PSI诱导的PC12细胞帕金森病模型的动态蛋白质组学研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
本研究采用MTT及荧光染色法检测PSI对PC12细胞的毒性作用;首次采用荧光差异凝胶电泳(DIGE)和质谱(MS)技术分析PSI引起的动态蛋白质组学改变。
     本研究成功地建立了PSI诱导的PC12细胞帕金森病(PD)模型,比较全面地复制出PD细胞自噬、细胞凋亡、包涵体形成的病理特征。首次发现并鉴定了PSI诱导的PC12细胞PD模型动态蛋白质组学改变。通过DIGE和MS,鉴定出39个差异表达的蛋白点,它们是27种不同基因的表达产物。本研究已鉴定的27种蛋白中氧调节蛋白150前体、热休克蛋白105、热休克蛋白4、结构型热休克同源蛋白70拟基因蛋白1、葡萄糖调节蛋白58、葡萄糖调节蛋白75、葡萄糖调节蛋白94、酪氨酸羟化酶、26S蛋白酶体亚单位ATP酶2、类26S蛋白酶体亚单位p40.5、醛还原酶、细胞角蛋白8、血红素加氧酶-1、泛素羧基末端水解酶PGP9.5、钙激活蛋白酶共15种蛋白表达上调(p<0.05或p<0.01),外周蛋白、F1-ATP合成酶、actin cytoplasmic 2、Asnal、14-3-3蛋白、T-complex protein 1 alpha subunit、α-微管蛋白6、过氧化物还原酶2、过氧化物还原酶4、ERp29共10种蛋白表达下降(p<0.05或p<0.01)。除此之外,点936、952同为p47蛋白,但点936表达升高(p<0.05或p<0.01),点952表达下降(p<0.05);点598、1316、1351都是HSP27蛋白,但点598表达增加(p<0.01)而点1316、1351表达下降(p<0.05),这可能与同种蛋白的不同修饰状态有关。
     已鉴定的蛋白质中有7种蛋白表现为两种或两种以上的同功体(isform)。如蛋白点36、37、38,点174、178、183、185、186,点252、253,点275、281,点471、475,点936、952,点598、1316、1351,它们的分子量相当,等电点间隔相同,而且各点的质谱鉴定结果完全一致,提示同一种蛋白的一系列蛋白点可能代表不同的翻译后修饰。
Part I. Establishment of the model of Parkinson’s disease induced by proteasome inhibitor PSI in PC12 cells
     Parkinson’s disease (PD) is the second most common neurodegenerative disorder. Ubiquitin-proteasome system (UPS) is the primary enzymes involved in the degradation of proteins within cells. It can clear mutated, damaged and misfolded proteins in eukaryotic cells and maintain cell homeostasis. Recently, investigations have shown that UPS dysfunction may be the core pathogenesis of PD, but the details of the pathogenetic mechanism of UPS dysfunction in PD is still unknown.
     Objective To study the effects of proteasome inhibition on PC12 cells. Methods PC12 cells were treated with different concentrations of PSI (1, 2.5, 5, 10, 20μM) for 12h, 24h, 36h, 48h, 60h, 72h, respectively. Cell viability was calculated by MTT conversion; apoptotic percentage was measured by fluorescent staining; morphological changes were determined by HE staining. Result 10μM and 20μM PSI significantly decreased the cell viability to 85.46% and 84.08% after 24 hours, compared with that of control groups (p<0.05 or p<0.01). The cell viability further decreased with time. After 72 hours the viability of PC12 cells decreased significantly to 18.27% and 21.09% in 10μM and 20μM PSI treated groups, compared with control group (p<0.01). At all time points, except 12th hour, the viability of PC12 cells treated with 10μM or 20μM PSI lowered significantly, compared with that with 1μM PSI (P<0.05 or P<0.01). Apoptotic percentage of PC12 cells exposed to vehicle or increasing concentrations of PSI was measured by fluorescent DNA-binding dyes. The apoptotic nuclei of PC12 cells increased in groups exposed to increasing concentrations of PSI. In 10μM PSI treated groups, the apoptotic rate of PC12 cells increased with time. The apoptotic rates at 12h, 24h, 36h, 48h in 10μM PSI treated groups were 5.13±1.74%, 12.80±3.92%, 23.64±5.79%, 13.37±4.02%, respectively, measured by Hoechst33342 stain and 5.21±1.78%, 14.47±4.55%, 25.53±5.45%, 14.86±2.98%, respectively, measured by AO and EB stains, significantly higher than that in control groups (p<0.01), except that at 12h. In PSI treated groups, the population of dead cells increased from 36h, the cell membrane was destroyed from 48h, cell structure was severely damaged and cell loss obviously at 60h and 72h. Thus the apoptotic rate can’t be calculated at 60h and 72h. Control and 10μM PSI treated PC12 cells were fixed and stained with HE. In cells of control groups, the nucleus was stained with blue and the cytoplasm was stained with pink. There were no inclusions and vacuoles in cells of control groups. Eosinophilic cytoplasmic inclusions appeared in cells exposed to PSI for 12 hours and presented in almost all cells from 24h. Cytoplasmic vacuoles, putative autophagic vacuoles, became apparently at 24h and 36h after PSI treatment. At 48h, the cell membrane was damaged or broke down and eosinophilic inclusions remained intact. Conclusion Proteasome inhibitor PSI has obvious toxicity to PC12 cells, which decreases the cell viability, induces apoptosis and promotes eosinophilic inclusion formation, mimics the pathological hallmarks of Parkinson’s disease. Significance A widely accepted cell model of Parkinson’s disease was established, which set up a reliable basis for the study of mechanisms of proteasome dysfunction in the pathogenesis of Parkinson’s disease at proteomic level.
     Part II Dynamic proteomic studies of model of Parkinson’s disease induced by proteasome inhibitor PSI in PC12 cells Previous investigations have shown that UPS dysfunction may be the core pathogenesis of PD,but the details of the pathogenetic mechanism of UPS dysfunction in PD is still unknown. We have established the model of Parkinson’s disease induced with PSI in PC12 cells, which can mimic the primary hallmarks of PD, including cell autophagy, cell apoptosis, and cytoplasmic inclusions.
