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.
Blood meal is a high protein, low value by‐product of the meat processing industry that can be converted into a thermoplastic material by extrusion with a combination of a surfactant, urea, a reducing agent, water, and plasticiser. Changes in protein structure after each processing step (mixing with additives, extrusion, injection molding, and conditioning) were explored using synchrotron FTIR microspectroscopy. Blood meal particles were found to have higher β‐sheet content around the perimeter with a randomly structured core. α‐Helices were either located near the core or were evenly distributed throughout the particle. Structural rearrangement consistent with consolidation into a thermoplastic was seen after extrusion with processing additives, resulting in reduced α‐helices and increased β‐sheets. Including triethylene glycol as a plasticiser reduced α‐helices and β‐sheets in all processing steps. At all processing stages, regions with increased β‐sheets could be identified suggesting blood meal‐based thermoplastics should be considered as a semicrystalline polymer where clusters of crystalline regions are distributed throughout the disordered material. © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

2.
Protein thermoplastics, like Novatein, typically comprise a polymer, additives, and plasticizers. During foaming, the plasticizers performed different functions; triethylene glycol (TEG) affected the Tg of the polymer, while water altered the shear viscosity and urea the extensional viscosity. Water and urea also functioned as blowing agents; increasing urea increased the expansion ratio, while an increase in TEG and water decreased it. It was concluded that at this level of plasticization, phase separation occurred and that the plasticizers are most influential during late bubble growth and stabilization. Contrary to previous thought, increasing water had no effect on the Tg, but did lower the shear viscosity, while increasing TEG had the opposite effect. This enables the properties of protein thermoplastics to be tailored during development. During foaming, lower viscosity allows more time for gases to diffuse before the viscosity in the surrounding polymer increases, restricting bubble growth. At higher TEG content, the material is processed further above its Tg, slowing down stabilization and decreasing expansion. This study clarified the role of plasticizers during foaming and showed a decrease in random coils and β-turns (measured by FT-IR) with increasing expansion ratio. Knowledge of these mechanisms enables tailoring properties of protein thermoplastics foams during development. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 47781.  相似文献   

3.
Novatein is a thermoplastic produced from blood meal and is used in different agricultural applications. Novatein has some unique processing challenges and its rheology was studied using screw-driven capillary rheometry, with a particular focus on sheet extrusion using ethylene glycol, glycerol, propylene glycol (PG), or triethylene glycol (TEG) as plasticizers. The entrance pressure drop contributed up to 44% of the total pressure drop (entrance and capillary pressure drop), but this was significantly reduced by plasticization or increased temperature. Polyol addition led to higher shear viscosities in comparison to no polyol plasticization, most likely due to improved chain mobility resulting in orientation effects. Elongational flow was dominated by primary plasticization of the protein-rich phase and changes in secondary structure, whereas secondary plasticization (phase separation into a polyol-rich phase) played a significant role in the reduction of the shear viscosity. Of the selected plasticizers, PG showed the most efficient plasticization in both shear and elongational flow. When combined with the beneficial secondary structural changes brought about by TEG, the sheet forming ability of Novatein was drastically improved.  相似文献   

