首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到15条相似文献,搜索用时 15 毫秒
1.
To obtain further information about the formation of cellulose I and I, cross polarization/magic angle spinning (CP/MAS)13C NMR spectroscopy was used to study the effects of polymeric additives, stirring and culture temperature on the I When xyloglucan (XG) or carboxymethyl cellulose sodium salt (CMC) was added to the incubation medium, the amount of cellulose I decreased markedly, from a normal level of 64% to as low as 30%, with the most additive giving the lowest levels of I. Moreover, stirring causes mixtures containing even small amounts of XG to have a large effect. These results suggest that CMC or XG interferes with the aggregation of fibrillar units into the normal ribbon assemblies. It may be that there is a strain associated with this aggregation that results in the higher-energy I form. Thus, cellulose I may grow preferentially when the strain caused by aggregation is not present. Lower temperatures (36–10 °C) gave an increase in I (from 56 to 72%).  相似文献   

2.
In this work, for the first time, in situ formation and transformation process of embryo calcium phosphate (Ca–P) minerals on three-dimensional bacterial cellulose nanofibers was investigated. Combined with XRD, X-ray absorption near-edge structure results revealed that the embryo precursor was amorphous calcium phosphate which was subsequently converted to β-tricalcium phosphate, octacalcium phosphate, and finally to the more thermodynamically stable form of hydroxyapatite. The methodology reported herein may be extended to the studies of Ca–P and other minerals on various substrates.  相似文献   

3.
The relative reactivity of the I and I phases of Valonia cellulose toward partial homogeneous acetylation was investigated by FT-IR and CP/MAS 13C-NMR spectroscopy. At the beginning of the acetylation and when only partial reaction was achieved, it was found that the reactivity of the I phase was substantially higher than that of the corresponding I component. At a later stage of acetylation, the difference in reactivity between the two phases was less pronounced. In correlation with previous ultrastructural observations (Sassi and Chanzy, 1995), it can be concluded that at equivalent accessibility, the I phase of cellulose is indeed more reactive toward acetylation than the I phase. The homogeneous acetylation of cellulose is essentially a surface reaction that affects only the accessible parts located at the surface of the microfibrils. The decrease in the rate of I phase disappearance with acetylation time confirms therefore that the microstructure of Valonia is made of domains that are distributed throughout the thickness of its microfibrils.  相似文献   

4.
Periodic planewave and molecular cluster density functional theory (DFT) calculations were performed on Iα and Iβ cellulose in four different conformations each. The results are consistent with the previous interpretation of experimental X-ray and neutron diffraction data that both Iα and Iβ cellulose are dominantly found in the tg conformation of the hydroxymethyl group with a H-bonding conformation termed “Network A”. Structural and energetic results of the periodic DFT calculations with dispersion corrections (DFT-D2) are consistent with observation suggesting that this methodology is accurate to within a few percent for modeling cellulose. The structural and energetic results were confirmed by comparison of calculated vibrational frequencies against observed infrared and Raman frequencies of Iα and Iβ cellulose. Structures extracted from the periodic DFT-D2 energy minimizations were used to calculate the 13C nuclear magnetic resonance chemical shifts (δ13C), and the tg/Network A conformations of both Iα and Iβ cellulose produced excellent correlations with observed δ13C values.  相似文献   

5.
Oriented films of cellulose prepared from algal cellulose were hydrothermally treated to convert them into highly crystalline cellulose Iβ. The lateral thermal expansion behavior of the prepared cellulose Iβ films was investigated using X-ray diffraction at temperatures from 20 to 300 °C. Cellulose Iβ was transformed into the high-temperature phase when the temperature was above 230 °C, allowing the lateral thermal expansion coefficient of cellulose Iβ and its high-temperature phase to be measured. For cellulose Iβ, the thermal expansion coefficients (TECs) of the a- and b-axes were αa = 9.8 × 10−5 °C−1 and αb = 1.2 × 10−5 °C−1, respectively. This anisotropic thermal expansion behavior in the lateral direction is ascribed to the crystal structure and to the hydrogen-bonding system of cellulose Iβ. For the high-temperature phase, the anisotropy was more conspicuous, and the TECs of the a- and b-axes were αa = 19.8 × 10−5 °C−1 and αb = −1.6 × 10−5 °C−1, respectively. Synchrotron X-ray fiber diffraction diagrams of the high-temperature phase were also recorded at 250 °C. The cellulose high-temperature phase is composed of a two-chain monoclinic unit cell, a = 0.819 nm, b = 0.818 nm, c (fiber repeat) = 1.037 nm, and γ = 96.4°, with space group = P21. The volume of this cell is 4.6% larger than that of cellulose Iβ at 30 °C.  相似文献   

