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.
By combining anisotropy of small-angle neutron scattering (SANS) and optical anisotropy (linear dichroism, l.d.) on flow-oriented RecA-DNA complexes, the average DNA-base orientation has been determined in RecA complexes with double-stranded (ds) as well as single-stranded (ss) DNA. From the anisotropy of the two-dimensional SANS intensity representation, the second moment orientation function S is obtained. Knowledge of S is crucial for the interpretation of l.d. spectra in terms of orientation of the DNA bases and the aromatic amino acid residues. The DNA-base planes are essentially perpendicular to the fibre axis of the complex between RecA and dsDNA in the presence of cofactor ATP gamma S. A somewhat tilted base geometry is found for the RecA-ATP gamma S complexes with single-stranded poly(dT) and poly(d epsilon A). This behaviour contrasts the RecA-ssDNA complex formed without cofactor which displays a poor orientation of the bases. Well-ordered bases in the ssDNA-RecA complex is possibly reflecting the role of RecA in preparing a nucleotide strand for base-pairing in the search-for-homology process. While the central SANS intensity is essentially independent of the pitch of the helical complex, a secondary intensity maximum, which becomes focused upon flow orientation, is found to be a sensitive measure of the pitch. The pitch values for the complexes compare well with cryo-electron microscopy results but are slightly larger than those seen for uranyl-stained samples.  相似文献   

2.
The interaction between RecA and DNA (in the form of unmodified single-stranded DNA, fluorescent single-stranded DNA and double-stranded DNA) is studied with linear dichroism and fluorescence spectroscopy. RecA is found to form a complex with single-stranded DNA with a binding stoichiometry of about four nucleotides per RecA monomer, in which the DNA bases appear to have a random orientation. Addition of ATP gamma S (a non-hydrolyzable analog of ATP) reduces the stoichiometry to about three nucleotides per RecA and causes the DNA bases to adopt an orientation preferentially perpendicular to the fiber axis. This complex can incorporate an additional strand of single-stranded DNA or double-stranded DNA, yielding a total stoichiometry of six nucleotides or three nucleotides and three base-pairs, respectively, per RecA. RecA, in the presence of ATP gamma S, is also found to interact with double-stranded DNA, with a stoichiometry of about three base-pairs per RecA. In all studied complexes, the tryptophan residues in the RecA protein are oriented with their planes preferentially parallel to the fiber axis, whereas in complexes involving ATP gamma S the planes of the DNA bases are oriented preferentially perpendicular to the fiber. This virtually excludes the possibility that the tryptophan residues are intercalated in the DNA helix. On the basis of these results, a model for the research of homology in the RecA-mediated, strand-exchange reaction in the genetic recombination process is proposed.  相似文献   

3.
The interaction of RecA protein with short single-stranded oligonucleotides is characterised by flow linear dichroism (LD), isoelectric focusing (IEF) and electron microscopy (EM). From LD and EM it is evident that RecA forms long filaments with at least some 50 oligonucleotides in a ‘train formation’. The tendency to form trains is substantially lower when an amino group is attached to the 5′ end of the oligonucleotide, suggesting that the modification impairs protein-protein interactions at the interface between two oligomers. From LD it is also evident that no bridging occurs between RecA–Oligonucleotide complexes containing more than one oligomer strand per RecA filament. This property make them manageable in polyacrylamide gels, hence allowing characterisation by IEF. RecA was found acidic with a pI of 5.0. The pI was not dependent on the presence of bound cofactor (ATPγS) and oligonucleotides suggesting that protonation of the protein readily occurs to compensate for the negative charges provided by bound cofactor and DNA.  相似文献   

4.
We have analyzed the nature of RecA protein-RecA protein interactions using an affinity column prepared by coupling RecA protein to an agarose support. When radiolabeled soluble proteins from Escherichia coli are applied to this column, only the labeled RecA protein from the extract was selectively retained and bound tightly to the affinity column. Efficient binding of purified 35S-labeled RecA protein required Mg2+, and high salt did not interfere with the binding of RecA protein to the column. Complete removal of the bound enzyme from the affinity column required treatment with guanidine HCl (5 M) or urea (8 M). These and other properties suggest that hydrophobic interactions contribute significantly to RecA protein subunit recognition in solution. Using a series of truncated RecA proteins synthesized in vitro, we have obtained evidence that at least some of the sequences involved in protein recognition are localized within the first 90 amino-terminal residues of the protein. Based on the observation that RecA proteins from three heterologous bacteria are specifically retained on the E. coli RecA affinity column, it is likely that this binding domain is highly conserved and is required for interaction and association of RecA protein monomers. Stable ternary complexes of RecA protein and single-stranded DNA were formed in the presence of the nonhydrolyzable ATP analog adenosine 5'-O-(thiotriphosphate) and applied to the affinity columns. Most of the complexes formed with M13 DNA could be eluted in high salt, whereas a substantial fraction of those formed with the oligonucleotide (dT)25-30 remained bound in high salt and were quantitatively eluted with guanidine HCl (5 M). The different binding properties of these RecA protein-DNA complexes likely reflect differences in the availability of a hydrophobic surface on RecA protein when it is bound to long polynucleotides compared to short oligonucleotides.  相似文献   

