• 沒有找到結果。

Catechol derivatives oxidative activity of copper-substituted PepD (CuCu-PepD)

Chapter 3. Results and Discussion

3.6 Catechol derivatives oxidative activity of copper-substituted PepD (CuCu-PepD)

In our previous study, the hydrolysis activity of PepD was found to be inhibited by dopamine or L-dopa. Interestingly, the peptide-bond hydrolysis activity of PepD can be converted to catechol oxidative activity by substitution of zinc ions with copper ions in active site (CuCu-PepD). However, even though CuCu-PepD can oxidize catechol derivatives, which are catecholamine hormones produced from tyrosine metabolism (dopamine, L-dopa, epinephrine and norepinephrine) (Fig. 3.13), it is unable to oxidize catechol or 3,5-di-tert-butylcatechol (DTC), which has distinguished by their non-polar side chains (Fig. 3.14). This result indicated that CuCu-PepD specifically reacts with only substrates having a polar tail. No other enzyme to date, to our knowledge, has been reported to have such a property of specialized substrate selectivity.

Furthermore, enzyme kinetics analysis revealed similar catalytic efficiency (kcat/Km) of CuCu-PepD for several catecholamine hormones. CuCu-PepD oxidation of epinephrine was more efficient than that of other related substrates (Table 3.4). This observation likely reflects the good binding that exists between enzyme and epinephrine (Km = 0.073 mM). Comparison of the enzyme kinetics between CuCu-PepD and copper-substituted SgAP (CuCu-SgAP), as reported by Ming et al.72, indicated that CuCu-PepD exhibited better substrate binding affinity for dopamine (Km = 0.6 mM) and DTC (Km = 0.44 mM) than did CuCu-SgAP. However, the low kcat of CuCu-PepD for several catechol derivatives leads to decreased oxidative efficiency (Table 3.4).

67

Figure 3.13. Catecholamine hormones produced from tyrosine metabolism.

Figure 3.14. Structures of catechol and its derivatives. The polar tail of catechol derivatives are indicated by red color.

68

Table 3.4: Comparison of enzyme kinetics of PepD, CuCu-PepD, and CuCu-SgAP*

for different substrates. *Referred from Ming et al.72 (gray background).

3.7 Protein-ligand docking between CuCu-PepD and catecholamine hormones (L-dopa, dopamine, epinephrine, and norepinephrine)

In the previous report by Ming et al., the active site of CuCu-SgAP provides the hydrophobic pocket for specific recognition in binding of DTC via its di-tert-butyl groups. However, CuCu-PepD was unable to oxidize catechol or DTC, but it could oxidize the catechol derivatives with polar tail. In order to understand the binding network for each catecholamine hormones with CuCu-PepD, a protein-ligand docking strategy was applied to investigate the molecular interaction that occurred between enzyme and substrate. Based on the results obtained, dopamine, L-dopa, epinephrine, and norepinephrine could be bound by CuCu-PepD in the enzyme’s metal ions binding center. The docking model of the CuCu-PepD-dopamine complex indicated that ortho-hydroxyl group of dopamine could be bound by Asp119, Glu149, Asp173 and His461, and the amine group could be bound by the Glu175 side chain via hydrogen bonding (Fig.

3.15A). In addition, the binding network of L-dopa with CuCu-PepD was found to be similar to that for dopamine (Fig. 3.15B). Since the amine group of L-dopa is distant

69

from Glu175, the combination between the hydroxyl group of Tyr181 and the carboxylate group of L-dopa is more likely (Fig. 3.15B). When the binding networks between epinephrine and norepinephrine were compared, they were found to be similar in that hydroxyl groups of both substrates were bound by Glu175 at a distance of 2.7 Å (Fig.

3.15 C and D). However, the end-methyl group of epinephrine was able to further interaction via the hydrophobic pocket produced by Ile432 and Ala434. Therefore, the protein-ligand docking results indicated that the better catalytic efficiency of CuCu-PepD for epinephrine was a result of epinephrine fitting better into the PepD active site.

