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.
As a prelude to studies of co-gelation with galatin, the gelation behaviour of Paselli maltodextrins SA-6 and SA-2 (DE ≈ 6 and 2, respectively) was mapped out over the experimentally-accessible range of temperature (T) and concentration (c), using a simple visual method to determine the time required for formation of a self-supporting network (tg). For both samples, log tg decreased linearly with log c and increased linearly with T. At equivalent temperatures and concentrations, SA-2 gelled between 20 and 60 times faster than SA-6.

Selected samples were monitored more rigorously by mechanical spectroscopy, taking tg as the time at which elastic response (G′) became greater than viscous response (G″). In all cases the values of tg obtained by this procedure were lower than those from visual inspection, by a constant factor of about 3·4.

The concentration-dependence of gel moduli (G′) for SA-2 and for gelatin (second-extract limed ossein; LO-2) fitted accurately to the form anticipated from cascade theory for normal polymer networks. For SA-6, by contrast, log G′ varied linearly with log c over the entire range at which measurements could be made, indicating a different mechanism of structure-formation (such as the agglomeration of short, aggregated helices).  相似文献   


2.
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   

3.
The relationship between the height of gels determined by a sag test and their elastic shear modulus (G′) has been both investigated experimentally and simulated using a finite element analysis for the inhomogeneous deformation of gels due to gravity. It was assumed in the simulations that gels can be modeled as incompressible linear elastic materials. General relationships between the sag of gels and their elastic modulus were obtained from the simulations for slip and no-slip conditions. The relationships were tested experimentally on pectin, gelatin and polyacrylamide gels with a range of concentrations and rigidities. The good agreement between the predictions and the results shows that these gels can be modeled accurately as incompressible elastic materials. A standard 150° SAG pectin gel, which sags 23.5% in the SAG test, has G′ moduli of 429 and 379 Pa under slip and no-slip conditions, respectively.  相似文献   

4.
Co-gels of potato maltodextrin Paselli SA-6 with gelatin were prepared by rapid quenching of mixed solutions from 90°C. At fixed setting temperature and fixed concentration of gelatin, the time required to form a self-supporting network showed an initial steady decrease with increasing concentration of SA-6 (as expected from polymer exclusion), but then increased dramatically before again decreasing. The interpretation of this behaviour as phase inversion from a gelatin-continuous network with SA-6 inclusions to a (more slowly-forming) SA-6 network with gelatin inclusions was confirmed by differential scanning calorimetry (showing both components melting separately, with no evidence of specific interaction), mechanical spectroscopy (showing that the mixed gel network was destroyed completely by melting of the gelatin component at low concentrations of SA-6, but only weakened at SA-6 concentrations above the inversion point) and by light microscopy (showing the expected changes in distribution of the two polymers).

In similar studies using the faster-gelling potato maltodextrin Paselli SA-2, microscopy and gel-melting profiles again showed phase-inversion from a gelatin-continuous network at low concentrations of SA-2 to a maltodextrin-continuous network at higher concentrations. Inversion, however, occurred at a lower concentration of maltodextrin than in the gelatin/SA-6 systems, and the accompanying change in gelation rate was confined to a sharp decrease in the dependence of gel-time on SA-2 concentration.  相似文献   


5.
The effect of thermodynamic incompatibility in mixed solutions of gelatin and Paselli maltodextrins SA-6 and SA-2 has been studied at a temperature (45°C) where the individual polymers are stable as disordered coils. Concentrated mixtures of SA-6 and gelatin showed classic phase separation into two co-existing liquid layers, with compositions lying along a well-defined binodal. On decreasing SA-6 concentration below the binodal, however, a substantial proportion (up to 60%) of the maltodextrin was precipitated, with normal single-phase solutions occurring only at much lower concentrations of both polymers. SA-2 showed a more extreme version of the same behaviour, with precipitation of up to 100% of the maltodextrin and no evidence of co-existing liquid phases at any accessible concentrations. In both cases, the amount of maltodextrin precipitated was proportional to the square of its initial concentration and to the first power of gelatin concentration, indicating that gelatin drives self-association and aggregation of maltodextrin when both polymers are present in a single liquid phase. 1H NMR showed the precipitated maltodextrin to be higher in molecular weight and in degree of branching than the material remaining in solution, and particlesize analysis indicated that the volume of the individual maltodextrin particles increased linearly with the total mass precipitated.  相似文献   

