首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
We report on the comprehension of novel europium activated hybrid organic Eu(dmh)3phen (Eu: europium, dmh: 2,6‐dimethyl‐3,5‐heptanedione, phen: 1,10 phenanthroline) organo‐metallic complexes, synthesized at different pH values by the solution technique. Photo physical properties of these complexes in various basic and acidic solvents were probed by UV–vis optical absorption and photoluminescence (PL) spectra. Minute differences in optical absorption peaks with variable optical densities were encountered with the variation in solvent from basic (chloroform, toluene, tetrahydrofuran) to acidic (acetic acid) media, revealing bathochromic shift in the absorption peaks. The PL spectra of the complex in various acidic and basic organic solvents revealed the position of the emission peak at 613 nm irrespective of the changes in solvents whereas the excitation spectrum almost matched with that of the UV–vis absorption data. The optical density was found to be maximum for the complex with pH 7.0 whereas it gradually decreased when pH was lowered to 6.0 or raised to 8.0 at an interval of 0.5, demonstrating its pH sensitive nature. Several spectroscopic parameters related to probability of transition such as absorbance A(λ), Napierian absorption coefficient α(λ), molecular absorption cross‐section σ(λ), radiative lifetime (τ0) and oscillator strength (f) were calculated from UV–vis spectra. The relative intensity ratio (R‐ratio), calculated from the emission spectra was found to be almost the same in all the organic solvents. The optical energy gap, calculated for the designed complexes were found to be well in accordance with the ideal acceptance value of energy gap of the emissive materials used for fabrication of red organic light‐emitting diode (OLED). The relation between Stoke's shift and solvent polarity function was established by Lippert–Mataga plot. This remarkable independence of the electronic absorption spectra of Eu complexes on the nature of the solvent with unique emission wavelength furnishes its potential to serve as a red light emitter for solution processed OLEDs, display panels and solid‐state lighting.  相似文献   

2.
The secondary structure of bovine αs-casein and chemically modified αs-casein in various solvents was investigated by infrared absorption spectrum and optical rotatory dispersion measurements. Amino groups of αs-casein were either succinylated or acetylated, and carboxyl groups were either methylated or ethylated. Acetylated- and ethylated-αs-caseins are insoluble in water. Water-soluble samples have unordered structure in water. In organic solvents, such as 2-chloroethanol and ethylene glycol, they have about 50% α-helical fraction. On the other hand, it was found that methylated-αs-casein had two infrared absorption peaks centered at 1625 and 1643 cm?1 in D2O-CH3OD mixed solvent. This fact may be connected with the presence of β-structure. In the case of solid film of this sample, cast from solution containing CH3OH, the presence of β-structure was indicated, too. The authors attempted to explain the formation of β-structure in methylated-αs-casein in terms of the electrostatic interactions due to the differences in the net charge between methylated and unmodified αs-caseins.  相似文献   

3.
The kinetics of the complexation of Ni(II) with 1,10-phenanthroline(phen), 4,7-dimethyl-1,10-phenanthroline(dmphen), and 5-nitro-1,10-phenanthroline(NO2phen) in acetonitrile-water mixed solvents of acetonitrile mole fraction xAN = 0, 0.05, 0.1, 0.2 and 0.3 at 288, 293, 298 and 303 K have been studied by stopped-flow method at ionic strength of 1.0 (NaClO4) and pH 7.4. The corresponding activation enthalpy, entropy, and free energy were determined from the observed rate constants. The complexation of Ni(II) with the three ligands has comparable observed rate constants; in pure water the observed rate constants are (×103 dm3 mol−1 s−1) 2.31, 2.57, and 1.38 for phen, dmphen and NO2phen, respectively. The corresponding activation parameters for the three ligands are, however, considerably different; in pure water the ΔHS (kJ mol−1/J K−1 mol−1) are 44.7/−30.2, 19.5/−114.1, and 32.2/−76.9 for phen, dmphen, and NO2phen, respectively. The effects of solvent composition on the kinetics are also markedly different for the three ligands. The ΔH and ΔS showed a minimum at xAN = 0.1 for phen; for dmphen and NO2phen, however, maxima at xAN = 0.2 were observed. Nevertheless, there is an effective enthalpy-entropy compensation for the ΔHS of all the three ligands, demonstrating the significant effects of the changes in solvation and solvent structure on the complexation kinetics. As the rate-determining step of Ni(II) complexation is the dissociation of a water molecule from Ni(II), the solvent and ligand dependencies in the Ni(II) complexation kinetics are ascribed to the change in solvation status of the ligands and the altered solvent structures upon changing solvent composition.  相似文献   

