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.
The theory of Martin-Synge distribution (MSD) was refined, with special attention being focused upon the derivation of the separation functions. The separation function for the fundamental distribution of MSD was obtained in the form v = t2k1 + 1)(αk1 + β)[(αk1 + 1)1/2 + (αk1 + β)1/2]2k1(β ? 1)2, where ν is the number of aliquots vm driven through the apparatus, t the abscissa of the standard normal distribution, α = vm/v8 the phase ratio, β = k1/k2≥ 1 the separation factor, and k1 the partition coefficient of the more rapidly moving component; ν was shown to have minima at given αk1 values. The separation function of the single withdrawal of MSD was presented in the form N = u + 1 = t2(2αk1 + β + 1)2/(β ? 1)2+ 1, where N is the number of partition units; N is minimal when αk1 = 0. The elution volumes and standard deviations of the two compounds to be separated were mathematically analyzed in a manner similar to that previously presented when dealing with the theory of counter-current distribution (CCD). As in CCD, the elution volumes in MSD were found to have minima at given αk1 values. However, the standard deviations of the elution curves also have minima in respect to αk1 in MSD, which is a different situation as compared to CCD. The selection of optimal operating conditions was found to be more critical in MSD than in CCD.  相似文献   

2.
The theory of countercurrent distribution (CCD) was reviewed and extended. The separation function for the fundamental distribution of CCD was presented in the form n = t2k1+β)2k1(β?1)2 where n is the number of transfers, t the abscissa of the standard normal distribution, α = vm/v8 the phase ratio, β = k1/k2≥ 1 the separation factor, and k1 the partition coefficient of the more radidly moving component; n was found to be minimal on the condition αk1 = β. The separation function for the single withdrawal of CCD was obtained in the form N = u + 1 = t2{(αk1 + 1)1/2 + [β(αk1 + β)]1/2}2/(β ? 1)2+ 1, where N is the number of partition units. From this equation it appears that N is minimal when αk1 = 0. Compared with the former separation functions presented in the literature, these separation functions have the advantage of giving directly the relationships among the phase ratio, the absolute partition coefficient, the separation factor, the resolution degree, and the number of transfers or partition units required. In addition, the dependencies of the elution volumes and the widths of the elution curves on α, β, and the partition coefficients were considered mathematically by means of differential calculus. The elution volumes were found to have minima at certain αk1 values. The standard deviations, on the contrary, did not have minima in respect to αk1. The theory presented can be used for selecting proper operating conditions while separating chemical compounds.  相似文献   

3.
From φ, ψ data on eleven proteins, a 20 × 20 × 20 Table of tripeptides has been computed to evaluate the influence of nearest neighbors (n ? 1) and (n + 1) on the φ, ψ angles of amino acid (n). From this Table, having removed values for horse cytochrome c and using the sequences of 18 cytochromes c and the procedure of Kabat &; Wu (1972), an attempt was made to select a set of φ, ψ angles for positions 2 to 103 of cytochrome c and compare them with the values obtained from the atomic co-ordinates. Agreement was good for 56, intermediate for 29 and poor for 17 residues. Eleven of the 17 with poor agreement were residues contacting the heme or adjacent to a contacting residue. Moreover, 6 of the 17 poor values were in regions of the φ, ψ plot for which no occurrences in the ten known proteins were reported, and for four others known values were minimal so that no basis for selection existed. Frequency distributions on the Ramachandran plot (Ramachandran &; Sasisekharan, 1968) of all φ, ψ values in the eleven known proteins are given as well as a contour plot for such frequencies. The uses and limitations of the procedure are discussed and the need for obtaining accurate estimates of errors in experimentally determined φ, ψ angles is emphasised.  相似文献   

