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.
Several beta replacement and alpha,beta elimination reactions catalyzed by tryptophanase from Escherichia coli are shown to proceed stereospecifically with retention of configuration. These conversions include synthesis of tryptophan from (2S,3R)- and (2s,3s)-[3(-3H)]serine in the presence of indole, deamination of these serines in D2O to pyruvate and ammonia, and cleavage of (2S,3R)-and (2S,3S)-[3(-3H)]tryptophan in D2O to indole, pyruvate, and ammonia. A coupled reaction with lactate dehydrogenase was used to trap the stereospecifically labeled [3-H,2H,3H]pryuvates as lactate, which was oxidized to acetate for chirality analysis of the methyl group. During deamination of tryptophan there is significant intramolecular transfer of the alpha proton of the amino acid to C-3 of indole. To determine the exposed face of the cofactor.substrate complex on the enzyme surface and to analyze its conformational orientation, sodium boro[3H]hydride was used to reduce tryptophanase-bound alaninepyridoxal phosphate Schiff's base. Degradation of the resulting pyridoxylalanine to (2S)-[2(-3H)]alanine and (4'S)-[4'(-3H)]pyridoxamine demonstrates that reduction occurs from the exposed si face at C-4' of the complex and that the ketimine double bond is trans.  相似文献   

2.
Swamy N  Xu W  Paz N  Hsieh JC  Haussler MR  Maalouf GJ  Mohr SC  Ray R 《Biochemistry》2000,39(40):12162-12171
We have combined molecular modeling and classical structure-function techniques to define the interactions between the ligand-binding domain (LBD) of the vitamin D nuclear receptor (VDR) and its natural ligand, 1alpha,25-dihydroxyvitamin D(3) [1alpha,25-(OH)(2)D(3)]. The affinity analogue 1alpha,25-(OH)(2)D(3)-3-bromoacetate exclusively labeled Cys-288 in the VDR-LBD. Mutation of C288 to glycine abolished this affinity labeling, whereas the VDR-LBD mutants C337G and C369G (other conserved cysteines in the VDR-LBD) were labeled similarly to the wild-type protein. These results revealed that the A-ring 3-OH group docks next to C288 in the binding pocket. We further mutated M284 and W286 (separately creating M284A, M284S, W286A, and W286F) and caused severe loss of ligand binding, indicating the crucial role played by the contiguous segment between M284 and C288. Alignment of the VDR-LBD sequence with the sequences of nuclear receptor LBDs of known 3-D structure positioned M284 and W286 in the presumed beta-hairpin of the molecule, thereby identifying it as the region contacting the A-ring of 1alpha, 25-(OH)(2)D(3). From the multiple sequence alignment, we developed a homologous extension model of the VDR-LBD. The model has a canonical nuclear receptor fold with helices H1-H12 and a single beta hairpin but lacks the long insert (residues 161-221) between H2 and H3. We docked the alpha-conformation of the A-ring into the binding pocket first so as to incorporate the above-noted interacting residues. The model predicts hydrogen bonding contacts between ligand and protein at S237 and D299 as well as at the site of the natural mutation R274L. Mutation of S237 or D299 to alanine largely abolished ligand binding, whereas changing K302, a nonligand-contacting residue, to alanine left binding unaffected. In the "activation" helix 12, the model places V418 closest to the ligand, and, consistent with this prediction, the mutation V418S abolished ligand binding. The studies together have enabled us to identify 1alpha,25-(OH)(2)D(3)-binding motifs in the ligand-binding pocket of VDR.  相似文献   

