首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到19条相似文献,搜索用时 187 毫秒
1.
研究了迷迭香酸对大肠杆菌L-天冬酰胺酶促反应动力学的影响, 结果表明, 迷迭香酸能降低天冬酰胺酶的表观米氏常数, 是该酶的激活剂; 亚铁离子能显著促进迷迭香酸的抗菌活性, 吸收光谱分析显示, 迷迭香酸与亚铁离子存在直接的相互作用, 二者的结合比为1:1。  相似文献   

2.
L-天冬酰胺酶用于治疗急性淋巴细胞白血病与淋巴瘤,但高昂的价格限制了临床上的广泛应用,重组表达解决了成本的难题。L-天冬酰胺酶所具有的低谷氨酰胺酶活性是其副作用的根源,同时细菌来源的L-天冬酰胺酶能导致危及生命的过敏反应,寻找副作用更低的优质L-天冬酰胺酶及其修饰方法,成为研究的重点。对L-天冬酰胺酶的重组表达,及具有更优性质的L-天冬酰胺酶开发进展进行了综述,并阐明其在临床治疗及食品工业领域的巨大应用潜力。  相似文献   

3.
【目的】从大肠杆菌Nissle1917中获得L-天冬酰胺酶Ⅱ基因,并研究其抗肿瘤活性。【方法】以大肠杆菌Nissle1917基因组为模板PCR扩增L-天冬酰胺酶Ⅱ基因,克隆至可诱导表达载体pET28a上。将L-天冬酰胺酶Ⅱ表达载体pET28a-asp转化至大肠杆菌BL21(DE3)中并通过IPTG诱导表达,经聚丙烯酰胺凝胶电泳(SDS-PAGE)和液相色谱-质谱(LC-MS)对表达的L-天冬酰胺酶Ⅱ进行鉴定,并通过镍柱亲和层析纯化收集表达出的L-天冬酰胺酶Ⅱ。用纯化定量以后的L-天冬酰胺酶Ⅱ作用小鼠乳腺癌4T1细胞、人肝癌Hep-3B细胞和人脐静脉内皮细胞HUVEC。【结果】来自于大肠杆菌Nissle1917的L-天冬酰胺酶Ⅱ基因可在大肠杆菌BL21中高效表达并通过LC-MS得到鉴定,细胞毒性实验结果表明L-天冬酰胺酶Ⅱ对4T1细胞和Hep-3B细胞的生长具有较强的抑制作用,而对人脐静脉内皮细胞HUVEC的生长无明显抑制效果。【结论】来源于大肠杆菌Nissle1917的L-天冬酰胺酶Ⅱ能显著抑制4T1细胞和Hep-3肿瘤细胞的生长,而对人正常组织细胞的生长无明显抑制效果,为进一步研究L-天冬酰胺酶Ⅱ特异性抗肿瘤作用机制和对实体瘤的应用研究奠定了重要基础。  相似文献   

4.
探索生物转化法制备L-天冬酰胺的技术与工艺。通过分子生物学方法,克隆来源于大肠杆菌(Escherichia coli, E.coli)JM109的天冬酰胺合成酶A基因asnA,并于E. coli BL21(DE3)中表达,利用构建的E.coli基因工程菌E.coli BL21(DE3)/pET28a(+)-asnA全细胞高密度催化L-天冬氨酸生产L-天冬酰胺,以PITC柱前衍生-高效液相检测底物和产物。表达的蛋白质分子质量约为37kDa,与预期大小相符,比酶活力为1786.6U/g。L-天冬氨酸转化率为95.8%,L-天冬酰胺产量可达126.5g/L,生产速率为15.81g/(L·h)。结果表明,已成功构建高效表达天冬酰胺合成酶A基因工程菌株,并用于催化L-天冬氨酸转化生产L-天冬酰胺,解决了L-天冬酰胺生物转化生产工艺中ATP成本过高的难题,为L-天冬酰胺制备提供新的绿色途径。  相似文献   

5.
本文报导了天冬酰胺酶及PEG_2-天冬酰胺酶对废物L-天冬酰胺、谷氨酰胺亲和性的研究,结果表明:PEG_2-天冬酰胺酶对谷氨酰胺的亲和性明显强于天冬酰胺酶(Km值分别为7.35×10~(-3)mol/L和7.14×10~(-2)mol/L),对天冬酰胺的亲和性略强于天冬酰胺酶(Km值分别为2.9×10~(-5)mol/L和4.0×10~(-5)mol/L)。天冬酰胺酶和PEG_2-天冬酰胺酶的CD光谱表明:天冬酰胺和谷氨酰胺对天冬酰胺酶和PEG_2-天冬酰胺酶的构象影响较大,但天冬酰胺酶和PEG_2-天冬酰胺酶的构象变化趋势有明显的不同。  相似文献   

