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 656 毫秒
1.
天津短杆菌T6-13谷氨酸脱氢酶的研究   总被引:1,自引:0,他引:1  
本文研究了天津短杆菌(Brevibacterium tianjinese)T6-13谷氨酸脱氢酶(GDH)[EC1.4.1.4]的纯化和性质。该酶以辅酶11(NADP)为其专一性辅酶,正、逆反应酶活力最适pH分别为7.5和8.9—9 9,对热较敏感。 该酶对还原型辅酶11(NADPH)、α-酮戊二酸(α-KG)、NH,、NADP和L-谷氨酸(GA)的Km值分别为0.076、3.23、4.0、0.02和120.48 mmol/L。该酶受反应产物的抑制,逆反应受NADPH、α-KG和NH+4的抑制,正反应受NADP和谷氨酸的抑制,但该酶所催化的逆反应既不受三羧酸循环代谢中间产物的抑制,也不受氨基酸的抑制和氨基酸的积累抑制。对发酵过程中谷氨酸脱氢酶活力变化的研究表明,前期酶活力逐渐上升,当发酵至16小时左右酶活力最高,其后酶活力逐渐下降;二级种子的酶活力与发酵过程中酶活力最高时相当。  相似文献   

2.
【背景】高效实现D-氨基酸的生物合成一直是人们关注的热点。内消旋-二氨基庚二酸脱氢酶(meso-diaminopimelate dehydrogenase,DAPDH)能够直接催化2-酮酸和氨合成D-氨基酸。【目的】提高DAPDH对烷基取代2-酮酸的催化活力,并解释其催化机制。【方法】以来源于嗜热共生杆菌(Symbiobacteriumthermophilum)的内消旋-二氨基庚二酸脱氢酶(StDAPDH)为模板,在前期结构分析结合被选择位点突变结果的基础上,确定对H227位进行定点饱和突变,并以D-丙氨酸、D-2-氨基丁酸、D-正缬氨酸、D-谷氨酸为底物进行筛选。【结果】获得突变体H227Q和H227N。突变体H227Q对丙酮酸、2-氧代丁酸、2-氧代戊酸、2-酮戊二酸的比活力比野生型分别提高了10.9、11.5、8.6和7.6倍。动力学参数表明,突变体H227Q同时提高了酶对底物的亲和力及催化常数,使其对丙酮酸的催化效率(k_(cat)/K_m)相较于野生型提高了9.4倍。利用分子模拟技术分析突变体H227Q与产物氨基酸之间的相互作用表明,227位的谷氨酰胺通过与氨基酸的羧酸形成氢键,使得氨基酸产物Cα上的氢和辅酶烟酰胺环C4原子之间的距离缩短。【结论】利用定向进化技术提高DAPDH对烷基取代2-酮酸的催化活力,有助于开发新型的高效生物催化剂,这些工作也为下一步继续进行更具挑战性的D-氨基酸研究提供了基础。  相似文献   

3.
丙氨酸脱氢酶(alanine dehydrogenase,ALD,EC 1.4.1.1)是一种以烟酰胺腺嘌呤(NAD)为辅酶的氨基酸脱氢酶.丙氨酸脱氢酶可逆催化丙氨酸氧化脱氨生成丙酮酸、氨及NADH.丙氨酸脱氢酶也是调节氨基酸代谢和糖代谢的重要酶类,其催化反应的产物丙酮酸广泛应用于医药、农药和食品等领域,具有良好的发展前景.主要介绍丙氨酸脱氢酶的纯化及活力检测、酶空间结构(底物结合住点),以及催化反应机理等方面的研究.  相似文献   

4.
自然界中依赖烟酰胺类辅酶(NAD+或NADP+)的脱氢酶是氧化还原酶中最重要的一类,基于此类酶的生物传感器应用前景广阔,近年来发展迅速。构建这类传感器需要两项关键技术,即氧化型辅酶在电极表面的再生和辅酶固定化。本文介绍了辅酶电化学再生的主要方法、辅酶固定化的常见手段,以及相关的研究进展。  相似文献   

