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 46 毫秒
1.
In the presence of DNA, the antitumor drug N2-methyl-9-hydroxyellipticinium (elliptinium; NMHE) [Le Pecq, J. B., Gosse, C., Dat-Xuong, N., & Paoletti, C. (1975) C. R. Seances Acad. Sci., Ser. D 281, 1365-1367] is oxidized by the horseradish peroxidase-hydrogen peroxide (HRP-H2O2) system to the quinone imine derivative N2-methyl-9-oxoellipticinium (NMOE) [Auclair, C., & Paoletti, C. (1981) J. Med. Chem. 24, 289-295], which interacts with DNA according to the intercalation mode. When excess H2O2 was used, the major part of the quinone imine was further oxidized to the o-quinone N2-methyl-9,10-dioxoellipticinium [Bernadou, J., Meunier, G., Paoletti, C., & Meunier, B. (1983) J. Med. Chem. 26, 574-579]. In the presence of stoichiometric amounts of H2O2 (H2O2/NMHE = 1), NMOE reacts with DNA, yielding a fluorescent compound irreversibly linked to the nucleic acid, which is related to the covalent binding of the ellipticinium chromophore. Under optimal reaction conditions, NMHE binding occurs according to a first-order process (k = 4.3 X 10(-3) min-1) with a linear increase with respect to drug to nucleotide ratio up to a maximum binding of 1 NMHE per 20 base pairs (r = 0.05). The fluorescence spectra (ex, 330 nm; em, 548 nm) of NMHE bound to DNA, the occurrence of energy transfer from the DNA to the drug, and the DNA length increase of the DNA-NMHE adduct suggest that the binding occurs at the intercalating site with limited denaturation of the DNA helix.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
An ethidium homodimer and acridine ethidium heterodimer have been synthesized (Gaugain, B., Barbet, J., Oberlin, R., Roques, B. P., & Le Pecq, J. B. (1978) Biochemistry 17 (preceding paper in this issue)). The binding of these molecules to DNA has been studied. We show that these dimers intercalate only one of their chromophores in DNA. At high salt concentration (Na+ greater than 1 M) only a single type of DNA-binding site exists. Binding affinity constants can then be measured directly using the Mc Ghee & Von Hippel treatment (Mc Ghee, J. D., & Von Hippel, P. H. (1974) J. Mol. Biol. 86, 469). In these conditions the dimers cover four base pairs when bound to DNA. Binding affinities have been deduced from competition experiments in 0.2 M Na+ and are in agreement with the extrapolated values determined from direct DNA-binding measurements at high ionic strength. As expected, the intrinsic binding constant of these dimers is considerably larger than the affinity of the monomer (ethidium dimer K = 2 X 10(8) M-1; ethidium bromide K = 1.5 X 10(5) M-1 in 0.2 M Na+). The fluorescence properties of these molecules have also been studied. The efficiency of the energy transfer from the acridine to the phenanthridinium chromophore, in the acridine ethidium heterodimer when bound to DNA, depends on the square of the AT base pair content. The large increase of fluorescence on binding to DNA combined with a high affinity constant for nucleic acid fluorescent probes. In particular, such molecules can be used in competition experiments to determine the DNA binding constant of ligands of high binding affinity such as bifunctional intercalators.  相似文献   

3.
Data relating to the effect of intercalating drugs on the winding of the DNA helix is re-considered. Analyses by Paoletti &; Le Pecq (1971) of fluorescence depolarization, X-ray diffraction and molecular model building are reappraised. It is concluded that the helix is unwound by ~12 ° as proposed by Fuller &; Waring (1964). This refutes the recent suggestion by Paoletti &; Le Pecq (1971) that the intercalation winds the helix by ~13 °.  相似文献   