Figure 3.15. Protein-ligand docking of CuCu-PepD with (A) dopamine, (B) L-dopa, (C) epinephrine, and (D) norepinephrine. A local view of the di-metal ions center of PepD is presented for each. Residues are shown as sticks and labeled accordingly. The copper ions of CuCu-PepD are depicted by two light blue spheres.

70

CHAPTER 4

Conclusions and Future Perspectives

In conclusion, despite the lack of detectable sequence homology, the PepD enzyme has clear structural homology to other di-zinc-dependent M20 and M28 family enzymes. The crystal structure of V. alginolyticus PepD reveals it to be a dimer with two domains in each subunit. The catalytic domain of PepD contains two zinc ions and is structurally homologous to other proteolytic enzymes with dinuclear zinc catalytic sites.

The lid domain, on the other hand, is structurally homologous to that of PepV, but having a distinct topology of β-sheet order. Interestingly, part of the lid domain of the PepD structure is also homologous to the lid domain of the dimeric proteins.

Nevertheless, PepV exists as a monomer, while the PepD and related di-zinc-dependent M20/M28 family of enzymes are determined to be dimers. Structural comparisons between PepD and related di-zinc-dependent metallopeptidases suggest that formation of the catalytically competent active site in the PepD family of enzymes may be associated with transition from an open to a closed enzyme conformation. In parallel, the site-directed mutation of the putative substrate C-terminal binding residues, N260A and R369A, resulted in complete loss or partial decrease of the enzymatic activity.

Furthermore, enzymatic assay of the truncated PepD catalytic domain, PepDCAT, further demonstrated the functional role of the lid domain in substrate binding and selectivity.

The structural data on PepD reported here may inspire strategies for the improvement of the PepD family of enzymes toward applications in biotechnology and allow the design of targeted disease peptidases or prodrugs with altered specificity.

71

On the other hand, it is important to consider the observation that substitution of zinc ions with copper ions in the PepD active site generated an enzyme, CuCu-PepD, preferred catechol derivatives containing polar tails. Based on the model of protein-ligand docking, the ortho-hydroxyl groups of each substrate are bound by Asp119, Glu149, Asp173 and His461 through hydrogen bonding as well as through polar tail interactions with Glu175 and Tyr181. The model of CuCu-PepD-epinephrine fitted better into the active site by additional hydrophobic interactions between the PepD Ile432 and Ala434 and the terminal methyl group of epinephrine. Enzyme kinetics studies indicated that oxidation of epinephrine produced notable catalytic efficiency, explaining its low Km value. In all, PepD is a unique enzyme, whose function can be effectively altered by simple metal ions substitution. This study not only unveiled a correlation of enzyme function between peptide hydrolysis and catechol oxidation, but also may provide a new direction by which divergent enzyme evolution occurs.

There are several possible directions that could be considered for future studies based upon the findings presented here. To better understand the PepD structure-function relationship and biological function, and to determine potential applications in enzymatic therapy, the following investigations are suggested:

I. Structure analysis of PepD-inhibitor complex by X-ray crystallography

In our previous report, the crucial residues in PepD for metal ions binding and substrate C-terminal and/or transition state binding were identified. We used PepD native structure and mutational analysis; however, the substrate binding network and substrates can induce conformational changes in the enzyme that were not addressed by our study. Therefore, co-crystallization of PepD with inhibitor (bestatin or dopamine) would provide further insights into the substrate binding sites and catalytic mechanism.

72

Furthermore, the enzyme conformational change from the open to closed form might also be determined by complex structure.

II. Proteomic study of V. alginolyticus pepD-deficient knockout strain

Proteomic study includes not only the identification and quantification of a particular protein panel under certain physiologic conditions, but also the determination of their localization, modifications, interactions, activities, and ultimately their function.

