用户名: 密码: 验证码:
米多霉素生物合成机理的研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
米多霉素是一类肽核苷类抗生素,具有强烈的抑制植物白粉病的活性。本研究利用来自杀稻瘟菌素生物合成基因簇上的胞嘧啶核苷单磷酸水解酶基因blsM及其同源基因设计兼并引物,从生裂链轮丝菌ZJU5119的基因组文库中筛选出六个相互重叠的柯斯质粒,其中柯斯质粒14A6可以使异源宿主变铅青链霉菌合成米多霉素,表明其包含了米多霉素生物合成所必须的全部基因。
     对14A6进行序列测定,其插入序列全长43,561bp,通过FramePlot beta4.0预测,它包含41个开放阅读框。其中milA编码了胞嘧啶核苷单磷酸羟甲基化酶,敲除milA的生裂链轮丝菌突变株只产生去羟甲基米多霉素。MilB是BlsM的同源蛋白,milB的突变株丧失了产生米多霉素和去羟甲基米多霉素的能力。大肠杆菌融合表达的MilA蛋白能够专一的在胞嘧啶核苷单磷酸的C5位上引入羟甲基基团,而不接受胞嘧啶胞嘧啶脱氧核苷单磷酸作为其底物;融合表达的MilB的酶动力学实验表明:它能够高效的水解羟甲基胞嘧啶核苷单磷酸成羟甲基胞嘧啶,同时也能够水解胞嘧啶核苷单磷酸形成胞嘧啶;在MilA和MilB共同作用下形成的羟甲基胞嘧啶胞嘧啶的比例约为9:1,而生裂链轮丝的发酵产物中米多霉素的产量与去羟甲基米多霉素比例却只有3:1,这种不一致在UDP-葡萄糖醛酸转移酶MilC的动力学反应参数中得到了解释,MilC催化UDP-葡萄糖醛酸与胞嘧啶和羟甲基胞嘧啶反应的Kcat/Km分别为: 0.5250±0.0019和0.2719±0.0012,表明其对胞嘧啶有着更高的催化效率,这部分抵消了MilB对羟甲基胞嘧啶强烈的偏好性。
     米多霉素生物合成中一个让人迷惑而又引人兴趣的问题是葡萄糖残基C6原子的来源。用稳定性同位素C13 6个碳原子全标记的精氨酸喂养生裂链轮丝菌,其发酵产物中发现了分子量为520的米多霉素,含量比本底水平增高5倍,而非标记的米多霉素分子量为514。同时,高分辨二级质谱数据中所有包含胍基侧链的断裂碎片都比未喂养的对应部分分子量增高了6,表明标记的6个碳原子在精氨酸或者其衍生物与葡萄糖醛酸缩合过程中全部保留了下来,脱羧反应去掉的是葡萄糖醛酸羧基碳。所以,米多霉素葡萄糖残基C6原子来源于精氨酸而不是葡萄糖。
     为了系统的提出米多霉素完整的生物合成途径,我们结合异源表达和系统突变,首先确定了米多霉素生物合成基因簇的左右两个边界分别位于orf-1-milA与milQ-orf+1之间,对其中包括的17个orf一一进行突变,除了milL突变没有影响之外,证明其它的16个基因与米多霉素的生物合成相关。
     milG编码的是一个新颖的Radical SAM家族蛋白,它是唯一一个氧化还原酶候选基因,负责羟甲基胞嘧啶葡萄糖醛酸C4位上羟基氧化成羰基,在milG突变株的发酵液中,中间产物羟甲基胞嘧啶葡萄糖醛酸大量积累,这充分证明了其功能就是负责这一步关键的催化反应,这也是第一次在核苷类抗生素生物合成中报道的新型氧化酶基因。分别敲除两个氨基糖苷类磷酸转移酶基因milE和milQ,突变株都丧失了产生米多霉素的能力,推测这两个磷酸转移酶可能负责糖上的C2-C3间双键的形成。可能编码天冬氨酸/酪氨酸/芳香族氨基酸类氨基转移酶的milM和编码二氢吡啶甲酸合酶的milN基因的突变株均不能够再产生米多霉素,MilM可能参与催化精氨酸α位的氨基转化成羰基而形成α-酮酸,而MilN则负责α-羰基与脱羧后的己糖碳负离子之间的缩合。编码degT/dnrJ/eryC1/strS类氨基转移酶的milD和编码可能的连接酶功能的milH突变株都丧失了产生米多霉素的能力,推测前者是负责C4位上羰基的转氨反应,后者负责将激活的丝氨酸连接到C4位的氨基上,形成类似肽键的酰胺键。而中断LuxR家族的转录调节因子基因milO也消除了米多霉素的生产,说明它可能是一个途径专一性的调节基因。MilJ显示与Ubiquitone的羟甲基化酶有一定的同源性,可能负责精氨酸的gamma羟化;而MilI可能负责了丝氨酸的活化;ABC transporter基因milP的突变株也不产米多霉素,它很可能是米多霉素的抗性基因;编码Major Facilitator Superfamily的基因milk的突变株还能够产生米多霉素,但产量降低,它可能与磷酸葡萄糖的转运有关。
Mildiomycin (MIL) is a peptidyl nucleoside antibiotic with strong activity against powdery mildew disease of plants. In present work, the degenerate primers were designed based on blsM, the gene of CMP hydrolase in the blsticidin S biosynthetic gene cluster, together with its homologues. With the primers, six overlapping cosmids were screened from the genomic library of Streptoverticillum rimofaciens ZJU5119. 14A6, one of them can confer Streptomyces lividans 1326 to produce MIL, implying that it contains all the essential genes for biosynthesis of MIL.
