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.
The primary structure of a 61-amino-acid residue peptide from the pancreas of the European eel (Anguilla anguilla) has been established as E E K S G(5)L Y R K P(10)S C G E M(15)S A M H A(20)C P M N F(25)A P V C G(30)T D G N T(35)Y P N E C(40)S L C F Q(45)R Q N T K(50)T D I L I(55)T K D D R(60)C. There was no indication of microheterogeneity. This peptide shows structural similarity to pancreatic secretory trypsin inhibitors from several mammalian species and to a cholecystokinin-releasing peptide isolated from rat pancreatic juice. A comparison of the amino acid sequences of the peptides has identified a domain in the central region of the molecules that has been strongly conserved during evolution. In contrast, the amino acid sequence in the region corresponding to the reactive centre of the mammalian trypsin inhibitors is very poorly conserved in the eel peptide. The P1-P1' reactive site lysine-isoleucine (or arginine-isoleucine) bond in the mammalian trypsin inhibitors is replaced by a methionine-asparagine bond. This region does, however, show limited homology to the reactive centre of human alpha 1-protease inhibitor suggesting that the eel peptide may function as an inhibitor of other proteolytic enzymes in the pancreas.  相似文献   

3.
A biochromatographic approach is developed to measure for the first time changes in enthalpy, heat capacity change and protonation for the binding of nor-NOHA to arginase in a wide temperature range. For this, the arginase enzyme was immobilized on a chromatographic support. It was established that this novel arginase column was stable during an extended period of time. The affinity of nor-NOHA to arginase is high and changes slightly with the pH, because the number of protons linked to binding is low. The determination of the enthalpy change at different pH values suggested that the protonated group in the nor-NOHA-arginase complex exhibits a heat protonation of approximately -33 kJ/mol. This value agrees with the protonation of an imidazole group. Our result confirmed that active-site residue Hist 141 is protonated as imidazolium cation. Hist 141 can function as a general acid to protonate the leaving amino group of L-ornithine during catalysis. The thermodynamic data showed that nor-NOHA-arginase binding, for low temperature (<15 degrees C), is enthalpically unfavourable and being dominated by a positive entropy change. This result suggests that dehydration at the binding interface and charge-charge interactions contribute to the nor-NOHA-arginase complex formation. The temperature dependence of the free energy of binding is weak because of the enthalpy-entropy compensation caused by a large heat capacity change, DeltaC(p)=-2.43 kJ/mol/K, of arginase. Above 15 degrees C, the thermodynamic data DeltaH and DeltaS became negative due to van der Waals interactions and hydrogen bonding which are engaged at the complex interface confirming strong enzyme-inhibitor hydrogen bond networks. As well, by the use of these thermodynamic data and known correlations it was clearly demonstrated that the binding of nor-NOHA to arginase produces slight conformational changes in the vicinity of the active site. Our work indicated that our biochromatographic approach could soon become very attractive for studying other enzyme-ligand binding.  相似文献   

4.
Virtual screening, a fast, computational approach to identify drug leads [Perola, E.; Xu, K.; Kollmeyer, T. M.; Kaufmann, S. H.; Prendergast, F. G. J. Med. Chem.2000, 43, 401; Miller, M. A. Nat. Rev. Drug Disc.2002, 1 220], is limited by a known challenge in crystallographically determining flexible regions of proteins. This approach has not been able to identify active inhibitors of the severe acute respiratory syndrome-associated coronavirus (SARS-CoV) using solely the crystal structures of a SARS-CoV cysteine proteinase with a flexible loop in the active site [Yang, H. T.; Yang, M. J.; Ding, Y.; Liu, Y. W.; Lou, Z. Y. Proc. Natl. Acad. Sci. U.S.A.2003, 100, 13190; Jenwitheesuk, E.; Samudrala, R. Bioorg. Med. Chem. Lett.2003, 13, 3989; Rajnarayanan, R. V.; Dakshanamurthy, S.; Pattabiraman, N. Biochem. Biophys. Res. Commun.2004, 321, 370; Du, Q.; Wang, S.; Wei, D.; Sirois, S.; Chou, K. Anal. Biochem.2005, 337, 262; Du, Q.; Wang, S.; Zhu, Y.; Wei, D.; Guo, H. Peptides2004, 25, 1857; Lee, V.; Wittayanarakul, K.; Remsungenen, T.; Parasuk, V.; Sompornpisut, P. Science (Asia)2003, 29, 181; Toney, J.; Navas-Martin, S.; Weiss, S.; Koeller, A. J. Med. Chem.2004, 47, 1079; Zhang, X. W.; Yap, Y. L. Bioorg. Med. Chem.2004, 12, 2517]. This article demonstrates a genome-to-drug-lead approach that uses terascale computing to model flexible regions of proteins, thus permitting the utilization of genetic information to identify drug leads expeditiously. A small-molecule inhibitor of SARS-CoV, exhibiting an effective concentration (EC50) of 23 microM in cell-based assays, was identified through virtual screening against a computer-predicted model of the cysteine proteinase. Screening against two crystal structures of the same proteinase failed to identify the 23-microM inhibitor. This study suggests that terascale computing can complement crystallography, broaden the scope of virtual screening, and accelerate the development of therapeutics to treat emerging infectious diseases such as SARS and Bird Flu.  相似文献   

