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 109 毫秒
1.
Author index     
Photoaffinity labeling techniques have recently demonstrated that mammalian β1- and β2-adrenergic receptors reside on peptides of Mr 62 000–64 000. These receptor peptides are susceptible to endogenous metalloproteinases which produce peptides of Mr 30 000–55 000. Several proteinase inhibitors markedly attenuate this process, specifically EDTA and EGTA. In this study we investigated the functional significance of this proteolysis (and its inhibition) in the β2-adrenergic receptor-adenylate cyclase system derived from rat lung membranes. Membrane preparations containing proteolytically derived fragments of the receptor of Mr 40000–55 000 are fully functional with respect to their ability to bind β-adrenergic antagonist radioligands such as [3H]dihydroalprenolol and β-adrenergic antagonist photoaffinity reagents such as p-azido-m-[125I]iodobenzylcarazolol. They retain the ability to form a high-affinity, agonist-promoted, guanine nucleotide-sensitive complex thought to represent a ternary complex of agonist, receptor and guanine nucleotide regulatory protein. Nonetheless, after proteolysis, GTP is less able to revert this high-affinity receptor complex to one of lower affinity, and all aspects of adenylate cyclase stimulation are reduced. In addition, the functional integrity of the N protein in membranes prepared without proteinase inhibitors is reduced as assessed by reconstitution studies with the cyc[su? variant of S49 lymphoma cell membranes. These results suggest that endogenous proteolysis does not directly impair the ability of β-adrenergic receptors to either bind ligands or interact with the guanine nucleotide regulatory protein. However, they imply that endogenous proteolysis likely impairs the functionality of other components of the adenylate cyclase system, such as the nucleotide regulatory protein.  相似文献   

2.
The catalytic component of adenylate cyclase and [3H]dopamine binding protein were solubilized with 2% Lubrol PX in the presence of NaF from the synaptic membranes of canine caudate nucleus and were separated into distinct fractions by gel exclusion chromatography on a Sephadex G-200 column. The dissociated adenylate cyclase was no longer responsive to dopamine but was considerably stimulated by 10 mm NaF. Dissociated [3H]-dopamine binding protein possessed the apparent dissociation constant of 3.2 μm for dopamine, almost identical to that of the particulate preparations. The affinities of [3H]-dopamine binding protein to catecholamines and neuroleptics were also very similar to those of particulate preparations. After the adenylate cyclase and [3H]dopamine binding protein were preincubated together at 4 °C for 30 min, the cyclase activity displayed a dose-dependent increase by dopamine with the Ka of 1.6 μm, the concentration of dopamine to stimulate half-maximally. Stimulation of the reconstituted adenylate cyclase by dopamine was maximally 2.7-fold and was strongly inhibited by neuroleptics such as chlorpromazine and haloperidol. These results suggest that [3H]dopamine binding protein is identical to the regulatory subunit of dopamine-sensitive adenylate cyclase in the synaptic membranes of canine caudate nucleus.  相似文献   

3.
The binding characteristics of the β-adrenergic agonist (±)-[3H]hydroxybenzylisoproterenol to rat adipocyte membranes were studied. Binding was rapid, reaching equilibrium within 10 min at 37°C (second order rate constant k1=1.37·107·M?1·min?1). Dissociation of specific binding by 0.5 mM (?)-isoproterenol suggested dissociation from two different sites with respective dissociation rate constants k2 of 0.106·min?1 and 0.011·min?1.[3H]Hydroxybenzylisoproterenol binding was saturable (Bmax=690±107 fmol/mg protein), yielding curvilinear Scatchard plots. Computer modeling of these data were consistent with the existence of two classes of [3H]hydroxybenzylisoproterenol binding sites, one having high affinity (KD=3.5±0.7 nM) but low binding capacity (10% of the total sites) and one haveing low affinity (KD=101±20 nM) but high binding capacity (90% of the sites). Adrenergic ligands competed with [3H]hydroxybenzylisoproterenol binding with the following order of potency=(?)-propranolol>(?)-isoproterenol>(?)-norepinephrine≈ (?)-epinephrine>>(+)-isoproterenol=(+)-propranolo, which is consistent with binding to β1-adrenergic receptors. Competition curves of [3H]hydroxybenzylisoproterenol binding by the β-agonist (?)-isoproterenol were shallow and modeled to two affinity states of binding, whereas, competition curves by β-antagonist (?)-propranolol were steeper with Hill number near to one. Gpp[NH]p severely reduced [3H]hydroxybenzyl-isoproterenol binding, an effect which apparently resulted from the reduction of the number of both the high and low affinity sites. In membranes which had been previously exposed to (?)-isoproterenol, then number of [3H]hydroxybenzylisoproterenol binding sites was reduced by 50%, an effect which apparently resulted from the loss of part of both the high and low affinity state binding sites. Finally, the ability of (?)-isoproterenol to stimulate adenylate cyclase correlate closely with the ability of (?)-isoproterenol to displace [3H]hydroxybenzylisoproterenol binding. Comparison of these findings with the binding characteristics of the β-antagonist [3H]dihydroalprenolol to rat adipocyte membranes, led to conclude that [3H]hydroxybenzylisoproterenol can be successfully used to label the β-adrenergic receptors of rat fat cells and suggests that it might be a better ligand than [3H]dihydroalprenolol in these cells.  相似文献   

