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 819 毫秒
1.
2.
R E Hurd  E Azhderian  B R Reid 《Biochemistry》1979,18(18):4012-4017
We have studied the effects of Co2+ and Mn2+ ions on the low-field nuclear magnetic resonance (NMR) spectra of pure class 1 transfer ribonucleic acid (tRNA) species. With 1.2 mM tRNA in the presence of 15 mM MgCl2 discrete paramagnetic effects were observed for Co2+ at concentrations in the range 0.02--0.1 mM and for Mn2+ in the range 0.002--0.01 mM, indicating fast exchange of these cations with tRNA. Both of these cations paramagnetically relax the s4U8--A14 resonance as well as other resonances from proximal base pairs. The Co2+ site appears to be the same site on G15 which was observed crystallographically [Jack, A., Ladner, J. E., Rhodes, D., Brown, R. S., & Klug, A. (1977) J. Mol. Biol. 111, 315-328]; the initially occupied tight Mn2+ site is the cation site involving the phosphate of U8. There are three base pairs within 10 A of both sites, namely, G15--C48, A14--s4U8, and C13--G22; this has led to the assignment of the G15--C48 and C13--G22 resonances in the NMR spectrum [Jack, A., Ladner, J. E., Rhodes, D., Brown, R. S., & Klug, A. (1977) J. Mol. Biol. 111, 315--328; Holbrook, S. R., Sussman, J. L., Warrant, R. W., Church, G. M., & Kim, Sung-Hou (1977) Nucleic Acids Res. 4, 2811--2820; Quigley, G. J., Teeter, M. M., & Rich, A. (1978) Proc. Natl. Acad. Sci. U.S.A. 75, 64--68].  相似文献   

3.
The contractile sheaths of five defective, PBS X-like bacteriophages from Bacillus subtilis and B. licheniformis were investigated by electron microscopy, dodecylsulphate gel electrophoresis and immunodiffusion. Electron microscope images of the extended and contracted sheaths were of similar appearance, although their lengths were different. The surface lattices of both the extended and the contracted sheaths were determined by optical diffraction. This showed that the quaternary structure of the sheaths of all five defective phages originated from identical surface lattices, which could be approximately expressed by the selection rules L = -2n' + 3m and L = 9N' + 17M for the extended and contracted sheaths respectively, in which 6n' = n with n = 0 or an integer multiple of 6. These results indicated that the packing of the protein subunits in these sheaths differed from those of other bacteriophages, for example T4 and millimicron [Amos and Klug, J. Mol. Biol. 99, 51--73 (1975); Admiraal and Mellema, J. Ultrastruct. Res. 56, 48--64 (1976)]. The molecular weight of the main sheath protein of the defective phages, as determined by dodecylsulphate gel electrophoresis, was approximately 50000. This value differed from that for T4, but was similar to that of millimicron [Admiraal and Mellema, J. Ultrastruct. Res. 56, 48--64 (1976); King and Laemmli, J. Mol. Biol, 75, 315--337 (1973)]. The results of immunodiffusion experiments, however, pointed to a chemical difference between the sheath proteins of the defective phages and millimicron, in addition to T4.  相似文献   

4.
The organization of Arabidopsis thaliana photosystem II (PSII) and its associated light-harvesting antenna (LHCII) was studied in isolated PSII-LHCII supercomplexes and native membrane-bound crystals by transmission electron microscopy and image analysis. Over 4000 single-particle projections of PSII-LHCII supercomplexes were analyzed. In comparison to spinach supercomplexes [Boekema, E.J., van Roon, H., van Breemen, J.F.L. & Dekker, J.P. (1999) Eur. J. Biochem. 266, 444-452] some striking differences were revealed: a much larger number of supercomplexes from Arabidopsis contain copies of M-type LHCII trimers. M-type trimers can also bind in the absence of the more common S-type trimers. No binding of l-type trimers could be detected. Analysis of native membrane-bound PSII crystals revealed a novel type of crystal with a unit cell of 25.6 x 21.4 nm (angle 77 degrees ), which is larger than any of the PSII lattices observed before. The data show that the unit cell is built up from C2S2M2 supercomplexes, rather than from C2S2M supercomplexes observed in native membrane crystals from spinach [Boekema, E.J., Van Breemen, J.F.L., Van Roon, H. & Dekker, J.P. (2000) J. Mol. Biol. 301, 1123-1133]. It is concluded from both the single particle analysis and the crystal analysis that the M-type trimers bind more strongly to PSII core complexes in Arabidopsis than in spinach.  相似文献   