5.
6.
Eason PD  Imperiali B 《Biochemistry》1999,38(17):5430-5437
Recent work has resulted in the development of potent inhibitors of oligosaccharyl transferase (OT), the enzyme that catalyzes the cotranslational glycosylation of asparagine [Hendrickson, T. L., Spencer, J. R., Kato, M., and Imperiali, B. (1996) J. Am. Chem. Soc. 118, 7636-7637; Kellenberger, C., Hendrickson, T. L., and Imperiali, B. (1997) Biochemistry 36, 12554-12559]. However, no specific OT inhibitors that function in the cellular environment have yet been reported. The peptide cyclo(hex-Amb-Cys)-Thr-Val-Thr-Nph-NH2 was previously shown to exhibit nanomolar inhibition (Ki = 37 nM) through slow tight binding kinetics [Hendrickson, T. L., Spencer, J. R., Kato, M., and Imperiali, B. (1996) J. Am. Chem. Soc. 118, 7636-7637]. Included herein is the redesign of this prototype inhibitor for achieving both passive and active translocation into model membrane systems representing the endoplasmic reticulum (ER). The strategy for passive transport involved the incorporation of a membrane permeable import function previously shown to carry various peptides across the outer as well as the interior cellular membranes [Rojas, M., Donahue, J. P., Tan, Z., and Lin, Y.-Z. (1998) Nat. Biotechnol. 16, 370-375]. Assessment of function in intact ER membranes revealed that the inhibitor targeted toward passive diffusion demonstrated concentration-dependent inhibition of two different glycosylation substrates. Thus, this modified inhibitor achieved potent inhibition of glycosylation after being successfully transported through the ER membrane. In the active translocation approach, the lead OT inhibitor and a corresponding substrate were redesigned to include features recognized by the transporter associated with antigen processing (TAP). This protein translocates peptides into the lumen of the ER [Heemels, M.-T., Schumacher, T. N. M., Wonigeit, K., and Ploegh, H. L. (1993) Science 262, 2059-2063]. However, although acceptance of the cyclized substrate by the TAP receptor was demonstrated via efficient transport and glycosylation, the modified inhibitor was not translocated by TAP machinery, and therefore, active translocation was achieved for the modified substrate only. Both of these ER transport methods afforded redesigned OT inhibitors that retained their inhibitor properties in vitro, regardless of the extensions to the carboxy-terminus of the root inhibitor. The above family of redesigned inhibitors provides a template for generating a transcellular pathway and represents the first step toward OT inhibition in intact cells.  相似文献   

7.
Arginase is a binuclear Mn(2+) metalloenzyme that catalyzes the hydrolysis of L-arginine to L-ornithine and urea. X-ray crystal structures of arginase complexed to substrate analogues N(omega)-hydroxy-L-arginine and N(omega)-hydroxy-nor-L-arginine, as well as the products L-ornithine and urea, complete a set of structural "snapshots" along the reaction coordinate of arginase catalysis when interpreted along with the X-ray crystal structure of the arginase-transition-state analogue complex described in Kim et al. [Kim, N. N., Cox, J. D., Baggio, R. F., Emig, F. A., Mistry, S., Harper, S. L., Speicher, D. W., Morris, Jr., S. M., Ash, D. E., Traish, A. M., and Christianson, D. W. (2001) Biochemistry 40, 2678-2688]. Taken together, these structures render important insight on the structural determinants of tight binding inhibitors. Furthermore, we demonstrate for the first time the structural mechanistic link between arginase and NO synthase through their respective complexes with N(omega)-hydroxy-L-arginine. That N(omega)-hydroxy-L-arginine is a catalytic intermediate for NO synthase and an inhibitor of arginase reflects the reciprocal metabolic relationship between these two critical enzymes of L-arginine catabolism.  相似文献   

