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 0 毫秒
1.
2.
To understand the structure-activity relationship of isoquinoline alkaloids, absorption, fluorescence, circular dichroism, and thermodynamics were employed to study the interaction of five C-9-ω-amino alkyl ether analogs from the plant alkaloid berberine with double-stranded calf thymus DNA. The C-9 derivatization resulted in dramatic enhancements in the fluorescence emission of these compounds. The most remarkable changes in the spectral and binding properties were in the BC4 and BC5 derivatives. Interactions of these analogs, which have an additional recognition motif with DNA, were evaluated through different spectroscopic and calorimetric titration experiments. The analogs remarkably enhanced the DNA binding affinity and the same was directly dependent on the alkyl chain length. The analog with six alkyl chains enhanced the DNA binding affinity by about 33 times compared with berberine. The binding became more entropically driven with increasing chain length. These results may be of potential use in the design of berberine derivatives and understanding of the structure-activity relationship for improved therapeutic applications.  相似文献   

3.
X-ray diffraction analyses of fibers of polydeoxyadenylic acid · polydeoxythymidylic acid show that this molecule exists as a 10-fold double-helix with axial rise per nucleotide h = 3.24 to 3.29 A?. The structure is very similar to B-DNA (h = 3.37 A?) in having C3-exo furanose rings and base-pairs positioned centrally on the helix axis, but distinctive enough to have two packing modes, neither of which has been observed for B-DNA. Although the triple-stranded poly(dT) · poly(dA) · poly(dT) also has a large value of h(3.26 Å), each of the chains is a 12-fold helix of the A-genus with C3-endo furanose rings and bases displaced several Angstrom units from the helix axis.  相似文献   

4.
The binding properties of [RuL2(mip)]2+ {where L is 1,10-phenanthroline (phen) or 4,7-dimethyl-1,10-phenanthrollne (4,7-dmp) and mip is 2′-(3″,4″-methylenedioxyphenyl)imidazo[4′,5′-f][1,10]phenanthroline} with regard to the triplex RNA poly(U)·poly(A)*poly(U) were investigated using various biophysical techniques and quantum chemistry calculations. In comparison with [Ru(4,7-dmp)2(mip)]2+, remarkably higher binding affinity of [Ru(phen)2(mip)]2+ for the triplex RNA poly(U)·poly(A)*poly(U) was achieved by changing the ancillary ligands. The stabilization of the Hoogsteen-base-paired third strand was improved by about 10.9 °C by [Ru(phen)2(mip)]2+ against 6.6 °C by [Ru(4,7-dmp)2(mip)]2+. To the best of our knowledge, [Ru(phen)2(mip)]2+ is the first metal complex able to raise the third-strand stabilization of poly(U)·poly(A)*poly(U) from 37.5 to 48.4 °C. The results reveal that the ancillary ligands have an important effect on third-strand stabilization of the triplex RNA poly(U)·poly(A)*poly(U) when metal complexes contain the same intercalative ligands.  相似文献   

5.
Deoxyribonucleic acid (DNA) is a vital molecule for life since it contains genetic information. However, DNA has recently been reported to have unique properties that make it suitable for bionanoelectronic applications, such as the possibility of electrical conductivity and self-organisation. Self-assembled DNA network structures have been observed on several substrates, but the detailed self-assembly mechanism has yet to be determined. The present study investigates self-assembled structures of DNA both theoretically and experimentally. We developed a reaction–diffusion model and used it to investigate pattern formations observed by atomic force microscopy. The computational results qualitatively replicate the network patterns of DNA molecules based on a quantitative agreement with the surface size and timescale. The model can account for the effect of the DNA concentration on pattern formation. Furthermore, peculiar geometric patterns are simulated for mica and highly oriented pyrolytic graphite surfaces.  相似文献   

6.
Abstract

X-ray diffraction in fibres revealed that the calcium salt of poly(dA) · poly(dT) is a 10-fold double helix with a pitch of 3.23 nm. The opposite sugar-phosphate chains in the refined model are characterized by a complete conformational equivalence and contain sugars in a conformation close to C2′-endo.