5.
The possibility suggested recently [Hider, R.C., Fern, E.B. and London, D.R. (1969) Biochem. J. 114, 171-178; Hider, R.C., Fern, E.B. and London, D.R. (1971) Biochem. J. 121, 817-827; van Venrooij, W.J., Poort, C., Kramer, M.F. and Jansen, M.T. (1972) Eur. J. Biochem. 30, 427-433; and Adamson, L.F., Herington, A.C. and Bornstein, J. (1972) Biochim. Biophys. Acta 282, 352-365] that protein synthesis takes place using amino acids directly from the membrane transport system and not from an intracellular pool has been investigated in rat heart. The tissue was perfused first for 30 min with either [14C]glycine or [14C]leucine and then for a further 30 min with identical medium containing [3H]glycine or [3H]leucine, respectively. After an initial lag, [14C]glycine was incorporated into protein at a linear rate up to 60 min. The [3H]glycine was accumulated into tissue water and incorporated just as readily as the [14C]glycine had been. The rate of total protein synthesis agrees with literature values only if intracellular and not extracellular specific activity values are used in the calculation. Some glycine was converted to serine or threonine. Leucine influx and efflux were very rapid in contrast to the relatively slow exchange reported for incubated tissues [Hider, R.C., Fern, E.B. and London, D.R. (1969) Biochem. J. 114, 171-178; Hider, R.C., Fern, E.B. and London, D.R. (1971) Biochem. J. 121, 817-827; van Venrooij, W.J., Poort, C., Kramer, M.F. and Jansen, M.T. (1972) Eur. J. Biochem. 30, 427-433]. The results are consistent with the existence of an intracellular precursor pool for glycine. Some possible reasons for the discrepancies between this and the other studies are discussed.  相似文献   

6.
R Rauhut  A Jger  C Conrad    G Klug 《Nucleic acids research》1996,24(7):1246-1251
The large subunit ribosomal RNA of the purple bacterium Rhodobacter capsulatus shows fragmentation into pieces of 14 and 16S, both fragments forming the functional equivalent of intact 23S rRNA. An RNA-processing step removes an extra stem-loop structure from the 23S rRNA [Kordes, E., Jock, S., Fritsch, J., Bosch, F. and Klug, G. (1994) J. Bacteriol., 176, 1121-1127]. Taking advantage of the fragmentation deficient mutant strain Fm65, we used genetic complementation to find the mutated gene responsible for this aberration. It was identified as the Rhodobacter homologue to mc from Escherichia coli encoding endoribonuclease III (RNase III). The predicted protein has 226 amino acids with a molecular weight of 25.5 kDa. It shares high homology with other known RNase III enzymes over the full length. In particular it shows the double-stranded RNA-binding domain (dsRBD) motif essential for binding of dsRNA substrates. The Fm65 mutant has a frame shift mutation resulting in complete loss of the dsRBD rendering the enzyme inactive. The cloned Rhodobacter enzyme can substitute RNase III activity in an RNase III deficient E. coli strain. Contrary to E. coli, the Rhodobacter mc is in one operon together with the lep gene encoding the leader peptidase.  相似文献   

7.
To evaluate their role in the active site of the UDP-N-acetylmuramoyl-L-alanine:D-glutamate ligase (MurD) from Escherichia coli, 12 residues conserved either in the Mur superfamily [Eveland, S. S., Pompliano, D. L., and Anderson, M. S. (1997) Biochemistry 36, 6223-6229; Bouhss, A., Mengin-Lecreulx, D., Blanot, D., van Heijenoort, J., and Parquet, C. (1997) Biochemistry 36, 11556-11563] or in the sequences of 26 MurD orthologs were submitted to site-directed mutagenesis. All these residues lay within the cleft of the active site of MurD as defined by its 3D structure [Bertrand, J. A., Auger, D., Fanchon, E., Martin, L., Blanot, D., van Heijenoort, J., and Dideberg, O. (1997) EMBO J. 16, 3416-3425]. Fourteen mutant proteins (D35A, K115A, E157A/K, H183A, Y194F, K198A/F, N268A, N271A, H301A, R302A, D317A, and R425A) containing a C-terminal (His)(6) extension were prepared and their steady-state kinetic parameters determined. All had a reduced enzymatic activity, which in many cases was very low, but no mutation led to a total loss of activity. Examination of the specificity constants k(cat)/K(m) for the three MurD substrates indicated that most mutations affected both the binding of one substrate and the catalytic process. These kinetic results correlated with the assigned function of the residues based on the X-ray structures.  相似文献   

