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 15 毫秒
1.
Electron paramagnetic resonance (EPR) spin-trapping experiments, employing the novel spin-trap DEPMPO, provide evidence for the formation of protein-peroxyl radicals from the reaction of bovine serum albumin (BSA) or lysozyme with HO · in the presence of O 2 . Spin-trapping leads to the detection of anisotropic spectra of partially immobilized protein-peroxyl spin-adducts; positive identification is based on a novel spectrum simulation approach (through which broadened anisotropic spectra are simulated and compared with experiment) and by comparison of results with those obtained when MeO 2 · is trapped and the adduct frozen in a solid matrix.  相似文献   

2.
The infrared, visible and nuclear magnetic resonance spectra of protochlorophyll a and vinylprotochlorophyll a in dry non-polar solvents (carbon tetrachloride, chloroform, cyclohexane) are presented and interpreted in terms of dimer interaction.

The infrared spectra in the 1600–1800 cm−1 region clearly show the existence of a coordination interaction between the C-9 ketone oxygen function of one molecule and the central magnesium atom of another molecule. Infrared spectra in the OH stretching region (3200–3800 cm−1) provide a valuable test of the water content in the samples.

The analysis of the absorption and circular dichroism spectra of protochlorophyll a and vinylprotochlorophyll a in carbon tetrachloride demonstrates the existence of a monomer-dimer equilibrium in the concentration range from 10−6 to 5 · 10−4 M. The dimerization constants are (6±2) · 105 1 · M−1 for protochlorophyll a and (4.5±2) · 105 1 · M−1 for vinylprotochlorophyll a at 20 °C. The deconvolution of visible spectra in the red region has been performed in order to obtain quantitative information on the dimer structure. Two models involving a parallel or a perpendicular arrangement of the associated molecules are considered.

From 1H NMR spectra, it appears that the region of overlap occurs near ring V, in agreement with the interpretation of the infrared spectra.  相似文献   


3.
It has been suggested that some techniques of tissue preparation for esr spectroscopy may artifactually generate radicals. We have investigated this, together with the possibility that the susceptibility of the tissue to preparation artifacts may be altered by ischaemia and reperfusion. Three different methods of tissue processing have been assessed: (i) freeze-clamping (- 196 °C), using grooved, aluminium tongs which produce frozen cylinders of tissue (3 mm diameter) which fit directly into esr tubes; (ii) grinding of freeze-clamped tissue with a porcelain pestle and mortar; (iii) lyophilisation of ground, freeze-clamped, tissue. Isolated rat hearts (n = 7 or n = 5/group) were subjected to aerobic perfusion (10 min, 37 °C), total, global ischaemia (15 min) and reperfusion (30 sec). Hearts were freeze-clamped at the end of each period. Tissue was prepared by each of the three methods and esr spectra recorded at - 100 °C. In spectra from tissue which had been freeze-clamped only, broad high- and low-spin iron III signals (g = 1.9, g = 2.2-2.9 and g = 4.6) were seen together with a narrow, well-defined signal (g = 2.005), possibly from a semiquinone radical. In spectra from ground samples, an anisotropic signal (g = 2.040 and g = 2.008), probably from a peroxyl radical, was observed in addition to the iron III signals. The intensity of the anisotropic signal varied with perfusion conditions; in ischaemic tissue it was decreased to 33 ± 10% of the control value and in reperfused tissue it was decreased to 76 ± 26%. In spectra from lyophilised samples, a narrow signal (g = 2.009), probably from a protein radical, was observed in addition to the iron III signals. The intensity of the signal at g = 2.009 was increased in ischaemic tissue to 170 ± 57% of the control value and in reperfused tissue to 241 ± 85%. In conclusion, artifactual generation of radicals can occur upon grinding (peroxyl radical) and lyophilisation (protein radical). Ischaemia and reperfusion may alter not only radical content per se but may also modify the susceptibility of the tissue to the artifactual production of radicals.  相似文献   

4.
The sulfonic acid function was introduced into chitosan by reacting it with 5-formyl-2-furansulfonic acid, sodium salt, under the mild conditions of the Schiff reaction, thus avoiding polymer degradation and O-substitution. The reaction of chitosan (degree of deacetylation 0·58) with 5-formyl-2-furansulfonic acid, sodium salt produced a viscous solution that, upon hydrogenation, yielded N-sulfofurfuryl chitosan sodium salt. Infrared spectrometry, alkalimetry and elemental analysis provided evidence that the degree of substitution was 0·26. Circular dichroism measurements on solutions showed multiple Cotton bands in the pH interval 7·1–8·3, while at lower and higher pH values just one negative band was observed, thus providing indication of the polyampholyte nature of N-sulfofurfuryl chitosan. The 13C-NMR and FTIR spectra showed typical signals of furane carbons. Metal ion solutions at concentrations in the range 0·1–5·0 m , pH 6, promoted precipitation of metal ion complexes of N-sulfofurfuryl chitosan, with most effective removal from the solutions for Cu(II), Pb(II) and Ni(II). Sulfoethyl N-carboxymethyl chitosan was also synthesized from 2-chloroethanesulfonic acid in organic media: the sulfur content was similar (3·7%) in both polymers.  相似文献   