4.
In this study, the melt structure of isotactic polypropylene (iPP) nucleated with α/β compounded nucleating agents (α/β‐CNA, composed of the α‐NA of 0.15 wt % Millad 3988 and the β‐NA of 0.05 wt % WBG‐II) was tuned by changing the fusion temperature Tf. In this way, the role of melt structure on the crystallization behavior and polymorphic composition of iPP were investigated by differential scanning calorimetry (DSC), wide‐angle X‐ray scattering (WAXD) and scanning electron microscopy (SEM). The results showed that when Tf = 200°C (iPP was fully molten), the α/β‐CNA cannot encourage β‐phase crystallization since the nucleation efficiency (NE) of the α‐NA 3988 was obviously higher than that of the β‐NA WBG‐II. Surprisingly, when Tf was in 179–167°C, an amount of ordered structures survived in the melt, resulting in significant increase of the proportion of β‐phase (achieving 74.9% at maximum), indicating that the ordered structures of iPP played determining role in β‐phase crystallization of iPP nucleated with the α/β‐CNA. Further investigation on iPP respectively nucleated with individual 3988 and WBG‐II showed that as Tf decreased from 200°C to 167°C, the crystallization peak temperature Tc of iPP/3988 stayed almost constant, while Tc of iPP/WBG‐II increased gradually when Tf < 189°C and became higher than that of iPP/3988 when Tf decreased to 179°C and lower, which can be used to explain the influence of ordered structure and α/β‐CNA on iPP crystallization. Using this method, the selection of α‐NA for α/β‐CNA can be greatly expanded even if the inherent NE of β‐NA is lower than that of the α‐NA. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 41355.  相似文献   

5.
A batch processing method is used to fabricate foams comprising of a blend of poly(lactic acid) (PLA) and Novatein, a protein‐based thermoplastic. Various compositions of Novatein/PLA are prepared with and without a compatibilizer, PLA grafted with itaconic anhydride (PLA‐g‐IA). Pure Novatein cannot form a cellular structure at a foaming temperature of 80 °C, however, in a blend with 50 wt % of PLA, microcells form with smaller cell sizes (3.36 µm) and higher cell density (8.44 × 1021 cells cm?3) compared to pure PLA and blends with higher amounts of PLA. The incorporation of 50 wt % of semicrystalline Novatein stiffens the amorphous PLA phase, which restrains cell coalescence and cell collapse in the blends. At a foaming temperature of 140 °C, NTP30–PLA70 shows a unique interconnected porous morphology which can be attributed to the CO2‐induced plasticization effect. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2017 , 134 , 45561.  相似文献   

6.
The effect of uniaxial drawing of biodegradable soy protein isolate (SPI) polymer film on mechanical properties was investigated to accelerate the efforts to develop SPI films with improved properties. The films containing 0–30 wt% glycerol were drawn uniaxially up to a draw ratio of 2.5. The mechanical properties of the SPI film increased significantly after uniaxial drawing. The tensile strength of the undrawn film (49.7 MPa) was approximately doubled by subjecting the film to uniaxial drawing to a D.R. of 2.5. Wide‐angle X‐ray diffraction and differential scanning calorimetry measurements did not show evidence of generation of a crystal phase in the drawn SPI films. ATR‐FTIR revealed that the protein film contained mainly α‐helix and β‐sheets secondary structures. Microwave molecular orientation analysis showed that birefringence increased with increasing draw ratios. Mechanical anisotropy of the SPI film via orientation of α‐helix and β‐sheets structure is thought to be responsible for the enhancement of mechanical properties with uniaxial drawing of the SPI films. POLYM. ENG. SCI., 47:374–380, 2007. © 2007 Society of Plastics Engineers.  相似文献   

7.
Interactions of the cationic amphipathic peptide KLALKLALKALKAALKLA‐NH2 (KLAL) and its double D ‐amino acid replacement analogues l11k12‐KLAL and k9a10‐KLAL with lipid monolayers of anionic POPG, zwitterionic POPC and mixtures thereof at the air/water interface were investigated by infrared reflection– absorption spectroscopy (IRRAS). At high surface pressure (>30 mN m?1) all peptides incorporated into lipid monolayers containing at least 25 % anionic POPG, and adopted an α‐helical conformation. Creation of free surface by expansion of the monolayers resulted in an additional adsorption of peptides from the subphase, but now in a β‐sheet conformation; this led to the coexistence of peptides in two distinctly different conformations within the lipid monolayer. The β‐sheets bound to the free surface could be squeezed out of the film by compressing the film to low surface areas, whereas the α‐helices remained bound to the lipids until the film collapsed. When bound to the lipid monolayer, the helical axis of the peptides is oriented almost parallel to the surface of the monolayer.  相似文献   