6.
L-天冬酰胺酶(L-asparaginase, L-ASN)广泛用于恶性肿瘤治疗及低丙烯酰胺食品生产,然而其较低的表达水平限制了应用推广。异源蛋白表达是提高目标酶表达水平的有效策略,芽胞杆菌广泛用于酶蛋白的高效生产,本研究拟通过表达元件及宿主优化提高芽胞杆菌(Bacillus)中L-天冬酰胺酶产量。首先,筛选了5种信号肽(SPSacC、SPAmyL、SPAprE、SPYwbN、SPWapA)用于L-天冬酰胺酶的分泌表达,其中SPSacC介导下L-天冬酰胺酶分泌效果最好,酶活达到157.61 U/mL。随后,选取了4种芽胞杆菌强启动子(P43、PykzA-P43、PUbay、Pbac A),其中串联启动子PykzA-P43介导的L-天冬酰胺酶表达量最高,较对照菌株提高了52.94%。最后,筛选了3种芽胞杆菌表达宿主:地衣芽胞杆菌(Bacillus licheniformi...  相似文献   

7.
蓖麻蚕丝腺L-天冬酰胺酶的纯化及其性质   总被引:1,自引:0,他引:1  
本文总结从蓖麻蚕5龄幼虫的丝腺经过匀浆、离心、氯化锰沉淀、硫酸铵盐析、加热处理、离子交换柱层析、吸附层析、凝胶过滤等步骤纯化L-天冬酰胺酶的过程,并对其生化性质进行了研究。纯化后L-天冬酰胺酶的比活力为原始上清液的246倍。该酶的分子量为112,000,由4个亚基组成;在溶液中以及冻于过程中,酶分子会相互聚合成多聚体。该酶在pH8.5,0.1M硼酸缓冲液中的Km值为1.1×10-4M,最适pH为8.5,在pH7.5-10.0的范围内,都具有较高的活力。PO4、SO4=、Cl-等离子对该酶有激活作用。蓖麻蚕丝腺L-天冬酰胺酶很不稳定,特别在偏碱性溶液中极易失去活力;L-天冬氨酸、二硫苏糖醇、乙二胺四乙酸对其有保护作用,其中以L-天冬氨酸的作用最为明显。另外,还分析了该酶的氨基酸组成等,并与不同来源的L-天冬酰胺酶进行了比较研究。  相似文献   

8.
氧载体对L—天冬酰胺酶发酵过程影响的研究   总被引:5,自引:0,他引:5  
以抗癌药物L天冬酰胺酶生产为应用背景,针对发酵过程中存在严重耗氧问题,研究了氧载体对发酵过程的影响。通过对几种氧载体的筛选,认为正十二烷最适合于该发酵过程。随后以产物L天冬酰胺酶活性、菌体浓度以及溶氧水平为主要指标,考察了氧载体在发酵过程中的作用,实验表明,发酵基质中5%正十二烷的添加量为最佳浓度,这种氧载体的加入,明显地提高了发酵介质中的溶氧水平,改善了供氧条件,增加了菌体浓度,提高了L天冬酰胺酶发酵水平,在优化条件下,可使发酵液最终酶活提高21%左右  相似文献   

9.
氧载体对L-天冬酰胺酶发酵过程影响的研究   总被引:3,自引:0,他引:3  
以抗癌药物L-天冬酰胺酶生产为应用背景,针对发酵过程中存在严重耗氧问题,研究了氧载体对发酵过程的影响。通过对几种氧载体的筛选,认为正十二烷最适合于该发酵过程。随后以产物L-天冬酰胺酶活性、菌体浓度以及溶氧水平为主要指标,考察了氧载体在发酵过程中的作用.实验表明,发酵基质中5%正十二烷的添加量为最佳浓度,这种氧载体的加入,明显地提高了发酵介质中的溶氧水平,改善了供氧条件,增加了菌体浓度,提高了L-天冬酰胺酶发酵水平,在优化条件下,可使发酵液最终酶活提高21%左右。  相似文献   

10.
单个L-天冬酰胺酶分子由四个相同的亚基组成,亚基分子量37,000道尔顿,电镜下可见四个球形亚基形成一个中间有小孔的平面,它们有一致的一级结构。用固定化亚基技术证实:亚基单独存在不表现催化能力,但固定化亚基具有“检起”可溶性初生亚基的能力,重组成寡聚的固定化酶之后酶活力恢复。  相似文献   

