首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
Tetraheme cytochrome c3 (13 kDa) and flavodoxin (16 kDa), are small electron transfer proteins that have been used to mimic, in vitro, part of the electron-transfer chain that operates between substract electron donors and respiratory electron acceptors partners in Desulfovibrio species (Palma, N., Moura, I., LeGall, J., Van Beeumen, J., Wampler, J., Moura, J. J. G. (1994) Biochemistry 33, 6394-6407). The electron transfer between these two proteins is believed to occur through the formation of a specific complex where electrostatic interaction is the main driving force (Stewart, D., LeGall, J., Moura, I., Moura, J.J.G., Peck, H.D., Xavier, A.V., Weiner, P.K. and Wampler, J.E. (1988) Biochemistry 27, 2444-2450, Stewart, D., LeGall, J., Moura, I., Moura, J.J.G., Peck, H.D., Xavier, A.V., Weiner, P., Wampler, J. (1989) Eur. J. Biochem. 185, 695-700). In order to obtain structural information of the pre-complex, a covalent complex between the two proteins was prepared. A water-soluble carbodiimide [EDC (1-ethyl-3(3 dimethylaminopropyl) carbodiimide hydrochloride] was used for the cross linking reaction. The reaction was optimized varying a wide number of experimental parameters such as ionic strength, protein and cross linker concentration, and utilization of different cross linkers and reaction time between the crosslinker and proteins.  相似文献   

3.
Previous crystallographic studies of the antibacterial trimethoprim in complexes with bacterial and avian dihydrofolate reductases have shown substantial differences in the mode of binding, providing plausible explanations for the origin of the remarkable species selectivity of this inhibitor (Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., Kaufman, B. T., Beddell, C. R., Champness, J. N., Stammers, D. K., and Kraut, J. (1985) J. Biol. Chem. 260, 381-391; Matthews, D. A., Bolin, J. T., Burridge, J. M., Filman, D. J., Volz, K. W., and Kraut, J. (1985) J. Biol. Chem. 260, 392-399). A major species difference between the active sites is that the only carboxylate present is always Glu in vertebrates and Asp in bacteria. Crystallographic studies of the wild-type and E30D mutant of the enzyme from mouse now reveal that in both cases trimethoprim is bound in an identical fashion to that observed with the bacterial enzyme, and there is no obvious single explanation for the origin of the 10(5)-fold selectivity of trimethoprim binding. In an earlier study of a mouse wild-type enzyme using more limited data it was proposed that trimethoprim bound in the avian mode (Stammers, D. K., Champness, J. N., Beddell, C. R., Dann, J. G., Eliopoulos, E. E., Geddes, A. J., Ogg, D., and North, A. C. T. (1987) FEBS Lett. 218, 178-184), but a re-examination indicates that the occupancy of the active site by trimethoprim is less than had been thought, and we are currently unable to make an unambiguous interpretation of the electron density maps and cannot confirm the avian mode of binding in those crystals.  相似文献   

4.
Expression of insulin metabolic effects can be obtained by anti-receptor antibodies without activation of the tyrosine kinase activity [O'Brien R. M., Soos M. A. and Siddle K. (1987) EMBO J. 6, 4003-4010; Forsayeth J. R., Caro J. F., Sinha M. K., Maddux B. A. and Goldfine I. D. (1987) Proc. natn. Acad. Sci. U.S.A. 84, 34,448-34,514; Ponzio G., Contreres J. O., Debant A., Baron V., Gautier N., Dolais-Kitabgi J. and Rossi B. (1988) EMBO J. 7, 4111-4117; Hawley D. M., Maddux B. A., Patel R. G., Wong K. Y., Mamula P. W., Firestone G. L., Brunetti A., Verspohl E. and Goldfine I. D. (1989) J. biol. Chem. 264, 2438-2444; Soos M. A., O'Brien R. M., Brindle N. P. J., Stigter J. M., Okamoto A. K., Whittaker J. and Siddle K. (1989) Proc. natn. Acad. Sci. U.S.A. 86, 5217-5221.]. Recently, we have proposed that receptor cross-linking is sufficient in itself to stimulate glycogen synthesis, even if aggregation was performed on receptors mutated on Tyr 1162 and Tyr 1163 and thus devoid of tyrosine kinase activity [Debant A., Ponzio G., Clauser E., Contreres J. O. and Rossi B. (1989) Biochemistry 28, 14-17]. The aim of this study was to gain information on the involvement of receptor clustering in the expression of the different insulin biological effects. To this end, we studied the mimetic effects of wheat-germ agglutinin, which is likely to induce receptor aggregation without interacting with the receptor protein moiety. Wheat-germ agglutinin failed to promote DNA synthesis, whereas the lectin behaved as a potent mimicker of insulin on tyrosine aminotransferase activity and amino-acid transport. However, this stimulatory effect did not parallel the activation of receptor autophosphorylation. Our data reinforce the idea that the expression of the metabolic effects of insulin are not strictly dependent on a general tyrosine kinase activation.  相似文献   