6.
In a continuation of the work reported in the preceding paper, the effect of chicory root inulin (Mr≈4.5 kD) and branched dextrans of Mr 67, 464 and 2000 kD on gelation of low methoxy pectin (DE 31; 2.0 wt%; pH≈2.9–3.0) on cooling from 90 to 5 °C in the presence of stoichiometric Ca2+ has been characterised by low amplitude oscillatory measurements of G′ and G″. As found for other polymeric cosolutes studied previously (oxidised starch, potato maltodextrin, guar gum and locust bean gum), increasing concentrations of dextran or inulin caused a progressive increase of G′ and G″ in the solution state at 90 °C (attributed to segregative interactions promoting formation of calcium-mediated ‘egg-box’ junctions between pectin chains) and a progressive reduction in final moduli at 5 °C (attributed to excessive association of calcium pectinate into large aggregated bundles), with cooling curves crossing those for calcium pectinate alone at 55 °C. For the dextran samples, the ability to promote initial association of pectin chains at high temperature decreased with increasing molecular weight (i.e. with the smallest molecules having the greatest effect, as was also found for guar gum), but the reduction in final modulus on completion of cooling to 5 °C became more pronounced. By the criterion of the decrease in the change in log G′ during cooling, the three dextran samples studied seem roughly comparable in their overall effectiveness in promoting self-association of pectin, with inulin having a substantially smaller effect. Both materials, however, are less effective than the partially depolymerised starches and galactomannans studied previously. It is tentatively suggested that this may be associated with the flexibility introduced by the presence of three single bonds in the predominant glycosidic linkages of both dextran and inulin.  相似文献   

7.
8.
9.
Recent studies on dilute solutions (Borgström et al. (1996), Int. J. Biol. Macromol. 18, 223) have shown that kappa-carrageenan helices associate into superhelical rigid rods in mixed 0.1 M aqueous solutions of NaI and CsI above a critical mole fraction (xCs = 0.4) of Cs. This work concerns the temperature-dependent rheology of more concentrated systems in mixed and pure solutions of the same salts. Gels with low moduli were even found in NaI alone, although this salt is known to impede the gelation of kappa-carrageenan, but only above 0.9% (w/w) of carrageenan. These gels were reminiscent of iota-carrageenan gels in two respects: the (low) magnitude of the shear storage modulus (G′), and the absence of hysteresis in the sol-gel transition. On the other hand, both the threshold concentration for gelation and the ratio between the loss and storage moduli were substantially higher for the kappa-carrageenan gels in NaI. In mixed solutions of CsI and NaI, two types of kappa-carrageenan gels could be distinguished, depending on the cesium content. The transition occurred at xCs = 0.4, as in the previous studies on dilute solutions. Below xCs = 0.4, the gels were similar to those in NaI alone. Above xCs = 0.4, the gels were similar to ‘conventional’ kappa-carrageenan gels, formed in salts such as KC1: a pronounced thermal hysteresis appeared in the sol-gel transition, the gels showed tendencies for syneresis, and G′ increased dramatically with increasing cesium content.  相似文献   

10.
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   

11.
Water-soluble (WSP) and insoluble non-starch polysaccharides (WIP) were isolated from wheat flour to evaluate the effects of WSP and WIP on starch gel properties. Isolated WSP and WIP were added to two types of isolated wheat starch with different amylose content at a concentration of 3% based on the dry weight of starch. 30% starch gels were prepared and stored at 5 °C for 1, 8, or 24 h. The dynamic viscoelasticity of 30% starch gels mixed with WSP and WIP was measured using parallel plate geometry, showing that WSP and WIP affected the elastic component of starch gels in opposite ways. Adding WIP increased the storage shear modulus (G′) of starch gels, while adding WSP decreased G′ and dramatically increased the loss tangent (tan δ=G″/G′).  相似文献   

12.
Products released by Luclia cuprina (sheep blowfly) during egg hatch and early larval development were found to contain a variable number of proteases, probably reflecting their different functions during these developmental phases. Moreover, a number of the developmentally regulated proteases produced only during egg hatch appeared to be strongly egg shell associated, indicating the egg shell proteins to be their specific substrates. Several proteases were tested for their ability to enhance the rate of egg hatch. The commercial proteases chymotrypsin and trypsin were able to significantly enhance (p≤0.01) egg hatch by 36% and 44%, respectively, when compared to an untreated control. Interestingly, fluids containing the egg shell associated proteases were able to significantly (p≤0.001) enhance L. cuprina egg hatch by up to 70%. This hatch enhancement activity could be significantly (p≤0.001) reversed by the addition of Phenylmethylsulphonyl fluoride (PMSF). Fractionation of the egg-associated proteases by gelatin affinity chromatography suggested that gelatin-binding molecules were responsible for the majority of the egg hatch enhancement activity in this preparation.  相似文献   

