嗜热β-1,4-葡聚糖内切酶和酰基肽水解酶突变体的晶体结构和功能研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
来源于嗜热菌的酶类具有极好的高温反应活性和稳定性,在能源、化工、食品和医药等行业展现出越来越广阔的应用潜力。解析嗜热酶及酶-底物复合物的晶体结构,研究其结构特征,将有助于揭示酶超常生物学稳定性以及高温催化反应机制的分子基础,为发掘嗜热酶的新功能、指导酶分子设计和改造等工作提供必要的理论依据。
     论文第一部分运用分子生物学技术克隆、表达、纯化了来源于嗜热菌Fervidobacterium nodosum Rt17-B1的嗜热β-1,4-葡聚糖内切酶FnCel5A及其N末端截短突变体FnCel5AND,并运用蛋白质晶体学的技术进行结晶,最终使用单波长反常散射法解析了分辨率达1.7 ?的FnCel5A和2.2 ?的复合物FnCel5AND:Glucose的晶体结构。结构分析表明,该酶具有糖基水解酶第五家族典型的(β/α)8形式TIM桶状结构;与中温和低温结构同源物比较,FnCel5A在氨基酸组成和分子内相互作用等方面具备高稳定性酶的结构特征;确定了Glu283-His126和Glu167-His226双催化活性中心,以及底物结合残基;结合分子模拟,提出了FnCel5A与底物结合的模型,阐明了该酶水解直链糖底物的催化机理。
     论文第二部分克隆、表达、表征了来源于嗜热古菌的酰基肽水解酶N末端截短突变体ApAPH-ΔN21;结晶并利用分子置换法解析了2.5 ?分辨率的突变体结构。结构分析和性质测定结果显示,ApAPH-ΔN21较野生型的最适反应温度降低18℃,对热和化学变性剂抗性降低;平均结构B因子由野生型的19.4 ?2升高到突变体的28.4 ?2。以上证据表明保守的N末端的对于保持ApAPH或家族中其他成员的稳定性具有重要作用。
Enzymes produced by thermophiles and hyperthermophiles are typically thermostable (resistant to irreversible inactivation at high temperatures) and thermophilic (i.e. optimally active at high temperatures) and are known as thermozymes. In view of these important advantages, thermozymes are attracting much industrial interest. The structure determination of thermostable enzymes and their complex with substrate or inhibitor, and investigation of the structural feature of stability will illuminate the structure base of stability and catalytic mechanisum at high temperature. This knowledge can lead to the development of new and more efficient protein engineering strategies and a wide range of biotechnological applications.
     Cellulose, the most abundant organic source of feed, fuel and chemicals consists of glucose units linked byβ-1,4-glycosidic bonds in a linear mode. The difference in the type of bond and the highly ordered crystalline form of the compound between starch and cellulose make cellulose more resistant to digest and hydrolyze. The enzymes required for the hydrolysis of cellulose include endoglucanases, exoglucanases andβ-glucosidases. While cellulase is an endoglucanase that hydrolyses cellulose randomly, producing oligosacaccharides, cellobiose and glucose, exoglucanases hydrolyzeβ-1,4-D-glucosidic linkages in cellulose releasing cellobiose from the non-reducing end. On the other hand,β-glycosidases of thermophilic origin, which have received renewed attention in the pharmaceutical industry hydrolyze cellobiose to glucose.
     In the current industrial processes, cellulolytic enzymes are employed in the color extractions of juices, in detergents causing color brightening and softening, in the biostoning of jeans, in the pretreatment of biomass that contains cellulose to improve nutritional quality of forage and in the pretreatment of industrial wastes. But cellulases, which we have already used, are still unsatisfactory. For instance, the biopolishing process of cotton in the textile industry requires cellulase stable at high temperature close to 100℃. Presently used enzymes for this purpose, however, are active only at 50–55℃. Cellulase production is also found to be the most expensive step during ethanol production from cellulosic biomass, and accounted for approximately 40% of the total cost. So, thermophilic cellulases with high activity and high stability are required to attack the native crystalline cellulose, which is water insoluble and occurs as fibers of densely packed structures.
     In our previous study, the thermophilic endo-β-1,4-glucanase FnCel5A was first cloned from Fervidobacterium nodosum Rt17-B1 and overexpressed in E. coli. The optimum temperature and pH of FnCel5A were 80℃and 5.5-5.8, respectively. The recombinant enzyme had a high endo-β-1,4-glucan activity. The enzyme was high thermal active with t1/2 was 48 h at 80℃, and can be thermal activated by 40% at 70℃、75℃、80℃for 2 h. The activity of FnCel5A was totally inhibited by Zn2+ (relative remain activity 3 %), and partially inhibited by Ni2+, Co2+and K+ (relative remain activity 73%, 79% and 88%). Structure determination of thermostable cellulases and their complex with substrate or inhibitor, and investigation of the structural feature of stability will illuminate the structure base of stability and catalytic mechanisum at high temperature. This knowledge not only can lead to find the new function of the enzyme, but also can direct molecular design and modification of present cellulases. It will accelerate the development of new and more efficient protein engineering strategies and start a wide range of biotechnological applications.
     In the first part of my thesis, the thermophilic endo-β-1,4-glucanase FnCel5A and the N-terminal truncated mutant FnCel5AND were cloned, overexpressed and purified from Fervidobacterium nodosum Rt17-B1. The recombinant has been crystallized and the crystal structure of FnCel5A and complex FnCel5AND:Glucose have been determined by Singlewavelenth Anamolous Dispersion and molecular replacement.to 1.7 ? and 2.5 ? respectively. Crystal structure shows that FnCel5A presents the (β/α)8-barrel structure typical of clan GH-5 of glycoside hydrolase families. Compare with the homologous from mesophiles and psychrophiles, it suggests that FnCel5A owns the structure feature of thermostable enzyme in amino acid composition and intermolecular interaction. Structure analysis of the complex FnCel5AND:Glucose lead us to make sure that Asn51, Trp61, His126, His127, Asn166, Glu167, Trp204, His226, Tyr228, Phe231, His235, Trp240, Glu283, Trp316 and Phe322 are important residues for catalytic function and substrate binding. Combined with the molecular simulations, the model of substrate binding was established. The model described the catalytic mechanisum of the straight chain substrate of glycosyl in detail.
