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 734 毫秒
1.
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   

2.
Ricin A chain is an N-glycosidase which removes a single adenine base from a conservative loop of 28S rRNA, thereby inactivating eukaryotic ribosomes. The mechanism of action has been proposed to include transition-state stabilization of an oxycarbonium ion on the substrate ribose by interaction with Glu 177. Conversion of Glu 177 to Gln reduces activity nearly 200-fold [Ready, M. P., Kim, Y., & Robertus, J. D. (1991) Proteins: Struct., Funct., Genet. 10, 270-278] while conversion to Ala (E177A) reduces activity only 20-fold [Schlossman, D., Withers, D., Welsh, P., Alexander, A., Robertus, J., & Frankel, A. (1989) Mol. Cell. Biol. 9, 5012-5021]. X-ray analysis of the latter mutant protein shows that a residue at the edge of the active site, Glu 208, rotates into the space left vacant by the mutation. Its rearranged carboxylate partially substitutes for that of Glu 177. This is equivalent to the rescue of enzyme activity by a second-site reversion. Kinetic analysis shows the E177A mutation affects kcat and not Km, consistent with the notion that the carboxylate serves in transition-state stabilization.  相似文献   

3.
The reaction between cytochrome c (Cc) and Rhodobacter sphaeroides cytochrome c oxidase (CcO) was studied using a cytochrome c derivative labeled with ruthenium trisbipyridine at lysine 55 (Ru-55-Cc). Flash photolysis of a 1:1 complex between Ru-55-Cc and CcO at low ionic strength results in electron transfer from photoreduced heme c to Cu(A) with an intracomplex rate constant of k(a) = 4 x 10(4) s(-1), followed by electron transfer from Cu(A) to heme a with a rate constant of k(b) = 9 x 10(4) s(-1). The effects of CcO surface mutations on the kinetics follow the order D214N > E157Q > E148Q > D195N > D151N/E152Q approximately D188N/E189Q approximately wild type, indicating that the acidic residues Asp(214), Glu(157), Glu(148), and Asp(195) on subunit II interact electrostatically with the lysines surrounding the heme crevice of Cc. Mutating the highly conserved tryptophan residue, Trp(143), to Phe or Ala decreased the intracomplex electron transfer rate constant k(a) by 450- and 1200-fold, respectively, without affecting the dissociation constant K(D). It therefore appears that the indole ring of Trp(143) mediates electron transfer from the heme group of Cc to Cu(A). These results are consistent with steady-state kinetic results (Zhen, Y., Hoganson, C. W., Babcock, G. T., and Ferguson-Miller, S. (1999) J. Biol. Chem. 274, 38032-38041) and a computational docking analysis (Roberts, V. A., and Pique, M. E. (1999) J. Biol. Chem. 274, 38051-38060).  相似文献   

4.
The present work describes the dynamics of the apo form of cytochrome b(562), a small soluble protein consisting of 106 amino acid residues [Itagaki, E., and Hager, L. P. (1966) J. Biol. Chem. 241, 3687-3695]. The presence of exchange in the millisecond time scale is demonstrated for the last part of helix IV (residues 95-105 in the holo form). The chemical shift index analysis [Wishart, D. S., and Sykes, B. D. (1994) J. Biomol. NMR 4, 171-180] based on H(alpha), C(alpha), C(beta), and C' chemical shifts suggests a larger helical content than shown in the NMR structure based on NOEs. These results indicate the presence of helical-like conformations participating in the exchange process. This hypothesis is consistent with amide deuterium exchange rates and the presence of some hydrogen bonds identified from amide chemical shift temperature coefficients [Baxter, N. J., and Williamson, M. P. (1997) J. Biomol. NMR 9, 359-369]. (15)N relaxation indicates limited mobility for the amide protons of this part of the helix in the picosecond time scale. A 30 ns stochastic dynamics simulation shows small fluctuations around the helical conformation on this time scale. These fluctuations, however, do not result in a significant decrease of the calculated order parameters which are consistent with the experimental (15)N relaxation data. These results resolve an apparent discrepancy in the NMR structures between the disorder observed in helix IV due to a lack of NOEs and the secondary structure predictions based on H(alpha) chemical shifts [Feng, Y., Wand, A. J., and Sligar, S. G. (1994) Struct. Biol. 1, 30-35].  相似文献   