4.
The binding of (±)-[3H]carazolol, a recently developed β-adrenergic antagonist of high potency, to rat cerebral cortical membranes is compared to the binding of (?)-[3H]dihydroalprenolol (3H-DHA). 3H-Carazolol binds saturably to cortical β-receptors with a KD of 0.15 nM, a value approximately four times lower than that for 3H-DHA. Considering that 3H-carazolol was used as the racemic mixture and 3H-DHA as the (?)-isomer, an equivalent formulation of 3H-carazolol would be 8–10 times more potent than 3H-DHA. This increased affinity can be explained by the observed two fold greater association rate constant and a two fold lower dissociation rate constant. The drug displacement profile of 3H-carazolol binding is very similar to that of 3H-DHA. 3H-Carazolol has equal displacements constants when binding is performed in calf cerebral cortex (which contains mainly β1 receptors) and calf cerebellum (which contains mainly β2 receptors), indicating that 3H-carazolol binds with equal affinity to β1 and β2 receptors. The percent free drug (i.e. unbound to serum proteins) for both carazolol and propranolol in rabbit serum is approximately 10%. At physiologically equivalent doses of carazolol and propranolol in the rabbit, there is no detectable free β-blocking activity at 15, 30 or 60 min after intravenous injection of carazolol, although substantial propranolol activity is detected.  相似文献   

5.
The effect of subchronic infusion of desipramine, a norepinephrine uptake inhibitor, and clenbuterol, a beta-adrenergic agonist, on the central beta receptor of the rat was determined using in vitro [3H]dihydroalprenolol binding. Desipramine produced significant decreases of the receptor in neocortex and hippocampal formation, and clenbuterol effected such decreases in corpus striatum and cerebellum. Both drugs caused a marked decrease in the activity of isoproterenol-sensitive adenylate cyclase in neocortex. The alpha2 receptor of neocortex and cerebellum was unchanged by either drug as assessed by in vitro[3H]p-aminoclonidine binding. The results are discussed in terms of the different mechanisms of action of desipramine and clenbuterol, and the efficacy of these two drugs in the treatment of depression.  相似文献   

6.
The binding characteristics of the β-adrenergic antagonist, [3H]dihydroalprenolol, to hamster white adipocyte membranes were studied. This binding occurred at two classes of sites, one having high affinity (Kd = 1.6±1.3 nM) but low capacity (32±17 fmol/mg membrane protein) and one having low affinity but high binding capacity. While the binding at the high-affinity sites was competitively and stereoselectively displaced by both β-antagonists and β-agonists, competition at the low-affinity sites occurred only with β-antagonists and was non-stereoselective. Thus, the β-agonist (?)-isoproterenol was further used to define nonspecific binding. Under these conditions, saturation studies showed a single class of high-affinity (Kd = 1.6±0.5 nM) binding sites with a binding capacity of 53 ± 13 fmol/mg membrane protein (corresponding to 4000 ± 980 sites per cell), and independent kinetic analysis provided a Kd value of 1.9 nM. Competition experiments showed that these binding sites had the characteristics of a β1-receptor subtype, yielding Kd values in good agreement with the Kact and the Ki values found for agonist-stimulation and for antagonist-inhibition of adenylate cyclase in membranes and of cyclic AMP accumulation and lipolysis in intact cells. Furthermore, the ability of β-agonists to compete with this binding was severely depressed by p[NH]ppG. These results thus support the contention that the specific [3H]dihydroalprenolol binding sites defined as the binding displaceable by (?)-isoproterenol represent the physiologically relevant β-adrenergic receptors of hamster white adipocytes. Finally, studies of the lipolytic response of these cells to (?)-norepinephrine showed that the inhibitory effect of the α2-component of this catecholamine was apparent only when the effects of endogenous adenosine were suppressed, a result which argues against an important regulatory role for the α2-receptors in the adrenergic control of lipolysis in hamster white adipocytes.  相似文献   

