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.
Four genomic DNAs of differing GC content (Micrococcus luteus, 72% GC; Escherichia coli, 50% GC; calf thymus, 42% GC; Clostridium perfringens, 27% GC) have been employed as targets of interaction by the cationic polyamines spermidine {[H3N(CH2)3NH2(CH2)4NH3]3+} and spermine {[(CH2)4(NH2(CH2)3NH3)2]4+}. In solutions containing 60 mM DNA phosphate (~20 mg DNA/ml) and either 1, 5 or 60 mM polyamine, only Raman bands associated with the phosphates exhibit large spectral changes, demonstrating that B-DNA phosphates are the primary targets of interaction. Phosphate perturbations, which are independent of base composition, are consistent with a model of non-specific cation binding in which delocalized polyamines diffuse along DNA while confined by the strong electrostatic potential gradient perpendicular to the helix axis. This finding provides experimental support for models in which polyamine-induced DNA condensation is driven by non-specific electrostatic binding. The Raman spectra also demonstrate that major groove sites (guanine N7 and thymine C5H3) are less affected than phosphates by polyamine–DNA interactions. Modest dependence of polyamine binding on genome base composition suggests that sequence context plays only a secondary role in recognition. Importantly, the results demonstrate that polyamine binding has a negligible effect on the native B-form secondary structure. The capability of spermidine or spermine to bind and condense genomic B-DNA without disrupting the native structure must be taken into account when considering DNA organization within bacterial nucleoids or cell nuclei.  相似文献   

2.
Sulforaphane (SFN) is an isothiocyanate found in cruciferous vegetables with anti-inflammatory, anti-oxidant and anti-cancer activities. However, the antioxidant and anticancer mechanism of sulforaphane is not well understood. In the present research, we reported binding modes, binding constants and stability of SFN–DNA and -RNA complexes by Fourier transform infrared (FTIR) and UV–Visible spectroscopic methods. Spectroscopic evidence showed DNA intercalation with some degree of groove binding. SFN binds minor and major grooves of DNA and backbone phosphate (PO2), while RNA binding is through G, U, A bases with some degree of SFN–phosphate (PO2) interaction. Overall binding constants were estimated to be K(SFN–DNA)=3.01 (± 0.035)×104 M-1 and K(SFN–RNA)= 6.63 (±0.042)×103 M-1. At high SFN concentration (SFN/RNA = 1/1), DNA conformation changed from B to A occurred, while RNA remained in A-family structure.  相似文献   

3.
The interaction of hexamminecobalt(III), Co(NH3)63+, with 160 and 3000–8000 bp length calf thymus DNA has been investigated by circular dichroism, acoustic and densimetric techniques. The acoustic titration curves of 160 bp DNA revealed three stages of interaction: (i) Co(NH3)63+ binding up to the molar ratio [Co(NH3)63+]/[P] = 0.25, prior to DNA condensation; (ii) a condensation process between [Co(NH3)63+]/[P] = 0.25 and 0.30; and (iii) precipitation after [Co(NH3)63+]/[P] = 0.3. In the case of 3000–8000 bp DNA only two processes were observed: (i) binding up to [Co(NH3)63+]/[P] = 0.3; and (ii) precipitation after this point. In agreement with earlier observations, long DNA aggregates without changes in its B-form circular dichroism spectrum, while short DNA demonstrates a positive B→Ψ transition after [Co(NH3)63+]/[P] = 0.25. From ultrasonic and densimetric measurements the effects of Co(NH3)63+ binding on volume and compressibility have been obtained. The binding of Co(NH3)63+ to both short and long DNA is characterized by similar changes in volume and compressibility calculated per mole Co(NH3)63+: ΔV = 9 cm3 mol–1 and Δκ = 33 × 10–4 cm3 mol–1 bar–1. The positive sign of the parameters indicates dehydration, i.e. water release from Co(NH3)63+ and the atomic groups of DNA. This extent of water displacement would be consistent with the formation of two direct, hydrogen bonded contacts between the cation and the phosphates of DNA.  相似文献   