     Objective To investigate the dynamic proteomic changes in PC12 cells treated with proteasome inhibitor PSI. Methods 2D-DIGE and MALDI-TOF MS techniques were used to detect dynamic proteomic changes in PC12 cells treated with 10μM PSI for 3h, 6h, 12h, 24h, 36h, and 48h. Results About 2000 protein spots were seen in 2D-DIGE images. In PSI treated groups, 6 protein points were significantly changed at 3h, 9 protein points were significantly changed at 6h, 49 protein points were significantly changed at 12h, 46 protein points were significantly changed at 24h, 122 protein points were significantly changed at 36h, 97 protein points were significantly changed at 48h, compared with those in control groups. 39 protein points were successfully identified by MALDI-TOF-MS and they are the products of 27 different genes. 7 proteins had two or more isforms. Protein point 36, 37, 38, protein point 174, 178, 183, 185, 186, protein point 252, 253, protein point 275, 281, protein point 471, 475, protein point 936 and 952, and protein point 598, 1316, 1351 showed almost identical molecular mass but different pI, suggesting different post-translation modifications. In PSI treated groups, ORP150, HSP105, HSP4, HSC70-ps1, GRP58, GRP75, GRP94, TH, PSMC2, 26S proteasome subunit p40.5, AR, CK8, HO-1, UCH-L1 PGP9.5, calpain were all significantly up-regulated (p<0.05 or p<0.01) and PRP, F1-ATPase, actin cytoplasic 2, Asnal, tyrosine-3-mono -oxygenase/ryptopha5-monooxygenase activation protein, TCP-1α,α-tubulin 6, Prx2, Prx4, ERp29 were all significantly down-regulated, compared with those in control groups (p<0.05 or p<0.01). Both of protein spot 936 and 952 were p47, but point 936 was up-regulated and point 952 was down-regulated in PSI treated groups, compared to those in control groups (p<0.05). All of protein spot 598, 1316, 1351 were HSP27, but spot 598 was significantly up-regulated (p<0.01) and spot 1316 and 1351 were significantly down-regulated (p<0.05). Significance For the first time, present study identified 39 protein points of 27 proteins which significantly changed in Parkinson’s disease model induced by proteasome inhibitor PSI in dopaminergic PC12 cells and revealed the dynamic proteomic changes. These findings provide new valuable clues for further study of the pathogenesis of Parkinson’s disease and present new candidates of treatment targets of Parkinson’s disease.
引文
1. Betarbet R, Sherer TB, Greenamyre JT. Ubiquitin-proteasome system and Parkinson’ s diseases[J]. Exp. Neurol. 2005; 191 (Suppl 1): S17-S27.
    2. McNaught KS, Olanow CW, Halliwell B, et al. Failure of the ubiquitin-proteasome system in Parkinson’s disease[J]. Nat Rev Neurosci. 2001; 2(8): 589-594.
    3. Rideout HJ, Larsen KE, Sulzer D, et al. Proteasomal inhibition leads to formation of ubiquitin/α-synuclein-immunoreactive inclusions in PC12 cells[J]. J Neurochem. 2001; 78(4): 899-908.
    4. McNaught KSP, Perl DP, Brownell AL, et al. Systemic exposure to proteasome inhibitors causes a progressive model of Parkinson’s disease [J]. Ann Neurol. 2004; 56(1): 149-162.
    5. Ross CA, Pickart CM. The ubiquitin-proteasome pathway in Parkinson’s disease and other neurodegenerative disease[J]. Trends Cell Biol. 2004; 14(12): 703-711.
    6. McNaught KS, Jackson T, JnoBaptiste R, et al. Proteasome dysfunction in sporadic Parkinson’s disease[J]. Neurology. 2006; 66(10Suppl4): S37-S49.
    7. Maruyama Y, Teraoka H, Iwata H, et al. Inhibitory effects of endogenous dopaminergic neurotoxin, norsalsolinol on dopamine secretion in PC12 rat pheochromocytoma cells[J]. Neurochem Int. 2001; 38(7): 567-572.
    8. Wang JJ, Zhang T, Niu DB, et al. Doxycycline-regulated co-expression of GDNF and TH in PC12 cells[J]. Neurosci Lett. 2006; 401(1-2): 142-145.
    9. Anglade P, Vyas S, Javoy-Agid F, et al. Apoptosis and autophagy in nigral neurons of patients with Parkinson’s disease[J]. Histol Histopathol. 1997; 12(1): 25-31.
    10. Menendez-Benito V, Verhoef LG, Masucci MG, et al. Endoplasmic reticulum stress compromises the ubiquitin-proteasome system[J]. Hum Mol Genet. 2005; 14(19): 2787 -2799.
    11. Bush KT, Goldberg AL, Nigam SK. Proteasome inhibition leads to a heat-shock response, induction of endoplasmic reticulum chaperones, and thermotolerance[J]. J Biol Chem. 1997; 272(14): 9086-9092.
    12. Bergeron JJM, Brenner MB, Thomas DY, et al. Calnexin: a membrane-bound chaperone of the endoplasmic reticulum[J]. Trends Biochem Sci. 1994; 19(3):124-128.
    13. Yamagishi N, Ishihara K, Saito Y, et al. Hsp105 but not Hsp70 family proteins suppress the aggregation of heat-denatured protein in the presence of ADP[J]. FEBS Lett. 2003; 555(2): 390-396.
    14. Yamagishi N, Ishihara K, Saito Y, et al. Hsp105 family proteins suppress staurosporine -induced apoptosis by inhibiting the translocation of Bax to mitochondria in HeLa cells[J]. Exp Cell Res. 2006; 312(17): 3215-3223.
    15. Mimohara M. Heat shock protein 105 in multiple sclerosis[J]. Nippon Rinsho. 2003; 61(8): 1317-1322.
    16. Fink AL. Chaperone-mediated protein folding[J]. Physiol Rev. 1999; 79(2): 425-449.
    17. Morrow G, Tanguay RM. Heat shock proteins and aging in Drosophila melanogaster[J]. Semin Cell Dev Biol. 2003; 14(5): 291-199.
    18. Dedmon MM, Christodoulou J, Wilson MR, et al. Heat shock protein 70 inhibits alpha-synuclein fibril formation via preferential binding to prefibrillar species[J]. J Biol Chem. 2005; 280(15): 14733-14740.
    19. Huang C, Cheng H, Hao S, et al. Heat shock protein 70 inhibits alpha-synuclein fibril formation via interactions with diverse intermediates[J]. J Mol Biol. 2006; 364(3): 323-336.
    20. Auluck PK, Youngah HY, Trojanowski JQ, et al. Chaperone suppression of α-synuclein toxicity in a Drosophila model for Parkinson’s disease[J]. Science. 2002; 295(5556): 865-868.
    21. Mytilineou C, McNaught KS, Shashidharan P, et al. Inhibition of proteasome activity sensitizes dopamine neurons to protein alterations and oxidative stress[J]. J Neural Transm. 2004; 111(10-11): 1237-1251.
    22. Ahn TB, Jeon BS. Protective role of heat shock and heat shock protein 70 in lactacystin-induced cell death both in the rat substantia nigra and PC12 cells[J]. Brain Res. 2006; 1087(1): 159-167.