3.
This study investigates the catalytic and allosteric roles of a flexible loop in the tryptophan synthase alpha 2 beta 2 complex. This loop connects helix 6 and strand 6 in the alpha subunit, an 8-fold alpha/beta barrel polypeptide. We have engineered three mutations in this disordered loop: a deletion of residues 185-187 and the replacement of threonine 183 by serine (T183S) or by alanine (T183A). Position 183 is a site of an inactivating mutation identified by Yanofsky's group (Yanofsky, C., Drapeau, G. R., Guest, J. R., and Carlton, B. C. (1967) Proc. Natl. Acad. Sci. U.S.A. 57, 296-298). The three engineered alpha subunits form stable, stoichiometric alpha 2 beta 2 complexes with the beta subunit which bind alpha and beta subunit ligands. Although changing threonine 183 to serine has little effect on the enzymatic properties, changing threonine 183 to alanine or deleting residues 185-187 results in a 50-fold reduction in the intrinsic activity of the alpha subunit alone and in the alpha site activity of the alpha 2 beta 2 complex. The latter two mutations profoundly alter the way in which the alpha subunit modulates the spectral properties and the activities of the wild-type beta subunit. These mutations also eliminate the effects of alpha subunit ligands on the beta subunit. Although the beta subunit ligand, L-serine, greatly stabilizes the wild-type alpha 2 beta 2 complex to dissociation and to proteolysis, L-serine stabilizes the T183A alpha 2 beta 2 complex weakly or not at all. Our findings suggest that the hydroxyl residue at position 183 and the adjacent residues in the alpha subunit loop play critical roles in the reciprocal communication between the alpha and beta subunits in the alpha 2 beta 2 complex. The results also help to explain how the wild-type alpha subunit or ammonium ion modulates the activities of the beta subunit.  相似文献   

4.
Voltage-gated L-type Ca(2+) channels from cardiac (alpha(1C)) and skeletal (alpha(1S)) muscle differ from one another in ion selectivity and permeation properties, including unitary conductance. In 110 mM Ba(2+), unitary conductance of alpha(1S) is approximately half that of alpha(1C). As a step toward understanding the mechanism of rapid ion flux through these highly selective ion channels, we used chimeras constructed between alpha(1C) and alpha(1S) to identify structural features responsible for the difference in conductance. Combined replacement of the four pore-lining P-loops in alpha(1C) with P-loops from alpha(1S) reduced unitary conductance to a value intermediate between those of the two parent channels. Combined replacement of four larger regions that include sequences flanking the P-loops (S5 and S6 segments along with the P-loop-containing linker between these segments (S5-6)) conferred alpha(1S)-like conductance on alpha(1C). Likewise, substitution of the four S5-6 regions of alpha(1C) into alpha(1S) conferred alpha(1C)-like conductance on alpha(1S). These results indicate that, comparing alpha(1C) with alpha(1S), the differences in structure that are responsible for the difference in ion conduction are housed within the S5-6 regions. Moreover, the pattern of unitary conductance values obtained for chimeras in which a single P-loop or single S5-6 region was replaced suggest a concerted action of pore-lining regions in the control of ion conduction.  相似文献   

5.
To map out the heavy metal binding sites of iduronic acid containing oligosaccharides isolated from human kidneys, we studied Zn(II) binding by nuclear magnetic resonance (NMR) and molecular modeling to two disaccharides isolated after nitrous acid depolymerization of heparin and two synthetic disaccharides representative of the heparin structure, namely, IdopA2S (alpha 1,4)AnManOH, 1 alpha, IdopA2S (alpha 1,4)AnManOH6S, 1b, IdopA2S-(alpha 1,4)GlcNS alpha Me, 2a, and IdopA2S (alpha 1,4)GlcNS6S alpha Me, 2b (see previous article in this series). A conformational analysis of the metal free and metal bound solutions was made by comparing calculated [(NOE)]s, [T1]s, and [J]s to experimental values. The 1C4, 4C1, and 2S0 conformations of the L-idopyranosiduronate ring and the 4E and 4T3 of the anhydro-D-mannitol ring are evaluated as are rotations about the C5-C6 hydroxymethylene of the AnManOH(6S) or GlcNS (6S) residues. The NOE between IdopA2S H1 and H3 and the known NOE between H2 and H5, as well as the T1 of IdopA2S H3, are introduced as NMR observables sensitive to the IdopA2S ring conformation. Similarly, a NOE between IdopA2S H5 and AnManOH(6S) or GlcNS(6S) H3 was observed that directly restricts the allowed interglycosidic conformational space. For all disaccharides, the Zn(II) bound spectral data are consistent with models in which these motions are partially "frozen" such that the 1C4 conformation of the IdopA2S is stabilized along with the 4T3 conformation of the AnManOH(6S) ring. The interglycosidic conformation is also stabilized in one of two minima. Electrostatic potential energy calculations gave the best overall agreement with experiment and suggest metal binding conformations with the carboxylate and ring oxygen of the IdopA2S residues (1C4 conformation) and either O3 of the GlcNS(6S) residues or the sulfate oxygens of the 6-sulphate for 2b providing additional chelating sites. These chelation models concur with the observation of marked 13C and 1H NMR chemical shifts for the IdopA2S resonances and of GlcNS H3 for 2 alpha and GlcNS6S C6 for 2b. This study of model compounds implicates the IdopA2S(alpha 1,4)GlcNS6S group as part of the heavy metal binding site in biologically important acidic oligosaccharides such as heparin.  相似文献   

