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 187 毫秒
1.
Conductance and relaxations of gelatin films in glassy and rubbery states   总被引:1,自引:0,他引:1  
The dielectric constant, ′, and the dielectric loss, ″, for gelatin films were measured in the glassy and rubbery states over a frequency range from 20 Hz to 10 MHz; ′ and ″ were transformed into M* formalism (M*=1/(′−i″)=M′+iM″; i, the imaginary unit). The peak of ″ was masked probably due to dc conduction, but the peak of M″, e.g. the conductivity relaxation, for the gelatin used was observed. By fitting the M″ data to the Havriliak–Negami type equation, the relaxation time, τHN, was evaluated. The value of the activation energy, Eτ, evaluated from an Arrhenius plot of 1/τHN, agreed well with that of Eσ evaluated from the DC conductivity σ0 both in the glassy and rubbery states, indicating that the conductivity relaxation observed for the gelatin films was ascribed to ionic conduction. The value of the activation energy in the glassy state was larger than that in the rubbery state.  相似文献   

2.
This report describes the effect of temperature on the mechanical viscoelastic properties such as: storage modulus (E′), loss modulus (E″), and loss tangent (tan δ) of the collagen sponges modified with hyaluronic acid (HA). In order to detect collagen–HA copolymer denaturation and to assess its thermal stability, the differential scanning calorimetry (DSC) supplemented by thermogravimetric (TG) measurements was used. The denaturation temperature (Td) of unmodified collagen samples increased from 69 to 86 °C for cross-linked samples, respectively. These temperature dependencies show remarkable changes in E′ and E″ at selected temperature up to 226 °C for all samples due to the release of loosely and strongly bound water. The influence of HA on the viscoelastic behavior of collagen is manifested by a shift of the tan δ peak associated with the process of decomposition towards higher temperatures resulting in a higher thermo-stability of the modified scaffolds.  相似文献   

3.
The shear piezoelectricity was observed in oriented films of poly-β-hydroxybutyrate (PHB) and copolymers of β-hydroxybutyrate (HB) and β-hydroxyvalerate (HV). The piezoelectric stress constant 314 = e14ie14 (polarization/strain), the piezoelectric strain constant d14 = d14id14 (polarization/stress), the elastic constant c = c′ + ic″ and the dielectric constant = ′ − i″ were determined at a frequency of 10 Hz over a temperature range from −150° to +150°C. Piezoelectric relaxations as well as elastic and dielectric relaxations were clearly observed at the glass transition temperature of about 15°C. In order to evaluate the piezoelectric constants (e2 and d2) for the piezoelectric phase which consists of the crystalline region and the oriented non-crystalline region, a spherical dispersion two phase model was utilized. Assuming the appropriate fixed values for the elastic and dielectric constants in the piezoelectric phase, d2 and d2 were calculated as a function of temperature. For a PHB and a copolymer (17 HV/83 HB), e2 and d2 showed relaxations, leading to a conclusion that the instantaneous piezoelectric constant in the crystalline phase is constant independent of temperature but the piezoelectric constant in the oriented non-crystalline phase is relaxational and has the opposite sign. For a copolymer (25 HV/75 HB) and a chloroform treated copolymer (17 HV/83 HB), e2 and d2 were constant independent of temperature, indicating that the oriented non-crystalline phase has disappeared owing to the increased molecular flexibility due to copolymerization or annealing in chloroform vapour.  相似文献   

4.
Cross-linked waxy maize (CWM) starch dispersions (STDs) of concentration 50 g kg−1 were heated in sucrose solutions containing 0–600 g kg−1 (g sucrose/kg dispersion) at 85 °C at low shear and in intermittently agitated cans at 110 °C. The STDs heated in 0–300 g kg−1 sucrose exhibited antithixotropic behavior, while those heated in 400–600 g kg−1 sucrose exhibited thixotropic behavior. The mean starch granule diameter of the starch dispersions did not show strong dependence on sucrose concentration. The dispersions, especially those with high sucrose concentrations and heated at 110 °C, exhibited G′ versus frequency (ω) profiles of gels. The STDs exhibited first normal stress differences that increased in magnitude with the concentration of sucrose. Values of the first normal stress coefficient of canned dispersions calculated from dynamic rheological data plotted against ω and experimental values plotted against shear rate of some of the STDs overlapped.  相似文献   

