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.
Benzene-1,2-, 1,3-, and 1,4-di-N-substituted carbamates (1-15) are synthesized as the conformationally constrained inhibitors of acetylcholinesterase and mimic gauche, eclipsed, and anti-conformations of acetylcholine, respectively. All carbamates 1-15 are characterized as the pseudo substrate inhibitors of acetylcholinesterase. For a series of geometric isomers, the inhibitory potencies are as follows: benzene-1,4-di-N-substituted carbamate (para compound) > benzene-1,3-di-N-substituted carbamate (meta compound) > benzene-1,2-di-N-substituted carbamate (ortho compound). Therefore, benzene-1,4-di-N-substituted carbamates (para compounds), with the angle of 180 degrees between two C(benzene)-O bonds, mimic the preferable anti C-O/C-N conformers of acetylcholine for the choline ethylene backbone in the acetylcholinesterase catalysis.  相似文献   

2.
Conformationally constrained 1,2-, 1,3-, and 1,4-benzene-di-N-octylcarbamates are potent reversible competitive inhibitors of Naja mocambique mocambique phospholipase A(2) with the K(i) values of 11, 4, and 15 microM, respectively. With the angle of 120(o) between two C(benzene)-O bonds, 1,3-benzene-di-N-octylcarbamate mimics the preferable eclipsed C(sn-2)-O/C(sn-3)-O conformer of phospholipid in the enzyme-phospholipid complex. Further, a three-step phospholipase A(2) inhibition mechanism by the inhibitor is proposed.  相似文献   

3.
The interaction of diacylglycerols, primarily 1,2-dilauroyl-sn-glycerol (1,2-DLG), with egg phosphatidylcholine (PC) bilayers was studied by NMR spectroscopy and other physical techniques. In the low proportions used (less than or equal to 20 mol % with respect to total lipid), 1,2-DLG formed bilayers with PC with no hexagonal phase separation, as assessed by light, polarizing and electron microscopy, and 31P and 13C NMR spectroscopy. The 13C-carbonyl chemical shift of 90% [13C]carbonyl 1,2-DLG was monitored in small unilamellar vesicles as a function of relative DLG content (1.5-20%) and temperature (10-55 degrees C). The chemically inequivalent sn-1 and sn-2 carbonyls gave a single, narrow resonance in vesicles, in contrast to neat 1,2-DLG and 1,2-DLG in organic solvents, whose spectra showed two well-separated carbonyl resonances. The chemical shift of 1,2-DLG in PC shows that the carbonyl groups are proximal to the aqueous interface, necessitating orientation of the DLG molecule along the normal to the bilayer. Both carbonyl groups are H-bonded to H2O, but the secondary ester (sn-2) carbonyl is relatively more hydrated than the primary ester (sn-1) carbonyl. The 13C-carbonyl chemical shift data further suggest that the interfacial conformation resembles that of crystalline and liquid crystalline lamellar 1,2-dilauroyl-sn-glycero-3-phosphatidylethanolamine and certain PCs, in which the glycerol backbone is perpendicular to the bilayer plane. This conformation is different from that of crystalline 1,2-dilauroyl-sn-glycerol, in which the glycerol backbone is parallel to the bilayer plane. Between 1.5 and 8% DLG in vesicles, the chemical shift of the 1,2-DLG carbonyl at a given temperature was constant. However, above 8% DLG the chemical shift at each temperature increased with increasing DLG concentration, suggesting increased hydration at higher DLG content. At low temperatures 13C NMR spectra of vesicles with the highest proportions of 1,2-DLG studied (15 and 20%) showed two DLG carbonyl resonances, which most likely represent 1,2-DLG on outer and inner leaflets of the vesicle bilayer. The two peaks collapsed into a single resonance by 38 degrees C, at which temperature the two environments equilibrate with a rate constant of approximately 60 s-1 (t1/2 approximately 10 ms). Thus, transbilayer movement of DLG is extremely fast compared with phospholipids. In vesicles the 1,3-isomer of DLG exhibited a narrow carbonyl peak slightly downfield from that of 1,2-DLG. Acyl chain migration from 1,2-DLG to 1,3-DLG was monitored directly in the vesicle by time-dependent NMR measurements.  相似文献   