As a result a new model of the sodium salt of poly(dA) · poly(dT)has been constructed, which is different from the Heteronomous DNA proposed earlier (S. Arnott et al., Nucl. Acids Res. 11, 4141 (1983)). The new model of Na-poly(dA) · poly(dT) has conformationally similar opposite chains; it is a structure of the B-type, rather like that of Ca-poty(dA) · poly(dT).  相似文献   

7.
The binding of a dimeric distamycin analog (Pt–bis–Dst) to poly[d(A–T)]poly[d(A–T)], poly(dA)poly(dT), and duplex O23 with the sequence 5’-GCCAATATATATATATTATTAGG-3’, which occurs at the origin of replication (OriS) of the herpes simplex virus, was studied via UV and CD spectroscopy. The synthetic polyamide differs from the natural antibiotic in having two distamycin moieties that are linked via a glycine cis-diamino platinum group. The Pt–bis–Dst binding to poly[d(A–T)]poly[d(A–T)] and poly(dA)poly(dT) reached saturation at approximately one ligand molecule per eight bp. As the ligand–base pair ratio further increased, the maximum wavelength band tended to shift toward longer wavelengths in the CD spectra of complexes with poly[d(A–T)]poly[d(A–T)] and a shoulder appeared in the 290–310 nm spectral region that was absent from the CD spectra of complexes with lower ligand coverages. At higher ligand–oligonucleotide molar ratios, Pt–bis–Dst could bind to poly[d(A–T)]poly[d(A–T)] in the form of hairpins or associations that result from interactions between the distamycin moieties of two neighbor Pt–bis–Dst molecules. The structures of the complexes were stabilized by interactions between the pirrolcarboxamide moieties of two Pt–bis–Dst molecules absorbed on adjacent overlapping binding sites. The interactions could also be responsible for the concentration-dependent spectral changes that were observed during the formation of a complex between Pt–bis–Dst and poly[d(A–T)]poly[d(A–T)]. Spectral changes were almost absent in the case of Pt–bis–Dst binding to poly(dA)poly(dT). The binding of Pt–bis–Dst to duplex O23 reached saturation at two ligand molecules per duplex, which contained a cluster of 18 AT pairs. At higher molar-concentration ratios, duplex CD spectra underwent changes similar to those that were observed for Pt–bis–Dst binding to poly[d(A–T)]poly[d(A–T)]. Testing Pt–bis–Dst for antiviral activity identified 1.5 μg/mL as a concentration that halved the cytopathic effect of the herpes simplex virus on Vero E6 cells; the selectivity index of antiviral action was 65; cytotoxicity was relatively low. The Pt–bis–Dst concentration that caused the death of approximately half of the cells was estimated at 100 μg/mL.  相似文献   

8.
The inhibitory effect of the polypeptide antibiotics netropsin and distamycin A on DNA dependent nucleic acid synthesis has been shown to be related to the base composition of the template DNA. A number of natural DNA's of quite different dA·dT content as well as poly (dI-dC)·poly (dI-dC), poly (dA-dT)·poly (dA-dT), poly (dA) · poly (dT) and poly (dG)·poly(dC) has been studied as templates in DNA and in part in RNA polymerase reaction. The highest binding efficiency of netropsin existing for (dA·dT)-containing DNA polymers and the less pronounced interaction with the (dI·dC)-containing polymer shown by the melting and CD spectral behaviour of the complexes are entirely reflected in the template inactivation. The same is evident for distamycin A. However, in contrast to netropsin the antibiotic distamycin A exhibits some binding tendency to poly (dG)·poly (dC). Binding effects of a netropsin derivative to DNA and (dA·dT)-containing polymers suggest the importance of hydrogen bonds of the peptide groups in the complex formation.  相似文献   

9.
Monomeric bis-benzimidazole (MB) is an analog of the Hoechst 33258 dye. The enthalpy and entropy of MB binding were evaluated by analyzing the calorimetric data on MB reverse titration with poly(dA) · poly(dT). A mathematical model was developed to estimate the thermodynamic parameters of binding on the basis of calorimetric data. The results agree well with spectrophotometric data on the binding of analogous compounds. The model was used to estimate the parameters of binding with poly(dA) · poly(dT) for dimeric bis-benzimidazole (DB), which consists of two bis-benzimidazole monomers linked via a flexible chain. The ligand was assumed to produce different types of complexes with the polymer.  相似文献   

