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 555 毫秒
1.
The physical–chemical properties, fatty acid composition and thermal properties of goat subcutaneous (SF), tallow (TF) and intestinal (IF) fats were determined. SF differed from other fat types with respect to its lower melting (41.6 °C), lower saponification (190.3 mg KOH/g) and higher iodine (40.4) values as compared to those of other fats. Goat fat types contained palmitic acid (C16:0), stearic acid (C18:0), oleic acid (C18:1ω9) and linoleic acid (C18:2ω6) as the major components of the fatty acid composition (23.06–23.52, 22.95–39.03, 21.94–36.16 and 1.96–2.22%, respectively). A differential scanning calorimetry (DSC) study revealed that two characteristic peaks were detected in both crystallization and melting curves. Major peaks (T peak) of TF and IF were similar and determined as 34.02–35.24 and 9.95–10.72 °C, respectively for the crystallization peaks and 15.11–18.26 and 50.70–52.76 °C, respectively for the melting peaks in the DSC curves; but those of SF (27.14 and 4.36 °C for crystallization peaks and 8.39 and 44.93 °C for melting peaks) differed remarkably from those of other fat types.  相似文献   

2.
Summary We report here on the apparent compatibility of a series of n-alkane/PDMPO blends. Three alkanes were examined [eicosane (C20), tetracosane (C24), and Polywax 2000 (nominally C142)] as blends of 5, 10, 15, and 20 wt.% alkane with PDMPO. Samples were judged to be compatible or incompatible according to three criteria: (a) visual appearance, i.e., transparent vs. opaque molded specimens, (b) the presence or absence of an alkane crystalline melting peak in a DSC thermogram, and (c) downward shifts in the PDMPO-rich phase glass transition temperatures from the T g range of 195–200°C for PDMPO homopolymer. The T u>Tm alkane component transition has not been observed in any of the blends prepared to date.  相似文献   

3.
The physical‐chemical properties, fatty acid composition and thermal properties of cattle subcutaneous, tallow and intestinal fats were determined. Subcutaneous fat differed from the other fat types with respect to its lower melting point (29.0 °C) and higher saponification (211.4 mg KOH/g) and iodine (50.55) values. The cattle fat types contained palmitic acid (16:0), stearic acid (18:0), oleic acid (18:1n9) and linoleic acid (18:2n6) as the major components of fatty acid composition (24.58–25.90%, 10.21–33.33%, 28.18–46.05%, 1.54–1.73%, respectively). A differential scanning calorimetry (DSC) study revealed that two characteristic peaks were detected in both crystallization and melting curves. Major peaks (Tpeak) of tallow and intestinal fats were similar and determined as 24.10–27.71 °C and 2.16–4.75 °C, respectively, for crystallization peaks and 7.09–9.39 °C and 43.28–46.49 °C, respectively, for melting peaks in DSC curves; however, those of subcutaneous fat (12.48 °C and –6.79 °C for crystallization peaks and 3.56 °C and 23.55 °C for melting peaks) differed remarkably from those of the other fat types.  相似文献   

4.
Thermal decomposition of pure HMX shows DSC endothermic peaks at 210°C from β to γ phase transformation and at 285°C from the HMX melting followed by an instantaneous exothermic decomposition leading to a strong peak at 290°C and a very strong DTG peak at 325°C with a mass loss of 95%. However, the GS-2 mixture shows two DSC exothermic peaks at 225°C and 270°C assigned to nitroguanidine etc. and HMX decomposition. Addition of NH4ClO4 to HMX causes the decomposition to take place before melting and decomposition peak is lowered to 230°C. However in the case of GS-2, a strong effect was observed and the amount of the mixture was to be decreased to obtain a reasonable DSC curve. This shows that the addition of NH4CIO4 to HMX increases the decomposition of HMX, however this effect is more pronounced in GS-2 mixture. Even the minimum concentration of NH4CIO4 tested (10%) shows strong effect. The addition of NH4NO3 to HMX and GS-2 shows DSC exothermic decomposition peaks at 272°C and 245°C, showing a contribution from NH4NO3 decomposes near to the decomposition of HMX, it does not show a strong effect on the decomposition of HMX/GS-2 as compared to NH4ClO4. Addition of KClO4, NaClO4 and KNO3 don't show any effect in these thermoanalytical studies.  相似文献   

