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 603 毫秒
1.
Combined multidimensional nuclear magnetic resonance spectroscopy and electrospray mass spectrometry was used to analyze the platinated DNA adduct of the phase II anticancer drug [{trans-PtCl(NH3)2}2-μ-{trans-Pt(NH3)2(NH2(CH2)6NH2)2}](NO3)4 (BBR3464) with [5′-d(ACG*TATACG*T)-3′]2. Two 1,2-interstrand cross-links were formed by concomitant binding of two trinuclear moieties to the oligonucleotide. The four DNA-bound platinum atoms coordinated in the major groove at N7 positions of guanines in the 3′ → 3′ direction and the central platinum unit is expected to lie in the DNA minor groove. This is the first report of such a DNA lesion. The melting temperature of the adduct is 76 °C and is 42 °C higher than that of the unplatinated DNA. The sugar residues of the platinated bases are in the N-type conformation and the G9 nucleoside is in the syn orientation, while the G3 nucleoside appears to retain the anti configuration. The secondary structure of DNA was significantly changed upon cross-linking of the two BBR3464 molecules. Base destacking occurs between A1/C2 and C2/G3 and weakened stacking is seen for the C8/G9 and G9/T10 bases. The lack of Watson–Crick base pairing is also seen for A1–T10 and C2–G9 base pairs, whereas Watson–Crick base pairs in the central sequence of the DNA (T4 → A7) are well maintained. While DNA repair proteins may “see” different platinated adducts as bulky “lesions”, the subtle differences involved in base pairing and stacking, as summarized here, may extend to their role as a substrate for repair enzymes. Thus, differences in protein recognition and repair efficiency among the various interstrand cross-links are likely and a subject worthy of detailed exploration. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

2.
The nonamer 5'd(CTCAGCCTC) 3' 1 has been reacted with cis-diamminediaquaplatinum(II) in water at pH 4.2. The major reaction product was shown by enzymatic digestion and 1H NMR to be the d(ApG)cis-Pt(NH3)2 chelate [cis-Pt(NH3)2[d(CTCAGCCTC)-N7(4),N7(5)]] 1-Pt. When mixed with its complementary strand 2, 1-Pt forms a B DNA type duplex 3-Pt with a Tm of 35 degrees C (versus 58 degrees C for the unplatinated duplex). The NMR study of the exchangeable protons of 3-Pt revealed that the helix distortion is localized on the CA*G*-CTG moiety (the asterisks indicating the platinum chelation sites) with a strong perturbation of the A*(4)T(15) base pair related to a large tilt of A*(4).  相似文献   

3.
 Reaction of [Pt(dien)Cl]+ (1) with the 14-mer oligonucleotide 5′-d(ATACATGGTACATA) (I) gave rise to two major species which corresponded to the 5′-G and 3′-G platinated monofunctional adducts, and a minor amount of the bis-platinated adduct formed during the later stages of the reaction. The reaction of (1) with the related octamer 5′-d(ATACATGG) (II) was also investigated. Kinetic data obtained by HPLC showed that the 5′-G and 3′-G bases of the 14-mer oligonucleotide were platinated at similar rates: the second-order rate constant is 53×10–2 M–1 s–1 at 298 K in 0.1 M NaClO4. However, the platination rate of 5′-G of the octamer (II) (k=69×10–2 M–1 s–1) was enhanced by a factor of three compared to the rate of platination at 3′-G (k=22×10–2 M–1 s–1). All the adducts were separated by HPLC and characterized by NMR spectroscopy, enzymatic digestion and MALDI-TOF mass spectrometry. 1H and 15N NMR shifts suggest that there are distinct conformational differences between 14-mer duplexes platinated at 5′-G (I5′ ds) and 3–G (I3′ ds). Molecular mechanics modelling indicates that rotation around the Pt-N7 bond is more restricted in the case of the 5′-G adduct than in that of the 3′-G adduct. The binding of {Pt(dien)}2+ to 5′-GN7 and 3′-GN7 in the monofunctional adducts of (I) was shown to be reversible upon the addition of high concentrations of chloride ions. Received: 3 July 1998 / Accepted: 10 November 1998  相似文献   