     The insertion sequence of 14A6 was determined and assembled into 43,561bp on which 41 ORFs were predicted by FramePlot beta4.0. The mutant of milA, a gene encoding CMP hydroxymethylase, could produce only dehydroxymethyl MIL (dHM-MIL), while neither of MIL and dHM-MIL can be biosynthesized when milB, homolog of blsM was disrupted. Recombinant MilA was purified in E. coli and shown to specifically introduce a C-5 hydroxymethyl group on CMP but not able to accept either cytosine or dCMP as substrate. The kinetic parameters of recombinant MilB showed that it hydrolyzed hydroxymethyl-CMP to hydroxymethyl-cytosine (HMC) more efficiently than did CMP to cytosine. With the CMP as substrate, the ratio of free HMC to cytosine generated by MilB and MilA was ca. 9:1 in in vitro assays while the ratio of MIL to dHM-MIL was 3:1 in the extracted broth of Streptoverticillum rimofaciens. The inconsistence was partly compensated by the substrate bias of MilC, which can catalyse the coupling of the cytosine or HMC with UDP-glucuronic acid into cytosylglucuronic acid (CGA) or HMCGA, respectively. The Kcat/Km of MilC to cytosine and HMC were 0.5250±0.0019 and 0.2719±0.0012, respectively.
     A puzzling and interesting question regarding MIL biosynthesis was the origin of C-6 in carboxyl group. Using the stable isotope labeled L-arginine (U-13C6) to feed Streptoverticillum rimofaciens, the productivity of MIL with a mass of 520, 6 increments than that without isotope incorporation, increased to 11% from 2% after feeding of labeled L-arginine. Q-TOF/MS analyses of the purified mildiomycin indicated that all the fragments containing the guanidino-side chain gave a mass increment of 6, indicating thet all of six carbon atoms of labeled L-arginine were kept after the coupling of L-arginine or its derivative with glucuronic acid moiety, namely, C6 was derived from L-arginine rather than sugar as the precursors of MIL.
     To propose the complete biosynthetic pathway of MIL, the boundaries were firstly determined via systamtic gene disruption as well as end sequencing of six overlapped cosmids. Two boundaries were located within orf-1-milA and milQ-orf+1. All included genes within the boundary were disrupted individually, and 16 genes were demonstrated to be involved in the biosynthesis of MIL except for milL, of which knock-out caused no effect on production of MIL.
     The gene of milG encodes a unique Radical SAM family protein and is the only oxidase candidate supposed to be in charge of the oxidation of the hydroxyl group to carbonyl group in C4 of the hydroxymethyl cytosylglucuronic acid. In accordance with this assumption, intermediate of HMCGA was accumulated highly in the milG mutant, strongly supporting the role of MilG as a novel oxidase.in this key step. This was the first report of Radical SAM protein as the oxidase in the biosynthesis of nucleoside antibiotics.
     Disruption of milE and milQ, two aminoglycoside phosphotransferase genes, abolished the production of MIL; they are presumably involved in the double bond formation between C2 and C3 of the sugar moiety. The gene of milM and milN, encoding an aspartate/tyrosine/aromatic aminotransferase and a dihydrodipicolinate synthetase respectively, were disrupted, and neither resultant mutant could produce MIL. MilM was supposed to convertα-amino group of L-arginine to carbonyl group, and MilN sequentially catalyze the coupling of the carbonyl group in theα-ketonic acid with the carboanion in the decarboxylated hexose. Another aminotransferase gene, namely milD, encoding a degT/dnrJ/eryC1/strS aminotransferase, was considered to transfer amino group to the carbonyl group in the C-4 of sugar residue. MilH, the homologue of BlsK, might be in charge of the attachment of activated L-serine to C4 group of sugar moiety to form amide-like bond in the MIL biosynthesis.
     The disruption of a LuxR family regulator gene, milO, abolished the production of MIL, suggesting MilO should be a pathway-specific positive regulator. MilJ shows some homology with the Ubiquitone hydroxylase and postulated to be responsible for the hydroxylation ofγ-carbon in arginine residue. MilI was supposed to be involved in the activivation of L-serine. The mutant of an ABC transporter gene, milP, which was considered as the resistance gene, could not produce the MIL; the mutant of milk, a Major Facilitator Superfamily gene, produced less MIL compared to the wild type. MilK was proposed to be involved in the transportation of phosphogluconate.
