the NADPH pool, the fungal GAPDH enzyme from Kluyveromyces lactis was overexpressed in the 424A(LNH-ST) strain. Four pentose phosphate pathway (PPP) genes, TKL1, TAL1, RKI1 and RPE1 from S. cerevisiae, were also overexpressed in 424A(LNH-ST). Overexpression of GAPDH lowered xylitol production by more than 40%. However, other strains carrying different combinations of XR and XD, as well as new strains containing the overexpressed PPP genes, did not yield any significant improvement in xylose fermentation.  相似文献   

12.
The NADPH‐cytochrome P450 oxidoreductase (CYPOR) enzyme is a membrane‐bound protein and contains both FAD and FMN cofactors. The enzyme transfers two electrons, one at a time, from NADPH to cytochrome P450 enzymes to function in the enzymatic reactions. We previously expressed in Escherichia coli the membrane‐bound CYPOR (flAnCYPOR) from Anopheles minimus mosquito. We demonstrated the ability of flAnCYPOR to support the An. minimus CYP6AA3 enzyme activity in deltamethrin degradation in vitro. The present study revealed that the flAnCYPOR purified enzyme, analyzed by a fluorometric method, readily lost its flavin cofactors. When supplemented with exogenous flavin cofactors, the activity of flAnCYPOR‐mediated cytochrome c reduction was increased. Mutant enzymes containing phenylalanine substitutions at leucine residues 86 and 219 were constructed and found to increase retention of FMN cofactor in the flAnCYPOR enzymes. Kinetic study by measuring cytochrome c–reducing activity indicated that the wild‐type and mutant flAnCYPORs followed a non‐classical two‐site Ping‐Pong mechanism, similar to rat CYPOR. The single mutant (L86F or L219F) and double mutant (L86F/L219F) flAnCYPOR enzymes, upon reconstitution with the An. minimus cytochrome P450 CYP6AA3 and a NADPH‐regenerating system, increased CYP6AA3‐mediated deltamethrin degradation compared to the wild‐type flAnCYPOR enzyme. The increased enzyme activity could illustrate a more efficient electron transfer of AnCYPOR to CYP6AA3 cytochrome P450 enzyme. Addition of extra flavin cofactors could increase CYP6AA3‐mediated activity supported by wild‐type and mutant flAnCYPOR enzymes. Thus, both leucine to phenylalanine substitutions are essential for flAnCYPOR enzyme in supporting CYP6AA3‐mediated metabolism. © 2010 Wiley Periodicals, Inc.  相似文献   

13.
Short‐chain dehydrogenase/reductase (SDR) is distributed in many organisms, from bacteria to humans, and has significant roles in metabolism of carbohydrates, lipids, amino acids, and other biomolecules. An important intermediate in acidic polysaccharide metabolism is 2‐keto‐3‐deoxy‐d ‐gluconate (KDG). Recently, two short and long loops in Sphingomonas KDG‐producing SDR enzymes (NADPH‐dependent A1‐R and NADH‐dependent A1‐R′) involved in alginate metabolism were shown to be crucial for NADPH or NADH coenzyme specificity. Two SDR family enzymes—KduD from Pectobacterium carotovorum (PcaKduD) and DhuD from Streptococcus pyogenes (SpyDhuD)—prefer NADH as coenzyme, although only PcaKduD can utilize both NADPH and NADH. Both enzymes reduce 2,5‐diketo‐3‐deoxy‐d ‐gluconate to produce KDG. Tertiary and quaternary structures of SpyDhuD and PcaKduD and its complex with NADH were determined at high resolution (approximately 1.6 Å) by X‐ray crystallography. Both PcaKduD and SpyDhuD consist of a three‐layered structure, α/β/α, with a coenzyme‐binding site in the Rossmann fold; similar to enzymes A1‐R and A1‐R′, both arrange the two short and long loops close to the coenzyme‐binding site. The primary structures of the two loops in PcaKduD and SpyDhuD were similar to those in A1‐R′ but not A1‐R. Charge neutrality and moderate space at the binding site of the nucleoside ribose 2′ coenzyme region were determined to be structurally crucial for dual‐coenzyme specificity in PcaKduD by structural comparison of the NADH‐ and NADPH‐specific SDR enzymes. The corresponding site in SpyDhuD was negatively charged and spatially shallow. This is the first reported study on structural determinants in SDR family KduD related to dual‐coenzyme specificity. Proteins 2016; 84:934–947. © 2016 Wiley Periodicals, Inc.  相似文献   

