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.
This study aimed to evaluate the ability of commercial soy protein isolate (SPI) to form cold-set gels under different pHs (5–11), pre-heating temperatures (60 °C, 80 °C), CaCl2 (0–15 mM) and SPI (5–15%, w/v) concentrations, and also select a formulation for the investigation of the effects of incorporating locust bean gum (LBG) (0–0.3%, w/v) and solid lipid microparticles (SLM) on gels rheological and microstructural properties. Gels were evaluated in terms of visual aspect, water-holding capacity, microstructure (using confocal laser scanning microscopy and cryo-scanning electronic microscopy) and rheological properties. SPI showed higher solubilities at pHs 7 (32.0%), 9 (51.6%) and 11 (100%). Self-supported gels were obtained under several conditions at alkaline pHs. At pH 7, only systems pre-heated to 80 °C with 15% (w/v) SPI and 10 or 15 mM CaCl2 gave self-supported gels. At neutral pH, samples showed relative structural instability, which was minimized with LBG incorporation. Formulations GSPI (pH 7, preheated to 80 °C, 15% (w/v) SPI, 10 mM CaCl2) and GMIX (pH 7, preheated to 80 °C, 15% (w/v) SPI, 0.2% (w/v) LBG, 15 mM CaCl2) were selected for emulsion-filled gels (EFG) production. Power law parameters (K′, K″), calculated from frequency sweep results, revealed that non-filled GMIX (K′: 472.1; K″: 77.6) was stronger than GSPI (K′: 170.4; K″: 33.6). Besides, GMIX showed microphase separation. SLM stabilized with Tween 80-Span 80 were active fillers in EFG, altering microstructures and increasing G’, G” and the Young’s modulus (1.8 to 2.1 kPa for GSPI and 1.4 to 2.2 kPa for GMIX).  相似文献   

2.
In the present work, we carried out density functional calculations of struvite – the main component of the so-called infectious urinary stones – to study its structural and elastic properties. Using a local density approximation and a generalised gradient approximation, we calculated the equilibrium structural parameters and elastic constants C ijkl . At present, there is no experimental data for these elastic constants C ijkl for comparison. Besides the elastic constants, we also present the calculated macroscopic mechanical parameters, namely the bulk modulus (K), the shear modulus (G) and Young's modulus (E). The values of these moduli are found to be in good agreement with available experimental data. Our results imply that the mechanical stability of struvite is limited by the shear modulus, G. The study also explores the energy-band structure to understand the obtained values of the elastic constants.  相似文献   

3.
Abstract

The impact of intramolecular stereoelectronic effects has been examined by comparison of the solution structures of natural oligo-DNA duplex, 5′(1C2G3C4G5A6A7T8T9C10G11C12G)2 3′, and its carbocyclic-nucleotide analogues in which the pentose sugar in 2′-dA residue is replaced with its carbocyclic counterpart (i.e. 2′-deoxyaristeromycin). Based on the NMR evidences, it has been shown, that 2′-deoxyaristeromycin analog exists in a dynamic equilibrium between the two forms of duplexes, one with W-C bp and the second with Hoogsteen bp in ca 1:1 ratio at lower temperature (below 35°C) and as hairpin at higher temperature (from ~40° – 60°C).  相似文献   

4.
The dynamics of the opening-closing of the constituent base-pairs as well as of the exchange kinetics of the base-paired imino and amino protons with water in a DNA-RNA hybrid, [5′r(G1A2U3U4G5A6A7)3′]:5′p[d(T8C9A10A11T12C13)]3′-Pzn] duplex (I), are reported here in details for the first time. The exchange kinetics of amino and imino protons in the DNA-RNA hybrid (duplex I) have been compared with identical studies on the following B-DNA duplexes: d(C1G2T3A4C5G6)2 (II), d[p(5′T1G2T3T4T5G6 G7C8)3′]:d[p(5′C9C10A11A12A13C14A15)3′] (III), d(C5G6C7G8A9A10T11T12C13G14C15G16)2 (IV) and d(C1G2C3G4C5G6C7G8A9A10T11T12C13G14C15G16C17G18C19G20)2 (V). This comparative study shows that the life-times τo of various base-pairs in the DNA-RNA hybrid (I) varies in the range of ∼ 1 ms, and they are quite comparable to those of the shorter B-DNA duplexes (II) and (III), but very different from the τo of the larger duplexes (IV) and (V): the τo for the base pair of T11 and T12 residues in the 20-mer (duplex V) are 2.9 ± 2.3 ms and 23.2 ± 8.9 ms, respectively, while the corresponding τo in the 12-mer (duplex IV) are 2.8 ± 2.2 ms and 17.4 ± 5.4 ms. It has also been shown that the total energy of activation (Ea) assessed from the exchange rates of both imino and amino protons, representing energetic contributions from both base-pair and helix opening-closing as well as from the exchange process of the imino protons from the open state with the bound water, is close to the Ea of the short B-DNA duplex (Ea ≈ 28–47 kcal/mol).  相似文献   