6.
Formation of bile acids from sitosterol in bile-fistulated female Wistar rats was studied with use of 4-14C-labeled sitosterol and sitosterol labeled with 3H in specific positions. The major part (about 75%) of the 14C radioactivity recovered as bile acids in bile after intravenous administration of [4-14C]sitosterol was found to be considerably more polar than cholic acid, and only trace amounts of radioactivity had chromatographic properties similar to those of cholic acid and chenodeoxycholic acid. It was shown that polar metabolites were formed by intermediate oxidation of the 3 beta-hydroxyl group (loss of 3H from 3 alpha-3H-labeled sitosterol) and that the most polar fraction did not contain a hydroxyl group at C7 (retention of 3H in 7 alpha,7 beta-3H2-labeled sitosterol). Furthermore, the polar metabolites had lost at least the terminal 6 or 7 carbon atoms of the side chain (loss of 3H from 22,23-3H2- and 24,28-3H2-labeled sitosterol). Experiments with 3H-labeled 7 alpha-hydroxysitosterol and 4-14C-labeled 26-hydroxysitosterol showed that none of these compounds was an efficient precursor to the polar metabolites. By analysis of purified most polar products of [4-14C] sitosterol by radio-gas chromatography and the same products of 7 alpha,7 beta-[2H2]sitosterol by combined gas chromatography-mass spectrometry, two major metabolites could be identified as C21 bile acids. One metabolite had three hydroxyl groups (3 alpha, 15, and unknown), and one had two hydroxyl groups (3 alpha, 15) and one keto group. Considerably less C21 bile acids were formed from [4-14C]sitosterol in male than in female Wistar rats. The C21 bile acids formed in male rats did not contain a 15-hydroxyl group. Conversion of a [4-14C]sitosterol into C21 bile acids did also occur in adrenalectomized and ovariectomized rats, indicating that endocrine tissues are not involved. Experiments with isolated perfused liver gave direct evidence that the overall conversion of sitosterol into C21 bile acids occurs in this organ. Intravenously injected 7 alpha,7 beta-3H-labeled campesterol gave a product pattern identical to that of 4-14C-labeled sitosterol. Possible mechanisms for hepatic conversion of sitosterol and campesterol into C21 bile acids are discussed.  相似文献   

7.
By using the large cytoplasmic domain of the nicotinic acetylcholine receptor (AChR) alpha4 subunit as a bait in the yeast two-hybrid system, we isolated the first cytosolic protein, 14-3-3eta, known to interact directly with neuronal AChRs. 14-3-3eta is a member of a family of proteins that function as regulatory or chaperone/ scaffolding/adaptor proteins. 14-3-3eta interacted with the recombinant alpha4 subunit alone in tsA 201 cells following activation of cAMP-dependent protein kinase by forskolin. The interaction of 14-3-3eta with recombinant alpha4 subunits was abolished when serine 441 of the alpha4 subunit was mutated to alanine (alpha4(S441A)). The surface levels of recombinant wild-type alpha4beta2 AChRs were approximately 2-fold higher than those of mutant alpha4(S441A)beta2 AChRs. The interaction significantly increased the steady state levels of the alpha4 subunit and alpha4beta2 AChRs but not that of the mutant alpha4(S441A) subunit or mutant alpha4(S441A)beta2 AChRs. The EC50 values for activation by acetylcholine were not significantly different for alpha4beta2 AChRs and alpha4(S441A)beta2 AChRs coexpressed with 14-3-3eta in oocytes following treatment with forskolin. 14-3-3 coimmunopurified with native alpha4 AChRs from brain. These results support a role for 14-3-3 in dynamically regulating the expression levels of alpha4beta2 AChRs through its interaction with the alpha4 subunit.  相似文献   