    23. Kermer P, Digicaylioglu MH, Kaul M, et al. BAG1 over-expression in brain protects against stroke[J]. Brain Pathol. 2003; 13(4): 495-506.
    24. Ahn SG, Kim SA, Yoon JH, et al. Heat-shock cognate 70 is required for the activation of heat-shock factor 1 in mammalian cells[J]. Biochem J. 2005; 392(Pt1): 145-152.
    25. Kawamoto Y, Akiguchi I, Shirakashi Y, et al. Accumulation of Hsc70 and Hsp70 in glial cytoplasmic inclusions in patients with multiple system atrophy[J]. Brain Res. 2007; 1136(1): 219-227.
    26. Donaire V, Niso M, Moran JM, et al. Heat shock proteins protect both MPP+ and paraquat neurotoxicity[J]. Brain Res Bull. 2005; 67(6): 509-514.
    27. Reynolds LP, Allen GV. A review of heat shock protein induction following cerebellar injury[J]. Cerebellum. 2003; 2(3): 171-177.
    28. Sharp FR, Bernamdim M, Bartels M, et al. Glial expression of heat shock proteins (HSPs) and oxygen-regulated proteins (ORPs)[J]. Prog Brain Res. 2001; 132: 427-440.
    29. Gorman MA, Szegezdi E, Quigney DJ, et al. Hsp27 inhibits 6-hydroxy-dopamine -induced cytochrome c release and apoptosis in PC12 cells[J]. Communications. 2005; 327(3): 801-810.
    30. Outeiro TF, Klucken J, Strathearn KE, et al. Small heat shock proteins protect against α-synuclein-induced toxicity and aggregation[J]. Biochem Biophys Res Commun. 2006; 351(3): 631-638.
    31. Whittier JE, Xiong Y, Rechsteiner MC, et al. Hsp90 enhances degradation of oxidized calmodulin by the 20S proteasome[J]. J Biol Chem. 2004; 279(44): 46135-46142.
    32. Fu Y, Lee AS. Glucose regulated proteins in cancer progression, drug resistance and immunotherapy[J]. Cancer Biol Ther. 2006; 5(7): 741-744.
    33. Reddy RK, Lu J, Lee AS. The endoplasmic reticulum chaperone glycoprotein GRP94 with Ca2+-binding and antiapoptotic properties is a novel proteolytic target of calpain during etoposide-induced apoptosis[J]. J Biol Chem. 1999; 274(40): 28476-28483.
    34. Wadhwa R, Taira K, Kaul SC. An Hsp70 family chaperone, mortalin/mthsp70 /PBP74/Grp75: what, when, and where?[J]. Cell Stress Chaperones. 2002; 7(3): 309-316.
    35. Savel AS, Novicova LA. ATP-dependent proteolysis in mutochodria, m-AAA protease and PIMI protease exert overlapping substrate specificities and cooperate with the mthsp70 system[J]. J Biol Chem. 1999; 273(32): 20596-20602.
    36. 刘雯, 左彶, 刘炎, 李锦燕, 郭锋. 缺糖损伤的PC12细胞中grp75基因的表达[J]. 中国优生与遗传杂志. 2005; 13(1): 16-17.
    37. 郑东航, 左彶, 杨增杰, 夏蓓莉, 张咸宁. Grp75对细胞缺糖损伤的保护作用[J]. 遗传学报. 2000; 27(8): 666-671.
    38. 左彶. grp75: 脑缺血性损伤得分子标记[J]. Chinese Bull Life Sci. 1998; 10(5): 216 -217.
    39. Jin J, Hulette G, Wang Y, et al. Proteomic identification of a stress protein, mortalin/mthsp70/GRP75: relevance to Parkinson disease[J]. Mol Cell Proteomics. 2006; 5(7): 1193-1204.
    40. Iwawaki T, Hosoda A, Okuda T, et al. Translational control by the ER transmembrane kinase/ribonuclease IRE1 under ER stress[J]. Nat Cell Biol. 2001; 3(2): 158-164.
    41. Bedard K, Szabo E, Michalak M, et al. Cellular functions of endoplasmic reticulumchaperones calreticulin, calnexin, and ERp57[J]. Int Rev Cytol. 2005; 245: 91-121.
    42. Grebenova D, Kuzelova K, Smetana K, et al. Mitochondrial and endoplasmic reticulum stress-induced apoptotic pathways are activated by 5-aminolevulinic acid-based photodynamic therapy in HL60 leukemia cells[J]. J Photochem Photobiol B. 2003; 69(2): 71-85.
    43. Mukhopadhyay S, Shah M, Patel K, et al. Monocrotaline pyrrole-induced megalocytosis of lung and breast epithelial cells: Disruption of plasma membrane and Golgi dynamics and an enhanced unfolded protein response[J]. Toxicol Appl Pharmacol. 2006; 211(3): 209-220.
    44. Erickson RR, Dunning LM, Olson DA, et al. In cerebrospinal fluid ER chaperones ERp57 and calreticulin bind β-amyloid[J]. Biochem Biophys Res Commun. 2005; 332(1): 50-57.
    45. Frenkel Z, Shenkman M, Kondratyev M, et al. Separate roles and different routing of calnexin and ERp57 in endoplasmic reticulum quality control revealed by interactions with asialoglycoprotein receptor chains[J]. Mol Bilo Cell. 2004; 15(5): 2133-2142.
    46. Kimura T, Imaishi K, Hagiwara Y, et al. ERp57 binds competitively to protein disulfide isomerase and calreticulin[J]. Biochem Biophys Res Commun. 2005; 331(1): 224-230.
    47. Aleshin AN, Sawa Y, Kitagawa-Sakakida S, et al. 150-kDa oxygen-regulated protein attenuates myocardial ischemia-reperfusion injury in rat heart[J]. J Mol Cell Cardiol. 2005; 38(3): 517-525.
    48. Miyazaki M, Ozawa K, Hori O, et al. Expression of 150-kd oxygen-regulated protein in the hippocampus suppreese delayed neuronal cell death[J]. J Cereb Blood Flow Metab. 2002; 22(8): 979-987.
    49. Kitano H, Nishimura H, Tachibana H, et al. ORP150 ameliorates ischemia/ reperfusion injury from middle cerebral artery occlusion in mouse brain[J]. Brain Res. 2004;1015(1-2): 122-128.
    50. Kitao Y, Ozawa K, Miyazaki M, et al. Expression of the endoplasmic reticulum molecular chaperone (ORP150) rescues hippocampal neurons from glutamate toxicity[J]. J Clin Invest. 2001; 108(10): 1439-1450.