5.
The structural, elastic and electronic properties of Co7M6 (M?=?W, Mo, Nb) μ phases were investigated by first-principles calculations based on the density functional theory (DFT). The calculated cohesive energy indicates that Co7M6 (M?=?W, Mo, Nb) μ phases are thermodynamically stable. Besides, Co7W6 owns a higher structural stability than that of Co7Mo6 and Co7Nb6. The obtained elastic constant demonstrates that Co7M6 (M?=?W, Mo, Nb) are mechanically stable. With Voigt-Reuss-Hill (VRH) approximation, the elastic bulk modulus (B), shear modulus (G), Young's modulus (E) and Poisson's ratio (ν) were derived. The ductility and plasticity as well as the elastic anisotropy of the three phases were discussed in details. Finally, the density of states and charge density difference were also analysed to reveal the underlying mechanism of structural stability and the elastic properties.  相似文献   

6.
The equilibrium structure, elastic constants Cij and thermodynamic functions of cubic titanium nitride (TiN) were calculated within the temperature range of 0–3100 K and under a pressure range 0–60 GPa. Properties were computed using the generalised gradient approximations (GGA) exchange-correlation functional. Calculated mechanical properties (Elastic constants, Young’s modulus and shear modulus) and phonon spectra of TiN obtained via robust DFT-QHA algorithm, were generally in a good agreement with available experimental and theoretical analogous values. In particular, a well-examined quasi-harmonic approximation method implemented in the Gibbs2 code is utilised herein to provide accurate estimation of thermal expansion coefficients, entropies, heat capacity values (at different combinations of temperature/volume/pressure) and Debye’s temperature. Parameters calculated herein shall be useful to elucidate the superior performance of TiN at harsh operational conditions encompassing elevated temperatures and pressures pertinent to cutting machineries and surface coatings.  相似文献   

7.
The interaction between 3‐spiro‐2′‐pyrrolidine‐3′‐spiro‐3″‐piperidine‐2,3″‐dione (PPD) and bovine serum albumin (BSA) in aqueous solution was studied using fluorescence and UV–vis spectroscopy. Fluorescence emission data revealed that BSA (1.00 × 10‐5 mol/L) fluorescence was statically quenched by PPD at various concentrations, which implies that a PPD–BSA complex was formed. The binding constant (KA), the number of binding sites (n) and the specific binding site of the PPD with BSA were determined. Energy‐transfer efficiency parameters were determined and the mechanism of the interaction discussed. The thermodynamic parameters, ΔG, ΔH and ΔS, were obtained according to van't Hoff's equation, showing the involvement of hydrophobic forces in these interactions. The effect of PPD acting on the BSA conformation was detected by synchronous fluorescence. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

8.
Abstract

The non-exchangeable 1H-NMR signals of the branch core trinucleotide of the lariat branch site (A2′p5′G 3′p5′C), 1) and its derivatives 2 and 3 are completely assigned using one- and two- dimensional NMR techniques including NOE, COSY, NOESY, 1H-1HINADEQUATE and 2D-J-resolved spectroscopy. From the vicinal coupling constants in the individual ribose rings, NOE data and T1 measurements, the following properties of the trimers are deduced.(i)The unique stacking behavior of the trimers is S1′N 3′N, and the sugar rings exist predominantly in the N-conformation (3′-endo-2′-exo).(ii)The sugar-base orientations appear to be anti.(iii) The branched trimers exist in solution as single-stranded right-handed conformations resembling A-RNA with stacking between the adenine and guanine residues in aqueous solution at 21°C and pH 7.2.(iv) The calculated values for the torsion angles εt andγ+ for the trimers are 201–203° and 71–86%, respectively, while the percent β1 values are higher for the guanine (87–92%) than the cytosine residues (73–77%). The computer generated depiction of the triribonucleotide 1 is also shown. These subtle structural features may act as recognition signals for this critical lariat branch site which is essential for the second step in yeast mRNA splicing.  相似文献   

