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 40 毫秒
1.
Chromophoric [4-(dimethylamino)cinnamoyl]imidazole reacts with the serine protease alpha-chymotrypsin to form an acyl enzyme. At pHs below 4.0, the acyl enzyme turns over very slowly to yield the free acid. During this slow deacylation it is possible to obtain a very good resonance Raman spectrum of the acyl intermediate by using the 350.7-nm line of the krypton laser. The resonance Raman carbonyl frequency of the covalently bonded substrate and its wavelength at maximum intensity in the absorption spectrum of the acyl enzyme have been taken and used to monitor the active site environment. A comparison has been made of the absorption and Raman spectra of the acyl enzyme and those of the corresponding chromophoric methyl ester, aldehyde, and imidazole model compounds. A linear correlation is found between the wavelength of maximum absorption and the Raman frequency of the carbonyl group over a wide range of solvent conditions for each of the model compounds. By combining the Raman carbonyl frequency with the absorption maximum, we can determine that the bond order changes in the carbonyl bond of the bound substrate are not due to changes in the solvent, since the carbonyl frequency and the absorption maximum of the acyl enzyme do not fall on any of the linear correlations for the model compounds. The unusual spectroscopic properties of the bound substrate appear to be due to some specific enzyme-induced change in the substrate when it is bound at the active site. Thermal unfolding of the acyl enzymes changes both the carbonyl frequency of the acyl enzyme and its absorption maximum to completely different values.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Despite the availability of many experimental data and some modeling studies, questions remain as to the precise mechanism of the serine proteases. Here we report molecular dynamics simulations on the acyl-enzyme complex and the tetrahedral intermediate during the deacylation step in elastase catalyzed hydrolysis of a simple peptide. The models are based on recent crystallographic data for an acyl-enzyme intermediate at pH 5 and a time-resolved study on the deacylation step. Simulations were carried out on the acyl enzyme complex with His-57 in protonated (as for the pH 5 crystallographic work) and deprotonated forms. In both cases, a water molecule that could provide the nucleophilic hydroxide ion to attack the ester carbonyl was located between the imidazole ring of His-57 and the carbonyl carbon, close to the hydrolytic position assigned in the crystal structure. In the "neutral pH" simulations of the acyl-enzyme complex, the hydrolytic water oxygen was hydrogen bonded to the imidazole ring and the side chain of Arg-61. Alternative stable locations for water in the active site were also observed. Movement of the His-57 side-chain from that observed in the crystal structure allowed more solvent waters to enter the active site, suggesting that an alternative hydrolytic process directly involving two water molecules may be possible. At the acyl-enzyme stage, the ester carbonyl was found to flip easily in and out of the oxyanion hole. In contrast, simulations on the tetrahedral intermediate showed no significant movement of His-57 and the ester carbonyl was constantly located in the oxyanion hole. A comparison between the simulated tetrahedral intermediate and a time-resolved crystallographic structure assigned as predominantly reflecting the tetrahedral intermediate suggests that the experimental structure may not precisely represent an optimal arrangement for catalysis in solution. Movement of loop residues 216-223 and P3 residue, seen both in the tetrahedral simulation and the experimental analysis, could be related to product release. Furthermore, an analysis of the geometric data obtained from the simulations and the pH 5 crystal structure of the acyl-enzyme suggests that since His-57 is protonated, in some aspects, this crystal structure resembles the tetrahedral intermediate.  相似文献   

