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.
Right-handed RNA duplexes of (CG)n sequence undergo salt-induced helicity reversal, forming left-handed RNA double helices (Z-RNA). In contrast to the thoroughly studied Z-DNA, no Z-RNA structure of natural origin is known. Here we report the NMR structure of a half-turn, left-handed RNA helix (CGCGCG)2 determined in 6 M NaClO4. This is the first nucleic acid motif determined at such high salt. Sequential assignments of non-exchangeable proton resonances of the Z-form were based on the hitherto unreported NOE connectivity path [H6(n)-H5′/H5″(n)-H8(n+1)-H1′(n+1)-H6(n+2)] found for left-handed helices. Z-RNA structure shows several conformational features significantly different from Z-DNA. Intra-strand but no inter-strand base stacking was observed for both CpG and GpC steps. Helical twist angles for CpG steps have small positive values (4–7°), whereas GpC steps have large negative values (−61°). In the full-turn model of Z-RNA (12.4 bp per turn), base pairs are much closer to the helix axis than in Z-DNA, thus both the very deep, narrow minor groove with buried cytidine 2′-OH groups, and the major groove are well defined. The 2′-OH group of cytidines plays a crucial role in the Z-RNA structure and its formation; 2′-O-methylation of cytidine, but not of guanosine residues prohibits A to Z helicity reversal.  相似文献   

2.
The solution structure and hydration of the chimeric duplex [d(CGC)r(aaa)d(TTTGCG)]2, in which the central hybrid segment is flanked by DNA duplexes at both ends, was determined using two-dimensional NMR, simulated annealing and restrained molecular dynamics. The solution structure of this chimeric duplex differs from the previously determined X-ray structure of the analogous B-DNA duplex [d(CGCAAATTTGCG)]2 as well as NMR structure of the analogous A-RNA duplex [r(cgcaaauuugcg)]2. Long-lived water molecules with correlation time τc longer than 0.3 ns were found close to the RNA adenine H2 and H1′ protons in the hybrid segment. A possible long-lived water molecule was also detected close to the methyl group of 7T in the RNA–DNA junction but not with the other two thymines (8T and 9T). This result correlates with the structural studies that only DNA residue 7T in the RNA–DNA junction adopts an O4′-endo sugar conformation, while the other DNA residues including 3C in the DNA–RNA junction, adopt C1′-exo or C2′-endo conformations. The exchange rates for RNA C2′-OH were found to be ~520 s–1. This slow exchange rate may be due to the narrow minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2, which may trap the water molecules and restrict the dynamic motion of hydroxyl protons. The minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2 is wider than its B-DNA analog but narrower than that of the A-RNA analog. It was further confirmed by its titration with the minor groove binding drug distamycin. A possible 2:1 binding mode was found by the titration experiments, suggesting that this chimeric duplex contains a wider minor groove than its B-DNA analog but still narrow enough to hold two distamycin molecules. These distinct structural features and hydration patterns of this chimeric duplex provide a molecular basis for further understanding the structure and recognition of DNA·RNA hybrid and chimeric duplexes.  相似文献   

3.
The stability of the 1:1 complex of sodium ion with the dianion of guanosine 5′-monophosphate has been determined by means of a potentiometric titration employing a specific ion electrode. The stability constant for the reaction Na+ + 5′-GMP2- Na(5′-GMP)- was found to be 2.85 ± 0.36 M-1 at 5°C and an ionic strength of 1.1 ± 0.1 M. Although 5′-GMP forms ordered self-structures at high concentration in the presence of sodium ions, in dilute solution and at low sodium ion concentrations the Na+ binding is weak and typical of that for other nucleotides.  相似文献   

4.
The crystal and molecular structure of 2′-O-Me(CGCGCG)2 has been determined at 1.19 Å resolution, at 100 K, using synchrotron radiation. The structure in space group P3212 is a half-turn right-handed helix that includes two 2-methyl-2,4-pentanediol (MPD) molecules bound in the minor groove. The structure deviates from A-form RNA. The duplex is overwound with an average value of 9.7 bp per turn, characterised as having a C3′-endo sugar pucker, very low base pair rise and high helical twist and inclination angles. The structure includes 65 ordered water molecules. Only a single row of water molecules is observed in the minor groove due to the presence of hydrophobic 2′-O-methyl groups. As many as five magnesium ions are located in the structure. Two are in the major groove and interact with O6 and N7 of guanosine and N4 of cytidine residues through their hydration spheres. This work provides the first example of molecular interactions of nucleic acids with MPD, which was used as a precipitant, cryo-solvent and resolution enhancing agent. The two MPD molecules intrude into the hydration network in the minor groove, each forming hydrogen bonds between their secondary hydroxyl group and exo-amino functions of guanosine residues. Comparison of the 2′-O-Me(CGCGCG)2 structure in the P3212 and P6122 crystals delineates stability of the water network within the minor groove to dehydration by MPD and is of interest for evaluating factors governing small molecule binding to RNA. Intrusion of MPD into the minor groove of 2′-O-Me(CGCGCG)2 is discussed with respect to RNA dehydration, a prerequisite of Z-RNA formation.  相似文献   