11.
Experiments using equilibrium dialysis and fluorescence quenching provided direct evidence that approximately four moles of L-aspartic acid were bound per mole of tetrameric L-asparaginase from Escherichia coli, with a dissociation constant on the order of 60-160 microM. In addition, a set of weaker binding sites with a dissociation constant in the millimolar range were detected. Kinetic studies also revealed that L-aspartic acid inhibited L-asparaginase competitively, with an inhibition constant of 80 microM at micromolar concentrations of L-asparagine; at millimolar concentrations of the amide, an increase in maximal velocity but a decrease in affinity for L-asparagine were observed. L-Aspartic acid at millimolar levels again displayed competitive inhibition. These and other observations suggest that L-aspartic acid binds not only to the active site but also a second site with lower intrinsic affinity for it. The observed "substrate activation" is most likely attributable to the binding of a second molecule of L-asparagine rather than negative cooperativity among the tight sites of the subunits of this tetrameric enzyme. Further support for L-aspartic acid binding to the active site comes from experiments in which the enzyme, when exposed to various group-specific reagents suffered parallel loss of catalytic activity and in its ability to bind L-aspartic acid. Different commercial preparations of Escherichia coli L-asparaginase were found to contain approximately 2-4 moles of L-aspartic acid; these were incompletely removed by dialysis, but could be removed by transamination or decarboxylation. Efficiency of dialysis increased with increasing pH. Taken together, this set of results is consistent with the existence of a covalent beta-aspartyl enzyme intermediate.  相似文献   

12.
Abstract

Experiments using equilibrium dialysis and fluorescence quenching provided direct evidence that approximately four moles of L-aspartic acid were bound per mole of tetrameric L-asparaginase from Escherichia coli, with a dissociation constant on the order of 60-160 μM. In addition, a set of weaker binding sites with a dissociation constant in the millimolar range were detected. Kinetic studies also revealed that L-aspartic acid inhibited L-asparaginase competitively, with an inhibition constant of 80 μM at micromolar concentrations of L-asparagine; at millimolar concentrations of the amide, an increase in maximal velocity but a decrease in affinity for L-asparagine were observed. L-Aspartic acid at millimolar levels again displayed competitive inhibition. These and other observations suggest that L-aspartic acid binds not only to the active site but also a second site with lower intrinsic affinity for it. The observed “substrate activation” is most likely attributable to the binding of a second molecule of L-asparagine rather than negative cooperativity among the tight sites of the subunits of this tetrameric enzyme. Further support for L-aspartic acid binding to the active site comes from experiments in which the enzyme, when exposed to various group-specific reagents suffered parallel loss of catalytic activity and in its ability to bind L-aspartic acid. Different commercial preparations of Escherichia coli L-asparaginase were found to contain ~ 2-4 moles of L-aspartic acid; these were incompletely removed by dialysis, but could be removed by transamination or decarboxylation. Efficiency of dialysis increased with increasing pH. Taken together, this set of results is consistent with the existence of a covalent β-aspartyl enzyme intermediate.  相似文献   

13.
Human glycoasparaginase (N4-(beta-N-acetyl-D-glucosaminyl)-L-asparaginase, EC 3.5.1.26) hydrolyzes a series of compounds that contain L-asparagine residue with free alpha-amino and alpha-carboxyl groups. Substrates include high mannose and complex type glycoasparagines as well as those that lack the di-N-acetylchitobiose moiety, L-aspartic acid beta-methyl ester and L-aspartic acid beta-hydroxamate. The enzyme is inactive toward L-asparagine and L-glutamine and glycoasparagines containing substituted alpha-amino and/or alpha-carboxyl groups. In the presence of the acyl acceptor hydroxylamine, glycoasparaginase catalyzes the synthesis of L-aspartic acid beta-hydroxamate from aspartyl-glucosamine, L-aspartic acid beta-methyl ester, and L-aspartic acid. 13C NMR studies using 18O-labeled L-aspartic acid demonstrate that glycoasparaginase catalyzes an oxygen exchange between water and the carboxyl group at C-4 of L-aspartic acid. These results indicate that glycoasparaginase reacts as an exo-hydrolase toward the L-asparagine moiety of the substrates and the free alpha-amino and alpha-carboxyl groups are required for the enzyme reaction. The results are consistent with an L-asparaginase-like reaction pathway which involves a beta-aspartyl enzyme intermediate. Since glycoasparaginase is active toward a series of structurally different glycoasparagines, we suggest the revised systematic name of N4-(beta-glycosyl)-L-asparaginase for the enzyme.  相似文献   

14.
The kinetic mechanisms of the reactions catalyzed by the two catalytic domains of aspartokinase-homoserine dehydrogenase I from Escherichia coli have been determined. Initial velocity, product inhibition, and dead-end inhibition studies of homoserine dehydrogenase are consistent with an ordered addition of NADPH and aspartate beta-semialdehyde followed by an ordered release of homoserine and NADP+. Aspartokinase I catalyzes the phosphorylation of a number of L-aspartic acid analogues and, moreover, can utilize MgdATP as a phosphoryl donor. Because of this broad substrate specificity, alternative substrate diagnostics was used to probe the kinetic mechanism of this enzyme. The kinetic patterns showed two sets of intersecting lines that are indicative of a random mechanism. Incorporation of these results with the data obtained from initial velocity, product inhibition, and dead-end inhibition studies at pH 8.0 are consistent with a random addition of L-aspartic acid and MgATP and an ordered release of MgADP and beta-aspartyl phosphate.  相似文献   