10.
Abstract

Oligodeoxyribonucleotides containing dA·dU base combinations were shown to form parallel stranded DNA. CD spectra and hyperchromicity profiles provide evidence that the structure is very similar to that of a related parallel stranded dA·oligomer. Thermal denaturation studies show that these parallel dAdU sequences are significantly less stable than their dA·analogues in either antiparallel or parallel stranded orientations. The stabilizing effect of the 5- methyl group is similar for parallel and antiparallel sequences. The minor groove binding drug Hoechst 33258 binds with similar affinity to APS dA·and APS dA·dU sequences. However, binding to the PS dA·hairpin is significantly impaired as a consequence of the different groove dimensions and the presence of thymine methyl groups at the binding site. This results in an 8.6 kJmoF reduced free energy of binding for the PS dA·sequence. Replacement of the bulky methyl group with a hydrogen (ie. T -> U) results in significantly stronger Hoechst 33258 binding to the parallel dA·dU sequences with a penalty of only 4.1 kJmol?1. Our data demonstrate that although Hoechst 33258 detects the altered groove, it is still able to bind a PS duplex containing dA·dU base pairs with high affinity, despite the large structural differences from its regular binding site in APS DNA.  相似文献   

11.
The binding of two sugar containing antibiotics viz. aristololactam-β-D-glucoside and daunomycin with single and double stranded poly(A) was investigated by spectroscopic and calorimetric studies. The binding affinity of daunomycin to ss poly(A) was of the order of 106 M− 1 and that to ds poly(A) was of the order of 105 M− 1. Aristololactam-β-D-glucoside showed a relatively weaker binding with an affinity of the order of 104 M− 1 with both the conformations of poly(A). Fluorescence studies showed maximum quenching for daunomycin-ss poly(A) complexes. The binding constants calculated from fluorescence spectroscopy were in good agreement with that obtained from UV spectroscopy. Moderate perturbation of circular dichroic spectra of both the conformations of poly(A) in presence of these molecules with concomitant formation of prominent extrinsic CD bands in the 300-450 nm region further revealed the association. Isothermal titration calorimetry results showed an overall entropy driven binding in all the four systems though the entropy change was maximum in daunomycin-ss poly(A) binding. The binding affinity was also maximum for daunomycin-ss poly(A) and varied as daunomycin-ds poly(A)>aristololactam-β-D-glucoside-ds poly(A)>aristololactam-β-D-glucoside-ss poly(A). A 1:1 binding stoichiometry was observed in all the cases, as confirmed by Job plot analysis, indicating the interaction to consist of a single binding mode. Ferrocyanide quenching studies showed good stacking interaction in all cases but was best for daunomycin-ss poly(A) interaction. No self-structure formation was observed in poly(A) with both daunomycin and aristololactam-β-D-glucoside suggesting the hindrance of the sugar moiety for such structural organization.  相似文献   

12.
Using the AMBER software package (Weiner and Kollman 1981) substantially modified for electrostatic contributions, the structural energies of the double-stranded oligonucleotides dA12·dT12 and d(GCTCGAAAAA)4·d(TTTTTCGAGC)4 were minimized. Using various starting structures for the molecule dA12·dT12, one final structure is obtained which possesses the experimentally determined properties of poly(dA)·poly(dT). This structure is an A-form-B-form-hybrid structure similar to that of Arnott et al. (1983). The dA-strand is similar to an A-form while the dT-strand is similar to normal B-form. This structure and separately optimized B-form sequence stretches were used to construct the double-stranded fragment d(GCTCGAAAAA)4 which again was optimized. This sequence, when imbedded in a DNA fragment as contiguous repeats, shows a gel migration anomaly which has been interpreted as stable curvature of the DNA (Diekmann 1986). The calculated structure of this sequence indeed has a curved helix axis and is discussed as a model for curved DNA. A theoretical formalism is presented which allows one to calculate the structural parameters of any nucleic acid double helix in two different geometrical representations. This formalism is used to determine the parameters of the base-pair orientations of the curved structure in terms of wedge as well as cylindrical parameters. In the structural model presented here, the curvature of the helix axis results from an alternation of two different DNA structures in which the base-pairs possess different angles with the helix axis (cylinder tilt). Resulting from geometric restraints, a negative cylinder tilt angle correlates strongly with the closing of the minor groove (wedge roll). The blocks with different structure are not exactly coincident with the dA5-blocks and the B-DNA stretches. Within the dA5 block, base-pair tilt and wedge roll adopt large values which proceed into the 3 flanking B-DNA sequence by about one base-pair. These properties of the structure calculated here are discussed in terms of different models explaining DNA curvature.  相似文献   