4.
An ethidium homodimer and an acridine ethidium heterodimer have been synthesized. The ethidium and the acridine chromophore were introduced in such bifunctional intercalators in order to allow the fluorometric study of the interaction of such molecules with DNA, which is reported in the companion paper (Gaugain, B., Barbet, J., Capelle, N., Roques, B.P., & Le Pecq, J.B.(1978) Biochemistry 17 (following paper in this issue)). In the preparation of the acridine-ethidium dimer, we report the use of acetyl groups as new protecting agents in the phenanthridine series. Conformational studies of these molecules by visible absorption and NMR spectroscopy indicate that these dimers exist in equilibrium between folded and unfolded conformations and that this equilibrium is pH and temperature dependent. Models for the geometry of the folded forms are proposed.  相似文献   

5.
T Y Teng  H W Huang  G A Olah 《Biochemistry》1987,26(25):8066-8072
A previous extended X-ray absorption fine structure (EXAFS) study of photolyzed carboxymyoglobin (MbCO) [Chance, B., Fischetti, R., & Powers, L. (1983) Biochemistry 22, 3820-3829; Powers, L., Sessler, J. L., Woolery, G. L., & Chance, B. (1984) Biochemistry 23, 5519-5523] has provoked much discussion on the heme structure of the photoproduct (MbCO). The EXAFS interpretation that the Fe-CO distance increases by no more than 0.05 A following photodissociation has been regarded as inconsistent with optical, infrared, and magnetic susceptibility studies [Fiamingo, F. G., & Alben, J. O. (1985) Biochemistry 24, 7964-7970; Sassaroli, M., & Rousseau, D. L. (1986) J. Biol. Chem. 261, 16292-16294]. The present experiment was performed with well-characterized dry film samples in which MbCO molecules were embedded in a poly(vinyl alcohol) matrix [Teng, T. Y., & Huang, H. W. (1986) Biochim. Biophys. Acta 874, 13-18]. The sample had a high protein concentration (12 mM) to yield adequate EXAFS signals but was very thin (40 micron) so that complete photolysis could be easily achieved by a single flash from a xenon lamp. Although the electronic state of MbCO resembles that of deoxymyoglobin (deoxy-Mb), direct comparison of EXAFS spectra indicates that structurally MbCO is much closer to MbCO than to deoxy-Mb.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Glycosylation can affect the physical and biochemical properties of the polypeptide chain in glycoproteins. Asparagine-N-linked polylactosaminyl glycosylation of the chymotryptic 44-kDa gelatin-binding domain from human placental fibronectin confers protease resistance [Zhu, B. C. R., Fisher, S. F., Panda, H., Calaycay, J., Shively, J. E. & Laine, R. A. (1984) J. Biol. Chem. 259, 3962-3970] and weaken the binding to gelatin [Zhu, B. C. R. & Laine, R. A. (1985) J. Biol. Chem. 260, 4041-4045]. Intrinsic tryptophan fluorescence of the gelatin-binding domain was used to probe glycosylation-dependent protein conformation changes. In gelatin-binding fragments containing incrementally smaller polylactosamine oligosaccharides, the fluorescence intensity progressively decreased and the emission spectrum shifted about 7 nm to the blue. Removal of the polylactosamine chains from a highly glycosylated fragment with endo-beta-galactosidase from Escherichia freundii also quenched the protein fluorescence. The fluorescence lifetimes did not appear to be affected by the extent of glycosylation, suggesting static quenching of the tryptophan emission in the low glycosylated fragments. Acrylamide quenching studies showed that the accessibility of the tryptophans to small solutes was not altered by glycosylation. The steady-state emission anisotropy increased with decreasing polylactosamine chain length. The results indicate that the polylactosamine chains alter the tryptophan environments in the gelatin-binding domain, probably by changing the polypeptide conformation. These putative protein conformation changes may be partially responsible for the altered gelatin binding, protease resistance, and cell adhesion functions of fetal tissue fibronectin.  相似文献   

