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.
The thermal behaviors of α-d-glucose in the melting and glass transition regions were examined utilizing the calorimetric methods of standard differential scanning calorimetry (DSC), standard temperature-modulated differential scanning calorimetry (TMDSC), quasi-isothermal temperature-modulated differential scanning calorimetry (quasi-TMDSC), and thermogravimetric analysis (TGA). The quantitative thermal analyses of experimental data of crystalline and amorphous α-d-glucose were performed based on heat capacities. The total, apparent and reversing heat capacities, and phase transitions were evaluated on heating and cooling. The melting temperature (Tm) of a crystalline carbohydrate such as α-d-glucose, shows a heating rate dependence, with the melting peak shifted to lower temperature for a lower heating rate, and with superheating of around 25 K. The superheating of crystalline α-d-glucose is observed as shifting the melting peak for higher heating rates, above the equilibrium melting temperature due to of the slow melting process. The equilibrium melting temperature and heat of fusion of crystalline α-d-glucose were estimated. Changes of reversing heat capacity evaluated by TMDSC at glass transition (Tg) of amorphous and melting process at Tm of fully crystalline α-d-glucose are similar. In both, the amorphous and crystalline phases, the same origin of heat capacity changes, in the Tg and Tm area, are attributable to molecular rotational motion. Degradation occurs simultaneously with the melting process of the crystalline phase. The stability of crystalline α-d-glucose was examined by TGA and TMDSC in the melting region, with the degradation shown to be resulting from changes of mass with temperature and time. The experimental heat capacities of fully crystalline and amorphous α-d-glucose were analyzed in reference to the solid, vibrational, and liquid heat capacities, which were approximated based on the ATHAS scheme and Data Bank.  相似文献   

2.
Melting and crystallization phenomena of amylose-lipid complexes, crystallized either from dilute solution or concentrated amylose ‘melts’ under various conditions, were studied using differential scanning calorimetry (DSC). The melting enthalpies (complex/H2O: 0·1–0·5 w/w) of the solution-grown crystalline complexes were 20·4±0·8 J g?1 for amylosemonopalmitin (AM-1-C16), 26·5±1·5 J g?1 for amylose-lysolecithin (AM-lys/in) and 26·6±1·6 J g?1 for amylose-lauric acid (AM-C12). While melting of the AM-lys/in complex showed a single transition for all concentrations studied, the melting behavior of the AM-1-C16 and the AM-C12 was rather complex at low or intermediate water contents. At a heating rate of 10°C min?1 two endothermic transitions with an intermediate exothermic peak were observed, indicative of non-equilibrium melting. A process of partial melting, followed by recrystallization and final melting, is suggested to explain such multiple-melting characteristics. These phenomena become less prominent with increasing water content; presumably due to the depression of the glass transition (Tg) and the melting temperatures (Tm). The annealing behavior of AM-1-C16 further suggested that the development of new structural order upon heating takes place primarily after partial melting of the initial crystalline structure. Overall, the DSC data are typical of those for semicrystalline polymers and consistent with a lamellar type of molecular organization in the crystalline regions of these materials.  相似文献   

3.
We have used differential scanning calorimetry (DSC) to determine the unfolding properties of commercial products of human serum albumin (HSA) prepared from pooled human blood, transgenic yeast, and transgenic rice. The initial melting temperatures (Tm1) for the unfolding transitions of the HSA products varied from 62°C to 75°C. We characterized the samples for purity, fatty acid content, and molecular weight. The effects of adding fatty acids, heat pasteurization, and a low pH defatting technique on the transition temperatures were measured. Defatted HSA has a structure with the lowest stability (Tm of ~62°C). When fatty acids are bound to HSA, the structure is stabilized (Tm of ~64–72°C), and prolonged heating (pasteurization at 60°C) results in a heat‐stabilized structural form containing fatty acids (Tm of ~75–80°C). This process was shown to be reversible by a low pH defatting step. This study shows that the fatty acid composition and bioprocessing history of the HSA commercial products results in the large differences in the thermal stability. © 2014 American Institute of Chemical Engineers Biotechnol. Prog., 31:62–69, 2015  相似文献   