8.
Alpha4 and beta2 nicotinic acetylcholine (nACh) receptor subunits expressed heterologously in Xenopus oocytes assemble into a mixture of receptors with high and low agonist sensitivity whose relative abundance is influenced by the heteropentamer subunit ratio. We have found that inhibition of protein kinase A by KT5720 decreased maximal [3H]cytisine binding and acetylcholine (ACh)-induced current responses, and increased the relative proportion of alpha4beta2 receptors with high agonist sensitivity. Mutation of serine 467, a putative protein kinase A substrate in a chaperone protein binding motif within the large cytoplasmic domain of the alpha4 subunit, to alanine or asparate decreased or increased, respectively, maximal [3H]cytisine binding and ACh response amplitude. Expression of alpha4S467A mutant subunits decreased steady levels of alpha4 and the relative proportion of alpha4beta2 receptors with low agonist sensitivity, whilst expression of alpha4S467D increased steady levels of alpha4 and alpha4beta2 receptors with low agonist sensitivity. Difopein, an inhibitor of chaperone 14-3-3 proteins, decreased [3H]cytisine binding and ACh responses and increased the proportion of alpha4beta2 with high sensitivity to activation by ACh. Thus, post-translational modification affecting steady-state levels of alpha4 subunits provides a possible means for physiologically relevant, chaperone-mediated variation in the relative proportion of high and low agonist sensitivity alpha4beta2 nACh receptors.  相似文献   

9.
Abstract: To study the regulation of the synthesis of quinolinic and kynurenic acids in vivo, we evaluated (a) the metabolism of administered kynurenine by measuring the content of its main metabolites 3-hydroxykynurenine, anthranilic acid, and 3-hydroxyanthranilic acid in blood and brain of mice; (b) the effects of ( m -nitrobenzoyl)alanine, a selective inhibitor of kynurenine hydroxylase and of ( o -methoxybenzoyl)alanine, a selective inhibitor of kynureninase, on this metabolism; and (c) the effects of ( o -methoxybenzoyl)alanine on liver kynureninase and 3-hydroxykynureninase activity. The conclusions drawn from these experiments are (a) the disposition of administered kynurenine preferentially occurs through hydroxylation in brain and through hydrolysis in peripheral tissues; (b) ( m -nitrobenzoyl)alanine, the inhibitor of kynurenine hydroxylase, causes the expected changes in brain kynurenine metabolism, such as a decrease of 3-hydroxykynurenine, and an increase of kynurenic acid; and (c) ( o -methoxybenzoyl)alanine, the kynureninase inhibitor, increases brain concentration of the cytotoxic compound 3-hydroxykynurenine, and unexpectedly does not reduce brain concentration of 3-hydroxyanthranilic acid, the direct precursor of quinolinic acid. Taken together, the experiments suggest that the systemic administration of a kynurenine hydroxylase inhibitor is a rational approach to increase the brain content of kynurenate and to decrease that of cytotoxic kynurenine metabolites, such as 3-hydroxykynurenine and quinolinic acid.  相似文献   