9.
Abstract

Synthetic oligoribonucleotides (RNA) are efficiently prepared with 2′-O-tert-butyldimethylsilyl nucleoside 3′-O-phosphoramidites with labile base-protection; Admf or APac, Gdmf, Cibu, U. After cleavage from the polystyrene support, the exocyclic amine protecting groups are removed with conc. NH4OH: ethanol/3:1 by heating at 55°C for 3–5 h. The 2′-O- silyl protecting groups are removed with tetra-n-butylammonium fluoride in THF or more conveniently with neat triethylamine trihydrofluoride. To gain the advantages of increased capacity on reverse phase HPLC and the convenience of cartridge based purification (OPC, Oligonucleotide Purification Cartridge), the 5′ trityl was left on the RNA as the final protecting group to be removed. The mild conditions which are effective for trityl removal are shown to preserve 3′-5′ phosphate linkage integrity in RNA. The absence of phosphate migration is demonstrated by model studies, utilizing N4 -isobutyryl-5′-O-DMT-3′-O-TBDMS-2′-O-(2-cyanoethyl-N,N-diisopropylphosphoramidite) as a control monomer and digestion by 3′-5′ selective P1 nuclease and alkaline phosphatase and HPLC analysis. Oligoribonucleotides were analyzed by Microgel capillary electrophoresis, anion-exchange HPLC, and the enzymatic digest/HPLC method.

  相似文献   

10.
Abstract

ID NOE 1H NMR spectroscopy at 500 MHz was employed to examine the structure of poly(dA)·poly(dT) in solution. NOE experiments were conducted as a function of presaturation pulse length (50, 30, 20 and 10 msec) and.power (19 and 20 db) to distinguish the primary NOEs from spin diffusion. The 10 msec NOE experiments took 49 hrs and over 55,000 scans for each case and the difference spectra were almost free from diffusion.

The spin diffused NOE difference spectra as well as difference NOE spectra in 90% H2O + 10% D2O in which TNH3 was presaturated enabled to make a complete assignment of the base and sugar protons. It is shown that poly(dA) ·poly(dT) melts in a fashion in which single stranded bubbles are formed with increasing temperature.

Extremely strong primary NOEs were observed at H2′/H2″ when AH8 and TH6 were presaturated. The observed NOEs at AH2′ and that AH2″ were very similar as were the NOEs at TH2′ and TH2″. The observed NOEs at AH2′ and AH2″when AH8 was presaturated were very similar to those observed at TH2′ and TH2″ when TH6 was presaturated. In addition, presaturation of H1′ of A and T residues resulted in similar NOEs at AH2′/H2″ and TH2′/H2″ region and these NOEs at H2′ and H2″ were distinctly asymmetric as expected in a C2′-endo sugar pucker. There was not a trace of NOE at AH8 and TH6 when AH3′ and TH3′ were presaturated indicating that C3′-endo, × = 30–40° conformation is not valid for this DNA. From these NOE data, chemical shift shielding calculations and stereochemistry based computer modellings, we conclude that poly(dA)·poly(dT) in solution adopts a right- handed B-DNA duplex in which both dA and dT strands are conformationally equivalent with C2′-endo sugar pucker and a glycosyl torsion, ×, of ?73°, the remaining backbone torsion angles being φ′ = 221°, ω′ = 212°, ω = 310°, φ = 149°, ψ = 42°, ψ′ = 139°. The experimental data are in total disagreement with the heteronomous DNA model of Arnott et. al. proposed for the fibrous state. (Arnott, S., Chandrasekaran, R., Hall, I.H., and Puigjaner, L.C., Nucl. Acid Res. 11, 4141, 1983).  相似文献   

