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 31 毫秒
1.
Fluorescence excitation and emission spectra, relative fluorescence quantum yield phi r and fluorescence lifetime tau of methyl 8-(2-anthroyl)-octanoate have been studied in a set of organic solvents covering a large scale of polarity and in the presence of water. In this probe, the 2-anthroyl chromophore exhibits quite remarkable and unique fluorescence properties. Thus, when going from n-hexane to methanol, the maximum emission wavelength lambda em max shifts from 404 nm to 492 nm while phi r and tau increase from 1 to 17.7 and from 0.91 ns to 13.5 ns, respectively. These increments are still more accentuated in the presence of water with estimated values of 526 nm for lambda em max, 27 for phi r and 20 ns for tau in this solvent. Because of the presence of a keto group which is a hydrogen bond acceptor and which can conjugate with the aromatic ring so as to provide the chromophore with a high dipole moment, the fluorescence properties of the probe strongly depend on the polarity of the surrounding medium. They can be accounted for in terms of general solvent effects (dipolar solute/solvent interactions) in the presence of aprotic solvents and in terms of specific solvent effects (hydrogen bonding) in protic solvents. Such properties of solvatochromism make the 2-anthroyl chromophore, after 8-(2-anthroyl)octanoic acid has been attached to phospholipids (E. Perochon and J.F. Tocanne (1991) Chem. Phys. Lipids 58, 7-17) a potential tool for studying microenvironmental polarity in biological membranes.  相似文献   

2.
An octapeptide containing a central -Aib-Gly- segment capable of adopting beta-turn conformations compatible with both hairpin (beta(II') or beta(I')) and helical (beta(I)) structures has been designed. The effect of solvent on the conformation of the peptide Boc-Leu-Val-Val-Aib-Gly-Leu-Val-Val-OMe (VIII; Boc: t-butyloxycarbonyl; OMe: methyl ester) has been investigated by NMR and CD spectroscopy. Peptide VIII adopts a well-defined beta-hairpin conformation in solvents capable of hydrogen bonding like (CD(3))(2)SO and CD(3)OH. In solvents that have a lower tendency to interact with backbone peptide groups, like CDCl(3) and CD(3)CN, helical conformations predominate. Nuclear Overhauser effects between the backbone protons and solvent shielding of NH groups involved in cross-strand hydrogen bonding, backbone chemical shifts, and vicinal coupling constants provide further support for the conformational assignments in different solvents. Truncated peptides Boc-Val-Val-Aib-Gly-Leu-Val-Val-OMe (VII), Boc-Val-Val-Aib-Gly-Leu-Val-OMe (VI), and Boc-Val-Aib-Gly-Leu-OMe (IV) were studied in CDCl(3) and (CD(3))(2)SO by 500 MHz (1)H-NMR spectroscopy. Peptides IV and VI show no evidence for hairpin conformation in both the solvents. The three truncated peptides show a well-defined helical conformation in CDCl(3). In (CD(3))(2)SO, peptide VII adopts a beta-hairpin conformation. The results establish that peptides may be designed, which are poised to undergo a dramatic conformational transition.  相似文献   

3.
The light-induced difference Fourier transform infrared spectrum between the L or N intermediate minus light-adapted bacteriorhodopsin (BR) was measured in order to examine the protonated states and the changes in the interactions of carboxylic acids of Asp-96 and Asp-115 in these intermediates. Vibrational bands due to the protonated and unprotonated carboxylic acid were identified by isotope shift and band depletion upon substitution of Asp-96 or -115 by asparagine. While the signal due to the deprotonation of Asp-96 was clearly observed in the N intermediate, this residue remained protonated in L. Asp-115 was partially deprotonated in L. The C = O stretching vibration of protonated Asp-96 of L showed almost no shift upon 2H2O substitution, in contrast to the corresponding band of Asp-96 or Asp-115 of BR, which shifted by 9-12 cm-1 under the same conditions. In the model system of acetic acid in organic solvents, such an absence of the shift of the C = O stretching vibration of the protonated carboxylic acid upon 2H2O substitution was seen only when the O-H of acetic acid is hydrogen-bonded. The non-hydrogen-bonded monomer showed the 2H2O-dependent shift. Thus, the O-H bond of Asp-96 enters into hydrogen bonding upon conversion of BR to L. Its increased hydrogen bonding in L is consistent with the observed downshift of the O-H stretching vibration of the carboxylic acid of Asp-96.  相似文献   