13.
Jiang Z 《Biomacromolecules》2011,12(5):1912-1919
Candida antarctica lipase B (CALB) was successfully used to promote synthesis of aliphatic poly(carbonate-co-ester) copolymers from dialkyl carbonate, diol, and lactone monomers. The polymerization reactions were carried out in two stages: first-stage oligomerization under low vacuum, followed by second-stage polymerization under high vacuum. Therefore, copolymerization of ω-pentadecalactone (PDL), diethyl carbonate (DEC), and 1,4-butanediol (BD) yielded PDL-DEC-BD copolymers with a M(w) of whole product (nonfractionated) up to 33?000 and M(w)/M(n) between 1.2 and 2.3. Desirable reaction temperature for the copolymerization was found to be ~80 °C. The copolymer compositions, in the range from 10 to 80 mol % PDL unit content versus total (PDL + carbonate) units, were effectively controlled by adjusting the monomer feed ratio. Reprecipitation in chloroform/methanol mixture allowed isolation of the purified copolymers in up to 92% yield. (1)H and (13)C NMR analyses, including statistical analysis on repeat unit sequence distribution, were used to determine the polymer microstructures. The synthesized PDL-DEC-BD copolymers possessed near random structures with all possible combinations of PDL, carbonate, and butylene units via either ester or carbonate linkages in the polymer chains and are more appropriately named as poly(PDL-co-butylene-co-carbonate).  相似文献   

14.
Abstract

Monte-Carlo simulation of poly(dA) · poly(dT) hydration by 30 water molecules per nucleotide pair has been performed. Two B-family conformations, both with a 36° helical twist but with different minor groove widths, were considered. One conformation is Arnott's standard B form, the other one is specific for poly(dA) · poly(dT) B′ form with a narrowed minor groove. The mean energies and the mean numbers of water-water and water-DNA hydrogen bonds are close for the two conformations. Nevertheless, the hydration shell of the B' form differs drastically from that of the standard B form. The water arrangement in the minor groove of the B′ form resembles the spine of hydration in the central part of Dickerson's dodecamer d(CGCGAATTCGCG). No such spine is formed in the hydration shell of the usual B form with a wider minor groove. In this conformation water bridges between adenine N3 or thymine O2 and oxygen of the sugar ring of the neighbouring nucleotide along the chain can be formed (“strings” in Dickerson's decamer d(CCAAGATTGG)).  相似文献   

15.
Conventional polyurethanes (PUs) are among biomaterials not intended to degrade but are susceptible to hydrolytic, oxidative and enzymatic degradation in vivo. Biodegradable PUs are typically prepared from polyester polyols, aliphatic diisocyanates and chain extenders. In this work we have developed a degradable monomer based on α-amino acid to accelerate hard segment degradation. Thus a new class of degradable poly(ether–urethane–urea)s (PEUUs) was synthesized via direct reaction of 4,4′-methylene-bis(4-phenylisocyanate) (MDI), l-leucine anhydride (LA) and polyethylene glycol with molecular weight of 1,000 (PEG-1000) as polyether soft segment. The resulting polymers are environmentally biodegradable and thermally stable. Decomposition temperatures for 5 % weight loss occurred above 300 °C by TGA in nitrogen atmospheres. Some structural characterization and physical properties of these polymers before and after degradation in soil, river water and sludge are reported. The environmental degradation of the polymer films was investigated by SEM, FTIR, TGA, DSC, GPC and XRD techniques. A significant rate of degradation occurred in PEUU samples under river water and sludge condition. The polymeric films were not toxic to E. coli (Gram negative), Staphylococcus aureus and Micrococcus (Gram positive) bacteria and showed good biofilm formation on polymer surface. Our results show that hard segment degraded selectively as much as soft segment and these polymers are susceptible to degradation in soil and water. Thus our study shows that new environment-friendly polyurethane, which can degrade in soil, river water and sludge, is synthesized.  相似文献   

16.
Objective: To compare the health status of groups of Primary Sjögren's and Xerostomia patients, using the Medical Outcomes Short Form 36 (SF‐36). The SF‐36 is a generic measure, divided into eight domains, used in the assessment of health‐related quality of life. Patients and methods : The SF‐36 was given to 2 groups: Group 1 comprised 43 patients diagnosed with Primary Sjögren's Syndrome (1SS) and an unstimulated whole salivary flow rate (UFR) of <0.1 ml/min). Group 2 (n = 40) reported Xerosiomia but had an UFR >0.2 ml/min. Sub groups of patients in Groups 1 and 2 were compared with community normative data, for the SF‐36 Results: There were trends to suggest lower SF36 scores for 1SS patients but there were no significant differences between the mean domain scores of Groups 1 and 2. 1SS and Xerostomia patients registered lower mean scores across all 8 domains, compared with normative community data. Conclusion: The SF‐36 was unable to detect significant differences between subjects with 1SS and Xerostomia but a larger sample size is required to confirm these findings. The results of this limited study suggest that a disease‐specific measure is required to assess the impact 1SS on health‐related Quality of life (QOL).  相似文献   