4.
The 1.7 Å X-ray crystal structure of the B-DNA dodecamer, [d(CGCGAATTCGCG)]2 (DDD)-bound non-covalently to a platinum(II) complex, [{Pt(NH3)3}2-µ-{trans-Pt(NH3)2(NH2(CH2)6NH2)2}](NO3)6 (1, TriplatinNC-A,) shows the trinuclear cation extended along the phosphate backbone and bridging the minor groove. The square planar tetra-am(m)ine Pt(II) units form bidentate N-O-N complexes with OP atoms, in a Phosphate Clamp motif. The geometry is conserved and the interaction prefers O2P over O1P atoms (frequency of interaction is O2P > O1P, base and sugar oxygens > N). The binding mode is very similar to that reported for the DDD and [{trans-Pt(NH3)2(NH2(CH2)6(NH3+)}2-µ-{trans-Pt(NH3)2(NH2(CH2)6NH2)2}](NO3)8 (3, TriplatinNC), which exhibits in vivo anti-tumour activity. In the present case, only three sets of Phosphate Clamps were found because one of the three Pt(II) coordination spheres was not clearly observed and was characterized as a bare Pt2+ ion. Based on the electron density, the relative occupancy of DDD and the sum of three Pt(II) atoms in the DDD-1 complex was 1:1.69, whereas the ratio for DDD-2 was 1:2.85, almost the mixing ratio in the crystallization drop. The high repetition and geometric regularity of the motif suggests that it can be developed as a modular nucleic acid binding device with general utility.  相似文献   

5.
The quadruplex structures of the human telomere sequences AG3(T2AG3)3 I and (T2AG3)4 II were investigated in the presence of Na+ and K+ ions, through the cross-linking of adenines and guanines by the cis- and trans-[Pt(NH3)2(H2O)2](NO3)2 complexes 1 and 2. The bases involved in chelation of the cis- and trans-Pt(NH3)2 moieties were identified by chemical and 3′-exonuclease digestions of the products isolated after denaturing gel electrophoresis. These are the four adenines of each sequence and four out of the 12 guanines. Two largely different structures have been reported for I: A from NMR data in Na+ solution and B from X-ray data of a K+-containing crystal. Structure A alone agrees with our conclusions about the formation of the A1–G10, A13–G22, A1–A13 platinum chelates at the top of the quadruplex and A7–A19, G4–A19 and A7–G20 at the bottom, whether the Na+ or K+ ion is present. At variance with a recent proposal that structures A and B could be the major species in Na+ and K+ solutions, respectively, our results suggest that structure A exists predominantly in the presence of both ions. They also suggest that covalent platinum cross-linking of a human telomere sequence could be used to inhibit telomerase.  相似文献   

6.
The trinuclear BBR3464 ([{trans-PtCl(NH3)2}2µ-(trans-Pt(NH3)2(H2N(CH2)6NH2)2)]4+) belongs to the polynuclear class of platinum-based anticancer agents. DNA adducts of this complex differ significantly in structure and type from those of clinically used mononuclear platinum complexes, especially, long-range (Pt, Pt) intrastrand and interstrand cross-links are formed in both 5′–5′ and 3′–3′ orientations. We show employing short oligonucleotide duplexes containing single, site-specific cross-links of BBR3464 and gel electrophoresis that in contrast to major DNA adducts of clinically used platinum complexes, under physiological conditions the coordination bonds between platinum and N7 of G residues involved in the cross-links of BBR3464 can be cleaved. This cleavage may lead to the linkage isomerization reactions between this metallodrug and double-helical DNA. Differential scanning calorimetry of duplexes containing single, site-specific cross-links of BBR3464 reveals that one of the driving forces that leads to the lability of DNA cross-links of this metallodrug is a difference between the thermodynamic destabilization induced by the cross-link and by the adduct into which it could isomerize. The rearrangements may proceed in the way that cross-links originally formed in one strand of DNA can spontaneously translocate from one DNA strand to its complementary counterpart, which may evoke walking of the platinum complex on DNA molecule.  相似文献   