5.
An extensive analysis of structural databases is carried out to investigate the relative flexibility of B-DNA and A-RNA duplexes in crystal form. Our results show that the general anisotropic concept of flexibility is not very useful to compare the deformability of B-DNA and A-RNA duplexes, since the flexibility patterns of B-DNA and A-RNA are quite different. In other words, ‘flexibility’ is a dangerous word for describing macromolecules, unless it is clearly defined. A few soft essential movements explain most of the natural flexibility of A-RNA, whereas many are necessary for B-DNA. Essential movements occurring in naked B-DNAs are identical to those necessary to deform DNA in DNA–protein complexes, which suggest that evolution has designed DNA–protein complexes so that B-DNA is deformed according to its natural tendency. DNA is generally more flexible, but for some distortions A-RNA is easier to deform. Local stiffness constants obtained for naked B-DNAs and DNA complexes are very close, demonstrating that global distortions in DNA necessary for binding to proteins are the result of the addition of small concerted deformations at the base-pair level. Finally, it is worth noting that in general the picture of the relative deformability of A-RNA and DNA derived from database analysis agrees very well with that derived from state of the art molecular dynamics (MD) simulations.  相似文献   

6.
We analyzed cleavage of precursor tRNAs with an LNA, 2′-OCH3, 2′-H or 2′-F modification at the canonical (c0) site by bacterial RNase P. We infer that the major function of the 2′-substituent at nt −1 during substrate ground state binding is to accept an H-bond. Cleavage of the LNA substrate at the c0 site by Escherichia coli RNase P RNA demonstrated that the transition state for cleavage can in principle be achieved with a locked C3′ -endo ribose and without the H-bond donor function of the 2′-substituent. LNA and 2′-OCH3 suppressed processing at the major aberrant m1 site; instead, the m+1 (nt +1/+2) site was utilized. For the LNA variant, parallel pathways leading to cleavage at the c0 and m+1 sites had different pH profiles, with a higher Mg2+ requirement for c0 versus m+1 cleavage. The strong catalytic defect for LNA and 2′-OCH3 supports a model where the extra methylene (LNA) or methyl group (2′-OCH3) causes a steric interference with a nearby bound catalytic Mg2+ during its recoordination on the way to the transition state for cleavage. The presence of the protein cofactor suppressed the ground state binding defects, but not the catalytic defects.  相似文献   

7.
1. The dephosphorylation of 3′-AMP, 3′-dAMP, 3′-CMP and 3′-dCMP was studied in the postmicrosomal supernatant of rat spleen and liver. In both organs 3′-AMP and 3′-dAMP were dephosphorylated at an appreciable rate, in both the presence and the absence of Mg2+. The pH optimum for this dephosphorylation was in the range 4.5–5.0. 3′-CMP and 3′-dCMP were very slowly degraded, though the activity towards 3′-dCMP increased somewhat in the presence of Mg2+. The optimum pH for this Mg2+-dependent dephosphorylation was 5.5–6.0. 2. The rate of dephosphorylation of 3′-AMP and 3′-dAMP per mg of protein was about 5 times as high in spleen as in liver. 3. The dephosphorylation of 3′-AMP could be ascribed to a single enzyme with pH optimum about 4.5. The activity towards 3′-dAMP could be resolved into one component coinciding with the 3′-dAMP-degrading enzyme, and one Mg2+-requiring component probably identical with the soluble deoxyinosine-activated nucleotidase. The dephosphorylation of 3′-dCMP seemed to be performed only by the latter enzyme. 4. The enzyme dephosphorylating 3′-AMP was purified 200-fold from the postmicrosomal supernatant and its physical and catalytic properties were compared with those of acid nucleotidase (EC 3.1.3.31) purified from rat liver lysosomes. The two enzymes were identical in all properties tested (substrate specificity, Km, molecular weight, response to phosphatase inhibitors), but some of the data differed from earlier reports on the acid nucleotidase. 5. The subcellular localization of the acid nucleotidase, its relationship to the acid phosphatase(s) and its role in the breakdown of nucleic acid constituents are discussed.  相似文献   