3.
I.r. difference spectra are presented for 3-(indol-3-yl)acryloyl-, cinnamoyl-, 3-(5-methylthien-2-yl)acryloyl-, dehydrocinnamoyl- and dihydrocinnamoyl-chymotrypsins at low pH, where the acyl-enzymes are catalytically inactive. At least two absorption bands are seen in each case in the ester carbonyl stretching region of the spectrum. Cinnamoyl-chymotrypsin substituted at the carbonyl carbon atom with 13C was prepared. A difference spectrum in which 13C-substituted acyl-enzyme was subtracted from [12C]acyl-enzyme shows two bands in the ester carbonyl region and thus confirms the assignment of the features to the single ester carbonyl group. The frequencies of the ester carbonyl bands are interpreted in terms of differential hydrogen-bonding. In each case a lower-frequency relatively narrow band is assigned to a productive potentially reactive binding mode in which the carbonyl oxygen atom is inserted in the oxyanion hole of the enzyme active centre. The higher-frequency band, which is broader, is assigned to a non-productive binding mode in each case, where a water molecule bridges from the carbonyl oxygen atom to His-57; this mode is equivalent to the crystallographically determined structure of 3-(indol-3-yl)acryloyl-chymotrypsin, i.e. the Henderson structure. A difference spectrum of dihydrocinnamoyl-chymotrypsin taken at higher pH shows resolution of a feature centred upon 1731 cm-1, which is assigned to a non-bonded conformer in which the carbonyl oxygen atom is not hydrogen-bonded. Perturbation of the protein spectrum in the presence of acyl groups is interpreted in terms of enhanced structural rigidity. It is reported that the ester carbonyl region of the difference spectrum of cinnamoyl-subtilisin is complicated by overlap of features that arise from protein perturbation. Measurements of carbonyl absorption frequencies in a number of solvents of the methyl esters of the acyl groups used to make acyl-enzymes have permitted determination of the apparent dielectric constants experienced by carbonyl groups in the enzyme active centre as well as a discussion of the effects of polarity. The ester carbonyl bond strengths of the various conformations were estimated by using simple harmonic oscillator theory and an empirical relation between the force constants and bond strengths. The fractional bond breaking induced by hydrogen-bonding was used to calculate rate enhancement factors by using absolute reaction rate theory.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

4.
We describe herein the design and in vitro biochemical evaluation of a novel class of mechanism-based inhibitors of human leukocyte elastase (HLE) that inactivate the enzyme via an unprecedented enzyme-induced sulfonamide fragmentation cascade. The inhibitors incorporate in their structure an appropriately functionalized saccharin scaffold. Furthermore, the inactivation of the enzyme by these inhibitors was found to be time-dependent and to involve the active site. Biochemical, HPLC, and mass spectrometric studies show that the interaction of these inhibitors with HLE results in the formation of a stable acyl complex and is accompanied by the release of (L) phenylalanine methyl ester. The data are consistent with initial formation of a Michaelis-Menten complex and subsequent formation of a tetrahedral intermediate with the active site serine (Ser(195)). Collapse of the tetrahedral intermediate with tandem fragmentation results in the formation of a highly reactive conjugated sulfonyl imine which can either react with water to form a stable acyl enzyme and/or undergo a Michael addition reaction with an active site nucleophilic residue (His(57)). It is also demonstrated herein that this class of compounds can be used in the design of inhibitors of serine proteases having either a neutral or basic primary substrate specificity. Thus, the results suggest that these inhibitors constitute a potential general class of mechanism-based inhibitors of (chymo)trypsin-like serine proteases.  相似文献   

5.
The crystal structure of the acyl complex of porcine pancreatic elastase with its peptidyl ester substrate N-acetyl-ala-ala-ala-methyl ester (Ac(Ala)3OMe) has been determined at 2.5 A resolution. The complex was stabilized by exploiting the "glass transition" in protein dynamics that occurs at around -53 degrees C (220 K). Substrate was flowed into the crystal in a cryoprotective solvent above this temperature, and then the crystal was rapidly cooled to a temperature below the transition to trap the species that formed. The use of a flow cell makes the experiment a kinetic one and means that the species prior to the rate determining transition state has a chance to accumulate. The resulting crystal structure shows an acyl-enzyme intermediate in which the leaving group is absent and the carbonyl carbon of the C-terminal alanine residue is covalently bound to the gamma oxygen of the active site serine. The ester carbonyl shows no significant distortion from planarity, with the carbonyl oxygen forming one hydrogen bond with the oxyanion hole. The tripeptide is bound in an extended antiparallel beta-sheet with main chain residues of the enzyme. The geometry and interactions of this acyl-enzyme suggest that it represents a productive intermediate. To test this hypothesis, the same crystal was then warmed above the glass transition temperature and a second data set was collected. The resulting electron density map shows no sign of the substrate, indicating hydrolysis of the intermediate followed by product release. This experiment provides direct evidence for the importance of dynamic properties in catalysis and also provides a blueprint for the stabilization of other short-lived species for direct crystallographic observation.  相似文献   