7.
Sequence-specific spin-labeled oligodeoxynucleotides with conformation-sensitive electron paramagnetic resonance (EPR) signals are synthesized and examined as solution-phase nucleic acid hybridization probes. Either a proxyl or tempo ring linked to the C(5) position of deoxyuridine (dU) by a nonrigid two-atom methylamino tether is incorporated within 15-mers by phosphotriester chemistry yielding stable spin-labeled probes with distinctive EPR specific activity (AEPR) values. The AEPR is greater for a proxyl-labeled than for a tempo-labeled probe and is consistent with EPR data of enzymatically labeled 26-mers [Bobst, A. M., Pauly, G. T., Keyes, R. S., and Bobst, E. V. (1988) FEBS Lett. 228, 33-36], after normalizing for percent labeling. The spectral characteristics of the free probes and the probe/target complexes are similar to those of enzymatically spin-labeled nucleic acids containing a different nonrigid two-atom-tethered spin label [Bobst, A. M., Kao, S.-C., Toppin, R. C., Ireland, J. C., and Thomas, I. E. (1984) J. Mol. Biol. 173, 63-70]. The presence of target DNA is detected in solution by EPR spectroscopy and the assay is based on the characteristic line-shape change associated with hybridization. The EPR spectra of free and bound probe reflect little interference from changes in global dynamics of the probe, and the line-shape change upon complexation results primarily from a change in local base dynamics. The presence or absence of hybridization can be detected in a loop-gap resonator with about 1 pmol of spin-labeled 15-mer within minutes.  相似文献   

8.
The coordination sphere of both the structural and catalytic zinc ions of Bacillus cereus phospholipase C has been probed by substitution of cobalt(II) for zinc and investigation of the resultant derivatives by a variety of spectroscopic techniques. The electronic absorption, circular dichroic, magnetic circular dichroic, and electron paramagnetic resonance spectra were found to be strikingly similar when cobalt(II) was substituted into either site and are consistent with a distorted octahedral environment for the metal ion in both sites. Octahedral coordination appears comparatively rare in zinc metalloenzymes but has been suggested for glyoxalase I [Sellin, S., Eriksson, L. E. G., Aronsson, A.-C., & Mannervik, B. (1983) J. Biol. Chem. 258, 2091-2093; Garcia-Iniguez, L., Powers, L., Chance, B., Sellin, S., Mannervik, B., & Mildvan, A. S. (1984) Biochemistry 23, 685-689], transcarboxylase [Fung, C.-H., Mildvan, A. S., & Leigh, J. S. (1974) Biochemistry 13, 1160-1169], and the regulatory binding site of Aeromonas aminopeptidase [Prescott, J. M., Wagner, F. W., Holmquist, B., & Vallee, B. L. (1985) Biochemistry 24, 5350-5356]. Phospholipase C is so far unique in having two such sites.  相似文献   

9.
W Bujalowski  T M Lohman 《Biochemistry》1987,26(11):3099-3106
We describe a general method for the analysis of ligand-macromolecule binding equilibria for cases in which the interaction is monitored by a change in a signal originating from the ligand. This method allows the absolute determination of the average degree of ligand binding per macromolecule without any assumptions concerning the number of modes or states for ligand binding or the relationship between the fractional signal change and the fraction of bound ligand. Although this method is generally applicable to any type of signal, we discuss the details of the method as it applies to the analysis of binding data monitored by a change in fluorescence of a ligand upon binding to a nucleic acid. We apply the analysis to the equilibrium binding of Escherichia coli single-strand binding (SSB) protein to single-stranded nucleic acids, which is monitored by the quenching of the intrinsic tryptophan fluorescence of the SSB protein. With this method, one can quantitatively determine the relationship between the fractional signal change of the ligand and the fraction of bound ligand, LB/LT, and rigorously test whether the signal change is directly proportional to LB/LT. For E. coli SSB protein binding to single-stranded nucleic acids in its (SSB)65 binding mode [Lohman, T. M., & Overman, L. B. (1985) J. Biol. Chem. 260, 3594; Chrysogelos, S., & Griffith, J. (1982) Proc. Natl. Acad. Sci. U.S.A. 79, 5803], we show that the fractional quenching of the SSB fluorescence is equal to the fraction of bound SSB.  相似文献   