5.
The role of the conserved histidine-187 located in the leucine intercalation loop of Escherichia coli uracil-DNA glycosylase (Ung) was investigated. Using site-directed mutagenesis, an Ung H187D mutant protein was created, overproduced, purified to apparent homogeneity, and characterized in comparison to wild-type Ung. The properties of Ung H187D differed from Ung with respect to specific activity, substrate specificity, DNA binding, pH optimum, and inhibition by uracil analogues. Ung H187D exhibited a 55000-fold lower specific activity and a shift in pH optimum from pH 8.0 to 7.0. Under reaction conditions optimal for wild-type Ung (pH 8.0), the substrate preference of Ung H187D on defined single- and double-stranded oligonucleotides (25-mers) containing a site-specific uracil target was U/G-25-mer > U-25-mer > U/A-25-mer. However, Ung H187D processed these same DNA substrates at comparable rates at pH 7.0 and the activity was stimulated approximately 3-fold relative to the U-25-mer substrate. Ung H187D was less susceptible than Ung to inhibition by uracil, 6-amino uracil, and 5-fluorouracil. Using UV-catalyzed protein/DNA cross-linking to measure DNA binding affinity, the efficiency of Ung H187D binding to thymine-, uracil-, and apyrimidinic-site-containing DNA was (dT20) = (dT19-U) >/= (dT19-AP). Comparative analysis of the biochemical properties and the X-ray crystallographic structures of Ung and Ung H187D [Putnam, C. D., Shroyer, M. J. N., Lundquist, A. J., Mol, C. D., Arvai, A. S., Mosbaugh, D. W., and Tainer, J. A. (1999) J. Mol. Biol. 287, 331-346] provided insight regarding the role of His-187 in the catalytic mechanism of glycosylic bond cleavage. A novel mechanism is proposed wherein the developing negative charge on the uracil ring and concomitant polarization of the N1-C1' bond is sustained by resonance effects and hydrogen bonding involving the imidazole side chain of His-187.  相似文献   

6.
Wessel PM  Graciet E  Douce R  Dumas R 《Biochemistry》2000,39(49):15136-15143
A three-dimensional structure comparison between the dimeric regulatory serine-binding domain of Escherichia coli D-3-phosphoglycerate dehydrogenase [Schuller, D. J., Grant, G. A., and Banaszak, L. J. (1995) Nat. Struct. Biol. 2, 69-76] and the regulatory domain of E. coli threonine deaminase [Gallagher, D. T., Gilliland, G. L., Xiao, G., Zondlo, J., Fisher, K. E., Chinchilla, D. , and Eisenstein, E. (1998) Structure 6, 465-475] led us to make the hypothesis that threonine deaminase could have two binding sites per monomer. To test this hypothesis about the corresponding plant enzyme, site-directed mutagenesis was carried out on the recombinant Arabidopsis thaliana threonine deaminase. Kinetic and binding experiments demonstrated for the first time that each regulatory domain of the monomers of A. thaliana threonine deaminase possesses two different effector-binding sites constituted in part by Y449 and Y543. Our results demonstrate that Y449 belongs to a high-affinity binding site whose interaction with a first isoleucine induces conformational modifications yielding a conformer displaying a higher activity and with enhanced ability to bind a second isoleucine on a lower-affinity binding site containing Y543. Isoleucine interaction with this latter binding site is responsible for conformational modifications leading to final inhibition of the enzyme. Y449 interacts with both regulators, isoleucine and valine. However, interaction of valine with the high-affinity binding site induces different conformational modifications leading to reversal of isoleucine binding and reversal of inhibition.  相似文献   