17.
In this study, we compared the interactions of invasive and non-invasive strains of E. coli with clinical and environmental isolates of Acanthamoeba. The environmental isolate of Acanthamoeba exhibited significantly higher association with E. coli compared with the clinical isolates of Acanthamoeba. The ratio of E. coli per amoebae was more than 8-fold higher in the environmental isolate compared with the clinical isolates of Acanthamoeba. Interestingly, non-pathogenic environmental Acanthamoeba showed uptake and/or survival of the non-invasive E. coli. In contrast, clinical isolates of Acanthamoeba did not support uptake and/or survival of non-invasive E. coli. Using several mutants derived from K1, we demonstrated that outer membrane protein A (OmpA) and lipopolysaccharide (LPS) are crucial bacterial determinants responsible for E. coli K1 interactions and in the intracellular survival of E. coli in Acanthamoeba. The use of Acanthamoeba as a model to study E. coli K1 pathogenesis and to understand bacterial immune evasion strategies is discussed further.  相似文献   

18.
Several new acridinium esters 2 – 9 having their central acridinium ring bearing a 9-(2,5-dimethylphenoxycarbonyl), 9-(2,6-bis(trifluoromethyl)phenoxycarbonyl) or 9-(2,6-dinitrophenoxycarbonyl) group, and a 10-methyl, 10-(3-(succinimidyloxycarbonyl)propyl), 10-(5-(succinimidyloxycarbonyl)pentyl), or 10-(10-(succinimidyloxycarbonyl)decyl) group, have been synthesized and their chemiluminescent properties have been tested. The 2,5-dimethylphenyl acridinium esters emit light slowly (glow) when treated with alkaline hydrogen peroxide, while the 2,6-dinitrophenyl and 2,6-bis(trifluoromethyl)phenyl esters emit light rapidly (flash). The substituent at the 10 position affects the hydrolytic stabilities of the compounds.  相似文献   

19.
Carbapenems are used to control the outbreak of β-lactamases expressing bacteria. The effectiveness of drugs is influenced by its interaction with human serum albumin (HSA). Strong binding of carbapenems to HSA may lead to decreased bioavailability of the drug. The non-optimal drug dosage will provide a positive selection pressure on bacteria to develop resistance. Here, we investigated the interaction between meropenem and HSA at physiological pH 7.5 (N-isoform HSA) and non-physiological pH 9.2 (B-isoform HSA). Results showed that meropenem quenches the fluorescence of both ‘N’ and ‘B’ isoforms of HSA (ΔG < 0 and binding constant ~104 M?1). Electrostatic interactions and van der Waal interactions along with H-bonds stabilized the complex of meropenem with ‘N’ and ‘B’ isoforms of HSA, respectively. Molecular docking results revealed that meropenem binds to HSA near Sudlow’s site II (subdomain IIIA) close to Trp-214 with a contribution of a few residues of subdomain IIA. CD spectroscopy showed a change in the conformation of both the isoforms of HSA upon meropenem binding. The catalytic efficiency of HSA (only N-isoform) on p-nitrophenyl acetate was increased primarily due to a decrease in Km and an increase in kcat values. This study provides an insight into the molecular basis of interaction between meropenem and HSA.  相似文献   

20.
Abstract

A normal coordinate analysis has been carried out on guanosine and cytidine residues appearing in oligo and polynucleotides by using a simplified valence force field that allows the vibrational spectra of 5′-dGMP and 2′-deoxycytidine molecules to be reproduced. The role of both C2′-endo and C3′-endo conformations on sugar pucker, as well as that of glycosidic torsion angle (χ), on several characteristic vibration modes of these residues have been studied. The present calculations based on a non-redundant set of internal coordinates preserving the harmonic approximation of the potential field, allows us to explain quite satisfactorily the modifications of the vibrational spectra in the 1550-1250 cm?1 and 785-500 cm?1 regions, when the right → left-handed conformational transition occurs.  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