引文
1. Waksman, S.A. and A.T. Henrici, The Nomenclature and Classification of the Actinomycetes. J Bacteriol, 1943. 46(4): p. 337-41.
    2. Redenbach, M., et al., A set of ordered cosmids and a detailed genetic and physical map for the 8 Mb Streptomyces coelicolor A3(2) chromosome. Mol Microbiol, 1996. 21(1): p. 77-96.
    3. Tewfik, E.M. and S.G. Bradley, Characterization of deoxyribonucleic acids from streptomycetes and nocardiae. J Bacteriol, 1967. 94(6): p. 1994-2000.
    4. Bentley, S.D., et al., Complete genome sequence of the model actinomycete Streptomyces coelicolor A3(2). Nature, 2002. 417(6885): p. 141-7.
    5. Kieser, T., M.J. Bibb, K.F. Chater, M.J. Butter, and D.A. Hopwood., Practical Streptomyces genetics. a laboratory manual. . 2000, Norwich, : John Innes Foundation, United Kingdom.
    6. Isono, K., Nucleoside antibiotics: structure, biological activity, and biosynthesis. J Antibiot (Tokyo), 1988. 41(12): p. 1711-39.
    7. Lacalle, R.A., J.A. Tercero, and A. Jimenez, Cloning of the complete biosynthetic gene cluster for an aminonucleoside antibiotic, puromycin, and its regulated expression in heterologous hosts. Embo J, 1992. 11(2): p. 785-92.
    8. Isono, K., Current progress on nucleoside antibiotics. Pharmacol Ther, 1991. 52(3): p. 269-86.
    9. Hirasawa, K. and K. Isono, Formation of 8-azaguanine from guanine by Streptomyces albus. J Antibiot (Tokyo), 1978. 31(6): p. 628-9.
    10. Drautz, H., et al., Metabolic products of microorganisms. 239. Bacimethrin isolated from Streptomyces albus identification, derivatives, synthesis and biological properties. J Antibiot (Tokyo), 1987. 40(10): p. 1431-9.
    11. McCarthy, J.R., Jr., R.K. Robins, and M.J. Robins, Purine nucleosides. XXII. The synthesis of angustmycin A (decoyinine) and related unsaturated nucleosides. J Am Chem Soc, 1968. 90(18): p. 4993-9.
    12. Cunningham, K.G., et al., Cordycepin, a metabolic product isolated from culturesof Cordyceps militaris (Linn.) Link. Nature, 1950. 166(4231): p. 949.
    13. Morton, G.O., et al., The structure of nucleocidin. 3. (A new structure). J Am Chem Soc, 1969. 91(6): p. 1535-7.
    14. Takeuchi, S., et al., Blasticidin S, a new antibiotic. J Antibiot (Tokyo), 1958. 11(1): p. 1-5.
    15. Roberts, W.P., M.E. Tate, and A. Kerr, Agrocin 84 is a 6-N-phosphoramidate of an adenine nucleotide analogue. Nature, 1977. 265(5592): p. 379-81.
    16. Sharma, C.B., et al., The exotoxin of Bacillus thuringiensis: a new C-mitotic agent. Experientia, 1976. 32(11): p. 1465-6.
    17. Hopwood, D.A., Genetic engineering of Streptomyces to create hybrid antibiotics. Curr Opin Biotechnol, 1993. 4(5): p. 531-7.
    18. Hopwood, D.A., Genetic Contributions to Understanding Polyketide Synthases. Chem Rev, 1997. 97(7): p. 2465-2498.
    19. Cundliffe, E. and K. McQuillen, Peptide bond formation: the effects of antibiotics on the puromycin reaction. J Gen Microbiol, 1968. 50(3): p. Suppl:x.
    20. Vara, J., et al., Cloning and expression of a puromycin N-acetyl transferase gene from Streptomyces alboniger in Streptomyces lividans and Escherichia coli. Gene, 1985. 33(2): p. 197-206.
    21. Vara, J.A., et al., Two genes in Streptomyces alboniger puromycin biosynthesis pathway are closely linked. Gene, 1988. 69(1): p. 135-40.
    22. Lacalle, R.A., et al., Molecular analysis of the pac gene encoding a puromycin N-acetyl transferase from Streptomyces alboniger. Gene, 1989. 79(2): p. 375-80.
    23. Lacalle, R.A., D. Ruiz, and A. Jimenez, Molecular analysis of the dmpM gene encoding an O-demethyl puromycin O-methyltransferase from Streptomyces alboniger. Gene, 1991. 109(1): p. 55-61.
    24. Tercero, J.A., et al., The biosynthetic pathway of the aminonucleoside antibiotic puromycin, as deduced from the molecular analysis of the pur cluster of Streptomyces alboniger. J Biol Chem, 1996. 271(3): p. 1579-90.
    25. Sanchez, M.B., et al., The pur3 gene from the pur cluster encodes a monophosphatase essential for puromycin biosynthesis in Streptomyces. FEBS Lett,2006. 580(7): p. 1807-11.
    26. Angel Rubio, M., et al., The pur6 gene of the puromycin biosynthetic gene cluster from Streptomyces alboniger encodes a tyrosinyl-aminonucleoside synthetase. FEBS Lett, 2004. 577(3): p. 371-5.
