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 718 毫秒
1.
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   

2.
The substitution-inert polynuclear platinum(II) complex (PPC) series, [{trans-Pt(NH3)2(NH2(CH2)nNH3)}2-μ-(trans-Pt(NH3)2(NH2(CH2)nNH2)2}](NO3)8, where n = 5 (AH78P), 6 (AH78 TriplatinNC) and 7 (AH78H), are potent non-covalent DNA binding agents where nucleic acid recognition is achieved through use of the ‘phosphate clamp'' where the square-planar tetra-am(m)ine Pt(II) coordination units all form bidentate N–O–N complexes through hydrogen bonding with phosphate oxygens. The modular nature of PPC–DNA interactions results in high affinity for calf thymus DNA (Kapp ∼5 × 107 M−1). The phosphate clamp–DNA interactions result in condensation of superhelical and B-DNA, displacement of intercalated ethidium bromide and facilitate cooperative binding of Hoechst 33258 at the minor groove. The effect of linker chain length on DNA conformational changes was examined and the pentane-bridged complex, AH78P, was optimal for condensing DNA with results in the nanomolar region. Analysis of binding affinity and conformational changes for sequence-specific oligonucleotides by ITC, dialysis, ICP-MS, CD and 2D-1H NMR experiments indicate that two limiting modes of phosphate clamp binding can be distinguished through their conformational changes and strongly suggest that DNA condensation is driven by minor-groove spanning. Triplatin-DNA binding prevents endonuclease activity by type II restriction enzymes BamHI, EcoRI and SalI, and inhibition was confirmed through the development of an on-chip microfluidic protocol.  相似文献   

3.
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   

4.
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   

5.
Metal ions, and magnesium in particular, are known to be involved in RNA folding by stabilizing secondary and tertiary structures, and, as cofactors, in RNA enzymatic activity. We have conducted a systematic crystallographic analysis of cation binding to the duplex form of the HIV-1 RNA dimerization initiation site for the subtype-A and -B natural sequences. Eleven ions (K+, Pb2+, Mn2+, Ba2+, Ca2+, Cd2+, Sr2+, Zn2+, Co2+, Au3+ and Pt4+) and two hexammines [Co (NH3)6]3+ and [Ru (NH3)6]3+ were found to bind to the DIS duplex structure. Although the two sequences are very similar, strong differences were found in their cation binding properties. Divalent cations bind almost exclusively, as Mg2+, at ‘Hoogsteen’ sites of guanine residues, with a cation-dependent affinity for each site. Notably, a given cation can have very different affinities for a priori equivalent sites within the same molecule. Surprisingly, none of the two hexammines used were able to efficiently replace hexahydrated magnesium. Instead, [Co (NH3)4]3+ was seen bound by inner-sphere coordination to the RNA. This raises some questions about the practical use of [Co (NH3)6]3+ as a [Mg (H2O)6]2+ mimetic. Also very unexpected was the binding of the small Au3+ cation exactly between the Watson–Crick sites of a G-C base pair after an obligatory deprotonation of N1 of the guanine base. This extensive study of metal ion binding using X-ray crystallography significantly enriches our knowledge on the binding of middleweight or heavy metal ions to RNA, particularly compared with magnesium.  相似文献   

6.
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   

7.
Rhodococcus sp. strain DN22 can convert hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) to nitrite, but information on degradation products or the fate of carbon is not known. The present study describes aerobic biodegradation of RDX (175 μM) when used as an N source for strain DN22. RDX was converted to nitrite (NO2) (30%), nitrous oxide (N2O) (3.2%), ammonia (10%), and formaldehyde (HCHO) (27%), which later converted to carbon dioxide. In experiments with ring-labeled [15N]-RDX, gas chromatographic/mass spectrophotometric (GC/MS) analysis revealed N2O with two molecular mass ions: one at 44 Da, corresponding to 14N14NO, and the second at 45 Da, corresponding to 15N14NO. The nonlabeled N2O could be formed only from -NO2, whereas the 15N-labeled one was presumed to originate from a nitramine group (15N-14NO2) in RDX. Liquid chromatographic (LC)-MS electrospray analyses indicated the formation of a dead end product with a deprotonated molecular mass ion [M-H] at 118 Da. High-resolution MS indicated a molecular formula of C2H5N3O3. When the experiment was repeated with ring-labeled [15N]-RDX, the [M-H] appeared at 120 Da, indicating that two of the three N atoms in the metabolite originated from the ring in RDX. When [U-14C]-RDX was used in the experiment, 64% of the original radioactivity in RDX incorporated into the metabolite with a molecular weight (MW) of 119 (high-pressure LC/radioactivity) and 30% in 14CO2 (mineralization) after 4 days of incubation, suggesting that one of the carbon atoms in RDX was converted to CO2 and the other two were incorporated in the ring cleavage product with an MW of 119. Based on the above stoichiometry, we propose a degradation pathway for RDX based on initial denitration followed by ring cleavage to formaldehyde and the dead end product with an MW of 119.  相似文献   