10.
Malonyl-CoA decarboxylase from the uropygial gland of goose decarboxylated (R,S)-methylmalonyl-CoA at a slow rate and introduced 3H from [3H]2O into the resulting propionyl-CoA. Carboxylation of this labeled propionyl-CoA by propionyl-CoA carboxylase from pig heart and acetyl-CoA carboxylase from the uropygial gland completely removed 3H. Repeated treatment of (R,S)-[methyl-14C]methylmalonyl-CoA with the decarboxylase converted 50% of the substrate into propionyl-CoA, whereas (S)-methylmalonyl-CoA, generated by both carboxylases, was completely decarboxylated. Radioactive (R)- (S), and (R,S)-methylmalonyl-CoA were equally incorporated into fatty acids by fatty acid synthetase from the uropygial gland. The residual methylmalonyl-CoA remaining after fatty acid synthetase reaction on (R,S)-methylmalonyl-CoA was also racemic. These results show that: (a) the decarboxylase is stereospecific, (b) replacement of the carboxyl group by hydrogen occurs with retention of configuration, (c) acetyl-CoA carboxylase of the uropygial gland generates (S)-methylmalonyl-CoA from propionyl-CoA, and (d) fatty acid synthetase is not stereospecific for methylmalonyl-CoA.  相似文献   

11.
A novel structural analogue of kynurenine, 2-amino-4-[3'-hydroxyphenyl]-4-hydroxybutanoic acid 6, was synthesised as an inhibitor of kynureninase. The compound had a significant inhibitory effect on kynureninase from both rat and human, giving a K(i) of 100 nM. It was thus found that removal of the aryl amino group coupled with a reduction of the carbonyl group at position 7 of the alanine side chain greatly enhanced potency of the inhibitor.  相似文献   

12.
B Badet  C Walsh 《Biochemistry》1985,24(6):1333-1341
An alanine racemase has been purified some 30 000-fold almost to homogeneity from Gram-positive Streptococcus faecalis NCIB 6459; the enzyme has been purified to the same extent (4000-fold) from an O-carbamyl-D-serine-resistant mutant with a 7-fold higher enzyme level in crude extract. The racemase has one pyridoxal phosphate molecule per 42-kDa subunit, has a Vmax of 3570 units/mg and a Km of 7.8 mM in the L to D direction, and has a Vmax of 1210 units/mg and a Km of 2.2 mM in the D to L direction. The Keq is 0.8 and kcat/Km values are ca. 3 X 10(5) M-1 s-1. The purified enzyme is inhibited in a time-dependent manner by both L- and D-(l-aminoethyl)phosphonates (Ala-P), confirming observations of Atherton et al. in crude extracts of this organism [Atherton, F. R., Hall, M. J., Hassal, C. H., Holmes, S. W., Lambert, R. W., Lloyd, W. J., & Ringrose, P. S. (1980) Antimicrob. Agents Chemother. 18, 897]. Studies with [1-2H]-, [1-3H]-, and [1,2-14C]Ala-P rule out enzymic activation and processing as the basis for irreversible inhibition. Thus, enzyme after exposure to [14C]Ala-P or [alpha-3H]Ala-P and gel filtration contains stoichiometric amounts of radioactive label, but denaturation quantitatively releases intact Ala-P into solution as revealed by high-performance liquid chromatography and cocrystallization with authentic material. The Ala-P isomers are slow binding inhibitors of this racemase as is the alpha,alpha'-dimethyl analogue but not the D or L isomers of the corresponding phosphinate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
1. (3RS,6R)-[6-2H1,6-3H1,6-14C], (3RS,6S)-[6-2H1,6-3H1,6-14C] and (3RS)-[6-3H1,6-14C]mevalonolactones were synthesised from R-[2H1,3H1,2-14C], S-[2H1,3H1,2-14C] and [3h1,2-14C]acetic acids respectively. 2. Each mevalonate was converted into cholesterol by a rat liver preparation. 3. Each cholesterol specimen was converted into androsta-1,4-diene-3,17-dione by incubation with Mycobacterium phlei in the presence of 2,2'.dipyridyl. Each specimen of androsta-1,4-diene-3,17-dione was converted into androsta-1,4-dien-3-one-17-ethylene ketail. 4. The samples of androsta-1,4-dien-3-one-17-ethylene ketal were each converted chemically into oestrones in which the methyl group at C-18 is the only carbon atom that originated from C-6 in mevalonolactone. 5. The oestrone from (3RS)-[6-3H1,6-14C]mevalonolactone was oxidised chemically to acetic acid which was converted into p-bromophenacyl acetate and the 3H/14C ratio was measured. 6. There was no overall loss of tritium from the methyl group of acetic acid, as measured by determining the 3H/14C ratios of the p-bromophenacyl esters, when the synthetic and degradative procedures 1 -- 5 were tested with [3H1,2-14C]acetic acid. 7. The oestrones derived from the 6R and 6S-mevalonolactones were oxidised. The chiralities of the resulting acetates were determined by an established procedure whereby the acetates were converted into 2S-malates which were examined for loss of tritium on equilibration with fumarate hydratase. 8. The oestrone from (3RS,6R)-[6-2H1,6-3H1,6-14C]mevalonate gave acetic acid which was converted into 2S-malate that retained 68.6% of its tritium after treatment with fumarate hydratase; the configuration of this acetic acid was R. 9. The oestrone from (3RS,6S)-E16-2H1,6-3H1,6-14C]mevalonate was oxidised to acetic acid which was converted into 2S-malate that retained 31.9% of its tritium after treatment with fumarate hydratase; the configuration of this acetic acid was S. 10. There was no overall change in the configuration of a chiral methyl group between C-6 of mevalonate and C-18 of oestrone. It is cncluded that the intramolecular migration of a chiral methyl group from C-15 in 2,3-oxidosqualene to C-13 in lanosterol is stereospecific and occurs with overall retention of configuration.  相似文献   