7.
[3H]Yohimbine, a potent α2-adrenergic antagonist, was used to label the α2-adrenergic receptors in membranes isolated from human platelets. Binding of [3H]yohimbine to platelet membranes appears to have all the characteristics of binding to α2-adrenergic receptors. Binding reached a steady state in 2–3 min at 37°C and was completely reversible upon the addition of excess phentolamine or yohimbine (both at 10?5 M;t12 = 2.37 min). [3H]Yohimbine bound to a single class of noncooperative sites with a dissociation constant of 1.74 nM. At saturation, the total number of binding sites was calculated to be 191 fmol/mg protein. [3H]Yohimbine binding was stereo-specifically inhibited by epinephrine: the (?) isomer was 11-times more potent than the (+) isomer. Cathecholamine agonists competed for the occupancy of the [3H]yohimbine-binding sites with an order of potency: clonidine > (?)-epinephrine > (?)-norepinephrine >> (?)-isoproterenol. The potent α2-adrenergic antagonist, phentolamine, competed for the sites whereas the β-antagonist, (±)-propanolol, was a very weak inhibitor. 0.1 mM GTP reduced the bindng affinity of the agonists, while producing no change in antagonist-binding affinity. Dopamine and serotonine competed only at very high concentrations. Similarly, muscarinic cholinergic ligands were also poor inhibitors of [3H]yohimbine binding. These results suggest tht [3H]yohimbine binding to human platelet membranes is specific, rapid, saturable, reversible and, therefore, can be successfully used to label α2-adrenergic receptors.  相似文献   

8.
In vitro, the accumulation and release of [methyl-3H]thymidine ([3H]thymidine) by the isolated choroid plexus, the anatomical locus of the blood-cerebrospinal fluid barrier, was studied. With concentrations of [3H]thymidine in the medium of 1.0 μm (or greater), the choroid plexus accumulated [3H]thymidine against a concentration gradient by a process that depended on intracellular energy production but did not depend on intracellular binding or metabolism of the [3H]thymidine. This transport process was inhibited (although differentially) by various nucleosides and low temperatures but not by 2-deoxyribose or pyrimidine bases. With concentrations of less than 1.0 μm [3H]thymidine in the medium, the choroid plexus accumulated [3H]thymidine against a concentration gradient. However, the majority of the [3H]thymidine within the choroid plexus was metabolized to [3H]thymidine nucleotides at low extracellular [3H]thymidine concentrations (3 nm). This accumulation process depended, in large part, on saturable intracellular phosphorylation. Thymidine was the principal form released from choroid plexuses that had been incubated for various times in media containing concentrations of thymidine from 3 to 1.0 mm. The release of thymidine from choroid plexus was depressed by cold temperatures and a very high (2.56 mmol/kg) intracellular thymidine concentration.  相似文献   

9.
HeLa cells contain receptors on their surface which are β-adrenergic in nature. The binding of (?)-[3H]dihydroalprenolol is rapid, reversible, stereo-specific and of relatively high affinity. The HeLa cells also contain an adenylate cyclase which is activated by (?)-isoproterenol > (?)-epinephrine > (?)-norepinephrine. The adenylate cyclase of HeLa is also activated by guanyl-5′-yl-imidodophosphate (Gpp(NH)p), a nonhydrolyzable analogue of GTP. Inclusion of both (?)-isoproterenol and Gpp(NH)p leads to approximately additive rathen than synergistic activation of adenylate cyclase. After treatment of HeLa cells with 5 mM sodium butyrate there is an increase in the number of β-adrenergic receptors, but not in their affinity, which is reflected in an increased ability of (?)-isoproterenol to activate adenylate cyclase. Other properties of the β-adrenergic receptor including association and dissociation rates, temperature optimum of adenylate cyclase and response to Gpp(NH)p are relatively unaffected by butyrate pretreatment of the cells.  相似文献   