8.
Polyamine analogues have been studied as potential inhibitors or substrates of barley leaf polyamine oxidase. NH2(CH2)3NH(CH2)10NH2 was particularly effective as an inhibitor of spermine oxidation at pH 4·5 (Ki = 5 × 10?6 M). Methylglyoxal-bis(guanylhydrazone) inhibited spermine oxidation only slightly (Ki = 10?4 M). Activity with the polyamine analogues as substrates was generally 10% or less of the activity with spermine. The Km for oxygen was 3 × 10?4 M. The Km for spermine oxidation was independent of oxygen concentration. Using the N-methyl-2-benzothiazolone hydrazine reagent, 1-(3-aminopropyl)pyrroline was shown to be formed stoichiometrically by the enzyme on oxidation of spermine. The enzyme will not function as a dehydrogenase in the presence of oxygen with either potassium ferricyanide or dichlorophenolindophenol as electron acceptors. Activity in the leaves increased with age, up to 4 weeks. In the leaves of 11-week-old plants activity was lower than in leaves of 1-week-old plants. The enzyme was mainly associated with an easily-sedimented particulate fraction, and relatively small proportions were found in the cell wall or soluble fractions.  相似文献   

9.
Sexual crosses were used to determine the genetic basis of resistance to the sterol 14 α-demethylase inhibitor fungicide prochloraz in the cereal eyespot pathogen Tapesia yallundae. Three different crosses between sensitive parental strains (22-432 and 22-433 [the concentration required to inhibit growth by 50% {IG50} for each was ≤0.03 mg/liter]) and field isolates from France and New Zealand with differing levels of resistance (PR11 [IG50 = 0.5 mg/liter], PR1 [IG50 = 1.0 mg/liter], and 11-3-18 [IG50 = 2.4 mg/liter]) yielded progeny showing a bimodal distribution, with an even number of sensitive and resistant progeny. This indicated the segregation of a single major gene for resistance in each cross, which was confirmed by the use of backcrosses, crosses between F1 progeny, and control crosses between sensitive parents. However, there was also evidence of additional quantitative genetic components responsible for the increased IG50s of the more resistant isolates. A further cross was made between isolate PR11 and an F1 progeny arising from isolate 11-3-18, and this also yielded progeny which were entirely prochloraz resistant. This suggested that resistance genes were allelic in these two isolates, with resistance conferred by a gene at the same locus (or closely linked loci), despite the fact that the isolates (PR11 and 11-3-18) originated from different continents.  相似文献   

10.
The potent bactericidal activity of sodium nitroprusside {SNP; Na2[Fe(CN)5(NO)]} towards Clostridium sporogenes has been investigated. SNP inhibited cell growth in the concentration range of 10 to 40 μM. Concentrations above 80 μM caused irreversible loss of cell viability and cell lysis. Inhibition of cell growth was similar in complex and in defined media. SNP was found to be unreactive towards individual components of the defined medium, with the exception of cysteine. The chemical characteristics responsible for the potency of SNP were investigated by synthesizing analogs of SNP in which the Fe was replaced by different metals. The inhibitory potency of the pentacyanonitrosyl complexes decreased in the order Fe > Cr > V, which correlates with N-O stretching frequency (vNO). In contrast, the Ru complex which had a vNO comparable to that of Fe was a poor inhibitor. Electron paramagnetic resonance spectroscopy showed that SNP was rapidly reduced to the paramagnetic Fe(I) compound [Fe(CN)4(NO)]2− on contact with cells. Analysis of fractions from SNP-treated cells showed 90% oxidation of thiols in the cell walls compared with those in control cells. The toxicity of SNP involves S-nitrosation and reduction, the lack of toxicity of the Ru analog being consistent with the fact that it has poor reactivity towards thiols. When C. sporogenes cells were exposed to sublethal concentrations of SNP and viewed under the electron microscope, they showed blisters on the surface. These results point to the cell wall surface as a primary point of attack of the nitrosyl complex.  相似文献   