    27. Espinosa, J.C., et al., The pur7 gene from the puromycin biosynthetic pur cluster of Streptomyces alboniger encodes a nudix hydrolase. J Bacteriol, 1999. 181(16): p. 4914-8.
    28. Tercero, J.A., R.A. Lacalle, and A. Jimenez, The pur8 gene from the pur cluster of Streptomyces alboniger encodes a highly hydrophobic polypeptide which confers resistance to puromycin. Eur J Biochem, 1993. 218(3): p. 963-71.
    29. Rubio, M.A., et al., The Pur10 protein encoded in the gene cluster for puromycin biosynthesis of Streptomyces alboniger is an NAD-dependent ATP dehydrogenase. FEBS Lett, 1998. 437(3): p. 197-200.
    30. Tercero, J.A., J.C. Espinosa, and A. Jimenez, Expression of the Streptomyces alboniger pur cluster in Streptomyces lividans is dependent on the bldA-encoded tRNALeu. FEBS Lett, 1998. 421(3): p. 221-3.
    31. Kirst, H.A., et al., The structure of A201A, a novel nucleoside antibiotic. J Antibiot (Tokyo), 1985. 38(5): p. 575-86.
    32. Saugar, I., et al., Identification of a set of genes involved in the biosynthesis of the aminonucleoside moiety of antibiotic A201A from Streptomyces capreolus. Eur J Biochem, 2002. 269(22): p. 5527-35.
    33. Barrasa, M.I., J.A. Tercero, and A. Jimenez, The aminonucleoside antibiotic A201A is inactivated by a phosphotransferase activity from Streptomyces capreolus NRRL 3817, the producing organism. Isolation and molecular characterization of the relevant encoding gene and its DNA flanking regions. Eur J Biochem, 1997. 245(1): p. 54-63.
    34. Barrasa, M.I., et al., The ard1 gene from Streptomyces capreolus encodes a polypeptide of the ABC-transporters superfamily which confers resistance to the aminonucleoside antibiotic A201A. Eur J Biochem, 1995. 228(3): p. 562-9.
    35. Delzer, J., et al., New nikkomycins by mutasynthesis and directed fermentation. JAntibiot (Tokyo), 1984. 37(1): p. 80-2.
    36. Mohrle, V., U. Roos, and C. Bormann, Identification of cellular proteins involved in nikkomycin production in Streptomyces tendae Tu901. Mol Microbiol, 1995. 15(3): p. 561-71.
    37. Bormann, C., V. Mohrle, and C. Bruntner, Cloning and heterologous expression of the entire set of structural genes for nikkomycin synthesis from Streptomyces tendae Tu901 in Streptomyces lividans. J Bacteriol, 1996. 178(4): p. 1216-8.
    38. Lauer, B., R. Russwurm, and C. Bormann, Molecular characterization of two genes from Streptomyces tendae Tu901 required for the formation of the 4-formyl-4-imidazolin-2-one-containing nucleoside moiety of the peptidyl nucleoside antibiotic nikkomycin. Eur J Biochem, 2000. 267(6): p. 1698-706.
    39. Lauer, B., et al., Molecular characterization of co-transcribed genes from Streptomyces tendae Tu901 involved in the biosynthesis of the peptidyl moiety and assembly of the peptidyl nucleoside antibiotic nikkomycin. Mol Gen Genet, 2001. 264(5): p. 662-73.
    40. Bruntner, C. and C. Bormann, The Streptomyces tendae Tu901 L-lysine 2-aminotransferase catalyzes the initial reaction in nikkomycin D biosynthesis. Eur J Biochem, 1998. 254(2): p. 347-55.
    41. Carrell, C.J., et al., NikD, an unusual amino acid oxidase essential for nikkomycin biosynthesis: structures of closed and open forms at 1.15 and 1.90 A resolution. Structure, 2007. 15(8): p. 928-41.
    42. Venci, D., G. Zhao, and M.S. Jorns, Molecular characterization of NikD, a new flavoenzyme important in the biosynthesis of nikkomycin antibiotics. Biochemistry, 2002. 41(52): p. 15795-802.
    43. Chen, H., et al., Formation of beta-hydroxy histidine in the biosynthesis of nikkomycin antibiotics. Chem Biol, 2002. 9(1): p. 103-12.
    44. Liu, G., et al., A pathway-specific transcriptional regulatory gene for nikkomycin biosynthesis in Streptomyces ansochromogenes that also influences colony development. Mol Microbiol, 2005. 55(6): p. 1855-66.
    45. Ginj, C., et al., 3'-Enolpyruvyl-UMP, a novel and unexpected metabolite innikkomycin biosynthesis. Chembiochem, 2005. 6(11): p. 1974-6.
    46. Huang, K.T., T. Misato, and H. Asuyama, Effect of Blasticidin S on Protein Synthesis of Piricularia Oryzae. J Antibiot (Tokyo), 1964. 17: p. 65-70.