5.
Density functional theory (DFT) computations at the B3LYP/Lanl2DZ level were used to elucidate the oxygen atom transfer (OAT) and coupled electron proton transfer (CEPT) reaction steps involved in the biomimetic catalytic cycle performed by polymer-supported MoVIO2(NN′)2 complexes [NN′ = phenyl-(pyrrolato-2-ylmethylene)-amine] with water as oxygen source, trimethyl-phosphane as oxygen acceptor and one-electron oxidising agents. The DFT method employed has been validated against experimental data [X-ray crystal structures of a NN′ ligand and a MoVIO2(NN′)2 complex as well as kinetic data]. The rate-limiting step in the forward-OAT from [MoVIO2] to PMe3 is the attack of PMe3 at an oxo ligand with ΔG (298 K) = 64.6 kJ mol−1. Dissociation of the product OPMe3 is facile with ΔG (298 K) = 26.3 kJ mol−1 giving a mono-oxo [MoIVO] complex which fills its coordination sphere with a further PMe3 substrate with ΔG (298 K) = 39.2 kJ mol−1. One-electron oxidation to a Mo(V) phosphane complex precedes the coordination of water/hydroxide. Additionally, the comproportionation of [MoVIO2] and [MoIVO] to dinuclear oxo-bridged [OMoV–O–MoVO] species has been calculated as the thermodynamic sink in this system and the back-OAT from dmso to mono-oxo [MoIVO] to give [MoVIO2] has been shown to involve an equilibrium between stereoisomeric [MoVIO2] complexes with an activation barrier of ΔG (298 K) = 113.1 kJ mol−1.  相似文献   

6.
[MnL](ClO4)2 (L = N,N′,N″-tris(2-hydroxypropyl)-1,4,7-triazacyclononane) has been tested for catalyzing sulfide oxidation. In the presence of this complex, ethyl phenyl sulfide, butyl sulfide and phenyl sulfide are completely oxidized to the corresponding sulfoxides and sulfones with H2O2 as the oxidant. 2-Chloroethyl phenyl sulfide oxidation yield 2-chloroethyl phenyl sulfone and phenyl vinyl sulfone. In ethyl phenyl sulfide oxidation, effects of complex and H2O2 concentration and temperature on the reaction rate have been discussed. Through controlling reaction conditions, ethyl phenyl sulfoxide and ethyl phenyl sulfone may be produced selectively. The UV–Vis and electron paramagnetic resonance (EPR) studies on catalyst solution indicate that metal centre of the complex is transformed from Mn(II) to Mn(IV) after the addition of H2O2. At 25 °C, rate constant for ethyl phenyl sulfide oxidation is 4.38 × 10−3 min−1.  相似文献   

7.
The reaction of thiamine with K2PtIICl4 and with PtIVCl4 in the presence of excess NaSCN in aqueous solution gave thiamine salts, (H-thiamine)[Pt(SCN)4] · 3H2O (1) and (H-thiamine)[Pt(SCN)6] · H2O (2), respectively, structures of which have been determined by X-ray diffraction. The thiamine molecule adopts the usual F conformation in each salt. In 1, [Pt(SCN)4]2− ions act as large planar spacers in the crystal lattice and interact scarcely with thiamine, except for a hydrogen bonding with the terminal hydroxy O(5γ). Instead, water molecules form two types of host–guest-like interactions with the pyrimidine and the thiazolium moieties of a thiamine molecule, one being a C(2)–Hwaterpyrimidine bridge and the other being an N(4′)–Hwaterthiazolium bridge. In 2, despite the much larger ion size, octahedral [Pt(SCN)6]2− ions form a C(2)–Hanionpyrimidine bridge and an N(4′)–Hanionthiazolium bridge. An additional hydrogen bonding between the anion and the terminal O(5γ) of thiamine creates a hydrogen-bonded macrocyclic ring {thiaminium–[Pt(SCN)6]2−}2, a supramolecule.  相似文献   

8.
The conformation of red bean globulin dispersions (≈10% in D2O or deuterated phosphate buffer pD 7.4) under the influence of pH, chaotropic salts, protein structure perturbants, and heating conditions was studied by Fourier-transform infrared (FTIR) spectroscopy. The FTIR spectrum of red bean globulin showed major bands from 1682 to 1637 cm−1 in the amide I′ region, corresponding to the four types of secondary structures, i.e. β-turns, β-sheets, -helix and random coils. At extreme pH conditions, there were changes in intensity in bands attributed to β-sheet (1637 and 1618 cm−1) and random coil (1644 cm−1) structures, and shifts of these bands to lower or higher wavenumbers, indicating changes in protein conformation. Chaotropic salts caused progressive increases in random coil structures and concomitant decreases in β-sheet bands, following the lyotrophic series of anions. In the presence of sodium dodecyl sulfate and ethylene glycol, pronounced increases in the random coil band were observed, accompanied by slight shifts of the β-sheet band. Addition of dithiothreitol and N-ethylmaleimide did not cause marked changes in the FTIR spectra. Heating at increasing temperature led to progressive decreases in the intensity of the -helix and β-sheet bands and increases in random coil band intensity, leveling off at around 60 °C. The data suggest that re-organization of protein structure occurred at temperatures well below the denaturation temperature of red bean globulin (86 °C) as determined by differential scanning calorimetry. This was accompanied by pronounced increases in the intensity of the two intermolecular β-sheet bands (1682 and 1619–1620 cm−1) associated with the formation of aggregated strands at higher temperatures (80–90 °C). Increases in intensity of the aggregation bands were also observed in the heat-induced buffer-soluble and insoluble aggregates.  相似文献   