4.
We present a method for improving the quality of nuclear magnetic resonance (NMR) spectra involving exchangeable protons near the base of the stem of RNA hairpin molecules. NMR spectra of five different RNA hairpins were compared. These hairpins consisted of a native RNA structure and four molecules each having different unpaired, or dangling, nucleotides at the 3′ end. NMR experiments were acquired in water for each construct and the quality of the imino proton spectral regions were examined. The imino resonances near the base of the stem of the wild type RNA structure were not observed due to breathing motions. However, a significant increase in spectral quality for molecules with dangling 3′ adenosine or guanosine nucleotides was observed, with imino protons detected in these constructs that were not observed in the wild type construct. A modest improvement in spectral quality was seen for the construct with a 3′ unpaired uridine, whereas no significant improvement was observed for a 3′ unpaired cytidine. This improvement in NMR spectral quality mirrors the increased thermodynamic stability observed for 3′ unpaired nucleotides which is dependant on the stacking interactions of these nucleotides against the base of the stem. The use of a dangling 3′ adenosine nucleotide represents an easy method to significantly improve the quality of NMR spectra of RNA molecules. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

5.
 An approach is presented which probes the possible use of trans-[(NH3)2PtCl]+-modified deoxyoligonucleotides in the antisense strategy. It consists of (1) the selective platination of an oligonucleotide containing 11 pyrimidine (T, C) bases as well as a single guanine (G) as a Pt-anchoring group at the 5′-end to give trans-[(NH3)2Pt{5′-d(GN7T2C2T2C2T2C}Cl]10– 1 ("antisense strand") and (2) subsequent hybridization with the purine 12-mer 5′-d(GA2G2A2G2A2G)11– ("sense strand"). According to HPLC, three major species 24 are formed during reaction (2), all of which are cross-linking adducts between 1 and the sense strand, as confirmed by ESI MS and melting temperature measurements. Only for the major product 3 can a structure be proposed on the basis of 1D and 2D NMR spectra. According to these, G1 of the antisense strand is cross-linked with G20 via trans-(NH3)2PtII. The complementary overhangs of the duplex represent "sticky ends" and are, in principle, capable of associating into multimers of the duplex. Received: 29 March 1999 / Accepted: 26 July 1999  相似文献   

6.
A high melting cis-[Pt(NH3)2[d(GpG)]]adduct of a decanucleotide duplex   总被引:2,自引:0,他引:2  
The [cis-Pt(NH3)2(d(GCCGGATCGC)-N7(4), N7(5))]-d(GCGATCCGGC) duplex has been prepared with Tm = 49 degrees C (vs 58 degrees C for the unplatinated form). NMR of the ten observable imino protons supports a kinked structure with intact base pairing of the duplex on the 3'-side of the d(GpG).cis-Pt chelate (relative to the platinated strand) The modification of the B-DNA type CD spectrum, due to the platinum chelate, is comparable to that observed for the platination (at a 0.05 Pt:base ratio) of the Micrococcus Lysodeikticus DNA (72% GC).  相似文献   

7.
The DNA duplex d(CTCTCG*AGTCTC).d(GAGAC-TC*GAGAG) containing a single trans- diammine-dichloroplatinum(II) interstrand cross-link (where G* and C* represent the platinated bases) has been studied by two-dimensional NMR. All the exchangeable and non-exchangeable proton resonance lines were assigned (except H5'/H5") and the NOE intensities were transformed into distances via the RELAZ program. By combining the NOESY and COSY data (330 constraints) and NMR-constrained molecular mechanics using JUMNA, a solution structure of the cross-linked duplex has been determined. The duplex is distorted over two base pairs on each side of the interstrand cross-link and exhibits a slight bending of its axis ( approximately 20 degrees ) towards the minor groove. The platinated guanine G* adopts a syn conformation. The rotation results in a Hoogsteen-type pairing between the complementary G(6)* and C(19)* residues which is mediated by the platinum moiety and is stabilized by a hydrogen bond between O6(G(6)*) and N4H(C(19)*). The rise between the cross-linked residues and the adjacent residues is increased owing to the interaction between these adjacent residues and the ammine groups of the platinum moiety. These results are discussed in relation to the slow rate of closure of the monofunctional adducts into interstrand cross-links.  相似文献   