14.
The stereochemistry of the hydrogen transfer to NAD catalyzed by (S)alanine dehydrogenase [ (S)alanine: NAD oxidoreductase (EC 1.4.1.1) ] from B. subtilis was investigated. The label at C-2 of (S) [2,3--3H] alanine was enzymatically transferred to NAD, and the [4--3H]NADH produced isolated and the stereochemistry at C-4 investigated. It was found that the label was exclusively located at the (R) position which indicates that (S)alanine dehydrogenase is an A-type enzyme. This result was confirmed in an alternate way by reducing enzymatically [4--3H]NAD with non labeled (S)alanine and (S)alanine dehydrogenase and investigating the stereochemistry of the ]4--3H]NADH produced. As expected, the label was now exclusively located at the (S) position. This proves that (S)alanine dehydrogenase isolated from B. subtilis should be classified as an A-enzyme with regard to the stereochemistry of the hydrogen transfer to NAD.  相似文献   

15.
Analysis of O2 binding equilibria by two independent groups has suggested that the affinity for binding the fourth O2 to Hb tetramers is very high, about 800-1200 cal/mol higher than that of dimers (Chu, A. H., Turner, B. W., and Ackers, G. K. (1984) Biochemistry 23, 604-167; Di Cera, E., Robert, C. H., and Gill, S. J. (1987) Biochemistry 26, 4003-4008). Recently, Gibson and Edelstein challenged the reality of the quaternary enhancement effect, based on kinetic data (Gibson, Q. H., and Edelstein, S. J. (1987) J. Biol. Chem. 262, 516-519). However, these studies failed to directly address the key issue of the relative affinities of dimers and alpha 2 beta 2(O2)3. Furthermore, the extent to which alpha/beta differences influence these results remains an open question. Using partial laser photolysis and O2/CO replacement techniques we have, for the first time, resolved the rates of O2 association and dissociation to both alpha and beta chains within "R state" tetramers and dimers. We find that the beta chains are faster than alpha for both O2 binding (approximately 2-fold) and release (approximately 3-fold). The kinetically determined O2 affinities derived from these data are essentially identical for dimers and alpha 2 beta 2(O2)3. That is, the data do not show significant quaternary enhancement and suggest that the equilibrium data have both overestimated the affinity of alpha 2 beta 2(O2)3 and underestimated the affinity of dimers. The significance of and possible origins for the discrepancy between equilibrium and kinetic data are discussed.  相似文献   