14.
Mycobacterium tuberculosis catalase‐peroxidase (KatG) is a bifunctional hemoprotein that has been shown to activate isoniazid (INH), a pro‐drug that is integral to frontline antituberculosis treatments. The activated species, presumed to be an isonicotinoyl radical, couples to NAD+/NADH forming an isoniazid‐NADH adduct that ultimately confers anti‐tubercular activity. To better understand the mechanisms of isoniazid activation as well as the origins of KatG‐derived INH‐resistance, we have compared the catalytic properties (including the ability to form the INH‐NADH adduct) of the wild‐type enzyme to 23 KatG mutants which have been associated with isoniazid resistance in clinical M. tuberculosis isolates. Neither catalase nor peroxidase activities, the two inherent enzymatic functions of KatG, were found to correlate with isoniazid resistance. Furthermore, catalase function was lost in mutants which lacked the Met‐Tyr‐Trp crosslink, the biogenic cofactor in KatG which has been previously shown to be integral to this activity. The presence or absence of the crosslink itself, however, was also found to not correlate with INH resistance. The KatG resistance‐conferring mutants were then assayed for their ability to generate the INH‐NADH adduct in the presence of peroxide (t‐BuOOH and H2O2), superoxide, and no exogenous oxidant (air‐only background control). The results demonstrate that residue location plays a critical role in determining INH‐resistance mechanisms associated with INH activation; however, different mutations at the same location can produce vastly different reactivities that are oxidant‐specific. Furthermore, the data can be interpreted to suggest the presence of a second mechanism of INH‐resistance that is not correlated with the formation of the INH‐NADH adduct.  相似文献   

15.
Pyrroline-5-carboxylic acid reductase from soybean leaves   总被引:1,自引:0,他引:1  
Pyrroline-5-carboxylic acid reductase was purified 40-fold from soybean leaves (Glycine max L. var Corsoy). The enzyme was fairly unstable, had a broad pH optimum, and was inactivated by heat and acid; NADH and NADPH both served as cofactors. It had a higher activity with NADH (about 4 ×) compared to NADPH, but a lower Km for NADPH. NADP+ inhibited both the NADH- and NADPH-dependent activity. Sulfhydryl group blocking agents reduced the activity as did the carbonyl blocking agent, NH2OH. Thiazolidine-4-carboxylic acid and phosphate inhibited the enzyme and proline inhibited only at high concentrations. ATP, GTP, and CTP were all effective inhibitors of both the NADH- and NADPH-dependent activity. Phosphorylated nucleotide inhibition was reversed by Mg2+ ions.  相似文献   

16.
Thermotoga maritima TM0298 is annotated as an alcohol dehydrogenase, yet it shows high identity and similarity to mesophilic mannitol dehydrogenases. To investigate this enzyme further, its gene was cloned and expressed in Escherichia coli. The purified recombinant enzyme was most active on fructose and mannitol, making it the first known hyperthermophilic mannitol dehydrogenase. T. maritima mannitol dehydrogenase (TmMtDH) is optimally active between 90 and 100 °C and retains 63% of its activity at 120 °C but shows no detectable activity at room temperature. Its kinetic inactivation follows a first-order mechanism, with half-lives of 57 min at 80 °C and 6 min at 95 °C. Although TmMtDH has a higher V max with NADPH than with NADH, its catalytic efficiency is 2.2 times higher with NADH than with NADPH and 33 times higher with NAD+ than with NADP+. This cofactor specificity can be explained by the high density of negatively charged residues (Glu193, Asp195, and Glu196) downstream of the NAD(P) interaction site, the glycine motif. We demonstrate that TmMtDH contains a single catalytic zinc per subunit. Finally, we provide the first proof of concept that mannitol can be produced directly from glucose in a two-step enzymatic process, using a Thermotoga neapolitana xylose isomerase mutant and TmMtDH at 60 °C.  相似文献   

17.
A simple polyether‐tethered pyrrole‐polyamide dimer 1 was synthesized in 50% yield from the reaction of 2,2,2‐trichloro‐1‐(1‐methyl‐4‐nitro‐1H‐pyrrol‐2‐yl)ethanone with 2,2′‐[1,2‐ethanediylbis(oxy)]bisethanamine, and fully characterized on the basis of 1H‐ and 13C‐NMR, MS, HR‐MS, and IR data. Agarose gel‐electrophoresis study of the cleavage of plasmid pBR322 DNA by the complexes of compound 1 with seven metal ions indicated that most of the metal complexes were capable of efficiently cleaving DNA at pH 7.0 and 37°. Among them, the CuII complex exhibited the highest activity, with the maximal catalytic rate constant kmax and Michaelis constant KM being 5.61 h?1 and 7.30 mM , respectively. Spectroscopic, ESI‐MS, ethidium‐bromide (EB) displacement, and viscosity experiments indicated that compound 1 could form a 1 : 1 complex with CuII ion, and that this complex showed moderate binding affinity toward calf‐thymus DNA.  相似文献   