6.
A nucleophilic group in the active site of aldehyde dehydrogenase, which covalently binds the aldehyde moiety during the enzyme-catalyzed oxidation of aldehydes to acids, was acylated with the chromophoric aldehyde trans-4-(N,N-dimethylamino)cinnamaldehyde (DACA). Acyl-enzyme trapped by precipitation with perchloric acid was digested with trypsin, and the peptide associated with the chromophoric group was isolated and shown to be Gln-Ala-Phe-Gln-Ile-Gly-Ser-Pro-Trp-Arg. After redigestion with thermolysin, the chromophore was associated with the C-terminal hexaresidue part. If the chromophore is attached to this peptide, serine would be expected to bind the aldehyde and lead to the required acylated derivative. Differential labeling experiments were performed in which all free thiol groups on the acylated enzyme were blocked by carboxymethylation. The acyl chromophore was then removed by controlled hydrolysis and the protein reacted with [14C]iodoacetamide. No 14C-labeled tryptic peptides were isolated, suggesting that the sulfur of a cysteine cannot be the acylated residue in the precipitated acyl-enzyme.  相似文献   

7.
The interaction of F with high and low spin ferric deuteroporphyrin IX dimethyl ester and a low spin model compound, bis(histidine methyl ester) deuterohemin IX has been studied in dimethylformamide solution by low-temperature EPR. The reaction of F with these complexes leads to high spin compounds. The structure of the EPR line at g = 2 is due to superhyperfine interactions with axial fluoride ligands. It allows their identification as mono- or difluoride complexes. Their optical absorption spectra are reported. In the particular cases of bis(imidazole) deuterohemin IX dimethyl ester and of the model compound, the variations of the EPR spectra as functions of concentration of ionic ligand are reported. Three new low spin complexes are thus obtained. They are characterized by a specific interaction of F with the NH group of the imidazole ring. This is proved following a second independent study in which we report the changes in g tensor principal values of low spin ferric porphyrins with the basicity (pKa) of various nitrogenous bases.  相似文献   

8.
L-Cystine diketopiperazine (1), L-cysteinyl-cysteine disulfide -HCl (2), L-cysteinyl-cysteine disulfide methyl ester -HCl (3), and t-butyloxycarbonyl-L-cysteinyl-cysteine disulfide methyl ester (4) are investigated by CD, ultraviolet, 13C NMR, infrared and laser Raman spectroscopy. The temperature dependence of the 13C NMR signals of 1 reveals an exceptionally high energy barrier of deltaGNo. = 15.8 +/- 0.2 kcal/mol for the reversible change in helicity of the inherently dissymmetric disulfide bridge of 1. The P-helical diastereomer predominates in dimethyl-sulfoxide at 25 degrees C, with 80-85% of the molecules having this configuration. The Cotton effects of 1 are larger and show smaller temperature coefficients than the conformationally more mobile cystine compounds 2 and 3. After dissolving crystals of 1 in 95% ethanol there is a time-dependent decrease of the ellipticity of the negative Cotton effect at 225 nm, indicating a conformational change in going from crystal to solution. Besides 1, 2 and 3 are at present the only known examples of cystine derivatives with C-S-S-C torsional angles around 90 degrees, which do not exhibit optical activity in the long wavelength disulfide absorption, as is predicted for 1 from the Linderberg-Michl model. At 305 nm a new weak Cotton effect was discovered for 1. The solvent dependence of the CD spectra is discussed and the infrared and Raman spectra are assigned.  相似文献   