10.
J A Grasby  B A Connolly 《Biochemistry》1992,31(34):7855-7861
The stereochemical course of the reaction catalyzed by the EcoRV restriction endonuclease has been determined. This endonuclease recognizes GATATC sequence and cuts between the central T and dA bases. The Rp isomer of d(GACGATsATCGTC) (this dodecamer contains a phosphorothioate rather than the usual phosphate group between the central T and dA residues, indicated by the s) was a substrate for the endonuclease. Performing this reaction in H2 18O gave [18O]dps(ATCGTC) (a pentamer containing an 18O-labeled 5'-phosphorothioate) which was converted to [18O]dAMPS with nuclease P1. This deoxynucleoside 5'-[18O]phosphorothioate was stereospecifically converted to [18O]dATP alpha S with adenylate kinase and pyruvate kinase [Brody, R. S., & Frey, P. A. (1981) Biochemistry 20, 1245-1251]. Analysis of the position of the 18O in this product by 31P NMR spectroscopy showed that it was in a bridging position between the alpha- and beta-phosphorus atoms. This indicates that the EcoRV hydrolysis proceeds with inversion of configuration at phosphorus. The simplest interpretation is that the mechanism of this endonuclease involves a direct in-line attack at phosphorus by H2O with a trigonal bipyramidal transition state. A covalent enzyme oligodeoxynucleotide species can be discounted as an intermediate. An identical result has been previously observed with the EcoR1 endonuclease [Connolly, B. A., Eckstein, F., & Pingoud, A. (1984) J. Biol. Chem. 259, 10760-10763]. X-ray crystallography has shown that both of these endonucleases contain a conserved array of amino acids at their active sites. Possible mechanistic roles for these conserved amino acids in the light of the stereochemical findings are discussed.  相似文献   

11.
Y Ohashi  M Iwamori  T Ogawa  Y Nagai 《Biochemistry》1987,26(13):3990-3995
The structures of long-chain bases are expressed as [CH2C(NH2) = CHR]+ (Z+) in the positive ion mode spectra obtained on fast atom bombardment (FAB) mass spectrometry or liquid-matrix-assisted secondary ion mass spectrometry (SIMS) [Benninghoven, A., Ed. (1983) Ion Formation from Organic Solids, Springer, Berlin]. This phenomenon is common to sphingolipids in general: glycosphingolipids [see reviews by Sweeley and Nunez [Sweeley, C. C., & Nunez, H. A. (1985) Annu. Rev. Biochem. 54, 765] and Kanfer and Hakomori [Kanfer, J. N., & Hakomori, S. (1983) Handb. Lipid Res. 3]] and phosphonosphingolipids [Hayashi, A., & Matsubara, T. (1982) in New Vistas in Glycolipid Research (Makita, A., Handa, S., Taketomi, T., & Nagai, Y., Eds.) p 103, Plenum, New York], inclusive. Phytosphingosine compounds show the same type of fragmentation without additional dehydration if a neutral matrix is used. A Z+ ion is easily detected in the lower mass region (m/z 200-400) as an even mass number fragment ion, and confirmation is made by means of B/E constant and B2/E constant linked scan techniques [Boyd, R. K., & Beynon, J. H. (1977) Org. Mass Spectrom. 12, 163; Boyd, R. K., & Shushan, B. (1981) Int. J. Mass Spectrom. Ion Phys. 37, 355; Macdonald, C. G., & Lacey, M. J. (1984) Org. Mass Spectrom. 19, 55]. [Principles of linked scannings are explicitly summarized by Jennings and Mason [Jennings, K. R., & Mason, R. S. (1983) in Tandem Mass Spectrometry (McLafferty, F. W., Ed.) p 197, Wiley, New York] besides the cited literature.]  相似文献   