8.
Tri‐ethylene glycol (TEG) is an effective plasticizer for many protein‐based thermoplastics because of its low volatility, however, partial miscibility with the protein matrix may still lead to some phase separation. Spatial variation of TEG concentration in bloodmeal‐based thermoplastics as a result of processing was investigated using synchrotron‐based FT‐IR micro‐spectroscopy. Although TEG forms strong hydrogen bonding with proteins, for the protein to fold into β‐sheets bound plasticizer must be released. TEG can then migrate, pooling into localized areas, rich in plasticizer. Further heating causes further migration towards the edge of plasticized bloodmeal particles where the TEG may evaporate. Thermo‐gravimetric analysis confirmed that loss of TEG by evaporation may occur at 120°C, given enough time for diffusion. Efficient mixing combined with a short residence time at elevated temperature mean significant plasticizer loss is unlikely during processing. However, it does limit long‐term use at elevated temperatures. © 2013 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 39969.  相似文献   

9.
Hydrolyzed proteins previously shown to aggregate in aqueous solution were compounded into synthetic polyisoprene rubber (IR). Modulus increases of up to 232% resulted from protein reinforcement of IR. Increased hydrogen bonding on amine groups and the presence of β‐sheets in the protein phase were observed via Fourier transform infrared (FTIR) spectroscopy. The total β‐sheet amount relative to the IR content strongly correlated to the modulus and varied with the protein concentration, protein aggregation state, and compounding conditions. Isotropic protein aggregates on the order of hundreds of nanometers were observed by scanning electron microscopy with energy dispersive x‐ray spectroscopy (SEM‐EDX). The aggregates were evenly dispersed throughout the rubber matrix after compounding. The composite glass transition temperature (Tg) was unchanged from the control, which indicated that the protein and rubber existed as two discrete phases. Remarkably, protein β‐sheet structures were observed in FTIR even after rubber compounding under harsh conditions. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2018 , 135, 46026.  相似文献   

10.
Peptide macrocycles are widely utilized in the development of high affinity ligands, including stapled α‐helices. The linear rigidity of a 1,3‐diynyl linkage provides an optimal distance (7 Å) between β‐carbons of the i,i+4 amino acid side chains, thus suggesting its utility in stabilizing α‐helical structures. Here, we report the development of an on‐resin strategy for an intramolecular Glaser reaction between two alkyne‐terminated side chains by using copper chloride, an essential bpy‐diol ligand, and diisopropylethylamine at room temperature. The efficiency of this ligation was illustrated by the synthesis of (i,i+4)‐, (i,i+5)‐, (i,i+6)‐, and (i,i+7)‐stapled BCL‐9 α‐helical peptides using the unnatural amino acid propargyl serine. Overall, this procedurally simple method relies on inexpensive and widely available reagents to generate low molecular weight 23‐, 26‐, 29‐, and 32‐membered peptide macrocycles.  相似文献   

11.
The nucleating ability of p‐cyclohexylamide carboxybenzene (β‐NA) towards isotactic polypropylene (iPP) was investigated by differential scanning calorimetry, X‐ray diffraction, polarized optical microscopy and scanning electron microscopy. β‐NA is identified to have dual nucleating ability for α‐iPP and β‐iPP under appropriate kinetic conditions. The formation of β‐iPP is dependent on the content of β‐NA. The content of β‐phase can reach as high as 96.96% with the addition of only 0.05 wt% β‐NA. Under non‐isothermal crystallization the content of β‐iPP increases with increasing cooling rate. The maximum β‐crystal content is obtained at a cooling rate of 40 °C min–1. The supermolecular structure of the β‐iPP is identified as a leaf‐like transcrystalline structure with an ordered lamellae arrangement perpendicular to the special surface of β‐NA. Under isothermal crystallization β‐crystals can be formed in the temperature range 80–140 °C. The content of β‐crystals reaches its maximum value at a crystallization temperature of 130 °C. © 2012 Society of Chemical Industry  相似文献   