4.
The effect of abscisic acid on the thermotropic properties of dipalmitoylphosphatidylcholine (DPPC) and on phosphatidylethanolamines (natural (PE) and dipalmitoylphosphatidylethanolamine (DPPE)) bilayers was investigated by differential scanning calorimetry (DSC). Abscisic acid eliminates the pretransition of DPPC, causes a downward shift of its temperature of melting (Tm) and broadens the melting peak without changing the enthalpy of melting. In natural PE bilayers interacting with abscisic acid a small decrease in the enthalpy of melting almost without change of Tm was detected, whereas in synthetic DPPE abscisic acid caused a small shift of Tm and small broadening of the melting peak without changing the enthalpy of melting. Abscisic acid increases the conductance to Na+ or K+ by three orders of magnitude in planar lipid membranes formed from PE monolayers and by less than two orders of magnitude in membranes formed from PC monolayers.  相似文献   

5.
《Carbohydrate polymers》1986,6(4):269-288
Thermal analysis of solution-grown crystalline amylose-V complexes of amylose, amylodextrin and β-cyclodextrin with a series of saturated 1-monoglycerides (C10–C18), lysolecithin, lauric acid and 1,3-dipalmitin was carried out using differential scanning calorimetry (DSC). At intermediate water contents (60%) and moderate heating rates (10°C min−1), some of these complexes exhibited two melting transitions separated by an exothermic effect. A mechanism of partial melting, followed by recrystallization and final melting, is proposed to account for such non-equilibrium melting. These phenomena are strongly influenced by the amount of water present in the system; the higher the moisture content, the more cooperative the melting and the lower the occurrence of secondary crystallization processes. The overall thermal behavior of the complex/H2O mixture can be explained in terms of both thermodynamic melting point depression due to the diluent and structural reorganization upon heating. The latter is related to the crystallite morphology/habit of the complex and appears to be dictated by the nature of the ligand molecule. Good amylose complexing agents induce metastable, less perfected crystalline structures that are inclined to reorganization upon heating in the DSC, presumably via a lamella thickening mechanism. Due to the non-equilibrium character of the melting process of these complexes, theoretical treatment of melting data using expressions such as the Flory-Huggins equation is not applicable.  相似文献   

6.
The effect of lauric acid and low-HLB sucrose esters (L-195, S170) on the isothermal crystallization of coconut oil was investigated by differential scanning calorimetry. The fundamental crystallization parameters, such as induction time of nucleation and crystallization rate, were obtained by using the Gompertz equation. The Gibb's free energy of nucleation was calculated via the Fisher–Turnbull equation based on the equilibrium melting temperature. All additives, investigated in this work, proved to have an inhibition effect on nucleation and crystallization kinetics of coconut oil. Our results revealed that the inhibition effect is related to the dissimilarity of the molecular characteristics between coconut oil and the additives. The equilibrium melting temperature (Tm°) of the coconut oil–additive mixtures estimated by the Hoffman–Weeks method was decreased with the addition of lauric acid and increased by using sucrose esters as additives. Micrographs showing simultaneous crystallization of coconut oil and lauric acid indicated that strong molecular interaction led to the increase in lamellar thickness resulting in the Tm° depression of coconut oil. The addition of L-195 modified the crystal morphology of coconut oil into large, dense, non-porous crystals without altering the polymorphic occurrence of coconut oil. The enhancement in lamellar thickness and crystal perfection supported the Tm° elevation of coconut oil.  相似文献   

7.
The thermal and microstructural properties of frozen hydrated gluten were studied by using differential scanning calorimetry (DSC), modulated DSC, and low-temperature scanning electron microscopy (cryo-SEM). This work was undertaken to investigate the thermal transitions observed in frozen hydrated gluten and relate them to its microstructure. The minor peak that is observed just before the major endotherm (melting of bulk ice) was assigned to the melting of ice that is confined to capillaries formed by gluten. The Defay–Prigogine theory for the depression of melting point of fluids confined in capillaries was put forward in order to explain the calorimetric results. The pore radius size of the capillaries was calculated by using four different empirical models. Kinetic analysis of the growth of the pore radius size revealed that it follows first-order kinetics. Cryo-SEM observations revealed that gluten forms a continuous homogeneous and not fibrous network. Results of the present investigation showed that is impossible to assign a T g value for hydrated frozen gluten because of the wide temperature range over which the gluten matrix vitrifies, and therefore the construction of state diagrams is not feasible at subzero temperatures for this material. Furthermore, the gluten matrix is deteriorated with two different mechanisms from ice recrystallization, one that results from the growth of ice that is confined in capillaries and the other from the growth of bulk ice.  相似文献   