11.
The cross-linking reaction described previously in the DNA and 2′-O-methyl RNA series is extended to RNA duplexes. A 17mer single-stranded RNA containing the 1,3-trans-{Pt(NH3)2[(GAG)-N7G,N7G]} intrastrand chelate, named G*AG* (* indicating a platinated base) gives, upon pairing with the complementary RNA strand, the G*AG/CUC* interstrand cross-link. The rate of the reaction in 200 mM NaClO4 is similar to that observed for DNA–RNA duplexes. It depends on the added Na+ or Mg2+ cation and on its concentration. RNA duplexes containing GA/GA or AG/AG tandem mismatches in the rearrangement triplet core were also studied. The major interstrand cross-links, G*AG/CGA* and G*AG/AGC*, are accompanied by a minor one involving the central G of the CGA or AGC complementary sequence G*AG/CG*A and G*AG/AG*C. In 200 mM NaClO4, the G*A/GA tandem mismatch does not modify the rate of the cross-linking rearrangement whereas the AG*/AG mismatch slows it down by a factor of four. Our results reflect the predominance of the local structure of the rearrangement core over the nucleophility of the cross-linking base. They also show that the reaction could be used to trap tertiary structures of naturally occurring RNAs, including those with the commonly encountered GA/GA mismatch.  相似文献   

12.
Abstract

Molecular dynamics simulations with simulated annealing are performed on polyamine-DNA systems in order to determine the binding sites of putrescine, cadaverine, spermidine and spermine on A- and B-DNA. The simulations either contain no additional counterions or sufficient Na+ ions, together with the charge on the polyamine, to provide 73% neutralisation of the charges on the DNA phosphates. The stabilisation energies of the complexes indicate that all four polyamines should stabilise A-DNA in preference to B-DNA, which is in agreement with experiment in the case of spermine and spermidine, but not in the case of putrescine or cadaverine. The major groove is the preferred binding site on A-DNA of all the polyamines. Putrescine and cadaverine tend to bind to the sugar-phosphate backbone of B-DNA, whereas spermidine and spermine occupy more varied sites, including binding along the backbone and bridging both the major and minor grooves.  相似文献   

13.
Techniques of compartmental (efflux) and kinetic influx analyses with the radiotracer 13NH4+ were used to examine the adaptation to hypoxia (15, 35, and 50% O2 saturation) of root N uptake and metabolism in 3-week-old hydroponically grown rice (Oryza sativa L., cv IR72) seedlings. A time-dependence study of NH4+ influx into rice roots after onset of hypoxia (15% O2) revealed an initial increase in the first 1 to 2.5 h after treatment imposition, followed by a decline to less than 50% of influx in control plants by 4 d. Efflux analyses conducted 0, 1, 3, and 5 d after the treatment confirmed this adaptation pattern of NH4+ uptake. Half-lives for NH4+ exchange with subcellular compartments, cytoplasmic NH4+ concentrations, and efflux (as percentage of influx) were unaffected by hypoxia. However, significant differences were observed in the relative amounts of N allocated to NH4+ assimilation and the vacuole versus translocation to the shoot. Kinetic experiments conducted at 100, 50, 35, and 15% O2 saturation showed no significant change in the Km value for NH4+ uptake with varying O2 supply. However, Vmax was 42% higher than controls at 50% O2 saturation, unchanged at 35%, and 10% lower than controls at 15% O2. The significance of these flux adaptations is discussed.  相似文献   