12.
The full characterization of chitin obtained from squid, shrimp, prawn, lobsters, and king crab is reported. Elemental analysis, including metals such as Ca, Mg, Zn, Cd, Hg, Cr, Mn, Cu, and Pb, was performed, which is quite relevant because the skeleton composition is slightly different for each species. The morphology was studied by means of TEM and their compositions were determined by energy‐dispersive X‐ray analysis. 13C cross polarization/magic angle spinning NMR was applied to determine the chemical shift of all the carbons and the difference between them. Chitin was isolated by using chemical methods, alternating hydrochloric acid and sodium hydroxide. The α‐chitin from shrimp, prawn, lobsters, and king crabs showed two signals at 73.7 and 75.6 ppm. Meanwhile, the β‐chitin from squid exhibited one signal at 75.2 ppm. FTIR studies were used to analyze α‐chitin from shrimp and β‐chitin from squid. The α‐chitin exhibited amide I vibration modes at 1660 and 1627 cm?1, whereas the β‐chitin showed one band at 1656 cm?1. X‐ray diffraction showed that α‐chitin is orthorhombic (a = 4.74 Å, b = 18.86 Å, and c = 10.32 Å) and β‐chitin had a monoclinic dihydrated form (a = 4.80 Å, b = 10.40 Å, c = 11.10 Å, and β = 97°). © 2004 Wiley Periodicals, Inc. J Appl Polym Sci 93: 1876–1885, 2004  相似文献   

13.
Peptidomimetic oligomers and foldamers have received considerable attention for over a decade, with β‐peptides and the so‐called peptoids (N‐alkylglycine oligomers) representing prominent examples of such architectures. Lately, hybrid or mixed backbones consisting of both α‐ and β‐amino acids (α/β‐peptides) have been investigated in some detail as well. The present Minireview is a survey of the literature concerning hybrid structures of α‐amino acids and peptoids, including β‐peptoids (N‐alkyl‐β‐alanine oligomers), and is intended to give an overview of this area of research within the field of peptidomimetic science.  相似文献   

14.
Nonisothermal cold‐ and melt‐crystallization of syndiotactic polystyrene (sPS) were carefully carried out by Perkin–Elmer Diamond differential scanning calorimetry, polarized optical microcopy (POM), and wide angle X‐ray diffraction. The experimental data subjected to the two types of processing were thoroughly analyzed on the basis of Avrami, Tobin, Ziabicki, and combination of Avrami and Ozawa models. Avrami, Tobin, and Ziabicki analyses indicate that nonisothermal cold‐crystallization (A) characterizes smaller Avami and Tobin exponent and larger Ziabicki kinetic crystallizability index G than those obtained from nonisothermal melt‐crystallization (B) possibly due to the existence of partially ordered structures in the quenched samples. Kissinger and the differential isoconversional method (DICM) of Friedman's were utilized to obtain effective energy barrier of A, in good agreement with that obtained by using Arrhenius equation to analyze the isothermal cold‐crystallization, indicating that Kissinger and Friedman equations can be applied to obtain activation energy from A of sPS. X‐ray diffraction analysis indicates that cold‐crystallization mainly produces α‐type crystal but for melt‐crystallization the contents of α‐type and β‐type crystals depend on the cooling rates. The POM also indicates the difference of end morphology of the sample between A and B. At the same time, the DICM of Friedman's was applied to analyze experimental data of B, which were divided into two groups with 20 K/min as the threshold, and it was found that the formation of β‐type crystal possesses larger absolute value of effective activation barrier than the formation of α‐type crystal. © 2006Wiley Periodicals, Inc. J Appl Polym Sci 103: 1311–1324, 2007  相似文献   