8.
Amino acid sequences were analyzed for two isoforms of myosin essential light chain, LC17a and LC17b [Hasegawa, Y., Ueno, H., Horie, K., & Morita, F. (1988) J. Biochem. 103, 15-18] from porcine aorta smooth muscle. Both LC17a and LC17b consisted of 150 amino acid residues and their N-terminal Cys residues were blocked by an acetyl group. The amino acid sequences of LC17a and LC17b were common from the N-terminal to Glu-141 and five amino acid substitutions were observed within the remaining C-terminal 9 residues. The amino acid sequences of LC17a and LC17b were identical to those deduced from the nucleotide sequences of bovine aortic cDNAs encoding the two isoforms [Lash, J. A., Helper, D.J., Klug, M., Nicolozakes, A.W., & Hathaway, D.R. (1990) Nucleic Acids Res. 18, 7176].  相似文献   

9.
Khan MA  Goss DJ 《Biochemistry》2005,44(11):4510-4516
Previous kinetic binding studies of wheat germ protein synthesis eukaryotic translational initiation factor eIFiso4F and its subunit, eIFiso4E, with m(7)GTP and mRNA analogues indicated that binding occurred by a two-step process with the first step occurring at a rate close to the diffusion-controlled rate [Sha, M., Wang, Y., Xiang, T., van Heerden, A., Browning, K. S., and Goss, D. J. (1995) J. Biol. Chem. 270, 29904-29909]. The kinetic effects of eIF4B, PABP, and wheat germ eIFiso4F with two mRNA cap analogues and the temperature dependence of this reaction were measured and compared. The Arrhenius activation energies for binding of the two mRNA cap analogues, Ant-m(7)GTP and m(7)GpppG, were significantly different. Fluorescence stopped-flow studies of the eIFiso4F.eIF4B protein complex with two m(7)G cap analogues show a concentration-independent conformational change. The rate of this conformational change was approximately 2.4-fold faster for the eIFiso4F.eIF4B complex compared with our previous studies of eIFiso4F [Sha, M., Wang, Y., Xiang, T., van Heerden, A., Browning, K. S., and Goss, D. J. (1995) J. Biol. Chem. 270, 29904-29909]. The dissociation rates were 3.7- and 5.4-fold slower for eIFiso4F.Ant-m(7)GTP and eIFiso4F.m(7)GpppG, respectively, in the presence of eIF4B and PABP. These studies show that eIF4B and PABP enhance the interaction with the cap and probably are involved in protein-protein interactions as well. The temperature dependence of the cap binding reaction was markedly reduced in the presence of either eIF4B or PABP. However, when both eIF4B and PABP were present, not only was the energy barrier reduced but the binding rate was faster. Since cap binding is thought to be the rate-limiting step in protein synthesis, these two proteins may perform a critical function in regulation of the overall protein synthesis efficiency. This suggests that the presence of both proteins leads to a rapid, stable complex, which serves as a scaffold for further initiation complex formation.  相似文献   