4.
SET domain lysine methyltransferases (KMTs) are S-adenosylmethionine (AdoMet)-dependent enzymes that catalyze the site-specific methylation of lysyl residues in histone and non-histone proteins. Based on crystallographic and cofactor binding studies, carbon-oxygen (CH · · · O) hydrogen bonds have been proposed to coordinate the methyl groups of AdoMet and methyllysine within the SET domain active site. However, the presence of these hydrogen bonds has only been inferred due to the uncertainty of hydrogen atom positions in x-ray crystal structures. To experimentally resolve the positions of the methyl hydrogen atoms, we used NMR (1)H chemical shift coupled with quantum mechanics calculations to examine the interactions of the AdoMet methyl group in the active site of the human KMT SET7/9. Our results indicated that at least two of the three hydrogens in the AdoMet methyl group engage in CH · · · O hydrogen bonding. These findings represent direct, quantitative evidence of CH · · · O hydrogen bond formation in the SET domain active site and suggest a role for these interactions in catalysis. Furthermore, thermodynamic analysis of AdoMet binding indicated that these interactions are important for cofactor binding across SET domain enzymes.  相似文献   

5.
The X-ray structures of cocrystals between 2,2'-dipyridyl-N,N'-dioxide (1) with fumaric acid (2), itaconic acid (3), succinic acid (4), and oxalic acid (5) were solved to determine if concurrent CH...O interactions were capable of orienting the bimolecular association of the two molecules. Cocrystals 1.2, 1.3 and 1.4 produce cyclic hydrogen bonded motifs employing pair-wise OH...O and CH...O hydrogen bonds, whereas cocrystal 1.5 forms analogous OH...O hydrogen bonds with a different set of intermolecular CH...O hydrogen bonds. Evidence of cocrystal formation was also observed for these complexes by differential scanning calorimetry and FT-IR spectroscopy. The structures of 1.2, 1.3 and 1.4 demonstrate the potential of the pair-wise OH...O and CH...O hydrogen bonding interactions and serve to illustrate their use as hydrogen bonding isosteres in crystal engineering, molecular recognition, and drug design.  相似文献   

6.
Sequence-specific protein-nucleic acid recognition is determined, in part, by hydrogen bonding interactions between amino acid side-chains and nucleotide bases. To examine the repertoire of possible interactions, we have calculated geometrically plausible arrangements in which amino acids hydrogen bond to unpaired bases, such as those found in RNA bulges and loops, or to the 53 possible RNA base-pairs. We find 32 possible interactions that involve two or more hydrogen bonds to the six unpaired bases (including protonated A and C), 17 of which have been observed. We find 186 "spanning" interactions to base-pairs in which the amino acid hydrogen bonds to both bases, in principle allowing particular base-pairs to be selectively targeted, and nine of these have been observed. Four calculated interactions span the Watson-Crick pairs and 15 span the G:U wobble pair, including two interesting arrangements with three hydrogen bonds to the Arg guanidinum group that have not yet been observed. The inherent donor-acceptor arrangements of the bases support many possible interactions to Asn (or Gln) and Ser (or Thr or Tyr), few interactions to Asp (or Glu) even though several already have been observed, and interactions to U (or T) only if the base is in an unpaired context, as also observed in several cases. This study highlights how complementary arrangements of donors and acceptors can contribute to base-specific recognition of RNA, predicts interactions not yet observed, and provides tools to analyze proposed contacts or design novel interactions.  相似文献   

7.
Semisynthetic zinc chlorins are shown for the first time to self-assemble in the absence of an intrinsic hydroxy group, which is always present in the chlorosomal bacteriochlorophylls (BChl's) c, d and e. Instead, the presently studied compounds have carbonyl groups. These cannot function as hydrogen bond donating groups. However due to interspacing water molecules bound to the zinc ion, double hydrogen bonding can occur to adjacent tetrapyrrolic macrocycles equipped with carbonyl recognition groups. Solution studies comprising UV-Vis absorption, electronic circular dichroism (ECD) and FT-IR show that different aggregates are formed in hydrated solvents in comparison to dry nonpolar solvents. Single crystal X-ray studies show variable supramolecular interactions either with interspacing water molecules coordinating the Zn ion within a porphyrin or with the 17(2) carbonyl group of a chlorin ligating the Zn ion. Our findings have implications for a minimalistic design of self-assembling chromophores, which can act as efficient light-harvesting units.  相似文献   