     Termini of proteins are usually the regions with highest thermal factors in a protein structure. They are likely to unfold first during thermal denaturation. Many proteins whose stability are modified by the N-terminal structure and find that extra extension, deletion of the terminus and disruption of the interaction between the terminus and the structure will alter the stability of the protein. Researchers speculate that some protein termini are the original nucleus of protein folding and the high thermostability is the result of the tight binding of the terminus with the protein. In our previous report, the APH gene from the thermophilic archaeon Aeropyrum pernix K1 has been cloned and overexpressed in Escherichia coli. The enzyme possesses both acylpeptide hydrolase and esterase activities. It is extremely thermostable. We determined the crystal structure feature of ApAPH, which include anα/βhydrolase domain, and aβ-propeller domain. The N-terminus of the POP family provides the outermost strand, which extends from theβ-propeller domain and packs against theα/βhydrolase domain. Homology analysis by sequence and structure reveals that ApAPH has the shortest N-terminal extension, which raised an intriguing question about the function of the most conserved N-terminus on protein stability and catalysis. Moreover, analysis of the ApAPH structure reveals that N-terminus anchoring is achieved through ion pairing and hydrophobic interactions. Since the N-terminal helix of APH is conserved in all members of the POP family, we speculate that this region may play an important role in either function or stability, or both.
     To investigate the thermostable mechanism of N-terminus, ApAPH and N-terminal truncated mutant ApAPH-ΔN21 were cloned, overexpressed and purified in the second part of the thesis. The recombinant ApAPH-ΔN21 has been crystallized and the crystal structure has been determined by molecular replacement. The main-chain conformations of them are very similar, except for theα-helix at the N-terminus. The root mean square (r.m.s.) deviation of all the main-chain atoms between the two structures is 0.954 ?. It means that there is no significant change between two structures.
     The character of enzymatic activity and stability were determined by enzymatic assay, spectroscopy and structural analysis. Compare with the wild type, the thermostable mechanism of N-terminus of ApAPH was elucidated.
     (1) The N-terminal helix is important for the enzyme's thermophilic function.
     Although apAPH shows high structure similarity with several members of the POP family, in particular, on their secondary structures, these enzymes have diverse N-terminal extensions. Previously, it has been suggested that a tight binding of the N-terminal region to the catalytic domain in hyperthermophiles is one of the mechanisms for their high thermostability. apAPH possesses the shortest N terminal extension among all known POP family members and the secondary structure element in this region is conserved among all members of the POP family. The purpose of this paper is to examine the role of the apAPH N-terminal extension in stability and enzymatic activity. As expected, deletion of the N-terminal region leads to the shift of the optimal temperature to a lower value by ~18°C, confirming that the N-terminal helix region is critical for the hyperthermophilic function of the enzyme. Since thermophilic archaea is at the bottom of the evolution tree, we anticipate that this terminal region may play an important role in all POP family members.
     (2) The salt bridges have minor effects on the stability of the enzyme.
     Analysis of apAPH crystal structure reveals that the two salt bridges Asp15-Arg355 and Arg18-Asp325 are involved in connecting the N-terminal extension to the catalytic domain at the surface of the protein. Further examination of the crystal structures of the members in the POP family reveals that the two salt bridges also exist in porcine POP numbering (R60-D695 and K64-E691). Based on our results, disruption of these highly conserved salt bridges has only minor effects on the conformation, stability, and catalytic properties of enzymes. The result illustrates that the lack of such salt bridges in the N terminal region of apAPH leads to only small changes in stability and catalytic function. In the case of mutant D15A/R18A, almost all the results about activity and stability of mutant D15A/R18A are closer to the wild type protein than other mutants. Molecular dynamic simulation reveals that two new salt bridges (Arg11-Asp325 and Arg14-Asp325) and more hydrophobic interactions are formed. Therefore the thermostability of the double mutant recovered. These results suggest that other factors such as hydrophobic interactions between the catalytic domain and the N-terminal region are likely to be responsible for the stabilization of the WT enzyme.
     (3) Deletion of N-helix increases the conformation flexibility and enzymatic activity at low temperature.
     The accumulation of crystallographic data on thermostable and hyperthermostable enzymes, together with statistic analyses permitted by the rapid advancement of genome sequencing projects, has substantially improved our understanding of protein stability determinants. Academic curiosity is now tackling the issue of the relationship among stability/activity/flexibility of thermophilic proteins. The main question is whether in enzymes stability and activity are strictly related or not. The answer is not clear-cut. An interesting observation from our experiments is that the enzymatic activity of theΔN21 mutant is higher than that of the WT enzyme at low temperatures. Previously, we have determined the crystal structure of the mutantΔN21 mutant (2QZP).We found that the B factors of the mutant are significantly higher than those of the wild type enzyme in most parts of the structure; the averaged B factors rise from19.4 ?2 for the wild type to 28.4 ?2 for the mutantΔN21. The analysis of the structural features of both wild type and the truncated mutant reveals that there is a decrease in hydrogen bonds and salt bridges in both domains of the mutant, although there are no significant changes in the static structures between the wild type and theΔN21 mutant. It is likely that the increased flexibility of the active site, which is located at the interface of the two domains, is caused by the disruption of ionic interactions and hydrogen bonds. The conformation flexibility of the wild type and theΔN21 mutant was also investigated by the FIRST method. Compared with the wild type, the mutant ApAPH-ΔN21 exhibited flexible changes, including propeller I, II, III and residues 497–502 and 553–567. Especially, N-terminal truncation destroyed the original secondary structure of residues 559–567, replaced with high flexible loops which are close to the active site. Such flexibility changes can contribute to the high activity at the low temperature range, but is also linked to a reduced stability of the molecular edifice.