5.
B J Rao  B Jwang  M Dutreix 《Biochimie》1991,73(4):363-370
During the directional strand exchange that is promoted by RecA protein between linear duplex DNA and circular single-stranded DNA, a triple-stranded DNA intermediate was formed and persisted even after the completion of strand transfer followed by deproteinization. In the deproteinized three-stranded DNA complexes, the sequestered linear third strand resisted digestion by E coli exonuclease I. In relation to polarity of strand exchange which defines the proximal and distal ends of the duplex DNA, when homology was restricted to the distal region of duplex substrate, the joints formed efficiently and were stable even upon complete deproteinization. Enzymatic probing of deproteinized distal joints with nuclease P1 revealed that the joints consist of long three-stranded structures that at neutral pH lack significant single-stranded character in any of the three strands. Instead of circular single-stranded DNA, when a linear single strand is recombined with partially homologous duplex DNA, in the presence of SSB, the formation of homologous joints by RecA protein, is significantly more efficient at distal end than at the proximal. Taken together, these observations suggest that with any single-stranded DNA (circular or linear), RecA protein efficiently promotes the formation of distal joints, from which, however, authentic strand exchange may not occur. Moreover, these joints might represent an intermediate which is trapped into a stable triple stranded state.  相似文献   

6.
In an attempt to understand the role of ATP as a cofactor at the interaction of the RecA protein with DNA, we have studied the orientation geometries of the cofactor analogs adenosine 5′-O-(3-thiotriphosphate) (ATPγS) and guanosine 5′-O-(3-thiotriphosphate) (GTPγS) in RecA-DNA complexes using flow linear dichroism spectroscopy. Both cofactors promote the formation of RecA-DNA complexes of similar structure as judged from similar orientations of DNA bases. The DNA orientation was probed through the dichroism of the long-wavelength absorption of a DNA analog, poly(dεA). In this way differences between the dichroic spectra of the ATPγS–RecA–DNA and GTPγS-RecA-DNA complexes, observed in the shorter-wavelength region, are related to orientation at variations of the cofactor chromophores. The results show that the guanine plane of GTPγS is oriented parallel with the principal axis of the complex in contrast to the more perpendicular orientation of the DNA bases. This observation directly excludes the possibility that the cofactor could be intercalated between the DNA bases. This observation directly excludes the possibility that the cofactor could be intercalated between the DNA bases. The orientation of the adenine base of ATPγS, which may be similar to that of guanine of GTPγS albeit not exactly the same, is also inconsistent with intercalation. The possibility that the cofactor bound to the protein could be intercalated in DNA had been speculated from the observation that some DNA intercalators can induce RecA binding to DNA in the absence of cofator. There are probably no direct interactions between the cofator and the DNA bases and the role of the cofactor is probably related to interaction with RecA and a modification of protein conformation.  相似文献   

7.
Cytoplasmic ribosomes have been isolated from wheat germ, and the structure of ribosomal subunits has been examined by electron microscopy of negatively stained preparations. Small (40 S) subunits show structural features generally regarded as characteristic of eukaryotic particles, while large (60 S) subunits show shapes that are equally well described by models of prokaryotic 50 S particles. Small subunit 18 S RNA contains 2 residues of N6,N6-dimethyladenosine 19 and 20 residues from the 3'-end (Hagenbüchle, O., Santer, M., Steitz, J. A., and Mans, R. J. (1978) Cell 13, 551-563). Nucleoside analysis by high performance liquid chromatography shows no other residues of this component in the RNA. Anti-dimethyladenosine immunoglobulins were reacted with wheat germ 40 S subunits, and the resulting complexes were studied by electron microscopy in order to localize the nucleoside. In about 90% of the complexes observed, antibody-subunit contact was consistent with a single binding site. We place the dimethyladenosine residues at or near the end of the platform of the 40 S particle in a position nearly equivalent to that previously identified in prokaryotic and chloroplast subunits (Trempe, M. R., and Glitz, D. G. (1981) J. Biol. Chem. 256, 11873-11879).  相似文献   