10.
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   

11.
An integrated magneto-biostratigraphic framework is presented for Middle Miocene sediments of DSDP Site 372 located in the Western Mediterranean. Detailed biostratigraphic analysis shows a nearly complete sequence of early Middle Miocene calcareous plankton bioevents in the Mediterranean, including the LCO (Last Common Occurrence) of the nannofossil Sphenolithus heteromorphus which has been astronomically dated in the Ras il Pellegrin (RIP) section on Malta Island [Abels, H.A., Hilgen, F.J., Krijgsman, W., Kruk, R.W., Raffi, I., Turco, E., Zachariasse, W.J., 2005. Long-period orbital control on middle Miocene global cooling: integrated stratigraphy and astronomical tuning of the Blue Clay Formation on Malta, Paleoceanography, 20, PA4012. doi: 10.1029/2004PA001129. 11 pp]. Thermal demagnetization of discrete samples revealed a characteristic low-temperature component with dual polarities despite a weak paleomagnetic signal. The resultant magnetostratigraphic record, combined with the calcareous plankton biostratigraphy, is straightforwardly correlated to the geomagnetic polarity time scale (CK95) of Cande and Kent [Cande, S.C., Kent, D.V., 1995. Revised calibration of the Geomagnetic Polarity Time Scale for the Late Cretaceous and Cenozoic, J. Geophys. Res., 100, 6093–6095] and the Astronomical Tuned Neogene Time Scale (ATNTS04) of Lourens et al. [Lourens, L.J., Hilgen, F.J., Laskar, J., Shackleton, N.J., Wilson, D., 2004. The Neogene Period. In: Gradstein F.M., Ogg J.G., Smith A.G. (Eds.), A Geologic Time Scale, Cambridge Univ. Press, pp. 409–440]. The subchrons recorded in Site 372 succession range from C5Br up to C5AAr. To confirm the magnetostratigraphic calibration, 40Ar/39Ar dating was performed on feldspar of two volcanic ash layers. The radio-isotopic dating indicates a younger age for these two ash layers compared to the magnetostratigraphic calibrated ages according to the CK95 and ATNTS04 age models. However, if the astronomically calibrated age of 28.21 ± 0.04 Ma is used for the Fish Canyon standard (FCs), the age for the older ash layer exactly matches its ATNTS04 age. Ages for bioevents were calculated assuming constant sedimentation rates between magnetostratigraphic age-tie points. The S. heteromorphus LCO has an age of 13.54 Ma and 13.63 Ma according to CK95 and ATNTS04, respectively, which is consistent with the astronomical tuned age of 13.65 Ma determined at RIP section [Abels, H.A., Hilgen, F.J., Krijgsman, W., Kruk, R.W., Raffi, I., Turco, E., Zachariasse, W.J., 2005. Long-period orbital control on middle Miocene global cooling: integrated stratigraphy and astronomical tuning of the Blue Clay Formation on Malta, Paleoceanography, 20, PA4012. doi:10.1029/2004PA001129. 11 pp]. We therefore conclude that the magnetostratigraphic calibration of DSDP Site 372 is correct and that, for the time being, this site can be considered as a reference section for the early Middle Miocene of the Mediterranean region.  相似文献   

12.
D E Robertson  F Daldal  P L Dutton 《Biochemistry》1990,29(51):11249-11260
Seven single-site mutants in six residues of the cyt b polypeptide of Rhodobacter capsulatus selected for resistance to the Qo site inhibitors stigmatellin, myxothiazol, or mucidin [Daldal, F., Tokito, M.K., Davidson, E., & Faham, M. (1989) EMBO J. 8, 3951-3961] have been characterized by using optical and EPR spectroscopy and single-turnover kinetic analysis. The strains were compared with wild-type strain MT1131 and with the Ps- strain R126 (G158D), which is dysfunctional in its Qo site [Robertson, D.E., Davidson, E., Prince, R.C., van den Berg, W.H., Marrs, B.L., & Dutton, P.L. (1986) J. Biol. Chem. 261, 584-591]. Mutants selected for stigmatellin resistance induced a weakening in the binding of the inhibitor without discernible loss of ubiquinone(Q)/ubiquinol(QH2) binding affinity to the Qo site or kinetic impairment to catalysis. Mutants selected for myxothiazol or mucidin resistance, inducing weakening of inhibitor binding, all displayed impaired rates of Qo site catalysis: The most severe cases (F144L, F144S) displayed loss of affinity for Q, and evidence suggests that parallel loss of affinity for the substrate QH2 was incurred in these strains. The results provide a view of the nature of the interaction of Q and QH2 of the Qpool with the Qo site. Consideration of the mutational substitutions and their structural positions along with comparisons with the QA and QB sites of the photosynthetic reaction center suggests a model for the structure of the Qo site.  相似文献   