7.
In eukaryotic Na+/Ca2+ exchangers (NCX) the Ca2+ binding CBD1 and CBD2 domains form a two-domain regulatory tandem (CBD12). An allosteric Ca2+ sensor (Ca3–Ca4 sites) is located on CBD1, whereas CBD2 contains a splice-variant segment. Recently, a Ca2+-driven interdomain switch has been described, albeit how it couples Ca2+ binding with signal propagation remains unclear. To resolve the dynamic features of Ca2+-induced conformational transitions we analyze here distinct splice variants and mutants of isolated CBD12 at varying temperatures by using small angle x-ray scattering (SAXS) and equilibrium 45Ca2+ binding assays. The ensemble optimization method SAXS analysis demonstrates that the apo and Mg2+-bound forms of CBD12 are highly flexible, whereas Ca2+ binding to the Ca3–Ca4 sites results in a population shift of conformational landscape to more rigidified states. Population shift occurs even under conditions in which no effect of Ca2+ is observed on the globally derived Dmax (maximal interatomic distance), although under comparable conditions a normal [Ca2+]-dependent allosteric regulation occurs. Low affinity sites (Ca1–Ca2) of CBD1 do not contribute to Ca2+-induced population shift, but the occupancy of these sites by 1 mm Mg2+ shifts the Ca2+ affinity (Kd) at the neighboring Ca3–Ca4 sites from ∼ 50 nm to ∼ 200 nm and thus, keeps the primary Ca2+ sensor (Ca3–Ca4 sites) within a physiological range. Thus, Ca2+ binding to the Ca3–Ca4 sites results in a population shift, where more constraint conformational states become highly populated at dynamic equilibrium in the absence of global conformational transitions in CBD alignment.  相似文献   

8.
The active component of phosphate solutions, in relation to promoter action on oxidising enzymes, is the PO4 '''''' ion. This is shown by the demonstration of a hyperbolic relationship between per cent production of CO2 (of Elodea) and pPO4, the measure of the phosphate ion potential. This is consistent with the rate of respiration as affected by changing pPO4 through change of total phosphate concentration while pH is kept constant. The equation for this relationship is (CO2a) (pPO4b)n = K where a, b, n, and K are constants and n = 1. The same relationship to phosphate ion concentration, expressed by the equation (Activity of enzyme) (pPO4)n = K, where n and K are constants and n varies from 1 to 6 under different conditions, appears to hold for some other enzyme actions, including those of peroxidase and pancreatic lipase.  相似文献   

9.
Nucleic acids exist in a dynamic equilibrium with a number of molecules that constantly interact with them and regulate the cellular activities. The inherent nature of the structure and conformational integrity of these macromolecules can lead to altered biological activity through proper targeting of nucleic acids binding ligands or drug molecules. We studied the interaction of naturally occurring methylxanthines such as theophylline, theobromine and caffeine with DNA, using UV absorption and Fourier transform infrared (FTIR) spectroscopic methods, and especially monitored their binding affinity in the presence of Mg2+ and during helix-coil transitions of DNA by temperature (Tm) or pH melting profiles. The study indicates that all these molecules effectively bind to DNA in a dose dependent manner. The overall binding constants of DNA-theophylline = 3.5×103 M−1, DNA-theobromine = 1.1×103 M−1, and DNA-Caffeine = 3.8×103 M−1. On the other hand Tm/pH melting profiles showed 24–35% of enhanced binding activity of methylxanthines during helix-coil transitions of DNA rather than to its native double helical structure. The FTIR analysis divulged that theophylline, theobromine and caffeine interact with all the base pairs of DNA (A-T; G-C) and phosphate group through hydrogen bond (H-bond) interaction. In the presence of Mg2+, methylxanthines altered the structure of DNA from B to A-family. However, the B-family structure of DNA remained unaltered in DNA-methylxanthines complexes or in the absence of Mg2+. The spectral analyses indicated the order of binding affinity as “caffeine≥theophylline>theobromine” to the native double helical DNA, and “theophylline≥theobromine>caffeine to the denatured form of DNA and in the presence of divalent metal ions.  相似文献   

