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.
G C Levy  D J Craik  A Kumar  R E London 《Biopolymers》1983,22(12):2703-2726
The nature of internal and overall motions in native (double-stranded) and denatured (single-stranded) DNA fragments 120–160 base pairs (bp) long is examined by molecular-dynamics modeling using 13C-nmr spin-relaxation data obtained over the frequency range of 37–125 MHz. The broad range of 13C frequencies is required to differentiate among various models. Relatively narrow linewidths, large nuclear Overhauser enhancements (NOEs), and short T1 values all vary significantly with frequency and indicate the presence of rapid, restricted internal motions on the nanosecond time scale. For double-stranded DNA monomer fragments (147 bp, 24 Å diam at 32°C), the overall motion is that of an axially symmetric cylinder (τx = ~10?6 s;τZ = ~1.8 × 10?8s), which is in good agreement with values calculated from hydrodynamic theory (τx = ~1.8 × 10?6 s; τZ = ~2.7 × 10?8 s). The DNA internal motion can be modeled as restricted amplitude internal diffusion of individual C? H vectors of deoxyribose methine carbons C1′, C3′, and C4′, either with conic boundary conditions (τw = ~4 × 10?9 s, θcone = ~21°) or as a bistable jump (τA = τB = ~2 × 10?9 s, θ = ~15°). We discuss the critical role in molecular-dynamics modeling played by the angle (β) that individual C? H vectors make with the long axis of the DNA helix. Heat denaturation brings about increases in both the rate and amplitude of the internal motion (described by the wobble model with τW = ~0.2 × 10?9 s, θcone = ~50°), and overall motion is affected by becoming essentially isotropic (τx = τZ = ~5 × 10?8 s) for the single-stranded molecules. Since 13C-nmr data obtained at various DNA concentrations for C2′ of the deoxyribose ring is not described well by the above models, a new model incorporating an additional internal motion is proposed to take into account the rapid, extensive, and weakly coupled motion of C2′.  相似文献   

2.
Primary and tertiary amine-initiated polymerizations of L -alanine-N-carboxyanhydride (L -Ala-NCA) were conducted at 20 or 100°C in a variety of solvents. The 75.5-MHz 13C-nmr CP/MAS spectra of the resulting poly(L -alanines) revealed that all samples contain both α-helix and pleated-sheet structures. Depending on the reaction conditions the α-helix content varied between ca. 1 and 99%. Reprecipitation from aprotic nonsolvents does not change the α-helix/β-sheet ratio, indicating that this ratio is thermodynamically controlled. Since relatively large amounts of oligopeptides of degree of polymerization (DP ) 4–6 can be extracted by means of acetic acid, it is concluded that (a) most poly(L -alanines) possess a bimodal molecular weight distribution, (b) the oligopeptide fraction with DP ? 11 is responsible for the β-sheet fraction of all samples, and (c) the two-stage crystal growth proposed by Komoto and Kawai is not correct. Solubilizing initiators such as poly(ethylene oxide) NH2 prevent the precipitation of oligoalanine and, thus, the formation of a β-sheet structure. 13C-nmr CP/MAS measurements also show that tri- and tetra-L -alanines form insoluble β-sheet structures.  相似文献   

3.
The helix-coil transitions for poly(L -glutamic acid) (PGA) in 0.2M NaCl and in its mixture with dioxane were studied by the methods of spectropolarimetry, viscometry, and potentiometric titration at different temperatures from 8 to 50°C. The enthalpy and entropy differences between the helical and coillike states of uncharged PGA molecules were determined from the curves of potentiometric titration. The temperature dependence of the cooperativity parameter σ was determined by two methods: from the sharpness of transition and from the dependence of the intrinsic viscosity on the helical content in the transition region. In 0.2MNaCl, σ= (2.5 ± 0.5) × 10?3 and practically does not depend on temperature, i.e., the cooperativity of the helix-coil transition is connected mainly with the entropy decrease in initiating helical regions (ΔSi ≈ ?12 is mole of helical regions). On the contrary, initiation of a helical region in the water-organic solvent mixture is accompanied by a considerable enthalpy increase.  相似文献   