A previous report by Brombacher et al., in which pepD gene was overexpressed in E.

coli, indicated that PepD was vital to reducing biofilm formation23. In a similar manner, a V. alginolyticus pepD-deficient strain might be used to investigate morphology and biofilm formation by comparing with wild-type. Two-dimensional protein electrophoresis and mass spectrometry would help to provide more detailed proteomic information about the biological effect on protein signaling pathways by PepD.

III. Potential application in antibody directed enzyme prodrug therapy (ADEPT) In ADEPT, the range of potential enzymes is limited by several constraints. For example, the enzyme’s function is to efficiently convert an inactive prodrug into the active drug; thus, its optimal pH must be near the pH of the tumor extracellular fluid.

V. alginolyticus PepD considered a promising candidate for use in ADEPT because it works at pH 6.8-7.4, which is similar to the environment of human body. In addition, it can hydrolyze a dipeptide into two amino acids, which is affectively the same mechanism of prodrug hydrolyzation. The analysis of V. alginolyticus PepD performed to date (including functional residues characterization, protein structure, mutagenesis analysis and enzyme kinetics analysis) would aid in the design of a PepD mutant with improved digestion efficiency for a particular prodrug with a dipeptide

73

skeleton. Therefore, it is feasible that enzymatic engineering of PepD can generate desired mutant proteins to hydrolyze a prodrug, such as benzoic acid mustard prodrug (Fig. 1.9). This type of PepD study will require the development of an animal model to determine the efficacy and safety of a PepD-based ADEPT strategy, prior to its use in humans.

IV. Molecular evolution of metal-substituted dipeptidase for protein plasticity

A unique enzyme catalytic promiscuity has recently been observed for the dinuclear aminopeptidase from Streptomyces griseus. SgAP exhibits a high efficiency of catalytic promiscuity toward phosphonate and phosphodiester hydrolysis under different physiological conditions67, 68. Moreover, the peptide hydrolysis activity of SgAP70, 71 could be converted to catechol oxidative activity by manipulation of its metal derivatives72.

In PepD, we have observed catechol oxidative activity induced by copper ions substitution in the active site. Using the crystal structure of PepD and attaining a better understanding of the critical role of individual amino-acid residues involved in metal ions binding and substrate recognition, it may be possible to create an artificial enzyme with phosphoester hydrolysis activity. This might be achieved by a combination of site-directed mutagenesis and metal substitution of PepD protein. Development of an artificial PepD exhibiting phosphoesterase function might generate new potentials in chemical warfare, such as degradation of the nerve agent Soman toxin89-91.

74

References

1. Wilcox, D. E., Binuclear Metallohydrolases. Chem Rev 1996, 96 (7), 2435-2458.

2. Dismukes, G. C., Manganese Enzymes with Binuclear Active Sites. Chem Rev 1996, 96 (7), 2909-2926.

3. Lipscomb, W. N.; Strater, N., Recent Advances in Zinc Enzymology. Chem Rev 1996, 96 (7), 2375-2434.

4. Munih, P.; Moulin, A.; Stamper, C. C.; Bennett, B.; Ringe, D.; Petsko, G. A.; Holz, R. C., X-ray crystallographic characterization of the Co(II)-substituted Tris-bound form of the aminopeptidase from Aeromonas proteolytica. J Inorg Biochem 2007, 101 (8), 1099-107.

5. Rawlings, N. D.; Barrett, A. J., Evolutionary families of metallopeptidases.

Methods Enzymol 1995, 248, 183-228.

6. Rawlings, N. D.; Morton, F. R.; Barrett, A. J., MEROPS: the peptidase database.

Nucleic Acids Res 2006, 34 (Database issue), D270-2.

7. Kwon, K.; Hasseman, J.; Latham, S.; Grose, C.; Do, Y.; Fleischmann, R. D.; Pieper, R.; Peterson, S. N., Recombinant expression and functional analysis of proteases from Streptococcus pneumoniae, Bacillus anthracis, and Yersinia pestis. BMC Biochem 2011, 12 (1), 17.