9.
Aspartate 145 plays an essential role in the active site of 4-chlorobenzoyl-CoA dehalogenase, forming a transient covalent link at the 4-position of the benzoate during the conversion of the substrate to 4-hydroxybenzoyl-CoA. Replacement of Asp 145 by residues such as alanine or serine results in total inactivation, and stable complexes can be formed with either substrate or product. The Raman spectroscopic characterization of some of the latter is described in the preceding publication (Dong et al.). The present work investigates complexes formed by D145N dehalogenase and substrate or product. Time-resolved absorption and Raman difference spectroscopic data show that these systems evolve rapidly with time. For the substrate complex, initially the absorption and Raman spectra show the signatures of the substrate bound in the active site of the asparagine 145 form of the enzyme but these signatures are accompanied by those for the ionized product. After several minutes these signatures disappear to be replaced with those closely resembling the un-ionized product in the active site of wild-type dehalogenase. Similarly, for the product complex, the absorption and Raman spectra initially show evidence for ionized product in the active site of D145N, but these are rapidly replaced by signatures closely resembling the un-ionized product bound to wild-type enzyme. It is proposed that product bound to the active site of asparagine 145 dehalogenase catalyzes the deamidation of the asparagine side chain to produce the wild-type aspartate 145. For the complexes involving substrate, the asparagine 145 enzyme population contains a small amount of the WT enzyme, formed by spontaneous deamidation, that produces product. In turn, these product molecules catalyze the deamidation of Asn 145 in the major enzyme population. Thus, conversions of substrate to product and of D145N to D145D dehalogenase go on simultaneously. The spontaneous deamidation of asparagine 145 has been characterized by allowing the enzyme to stand at RT in Hepes buffer at pH 7.5. Under these conditions deamidation occurs with a rate constant of 0.0024 h-1. The rate of product-catalyzed deamidation in Hepes buffer at 22 degrees C was measured by stopped-flow kinetics to be 0.024 s-1, 36000 times faster than the spontaneous process. A feature near 1570 cm-1 could be observed in the early Raman spectra of both substrate and product-enzyme complexes. This band is not associated with either substrate or product and is tentatively assigned to an ester-like species formed by the attack of the product's 4-O- group on the carbonyl of asparagine's side chain and the subsequent release of ammonia. A reaction scheme is proposed, incorporating these observations.  相似文献   

10.
The 1H-NMR spectra and the resonance Raman spectra of intermediate spin complex, octaethylporphyrinatoiron (III) perchlorate (OEP-Fe(III)ClO4) and its mono imidazole adduct have been recorded and analyzed. The perchlorate complex was determined to be an intermediate-spin state (S = 3/2) in dichloromethane. The mono imidazole and 2-methylimidazole adducts of OEP-Fe(III)ClO4 were of the high-spin state in dichloromethane, which is a good model for the ferrihemoproteins such as metmyoglobins. The spin state of OEP-Fe(III)ClO4 varies the polarity of solvent from typical high-spin (S = 5/2) to typical low-spin (S = 1/2) state including intermediate-spin state (S = 3/2). The resonance Raman studies of the intermediate-spin complex in various solvents indicate that the complex is a plausible model to reproduce anomalous physico-chemical properties of the ferricytochrome c' at physiological condition.  相似文献   

11.
The crystal structure of the acyl complex of porcine pancreatic elastase with its peptidyl ester substrate N-acetyl-ala-ala-ala-methyl ester (Ac(Ala)3OMe) has been determined at 2.5 Å resolution. The complex was stabilized by exploiting the “glass transition” in protein dynamics that occurs at around −53 °C (220 K). Substrate was flowed into the crystal in a cryoprotective solvent above this temperature, and then the crystal was rapidly cooled to a temperature below the transition to trap the species that formed. The use of a flow cell makes the experiment a kinetic one and means that the species prior to the rate determining transition state has a chance to accumulate. The resulting crystal structure shows an acyl-enzyme intermediate in which the leaving group is absent and the carbonyl carbon of the C-terminal alanine residue is covalently bound to the gamma oxygen of the active site serine. The ester carbonyl shows no significant distortion from planarity, with the carbonyl oxygen forming one hydrogen bond with the oxyanion hole. The tripeptide is bound in an extended antiparallel β-sheet with main chain residues of the enzyme. The geometry and interactions of this acyl-enzyme suggest that it represents a productive intermediate. To test this hypothesis, the same crystal was then warmed above the glass transition temperature and a second data set was collected. The resulting electron density map shows no sign of the substrate, indicating hydrolysis of the intermediate followed by product release. This experiment provides direct evidence for the importance of dynamic properties in catalysis and also provides a blueprint for the stabilization of other short-lived species for direct crystallographic observation.  相似文献   