7.
J H Shim  S J Benkovic 《Biochemistry》1999,38(31):10024-10031
Site-directed mutagenesis followed by studies of the pH dependence of the kinetic parameters of the mutants has been used to probe the role of the active site residues and loops in catalysis by glycinamide ribonucleotide transformylase (EC 2.1.2.2). The analysis of the mutants of the strictly conserved active site residues, His108 and Asp144, revealed that His108 acts in a salt bridge with Asp144 as a general acid catalyst with a pK(a) value of 9.7. Asp144 also plays a key role in the preparation of the active site geometry for catalysis. The rate-limiting step in the pH range of 6-10 appears to be the catalytic steps involving tetrahedral intermediates, supported by the observation of a pL (L being H or D)-independent solvent deuterium isotope effect of 2. The ionization of the amino group of glycinamide ribonucleotide both as a free and as a bound form dominates the kinetic behavior at low pH. The analysis of a mutation, H121Q, within the loop spanning amino acids 111-131 suggests the closure of the loop is involved in the binding of the substrate. The kinetic behavior parallels pH effects revealed by a series of X-ray crystallographic structures of the apoenzyme and inhibitor-bound enzyme [Su, Y., Yamashita, M. M., Greasley, S. E. , Mullen, C. A., Shim, J. H., Jennings, P. A., Benkovic, S. J., and Wilson, I. A. (1998) J. Mol. Biol. 281, 485-499], permitting a more exact formulation of the probable catalytic mechanism.  相似文献   

8.
We have previously shown that Phe(120), Glu(216), and Asp(301) in the active site of cytochrome P450 2D6 (CYP2D6) play a key role in substrate recognition by this important drug-metabolizing enzyme (Paine, M. J., McLaughlin, L. A., Flanagan, J. U., Kemp, C. A., Sutcliffe, M. J., Roberts, G. C., and Wolf, C. R. (2003) J. Biol. Chem. 278, 4021-4027 and Flanagan, J. U., Maréchal, J.-D., Ward, R., Kemp, C. A., McLaughlin, L. A., Sutcliffe, M. J., Roberts, G. C., Paine, M. J., and Wolf, C. R. (2004) Biochem. J. 380, 353-360). We have now examined the effect of mutations of these residues on interactions of the enzyme with the prototypical CYP2D6 inhibitor, quinidine. Abolition of the negative charge at either or both residues 216 and 301 decreased quinidine inhibition of bufuralol 1'-hydroxylation and dextromethorphan O-demethylation by at least 100-fold. The apparent dissociation constants (K(d)) for quinidine binding to the wild-type enzyme and the E216D and D301E mutants were 0.25-0.50 microm. The amide substitution of Glu(216) or Asp(301) resulted in 30-64-fold increases in the K(d) for quinidine. The double mutant E216Q/D301Q showed the largest decrease in quinidine affinity, with a K(d) of 65 microm. Alanine substitution of Phe(120), Phe(481),or Phe(483) had only a minor effect on the inhibition of bufuralol 1'-hydroxylation and dextromethorphan O-demethylation and on binding. In contrast to the wild-type enzyme, a number of the mutants studied were found to be able to metabolize quinidine. E216F produced O-demethylated quinidine, and F120A and E216Q/D301Q produced both O-demethylated quinidine and 3-hydroxyquinidine metabolites. Homology modeling and molecular docking were used to predict the modes of quinidine binding to the wild-type and mutant enzymes; these were able to rationalize the experimental observations.  相似文献   