5.
目的:将来源于简单节杆菌的3-甾酮-△~1-脱氢酶(3-ketosteroid-Delta(1)-dehydrogenase,KSDD)在大肠杆菌中进行表达,获得具有活性的脱氢酶;利用计算机预测KSDD的三级结构,并通过定点突变确定酶的关键位点以期优化脱氢酶的活性及性质。方法:克隆简单节杆菌编码KSDD的基因ksdd构建原核表达载体,以Escherichia coli BL21(DE3)为表达宿主构建重组菌并诱导表达,HPLC法检测重组酶催化4-AD脱氢的转化率;通过SWISS-MODEL同源建模分析KSDD结构,对预测的催化关键位点氨基酸残基进行定点突变并研究突变后重组酶的活性变化。结果:成功构建了表达脱氢酶KSDD的重组菌E.coli pET-22-ksdd,21℃下诱导表达后,重组酶对4-AD的转化率为27%;通过SWISS-MODEL同源建模预测出脱氢酶结构并对4个关键位点进行定点突变设计,获得突变子Y120R、Y320L、Y488F和G492Y。突变子Y120R和Y488F失活,证明其为酶的活性位点;突变子Y320L的转化率与野生型基本一致,但37℃反应条件下稳定性有所提高;突变子G492Y对4-AD的转化率是野生型的1.2倍,37℃条件下稳定性有所提高,是突变后氨基酸位点疏水性增加和周围静电作用改变所导致。结论:目前对简单节杆菌3-甾酮-△~1-脱氢酶结构分析及催化机理相关的研究较少,本研究验证了KSDD的活性位点,优化了酶的稳定性,为进一步对酶的性质进行定向改造打下了基础。  相似文献   

6.
过量表达苹果酸脱氢酶对大肠杆菌NZN111产丁二酸的影响   总被引:2,自引:1,他引:1  
大肠杆菌NZN111是敲除了乳酸脱氢酶的编码基因 (ldhA) 和丙酮酸-甲酸裂解酶的编码基因 (pflB) 的工程菌,厌氧条件下由于辅酶NAD(H) 的不平衡导致其丧失了代谢葡萄糖的能力。构建了苹果酸脱氢酶的重组菌大肠杆菌NZN111/pTrc99a-mdh,在厌氧摇瓶发酵过程中通过0.3 mmol/L的IPTG诱导后重组菌的苹果酸脱氢酶 (Malate dehydrogenase,MDH) 酶活较出发菌株提高了14.8倍,NADH/NAD+的比例从0.64下降到0.26,同时NAD+和NADH浓度分别  相似文献   

7.
6-磷酸葡萄糖脱氢酶催化6-磷酸葡萄糖生成6-磷酸葡萄糖酸,并生成NADPH,是微生物胞内磷酸戊糖途径(PPP)的关键酶。本研究以食品安全菌米曲霉CICC2012为材料,克隆获得6-磷酸葡萄糖脱氢酶基因(GenBank登录号:JN123468)。序列分析表明,该酶是由222个氨基酸组成的亲水性蛋白;128~134位氨基酸序列DHYLGKE为活性区域;170~176位氨基酸序列GTEGRGG可能为辅因子结合位点。进化树分析表明,米曲霉6-磷酸葡萄糖脱氢酶同其他丝状真菌及酵母的G6PDH较相似。  相似文献   

8.
【目的】构建己糖激酶与葡萄糖-6-磷酸脱氢酶的大肠杆菌共表达体系,以葡萄糖为底物实现辅酶NADPH的高效再生。【方法】通过分子生物学方法,克隆己糖激酶HKgs、HKpp基因,并于Escherichia coli BL21(DE3)中表达,再将己糖激酶HKgs、HKpp分别与葡萄糖-6-磷酸脱氢酶Gpd PP共表达,实现NADPH的原位再生。比较两个共表达工程菌的辅酶再生效果,并针对催化活力较高的工程菌BL21(HKgs+Gpd PP)进行表达条件优化。【结果】NADPH再生活力达到856 U/L。该辅酶再生体系与醇脱氢酶Adh R联合催化,使不对称还原4-氯乙酰乙酸乙酯的催化活力提高至原始值的2.5倍。【结论】通过己糖激酶与葡萄糖-6-磷酸脱氢酶在大肠杆菌中的共表达,构建了一个新的NADPH高效再生体系,并用于醇脱氢酶催化的不对称还原反应。  相似文献   