8.
Calorimetric heat effects and structural rearrangements assessed by means of Fourier transform infrared (FTIR) amide I spectra were followed by immersing dry human serum albumin and bovine pancreatic alpha-chymotrypsin in low water organic solvents and in pure water at 298 K. Enthalpy changes upon immersion of the proteins in different media are in a good linear correlation with the corresponding IR absorbance changes. Based on calorimetric and FTIR data the solvents were divided into two groups. The first group includes carbon tetrachloride, benzene, nitromethane, acetonitrile, 1,4-dioxane, n-butanol, n-propanol and pyridine where no significant heat evolution and structural changes were found during protein immersion. Due to kinetic reasons no significant protein-solvent interactions are expected in such systems. The second group of solvents includes dimethyl sulfoxide, methanol, ethanol, and water. Immersion of proteins in these media results in protein swelling and involves significant exothermic heat evolution and structural changes in the protein. Dividing of different media in the two groups is in a qualitative correlation with the solvent hydrophilicity defined as partial excess molar Gibbs free energy of water at infinite dilution in a given solvent. The first group includes the solvents with hydrophilicity exceeding 2.7 kJ/mol. More hydrophilic second group solvents have this energy values less than 2.3 kJ/mol. The hydrogen bond donating ability of the solvents also assists in protein swelling. Hydrogen bonding between protein and solvent is assumed to be a main factor controlling the swelling of dry solid proteins in the studied solvents.  相似文献   

9.
Pectinic acid and calcium pectate gels condensed into uniaxially oriented fibers have been studied by X-ray diffraction. Although the diffraction patterns correspond to systems of only limited order, they show that both systems conserve the 1.3 nm axial period and 0.43 nm pseudo-period observed in sodium pectate. Pectinic acid further resembles sodium pectate in packing isometrically in a hexagonal net of side 0.84 nm. On the other hand, calcium pectate fibers contain the 1.2 nm lateral spacing observed in pectic acid. Speculative models for pectinic acid and calcium pectate have been developed. The former structure could be stabilized by hydrophobic binding from columns of methyl groups as well as by specific intermolecular hydrogen bonds. In the latter, the main interactions between pairs of chains could be bridges formed by calcium ions, which incorporate into their co-ordination shells two polyanion oxygen atoms from one chain and three from another. These model-building studies provide plausible visualizations of two different kinds of junction zones that may exist in pectic gels.  相似文献   

10.
Circular dichroism, optical rotatory dispersion, and viscosity of hyaluronic acid at various solvents compositions, concentrations, and pH values have been studied. The data show a large change in the molecular properties in organic/water solvents such as ethanol, p-dioxane, or acetonitrile/water at pH ? pKa. At this pH range of aqueous solution, hyaluronic acid shows a CD minimum near 210 nm whereas in the presence of organic solvent it exhibits a strong negative dichroism (below 200 nm) and a positive band near 226 nm. It undergoes a sharp, cooperative transition with respect to pH and solvent. The observed CD features are assigned to the π-π* and n-π* transitions of the amide and carboxyl chromophores. The ORD results show a gradual blue shift of trough at 220 nm with increasing magnitude of rotation when the organic solvents and hydrogen ion concentrations are increased. A one-term Drude's equation was used to analyze the ORD data, and the result show a variation of dispersion parameters with different solvents in accordance with the observed CD changes. The intrinsic viscosity of hyaluronic acid in mixed solvent at pH 2.6 is lower than that of aqueous solution. All the observed property changes of hyaluronic acid are reversed on addition of foramide in mixed solvents indicating that the hydrogen bonds are involved in this transition. The observed spectroscopic and hydrodynamic features are attributed to a conformational change of hyaluronic acid in a mixed solvent involving intramolecular hydrogen bonding between the acetamido and carboxyl groups. The possible conformational state of hyaluronic acid in solution under various conditions is discussed in terms of the reported helical structure of hyaluronic acid from x-ray diffraction studies.  相似文献   

11.
A method for the rapid estimation of the extent of complex formation in molecular imprinting prepolymerization mixtures is described. By the use of a UV spectroscopy titration procedure, apparent binding constants for such self-assembly processes have been obtained. This method was used for comparison of the interactions between a dipeptide template (N-acetyl-L-phenylalaninyl-L-tryptophanyl methyl ester) and the functional monomer methacrylic acid, and the monomer analogues acetic acid and trifluoroacetic acid. The importance of template-monomer association during the molecular imprinting prepolymerization phase is discussed with respect to the systems studied.  相似文献   