9.
Lactate monooxygenase (LMO) catalyzes the conversion of L-lactate to acetate, CO(2), and water with the incorporation of molecular oxygen. Arginine 187 of LMO is highly conserved within the family of L-alpha-hydroxyacid oxidizing enzymes (Lê, K. H. D., and Lederer, F. (1991) J. Biol. Chem. 266, 20877-20881). By comparison with the equivalent residue in flavocytochrome b(2) from Saccharomyces cerevisiae (Pike, A. D., Chapman, S. K, Manson, F. D. C,. Reid, G. A. , Gondry, M., and Lederer, F. (1996) in Flavins and Flavoproteins (Stevenson, K. J., Massey, V., and Williams, C. H., Jr., eds) pp. 571-574, University of Calgary Press, Calgary, AB, Canada), arginine 187 might be expected to have an important role in catalytic efficiency and substrate binding in LMO. Histidine 240 is predicted to be close to the substrate binding site of LMO, although it is not conserved within the enzyme family. Arginine 187 has been replaced with methionine (R187M), and histidine 240 has been replaced with glutamine (H240Q). L-Lactate oxidation by R187M is very slow. The binding of L-lactate to the mutant enzyme appears to be very weak, as is the binding of oxalate, a transition state analogue. The binding of pyruvate to the reduced enzyme is also very weak, resulting in complete uncoupling of enzyme turnover, with H(2)O(2) and pyruvate as the final products. In addition, anionic forms of the flavin are unstable. The K(d) for sulfite is increased nearly 400-fold by this mutation. The semiquinone form of R187M is also thermodynamically unstable, although the overall midpoint potential for the two-electron reduction of R187M is only 34 mV lower than for the wild-type enzyme. H240Q more closely resembles the wild-type enzyme. The steady-state activity of H240Q is completely coupled. The k(cat) is similar to that for the wild-type enzyme.  相似文献   

10.
Filopodia are actin-rich fingerlike protrusions found at the leading edge of migrating cells and are believed to play a role in directional sensing. Previous studies have shown that myosin-X (myoX) promotes filopodia formation and that this is mediated through its ability to deliver specific cargoes to the cell periphery (Tokuo, H., and M. Ikebe. 2004. Biochem Biophys. Commun. 319:214-220; Zhang, H., J.S. Berg, Z. Li, Y. Wang, P. Lang, A.D. Sousa, A. Bhaskar, R.E. Cheney, and S. Stromblad. 2004. Nat. Cell Biol. 6:523-531; Bohil, A.B., B.W. Robertson, and R.E. Cheney. 2006. Proc. Natl. Acad. Sci. USA. 103:12411-12416; Zhu, X.J., C.Z. Wang, P.G. Dai, Y. Xie, N.N. Song, Y. Liu, Q.S. Du, L. Mei, Y.Q. Ding, and W.C. Xiong. 2007. Nat. Cell Biol. 9:184-192). In this study, we show that the motor function of myoX and not the cargo function is critical for initiating filopodia formation. Using a dimer-inducing technique, we find that myoX lacking its cargo-binding tail moves laterally at the leading edge of lamellipodia and induces filopodia in living cells. We conclude that the motor function of the two-headed form of myoX is critical for actin reorganization at the leading edge, leading to filopodia formation.  相似文献   

11.
The amino acid sequence of the largest fragment, CNBr Ia (203 residues) has been reported (Yokota, E., and Riggs, A. F. (1984) J. Biol. Chem. 259, 4739-4749). The amino acid sequences of the second largest fragment, CNBr Ib (142 residues), and of the 12 smaller fragments are reported in accompanying papers (Moore, M. D., Behrens, P. Q., and Riggs, A. F. (1986) J. Biol. Chem. 261, 10511-10519; Behrens, P. Q., Nakashima, H., and Riggs, A. F. (1986) J. Biol. Chem. 261, 10520-10525). The complete amino acid sequence of hemocyanin component II has been established by isolation and analysis of 13 methionine-containing peptides from either a tryptic digest or a Staphylococcus aureus strain V8 protease digest of whole carboxamidomethylated hemocyanin II. Hemocyanin II is composed of 628 residues and has a molecular weight with two copper atoms of 72,946.  相似文献   