11.
An engineered reversal of the β-oxidation cycle was exploited to demonstrate its utility for the synthesis of medium chain (6–10-carbons) ω-hydroxyacids and dicarboxylic acids from glycerol as the only carbon source. A redesigned β-oxidation reversal facilitated the production of medium chain carboxylic acids, which were converted to ω-hydroxyacids and dicarboxylic acids by the action of an engineered ω-oxidation pathway. The selection of a key thiolase (bktB) and thioesterase (ydiI) in combination with previously established core β-oxidation reversal enzymes, as well as the development of chromosomal expression systems for the independent control of pathway enzymes, enabled the generation of C6–C10 carboxylic acids and provided a platform for vector based independent expression of ω-functionalization enzymes. Using this approach, the expression of the Pseudomonas putida alkane monooxygenase system, encoded by alkBGT, in combination with all β-oxidation reversal enzymes resulted in the production of 6-hydroxyhexanoic acid, 8-hydroxyoctanoic acid, and 10-hydroxydecanoic acid. Following identification and characterization of potential alcohol and aldehyde dehydrogenases, chnD and chnE from Acinetobacter sp. strain SE19 were expressed in conjunction with alkBGT to demonstrate the synthesis of the C6–C10 dicarboxylic acids, adipic acid, suberic acid, and sebacic acid. The potential of a β-oxidation cycle with ω-oxidation termination pathways was further demonstrated through the production of greater than 0.8 g/L C6–C10 ω-hydroxyacids or about 0.5 g/L dicarboxylic acids of the same chain lengths from glycerol (an unrelated carbon source) using minimal media.  相似文献   

12.
The influence of one DNA region on the stability of an adjoining region (telestability) was examined. Melting curves of three block DNA's, d(C15A15)·d(T15G15), d(C20A15)·d(T15G20), and d(C20A10)·d(T10G20) were analyzed in terms of the nearest neighbor Ising model. Comparisons of predicted and experimental curves were made in 0.01 M and 0.1 M sodium ion solutions. The nearest neighbor formalism was also employed to analyze block DNA transition in the presence of actinomycin, a G·C specific molecule. The results show that nearest neighbor base-pair interaction cannot predict the melting curves of the block DNA's. Adjustments in theoretical parameters to account for phosphate repulsion assuming a B conformation throughout the DNA's do not alter this conclusion. Changes in the theoretical parameters, which provide good overall agreement, are consistent with a substantial stabilization of the A·T region nearest the G·C block. The melting temperature T A·T for the average A·T pari in d(C20A10)·d(T10G20), with 10 A·T pairs, appears to be 4°C greater than TA·T for d(C15A15)·d(T15G15) and d(C20A15)·d(T15G20), both with 15 A·T pairs. Actinomycin bound to the G·C end effectively stabilizes the A·T end by 9°C. These results indicate a long-range contribution to the interactions governing DNA stability. A possible mechanism for these interactions will be discussed.  相似文献   

13.
The effects of heating temperature on gel properties and conformational changes were investigated in glycinin and β-conglycinin gels using Theological and Fourier transform ir (FTIR) methods. Solutions of 15 wt % glycinin or β-conglycinin in 35 mM phosphate buffer at pH 7.6 were heated at various temperatures for 30 min and rheological properties were measured at 20°C. The storage modulus G′ as a function of frequency changed from a monotonical decrease with decreasing frequency to a plateau in the range from 0.0018 to 40 Hz by heating at temperatures higher than 80°C for glycinin and 65°C for β-conglycinin. A band at 1618 cm?1 (associated with the β-sheet structure) on ir spectra increased with the formation of heat-induced gels. The value of the storage modulus G′ correlated well with the increase in absorbance at 1618 cm?1. These results suggest that the formation of a β-sheet structure may be closely related to the value of the storage modulus G′ for heat-induced gels in soybean proteins and that heat-induced gels of glycinin and β-conglycinin are formed by cross-links with intermolecular β-sheet structures. © 1994 John Wiley & Sons, Inc.  相似文献   