8.
The protomeric tautomerizm and conformation of the 2-methyl-4-pyridin-2′-yl-1,5-benzodiazepine molecule were investigated, and its three neutral tautomers (B1,B2,B3) and their rotamers (C1,C2,C3) were considered. Full geometry optimizations were carried out at the HF/6-31G* and B3LYP/6-31G* levels in gas phase and in water. The tautomerization processes in water (ɛ = 78.54) were studied by using self-consistent reaction field theory. The calculation showed that the boat conformation is dominant for the seven-membered diazepine ring in all of the structures, even with different double bond positions. The calculated relative free energies (ΔG) showed that the tautomer C1 was the most stable structure, and its conformer B1 was the second most stable in the gas phase and in water. Figure 2-Methyl-4-pyridin-2′-yl-1,5-benzodiazepine  相似文献   

9.
The guanine-N7 monoadduct of [Pt(en)Cl(ACRAMTU)](NO3)2 (PT-ACRAMTU; en=ethane-1,2-diamine, ACRAMTU=1-[2-(acridin-9-ylamino)ethyl]-1,3-dimethylthiourea), a dual metalating/intercalating cytotoxic agent, was generated in a double-stranded dodecamer, d(CCTCTCG*TCTCC/GGAGACGAGAGG) (III*), and isolated by preparative reverse-phase high-performance liquid chromatography (HPLC). The adduct was characterized using matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS), circular-dichroism spectropolarimetry (CD), UV-melting curves, and NMR spectroscopy. In addition, a molecular mechanics/restrained molecular dynamics (MM/rMD) study was performed for this adduct using the AMBER force field. Monoadduction of the sequence leads to a pronounced increase in melting temperature, Tm=Tm(III*)–Tm(III)=9.7 °C. Because there is complete enthalpy–entropy compensation, binding occurs without noticeable thermodynamic destabilization. This feature and the CD (induced-ligand circular dichroism) and NMR (upfield shifts of aromatic acridine proton signals) data are indicative of a unique, nondenaturing dual-binding mode that involves partial intercalation of the acridine chromophore. An energy-minimized AMBER model of III* demonstrates that platination of G7-N7 of guanine in the major groove and partial insertion of the acridine moiety into the C6G19/G7C18 base step on the 5 face of the modified purine base is feasible and supportive of the experimental results. Differences in the biophysical properties between III* and duplexes containing adducts of the clinical-drug cisplatin are outlined, and possible biological consequences are discussed.Electronic Supplementary Material Supplementary material is available in the online version of this article at http://dx.doi.org/10.1007/s00775-004-0534-3Abbreviations ACRAMTU 1-[2-(acridin-9-ylamino)ethyl]-1,3-dimethylthiourea - dGuo 2-deoxyguanosine - dGuo* [Pt(en)(ACRAMTU-S)(dGuo-N7)]3+ - en ethane-1,2-diamine - ICD Induced circular dichroism - MALDI-TOF MS Matrix-assisted laser desorption ionization time-of-flight mass spectrometry - MM Molecular mechanics - PIPES 1,4-piperazinediethanesulfonic acid - PT-ACRAMTU [Pt(en)Cl(ACRAMTU)](NO3)2 - rMD Restrained molecular dynamics  相似文献   