9.
A method for determination of carboxymethyl cellulose (CMC) molecular weight (MW) and chemical heterogeneity (degree of oxidation (DO)) using a bi-detector HPSEC (UV-detector online with refractometer) has been developed. It has been found that the use of 0.5 N NaOH or 0.4 M acetate buffer as the eluent ensures CMC separation according to MW. It has been revealed that the universal calibration for the polyelectrolyte CMC and the neutral polymer dextran is valid under the conditions applied. The Mark–Houwink equations for CMC in 0.5 N NaOH and 0.4 M acetate buffer have been estimated to be [η]=5.37×10−4 MW0.73 and [η] =2.24×10−4 MW0.83 dl g−1, respectively. The equation log K=1.64−4.00 ml g−1 for CMC has been estimated. An approach for determining DO from adsorption at 290 or 313 nm has been developed.  相似文献   

10.
A [CoIII(N2S2)]NEt4 complex, with two carboxamido nitrogens and two alkylthiolato sulfurs, was prepared from N,N′-(2-thioacetyl-isobutyryl)-2-aminobenzylamine, and characterized. It crystallizes with a distorted square planar structure including two short Co–N bonds (≈1.882 Å) and two short Co–S bonds (≈2.134 Å). The ligand defines an 11-atom chelate, which may be Co ligands in the mean plane of Co-containing nitrile hydratase. The CoIII oxidation state, reversibly reduced at −1.13 V (vs. SCE) and irreversibly oxidized at +1.29 V (vs. SCE) in DMF, is stable over a 2 V potential range. From the temperature dependence of its magnetic susceptibility, cobalt(III) was found to be in an S=1 triplet ground state, in agreement with the broad resonances observed in its 1H-NMR spectrum. Preliminary spectral studies showed that this complex does not interact with imidazole, H2O or HO, but binds two CN anions or two NO molecules. The IR spectrum of the dinitrosyl complex exhibits two NO stretches at 1765 and 1820 cm−1, in the range previously observed for dinitrosylated complexes derived from cobalt(I). This result suggests that, similarly to Fe NHases, Co NHases might readily bind NO.  相似文献   

11.
Synthesis of chitosan sulfates with low molecular weight (Mv 9000–35,000 Da) was carried out by sulfation of low molecular weight chitosan (Mv 10,000–50,000 Da). The oleum was used as sulfating agent and dimethylfornamide as medium. The chitosans were prepared by enzymatic and acidic hydrolysis of initial high molecular weight chitosan as well as by extrusion solid-state deacetylation of chitin. As was shown by FT-IR and NMR-methods and elemental analysis, the sulfation occurred at C-6 and C-3 positions and substitution degree is 1.10–1.63. The molecular weight sulfated chitosan was determined by viscometric method and the Mark–Houwink equation [η]=10−5 4.97 M0.77. Study of anticoagulant activity showed that chitosan sulfates with lowered molecular weight demonstrated a regular increase of anti-Xa activity like heparins.  相似文献   

12.
Several constitutive models have been discussed to explain data for some foods in diluted and concentrated systems. Firstly, the theories of Rouse and Zimm, as well as rod-like theory, were used to study the conformation of the pectins in dilute solution. Among the dilute theories, the random coil theory of Zimm best explained the experimental data and suggested a certain level of intermolecular interaction present in the dilute pectin solution.

The Bird-Carreau constitutive theory with four empirical constants and zero shear limiting viscosity was used to describe the viscoelastic properties of the solutions of the guar, CMC/guar, glutenin, gluten and wheat flour doughs. The Bird-Carreau model was able to predict η and η′ in the high and low frequency regions for 1% guar solution. In the case of CMC/guar blend, the Bird-Carreau model explained steady shear and dynamic properties very well in the higher shear rate or frequency region of 1–100 s−1. However, η″/ω does not tend to a zero shear constant value. The Bird-Carreau model also gave good predictions on the rheological properties of gluten and glutenin biopolymers in the free-flow region.