5.
Bone morphogenetic proteins (BMPs) are members of the transforming growth factor-beta superfamily of multifunctional ligands that transduce their signals through type I and II serine/threonine kinase receptors and intracellular Smad proteins. Recently, we identified the glycosylphosphatidylinositol-anchored repulsive guidance molecules RGMa, DRAGON (RGMb), and hemojuvelin (RGMc) as coreceptors for BMP signaling (Babbit, J. L., Huang, F. W., Wrighting, D. W., Xia, Y., Sidis, Y., Samad, T. A., Campagna, J. A., Chung, R., Schneyer, A., Woolf, C. J., Andrews, N. C., and Lin, H. Y. (2006) Nat. Genet. 38, 531-539; Babbit, J. L., Zhang, Y., Samad, T. A., Xia, Y., Tang, J., Schneyer, A., Woolf, C. J., and Lin, H. Y. (2005) J. Biol. Chem. 280, 29820-29827; Samad, T. A., Rebbapragada, A., Bell, E., Zhang, Y., Sidis, Y., Jeong, S. J., Campagna, J. A., Perusini, S., Fabrizio, D. A., Schneyer, A. L., Lin, H. Y., Brivanlou, A. H., Attisano, L., and Woolf, C. J. (2005) J. Biol. Chem. 280, 14122-14129). However, the mechanism by which RGM family members enhance BMP signaling remains unknown. Here, we report that RGMa bound to radiolabeled BMP2 and BMP4 with Kd values of 2.4+/-0.2 and 1.4+/-0.1 nm, respectively. In KGN human ovarian granulosa cells and mouse pulmonary artery smooth muscle cells, BMP2 and BMP4 signaling required BMP receptor type II (BMPRII), but not activin receptor type IIA (ActRIIA) or ActRIIB, based on changes in BMP signaling by small interfering RNA inhibition of receptor expression. In contrast, cells transfected with RGMa utilized both BMPRII and ActRIIA for BMP2 or BMP4 signaling. Furthermore, in BmpRII-null pulmonary artery smooth muscle cells, BMP2 and BMP4 signaling was reduced by inhibition of endogenous RGMa expression, and RGMa-mediated BMP signaling required ActRIIA expression. These findings suggest that RGMa facilitates the use of ActRIIA by endogenous BMP2 and BMP4 ligands that otherwise prefer signaling via BMPRII and that increased utilization of ActRIIA leads to generation of an enhanced BMP signal.  相似文献   

6.
Beta-actinin is equivalent to Cap Z protein   总被引:2,自引:0,他引:2  
Chicken skeletal muscle beta-actinin, previously reported to bind the slow-exchanging (pointed) ends of actin filaments was purified to homogeneity. By two dimensional gel electrophoresis, it consists of two subunits, beta I (35 kDa) and beta II (32 kDa), and each subunit has two isoforms. The amino acid sequences of V8 protease-digested peptides of beta I were nearly identical with those of portions of the muscle barbed end-blocking protein Cap Z alpha, although several amino acids were different from those deduced from cDNA sequences (Casella, J.F., Casella, S.J., Hollands, J.A., Caldwell, J.E., and Cooper, J.A. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 5800-5804). The amino acid sequences of two peptides from beta II were completely identical with portions of Cap Z beta deduced from cDNA sequences (Caldwell, J.E., Waddle, J.A., Cooper, J.A., Hollands, J.A., Casella, S.J., and Casella, J.F. (1989) J. Biol. Chem. 264, 12648-12652). beta-Actinin capped the barbed end of an actin filament as evidenced by actin assembly of myosin S1-decorated filaments and specifically its impairment of growth in the "barbed" direction. Thus it is concluded that highly purified beta-actinin is identical with the more recently described Cap Z, an actin barbed-end capping protein of chicken skeletal muscle.  相似文献   