13.
E I Hyde  B R Reid 《Biochemistry》1985,24(16):4315-4325
The effects of magnesium, spermine, and temperature on the conformation of Escherichia coli tRNAPhe have been examined by proton and phosphorus nuclear magnetic resonance spectroscopy. In the low-field proton NMR spectra we have characterized two slowly interconverting conformations of this tRNA at low magnesium ion concentrations. The relative proportion of the conformers is ion dependent but not ion specific. Magnesium affects protons in all the stems of tRNA while spermine effects are localized near the s4U-8-A-14 and G-15-C-48 tertiary bonds. The effects seen in the proton NMR spectra are compared and correlated with those observed in the phosphorus spectra to give assignments of some of the resolved signals from the phosphate groups. The phosphorus spectra are compared with those of yeast tRNAPhe [Gorenstein, D. G., Goldfield, E. M., Chen, R., Kovar, K., & Luxon, B. A. (1981) Biochemistry 20, 2141; Salemink, P. J. M., Reijerse, E. J., Mollevanger, L., & Hilbers, C. W. (1981) Eur. J. Biochem. 115, 635], and the ion effects are discussed with reference to the magnesium and spermine sites found in the crystal structures of yeast tRNAPhe [Holbrook, S. R., Sussman, J. L., Warrant, R. W., Church, G. M., & Kim, S.-H. (1977) Nucleic Acids Res. 4, 2811; Quigley, G. J., Teeter, M. M., & Rich, A. (1978) Proc. Natl. Acad. Sci. U.S.A. 75, 64; Jack, A., Ladner, J. E., Rhodes, D., Brown, R. S., & Klug, A. (1977) J. Mol. Biol. 111, 315].  相似文献   

14.
Goldfarb NE  Lam MT  Bose AK  Patel AM  Duckworth AJ  Dunn BM 《Biochemistry》2005,44(48):15725-15733
Human cathepsin D (hCatD) is an aspartic peptidase with a low pH optimum. X-ray crystal structures have been solved for an active, low pH (pH 5.1) form (CatD(lo)) [Baldwin, E. T., Bhat, T. N., Gulnik, S., Hosur, M. V., Sowder, R. C., Cachau, R. E., Collins, J., Silva, A. M., and Erickson, J. W. (1993) Proc. Natl. Acad. Sci. U.S.A. 90, 6796-6800] and an inactive, high pH (pH 7.5) form (CatD(hi)) [Lee, A. Y., Gulnik, S. V., and Erickson, J. W. (1998) Nat. Struct. Biol. 5, 866-871]. It has been suggested that ionizable switches involving the carboxylate side chains of E5, E180, and D187 may mediate the reversible interconversion between CatD(hi) and CatD(lo) and that Y10 stabilizes CatD(hi) [Lee, A. Y., Gulnik, S. V., and Erickson, J. W. (1998) Nat. Struct. Biol. 5, 866-871]. To test these hypotheses, we generated single point mutants in "short" recombinant human pseudocathepsin D (srCatD), a model kinetically similar to hCatD [Beyer, B. M., and Dunn, B. M. (1996) J. Biol. Chem. 271, 15590-15596]. E180Q, Y10F, and D187N exhibit significantly higher kcat/Km values (2-, 3-, and 6-fold, respectively) at pH 3.7 and 4.75 compared to srCatD, indicating that these residues are important in stabilizing the CatD(hi). E5Q exhibits a 2-fold lower kcat/Km compared to srCatD at both pH values, indicating the importance of E5 in stabilizing the CatD(lo). Accordingly, full time-course "pH-jump" (pH 5.5-4.75) studies of substrate hydrolysis indicate that E180Q, D187N, and Y10F have shorter kinetic lag phases that represent the change from CatD(hi) to CatD(lo) compared to srCatD and E5Q. Intrinsic tryptophan fluorescence reveals that the variants have a native-like structure over the pH range of our assays. The results indicate that E180 and D187 participate as an electrostatic switch that initiates the conformational change of CatD(lo) to CatD(hi) and Y10 stabilizes CatD(hi) by hydrogen bonding to the catalytic Asp 33. E5 appears to play a less significant role as an ionic switch that stabilizes CatD(lo).  相似文献   