5.
Multiple melting characteristics of a highly isotactic polypropylene (iPP) were studied by means of differential scanning calorimetry (DSC). Double melting characteristics were observed on melting iPP crystallized isothermally at temperatures ranging from 110 to 140°C. iPP crystallized below and above 125°C exhibited different double melting characteristics from each other. For iPP crystallized below 125°C, the single melting peak split into two peaks during slow DSC heating scans without changing the total crystallinity in the polymer. On the other hand, the double melting endotherm of iPP crystallized above 125°C seemed to come from two preexisting crystal fractions having different Tm. There existed an optimum annealing temperature range where the five-minute annealing of iPP raised Tm of the polymer significantly. The treatment also increased the crystallinity of iPP crystallized isothermally at 110°C by 12%.  相似文献   

6.
Surface Melting in Alkane Emulsion Droplets as Affected by Surfactant Type   总被引:1,自引:1,他引:0  
The influence of surfactant type (Tween 20, Tween 40, Tween 60, Tween 80, Brij 58, Triton X-100, SDS, STS) on the crystallization and melting characteristics of emulsified (mean droplet diameter 0.52 μm) n-octadecane and n-eicosane were studied using microcalorimetry. The melting point (~37 °C) of the eicosane droplets was higher than the crystallization point (~24 °C) and was not affected by the surfactant selected. There was a similar separation between the crystallization (~14 °C) and melting (~28 °C) point of the emulsified octadecane however the details of the transitions was affected by the surfactant selected. For Tween 40 and Brij 58-stabilized droplets there was a split peak on crystallization which we attribute to a surface heterogeneous nucleation mechanism. Only these surfactant-alkane combinations had a split peak on melting. The size of the lower temperature fraction decreased with droplet size suggesting another surface effect. However, the size of the surface layer was calculated to be many times the length of the surfactant tail suggesting the crystal structure was modified by the nucleation mechanism.  相似文献   

7.
Two series of poly(ε‐caprolactone)‐b‐poly(ethylene glycol)‐b‐poly(ε‐caprolactone) triblock copolymers were prepared by the ring opening polymerization of ε‐caprolactone in the presence of poly(ethylene glycol) and dibutylmagnesium in 1,4‐dioxane solution at 70°C. The triblock structure and molecular weight of the copolymers were analyzed and confirmed by 1H NMR, 13C NMR, FTIR, and gel permeation chromatography. The crystallization and thermal properties of the copolymers were investigated by wide‐angle X‐ray diffraction (WAXD) and differential scanning calorimetry (DSC). The results illustrated that the crystallization and melting behaviors of the copolymers were depended on the copolymer composition and the relative length of each block in copolymers. Crystallization exothermal peaks (Tc) and melting endothermic peaks (Tm) of PEG block were significantly influenced by the relative length of PCL blocks, due to the hindrance of the lateral PCL blocks. With increasing of the length of PCL blocks, the diffraction and the melting peak of PEG block disappeared gradually in the WAXD patterns and DSC curves, respectively. In contrast, the crystallization of PCL blocks was not suppressed by the middle PEG block. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

8.
Films iodinated at solution before casting (IBC films) were prepared by casting aqueous solutions of 10 wt % poly(vinyl alcohol) (PVA) containing selected quantities of I2/KI. The quantity of I2/KI was controlled to obtain 15.2, 39.8, 83.2, 117.0, and 140.1%. The Thermogravimetry (TG) curves of the IBC film exhibited three distinct zones corresponding to the evaporation of H2O and I2 molecules (zone I), evaporation of I2 and partial decomposition of side groups (? OH) (zone II), degradation of the remaining side groups and partial degradation of the main chain (zone III‐1), and degradation of the remaining main chain and the char zone corresponding to KI. The crystalline structure of the film with a weight gain of 15.2% was almost the same as that of the pure PVA, and the film with the weight gain of 140% was almost amorphous. The differential scanning calorimetry (DSC) thermograms of the IBC films with a weight gain of 15.2% and 39.8% indicated endothermic single or double peaks at around 180°C, corresponding to the crystal melting and degradation of side groups; those with weight gains of 83.2% and above indicated exothermic peaks at around 170°C, corresponding to crystallization, and broad endothermic peak at around 180–200°C, corresponding to the crystal melting and degradation of side groups. The dynamic mechanical αa transition of the IBC film with the weight gain of 140.1% appeared at around 20°C. X‐ray diffraction and DSC analysis of deiodinated films show that the crystal structure, on deiodination of all the IBC films, regardless of crystallinity, returned to that of the pure PVA. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3497–3502, 2006  相似文献   