    47. Otake, N., et al., The structure of blasticidin S. Tetrahedron Lett, 1965. 19: p. 1411-9.
    48. Seto, H. and H. Yonehara, Studies on the biosynthesis of blasticidin S. VI. The isolation and structure of blasticidin H. J Antibiot (Tokyo), 1977. 30(11): p. 1019-21.
    49. Seto, H., K. Furihata, and H. Yonehara, Studies on the biosynthesis of blasticidin S. V. Isolation and structure of pentopyranic acid. J Antibiot (Tokyo), 1976. 29(5): p. 595-6.
    50. Seto, H., I. Yamaguchi, and N. Otake, Studies on the Biosynthesis of Blasticidin S Part I. Precursors of Blasticidin S Biosynthesis. Agr Biol Chem, 1968. 32(10): p. 1292-1289.
    51. Cone, M.C., et al., Cloning and heterologous expression of blasticidin S biosynthetic genes from Streptomyces griseochromogenes. J Antibiot (Tokyo), 1998. 51(6): p. 570-8.
    52. Cone, M.C., et al., The blasticidin S biosynthesis gene cluster from Streptomyces griseochromogenes: sequence analysis, organization, and initial characterization. Chembiochem, 2003. 4(9): p. 821-8.
    53. Grochowski, L.L. and T.M. Zabriskie, Characterization of BlsM, a nucleotide hydrolase involved in cytosine production for the biosynthesis of blasticidin S. Chembiochem, 2006. 7(6): p. 957-64.
    54. Waksman, S.A., Production and Activity of Streptothricin. J Bacteriol, 1943. 46(3): p. 299-310.
    55. Kusumoto, S., et al., Total chemical structure of streptothricin. J Antibiot (Tokyo), 1982. 35(7): p. 925-7.
    56. Haupt, I., et al., Action of streptothricin F on ribosomal functions. J Antibiot (Tokyo), 1980. 33(6): p. 636-41.
    57. Gold, S.E., et al., Three selectable markers for transformation of Ustilago maydis.Gene, 1994. 142(2): p. 225-30.
    58. Kobayashi, T., et al., Isolation and characterization of a pock-forming plasmid pTA4001 from Streptomyces lavendulae. J Antibiot (Tokyo), 1984. 37(4): p. 368-75.
    59. Fernandez-Moreno, M.A., C. Vallin, and F. Malpartida, Streptothricin biosynthesis is catalyzed by enzymes related to nonribosomal peptide bond formation. J Bacteriol, 1997. 179(22): p. 6929-36.
    60. Grammel, N., et al., A beta-lysine adenylating enzyme and a beta-lysine binding protein involved in poly beta-lysine chain assembly in nourseothricin synthesis in Streptomyces noursei. Eur J Biochem, 2002. 269(1): p. 347-57.
    61. Iwasa, T., K. Suetomi, and T. Kusaka, Taxonomic study and fermentation of producing organism and antimicrobial activity of mildiomycin. J Antibiot (Tokyo), 1978. 31(6): p. 511-8.
    62. Harada, S. and T. Kishi, Isolation and characterization of mildiomycin, a new nucleoside antibiotic. J Antibiot (Tokyo), 1978. 31(6): p. 519-24.
    63. Harada, S., E. Mizuta, and T. Kishi, Structure of mildiomycin, a new antifungal nucleoside antibiotic. tetrahedron, 1981. 37: p. 1317-1327.
    64. Harada, S., E. Mizuta, and T. Kishi, Structure of Mildiomycin, a New Antifungal Nucleoside Antibiotic. J Am. Chem. Soc., 1978. 100(15): p. 4895-4897.
    65.吴耀荣,赵双宜, and夏光敏,植物抗白粉病的分子机理.中国生物工程杂志, 2002. 22(3): p. 54-57.
    66.周益林,段霞瑜, and盛宝钦,植物白粉病的化学防治进展.农药学学报, 2001. 3(2): p. 12-18.
    67. Feduchi, E., M. Cosin, and L. Carrasco, Mildiomycin: a nucleoside antibiotic that inhibits protein synthesis. J Antibiot (Tokyo), 1985. 38(3): p. 415-9.
    68. Kishimoto, K., et al., Effect of ferrous ion on mildiomycin production by Streptoverticillium rimofaciens. J Antibiot (Tokyo), 1996. 49(8): p. 770-4.
    69. Kishimoto, K., et al., Effect of phosphate ion on mildiomycin production by Streptoverticillium rimofaciens. J Antibiot (Tokyo), 1996. 49(8): p. 775-80.
    70. Kishimoto, K. and S. Akiyama, Stimulatory effect of ferrous ion on mildiomycinproduction by streptoverticillium rimofaciens. biotechnology letters, 1997. 19(7): p. 699-702.
    71. Kishimoto, K., et al., Effect of ferrous ion on amino acid metabolism in mildiomycin production by Streptoverticillium rimofaciens. J Antibiot (Tokyo), 1997. 50(3): p. 206-11.