14.
Double-stranded synthetic polydeoxynucleotides of the general form poly[d(GnCn)] · poly[d(GnCn)], poly[d(GnC)] · poly[d(GCn)], and poly[d(AnTn)] · poly[d(AnTn)] have been synthesized. When n = 4 or larger, the CD spectra of polymers of the form poly[d(GnCn)] · poly[d(GnCn)] or poly[d(GnC)] · poly[d(GCn)] closely resemble the spectrum of poly[dG] · poly[dC], suggesting that a string of four continguous guanosine residues is sufficient to induce a conformation resembling that of the polypurine · polypyrimidine. With polymers of the form poly[d(AnTn)] · poly[d(AnTn)], however, the CD spectrum only gradually approaches that of poly[dA] · poly[dT].  相似文献   

15.
A putative carotenoid oxygenase from Novosphingobium aromaticivorans was purified with a specific activity of 0.8?U/mg by His-Trap affinity chromatography. The native enzyme was estimated to be a 52?kDa monomer. Enzyme activity for β-apo-8′-carotenal was maximal at pH 8.0 and 45?°C, with a half life of 15.3?h, K m of 21?μM, and k cat of 25?l/min. The enzyme exhibited cleavage activity only for carotenoids containing one β-ionone ring and its catalytic efficiency (k cat/K m) followed the order β-apo-8′-carotenal?>?β-apo-4′-carotenal?>?γ-carotene. The enzyme converted these carotenoids to β-apo-13-carotenones by cleaving their C13–C14 double bonds. The oxygen atom of β-apo-13-carotenone originated not from water but from molecular oxygen. Thus, the enzyme was an apo-carotenoid 13,14-dioxygenase.  相似文献   

16.
The mechanical properties of gelatin films were studied in relation to the effect of water, and compared with those of collagen films. The S-shaped sorption isotherm was separated into an adsorption curve C1 and dissolution curve C2. From the C2 curve, the interaction parameter χ1 of Flory–Huggins' equation was calculated. The χ1 of gelatin were larger than those of collagen at low relative humidities (RH), while they coincided with each other at high RH. It was found that a composite curve was made by shifting stress relaxation curves obtained at different humidities along the log time axis. The shift factor for the formation of the composite curve was analyzed by Fujita–Kishimoto's equation, which was based on the free volume theory. The parameter β, which expressed the extent of the contribution of sorbed water to the increase in the free volume of the system, was 0.05 in the range of C2 from 0 to 0.08 (0–65% RH). This value was much smaller than 0.16 for collagen. The value was 0.16 in the range of C2 higher than 0.08, which was equal to that of the collagen. Dynamic shear modulus G′, loss modulus G″, and tan δ were determined as functions of RH. The gelatin film extended more than 100% at 73% RH under the very small stress of about 107 dyn/cm2. This corresponds to the region where β changes from 0.05 to 0.16, although such a phenomenon was not observed in the collagen film. The wide-angle X-ray pattern of extended gelatin was similar to that of renatured collagen fiber.  相似文献   

17.
ABSTRACT

Pressure dependence of stability, phonon, Debye temperature, physical, mechanical and thermodynamic properties of Rh3Al intermetallic compound were investigated by first-principles The calculated cohesive energy (Ec), formation enthalpy (ΔH) show that Rh3Al is a thermodynamically stable compound. Properties related to the phonons of Rh3Al were also obtained. In addition, the transverse sound velocity (νs), longitudinal sound velocity (νl), average sound velocity (νm) and Debye temperature (ΘD) of Rh3Al were calculated by using the VRH method along with pressure range from 0 to 60?GPa. The values of lattice parameters, bulk modulus and its first-order pressure derivative are consistent well with other works. The band structure indicates that Rh3Al compound exhibits a metallic character. Moreover, the total density of states, partial density of states, Mulliken charges and electron density difference have been analysed to explain the physical properties. Based on the stress–strain approach and the Born stability criteria, the mechanical properties were evaluated by elastic constants (Cij), other modulus (B, E, G), (B/G) ratio, Poisson’s ratio (ν), the anisotropic index (A), hardness (H) and compressibility (K) for this intermetallic compound. Finally, the thermodynamic properties, including enthalpy, free energy, entropy and heat capacity are discussed range from 0 to 1000?K.  相似文献   