12.
The properties of a-chymotrypsin methylated at histidine-57 were examined to explain the mechanism of this enzyme which is about 105 times less active than chymotrypsin. Studies on the protein showed (i) an alteration in the acyl and leaving group specificity, (ii) decreased binding of some protein protease inhibitors by methyl chymotrypsin, (iii) lack of dimerization of methyl chymotrypsin at low pH, (iv) decreased stability of methyl chymotrypsin in urea, (v) a larger solvent deuterium isotope effect with methyl chymotrypsin, and (vi) decreased binding of a tetrahedral intermediate analog to methyl chymotrypsin. These properties suggest that while only subtle alterations occur in the active site upon methylation of His-57, the transition state and the tetrahedral intermediate are destabilized but not to the same extent. General base catalysis remains an integral feature of the hydrolytic mechanism of the modified chymotrypsin, and the base appears to be the methylated nitrogen of the imidazole moiety of His-57.  相似文献   

13.
13C NMR spectroscopy has been used to demonstrate that 13CN-labeled benzoylamidoacetonitrile forms a covalent adduct with the thiol group of cysteine 25 in the active site of papain. Spectral comparison with model compounds indicates that the adduct is a thioimidate. On the basis of a proposed mechanism for the formation of the thioimidate, it is concluded that the -CH2C(= NH)S--imino nitrogen does not sit in the active site in the same manner as the thiol ester carbonyl oxygen of the thiol acyl enzyme (or the oxyanion of the tetrahedral intermediate). Thus, in this sense the stabilization of the thioimidate does not reflect a similarity in structure between the bound thioimidate and the transition state.  相似文献   

14.
4-trans-(N,N-Dimethylamino)cinnamaldehyde (DACA) is a chromophoric substrate of aldehyde dehydrogenase (EC 1.2.1.3) whose fate can be followed during catalysis. During this investigation we found that DACA also fluoresces and that this fluorescence is enhanced and blue-shifted upon binding to aldehyde dehydrogenase. Binding of DACA to aldehyde dehydrogenase also occurs in the absence of coenzyme. Benzaldehyde (a substrate), acetophenone (a substrate-competitive inhibitor), and the substrate-competitive affinity reagent bromoaceto-phenone interfere with DACA binding. Thus, DACA binds to the active site and can be employed for titration of active aldehyde dehydrogenase. Both E1 and E2 isozymes, which are homotetramers, bind DACA with dissociation constants of 1–4 M with a stoichiometry of 2 mol DACA/mol enzyme. The stoichiometry of enzyme–acyl intermediate was also found to be 2 mol DACA/mol enzyme for both E1 and E2 isozymes. Thus, both enzymes appear to have only two substrate-binding sites which participate in catalysis. The level of enzyme–acyl intermediate remained constant at different pH values, showing that enhancement of velocity with pH was not due to altered DACA–enzyme levels. When the reaction velocity was increased even further by using 150 M Mg2+ the intermediate level was decreased, suggesting that both increased pH and Mg2+ promote decomposition of the DACA–enzyme intermediate. Titration with DACA permits study of aldehyde substrate catalysis before central complex interconversion.  相似文献   