8.
The newly identified specific V-ATPase inhibitor, salicylihalamide A, is distinct from any previously identified V-ATPase inhibitors in that it inhibits only mammalian V-ATPases, but not those from yeast or other fungi (Boyd, M. R., Farina, C., Belfiore, P., Gagliardi, S., Kim, J. W., Hayakawa, Y., Beutler, J. A., McKee, T. C., Bowman, B. J., and Bowman, E. J. (2001) J. Pharmacol. Exp. Ther. 297, 114-120). In addition, salicylihalamide A does not compete with concanamycin or bafilomycin for binding to V-ATPase, indicating that it has a different binding site from those classic V-ATPase inhibitors (Huss, M., Ingenhorst, G., Konig, S., Gassel, M., Drose, S., Zeeck, A., Altendorf, K., and Wieczorek, H. (2002) J. Biol. Chem. 277, 40544-40548). By using purified bovine brain V-pump and its dissociated V(1) and V(0) sectors, we identified the recognition and binding site for salicylihalamide to be within the V(0) domain. Salicylihalamide does not inhibit the ATP hydrolysis activity of the dissociated V(1)-ATPase but inhibits the ATPase activity of the holoenzyme by inhibiting the V(0) domain. Salicylihalamide causes a dramatic redistribution of cytosolic V(1) from soluble to membrane-associated form, a change not observed in cells treated with either bafilomycin or NH(4)Cl. By synthesizing and characterizing a series of salicylihalamide derivatives, we investigated the structural determinants of salicylihalamide inhibition in terms of potency and reversibility, and used this information to suggest a possible binding mechanism.  相似文献   

9.
Cyclooxygenase (COX) is a key enzyme in the biosynthetic pathway leading to the formation of prostaglandins, which are mediators of inflammation [D.L. Dewitt, W.L. Smith, Primary structure of prostaglandin G/H synthase from sheep vesicular gland determined from the complementary DNA sequence, Proc. Natl. Acad. Sci. USA 85 (1988) 1412-1416, 1]. It exists mainly in two isoforms COX-1 and COX-2 [A. Raz, A. Wyche, N. Siegel, P. Needleman, Regulation of fibroblast cyclooxygenase synthesis by interleukin-1, J. Biol. Chem. 263 (1988) 3022-3028, 2]. The conventional non-steroidal anti-inflammatory drugs (NSAIDs) have adverse gastrointestinal side-effects, because they inhibit both isoforms [T.D. Warner, F. Guiliano, I. Vojnovic, A. Bukasa, J.A. Mitchell, J.P. Vane, Nonsteroid drug selectivities for cyclo-oxygenase-1 rather than cyclo-oxygenase-2 are associated with human gastrointestinal toxicity: a full in vitro analysis, Proc. Natl. Acad. Sci. USA 96 (1999) 7563-7568, 3; L.J. Marnett, A.S. Kalgutkar, Cyclooxygenase 2 inhibitors: discovery, selectivity and the future, Trends Pharmacol. Sci. 20 (1999) 465-469, 4; J.R. Vane, NSAIDs, Cox-2 inhibitors, and the gut, Lancet 346 (1995) 1105-1106, 5]. Therefore drugs which selectively inhibit COX-2, known as coxibs were developed. Recent reports on the harmful cardiovascular and renovascular side-effects of the anti-inflammatory drugs have led to the quest for a novel class of COX-2 selective inhibitors. Keeping this in mind, we have used the X-ray crystal structures of the complexes of the COX-1 and COX-2 with the known inhibitors for a rational, structure based approach to design a small peptide, which is potent inhibitor for COX-2. The peptides have been checked experimentally by in-vitro kinetic studies using surface plasmon resonance (SPR) and other biochemical methods. We have identified a tripeptide inhibitor which is a potential lead for a new class of COX-2 inhibitor. The dissociation constant (K(D)) determined for COX-2 with peptide WCS is 1.90x10(-10)M, the kinetic constant (K(i)) determined by spectrophotometry is 4.85x10(-9)M and the IC(50) value is 1.5x10(-8)M by ELISA test.  相似文献   