8. Chen, S. L.; Marino, T.; Fang, W. H.; Russo, N.; Himo, F., Peptide hydrolysis by the binuclear zinc enzyme aminopeptidase from Aeromonas proteolytica: a density functional theory study. J Phys Chem B 2008, 112 (8), 2494-500.

9. Rawlings, N. D.; Barrett, A. J., Evolutionary families of peptidases. Biochem J 1993, 290 ( Pt 1), 205-18.

10. Lindner, H. A.; Alary, A.; Boju, L. I.; Sulea, T.; Menard, R., Roles of dimerization

75

domain residues in binding and catalysis by aminoacylase-1. Biochemistry 2005, 44 (48), 15645-51.

11. Henrich, B.; Monnerjahn, U.; Plapp, R., Peptidase D gene (pepD) of Escherichia coli K-12: nucleotide sequence, transcript mapping, and comparison with other peptidase genes. J Bacteriol 1990, 172 (8), 4641-51.

12. Savijoki, K.; Palva, A., Purification and molecular characterization of a tripeptidase (PepT) from Lactobacillus helveticus. Appl Environ Microbiol 2000, 66 (2), 794-800.

13. Hellendoorn, M. A.; Franke-Fayard, B. M.; Mierau, I.; Venema, G.; Kok, J., Cloning and analysis of the pepV dipeptidase gene of Lactococcus lactis MG1363. J Bacteriol 1997, 179 (11), 3410-5.

14. Hershcovitz, Y. F.; Gilboa, R.; Reiland, V.; Shoham, G.; Shoham, Y., Catalytic mechanism of SGAP, a double-zinc aminopeptidase from Streptomyces griseus. FEBS J 2007, 274 (15), 3864-76.

15. Teufel, M.; Saudek, V.; Ledig, J. P.; Bernhardt, A.; Boularand, S.; Carreau, A.;

Cairns, N. J.; Carter, C.; Cowley, D. J.; Duverger, D.; Ganzhorn, A. J.; Guenet, C.;

Heintzelmann, B.; Laucher, V.; Sauvage, C.; Smirnova, T., Sequence identification and characterization of human carnosinase and a closely related non-specific dipeptidase. J Biol Chem 2003, 278 (8), 6521-31.

16. Sherwood, R. F.; Melton, R. G.; Alwan, S. M.; Hughes, P., Purification and properties of carboxypeptidase G2 from Pseudomonas sp. strain RS-16. Use of a novel triazine dye affinity method. Eur J Biochem 1985, 148 (3), 447-53.

17. Gojkovic, Z.; Sandrini, M. P.; Piskur, J., Eukaryotic beta-alanine synthases are functionally related but have a high degree of structural diversity. Genetics 2001, 158 (3), 999-1011.

18. Born, T. L.; Zheng, R.; Blanchard, J. S., Hydrolysis of

76

N-succinyl-L,L-diaminopimelic acid by the Haemophilus influenzae dapE-encoded desuccinylase: metal activation, solvent isotope effects, and kinetic mechanism.

Biochemistry 1998, 37 (29), 10478-87.

19. Meinnel, T.; Schmitt, E.; Mechulam, Y.; Blanquet, S., Structural and biochemical characterization of the Escherichia coli argE gene product. J Bacteriol 1992, 174 (7), 2323-31.

20. van der Drift, C.; Ketelaars, H. C., Carnosinase: its presence in Pseudomonas aeruginosa. Antonie Van Leeuwenhoek 1974, 40 (3), 377-84.

21. Schroeder, U.; Henrich, B.; Fink, J.; Plapp, R., Peptidase D of Escherichia coli K-12, a metallopeptidase of low substrate specificity. FEMS Microbiol Lett 1994, 123 (1-2), 153-9.