9.
孔雀绿定磷法测定植物NADP磷酸酶活性   总被引:5,自引:0,他引:5  
AMethodofPhosphateDeterminationUsingMalachiteGreenFitfortheMeasurementofNADPPhosphataseActivityYANGWan-Nian,HEZhi-Chang(CollegeofLifeSciences,WuhanUniversity,Wuhan430072)NADP磷酸酶催化NADP水解生成NAD和磷酸:NADP+H2O→NAD+Pi。该酶与NAD激酶一起参与NAD和NADP水平的调节。其活性通过乙醇脱氢酶循环反应生成的NAD确定[1]。该方法虽然比较灵敏,但操作比较繁琐,反应条件不易控制。孔雀绿定磷法是一种灵敏度很高的测定无机磷的方法[2,3]已用于ATP酶[2,3,4]活性及钙调素含量[3]的…  相似文献   

10.
[目的] 从芽胞杆菌(Bacillus sp.)YX-1基因组中克隆出一种有机溶剂耐受型的葡萄糖脱氢酶基因,实现了该基因在大肠杆菌中的高效表达,研究了重组蛋白的酶学性质.[方法] 依据芽胞杆菌属中葡萄糖脱氢酶氨基酸序列的保守性,设计合理引物,钓取来源于Bacillus sp.YX-1的葡萄糖脱氢酶基因,构建诱导型表达载体pET28a-gdh,于大肠杆菌中进行表达.镍柱亲和层析法纯化重组蛋白,考察了重组蛋白的酶学性质.[结果] 葡萄糖脱氢酶基因全长为786 bp,编码261个氨基酸.酶学研究结果表明:该酶最适反应温度为45℃,最适pH值为8.0;具有良好的有机溶剂耐受性,于50%的辛烷、环己烷、癸烷中室温放置1h后,酶活仍能保持90%以上;具有较宽的底物谱,对多种糖均具有一定的催化活性,其中催化D-葡萄糖的活力最高,产生还原型辅酶因子;对辅酶NADH和NADPH具有相似的依赖性,对NAD+和NADP+的催化比活分别为8.37 U/mg和8.62 U/mg.[结论]利用生物信息学成功地挖掘出Bacillus sp.YX-1一种耐有机溶剂的葡萄糖脱氢酶,为氧化还原酶在有机相反应中的的辅酶再生循环提供了新型的生物催化剂.  相似文献   

11.
NAD+-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short chain dehydrogenase/reductase (SDR) family, is responsible for the biological inactivation of prostaglandins. Sequence alignment within SDR coupled with molecular modeling analysis has suggested that Gln-15, Asp-36, and Trp-37 of 15-PGDH may determine the coenzyme specificity of this enzyme. Site-directed mutagenesis was used to examine the important roles of these residues. Several single mutants (Q15K, Q15R, W37K, and W37R), double mutants (Q15K-W37K, Q15K-W37R, Q15R-W37K, and Q15R-W37R), and triple mutants (Q15K-D36A-W37R and Q15K-D36S-W37R) were prepared and expressed as glutathione S-transferase (GST) fusion proteins in Escherichia coli and purified by GSH-agarose affinity chromatography. Mutants Q15K, Q15R, W37K, W37R, Q15K-W37K, and Q15R-W37K were found to be inactive or almost inactive with NADP+ but still retained substantial activity with NAD+. Mutant Q15K-W37R and mutant Q15R-W37R showed comparable activity for NAD+ and NADP+ with an increase in activity nearly 3-fold over that of the wild type. However, approximately 30-fold higher in K(m) for NADP+ than that of the wild type enzyme for NAD+ was found for mutants Q15K-W37R and Q15R-W37R. Similarly, the K(m) values for PGE(2) of mutants were also shown to increase over that of the wild type. Further mutation of Asp-36 to either an alanine or a serine of the double mutant Q15K-W37R (i.e., triple mutants Q15K-D36A-W37R and Q15K-D36S-W37R) rendered the mutants exhibiting exclusive activity with NADP+ but not with NAD+. The triple mutants showed a decrease in K(m) for NADP+ but an increase in K(m) for PGE(2). Further mutation at Ala-14 to a serine of a triple mutant (Q15K-D36S-W37R) decreased the K(m) values for both NADP+ and PGE(2) to levels comparable to those of the wild type. These results indicate that the coenzyme specificity of 15-PGDH can be altered from NAD+ to NADP+ by changing a few critical residues near the N-terminal end.  相似文献   