8.
An amorphous solid of cyclomaltoheptaose (β-cyclodextrin, β-CD) was formed by milling its crystalline form using a high-energy planetary mill at room temperature. The glass transition of this amorphous solid was found to occur above the thermal degradation point of the material preventing its direct observation and thus its full characterization. The corresponding glass transition temperature (Tg) and the ΔCp at Tg have, however, been estimated by extrapolation of Tg and ΔCp of closely related amorphous compounds. These compounds include methylated β-CD with different degrees of substitution and molecular alloys obtained by co-milling β-CD and methylated β-CD (DS 1.8) at different ratios. The physical characterization of the amorphous states have been performed by powder X-ray diffraction and differential scanning calorimetry, while the chemical integrity of β-CD upon milling was checked by NMR spectroscopy and mass spectrometry.  相似文献   

9.
Ca2+/calmodulin‐dependent protein kinase II (CaMKII) is a Ser/Thr kinase necessary for long‐term memory formation and other Ca2+‐dependent signaling cascades such as fertilization. Here, we investigated the stability of CaMKIIα using a combination of differential scanning calorimetry (DSC), X‐ray crystallography, and mass photometry (MP). The kinase domain has a low thermal stability (apparent Tm = 36°C), which is slightly stabilized by ATP/MgCl2 binding (apparent Tm = 40°C) and significantly stabilized by regulatory segment binding (apparent Tm = 60°C). We crystallized the kinase domain of CaMKII bound to p‐coumaric acid in the active site. This structure reveals solvent‐exposed hydrophobic residues in the substrate‐binding pocket, which are normally buried in the autoinhibited structure when the regulatory segment is present. This likely accounts for the large stabilization that we observe in DSC measurements comparing the kinase alone with the kinase plus regulatory segment. The hub domain alone is extremely stable (apparent Tm ~ 90°C), and the holoenzyme structure has multiple unfolding transitions ranging from ~60°C to 100°C. Using MP, we compared a CaMKIIα holoenzyme with different variable linker regions and determined that the dissociation of both these holoenzymes occurs at a higher concentration (is less stable) compared with the hub domain alone. We conclude that within the context of the holoenzyme structure, the kinase domain is stabilized, whereas the hub domain is destabilized. These data support a model where domains within the holoenzyme interact.  相似文献   

10.
L H Chang  A G Marshall 《Biopolymers》1986,25(7):1299-1313
The unfolding of B. subtilis 5S RNA is examined by direct calorimetric measurement in the presence of various concentrations of Na+ and Mg2+. The composite differential scanning calorimetry (DSC) curve is analyzed into 3–5 individual two-state melting transitions. In the absence of added Na+ or Mg2+, the 5S RNA segments melt together at Tm = 40°C. Addition of Na+ stabilizes the molecular structure (Tm = 56°C) and widens the melting temperature range, so that up to five component transitions are observed. Addition of Mg2+ alone produces a very stable structure (Tm = 75°C) with highly cooperative melting. Finally, addition of both Na+ and Mg2+ produces the highest stability (Tm = 76°C). The results are interpreted according to hypothetical secondary and tertiary base-pairing schemes. The conformational changes demonstrated here may facilitate the movement of the protein synthesis machinery during RNA translation.  相似文献   