12.
Javier Miragaya 《Steroids》2009,74(9):735-361
A norbornyl-2-acetyl derivative of cholic acid ([3β,5β,7α,12α]-3[(norbornyl-2-acetyl)-amino]-7,12-dihydroxycholan-24-oic acid -NbCH2CA-) was synthesized and recrystallized in two dipolar aprotic solvents (acetone, DMSO) and in one protic solvent (2-propanol). In DMSO and acetone the crystals are orthorhombic, P212121 (all their parameters being very similar) while in 2-propanol the crystal is monoclinic, P21. The inclusion complexes with the solvent have a 1:1 stochiometry with DMSO and acetone and 1:2 with 2-propanol. All solvents are forming a hydrogen bond with the amide bond of the bridge between the norbornyl residue and the steroid nucleus of the bile acid. In DMSO and acetone the β side of the steroid groups lies in the same region facilitating hydrophobic interactions, and the molecules are disposed in an antiparallel orientation (the methyl groups having a β interdigitation) forming bilayers. The width of the bilayers is 9.231 Å and 8.859 Å in DMSO and acetone, respectively. A lamellar structure is also evident for the crystal in 2-propanol (the width being 11.908 Å), but the packing is different from the previous one since a sliding between the steroid groups is observed and the methyl groups are not interdigitated. Four different hydrogen bonds are established by every steroid molecule in the NbCH2CA/DMSO (or acetone) crystal. This hydrogen bond network interconnects the hydrophilic regions of the lamellar structure. The hydrogen bond network of the NbCH2CA:2-propanol crystal is different because of the different abilities of 2-propanol to form hydrogen bonds. The side chain has a ttti conformation in the two orthorhombic crystals, and a tgtg one in the monoclinic crystal.  相似文献   

13.
(R)-(-)-α-Methoxy phenyl acetic acid, (S)-(-)-1,1'-(2-naphthol), and (R)-(+)-α-methoxy-α-trifluoromethyl phenyl acetic acid were evaluated as chiral shift reagents (CSRs) for (1)H NMR spectroscopic resolution and determination of R and S enantiomers of modafinil (MDL) in bulk drugs and formulations. Effects of the nature of CSR and the weight ratio of substrate to shift reagent on enantiomeric discrimination were investigated. Intramolecular and intermolecular hydrogen bonding interactions between the drug and the CSR seem to be the driving force for desired resolution. A mechanism was proposed to explain the interactions between (R, S)-enantiomers of MDL and (R)-(-)-α-methoxy phenyl acetic acid. The method was validated and applied successfully to determine the enantiomeric purity of MDL in tablet formulations.  相似文献   

14.
Hansen solubility parameters (HSPs), often used to predict the miscibility between two compounds, are an alternative tool in evaluating the ability of the solvent to interact via dispersion, dipole-dipole, and hydrogen bonding interactions. The aim of this paper is to find a simple way to calculate HSPs for complex mixtures of triglycerides (TAGs). HSPs were calculated using two approaches: the first assumes that the contributions to the dispersion, dipole-dipole, and hydrogen bonding interactions may be subdivided into larger functional moieties (i.e., fatty acids and fatty acid methyl esters) that are additive, while the second approach assumes that vegetable oils are comprised of mixtures of simple TAGs in the same mass fractions as the fatty acids. The HSPs obtained using the two approaches are compared to reference values determined using the “Hansen Solubility Parameters in Practice” software (HSPiP) considering the complex TAG profile for each vegetable oil.HSPs for vegetable oils, obtained with the HSPiP software, did not correspond well to the HSPs obtained from the group contribution approach, when using fatty acids, fatty acids + glycerol or fatty acid methyl esters. In contrast, the HSPs calculated for vegetable oils, assuming that all TAGs are simple and in the same mass fractions as the fatty acids, provide similar values to the HSPs obtained from the HSPiP software. Therefore, it is possible to calculate the HSPs for complex oils by simply knowing the fatty acid composition. Knowledge of HSPs may be used to rationalize the ability of certain low molecular weight molecules to develop organogels in vegetable oils as well as the crystallization of triglycerides.  相似文献   