10.
 The interaction of the two chiral isomers of the new anticancer agent [Pt(ampyr)(cbdca)] (ampyr=aminomethylpyrrolidine, cbdca=cyclobutanedicarboxylate) with 5′-GMP and with short G-containing oligonucleotides has been studied using 1H and 31P NMR, UV-vis spectroscopy and molecular modelling. Each isomer loses the cbdca ligand upon binding to the DNA fragments. Two geometrical isomers of the DNA adducts are formed owing to the presence of the unsymmetric ampyr ligand. These isomers prove to be GG-N7,N7 chelates for d(GpG), d(pGpG) and d(CpGpG). A slight preference for the formation of one geometrical isomer is found in the case of DNA fragments having a phosphate moiety and/or a C base at the 5′-site of the GG sequence. H-bonding interactions from the NH2 moiety towards the 5′-phosphate group and/or the O atom of the C base clearly favour the formation of one geometrical isomer. The presence of these H-bonds, together with the bulky pyrrolidine ring, has resulted in the unique observation (by 1H NMR) of NH protons of coordinated amines that do not rapidly exchange in a 99.95% D2O solution. Temperature-dependence studies show an extremely slow stack ⇄ destack conformational change for the CGG adducts of the S isomer, which could be related to these stable H-bonds of the amine protons towards the oligonucleotide. For the R isomer this stack ⇄ destack conformational change is faster, probably owing to more steric hindrance of the pyrrolidine ring as deduced from the NOESY data, and as also suggested by molecular modelling. The observation of extremely slow rotation around the Pt-N7 bond for [Pt(R-ampyr)(GMP-N7)2] provides further evidence for increased steric hindrance of the R isomer compared to the S isomer. The rate of binding of the drug to G bases proved to be second order for both isomers; in fact the (toxic) S isomer is about two times more reactive than the (non-toxic) R isomer, as seen from k 2 values of 0.17±0.01 M–1 s–1 for [Pt(S-ampyr)(cbdca)] and 0.09±0.01 M–1 s–1 for [Pt(R-ampyr)(cbdca)]. No solvent-assisted pathway is involved in these reactions, since the complexes prove to be stable in solution for weeks and therefore only a direct attack of the G base on the Pt must be involved. Because hardly any intermediate species can be detected during the reaction, coordination of the second G base must occur much faster than the binding of the first G base. Since direct attack of the nucleobases takes place, steric interactions become extremely important and therefore are likely to determine the reactivity, activity and even the toxicity of such Pt complexes. Received: 12 January 1999 / Accepted: 17 June 1999  相似文献   

11.
The stability of the tri–μ–hydrido–bis[(η5–C5Me5)aluminum], Cp*2Al2H3, 1 is studied at B3LYP/6–311+G(d,p), CCSD(T)//B3LYP/6–311+G(d,p) and MP4//B3LYP/6–311+G(d,p) levels. The coordination between Al2H3 entity and both C5(CH3)5 groups is ensured by strong electrostatic and orbital interactions. The orbital analysis of the interacting fragments shows that Al2H3 acceptor, which keeps its tribridged structure, implies the vacant ( \texta1¢ ) \left( {{\text{a}}_1^\prime } \right) and five antibonding (a2¢¢ a_2^{\prime \prime } , e′ and e″) molecular orbitals to interact with two orbitals mixtures, b1 and e" of the donors (C5Me5). When we take into account the solvent effect, the computation shows that 1 seems to be stable in condensed phase with a tribridged bond between the Al atoms [Cp*Al(μ-H)3AlCp*], whereas in the gas phase, the monobridged Cp*AlH(μ-H)AlHCp* 4 is slightly favored (4 kcal mol−1). We propose that 1 could be prepared thanks to Cp*Al (2) and Cp*AlH2 (3) reaction in acidic medium. The experimental treatment of this type of metallocenes would contribute to the development of the organometallic chemistry of 13th group elements.   相似文献   