    72. Sawada, H., et al., Strain improvement of a mildiomycin producer, Streptoverticillium rimofaciens B-98891 Appl Microbiol Biotechnol, 1986. 25: p. 132-136.
    73.徐志南, et al.,米多霉素产生菌原生质体融合研究.浙江大学学报(工学版), 2006. 40(7): p. 1262-1266.
    74. Sawada, H., et al., Microbial production of Mildiomycin Analogues by the addition of cytosine analogues. J Ferment Technol, 1984. 62(6): p. 537-543.
    75.梁立峰, et al.,米多霉素衍生物的生物合成及结构鉴定.浙江大学学报(工学版), 2006. 40(11): p. 1909-1913.
    76. Ye, D., Z.N. Xu, and P.L. Cen, Medium optimization for enhanced production of cytosine-substituted mildiomycin analogue (MIL-C) by Streptoverticillium rimofaciens ZJU 5119. J Zhejiang Univ Sci B, 2008. 9(1): p. 77-84.
    77. Wyatt, G.R. and S.S. Cohen, The bases of the nucleic acids of some bacterial and animal viruses: the occurrence of 5-hydroxymethylcytosine. Biochem J, 1953. 55(5): p. 774-82.
    78. Sawada, H., et al., Formation of 5-Hydroxymethylcytosine from Cytidine 5'-Mono-Phosphate by Streptoverticillium rimofaciens. J. Ferment. Technol.,, 1985. 63(1): p. 17-21.
    79. Flaks, J.G. and S.S. Cohen, The enzymic synthesis of 5-hydroxymethyldeoxycytidylic acid. Biochim Biophys Acta, 1957. 25(3): p. 667-8.
    80. Sawada, H., et al., Biosynthetic Pathway of 5-Hydroxymethylcytosine by Streptoverticillium rimofaciens. J Ferment Technol, 1985. 63(1): p. 23-27.
    81. Li, L., et al., The Mildiomycin Biosynthesis: Initial Steps for Sequential Generation of 5-Hydroxymethylcytidine 5'-Monophosphate and 5-Hydroxymethylcytosine in Streptoverticillium rimofaciens ZJU5119. Chembiochem, 2008. 9(8): p.1286-1294.
    82. Hopwood, D.A., et al., Plasmids, recombination and chromosome mapping in Streptomyces lividans 66. J Gen Microbiol, 1983. 129(7): p. 2257-69.
    83. Paget, M.S., et al., Evidence that the extracytoplasmic function sigma factor sigmaE is required for normal cell wall structure in Streptomyces coelicolor A3(2). J Bacteriol, 1999. 181(1): p. 204-11.
    84. Gust, B., et al., PCR-targeted Streptomyces gene replacement identifies a protein domain needed for biosynthesis of the sesquiterpene soil odor geosmin. Proc Natl Acad Sci U S A, 2003. 100(4): p. 1541-6.
    85. Bierman, M., et al., Plasmid cloning vectors for the conjugal transfer of DNA from Escherichia coli to Streptomyces spp. Gene, 1992. 116(1): p. 43-9.
    86. Short, J.M., et al., Lambda ZAP: a bacteriophage lambda expression vector with in vivo excision properties. Nucleic Acids Res, 1988. 16(15): p. 7583-600.
    87. Sambrook, J., E.F. Fritsch, and T. Maniatis, Molecular cloning: a laboratory manual, 2nd ed. . 1989, N.Y.: Cold Spring Harbor Laboratory, Cold Spring Harbor,.
    88. Tatusov, R.L., et al., The COG database: new developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res, 2001. 29(1): p. 22-8.
    89. Tatusov, R.L., et al., The COG database: a tool for genome-scale analysis of protein functions and evolution. Nucleic Acids Res, 2000. 28(1): p. 33-6.
    90. Kaminski, P.A., et al., In Vivo Reshaping the Catalytic Site of Nucleoside 2'-Deoxyribosyltransferase for Dideoxy- and Didehydronucleosides via a Single Amino Acid Substitution. J Biol Chem, 2008. 283(29): p. 20053-9.
    91. Cook, W.J., S.A. Short, and S.E. Ealick, Crystallization and preliminary X-ray investigation of recombinant Lactobacillus leichmannii nucleoside deoxyribosyltransferase. J Biol Chem, 1990. 265(5): p. 2682-3.
    92. Rose, T.M., et al., Consensus-degenerate hybrid oligonucleotide primers for amplification of distantly related sequences. Nucleic Acids Res, 1998. 26(7): p. 1628-35.
    93. Smokvina, T., et al., Construction of a series of pSAM2-based integrative vectors for use in actinomycetes. Gene, 1990. 94(1): p. 53-9.
    94. Combes, P., et al., The streptomyces genome contains multiple pseudo-attB sites for the (phi)C31-encoded site-specific recombination system. J Bacteriol, 2002. 184(20): p. 5746-52.
    95. Ishikawa, J. and K. Hotta, FramePlot: a new implementation of the frame analysis for predicting protein-coding regions in bacterial DNA with a high G + C content. FEMS Microbiol Lett, 1999. 174(2): p. 251-3.