7.
We have previously discovered the opium alkaloid noscapine as a microtubule interacting agent that binds to tubulin, alters the dynamics of microtubule assembly, and arrests mammalian cells at mitosis (Ye, K., Ke, Y., Keshava, N., Shanks, J., Kapp, J. A., Tekmal, R. R., Petros, J., and Joshi, H. C. (1998) Proc. Natl. Acad. Sci. U. S. A. 95, 1601-1606; Ye, K., Zhou, J., Landen, J. W., Bradbury, E. M., and Joshi, H. C. (2001) J. Biol. Chem. 276, 46697-46700; Zhou, J., Panda, D., Landen, J. W., Wilson, L., and Joshi, H. C. (2002) J. Biol. Chem. 277, 17200-17208). Here we show that noscapine does not compete with paclitaxel for tubulin binding and can efficiently inhibit the proliferation of both paclitaxel-sensitive and paclitaxel-resistant human ovarian carcinoma cells (i.e. the parental cell line 1A9 and two derivative cell lines, 1A9PTX10 and 1A9PTX22, which harbor beta-tubulin mutations that impair paclitaxel-tubulin interaction (Giannakakou, P., Sackett, D. L., Kang, Y. K., Zhan, Z., Buters, J. T., Fojo, T., and Poruchynsky, M. S. (1997) J. Biol. Chem. 272, 17118-17125). Strikingly, these cells undergo apoptotic death upon noscapine treatment, accompanied by activation of the c-Jun NH(2)-terminal kinases (JNK). Furthermore, inhibition of JNK activity by treatment with antisense oligonucleotide or transfection with dominant-negative JNK blocks noscapine-induced apoptosis. These findings thus indicate a great potential for noscapine in the treatment of paclitaxel-resistant human cancers. In addition, our results suggest that the JNK pathway plays an essential role in microtubule inhibitor-induced apoptosis.  相似文献   

8.
The phylogenetic problem of Huia (Amphibia: Ranidae)   总被引:1,自引:0,他引:1  
A taxonomic consensus for the diverse and pan-global frog family Ranidae is lacking. A recently proposed classification of living amphibians [Frost, D.R., Grant, T., Faivovich, J., Bain, R. H., Haas, A., Haddad, C.F.B., de Sá, R.O., Channing, A., Wilkinson, M., Donnellan, S.C., Raxworthy, C.J., Campbell, J.A., Blotto, B.L., Moler, P., Drewes, R.C., Nussbaum, R.A., Lynch, J.D., Green, D.M., Wheeler, W.C., 2006. The amphibian tree of life. B. Am. Mus. Nat. Hist. 297, 1-370] included expansion of the Southeast Asian ranid frog genus Huia from seven to 47 species, but without having studied the type species of Huia. This study tested the monophyly of this concept of Huia by sampling the type species and putative members of Huia. Molecular phylogenetic analyses consistently recovered the type species H. cavitympanum as the sister taxon to other Bornean-endemic species in the genus Meristogenys, rendering all previously published concepts of Huia as polyphyletic. Members of Huia sensu [Frost, D.R., Grant, T., Faivovich, J., Bain, R. H., Haas, A., Haddad, C.F.B., de Sá, R.O., Channing, A., Wilkinson, M., Donnellan, S.C., Raxworthy, C.J., Campbell, J.A., Blotto, B.L., Moler, P., Drewes, R.C., Nussbaum, R.A., Lynch, J.D., Green, D.M., Wheeler, W.C., 2006. The amphibian tree of life. B. Am. Mus. Nat. Hist. 297, 1-370.] appear in four places within the family Ranidae. A clade containing the type species of Odorrana is phylogenetically unrelated to the type species of Huia, and Odorrana is removed from synonymy with Huia. These findings underscore the need to include relevant type species in phylogenetic studies before proposing sweeping taxonomic changes. The molecular phylogenetic analyses revealed a high degree of homoplasy in larval and adult morphology of Asian ranid frogs. Detailed studies are needed to identify morphological synapomorphies that unite members in these major clades of ranid frogs.  相似文献   