10.
(1) H+/electron acceptor ratios have been determined with the oxidant pulse method for cells of denitrifying Paracoccus denitrificans oxidizing endogenous substrates during reduction of O2, NO?2 or N2O. Under optimal H+-translocation conditions, the ratios H+O, H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were 6.0–6.3, 4.02, 5.79 and 3.37, respectively. (2) With ascorbate/N,N,N′,N′-tetramethyl-p-phenylenediamine as exogenous substrate, addition of NO?2 or N2O to an anaerobic cell suspension resulted in rapid alkalinization of the outer bulk medium. H+N2O, H+NO?2 for reduction to N2 and H+NO?2 for reduction to N2O were ?0.84, ?2.33 and ?1.90, respectively. (3) The H+oxidant ratios, mentioned in item 2, were not altered in the presence of valinomycinK+ and the triphenylmethylphosphonium cation. (4) A simplified scheme of electron transport to O2, NO?2 and N2O is presented which shows a periplasmic orientation of the nitrite reductase as well as the nitrous oxide reductase. Electrons destined for NO?2, N2O or O2 pass two H+-translocating sites. The H+electron acceptor ratios predicted by this scheme are in good agreement with the experimental values.  相似文献   

11.
Adenylosuccinate synthase (EC 6.3.4.4.) (l-aspartate + GTP + IMPMg2+adenylosuccinate + GDP + Pi) is an important site for the regulation of adenylate biosynthesis. A partially purified preparation of the enzyme from Escherichia coli B showed feedback inhibition by ADP and AMP, weak positive response to the adenylate energy charge, and weak positive response to the mole fraction of GTP in the GTP + GDP pool. These responses seem to ensure that the synthesis of adenine nucleotides will be controlled appropriately in response to the level of end products and to the energy state of the cell, and to avoid the potential difficulties arising from the fact that the end products of this sequence and the indicators of the energy state of the cell are the same compounds.  相似文献   

12.
The in vitro reaction of bacteriophage T7-DNA with the radioactive diastereomeric benzo(a)pyrene-diol-epoxides, (±) [3H9, 3H10]-7β,8α-dihydroxy-9α,10β-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene, and (±) [3H9, 3H10]-7β,8α-dihydroxy-9β,19β-epoxy-7,8,9,10-tetrahydrobenzo(1)pyrene, was investigated. Chromatographic analysis of digests of the DNA allowed the distinction of characteristic deoxynucleoside adduct peaks for the two benzo(a)pyrene-diol-epoxides. Our results, together with data from the literature, allow the identification of these adducts as mostly N2-(10-7β,8α,9α-trihydroxy-7,8,9,10-tetrahydrobenzo(a)pyreney1)deoxyguanosine and N2-(10-7β,8α,9β-trihydroxy-7,8,9,10-tetrahydrobenzo(a)pyreney1)deoxyguanosine, respectively. DNA-benzo(a)pyrene adducts with the same chromatographic properties were formed in mouse embryo fibroblasts upon treatment with benzo(a)pyrene.  相似文献   

13.
beta-Endorphin: characteristics of binding sites in the rat brain.   总被引:3,自引:0,他引:3  
Stereospecific binding of human β-endorphin to rat membrane preparations is described for the first time using [3H-Tyr27]-βh-endorphin as the ligand. The binding is time dependent and saturable with respect to βh-endorphin with an apparent dissociation constant of 0.3 nM. Sodium ion (100 mM) elevates this value to 2.5 nM but has no effect on the total number of binding sites present in the membrane preparation. The ability of certain β-endorphin analogs, opiate agonists as well as antagonists to inhibit the binding of βh-endorphin, is presented.  相似文献   

14.
Liposomes prepared with cholesterol and dipalmitoyl phosphatidylcholine were incubated with a clone of normal rat kidney fibroblast of cells in culture. The cells took up [14C]cholesterol in proportion to the concentration of liposomes in the incubation medium, and the uptake increased with time over the four hours of study. Two cell membrane enzymes, adenylate cyclase and (Na+ + K+)-ATPase, exhibited decreased activity after treatment with cholesterol-containing liposomes. The decrease in adenylate cyclase activity was directly proportional to the uptake of [14C]cholesterol. When a variety of subclones of NRK 5W were examined some were found to respond to cholesterol treatment and some did not. These data are consistent with the view that membrane cholesterol content plays a role in controlling the activity of some plasma membrane enzymes.  相似文献   