4.
Fourier transform infrared spectroscopy was used to characterize the lamellar phases of 1,3-dipalmitoyl-sn-glycero-2-phosphocholine (1,3-DPPC), a positional isomer of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (1,2-DPPC). The molecule exists in three distinct phases over the temperature interval 0-70 degrees C. In the low-temperature (LC) phase, the spectra are indicative of acyl chains packed in an orthorhombic subcell, while the carbonyl groups and phosphate ester at the head group show evidence of only partial hydration. The transition from the low-temperature (LC) phase to the intermediate-temperature (L beta) phase at 25 degrees C corresponds to a temperature-induced head-group hydration in which the hydration of the phosphate and carbonyl ester groups results in the reorganization of the hydrocarbon chain-packing subcell from orthorhombic to hexagonal. The transition from the intermediate (L beta) to the high-temperature (L alpha) phase at 37 degrees C is a gel-to-liquid-crystalline phase transition analogous to the 41.5 degrees C transition of 1,2-DPPC. The spectra of the acyl-chain carbonyl groups show evidence of significant differences in molecular conformation at the carbonyl esters in the LC phase. In the L beta and L alpha phases, the carbonyl band contour becomes much more symmetric. However, two components are clearly present in the spectra indicating that the sn-1 and sn-3 carbonyls experience slightly different environments. The observed differences are likely due to a preferred conformation of the phosphocholine group relative to the glycerol backbone. Indications from the infrared spectra of differences in the structure of the C = O groups provide a possible explanation for the selection of the sn-1 chain of 1,3-DPPC by phospholipase A2 on the basis of a preferred head group conformation.  相似文献   

5.
The ortho-, meta-, and para-trifluoromethylphenylhydrazine inhibitors of porcine plasma amine oxidase were synthesized. Titrations of plasma amine oxidase with these inhibitors demonstrated that 1 mol of trifluoromethylphenylhydrazine completely and irreversibly modifies 1 mol of enzyme by covalently binding to the active carbonyl cofactor. NMR relaxation measurements on the fluorine nuclei were obtained at 188.22 and 74.84 MHz for each inhibitor-enzyme adduct. These measurements were used to calculate the exact distance and orientation between the inhibitor-binding site and the copper cofactor. The copper lies in the plane of the aromatic ring of the inhibitor at distances of 10.9, 14.3, and 15.5 A from the fluorines in the ortho-, meta-, and para- positions of the ring, respectively. Since the inhibitors react with the active carbonyl cofactor, this defines the relationship between the copper and the active carbonyl cofactor in the enzyme, and provides a basis for choosing between mechanisms for the transfer of electrons from the amine substrate to oxygen.  相似文献   

6.
The aim of this study is to probe the glycerol backbone conformation of the substrate (or inhibitor) in the active site of Pseudomonas species lipase by the 1,2-cyclopentandiol analogues of the ethylene glycerol carbamate inhibitors. Cyclopentane-carbamates, cis-1,2-di-N-n- butylcarbamyl-cyclopentane (1) and trans-1,2-di-N-n-butylcarbamyl-cyclopentane (2), are the conformationally constrained analogues of 1,2-di-N-n-butylcarbamyl ethane (3). All carbamates are synthesized and characterized as the pseudo-substrate inhibitors of the enzyme. Cis-cyclopentane-di-carbamate (1) is a more potent inhibitor than both ethane-di-carbamate (3) and trans-cyclopentane-di-carbamate (2) probably because the glycerol backbone conformations of cis-cyclopentane-di-carbamate (1) are constrained by the cyclopentane ring and cis-cyclopentane-di-cabamate (1) is a meso compound but trans-cyclopentane-di-carbamate (2) is a racemate.  相似文献   