8.
A series of ternary copper(II) complexes have been derived using levofloxacin and five phenanthroline derivatives. Complexes were characterized using infrared spectroscopy, Thermogravimetric (TG)-analysis, fast atom bombardment mass spectroscopy and reflectance spectra. Synthesized complexes exhibit the only d-d band at ~ 666?nm points toward a distorted square pyramidal geometry at metal centre with one unpaired electron responsible for paramagnetic behaviour of whole moiety. Binding behaviour of the complexes toward Herring Sperm DNA were determined using ultraviolet-Vis (UV-Vis) absorption titration and viscometric titration experiment, where as the cleavage efficacy of the complexes toward pUC19 DNA was determined by electrophoresis in presence of ethidium bromide. Complexes exhibit superoxide dismutase-like activity with their IC(50) values ranging from 0.7917 to 1.7432 μM.  相似文献   

9.
A series of ternary copper(II) complexes have been derived using levofloxacin and five phenanthroline derivatives. Complexes were characterized using infrared spectroscopy, Thermogravimetric (TG)-analysis, fast atom bombardment mass spectroscopy and reflectance spectra. Synthesized complexes exhibit the only d-d band at ~ 666?nm points toward a distorted square pyramidal geometry at metal centre with one unpaired electron responsible for paramagnetic behaviour of whole moiety. Binding behaviour of the complexes toward Herring Sperm DNA were determined using ultraviolet-Vis (UV-Vis) absorption titration and viscometric titration experiment, where as the cleavage efficacy of the complexes toward pUC19 DNA was determined by electrophoresis in presence of ethidium bromide. Complexes exhibit superoxide dismutase–like activity with their IC50 values ranging from 0.7917 to 1.7432 µM.  相似文献   

10.
Sedimentation analysis of polyadenylation-specific complexes.   总被引:21,自引:11,他引:10  
Precursor RNA containing the adenovirus L3 polyadenylation site is assembled into a 50S complex upon incubation with HeLa nuclear extract at 30 degrees C. The cofactor and sequence requirements for 50S complex formation are similar to those of the in vitro polyadenylation reaction. Assembly of this complex requires ATP but is not dependent upon synthesis of a poly(A) tract. In addition, a 50S complex does not form on substrate RNA in which the AAUAAA hexanucleotide upstream of the poly(A) site has been mutated to AAGAAA or on RNA in which sequences between +5 and +48 nucleotides downstream of the site have been removed. These mutations also prevent in vitro processing of substrate RNA. Kinetic studies suggest that the 50S complex is an intermediate in the polyadenylation reaction. It forms at an early stage in the reaction and at later times contains both poly(A)+ RNA as well as unreacted precursor. U-type small nuclear ribonucleoprotein particles are components of the 50S complex, as shown by immunoprecipitation with antiserum specific to the trimethyl cap of these small nuclear RNAs.  相似文献   