8.
We have investigated the anomalously weak binding of human papillomavirus (HPV) regulatory protein E2 to a DNA target containing the spacer sequence TATA. Experiments in magnesium (Mg2+) and calcium (Ca2+) ion buffers revealed a marked reduction in cutting by DNase I at the CpG sequence in the protein-binding site 3′ to the TATA spacer sequence, Studies of the cation dependence of DNA-E2 affinities showed that upon E2 binding the TATA sequence releases approximately twice as many Mg2+ ions as the average of the other spacer sequences. Binding experiments for TATA spacer relative to ATAT showed that in potassium ion (K+) the E2 affinity of the two sequences is nearly equal, but the relative dissociation constant (Kd) for TATA increases in the order K+ < Na+ < Ca2+ < Mg2+. Except for Mg2+, Kd for TATA relative to ATAT is independent of ion concentration, whereas for Mg2+ the affinity for TATA drops sharply as ion concentration increases. Thus, ions of increasing positive charge density increasingly distort the E2 binding site, weakening the affinity for protein. In the case of Mg2+, additional ions are bound to TATA that require displacement for protein binding. We suggest that the TATA sequence may bias the DNA structure towards a conformation that binds the protein relatively weakly.  相似文献   

9.
Experimental characterization of the structural couplings in free B-DNA in solution has been elusive, because of subtle effects that are challenging to tackle. Here, the exploitation of the NMR measurements collected on four dodecamers containing a substantial set of dinucleotide sequences provides new, consistent correlations revealing the DNA intrinsic mechanics. The difference between two successive residual dipolar couplings (ΔRDCs) involving C6/8-H6/8, C3′-H3′ and C4′-H4′ vectors are correlated to the 31P chemical shifts (δP), which reflect the populations of the BI and BII backbone states. The δPs are also correlated to the internucleotide distances (Dinter) involving H6/8, H2′ and H2″ protons. Calculations of NMR quantities on high resolution X-ray structures and controlled models of DNA enable to interpret these couplings: the studied ΔRDCs depend mostly on roll, while Dinter are mainly sensitive to twist or slide. Overall, these relations demonstrate how δP measurements inform on key inter base parameters, in addition to probe the BI↔BII backbone equilibrium, and shed new light into coordinated motions of phosphate groups and bases in free B-DNA in solution. Inspection of the 5′ and 3′ ends of the dodecamers also supplies new information on the fraying events, otherwise neglected.  相似文献   

10.
Clostridium thermocellum polynucleotide kinase (CthPnk), the 5′ end-healing module of a bacterial RNA repair system, catalyzes reversible phosphoryl transfer from an NTP donor to a 5′-OH polynucleotide acceptor. Here we report the crystal structures of CthPnk-D38N in a Michaelis complex with GTP•Mg2+ and a 5′-OH oligonucleotide and a product complex with GDP•Mg2+ and a 5′-PO4 oligonucleotide. The O5′ nucleophile is situated 3.0 Å from the GTP γ phosphorus in the Michaelis complex, where it is coordinated by Asn38 and is apical to the bridging β phosphate oxygen of the GDP leaving group. In the product complex, the transferred phosphate has undergone stereochemical inversion and Asn38 coordinates the 5′-bridging phosphate oxygen of the oligonucleotide. The D38N enzyme is poised for catalysis, but cannot execute because it lacks Asp38—hereby implicated as the essential general base catalyst that abstracts a proton from the 5′-OH during the kinase reaction. Asp38 serves as a general acid catalyst during the ‘reverse kinase’ reaction by donating a proton to the O5′ leaving group of the 5′-PO4 strand. The acceptor strand binding mode of CthPnk is distinct from that of bacteriophage T4 Pnk.  相似文献   