7.
A detailed structure/function analysis of the substrate specificity of Escherichia coli sn-1,2-diacylglycerol kinase was performed with three goals in mind: (a) to define the substrate specificity; (b) to discover inhibitors; and (c) to elucidate the specificity of diacylglycerol-dependent inactivation. Forty-seven structural analogues of sn-1,2-diacylglycerol were prepared and examined as substrates, inhibitors, and irreversible inactivators of the enzyme using mixed micellar assay methods. Modification of the acyl chains or the sn-2 ester affected the apparent Km but had only small effects on Vm; modifications of the sn-1 ester, sn-3 methylene, or sn-3 hydroxyl had large effects on the apparent Vm and smaller effects on Km. Consistent with these observations, diacylglycerol analogues modified only in the acyl chains or sn-2 ester were not diacylglycerol kinase inhibitors, whereas analogues with substitutions of the sn-1 ester or sn-3 hydroxyl frequently caused inhibition. A hydrogen bond-donating group was required for an analogue to be a diacylglycerol kinase inhibitor. Studies of diacylglycerol kinase inactivation by the various analogues were consistent with the previous conclusion that this process involves an interaction of diacylglycerols with an enzyme conformation different from that active in catalysis (Walsh, J. P., and Bell, R. M. (1986) J. Biol. Chem. 261, 15062-15069). Studies with a water-soluble diacylglycerol, sn-1,2-dibutyrylglycerol, allowed direct comparison of diacylglycerol kinase activity in mixed micelles with that in native membranes. The results are discussed in relation to the structural requirements of other diacylglycerol-dependent enzymes.  相似文献   

8.
G Lin  C T Shieh  H C Ho  J Y Chouhwang  W Y Lin  C P Lu 《Biochemistry》1999,38(31):9971-9981
Alkyl-N-phenyl carbamates (2-8) (see Figure 1), alkyl-N-phenyl thiocarbamates (9-15), 2,2'-biphenyl-2-ol-2'-N-substituted carbamates (16-23), and 2, 2'-biphenyl-2-N-octadecylcarbamate-2'-N-substituted carbamates (24-31) are prepared and evaluated for their inhibition effects on porcine pancreatic cholesterol esterase and Pseudomona species lipase. All inhibitors are characterized as transient or pseudo substrate inhibitors for both enzymes. Both enzymes are not protected from inhibition and further inactivated by carbamates 2-8 and thiocarbamates 9-15 in the presence of trifluoroacetophenone. Therefore, carbamates 2-8 and thiocarbamates 9-15 are exceptions for active site binding inhibitors and are probably the second alkyl-chain binding-site-directed inhibitors for both enzymes. The inhibition data for carbamates 2-8 and thiocarbamates 9-15 are correlated with the steric constant, E(s), and the hydrophobicity constant, pi; however, the inhibition data are not correlated with the Taft substituent constant, sigma. A comparison of the inhibition data for carbamates 2-8 and thiocarbamates 9-15 toward both enzymes indicates that thiocarbamates 9-15 are more potent inhibitors than carbamates 2-8. A comparison of the inhibition data for cholesterol esterase and Pseudomona species lipase by carbamates 2-8 or thiocarbamates 9-15 indicates that cholesterol esterase is more sensitive to the E(s) and pi values than Pseudomona species lipase. The negative slope values for the logarithms of inhibition data for Pseudomona species lipase by carbamates 2-8 and thiocarbamates 9-15 versus E(s) and pi indicate that the second alkyl-chain-binding site of Pseudomona species lipase is huge, hydrophilic, compared to that of cholesterol esterase, and prefers to interact with a bulky, hydrophilic inhibitor rather than a small, hydrophobic one. On the contrary, the second alkyl-chain-binding site of cholesterol esterase prefers to bind to a small, hydrophobic inhibitor. Both enzymes are protected from inhibition by carbamates 16-23 in the presence of trifluoroacetophenone. Therefore, carbamates 16-23 are characterized as the alkyl chain binding site, esteratic site oxyanion active site directed pseudo substrate inhibitors for both enzymes. Both enzyme inhibition data for carbamates 16-22 are well-correlated with sigma alone. The negative rho values for these correlations indicate that the serine residue of both enzymes and carbamates 16-22 forms the tetrahedral species with more positive charges than inhibitors and the enzymes and follow the formation of the carbamyl enzymes with more positive charges than the tetrahedral species. Carbamates 24-31 are also exceptions for active site binding inhibitors and probably the second alkyl chain binding site-directed inhibitors for both enzymes. However, the enzyme inhibition constants for carbamates 24-31 are correlated with values of sigma, E(s), and pi. The negative rho values for these correlations indicate that both enzymes and carbamates 24-31 form the tetrahedral species with more positive charges than inhibitors and the enzymes and follow the formation of the carbamyl enzymes with more positive charges than those tetrahedral species. Therefore, carbamates 24-31 may bind to both the active sites and the second alkyl chain binding site and follow the evacuation of the active sites. A comparison of the rho values for cholesterol esterase and Pseudomona species lipase by carbamates 24-31 indicates that cholesterol esterase is much more sensitive to the sigma values than Pseudomona species lipase. The negative sensitivity values, delta, for the cholesterol esterase inhibitions by carbamates 24-31 indicate that the enzyme prefers to bind to a bulky carbamyl group rather than bind to a small one. The hydrophobicity of carbamates 24-31 does not play a major role in both enzyme inhibitions.  相似文献   