4.
The infrared (IR), vibrational circular dichroism (VCD), and electronic circular dichroism (ECD) spectra of short cationic sequential peptides (L ‐Lys‐L ‐Ala‐L ‐Ala)n (n = 1, 2, and 3) were measured over a range of temperatures (20–90 °C) in aqueous solution at near‐neutral pH values in order to investigate their solution conformations and thermally induced conformational changes. VCD spectra of all three oligopeptides measured in the amide I′ region indicate the presence of extended helical polyproline II (PPII)‐like conformation at room temperature. UV‐ECD spectra confirmed this conclusion. Thus, the oligopeptides adopt a PPII‐like conformation, independent of the length of the peptide chain. However, the optimized dihedral angles ? and ψ are within the range ?82 to ?107° and 143–154°, respectively, and differ from the canonical PPII values. At elevated temperatures, the observed intensity and bandshape variations in the VCD and ECD spectra show that the PPII‐like conformation of the Lys‐Ala‐Ala sequence is still preferred, being in equilibrium with an unordered conformer at near‐neutral pH values within the range of temperatures from 20 to 90 °C. This finding was obtained from analysis of the temperature‐dependent spectra using the singular value decomposition method. The study presents KAA‐containing oligopeptides as conformationally stable models of biologically important cationic peptides and proteins. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

5.
The growth of Nile tilapia Oreochromis niloticus (0·02–20·00 g) was measured when fed to excess during the hours of light, following their exposure to five thermal regimes fluctuating around the thermal optimum for growth (Topt = 30° C) over the diel cycle of day (light, L) and night (dark, N), i.e. 27° C(L):33° C(N), 28·5° C(L):31·5° C(N), 30° C(L):30° C(N), 31·5° C(L):28·5° C(N) and 33° C(L):27° C(N) (two replicates per treatment, six weeks' rearing, growth measurements at weekly intervals). A model constructed with a stepwise multiple‐regression analysis accounted for 87·4% of the variation of the specific growth rate (G, % M day?1) from the variations of wet mass (M), the extent of the thermal fluctuation (FT) and their interactions, i.e. log10G = 1·7686 ? 0·2136 log10M + 0·0806 [log 10M× log 10 (1 + FT)] ? 0·0394 [log10M× log 10 (1 + FT)]2. Based on this model, the thermal fluctuation that produces the fastest growth ( ,°C) decreases in a curvilinear way, from 5·1° C at 20 mg to c. 0·7° C at 20 g. Thermal regimes that produce the slowest growth also produce the highest size heterogeneity. Functional hypotheses behind the size‐dependent effects of thermal fluctuations are discussed, together with their implications in natural habitats and aquaculture systems with in different contexts of food availability.  相似文献   

6.
The crystal structure of a dipeptide L -leucyl–L -leucine (C12H24N2O3) has been determined. The crystals are monoclinic, space group P21, with a = 5.434(4) Å, b = 15.712(7) Å, c = 11.275(2) Å, β = 100.41(1)°, and Z = 2. The crystals contain one molecule of dimethyl sulfoxide (DMSO) as solvent of crystallization for each dipeptide molecule. The structure has been solved by direct methods and refined to a final R index of 0.059 for 920 reflections (sinθ/λ ? 0.60 Å?1) with I ? 2σ (I). The trans peptide unit shows substantial degree of non-planarity (Δω = 14°). The peptide backbone adopts an extended conformation with torsion angles of ψ1 = 138(1)°, ω1 = 166(1)°, ?2 = ? 149.3(7)°, ψ21 = 164.2(7)°, and ψ22 = ? 15(1)°. For the first leucyl residue, the side-chain conformation is specified by the torsion angles 1χ1 = 176.7(7)°, 1χ21 = 62(1)°, 1χ22 = ? 177.4(8)°; the second leucyl residue adopts a Sterically unfavorable conformation with 2χ1 = 61(1)°, 2χ21 = 97(1)°, and 2χ22 = ?151(1)°. The packing involves head-to-tail interaction of peptide molecules and segregation of polar and nonpolar regions. The DMSO molecule is strongly hydrogen bonded to the terminal NH group. © 1994 John Wiley & Sons, Inc.  相似文献   