22. Klein, J.; Henrich, B.; Plapp, R., Cloning and expression of the pepD gene of Escherichia coli. J Gen Microbiol 1986, 132 (8), 2337-43.

23. Brombacher, E.; Dorel, C.; Zehnder, A. J.; Landini, P., The curli biosynthesis regulator CsgD co-ordinates the expression of both positive and negative determinants for biofilm formation in Escherichia coli. Microbiology 2003, 149 (Pt 10), 2847-57.

24. Miller, C. G.; Schwartz, G., Peptidase-deficient mutants of Escherichia coli. J Bacteriol 1978, 135 (2), 603-11.

25. Kirsh, M.; Dembinski, D. R.; Hartman, P. E.; Miller, C. G., Salmonella typhimurium peptidase active on carnosine. J Bacteriol 1978, 134 (2), 361-74.

26. Jozic, D.; Bourenkow, G.; Bartunik, H.; Scholze, H.; Dive, V.; Henrich, B.; Huber, R.; Bode, W.; Maskos, K., Crystal structure of the dinuclear zinc aminopeptidase PepV from Lactobacillus delbrueckii unravels its preference for dipeptides. Structure 2002, 10 (8), 1097-106.

27. Albrecht, A. M.; Boldizsar, E.; Hutchison, D. J., Carboxypeptidase displaying

77

differential velocity in hydrolysis of methotrexate, 5-methyltetrahydrofolic acid, and leucovorin. J Bacteriol 1978, 134 (2), 506-13.

28. Tucker, A. D.; Roswell, S.; Melton, R. G.; Paupitt, R. A., A new crystal form of carboxypeptidase G2 from Pseudomonas sp. strain RS-16 which is more amenable to structure determination. Acta Crystallogr D Biol Crystallogr 1996, 52 (Pt 4), 890-2.

29. Rowsell, S.; Pauptit, R. A.; Tucker, A. D.; Melton, R. G.; Blow, D. M.; Brick, P., Crystal structure of carboxypeptidase G2, a bacterial enzyme with applications in cancer therapy. Structure 1997, 5 (3), 337-47.

30. Vongerichten, K. F.; Klein, J. R.; Matern, H.; Plapp, R., Cloning and nucleotide sequence analysis of pepV, a carnosinase gene from Lactobacillus delbrueckii subsp.

lactis DSM 7290, and partial characterization of the enzyme. Microbiology 1994, 140 ( Pt 10), 2591-600.

31. Biagini, A.; Puigserver, A., Sequence analysis of the aminoacylase-1 family. A new proposed signature for metalloexopeptidases. Comp Biochem Physiol B Biochem Mol Biol 2001, 128 (3), 469-81.

32. Hakansson, K.; Miller, C. G., Structure of peptidase T from Salmonella typhimurium. Eur J Biochem 2002, 269 (2), 443-50.

33. Lundgren, S.; Gojkovic, Z.; Piskur, J.; Dobritzsch, D., Yeast beta-alanine synthase shares a structural scaffold and origin with dizinc-dependent exopeptidases. J Biol Chem 2003, 278 (51), 51851-62.

34. Agarwal, R.; Burley, S. K.; Swaminathan, S., Structural analysis of a ternary complex of allantoate amidohydrolase from Escherichia coli reveals its mechanics. J Mol Biol 2007, 368 (2), 450-63.

35. Unno, H.; Yamashita, T.; Ujita, S.; Okumura, N.; Otani, H.; Okumura, A.; Nagai, K.;

Kusunoki, M., Structural basis for substrate recognition and hydrolysis by mouse

78

carnosinase CN2. J Biol Chem 2008, 283 (40), 27289-99.

36. Perry, T. L.; Hansen, S.; Love, D. L., Serum-carnosinase deficiency in carnosinaemia. Lancet 1968, 1 (7554), 1229-30.