10.
A emm53 subclass of Group A Streptococcus pyogenes (GAS) interacts tightly with human plasma plasminogen (hPg) and plasmin (hPm) via the kringle 2 (K2hPg) domain of hPg/hPm and the N-terminal a1a2 regions of a GAS coiled-coil M-like protein (PAM). Previous studies have shown that a monomeric PAM fragment, VEK30 (residues 97–125 + Tyr), interacted specifically with isolated K2hPg. However, the binding strength of VEK30 (KD = 56 nm) was ∼60-fold weaker than that of full-length dimeric PAM (KD = 1 nm). To assess whether this attenuated binding was due to the inability of VEK30 to dimerize, we defined the minimal length of PAM required to dimerize using a series of peptides with additional PAM residues placed at the NH2 and COOH termini of VEK30. VEK64 (PAM residues 83–145 + Tyr) was found to be the smallest peptide that adopted an α-helical dimer, and was bound to K2hPg with nearly the same affinity as PAM (KD = 1–2 nm). However, addition of two PAM residues (Arg126-His127) to the COOH terminus of VEK30 (VEK32) maintained a monomeric peptidic structure, but exhibited similar K2hPg binding affinity as full-length dimeric PAM. We identified five residues in a1a2 (Arg113, His114, Glu116, Arg126, His127), mutation of which reduced PAM binding affinity for K2hPg by ∼1000-fold. Replacement of these critical residues by Ala in the GAS genome resulted in reduced virulence, similar to the effects of inactivating the PAM gene entirely. We conclude that rather than dimerization of PAM, the five key residues in the binding domain of PAM are essential to mediate the high affinity interaction with hPg, leading to increased GAS virulence.  相似文献   

11.
Complexes of cationic liposomes with DNA are promising tools to deliver genetic information into cells for gene therapy and vaccines. Electrostatic interaction is thought to be the major force in lipid–DNA interaction, while lipid-base binding and the stability of cationic lipid–DNA complexes have been the subject of more debate in recent years. The aim of this study was to examine the complexation of calf-thymus DNA with cholesterol (Chol), 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP), dioctadecyldimethylammoniumbromide (DDAB) and dioleoylphosphatidylethanolamine (DOPE), at physiological condition, using constant DNA concentration and various lipid contents. Fourier transform infrared (FTIR), UV-visible, circular dichroism spectroscopic methods and atomic force microscopy were used to analyse lipid-binding site, the binding constant and the effects of lipid interaction on DNA stability and conformation. Structural analysis showed a strong lipid–DNA interaction via major and minor grooves and the backbone phosphate group with overall binding constants of KChol = 1.4 (±0.5) × 104 M−1, KDDAB = 2.4 (±0.80) × 104 M−1, KDOTAP = 3.1 (±0.90) × 104 M−1 and KDOPE = 1.45 (± 0.60) × 104 M−1. The order of stability of lipid–DNA complexation is DOTAP>DDAB>DOPE>Chol. Hydrophobic interactions between lipid aliphatic tails and DNA were observed. Chol and DOPE induced a partial B to A-DNA conformational transition, while a partial B to C-DNA alteration occurred for DDAB and DOTAP at high lipid concentrations. DNA aggregation was observed at high lipid content.  相似文献   