6.
Periodic and molecular cluster density functional theory calculations were performed on the Iα (001), Iα (021), Iβ (100), and Iβ (110) surfaces of cellulose with and without explicit H2O molecules of hydration. The energy-minimized H-bonding structures, water adsorption energies, vibrational spectra, and 13C NMR chemical shifts are discussed. The H-bonded structures and water adsorption energies (ΔEads) are used to distinguish hydrophobic and hydrophilic cellulose–water interactions. O–H stretching vibrational modes are assigned for hydrated and dry cellulose surfaces. Calculations of the 13C NMR chemical shifts for the C4 and C6 surface atoms demonstrate that these δ13C4 and δ13C6 values can be upfield shifted from the bulk values as observed without rotation of the hydroxymethyl groups from the bulk tg conformation to the gt conformation as previously assumed.  相似文献   

7.
8.
The partial molar volumes (Va) of 1-alkanols (carbon number, m=5, 6, 7) in - and -cyclodextrin (CD) solutions at 5.00 mmol kg–1 have been determined as a function of alkanol concentration (Ca) between 293.2 and 308.2 K by using a dilatometer. It has been observed that with an increase in Ca, Va increased in -CD solution but decreased in -CD solution, asymptotically to a value of Va in CD-free water. The dependence of Va on Ca provided the binding constant (K) of 1:1 complex, the volume change in complex formation, and the partial molar volume of complex itself. The complex formation mechanism has been discussed on the basis of these values and their carbon number dependences in the respect of geometric behavior, hydrophobic interaction, and van der Waals interaction. It is concluded that the CD cavity in water is not rigid but flexible for fitting in nicely with guest molecule.  相似文献   

9.
The formation of Pd–In catalysts synthesized from the heteronuclear acetate complex PdIn(CH3COO)5 was studied by temperature-programmed reduction, electron microscopy, IR spectroscopy of adsorbed CO and hydrogen temperature-programmed desorption (H2-TPD). IR spectroscopy of adsorbed CO and H2-TPD confirmed the formation of bimetallic Pd–In nanoparticles. It was found that the Pd–In nanoparticle surface contains predominantly Pd atoms separated from one another by indium atoms, which is evidenced by the disappearance of the CO band shift resulting from the lateral dipole–dipole interaction between adsorbed CO molecules and by a significant decrease in the band intensity of CO adsorbed in bridged form. Almost complete inhibition of palladium hydride (PdHx) provides additional evidence of the formation of Pd–In bimetallic particles.  相似文献   

10.
11.
A theory is presented to predict the intensity of satellites in the compounds through second order in perturbation theory. It was found that the satellite intensity is proportional to the square of the orbital-relaxation energy. We have applied the theory to the interpretation of chemical effects of Kα parasites—satellites within the natural width of main lines—of Cl compounds. The relaxation energies of the compounds have been calculated with the use of an SCC-DV-Xα MO method. The theory is in good agreement with the experimental results except for CsCl and RbCl. A reason for the exceptions is also presented.  相似文献   

12.
《Polyhedron》1999,18(5):657-667
Three cobalt (II)hexanedioate complexes [Co(H2O)4(H2L)]n 1 (H2L=hexanedioic acid), Co(imidazole)4 (H2L)]n 2 and [Co(pyridine)2 (H2O)4][H2L] 3 are synthesized and structurally characterized to study the effect of N-donor substituents coordinated to the metal center on the polymeric network. Complex 1 is an extended linear polymer; Co(H2O)4 units are linked by the monodendate carboxylate from either end of the extended deprotonated hexanedioic acid. There are intra- and interchain H-bonding interactions between the coordinated water molecules and the end carboxylate O atoms, the uncoordinated O atom creates two dimensional hydrogen bonding pattern. Complex 2 also is a linear polymer; Co(imidazole)4 units are linked by monodentate dibasic acid at the either end but with S shaped conformation of the hexanedioic acid, not as fully extended as in 1. The effect of bulkier N-donor substitution is seen in the distortion of the octahedral coordination polyhedron of Co(II). The noncordinated carboxylate oxygen makes one intra and one interchain H-bonding interaction with the imidazole N–H group making a two-dimensional H-bonded network as in 1. In 3 with the two strong N-donor pyridines coordinated to the metal center, the hexanedioate is out of the coordination sphere and acts as a counter ion. The Co(pyridine)2(H2O)4 units are linked by H-bonding in both the dimensions by extensively folded adipate dianion forming a sheet structure parallel to ab plane. According to our knowledge this is the first example showing a strong H-bonding network in which a tetraaquaCo(II) center forms an eight-membered ring with bidendate H-bonding interactions. None of the coordination polymeric structures form any channels in their molecular packing, even to include a small entity as a water molecule.  相似文献   