    51. Hubbard MJ, Mangum JE, Mchugh NJ. Purification and biochemical characterization of native ERp29 from rat liver[J]. Biochem J. 2004; 383(3): 589-597.
    52. Hubbard MJ. Functional proteomics: The goalposts are moving[J]. Proteomics. 2002; 2(9): 1069-1078.
    53. Morand JP, Macri J, Adeli K. Proteomic profiling of hepatic endoplasmic reticulum -associated proteins in an animal model of insulin resistance and metabolic dyslipidemia[J]. J Biol Chem. 2005; 280(18): 17626-17633.
    54. Toronen P, Storvik M, Linden AM, et al. Expression profiling to understand actions of NMDA/glutamate receptor antagonists in rat brain[J]. Neurochem Res. 2002; 27(10): 1209-1220.
    55. Bo Z, Yongping S, Fengchao W, et al. Identification of differentially expressed proteins of gamma-ray irradiated rat intestinal epithelial IEC-6 cells by two-dimensional gel electrophoresis and matrix-assisted laser desorption/ ionization-time of flight mass spectrometry[J]. Proteomics. 2005; 5(2): 426-432.
    56. Huang YH, Chang AY, Huang CM, et al. Proteomic analysis of lipopolysaccharide -induced apoptosis in PC12 cells [J]. Proteomics. 2002; 2(9): 1220-1228.
    57. Llorca O, Martín-Benito J, Grantham J, et al. The ‘sequential allosteric ring’ mechanism in the eukaryotic chaperonin-assisted folding of actin and tubulin[J]. EMBO J. 2001; 20(15): 4065-4075.
    58. Yokota S, Kayano T, Ohta T, et al. Proteasome-dependent degradation of cytosolic chaperonin CCT[J]. Biochem Biophys Res Commun. 2000; 279(2): 712-717.
    59. Baranano DE, Snyder SH. Neural roles for heme oxygenase: contrasts to nitric oxide synthase[J]. Proc Natl Acad Sci USA. 2001; 98(20): 10996-11002.
    60. Dore S, Takahashi M, Ferris CD, et al. Bilirubin formed by activation of heme oxygenase-2, protects neurons against oxidative stress injury[J]. Proc Natl Acad Sci USA. 1999; 96(5): 2445-2450.
    61. Fukuda K, Richmon JD, Sato M, et al. Induction of heme oxygenase-1 (HO-1) in glia after traumatic brain injury[J]. Brain Res. 1996; 736(1-2): 68-75.
    62. Panahian N, Yoshiura M, Maines MD. Overexpression of heme oxygenase-1 is neuroprotective in a model of permanent middle cerebral artery occlusion in transgenic mice[J]. Neurochem. 1999; 72(3): 1187-1203.
    63. Kadoya C, Domino EF, Yang GY, et al. Preischemic but not postischemic zinc protoporphyrin treatQment reduces infarct size and edema accumulation after temporary focal cerebral ischemia in rats[J]. Stroke. 1995; 26(6): 1035-1038.
    64. Saavedra A, Baltazar G, Carvalho CM, et al. GDNF modulates HO-1 expression in substantia nigra postnatal cell cultures[J]. Free Radic Biol Med. 2005; 39(12): 1611-1619.
    65. Castellani R, Smith MA, Richey PL, et al. Glycoxidation and oxidative stress in Parkinson disease and diffuse Lewy body disease[J]. Brain Res. 1996; 737(1-2): 195-200
    66. Fernandez-Gonzalez A, Perez-Otano I, Morgan JI. MPTP selectively induces haem oxygenase-1 expression in striatal astrocytes[J]. Eur J Neurosci. 2000; 12(5): 1573-1583.
    67. Darby N, Morin PE, Tallbo G, et al. Refolding of bovine pancreatic trypain inhibitor via non-native disulphide intermediates[J]. J Mol Bio. 1995; 249(2): 463-477.
    68. Kim H, Lee TH, Park ES, et al. Role of peroxiredoxins in regulating intracellularhydrogen peroxide and hydrogen peroxide-induce apoptosis in thyroid cells[J]. J Biol Chem. 2000; 275(24):18266-18270.
    69. Chang JW,Jeon HB,Lee JH,et al. Augmented expression of perioxiredoxin in lung cancer[J]. Biochem Biophys Res Commun. 2001; 289(2): 507-512.
    70. Shen C, Nathan C. Nonredundant antioxidant defense by multiple two-cysteine peroxiredoxins in human prostate cancer cells[J]. Mol Med. 2002; 8(2): 95-102.
    71. Beaulieu JM, Robertson J, Julien JP. Interactions between peripherin and neurofilaments in cultured cells: disruption of peripherin assembly by the NF-M and NF-H subunits[J]. Biochem Cell Biol. 1999; 77(1): 41-45.
    72. Beaulieu JM, Kriz J, Julien JP. Induction of peripherin expression in subsets of brain neurons after lesion injury or cerebral ischemia[J]. Brain Res. 2002; 946(2): 153-161.
    73. Strong MJ, Leystra-Lantz G, Ge WW. Intermediate filament steady-state mRNA levels in amyotrophic lateral sclerosis[J]. Biochem Biophys Research Commun. 2004; 316(2): 317-322.
    74. Beaulieu JM, Nguyen MD, Julien JP. Late onset death of motor neurons in mice overexpressing wild-type peripherin[J]. J Cell Biol. 1999; 147(3): 531-544.
    75. He CZ and Hays AP. Expression of peripherin in ubiquinated inclusions of amyotrophic lateral sclerosis[J]. J Neurol Sci. 2004; 217(1): 47-54.
    76. Robertson J, Beaulieu JM, Doroudchi MM, et al. Apoptotic death of neurons exhibiting peripherin aggregates is mediated by the proinflammatory cytokine tumor necrosis factor-alpha[J]. J Cell Biol. 2001; 155(2): 217-226.
    77. Bernhard D, Csordas A, Henderson B, et al. Cigarette smoke metal-catalyzed protein oxidation leads to vascular endothelial cell contraction by depolymerization of microtubules[J]. FASEB J. 2005; 19(9): 1096-1107.
    78. Ren Y, Zhao J, Feng J. Parkin binds to alpha/beta tubulin and increases theirubiquitination and degradation[J]. Neuroscience. 2003; 23(8): 3316-3324.
    79. Ardley HC, Scott GB, Rose SA, et al. Inhibition of proteasomal activity causes inclusion formation in neuronal and non-neuronal cells overexpressing Parkin[J]. Mol Biol Cell. 2003; 14(11): 4541-4556.