15.
Glycosynthases—retaining glycosidases mutated at their catalytic nucleophile—catalyze the formation of glycosidic bonds from glycosyl fluorides as donor sugars and various glycosides as acceptor sugars. Here the first glycosynthase derived from a family 35 β‐galactosidase is described. The Glu→Gly mutant of BgaC from Bacillus circulans (BgaC‐E233G) catalyzed regioselective galactosylation at the 3‐position of the sugar acceptors with α‐galactosyl fluoride as the donor. Transfer to 4‐nitophenyl α‐D ‐N‐acetyl‐glucosaminide and α‐D ‐N‐acetylgalactosaminide yielded 4‐nitophenyl α‐lacto‐N‐biose and α‐galacto‐N‐biose, respectively, in high yields (up to 98 %). Kinetic analysis revealed that the high affinity of the acceptors contributed mostly to the BgaC‐E233G‐catalyzed transglycosylation. BgaC‐E233G showed no activity with β‐(1,3)‐linked disaccharides as acceptors, thus suggesting that this enzyme can be used in “one‐pot synthesis” of LNB‐ or GNB‐containing glycans.  相似文献   

16.
N‐Acylethanolamine acid amidase (NAAA) is a cysteine amidase that preferentially hydrolyzes saturated or monounsaturated fatty acid ethanolamides (FAEs), such as palmitoylethanolamide (PEA) and oleoylethanolamide (OEA), which are endogenous agonists of nuclear peroxisome proliferator‐activated receptor‐α (PPAR‐α). Compounds that feature an α‐amino‐β‐lactone ring have been identified as potent and selective NAAA inhibitors and have been shown to exert marked anti‐inflammatory effects that are mediated through FAE‐dependent activation of PPAR‐α. We synthesized and tested a series of racemic, diastereomerically pure β‐substituted α‐amino‐β‐lactones, as either carbamate or amide derivatives, investigating the structure–activity and structure–stability relationships (SAR and SSR) following changes in β‐substituent size, relative stereochemistry at the α‐ and β‐positions, and α‐amino functionality. Substituted carbamate derivatives emerged as more active and stable than amide analogues, with the cis configuration being generally preferred for stability. Increased steric bulk at the β‐position negatively affected NAAA inhibitory potency, while improving both chemical and plasma stability.  相似文献   

17.
Morphological characteristics of poly(vinylidene fluoride) (PVDF) films, filled with mass fractions (w ≤ 20%) of Barium Chloride (BaCl2), were investigated by X‐ray diffraction (XRD), Fourier transform infrared (FTIR) absorption spectra and differential scanning calorimetry (DSC) measurements. The dielectric properties of films were measured from 250 Hz to 1 MHz range between 100 and 400 K as a function of frequency and temperature. Spectroscopic data revealed that the filled and unfilled films include α‐, β‐, and γ‐crystalline phases. By a 20% filling, 73% increase was obtained in the total degree of crystallinity. Since the BaCl2 formed fluorine bridges over the chain segments on the crystal lamellar surface, the γ‐crystalline phase decreased, whereas the total degree of crystallinity increased. Dielectric measurements showed that maximum of the dielectric loss factor belonging to β‐relaxation transition decreased linearly with filling level. The filling process did not have any effect on the real dielectric constant till α‐relaxation transition region. However, in the α‐relaxation transition region, it was determined that the real dielectric constant increased linearly with filling level. POLYM. COMPOS., 31:1782–1789, 2010. © 2010 Society of Plastics Engineers.  相似文献   

18.
The effects of α‐form and β‐form nuclei on polymorphic morphology of poly(butylene adipate) (PBA) upon recrystallization from the molten state up to various Tmax values were examined by differential scanning calorimetry (DSC), wide‐angle X‐ray diffraction (WAXD) and polarized light microscopy (PLM). In this study, PBA with complex melting and polymorphism behaviour was used as a model for examining different types and extents of residual nuclei. As the PBA initially containing the sole α‐crystal was brought to a molten state of various Tmax, the extents of trace α‐form crystal nuclei varied and were dependent on Tmax. Furthermore, it did not matter whether, initially, the PBA contained α‐ or β‐form crystals (or both) because only a single type of α‐nuclei could be left upon treatment to the molten liquid state at Tmax. Therefore, only the α‐crystal in PBA had ‘memory capacity’ in the molten liquid state while the β‐crystal did not. This was so because the latter had been completely transformed into the solid state prior to being heated into a liquid. PBA crystallized before α‐nuclei could be packed into α‐crystal, regardless of the crystallization temperature (Tc). For recrystallization from molten PBA without any nuclei, the crystalline polymorphism was correspondingly influenced by Tc. Copyright © 2005 Society of Chemical Industry  相似文献   