4.
Abstract

Ketoreductases (KREDs) were employed for enantioselective reduction of 7-hydroxy-2-tetralone 1a and adduct 7-methoxy-2-tetralonbisulfite 2a to their corresponding (S)-/(R)-alcohols. In addition, the effect of additives such as organic solvents and β-cyclodextrin derivatives on the enzyme reductions was investigated. The changes in enzyme activity as a function of additives were correlated to structural alterations of the KREDs using circular dichroism and fluorescence spectrophotometric measurements. The effects of both the organic solvents and β-cyclodextrin derivatives on substrate solubility and equilibrium binding constants (log K) of β-cyclodextrin-substrate complexes were determined.  相似文献   

5.
In order to conclude about the feasibility of using water-immiscible organic solvents in biological waste-gas treatment, a theoretical study was done in which different types of organic-solvent-containing systems are compared with systems where the pollutant is transferred directly to the water phase. For each system the total equipment volume needed to remove 99% of a pollutant from a waste-gas stream is calculated. Three different pollutants with a different solubility in water are considered: Hexane (m gw =71), dichloromethane (m gw =0.1) and acetone (m gw =0.0016), withm gw the partition coefficient (kg/m3 gas/kg/m3 water) of the pollutant between the gas and the water phase. From the results it is concluded that the use of organic solvents is only advantageous in case the specific area for mass transfer between solvent and water is large enough to compensate for the additional transport resistance introduced by the solvent, and secondly if the solvent shows a sufficiently high affinity for the pollutants.  相似文献   

6.
Many solution processing methods of exfoliation of layered materials have been studied during the last few years; most of them are based on organic solvents or rely on surfactants and other funtionalization agents. Pure water should be an ideal solvent, however, it is generally believed, based on solubility theories that stable dispersions of water could not be achieved and systematic studies are lacking. Here we describe the use of water as a solvent and the stabilization process involved therein. We introduce an exfoliation method of molybdenum disulfide (MoS2) in pure water at high concentration (i.e., 0.14 ± 0.01 g L−1). This was achieved by thinning the bulk MoS2 by mechanical exfoliation between sand papers and dispersing it by liquid exfoliation through probe sonication in water. We observed thin MoS2 nanosheets in water characterized by TEM, AFM and SEM images. The dimensions of the nanosheets were around 200 nm, the same range obtained in organic solvents. Electrophoretic mobility measurements indicated that electrical charges may be responsible for the stabilization of the dispersions. A probability decay equation was proposed to compare the stability of these dispersions with the ones reported in the literature. Water can be used as a solvent to disperse nanosheets and although the stability of the dispersions may not be as high as in organic solvents, the present method could be employed for a number of applications where the dispersions can be produced on site and organic solvents are not desirable.  相似文献   

7.
Thermolysin-catalyzed peptide synthesis using N-benzyloxycarbonyl)-l-phenylalanine (Z-Phe) and l-phenylalanine methyl ester (Phe-OMe) as substrates was done mainly in a water-organic one phase solvent system. The organic solvent content used was less than the saturation concentration in buffer. With organic solvents with high log P values, the enzymatic activity increased as the organic solvent content increased; but further increases in the organic solvent content decreased the enzymatic activity, showing an “organic activity” profile. On the other hand, with organic solvents of low log P values, the enzymatic reaction was inhibited even by the initial addition of organic solvents. When a correlation between maximum activities and logP values or Hildebrand solubility parameters was investigated, a linear correlation was obtained among the same category of organic solvents, but not between categories. This suggests that the direct effect of organic solvents on the microenvironment of the enzyme largely depends on the molecular structure of the solvents.  相似文献   

8.
A series of new heteroleptic, tris(polypyridyl)chromium(III) complexes, [Cr(phen)2L]3+ (L = substituted phenanthrolines or bipyridines), has been prepared and characterized, and their photophyical properties in a number of solvents have been investigated. X-ray crystallography measurements confirmed that the cationic (3+) units contain only one ligand L plus two phenanthroline ligands. Electrochemical and photophysical data showed that both ground state potentials and lifetime decays are sensitive to ligand structure and the nature of the solvent with the exception of compounds containing L = 5-amino-1,10-phenanthroline (aphen) and 2,2′-bipyrimidine (bpm). Addition of electron-donating groups in the ligand structure shifts redox potentials to more negative values than those observed for the parent compound, [Cr(phen)3]3+. Emission decays show a complex dependence with the solvent. The longest lifetime was observed for [Cr(phen)2(dip)]3+ (dip = 4,7-diphenylphenanthroline) in air-free aqueous solutions, τ = 273 μs. Solvent effects are explained in terms of the affinity of hydrophobic complexes for non-polar solvent molecules and the solvent microstructure surrounding chromium units.  相似文献   