9.
The stereospecificity of monoacylglycerol acyltransferase from rat intestinal mucosa and suckling rat liver microsomes was examined using sn-1,2-diacylglycerol kinase from Escherichia coli. With 2-monooleoyl glycerol and palmitoyl-CoA, 88 and 87.9% of the diacylglycerol synthesized by the intestinal mucosa and suckling liver, respectively, was demonstrated to be the sn-1,2-isomer. Analysis of similar preparations of these diacylglycerol products by gas-liquid chromatography-mass spectrometry indicated that most of the remaining diacylglycerol was the 1,3-isomer that probably arose via acyl-migration. These results indicate that monoacylglycerol acyltransferase is stereospecific. Measurement of acyltransferase activities in microsomes using 1- and 2-monoacyl- and monoalkylglycerols as substrates indicated that the monoacylglycerol acyltransferases from suckling liver and intestinal mucosa have different substrate specificities.  相似文献   

10.
Streptococcus mutans BHT metabolizes radioactive 3-dodecyl-sn-glycerol (sn-3-DDG) almost exclusively to lysophosphatidic acid, phosphatidic acid and 1,3-diradyl-sn-glycerol, whereas the cells of this organism metabolize 1-dodecyl-sn-glycerol (sn-1-DDG) to all of the glycerol lipids of S. mutans BHT, with the largest amounts incorporated into phosphatidylglycerol and diradylglycerol (mostly the 1,2- but also the 1,3-isomer). (The common names of lipids, such as phosphatidic acid, are used in the broader sense to mean that the lipid may contain alkyl as well as acyl groups.) The addition of an equivalent amount of nonradioactive sn-3-DDG to radioactive sn-1-DDG causes more of the radioactivity to accumulate at phosphatidic acid. These results indicate that the monoglyceride kinase (EC 2.7.1.94), lysophosphatidic acid acyltransferase (EC 2.3.1.40) and the monoglyceride acyltransferase (EC 2.3.1.22) enzymatic reactions are not stereospecific, and that the CDP-diglyceride synthase (EC 2.7.7.41) and phosphatidic acid phosphatase (EC 3.1.3.4) metabolic steps are stereospecific in S. mutans BHT. The synthesis of phosphatidic acid and lysophosphatidic acid from sn-3-DDG provides a unique method for synthesizing these glycerol lipids with the uncommon stereochemical configuration in which the phosphate moiety is in the sn-1 position.  相似文献   