7.
Abstract

1,3,5-Trimethyl-N4-hydroxycytosine, an analogue of the promutagenic N4-hydroxycytosine and 5-methyl-N4-hydroxycytosine nucleosides, crystallizes in the monoclinic space group P 21/n with cell dimensions at ?147°C: a = 7.1481(7), b = 9.2565(5), c = 13.3086(12) Å, β = 97.90(2)°, V = 872.24(13) Å3, ρc = 1.426 Mg m?3, Z = 4, F(000) = 401.39, μ = 0.91 mm?1, λ(Cu) = 1.54056 Å, 20(max) = 139.3°. The crystal structure has been solved by X-ray difraction and refined to R = 3.7 % for 1457 reflections. Notwithstandin the steric hindrance imposed by methyl groups at both N(3) and C(5), the exocyclic N4-OH group is located essentially in the plane of the ring, giving rise to an “overcrowded” molecule, like that of 1,5,N4,N4-tetramethylcytosine. The conformational parameters have also been compared with those of a number of related and previously reported N(1)-substituted cytosines. In the present compound the N4-OH rotamer is in the anti conformation relative to the ring N(3), hence similar to that of one of the rotamers in N(1)-substituted N4-hydroxycytosine, which permits normal Watson-Crick base pairing of the latter, relevant to the mechanism of hydroxylamine mutagenesis.  相似文献   

8.
A Yaron  N Tal  A Berger 《Biopolymers》1972,11(12):2461-2481
The sequence-ordered copolymer poly-(Lys-Ala-Ala) was synthesized by polycondensation of the N-hydroxysuccinimide ester of ε,Z-Lys-Ala-Ala and deprotection of the polymerization product. A fraction of molecular weight 13,000 obtained by ion-exchange chromatography was investigated. The polymer is freely soluble in water at all pH values, and is completely digested by trypsin and elastase. From CD and ORD data it was concluded that in water at 1°C the ionized form (at pH 6.5) of the polymer is helical. On heating, helix-coil transition curves were obtained with a midpoint, Tm, depending on salt concentration. In salt-free water Tm = 12.3°C and in 0.2M NaCl Tm = 28.5°C. Adding MeOH, causes an increase in the helical content of the polymer (half helicity at 20% MeOH, without salt, at 29°C). Guanidine·HCl was shown to decrease the helicity. At 1°C half helicity. The nonionized polymer helix is more stable (Tm~90°C). At the high pH, at 60°C, when concentration of the polymer is higher than 1.9 × 10-2M, a precipitate is formed which redissolves on cooling with the original helicity. This does not occur in the presence of 50% MeOH. By comparison with polylysine it was concluded that replacing two-thirds of the lysine residues in polylysine by alanine leads to a polymer forming a more stable α-helix, when fully ionized. This is essentially due to the diminished coulombic repulsion. Uncharged lysine residues are comparable to alanine residues in their helix-forming tendency since the sequential polymer as well as one-third ionized polylysine are helical to approximately the same extent at room temperature.  相似文献   

9.
H. N. Cheng  F. A. Bovey 《Biopolymers》1977,16(7):1465-1472
By means of carbon-13 nmr (at 25 MHz) the trans/cis conformer ratio in glycyl-L -proline has been measured in aqueous (D2O) solution over the temperature range 33–96°C. It is found that ΔH0 = ?4.2 kJ/mole and ΔS0 = ?9.7 J/mole/K. Measurements of the T1 values for the proline ring carbons yielded values consistent with a fast puckering process involving both the β- and γ-carbons. Measurements of the rate of cis-trans conformational interconversion in glycyl-L -proline, using complete line-shape analysis for the glycyl α-carbon resonance, gave values for the transcis isomerization as follows: ΔH = 83.5 ± 0.2 kJ/mole; ΔS = 0.0 ± 10 J/mole/K. A more approximate determination from coalescence temperature observations gave a value of ΔG of 82.0 ± 0.4 kJ/mole for this process in acetyl-L -proline in aqueous solution. The presence of 12M NaSCN lowered this barrier by ca. 2.6 kJ/mole. Such measurements are relevant to present theoretical models of the denaturation-renaturation processes in proteins, in which proline residues may play a key role.  相似文献   