9.
The free energy of transfer, ΔG°tr, for 21 univalent ions are determined from water to methanol, acetonitrile, dimethylsulfoxide (DMSO), pyridine, tetrahydrothiophene and liquid ammonia. These solvents show a wide range of donor properties, whereby water and methanol are regarded as hard donors, dimethylsulfoxide and acetonitrile are on the borderline between hard and soft, and the remaining solvents are regarded as typical soft donors. The ΔG°tr values of ionic compounds are calculated from solubility product measurements of 1:1 salts. The extrathermodynamic tetraphenylarsonium tetraphenylborate (TATB) assumption has been applied in order to calculate the contributions from the single ions. The TATB assumption implies that the two large ions Ph4As+ and BPh4 are equally solvated, thus ΔG°tr(AsPh4+)=ΔG°tr(BPh4), for all solvent pairs. Standard electrode potentials in non-aqueous solvents can be calculated from the standard electrode potentials in water and the ΔG°tr values. The standard electrode potentials calculated from the solubility product measurements, and the potentiometrically determined ones were found to be in excellent agreement. The extrathermodynamic assumption has thereby been experimentally shown to be close to the truth.  相似文献   

10.
In this study, the effect of various organic solvents on enzyme activity and substrate enantiomeric excess (ees) of the lipase from Burkholderia cenocepacia (BCL) was investigated in the enantioselective transesterification of 1-phenylethanol. Secondary structure analysis by Fourier transform-infrared spectroscopy (FT-IR) showed that the variations in secondary structure element content (α-helix, β-sheet, β-turn and random coil) were probably responsible for the changes in enzyme activity and ees. Furthermore, the change in fluorescence intensity indicated, to some extent, the alteration in tertiary structure, which may also explain why organic solvents affect enzyme activity and ees. Moreover, response surface methodology (RSM) was employed to optimize the reaction parameters. The optimized reaction conditions were: substrate molar ratio 4.7:1; reaction time 18.6 h, and reaction temperature 53.4 °C. Under the optimal reaction conditions, the ees and eep were respectively 99.22% and 98.74%, and the corresponding enzyme activity was 1392.2 U/min/g protein. Compared with other lipases, BCL exhibited better catalytic efficiency and has significant potential in industrial applications.  相似文献   

11.
Second-order rate constants, k2, for the substitution of the ferrocene-containing β-diketonato ligands FcCOCHCOR with R=CF3 (ferrocenoyltrifluroacetonato, fctfa, pKa 6.56), CCl3 (ferrocenoyltrichloroacetonato, fctca, 7.13), CH3 (ferrocenoylacatonato, fca, 10.01), Ph (anion of benzoylferrocenoylmethane, bfcm, 10.41) and Fc (anion of diferrocenoylmethane, dfcm, 13.1) (Ph=phenyl, Fc=ferrocenyl, values in brackets are the pKa values of the free β-diketones) from the complexes [Rh(cod)(FcCOCHCOR)] with 1,10-phenanthroline (phen, cod=1,5-cyclooctadiene) at 25 °C were found to be 560 (R=CF3), 1370 (CCl3), 30 (Ph), 18 (CH3) and 7.0 dm3 mol−1 s−1 (Fc), respectively. The temperature dependence of each reaction was determined and the large negative values obtained for activation, ΔS#<−100 J K−1 mol−1 for all but R=CCl3S#CCl3=−81 J K−1 mol−1), suggests an associative substitution mechanism. The rate law of the reaction was found to be R={ks+k2[phen]}[Rh(cod)(FcCOCHCOR)]. Since the solvent-associated rate constant ks≈0 for all R except Ph (ks,RPh=0.06 s−1) the solvent, methanol, plays a limited role in the reaction. Results are interpreted to imply that the rate-determining step during substitution is breaking of an RhO bond and not the formation of an RhN bond. The role of β-diketone pKa and group electronegativity, χ, of each R group on the rate of substitution are also discussed.  相似文献   