12.
Despite considerable interest and investigations on cationic lipid–DNA complexes, reports on lipid–RNA interaction are very limited. In contrast to lipid–DNA complexes where lipid binding induces partial B to A and B to C conformational changes, lipid–tRNA complexation preserves tRNA folded state. This study is the first attempt to investigate the binding of cationic lipid with transfer RNA and the effect of lipid complexation on tRNA aggregation and condensation. We examine the interaction of tRNA with cholesterol (Chol), 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP), dioctadecyldimethylammoniumbromide (DDAB) and dioleoylphosphatidylethanolamine (DOPE), at physiological condition, using constant tRNA concentration and various lipid contents. FTIR, UV-visible, CD spectroscopic methods and atomic force microscopy (AFM) were used to analyze lipid binding site, the binding constant and the effects of lipid interaction on tRNA stability, conformation and condensation. Structural analysis showed lipid–tRNA interactions with G–C and A–U base pairs as well as the backbone phosphate group with overall binding constants of KChol = 5.94 (± 0.8) × 104 M–1, KDDAB = 8.33 (± 0.90) × 105 M–1, KDOTAP = 1.05 (± 0.30) × 105 M–1 and KDOPE = 2.75 (± 0.50) × 104 M–1. The order of stability of lipid–tRNA complexation is DDAB > DOTAP > Chol > DOPE. Hydrophobic interactions between lipid aliphatic tails and tRNA were observed. RNA remains in A-family structure, while biopolymer aggregation and condensation occurred at high lipid concentrations.  相似文献   