15.
Proton NMR studies at 300 MHz and 500 MHz in aqueous solution were carried out on the modified self-complementary tetranucleoside trisphosphate d(T-A'-T-A'), in which d(A') represents 2-aminodeoxyadenosine. The NMR spectra were observed at two sample concentrations over the temperature range 2-70 degrees C. Assignments, based on homonuclear decoupling and nuclear Overhauser enhancement (NOE) experiments, are given. The concentration dependence of the chemical shift vs temperature profiles was used to extract information concerning duplex formation. The proton-proton and proton-phosphorus coupling constants were obtained at four temperatures and yielded accurate conformational data on the sugar ring and on the back-bone angles beta, gamma and delta. From the observed line broadening, shift profiles, NOEs, and from the observation of imino proton resonances, it is concluded that the compound exists as a miniduplex at low temperature. Comparison of these observations with similar observations on the parent compound d(T-A-T-A) indicate that substitution of dA by dA' increases the tendency towards duplex formation. At low temperature the compound adopts a stacked B-DNA type structure: destacking occurs on raising the temperature. The sequence-dependent sugar ring geometry [Mellema, J.-R., Pieters, J. M. L., van der Marel, G. A., van Boom, J. H., Haasnoot, C. A. G. & Altona, C. (1984) Eur. J. Biochem. 143, 285-301] is present in this molecule. The conformational parameters of d(T-A'-T-A') and d(T-A-T-A) are quite similar, thus substitution of dA by dA' has no measurable influence on the geometry of the sugar ring or on the backbone angles beta, gamma and delta.  相似文献   

16.
The complete amino acid sequence of a basic non-histone protein, H6, isolated from the chromatin of rainbow trout (Salmo gairdnerii) testis cells, has been determined. Protein H6, first described by D. T. Wigle and G. H. Dixon [J. Biol. Chem. 246, 5636--5644 (1971)] was extracted with 5% trichloracetic acid and purified by ion-exchange chromatography on carboxymethyl-cellulose (CM-52). Sequence analysis was performed by automatic Edman degradation of the amino terminus of the intact protein and a series of large fragments derived by cleavage with chymotrypsin, staphylococcal protease and with mild acid to cleave at aspartic acid residues. Protein H6 possesses 69 residues and shows considerable similarities to the 89-residue calf thymus HMG-17 protein previously sequenced [Walker, J. M., Hastings, J. R. B. & Johns, E. W. (1977) Eur. J. Biochem. 76, 461--468]. B. Levy W. and G. H. Dixon [Proc. Natl Acad. Sci. U.S.A. 74, 2810--2814 (1977)] have shown that H6 is selectively solubilized when trout testis nuclei (or chromatin) are digested with DNase I under conditions which preferentially hydrolyze that portion of DNA enriched in transcribed sequences [Levy, W. B. & Dixon, G. H. (1977) Nucleic Acids Res. 4, 883--898]. Recently H6 has been located as a stoichiometric component of a distinct subset of trout testis nucleosomes that are complexed with a core nucleosome comprising 140 base pairs of DNA and the inner histones H2A, H2B, H3 and H4 [Levy, W. B., Connor, W. & Dixon, G. H. (1979) J. Biol. Chem., in the press].  相似文献   

17.
The reduction of nitrite to nitric oxide in respiratory denitrification is catalyzed by a cytochrome cd(1) nitrite reductase in Paracoccus pantotrophus (formerly known as Thiosphaera pantotropha LMD 92.63). High-resolution structures are available for the fully oxidized [Fül?p, V., Moir, J. W., Ferguson, S. J., and Hajdu, J. (1995) Cell 81, 369-377; Baker, S. C., Saunders, N. F., Willis, A. C., Ferguson, S. J., Hajdu, J., and Fül?p, V. (1997) J. Mol. Biol. 269, 440-455] and fully reduced forms of this enzyme, as well as for various intermediates in its catalytic cycle [Williams, P. A., Fül?p, V., Garman, E. F., Saunders, N. F., Ferguson, S. J., and Hajdu, J. (1997) Nature 389, 406-412]. On the basis of these structures, quantum mechanical techniques (QM), including density functional methods (DFT), were combined with simulated annealing (SA) and molecular mechanics techniques (MM) to calculate the electronic distribution of molecular orbitals in the active site during catalysis. The results show likely trajectories for electrons, protons, substrates, and products in the process of nitrite reduction, and offer an interpretation of the reaction mechanism. The calculations indicate that the redox state of the d(1) heme and charges on two histidines in the active site orchestrate catalysis locally. Binding of nitrite to the reduced iron is followed by proton transfer from His345 and His388 to one of the oxygens of nitrite, creating a water molecule and an [Fe(II)-NO(+)] complex. Valence isomerization within this complex gives [Fe(III)-NO]. The release of NO from the ferric iron is influenced by the protonation state of His345 and His388, and by the orientation of NO on the d(1) heme. Return of Tyr25 to a hydrogen-bonding position between His345 and His388 facilitates product release, but a rebinding of Tyr25 to the oxidized iron may be bypassed in steady-state catalysis.  相似文献   