引文
1. Rothschild L J, Manicinelli R L. Life in extreme environments, Nature, 2001, 409: 1092-101.
    2. Vieille C, Zeikus G J. Hyperthermophilic enzymes: sources, uses, and molecular mechanisms for thermostability. Microbiol Mol Biol Rev 2001, 65: 1–43.
    3.和致中,彭谦,陈俊莫,高温菌生物学,北京:科学教育出版社,2000,60-65。
    4. Kato C, Inoue A, Horikoshi K. Isolating and characterizing deep-sea marine microorganisms. Trends Biotechnol. 1996, 14 (1): 6-12.
    5. Spain J C. Biodegradation of nitroaromatic compounds. Annu Rev Microbiol. 1995, 49: 523-55.
    6. Adams M W, Perler F B, Kelly R M. Extremozymes: expanding the limits of biocatalysis. Biotechnology (N Y). 1995, 13 (7): 662-8.
    7. Blochl E, Rachel R, Burggraf S, Hafenbrandl D, Jannasch H W, Stetter K O. Pyrolobus fumarii, gen. and sp. nov. represents a novel group of Archaea extending the upper temperature limit for life to 113°C. Extemophiles. 1997, 1: 14-21.
    8.刘志恒主编,现代微生物学,北京:科学出版社,2002,476-477。
    9. Vieille C, Burdette D S, Zeikus J G. Thermozymes. Biotechnol Annu Rev 1996, 2: 1–83.
    10. Becker P, Abu-Reesh I, Markossian S, Antranikian G, Markl H. Determination of the kinetic parameters during continuous cultivation of the lipase producing thermophile Bacillus sp IHI-91 on olive oil. Appl Microbiol Biotechnol 1997, 48: 184–90.
    11. Fields P A. Protein function at thermal extremes: balancing stability and flexibility. Comp Biochem Physiol, Part A Mol Integr Physiol 2001, 129: 417–31.
    12. Scandurra R, Consalvi V, Chiaraluce R, Politi L, Engel P C. Protein thermostability in extremophiles. Biochimie 1998, 80: 933–41.
    13. Fontana A. Analysis and modulation of protein stability. Curr Opin Biotechnol 1991, 2: 551–60.
    14. Ladenstein R, Antranikian G. Proteins from hyperthermophiles: stability and enzymatic catalysis close to the boiling point of water. Adv Biochem Eng Biotechnol 1998, 61: 37–85.
    15. Haney P J, Badger J H, Buldak G L, Reich C I, Woese C R and Olsen G J. Thermal adaptation analyzed by comparison of protein sequences from mesophilic and extremely thermophilic Methanococcus species, Proc. Natl. Acad. Sci. US , 1999, 96: 3578–3583.
    16. Cambillau C, and Claverie J M. Structural and genomic correlates of hyperthermostability, J. Biol. Chem., 2000, 275: 32383–32386.
    17. Chakravarty S and Varadarajan R. Elucidation of determinants of protein stability through genome sequence analysis, FEBS Lett., 2000, 470: 65–69.
    18. Thompson M J and Eisenberg D. Transproteomic evidence of a loop-deletion mechanism for enhancing protein thermostability, J. Mol. Biol., 1999, 290: 595–604.
    19. Reinhard S and Wolfgang L. Thermophilic adaptation of proteins, Critical Reviews in Biochemistry and Molecular Miology, 2001, 36: 39-106.
    20. Smith D R, Stamm L A, Deloughery C, Lee H M, Dubois J, Aldredge T, Bashirzadeh R, Blakely D, Cook R, Gilbert K. Complete genome sequence of Methanobacterium thermo- autotrophicum delta H: functional analysis and comparative genomics. J Bacteriol. 1997, 179: 7135-7155.
    21. Cacciapuoti G, Fusco S, Caiazzo N, Zappia V, Porcelli M. Hetero- logous expression of
    59-methylthioadenosine phosphorylase from the archaeon Sulfolobus solfataricus: characterization of the recombinant protein and involvement of disulfide bonds in thermo- philicity and thermostability. Protein Expression Purif. 1999, 16: 125–13.
    22. Schultes V, Deutzmann R, Jaenicke R. Complete amino-acid sequence of glyceraldehyde-3-phosphate dehydrogenase from the hyperthermophilic eubacterium Thermotoga maritima. Eur. J. Biochem. 1990, 192: 25–35.
    23. Vogt G, Woell S, Argos P. Protein thermal stability, hydrogen bonds, and ion pairs. J Mol Biol 1997, 269: 631–43.
    24. Suvd D, Fujimoto Z, Takase K, Matsumura M, Mizuno H. Crystal structure of Bacillus stearothermophilus aamylase: possible factors determining the thermostability. J Biochem 2001, 129: 461–8.
    25. Eijsink V G, Vriend G, Van der Zee J R, Van den Burg B, Venema G. Increasing the thermostability of the neutral proteinase of Bacillus stearothermophilus by improvement of internal hydrogen-bonding. Biochem J 1992b, 285: 625–8.
    26. Tanner J J, Hecht R M, Krause K L. Determinants of enzyme thermostability observed in the molecular structure of Thermus aquaticus d-glyceraldehyde-3-phosphate dehydrogenase at 2.5 ? resolution. Biochemistry 1996, 35: 2597–609.
