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 78 毫秒
1.
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   

2.
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   

3.
4.
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   

5.
Ca transients measured in failing human ventricular myocytes exhibit reduced amplitude and slowed relaxation [Beuckelmann, D.J., Nabauer, M., Erdmann, E., 1992. Intracellular calcium handling in isolated ventricular myocytes from patients with terminal heart failure. Circulation 85, 1046-1055; Gwathmey, J.K., Copelas, L., MacKinnon, R., Schoen, F.J., Feldman, M.D., Grossman, W., Morgan, J.P., 1987. Abnormal intracellular calcium handling in myocardium from patients with end-stage heart failure. Circ. Res. 61, 70-76; Kaab, S., Nuss, H. B., Chiamvimonvat, N., O'Rourke, B., Pak, P.H., Kass, D.A., Marban, E., Tomaselli, G.F., 1996. Ionic mechanism of action potential prolongation in ventricular myocytes from dogs with pacing-induced heart failure. Circ. Res. 78(2); Li, H.G., Jones, D.L., Yee, R., Klein, G.J., 1992. Electrophysiologic substrate associated with pacing-induced hert failure in dogs: potential value of programmed stimulation in predicting sudden death. J. Am. Coll. Cardiol. 19(2), 444-449; Vermeulen, J.T., McGuire, M.A., Opthof, T., Colonel, R., Bakker, J.M.T.d., Klopping, C., Janse, M.J., 1994. Triggered activity and automaticity in ventricular trabeculae of failing human and rabbit hearts. Cardiovasc. Res. 28, 1547-1554.] and blunted frequency dependence [Davies, C.H., Davia, K., Bennett, J.G., Pepper, J.R., Poole-Wilson, P.A., Harding, S.E., 1995. Reduced contraction and altered frequency response of isolated ventricular myocytes from patients with heart failure. Circulation, 92, 2540-2549; Hasenfuss, G., Reinecke, H., Studer, R., Meyer, M., Pieske, B., Holtz, J., Holubarsch, C., Posival, H., Just, H., Drexler, H., 1994. Relation between myocardial function and expression of sarcoplasmic reticulum Ca-ATPase in failing and nonfailing human myocardium. Circ. Res. 75, 434-442; Hasenfuss, G., Reinecke, H., Studer, R., Pieske, B., Meyer, M., Drexler, H., Just, H., 1996. Calcium cycling proteins and force-frequency relationships in heart failure. Basic Res. Cardiol. 91, 17-22; Monte, F.D., O'Gara, P., Poole-Wilson, P.A., Yacoub, M., Harding, S.E., 1995. Cell geometry and contractile abnormalities of myocytes from failing human left ventricle. Cardiovasc. Res. 30, 281-290; Philips, P.J., Gwathmey, J.K., Feldman, M.D., Schoen, F.J., Grossman, W., Morgan, J.P., 1990. Post-extrasystolic potentiation and the force-frequency relationships: differential augmentation of myocardial contractility in working myocardium from patients with end-stage heart failure. J. Mol. Cell. Cardiol. 22, 99-110; Pieske, B., Hasenfuss, G., Holubarsch, C., Schwinger, R., Bohm, M., Just, H., 1992. Alterations of the force-frequency relationship in the failing human heart depend on the underlying cardiac disease. Basic Res. Cardiol. 87, 213-221.]. Analyses of protein levels in these failing hearts reveal that the SR Ca-ATPase is down-regulated on average by 50% and that the Na/Ca exchanger is up-regulated on average by a factor of two. In this paper, we test the hypothesis that this altered pattern of expression of Ca handling proteins is sufficient to account for changes in excitation-contraction coupling properties measured experimentally at the cellular level. To do this, we present an integrated model of excitation-contraction coupling in the guinea pig ventricular cell. The model is used to determine the effects of SR Ca-ATPase down-regulation and Na/Ca exchanger up-regulation on action potential duration, Ca transient shape and amplitude, and isometric force. Model analyses demonstrate that changes in Ca handling proteins play a direct and critical role in prolongation of action potential duration, and in reduction of contractile force in heart failure.  相似文献   

6.
L Powers  B M Kincaid 《Biochemistry》1989,28(10):4461-4468
Differences in the methods of analysis of X-ray absorption data used by Powers et al. [Powers, L., Blumberg, W. E., Chance, B., Barlow, C., Leigh, J., Jr., Smith, J., Yonetani, T., Vik, S., & Peisach, J. (1979) Biochim. Biophys. Acta 547, 520-538; Powers, L., Chance, B., Ching, Y., & Angiolillo, P. (1981) Biophys. J. 34, 465-498] and Scott et al. [Scott, R., Schwartz, J., & Cramer S. (1986) Biochemistry 25, 5546-5555] are clarified. In addition, we compare the X-ray absorption data and results for resting cytochrome c oxidase reported by both groups using the same analysis method and conclude apart from any assumptions that the data are not identical.  相似文献   

7.
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   

8.
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   

9.
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   

10.
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   

11.
Morillas-Marquez, F., Martin-Sanchez, J., Acedo-Sanchez, C., Pineda, J. A., Macias, J., and Sanjuan-Garcia, J. Leishmania infantum (Protozoa, kinetoplastida): Transmission from infected patients to experimental animal under conditions that simulate needle-sharing. Experimental Parasitologym100, 71-74.  相似文献   

12.
Learning module networks from genome-wide location and expression data   总被引:6,自引:0,他引:6  
Xu X  Wang L  Ding D 《FEBS letters》2004,578(3):297-304
  相似文献   