13.
The influence of pressure, temperature, of “matrix gases” N2, Ar, H2 and of the pretreatment of the vessel wall on the rate of reaction from 60Co γ-radiolysis of hydrogen—oxygen-mixtures, in the region of slow reaction, was investigated. The G(-H2)-value2 of H2/O2-mixtures (H2:O2 = 1:9−2:1) ranges from 1 to 14 with only slight dependence on pressure, temperature, H2/O2-ratio, and surface/volume ratio (S/V). The temperature has little influence (35–210°C). Replacing most of the O2 in the H2:O2 (1:9)-mixtures with N2, Ar or excess H2 at higher temperature, causes the G(-H2)-values to increase. The influence of these matrix gases increases with increasing temperature (35–210°C) and decreasing S/V ratio (0.59 and 3.8 cm-1) of the reaction vessel; it depends also on the pretreatment of the wall surface. Varying the total pressure, the G(-H2)-values show a temperature and gas mixture dependent maximum between about 20 and 200 mb. At higher temperature (210°C) we observed an influence of dose for 50 mb H2/air-mixtures, whereas at 1 b and 35–90°C no influence of the dose on the rate of reaction of such mixtures was found.The activation by N2, Ar, H2 is discussed on the base of the H2/O2-reaction being a radical-chain reaction, built up by at least 38 coupled elementary steps (Ref(1) or see part 2). O2 reacts with H2, at increased rates of conversions (> 25%), in the expected stoichiometric ratio of 2:1. Oxygen may however also be converted in non-stiochiometric amounts under certain conditions.  相似文献   

14.
The mixed-ligand complexes, [Cux(phen)yLz] (where phen stands for 1,10-phenanthroline and L for some aliphatic acids, aromatic acids, amino acids, phenols and ethylenediamine in the ratio 1:1:1 and 1:2:1) undergo two-step, diffusion-controlled, irreversible electro-reduction in 0.5 M KNO3 at the dropping mercury electrode. The double-wave nature of these complexes may be attributed to the adsorption effect of 1,10-phenanthroline. From the characteristics of the irreversible wave, the rate constant kf, for each complex has been calculated at its formal potential. The driving force for the formation of ternary systems has been analysed. The parent complex, [Cu(phen)] is found to show a discriminating nature towards the incoming secondary ligand. Additional stability effects compared to such systems with 2,2′-bipyridyl as the parent ligand have been explained on the basis of comparatively greater π-effects of 1,10-phenanthroline. Such effects are found to become more pronounced, due to the cooperative effect, if the secondary ligand also has a π-system.  相似文献   

15.
The mixed-valence Cu3Cl4 · 2C3N3(OC3H5)3 π-complex (I) is synthesized from 2,4,6-triallyloxy-1,3,5-triazine (L) and copper(II) chloride by the alternating-current electrochemical method in an ethanolic solution. Single crystals of complex I are studied by X-ray diffraction analysis: space group I41/a, a = 25.39(8), c = 10.14(6) Å, V = 6537(48) Å3, Z = 8, R = 0.0814. The copper(I) and chlorine atoms form unique cyclic tetramers Cu4Cl4. The coordination sphere of each copper(I) atom includes the C=C bond of the allyl group of the ligand molecule L in addition to the two bridging chlorine atoms. Two nearest triazine rings are coordinated through the nitrogen atoms to the copper(II) atoms, whose environment is completed to a square with two chlorine atoms. An intricate three-dimensional structure is formed due to the μ4-bridging function of the cupro(I) tetramer and the μ2-bridging function of the ligand molecule L.  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