    27. Ribeiro S M, Darimont B, Sterner R, Huber R. Small structural changes account for the high thermostability of 1[4Fe-4S] ferredoxin from the hyper thermophilic bacterium Thermotoga maritima. Structure. 1996, 4: 1290–1301.
    28. Dill K A. Dominant forces in protein folding. Biochemistry1990, 29: 7133–7155.
    29. Perutz M F. Electrostatic effects in proteins. Science. 1978, 201: 1187–1191.
    30. Alsop E, Silver M, Livesay D R. Optimized electrostatic surfaces parallel increased thermostability: a structural bioinformatic analysis. Protein Eng 2003, 16: 871–4.
    31. Hakamada Y, Hatada Y, Ozawa T, Ozaki K, Kobayashi T, Ito S. Identification of thermostabilizing residues in a Bacillus alkaline cellulase by construction of chimeras from mesophilic and thermostable enzymes and sitedirected mutagenesis. FEMS Microbiol Lett 2001, 195: 67–72.
    32. Karshikoff A, Ladenstein R. Ion pairs and the thermotolerance of proteins from hyperthermophiles: a“traffic rule”for hot roads. Trends Biochem Sci 2001, 26: 550–6.
    33. Das R, Gerstein M. The stability of thermophilic proteins: a study based on comprehensive genome comparison. Funct Integr Genomics 2000, 1: 76–88.
    34. Goodenough P W, Jenkins J A. Protein engineering to change thermal stability for food enzymes. Biochem Soc Trans 1991, 19: 655–62.
    35. Pace C N. Contribution of the hydrophobic effect to globular protein stability. J Mol Biol 1992, 226: 29–35.
    36. Kirino H, Aoki M, Aoshima M, Hayashi Y, Ohba M, Yamagishi A, Wakagi T, Oshima T. Hydrophobic interaction at the subunit interface contributes to the thermostability of 3-isopropylmalate dehydrogenase from an extreme thermophile, Thermus thermophilus. Eur. J. Biochem. 1994, 220: 275–281.
    37. Matsumura M, Signor G, Matthews B W. Substantial increase of protein stability by multiple disulphide bonds. Nature 1989, 342: 291– 3.
    38. Takagi H, Takahashi T, Momose H, Inouye M, Maeda Y, Matsuzawa H, et al. Enhancement of the thermostability of subtilisin E by introduction of a disulfide bond engineered on the basis of structural comparison with a thermophilic serine protease. J Biol Chem 1990, 265: 6874–8.
    39. Li W T, Grayling R A, Sandman K, Edmondson S, Shriver J W, Reeve J N. Thermodynamic stability of archaeal histones. Biochemist 1998, 37: 10563–72.
    40. Choi I G, Bang W G, Kim S H, Yu Y G. Extremely thermostable serine-type protease from Aquifex pyrophilus. Molecular cloning, expression, and characterization. J. Biol. Chem. 1999, 274: 881–888.
    41. Burley S K, Petsko G A. Aromatic-aromatic interaction: a mechanism of protein structure stabilization. Science. 1985, 229: 23–28.
    42. Teplyakov A V, Kuranova I P, Harutyunyan E H, Vainshtein B K, Frommel C, Hohne W E and Wilson K S. Crystal structure of thermitase at 1.4 ? resolution. J. Mol. Biol. 1990, 214: 261–279.
    43. Serrano L, Bycroft M and Fersht A R. Aromatic-aromatic interactions and protein stability. Investigation by double-mutant cycles. J. Mol. Biol. 1991, 218: 465–475.
    44. Marg G A and Clark D S. Activation of the glucose isomerase by divalent cations: evidence for two distinct metal-binding sites. Enzyme Microb. Technol. 1990, 12: 367–373.
    45. Vieille C, Epting K L, Kelly R M, Zeikus J G. Bivalent cations and amino-acid composition contribute to the thermostability of Bacillus licheniformis xylose isomerase. Eur J Biochem 2001, 268: 6291–301.
    46. Radfar R, Leaphart A, Brewer J M, Minor W, Odom J D, Dunlap R B, et al. Cation binding and thermostability of FTHFS monovalent cation binding sites and thermostability of N10-formyltetrahydrofolate synthetase from Moorella thermoacetica. Biochemist 2000, 39: 14481–6.
    47. Bonisch H, Backmann J, Kath T, Naumann D, Schfer G. Adenylate kinase from Sulfolobus acidocaldarius: expression in Escherichia coli and characterization by Fourier transform infrared spectroscopy. Arch Biochem Biophys 1996, 333: 75–84.
    48. Leone M, Di Lello P, Ohlenschlager O, Pedone E M, Bartolucci S, Rossi M, et al. Solution structure and backbone dynamics of the K18G/R82E Alicyclobacillus acidocaldarius thioredoxin mutant: a molecular analysis of its reduced thermal stability. Biochemist 2004, 43: 6043–58.
    49. Russell R J M, Ferguson J M, Hough D W, Danson M J, Taylor G L. The crystal structure of citrate synthase from the hyperthermophilic Archaeon Pyrococcus furiosus at 19 angstrom resolution. Biochemist 1997, 36: 9983–94.
    50. Knapp S, de Vos W M, Rice D, Ladenstein R. Crystal structure of glutamate dehydrogenase from the hyperthermophilic eubacterium Thermotoga maritima at 3.0 ? resolution. J Mol Biol 1997, 267: 916–32.
    51. Kirino H, Aoki M, Aoshima M, Hayashi Y, Ohba M, Yamagishi A, Wakagi T, and Oshima T. Hydrophobic interaction at the subunit interface contributes to the thermostability of 3-isopropylmalate dehydrogenase from an extreme thermophile, Thermus thermophilus. Eur. J. Biochem. 1994, 220: 275–281.