9.
UV and ionizing radiation (IR) activate DNA damage checkpoints and induce Cdc25A degradation (Mailand, N., Falck, J., Lukas, C., Syljuasen, R. G., Welcker, M., Bartek, J., and Lukas, J. (2000) Science 288, 1425-1429; Falck, J., Mailand, N., Syljuasen, R. G., Bartek, J., and Lukas J. (2001) Nature 410, 842-847). The degradation of Cdc25A is abrogated by caffeine, which implicates Chk1 as the potential mediator (Mailand, N., Falck, J., Lukas, C., Syljuasen, R. G., Welcker, M., Bartek, J., and Lukas, J. (2000) Science 288, 1425-1429). However, the involvement of Chk1 is far from clear, because caffeine is a rather nonspecific inhibitor of the ATR/Chk1 signaling pathway. Additionally, it is not known whether DNA-damaging drugs commonly used in chemotherapy, which may activate different signal transduction pathways than UV or IR, also confer Cdc25A degradation. Herein, we show that camptothecin and doxorubicin, two widely used topoisomerase inhibitors conferring S and G2 arrest, respectively, cause the degradation of Cdc25A. Using a small interfering RNA that enables the specific elimination of Chk1 expression, we show that the observed proteolysis of Cdc25A is mediated through Chk1. Moreover, Cdc25A overexpression abrogates the Chk1-mediated degradation and overcomes the doxorubicin-induced G2 arrest through dephosphorylation and activation of Cdc2/Cdk1 in a dose-dependent manner. These results suggest that: (a) Cdc25A is involved in the G2/M transition in addition to its commonly accepted effect on G1/S progression, and (b) Chk1 mediates both S and G2 checkpoint and is thus a more ubiquitous cell cycle checkpoint mediator than previously thought.  相似文献   

10.
HFE is the protein product of the gene mutated in the autosomal recessive disease hereditary hemochromatosis (Feder, J. N., Gnirke, A., Thomas, W., Tsuchihashi, Z., Ruddy, D. A., Basava, A., Dormishian, F., Domingo, R. J., Ellis, M. C., Fullan, A., Hinton, L. M., Jones, N. L., Kimmel, B. E., Kronmal, G. S., Lauer, P., Lee, V. K., Loeb, D. B., Mapa, F. A., McClelland, E., Meyer, N. C., Mintier, G. A., Moeller, N., Moore, T., Morikang, E., Prasss, C. E., Quintana, L., Starnes, S. M., Schatzman, R. C., Brunke, K. J., Drayna, D. T., Risch, N. J., Bacon, B. R., and Wolff, R. R. (1996) Nat. Genet. 13, 399-408). At the cell surface, HFE complexes with transferrin receptor (TfR), increasing the dissociation constant of transferrin (Tf) for its receptor 10-fold (Gross, C. N., Irrinki, A., Feder, J. N., and Enns, C. A. (1998) J. Biol. Chem. 273, 22068-22074; Feder, J. N., Penny, D. M., Irrinki, A., Lee, V. K., Lebron, J. A., Watson, N. , Tsuchihashi, Z., Sigal, E., Bjorkman, P. J., and Schatzman, R. C. (1998) Proc. Natl. Acad. Sci. U S A 95, 1472-1477). HFE does not remain at the cell surface, but traffics with TfR to Tf-positive internal compartments (Gross et al., 1998). Using a HeLa cell line in which the expression of HFE is controlled by tetracycline, we show that the expression of HFE reduces 55Fe uptake from Tf by 33% but does not affect the endocytic or exocytic rates of TfR cycling. Therefore, HFE appears to reduce cellular acquisition of iron from Tf within endocytic compartments. HFE specifically reduces iron uptake from Tf, as non-Tf-mediated iron uptake from Fe-nitrilotriacetic acid is not altered. These results explain the decreased ferritin levels seen in our HeLa cell system and demonstrate the specific control of HFE over the Tf-mediated pathway of iron uptake. These results also have implications for the understanding of cellular iron homeostasis in organs such as the liver, pancreas, heart, and spleen that are iron loaded in hereditary hemochromatotic individuals lacking functional HFE.  相似文献   

11.
The formyl peptide receptor (FPR) is a chemotactic G protein-coupled receptor found on the surface of phagocytes. We have previously shown that the formyl peptide binding site maps to the membrane-spanning region (Miettinen, H. M., Mills, J. S., Gripentrog, J. M., Dratz, E. A., Granger, B. L., and Jesaitis, A. J. (1997) J. Immunol. 159, 4045-4054). Recent reports have indicated that non-formylated peptides, such as MMWLL can also activate this receptor (Chen, J., Bernstein, H. S., Chen, M., Wang, L., Ishi, M., Turck, C. W., and Coughlin, S. R. (1995) J. Biol. Chem. 270, 23398-23401.) Here we show that the selectivity for the binding of different NH(2)-terminal analogs of MMWLL or MLF can be markedly altered by mutating Asp-106 to asparagine or Arg-201 to alanine. Both D106N and R201A produced a similar change in ligand specificity, including an enhanced ability to bind the HIV-1 peptide DP178. In contrast, the mutation R205A exhibited altered specificity at the COOH terminus of fMLF, with R205A binding fMLF-O-butyl > fMLF-O-methyl > fMLF, whereas wt FPR bound fMLF > fMLF-O-methyl approximately fMLF-O-butyl. These data, taken together with our previous finding that the leucine side chain of fMLF is probably bound to FPR near FPR (93)VRK(95) (Mills, J. S., Miettinen, H. M., Barnidge, D., Vlases, M. J., Wimer-Mackin, S., Dratz, E. A., and Jesaitis, A. J. (1998) J. Biol. Chem. 273, 10428-10435.), indicate that the most likely positioning of fMLF in the binding pocket of FPR is approximately parallel to the fifth transmembrane helix with the formamide group of fMLF hydrogen-bonded to both Asp-106 and Arg-201, the leucine side chain pointing toward the second transmembrane region, and the COOH-terminal carboxyl group of fMLF ion-paired with Arg-205.  相似文献   