12.
Charge recombination of the primary radical pair in D1/D2 reaction centers from photosystem 2 has been studied by time-resolved fluorescence and absorption spectroscopy. The kinetics of the primary radical pair are multiexponential and exhibit at least two lifetimes of 20 and 52 ns. In addition, a third lifetime of approximately 500 ps also appears to be present. These multiexponential charge-recombination kinetics reflect either different conformational states of D1/D2 reaction centers, with the different conformers exhibiting different radical pair lifetimes, or relaxations in the free energy of the radical pair state. Whichever model is invoked, the free energies of formation of the different radical pair states exhibit a linear temperature dependence from 100 to 220 K, indicating that they are dominated by entropy with negligible enthalpy contributions. These results are in agreement with previous determinations of the thermodynamics that govern primary charge separation in both D1/D2 reaction centers [Booth, P.J., Crystall, B., Giorgi, L. B., Barber, J., Klug, D.R., & Porter, G. (1990) Biochim. Biophys. Acta 1016, 141-152] and reaction centers of purple bacteria [Woodbury, N.W.T., & Parson, W.W. (1984) Biochim. Biophys. Acta 767, 345-361]. It is possible that these observations reflect structural changes that accompanying primary charge separation and assist in stabilization of the radical pair state thus optimizing the efficiency of primary electron transfer.  相似文献   

13.
Phased psoralen cross-links do not bend the DNA double helix   总被引:1,自引:0,他引:1  
T E Haran  D M Crothers 《Biochemistry》1988,27(18):6967-6971
Although the chemical reaction of psoralens with nucleic acids is well understood, the structure of psoralen-DNA cross-linked products is still not clear. Model building studies base on the crystal structure of the psoralen-thymine monoadduct suggest that each cross-link bends the DNA double helix by 46.5 degrees [Pearlman, D. A., Holbrook, S. R., Pirkle, D. H., & Kim, S.-H. (1985) Science (Washington, D.C.) 227, 1304-1308]. On the other hand, Sinden and Hagerman [Sinden, R. R., & Hagerman, P. J. (1984) Biochemistry 23, 6299-6303] find that, in solution, psoralen cross-linked DNA is not bent. Here we use gel electrophoresis to test the validity of the current models. We have synthesized a series of DNA fragments (21-24 base pairs in length), each containing one unique T-A site for 4'-(hydroxymethyl)-4,5',8-trimethylpsoralen (HMT) cross-linking. Because of an estimated 28 degrees unwinding of the helix by HMT [Wiesehahn, G., & Hearst, J. E. (1978) Proc. Natl. Acad. Sci. U.S.A. 75, 2703-2707], one expects that the 22-bp cross-linked fragment will be repeated nearly in phase with the average helical screw when multimerized. In that sequence ligation will maximally amplify any deformation to the double helix. We find that the ligated multimers of cross-linked DNA migrate close to the multimers of non-cross-linked DNA on polyacrylamide gels. Our observations place an upper limit of 10 degrees on DNA bending induced by psoralen cross-linking and indicate unwinding by about 1 bp, as well as stiffening of the double helix. These properties are not unexpected for classical intercalators.  相似文献   