9.
The structure and thermal properties of polyamide‐1010 (PA1010), treated at 250°C for 30 min under pressures of 0.7–2.5 GPa, were studied with wide‐angle X‐ray diffraction (WAXD), infrared (IR), differential scanning calorimetry (DSC), and thermogravimetric analysis (TGA). Crystals were formed when the pressures were less than 1.0 GPa or greater than 1.2 GPa. With increasing pressure, the intensity of the diffraction peak at approximately 24° was enhanced, whereas the peak at approximately 20° was depressed. The triclinic crystal structure of PA1010 was preserved. The highest melting temperature of the crystals obtained in this work was 208°C for PA1010 treated at 1.5 GPa. Crosslinking occurred under pressures of 1.0–1.2 GPa. Only a broad diffraction peak centered at approximately 20° was observed on WAXD patterns, and no melting and crystallization peaks were found on DSC curves. IR spectra of crosslinked PA1010 showed a remarkable absorption band at 1370 cm?1. The N? H stretching vibration band at 3305 cm?1 was weakened. Crystallized PA1010 had a higher thermal stability than crosslinked PA1010, as indicated on TGA curves by a higher onset temperature of decomposition. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 83: 2522–2527, 2002  相似文献   

10.
Highly saturated (HS) soybean oil (SBO), which contained 23.3% palmitic acid (C16:0) and 20.0% stearic acid (C18:0), was interesterified at 70°C in preparation for the processing of a trans-free margarine. High-performance liquid chromatography analysis of the triacylglycerides and analysis of the sn-2 fatty acid composition showed no further change after 10 min of interesterification. The interesterified HS SBO had a slip melting point of 34.5°C, compared with 9.5°C in the non-interesterified HS SBO, and increased melting and crystallization temperatures were found using differential scanning calorimetry. Analysis of solid-fat content by nuclear magnetic resonance revealed the presence of only a small amount of solids above 33°C. A 50:50 blend of interesterified HS SBO and SBO with a typical fatty acid composition was used to make the margarine. Compared to commercial soft-tub margarine, the maximal peak force on the texture analyzer of this blended margarine was about 2.3 times greater, the hardness about 2.6 times greater, and adhesiveness about 1.5 times greater. There were small but statistically significant differences (α=0.05) in the sensory properties of spreadability, graininess, and waxiness between the commercial and blended margarines at 4.5°C and, except for graininess, at 11.5°C. These very small differences suggest a potential use for HS SBO in margarine products.  相似文献   

11.
Esters of castor oil and hydrogenated castor oil were prepared with C6, C12, C16, C18 fatty acids, using tetra‐n‐butyl titanate as a catalyst and n‐butyl benzene as a water entrainer. Physical properties such as melting point, refractive index, viscosity, and specific gravity of these esters were measured. Slip melting points of the esters were very low in both cases. These esters did not crystallize even at low temperature. The highest slip melting point obtained was 21 °C with stearoyl hydrogenated castor oil ester and lowest slip melting point obtained was —6 °C with hexanoyl castor oil ester.  相似文献   

12.
The melting, isothermal and nonisothermal crystallization behaviors of poly(3‐hydroxybutyrate) (PHB) have been studied by means of temperature modulated differential scanning calorimetry (TMDSC) and conventional DSC. Various experimental conditions including isothermal/annealing temperatures (80, 90, 100, 105, 110, 120, 130, and 140°C), cooling rates (2, 5, 10, 20, and 50°C/min) and heating rates (5, 10, 20, 30, 40, and 50°C/min) have been investigated. The lower endothermic peak (Tm1) representing the original crystals prior to DSC scan, while the higher one (Tm2) is attributed to the melting of the crystals formed by recrystallization. Thermomechanical analysis (TMA) was used to evaluate the original melting temperature (Tmelt) and glass transition temperature (Tg) as comparison to DSC analysis. The multiple melting phenomenon was ascribed to the melting‐recrystallization‐remelting mechanism of the crystallites with lower thermal stability showing at Tm1. Different models (Avrami, Jeziorny‐modified‐Avrami, Liu and Mo, and Ozawa model) were utilized to describe the crystallization kinetics. It was found that Liu and Mo's analysis and Jeziorny‐modified‐Avrami model were successful to explain the nonisothermal crystallization kinetic of PHB. The activation energies were estimated in both isothermal and nonisothermal crystallization process, which were 102 and 116 kJ/mol in respective condition. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42412.  相似文献   