    52. Moriyama H, Onodera K, Sakurai M, Tanaka N, Kirino-Kagawa H, Oshima T and Katsube Y. The crystal structures of mutated 3-isopropylmalate dehydrogenase from Thermus thermophilus HB8 and their relationship to the thermostability of the enzyme. J. Biochem. 1995, 117: 408–413.
    53. Thoma R, Hennig M, Sterner R and Kirschner K. Structure and function of mutationally generated monomers of dimeric phosphor- ribosylanthranilate isomerase from Thermotoga maritime. Structure. 2000, 8: 265–276.
    54. Matthews B W, Nicholson H, Becktel W J. Enhanced protein thermostability from site-directed mutations that decrease the entropy of unfolding. Proc. Natl. Acad. Sci. USA 1987, 84: 6663–6667.
    55. Eijsink V G, Vriend G, Hardy F, Veltman O R, Van der Vinne B, Van der Burg B, Dijkstra B W, Van der Zee J R and Venema G. Structural determinants of the thermostability of thermolysin-like Bacillus neutral proteases, p. 91–99.
    56. Takano K, Yamagata Y, Yutani K. Role of amino acid residues in left-handed helical conformation for the conformational stability of a protein. Proteins 2001a, 45: 274–80.
    57. Facchiano A M, Colonna G, Ragone R. Helix stabilizing factors and stabilization of thermophilic proteins: an X-ray based study. Protein Eng 1998, 11: 753– 60.
    58. Hol W G. The role of the a-helix dipole in protein function and structure. Prog Biophys Mol Biol 1985, 45: 149–95.
    59. Eijsink V G, Vriend G, van den Burg B, van der Zee J R, Venema G. Increasing the thermostability of a neutral protease by replacing positively charged amino acids in the N-terminal turn of a-helices. Protein Eng 1992a, 5: 165–70.
    60. Sheridan R P, Levy R M, Salemme F R. a-Helix dipole model and electrostatic stabilization of 4-a-helical proteins. Proc Natl Acad Sci 1982, 79: 4545–9.
    61. Vieille C, Zeikus G J. Hyperthermophilic Enzymes: Sources, Uses, and Molecular Mechanisms for Thermostability .Microbiol. Mol. Biol. Rev., Mar 2001; 65: 1 - 43.
    62. Daniel R M, Dines M, Petach H H. The denaturation and degradation of stable enzymes at high temperatures. Biochem J. 1996, 1; 317 ( Pt 1): 1-11.
    63.徐恒平,张树政嗜极菌的极端酶生物工程进展1997, 17(1): 2-5.
    64.王柏婧,冯雁,王师钰等嗜热酶及其应用微生物学报2002, 42 (2): 259-262.
    65.林影,卢嶸德极端酶及其工业应用工业微生物2000, 30(2):51-53
    66. Haki G D, Rakshit S K. Developments in industrially important thermostable enzymes: a review. Bioresour Technol. 2003, 89(1): 17-34.
    67. Khandke K M, Vithayathil P J, Murthy S K. Degradation of larchwood xylan by enzymes of a thermophilic fungus, Thermoascus aurantiacus. Arch Biochem Biophys. 1989, 274 (2): 501-10.
    68. Bartolucci S, Rella R, Guagliardi A, Raia C A, Gambacorta A, Rosa, M D, Rossi M. Malic enzyme from archaebacterium Sulfolobus solfataricus. Purification, structure, and kinetic properties.J. Biol. Chem.,1987, 262: 7725-7731.
    69. Dong F M, Wang L L, Wang C M, Cheng J P, He Z Q, Sheng Z J, Shen R Q. Molecular cloning and mapping of phenol degradation genes from Bacillus stearothermophilus FDTP-3 and their expression in Escherichia coli. Appl. Envir. Microbiol.1992, 58: 2531-2535.
    70. Gokcay C F and Yurteri R N. Fuel. 1983, 1223.
    71. Cutchen M, et al. Biotechnol. Bioent.1996, 52: 332-339.
    72. Demirbas A, Biomass resource facilities and biomass processing for Fuels and chemicals, Energy Conversion and management, 2001, 4(2): 1357-1378.
    73. Richmond P A, In Biosynthesis and Biodegradation of Cellulose, New York Marcel Dekker, 1991: 5-23
    74. David N S H. Polymer NEWS 1988, 13: 134-140.
    75. Katie Heaton, Industrial Uses of Thermophilic Cellulase Protein Structure and Function, 2004 Spring: 645.
    76.陈国符,邬义明,《植物纤维化学》北京:轻工业出版社,1991.
    77.谢占玲,吴润,纤维素酶的研究进展,草业学,2004,21 (4): 72-76.
    78. Zhang Y H P, Himmel Mielenz J R. Outlook for cellulase improvement: screening and selection strategies. Biotechnol Advan, 2006, 4: 452-481.
    79. Wood T M, Properties and mode of action of cellulose, Biochem. Soc.Trans. 1985, 3: 407-410.
    80. Wood T M, Fungalcellulases, Biochem. Soc. Trans, 1992, 20: 45-53.
    81. Teeri T T, Crystalline cellulose degradation: new insight into the function of cellobiohydrolas, Biotechnol Trends, 1997, 15(5): 160-167.
    82. Henrissat B, Claeyssens M, Tomme P, et al. Cellulase families revealed by hydrophobic cluster analysis, Gene, 1989, 83-95.
    83. Henrissat B. A classification of glycosyl hydrolases based on amino acid sequence similarities. Biochem,1991, J 280: 309-316.