14.
B H Oh  J L Markley 《Biochemistry》1990,29(16):3993-4004
Complete sequence-specific assignments were determined for the diamagnetic 1H resonances from Anabaena 7120 ferredoxin (Mr = 11,000). A novel assignment procedure was followed whose first step was the identification of the 13C spin systems of the amino acids by a 13C(13C) double quantum correlation experiment [Oh, B.-H., Westler, M. W., Darba, P., & Markley, J. L. (1988) Science 240, 908-911]. Then, the 1H spin systems of the amino acids were identified from the 13C spin systems by means of direct and relayed 1H(13C) single-bond correlations [Oh, B.-H., Westler, W. M., & Markley, J. L. (1989) J. Am. Chem. Soc. 111, 3083-3085]. The sequential resonance assignments were based mainly on conventional interresidue 1H alpha i-1HNi + 1 NOE connectivities. Resonances from 18 residues were not resolved in two-dimensional 1H NMR spectra. When these residues were mapped onto the X-ray crystal structure of the homologous ferredoxin from Spirulina platensis [Fukuyama, K., Hase, T., Matsumoto, S., Tsukihara, T., Katsube, Y., Tanaka, N., Kakudo, M., Wada, K., & Matsubara, H. (1980) Nature 286, 522-524], it was found that they correspond to amino acids close to the paramagnetic 2Fe.2S* cluster. Cross peaks in two-dimensional homonuclear 1H NMR spectra were not observed for any protons closer than about 7.8 A to both iron atoms. Secondary structural features identified in solution include two antiparallel beta-sheets, one parallel beta-sheet, and one alpha-helix.  相似文献   

15.
T M Lohman 《Biochemistry》1984,23(20):4665-4675
The dissociation kinetics of bacteriophage T4 coded gene 32 protein-single-stranded nucleic acid complexes have been examined as a function of monovalent salt concentration, temperature, and pH in order to investigate the details of the dissociation of cooperatively bound protein. Fluorescence stopped-flow techniques were used, and irreversible dissociation was induced by a combination of [NaCl] jumps and mixing with excess nucleic acid competitor. This made it possible to directly investigate the irreversible dissociation process over a wide range of NaCl concentrations [e.g., from 50 mM to 0.60 M for the gene 32 protein-poly(A) complex], in the absence of reassociation. Over the entire salt range, the only dissociable species observed is the singly contiguously bound gene 32 protein which dissociates from the ends of protein clusters. However, the [NaCl] dependence of the dissociation rate constant suggests that two competing pathways exist for dissociation of cooperatively bound gene 32 protein from the ends of protein clusters. At high monovalent salt concentrations, dissociation is dominated by a single-step process, with log ke/log [NaCl] = 6.5 +/- 0.5; i.e., the dissociation rate constant increases with increasing NaCl concentration due to the uptake of approximately six monovalent ions upon dissociation. This indicates that singly contiguous protein dissociates directly into solution. However, at much lower [NaCl] the data suggest that gene 32 protein, when bound at the end of a protein cluster, dissociates by first sliding off the end to form a noncooperatively bound intermediate which subsequently dissociates. A quantitative model which incorporates the sliding pathway [Berg, O. G., Winter, R. B., & von Hippel, P. H. (1981) Biochemistry 20, 6929-6948] in the dissociation mechanism fits the data reasonably well and suggests that noncooperatively bound monomers of gene 32 protein may be capable of one-dimensional translocation along single-stranded nucleic acids as suggested by independent kinetic data on the association reaction [Lohman, T. M., & Kowalczykowski, S. C. (1981) J. Mol. Biol. 152, 67-109]. It is also observed that both the absolute dissociation rate constant for T4 gene 32 protein and its salt dependence are sensitive to the average molecular weight and polydispersity of the nucleic acid sample used. This is a general phenomenon exhibited by proteins that bind to nucleic acids in a highly cooperative manner.  相似文献   