15.
beta-decarboxylation of L-aspartic acid was observed in the system, pyridoxal: L-aspartic acid:aluminum(III), 1:100:1 when heated at 80 degrees for three hours. This reaction was followed by electronic spectroscopy and showed quantitative conversion of pyridoxal to pyridoxamine indicating decarboxylation of the ketimine. alpha-Methyl-L-aspartic acid was not decarboxylated indicating the presence of the alpha-proton and prior transamination as requirements for decarboxylation. When pyridoxamine and oxalo-2-propionic acid were reacted at pD 4.60, product analysis by nmr showed the presence of pyridoxamine and alpha-ketobutyric acid, indicating hydrolysis of the decarboxylated ketimine. Decarboxylation was fast compared to spontaneous decarboxylation. A mechanism is proposed for non-enzymatic decarboxylation and the previously suggested mechanism for the inactivation of the enzyme aspartate beta-decarboxylase is discussed.  相似文献   

16.
T Tanaka  M Ito  T Ohmura  H Hidaka 《Biochemistry》1985,24(19):5281-5284
Ca2+-dependent cyclic nucleotide phosphodiesterase (Ca2+-PDE) activity was stimulated by poly(L-aspartic acid) but not by poly(L-glutamic acid), poly(L-arginine), poly(L-lysine), and poly(L-proline). This activation was Ca2+ independent and did not further enhance the activation of Ca2+-PDE by Ca2+-calmodulin (CaM). Poly(L-aspartic acid) produced an increase in the Vmax of the phosphodiesterase, associated with a decrease in the apparent Km for the substrate, such being similar to results obtained with Ca2+-CaM. Poly(L-aspartic acid) did not significantly stimulate myosin light chain kinase and other types of cyclic nucleotide phosphodiesterase. CaM antagonists such as N-(6-aminohexyl)-5-chloro-1-naphthalenesulfonamide (W-7), trifluoperazine, and chlorpromazine selectively antagonized activation of the enzyme by poly(L-aspartic acid). Kinetic analysis of W-7-induced inhibition of activation of phosphodiesterase by poly(L-aspartic acid) was in a competitive fashion, and the Ki value was 0.19 mM. On the other hand, prenylamine, another type of calmodulin antagonist that binds to CaM at sites different from the W-7 binding sites, did not inhibit the poly(L-aspartic acid)-induced activation of Ca2+-dependent cyclic nucleotide phosphodiesterase. These results imply that poly(L-aspartic acid) is a calcium-independent activator of Ca2+-dependent phosphodiesterase and that aspartic acids in the CaM molecule may play an important role in the activation of Ca2+-PDE.  相似文献   

17.
The enzyme L-aspartase from Escherichia coli has an absolute specificity for its amino acid substrate. An examination of a wide range of structural analogues of L-aspartic acid did not uncover any alternate substrates for this enzyme. A large number of competitive inhibitors of the enzyme have been characterized, with inhibition constants ranging over 2 orders of magnitude. A divalent metal ion is required for enzyme activity above pH 7, and this requirement is met by many transition and alkali earth metals. The binding stoichiometry has been established to be one metal ion bound per subunit. Paramagnetic relaxation studies have shown that the divalent metal ion binds at the recently discovered activator site on L-aspartase and not at the enzyme active site. Enzyme activators are bound within 5 A of the enzyme-bound divalent metal ion. The activator site is remote from the active site of the enzyme, since the relaxation of inhibitors that bind at the active site is not affected by paramagnetic metal ions bound at the activator site.  相似文献   

18.
The use of Escherichia coli asparaginase II as a drug for the treatment of acute lymphoblastic leukemia is complicated by the significant glutaminase side activity of the enzyme. To develop enzyme forms with reduced glutaminase activity, a number of variants with amino acid replacements in the vicinity of the substrate binding site were constructed and assayed for their kinetic and stability properties. We found that replacements of Asp248 affected glutamine turnover much more strongly than asparagine hydrolysis. In the wild-type enzyme, N248 modulates substrate binding to a neighboring subunit by hydrogen bonding to side chains that directly interact with the substrate. In variant N248A, the loss of transition state stabilization caused by the mutation was 15 kJ mol(-1) for L-glutamine compared to 4 kJ mol(-1) for L-aspartic beta-hydroxamate and 7 kJ mol(-1) for L-asparagine. Smaller differences were seen with other N248 variants. Modeling studies suggested that the selective reduction of glutaminase activity is the result of small conformational changes that affect active-site residues and catalytically relevant water molecules.  相似文献   

19.
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