18.
Intense efforts are underway to identify inhibitors of the enzyme gamma‐glutamyl transpeptidase 1 (GGT1) which cleaves extracellular gamma‐glutamyl compounds and contributes to the pathology of asthma, reperfusion injury and cancer. The glutamate analog, 6‐diazo‐5‐oxo‐norleucine (DON), inhibits GGT1. DON also inhibits many essential glutamine metabolizing enzymes rendering it too toxic for use in the clinic as a GGT1 inhibitor. We investigated the molecular mechanism of human GGT1 (hGGT1) inhibition by DON to determine possible strategies for increasing its specificity for hGGT1. DON is an irreversible inhibitor of hGGT1. The second order rate constant of inactivation was 0.052 mM ?1 min?1 and the K i was 2.7 ± 0.7 mM . The crystal structure of DON‐inactivated hGGT1 contained a molecule of DON without the diazo‐nitrogen atoms in the active site. The overall structure of the hGGT1‐DON complex resembled the structure of the apo‐enzyme; however, shifts were detected in the loop forming the oxyanion hole and elements of the main chain that form the entrance to the active site. The structure of hGGT1‐DON complex revealed two covalent bonds between the enzyme and inhibitor which were part of a six membered ring. The ring included the OG atom of Thr381, the reactive nucleophile of hGGT1 and the α‐amine of Thr381. The structure of DON‐bound hGGT1 has led to the discovery of a new mechanism of inactivation by DON that differs from its inactivation of other glutamine metabolizing enzymes, and insight into the activation of the catalytic nucleophile that initiates the hGGT1 reaction.  相似文献   

19.
During growth of Saccharomyces cerevisiae on glucose, the redox cofactors NADH and NADPH are predominantly involved in catabolism and biosynthesis, respectively. A deviation from the optimal level of these cofactors often results in major changes in the substrate uptake and biomass formation. However, the metabolism of xylose by recombinant S. cerevisiae carrying xylose reductase and xylitol dehydrogenase from the fungal pathway requires both NADH and NADPH and creates cofactor imbalance during growth on xylose. As one possible solution to overcoming this imbalance, the effect of overexpressing the native NADH kinase (encoded by the POS5 gene) in xylose-consuming recombinant S. cerevisiae directed either into the cytosol or to the mitochondria was evaluated. The physiology of the NADH kinase containing strains was also evaluated during growth on glucose. Overexpressing NADH kinase in the cytosol redirected carbon flow from CO2 to ethanol during aerobic growth on glucose and to ethanol and acetate during anaerobic growth on glucose. However, cytosolic NADH kinase has an opposite effect during anaerobic metabolism of xylose consumption by channeling carbon flow from ethanol to xylitol. In contrast, overexpressing NADH kinase in the mitochondria did not affect the physiology to a large extent. Overall, although NADH kinase did not increase the rate of xylose consumption, we believe that it can provide an important source of NADPH in yeast, which can be useful for metabolic engineering strategies where the redox fluxes are manipulated.  相似文献   

20.
The estrogen synthetase present in human placental microsomes appears to be dependent on the cooperative interaction of the reduced cofactors NADPH and NADH for optimal activity. Using steady-state concentrations of either cofactor, it was found that while the estrogen synthetase activity followed hyperbolic saturation kinetics with NADPH (Kmapp = 14 μM), the enzyme followed sigmoidal saturation kinetics when the cofactor was NADH, with the half-maximum velocity attained at a cofactor concentration of 1.1 mm. The maximum velocity obtained with NADPH as the cofactor was greater than with corresponding concentrations of NADH. Estrogen synthetase activity in the presence of NADH was not due to NADPH contamination. NADH, in the presence of small concentrations of NADPH (0.5 to 5 μm), stimulated significantly the rate of estrogen formation from androstenedione by placental microsomes and, in addition, the enzyme saturation kinetics changed from sigmoidal to hyperbolic, thus mimicking the effect of NADPH. Estrogen synthetase activity, measured in the presence of 1 mm NADH, was stimulated in a dose-dependent manner by NADPH (Kmapp = 0.4 μM NADPH) and, when the enzyme was measured in the presence of 5 μm NADPH, the activity was stimulated in a dose-dependent manner by NADH (Kmapp = 45 μM NADH). Estrogen synthetase activity measured in the presence of NADH, without and with NADPH (1 μm) remained linear both with time of incubation for approximately 15 min and with microsomal protein concentration up to 3 mg/ml. The apparent Km of estrogen synthetase for androstenedione, when measured in the presence of NADH, was 1 μm. The synergistic interaction between NADH and NADPH in stimulating placental estrogen synthetase activity observed in vitro may, conceivably, take place in vivo in the intact placenta.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号