13.
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   

14.
ARF-mediated hydrolysis of GTP has been demonstrated to regulate coat disassembly of Golgi-derived COPI transport vesicles (Tanigawa, G., Orci, L., Amherdt, M., Ravazzola, M., Helms, J.B. and Rothman, J.E. (1993) J. Cell Biol. 123, 1365-1371). In addition, a requirement for GTP hydrolysis at an early stage of COPI vesicle biogenesis has been established since cargo uptake is impaired in the presence of GTPgammaS (Nickel, W., Malsam, J., Gorgas, K., Ravazzola, M., Jenne, N., Helms, J.B. and Wieland, F.T. (1998) J. Cell Sci. 111, 3081-3090), a non-hydrolyzable analogue of GTP. We now demonstrate that the GTPase involved in the regulation of cargo uptake is ARF, revealing a multi-functional role of this GTPase in COPI-mediated vesicular transport. The molecular mechanism of cargo uptake as well as the functional implications of these findings on the overall process of COPI vesicle biogenesis are discussed.  相似文献   

15.
Chemical, genetic, and structural studies have defined a critical role for Asp-49 in the calcium-mediated activation of extracellular phospholipases A2 (PLA2). In 1984, a new class of PLA2 was isolated in which this invariant aspartate was replaced with a lysine (Maragnore, J.M., Merutka, G., Cho, W., Welches, W., Kezdy, F.J., and Heinrikson, R.L. (1984) J. Biol. Chem. 259, 13839-13843; Maragnore, J.M., and Heinrikson, R.L. (1986) J. Biol. Chem. 261, 4797-4804). The enzymatic activity of Lys-49 PLA2s has been questioned based on biochemical, mutational, and structural studies (van den Bergh, C.J., Slotboom, A.J., Verheij, H.M., and de Haas, G.H. (1988) Eur. J. Biochem. 176, 353-357). In this paper, we describe the structures of two crystal forms of the Lys-49 PLA2 isolated from the venom of Agkistridon piscivorus piscivorus. The refined models, along with complementary biochemical analysis, clarify the structural basis for the enzymatic inactivity of Lys-49 proteins.  相似文献   

16.
We have developed an assay based on rice embryogenic callus for rapid functional characterization of metabolic genes. We validated the assay using a selection of well‐characterized genes with known functions in the carotenoid biosynthesis pathway, allowing rapid visual screening of callus phenotypes based on tissue color. We then used the system to identify the functions of two uncharacterized genes: a chemically synthesized β–carotene ketolase gene optimized for maize codon usage, and a wild‐type Arabidopsis thaliana ortholog of the cauliflower Orange gene. In contrast to previous reports (Lopez, A.B., Van Eck, J., Conlin, B.J., Paolillo, D.J., O'Neill, J. and Li, L. ( 2008 ) J. Exp. Bot. 59, 213–223; Lu, S., Van Eck, J., Zhou, X., Lopez, A.B., O'Halloran, D.M., Cosman, K.M., Conlin, B.J., Paolillo, D.J., Garvin, D.F., Vrebalov, J., Kochian, L.V., Küpper, H., Earle, E.D., Cao, J. and Li, L. ( 2006 ) Plant Cell 18, 3594–3605), we found that the wild‐type Orange allele was sufficient to induce chromoplast differentiation. We also found that chromoplast differentiation was induced by increasing the availability of precursors and thus driving flux through the pathway, even in the absence of Orange. Remarkably, we found that diverse endosperm‐specific promoters were highly active in rice callus despite their restricted activity in mature plants. Our callus system provides a unique opportunity to predict the effect of metabolic engineering in complex pathways, and provides a starting point for quantitative modeling and the rational design of engineering strategies using synthetic biology. We discuss the impact of our data on analysis and engineering of the carotenoid biosynthesis pathway.  相似文献   

17.
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   

18.
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   

19.
20.
Agonist-elicited receptor sequestration is strikingly different for the alpha(2A)- versus alpha(2B)-adrenergic receptor (alpha(2)-AR) subtypes; the alpha(2B)-AR undergoes rapid and extensive disappearance from the HEK 293 cell surface, whereas the alpha(2A)-AR does not (Daunt, D. A., Hurt, C., Hein, L., Kallio, J., Feng, F., and Kobilka, B. K. (1997) Mol. Pharmacol. 51, 711-720; Eason, M. G., and Liggett, S. B. (1992) J. Biol. Chem. 267, 25473-25479). Since recent reports suggest that endocytosis is required for some G protein-coupled receptors to stimulate the mitogen-activated protein (MAP) kinase cascade (Daaka, Y., Luttrell, L. M., Ahn, S., Della Rocca, G. J., Ferguson, S. S., Caron, M. G., and Lefkowitz, R. J. (1998) J. Biol. Chem. 273, 685-688; Luttrell, L. M., Daaka, Y., Della Rocca, G. J., and Lefkowitz, R. J. (1997) J. Biol. Chem. 272, 31648-31656; Ignatova, E. G., Belcheva, M. M., Bohn, L. M., Neuman, M. C., and Coscia, C. J. (1999) J. Neurosci. 19, 56-63), we evaluated the differential ability of these two subtypes to activate MAP kinase. We observed no correlation between subtype-dependent agonist-elicited receptor redistribution and receptor activation of the MAP kinase cascade. Furthermore, incubation of cells with K(+)-depleted medium eliminated alpha(2B)-AR internalization but did not eliminate MAP kinase activation, suggesting that receptor internalization is not a general prerequisite for activation of the MAP kinase cascade via G(i)-coupled receptors. We also noted that neither dominant negative dynamin (K44A) nor concanavalin A treatment dramatically altered MAP kinase activation or receptor redistribution, indicating that these experimental tools do not universally block G protein-coupled receptor internalization.  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