12.
ElaC is a widespread gene found in eubacteria, archaebacteria, and mammals with a highly conserved sequence. Two human ElaC variants were recently associated with cancer (Tavtigian, S. V., Simard, J., Teng, D. H., Abtin, V., Baumgard, M., Beck, A., Camp, N. J., Carillo, A. R., Chen, Y., Dayananth, P., Desrochers, M., Dumont, M., Farnham, J. M., Frank, D., Frye, C., Ghaffari, S., Gupte, J. S., Hu, R., Iliev, D., Janecki, T., Kort, E. N., Laity, K. E., Leavitt, A., Leblanc, G., McArthur-Morrison, J., Pederson, A., Penn, B., Peterson, K. T., Reid, J. E., Richards, S., Schroeder, M., Smith, R., Snyder, S. C., Swedlund, B., Swensen, J., Thomas, A., Tranchant, M., Woodland, A. M., Labrie, F., Skolnick, M. H., Neuhausen, S., Rommens, J., and Cannon-Albright, L. A. (2001) Nat. Genet. 27, 172-180; Yanaihara, N., Kohno, T., Takakura, S., Takei, K., Otsuka, A., Sunaga, N., Takahashi, M., Yamazaki, M., Tashiro, H., Fukuzumi, Y., Fujimori, Y., Hagiwara, K., Tanaka, T., and Yokota, J. (2001) Genomics 72, 169-179). Analysis of the primary sequence indicates homology to an arylsulfatase and predicts a metallo-beta-lactamase fold. At present, no ElaC gene product has been investigated. We cloned the Escherichia coli ElaC gene and purified the recombinant gene product. An enzymatic analysis showed that ElaC does not encode an arylsulfatase but rather encodes a phosphodiesterase that hydrolyzes bis(p-nitrophenyl)phosphate with a k(cat) of 59 s(-1) and K' of 4 mm. Kinetic analysis of the dimeric enzyme revealed positive cooperativity for the substrate bis(p-nitrophenyl)phosphate with a Hill coefficient of 1.6, whereas hydrolysis of the substrate thymidine-5'-p-nitrophenyl phosphate followed Michaelis-Menten kinetics. Furthermore, the enzyme is capable of binding two zinc or two iron ions. However, it displays phosphodiesterase activity only in the zinc form. The metal environment characterized by zinc K-edge x-ray absorption spectroscopy was modeled with two histidine residues, one carboxylate group, and 1.5 oxygen atoms. This corresponds to the coordination found in other metallo-beta-lactamase domain proteins. Phosphodiesterase activity is strongly dependent on the presence of zinc. These results identify the currently unassigned gene product ElaC to be a novel binuclear zinc phosphodiesterase.  相似文献   

13.
The nucleotide sequences of partial cDNA clones coding for the core protein of a human polymorphic epithelial mucin were determined, and a large domain was found to consist of a 60-base pair tandem repeat sequence. The cDNA clones were originally selected (Gendler, S. J., Burchell, J. M., Duhig, T., Lamport, D., White, R., Parker, M., and Taylor-Papadimitriou, J. (1987) Proc. Natl. Acad. Sci. U. S. A. 84, 6060-6064) using three monoclonal antibodies which show differential reactivity with the mucin produced by normal and malignant breast. Two of the epitopes are exposed in the normally processed and cancer-associated mucin, while one epitope is unmasked only in the cancer-associated mucin (Burchell, J. M., Durbin, H., and Taylor-Papadimitriou, J. (1983) J. Immunol. 131, 508-513; Burchell, J., Gendler, S., Taylor-Papadimitriou, J., Girling, A., Lewis, A., Millis, R., and Lamport, D. (1987) Cancer Res. 47, 5476-5482). We show here that all three antibodies react with a synthetic peptide with an amino acid sequence corresponding to that predicted by the tandem repeat. Identification of the epitopes preferentially expressed on the cancer-associated mucin should allow a directed approach to the development of tumor-specific antibodies using synthetic peptides as immunogens.  相似文献   