12.
Heteronuclear NMR spin relaxation studies of conformational dynamics are coming into increasing use to help understand the functions of ribozymes and other RNAs. Due to strong magnetic interactions within the ribose ring, however, these studies have thus far largely been limited to 13C and 15N resonances on the nucleotide base side chains. We report here the application of the alternate-site 13C isotopic labeling scheme, pioneered by LeMaster for relaxation studies of amino acid side chains, to nucleic acid systems. We have used different strains of E. coli to prepare mononucleotides containing 13C label in one of two patterns: Either C1′ or C2′ in addition to C4′, termed (1′/2′,4′) labeling, or nearly complete labeling at the C2′ and C4′ sites only, termed (2′,4′) labeling. These patterns provide isolated H spin systems on the labeled carbon atoms and thus allow spin relaxation studies without interference from scalar or dipolar coupling. Using relaxation studies of AMP dissolved in glycerol at varying temperature to produce systems with correlation times characteristic of different size RNAs, we demonstrate the removal of errors due to interaction in T 1 measurements of larger nucleic acids and in T measurements in RNA molecules. By extending the applicability of spin relaxation measurements to backbone ribose groups, this technology should greatly improve the flexibility and completeness of NMR analyses of conformational dynamics in RNA.  相似文献   

13.
Phosphoimidazolide-activated ribomononucleotides (*pN; see Scheme I) are useful substrates for the nonenzymatic synthesis of oligonucleotides. In the presence of metal ions dilute neutral aqueous solutions of *pN (0.01 M) typically yield only small amounts of dimers and traces of oligomers; most of *pN hydrolyzes to yield nucleoside 5′-monophosphate (5′NMP). An earlier investigation of *pN reactions in highly concentrated aqueous solutions (up to 1.4 M) showed, as expected, that the percentage yield of the condensation products increases and the yield of the hydrolysis product correspondingly decreases with *pN concentration (Kanavarioti 1997). Here we report product distributions in reactions with one, two, or three reactive components at the same total nucleotide concentration. *pN used as substrates were the nucleoside 5′-phosphate 2-methylimidazolides, 2-MeImpN, with N= cytidine (C), uridine (U), or guanosine (G). Reactions were conducted as self-condensations, i.e., one nucleotide only, with two components in the three binary U,C, U,G, and C,G mixtures, and with three components in the ternary U,C,G mixture. The products are 5′NMP, 5′,5′-pyrophosphate-, 2′,5′-, 3′,5′-linked dimers, cyclic dimers, and a small percentage of longer oligomers. The surprising finding was that, under identical conditions, including the same total monomer concentration, the product distribution differs substantially from one reaction to another, most likely due to changing intermolecular interactions depending on the constituents. Even more unexpected was the observed trend according to which reactions of the U,C,G mixture produce the highest yield of internucleotide-linked dimers, whereas the self-condensations produce the least and the reactions with the binary mixtures produce yields that fall in between. What is remarkable is that the approximately two-fold increase in the percentage yield of internucleotide-linked dimers is not due to a concentration effect or a catalyst, but to the increased complexity of the system from a single to two and three components. These observations, perhaps, provide an example of how increased complexity in relatively simple chemical systems leads to organization of the material and consequently to chemical evolution. A possible link between prebiotic chemistry and the postulated RNA world is discussed. Received: 12 September 1997 / Accepted: 24 November 1997  相似文献   