The polydisperse type, Doi-Edwards model, fits the experimental G′ and G″ better than the monodisperse model for 5% apple pectin dispersion. However, there is still a discrepancy between experimental and predicted values.  相似文献   


13.
The ability of the yeast Issatchenkia orientalis CECT 10688 to secrete lipolytic activity in submerged culture was investigated. The yeast was grown in a complex medium supplemented with a fixed concentration of several lipidic compounds (triglycerides, fatty acids). Maximum enzyme activity around 70–80 U cm−3 was produced in cultures supplemented with tributyrin. An optimum tributyrin concentration of 10 g dm−3 was selected. Several surfactants were added to the cultures, but no significant increase in activity was detected. Finally, the effect of the type of carbon source on lipolytic enzyme production was studied. The best results were obtained with glucose and fructose (60–80 U cm−3), while rather low enzyme activity was found in cultures grown on lactose and maltose (about 20 U cm−3).  相似文献   

14.
The interaction between Ac-AMP2, a lectin-like small protein with antimicrobial and antifungal activity isolated from Amaranthus caudatus, and N,N′,N″-triacetyl chitotriose was studied using 1H NMR spectroscopy. Changes in chemical shift and line width upon increasing concentration of N,N′,N″-triacetyl chitotriose to Ac-AMP2 solutions at pH 6.9 and 2.4 were used to determine the interaction site and the association constant Ka. The most pronounced shifts occur mainly in the C-terminal half of the sequence. They involve the aromatic residues Phe18, Tyr20 and Tyr27 together with their surrounding residues, as well as the N-terminal Val-Gly-Glu segment. Several NOEs between Ac-AMP2 and the N,N′,N″-triacetyl chitotriose resonances are reported.  相似文献   

15.
Mixed gels of κ-carrageenan (κ-car) from Hypnea musciformis and galactomannans (Gal) from Cassia javanica (CJ) and locust bean gum (LBG) were compared using dynamic viscoelastic measurements and compression tests. Mixed gels at 5 g/l of total polymer concentration in 0.1 M KCl showed a synergistic maximum in viscoelastic measurements for κ-car/CJ and κ-car/LBG at 2:1 and 4:1 ratios, respectively. The synergistic maximum obtained from compression tests carried out for mixed gels at 10 g/l of total polymer concentration in 0.25 M KCl was the same for both κ-car/CJ and κ-car/LBG gels. An enhancement in the storage modulus (G′) and the loss modulus (G″) was observed in the mechanical spectra for the mixtures in relation to κ-car. The proportionally higher increase in G″ compared with G′, as indicated by the values of the loss tangent (tan δ), suggests that the Gal adhere non-specifically to the κ-car network.  相似文献   

16.
A detailed study of the interaction of chitin molecular species with the solvent system N,N-dimethylacetamide (DMAc)/lithium chloride (LiCl) allowed the development of a new method for chitin fractionation by coacervate extraction. The controlled increase of the extracting power of the solvent was carried out using slight modification of the solvent composition. Partial extractions of molecular species were done between coacervation and complete dissolution limits using different mixtures of DMAc/LiCl of increasing extracting power. Fractions were characterized in DMAc/LiCl 5% (w/w) by viscometry and size exclusion chromatography with refractive index and multi-angle laser light scattering detectors. Fractions obtained by coacervate extraction range from 80,000 to 710,000 g mol−1 with polydispersity index between 1.28 and 1.44. The Mark–Houwink–Sakurada equation constants a and K for chitin in DMAc/LiCl 5% (w/w) were found to be 0.95 and 7.6×10−5 dl g−1, respectively.  相似文献   