15.
P E Blatz  J H Mohler 《Biochemistry》1975,14(11):2304-2309
The specific conteranion and the solvent have been shown to regulate the electronic excitation energy of the N-retinylidene-n-butylammonium cation. Halogenated hydrocarbon solvents which can hydrogen bond with the anion shift the lambda-max to longer wavelengths, whereas the solvent dipole, acting as a bulk effect, shifts the wavelength-max to shorter wavelength. Here solvents which can donate two hydrogens for hydrogen bonding, such as cis- and trans-1,2-dichloroethylene and cis- and trans-1,2-dichlorocyclohexane, are used as solvents for the Cl-, Br-, and I- salts. As expected the cis solvents allow longer wavelengths than do the trans solvents. Results of nuclear magnetic resonance spectroscopy are shown to be in agreement with electronic absorption spectroscopy. The C-11 proton and the C-13 and C-9 methyl protons show a considerable downfield shift in the salts with respect to the Schiff base. Furthermore the same protons show a continuing downfield shift as the anions are exchanged from Cl-, Br- to I-. This is an agreement with the interpretation of greater positive charge delocalization as the anions are changed in the above manner. The infrared absorptions of the C - N group in the Schiff base and the protonated form are shown to be almost similar. This is rationalized by showing that the force constant can remain constant as the highly related factors bond order, bond distance, and the effective electronegativity are changed in a self-compensating manner.  相似文献   

16.
We report on molecular dynamics simulations of a medium-sized protein, a lipase from Rhizomucor miehei, in vacuum, in water, and in a nonpolar solvent, methyl hexanoate. Depending on force field and solvent, the molecular dynamics structures obtained as averages over 150 ps had root-mean-square deviations in the range of 1.9 to 3.6 A from the crystal structure. The largest differences between the structures were in hydrogen bonding and exposed surface areas of the protein. The surface area increased in both solvents and became smaller in vacuum. The change of surface exposure varied greatly between different residues and occurred in accordance with the hydrophobicity of the residue and the nature of the solvent. The fluctuations of the atoms were largest in the external loops and agreed well with crystallographic temperature factors. Root-mean-square fluctuations were significantly smaller in the nonpolar solvents than they were in water, which is in accordance with the notion that proteins become more rigid in nonpolar solvents. In methyl hexanoate a partial opening of the lid covering the active site occurred, letting a methyl hexanoate molecule approach the active site.  相似文献   

17.
Temperature induced spectral shifts of the 4-aminophthalimide (4-AP) emission spectra have been measured and compared to the predictions of the McRae solvent induced shift theory (J. Phys. Chem., 1957, 61, 562-572). Three moderately polar chloroalkanes selected as nonspecifically interacting media, and six hydrogen accepting or/and electron pair donating solvents have been used as the media in which the temperature influence on 4-AP-solvent interactions has been studied in the range of 180-320 K. Using the ab initio determined 4-AP ground state dipole moment and fitting appropriate expression originating from the mentioned theory to the shifts found in the chloroalkanes it has been possible to estimate the 4-AP excited state dipole moment, the probe excited state Onsager radius and its gas phase emission spectrum position. Using these values the thermochromic shifts of 4-AP emission spectra in hydrogen bond forming solvents have been predicted and compared to the experimental one. Temperature has been found to have different impact on the changes, upon excitation of the probe, in the mean values of the energies of different hydrogen bonds formed by 4-AP with solvents molecules.  相似文献   

18.
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   

19.
In production of succinic acid by fermentation, succinic acid and acetic acid are co-produced. To purify the succinic acid from binary-acid mixture of succinic acid and acetic acid, the tertiary amine-based extraction was used. In 1-octanol, the selectivity for succinic acid was proportional to the chain length of tertiary amine. But, the distribution of acids into organic phase was low in n-heptane. These results are due to the different degree of intramolecular hydrogen bonding of succinic acid and hydrophobicity of each acid.  相似文献   

20.
Conformational possibilities of pirrolidine analogues of acetylcholine beta-(N-methyl pirrolidinium)-ethyl ester of acetic acid and beta-(N-ethyl pirrolidinium)-ethyl ester of acetic acid and beta-(N-ethyl pirrolidinium)-ethyl ester of acetic acid were investigated by the method of atomic potentials. The conformational energy was considered as a sum of non-bonded and electrostatical interactions, torsional energy and distortions of bond angles. It has been shown that the replacement of the nitrogen methyl group to ethyl group results in decrease of the average barrier height between two gauche conformations of the O--C--C--N fragment. Comparison of conformational properties of some cholinesterase substrates permit to draw a suggestion that the barrier height influences the rate of the enzymatic hydrolysis.  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