16.
Systematic studies with purified alanine racemase and a number of substrate analogs permit the generalization that effective competitive inhibition is limited to 2- and 3-carbon compounds. A free α-amino group was not necessary for relatively tight binding; compounds lacking an amino group, or with an α-amino group acylated even by a bulky substituent, were bound as tightly as alanine. Substitution at the α-carbon of alanine (i.e., replacement of the α-H) eliminated binding, while substitution at the β-carbon generally reduced binding. Of several inhibitory compounds tested for substrate activity by H exchange with 3H2O, only glycine appeared active. Covalent binding to the enzyme by halo analogs was not demonstrated.  相似文献   

17.
A series of blood group H antigens reacting with monoclonal antibody MBrl has been found in human blood group A and AB erythrocytes, but not in O or B erythrocytes. These H antigens are clearly different from the globo-H structure (Fuc alpha 1----2Gal beta 1----3GalNAc beta 1----3Gal alpha 1----4Gal beta 1----4Glc beta 1----1Cer), which was previously isolated from O erythrocytes and is also reactive with the MBrl antibody. The new series of H antigens associated with blood group A has been characterized as having TLC mobilities which approximately coincide with those of H2, H3, and H4 glycolipids. One of these A-associated H antigens, having a similar TLC mobility as the H2 glycolipid, was isolated from A erythrocytes and was characterized by 1H NMR spectroscopy, methylation analysis, and enzymatic degradation as having the structure shown below: (formula, see text). The structure represents a precursor of the repetitive A epitope attached to type 2 chain, previously called type 3 chain A (Clausen, H., Levery, S. B., Nudelman, E., Tsuchiya, S., and Hakomori, S. (1985) Proc. Natl. Acad. Sci. U.S.A. 82, 1199-1203). This A-associated H structure is hereby called type 3 chain H.  相似文献   

18.
Dissociation of alpha beta and alpha gamma dimers of human hemoglobins (Hb) A and F into monomers was studied by alpha chain exchange (Shaeffer, J. R., McDonald, M. J., Turci, S. M., Dinda, D. M., and Bunn, H. F. (1984) J. Biol. Chem. 259, 14544-14547). Unlabeled carbonmonoxy-Hb A was incubated with trace amounts of preparatively purified, native, 3H-alpha subunits in 10 mM sodium phosphate, pH 7.0, at 25 degrees C. At appropriate times, free alpha monomers were separated from Hb A tetramers by anion exchange high performance liquid chromatography. Transfer of radioactivity from the alpha chain pool into Hb A was measured, yielding a first order dimer dissociation rate constant, k2 = (3.2 +/- 0.3) X 10(-3) h-1. The Arrhenius plot of k2 was linear between 7 and 37 degrees C, yielding an enthalpy of activation of 23 kcal/alpha beta dimer. As the chloride concentration was raised from 0 to 0.2 M, the dissociation rate increased 3-fold; with higher salt concentrations, however, the rate gradually returned to baseline. This rate was not altered by raising the pH from 6.5 to 7.2, but as pH was further raised to 8.4, kappa 2 increased about 3-fold. Hb F, which has an increased stability at alkaline pH, dissociated into alpha and gamma monomers 3 times more slowly than Hb A. Moreover, the dimer-monomer dissociation of Hb F was characterized by a significantly reduced pH dependence. These results demonstrate that both alpha beta and alpha gamma dimers of Hb A and Hb F dissociate reversibly into monomers under physiologic conditions. The differential pH dependence for dimer dissociation between Hb A and Hb F suggests that specific amino acid replacement at the alpha 1 gamma 1 interface confers increased resistance to alkaline denaturation.  相似文献   