11.
RecJ exonuclease: substrates, products and interaction with SSB   总被引:4,自引:0,他引:4  
The RecJ exonuclease from Escherichia coli degrades single-stranded DNA (ssDNA) in the 5′–3′ direction and participates in homologous recombination and mismatch repair. The experiments described here address RecJ's substrate requirements and reaction products. RecJ complexes on a variety of 5′ single-strand tailed substrates were analyzed by electrophoretic mobility shift in the absence of Mg2+ ion required for substrate degradation. RecJ required single-stranded tails of 7 nt or greater for robust binding; addition of Mg2+ confirmed that substrates with 5′ tails of 6 nt or less were poor substrates for RecJ exonuclease. RecJ is a processive exonuclease, degrading ~1000 nt after a single binding event to single-strand DNA, and releases mononucleotide products. RecJ is capable of degrading a single-stranded tail up to a double-stranded junction, although products in such reactions were heterogeneous and RecJ showed a limited ability to penetrate the duplex region. RecJ exonuclease was equally potent on 5′ phosphorylated and unphosphorylated ends. Finally, DNA binding and nuclease activity of RecJ was specifically enhanced by the pre-addition of ssDNA-binding protein and we propose that this specific interaction may aid recruitment of RecJ.  相似文献   

12.
Magnesium ions play important roles in the structure and function of nucleic acids. Whereas the tertiary folding of RNA often requires magnesium ions binding to tight places where phosphates are clustered, the molecular basis of the interactions of magnesium ions with RNA helical regions is less well understood. We have refined the crystal structures of four decamer oligonucleotides, d(ACCGGCCGGT), r(GCG)d(TATACGC), r(GC)d(GTATACGC) and r(G)d(GCGTATACGC) with bound hexahydrated magnesium ions at high resolution. The structures reveal that A-form nucleic acid has characteristic [Mg(H2O)6]2+ binding modes. One mode has the ion binding in the deep major groove of a GpN step at the O6/N7 sites of guanine bases via hydrogen bonds. Our crystallographic observations are consistent with the recent NMR observations that in solution [Co(NH3)6]3+, a model ion of [Mg(H2O)6]2+, binds in an identical manner. The other mode involves the binding of the ion to phosphates, bridging across the outer mouth of the narrow major groove. These [Mg(H2O)6]2+ ions are found at the most negative electrostatic potential regions of A-form duplexes. We propose that these two binding modes are important in the global charge neutralization, and therefore stability, of A-form duplexes.  相似文献   

13.
Conformational properties of trimeric and tetrameric 2′,5′-linked oligonucleotides, 3′-MOE-A32′,5′ (1) and 3′-MOE-A42′,5′ (2), and their 3′,5′-linked analogs, 2′-MOE-A33′,5′ (3) and 2′-MOE-A43′,5′ (4), were examined with the use of heteronuclear NMR spectroscopy. The temperature-dependent 3JHH, 3JHP and 3JCP coupling constants, acquired in the range of 273–343 K, gave insight into the conformation of sugar rings in terms of a two-state North ↔ South (N ↔ S) pseudorotational equilibrium and into the conformation of the sugar–phosphate backbone in the model antisense oligonucleotides 1–4. 2′,5′-linked oligomers 3′-MOE-A32′,5′ (1) and 3′-MOE-A42′,5′ (2) show preference for N-type conformers and indication of A-type conformational features, which is prerequisite for antisense hybridization. The drive of N ↔ S equilibrium in 1–4 has been rationalized with the competing gauche effects of 2′/3′-phosphodiester and 3′/2′-MOE groups, anomeric and steric effects. Furthermore, the pairwise comparisons of 3′-MOE with 3′-OH and 3′-deoxy 2′,5′-linked adenine trimers emphasized the fine tuning of N ↔ S equilibrium in 3′-MOE-A32′,5′ (1) and 3′-MOE-A42′,5′ (2) by the steric effects of 3′-MOE group and the possibility of water-mediated H-bonds with vicinal phosphodiester functionality. In full correspondence, the drive of N ↔ S equilibrium towards N by 2′-MOE in 3′,5′-linked analogs 2′-MOE-A33′,5′ (3) and 2′-MOE-A43′,5′ (4) is weaker in comparison with 3′-OH group in the corresponding ribo analogs. βt, γ+ and ε rotamers are preferred in both 2′,5′- and in 3′,5′-linked oligonucleotides 1–4.  相似文献   