    80. Connor B, Kozlowski DA, Unnerstall JR, et al. Glial cell line-derived neurotrophic factor(GDNF) gene delivery protects dopaminergic terminals from degeneration[J]. Exp Neurol. 2001; 169(1): 83-95.
    81. De Iuliis A, Grigoletto J, Recchia A, et al. A proteomic approach in the study of an animal model of Parkinson's disease[J]. Clin Chim Acta. 2005; 357(2): 202-209.
    82. Huynh DP, Scoles DR, Ho TH, et al. Parkin is associated with actin filaments in neuronal and nonneural cells[J]. Ann Neurol. 2000; 48(5): 737-744.
    83. Ben-Shachar D, Zuk R, Gazawi H, et al. Dopamine toxicity involves mitochondrial complex I inhibition: implications to dopamine-related neuropsychiatric disorders[J]. Biochem Pharmacol. 2004; 67(10): 1965-1974.
    84. Dela Fuente N, Maldonado AM, Portillo F. Glucose activation of the yeast plasma membrane H+-ATPase requires the ubiquitin-proteasome proteolytic pathway[J]. FEBS Lett. 1997; 411(2-3): 308-312.
    85. Davydov VV, Dobaeva NM, Bozhkov AI. Possible role of alteration of aldehyde’s scavenger enzymes during aging[J]. Exp Gerontol. 2004; 39(1): 11-16.
    86. Srivastava S, Chandra A, Ansari NH, et al. Identification of cardiac oxidoreductase(s) involved in the metabolism of the lipid peroxidation-derived aldehyde-4 –hydroxynonenal[J]. Biochem J. 1998; 329(Pt3): 469-475.
    87. Keightley JA, Shang L, Kinter M. Proteomic analysis of oxidative stress-resistant cells: a specific role for aldose reductase overexpression in cytoprotection[J]. Mol. Cell. Proteomics. 2004; 3(2): 167-175.
    88. Cui F, Wang Y, Wang J, et al. The up-regulation of proteasome subunits and lysosomal proteases in hepatocellular carcinomas of the HBx gene knockin transgenic mice[J]. Proteomics. 2006; 6(2): 498-504.
    89. Hori T, Kato S, Saeki M, et al. cDNA cloning and functional analysis of p28 (Nas6p) and p40.5 (Nas7p), two novel regulatory subunits of the 26S proteasome[J]. Gene. 1998; 216(1): 113-122.
    90. Kawamoto Y, Akiguchi I, Nakamura S, et al. 14-3-3 proteins in Lewy bodies in Parkinson disease and diffuse Lewy body disease brains[J]. J Neuropathol Exp Neurol. 2002; 61(3): 245-253.
    91. Luo GR, Chen R, Le WD. Are heat shock proteins therapeutic target for Parkinson's disease?[J]. Int J Biol Sci. 2007; 3(1): 20-26.
    92. Free TB, Olanow CW, Hauser RA, et al. Bilateral fetal nigral transplantation into the postcommissural putamen in Parkinson’s disease[J]. Ann Neurol. 1995; 38(3): 379-388.
    93. Goodwill KE, Sabatier C, Marks C, et al. Crystal structure of tyrosine hydroxylase at 2.3 A and its implications for inherited neurodegenerative diseases[J]. Nat Struct Biol. 1997; 4(7): 578-585.
    94. Ghribi O, Herman MM, Pramoonjago P, et al. MPP+ induces the endoplasmic reticulum stress response in rabbit brain involving activation of the ATF-6 and NF-kappaB signaling pathways[J]. J Neuropathol Exp Neurol. 2003; 62(11): 1144-1153.
    95. Ara J, Przedborski S, Naini AB, et al. Inactivation of tyrosine hydroxylase by nitration following exposure to peroxynitrite and 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine(MPTP)[J]. Proc Natl Acad Sci USA. 1998; 95(13): 7659-7663.
    96. Iwaya K, Ogawa H, Mukai Y, et al. Ubiquitin-immunoreactive degradation products of cytokeratin 8/18 correlate with aggressive breast cancer[J]. Cancer Sci. 2003; 94(10): 864-870.
    97. Wootz H, Hansson I, Korhonen L, et al. XIAP decreases caspase-12 cleavage and calpain activity in spinal cord of ALS transgenic mice[J]. Exp Cell Res. 2006; 312(10): 1890-1898.
    98. 徐冬梅,刘正湘. Calpain抑制剂I对大鼠急性缺血再灌注诱导心肌细胞凋亡的影响[J]. 中国组织化学与细胞化学杂志. 2001; 10(2): 160-163.
    99. Stefanis L, Larse KE, Rideout HJ, et al. Expression of A53T mutant but not wild-type α-synuclein in PC12 cells induces alterations of the ubiquitin-dependent degradation system, loss of dopamine release, and autophagic cell death[J]. J Neurosci. 2001; 21(24): 9549-9560.
    100. Sanges D, Comitato A, Tammaro R, et al. Apoptosis in retinal degeneration involves cross-talk between apop -tosin-inducing factor(AIF) and caspase-12 and is blocked by calpain inhibitors[J]. Proc Natl Acad Sci USA. 2006; 103(46): 17366-17371.
    101. Tizivion G, Avruch J. 14-3-3 proteins: active cofactors in cellular regulation by serine/threonine phosphorylation[J]. J Biol Chem. 2002; 277(5): 3061-3064.
    102. Masters SC, Yang HZ, Datta SR, et al. 14-3-3 inhibitors Bad-induced cell death throught interaction with Serine-136[J]. Mol Pharmacol. 2001; 60(6): 1325-1331.
    103. Woods YL, Ren G. Effect of multiple phosphorylation events on the transcription factors FKHR, FKHRL1 and AFX[J]. Biochem Soc Trans. 2002; 30(4): 391-397.
    104. Zhang R, He XR, Lir WM, et al. AIPI mediates TNF-α-induced ASK1 activation by facilitating dissociation of ASK1 from its inhibitor 14-3-3[J]. J Clin Invest. 2003; 111(12): 1933-1943.
    105. Vincenz C, Dixit VM. 14-3-3 protein associate with A20 in an isoform-specific manner and function both as chaperone and adapter molecules[J]. Cell Death Differ. 2001; 8(3): 265-272.
    106. Perez RG, Waymire JG, Lin E, et al. A role for alpha-synuclein in the regulation ofdopamine biosynthesis[J]. J Neurosci. 2002; 22(8): 3090-3099
    107. Xu J, Kao SY, Lee FJ, et al. Dopamine-dependent neurotoxicity of alpha-synuclein: a mechanism for selective neurodegeneration in Parkinson disease[J]. Nat Med. 2002; 8(6): 600-606.