    96. Finn, R.D., et al., Pfam: clans, web tools and services. Nucleic Acids Res, 2006. 34(Database issue): p. D247-51.
    97. Benkovic, S.J., On the mechanism of action of folate- and biopterin-requiring enzymes. Annu Rev Biochem, 1980. 49: p. 227-51.
    98. Ross, P., F. O'Gara, and S. Condon, Cloning and characterization of the thymidylate synthase gene from Lactococcus lactis subsp. lactis. Appl Environ Microbiol, 1990. 56(7): p. 2156-63.
    99. Hardy, L.W., et al., Atomic structure of thymidylate synthase: target for rational drug design. Science, 1987. 235(4787): p. 448-55.
    100. Kaneda, S., et al., Structural and functional analysis of the human thymidylate synthase gene. J Biol Chem, 1990. 265(33): p. 20277-84.
    101. Guo, J. and S.J. Gould, Biosynthesis of Blasticidin S from Cytosylglucuronic Acid(CGA). Isolation of cytosine/UDPglucuronosyltransferase and Incorporation of CGA by Streptomyces griseochromogenes. J. Am. Chem. Soc., 1991. 113: p. 5898-5899.
    102. Gould, S.J. and J. Guo, Cytosylglucuronic acid synthase (cytosine: UDP-glucuronosyltransferase) from Streptomyces griseochromogenes, the first prokaryotic UDP-glucuronosyltransferase. J Bacteriol, 1994. 176(5): p. 1282-6.
    103. Marchler-Bauer, A. and S.H. Bryant, CD-Search: protein domain annotations on the fly. Nucleic Acids Res, 2004. 32(Web Server issue): p. W327-31.
    104. Altschul, S.F., et al., Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res, 1997. 25(17): p. 3389-402.
    105. Zhou, J., J. Zhang, and W. Xie, Xenobiotic nuclear receptor-mediated regulation of UDP-glucuronosyl-transferases. Curr Drug Metab, 2005. 6(4): p. 289-98.
    106. Wackett, L.P. and D.T. Gibson, Metabolism of xenobiotic compounds by enzymes in cell extracts of the fungus Cunninghamella elegans. Biochem J, 1982. 205(1): p. 117-22.
    107. Szyperski, T., 13C-NMR, MS and metabolic flux balancing in biotechnology research. Q Rev Biophys, 1998. 31(1): p. 41-106.
    108. Lundberg, P., et al., Nuclear magnetic resonance studies of cellular metabolism. Anal Biochem, 1990. 191(2): p. 193-222.
    109. McLafferty, F.W., Studies of unusual simple molecules by neutralization-reionization mass spectrometry. Science, 1990. 247(4945): p. 925-9.
    110. Sofia, H.J., et al., Radical SAM, a novel protein superfamily linking unresolved steps in familiar biosynthetic pathways with radical mechanisms: functional characterization using new analysis and information visualization methods. Nucleic Acids Res, 2001. 29(5): p. 1097-106.
    111. Wang, S.C. and P.A. Frey, S-adenosylmethionine as an oxidant: the radical SAM superfamily. Trends Biochem Sci, 2007. 32(3): p. 101-10.
    112. Friedel, M.G., et al., Synthesis and stereochemical assignment of DNA spore photoproduct analogues. Chemistry, 2006. 12(23): p. 6081-94.
    113. Frey, P.A. and O.T. Magnusson, S-Adenosylmethionine: a wolf in sheep's clothing, or a rich man's adenosylcobalamin? Chem Rev, 2003. 103(6): p. 2129-48.
    114. Ahlert, J., et al., Identification of stsC, the gene encoding the L-glutamine:scyllo-inosose aminotransferase from streptomycin-producing Streptomycetes. Arch Microbiol, 1997. 168(2): p. 102-13.
    115. Sung, M.H., et al., Thermostable aspartate aminotransferase from a thermophilic Bacillus species. Gene cloning, sequence determination, and preliminary x-ray characterization. J Biol Chem, 1991. 266(4): p. 2567-72.
    116. Ko, T.P., et al., Crystallization and preliminary crystallographic analysis of the Escherichia coli tyrosine aminotransferase. Acta Crystallogr D Biol Crystallogr, 1999. 55(Pt 8): p. 1474-7.
    117. Nakai, T., et al., Structure of Thermus thermophilus HB8 aspartate aminotransferase and its complex with maleate. Biochemistry, 1999. 38(8): p. 2413-24.
    118. Mirwaldt, C., I. Korndorfer, and R. Huber, The crystal structure of dihydrodipicolinate synthase from Escherichia coli at 2.5 A resolution. J Mol Biol, 1995. 246(1): p. 227-39.
    119. Aisaka, K. and T. Uwajima, Cloning and constitutive expression of the N-acetylneuraminate lyase gene of Escherichia coli. Appl Environ Microbiol, 1986. 51(3): p. 562-5.
    120. Murphy, P.J., et al., The Rhizobium meliloti rhizopine mos locus is a mosaic structure facilitating its symbiotic regulation. J Bacteriol, 1993. 175(16): p. 5193-204.