14.
Calcium and thiol reactivity of human plasma clotting factor XIII   总被引:4,自引:3,他引:1       下载免费PDF全文
1. The reaction of iodoacetate, 2-chloromercuri-4-nitrophenol and 5,5′-dithiobis-(2-nitrobenzoate) with thrombin-cleaved Factor XIII (i.e. Factor XIIIa) was accompanied by enzyme inhibition. 2. The reaction with iodoacetate and 5,5′-dithiobis-(2-nitrobenzoate) was absolutely dependent on Ca2+, and the rate of reaction increased with the Ca2+ concentration up to very high, non-physiological concentrations. 3. 2-Chloromercuri-4-nitrophenol reacted with Factor XIIIa in the absence of Ca2+, but at a much slower rate. 4. Stopped-flow methods were used to quantify the reaction with 5,5′-dithiobis-(2-nitro-benzoate) because of the Ca2+-dependent dissociation of Factor XIIIa (a2b2) and subsequent aggregation of the a′ chains into turbid precipitates. 5. The 3-carboxy-4-nitrothio-phenolate released was consistent with the reaction of 2 thiol groups/molecule of Factor XIIIa. The isolated b chains of Factor XIII did not react with either of the chromophoric reagents. This indicated that the a′ chains of Factor XIIIa were responsible for the thiol reactivity of the enzyme. 6. The Ca2+ dependence of the enzyme inhibition by these thiol reagents was very dependent on protein concentration. This is discussed in relation to the Ca2+-induced dissociation of Factor XIIIa. 7. The acceptor substrate, casein, decreased the Ca2+ concentration required for enzyme inhibition by both the mercurial and the aromatic disulphide compounds. Dansylcadaverine did not affect Ca2+ dependence of inhibition.  相似文献   

15.
Propofol, a widely used intravenous general anesthetic, acts at anesthetic concentrations as a positive allosteric modulator of γ-aminobutyric acid type A receptors and at higher concentration as an inhibitor of nicotinic acetylcholine receptors (nAChRs). Here, we characterize propofol binding sites in a muscle-type nAChR by use of a photoreactive analog of propofol, 2-isopropyl-5-[3-(trifluoromethyl)-3H-diazirin-3-yl]phenol (AziPm). Based upon radioligand binding assays, AziPm stabilized the Torpedo nAChR in the resting state, whereas propofol stabilized the desensitized state. nAChR-rich membranes were photolabeled with [3H]AziPm, and labeled amino acids were identified by Edman degradation. [3H]AziPm binds at three sites within the nAChR transmembrane domain: (i) an intrasubunit site in the δ subunit helix bundle, photolabeling in the nAChR desensitized state (+agonist) δM2-18′ and two residues in δM1 (δPhe-232 and δCys-236); (ii) in the ion channel, photolabeling in the nAChR resting, closed channel state (−agonist) amino acids in the M2 helices (αM2-6′, βM2-6′ and -13′, and δM2-13′) that line the channel lumen (with photolabeling reduced by >90% in the desensitized state); and (iii) at the γ-α interface, photolabeling αM2-10′. Propofol enhanced [3H]AziPm photolabeling at αM2-10′. Propofol inhibited [3H]AziPm photolabeling within the δ subunit helix bundle at lower concentrations (IC50 = 40 μm) than it inhibited ion channel photolabeling (IC50 = 125 μm). These results identify for the first time a single intrasubunit propofol binding site in the nAChR transmembrane domain and suggest that this is the functionally relevant inhibitory binding site.  相似文献   

16.
The 1H n.m.r. study of the DNA-dependent RNA polymerase from Escherichia coli has revealed that the holoenzyme (ββ′α2σ) displays two mobile regions: one, observable also in the core enzyme (ββ′α2), is characterized by basic amino acids and its appearance and form depend on ionic strength; the other, specific to the holoenzyme, is characterized by threonine residues and its appearance does not depend on ionic strength.  相似文献   

17.
The interactions of self-complementary oligonucleotides with a group of metal-mediated DNA-binding drugs, including chromomycin A3, mithramycin and the novel compound UK-1, were examined via electrospray ionization quadrupole ion trap mass spectrometry. Both chromomycin and mithramycin were shown to bind preferentially to GC-rich oligonucleotide duplexes in a 2:1 drug:metal ratio, while UK-1 was shown to bind in a 1:1 drug:metal stoichiometric ratio without a strong sequence preference. These trends were observed in the presence of Co2+, Ni2+ and Zn2+, with the exception that chromomycin–Zn2+ complexes were not readily observed. The binding stoichiometries as well as the sequence specificities are in agreement with literature reports for solution studies. Binding selectivities and stabilities of the complexes were also probed using electrospray ionization mass spectrometry. Both of the GC-rich oligomers 5′-GCGCGC-3′ and 5′-GCGCATGCGC-3′ exhibited a binding preference for chromomycin over mithramycin in the presence of Co2+ and Ni2+. Energy-variable collisionally activated dissociation of the complexes was employed to determine the stabilities of the complexes. The relative metal-dependent binding energies were Ni2+ > Zn2+ > Co2+ for UK-1–oligomer complexes and Ni2+ > Co2+ for both mithramycin and chromomycin complexes.  相似文献   