    108. Wójcik C, Rowicka M, Kudlicki A, et al. Valosin-containing Protein (p97) is a regulator of endoplasmic reticulum stress and of the degradation of N-End rule and ubiquitin-fusion degradation pathway substrates in mammalian cells[J]. Mol Biol Cell. 2006; 17(11): 4606–4618.
    109. Hartmann-Petersen R, Wallace M, Hofmann K, et al. The ubx2 and ubx3 cofactors direct cdc48 activity to proteolytic and nonproteolytic ubiquitin- dependent processes[J]. Curr Biol. 2004; 14(9): 824-828.
    110. Meyer HH, Wang Y, Warren G. Direct binding of ubiquitin conjugates by the mammalian p97 adaptor complexes, p47 and Ufd1–Npl4[J]. EMBO J. 2002; 21(21): 5645-5652.
    111. Wojcik C, Yano M, DeMartino GN . RNA interference of valosin-containing protein (VCP/p97) reveals multiple cellular roles linked to ubiquitin/ proteasome-dependent proteolysis[J]. J Cell Sci. 2004; 117(Pt2): 281-292.
    112. Aguilar RC, Wendland B. Ubiquitin: not just for proteasomes anymore[J]. Curr Opin Cell Biol. 2003; 15(2): 184-190.
    113. Petersen R, Seeger M, Cordon C. Transferring substrates to the 26S proteasome[J]. Trends Biochem Sci. 2003; 28(1): 26-31.
    114. Bence NF, Sampat RM, Kopito RR. Impairment of the ubiquitin –proteasome system by protein aggregation[J]. Science. 2001; 292(5521): 1552-1555.
    115. Bennett EJ, Bence NF, Jayakumar R, er al. Global impairment of the ubiquitin -proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusionbody formation[J]. Mol Cell. 2005; 17(3): 351-365.
    116. Grune T, Jung T, Merker K, Davies KJ. Decreased proteolysis caused by protein aggregates, inclusion bodies, plaques, lipofuscin, ceroid, and ‘aggresomes’ during oxidative stress, aging, and disease[J]. Int J Biochem Cell Biol. 2004; 36(12): 2519-2530.
    117. Keller JN, Dimayuga E, Chen Q, et al. Autophagy, proteasomes, lipofuscin, and oxidative stress in the aging brain[J]. Int J Biochem Cell Biol. 2004; 36(12): 2376-2391.
    118. Keller JN, Hanni KB, Markesbery WR. Possible involvement of proteasome inhibition in aging: implications for oxidative stress[J]. Mech Ageing Dev. 2000; 113(1): 61-70.
    119. Zeng BY, Medhurst AD, Jackson M, et al. Proteasomal activity in brain differs between species and brain regions and changes with age[J]. Mech Ageing Dev. 2005; 126(6-7): 760-766.
    120. Floor E, Wetzel MG. Increased protein oxidation in human substantia nigra pars compacta in comparison with basal ganglia and prefrontal cortex measured with an improved dinitrophenylhydrazine assay[J]. J Neurochem. 1998; 70(1):268-75.
    121. Moss A, Blackburn-Munro G, Garry EM, et al. A role of the ubiquitin proteasome system in neuropathic pain[J]. J Neurosci. 2002; 22(4): 1363-1372.
    122. Lopez-Salon M, Alonso M, Vianna MR, et al. The ubiquitin proteasome cascade is required for mammalian long-term memory formation[J]. Eur J Neurosci. 2001; 14(11): 1820-1826.
    123. Chain DG, Schwartz JH, Hegde AN. Ubiquitin-mediated proteolysis in learning and memory[J]. Mol Neurobiol. 2000; 20(2-3): 125-142.
    124. Hegde AN, DiAntonio A. Ubiquitin and the synapse[J]. Nat Rev Neurosci. 2002; 3(11): 854-861.
    125. Hegde AN, Inokuchi K, Kandel ER, et al. Ubiquitin C-terminal hydrolase is an immediate-early gene essential for long-term facilitation in Aplysia[J]. Cell. 1997; 89(1): 115-126.
    126. McNaught KS, Olanow CW. Proteolytic stress: a unifying concept in the etiopathogenesis of familial and sporadic Parkinson’s disease[J]. Ann Neurol. 2003; 53(Suppl 3): S73–86.
    127. Gandhi S, Wood NW. Molecular pathophysiology of Parkinson’s disease[J]. Hum Mol Genet. 2005; 14(18): 2749-2755.
    128. Bence NF, Sampat RM, Kopito RR. Impairment of the ubiquitin–proteasome system by protein aggregation[J]. Science. 2001; 292(5521): 1552-1555.
    129. McNaught KS, Jenner P. Proteasomal function is impaired in substantia nigra in Parkinson’s disease[J]. Neurosci Lett. 2001; 297(3): 191-194.
    130. McNaught KS, Belizaire R, Isacson O, et al. Altered proteasomal function in sporadic Parkinson’s disease[J]. Exp Neurol. 2003; 179(1): 38-46.
    131. El-Khodor BF, Kholodilov NG, Yarygina O, et al. The expression of mRNAs for the proteasome complex is developmentally regulated in the rat mesencephalon[J]. Brain Res Dev Brain Res. 2001; 129(1): 47-56.
    132. DiDonato J, Mercurio F, Rosette C, et al. Mapping of the inducible IkappaB phosphorylation sites that signal its ubiquitination and degradation[J]. Mol Cell Biol. 1996; 16(4): 1295-1304.
    133. Conway KA, Rochet JC, Bieganski RM, et al. Kinetic stabilization of the alpha -synuclein protofibril by a dopamine alpha-synuclein adduct[J]. Science. 2001; 294(5545): 1346-1349.
    134. Cappai R, Leck SL, Tew DJ, et al. Dopamine promotes alpha-synuclein aggregation into SDS-resistant soluble oligomers via a distinct folding pathway[J]. FASEB J. 2005;19(10): 1377-1379.
    135. Golbe LI, Di Iorio G, Bonavita V, et al. A large kindred with autosomal dominant Parkinson’s disease[J]. Ann Neurol. 1990; 27(3): 276-282.
    136. Lee M, Hyun D, Halliwell B, et al. Effect of the overexpression of wild-type or mutant alpha-synuclein on cell susceptibility to insult[J]. J Neurochem. 2001; 76(4): 998 -1009.
    137. Feany MB, Bender WW. A Drosophila model of Parkinson’s disease[J]. Nature. 2000; 404(6776): 394-398.
    138. Kirik D, Annett LE, Burger C, et al. Nigrostriatal alpha-synucleinopathy induced by viral vector-mediated overexpression of human alpha-synuclein: a new primate model of Parkinson’s disease[J]. Proc Natl Acad Sci USA. 2003; 100(5): 2884-2889
    139. Conway KA, Harper JD, Lansbury PT. Accelerated in vitro fibril formation by a mutant alpha-synuclein linked to early-onset Parkinson disease[J]. Nat Med. 1998; 4(11): 1318-1320.