15.
Although known to be important factors in promoting catalysis, electric field effects in enzyme active sites are difficult to characterize from an experimental standpoint. Among optical probes of electric fields, Raman spectroscopy has the advantage of being able to distinguish electronic ground-state and excited-state effects. Earlier Raman studies on acyl derivatives of cysteine proteases [Doran, J. D., and Carey, P. R. (1996) Biochemistry 35, 12495-502], where the acyl group has extensive pi-electron conjugation, showed that electric field effects in the active site manifest themselves by polarizing the pi-electrons of the acyl group. Polarization gives rise to large shifts in certain Raman bands, e.g. , the C=C stretching band of the alpha,beta-unsaturated acyl group, and a large red shift in the absorption maximum. It was postulated that a major source of polarization is the alpha-helix dipole that originates from the alpha-helix terminating at the active-site cysteine of the cysteine protease family. In contrast, using the acyl group 5-methylthiophene acryloyl (5-MTA) as an active-site Raman probe, acyl enzymes of thiol- or selenol-subtilisin exhibit no polarization even though the acylating amino acid is at the terminus of an alpha-helix. Quantum mechanical calculations on 5-MTA ethyl thiol and selenol ethyl esters allowed us to identify the conformational states of these molecules along with their corresponding vibrational signatures. The Raman spectra of 5-MTA thiol and selenol subtilisins both showed that the acyl group binds in a single conformation in the active site that is s-trans about the =C-C=O single bond. Moreover, the positions of the C=C stretching bands show that the acyl group is not experiencing polarization. However, the release of steric constraints in the active site by mutagenesis, by creating the N155G form of selenol-subtilisin and the P225A form of thiol-subtilisin, results in the appearance of a second conformer in the active sites that is s-cis about the =C-C=O bond. The Raman signature of this second conformer indicates that it is strongly polarized with a permanent dipole being set up through the acyl group's pi-electron chain. Molecular modeling for 5-MTA in the active sites of selenol-subtilisin and N155G selenol-subtilisin confirms the findings from Raman spectroscopic studies and identifies the active-site features that give rise to polarization. The determinants of polarization appear to be strong electron pull at the acyl carbonyl group by a combination of hydrogen bonds and the field at the N-terminus of the alpha-helix and electron push from a negatively charged group placed at the opposite end of the chromophore.  相似文献   

16.
4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) is a chromophoric and fluorogenic substrate of aldehyde dehydrogenase. Fluorescence of DACA is enhanced by binding to aldehyde dehydrogenase in the absence of catalysis both in the presence and absence of the coenzyme analogue 5′AMP. DACA binds to aldehyde dehydrogenase with a dissociation constant of 1–3 μM and stoichiometry of 2 mol mol−1 enzyme. Incorporation of DACA during catalysis was also investigated and found to be 2 mol DACA mol−1 enzyme. Effect of pH on the stoichiometry of DACA incorporation during catalysis has shown that DACA incorporation remained constant at 2 mol DACA mol−1 enzyme, despite a 74-fold velocity enhancement between pH 5.0 and 9.0. Increase of pH increased decomposition of enzyme–acyl intermediate without affecting the rate-limiting step of the reaction. At pH 7.0 the pH stimulated velocity enhancement was 10-fold over that at pH 5.0; further velocity enhancement (11.5-fold that of pH 7.0) was achieved by 150 μM Mg2+ ions. The velocity at pH 7.0 with Mg2+ exceeded that of pH 9.0, and that at maximal pH stimulation at pH 9.5. It was observed that level of intermediate decreased to about 1 mol mol−1 enzyme, indicating that Mg2+ ions increased the rate of decomposition of the enzyme–acyl intermediate and shifted the rate-limiting step of the reaction to another step in the reaction sequence.  相似文献   