14.
Book Reviews     
M. S. Ridout 《Biometrics》2001,57(2):644-658
Books reviewed in this article:
BERNARDO, J. M., BERGER, J. O., DAWID, A. P., and SMITH, A. F. M. (editors). Bayesian Statistics 6.
VAN DER VAART, A. W. Asymptotic Statistics.
NOLAN, D. and SPEED, T. StatLabs: Mathematical Statistics Through Applications.
BINNS, M. R., NYROP, J. P., and VAN DER WERF, W. Sampling and Monitoring in Crop Protection: The Theoretical Basis for Developing Practical Decision Guides.
DIECKMANN, U., LAW, R., and METZ, J. A. J. (editors). The Geometry of Ecological Interactions: Simplifying Spatial Complexity.
LAWSON, A., BIGGERI, A., BOHNING, D., LESAFFRE, E., VIEL, 3.-F., and BERTOLLINI, R. Disease Mapping and Risk Assessment for Public Health.
BLAND, M. and PEACOCK, J. Statistical Questions in Evidence Based Medicine. Oxford University Press, Oxford, 2000.
MANLY, B. F. J. Statistics for Environmental Science and Management.
RÍOS INSUA, D. and RUGGERI, F. (editors) Robust Bayesian Analysis.
BARNDORFF-NIELSEN, 0. E., COX, D. R., and KLUP-PELBERG, C. Complex Stochastic Systems.
PETRIE, A. and SABIN, C. Medical Statistics at a Glance.
THERNEAU, T. M. and GRAMBSCH, P. M. Modeling Survival Data: Extending the Cox Model.
TAN, W.-Y. Stochastic Modeling of AIDS Epidemiology and HIV Pathogenesis.
CHATFIELD, C. Time-Series Forecasting.
MATTHEWS, J. N. S. An Introduction to Randomized Controlled Clinical Trials.
CLARK, I. and HARPER, W. V. Practical Geostatistics 2000.
TANIGUCHI, M. and KAKIZAWA, Y. Asymptotic Theory of Statistical Inference for Time Series.
KARIAN, Z. A. and DUDEWICZ, E. J. Fitting Statistical Distributions: The Generalized Lambda Distribution and Generalized Bootstrap Methods.  相似文献   

15.
《Journal of Phytopathology》1982,104(3):284-288
Book reviewed in this articles:
Bailey, J. A., and J. W. Mansfield (Eds.) , Phytoalexins.
Blakeman, J. P. (ed.) , Microbial ecology of the phylloplane.
Criigcr, G., und H. Ehle , Günstige Bcfallsbedingungen bei Prüfungen von Pungizidcn gegen Schadpilze in Gemiisekulturen.
Franz, J. M., und A. Krleg , Biologische Schadlingsbekampfung unter Beriick-sichtigung integrierter Verfahren.
Heun, A., und H. Mielke , Resistenzzüchtung gegen Pseudocercosporella her-potrichoides und Gaeumannomyces graminis beim Weizen.
Wareing, P. F., and I. D. J. Phillips , Growth and Differentiation in Plants.  相似文献   