10.
Book Reviewed in this article:
Canning, Elizabeth U., ed. 1981. Parasitological Topics: a Presentation Volume to P. C. C. Garnham F.R.S. on the Occasion of His 80th Birthday 1981
Krylov, M. V. 1981. Piroplazmidy. [Piroplasms.]
Baker, J. R. 1982. The Biology of Parasitic Protozoa
Barriga, Omar O. 1981. The Immunology of Parasitic Infections: a Handbook for Physicians, Veterinarians, and Biologists
Beyer, T. V., Bezukladnikova, N. A., Galuzo, I. G., Konovalova, S. I. & Pak, S. M., eds. 1979. Toksoplasmidy. [The Toxoplasmids
Geltzer, Ya. G., Korganova, G. A., Mavlyanova, M. I. & Nikolyuk, V. I., eds. 1980. Pochvennye Prosteyshie. [The Soil Protozoa.] (Protozoologiya
Beyer, T. V., Kazakova, I. I., Lakhonina, G. M., Roigas, E. M. & Teras, J. H., eds. 1981. Vzaimootnosheniya Prosteyshikh s Virusami. [The Interaction between Protozoa and Viruses.] (Protozoologiya
Ogimoto, Keiji & Imai, Soichi 1981 Allas of Rumen Microbiology
Long, Peter L., ed. 1982. The Biology of the Coccidia
Lloyd, David, Poole, Robert & Edwards, Steven W. 1982. The Cell Division Cycle: Temporal Organization and Control of Cellular Growth and Reproduction
Frederick, J. F., ed. 1981. Origins and Evolution of Eukaryotic Intracellular Organelles. [Ann. N.Y. Acad. Sci.
Hayat, M. A. 1981. Fixation for Electron Microscopy
Buetow, D. E., ed. 1982. The Biology of Euglena.
Ogden, C. G. & Hedley, R. H. 1980. An Atlas of Freshwater Testate Amoebae
Parker, S. P., ed. 1982. Synopsis and Classification of Living Organisms
Margulis, L. & Schwartz, K. V. 1982. Five Kingdoms: an Illustrated Guide to the Phyla of Life on Earth
Cairns, J., Jr., ed. 1982. Artificial Substrates
Curds, C. R. 1982. British and Other Freshwater Ciliated Protozoa  相似文献   

11.
Previous results, based on inhibition of fertilization by an anti-alpha6 integrin mAb (GoH3), suggest that the alpha6beta1 integrin on mouse eggs functions as the receptor for sperm (Almeida, E.A., A.P. Huovila, A.E. Sutherland, L.E. Stephens, P.G. Calarco, L. M. Shaw, A.M. Mercurio, A. Sonnenberg, P. Primakoff, D.G. Myles, and J.M. White. 1995. Cell. 81:1095-1104). Because the egg surface tetraspanin CD9 is essential for gamete fusion (Kaji, K., S. Oda, T. Shikano, T. Ohnuki, Y. Uematsu, J. Sakagami, N. Tada, S. Miyazaki, and A. Kudo. 2000. Nat. Genet. 24:279-282; Le Naour, F., E. Rubinstein, C. Jasmin, M. Prenant, and C. Boucheix. 2000. Science. 287:319-321; Miyado, K., G. Yamada, S. Yamada, H. Hasuwa, Y. Nakamura, F. Ryu, K. Suzuki, K. Kosai, K. Inoue, A. Ogura, M. Okabe, and E. Mekada. 2000. Science. 287:321-324) and CD9 is known to associate with integrins, recent models of gamete fusion have posited that egg CD9 acts in association with alpha6beta1 in fusion (Chen, M.S., K.S. Tung, S.A. Coonrod, Y. Takahashi, D. Bigler, A. Chang, Y. Yamashita, P.W. Kincade, J.C. Herr, and J.M. White. 1999. Proc. Natl. Acad. Sci. USA. 96:11830-11835; Kaji, K., S. Oda, T. Shikano, T. Ohnuki, Y. Uematsu, J. Sakagami, N. Tada, S. Miyazaki, and A. Kudo. 2000. Nat. Genet. 24:279-282; Le Naour, F., E. Rubinstein, C. Jasmin, M. Prenant, and C. Boucheix. 2000. Science. 287:319-321; Miyado, K., G. Yamada, S. Yamada, H. Hasuwa, Y. Nakamura, F. Ryu, K. Su- zuki, K. Kosai, K. Inoue, A. Ogura, M. Okabe, and E. Mekada. 2000. Science. 287:321-324). Using eggs from cultured ovaries of mice lacking the alpha6 integrin subunit, we found that the fertilization rate, fertilization index, and sperm binding were not impaired compared with wild-type or heterozygous controls. Furthermore, a reexamination of antibody inhibition, using an assay that better simulates in vivo fertilization conditions, revealed no inhibition of fusion by the GoH3 mAb. We also found that an anti-CD9 mAb completely blocks sperm fusion with either wild-type eggs or eggs lacking alpha6beta1. Based on these results, we conclude that the alpha6beta1 integrin is not essential for sperm-egg fusion, and we suggest a new model in which CD9 acts by itself, or interacts with egg protein(s) other than alpha6beta1, to function in sperm-egg fusion.  相似文献   