13.
We took a discovery approach to explore the actions of cAMP and two of its analogs, one a cAMP mimic ((Sp)-adenosine cyclic 3′:5′-monophosphorothioate ((Sp)-cAMPS)) and the other a diastereoisomeric antagonist ((Rp)-cAMPS), on a model system of the type Iα cyclic AMP-dependent protein kinase holoenzyme, RIα(91–244)·C-subunit, by using fluorescence spectroscopy and amide H/2H exchange mass spectrometry. Specifically, for the fluorescence experiments, fluorescein maleimide was conjugated to three cysteine single residue substitution mutants, R92C, T104C, and R239C, of RIα(91–244), and the effects of cAMP, (Sp)-cAMPS, and (Rp)-cAMPS on the kinetics of R-C binding and the time-resolved anisotropy of the reporter group at each conjugation site were measured. For the amide exchange experiments, ESI-TOF mass spectrometry with pepsin proteolytic fragmentation was used to assess the effects of (Rp)-cAMPS on amide exchange of the RIα(91–244)·C-subunit complex. We found that cAMP and its mimic perturbed at least parts of the C-subunit interaction Sites 2 and 3 but probably not Site 1 via reduced interactions of the linker region and αC of RIα(91–244). Surprisingly, (Rp)-cAMPS not only increased the affinity of RIα(91–244) toward the C-subunit by 5-fold but also produced long range effects that propagated through both the C- and R-subunits to produce limited unfolding and/or enhanced conformational flexibility. This combination of effects is consistent with (Rp)-cAMPS acting by enhancing the internal entropy of the R·C complex. Finally, the (Rp)-cAMPS-induced increase in affinity of RIα(91–244) toward the C-subunit indicates that (Rp)-cAMPS is better described as an inverse agonist because it decreases the fractional dissociation of the cyclic AMP-dependent protein kinase holoenzyme and in turn its basal activity.Cyclic AMP-dependent protein kinase (PKA)1 plays a crucial role in a plethora of cellular functions. All isoforms of PKA are composed of two catalytic (C) subunits and homodimeric regulatory (R) subunits (13). As the name implies, cAMP is a major PKA regulator (4). Much progress has been made in the last decade in delineating the molecular basis of action of cAMP. An important tactic in this endeavor has been through the comparison of the effects of cAMP with those of two phosphorothioate cAMP analogs: (Sp)-cAMPS (a cAMP mimic) and (Rp)-cAMPS (an antagonist and a diastereoisomer of (Sp)-cAMPS). Although the importance of geometry of the sulfur substitution is critical in determining the pharmacological properties of the two phosphorothioate cAMP analogs, the molecular basis for this behavior is not fully understood. To date, these comparisons have only been made using either wild-type or truncated mutants of the type Iα regulatory subunit (RIα) that are free in solution, not complexed to the C-subunit. X-ray spectroscopic examination of ligand-bound RIα(92–379) complexes reveals few differences between ligand-bound complexes, but the (Rp)-cAMPS complex is structurally “looser” with higher thermal factors than complexes formed with either cAMP or (Sp)-cAMPS (5). This is consistent with the observation that both cAMP and (Sp)-cAMPS, but not (Rp)-cAMPS, raise the urea concentration required for wild-type RIα unfolding (6). Further insight into the structural basis of cAMP action stems from NMR spectroscopic comparison of the effects of (Rp)-cAMPS, cAMP, and (Sp)-cAMPS on chemical shifts and 15N relaxation of the RIα(119–244) mutant (7). In addition to producing fewer significant chemical shift changes than either cAMP or (Sp)-cAMPS, (Rp)-cAMPS binding is associated with enhanced millisecond to microsecond time scale backbone motions of a β-turn (β2,3 loop) and around the phosphate-binding cassette (PBC) (7).Further insight into the molecular basis of actions of cAMP and its analogs should come from the analysis of ligand-bound R·C complexes. Unfortunately, the large size of even the heterodimeric R·C complex (∼95 kDa) and the difficulty of preparing (Rp)-cAMPS·R·C-subunit crystals currently preclude the use of both NMR spectroscopy and x-ray crystallography. Consequently, we took two alternative lower resolution approaches to this issue. One approach involves the use of site-directed labeling combined with fluorescence spectroscopy to examine both the effects of cAMP and its analogs on R-C subunit binding kinetics and on the conformational dynamics of RIα(91–244). RIα(91–244) includes the “A” cyclic nucleotide binding (CNB) domain, the pseudosubstrate, and linker domains and represents the minimal segments necessary for high affinity C-subunit binding (Fig. 1) (8). The other approach involves an examination of the effects of cAMP and its analogs on solvent exposure/conformational flexibility of RIα(91–244)·C-subunit complex using H/2H amide exchange measured with a combination of mass spectrometry (ESI-Q-TOF) and proteolytic fragmentation. In the first approach, fluorescein maleimide (FM) was conjugated to three cysteine substitution mutants with the substitution sites located near or within the pseudosubstrate sequence, the linker domain, or αC (R92C, T104C, and R239C, respectively) of RIα(91–244) (Fig. 1). The time-resolved fluorescence anisotropy results suggest that cAMP and (Sp)-cAMPS reduce the interaction of the RIα linker domain and αC with the two peripheral R-C interaction sites on the C-subunit (so-called Sites 2 and 3) without affecting the interaction of the pseudosubstrate sequence with the active site cleft (so-called Site 1). Because of limitations of the amide H/2H exchange experiments, only the effects of (Rp)-cAMPS on H/2H amide exchange in RIα(91–244)·C-subunit complex could be investigated. The results showed that (Rp)-cAMPS induces a relatively widespread increase in amide exchange, indicating limited unfolding and/or enhanced conformational flexibility that is propagated almost globally through the C-subunit and, at least, part of RIα. These conformational changes were accompanied by a 5-fold increase in the affinity of RIα(91–244) toward C-subunit, suggesting that, at least, some of the (Rp)-cAMPS effects are mediated by an increase in internal entropy. Finally, the (Rp)-cAMPS-induced increase in R-C affinity indicates that (Rp)-cAMPS is better described as an inverse agonist because the basal activity of the PKA holoenzyme should be decreased by (Rp)-cAMPS.Open in a separate windowFig. 1.Overview of PKA structure and cAMP analogs. A, domain organization of RIα showing the domain boundaries of RIα(91–244) where the pseudosubstrate in green is connected to CNB-A domain in blue by a linker segment. B, structure of RIα(91–244) in the C-subunit-bound conformation (Protein Data Bank code 1U7E (23)) showing the pseudosubstrate in green, linker in yellow, and helical subdomain comprising helices αN, αA, αB, and αC in blue and β-subdomain in tan. The PBC is in red. C, structure of the C·RIα(91–244) holoenzyme showing the C-subunit in tan and RIα(91–244) in blue. Sites for introduction of cysteines by site-directed mutagenesis are represented by red circles. The cAMP binding site (PBC) is in red. D, structure of cAMP showing the 2′-OH group and 3′–5′ phosphodiester bond. The exocyclic oxygens upon replacement with sulfur atoms to generate the (Sp)-cAMPS and (Rp)-cAMPS diastereomers are highlighted.  相似文献   