12.
The R‐specific alcohol dehydrogenase from Lactobacillus brevis (Lb‐ADH) catalyzes the enantioselective reduction of prochiral ketones to the corresponding secondary alcohols. It is stable and has broad substrate specificity. These features make this enzyme an attractive candidate for biotechnological applications. A drawback is its preference for NADP(H) as a cofactor, which is more expensive and labile than NAD(H). Structure‐based computational protein engineering was used to predict mutations to alter the cofactor specificity of Lb‐ADH. Mutations were introduced into Lb‐ADH and tested against the substrate acetophenone, with either NAD(H) or NADP(H) as cofactor. The mutant Arg38Pro showed fourfold increased activity with acetophenone and NAD(H) relative to the wild type. Both Arg38Pro and wild type exhibit a pH optimum of 5.5 with NAD(H) as cofactor, significantly more acidic than with NADP(H). These and related Lb‐ADH mutants may prove useful for the green synthesis of pharmaceutical precursors.  相似文献   

13.
Chloroplast ferredoxin-NADP(+) reductase has a 32,000-fold preference for NADPH over NADH, consistent with its main physiological role of NADP(+) photoreduction for de novo carbohydrate biosynthesis. Although it is distant from the 2'-phosphoryl group of NADP(+), replacement of the C-terminal tyrosine (Tyr(308) in the pea enzyme) by Trp, Phe, Gly, and Ser produced enzyme forms in which the preference for NADPH over NADH was decreased about 2-, 10-, 300-, and 400-fold, respectively. Remarkably, in the case of the Y308S mutant, the k(cat) value for the NADH-dependent activity approached that of the NADPH-dependent activity of the wild-type enzyme. Furthermore, difference spectra of the NAD(+) complexes revealed that the nicotinamide ring of NAD(+) binds at nearly full occupancy in the active site of both the Y308G and Y308S mutants. These results correlate well with the k(cat) values obtained with these mutants in the NADH-ferricyanide reaction. The data presented support the hypothesis that specific recognition of the 2'-phosphate group of NADP(H) is required but not sufficient to ensure a high degree of discrimination against NAD(H) in ferredoxin-NADP(+) reductase. Thus, the C-terminal tyrosine enhances the specificity of the reductase for NADP(H) by destabilizing the interaction of a moiety common to both coenzymes, i.e. the nicotinamide.  相似文献   

14.
Previously we found that replacement of seven amino acid residues in a loop region markedly shifted the coenzyme specificity of malate dehydrogenase from NAD(H) toward NADP(H). In the present study, we replaced the seven amino acid residues in the corresponding region of an NAD(H)-dependent lactate dehydrogenase with those of NADP(H)-dependent malate dehydrogenase, and examined the coenzyme specificity of the resulting mutant enzyme. Coenzyme specificity was significantly shifted by 399-fold toward NADPH when k cat?K m coenzyme was used as the measure of coenzyme specificity. The effect of the replacements on coenzyme specificity is discussed based on in silico simulation of the three-dimensional structure of the lactate dehydrogenase mutant.  相似文献   

15.
Phenylacetaldehyde dehydrogenase (PAD) and lactaldehyde dehydrogenase (ALD) share some structural and kinetic properties. One difference is that PAD can use NAD+ and NADP+, whereas ALD only uses NAD+. An acidic residue has been involved in the exclusion of NADP+ from the active site in pyridine nucleotide-dependent dehydrogenases. However, other factors may participate in NADP+ exclusion. In the present work, analysis of the sequence of the region involved in coenzyme binding showed that residue F180 of ALD might participate in coenzyme specificity. Interestingly, F180T mutation rendered an enzyme (ALD-F180T) with the ability to use NADP+. This enzyme showed an activity of 0.87 micromol/(min * mg) and K(m) for NADP+ of 78 microM. Furthermore, ALD-F180T exhibited a 16-fold increase in the V(m) /K(m) ratio with NAD+ as the coenzyme, from 12.8 to 211. This increase in catalytic efficiency was due to a diminution in K(m) for NAD+ from 47 to 7 microM and a higher V(m) from 0.51 to 1.48 micromol/(min * mg). In addition, an increased K(d) for NADH from 175 (wild-type) to 460 microM (mutant) indicates a faster product release and possibly a change in the rate-limiting step. For wild-type ALD it is described that the rate-limiting step is shared between deacylation and coenzyme dissociation. In contrast, in the present report the rate-limiting step in ALD-F180T was determined to be exclusively deacylation. In conclusion, residue F180 participates in the exclusion of NADP+ from the coenzyme binding site and disturbs the binding of NAD+.  相似文献   