12.
Rate constants and activation parameters (ΔH and ΔS)are reported for the oxidation of ferrocene by the tris-1,10-phenanthrolinecobalt(III) cation in t-butyl alcoholwater and in acetonewater solvent mixtures. Solvent effects on reactivity trends for these systems, for this same reaction in methanolwater mixtures, and for cobalt(II)-catalysed racemisation of Co(phen)33+ in t-butyl alcoholwater solvent mixtures are analysed into initial state and transition state contributions. The dependences of solubilities on solvent composition for ferrocene and for [Co(phen)3](ClO4)3 in methanol, t-butyl alcohol, and acetonewater mixtures are also reported; these results are needed in order to establish solvent effects on the initial states of the reactions studied.  相似文献   

13.
The series of new zirconium(IV) and hafnium(IV) phthalocyanines with various β-dicarbonyl ligands were prepared via direct interaction between di(chloro)zirconium(IV) or hafnium(IV) phthalocyanines and free β-diketones and also with 4-benzoyl-3-methyl-1-phenyl-2-pyrazolin-5-one. The structure of the obtained bis(β-dicarbonilato) zirconium(IV) and hafnium(IV) phthalocyanines was studied by two dimension 1H NMR spectroscopy (COSY, NOESY, ROESY). Absorption and fluorescence spectroscopic studies have been investigated in various solvents. Analyzed compounds of concentration range below 10?5 mol/dm3 do not aggregate in the organic solvents. Fluorescence quantum yields (ΦF) and natural life times (τ) of zirconium phthalocyanine complexes have been calculated in toluene, DMSO and THF.  相似文献   

14.
Three phenylalanine derivatives, Ac-Phe-NHMe, H-Phe-NHMe, and Ac-Phe-OH, were selected as models of Phe residues situated at the internal, the N-terminal, and the C-terminal positions of peptide chains, respctively. The side-chain conformations of the three compounds were analyzed from the vicnal coupling constants 3JαβR and 3JαβS, of their 1H- nmr spectra measured in various organic sovlent. The two β-protons were unambiguously assined by use of sterospecifically β-monodeuterated phenylalanines. The pro-S β-proton was always situated at lower field than the pro-R one when they were observed separately. The results of a solvent effect on the conformation of the tree compounds demonstrated that the rotamer populations are remarkable sensitive of the three compounds demonstrated that the rotamer populations are remarkably sensitive to solvent polarity and that the tendencies of the solvent effects are quite different from each other. Ac-Phe-OH Showed a trend similar to that of Ac-Phe-OEt reported by early workers. The rotamer populations of other derivatives (Ac-Phe-NMe2, Ac-Phe-NH2, Ac-Phe-OBut, and Ac-Phe-OBzl) and of Phe residues in some N-acetyl dipeptde esters (Ac-Phe-Gly-OMe, Ac-Phe-Val-OMe, and Ac-Gly-Phe-OMe) were also examined in several sovent, and it was found that substituents of the Phe carboxyl group—amides or esters—determine the tendency of the solvent effect. These results are interesting in the side-chain conformations of Phe residues in peptides and proteins in an environment of low polarity can be disscussed on this experimental basis. Factors responsible for the solvent effect are discussed from (1) a structural comparison of the compunds with various carboxylic substituents, (2) an expriment with cyclohexylalanine derivatives, and (3) the measurement in mixed solvents wiht similar polarity.  相似文献   

15.
Steroids are generally sparingly soluble in water. Thus, for in vitro studies of steroid metabolism or enzymology it is common practice to solubilize steroids by the addition of a small amount (2–10%, v/v) of an organic cosolvent. Methanol, ethanol, and 1,2-propanediol, singly or in combination, have been widely used (1). Effects of organic solvents on the kinetic parameters, Km and Vmax, of steroid-metabolizing enzymes with various substrates have been demonstrated (2,3), and the results are consistent with the conclusion that organic solvent influences on catalytic activity reflect, in part, effects on the aggregation state and solubility of steroid substrates.Light-scattering measurements have been applied extensively in studies of macromolecular structure (4) and micelle formation by a large variety of amphiphilic substances [reviewed in Ref. (5)]. Jones and Gordon (6) used a commercial instrument, designed specifically for light-scattering measurements, to characterize micelle formation in aqueous solutions by Δ5-3-ketosteroids containing various substituents at the 17β position. They showed that turbidity versus concentration plots were of the form seen in studies of micelle formation (5) and that steroids can exist in solution in monomeric or micellar forms, their aggregation state being a function of the polarity of the steroid solute and the composition of the solvent.To estimate solubility quantitatively 3H- or 14C-labeled steroids have been used in conjunction with centrifugation (3), dialysis (7), or filtration (8). These techniques allow for accurate estimates of solubility, but one may encounter problems due to nonspecific absorption on membranes or the unavailability of the labeled steroid of interest.We have observed that steroid aggregation and solubility can be estimated easily and with high sensitivity with a commercially available fluorometer. In this report the method is described and examples demonstrating the reproducibility and sensitivity of the technique are presented.  相似文献   