37. Perry, T. L.; Hansen, S.; Tischler, B.; Bunting, R.; Berry, K., Carnosinemia. A new metabolic disorder associated with neurologic disease and mental defect. N Engl J Med 1967, 277 (23), 1219-27.

38. Terplan, K. L.; Cares, H. L., Histopathology of the nervous system in carnosinase enzyme deficiency with mental retardation. Neurology 1972, 22 (6), 644-55.

39. Murphey, W. H.; Lindmark, D. G.; Patchen, L. I.; Housler, M. E.; Harrod, E. K.;

Mosovich, L., Serum carnosinase deficiency concomitant with mental retardation.

Pediatr Res 1973, 7 (7), 601-6.

40. Lunde, H.; Sjaastad, O.; Gjessing, L., Homocarnosinosis: hypercarnosinuria. J Neurochem 1982, 38 (1), 242-5.

41. Karita, M.; Etterbeek, M. L.; Forsyth, M. H.; Tummuru, M. K.; Blaser, M. J., Characterization of Helicobacter pylori dapE and construction of a conditionally lethal dapE mutant. Infect Immun 1997, 65 (10), 4158-64.

42. Minton, N. P.; Atkinson, T.; Bruton, C. J.; Sherwood, R. F., The complete nucleotide sequence of the Pseudomonas gene coding for carboxypeptidase G2. Gene 1984, 31 (1-3), 31-8.

43. Khan, T. H.; Eno-Amooquaye, E. A.; Searle, F.; Browne, P. J.; Osborn, H. M.;

Burke, P. J., Novel inhibitors of carboxypeptidase G2 (CPG2): potential use in antibody-directed enzyme prodrug therapy. J Med Chem 1999, 42 (6), 951-6.

44. Krause, A. S.; Weihrauch, M. R.; Bode, U.; Fleischhack, G.; Elter, T.; Heuer, T.;

Engert, A.; Diehl, V.; Josting, A., Carboxypeptidase-G2 rescue in cancer patients with delayed methotrexate elimination after high-dose methotrexate therapy. Leuk Lymphoma

79

2002, 43 (11), 2139-43.

45. Bagshawe, K. D., Antibody directed enzymes revive anti-cancer prodrugs concept.

Br J Cancer 1987, 56 (5), 531-2.

46. Bagshawe, K. D., Antibody-directed enzyme/prodrug therapy (ADEPT). Biochem Soc Trans 1990, 18 (5), 750-2.

47. Bagshawe, K. D.; Sharma, S. K.; Springer, C. J.; Rogers, G. T., Antibody directed enzyme prodrug therapy (ADEPT). A review of some theoretical, experimental and clinical aspects. Ann Oncol 1994, 5 (10), 879-91.

48. Adams, G. P.; Weiner, L. M., Monoclonal antibody therapy of cancer. Nat Biotechnol 2005, 23 (9), 1147-57.

49. Springer, C. J.; Bagshawe, K. D.; Sharma, S. K.; Searle, F.; Boden, J. A.; Antoniw, P.; Burke, P. J.; Rogers, G. T.; Sherwood, R. F.; Melton, R. G., Ablation of human choriocarcinoma xenografts in nude mice by antibody-directed enzyme prodrug therapy (ADEPT) with three novel compounds. Eur J Cancer 1991, 27 (11), 1361-6.

50. Antoniw, P.; Springer, C. J.; Bagshawe, K. D.; Searle, F.; Melton, R. G.; Rogers, G.

T.; Burke, P. J.; Sherwood, R. F., Disposition of the prodrug 4-(bis (2-chloroethyl) amino) benzoyl-L-glutamic acid and its active parent drug in mice. Br J Cancer 1990, 62 (6), 909-14.

51. Eccles, S. A.; Court, W. J.; Box, G. A.; Dean, C. J.; Melton, R. G.; Springer, C. J., Regression of established breast carcinoma xenografts with antibody-directed enzyme prodrug therapy against c-erbB2 p185. Cancer Res 1994, 54 (19), 5171-7.