18.
We used a combination of eddy flux, chamber and environmental measurements with an integrated suite of models to analyse the seasonality of net ecosystem carbon uptake (FCO2) in an 8-year-old, closed canopy Pinus radiata D.Don plantation in New Zealand (42°52′ S, 172°45′ E). The analyses utilized a biochemically based, big-leaf model of tree canopy photosynthesis (Ac), coupled to multiplicative environmental-constraint functions of canopy stomatal conductance (Gc) via environmental measurements, a temperature-dependent model of ecosystem respiration (Reco), and a soil water balance model. Available root zone water storage capacity at the measurement site is limited to about 50 mm for the very stony soil, and annual precipitation is only 660 mm, distributed evenly throughout the year. Accordingly the site is prone to soil moisture deficit throughout the summer. G c and Ac obtained maximum rates early in the growing season when plentiful soil water supply was associated with sufficient quantum irradiance (Qabs), and moderate air saturation deficit (D) and temperature (T). From late spring onwards, soil water deficit and D confined Gc and Ac congruously, which together with the solely temperature dependency of Reco resulted in the pronounced seasonality in FCO2. Reflecting a light-limitation of Ac in the closed canopy, modelled annual carbon (C) uptake was most sensitive to changes in Qabs. However, Qabs did not vary significantly between years, and changes in annual FCO2 were mostly due to variability in summer rainfall and D. Annual C-uptake of the forest was 717 g C m–2 in a near-average rainfall year, exceeding by one third the net uptake in a year with 20% less than average rainfall (515 g C m–2).  相似文献   

19.
Molecular-mechanics calculations have been carried out on the base-paired hexanucleoside pentaphosphates d(TATATA)2, d(ATATAT)2, d(A6)·d(T6), d(CGCGCG)2, d(GCGCGC)2, and d(C6)·d(G6) in both A- and B-DNA geometries. The calculated relative energies of these polymers are consistent with the relative stabilities of the polymers found experimentally. In particular, the results of our calculations support the observation that the homopolymer d(A)n·d(T)n is more stable in a B-DNA conformation, while the homopolymer d(G)n·d(C)n is more stable in an A-DNA conformation. The molecular interactions responsible for these differential stabilities include both inter- and intrastrand base stacking, as well as base–phosphate interactions. While definitive experiments on the heteropolymer stabilities have not yet been carried out, the results of our calculations also suggest a greater stability of the purine-3′,5′-pyrimidine sequence over the pyrimidine-3′,5′-purine sequence in both the A- and B-conformations. The reason for this greater stability lies in the importance of the inherent directionality (5′ → 3′ vs 3′ → 5′) of phosphate–base and base–base interactions. The largest conformation change observed on energy refinement is sugar repuckering, which occurs mainly on pyrimidine-attched sugars and only in the B-DNA geometry. We suggest a molecular mechanism, specifically, differential base–sugar steric interactions involving neighboring sugars, to explain why this repuckering occurs more with d(A6)·d(T6) than with other isomers.  相似文献   

20.
The proliferation kinetics of cells of the line NHIK 1922 grown in vitro and as solid tumours in the athymic mutant nude mouse has been studied. In vitro, growth curves were determined for exponentially growing populations and for populations synchronized by mitotic selection. The phase durations for these populations were determined by flow cytofluorometric measurements of DNA-histograms and pulsed incorporation of [3H]TdR respectively. The generation time and the phase durations for synchronized populations were found to be about equal to those for exponentially growing populations. The duration of the phases G1, S and G2+ M was found to be 8·5–9·5, 11·0–12·0 and 6·0–6·5 hr respectively, i.e. the generation time was 26·5–27·0 hr. The proliferation kinetics in vivo were studied by flow cytofluorometry and by the technique of percentage labelled mitoses. The median duration of S-phase and (G2+ M)-phase in vivo was found to be approximately the same as that observed in vitro, while the median duration of G1-phase was found to be approximately 5 hr longer in vivo than under the present in vitro growth conditions. The growth fraction in vivo was estimated to be approximately 50%. The non-proliferative compartment of the tumour cells was found to consist mainly of cells with the DNA-content of cells in G1-phase. It is concluded that the reduced rate of proliferation of NHIK 1922 cells in vivo is correlated with alterations in the duration of G1-phase and, hence, the proportion of cells in G1-phase.  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