11.
Polyhydroxyalkanoate (PHA) synthesis genes phaPCJ Ac cloned from Aeromonas caviae were transformed into Pseudomonas putida KTOY06ΔC, a mutant of P. putida KT2442, resulting in the ability of the recombinant P. putida KTOY06ΔC (phaPCJ A.c ) to produce a short-chain-length and medium-chain-length PHA block copolymer consisting of poly-3-hydroxybutyrate (PHB) as one block and random copolymer of 3-hydroxyvalerate (3HV) and 3-hydroxyheptanoate (3HHp) as another block. The novel block polymer was studied by differential scanning calorimetry (DSC), nuclear magnetic resonance, and rheology measurements. DSC studies showed the polymer to possess two glass transition temperatures (T g), one melting temperature (T m) and one cool crystallization temperature (T c). Rheology studies clearly indicated a polymer chain re-arrangement in the copolymer; these studies confirmed the polymer to be a block copolymer, with over 70 mol% homopolymer (PHB) of 3-hydroxybutyrate (3HB) as one block and around 30 mol% random copolymers of 3HV and 3HHp as the second block. The block copolymer was shown to have the highest tensile strength and Young’s modulus compared with a random copolymer with similar ratio and a blend of homopolymers PHB and PHVHHp with similar ratio. Compared with other commercially available PHA including PHB, PHBV, PHBHHx, and P3HB4HB, the short-chain- and medium-chain-length block copolymer PHB-b-PHVHHp showed differences in terms of mechanical properties and should draw more attentions from the PHA research community.  相似文献   

12.
The thermal stability of a recombinant α-amylase from Bacillus halmapalus α-amylase (BHA) has been investigated using circular dichroism spectroscopy (CD) and differential scanning calorimetry (DSC). This α-amylase is homologous to other Bacillus α-amylases where crystallographic studies have identified the existence of three calcium binding sites in the structure. Denaturation of BHA is irreversible with a Tm of approximately 89 °C and DSC thermograms can be described using a one-step irreversible model. A 5 °C increase in Tm in the presence of 10-fold excess CaCl2 was observed. However, a concomitant increase in the tendency to aggregate was also observed. The presence of 30–40-fold excess calcium chelator (ethylenediaminetetraacetic acid (EDTA) or ethylene glycol-bis[β-aminoethyl ether] N,N,N′,N′-tetraacetic acid (EGTA)) results in a large destabilization of BHA, corresponding to about 40 °C lower Tm as determined by both CD and DSC. Ten-fold excess EGTA reveals complex DSC thermograms corresponding to both reversible and irreversible transitions, which probably originate from different populations of BHA/calcium complexes. Combined interpretation of these observations and structural information on homologous α-amylases forms the basis for a suggested mechanism underlying the inactivation mechanism of BHA. The mechanism includes irreversible thermal denaturation of different BHA/calcium complexes and the calcium binding equilibria. Furthermore, the model accounts for a temperature-induced reversible structural change associated with calcium binding.  相似文献   

13.
The effect of tripalmitin (TP) crystallization on the thermomechanical properties of organogels developed with candelilla wax (CW) was investigated using safflower oil high in triolein (HOSFO) as the liquid phase. Factorial combinations of CW (i.e., 0–3%) and TP (i.e., 0–1%) in the HOSFO were used to develop organogels at three different temperatures (T set). The onset of crystallization (T g) during the cooling stage (10 °C/min), the melting temperature (T M), and the corresponding heat of melting (ΔH M) of the organogels were determined by differential scanning calorimetry. Results showed that, without CW, the crystallization of TP in the HOSFO at the concentrations and T set investigated (i.e., −10 °C to 25 °C) did not develop a three-dimensional network that provided significant viscoelasticity (i.e., solid-like behavior) to the HOSFO. The CW developed organogels in the HOSFO with T M’s that increased from ≈30.5 °C up to ≈42.5 °C as a function of CW concentration. In contrast, in the CW–1% TP system, the co-crystallization of TP and CW resulted in organogels with TM’s that varied just between 36 °C and 38 °C, independent of the CW concentration. Higher elastic modulus (G′) and yield stress (σ*) were obtained with 3% CW–1.0% TP organogels than with organogels developed just by CW, particularly at T set’s of −5 °C and 15 °C. This research showed that co-crystallization of TP and CW, occurring at different extent as a function of T set, resulted in organogels with thermomechanical properties different from the ones showed by CW organogels. The results showed that co-crystallization of triacylglycerides with CW might be a useful alternative to tailor particular physicochemical properties associated to a specific functionality (i.e., melting profile and texture). Organogelation of vegetable oil might be used to develop trans-free vegetable-oil-based spreads and coatings and also novel food products with new textural perceptions for the consumers.  相似文献   