19.
Mutatu W  Klettke KL  Foster C  Walker KD 《Biochemistry》2007,46(34):9785-9794
The phenylalanine aminomutase from Taxus catalyzes the vicinal exchange of the amino group and the pro-3S hydrogen of (2S)-alpha-phenylalanine to make (3R)-beta-phenylalanine. While the migration of the amino group from C2 of the substrate to C3 of the product is already known to proceed intramolecularly with retention of configuration, the stereochemistry of the hydrogen transfer remained unknown, until now. The chemical shifts of the prochiral hydrogens of authentic (3R)-beta-phenylalanine were established by 1H NMR, and the configuration of each hydrogen was assigned by 2H NMR analysis of a racemic mixture of [2,3-2H2]-(2S,3R)- and (2R,3S)-beta-phenylalanines synthesized via syn addition of deuterium gas with palladium catalyst to stereospecifically reduce the double bond of an N-acetyl enamine. After the aminomutase was incubated with [3,3-2H2]-(2S)-alpha-phenylalanine, the derived deuterium-labeled beta-diastereoisomer product, derivatized as the N-acetyl methyl ester, was analyzed by 2H NMR, which revealed that the mutase shuttles the pro-3S hydrogen to C2 of the beta-isomer product (designated 2S,3R) with retention of configuration. Retention of configuration at both reaction termini is unique among all aminomutase mechanisms examined so far. Furthermore, the dynamics of the Cbeta-H bond of the substrate were measured in a competitive experiment with deuterium-labeled substrate to calculate a primary kinetic isotope effect on Vmax/KM of 2.0 +/- 0.2, indicating that C-H bond cleavage is likely rate limiting. Isotope exchange data indicate that the migratory deuterium of [2H8]-(2S)-alpha-phenylalanine, at saturation, dynamically exchanges up to 75%, with protons from the solvent during the reaction after the first 10% of product is formed. The calculated equilibrium constant of 1.1 indicates that the beta-isomer was slightly favored relative to the alpha-isomer at 30 degrees C.  相似文献   

20.
The specific volumes of six 1,2-diacylphosphatidylcholines with monounsaturated acyl chains (diCn:1PC, n=14-24 is the even number of acyl chain carbons) in fluid bilayers in multilamellar vesicles dispersed in H(2)O were determined by the vibrating tube densitometry as a function of temperature. From the data obtained with diCn:1PC (n=14-22) vesicles in combination with the densitometric data from Tristram-Nagle et al. [Tristram-Nagle, S., Petrache, H.I., Nagle, J.F., 1998. Structure and interactions of fully hydrated dioleoylphosphatidylcholine bilayers. Biophys. J. 75, 917-925.] and Koenig and Gawrisch [Koenig, B.W., Gawrisch, K., 2005. Specific volumes of unsaturated phosphatidylcholines in the liquid crystalline lamellar phase. Biochim. Biophys. Acta 1715, 65-70.], the component volumes of phosphatidylcholines in fully hydrated fluid bilayers at 30 degrees C were obtained. The volume of the acyl chain CH and CH(2) group is V(CH)=22.30 A(3) and V(CH2) =A(3), respectively. The volume of the headgroup including the glyceryl and acyl carbonyls, V(H), and the ratio of acyl chain methyl and methylene group volumes, r=V(CH3):V(CH2) are linearly interdependent: V(H)=a-br, where a=434.41 A(3) and b=-55.36 A(3) at 30 degrees C. From the temperature dependencies of component volumes, their isobaric thermal expansivities (alpha(X)=V(X)(-1)(partial differential V(X)/ partial differential T) where X=CH(2), CH, or H were calculated: alpha(CH2)=118.4x10(-5)K(-1), alpha(CH)=71.0x10(-5)K(-1), alpha(H)=7.9x10(-5)K(-1) (for r=2) and alpha(H)=9.6x10(-5)K(-1) (for r=1.9). The specific volume of diC24:1PC changes at the main gel-fluid phase transition temperature, t(m)=26.7 degrees C, by 0.0621 ml/g, its specific volume is 0.9561 and 1.02634 ml/g at 20 and 30 degrees C, respectively, and its isobaric thermal expansivity alpha=68.7x10(-5) and 109.2x10(-5)K(-1) below and above t(m), respectively. The component volumes and thermal expansivities obtained can be used for the interpretation of X-ray and neutron scattering and diffraction experiments and for the guiding and testing molecular dynamics simulations of phosphatidylcholine bilayers in the fluid state.  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