12.
13.
The present paper describes the structures of the N-linked oligosaccharides of the human-immunodeficiency-virus (HIV) envelope glycoprotein gp120 (cloned from the HTLV-III B isolate and expressed as a secreted fusion protein after transfection of Chinese-hamster ovary cells), which is known to bind with high affinity to human T4-lymphocytes. Oligosaccharides were released from peptide by hydrazinolysis, fractionated by paper electrophoresis, high-performance lectin-affinity chromatography and Bio-Gel P-4 column chromatography, and their structures determined by sequential exoglycosidase digestions in conjunction with methylation analysis. The glycoprotein was found to be unique in its diversity of oligosaccharide structures. These include high-mannose type and hybrid type, as well as four categories of complex-type chains: mono-, bi-, tri- and tetra-antennary, with or without N-acetyl-lactosamine repeats, and with or without a core-region fucose residue. Among the sialidase-treated oligosaccharides, no less than 29 structures were identified as follows: (formula; see text) where G is galactose, GN is N-acetylglucosamine, M is mannose, F is fucose, and '+/- ' means that residues are present in a proportion of chains. The actual number of oligosaccharide structures is much greater, since before desialylation there was evidence that, among the hybrid and complex-type chains, all but 6% contained sialic acid at the C-3 position of terminal galactose residues, and partially sialylated forms of the bi- and multi-antennary chains were present. Detailed evidence for the proposed oligosaccharide sequences will be published as a supplementary paper [T. Mizuochi, M. W. Spellman, M. Larkin, J. Solomon, L. J. Basa & T. Feizi (1988) Biomed. Chromatogr., in the press].  相似文献   

14.
Janc JW  Clark JM  Warne RL  Elrod KC  Katz BA  Moore WR 《Biochemistry》2000,39(16):4792-4800
Serine proteases play a role in a variety of disease states and thus are attractive targets for therapeutic intervention. We report the kinetic characterization of a class of serine protease inhibitors whose potencies and selectivities are dramatically enhanced in the presence of Zn(II). The structural basis for Zn(II)-mediated inhibition of trypsin-like proteases has recently been reported [Katz, B. A., Clark, J. M., Finer-Moore, J. S., Jenkins, T. E., Johnson, C. R., Ross, M. J., Luong, C., Moore, W. R., and Stroud, R. M. (1998) Nature 391, 608-612]. A case study of the kinetic behavior of human tryptase inhibitors is provided to illustrate the general phenomenon of Zn(II)-mediated inhibition. Tryptase, Zn(II), and the inhibitor form a ternary complex which exhibits classic tight-binding inhibition. The half-life for release of inhibitor from the tryptase-Zn(II)-inhibitor complex has been measured for a number of inhibitors. Consistent with tight-binding behavior, potent tryptase inhibitors are characterized by extremely slow rates of dissociation from the ternary complex with half-lives on the order of hours. A model of human serum, designed to reproduce physiological levels of Zn(II), has been employed to evaluate the performance of Zn(II)-potentiated tryptase inhibitors under physiological conditions. We demonstrate that Zn(II)-mediated inhibition can be achieved at physiological Zn(II) levels.  相似文献   