    84. Henrissat B, Davies G. Structure and sequence–based classify cation of glycoside hydrolases, Curr Opin Struct Bio, 1997, l7: 637-644.
    85.博士学位论文:兔子盲肠未培养微生物纤维素酶基因的克隆﹑表达及产物的酶学特性研究,第6页.
    86.阎伯旭,孙迎庆,高培基,拟康氏木霉纤维素酶分子研究其结构域的结构与功能,纤维素科学与技术,1998,6(30): 1291.
    87. Zhang Y H P, Lynd L R. Toward an Aggregated Understanding of Enzymatic Hydrolysis of Cellulose: Noncomplexed Cellulase Systems, Biotechnology and Bioengineering, 2004, 88 (7): 797- 824.
    88. Rye C S, Withers S G., Glycosidases mechanisms, Curr Opin Chem Biol, 2000, 4: 573-580.
    89. Ducros V, Czjzek M, Belaich A, et al. Crystal structure of the catalytic domain of a bacterial cellulase belonging to family 5, Structure, 1995, 3: 939-949.
    90. Sinnott M L, Catalytic mechanisms of enzymatic glycosyl trfanser, Chem Rev, 1990, 90: 1171-1202.
    91. McCarter J D, Withers S G. Mechanisms of enzymatic glucoside hydrolysis. Curr Opin Struct Biol, 1994, 4: 885-892.
    92. Ly H D, Withers S G. Mutagenesis of glycosidases, Annu Rev Biochem, 1999, 68: 487-522.
    93. Davies G J. Structural studies on cellulases, Biochem.Soc.Trans, 1998, 26: 167-173.
    94. Tillbeurgh M, Tomme P, CLAETSSENS, et al. Limited proteoly sis of the cellobiohydrolase I from T.reesei, FEBS Lett, 1986: 223-227.
    95. Azevedo H, et al. Mictob Tech, 2000, 27: 325-329.
    96.周文龙,《酶在纺织中的应用》,中国纺织出版社,2001第1版.
    97. Henrissat B, Callebaut I, Fabrega S, Lehn P, Mornon J P & Davies G. Conserved catalytic machinery and the prediction of a common fold for several families of glycosyl hydrolases. Proc. Natl. Acad. Sci. USA 1995, 92, 7090-7094.
    98. Jenkins J, Lo Leggio L, Harris G & Pickersgill R,β-glucosidase,β-galactosidase, family A cellulases, family F xylanases and two barley glycanases form a superfamily with 8-foldβ/αarchitecture and with two conserved glutamates near the carboxy-terminal ends ofβ-strands four and seven. FEBS Lett. 1995, 362, 281-285.
    99. Warren R A J, Annual Review of Microbiology, Annual Reviews 1996, 50: 183-212.
    100. Lynd L R, Weimer P J, van Zyl W H, Pretorius I S, Microbial Cellulose Utilization: Fundamentals and Biotechnology, Microbiology and Molecular Biology Reviews, 2002, 66(3): 506–577.
    101. Ahsan M, Matsumoto M, Karita S, Kimura T, Sakka K, and Ohmiya K, Purification and characterization of the family J catalyticdomain derived from the Clostridium thermocellum endoglucanase CelJ, Biosci. Biotechnol. Biochem, 1997, 61: 427–431.
    102. Ait N, Creuzet N, Cattaneo J, Properties of glucosidase purified from Clostridium thermocellum. J. Gen. Microbiol, 1982, 128: 569–577.
    103. Kamrat T, Nidetzky B, Entrapment in E. coli improved the operational stability of recombinant-glucosidase CelB from Pyrococcus furiosus and facilitates biocatalyst recovery, Journal of Biotechnology, 2007, 129: 69-76.
    104. Bok J D, Yernool D A, Purification, characterization, and molecular analysis of thermostable cellulases CelA and CelB from Thermotoga neapolitana, Applied and Environmental Microbiology, 1998, 64: 4774-4781.
    105. Halldórsdóttir S, Thórólfsdóttir E, Cloning, sequencing and overexpression of a Rhodothermus marinus gene encoding a thermostable cellulase of glycosilhydrolase family 12, Applied Microbiology and Biotechnology, 1998, 49: 277-284.
    106. Huang Y, Krauss G, Cottaz S, Driguez H, Lipps G., A highly acid-stable and thermostable endo-β-glucanase from the thermoacidophilic archaeon Sulfolobus solfataricus, Biochem. J., 2005, 385: 581–588.
    107. Adney W S, Tucker M P, Grohmann K, Himmel M E, Thermostable purified endoglucanas from Acidothermus cellulolyticus ATCC 43068. United States Patent 5275944.
    108. Ando S, Ishida H, Kosugi Y, et al, Hyperthermostable Endoglucanase from Pyrococcus horikoshii, Appl Environ Microbiol, 2002, 68: 430-433.
    109. Kashima Y, Mori K, Fukada H et al, Analysis of the function of a hyperthermophilic endoglucanase from Pyrococcus horikoshii that hydrolyzes crystalline cellulose, Extremophilies, 2005, 9(1): 37-43.
    110. Feng Y, Duang C J, Cloning and identification of novel cellulase genes from uncultured microorganisms in rabbit cecum and characterization of the expressed cellulased, Appl Microbiol Biotechnol, 2006, l75: 319-328.
    111. Ni J f, Heterrologous, Overexpression of a Mutant Termite Cellulase Gene in Escherichia coli by DNA Shuffling of Four Orrhologous Parental cDNAs, Bioscl.Biotechnol Biochem, 2005, 69(9), 1711-1720.
    112. Wan T, Directed evolution for engineering pH profile of endoglucanase III from Trichoderma reesei Biomolecular Engineering, 2005, 22: 89-94.
    113.黎海彬,褐色高温单孢菌PE-211纤维素酶的纯化及固定化研究,现代化工,2007(1): 36.