10.
The synthesis of the tetrapeptide benzyloxycarbonyl(α-aminoisobutyryl-L -prolyl)2-methyl ester (Z-(Aib-Pro)2-OMe) and an analysis of its conformation in solution and the solid state are reported. Stepwise synthesis using dicyclohexylcarbodiimide leads to racemization at Pro(2). Evidence for the presence of diastereomeric tetrapeptides is obtained from 270-MHz1H-nmr and 67.89-MHz 13C-nmr. The all-L tetrapeptide is obtained by fractional crystallization from ethyl acetate. The NH of Aib(3) is shown to be involved in an intramo-lecular hydrogen bond by variable-temperature 1H-nmr and the solvent dependence of NH chemical shifts. The results are consistent with a β-turn conformation with Aib(1) and Pro(2) at the corners stabilized by a 4 → 1 hydrogen bond. The molecule crystallizes in the space group P212121, with a = 8.839, b = 14.938, and c = 22.015 Å. The structure has been refined to an R value of 0.051. The peptide backbone is all-trans, and a 4 → 1 hydrogen bond, between the CO group of the urethane moiety and Aib(3) NH, is observed. Aib(1) and Pro(2) occupy the corner positions of a type I β-turn with ? = ?55.4°, Ψ = ?31.3° for Aib(1) and ? = ?71.6°, Ψ = ?38° for Pro(2). The tertiary amide unit linking Pro(2) and Aib(3) is significantly distorted from planarity (Δω = 14.3°).  相似文献   

11.
The crystal structure of the nonapeptide Boc-D -Phe-Aib-Aib-Aib-Aib-Gly-Leu-Aib-AibOMe (I), which is an analogue of the N-terminal sequence of antiamoebins and emerimicins, establishes a completely 310-helical conformation with seven successive intramolecular 4 → 1 hydrogen bonds. The average, ?, ψ values for residues 1–8 are ?59° and ?32°, respectively. Crystal parameters are C47H77N9O12, space group P1, a = 10.636(4) Å, b = 11.239(4) Å, c = 12.227(6) Å, α = 101.17(4)°, β = 97.22(4)°, γ = 89.80(3)°, Z = 1, R = 5.95% for 3018 data with |F0| > 3α(F), resolution 0.93 Å. The use of the torsion angle κ = C(i ? 1)N(i)Cα(i)Cβ(i), where κ = 68° for D -Phe and κ = 164° for L -Leu, confirms the opposite configurations of these residues. The ?, ψ values of ?62° and ?32° at D -Phe are unusual, since this region is characteristic of residues with L configurations. Peptide I possesses only two chiral residues of opposing configuration. The observed right-handed 310-helical structure suggests that helix sense has probably been determined by the stereo-chemical preferences of the Leu residue. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
In the acridine orange–dermatan sulfate system, free and bound dye can be distinguished from each other spectroscopically. This permits the use of fluorometric methods to study the binding of acridine orange to the acid mucopolysaccharide dermatan sulfate. Experiments were conducted at 24°C in 10?3 M citrate/phosphate buffer at pH = 7.0. The binding of the dye is highly cooperative, as evidenced by considerable interaction between adjacent bound dye molecules. Analysis of the data indicates that dermatan sulfate binds 2.3 ± 0.3 mol of acridine orange per dermatan sulfate uronic acid residue with a cooperative binding constant, Kq ranging from 4.9 to 6.0 × 105 M?1 which corresponds to a free energy of 7.74 ? ΔG° ? 7.86. The cooperativity parameter q apparently increases with increasing polymer-to-dye ratio.  相似文献   