13.
Changes in DSC melting properties of palm oil (PO), sunflower oil (SFO), palm kernel olein (PKOo), and their belends in various ratios were studied by using a combination of blending, and chemical interesterification (CIE) techniques and determining total melting (ΔH f ) and partial melting (ΔH i°C ) enthalpies. Blending and CIE significantly modified the DSC melting properties of the PO/SFO/PKOo blends. PO and blends containing substantial amounts of PO and PKOo experienced an increase in their DSC ΔH f and ΔH i°C following CIE. The DSC ΔH f and ΔH i°C of PKOo, blends of PO/SFO at 1∶1 and 1∶3 ratios, and all blends of PKOo/SFO significantly decreased after CIE. The DSC ΔH f and ΔH i°C of SFO changed little following CIE. Randomization of FA distribution within and among TAG molecules of PO and PKOo led to modification in TAG composition of the PO/PKOo blends and improved miscibility between the two fats and consequently diminished the eutectic interaction that occurred between PO and PKOo.  相似文献   

14.
This work investigated how pre‐melting temperature (Tmax) and cooling rate (C) affected the non‐isothermal melt crystallization, melting behavior and crystal structure of syndiotactic polystyrene (sPS) by using differential scanning calorimetry (DSC) and wide angle X‐ray diffraction (WAXD) techniques. Experimental results indicated that raising Tmax or C decreased the crystallization peak temperature (Tp) and crystallization initiating temperature (Ti). The crystallization kinetics was analyzed through the Ozawa equation. Although the Ozawa exponent n and cooling function K(T) were determined for Tmax = 340°C and Tmax = 315°C specimens, for Tmax = 290°C specimens, the Ozawa equation was not applicable. Activation energies for the non‐isothermal crystallization processes of different Tmax specimens were estimated to be approximately 418 kJ/mol. As Tmax was raised the nucleation rate of sPS became slower. The multiple melting peaks were associated with different polymorphs as well as recrystallized crystals that formed during heating scans. The percentage content of α polymorph formed in the crystals under various crystallization conditions was estimated through WAXD experiments.  相似文献   

15.
Several pilot-scale trials reported in this paper, using palm stearin-rice bran oil (PS-RBO) blends, obviously did not contain trans FA (TFA), whereas the commercial products were found to contain 18–27% TFA. The effects of processing conditions such as rate of agitation, crystallization temperature, and composition of the blends on the crystal structure of shortenings were studied. The products were evaluated for their physicochemical characteristics using DSC, X-ray diffraction (XRD), HPLC, and FTIR techniques. The formulation containing 50% PS and 50% RBO showed melting and cooling characteristics similar to those of hydrogenated commercial “vanaspati” samples. Analysis of the FA composition revealed that the formulated shortenings contained 15–19% C18∶2 PUFA. Tocopherol and tocotrienol contents of the experimental shortenings were in the range of 850–1000 ppm with oryzanol content up to 0.6%. XRD studies demonstrated that the crystal form in the shortenings was predominantly the most stable β′ form, and there was less of the undesirable β form.  相似文献   