11.
We have reported that benzene‐1,2‐, 1,3‐, and 1,4‐di‐N‐substituted carbamates ( 1–15 ) are characterized as the conformationally constrained inhibitors of acetylcholinesterase and mimic gauche, eclipsed, and anti‐conformations of acetylcholine, respectively (J Biochem Mol Toxicol 2007;21:348–353). We further report the inhibition of butyrylcholinesterase by these inhibitors. Carbamates 1–15 are also characterized as the pseudosubstrate inhibitors of butyrylcholinesterase as in the acetylcholinesterase catalysis. Benzene‐1,4‐di‐Nn‐hexylcarbamate ( 12 ) and benzene‐1,4‐di‐Nn‐octylcarbamate ( 13 ) are the two most potent inhibitors of butyrylcholinesterase among inhibitors 1–15 . These two para compounds, with the angle of 180° between two C(benzene)? O bonds, mimic the preferable anti C? O/C? N conformers for the choline ethylene backbone of butyrylcholine during the butyrylcholinesterase catalysis. The second n‐hexylcarbamyl or n‐octylcarbamyl moiety of inhibitors 12 and 13 is proposed to bind tightly to the peripheral anionic site of butyrylcholinesterase from molecular modeling. Butyrylcholinesterase prefers para‐carbamates to ortho‐ and meta‐carbamates, whereas acetylcholinesterase prefers para‐ and meta‐carbamates to ortho‐carbamates. This result implies that the anionic site of butyrylcholinesterase is relatively smaller than that of acetylcholinesterase because meta‐carbamates, which may bind to the anionic sites of both enzymes, are not potent inhibitors of butyrylcholinesterase. © 2009 Wiley Periodicals, Inc. J Biochem Mol Toxicol 23:303–308, 2009; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/jbt.20286  相似文献   

12.
beta-1,4-Galactosyltransferase-I (beta4Gal-T1) transfers galactose from UDP-galactose to N-acetylglucosamine (GlcNAc) residues of the branched N-linked oligosaccharide chains of glycoproteins. In an N-linked biantennary oligosaccharide chain, one antenna is attached to the 3-hydroxyl-(1,3-arm), and the other to the 6-hydroxyl-(1,6-arm) group of mannose, which is beta-1,4-linked to an N-linked chitobiose, attached to the aspargine residue of a protein. For a better understanding of the branch specificity of beta4Gal-T1 towards the GlcNAc residues of N-glycans, we have carried out kinetic and crystallographic studies with the wild-type human beta4Gal-T1 (h-beta4Gal-T1) and the mutant Met340His-beta4Gal-T1 (h-M340H-beta4Gal-T1) in complex with a GlcNAc-containing pentasaccharide and several GlcNAc-containing trisaccharides present in N-glycans. The oligosaccharides used were: pentasaccharide GlcNAcbeta1,2-Manalpha1,6 (GlcNAcbeta1,2-Manalpha1,3)Man; the 1,6-arm trisaccharide, GlcNAcbeta1,2-Manalpha1,6-Manbeta-OR (1,2-1,6-arm); the 1,3-arm trisaccharides, GlcNAcbeta1,2-Manalpha1,3-Manbeta-OR (1,2-1,3-arm) and GlcNAcbeta1,4-Manalpha1,3-Manbeta-OR (1,4-1,3-arm); and the trisaccharide GlcNAcbeta1,4-GlcNAcbeta1,4-GlcNAc (chitotriose). With the wild-type h-beta4Gal-T1, the K(m) of 1,2-1,6-arm is approximately tenfold lower than for 1,2-1,3-arm and 1,4-1,3-arm, and 22-fold lower than for chitotriose. Crystal structures of h-M340H-beta4Gal-T1 in complex with the pentasaccharide and various trisaccharides at 1.9-2.0A resolution showed that beta4Gal-T1 is in a closed conformation with the oligosaccharide bound to the enzyme, and the 1,2-1,6-arm trisaccharide makes the maximum number of interactions with the enzyme, which is in concurrence with the lowest K(m) for the trisaccharide. Present studies suggest that beta4Gal-T1 interacts preferentially with the 1,2-1,6-arm trisaccharide rather than with the 1,2-1,3-arm or 1,4-1,3-arm of a bi- or tri-antennary oligosaccharide chain of N-glycan.  相似文献   