    114.王顺民,董文宾,嗜热酶的研究及其在食品等相关工业的应用,自然杂志,2004,26(6): 339-342.
    115.刘颖,林亲录,纤维素酶制取与应用研究进展,中国食物与营养,2006,4: 33-35.
    116. Chhabra S R, Shockley K R, Ward D E and Kelly R M, Regulation of Endo-Acting Glycosyl Hydrolases in the Hyperthermophilic Bacterium Thermotoga maritima Grown on Glucan- and Mannan-Based Polysaccharides Applied and environmental microbiology, 2002, 68(2): 545–554.
    117. Isabelle Bortoli-German, Jacques Haiech, Marc Chippaux, Frederic Barras, Informational Suppression to Investigate Structural Functional and Evolutionary Aspects of the Erwinia chrysanthemi Cellulase EGZ J, Mol. Biol. 1995, 246: 82–94.
    118. Schneider T R and Sheldrick G M. Substructure solution with SHELXD. Acta Crystallogr D Biol Crystallogr, 2002, 58 (Pt 10 Pt 2): 1772-9.
    119. Terwilliger T C and Berendzen J. Automated MAD and MIR structure solution. Acta Crystallogr D Biol Crystallogr, 1999, 55 (Pt 4): 849-61.
    120. Terwilliger T C. Maximum-likelihood density modification. Acta Crystallogr D Biol Crystallogr, 2000, 56 (Pt 8): 965-72.
    121. Emsley P and Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr, 2004, 60 (Pt 12 Pt 1): 2126-32.
    122. Brunger A T, Adams P D, Clore G M et al. Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallogr D Biol Crystallogr, 1998, 54 (Pt 5): 905-21.
    123. Laskowski R A, MacArthur M W, Moss D S et al. PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Cryst. , 1993, 26: 283-291.
    [1] Venalainen J I, Juvonen R O, Mannisto P T. Evolutionary relationships of the prolyl oligopeptidase family enzymes. Eur J Biochem. 2004, 271: 2705-15.
    [2] Rawlings N D, Polgár L and Barrett A J. A new family of serine-type peptidases related to prolyl oligopeptidase Biochem. J.1991, 279: 907–908.
    [3] Kanatani A, Masuda T, Shimoda T, Misoka F, Lin X S, Yoshimoto T. Protease II from Escherichia coli: sequencing and expression of the enzyme gene and characterization of the expressed enzyme. J. Biochem. 1991, 110: 315–320.
    [4] Polgár L. Structural relationship between lipases and peptidases of the prolyl oligopeptidase family. FEBS Lett.1992, 311: 281–284.
    [5] Wilk S. Prolyl endopeptidase. Life Sci. 1983, 33:2149– 2157.
    [6] Mentlein R. Proline residues in the maturation and degradation of peptide-hormones and neuropeptides. FEBS Lett. 1988, 234: 251–2566.
    [7] Cunningham D F andóConnor B. Proline specific peptidases. Biochim. Biophys. Acta 1997, 1343: 160–186.
    [8] Maes M, Goossens F, ScharpéS, Meltzer H Y, ? Hondt P and Cosyns P. Lower serum prolyl endopeptidase enzyme activity in major depression: further evidence that pep-tidases play a role in the pathophysiology of depression. Biol.Psychiatry. 1994, 35: 545–552.
    [9] Maes M, Goossens F, Scharpe S, Calabrese J, Desnyder R and Meltzer H Y. Alteration in plasma prolyl oligopeptidase activity in depression, mania, and schizophrenia: effects of antidepressants, mood stabilizers and antipsychotic drugs. Psychiatry Res. 1995, 58: 217–225.
    [10] Maes M, Monteleone P, Bencivenga R, Goossens F, Maj M, van Wes D. Lower serum activity of prolyl oligopeptidase in anorexia and bulimia nervosa. Psychoneuroendocrinology 2001, 26: 17–26.
    [11] Wilk S and Orlowski M. Inhibition of rabbit brain prolyl endopeptidase by N-benzyloxycarbonyl-prolyl-prolinal, a transition state aldehyde inhibitor. J. Neurochem. 1983, 41: 69–75.
    [12] Stone S R, Rennex D, Wikstrom P, Shaw E and Hofsteenge J. Inactivation of prolyl endopeptidase by a peptidylchloromethane. Kinetics of inactivation and identification of sites of modification. Biochem. J. 1991, 276: 837–840.
    [13] Polgár L. Structural relationship between lipases and peptidases of the prolyl oligopeptidase family. FEBS Lett.1992, 311: 281–284.
    [14] Kahyaoglu A, Haghjoo M, Fusheng G, Jordan F, Kettner C and Felf?ldi F. Low-barrier hydrogen bond is absent in the catalytic triads in the ground state but is present in a transition state complex in prolyl oligopeptidase family of serine proteases. J. Biol. Chem. 1997, 272: 25547–25554.
    [15] Ollis D I, Cheah E, Cygler M, Dijkstra B, Frolow F, Franken S M. Theα/βhydrolase fold. Protein Eng. 1992, 5: 197–211.
    [16] Polgár L. Structure and function of serine proteases. New Comp. Biochem. 1987, 16: 159–200.
    [17] Polgár L. Mechanisms of Protease Action, CRC Press 1989, 6: 87–122.
    [18] Polgár L. Structure and function of serine proteases. New Comp. Biochem. 1987, 16: 159–200.
    [19] Perona J J and Craik C S. Structural basis of substrate specificity in the serine proteases. Protein Sci. 1995, 4: 337–360.
    [20]孙之荣,王钰,胡胜民,郭青,贺福初.丝氨酸蛋白酶超家族分子结构进化研究.生物物理学报, 1999, 15: 530-535.