14.
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   

15.
The uptake and degradation of nanomolar levels of [methyl-14C]choline in estuarine water samples and in seawater filtrate cultures composed mainly of natural free-living bacteria was studied. Uptake of [14C]choline exhibited Michaelis-Menten kinetics, with Kt + Sn values of 1.7 to 2.9 nM in filtrate cultures and 1.7 to 4.1 nM in estuarine-water samples. Vmax values ranged from 0.5 to 3.3 nM · h−1. The uptake system for choline in natural microbial assemblages therefore displays very high affinity and appears able to scavenge this compound at the concentrations expected in seawater. Uptake of choline was inhibited by some natural structural analogs and p-chloromercuribenzoate, indicating that the transporter may be multifunctional and may involve a thiol binding site. When 11 nM [14C]choline was added to water samples, a significant fraction (>50%) of the methyl carbon was respired to CO2 in incubations lasting 10 to 53 h. Cells taking up [14C]choline produced [14C]glycine betaine ([14C]GBT), and up to 80% of the radioactivity retained by cells was in the form of GBT, a well-known osmolyte. Alteration of the salinity in filtrate cultures affected the relative proportion of [14C]choline degraded or converted to [14C]GBT, without substantially affecting the total metabolism of choline. Increasing the salinity from 14 to 25 or 35 ppt caused more [14C]GBT to be produced from choline but less 14CO2 to be produced than in the controls. Lowering the salinity to 7 ppt decreased [14C]GBT production and increased 14CO2 production slightly. Intracellular accumulations of [14C]GBT in the salt-stressed cultures were osmotically significant (34 mM). Choline may be used as an energy substrate by estuarine bacteria and may also serve as a precursor of the osmoprotectant GBT, particularly as bacteria are mixed into higher-salinity waters.  相似文献   

16.
We examined the rates and sustainability of methyl bromide (MeBr) oxidation in moderately low density cell suspensions (~6 × 107 cells ml−1) of the NH3-oxidizing bacterium Nitrosomonas europaea. In the presence of 10 mM NH4+ and 0.44, 0.22, and 0.11 mM MeBr, the initial rates of MeBr oxidation were sustained for 12, 12, and 24 h, respectively, despite the fact that only 10% of the NH4+, 18% of the NH4+, and 35% of the NH4+, respectively, were consumed. Although the duration of active MeBr oxidation generally decreased as the MeBr concentration increased, similar amounts of MeBr were oxidized with a large number of the NH4+-MeBr combinations examined (10 to 20 μmol mg [dry weight] of cells−1). Approximately 90% of the NH3-dependent O2 uptake activity and the NO2-producing activity were lost after N. europaea was exposed to 0.44 mM MeBr for 24 h. After MeBr was removed and the cells were resuspended in fresh growth medium, NO2 production increased exponentially, and 48 to 60 h was required to reach the level of activity observed initially in control cells that were not exposed to MeBr. It is not clear what percentage of the cells were capable of cell division after MeBr oxidation because NO2 accumulated more slowly in the exposed cells than in the unexposed cells despite the fact that the latter were diluted 10-fold to create inocula which exhibited equal initial activities. The decreases in NO2-producing and MeBr-oxidizing activities could not be attributed directly to NH4+ or NH3 limitation, to a decrease in the pH, to the composition of the incubation medium, or to toxic effects caused by accumulation of the end products of oxidation (NO2 and formaldehyde) in the medium. Additional cooxidation-related studies of N. europaea are needed to identify the mechanism(s) responsible for the MeBr-induced loss of cell activity and/or viability, to determine what percentages of cells damaged by cooxidative activities are culturable, and to determine if cooxidative activity interferes with the regulation of NH3-oxidizing activity.  相似文献   

17.
Degradation of glucose has been implicated in acetate production in rice field soil, but the abundance of glucose, the temporal change of glucose turnover, and the relationship between glucose and acetate catabolism are not well understood. We therefore measured the pool sizes of glucose and acetate in rice field soil and investigated the turnover of [U-14C]glucose and [2-14C]acetate. Acetate accumulated up to about 2 mM during days 5 to 10 after flooding of the soil. Subsequently, methanogenesis started and the acetate concentration decreased to about 100 to 200 μM. Glucose always made up >50% of the total monosaccharides detected. Glucose concentrations decreased during the first 10 days from 90 μM initially to about 3 μM after 40 days of incubation. With the exception at day 0 when glucose consumption was slow, the glucose turnover time was in the range of minutes, while the acetate turnover time was in the range of hours. Anaerobic degradation of [U-14C]glucose released [14C]acetate and 14CO2 as the main products, with [14C]acetate being released faster than 14CO2. The products of [2-14C]acetate metabolism, on the other hand, were 14CO2 during the reduction phase of soil incubation (days 0 to 15) and 14CH4 during the methanogenic phase (after day 15). Except during the accumulation period of acetate (days 5 to 10), approximately 50 to 80% of the acetate consumed was produced from glucose catabolism. However, during the accumulation period of acetate, the rate of acetate production from glucose greatly exceeded that of acetate consumption. Under steady-state conditions, up to 67% of the CH4 was produced from acetate, of which up to 56% was produced from glucose degradation.  相似文献   