13.
4,4'-Biphenyl-di-N-butylcarbamate (1), (S)-1,1'-bi-2-naphthyl-2, 2'-di-N-butylcarbamate (S-2), (S)-1, 1'-bi-2-naphthyl-2-N-butylcarbamate-2'-butyrate (S-3), 2, 2'-biphenyl-di-N-butylcarbamate (4), 2, 2'-biphenyl-2-N-octadecylcarbamate-2'-N-octylcarbamate (5), 2, 2'-biphenyl-2-N-octadecylcarbamate-2'-N-phenylcarbamate (6), 2, 2'-biphenyl-2-N-butylcarbamate-2'-butyrate (7), 2, 2'-biphenyl-2-N-butylcarbamate-2'-ol (8), 2, 2'-biphenyl-2-N-octylcarbamate-2'-ol (9), (R)-1, 1'-bi-2-N-naphthyl-2-butylcarbamate-2'-ol (R-10), and glyceryl-1,2, 3-tri-N-butylcarbamate (11) are prepared and evaluated for their inhibition effects on porcine pancreatic cholesterol esterase. All inhibitors are irreversible inhibitors of the enzyme. Carbamates 1-3 and 7-10 are the first alkyl chain and esteratic binding site-directed irreversible inhibitors due to the fact that the reactivity of the enzyme is protected by the irreversible inhibitor, trifluoroacetophenone in the presence of these carbamates. Carbamate 1 is the least potent inhibitor for the enzyme probably due to the fact that the inhibitor molecule adopts a linear conformation and one of the carbamyl groups of the inhibitor molecule covalently interacts with the first alkyl chain binding site of the enzyme while the other carbamyl group of the inhibitor molecule exposes outside the active site. With near orthogonal conformations at the pivot bond of biaryl groups, one carbamyl group of carbamates S-2, S-3, and R-10 covalently binds to the first alkyl chain binding site of the enzyme while the other carbamyl, butyryl, or hydroxy group can not bind covalently to the second alkyl chain binding site probably due to the orthogonal conformations. Carbamates 4-9 and 11 are very potent inhibitors for the enzyme probably due to the fact that all these molecules freely rotate at the pivot bond of the biphenyl or glyceryl group and therefore can fit well into the bent-shaped first and second alkyl chains binding sites of the enzyme. Although, carbamates 4-6 and 11 are irreversible inhibitors of cholesterol esterase, the enzyme is not protected but further inhibited by trifluoroacetophenone in the presence of these carbamates. Therefore, carbamates 4-6 and 11 covalently bind to the first alkyl chain binding site of the enzyme by one of the carbamyl groups and may also bind to the second alkyl chain binding site of the enzyme by the second carbamyl group. Besides the bent-shaped conformation, the inhibition by carbamate 6 is probably assisted by a favorable pi-pi interaction between Phe 324 at the second alkyl chain binding site of the enzyme and the phenyl group of the inhibitor molecule. For cholesterol esterase, carbamates 8-10 are more potent than carbamates S-2, 4, and 5 probably due to the fact that the inhibitor molecules interact with the second alkyl chain binding site of the enzyme through a hydrogen bond between the phenol hydroxy group of the inhibitor molecules and the His 435 residue in that site.  相似文献   

14.
Diuron, a chlorine-substituted dimethyl herbicide, is widely used in agriculture. Though the degradation of diuron in water has been studied much with experiments, little is known about the detailed degradation mechanism from the molecular level. In this work, the degradation mechanisms for OH-induced reactions of diuron in water phase are investigated at the MPWB1K/6–311+G(3df,2p)//MPWB1K/6–31+G(d,p) level with polarizable continuum model (PCM) calculation. Three reaction types including H-atom abstraction, addition, and substitution are identified. For H-atom abstraction reactions, the calculation results show that the reaction abstracting H atom from the methyl group has the lowest energy barrier; the potential barrier of ortho- H (H1’) abstraction is higher than the meta- H abstraction, and the reason is possibly that part of the potential energy is to overcome the side chain torsion for the H1’ abstraction reaction. For addition pathways, the ortho- site (C (2) atom) is the most favorable site that OH may first attack; the potential barriers for OH additions to the ortho- sites (pathways R7 and R8) and the chloro-substituted para- site (R10) are lower than other sites, indicating the ortho- and para- sites are more favorable to be attacked, matching well with the -NHCO- group as an ortho-para directing group.
Figure
Representative pathways including abstraction, addition and substitution for OH and diuron reactions  相似文献   