14.
The spectral (UV-Vis and IR) and electrochemical behavior of the nitrile bonded complexes [Ru(NH3)5L]2+ (L = 1,4-dicyanobenzene (1,4-dcb), 1,2-dicyanobenzene (1,2-dcb)), [Ru(NH3)5(NHC(OH)-bz-4-CN)]3+, [Ru(NH3)5(NHC(O)-bz-2-CN)]2+ and [Ru(NH3)5(NH(C)NHC(O)bz)]3+ (NH(C)NHC(O)-bz = 3-imino-1-oxo-isoindoline) are described. Oxidation of [Ru(NH3)5L]2+, at 0 ? pH ? 6, is followed by hydrolysis of the coordinated nitrile to give amide complexes in which the amide is through the nitrogen, with pH-dependent rate constants. The estimated values of the rate constant of hydrolysis (kobs) at 25 °C are 2.9 × 10−3 s−1 for [Ru(NH3)5(1,4-dcb)]3+ and 5.6 × 10−3 s−1 for [Ru(NH3)5(1,2-dcb)]3+ at pH 4.65. Reduction of [Ru(NH3)5(NHC(O)-bz-4-CN)]2+ and [Ru(NH3)5(NHC(O)-bz-2-CN)]2+ is followed by two reactions, one is an aquation forming [Ru(NH3)5(OH2)]2+ and free ligand, and the other an intramolecular linkage isomerization forming [Ru(NH3)5(NC-bz-4-NH2C(O))]2+ and [Ru(NH3)5(NC-bz-2-NH2C(O))]2+. The oxidized1,2-cyanobenzamide complex [Ru(NH3)5(NHC(OH)-bz-2-CN)]3+ undergoes an amide to nitrile intramolecular linkage isomerization, followed by a cyclization reaction resulting in [Ru(NH3)5(NH-(C)(HN-C(O)-2-bz))]3+ ((NH-(C)(HN-C(O)-2-bz)) = 3-imino-1-oxo-isoindoline bonded through the exocyclic nitrogen) (pKa = 4.3). The rates of these reactions, which occur with neighboring group participation, increase with acidity. The reduced form, [Ru(NH3)5(NH-(C)(HN-C(O)-2-bz))]2+, is relatively substitution inert.  相似文献   

15.
We present here molecular dynamics simulations and DNA conformational dynamics for a series of trinuclear platinum [Pt3(HPTAB)]6+-DNA adducts [HPTAB = N,N,N′,N′,N′′,N′′-hexakis (2-pyridyl-methyl)-1,3,5-tris(aminomethyl) benzene], including three types of bifunctional crosslinks and four types of trifunctional crosslinks. Our simulation results reveal that binding of the trinuclear platinum compound to a DNA duplex induces the duplex unwinding in the vicinity of the platination sites, and causes the DNA to bend toward the major groove. As a consequence, this produces a DNA molecule whose minor groove is more widened and shallow compared to that of an undamaged bare-DNA molecule. Notably, for trifunctional crosslinks, we have observed extensive DNA conformational distortions, which is rarely seen for normal platinum–DNA adducts. Our findings, in this study, thus provide further support for the idea that platinum compounds with trifunctional intra-strand or long-range-inter-strand cross-linking modes can generate larger DNA conformational distortions than other types of cross-linking modes.  相似文献   