52. Francis, R. J.; Sharma, S. K.; Springer, C.; Green, A. J.; Hope-Stone, L. D.; Sena, L.; Martin, J.; Adamson, K. L.; Robbins, A.; Gumbrell, L.; O'Malley, D.; Tsiompanou, E.;

Shahbakhti, H.; Webley, S.; Hochhauser, D.; Hilson, A. J.; Blakey, D.; Begent, R. H., A phase I trial of antibody directed enzyme prodrug therapy (ADEPT) in patients with

80

advanced colorectal carcinoma or other CEA producing tumours. Br J Cancer 2002, 87 (6), 600-7.

53. Gasparetti, C.; Faccio, G.; Arvas, M.; Buchert, J.; Saloheimo, M.; Kruus, K., Discovery of a new tyrosinase-like enzyme family lacking a C-terminally processed domain: production and characterization of an Aspergillus oryzae catechol oxidase. Appl Microbiol Biotechnol 2009, 86 (1), 213-26.

54. Chen, Q. X.; Liu, X. D.; Huang, H., Inactivation kinetics of mushroom tyrosinase in the dimethyl sulfoxide solution. Biochemistry (Mosc) 2003, 68 (6), 644-9.

55. Lerch, K., Neurospora tyrosinase: structural, spectroscopic and catalytic properties.

Mol Cell Biochem 1983, 52 (2), 125-38.

56. Cary, J. W.; Lax, A. R.; Flurkey, W. H., Cloning and characterization of cDNAs coding for Vicia faba polyphenol oxidase. Plant Mol Biol 1992, 20 (2), 245-53.

57. Deverall, B. J., Phenolase and pectic enzyme activity in the chocolate spot disease of beans. Nature 1961, 189 (28), 311.

58. Tief, K.; Hahne, M.; Schmidt, A.; Beermann, F., Tyrosinase, the key enzyme in melanin synthesis, is expressed in murine brain. Eur J Biochem 1996, 241 (1), 12-6.

59. Tief, K.; Schmidt, A.; Aguzzi, A.; Beermann, F., Tyrosinase is a new marker for cell populations in the mouse neural tube. Dev Dyn 1996, 205 (4), 445-56.

60. Klabunde, T.; Eicken, C.; Sacchettini, J. C.; Krebs, B., Crystal structure of a plant catechol oxidase containing a dicopper center. Nat Struct Biol 1998, 5 (12), 1084-90.

61. Pauling, L., Chemical achievement and hope for the future. Am Sci 1948, 36 (1), 51-8.

62. O'Brien, P. J.; Herschlag, D., Catalytic promiscuity and the evolution of new enzymatic activities. Chem Biol 1999, 6 (4), R91-R105.

63. Copley, S. D., Enzymes with extra talents: moonlighting functions and catalytic

81

promiscuity. Curr Opin Chem Biol 2003, 7 (2), 265-72.

64. Bornscheuer, U. T.; Kazlauskas, R. J., Catalytic promiscuity in biocatalysis: using old enzymes to form new bonds and follow new pathways. Angew Chem Int Ed Engl 2004, 43 (45), 6032-40.

65. Kazlauskas, R. J., Enhancing catalytic promiscuity for biocatalysis. Curr Opin Chem Biol 2005, 9 (2), 195-201.

66. Senior, S. Z.; Mans, L. L.; VanGuilder, H. D.; Kelly, K. A.; Hendrich, M. P.; Elgren, T. E., Catecholase activity associated with copper-S100B. Biochemistry 2003, 42 (15), 4392-7.

67. Park, H. I.; Ming, L. J., A 10(10) Rate Enhancement of Phosphodiester Hydrolysis by a Dinuclear Aminopeptidase-Transition-State Analogues as Substrates? Angew Chem Int Ed Engl 1999, 38 (19), 2914-2916.