16.
Microstructure of dibenzo-18-crown-6 (DB18C6) and DB18C6/Li+ complex in different solvents (water, methanol, chloroform, and nitrobenzene) have been analyzed using radial distribution function (RDF), coordination number (CN), and orientation profiles, in order to identify the role of solvents on complexation of DB18C6 with Li+, using molecular dynamics (MD) simulations. In contrast to aqueous solution of LiCl, no clear solvation pattern is found around Li+ in the presence of DB18C6. The effect of DB18C6 has been visualized in terms of reduction in peak height and shift in peak positions of gLi-Ow. The appearance of damped oscillations in velocity autocorrelation function (VACF) of complexed Li+ described the high frequency motion to a “rattling” of the ion in the cage of DB18C6. The solvent-complex interaction is found to be higher for water and methanol due to hydrogen bond (HB) interactions with DB18C6. However, the stability of DB18C6/Li+ complex is found to be almost similar for each solvent due to weak complex-solvent interactions. Further, Li+ complex of DB18C6 at the liquid/liquid interface of two immiscible solvents confirm the high interfacial activity of DB18C6 and DB18C6/Li+ complex. The complexed Li+ shows higher affinity for water than organic solvents; still they remain at the interface rather than migrating toward water due to higher surface tension of water as compared to organic solvents. These simulation results shed light on the role of counter-ions and spatial orientation of species in pure and hybrid solvents in the complexation of DB18C6 with Li+. Graphical Abstract
DB18C6/Li+ complex in pure solvents (water, methanol, chloroform, and nitrobenzene) and water/nitrobenzene interface  相似文献   

17.
The study of the solubility of unstructured proteins (αS1-, β-, and κ-casein) and well-structured globulin (β-lactoglobulin) in low water binary solvent systems demonstrated the crucial importance of solvent polarity and neutralization of protein polar functions on the final outcome of solubility experiments. The solubilities up to 38, 56, and 96% in CHCl3/CH3OH (1/1, v/v) acidified with HCl and up to 5, 10, and 25% in CHCl3/CH3OH (1/1, v/v) in the presence of triethylamine (TEA) were obtained for κ-, αS1-, and β-casein, respectively. The importance of protein charge neutralization was apparent when the solubilization was performed in basified CHCl3/CH3OH media, giving the optimal results when the studied proteins were brought before to their isoionic point. The maximum solubility of β-casein at its pI in 30–70% methanol in CHCl3 was reaching 50–60% with triethylamine (TEA) added. β-lactoglobulin could be solubilized up to 70% in CHCl3/CH3OH (7/3, v/v) acidified with HCl and up to 40% in CHCl3/CH3OH (3/7, v/v) in the presence of TEA. The observed yield of reductive alkylation of β-lactoglobulin was much higher (98%) when performed in studied solvent system than in aqueous conditions (75%). Apparently, steric hindrance of the well-folded β-barrel (in aqueous conditions) structure masks the portion of ε-NH2 groups. In the case of unstructured aqueous media β-casein, 90% alkylation yields were obained in organic and aqueous conditions.  相似文献   

18.
Techniques for the solubilization and fractionation of integral membrane proteins have been developed in recent years. A small portion of membrane protein (about 2%, proteolipid fraction) will partition into chloroform or 1-butanol, and, in several cases, these proteins retain functional activity. A virtually complete solubilization can be achieved at neutral pH by use of aprotic solvents, like hexamethylphosphoric triamide or N-methylpyrrolidone. At relatively low concentrations (< 3 M) aprotic solvents inhibited β-D-galactoside transport by whole cells and the derivative membrane vesicles of Escherichia coli, but this inhibition could be largely reversed by a simple washing procedure. At higher concentrations of aprotic solvent (5–6 M), 50–80% of the total protein of lactose transport-positive membrane vesicles was solubilized. When these extracts were added to intact lactose transport-negative membrane vesicles, lactose transport was reconstituted, the required energy being provided by either respiration (e.g., addition of D-lactate) or by a K+ diffusion potential established with the aid of valinomycin. The dicyclohexylcarbodiimide (DCCD)-reactive subunit of the E. coli ATPase complex was found to partition into chloroform, and to be amenable to further purification in organic solvent. Ether precipitation and chromatography on DEAE-cellulose and hydroxypropyl-Sephadex G-50 yielded an homogeneous polypeptide of an apparent molecular weight of 9,000. The purified and unlabeled DCCD-reactive protein was incorporated into K+-loaded liposomes, and a membrane potential was generated by the addition of valinomycin. There are indications that the DCCD-reactive protein alone made the membrane specifically permeable for protons.  相似文献   