15.
16.
Multicomponent high-resolution 1H and 13C NMR analysis has been employed for the purpose of detecting and quantifying a wide range of fatty acids (as triacylglycerols or otherwise) in encapsulated marine cod liver oil supplements. The 1H NMR technique provided quantitative data regarding the docosahexaenoic acid content of these products, which serves as a valuable index of fish oil quality, and a combination of both 1H and 13C spectroscopies permitted the analysis of many further components therein, including sn-1 monoacylglycerols, sn-1,2 and -1,3 diacylglycerol adducts, together with a range of minor components, such as trans-fatty acids, free glycerol and cholesterol, and added vitamins A and E. The identities of each of the above agents were confirmed by the application of two-dimensional 1H-1H spectroscopies. The NMR techniques employed also uniquely permitted determinations of the content of nonacylglycerol forms of highly unsaturated (or other) fatty acids in these products (i.e., ethyl esters), and therefore served as a means of distinguishing "natural" sources of cod liver oils from those subjected to chemical modification to and/or supplementation with synthetic derivatives such as ethyl docosahexaenoate or eicosopentaenoate. The analytical significance and putative health effects of the results acquired are discussed.  相似文献   

17.
The phosphatidylcholine (PC)-preferring phospholipase C (PLC) from Bacillus cereus (PLC(Bc)) hydrolyzes various 1,2-diacyl derivatives of PC at different rates. Substrates with side chains having eight or more carbons are present in micellular form in aqueous media and are processed most rapidly. The catalytic efficiency (k(cat)/K(m)) for the hydrolyses of short-chain PCs at concentrations below their respective critical micelle concentrations also decreases as the side chains become shorter, and this loss of efficiency owes its origin to increases in K(m). In order to ascertain whether the observed increases in K(m) might arise from conformational changes in the glycerol backbone, nuclear magnetic resonance (NMR) experiments were performed in D(2)O to determine the (3)J(HH) and (3)J(CH) coupling constants along the glycerol subunit of 1, 2-dipropanoyl-sn-glycero-3-phosphocholine (K(m)=61 mM), 1, 2-dibutanoyl-sn-glycero-3-phosphocholine (K(m)=21.2 mM) and 1, 2-dihexanoyl-sn-glycero-3-phosphocholine (K(m)=2.4 mM). Using these coupling constants, the fractional populations for each rotamer about the backbone of each of substrate were calculated. Two rotamers, which were approximately equally populated, about the sn-1-sn-2 bond of each substrate were significantly preferred, and in these conformers, the oxygens on the sn-1 and sn-2 carbons of the backbone were synclinal to optimize intramolecular hydrophobic interactions between the acyl side chains. There was greater flexibility about the sn-2-sn-3 bond, and each of the three possible staggered conformations was significantly populated, although there was a slight preference for the rotamer in which the oxygen bearing the phosphate head group was synclinal to the oxygen at the sn-2 carbon and to the sn-1 carbon; in this orientation, the head group is folded back relative to the side chains. These studies demonstrate that there is no significant change in the conformation about the glycerol backbone as a function of side chain length in short-chain phospholipids. Thus, prior organization of the substrate seems an unlikely determinant of the catalytic efficiency of PLC(Bc), and other factors such as hydrophobic interactions or differential solvation/desolvation effects associated with the complexation of the substrate with PLC(Bc) may be involved.  相似文献   