    140. Weinreb PH, Zhen W, Poon AW, et al. NACP, a protein implicated in Alzheimer’s disease and learning, is natively unfolded[J]. Biochemistry. 1996; 35(43): 13709 -13715.
    141. Conway KA, Lee SJ, Rochet JC, et al. Acceleration of oligomerization, not fibri -llization, is a shared property of both alpha-synuclein mutations linked to early-onset Parkinson’s disease: implications for pathogenesis and therapy[J]. Proc Natl Acad Sci USA. 2000; 97(2): 571-576.
    142. Li J, Uversky VN, Fink AL. Effect of familial Parkinson’s disease point mutations A30P and A53T on the structural properties, aggregation, and fibrillation of human alpha-synuclein[J]. Biochemistry. 2001; 40(38): 11604-11613.
    143. Lashuel HA, Petre BM, Wall J, et al. Alpha-synuclein, especially the Parkinson’sdisease-associated mutants, forms pore-like annular and tubular protofibrils[J]. J Mol Biol. 2002; 322(5): 1089-1102.
    144. Caughey B, Lansbury PT. Protofibrils, pores, fibrils, and neurodegener -ation: separating the responsible protein aggregates from the innocent bystanders[J]. Annu Rev Neurosci. 2003; 26: 267-298.
    145. Bennett MC, Bishop JF, Leng Y, et al. Degradation of alpha–synuclein by proteasome[J]. J Biol Chem. 1999; 274(48): 33855-33858.
    146. Tofaris GK, Layfield R, Spillantini MG. Alpha-synuclein metabolism and aggregation is linged to ubiqutin-independent degradation by the proteasome[J]. FEBS Lett. 2001; 509(1): 22-26.
    148. Lindersson E, Beedholm R, Hojrup P, et al. Proteasomal inhibition by alpha-synuclein filaments and oligomers[J]. J Biol Chem. 2004; 279(13): 12924-12934.
    149. Liu CW, Giasson BI, Lewis KA, et al. A precipitating role for truncated alpha -synuclein and the proteasome in alpha-synuclein aggregation: implications for pathogenesis of Parkinson’s disease[J]. J Biol Chem. 2005; 280(24): 22670 -22678.
    147. Tanaka Y, Engelender S, Igarashi S, et al. Inducible expression of mutant alpha -synuclein decreases proteasome activity and increases sensitivity to mitochondria -dependent apoptosis[J]. Hum Mol Genet. 2001; 10(9): 919-926.
    148. Stefanis L, Larsen KE, Rideout HJ, et al. Expression of A53T mutant but not wild-type alpha–synuclein in PC12 cells induces alterations of the ubiquitin-dependent degradation system, loss of dopamine release, and autophagic cell death[J]. J Neurosci. 2001; 21(24): 9549-9560.
    149. Bennett EJ, Bence NF, Jayakumar R, et al. Global impairment of the ubiquitin –proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation[J]. Mol Cell. 2005; 17(3): 351–365.
    150. Snyder H, Mensah K, Theisler C, et al. Aggregated and monomeric alpha–synuclein bind to the S6’ proteasomal protein and inhibit proteasomal function[J]. J Biol Chem. 2003; 278(14): 11753-11759.
    151. Yuji Tanaka, Simone Engelender, Shuichi Igarashi, et al. Inducible expression of mutant α-synuclein decreases proteasome activity and increases sensitivity to mitochondria-dependent apoptosis[J]. Hum Mol Genet. 2001; 10(9): 919-926.
    152. Chung KK, Zhang Y, Lim KL, et al. Parkin ubiquitinates the alpha-synuclein -interacting protein, synphilin-1: implications for Lewy-body formation in Parkinson disease[J]. Nature Med. 2001; 7(10): 1144-1150
    153. Itier JM, Ibanez P, Mena MA, et al. Parkin gene inactivation alters behaviour and dopamine neurotransmission in the mouse[J]. Hum Mol Genet. 2003; 12(18): 2277 -2291.
    154. Lincoln SJ, Maraganore DM, Lesnick TG, et al. Parkin variants in North American Parkinson’s disease: cases and controls[J]. Mov Disord. 2003; 18(11): 1306-1311.
    155. Petrucelli L, O’Farrell C, Lockhart PJ, et al. Parkin protects against the toxicity associated with mutant alpha-synuclein: proteasome dysfunction selectively affects catecholaminergic neurons[J]. Neuron. 2002; 36(6): 1007-1019.
    156. Pickart CM. Mechanisms underlying ubiquitination[J]. Annu Rev Biochem. 2001; 70: 503-533.
    157. Sakata E, Yamaguchi Y, Kurimoto E, et al. Parkin binds the Rpn10 subunit of 26S proteasomes through its ubiquitin-like domain[J]. EMBO Rep. 2003; 4(3): 301-306.
    158. Imai J, Yashiroda H, Maruya M, et al. Proteasomes and molecular chaperones: cellular machinery responsible for folding and destruction of unfolded proteins[J]. Cell Cycle. 2003; 2(6): 585-90.
    159. Imai Y, Soda M, Hatakeyama S, et al. CHIP is associated with Parkin, a generesponsible for familial Parkinson’s disease, and enhances its ubiquitin ligase activity[J]. Mol Cell. 2002; 10(1): 55- 67.
    160. Cyr DM, Hohfeld J, Patterson C. Protein quality control: U-box–containing E3 ubiquitin ligases join the fold[J]. Trends Biochem Sci. 2002; 27(7): 368-375.
    161. Imai Y, Soda M, Inoue H, et al. An unfolded putative transmembrane polypeptide, which can lead to endoplasmic reticulum stress, is a substrate of parkin[J]. Cell. 2001; 105(7): 891-902.
    162. Yang Y, Nishimura I, Imai Y, et al. Parkin suppresses dopaminergic neuron–selective neurotoxicity induced by Pael-R in Drosophila[J]. Neuron. 2003; 37(6): 911-924.
    163. Shimura H, Hattori N, Kubo S, et al. Familial Parkinson’s disease gene product, parkin, is aubiquitin–protein ligase[J]. Nat Genet. 2000; 25(3): 302-305.
    164. Shimura H, Schlossmacher MG, Hattori N, et al. Ubiquitination of a new form of {alpha}-synuclein by parkin from human brain: implications for Parkinson’s disease[J]. Science. 2001; 293(5528): 263-269.
    165. Leroy E, Boyer R, Auburger G, et al. The ubiquitin pathway in Parkinson’s disease[J]. Nature. 1998; 395(6701): 451- 452.