14.
Conformational preferences of the modified nucleosides N2-methylguanosine (m2G) and N2, N2-dimethylguanosine (m22G) have been studied theoretically by using quantum chemical perturbative configuration interaction with localized orbitals (PCILO) method. Automated complete geometry optimization using semiempirical quantum chemical RM1, along with ab initio molecular orbital Hartree–Fock (HF-SCF), and density functional theory (DFT) calculations has also been made to compare the salient features. Single-point energy calculation studies have been made on various models of m2G26:C/A/U44 and m22G26:C/A/U44. The glycosyl torsion angle prefers “syn” (χ = 286°) conformation for m2G and m22G molecules. These conformations are stabilized by N(3)–HC2′ and N(3)–HC3′ by replacing weak interaction between O5′–HC(8). The N2-methyl substituent of (m2G26) prefers “proximal” or s-trans conformation. It may also prefer “distal” or s-cis conformation that allows base pairing with A/U44 instead of C at the hinge region. Thus, N2-methyl group of m2G may have energetically two stable s-trans m2G:C/A/U or s-cis m2G:A/U rotamers. This could be because of free rotations around C–N bond. Similarly, N2, N2-dimethyl substituent of (m22G) prefers “distal” conformation that may allow base pairing with A/U instead of C at 44th position. Such orientations of m2G and m22G could play an important role in base-stacking interactions at the hinge region of tRNA during protein biosynthesis process.  相似文献   

15.
The interaction of the diastereomeric complexes Λ-[Ru(bpy)2(m-GHK)]Cl2 and Δ-[Ru(bpy)2(m-GHK)]Cl2 (bpy is 2,2′-bipyridine, GHK is glycine–l-histidine–l-lysine) with the deoxynucleotide duplex d(5′-CGCGATCGCG)2 was studied by means of 1H NMR spectroscopy. At a Δ-isomer to DNA ratio of 1:1, significant shifts for the metal complex are observed, whereas there is negligible effect on the oligonucleotide protons and only one intermolecular nuclear Overhauser effect (NOE) is present at the 2D nuclear Overhauser enhancement spectroscopy spectrum. The 1Η NMR spectrum at ratio 2:1 is characterized by a slight shift for the Δ-isomer’s bpy aromatic protons as well as significant shifts for the decanucleotide G4 H1′ and Η2″, A5 H2, G10 H1′, T6 NH and G2 NH protons. Furthermore, at ratio 2:1, 11 intermolecular NOEs are observed. The majority of the NOEs involve the sugar Η2′ and Η2″ protons sited in the major groove of the decanucleotide. Increasing the Δ-isomer to d(CGCGATCGCG)2 ratio to 5:1 results in noteworthy spectral changes. The Δ-isomer’s proton shifts are reduced, whereas significant shifts are observed for the decanucleotide protons, especially the sugar protons, as well as for the exchangeable protons. Interaction is characterized by the presence of only one intermolecular NOE. Furthermore, there is significant broadening of the imino proton signals as the ratio of the Δ-isomer to DΝΑ increases, which is attributed to the opening of the two strands of the duplex. The Λ-isomer, on the other hand, approaches the minor groove of the oligonucleotide and interacts only weakly, possibly by electrostatic interactions. Photocleavage studies were also conducted with the plasmid pUC19 and a 158-bp restriction fragment, showing that both diastereomers cleave DNA with similar efficiency, attacking mainly the guanines of the sequence probably by generating active oxygen species. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users. An erratum to this article can be found at  相似文献   

16.
It has been reported that stromal cell-derived factor-1 (SDF1), currently also designated CXCL12, plays a significant role in the development of nephritis and death in the lupus mice model. Using restriction length fragment polymorphism (RFLP) analysis we assessed the frequencies of SDF1-3′ G801A (rs 1801157) polymorphic variants between systemic lupus erythematosus (SLE) patients (n = 150) and controls (n = 300). There were no significant differences in the prevalence of SDF1-3′ G801A polymorphic variants in SLE patients and healthy individuals. However, we observed that the SDF1-3′ A/A and G/A genotypes (recessive model) contributed to renal manifestations of SLE OR = 3.042 (95% CI = 1.527–6.058, P = 0.002), and the p value stayed statistically significant after Bonferroni correction (p corr = 0.032) in SLE patients. We also found an association of the SDF1-3′ A/A and G/A genotypes (recessive model) with dermal manifestations of SLE OR = 2.510 (95% CI = 1.247–5.052, P = 0.0122), (p corr = 0.1952) but this did not remain statistically significant after Bonferroni correction. Our observations suggest that the SDF1-3′ G801A genotype may be associated with some clinical manifestations in patients with SLE.  相似文献   