16.
Poly(ethylene terephthalate) (PET) fibers of different draw ratios (as spun and drawn 2.8×) were submitted to different heat settings. Subsequently, they were treated in benzoic acid solutions at different concentrations and for different times of exposition. The plas-ticizing effect on the morphology of these heat-setted fibers due to the different treatments in benzoic acid solutions was studied by differential scanning calorimetry (DSC) and dynamic and mechanical thermal analysis (DMTA). The DSC experiments revealed the appearance of a premelting peak (pm1) around 130°C, when the applied heat settings were carried out for 8 h in the presence of boiling water, and around 150–160°C for the dry heat setting for 7 h at 130°C followed by 1 h in the presence of boiling water. These premelting peaks have been associated with the melting of smaller and imperfect crystals present in the amorphous region. Also, the analysis of the main melting peaks revealed that the undrawn fiber presented a wider distribution of smaller crystals than the drawn one. Their subsequent treatments in the benzoic acid solutions revealed through the DSC thermograms that this first premelting peak tends to disappear as the concentration and time of benzoic acid treatment increase. Also, a new premelting peak (pm2) appears in the DSC thermograms at temperature around 180°C. It seems that a dislocation of the pm 1 peak is occurring towards higher temperatures and is being transformed in this new premelting peak. The DMTA revealed the appearance of an αc transition in the temperature range from 140 to 160°C and this αc transition seems to be related to the relaxations of the new crystallities formed due to the plasticization effect of the benzoic acid. Also, the second premelting peak seems to be related to the fusion of such crystals. Therefore, a complex change in the morphology of the studied PET fibers due to the benzoic acid action has been revealed. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
Solubility data for gases in alkane solvents are frequently required in the petroleum and petrochemical industries, data for H2S (for example) beinThe solubilities of CO2 and H2S have been measured as functions of temperature in the n-alkanes C6 to C16 at normal pressure. LessThe terminal activity coefficients for CO2 in the alkane series fell from 2.3 for CO2 in C6H14 to 1.9 in C16H34 at 25°C. The  相似文献   

18.
Cocrystallization behavior of comb‐like poly(n‐docosyl acrylate) (PDA) with n‐docosanoic acid (C22) has been studied by means of differential scanning calorimetry (DSC) and X‐ray diffraction (XRD) methods. The DSC curves of blended samples of neat PDA with C22 show the characteristic melting endotherms that correspond to the melting of the crystallites. DSC measurements of PDA/C22 blends also suggest the existence of another crystalline form induced by the addition of the C22. From the XRD measurements, the existence of hexagonally‐packed crystalline lattice and the phase behavior of PDA/C22 blends at different mole percent are confirmed. Thermal degradation behavior of PDA and its corresponding blends with C22 is studied by thermogravimetric analysis. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 2140–2146, 2005  相似文献   

19.
Criteria for authentic olive oils were developed from isocratic high-performance liquid chromatography analyses of 99 olive oils from the major Mediterranean producers in the 1983–1986 crop years. Authentic olive oils include extra virgin, virgin and pure or refined oils, but exclude all reesterified and adulterated oils. The extra virgin through pure grades will have a combined area for the LOO (C18:2C18:1C18:1), LOP (C18:2C18:1C16:0), OOO (C18:1C18:1C18:1), POO (C16:0C18:1C18:1), POP (C16:0C18:1C18:1), and SOO (C18:0C18:1C18:1) peaks between 82.0 and 92.6% of the total area (L, linoleic; O, oleic; P, palmitic; S, stearic). Authentic oils will have ratios of LOO/LOP and OOO/POO that coincide with a line defined by OOO/POO=0.7844(LOO/LOP)+0.0968; correlation coefficient is 0.885. Authentic oils will not have a trilinolein (LLL) peak over 0.5% in area. Neither triolein (OOO) nor any other single peak suffices to characterize an olive oil sample as one of the authentic grades.  相似文献   

20.
This work was aimed at studying supercritical carbon dioxide (scCO2)‐induced melting temperature depression and crystallization of a syndiotactic polypropylene (sPP). Under scCO2, the melting temperature of the sPP could be significantly reduced depending on the CO2 pressure. The scCO2‐induced crystallization of sPP was investigated using differential scanning calorimeter (DSC) and Fourier transform infrared spectroscopy. Two melting peaks were observed in DSC. The one at lower melting temperature referred to the melting of the sPP crystals induced by scCO2 in its amorphous phase. Its location was shifted to higher temperature, and its area increased with increasing scCO2 treatment time, temperature, and pressure. The melting peak at higher temperature corresponded to the melting of the sPP crystals that already existed before scCO2 treatment. Its location and area remained almost unaffected by the scCO2 treatment. The scCO2‐induced crystallization was related to scCO2‐promoted transformation of the mesophase form III of the sPP to the stable form I. POLYM. ENG. SCI., 2008. © 2008 Society of Plastics Engineers  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