13.
The acylation of hydrazine, to afford the N,N′-diacyl derivatives, was catalysed by a number of lipases. The rates of the first and second steps depended on the lipase and the type of solvent used. Water, up to 0.4 M, had no detrimental effect on the yield and complete conversion to the N,N′-diacyl derivative was accomplished with some lipases. The hydrazide of 2-(4-isobutylphenyl)propanoic acid (ibuprofen), prepared by non-enzymatic reaction of ibuprofen methyl ester with hydrazine, acted as a nucleophile towards several lipases that do not accept ibuprofen derivatives as the acyl donor, but the enantiomer differentiation was inefficient in most cases. The best result was obtained with Pseudomonas lipoprotein lipase on EP 100 which formed the (R) enantiomer of the product (N-octanoyl-N′-2-(4-isobutylphenyl)propanoylhydrazine) with an enantiomeric ratio E of 26.  相似文献   

14.
Addition of gum arabic (average Mr≈450 kDa; 0.5–2.0 wt%) to solutions of low methoxy pectin (DE 31; 2.0 wt%; pH≈2.9–3.0) with stoichiometric Ca2+ caused massive increases in G′ and G″ in the pre-gel state at 90 °C (attributed to segregative interactions promoting formation of calcium-mediated ‘egg-box’ junctions between pectin chains) but had little effect on the gels formed on cooling to 5 °C. This is in marked contrast to the behaviour of other polymeric cosolutes studied in the investigations reported in the two preceding papers, which caused large reductions in gel moduli (attributed to excessive association of calcium pectinate into large aggregated bundles); the difference is tentatively ascribed to strengthening of the calcium pectinate network by divalent counterions to the uronate residues in gum arabic. When the complication of cation exchange was eliminated by extensive dialysis of gum arabic against 100 mM Na+ and use of the final dialysate in preparation of mixtures with calcium pectinate, massive increases in G′ and G″ at high temperature were again observed, but with accompanying reductions in moduli at low temperature, which, at gum arabic concentrations above 1.0 wt%, arose from collapse of the developing calcium pectinate network during cooling. The tentative conclusion from this work, and from the two preceding papers, is that enthalpically unfavourable (segregative) interactions between low methoxy pectin and polymeric cosolutes can be relieved in two ways: (i) Ca2+-mediated self-association of pectin into compact ordered assemblies which occupy less of the total volume, and (ii) conformational rearrangement of the cosolute molecules to minimise segmental interactions with pectin; conformational rearrangement is inhibited by chain stiffness and by branching; thus polymeric cosolute molecules of limited flexibility are more effective in promoting self-association of pectin than more flexible molecules of comparable size, and branched molecules are more effective than linear chains of comparable stiffness.  相似文献   

15.
The viscoelastic properties of aqueous solutions of the exocellular polysaccharide of Cyanospira capsulata have been studied, over a wide range of polymer concentrations, using small deformation oscillatory, steady and transient shear methods. The viscoelastic spectra generally resemble those of an entangled network, although notable deviations can be observed in the low frequency dependence of G′ and G″. At higher polymer concentrations, the viscoelastic spectrum shows solid-like behaviour over a wide range of frequencies. The superposition of η*(ω) and η( ) curves occurs only at low frequencies, at higher frequencies the slope of η*(ω) is lower than that of η( ). By studying the time evolution of shear stress after the inception of a steady shear rate (stress overshoot), the recovery of non-linear properties after steady shearing flow is seen to occur after times of c. 103 s (in the case of 1·1% w/v solutions).

The overall viscoelastic properties appear original in comparison with those of the two structurally limiting types of polysaccharide, the ‘ordered’ chain xanthan and the ‘random coil’ guar. A rationale for this ‘anomalous’ viscoelastic behaviour can be tentatively proposed in terms of flickering intermolecular cross-interactions between semi-flexible segments, which occur in addition to the usual topological constraints.  相似文献   


16.
Enantiomerically pure alkylphosphonate compounds RR′P(O)PNP (R=CnH2n+1, R′=OY with Y=CnH2n′+1 with n=n′ or nn′; PNP=p-nitrophenoxy) noted (RY), mimicking the transition state occurring during the carboxyester hydrolysis were synthesized and investigated as potential inhibitors of human gastric lipase (HGL) and human pancreatic lipase (HPL). The inhibitory properties of each enantiomer have been tested with the monomolecular films technique in addition to an enyzme linked immunosorbent assay (ELISA) in order to estimate simultaneously the residual enzymatic activity as well as the interfacial lipase binding. With both lipases, no obvious correlation between the inhibitor molar fraction (50) leading to half inhibition, and the chain length, R or Y was observed. (R11Y16)s were the best inhibitor of HPL and (R10Y11)s were the best inhibitors of HGL. We observed a highly enantioselective discrimination, both with the pure enantiomeric alkylphosphonate inhibitors as well as a scalemic mixture. We also showed, for the first time, that this enantioselective recognition can occur either during the catalytic step or during the initial interfacial adsorption step of the lipases. These experimental results were analyzed with two kinetic models of covalent as well as pseudo-competitive inhibition of lipolytic enzymes by two enantiomeric inhibitors.  相似文献   