17.
Reaction of the oxo-molybdenum(V) precursor [MoTp*(O)Cl2] [Tp* = hydrotris(3,5-dimethyl-1-pyrazolyl)borate] with H2NC6H4R-4 (R = OEt; OPr) in refluxing toluene in the presence of Et3N afforded the binuclear oxo-bridged oxo(arylimido) molybdenum(V) complexes [Tp*Mo(O)Cl](μ-O)[Tp*Mo(NC6H4OR-4)Cl]. Surprisingly, a similar reaction between [MoTp*(O)Cl2] and C6H5NH2 yielded the previously reported compound [{MoTp*(O)Cl}2(μ-O)] as the only product. The new compounds were characterized by microanalytical data, mass spectrometry, IR and 1H NMR spectroscopy. Cyclic voltammetric studies of the new compounds, of the previously reported compounds [Tp*Mo(O)Cl](μ-O)[Tp*Mo(NAr)Cl] (Ar = C6H4OMe-4, C6H4F-3, C6H4Cl-4, C6H4Br-4, and C6H4I-3), and of [{MoTp*(O)Cl}2(μ-O)] revealed a reversible one-electron oxidation process that is little affected by the nature of the substituent on the aryl group, whereas it is greatly affected by replacement of the imido ligand with an oxo ligand. The [{MoTp*(O)Cl}2(μ-O)] compound also shows a one-electron reduction process.  相似文献   

18.
Summary

Highly polymorphic DNA triplet repeats, (CAG)n, are located inside the first exon of the Huntington's disease gene. Inordinate expansion of this repeat is correlated with the onset and progression of the disease. NMR spectroscopy, gel electrophoresis, digestion by single-strand specific PI enzyme, and in vitro replication assay have been used to investigate the structural basis of (CAG)n expansion. Nondenaturing gel electrophoresis and ID 1H NMR studies of (CAG)5 and (CAG)6 reveal the presence of hairpins and mismatched duplexes as the major and minor populations respectively. However, at high DNA concentrations (i.e., 1.0–2.0 mM that is typically required for 2D NMR experiments) both (CAG)5 and (CAG)6 exist predominantly in mismatched duplex forms. Mismatched duplex structures of (CAG)5and (CAG)6 are useful, because they adequately model the stem of the biologically relevant hairpins formed by (CAG).,. We, therefore, performed detailed NMR spectroscopic studies on the duplexes of (CAG)5 and (CAG)6. We also studied a model duplex, (CGCAGCG)2 that contains the underlined building block of the duplex. This duplex shows the following structural characteristics: (i) all the nucleotides are in (C2′-endo, anti) conformations, (ii) mismatched A?A base pairs are flanked by two Watson-Crick G?C base pairs and (iii) A?A base pairs are stably stacked (and intra-helical) and are formed by a single N6-H—N1 hydrogen bond. The nature of A?A pairing is confirmed by temperature-dependent HMQC and HMQC-NOESY experiments on the [(CA*G)5]2 duplex where the adenines are 15N-labeled at N6. Temperature-and pH-dependent imino proton spectra, nondenaturing electrophoresis, and PI digestion data demonstrate that under a wide range of solution conditions longer (CAG)n repeats (n>10) exist exclusively in hairpin conformation with two single-stranded loops. Finally, an in vitro replication assay with (CAG)821 inserts in the Ml3 single-stranded DNA templates shows a replication bypass for the (CAG)21 insert but not for the (CAG)8 insert in the template. This demonstrates that for a sufficiently long insert (n=21 in this case), a hairpin is formed by the (CAG)., even in presence of its complementary strand. This observation implies that the formation of hairpin by the (CAG)n may cause slippage during replication and thus may explain the observed length polymorphism.  相似文献   