19.
As part of continuous efforts to understand the surprising synergetic effect between β‐nucleating agent and pre‐ordered structures of isotactic polypropylene (iPP) in significant enhancement of β‐crystallization (Ordered Structure Effect, OSE), two β‐nucleated iPP with different uniformities of stereo‐defect distribution (WPP‐A and WPP‐B) were prepared, their crystallization behaviors with variation of melt structures were studied in detail. The results revealed that β‐phase can hardly form in WPP‐A (whose stereo‐defect distribution is less uniform) because of its strong tendency of α‐nucleation caused by its less uniform stereo‐defect distribution, while WPP‐B is more favorable for β‐crystallization; As fusion temperature decreases, similar variation trends of crystallization temperature and β‐phase proportion can be observed from WPP‐A and WPP‐B, indicating the occurrence of OSE behavior, which provides unsurpassed β‐nucleation efficiency and induces β‐crystallization even in WPP‐A which is less favorable for β‐crystallization; moreover, the upper and lower limiting temperatures of Region II of WPP‐A and WPP‐B are identical, suggesting the uniformity of stereo‐defect distribution has little influence on temperature window for OSE (denoted as Region II). To explore the physical nature of Region II, self‐nucleation behavior and equilibrium melting temperature of PP‐A and PP‐B were studied. The lower limiting temperatures of exclusive self‐nucleation domain of both PP‐A and PP‐B are identical with the lower limiting temperatures of Region II in OSE (168°C); moreover, the Tm0 of both PP‐A and PP‐B are close to their upper limiting temperatures of Region II in OSE behavior (189°C). The possible explanation was proposed. © 2015 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2015 , 132, 42632.  相似文献   

20.
Lanthanides have interesting chemical properties; these include luminescent, magnetic, and catalytic functions. Toward the development of proteins incorporating novel functions, we have designed a new lanthanide‐binding motif, lanthanide fingers. These were designed based on the Zif268 zinc finger, which exhibits a ββα structural motif. Lanthanide fingers utilize an Asp2Glu2 metal‐coordination environment to bind lanthanides through a tetracarboxylate peptide ligand. The iterative design of a general lanthanide‐binding peptide incorporated the following key elements: 1) residues with high α‐helix and β‐sheet propensities in the respective secondary structures; 2) an optimized big box α‐helix N‐cap; 3) a Schellman α‐helix C‐cap motif; and 4) an optional D ‐Pro‐Ser type II’ β‐turn in the β‐hairpin. The peptides were characterized for lanthanide binding by circular dichroism (CD), NMR, and fluorescence spectroscopy. In all instances, stabilization of the peptide secondary structures resulted in an increase in metal affinity. The optimized protein design was a 25‐residue peptide that was a general lanthanide‐binding motif; this binds all lanthanides examined in a competitive aqueous environment, with a dissociation constant of 9.3 μM for binding Er3+. CD spectra of the peptide‐lanthanide complexes are similar to those of zinc fingers and other ββα proteins. Metal binding involves residues from the N‐terminal β‐hairpin and the C terminal α‐helical segments of the peptide. NMR data indicated that metal binding induced a global change in the peptide structure. The D ‐Pro‐Ser type II’ β‐turn motif could be replaced by Thr–Ile to generate genetically encodable lanthanide fingers. Replacement of the central Phe with Trp generated genetically encodable lanthanide fingers that exhibited terbium luminescence greater than that of an EF‐hand peptide.  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