18.
Divalent metal ions promote hydrolysis of RNA backbones generating 5′OH and 2′;3′P as cleavage products. In these reactions, the neighboring 2′OH act as the nucleophile. RNA catalyzed reactions also require divalent metal ions and a number of different metal ions function in RNA mediated cleavage of RNA. In one case, the LZV leadzyme, it was shown that this catalytic RNA requires lead for catalysis. So far, none of the naturally isolated ribozymes have been demonstrated to use lead to activate the nucleophile. Here we provide evidence that RNase P RNA, a naturally trans-acting ribozyme, has leadzyme properties. But, in contrast to LZV RNA, RNase P RNA mediated cleavage promoted by Pb2+ results in 5′ phosphate and 3′OH as cleavage products. Based on our findings, we infer that Pb2+ activates H2O to act as the nucleophile and we identified residues both in the substrate and RNase P RNA that most likely influenced the positioning of Pb2+ at the cleavage site. Our data suggest that Pb2+ can promote cleavage of RNA by activating either an inner sphere H2O or a neighboring 2′OH to act as nucleophile.  相似文献   

19.
The ribosome is an ancient macromolecular machine responsible for the synthesis of all proteins in all living organisms. Here we demonstrate that the ribosomal peptidyl transferase center (PTC) is supported by a framework of magnesium microclusters (Mg2+-μc's). Common features of Mg2+-μc's include two paired Mg2+ ions that are chelated by a common bridging phosphate group in the form Mg(a)2+–(O1P-P-O2P)–Mg(b)2+. This bridging phosphate is part of a 10-membered chelation ring in the form Mg(a)2+–(OP-P-O5′-C5′-C4′-C3′-O3′-P-OP)–Mg(a)2+. The two phosphate groups of this 10-membered ring are contributed by adjacent residues along the RNA backbone. Both Mg2+ ions are octahedrally coordinated, but are substantially dehydrated by interactions with additional RNA phosphate groups. The Mg2+-μc's in the LSU (large subunit) appear to be highly conserved over evolution, since they are unchanged in bacteria (Thermus thermophilus, PDB entry 2J01) and archaea (Haloarcula marismortui, PDB entry 1JJ2). The 2D elements of the 23S rRNA that are linked by Mg2+-μc's are conserved between the rRNAs of bacteria, archaea and eukarya and in mitochondrial rRNA, and in a proposed minimal 23S-rRNA. We observe Mg2+-μc's in other rRNAs including the bacterial 16S rRNA, and the P4–P6 domain of the tetrahymena Group I intron ribozyme. It appears that Mg2+-μc's are a primeval motif, with pivotal roles in RNA folding, function and evolution.  相似文献   

20.
Chen FM  Sha F  Chin KH  Chou SH 《Nucleic acids research》2003,31(14):4238-4246
Actinomycin D (ACTD) has been shown to bind weakly to the sequence -GGCC-, despite the presence of a GpC site. It was subsequently found, however, that d(CATGGCCATG) binds relatively well to ACTD but exhibits unusually slow association kinetics, contrary to the strong-binding -XGCY- sites. In an effort to elucidate the nature of such binding and to delineate the origin of its interesting kinetic behavior, studies have now been extended to include oligomers with the general sequence motifs of d(CXYGGCCY′X′G)2. It was found that analogous binding characteristics are observed for these self-duplex decamers and comparative studies with progressively base-truncated oligomers from the 5′-end led to the finding that d(GGCCY′X′G) oligomers bind ACTD considerably stronger than their parent decamers and exhibit 1:1 drug/strand binding stoichiometry. Melting profiles monitored at the drug spectral region indicated additional drug binding prior to the onset of eventual complex disruptions with near identical melting temperatures for all the oligomers studied. These results are consistent with the notion that the related oligomers share a common strong binding mode of a hairpin-type, with the 3′-terminus G folding back to base-pair with the C base of GGC. A binding scheme is proposed in which the oligomers d(CXYGGCCY′X′G) exist predominantly in the duplex form and bind ACTD initially at the central GGCC weak site but subsequently disrupt to accommodate the stronger hairpin binding and thus the slow association kinetics. Such a mechanism is supported by the observation of distinct biphasic fluorescence kinetic traces in the binding of 7-amino-ACTD to these duplexes.  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