19.
The characterization of T. vulgaris plant material for quality control purposes was performed by NMR-based methods. Direct extraction of 141 T. vulgaris samples with DMSO-d 6 enabled the obtainment of crude extracts with a representative composition in terms of both volatile and non-volatile constituents. The acquisition of 600 MHz 1H NMR spectra resulted in a dataset which was analyzed by a combination of metabolic profiling and target analysis approaches. Preliminary analysis of the 1H NMR spectra was performed by principal component analysis, which revealed sample discrimination on a chemotype basis (thymol, carvacrol and linalool chemotypes). Further minor discriminative constituents were identified as p-cymene, γ-terpinene, rosmarinic acid, and 3,4,3′,4′-tetrahydroxy-5,5′-diisopropyl-2,2′-dimethylbiphenyl. Metabolite identification was accomplished by 1D and 2D NMR techniques and supported by spiking experiments. Fast dereplication of constituents not available as reference compounds was performed by HPLC–SPE–NMR experiments. A targeted approach based on qHNMR was validated for quantification of the identified secondary metabolites. Validation was performed in terms of precision (intra-day RSD ≤ 4.51%, inter-day RSD ≤ 4.18%), repeatability (RSD ≤ 2.30%), accuracy (recovery rates within 93.4 and 103.4%), linearity (correlation coefficients ≥ 0.9990), robustness, and stability. The amount of the dominant monoterpene in thymol, carvacrol, and linalool chemotypes was respectively found to be within 0.4–2.6, 0.7–2.3, and 1.1–3.6% (w/w). Variable amounts of the precursors p-cymene and γ-terpinene were found in thymol and carvacrol chemotypes. The highest amount of rosmarinic acid and 3,4,3′,4′-tetrahydroxy-5,5′-diisopropyl-2,2′-dimethylbiphenyl in the analyzed samples was respectively 4.6 and 0.4% (w/w). Since quantification is performed on a weight basis, the essential oil content can be estimated based on the sum of the quantified monoterpenes. The NMR-based analysis of T. vulgaris represents a more comprehensive approach in comparison to traditional chromatographic methods such as GC and LC, respectively employed for the analysis of volatile and non-volatile constituents. Further advantages lie in the simple sample preparation, rapidity and reproducibility of the NMR analysis.  相似文献   

20.
The binding of the stereoisomers of [{Ru(phen)2}2(μ-bpm)]4+, [{Ru(phen)2}2(μ-dppm)]4+ and [{Ru(phen)2}2(μ-bb)]4+ {phen is 1,10-phenanthroline; bpm is 2,2′-bipyrimidine, dppm is 4,6-bis(2-pyridyl)pyrimidine, bb is 1,2-bis[4-(4′-methyl-2,2′-bipyridyl)]ethane} to an oligonucleotide duplex [d(GCATCGAAAGCTACG)•d(CGTAGCCGATGC)] containing a three-base bulge has been studied using a fluorescence intercalator displacement assay. Of the dinuclear ruthenium complexes, the dppm-linked species showed the strongest binding to the oligonucleotide, with the ΔΔ isomer binding slightly more strongly than the meso isomer and the ΛΛ isomer exhibiting the weakest binding. In order to determine whether the ΔΔ-[{Ru(phen)2}2(μ-dppm)]4+ metal complex specifically bound at the three-base bulge site, a 1H NMR study of the binding of the metal complex to the oligonucleotide duplex d(GCATCGAAAGCTACG)•d(CGTAGCCGATGC) was carried out. Although a detailed picture of the metal complex–oligonucleotide association could not be determined from the NMR results owing to the broadening of the resonances from the metal complex and nucleotide residues at the bulge site, the NMR results do indicate that the metal complex specifically binds at the three-base bulge site. The combined results of this study suggest that the dppm-bridged dinuclear ruthenium complexes have considerable potential as probes for the unusual secondary structure obtained by the insertion of a three-base bulge within duplex DNA.  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