16.
NAD(P)H regeneration is important for biocatalytic reactions that require these costly cofactors. A mutant phosphite dehydrogenase (PTDH-E175A/A176R) that utilizes both NAD and NADP efficiently is a very promising system for NAD(P)H regeneration. In this work, both the kinetic mechanism and practical application of PTDH-E175A/A176R were investigated for better understanding of the enzyme and to provide a basis for future optimization. Kinetic isotope effect studies with PTDH-E175A/A176R showed that the hydride transfer step is (partially) rate determining with both NAD and NADP giving (D)V values of 2.2 and 1.7, respectively, and (D)V/K(m,phosphite) values of 1.9 and 1.7, respectively. To better comprehend the relaxed cofactor specificity, the cofactor dissociation constants were determined utilizing tryptophan intrinsic fluorescence quenching. The dissociation constants of NAD and NADP with PTDH-E175A/A176R were 53 and 1.9 microm, respectively, while those of the products NADH and NADPH were 17.4 and 1.22 microm, respectively. Using sulfite as a substrate mimic, the binding order was established, with the cofactor binding first and sulfite binding second. The low dissociation constant for the cofactor product NADPH combined with the reduced values for (D)V and k(cat) implies that product release may become partially rate determining. However, product inhibition does not prevent efficient in situ NADPH regeneration by PTDH-E175A/A176R in a model system in which xylose was converted into xylitol by NADP-dependent xylose reductase. The in situ regeneration proceeded at a rate approximately fourfold faster with PTDH-E175A/A176R than with either WT PTDH or a NADP-specific Pseudomonas sp.101 formate dehydrogenase mutant with a total turnover number for NADPH of 2500.  相似文献   

17.
Gastric tissues from amphibian Rana perezi express the only vertebrate alcohol dehydrogenase (ADH8) that is specific for NADP(H) instead of NAD(H). In the crystallographic ADH8-NADP+ complex, a binding pocket for the extra phosphate group of coenzyme is formed by ADH8-specific residues Gly223-Thr224-His225, and the highly conserved Leu200 and Lys228. To investigate the minimal structural determinants for coenzyme specificity, several ADH8 mutants involving residues 223 to 225 were engineered and kinetically characterized. Computer-assisted modeling of the docked coenzymes was also performed with the mutant enzymes and compared with the wild-type crystallographic binary complex. The G223D mutant, having a negative charge in the phosphate-binding site, still preferred NADP(H) over NAD(H), as did the T224I and H225N mutants. Catalytic efficiency with NADP(H) dropped dramatically in the double mutants, G223D/T224I and T224I/H225N, and in the triple mutant, G223D/T224I/H225N (kcat/KmNADPH = 760 mm-1 min-1), as compared with the wild-type enzyme (kcat/KmNADPH = 133330 mm-1 min-1). This was associated with a lower binding affinity for NADP+ and a change in the rate-limiting step. Conversely, in the triple mutant, catalytic efficiency with NAD(H) increased, reaching values (kcat/KmNADH = 155000 mm-1 min-1) similar to those of the wild-type enzyme with NADP(H). The complete reversal of ADH8 coenzyme specificity was therefore attained by the substitution of only three consecutive residues in the phosphate-binding site, an unprecedented achievement within the ADH family.  相似文献   