17.
Bacillus species carrying subC gene encoding serine alkaline protease (SAP) enzyme were developed in order to increase the yield and selectivity in the bioprocess for SAP production. For this aim, subC gene was cloned into pHV1431 Escherichia coliBacillus shuttle vector, and transferred into nine host Bacillus species, i.e. B. alvei, B. amyloliquefaciens, B. badius, B. cereus, B. coagulans, B. firmus, B. licheniformis, B. sphaericus and B. subtilis. The influence of the host Bacillus species on SAP production on a defined medium with glucose was investigated in bioreactor systems. For each of the recombinant (r-) Bacillus species, effects of initial glucose concentration on cell growth and SAP production were investigated; and, physiological differences and similarities between the wild-type and r-Bacillus species are discussed. The highest biomass concentration was obtained with r-B. coagulans as 3.8 kg m−3 at the initial glucose concentration of CGo=20 kg m−3 and the highest volumetric SAP activity was obtained with r-B. amyloliquefaciens as 1650 U cm−3 at CGo=20 kg m−3. Overall SAP activity per amount of substrate consumed was the highest for r-B. sphaericus (137 U g−1 cm−3) and r-B. licheniformis (130 U g−1 cm−3). Among the r-Bacillus species the highest activity increase compared to the wild types was obtained with r-B. sphaericus while the lowest increase was obtained with r-B. amyloliquefaciens and r-B. licheniformis due to high SAP production potential of the wild-type strains. During storage of the host microorganisms, r-B. alvei and r-B. amyloliquefaciens were not able to bear the recombinant plasmid, probably, due to the restriction enzymes synthesized. Due to the highest stable volumetric activities r-B. licheniformis (950 U cm−3) and r-B. sphaericus (820 U cm−3) appear to be the favorable hosts for the production of SAP. All the r-Bacillus species excreted organic acids oxaloacetic and succinic acids, but, none excreted the amino acid valine. The variations in by-product distributions with each recombinant organism were also discussed.  相似文献   

18.
The reaction of N,N,N′,N′-tetrakis(2-pyridylmethyl)ethylenediamine (tpen) with VCl3 in CH3CN yields Cl3V(tpen)VCl3 which was hydrolyzed in water in the presence of oxygen affording [V2O2(μ-OH)2(tpen)]I2·2H2O, the crystal structure of which has been determined. Asyn-{OV(μ-OH)2VO}2+ core has been identified where the V(IV) centers are antiferromagnetically coupled (J = −150 cm−1;g = 1.80).  相似文献   

19.
Human fibroblast culture on a crosslinked dermal porcine collagen matrix   总被引:1,自引:0,他引:1  
The use of a novel porcine-derived collagen biomaterial as a dermal tissue engineering matrix was examined. The matrix is derived from porcine dermis, and is processed to retain the native collagen (Type 1) and elastin structure. Human primary fibroblasts were cultured on the matrix to examine its potential for creating a dermal replacement. Attachment of fibroblasts on the collagen was compared to tissue culture plastic and PET membranes. Cell proliferation was assessed using the MTT assay and DAPI staining. For seeding densities of 5×104 and 1×105 cells cm−2, PET and plastic demonstrated >95% attachment of seeded numbers after 3 h. The collagen matrix reached levels >80% after 3–4 h with no influence of the seeding density. Matrix samples with perforating pores of 40 μm diameter were also studied. After 216 h culture in static culture, with media replacement every 3 days, the final cell numbers reached 2.1×105 (perforated) and 2.0×105 cells cm−2 (unperforated). In comparison fibroblast culture in a perfusion bioreactor, with continuous media replacement, reached 2.3×105 (unperforated) and 2.5×105 cells cm−2 (perforated) after 216 h.  相似文献   

20.
Because of its novel bioactive properties the production of gymnodimine for use as a pharmaceutical precursor has aroused interest. The dinoflagellate, Karenia selliformis produces gymnodimine when grown in bulk culture using GP + selenium medium but the growth rates (μ) and levels of gymnodimine are low (μ, 0.05 days−1; gymnodimine 250 μg L−1 max). We describe the effects of organic acid additions (acetate, glycolate, alanine and glutamate additions and combinations of these) in enhancing growth and gymnodimine production in axenic cultures. The most effective organic acid combinations in decreasing order were: glycolate/alanine > acetate > glycolate. Glycolate/alanine optimised gymnodimine production by prolonging growth (maximum cell yield, 1.76 × 105 cells mL−1; gymnodimine, 1260 μg L−1; growth rate (μ), 0.2 days−1) compared to the control (growth maximum cell yield, 7.8 × 104 cells mL−1; gymnodimine, 780 μg L−1; μ, 0.17 days−1). Acetate enhanced gymnodimine by stimulating growth rate (μ, 0.23 days−1) and the large concentration of gymnodimine per cell (16 pg cell−1 cf. 9.8 pg cell−1 for the control) suggests a role for this compound in gymnodimine biosynthesis. Amending culture media with Mn2+ additions resulted in slightly decreased growth in control cultures and increased the gymnodimine while in glycolate/alanine cultures growth was stimulated but gymnodimine production decreased. The results suggest that the organic acid can enhance gymnodimine production by either enhancing growth maximum or the biosynthetic pathway.  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