12.
The structure of bovine odorant-binding protein (bOBP) revealed a striking feature of a dimer formed by domain swapping [Tegoni, M., Ramoni, R., Bignetti, E., Spinelli, S. & Cambillau, C. (1996) Nat. Struct. Biol.3, 863-867; Bianchet, M.A., Bains, G., Pelosi, P., Pevsner, J., Snyder, S.H., Monaco, H.L. & Amzel, L.M. (1996) Nat. Struct. Biol.3, 934-939] and the presence of a naturally occuring ligand [Ramoni, R., Vincent, F., Grolli, S., Conti, V., Malosse, C., Boyer, F.D., Nagnan-Le Meillour, P., Spinelli, S., Cambillau, C. & Tegoni, M. (2001) J. Biol. Chem.276, 7150-7155]. These features led us to investigate the binding of odorant molecules with bOBP in solution and in the crystal. The behavior of odorant molecules in bOBP resembles that observed with porcine OBP (pOBP), although the latter is monomeric and devoid of ligand when purified. The odorant molecules presented K(d) values with bOBP in the micromolar range. Most of the X-ray structures revealed that odorant molecules interact with a common set of residues forming the cavity wall and do not exhibit specific interactions. Depending on the ligand and on the monomer (A or B), a single residue--Phe89--presents alternate conformations and might control cross-talking between the subunits. Crystal data on both pOBP and bOBP, in contrast with binding and spectroscopic studies on rat OBP in solution, reveal an absence of significant conformational changes involving protein loops or backbone. Thus, the role of OBP in signal triggering remains unresolved.  相似文献   

13.
14.
Mutation of aspartic acid 187 to asparagine (D187N) or tyrosine (D187Y) in domain 2 of the actin-modulating protein gelsolin causes the neurodegenerative disease familial amyloidosis of Finnish type (FAF). These mutations render plasma gelsolin susceptible to aberrant proteolysis by furin in the trans-Golgi network, the initial proteolytic event in the formation of 71 and 53 residue fragments that assemble into amyloid fibrils. Ca(2+) binding stabilizes wild-type domain 2 gelsolin against denaturation and proteolysis, but the FAF variants are unable to bind and be stabilized by Ca(2+). Though the chain of events initiating FAF has been elucidated recently, uncertainty remains about the mechanistic details that allow the FAF variants to be processed. To test the hypothesis that impaired Ca(2+) binding in the D187 variants, but not other factors specific to residue 187, increases susceptibility to aberrant proteolysis and subsequent amyloidogenesis, we designed the gelsolin variant E209Q to remove a different Ca(2+) ligand from the same Ca(2+) site that is affected in the FAF variants. Here, we show that E209Q domain 2 does not bind Ca(2+) and is not stabilized against denaturation or furin proteolysis, analogous to the behavior exhibited by the FAF variants. Transfection of full-length E209Q into COS cells results in secretion of both the full-length and furin-processed fragments, as observed with D187N and D187Y. Mutation of the furin consensus sequence in D187N and E209Q gelsolin prevents cleavage during secretion, indicating that inhibition of proprotein convertases (furin) represents a viable therapeutic approach for the treatment of FAF. Mutations that diminish domain 2 Ca(2+) binding allow furin access to an otherwise protected cleavage site, initiating the proteolytic cascade that leads to gelsolin amyloidogenesis and FAF.  相似文献   