4.
Model-building of neurohypophyseal hormones   总被引:6,自引:0,他引:6  
An attempt is made to construct models of the polypeptide backbone of the neurohypophyseal hormones. Using the sequences of six hormones and the empirical statistical approach (Kabat &; Wu, 1972; Wu &; Kabat, 1973), an initial selection is made of a set of φ, ψ angles satisfying all six sequences from a 20 × 20 × 20 table of φ, ψ values of tripeptides in 11 known proteins. Because of insufficient data at several positions, φ, ψ angles in the α-helical and β-sheet domains were considered as possible alternatives and the final selection was made to bring the disulfide bond as close as possible (model I). These initial values are then subjected to an energy minimization to refine the model; reasonable agreement with nuclear magnetic resonance data for possible φ values is obtained. Model II is constructed to be consistent with the available nuclear magnetic resonance data and the suggestion of a sequential Cu2+-binding site. Both models contain similar structural features and may be converted to one another with rotation of only a few angles. Interactive computer graphics was applied to construct model II, and its use in problems in this type is described. The relation of the experimental data to the various models that have been proposed for the neurohypophyseal hormones is discussed and an experiment is suggested that may permit a choice between them.  相似文献   

5.
Coiling of beta-pleated sheets   总被引:4,自引:0,他引:4  
To form strongly twisted β-sheets, strands have to be coiled as well as twisted (Nishikawa &; Scherga, 1976). I show that strands coil in the appropriate right-handed direction if their main-chain torsion angles fulfil the following conditions: ψi ? ?φi + 1, ψi + 1 > ?φi + 2, ψi + 2 ? ?φi + 3, ψi + 3 > ?φi + 4…Lactate dehydrogenase, pancreatic trypsin inhibitor, thermolysin and concanavalin A contain strongly twisted β-sheets and in each case the strands are coiled by their φ, ψ values fulfilling these conditions.  相似文献   

6.
The αII-helix (? = ?70.47°, ψ = ?35.75°) is a structure having the same n and h as the (standard) αI-helix (? = ?57.37°, ψ = ?47.49°). Its conformational angles are commonly found in proteins. Using an improved α-helix force field, we have compared the vibrational frequencies of these two structures. Despite the small conformational differences, there are significant predicted differences in frequencies, particularly in the amide A, amide I, and amide II bands, and in the conformation-sensitive region below 900 cm?1. This analysis indicates that αII-helices are likely to be present in bacteriorhodopsin [Krimm, S. & Dwivedi, A. M. (1982) Science 216 , 407–408].  相似文献   

7.
The crystal structure of potassium hydroxide complexed amylose, obtained by heterogeneous deacetylation of amylose triacetate, has been determined through a combined stereochemical structure-refinement and X-ray diffraction-analysis. The structure crystallizes in an orthorhombic unit-cell with parameters a  8.84, b  12.31, and c (fiber repeat)  22.41 Å, and with P212121 symmetry. The conformation of the amylose chain is a distorted, left-handed helix with 6 d-glucose residues per turn. Each three-residue asymmetric unit is complexed with one molecule of potassium hydroxide and three molecules of water. The K+ ion coordinates with four oxygen atoms of the amylose chain and with two other oxygen atoms, and this coordination is probably the cause for the more-extended amylose chain-conformation than would be predicted from a φ, ψ map. The distortions in the chain are primarily manifested by different O-6 rotations and by slightly different bridge and φ, ψ angles for the individual residues. The structure is extensively hydrogen bonded, although largely through water molecules, which accounts for its ready water solubility. The left-handed conformation of the chain in this structure is consistent with the conformations of amylose triacetate and V-amylose, both of which are left-handed.  相似文献   

8.
In examining orientations of glycosidic linkages, measurements of three-bond coupling between 13C-1 and 1H-4′, or 13C-4′ and 1H-1, have been made from natural abundance, 1H-coupled, 13C-n.m r. spectra of maltose, cyclohexaamylose, and related compounds. Maltose and cyclohexaamylose in water exhibit inter-residue 13COC1H couplings of close to 3 Hz. In terms of torsional angles, φ and ψ, these findings suggest that, in aqueous solution, the molecules favor conformations that are appreciably more staggered than those known to exist in the solid state. Analogous measurements on O-acetyl derivatives suggest that φ is smaller, and ψ larger, than in maltose. Data are also presented for sucrose, maltosan, and α,α-trehalose.  相似文献   