15.
In analogy with studies previously reported for myeloperoxidase (Kooter, I. M.; Moguilevsky, N.; Bollen, A.; Van der Veen, L. A.; Otto, C.; Dekker, H. L.; Wever, R. J. Biol. Chem. 1999, 274, 26794), we examined for bovine lactoperoxidase the effect of mutation of Asp225 and Glu375, the residues thought to be responsible for the covalent binding of the heme group to the apoprotein. Starting from the plasmid encoding rbLPO (Watanabe, S.; Varsalona, F.; Yoo, Y.; Guillaume, J. P.; Bollen, A.; Shimazaki, K.; Moguilevsky, N. FEBS Letters 1998, 441, 476), which was engineered to carry mutations in correspondence of those residues, the mutants Asp225Val and Glu375Gln were expressed in CHO cells and their products purified and characterized. Unequivocal evidence about the existence of ester linkages as well as their relative contribution to the specific spectroscopic and catalytic properties of bLPO is here discussed.  相似文献   

16.
Inappropriate activity of p90 ribosomal S6 kinase (RSK) has been implicated in various human cancers as well as other pathologies. We previously reported the isolation, characterization, and synthesis of the natural product kaempferol 3-O-(3',4'-di-O-acetyl-alpha-l-rhamnopyranoside), termed SL0101 [Smith, J. A.; Poteet-Smith, C. E.; Xu, Y.; Errington, T. M.; Hecht, S. M.; Lannigan, D. A. Cancer Res., 2005, 65, 1027-1034: Xu, Y.-M; Smith, J. A.; Lannigan, D. A.; Hecht, S. M. Bioorg. Med. Chem., 2006, 14, 3974-3977: Maloney, D. J.; Hecht, S. M. Org. Lett., 2005, 7, 1097-1099]. SL0101 is a potent and specific inhibitor of RSK; therefore, we performed an analysis of the structural basis for the inhibitory activity of this lead compound. In in vitro kinase assays we found that acylation of the rhamnose moiety and the 4', 5, and 7-hydroxyl groups are responsible for maintaining a high affinity interaction of RSK with SL0101. It is likely that the hydroxyl groups facilitate RSK binding through their ability to form hydrogen bonds. To determine whether the SL0101 derivatives were specific for inhibition of RSK we analyzed their ability to preferentially inhibit the growth of the human breast cancer line, MCF-7, compared to the normal human breast line, MCF-10A. We have previously validated this differential growth assay as a convenient readout for analyzing the specificity of RSK inhibitors [Smith, J. A.; Maloney, D. J.; Clark, D. E.; Xu, Y.-M.; Hecht, S. M.; Lannigan, D. A. Bioorg. Med. Chem., 2006, 14, 6034-6042]. We found that acylation of the rhamnose moiety was essential for maintaining the selectivity for RSK inhibition in intact cells. Further, the efficacy of SL0101 in intact cells is limited by cellular uptake as well as possible hydrolysis of the acetyl groups on the rhamnose moiety by ubiquitous intracellular esterases. These studies should facilitate the development of a RSK inhibitor, based on the SL0101 pharmacophore, as an anti-cancer chemotherapeutic agent.  相似文献   

17.
Arginase is a binuclear manganese metalloenzyme that catalyzes the hydrolysis of l-arginine to form l-ornithine and urea. The X-ray crystal structure of a fully active, truncated form of human arginase II complexed with a boronic acid transition state analogue inhibitor has been determined at 2.7 A resolution. This structure is consistent with the hydrolysis of l-arginine through a metal-activated hydroxide mechanism. Given that human arginase II appears to play a role in regulating l-arginine bioavailability to NO synthase in human penile corpus cavernosum smooth muscle, the inhibition of human arginase II is a potential new strategy for the treatment of erectile dysfunction [Kim, N. N., Cox, J. D., Baggio, R. F., Emig, F. A., Mistry, S., Harper, S. L., Speicher, D. W., Morris, S. M., Ash, D. E., Traish, A. M., and Christianson, D. W. (2001) Biochemistry 40, 2678-2688]. Since NO synthase is found in human clitoral corpus cavernosum and vagina, we hypothesized that human arginase II is similarly present in these tissues and functions to regulate l-arginine bioavailability to NO synthase. Accordingly, hemodynamic studies conducted with a boronic acid arginase inhibitor in vivo are summarized, suggesting that the extrahepatic arginase plays a role in both male and female sexual arousal. Therefore, arginase II is a potential target for the treatment of male and female sexual arousal disorders.  相似文献   