11.
W J Ray  J M Puvathingal  Y W Liu 《Biochemistry》1991,30(28):6875-6885
Crystals of phosphoglucomutase, grown in 2.1 M ammonium sulfate, "desalted", and suspended in a 30% polyoxyethylene-8000/1 M glycine solution as described in the accompanying paper [Ray, W. J., Jr., Puvathingal, J. M., Bolin, J. T., Minor, W., Liu, Y., & Muchmore, S. W. (1991) Biochemistry 30 (preceding paper in this issue)], were treated with glucose phosphates to form an equilibrium mixture of the catalytically active substrate/product complexes. However, this treatment extensively fractured the crystals, even when very dilute solutions of glucose phosphates were used. But formation of the desired complexes was achieved, without fracturing, by introducing the glucose phosphates at high salt concentration, where they do not bind significantly to the enzyme, and maintaining their presence during subsequent sulfate-removal steps, in order to obtain essentially uniform binding throughout the crystal at all times. Although this procedure produced unfractured crystals of the catalytically active complexes, an adjustment in water activity was required to prevent the crystals from slowly liquefying in the presence of the added glucose phosphates. After this adjustment, the quality of diffraction-grade crystals subjected to this treatment was not significantly altered. An even larger adjustment in water activity was required to stabilize crystals that had been largely converted into a mixture of vanadate-based transition-state analogue complexes [cf. Ray, W. J., Jr., & Puvathingal, J. M. (1990) Biochemistry 29, 2790-2801] by means of an analogous procedure. The rationale for, and the implications of, this adjustment of water activity are discussed. The phenomenon of lattice-based binding cooperativity also is discussed together with a possible role for such cooperativity in the fracturing of protein crystals during formation of ligand complexes and possible ways to circumvent such fracturing based on the annealing of crystals at fractional saturation. An assay for quantifying the extent of formation of the vanadate-based transition-state analogue complexes in crystals of phosphoglucomutase is described. A solution to problems associated with producing and maintaining a steady-state in treated crystals is discussed within the context of maximizing the fraction of the crystalline enzyme present as a complex with one such inhibitor, glucose alpha-1-phosphate-6-vanadate. One of these problems, achieving a substantial reduction in sulfate concentration, could not be successfully addressed by employing the desalting procedure used to produce the substrate/product complexes, because of reduced diffusional rates in the final solution.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
A series of multinuclear macrocyclic polyamine metal (Zn(2+), Cu(2+), Co(2+)) complexes containing chiral dipeptide linkage were synthesized and used as artificial nuclease enzyme model. The interaction between the complexes and plasmid DNA (pUC19) was studied, and the results revealed that these complexes could act as powerful catalysts for the cleavage of plasmid DNA under physiological conditions.  相似文献   

13.
Enzymatic synthesis of polyuridylic acid containing modified bases.   总被引:2,自引:1,他引:1       下载免费PDF全文
Y K Ho  J Aradi    T J Bardos 《Nucleic acids research》1980,8(14):3175-3191
5'Mercaptouridine-5'-diphosphate (hs5UDP) has been synthesized and investigated as a substrate of the polynucleotide phosphorylase of Micrococcus luteus. While hs5UDP is not utilized alone, it can be copolymerized with UDP; however, unusually for this enzyme, the ratio of 5'mercaptouridylate vs. uridylate residues in the polynucleotide product (MPU) is always lower than the ratio of hs5UDP v. UDP in the substrate mixture. Furthermore, hs5UDP decreases the rate of the enzymic polymerization reaction. The MPU product forms two-stranded and three-stranded complexes with poly(A). The circular dichroic spectra of these complexes are similar to those formed between poly(U) and poly(A), but their melting profiles indicate somewhat lower stability. The physicochemical and biochemical properties of the enzymic product are qualitatively similar to those of MPU prepared by chemical modification; both are potent inhibitors of a DNA-dependent RNA polymerase.  相似文献   

14.
Solution structure of complexes between DNA and recombinase RecA from Escherchia coli, in the presence of the physiological cofactor ATP, is probed by flow linear dichroism (LD) spectroscopy. A problem of ADP accumulation which promotes dissociation of DNA-RecA is circumvented by using an ATP-regenerating system. The LD features indicate that the local structure of the complex is very similar to that found in the presence of the non-hydrolysable analog of ATP, adenosine-5'-O-[gamma-thio]triphosphate (ATP[gamma S]); the DNA bases are oriented with their planes preferentially perpendicular to the long axis of the filament, while the indole chromophores of the two tryptophan residues of RecA are rather parallel to this reference direction. A much smaller overall amplitude of the LD spectrum, compared to ATP[gamma S], is interpreted as a result of fast dissociation of RecA due to hydrolysis of ATP, producing transiently naked DNA regions which act like flexible joints, diminishing the macroscopic orientation of the RecA filaments. However, the ATP hydrolysis is not found to prevent simultaneous accommodation of two non-complementary DNA molecules in the RecA complex, as judged from the LD behaviour upon successive addition of two different polynucleotides or modified DNA strands. A notable difference from corresponding complexes formed with ATP[gamma S] is that, in the presence of ATP hydrolysis, the order in which the two DNA molecules have been added is insignificant as judged from virtually identical resulting structures; this observation indicates that exchange of DNA occurs between the two DNA accommodation sites within the RecA filament.  相似文献   