16.
The solution structures of two 27 nt RNA hairpins and their complexes with cobalt(III)-hexammine [Co(NH3)63+] were determined by NMR spectroscopy. The RNA hairpins are variants of the P4 region from Escherichia coli RNase P RNA: a U-to-A mutant changing the identity of the bulged nucleotide, and a U-to-C, C-to-U double mutant changing only the bulge position. Structures calculated from NMR constraints show that the RNA hairpins adopt different conformations. In the U-to-C, C-to-U double mutant, the conserved bulged uridine in the P4 wild-type stem is found to be shifted in the 3′-direction by one nucleotide when compared with the wild-type structure. Co(NH3)63+ is used as a spectroscopic probe for Mg(H2O)62+ binding sites because both complexes have octahedral symmetry and have similar radii. Intermolecular NOE crosspeaks between Co(NH3)63+ and RNA protons were used to locate the site of Co(NH3)63+ binding to both RNA hairpins. The metal ion binds in the major groove near a bulge loop in both mutants, but is shifted 3′ by about one base pair in the double mutant. The change of the metal ion binding site is compared with results obtained on corresponding mutant RNase P RNA molecules as reported by Harris and co-workers (RNA, 1, 210–218).  相似文献   

17.
In molecular biology, understanding the functional and structural aspects of DNA requires sequence-specific DNA binding probes. Especially, sequence-specific fluorescence probes offer the advantage of real-time monitoring of the conformational and structural reorganization of DNA in living cells. Herein, we designed a new class of D2A (one-donor-two-acceptor) near-infrared (NIR) fluorescence switch-on probe named quinone cyanine–dithiazole (QCy–DT) based on the distinctive internal charge transfer (ICT) process for minor groove recognition of AT-rich DNA. Interestingly, QCy–DT exhibited strong NIR-fluorescence enhancement in the presence of AT-rich DNA compared to GC-rich and single-stranded DNAs. We show sequence-specific minor groove recognition of QCy–DT for DNA containing 5′-AATT-3′ sequence over other variable (A/T)4 sequences and local nucleobase variation study around the 5′-X(AATT)Y-3′ recognition sequence revealed that X = A and Y = T are the most preferable nucleobases. The live cell imaging studies confirmed mammalian cell permeability, low-toxicity and selective staining capacity of nuclear DNA without requiring RNase treatment. Further, Plasmodium falciparum with an AT-rich genome showed specific uptake with a reasonably low IC50 value (<4 µM). The ease of synthesis, large Stokes shift, sequence-specific DNA minor groove recognition with switch-on NIR-fluorescence, photostability and parasite staining with low IC50 make QCy–DT a potential and commercially viable DNA probe.  相似文献   

18.
We present a simple and novel assay—employing a universal molecular beacon (MB) in the presence of Hg2+—for the detection of single nucleotide polymorphisms (SNPs) based on Hg2+–DNA complexes inducing a conformational change in the MB. The MB (T7-MB) contains a 19-mer loop and a stem of a pair of seven thymidine (T) bases, a carboxyfluorescein (FAM) unit at the 5′-end, and a 4-([4-(dimethylamino)phenyl]azo)benzoic acid (DABCYL) unit at the 3′-end. Upon formation of Hg2+–T7-MB complexes through T–Hg2+–T bonding, the conformation of T7-MB changes from a random coil to a folded structure, leading to a decreased distance between the FAM and DABCYL units and, hence, increased efficiency of fluorescence resonance energy transfer (FRET) between the FAM and DABCYL units, resulting in decreased fluorescence intensity of the MB. In the presence of complementary DNA, double-stranded DNA complexes form (instead of the Hg2+–T7-MB complexes), with FRET between the FAM and DABCYL units occurring to a lesser extent than in the folded structure. Under the optimal conditions (20 nM T7-MB, 20 mM NaCl, 1.0 μM Hg2+, 5.0 mM phosphate buffer solution, pH 7.4), the linear plot of the fluorescence intensity against the concentration of perfectly matched DNA was linear over the range 2–30 nM (R2 = 0.991), with a limit of detection of 0.5 nM at a signal-to-noise ratio of 3. This new probe provides higher selectivity toward DNA than that exhibited by conventional MBs.  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.
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   

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