68. Ercan, A.; Tay, W. M.; Grossman, S. H.; Ming, L. J., Mechanistic role of each metal ion in Streptomyces dinuclear aminopeptidase: PEPTIDE hydrolysis and 7x10(10)-fold rate enhancement of phosphodiester hydrolysis. J Inorg Biochem 2009, 104 (1), 19-29.

69. Ercan, A.; Park, H. I.; Ming, L. J., Remarkable enhancement of the hydrolyses of phosphoesters by dinuclear centers: Streptomyces aminopeptidase as a 'natural model system'. Chemical Communications 2000, (24), 2501-2502.

70. Spungin, A.; Blumberg, S., Streptomyces griseus aminopeptidase is a calcium-activated zinc metalloprotein. Purification and properties of the enzyme. Eur J Biochem 1989, 183 (2), 471-7.

71. Lin, L. Y.; Park, H. I.; Ming, L. J., Metal Binding and Active Site Structure of Di-Zinc Streptomyces griseus Aminopeptidase. J Biol Inorg Chem 1997, 2, 744-49.

72. da Silva, G. F.; Ming, L. J., Catechol oxidase activity of di-Cu2+-substituted aminopeptidase from Streptomyces griseus. J Am Chem Soc 2005, 127 (47), 16380-1.

82

73. da Silva, G. F. Z.; Ming, L. J., Catechol oxidase activity of Di-Cu2+-substituted aminopeptidase from Streptomyces griseus. J Am Chem Soc 2005, 127 (47), 16380-1.

74. Rompel, A.; Fischer, H.; Meiwes, D.; Buldt-Karentzopoulos, K.; Magrini, A.;

Eicken, C.; Gerdemann, C.; Krebs, B., Substrate specificity of catechol oxidase from Lycopus europaeus and characterization of the bioproducts of enzymic caffeic acid oxidation. FEBS Lett 1999, 445 (1), 103-10.

75. Torelli, S.; Belle, C.; Hamman, S.; Pierre, J. L.; Saint-Aman, E., Substrate binding in catechol oxidase activity: biomimetic approach. Inorg Chem 2002, 41 (15), 3983-9.

76. Wang, T. Y.; Chen, Y. C.; Kao, L. W.; Chang, C. Y.; Wang, Y. K.; Liu, Y. H.; Feng, J.

M.; Wu, T. K., Expression and characterization of the biofilm-related and carnosine-hydrolyzing aminoacylhistidine dipeptidase from Vibrio alginolyticus. FEBS J 2008, 275 (20), 5007-20.

77. Chang, C. Y.; Hsieh, Y. C.; Wang, T. Y.; Chen, C. J.; Wu, T. K., Purification, crystallization and preliminary X-ray analysis of an aminoacylhistidine dipeptidase (PepD) from Vibrio alginolyticus. Acta Crystallogr Sect F Struct Biol Cryst Commun 2009, 65 (Pt 3), 216-8.

78. Chang, C. Y.; Hsieh, Y. C.; Wang, T. Y.; Chen, Y. C.; Wang, Y. K.; Chiang, T. W.;

Chen, Y. J.; Chang, C. H.; Chen, C. J.; Wu, T. K., Crystal structure and mutational analysis of aminoacylhistidine dipeptidase from Vibrio alginolyticus reveal a new architecture of M20 metallopeptidases. J Biol Chem 2010, 285 (50), 39500-10.

79. Otwinowsk, Z.; Minor, W., Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol 1997, 276, 307-26.

80. Weiss, M. S., Global indicators of X-ray data quality. J Appl Cryst 2001, 34, 130-5.

81. Evans, P., Scaling and assessment of data quality. Acta Crystallogr D Biol Crystallogr 2006, 62 (Pt 1), 72-82.

83

82. Emsley, P.; Cowtan, K., Coot: model-building tools for molecular graphics. Acta

82. Emsley, P.; Cowtan, K., Coot: model-building tools for molecular graphics. Acta

相關文件