15.
Escherichia coli small ribosomal subunits have been reconstituted from RNA and high performance liquid chromatography-purified proteins including protein S19 that had been modified at its amino-terminal proline residue with 1-fluoro-2,4-dinitrobenzene. As detailed in the accompanying paper (Olah, T. V., Olson, H. M., Glitz, D. G., and Cooperman, B. S. (1988) J. Biol. Chem. 263, 4795-4800), dinitrophenyl (DNP)-S19 was efficiently incorporated into the site ordinarily occupied by S19. Antibodies to DNP bound effectively to the reconstituted subunits and did not cause dissociation of the modified protein from the subunit. Electron microscopy of the immune complexes was used to localize the modified protein on the subunit surface. More than 95% of the antibody binding sites seen were consistent with a single location of protein S19 on the upper portion or head of the subunit, on the surface that faces the 50 S particle in a 70 S ribosome, and in an area relatively distant from the subunit platform. The S19 site is close to the region in which 30 S subunits are photoaffinity labeled with puromycin. Protein S19 is thus near protein S14 in the small subunit and in proximity to the peptidyl transferase center of the 70 S ribosome.  相似文献   

16.
Binding of the recA gene product from Escherichia coli to single-stranded polynucleotides has been investigated using poly(dA) that have been modified by chloroacetaldehyde to yield fluorescent 1,N6-ethenoadenine (epsilon A) bases. A strong enhancement of the fluorescent quantum yield of poly(d epsilon A) is induced upon RecA protein binding. A 4-fold increase is observed in the absence of ATP or ATP gamma S and a 7-fold increase in the presence of either nucleoside triphosphate. RecA protein can bind to poly(d epsilon A) in the absence of both Mg2+ ions and ATP (or ATP gamma S) but Mg2+ ions are required to observe RecA protein binding in the presence of ATP (or ATP gamma S) at pH 7.5. ATP binding to the RecA-poly(d epsilon A) complex induces a dissociation of RecA from the polynucleotide followed by re-binding of [RecA-ATP-Mg2+] ternary complex. Whereas ATP-induced dissociation of RecA-poly(d epsilon A) complexes is a fast process, the subsequent binding reaction of [RecA-ATP-Mg2+] is slow. A model is proposed whereby [RecA-ATP-Mg2+] binding to poly(d epsilon A) involves slow nucleation and elongation processes along the polynucleotide backbone. The nucleation reaction is shown to involve at least a trimer or a tetramer. Polymerization of the [RecA-ATP-Mg2+] ternary complex stops when the polynucleotide is entirely covered with 6 +/- 1 nucleotides per RecA monomer. ATP hydrolysis then induces a release of RecA-ADP complexes from the polynucleotide template.  相似文献   

17.
Previous studies have shown that the region of human urokinase-type plasminogen activator (uPA) responsible for receptor binding resides in the amino-terminal fragment (ATF, residues 1-135) (Stoppelli, M.P., Corti, A., Soffientini, A., Cassani, G., Blasi, F., and Assoian, R.K. (1985) Proc. Natl. Acad. Sci. U.S. A. 82, 4939-4943). The area within ATF responsible for specific receptor binding has now been identified by the ability of different synthetic peptides corresponding to different regions of the amino terminus of uPA to inhibit receptor binding of 125I-labeled ATF. A peptide corresponding to human [Ala19]uPA-(12-32) resulted in 50% inhibition of ATF binding at 100 nM. Peptides uPA-(18-32) and [Ala13]uPA-(9-20) inhibit at 100 and 2000 microM, respectively. The human peptide uPA-(1-14) and the mouse peptide [Ala20]uPA-(13-33) have no effect on ATF receptor binding. This region of uPA is referred to as the growth factor module since it shares partial amino acid sequence homology (residues 14-33) to epidermal growth factor (EGF). Furthermore, this region of EGF is responsible for binding of EGF to its receptor (Komoriya, A. Hortsch, M., Meyers, C., Smith, M., Kanety, H., and Schlessinger, J. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 1351-1355). However, EGF does not inhibit ATF receptor binding. Comparison of the sequences responsible for receptor binding of uPA and EGF indicate that the region of highest homology is between residues 13-19 and 14-20 of human uPA and EGF, respectively. In addition, there is a conservation of the spacings of four cysteines in this module whereas there is no homology between residues 20-30 and 21-33 of uPA and EGF. Thus, residues 20-30 of uPA apparently confer receptor binding specificity, and residues 13-19 provide the proper conformation to the adjacent binding region.  相似文献   