13.
The three-dimensional structure was determined by x-ray crystallography for d(T[p](CE)T), a uv photoproduct of the cyanoethyl (CE) derivative of d(TpT), having the cis-syn cyclobutane (CB) geometry and the S-configuration at the chiral phosphorus atom. The crystals of C23H30N5O12P · 2H2O belong to the orthorhombic space group P212121 (Z = 4), with cell dimensions a = 11.596 Å, b = 14.834 Å, and c = 15.946 Å, containing two water molecules per asymmetric unit. The CB ring is puckered with a dihedral angle of 151°. The two pyrimidine bases are rotated by –29° from the position of direct overlap of their corresponding atoms. This represents a major distortion of DNA, since in DNA adjacent thymines are rotated by +36°. The pyrimidine rings are puckered with Cremer–Pople parameters for T[p] and in parentheses [p]T: Q: 0.24 Å (0.31 Å); θ: 123° (120°); ?: 141° (86°). These represent half-chairs designated as 6H1 (T[p]) and 6H5 ([p]T). The CB and pyrimidine ring conformations are interrelated, and we postulate that they execute a coupled interconversion in solution. The T[p] segment has the syn glycosyl conformation, a 2T3 sugar pucker, and gauche? conformation at C4′-C5′; the [p]T segment is anti, 3T4, trans. The C5′-O5′ torsion of the [p]T unit is –124.5°, and the C3′-O3′ torsion of the T[p] unit is –152.9°. Bond angles and bond lengths involving the phosphorus atom are similar to those of other phosphotriesters. The P-O3′ and P-05′ torsion angles are –138.1° and 58.6°, respectively. Several intermolecular (but no intramolecular) hydrogen bonds are found in the crystal.  相似文献   

14.
Many of the theoretical methods used for predicting the occurrence of α-helices in peptides are based on the helical preferences of amino acid monomer residues. In order to check whether the helix-forming tendencies are based on helical preferences of monomers only or also on their sequence contexts, we synthesized permuted sequences of the tripeptides GAP, GAV, and GAL that formed crystalline helices with near α-helical conformation. The tripeptides AFG and FAG formed good crystals. The x-ray crystallographic studies of AFG and FAG showed that though they contain the same amino acids as GAF but in different sequences, they do not assume a helical conformation in the solid state. On the other hand, AFG and FAG, which contain the same amino acids but in a different sequence, exhibit nearly the same backbone torsion angles corresponding to an incipient formation of a β-bulge, and exhibit nearly identical unit cells and crystal structures. Based on these results, it appears that the helix-forming tendencies of amino acids depend on the sequence context in which it occurs in a polypeptide. The synthetic peptides AFG (L -Ala-L -Phe-Gly) and FAG (L -Phe-L -Ala-Gly), C14H19N3O4, crystallize in the orthorhombic space group P212121, with a = 5. 232(1), b = 14. 622(2), c = 19. 157(3) Å, Dx = 1.329 g cm?3, Z = 4, R = 0.041 for 549 reflections for AFG, and with a = 5. 488(2), b = 14.189 (1), c = 18.562(1) Å, Dx = 1.348 g cm?3, Z = 4, R = 0.038 for 919 reflections for FAG. Unlike the other tripeptides GAF, GGV, GAL, and GAI, the crystals of AFG and FAG do not contain water molecule, and the molecules of AFG or FAG do not show the helical conformation. The torsion angles at the backbone of the peptide are ψ1 = 144. 5(5)°; ?2, ψ2 = ?98.1(6)°, ?65.2(6)° ?3, ψ13, ψ31 = 154.1(6)°, ?173.6(6)°, 6.9(8)° for AFG; and ψ1 = 162.6(3)°; ?2, ψ2 = ?96.7(4)°, ?46.3(4)°; ?3, ψ13, ψ31 = 150.1(3)°, ?168.7(3)°, 12.2(5)° for FAG. The conformation angles (? ψ) for residues 2 and 3 for both AFG and FAG show incipient formation of an β-bulge. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
Y C Fu  H V Wart  H A Scheraga 《Biopolymers》1976,15(9):1795-1813
The enthalpy change associated with the isothermal pH-induced uncharged coil-to-helix transition ΔHh° in poly(L -ornithine) in 0.1 N KCl has been determnined calorimetrically to be ?1530 ± 210 and ?1270 ± 530 cal/mol at 10° and 25°C, respectively. Titration data provided information about the state of charge of the polymer in the calorimetric experiments, and optical rotatory dispersion data about its conformation. In order to compute ΔHh°, the observed calorimetric heat was corrected for the heat of breaking the sample cell, the heat of dilution of HCl, the heat of neutralization of the OH? ion, and the heat of ionization of the δ-amino group in the random coil. The latter was obtained from similar calorimetric measurements on poly(D ,L -ornithine). Since it was discovered that poly(L -ornithine) undergoes chain cleavage at high pH, the calorimetric measurements were carried out under conditions where no degradation occurred. From the thermally induced uncharged helix–coil transition curve for poly(L -ornithine) at pH 11.68 in 0.1 N KCl in the 0°–40°C region, the transition temperature Ttr and the quantity (?θh/?T)Ttr have been obtained. From these values, together with the measured values of ΔHh°, the changes in the standard free energy ΔGh° and entropy ΔGh°, associated with the uncharged coil-to-helix transition at 10°C have been calculated to be ?33 cal/mol and ?5.3 cal/mol deg, respectively. The value of the Zimm–Bragg helix–coil stability constant σ has been calculated to be 1.4 × 10?2 and the value of s calculated to be 1.06 at 10°C, and between 0.60 and 0.92 at 25°C.  相似文献   