18.
Previous X-ray studies (2.8-A resolution) on the crystals of tobacco mosaic virus protein (TMVP) grown from solutions containing high salt have characterized the structure of the protein aggregate as a bilayered cylindrical disk formed by 34 identical subunits [Bloomer, A.C., Champness, J.N., Bricogne, G., Staden, R., & Klug, A. (1978) Nature (London) 276, 362-368]. Under low-salt conditions, 20S aggregates are in equilibrium with 4S species and involved in the efficient nucleation of TMV assembly in vitro [Butler, P.J.G. (1984) J. Gen. Virol. 65, 253-279]. We have investigated by sedimentation velocity and near-UV circular dichroism (CD) measurements the structure of 20S aggregates in low salt (I = 0.1 potassium phosphate at pH 7.0 and 20 degrees C) and the aggregates in high salt [0.2 M (NH4)2SO4 in I = 0.1 tris(hydroxymethyl)aminomethane hydrochloride at pH 8.0 and 20 degrees C, close to the conditions under which TMVP crystallizes as disk aggregates]. At high salt, we observe structures (presumably stacks of disks) having s20,w values around 40, 45, and 50 S, but not the 20S species present in low-salt buffers. The near-UV CD spectrum of 20S aggregates has been obtained for the first time, using computer techniques, from the spectra of the 4S-20S equilibrium mixture and the 4S species. This spectrum of 20S aggregates differs dramatically from that of the stacks of disks examined at both high and low salt (into which the stacks can be returned by dialysis), indicating that the difference is not a solvent effect.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
20.
The electron-transfer site of spinach plastocyanin   总被引:2,自引:0,他引:2  
J D Rush  F Levine  W H Koppenol 《Biochemistry》1988,27(16):5876-5884
Two sites for electron transfer have been proposed for plastocyanin: one near the copper ion and the other close to the acid patch formed by residues 42-45. Calculations of electrostatic properties of spinach plastocyanin and ionic strength dependences of electron-transfer reactions of this protein have been used to distinguish between these two sites. Calculations show that the electric potential field of spinach plastocyanin is highly asymmetric and that the protein has a dipole moment of 360 D. The negative end of the dipole axis emerges between the negative patches formed by residues 42-45, which is though to be the cation binding site, and residues 59-61. The angles between the dipole vector and vectors from the center of mass to the copper ion and to the acid patch are 90 degrees and 30 degrees, respectively. The angle between the dipole vector and a line from the center of mass to the site of electron transfer is evaluated from the ionic strength dependence of electron-transfer rates at pH 7.8 with the help of equations developed by Van Leeuwen et al. [van Leeuwen, J.W., Mofers, F.J.M., & Veerman, E.C.I. (1981) Biochim. Biophys. Acta 635, 434] and Van Leeuwen [van Leeuwen, J.W. (1983) Biochim. Biophys. Acta 743, 408]. The angles found are 85 degrees, 110 degrees, and 75 +/- 15 degrees for reactions with tris(1,10-phenanthroline)cobalt(III), hexacyanoferrate(III), and ferrocytochrome c, respectively. The electric potential field calculations suggest that the hexacyanoferrate(III) interaction angle corresponds to a unique site of minimum repulsion at the hydrophobic region of the protein surface, close to the copper ion.(ABSTRACT TRUNCATED AT 250 WORDS)  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