9.
3-O-(6-O-Acetyl-2,3-anhydro-4-deoxy-α-l-ribo-hexopyranosyl)-1,2:5,6-di-O-isopropylidene-α-d-glucofuranose has been synthesised and its monocrystal investigated by X-ray diffraction methods. The compound crystallises in the orthorhombic system, space group P212121, with cell constants a = 8.790(7), b = 11.678(4), and c = 21.457(10) Å. The intensity data were collected with a four-circle CAD-4 diffractometer. From a total of 1684 intensities, 1275 were of I > 2σI. The structure was solved by direct methods and refined by the full-matrix, least-squares procedure, resulting in R 0.057. The 4-deoxy-2,3-anhydropyranose ring is characterised by a sofa conformation (5E), the 1,2-O-isopropylidene ring has a hybrid conformation (E + T), and the 5,6-O-isopropylidene and the α-d-glucofuranose rings have twist (T) conformations. The φ and ψ torsion angles for the glycosidic linkage are 54(4)° and 29(4)°, respectively.  相似文献   

10.
Abstract

We generated φ -ψ conformational energy contour maps for the of N-acetyl alanine N'-methyl amide using the molecular mechanics forcefields AMBER, AMBER3, BI085, CFF91, CVFF, MM2, MM3, MM+, and SYBYL. With MM2, MM3, and MM+, we used a dielectric constant of ? = 1.5, the default effective value for these forcefields. With the other forcefields we used ? = 1 and 4, except with SYBYL, which, in Spartan 3.1, has no electrostatic term. All forcefields yielded the Ceq 7 conformation as a low-energy minimum or the global minimum. Most of the forcefields also yielded a minimum-energy conformation in the C5R, and αt. regions of the φ -ψ contour map. Fewer of the forcefields yielded a minimum in the Cax 7 region; however, adiabatic relaxation frequently lowered the relative energy of this region. Based on the appearance of the φ -ψ maps, the following pairs of forcefields were broadly similar (but not identical) to each other but dissimilar to the other pairs: AMBER3 and AMBER, BI085 and CHARMM, MM+ and MM2, SYBYL and ECEPP, and CFF91 and MM3. We used the data from the φ -ψ contour maps to compute the characteristic ratio of poly-L-alanine. Most of the computed values deviated significantly from the experimental value. Only the computed characteristic ratio of CFF91 without adiabatic relaxation at ? = 4 and MM3 without adiabatic relaxation at ? = 1.5 agreed with the experimental value.  相似文献   

11.
Cratoxylum cochinchinense displayed significant inhibition against protein tyrosine phosphatase 1B (PTP1B) and α-glucosidase, both of which are key target enzymes to attenuate diabetes and obesity. The compounds responsible for both enzymes inhibition were identified as twelve xanthones (112) among which compounds 1 and 2 were found to be new ones. All of them simultaneously inhibited PTP1B with IC50s of (2.4–52.5?µM), and α-glucosidase with IC50 values of (1.7–72.7?µM), respectively. Cratoxanthone A (3) and γ-mangostin (7) were estimated to be most active inhibitors against both PTP1B (IC50?=?2.4?µM for 3, 2.8?µM for 7) and α-glucosidase (IC50?=?4.8?µM for 3, 1.7?µM for 7). In kinetic studies, all isolated xanthones emerged to be mixed inhibitors of α-glucosidase, whereas they behaved as competitive inhibitors of PTP1B. In time dependent experiments, compound 3 showed isomerization inhibitory behavior with following kinetic parameters: Kiapp?=?2.4?µM; k5?=?0.05001?µM?1?S?1 and k6?=?0.02076?µM?1?S?1.  相似文献   