18.
Ferredoxin-NADP(+) reductase catalyses NADP(+) reduction, being specific for NADP(+)/H. To understand coenzyme specificity determinants and coenzyme specificity reversion, mutations at the NADP(+)/H pyrophosphate binding and of the C-terminal regions have been simultaneously introduced in Anabaena FNR. The T155G/A160T/L263P/Y303S mutant was produced. The mutated enzyme presents similar k(cat) values for NADPH and NADH, around 2.5 times slower than that reported for WT FNR with NADPH. Its K(m) value for NADH decreased 20-fold with regard to WT FNR, whereas the K(m) for NADPH remains similar. The combined effect is a much higher catalytic efficiency for NAD(+)/H, with a minor decrease of that for NADP(+)/H. In the mutated enzyme, the specificity for NADPH versus NADH has been decreased from 67,500 times to only 12 times, being unable to discriminate between both coenzymes. Additionally, giving the role stated for the C-terminal Tyr in FNR, its role in the energetics of the FAD binding has been analysed.  相似文献   

19.
Chloroplast glyceraldehyde-3-phosphate dehydrogenase (GAPDH) of higher plants uses both NADP(H) and NAD(H) as coenzyme and consists of one (GapA) or two types of subunits (GapA, GapB). AB-GAPDH is regulated in vivo through the action of thioredoxin and metabolites, showing higher kinetic preference for NADPH in the light than in darkness due to a specific effect on kcat(NADPH). Previous crystallographic studies on spinach chloroplast A4-GAPDH complexed with NADP or NAD showed that residues Thr33 and Ser188 are involved in NADP over NAD selectivity by interacting with the 2'-phosphate group of NADP. This suggested a possible involvement of these residues in the regulatory mechanism. Mutants of recombinant spinach GapA (A4-GAPDH) with Thr33 or Ser188 replaced by Ala (T33A, S188A and double mutant T33A/S188A) were produced, expressed in Escherichia coli, and compared to wild-type recombinant A4-GAPDH, in terms of crystal structures and kinetic properties. Affinity for NADPH was decreased significantly in all mutants, and kcat(NADPH) was lowered in mutants carrying the substitution of Ser188. NADH-dependent activity was unaffected. The decrease of kcat/Km of the NADPH-dependent reaction in Ser188 mutants resembles the behaviour of AB-GAPDH inhibited by oxidized thioredoxin, as confirmed by steady-state kinetic analysis of native enzyme. A significant expansion of size of the A4-tetramer was observed in the S188A mutant compared to wild-type A4. We conclude that in the absence of interactions between Ser188 and the 2'-phosphate group of NADP, the enzyme structure relaxes to a less compact conformation, which negatively affects the complex catalytic cycle of GADPH. A model based on this concept might be developed to explain the in vivo light-regulation of the GAPDH.  相似文献   

20.
Pichia stipitis NAD(+)-dependent xylitol dehydrogenase (XDH), a medium-chain dehydrogenase/reductase, is one of the key enzymes in ethanol fermentation from xylose. For the construction of an efficient biomass-ethanol conversion system, we focused on the two areas of XDH, 1) change of coenzyme specificity from NAD(+) to NADP(+) and 2) thermostabilization by introducing an additional zinc atom. Site-directed mutagenesis was used to examine the roles of Asp(207), Ile(208), Phe(209), and Asn(211) in the discrimination between NAD(+) and NADP(+). Single mutants (D207A, I208R, F209S, and N211R) improved 5 approximately 48-fold in catalytic efficiency (k(cat)/K(m)) with NADP(+) compared with the wild type but retained substantial activity with NAD(+). The double mutants (D207A/I208R and D207A/F209S) improved by 3 orders of magnitude in k(cat)/K(m) with NADP(+), but they still preferred NAD(+) to NADP(+). The triple mutant (D207A/I208R/F209S) and quadruple mutant (D207A/I208R/F209S/N211R) showed more than 4500-fold higher values in k(cat)/K(m) with NADP(+) than the wild-type enzyme, reaching values comparable with k(cat)/K(m) with NAD(+) of the wild-type enzyme. Because most NADP(+)-dependent XDH mutants constructed in this study decreased the thermostability compared with the wild-type enzyme, we attempted to improve the thermostability of XDH mutants by the introduction of an additional zinc atom. The introduction of three cysteine residues in wild-type XDH gave an additional zinc-binding site and improved the thermostability. The introduction of this mutation in D207A/I208R/F209S and D207A/I208R/F209S/N211R mutants increased the thermostability and further increased the catalytic activity with NADP(+).  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