16.
The structure of tetra-O-methyl- (+) -catechin has been determined in the crystalline state. Two independent molecules, denoted structure A and structure B, exist in the unit cell. Crystals are triclinic, space group P1, a = 4.8125(2) Å, b = 12.9148(8) Å, c = 13.8862(11) Å, α = 86.962(6) °, β = 89.120(5)°, γ = 88.044(5)°, Z = 2, Dc = 1.336 g cm?3, R = 0.033 for 6830 observations. The heterocyclic rings of the crystal structures are compared to previous results for 8-bromotetra-O-methyl-(+)-catechin, penta-O-acetyl-(+)-catechin, and (?) -epicatechin. One of the two molecules has a heterocyclic ring conformation similar to that observed previously for (?)-epicatechin, and the other has a heterocyclic ring conformation similar to one predicted earlier in a theoretical analysis of dimers of (+)-catechin and (?) -epicatechin. Both structure A and structure B in the crystal have heterocyclic ring conformations that place the dimethoxyphenyl substituent at C(2) in the equatorial position. However, this heterocyclic ring conformation does not explain the proton nmr coupling constant measured in solution. Molecular dynamics simulations show an equatorial ? axial interconversion of the heterocyclic ring, which can explain the nmr results. © 1993 John Wiley & Sons, Inc.  相似文献   

17.
Intrinsic viscosities of cyclic and linear lamda DNA   总被引:3,自引:0,他引:3  
The ratio of the intrinsic viscosities of the linear and circular forms of λ DNA, [η]L /[η]c, has been measured as a function of ionic strength in the range [Na+] = 0.6. M–0.03MCorrections were made for the presence of uncyclizable linear contaminant in circular preparations. By combining data in the literature on the ionic strength dependence of linear DNA of various molecular weights with that obtained here, it was possible to determine the expansion parameter εL as a function of [Na+]. εL is defined by the relation 〈L2〉 = b2N1+εL, where 〈L1〉 is the mean-square end-to-end distance of a chain of N segments of length b. The empirical relation εL = 0.05 ? 0.11 log [Na+] for native NaDNA at 25°C is found. When εL = 0, [η]L /[η]c extrapolates to 1.6, in good agreement with the theoretical prediction of 1.55. As εL increases, [η]L /[η]c increases, in agreement with a theory of Bloomfield and Zimm.  相似文献   

18.
Abstract

Phase transitions were studied of the sodium salt of poly(rA) ?poly(rU) induced by elevated temperature without Ni2+ and with Ni2+ in 0.07 M concentration in D2O (~0.4 [Ni]/[P]). The temperature was varied from 20° C to 90° C. The double-stranded conformation of poly(rA)?poly(rU) was observed at room temperature (20° C—23° C) with and without Ni2+ ions. In the absence of Ni2+ ions, partial double- to triple-strand transition of poly(rA) ?poly(rU) occurred at 58° C, whereas only single-stranded molecules existed at 70° C. While poly(rU) did not display significant helical structure, poly(rA) still maintained some helicity at this temperature. Ni2+ ions significantly stabilized the triple-helical structure. The temperature range of the stable triple-helix was between 45° C and 70° C with maximum stability around 53° C. Triple-to single-stranded transition of poly(rA) ?poly(rU) occurred around 72° C with loss of base stacking in single-stranded molecules. Stacked or aggregated structures of poly(rA) formed around 86° C. Hysteresis took place in the presence of Ni2+ during the reverse transition from the triple-stranded to the double-stranded form upon cooling. Reverse Hoogsteen type of hydrogen-bonding of the third strand in the triplex was suggested to be the most probable model for the triple-helical structure. VCD spectroscopy demonstrated significant advantages over infrared absorption or the related electronic CD spectroscopy.  相似文献   