15.
Using guanidinium and n-butylammonium cations (C+) as models for the positively charged side chains in arginine and lysine, we have determined the association constants with various oxyanions by potentiometric titration. For a dibasic acid, H2A, three association complexes may exist: K1M = [CHA][C+] [HA?]; K1D = [CA?][C+] [A2?]; K2D = [C2A][C+] [CA?]. For guanidinium ion and phosphate, K1M = 1.4, K1D = 2.6, and K2D = 5.1. The data for carboxylates indicate that the basicity of the oxyanion does not affect the association constant: acetate, pKa = 4.8, K1M = 0.37; formate, pKa = 3.8, K1M = 0.32; and chloroacetate, pKa = 2.9, K1M = 0.43, all with guanidinium ion. Association constants are also reported for carbonate, dimethylphosphinate, benzylphosphonate, and adenylate anions.  相似文献   

16.
[3H]Dihydroalprenolol, a potent ß-adrenergic antagonist, was used to identify the adenylate cyclase-coupled ß-adrenoceptors in isolated membranes of rat skeletal muscle. The receptor sites, as revealed [3H]dihydroalprenolol binding, were predominantly localized in plasmalemmal fraction. That skeletal muscle fraction may also contain the plasmalemma of other intramuscular cells, especially that of blood vessels. Hence, the [3H]dihydroalprenolol binding observed in that fraction may be due partly to its binding to the plasmalemma of blood vessels. Small but consistent binding was also observed in sarcoplasmic reticulum and mitochondria. The level of [3H]dihydroalprenolol binding in different subcellular fractions closely correlated with the level of adenylate cyclase present in those fractions.The binding of [3H]dihydroalprenolol to plasmalemma exhibited saturation kinetics. The binding was rapid, reaching equilibrium within 5 min, and it was readily dissociable. From the kinetics of binding, association (K1) and dissociation (K2) rate constants of 2.21 · M? · min?1 and 3.21 · 10?1, respectively, were obtained. The dissociation constant (Kd) of 15 nM for [3H]dihydroalprenolol obtained from saturation binding data closely agreed with the (Kd) derived from the ratio of dissociation and association rate constants (K2/K1).Several β-adrenergic agents known to be active on intact skeletal muscle also competed for [3H]dihydroalprenolol binding sites in isolated plasmalemma with essentially similar selectivity and stereospecificity. Catecholamines competed for [3H]dihydroalprenolol binding sites with a potency of isoproterenol > epinephrine > norepinephrine. A similar order of potency was noted for catecholamines in the activation of adenylate cyclase. Effects of catecholamines were stereospecific, (?)-isomers being more than potent than (+)-isomers. Phenylephrine, an α-adrenergic agonist, showed no effect either on [3H]dihydroalprenolol binding or on adenylate cyclase. Known ß-adrenergic antagonists, propranolol and alprenolol, stereospecifically inhibited the [3H]dihydroalprenolol binding and the isoproterenol-stimulated adenylate cyclase. The (Ki) values for the antagonists determined from inhibition of [3H]dihydroalprenolol binding agreed closely with the (Ki) values obtained from the inhibition of adenylate cyclase. The data suggest that the binding of [3H]dihydroalprenolol in skeletal muscle membranes possess the characteristics of a substance binding to the ß-adrenergic receptor.  相似文献   

17.
Ammonium[2-3H,1-14C]isobutyrate was converted by Pseudomonas putida ATCC 21244 into S(+)-β-hydroxyisobutyric acid (β-HIBA) with loss of the α-tritium atom. The recovered isobutyrate had the same 3H14C as the starting material. Ammonium (2S)-[3-13C]isobutyrate was synthesized and converted by P. putida into β-HIBA. The 13C-nmr of the corresponding methyl ester benzoate showed 13C enrichment in the hydroxymethyl carbon atom. The results therefore indicate that isobutyrate metabolism in this organism proceeds via an unsaturated intermediate (probably methacrylyl-CoA) formed by dehydrogenation of the 2-pro-S-methyl group of the precursor (isobutyryl-CoA). Hydration of the intermediate proceeds with addition of a proton at C-2 from the same side as the hydrogen removed in the dehydrogenation.  相似文献   

18.
In efforts to understand the molecular properties of ion channels in biomembranes, we have investigated the interaction of substituted guanidines with the Na+ channel site in membranes isolated from Electrophorus electricus. This interaction was measured by equilibrium competitive binding studies with [3H]tetrodotoxin ([3H]TTX); TTX has been shown to bind specifically to the Na+ channel in electrically excitable membranes. Although guanidine and small substituted guanidines such as methylguanidine or aminoguanidine competed with [3H]TTX for the membrane binding site, the apparent KI values for these derivatives were nearly seven orders of magnitude higher than the Kd for TTX. On the other hand, the binding of the guanidines was considerably enhanced by introducing a substituent aromatic ring or aliphatic chain. Detailed analysis of the binding of aliphatic guanidines of varying chain length clearly demonstrated the contribution made by hydrophobic interactions. These results suggest that the channel site may include a hydrophobic region in close proximity to the carboxylate previously postulated to be involved in TTX binding.  相似文献   