19.
The excellent optoelectronic properties demonstrated by hybrid organic/inorganic metal halide perovskites are all predicated on precisely controlling the exact nucleation and crystallization dynamics that occur during film formation. In general, high‐performance thin films are obtained by a method commonly called solvent engineering (or antisolvent quench) processing. The solvent engineering method removes excess solvent, but importantly leaves behind solvent that forms chemical adducts with the lead‐halide precursor salts. These adduct‐based precursor phases control nucleation and the growth of the polycrystalline domains. There has not yet been a comprehensive study comparing the various antisolvents used in different perovskite compositions containing cesium. In addition, there have been no reports of solvent engineering for high efficiency in all‐inorganic perovskites such as CsPbI3. In this work, inorganic perovskite composition CsPbI3 is specifically targeted and unique adducts formed between CsI and precursor solvents and antisolvents are found that have not been observed for other A‐site cation salts. These CsI adducts control nucleation more so than the PbI2–dimethyl sulfoxide (DMSO) adduct and demonstrate how the A‐site plays a significant role in crystallization. The use of methyl acetate (MeOAc) in this solvent engineering approach dictates crystallization through the formation of a CsI–MeOAc adduct and results in solar cells with a power conversion efficiency of 14.4%.  相似文献   

20.
The enthalpy of hydrogen-bond formation between guanine (G) and cytosine (C) in o-dichlorobenzene and in chloroform at 25°C has been determined by direct calorimetric measurement. We derivatized 2′-deoxyguanosine and 2′-deoxycytidine at the 5′- and 3′-hydroxyls with triisopropylsilyl groups; these groups increase the solubility of the nucleic acid bases in nonaqueous solvents. Such derivatization also prevents the ribose hydroxyls from forming hydrogen bonds. Consequently, hydrogen-bond formation in our system is primarily between the bases, and to a lesser extent, between base and solvent, and can be measured directly with calorimetry. To obtain the data on base-pair formation, we first took into account the contributions from self-association of each base, and where possible, have determined the ΔH of self-association. From isoperibolic titration calorimetry, our measured ΔH of C2 formation in chloroform is ?1.7 kcal/mol of C. Our measured ΔH of C:G base-pair formation in o-dichlorobenzene is ?6.65 ± 0.32 kcal/mol. Since o-dichlorobenzene does not form hydrogen bonds, the ΔH of C:G base-pair formation in this solvent represents the ΔH of the hydrogen-bonding interaction of C with G in a nonassociating solvent. In contrast, our measured ΔH of C:G base-pair formation in chloroform is ?5.77 ± 0.20 kcal/mol; thus, the absolute value of the enthalpy of hydrogen bonding in the C:G base pair is greater in o-dichlorobenzene than in chloroform. Since chloroform is a solvent known to form hydrogen bonds, the decrease in enthalpic contribution to C:G base pairing in chloroform is due to the formation of hydrogen bonds between the bases and the solvent. The ΔH of hydrogen bonding of G with C reported here differs from previous indirect estimates: Our measurements indicate the ΔH is 50% less in magnitude than the ΔH based on spectroscopic measurements of the extent of interaction. We have also observed that the enthalpy of hydrogen bonding of C with G in chloroform is greater when G is in excess than when C is in excess. This increased heat is due to the formation of C:Gn > 1 complexes that we have observed using 1H-nmr. Although C:G2 structures have previously been observed in triple-stranded polymeric nucleic acids, higher order structures have not been observed between C and G monomers in nonaqueous solvents until now. By using monomers as a model system to investigate hydrogen-bonding interactions in DNA and RNA, we have obtained the following results: A direct measurement of the ΔH of hydrogen bonding in the C:G complex in two nonaqueous solvents, and the first observation of C:Gn > 1 complexes between monomers. These results reinforce the importance of hydrogen bonding in the stabilization of various nucleic acid secondary and tertiary structures.  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