19.
Abstract

The effect of pH and temperature on the apparent association equilibrium constant (Ka) for the binding of the bovine and porcine pancreatic secretory trypsin inhibitor (Kazal-type inhibitor, PSTI) to human leukocyte elastase has been investigated. At pH8.0, values of the apparent thermodynamic parameters for human leukocyte elastase: Kazal-type inhibitor complex formation are: bovine PSTT – Ka = 6.3 × 104M?1, δ5G° = -26.9kJ/mol, δH° = +11.7kJ/mol, and δS° = +1.3 × 102 entropy units; porcine PSTI –Ka = 7.0 × 103M?1,δG° = -21.5kJ/mol, δH° = +13.0kJ/mol, and δS° = +1.2 × 102 entropy units (values of Ka δG° and δS° were obtained at 21.0°C; values of δH° were temperature independent over the range (between 5.0°C and 45.0°C) explored). On increasing the pH from 4.5 to 9.5, values of Ka for bovine and porcine PSTI binding to human leukocyte elastase increase thus reflecting the acidic pK-shift of the His57 catalytic residue from ?7.0, in the free enzyme, to ?5.1, in the serine proteinase: inhibitor complexes. Thermodynamics of bovine and porcine PSTI binding to human leukocyte elastase has been analyzed in parallel with that of related serine (pro)enzyme/Kazal-type inhibitor systems. Considering the known molecular models, the observed binding behaviour of bovine and porcine PSTI to human leukocyte elastase was related to the inferred stereochemistry of the serine proteinase/inhibitor contact region(s).  相似文献   

20.
β-N-Acetvlhexosaminidase (EC 3.2.1.52) was purified from the liver of a prawn, Penaeus japonicus, by ammonium sulfate fractionation and chromatography with Sephadex G-100, hydroxylapatite, DEAE-Cellulofine, and Cellulofine GCL-2000-m. The purified enzyme showed a single band keeping the potential activity on both native PAGE and SDS–PAGE. The apparent molecular weight was 64,000 and 110,000 by SDS–PAGE and gel filtration, respectively. The pI was less than 3.2 by chromatofocusing. The aminoterminal amino acid sequence was NH2-Thr-Leu-Pro-Pro-Pro-Trp-Gly-Trp-Ala-?-Asp-Gln-Gly-VaI-?-Val-Lys-Gly-Glu-Pro-. The optimum pH and temperature were 5.0 to 5.5 and 50°C, respectively. The enzyme was stable from pH 4 to 11, and below 55°C. It was 39% inhibited by 10mM HgCl2.

Steady-state kinetic analysis was done with the purified enzyme using N-acetylchitooligosaccharides (GlcNAcn, n = 2 to 6) and p-nitrophenyl N-acetylchitooligosaccharides (pNp-β-GlcNAcn, n= 1 to 3) as the substrates. The enzyme hydrolyzed all of these substrates to release monomeric GlcNAc from the non-reducing end of the substrate. The parameters of Km and kcat at 25°C and pH 5.5 were 0.137 mM and 598s–1 for pNp-β-GlcNAc, 0.117 mM and 298s–1 for GlcNAc2, 0.055 mM and 96.4s–1 for GlcNAc3, 0.044 mM and 30.1 s–1 for GlcNAc4, 0.045 mM and 14.7 s–1 for GlcNAc5, and 0.047 mM and 8.3 s–1 for GlcNAc6, respectively. These results suggest that this β-N-acetylhexosaminidase is an exo-type hydrolytic enzyme involved in chitin degradation, and prefers the shorter substrates.  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