    [21] Fül?p V, Szeltner Z and Polgár L. Catalysis of serine oligopeptidases is controlled by a gating filter mechanism. EMBO Reports 2000, 1: 277–281.
    [22] Jones W M, Manning L R and Manning J M. Enzymic cleavage of the blocked amino terminal residues of peptides. Biochem. Biophys. Res. Commun. 1986. 139: 244–250.
    [23] Jones W M, Scaloni A, Bossa F, Popowicz A M, Schneewind O and Manning J M. Genetic relationship between acylpeptide hydrolase and acylase, two hydrolytic enzymes with similar binding but different catalytic specificities. Proc. Natl. Acad. Sci. USA 1991, 88: 2194–2198.
    [24] Farries T C, Harris A, Auffret A D and Aitken A. Removal of N-acetyl groups from blocked peptides with acylpeptide hydrolase. Stabilization of the enzyme and its application to protein sequencing. Eur. J. Biochem. 1991, 196: 679–685.
    [25] Gade W and Brown J L. Purification and partial characterization of a-N-acylpeptide hydrolase from bovine liver. J. Biol. Chem. 1978, 253: 5012–5018.
    [26] Radhakrishna G and Wold F. Purification and characterization of a-N-acylpeptide hydrolase from rabbit muscle. J. Biol. Chem. 1989, 264: 11076–11081.
    [27] Sharma K K and Ortwerth B J. Bovine lens acylpeptide hydrolase. Purification and characterization of a tetrameric enzyme resistant to urea denaturation and proteolytic inactivation. Eur. J. Biochem. 1993, 216: 631–637.
    [28] Raphel V, Giardina T, Guevel L, Perrier J, Dupuis L, Gou X J.Cloning, sequencing and further characterization of acylpeptide hydrolase from porcine intestinal mucosa. Biochim. Biophys. Acta 1999, 1432: 371–381.
    [29] Scaloni A, Jones W M, Pospischil M, Schneewind O, Popowicz A, Bossa F. Deficiency of acylpeptide hydrolase in small-cell lung carcinoma cell lines. J. Lab. Clin. Med 1992, 120: 546–552.
    [30] Mitta M, Asada K, Uchimura Y, Kimizuka F, Kato I, Sakiyama F. The primary structure of porcine liver acylamino acid-releasing enzyme deduced from cDNA sequences. J. Biochem. 1989, 106: 548–551.
    [31] Kobayashi K, Lin L, Yeadon J, Klickstein L and Smith J. Cloning and sequence analysis of a rat liver cDNA encoding acyl-peptide hydrolase. J. Biol. Chem. 1989, 264: 8892–8899.
    [32] Scaloni A, Jones W M, Barra D, Pospischil M, Sassa S, Popowicz A. Acylpeptide hydrolase: inhibitors and some active site residues of the human enzyme. J. Biol.Chem. 1992, 267: 3811–3818.
    [33] Mitta M, Miyagi M, Kato I and Tsunasawa S. Identification of the catalytic triad residues of porcine liver acylamino acid-releasing enzyme. J. Biochem. 1998, 123: 924–931.
    [34] Feese M, Scaloni A, Jones W M, Manning J M and Remington S J. Crystallization and preliminary X-ray studies of human erythrocyte acylpeptide hydrolase. J. Mol. Biol. 1993, 223: 546–549.
    [35] Ishikawa K, Ishida H, Koyama Y, Kawarabayasi Y, Kawahara J, Matsui E. Acylamino acid-releasing enzyme from the thermophylic archaeon Pyrococcus horikoshii. J. Biol. Chem. 1998, 273: 17726–17731.
    [36] Jones W M, Scaloni A, Bossa F, Popowicz A M, Schneewind O and Manning J M. Genetic relationship between acylpeptide hydrolase and acylase, two hydrolytic enzymes with similar binding but different catalytic specificities. Proc. Natl. Acad. Sci. USA 1991, 88: 2194–2198.
    [37] Naylor S L, Marshall A, Hensel C, Martinez P F, Holley B and Sakaguchi A Y. The DNF15S2 locus at 3p21 is transcribed in normal lung small cell lung cancer. Genomics 1989, 4: 355–361.
    [38] Erlandsson R, Boldog F, Persson B, Zaborovsky E R, Allikmets R L, Sumegi J. The gene from the short arm of chromosome 3, at D3F15S2, frequently deleted in renal cell carcinoma, encodes acylpeptide hydrolase. Oncogene 1991, 6: 1293–1295.
    [39] Scaloni A, Jones W M, Pospischil M, Schneewind O, Popowicz A, Bossa F. Deficiency of acylpeptide hydrolase in small-cell lung carcinoma cell lines. J. Lab. Clin.Med. 1992, 120: 546–552.
    [40] Bartlam M, Wang G G, Yang H, Gao R J, Zhao X D, Xie G Q, Cao S G, Feng Y, Rao Z H. Crystal structure of an acylpeptide hydrolase/esterase from Aeropyrum pernix K1, Structure 2004, 12: 1481–1488.
    [41] Gao R J, Feng Y, Ishikaw K, Ishida H, Andob S, Kosugi Y, Cao S G. Cloning, purification and properties of a hyperthermophilic esterase from archaeon Aeropyrum pernix K1, Journal of Molecular Catalysis B: Enzymatic 2003, 24–25: 1–8.
    [42] Zhang Z M, Zheng B S, Wang Y P, Chen Y Q, Manco G, Feng Y. The conserved N-terminal helix of acylpeptide hydrolase from archaeon Aeropyrum pernix K1 is important for its hyperthermophilic activity, Biochimica et Biophysica Acta (BBA) - Proteins & Proteomics, 2008, 1784(9): 1176-1183.

© 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