14.
We previously studied a 16‐amino acid‐residue fragment of the C‐terminal β‐hairpin of the B3 domain (residues 46–61), [IG(46–61)] of the immunoglobulin binding protein G from Streptoccocus, and found that hydrophobic interactions and the turn region play an important role in stabilizing the structure. Based on these results, we carried out systematic structural studies of peptides derived from the sequence of IG (46–61) by systematically shortening the peptide by one residue at a time from both the C‐ and the N‐terminus. To determine the structure and stability of two resulting 12‐ and 14‐amino acid‐residue peptides, IG(48–59) and IG(47–60), respectively, we carried out circular dichroism, NMR, and calorimetric studies of these peptides in pure water. Our results show that IG(48–59) possesses organized three‐dimensional structure stabilized by hydrophobic interactions (Tyr50–Phe57 and Trp48–Val59) at T = 283 and 305 K. At T = 313 K, the structure breaks down because of increased chain entropy, but the turn region is preserved in the same position observed for the structure of the whole protein. The breakdown of structure occurs near the melting temperature of this peptide (Tm = 310 K) measured by differential scanning calorimetry (DSC). The melting temperature of IG(47–60) determined by DSC is Tm = 330 K and its structure is similar to that of the native β‐hairpin at all (lower) temperatures examined (283–313 K). Both of these truncated sequences are conserved in all known amino acid sequences of the B domains of the immunoglobulin binding protein G from bacteria. Thus, this study contributes to an understanding of the mechanism of folding of this whole family of proteins, and provides information about the mechanism of formation and stabilization of a β‐hairpin structural element. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

15.
Purified bovine rhodopsin was reconstituted into vesicles consisting of 1-stearoyl-2-oleoyl phosphatidylcholine or 1-stearoyl-2-docosahexaenoyl phosphatidylcholine with and without 30 mol % cholesterol. Rhodopsin stability was examined using differential scanning calorimetry (DSC). The thermal unfolding transition temperature (Tm) of rhodopsin was scan rate-dependent, demonstrating the presence of a rate-limited component of denaturation. The activation energy of this kinetically controlled process (Ea) was determined from DSC thermograms by four separate methods. Both Tm and Ea varied with bilayer composition. Cholesterol increased the Tm both the presence and absence of docosahexaenoic acid acyl chains (DHA). In contrast, cholesterol lowered Ea in the absence of DHA, but raised Ea in the presence of 20 mol % DHA-containing phospholipid. The relative acyl chain packing order was determined from measurements of diphenylhexatriene fluorescence anisotropy decay. The Tm for thermal unfolding was inversely related to acyl chain packing order. Rhodopsin kinetic stability (Ea) was reduced in highly ordered or disordered membranes. Maximal kinetic stability was found within the range of acyl chain order found in native bovine rod outer segment disk membranes. The results demonstrate that membrane composition has distinct effects on the thermal versus kinetic stabilities of membrane proteins, and suggests that a balance between membrane constituents with opposite effects on acyl chain packing, such as DHA and cholesterol, may be required for maximum protein stability.  相似文献   

16.
The rate of precipitation of the retrograded amylose product from a dil. amylose solution was determined by the centrifugal method. The results showed that the relation of the quantity of precipitate vs. time did not fit the typical second order reaction for the coalescence of colloidal particles but fitted the crystallization formula, in appearance.