18.
M-DNA is a complex between the divalent metal ions Zn2+, Ni2+ and Co2+ and duplex DNA which forms at a pH of ~8.5. The stability and formation of M-DNA was monitored with an ethidium fluorescence assay in order to assess the relationship between pH, metal ion concentration, DNA concentration and the base composition. The dismutation of calf thymus DNA exhibits hysteresis with the formation of M-DNA occurring at a higher pH than the reconversion of M-DNA back to B-DNA. Hysteresis is most prominent with the Ni form of M-DNA where complete reconversion to B-DNA takes several hours even in the presence of EDTA. Increasing the DNA concentration leads to an increase in the metal ion concentration required for M-DNA formation. Both poly(dG)•poly(dC) and poly(dA)•poly(dT) formed M-DNA more readily than the corresponding mixed sequence DNAs. For poly(dG)•(poly(dC) M-DNA formation was observed at pH 7.4 with 0.5 mM ZnCl2. Modified bases were incorporated into a 500 bp fragment of phage λ DNA by polymerase chain reaction. DNAs in which guanine was replaced with hypoxanthine or thymine with 5-fluorouracil formed M-DNA at pHs below 8 whereas substitutions such as 2-aminoadenine and 5-methylcytosine had little effect. Poly[d(A5FU)] also formed a very stable M-DNA duplex as judged from Tm measurements. It is evident that the lower the pKa of the imino proton of the base, the lower the pH at which M-DNA will form; a finding that is consistent with the replacement of the imino proton with the metal ion.  相似文献   

19.
The effects of macrocyclic polyamines and polymethylenediamines on various reactions influenced by polyamines have been studied. Among the amines tested, 2,3,4,3- and 3,3,3,4-cyclic polyamines, NH2(CH2)6NH2 and NH2(CH2)8NH2 had some ability to stimulate polyphenylalanine synthesis, globin synthesis and rat liver isoleucyl-tRNA formation. The degree of stimulation was at most 40% of that obtained by polyamines. In the degradation of poly(C) by bovine pancreatic RNAase A, all tested amines stimulated the degradation. In the NADPH-dependent lipid peroxidation of rat liver microsomes, the degree of inhibition by 2,3,2,3- or 2,3,3,3-cyclic polyamine was greater than that by spermine. The hydrolysis of ATP by an oligomycin-sensitive ATPase was inhibited by 2,3,4,3- and 3,3,3,4-cyclic polyamines, NH2(CH2)10NH2 and spermine at somewhat comparable levels. None of the macrocyclic polyamines or polymethylenediamines stimulated the growth of a polyamine-requiring mutant of Escherichia coli. Possible explanations for the differences in the effects of amines on the various reactions are discussed.  相似文献   

20.
Monoclonal antibodies specific for the cyclobutane pyrimidine dimer (CPD) are widely used for detection and quantification of DNA photolesions. However, the mechanisms of antigen binding by anti-CPD antibodies are little understood. Here we report NMR analyses of antigen recognition by TDM-2, which is a mouse monoclonal antibody specific for the cis-syn-cyclobutane thymine dimer (T[c,s]T). 31P NMR and surface plasmon resonance data indicated that the epitope recognized by TDM-2 comprises hexadeoxynucleotides centered on the CPD. Chemical shift perturbations observed for TDM-2 Fab upon binding to d(T[c,s]T) and d(TAT[c,s]TAT) were examined in order to identify the binding sites for these antigen analogs. It was revealed that d(T[c,s]T) binds to the central part of the antibody-combining site, while the CPD-flanking nucleotides bind to the positively charged area of the VH domain via electrostatic interactions. By applying a novel NMR method utilizing a pair of spin-labeled DNA analogs, the orientation of DNA with respect to the antigen-binding site was determined: CPD-containing oligonucleotides bind to TDM-2 in a crooked form, draping the 3′-side of the nucleotides onto the H1 and H3 segments, with the 5′-side on the H2 and L3 segments. These data provide valuable information for antibody engineering of TDM-2.  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