    121. Izard, T., et al., The three-dimensional structure of N-acetylneuraminate lyase from Escherichia coli. Structure, 1994. 2(5): p. 361-9.
    122. Zhang, R.G., et al., Structure of a bacterial quorum-sensing transcription factor complexed with pheromone and DNA. Nature, 2002. 417(6892): p. 971-4.
    123. Maris, A.E., et al., Dimerization allows DNA target site recognition by the NarL response regulator. Nat Struct Biol, 2002. 9(10): p. 771-8.
    124. Birck, C., et al., Insights into signal transduction revealed by the low resolution structure of the FixJ response regulator. J Mol Biol, 2002. 321(3): p. 447-57.
    125. Pappas, K.M., C.L. Weingart, and S.C. Winans, Chemical communication in proteobacteria: biochemical and structural studies of signal synthases and receptors required for intercellular signalling. Mol Microbiol, 2004. 53(3): p. 755-69.
    126. Ducros, V.M., et al., Crystal structure of GerE, the ultimate transcriptional regulator of spore formation in Bacillus subtilis. J Mol Biol, 2001. 306(4): p. 759-71.
    127. Schlegel, A., et al., Network regulation of the Escherichia coli maltose system. J Mol Microbiol Biotechnol, 2002. 4(3): p. 301-7.
    128. Trower, M.K. and K.G. Clark, PCR cloning of a streptomycin phosphotransferase(aphE) gene from Streptomyces griseus ATCC 12475. Nucleic Acids Res, 1990. 18(15): p. 4615.
    129. Sarwar, M. and M. Akhtar, Cloning of aminoglycoside phosphotransferase (APH) gene from antibiotic-producing strain of Bacillus circulans into a high-expression vector, pKK223-3. Purification, properties and location of the enzyme. Biochem J, 1990. 268(3): p. 671-7.
    130. Pugh, E.L. and S.J. Wakil, Studies on the mechanism of fatty acid synthesis. XIV. The prosthetic group of acyl carrier protein and the mode of its attachment to the protein. J Biol Chem, 1965. 240(12): p. 4727-33.
    131. Hung, L.W., et al., Crystal structure of the ATP-binding subunit of an ABC transporter. Nature, 1998. 396(6712): p. 703-7.
    132. Walmsley, A.R., et al., Sugar transporters from bacteria, parasites and mammals: structure-activity relationships. Trends Biochem Sci, 1998. 23(12): p. 476-81.
    133. Pao, S.S., I.T. Paulsen, and M.H. Saier, Jr., Major facilitator superfamily. Microbiol Mol Biol Rev, 1998. 62(1): p. 1-34.
    134. Frey, P.A., A.D. Hegeman, and F.J. Ruzicka, The Radical SAM Superfamily. Crit Rev Biochem Mol Biol, 2008. 43(1): p. 63-88.
    135. Fox, J.J., et al., Nucleosides. Xxv. Chemistry of Gougerotin. Antimicrob Agents Chemother (Bethesda), 1964. 10: p. 518-29.
    136. Takahashi, A., et al., Bagougeramines A and B, new nucleoside antibiotics produced by a strain of Bacillus circulans. I. Taxonomy of the producing organism and isolation and biological properties of the antibiotics. J Antibiot (Tokyo), 1986. 39(8): p. 1033-40.
    137. Hammill, R.L. and M. Hoehn, Anthelmycin, a New Antibiotic with Anthelmintic Properties. J Antibiot (Tokyo), 1964. 17: p. 100-3.
    138. Argoudelis, A.D., et al., Arginomycin: production, isolation, characterization and structure. J Antibiot (Tokyo), 1987. 40(6): p. 750-60.
    139. Watanabe, K.A., et al., Nucleosides. LXXXVII. Total synthesis of pentopyranine A, an alpha-L cytosine nucleoside elaborated by Streptomyces griseochromogenes. J Org Chem, 1974. 39(17): p. 2482-6.
    140. Chiu, T.M., et al., Nucleosides. LXXXIV. Total synthesis of pentopyranine C, a nucleoside elaborated by Streptomyces griseochromogenes. J Org Chem, 1973. 38(20): p. 3622-4.
    141. Hinuma, Y., et al., An amicetin (sacromycin) from Streptomyces sp. No. 5223. Studies on Streptomyces antibiotics. XXXVI. J Antibiot (Tokyo), 1955. 8(5): p. 148-52.
    142. Gonzalez, A., D. Vazquez, and A. Jimenez, Inhibition of translation in bacterial and eukaryotic systems by the antibiotic anthelmycin (hikizimycin). Biochim Biophys Acta, 1979. 561(2): p. 403-9.
    143. Stevens, C.L., J. Nemec, and G.H. Ransford, Total synthesis of the amino sugar nucleoside antibiotic, plicacetin. J Am Chem Soc, 1972. 94(9): p. 3280-1.
    144. Konishi, M., et al., Oxamicetin, a new antibiotic of bacterial origin. II. Structure of oxamicetin. J Antibiot (Tokyo), 1973. 26(12): p. 757-64.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700