16.
The side group of an invariant Gln in cGMP- and cAMP-specific phosphodiesterases (PDE) is held in different orientations by bonds with other amino acids and purportedly discriminates between guanine and adenine in cGMP and cAMP. In cGMP-specific PDE5, Gln(775) constrains the orientation of the invariant Gln(817) side chain, which forms bidentate bonds with 5'-GMP, vardenafil, sildenafil, and 3-isobutyl-1-methylxanthine (IBMX) (Sung, B. J., Hwang, K. Y., Jeon, Y. H., Lee, J. I., Heo, Y. S., Kim, J. H., Moon, J., Yoon, J. M., Hyun, Y. L., Kim, E., Eum, S. J., Park, S. Y., Lee, J. O., Lee, T. G., Ro, S., and Cho, J. M. (2003) Nature 425, 98-102; Huai, Q., Liu, Y., Francis, S. H., Corbin, J. D., and Ke, H. (2004) J. Biol. Chem. 279, 13095-13101; Zhang, K. Y., Card, G. L., Suzuki, Y., Artis, D. R., Fong, D., Gillette, S., Hsieh, D., Neiman, J., West, B. L., Zhang, C., Milburn, M. V., Kim, S. H., Schlessinger, J., and Bollag, G. (2004) Mol. Cell 15, 279-286). PDE5(Q817A) and PDE5(Q775A) were generated to test the hypotheses that Gln(817) is critical for cyclic nucleotide or inhibitor affinity and that Gln(775) immobilizes the Gln(817) side chain to provide cGMP/cAMP selectivity. Allosteric cGMP binding and the molecular mass of the mutant proteins were unchanged compared with PDE5(WT). For PDE5(Q817A), K(m) for cGMP or cAMP was weakened 60- or 2-fold, respectively. For PDE5(Q775A), K(m) for cGMP was weakened approximately 20-fold but was unchanged for cAMP. For PDE5(Q817A), vardenafil, sildenafil, and IBMX inhibitory potencies were weakened 610-, 48-, and 60-fold, respectively, indicating that Gln(817) is a major determinant of potency, especially for vardenafil, and that binding of vardenafil and sildenafil differs substantially. Sildenafil and vardenafil affinity were not significantly affected in PDE5(Q775A). It is concluded that Gln(817) is a positive determinant for PDE5 affinity for cGMP and several inhibitors; Gln(775), which perhaps restricts rotation of Gln(817) side chain, is critical for cGMP affinity but has no measurable effect on affinity for cAMP, sildenafil, or vardenafil.  相似文献   

17.
All eukaryotic forms of DNA topoisomerase I contain an extensive and highly charged N-terminal domain. This domain contains several nuclear localization sequences and is essential for in vivo function of the enzyme. However, so far no direct function of the N-terminal domain in the in vitro topoisomerase I reaction has been reported. In this study we have compared the in vitro activities of a truncated form of human topoisomerase I lacking amino acids 1-206 (p67) with the full-length enzyme (p91). Using these enzyme forms, we have identified for the first time a direct role of residues within the N-terminal domain in modulating topoisomerase I catalysis, as revealed by significant differences between p67 and p91 in DNA binding, cleavage, strand rotation, and ligation. A comparison with previously published studies showing no effect of deleting the first 174 or 190 amino acids of topoisomerase I (Stewart, L., Ireton, G. C., and Champoux, J. J. (1999) J. Biol. Chem. 274, 32950-32960; Bronstein, I. B., Wynne-Jones, A., Sukhanova, A., Fleury, F., Ianoul, A., Holden, J. A., Alix, A. J., Dodson, G. G., Jardillier, J. C., Nabiev, I., and Wilkinson, A. J. (1999) Anticancer Res. 19, 317-327) suggests a pivotal role of amino acids 191-206 in catalysis. Taken together the presented data indicate that at least part(s) of the N-terminal domain regulate(s) enzyme/DNA dynamics during relaxation most probably by controlling non-covalent DNA binding downstream of the cleavage site either directly or by coordinating DNA contacts by other parts of the enzyme.  相似文献   

18.
Aldehyde oxidoreductase (AOR) activity has been found in different sulfate reducing organisms (Moura, J. J. G., and Barata, B. A. S. (1994) in Methods in Enzymology (Peck, H. D., Jr., and LeGall, J., Eds.), Vol. 243, Chap. 4. Academic Press; Rom?o, M. J., Kn?blein, J., Huber, R., and Moura, J. J. G. (1997) Prog. Biophys. Mol. Biol. 68, 121-144). The enzyme was purified to homogeneity from extracts of Desulfovibrio desulfuricans (Dd) ATCC 27774, a sulfate reducer that can use sulfate or nitrate as terminal respiratory substrates. The protein (AORDd) is described as a homodimer (monomer, circa 100 kDa), contains a Mo-MCD pterin, 2 x [2Fe-2S] clusters, and lacks a flavin group. Visible and EPR spectroscopies indicate a close similarity with the AOR purified from Desulfovibrio gigas (Dg) (Barata, B. A. S., LeGall, J., and Moura, J. J. G. (1993) Biochemistry 32, 11559-11568). Activity and substrate specificity for different aldehydes were determined. EPR studies were performed in native and reduced states of the enzyme and after treatment with ethylene glycol and dithiothreitol. The AORDd was crystallized using ammonium sulfate as precipitant and the crystals belong to the space group P6(1)22, with unit cell dimensions a = b = 156.4 and c = 177.1 A. These crystals diffract to beyond 2.5 A resolution and a full data set was measured on a rotating anode generator. The data were used to solve the structure by Patterson Search methods, using the model of AORDg.  相似文献   