12.
In platelets, the glycoprotein (GP) Ib-IX receptor complex senses blood shear flow and transmits the mechanical signals into platelets. Recently, we have discovered a juxtamembrane mechanosensory domain (MSD) within the GPIbα subunit of GPIb-IX. Mechanical unfolding of the MSD activates GPIb-IX signaling into platelets, leading to their activation and clearance. Using optical tweezer-based single-molecule force measurement, we herein report a systematic biomechanical characterization of the MSD in its native, full-length receptor complex and a recombinant, unglycosylated MSD in isolation. The native MSD unfolds at a resting rate of 9 × 10?3 s?1. Upon exposure to pulling forces, MSD unfolding accelerates exponentially over a force scale of 2.0 pN. Importantly, the unfolded MSD can refold with or without applied forces. The unstressed refolding rate of MSD is ~17 s?1 and slows exponentially over a force scale of 3.7 pN. Our measurements confirm that the MSD is relatively unstable, with a folding free energy of 7.5 kBT. Because MSD refolding may turn off GPIb-IX’s mechanosensory signals, our results provide a mechanism for the requirement of a continuous pulling force of >15 pN to fully activate GPIb-IX.  相似文献   

13.
Potential energies of conformation of a dipeptide unit with butyl, seryl, threonyl, eysteinyl, and valyl side groups have been computed by using classical energy expressions. The presence of a γ-atom introduces characteristic restrictions on the backbone rotational angles ? and ψ the γ-atom itself is restricted to three staggered positions about the Cα—Cβ bond. The important results are that a γ-carbon in position I (χ1 ? 60°) cannot be accommodated in the standard right-and left-handed α-helices, whereas a γ-oxygen or sulfur could easily be accommodated in the right-handed α-helix. Further, a γ-carbon or a heteroatom in position II (χ1 ? 180°) does not favor a conformation ψ ? 180°, compared to two other positions. The valyl side group significantly reduces the allowed ? and ψ values and energetically prefers a β-conformation compared to right-or left-handed α-helical conformations. The less favorable α-helical conformation is possible only for γ (III, II) combination of the valyl residue. The observed ?, ψ, and χ1 values of all the amino acid residues in the three protein molecules, lysozyme, myoglobin, and chymotrypsin are compared with the theoretical predictions and the agreement is excellent. The results bring out the important fact that even in large molecules, the conformation of local segments are predominantly governed by the short-range intramolecular interactions.  相似文献   

14.
Quenching of singlet molecular oxygen (1ΔgO2) by α-tocopherol (I) involves the hydroxy function of the chromanol ring of I. In phosphatidylcholine (PC) uni- and multilamellar vesicles this structural element of I is localized at the interface polar headgroup/hydrophobic core. A dielectric constant of ? ~ 25 was determined for this special region of the PC bilayer. The ratio kQ/kR of rate constants of quenching processes (kQ) and irreversible reactions (kR) of I with 1ΔgO2 increases with decreasing polarity of the solvent. In ethanolic solutions where ? = 25.5, kQ/kR is about 40. Extrapolation of these results to phospholipid bilayers suggests that at the nearness of the ester carbonyl oxygen of the PC fatty acid moieties, α-tocopherol can deactivate approximately 40 1ΔgO2 molecules before being destroyed. It is concluded that in vivo, one may expect to find a higher kQ/kR ratio if the chromanol ring of I hides within the more hydrophobic interiors of the membrane surface peptides.  相似文献   

15.
The stoichiometry of Na coupling to amino acid movement across the brush border membrane of the rabbit distal ileum has been determined under initial rate conditions.The coupling ratio, defined as the amino acid-dependent Na influx/the Na-dependent amino acid influx, was equal to unity for alanine, measured over a 10-fold range of Na and alanine concentrations. Coupling ratio values determined under a single set of conditions for a number of amino acids varied from 1 for serine to 4.6 for methionine. Reducing the methionine concentration from 12.5 to 1.5 mM caused the coupling ratio value to fall from 4.6 to 1.2.These results are explained by assuming a fixed stoichiometry of 1 : 1 under all conditions, with initial binding of the amino acid (A) to the Na-dependent carrier (E) but with some amino acids being able to cross on the Na-dependent carrier in the absence of Na.The variation in coupling ratio values can be used to calculate KA, the apparent dissociation constant of amino acid from the Na-dependent carrier in the absence of Na, and the ratio k1k2, where k1 and k2 are first-order rate constants for translocation of the complexes EA and EANa, respectively. This method of processing results has been defined as delta analysis. The value of KA for methionine is 3.6 ± 1.1 mM and the k1k2 ratio is 1.01 ± 0.07. The constant coupling ratio value of 1 for alanine indicates that the value for KA is extremely high or that the k1 value is extremely low.  相似文献   