17.
C R Cremo  J C Grammer  R G Yount 《Biochemistry》1988,27(22):8415-8420
Myosin subfragment 1 (S1) can be specifically photomodified at the active site without polypeptide chain cleavage by irradiating the stable MgADP-orthovanadate-S1 complex with UV light above 300 nm [Grammer, J. C., Cremo, C. R., & Yount, R. G. (1988) Biochemistry (preceding paper in this issue)]. Here, the UV spectral properties of photomodified S1 were used to determine the nature and location of the photomodified residue(s) within S1. By comparison of the unusual pH dependence of the UV absorption spectrum of the photomodified S1 to that of the S1-MgADP-Vi complex as a control, the photomodified residue(s) was (were) localized to the 23-kDa NH2-terminal tryptic peptide of the heavy chain. NaBH4 reduced the photomodified S1, but not the control, to regenerate the original spectral properties and ATPase activities of the unmodified S1. Amino acid analysis of photomodified S1 reduced with NaB3H4 gave only [3H]serine, suggesting the hydroxyl group of serine had been oxidized to a "serine aldehyde". The pH dependence of the absorption spectrum of the photomodified enzyme can be explained by an equilibrium between a chromophoric enolate anion of the serine aldehyde (favored in base) and less chromophoric keto and enol forms (favored in acid). The oxidized serine(s) was (were) shown to be directly involved with the vanadate-dependent photocleavage of the S1 heavy chain previously described by Grammer et al. (1988). This serine(s) is (are) likely to be important to the binding and hydrolysis of the gamma-PO4 of ATP at the active site of S1.  相似文献   

18.
Raman spectroscopic data indicate that the conformations of the two hydrocarbon chains of dipalmitoyl phosphatidylcholine in aqueous dispersions of the lipid differ signficantly. The compounds 1-palmitoyl, 2-palmitoyl-d31-3-sn-phosphatidylcholine and 1-palmitoyl-d31, 2-palmitoyl-3-sn-phosphatidylcholine were synthesized. Aqueous dispersions of these phospholipids display very similar phase behavior, with both premelting and melting transitions at nearly identical temperatures, midway between the comparable transition temperatures of undeuterated and completely deuterated dipalmitoyl phosphatidylcholine. We have monitored the state of chains 1 and 2 of these molecules simultaneously and independently by Raman spectroscopy. Raman difference spectra taken between samples of the two compounds under identical conditions show significant features. We attribute these spectral differences to nonequivalent conformations of the fatty acyl chains attached at positions 1 and 2 on the glycerol backbone. Below the pretransition the conformation of chain 2 is, on average, slightly less all-trans than is the chain at position 1. There is some evidence that the conformations of the terminal methyl group of the two chains are significantly different at low temperatures.  相似文献   

19.
The spectroscopic properties and dynamics of the lowest excited singlet states of peridinin and two derivatives have been studied by steady-state absorption and fast-transient optical spectroscopic techniques. One derivative denoted PerOlEs, possesses a double bond and a methyl ester group instead of the r-ylidenebutenolide of peridinin. Another derivative denoted PerAcEs, is the biosynthetic precursor of peridinin and possesses a triple bond and a methyl ester group corresponding to the r-ylidenbutenolide function. Ultrafast time-resolved spectroscopic experiments in the visible and near-infrared regions were performed on the molecules and reveal the energies and regarding the structural features and interactions responsible for the unusual solvent-induced changes in the steady-state and transient absorption spectra and dynamics of dynamics of the excited electronic states. The data also provide information peridinin.  相似文献   

20.
Ferric iron protoporhyrin IX derivatives in SDS micelles have been investigated by means of visible absorption, resonance Raman, and XANES spectroscopies to establish specific correlations between the marker bands of the pentacoordinate derivatives obtained from the three different techniques. Hydroxyl and 1,2-dimethyl imidazole coordinated hemins display the typical spectroscopic marker bands of a pentacoordinate high-spin ferric iron derivative in both Raman and XANES spectra. In turn, the optical absorption spectra of these two derivatives are very different. This difference is in line with the assignment of hydroxyl as the fifth coordination ligand to free hemin in SDS micelles, as demonstrated by the isotopic shift of the frequency of Fe-OH bond with H(2)(18)O. The present assignments are relevant to the identification of the coordination state and the nature of the fifth ligand in ferric heme proteins.  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