18.
Vagin O  Denevich S  Munson K  Sachs G 《Biochemistry》2002,41(42):12755-12762
Inhibition of the gastric H,K-ATPase by the imidazo[1,2-alpha]pyridine, SCH28080, is strictly competitive with respect to K+ or its surrogate, NH4+. The inhibitory kinetics [V(max), K(m,app)(NH4+), K(i)(SCH28080), and competitive, mixed, or noncompetitive] of mutants can define the inhibitor binding domain and the route to the ion binding region within M4-6. While mutations Y799F, Y802F, I803L, S806N, V807I (M5), L811V (M5-6), Y928H (M8), and Q905N (M7-8) had no effect on inhibitor kinetics, mutations P798C, Y802L, P810A, P810G, C813A or -S, I814V or -F, F818C, T823V (M5, M5-6, and M6), E914Q, F917Y, G918E, T929L, and F932L (M7-8 and M8) reduced the affinity for SCH28080 up to 10-fold without affecting the nature of the kinetics. In contrast, the L809F substitution in the loop between M5 and M6 resulted in an approximately 100-fold decrease in inhibitor affinity, and substitutions L809V, I816L, Y925F, and M937V (M5-6, M6, and M8) reduced the inhibitor affinity by 10-fold, all resulting in noncompetitive kinetics. The mutants L811F, Y922I, and I940A also reduced the inhibitor affinity up to 10-fold but resulted in mixed inhibition. The mutations I819L, Q923V, and Y925A also gave mixed inhibition but without a change in inhibitor affinity. These data, and the 9-fold loss of SCH28080 affinity in the C813T mutant, suggest that the binding domain for SCH28080 contains the surface between L809 in the M5-6 loop and C813 at the luminal end of M6, approximately two helical turns down from the ion binding region, where it blocks the normal ion access pathway. On the basis of a model of the Ca-ATPase in the E2 conformation (PDB entry 1kju), the mutants that change the nature of the kinetics are arranged on one side of M8 and on the adjacent side of the M5-6 loop and M6 itself. This suggests that mutations in this region modify the enzyme structure so that K+ can access the ion binding domain even with SCH28080 bound.  相似文献   

19.
关于顶生金花茶分类地位的探讨   总被引:1,自引:0,他引:1  
戴月  薛跃规 《广西植物》2007,27(6):840-843
根据近年来对顶生金花茶和平果金花茶在其原产地的进一步调查和野外观察并参考过去的研究成果,发现:顶生金花茶与平果金花茶在分布、形态特征、物候期、生态习性及细胞学特性,尤其是繁殖器官和生活史等方面都有明显差异。因此,认为将其降为平果金花茶的变种颇为不妥,主张恢复顶生金花茶作为一个独立种的地位,其名称为顶生金花茶。  相似文献   

20.
Enhanced vascular arginase activity impairs endothelium-dependent vasorelaxation by decreasing l-arginine availability to endothelial nitric oxide (NO) synthase, thereby reducing NO production. Elevated angiotensin II (ANG II) is a key component of endothelial dysfunction in many cardiovascular diseases and has been linked to elevated arginase activity. We determined signaling mechanisms by which ANG II increases endothelial arginase function. Results show that ANG II (0.1 μM, 24 h) elevates arginase activity and arginase I expression in bovine aortic endothelial cells (BAECs) and decreases NO production. These effects are prevented by the arginase inhibitor BEC (100 μM). Blockade of ANG II AT(1) receptors or transfection with small interfering RNA (siRNA) for Gα12 and Gα13 also prevents ANG II-induced elevation of arginase activity, but siRNA for Gαq does not. ANG II also elevates active RhoA levels and induces phosphorylation of p38 MAPK. Inhibitors of RhoA activation (simvastatin, 0.1 μM) or Rho kinase (ROCK) (Y-27632, 10 μM; H1152, 0.5 μM) block both ANG II-induced elevation of arginase activity and phosphorylation of p38 MAPK. Furthermore, pretreatment of BAECs with p38 inhibitor SB-202190 (2 μM) or transfection with p38 MAPK siRNA prevents ANG II-induced increased arginase activity/expression and maintains NO production. Additionally, inhibitors of p38 MAPK (SB-203580, 5 μg·kg(-1)·day(-1)) or arginase (ABH, 8 mg·kg(-1)·day(-1)) or arginase gene knockout in mice prevents ANG II-induced vascular endothelial dysfunction and associated enhancement of arginase. These results indicate that ANG II increases endothelial arginase activity/expression through Gα12/13 G proteins coupled to AT(1) receptors and subsequent activation of RhoA/ROCK/p38 MAPK pathways leading to endothelial dysfunction.  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