16.
A single-crystal x-ray diffraction analysis of Boc-L -Ala-D -aIle-L -Ile-OMe has been carried out. The analysis has shown (a) that the tripeptide molecules have in part an α-extended conformation, the torsion angles of the L -Ala and D -aIle residues being φ1 = ?75.1° and ψ1 = ?25.8° and φ2 = 67.3° and ψ2 = 44.1°, respectively, and (b) that the molecules are organized in rippled planes where they occur in relative antiparallel orientation linked together side by side by H bonds. This molecular organization of the tripeptide corresponds closely to that of an antiparallel α-pleated sheet, and likely constitutes the first example of a structure of this kind for which a characterization at the atomic level has been achieved. A molecular dynamics study has shown that the molecular conformation of the tripeptide in the crystalline state is determined primarily by intermolecular interactions. © 1994 John Wiley & Sons, Inc.  相似文献   

17.
Monte-Carlo calculations of geometric and thermodynamic characteristics of the α-helix and the β-structure of polypeptides have been carried out. To describe a hydrogen bond both the Lippincott–Schroeder and Morse potentials were used. The internal rotation angles φ and ψ in the α-helix have been shown to fluctuate in the range of ±7°. The distribution functions on angles φ and ψ and on hydrogen bond lengths and angles in the α-helix have been computed and compared with those in myoglobin and lysozyme. Thermodynamic characteristics of the α-helix calculated in different approximations with the two forms of the hydrogen bond potentials have also been compared. The data obtained are close to the experimental values for polypeptides in neutral solution. Some geometric and thermodynamic characteristics of the regular parallel and antiparallel and irregular antiparallel β-structure have been found. In the β-structure the internal rotation angles vary within the interval ±15–20°. An increase in the cross and longitudinal dimensions of the β-structure only slightly influence both the geometric and thermodynamic characteristics.  相似文献   

18.
Phenolic acid decarboxylase (PAD) catalyzes the non-oxidative decarboxylation of p-coumaric acid (pCA) to p-hydroxystyrene (pHS). PAD from Bacillus amyloliquefaciens (BAPAD), which showed k cat/K m value for pCA (9.3?×?103?mM?1?s?1), was found as the most active one using the “Subgrouping Automata” program and by comparing enzyme activity. However, the production of pHS of recombinant Escherichia coli harboring BAPAD showed only a 22.7 % conversion yield due to product inhibition. Based on the partition coefficient of pHS and biocompatibility of the cell, 1-octanol was selected for the biphasic reaction. The conversion yield increased up to 98.0 % and 0.83 g/h/g DCW productivity was achieved at 100 mM pCA using equal volume of 1-octanol as an organic solvent. In the optimized biphasic reactor, using a three volume ratio of 1-octanol to phosphate buffer phase (50 mM, pH 7.0), the recombinant E. coli produced pHS with a 88.7 % conversion yield and 1.34 g/h/g DCW productivity at 300 mM pCA.  相似文献   

19.
Phthalic anhydride has been detected spectrophotometrically in the hydrolysis of phthalamic acid and N-phenylphthalamic acid in solutions which were made up to 5 M with sodium perchlorate. In solutions of lower ionic strength the variation of kobs with acid concentration follows the equation, kobs=(k1 + k2 [H3O+])/(1 + Ka/[H3O+]), and the values of k1 and k2 are both enhanced. The p values for the variation of k1 and k2 with the aryl group for the hydrolysis of N-arylphthalamic acids are ?1.23 and ?0.94. A mechanism involving nucleophilic catalysis by the carboxyl group in which breakdown of the tetrahedral intermediate is rate-limiting was proposed.  相似文献   

20.
Highlights? Two structures of the RGS2-Gαq complex were determined by X-ray crystallography ? RGS2 binds Gαq in a manner distinct from how other RGS proteins bind Gαi/o ? In its distinct pose, RGS2 forms extensive contacts with the α-helical domain of Gαq ? Helical domain contacts contribute to binding affinity and GAP potency of RGS2  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