16.
Raman spectroscopy of Z-form poly[d(A-T)].poly[d(A-T)   总被引:3,自引:0,他引:3  
Helical structures of double-stranded poly[d(A-T)] in solution have been studied by Raman spectroscopy. While the classical right-handed conformation B-type spectra are obtained in the case of sodium chloride solutions, a Z-form Raman spectrum is observed by addition of nickel ions at high sodium concentration, conditions in which the inversion of the circular dichroic spectrum of poly[d(A-T)] is detected, similar to that observed for high-salt poly[d(G-C)] solutions [Bourtayre, P., Liquier, J., Pizzorni, L., & Taillandier, E. (1987) J. Biomol. Struct. Dyn. 5, 97-104]. The characterization of the Z-form spectrum of poly[d(A-T)] is proposed by comparison with previously obtained characteristic Raman lines of Z-form poly[d(G-C)] and poly[d(A-C)].poly[d(G-T)] solutions and of d(CG)3 and d(CGCATGCG) crystals [Thamann, T. J., Lord, R. C., Wang, A. H.-J., & Rich, A. (1981) Nucleic Acids Res. 9, 5443-5457; Benevides, J. M., Wang, A. H.-J., van der Marel, G. A., van Boom, J. H., Rich, A., & Thomas, G. J., Jr. (1984) Nucleic Acids Res. 14, 5913-5925]. Detailed spectroscopic data are presented reflecting the reorientation of the purine-deoxyribose entities (C2'-endo/anti----C3'-endo/syn), the modification of the phosphodiester chain, and the adenosine lines in the 1300-cm-1 region. The role played by the hydrated nickel ions in the B----Z transition is discussed.  相似文献   

17.
R T Jiang  Y J Shyy  M D Tsai 《Biochemistry》1984,23(8):1661-1667
Separate diastereomers of 1,2-dipalmitoyl-sn-glycero-3- thiophosphoethanolamine ( DPPsE ) were prepared in 97% diastereomeric purity and characterized by 31P, 13C, and 1H nuclear magnetic resonance (NMR). The isomers hydrolyzed by phospholipases A2 and C specifically were designated as isomer B (31P NMR delta 59.13 in CDCl3 + Et3N ) and isomer A (59.29 ppm), respectively, analogous to the isomers B and A of 1,2-dipalmitoyl-sn-glycero-3- thiophosphocholine ( DPPsC ) [ Bruzik , K., Jiang , R.-T., & Tsai, M.-D. (1983) Biochemistry 22, 2478-2486]. Phospholipase D from cabbage was shown to be specific to isomer A of DPPsC in transphosphatidylation . The product DPPsE was shown to be isomer A. The absolute configuration of chiral DPPsE at phosphorus was elucidated by bromine-mediated desulfurization in H2 18O to give chiral 1,2-dipalmitoyl-sn-glycero-3-[18O]phosphoethanolamine ( [18O]DPPE) followed by 31 P NMR analysis [ Bruzik , K., & Tsai, M.-D. (1984) J. Am. Chem. Soc. 106, 747-754]. The absolute configuration of chiral DPPsC was elucidated by desulfurization in H2 18O mediated by bromine or cyanogen bromide to give chiral 1,2-dipalmitoyl-sn-glycero-3-[18O]phosphocholine ( [18O]DPPC), which was then converted to [18O]DPPE by phospholipase D with retention of configuration [ Bruzik , K., & Tsai, M.-D. (1984) Biochemistry (preceding paper in this issue)]. The results indicate that isomer A of both DPPsE and DPPsC is SP whereas isomer B is RP.  相似文献   

18.
A Louie  F Jurnak 《Biochemistry》1985,24(23):6433-6439
A new method for measuring the dissociation rate of the Escherichia coli elongation factor Tu-GTP--aminoacyl-tRNA complex has been developed and applied to the determination of the dissociation rates of ternary complexes formed between E. coli EF-Tu-GTP and a set of E. coli aminoacyl-tRNAs. The set of aminoacyl-tRNAs includes at least one tRNA coding for each of the 20 amino acids as well as purified isoacceptor tRNA species for arginine, glycine, leucine, lysine, and tyrosine. The results reveal that the dissociation rates vary for each ternary complex. Tu-GTP-Gln-tRNA dissociates the slowest and Tu-GTP-Val-tRNA the fastest of all noninitiator ternary complexes at 4 degrees C, pH 7.4. The equilibrium dissociation constant for Tu-GTP-Thr-tRNA has been determined to be 1.3 (0.4) X 10(-9) M under identical reaction conditions, and the absolute value of the equilibrium dissociation constant has been calculated for 28 ternary complexes from the relative equilibrium dissociation constant ratios previously measured [Louie, A., Ribeiro, N. S., Reid, B. R., & Jurnak, F. (1984) J. Biol. Chem. 259, 5010-5016]. The association rate of each ternary complex has been estimated from the ratio of the dissociation rate relative to the equilibrium dissociation constant. Tu-GTP-His-tRNA associates the fastest and Tu-GTP-Leu-tRNA1Leu the slowest. By inclusion of Tu-GTP-Met-tRNAfMet in the studies, evidence has been obtained that suggests that the initiator ternary complex does not function in the elongation cycle because the dissociation rate of the complex is very fast.  相似文献   