19.
The effects of absolute temperature (T), ionic strength (μ), and pH on the polymerization of tobacco mosaic virus protein from the 4 S form (A) to the 20 S form (D) were investigated by the method of sedimentation velocity. The loading concentration in grams per liter (C) was determined at which a just-detectable concentration (β) of 20 S material appeared. It was demonstrated experimentally that under the conditions employed herein, an equilibrium concentration of 20 S material was achieved in 3 h at the temperature of the experiment and that 20 S material dissociated again in 4 h or less to 4 S material either upon lowering the temperature or upon dilution. Thus, the use of thermodynamic equations for equilibrium processes was shown to be valid. The equation used to interpret the results, log (C?β) = constant + (ΔH12.3RT) + (ΔW1el2.3RT) ? K′ + ζpH, was derived from three separate models of the process, the only difference being in the anatomy of the constant; thus, the method of analysis is essentially independent of the model. ΔH1 and ΔW1el are the enthalpy and the change in electrical work per mole of A protein (the trimer of the polypeptide chain), Ks is the salting-out constant on the ionic strength basis, ζ is the number of moles of hydrogen ion bound per mole of A protein in the polymerization, and R is the gas constant. The three models leading to this equation are: a simple 11th-order equilibrium between A1 (the trimer of the polypeptide chain) and D, either the double disk or the double spiral of approximately the same molecular weight, designated model A; a second model, designated B, in which A1 was assumed to be in equilibrium with D at the same time that it is in equilibrium with A2, A3, etc., dimers and trimers, etc., of A1 in an isodesmic system; and a phase-separation model, designated model C, in which A protein is treated as a soluble material in equilibrium with D, considered as an insoluble phase. From electrical work theory, ΔWel1/T was shown to be essentially independent of T; therefore, in experiments at constant μ and constant pH the equation of log (C ? β) versus 1/T is linear with a slope of ΔH1/2.3R. The results fit such an equation over nearly a 20 °C-temperature range with a single value of ΔH1 of +32 kcal/mol A1. Results obtained when T and pH were held constant but μ was varied did not fit a straight line, which shows that more than simple salting-out is involved. When the effect of ionic strength on the electrical work contribution was considered in addition to salting-out, the data were interpreted to indicate a value of ΔW1el of 1.22 kcal/mol A1 at pH 6.7 and a value of 4.93 for Ks. When μ and T were held constant but pH was varied, and when allowance was made for the effect of pH changes on the electrical work contribution, a value of 1.1 was found for ζ. This means that something like 1.1 mol of hydrogen ion must be bound per mole of A1 protein in the formation of D. When this is added to the small amount of hydrogen ion bound per A1 before polymerization, at the pH values used, it turned out that for D to be formed, 1.5 H+ ions must be bound per A1 or 0.5 per protein polypeptide chain. This amounts to 1 H+ ion per polypeptide chain for half of the protein units, presumably those in one but not the other layer of the double disk or turn of the double spiral. When polymerization goes beyond the D stage, as shown by previously published data, additional H+ ions are bound. Simultaneous osmotic pressure studies and sedimentation studies were carried out, in both cases as a function of loading concentration C. These results were in complete disagreement with models A and C but agreed reasonably well with model B. The sedimentation studies permitted evaluation of the constant, β, to be 0.33 g/liter.  相似文献   

20.
Acid dissociation constants of aqueous cyclohexaamylose (6-Cy) and cycloheptaamylose (7-Cy) have been determined at 10–47 and 25–55°C, respectively, by pH potentiometry. Standard enthalpies and entropies of dissociation derived from the temperature dependences of these pKa's are ΔH0 = 8.4 ± 0.3 kcal mol?1, ΔS0 = ?28. ± 1 cal mol?10K?1 for 6-Cy and ΔH0 = 10.0 ± 0.1 kcal mol?1, ΔS0 = ?22.4 ±0.3 cal mol?10K?1 for 7-Cy. Intrinsic 13C nmr resonance displacements of anionic 6- and 7-Cy were measured at 30°C in 5% D2O (vv). These results indicate that the dissociation of 6- and 7-Cy involves both C2 and C3 20-hydroxyl groups. The thermodynamic and nmr parameters are discussed in terms of interglucosyl hydrogen bonding.  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