18.
Benzene-1,3-di-N-n-octylcarbamate (1), benzene-1-hydroxyl-3-N-n-octylcarbamate (2), benzene-1,3-di-N-n-ocztylthiocarbamate (3), and benzene-1-hydroxyl-3-N-n-octylthiocarbamate (4) are synthesized from 1,3-benzene-diol and are characterized as the pseudo-substrate inhibitors of acetylcholinesterase, butyrylcholinesterase, cholesterol esterase, lipase, trypsin, and chymotrypsin. For these six enzyme inhibitions by 1-4, the pKi values are linearly correlated with their log ki values - Br?nsted plots. Therefore, 1-4 inhibit these enzymes through a common mechanism. Moreover, both pKi and log ki values for the inhibitions by 1,3, and 4 are linearly correlated with both pKi and log ki values for the inhibitions by 2, respectively. Thus, the pKi values for the inhibitions by 2 are defined as the nucleophilicity constants of these enzymes (nenzyme). The log k2 values for the inhibitions by 1-4 are also linearly correlated with the nenzyme values. Therefore, the nucleophilicity for serine hydrolases and proteases toward 1-4 also applies the Swain-Scott correlations.  相似文献   

19.
The conformation of phosphatidylcholine in liquid-crystalline bilayers was studied with a novel, high-resolution method employing phosphatidylcholine species containing pyrenyl moieties in both acyl chains of variable length. Analysis of the intramolecular pyrene-pyrene collision data obtained for 30 such species in terms of a simple geometrical model showed that the sn-1 acyl chain penetrates, on the average, 0.84 +/- 0.11 methylene units (0.8 A) deeper into the bilayer than the sn-2 chain at 22 degrees C. A similar value was obtained at 37 degrees C. Since the penetration difference of the sn-1 and sn-2 acyl chains is inherently coupled to the conformation of the glycerol moiety, these data mean that the glycerol moiety of phosphatidylcholine is, on the average, only moderately tilted with respect to the bilayer plane in the liquid-crystalline state. This contrasts the perpendicular orientation observed previously for phosphatidylcholine crystals [Pearson, R. H., & Pascher, I. (1979) Nature 281, 499-501]. Importantly, addition of 50 mol % cholesterol, which is known to reduce dramatically the interactions between phosphatidylcholine molecules in bilayers, had only a small effect on the penetration difference of the acyl chains, strongly suggesting that the conformation of phosphatidylcholine in the liquid-crystalline state is determined largely by intramolecular, rather than intermolecular, interactions.  相似文献   

20.
Extensive work has been reported on the conformation in membranes of sn-3-phosphatidylcholines, -ethanolamines, -glycerols and -serines (sn-3-phospholipids), where the headgroup is linked to the third carbon atom in the glycerol backbone. One important feature common to all these naturally occurring phospholipids is that the glycerol moiety is oriented almost perpendicular to the membrane surface, with the sn-1 chain continuing in this direction, whereas the sn-2 chain starts first in a direction parallel to the layer and then bends sharply at the second carbon atom. We present here neutron diffraction results on 1,3-dipalmitoyl-glycero-2-phosphocholine (1,3-DPPC) in which the headgroup is attached to the second carbon atom in the middle of the glycerol part and the two other adjacent carbon atoms are linked to the paraffin chains. Two 1,3-DPPC samples. 2H-labelled at different positions, were studied. One sample was deuterated at the first methylene segment in each fatty acyl chain, and the other at the first segment in one chain and at the second segment in the other chain. The Patterson maps as well as the neutron density profiles show that both fatty acyl chains in 1,3-DPPC have the same conformation introduced by this symmetric chemical bond pattern. It is concluded from this that the C(1)C(3) vector of the glycerol backbone part is oriented parallel to the bilayer surface and the 2H nuclear magnetic resonance signals indicate that both chains have a conformation similar to that observed for the bent chain in sn-3-phospholipids. These findings indirectly confirm the idea that an intramolecular energy minimum, together with the packing geometry of the lipids in the membrane, produce the characteristic conformation around the glycerol backbone as is found for all naturally occurring phospholipids that have been studied so far.  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