15.
Klebsiella aerogenes urease was purified 1,070-fold with a 25% yield by a simple procedure involving DEAE-Sepharose, phenyl-Sepharose, Mono Q, and Superose 6 chromatographies. The enzyme preparation was comprised of three polypeptides with estimated Mr = 72,000, 11,000, and 9,000 in a alpha 2 beta 4 gamma 4 quaternary structure. The three components remained associated during native gel electrophoresis, Mono Q chromatography, and Superose 6 chromatography despite the presence of thiols, glycols, detergents, and varied buffer conditions. The apparent compositional complexity of K. aerogenes urease contrasts with the simple well-characterized homohexameric structure for jack bean urease (Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gazzola, C., Winzor, D. J., Blakeley, R. L., and Zerner, B. (1980) Can. J. Biochem. 58, 1323-1334); however, heteromeric subunit compositions were also observed for the enzymes from Proteus mirabilis, Sporosarcina ureae, and Selemonomas ruminantium. K. aerogenes urease exhibited a Km for urea of 2.8 +/- 0.6 mM and a Vmax of 2,800 +/- 200 mumol of urea min-1 mg-1 at 37 degrees C in 25 mM N-2-hydroxyethylpiperazineN'-2-ethanesulfonic acid, 5.0 mM EDTA buffer, pH 7.75. The enzyme activity was stable in 1% sodium dodecyl sulfate, 5% Triton X-100, 1 M KCl, and over a pH range from 5 to 10.5, with maximum activity observed at pH 7.75. Two active site groups were defined by their pKa values of 6.55 and 8.85. The amino acid composition of K. aerogenes urease more closely resembled that for the enzyme from Brevibacter ammoniagenes (Nakano, H., Takenishi, S., and Watanabe, Y. (1984) Agric. Biol. Chem. 48, 1495-1502) than those for plant ureases. Atomic absorption analysis was used to establish the presence of 2.1 +/- 0.3 mol of nickel per mol of 72,000-dalton subunit in K. aerogenes urease.  相似文献   

16.
To determine the interaction site for cytochrome c (Cc) on cytochrome c oxidase (CcO), a number of conserved carboxyl residues in subunit II of Rhodobacter sphaeroides CcO were mutated to neutral forms. A highly conserved tryptophan, Trp(143), was also mutated to phenylalanine and alanine. Spectroscopic and metal analyses of the surface carboxyl mutants revealed no overall structural changes. The double mutants D188Q/E189N and D151Q/E152N exhibit similar steady-state kinetic behavior as wild-type oxidase with horse Cc and R. sphaeroides Cc(2), showing that these residues are not involved in Cc binding. The single mutants E148Q, E157Q, D195N, and D214N have decreased activities and increased K(m) values, indicating they contribute to the Cc:CcO interface. However, their reactions with horse and R. sphaeroides Cc are different, as expected from the different distribution of surface lysines on these cytochromes c. Mutations at Trp(143) severely inhibit activity without changing the K(m) for Cc or disturbing the adjacent Cu(A) center. From these data, we identify a Cc binding area on CcO with Trp(143) and Asp(214) close to the site of electron transfer and Glu(148), Glu(157), and Asp(195) providing electrostatic guidance. The results are completely consistent with time-resolved kinetic measurements (Wang, K., Zhen, Y., Sadoski, R., Grinnell, S., Geren, L., Ferguson-Miller, S., Durham, B., and Millett, F. (1999) J. Biol. Chem. 274, 38042-38050) and computational docking analysis (Roberts, V. A., and Pique, M. E. (1999) J. Biol. Chem. 274, 38051-38060).  相似文献   

17.
A procedure is described for reconstructing three-dimensional objects from two-dimensional projections. The method is based both on the original Crowther, DeRosier, and Klug (DeRosier, D.J., and Klug, A. (1968), Nature 217, 130-134; Crowther, R.A., DeRosier, D.J., and Klug, A. (1970) Proc. R. Soc. London A 317, 319-340) work on image reconstructions of icosahedral viruses and on the concept of noncrystallographic symmetry (Rossmann, M.G., 1995, Curr. Opin. Struct. Biol. 5, 650-655). The procedure has been applied so far only to test data where the objective has been the determination of particle orientation, both ab initio and through the use of model data.  相似文献   