The rate of precipitation was in proportion to (c-ca)1.5, where c is the amylose concentration and ca the concentration of the dil. solution phase in the phase-separated solution. When the temperature dependence of the rate was treated according to the crystallization of polymers, it was found that the rate was in proportion to Tm2/T(ΔT)2, where Tm is the melting point of the polymer in solution and ΔT is (Tm?T). The Tm thus obtained was 120°C for an amylose solution. These results suggested a certain correlation between the amylose retrogradation and the crystallization.  相似文献   

17.
1,3-dipropyl-8-cyclopentylxanthine (CPX) has been shown to stimulate in vitro CFTR activity in ∆F508 cells. Data from a phase I study demonstrated erratic bioavailability and no measurable clinical response to oral CPX. One cause for its poor bioavailability may have been dissolution rate limited absorption, but there is little published physicochemical data on which to base an analysis. The objective of this study was to determine the solubility and solid-state characteristics of CPX. CPX is a weak acid with pKa of 9.83 and water solubility at pH 7.0 of 15.6 μM. Both laureth-23 and poloxamer 407 increased the apparent water solubility linearly with increasing concentrations. CPX exists in two crystal forms, one of which (form II) has been solved. Form II is a triclinic crystal with space group P1 and calculated density of 1.278 g/cm3. X-ray powder diffraction and differential scanning calorimetry studies (DSC) indicated that CPX crystals prepared at room temperature were mixtures of forms I and II. DSC results indicated a melting point of approximately 195°C for form I and 198°C for form II. Thermogravimetric analysis indicated no solvent loss upon heating. Dynamic water vapor sorption data indicated no significant water uptake by CPX up to 90% RH. Analysis of the data indicates that CPX may not be amenable to traditional formulation approaches for oral delivery.  相似文献   

18.
The granule morphology, microstructure, and thermal properties of micronized cassava starch prepared by a vacuum ball-grinding machine were investigated. Scanning electron microscopy (SEM) analysis indicated that the morphology of starch granule changes during the ball-grinding treatment. Differential scanning calorimetry (DSC) analysis indicated that the maximum peak temperature (Tp) of the gelatinization process, the glass transition (Tg), and peak height index (PHI) for the starch granules decreased when the size of micronized starch granules was reduced. When the size of starch granules was reduced beyond 9.11 μm, they have a tendency to agglomerate and their ΔH were increased. The granule size has a significant effect on the gelatinization properties of cassava starch. This study will provide useful information of the micronized starch for its potential industrial application.  相似文献   

19.
The passage of a phospholipid through the gel to liquid crystal phase transition is associated with an increase in the motional freedom of its fatty acyl chains as measured by spectroscopic techniques and an essentially isothermal absorption of heat as measured by differential scanning calorimetry (DSC). In addition, bilayers formed from that phospholipid display a permeability maximum for both non-electrolytes and electrolytes in the temperature region of the phase transition. In this study the sodium (and in some cases glucose) permeabilities of liposomes composed of either dimyristoyl or dipalmitoyl phosphatidylcholine plus dicetylphosphate were measured in the presence of a group of benzene and adamantane derivatives known to increase fatty acyl chain motion below the lipid transition temperature (Tc) and in the case of the adamantanes to also lower the Tc as measured by DSC. None of these compounds change the temperature at which the permeability maximum occurs despite their lowering of the phospholipid Tc. That is, in the presence of these additives there is observed an apparent dissociation between the phase transition and the permeability maximum. It is proposed that the permeability maximum normally observed in the temperature region of the Tc is associated with the completion of the ‘melting’ process. Hence a compound could cause early ‘melting’ of the bilayer but not change its permeability properties if the temperature at which the ‘melting’ process neared completion was not changed.  相似文献   

20.
The thermal stability of acid-soluble collagens was studied by circular dichroism (CD) spectroscopy. Adult bovine dermal collagen (BDC), rat-tail tendon collagen (RTC), and calf skin collagen (CSC) were compared. Despite some variability in amino acid composition and apparent molecular weight, the CD spectra for helical and unordered collagen structures were essentially the same for all the sources. The melting of these collagens occurs as a two-stage process characterized by a pretransition (T p) followed by complete denaturation (T d). The characteristic temperatures vary with the source of the collagen; for mature collagens (BDC, RTC) T p = 30°C and T d = 36deg;C, and for CSC T p = 34°C and T d = 40°C. Neutral salts, NaCl or KCl, at low concentrations (0.02–0.2 M) appear to bind to the collagens and shift the thermal transitions of these collagens to lower temperatures.  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