18.
The crystal structure of mandelate racemase (MR) has been solved at 3.0-A resolution by multiple isomorphous replacement and subsequently refined against X-ray diffraction data to 2.5-A resolution by use of both molecular dynamics refinement (XPLOR) and restrained least-squares refinement (PROLSQ). The current crystallographic R-factor for this structure is 18.3%. MR is composed of two major structural domains and a third, smaller, C-terminal domain. The N-terminal domain has an alpha + beta topology consisting of a three-stranded antiparallel beta-sheet followed by an antiparallel four alpha-helix bundle. The central domain is a singly wound parallel alpha/beta-barrel composed of eight central strands of beta-sheet and seven alpha-helices. The C-terminal domain consists of an irregular L-shaped loop with several short sections of antiparallel beta-sheet and two short alpha-helices. This C-terminal domain partially covers the junction between the major domains and occupies a region of the central domain that is filled by an eight alpha-helix in all other known parallel alpha/beta-barrels except for the barrel domain in muconate lactonizing enzyme (MLE) [Goldman, A., Ollis, D. L., & Steitz, T. A. (1987) J. Mol. Biol. 194, 143] whose overall polypeptide fold and amino acid sequence are strikingly similar to those of MR [Neidhart, D. J., Kenyon, G. L., Gerlt, J. A., & Petsko, G. A. (1990) Nature 347, 692]. In addition, the crystal structure reveals that, like MLE, MR is tightly packed as an octamer of identical subunits. The active site of MR is located between the two major domains, at the C-terminal ends of the beta-strands in the alpha/beta-barrel domain. The catalytically essential divalent metal ion is ligated by three side-chain carboxyl groups contributed by residues of the central beta-sheet. A model of a productive substrate complex of MR has been constructed on the basis of difference Fourier analysis at 3.5-A resolution of a complex between MR and (R,S)-p-iodomandelate, permitting identification of residues that may participate in substrate binding and catalysis. The ionizable groups of both Lys 166 and His 297 are positioned to interact with the chiral center of substrate, suggesting that both of these residues may function as acid/base catalysts.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
We have already determined the primary structure of the endogenous inhibitor for calcium-dependent protease (CANP inhibitor, calpastatin) from the cDNA sequence and revealed that the CANP inhibitor contains four internally repeating units which could be responsible for its multiple reactive sites (Emori, Y., Kawasaki, H., Imajoh, S., Imahori, K., and Suzuki, K. (1987) Proc. Natl. Acad. Sci. U.S.A. 84, 3590-3594). Restriction fragments of the cDNA corresponding to each of the four domains (encoding 104-156 amino acid residues of the total 718 residues) were subcloned into the multicloning site of pUC9 or pUC18 in a direction and frame matched to the lacZ' open reading frame of the vector. Under the lac operator-promoter system, we succeeded in producing truncated fragments of the CANP inhibitor in Escherichia coli. The CANP inhibitor fragments were partially purified, and the inhibitory activities toward calcium-dependent protease (CANP) were examined. All fragments containing well conserved regions of about 30 amino acid residues (domains I-IV) located in the middle of the four units exhibited the inhibitory activity. However, their inhibitory activities varied considerably. Further truncation experiments revealed that small fragments containing 30-70 amino acid residues of the CANP inhibitor still retained inhibitory activity. From these experimental results the following conclusions can be drawn: 1) each of the four repeating units of the CANP inhibitor (about 140 amino acid residues) is a real functional unit and can inhibit CANP activity independently; and 2) domains corresponding to well conserved sequences of about 30 amino acid residues containing a consensus Thr-Ile-Pro-Pro-X-Tyr-Arg sequence are essential for the inhibitory activity, and the bordering regions are important for its modulation.  相似文献   

20.
Nucleoside analysis of the RNA from the small subunit of wheat germ cytoplasmic ribosomes shows 1 mol each of N7-methylguanosine and N6-methyladenosine/mol of RNA. Antibodies directed against each methylated nucleoside were used to localize these residues within the subunit by electron microscopy of immune complexes. Antibodies to 7-methylguanosine bound 40 S subunits at a single site, at or slightly above the division between the upper and lower segments of the particle and on the surface furthest from the platform (or large lobe) of the subunit. This site is essentially equivalent to that previously seen with Escherichia coli and chloroplast 30 S subunits (Trempe, M. R., Ohgi, K., and Glitz, D. G. (1982) J. Biol. Chem. 257, 9822-9829). Antibodies to N6-monomethyladenosine were induced in rabbits with a nucleoside-albumin conjugate and shown to be specific for the modified nucleoside. Electron microscopy of antibody-subunit complexes placed the methyladenosine residue in a position that is essentially indistinguishable from that of 7-methylguanosine.  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