17.
Differential scanning calorimetry (DSC), rheological measurements and granule size analyses were performed to characterize the influence of phosphorylation substitution levels on the properties of cross-linked potato starch. Phosphorus oxychloride (POCl3) was used to produce the cross-linked potato starch. The levels of the reagent used for the reaction ranged between 40 and 5000 ppm (dwb). Storage (G′) and loss (G″) moduli were measured for a 5% (w/w) gelatinized starch dispersion stored at 20 °C for 24 h after heating at 85 °C for 30 min. The samples from 80 to 500 ppm were recognized as ‘strong gel'systems, whereas native potato starch showed ‘weak gel'behavior. Steady shear and dynamic viscoelastic properties of gelatinized starch dispersion were compared. Furthermore, granule mean diameter was measured by laser scattering for a 1% (w/w) dispersion heated at 85 °C for 30 min. The granules in the 100 ppm sample swelled to a maximum of about 2.6 times the native starch granule mean diameter.  相似文献   

18.
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   

19.
As an artificial ligand, a glycoconjugate polymer carrying carbohydrate moiety of lactosyl ceramide or globotriaosyl ceramide (Gb3) was synthesized. Gb3 is known as the receptor of Shiga toxin-producing Escherichia coli O157: H7. The preparation of the glycoconjugate polymer initially involves the construction of the carbohydrate moiety of Gb3 derivative which has n-pentenyl group as polymerizable group. In addition, the n-pentenyl group of the Gb3 derivative was modified and different polymerizable groups such as acrylamide group were introduced at ω-position of the aglycon. Radical polymerization of the synthesized glycosyl monomers with or without acrylamide proceeded smoothly in water using ammonium persulfate and N, N, N′, N′-tetramethylethylenediamine as usual initiator system and gave water-soluble glycoconjugate polymers having various polymer compositions. These polymers have the potential to neutralize Shiga toxin by reason of cluster effect and multivalency.  相似文献   

20.
1,10-Phenanthroline-5,6-dione (C12H6N2O2 (1)) reacts with V(η6-mesitylene)2 and Ti(η6-toluene)2 affording coordination compounds of general formula M(O,O′---C12H6N2O2)3 (M=Ti (2); M=V (3)) which further react with TiCl4 or TiCp2(CO)2 yielding the tetrametallic species M(O,O′---C12H6N2O2---N,N′)3(M′Ln)3 (M=V, M′Ln=TiCl4 (4); M=Ti, M′Ln=TiCp2 (5); M=V, M′Ln=TiCp2 (6)). The complex salt [Fe(N,N′---C12H6N2O2)3][PF6]2 (7) has been obtained from iron(II) chloride tetrahydrate and 1 in the presence of NH4PF6. The reaction of 7 with TiCp2(CO)2 affords the tetrametallic derivative [Fe(N,N′---C12H6N2O2---O,O′)3(TiCp2)3][PF6]2 (8). TiCl2(THF)2 reacts with MCp2(O,O′---C12H6N2O2) to give MCp2(O,O′---C12H6N2O2---N,N′)TiCl2 (M=Ti (9); M=V (10)). By reaction of TiCp2(O,O′---C12H6N2O2---N,N′)TiCl2 (9) with C12H6N2O2, the bimetallic derivative TiCp2(O,O′---C12H6N2O2---N,N′)TiCl2(O,O′---C12H6N2O2) (11) has been prepared, which readily adds to TiCl4, to give the trimetallic titanium derivative TiCp2(O,O′---C12H6N2O2---N,N′)TiCl2(O,O′---C12H6N2O2---N,N′)TiCl4 (12). VCp2(O,O′---C12H6N2O2---N,N′)TiCl2 (10) reacts with the tris-chelate iron(II) cation 7 affording the heptametallic cationic complex [Fe(N,N′---C12H6N2O2---O,O′)TiCl2(N,N′---C12H6N2O2---O,O′)VCp2]3 +2 isolated as the hexafluorophosphate 13.  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