19.
K Y Xu 《Biochemistry》1989,28(14):5764-5772
Determinations of reaction stoichiometry demonstrate that the covalent incorporation of one molecule of 5'-isothiocyanatofluorescein can inactivate one molecule of sodium and potassium ion activated adenosinetriphosphatase in agreement with earlier determination of this stoichiometry. Several different modified peptides are produced, however, when the modified enzyme is digested with trypsin. One of these peptides has been identified as HLLVMK (thioureidylfluorescein)GAPER by use of a specific immunoadsorbent. The modified lysine is lysine 501 in the amino acid sequence of the alpha polypeptide of (Na+ + K+)-ATPase. This peptide has been previously isolated from such digests [Farley, R. A., Tran, C. M., Carilli, C. T., Hawke, D., & Shively, J. E. (1984) J. Biol. Chem. 259, 9532-9535]. The other specifically modified peptides have been purified and identified by amino acid sequencing. Their sequences identify lysine 480 and lysine 766 from the alpha polypeptide as amino acids modified by 5'-isothiocyanatofluorescein in reactions sensitive to the addition of ATP and responsible for inactivation of the enzyme.  相似文献   

20.
Many important human genes have been cloned during the last ten years. In some cases, using reverse genetic techniques [Orkin, S. H. (1986) Cell 47, 845-850], disease-causing genes have been isolated whose product was previously unknown. Important examples include the dystrophin protein which, when mutated, gives rise to either Duchenne or Becker muscular dystrophy [Koenig, M., Hoffman, E. P., Bertelson, C. J., Monaco, A. P., Feener, C. and Kunkel, L. M. (1987) Cell 50, 509-517; Monaco, A. P., Bertelson, C. J., Liechti-Gallati, S. & Kunkel, L. M. (1988) Genomics 2, 90-95; Koenig, M., Monaco, A. P. & Kunkel, L. M. (1988) Cell 53, 219-228] and the cystic fibrosis transmembrane conductance regulator (CFTR) [Riordan, J. R., Rommens, J. M., Kerem, B.-S., Alon, N., Rozmahel, R., Grzelczak, Z., Zielenski, J., Lok, S., Plavsic, N., Chou, J.-L., Drumm, M. L., Ianuzzi, M. C., Collins, F. S. & Tsui, L.-C. (1989) Science 245, 1066-1073]. Recently the technology for systematically detecting single base-pair changes by chemical methods, enzymatic methods or direct DNA sequencing has greatly expanded and simplified. In addition to providing structural information about these clinically important genes and information on disease-causing mutations, these studies have led to an increased understanding of mechanisms of mutation, to the discovery of novel genetic mechanisms and to important clinical applications of carrier detection and pre-natal diagnosis. The recent rapid progress has been made possible by the development of DNA amplification using the polymerase chain reaction (pcr) invented by Saiki and colleagues [Saiki, R. K., Chang, C-A., Levenson, C. H., Warren, T. C., Boehm, C. D., Kazazian, H. H. & Ehrlich, H. A. (1988) N. Engl. J. Med. 319, 537-541].  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