19.
Several clinical trials have revealed that individuals who were given beta-carotene and vitamin A did not have a reduced risk of cancer compared to those given placebo; rather, vitamin A could actually have caused an adverse effect in the lungs of smokers [Omenn, G.S., Goodman, G.E., Thornquist, M.D., Balmes, J., Cullen, M.R., Glass, A., Keogh, J.P., Meyskens, F.L., Valanis, B., Williams, J.H., Barnhart, S. & Hammar, S. N. Engl. J. Med (1996) 334, 1150-1155; Hennekens, C.H., Buring, J.E., Manson, J.E., Stampfer, M., Rosner, B., Cook, N.R., Belanger, C., LaMotte, F., Gaziano, J.M., Ridker, P.M., Willet, W. & Peto, R. (1996) N. Engl. J. Med. 334, 1145-1149]. Using differential display techniques, an initial survey using rats showed that liver RNA expression of c-H-Ras was decreased and p53 increased in rats with chronic vitamin A deficiency. These findings prompted us to evaluate the expression of c-Jun, p53 and p21WAF1/CIF1 (by RT-PCR) in liver and lung of rats. This study showed that c-Jun levels were lower and that p53 and p21WAF1/CIF1 levels were higher in chronic vitamin A deficiency. Vitamin A supplementation increased expression of c-Jun, while decreasing the expression of p53 and p21WAF1/CIF1. Western-blot analysis demonstrated that c-Jun and p53 showed a similar pattern to that found in the RT-PCR analyses. Binding of retinoic acid receptors (RAR) to the c-Jun promoter was decreased in chronic vitamin A deficiency when compared to control hepatocytes, but contrasting results were found with acute vitamin A supplementated cells. DNA fragmentation and cytochrome c release from mitochondria were analyzed and no changes were found. In lung, an increase in the expression of c-Jun produced a significant increase in cyclin D1 expression. These results may explain, at least in part, the conflicting results found in patients supplemented with vitamin A and illustrate that the changes are not restricted to lung. Furthermore, these results suggest that pharmacological vitamin A supplementation may increase the risk of adverse effects including the risk of oncogenesis.  相似文献   

20.
Aggrecanase-1 (ADAMTS-4) is a member of the a disintegrin and metalloprotease with thrombospondin motifs (ADAMTS) protein family that was recently identified. Aggrecanase-1 is one of two ADAMTS cartilage-degrading enzymes purified from interleukin-1-stimulated bovine nasal cartilage (Tortorella, M. D., Burn, T. C., Pratta, M. A. , Abbaszade, I., Hollis, J. M., Liu, R., Rosenfeld, S. A., Copeland, R. A., Decicco, C. P., Wynn, R., Rockwell, A., Yang, F., Duke, J. L., Solomon, K., George, H., Bruckner, R., Nagase, H., Itoh, Y., Ellis, D. M., Ross, H., Wiswall, B. H., Murphy, K., Hillman, M. C., Jr., Hollis, G. F., and Arner, E.C. (1999) Science 284, 1664-1666; 2 Abbaszade, I., Liu, R. Q., Yang, F., Rosenfeld, S. A., Ross, O. H., Link, J. R., Ellis, D. M., Tortorella, M. D., Pratta, M. A., Hollis, J. M., Wynn, R., Duke, J. L., George, H. J., Hillman, M. C., Jr., Murphy, K., Wiswall, B. H., Copeland, R. A., Decicco, C. P., Bruckner, R., Nagase, H., Itoh, Y., Newton, R. C., Magolda, R. L., Trzaskos, J. M., and Burn, T. C. (1999) J. Biol. Chem. 274, 23443-23450). The aggrecan products generated by this enzyme are found in cartilage cultures stimulated with cytokines and in synovial fluid from patients with arthritis, suggesting that aggrecanase-1 may be important in diseases involving cartilage destruction. Here we demonstrate that the thrombospondin type-1 (TSP-1) motif located within the C terminus of aggrecanase-1 binds to the glycosaminoglycans of aggrecan. Data from several studies indicate that this binding of aggrecanase-1 to aggrecan through the TSP-1 motif is necessary for enzymatic cleavage of aggrecan. 1) A truncated form of aggrecanase-1 lacking the TSP-1 motif was not effective in cleaving aggrecan. 2) Several peptides representing different regions of the TSP-1 motif effectively blocked aggrecanase-1 cleavage of aggrecan by preventing the enzyme from binding to the substrate. 3) Aggrecanase-1 was not effective in cleaving glycosaminoglycan-free aggrecan. Taken together, these data suggest that the TSP-1 motif of aggrecanase-1 is critical for substrate recognition and cleavage.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号