    166. Liu Y, Fallon L, Lashuel HA, et al. The UCH-L1 gene encodes two opposing enzymatic activities that affect alpha-synuclein degradation and Parkinson’s disease susceptibility[J]. Cell. 2002; 111(2): 209-218.
    167. Luo GR, Chen S, Le WD. Are heat shock proteins therapeutic target for Parkinson's disease?[J]. Int J Biol Sci. 2007; 3(1): 20-26.
    168. McNaught KSP, Mytilineou C, JnoBaptiste R, et al. Impairment of the ubiquitin -proteasome system causes dopaminergic cell death and inclusion body formation in ventral mesencephalic cultures[J]. J Neurochem. 2002; 81(2): 301–306.
    169. Saigoh K, Wang YL, Suh JG, et al. Intragenic deletion in the gene encoding ubiquitincarboxy-terminal hydrolase in gad mice[J]. Nat Genet. 1999; 23(1): 47-51.
    170. Yokota T, Sugawara K, Ito K, et al. Down regulation of DJ-1 enhances cell death by oxidative stress, ER stress, and proteasome inhibition[J]. Biochem Biophys Res Commun. 2003; 312(4): 1342-1348.
    171. Taira T, Saito Y, Niki T, et al. DJ-1 has a role in antioxidative stress to prevent cell death[J]. EMBO Rep. 2004; 5(2): 213-218.
    172. Moore DJ, Zhang L, Troncoso J, et al. Association of DJ-1 and parkin mediated by pathogenic DJ-1 mutations and oxidative stress[J]. Hum Mol Genet. 2005; 14(1): 71-84.
    173. Olzmann JA, Brown K, Wilkinson KD, et al. Familial Parkinson’s disease -associated L166P mutation disrupts DJ-1 protein folding and function[J]. J Biol Chem. 2004; 279(9): 8506-8515.
    174. Moore DJ, Zhang L, Dawson TM, et al. A missense mutation (L166P) in DJ-1, linked to familial Parkinson’s disease, confers reduced protein stability and impairs homo-oligomerization[J]. J Neurochem. 2003; 87(6): 1558-1567.
    175. Shendelman S, Jonason A, Martinat C, et al. DJ-1 is a redox-dependent molecular chaperone that inhibits alpha-synuclein aggregate formation[J]. PLoS Biol. 2004; 2(11): e362.
    176. Goldberg MS, Pisani A, Haburcak M, et al. Nigrostriatal dopaminergic deficits and hypokinesia caused by inactivation of the familial parkinsonism-linked gene DJ-1[J]. Neuron. 2005; 45(4): 489-496.
    177. Valente EM, Abou-Sleiman PM, Caputo V, et al. Hereditary early-onset Parkinson’s disease caused by mutations in PINK1[J]. Science. 2004; 304(5674): 1158-1160.
    178. Rogaeva E, Johnson J, Lang AE, et al. Analysis of the PINK1 gene in a large cohort of cases with Parkinson disease[J]. Arch Neurol. 2004; 61(12): 1898-1904.
    179. Zimprich A, Biskup S, Leitner P, et al. Mutations in LRRK2 cause autosomal -dominant parkinsonism with pleomorphic pathology[J]. Neuron. 2004; 44(4): 601-607
    180. Gilks WP, Abou-Sieiman PM, Gandhi S, et al. A common LRRK2 mutation in idiopathic Parkinson’s disease[J]. Lancet. 2005; 365(9457): 415-416.
    181. Di Fonzo A, Rohe CF, Ferreira J, et al. A frequent LRRK2 gene mutation associated with autosomal dominant Parkinson’s disease[J]. Lancet. 2005; 365(9457): 412- 415.
    182. Smith WW, Pei Z, Jiang H, et al. Leucine-rich repeat kinase 2 (LRRK2) interacts with parkin, and mutant LRRK2 induces neuronal degeneration[J]. Proc Natl Acad Sci USA. 2005; 102(51): 18676-18681.
    183. West AB, Moore DJ, Biskup S, et al. Parkinson’s disease-associated mutations in leucine-rich repeat kinase 2 augment kinase activity[J]. Proc Natl Acad Sci USA. 2005; 102(46): 16842-16847.
    184. Tofaris GK, Razzaq A, Ghetti B, et al. Ubiquitination of alpha–synuclein in Lewy bodies is a pathological event not associated with impairment of proteasome function[J]. J Biol Chem. 2003; 278(45): 44405-44411.
    185. Grunblatt E, Mandel S, Jacob-Hirsch J, et al. Gene expression profiling of Parkinsonian substantia nigra pars compacta; alterations in ubiquitin–proteasome, heat shock protein, iron and oxidative stress regulated proteins, cell adhesion/cellular matrix and vesicle trafficking genes[J]. J Neural Transm. 2004; 111(12): 1543-73.
    186. Furukawa Y, Vigouroux S, Wong H, et al. Brain proteasomal function in sporadic Parkinson’s disease and related disorders[J]. Ann Neurol. 2002; 51(6): 779-782.
    187. Pickart CM, Cohen RE. Proteasomes and their kin: proteases in the machine age[J]. Nat Rev Mol Cell Biol. 2004; 5(3): 177-187.
    188. Fornai F, Lenzi P, Gesi M, et al. Fine structure and biochemical mechanisms underlying nigrostriatal inclusions and cell death after proteasome inhibition[J]. JNeurosci. 2003; 239(26): 8955-8966.
    189. Fenteany G, Schreiber SL. Lactacystin, proteasome function, and cell fate[J]. J Biol Chem. 1998; 273(15): 8545-8548.
    190. Sin N, Kim KB, Elofsson M, et al. Total synthesis of the potent proteasome inhibitor epoxomicin: a useful tool for understanding proteasome biology[J]. Bioorg Med Chem Lett. 1999; 9(15): 2283-2288.
    191. Kisselev AF, Goldberg AL. Proteasome inhibitors: from research tools to drug candidates[J]. Chem Biol. 2001; 8(8): 739-758.
    192. Zhou Y, Shie FS, Piccardo P, et al. Proteasomal inhibition induced by manganese ethylene-bis-dithiocarbamate: relevance to Parkinson’s disease[J]. Neuroscience. 2004; 128(2): 281-291.
    193. Ensign JC, Normand P, Burden JP, et al. Physiology of some actinomycete genera[J]. Res Microbiol. 1993; 144(8): 657-660.
    194. Cross T. Aquatic actinomycetes: a critical survey of the occurrence, growth and role of actinomycetes in aquatic habitats[J]. J Appl Bacteriol. 1981; 50(3): 397-423.

© 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