18.
The high-potential iron-sulfur protein (HiPIP) from Rhodospirillum tenue (strain 3761) shows only a weak (20-25%) sequence similarity to HiPIPs from Chromatium vinosum, Ectothiorhodospira halophila and Ectothiorhodospira vacuolata, including the strict conservation of only two of the twelve residues assumed to be in the 4Fe-4S cluster packing region [Tedro, S. M., Meyer, T. E. and Kamen, M. D. (1979) J. Biol. Chem. 254, 1495-1500]. In spite of these differences, the general range and distribution of hyperfine-shifted 1H-NMR peaks of oxidized and reduced R. tenue HiPIP resemble those of E. halophila HiPIP I [Krishnamoorthi, R., Markley, J. L., Cusanovich, M. A., Pryzycieki, C. T. and Meyer, T. E. (1986) Biochemistry 25, 60-67]. Temperature- and pH-dependence and longitudinal relaxation behavior were determined for hyperfine-shifted peaks of the oxidized protein. Tentative assignments of peaks to ligands and aromatic residues suggest the presence of common apoprotein-active-site interactions in these proteins. Differences occur in the pattern of paramagnetically shifted peaks attributed to hydrogens bonded to the 4Fe-4S cluster. Hyperfine-shifted peaks of R. tenue HiPIP are not perturbed by pH changes in the range 5-9. In contrast, those of the C. vinosum protein exhibit a pH-dependence of chemical shifts that has been attributed to the titration of His42 [Nettesheim, D. G., Meyer, T. E., Feinberg, B. A. and Otvos, J. D. (1983) J. Biol. Chem. 258, 8235-8239]. Since R. tenue HiPIP contains no histidine, the present observation confirms the above hypothesis.  相似文献   

19.
Fourteen fragments have been isolated from hemocyanin component II of Limulus polyphemus by cleavage with CNBr. The amino acid sequence of the largest fragment, CNBr Ia has been reported (Yokota, E., and Riggs, A. F. (1984) J. Biol. Chem. 259, 4739-4749). The amino acid sequence of the 12 smaller fragments is reported in an accompanying paper (Moore, M. D., Behrens, P. Q., and Riggs, A. F. (1985) J. Biol. Chem. 261, 10511-10519). We have determined the amino acid sequence of the second largest fragment, CNBr Ib. The fragment contains 142 residues and has a molecular weight of 16,095.  相似文献   

20.
The carboxyl-terminal domains of secretin family peptides have been shown to contain key determinants for high affinity binding to their receptors. In this work, we have examined the interaction between carboxyl-terminal residues within secretin and the prototypic secretin receptor. We previously utilized photoaffinity labeling to demonstrate spatial approximation between secretin residue 22 and the receptor domain that includes the first 30 residues of the amino terminus (Dong, M., Wang, Y., Pinon, D. I., Hadac, E. M., and Miller, L. J. (1999) J. Biol. Chem. 274, 903-909). Here, we further refined the site of labeling with the p-benzoyl-phenylalanine (Bpa(22)) probe to receptor residue Leu(17) using progressive cleavage of wild type and mutant secretin receptors (V13M and V16M) and sequence analysis. We also developed a new probe incorporating a photolabile Bpa at position 26 of secretin, closer to its carboxyl terminus. This analogue was also a potent agonist (EC(50) = 72 +/- 6 pm) and bound to the secretin receptor specifically and with high affinity (K(i) = 10.3 +/- 2.4 nm). It covalently labeled the secretin receptor at a single site saturably and specifically. This was localized to the segment between residues Gly(34) and Ala(41) using chemical and enzymatic cleavage of labeled wild type and A41M mutant receptor constructs and immunoprecipitation of epitope-tagged receptor fragments. Radiochemical sequencing identified the site of covalent attachment as residue Leu(36). These new insights, along with our recent report of contact between residue 6 within the amino-terminal half of secretin and this same amino-terminal region of this receptor (Dong, M., Wang, Y., Hadac, E. M., Pinon, D. I., Holicky, E. L., and Miller, L. J. (1999) J. Biol. Chem. 274, 19161-19167), support a key role for this region, making the molecular details of this interaction of major interest.  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