5.
Studies documenting spin trapping of lipid radicals in defined model systems have shown some surprising solvent effects with the spin trap DMPO. In aqueous reactions comparing the reduction of H2O2 and methyl linoleate hydroperoxide (MLOOH) by Fez+, hydroxyl (HO·) and lipid alkoxyl (LO·) radicals produce identical four-line spectra with line intensities 1:2:2:1. Both types of radicals react with commonly-used HO· scavengers, e.g. with ethanol to produce ·C(CH3)HOH and with dirnethylsulfoxide (DMSO)togive ·CH3. However, DMSO radicals (either ·CH3or ·OOCH3) react further with lipids, and when radicals are trapped in these MLOOH systems, multiple adducts are evident. When acetonitrile is added to the aqueous reaction systems in increasing concentrations, ·CH2CN radicals resulting from HO· attack on acetonitrile are evident, even with trace quantities of that solvent. In contrast, little, if any, reaction of LO· with acetonitrile occurs, even in 100% acetonitrile. A single four-line signal persists in the lipid systems as long as any water is present, although the relative intensity of the two center lines decreases as solvent-induced changes gradually dissociate the nitrogen and β-hydrogen splitting constants. Extraction of the aqueous-phase adducts into ethyl acetate shows clearly that the identical four-line spectra in the H202 and MLOOH systems arise from different radical species in this study, but the lack of stability of the adducts to phase transfer may limit the use of this technique for routine adduct identification in more complex systems. These results indicate that the four-line 1:2:2:1. aN = aH = 14.9G spectrum from DMPO cannot automatically be assigned to the HO· adduct in reaction systems where lipid is present, even when the expected spin adducts from ethanol or DMSO appear confirmatory for HO-. Conclusive distinction between HO· and LO· ultimately will require use of 13C-labelled DMPO or HPLC-MS separation and specific identification of adducts when DMPO is used as the spin trap.  相似文献   

6.
UGT1A10 was recently identified as the major isoform that conjugates estrogens. In this study, real-time PCR revealed high levels of UGT1A10 and UGT2B7 mRNA in human breast tissues. The expression of UGT1A10 in breast was a novel finding. UGT1A10 and UGT2B7 mRNAs were differentially expressed among normal and malignant specimens. Their overall expression was significantly decreased in breast carcinomas as compared to normal breast specimens (UGT1A10: 68 ± 26 vs. 252 ± 86, respectively; p < 0.05) and (UGT2B7: 1.4 ± 0.7 vs. 12 ± 4, respectively; p < 0.05). Interestingly, in African American women, UGT1A10 expression was significantly decreased in breast carcinomas in comparison to normals (57 ± 35 vs. 397 ± 152, respectively; p < 0.05). Among Caucasian women, UGT2B7 was significantly decreased in breast carcinomas in comparison to normals (1.1 ± 0.5 vs. 13.5 ± 6, respectively; p < 0.05). Glucuronidation of 4-hydroxylated estrone (4-OHE1) was significantly reduced in breast carcinomas compared to normals (30 ± 15 vs. 106 ± 31, respectively; p < 0.05). Differential down-regulation of UGT1A10 and UGT2B7 mRNAs, protein, and activity in breast carcinomas compared to the adjacent normal breast specimens from the same donor were also found. These data illustrate the novel finding of UGT1A10 in human breast and confirm the expression of UGT2B7. Significant individual variation and down-regulation of expression in breast carcinomas of both isoforms were also demonstrated. These findings provide evidence that decreased UGT expression and activity could result in the promotion of carcinogenesis.  相似文献   

7.
Nitric oxide reacts with nitronyl nitroxides (NNO) to form imino nitroxides (INO) and this transformation can be monitored using electron spin resonance spectroscopy. Recently, Akaike et al., reported that NNO such as 2-phenyl-4,4,5,5-tetramethylimidazoline-3-oxide-1-oxyl (PTIO) and its derivatives (e.g., carboxy-PTIO) react with nitric oxide (·NO) in a 1:1 stoichiometry forming 2-phenyl-4,4,5,5-tetra-methylimidazoline-1-oxyl (PTI) or the respective product (e.g., carboxy-PTI) together with nitrite and nitrate (Akaike et al., Biochemistry 32, 827-832, 1993). In this paper, we reevaluate their results and show that the stoichiometry of the reaction between PTIO and ·NO is 0.63 ± 0.06:1.0. The reason for this discrepancy is due to an erroneous assumption by Akaike et al., that the stoichiometry for the reaction between ·NO and O2 is 2:1 in aqueous solution. If the data reported by Akaike et al., were recalculated using a 4:1 stoichiometry established for the aqueous oxidation of ·NO, the reaction between ·NO and PTIO would give a stoichiometry of 0.5:1.0 in closer agreement with our data. We propose a mechanism for the reaction between PTIO and ·NO in aqueous solution. This mechanism predicts that the stoichiometry between carboxy-PTIO and ·NO is dependent on the rate of generation of ·NO and is 1:1 only at low rates of ·NO generation (i.e., 10-13 M/s). However the stoichiometry approaches 0.5:1.0 at higher rates of ·NO production or when it is added as a bolus. The ratio between nitrite and nitrate also varies as a function of the rate of generation of ·NO. The model agrees with previous experimental observations that the aqueous oxidation of ·NO in air saturated solutions will exclusively form nitrite and predicts that ·NO will only generate substantial amounts of nitrate if it is released at a rate less than 10-17 M/s. This may have important consequences in cellular systems where the concentration of ·NO is typically measured from nitrite production.  相似文献   

8.
Pretreatment of beet molasses to increase pullulan production   总被引:2,自引:0,他引:2  
Pretreatment of beet molasses with cation exchange resin, sulphuric acid, tricalcium phosphate, potassium ferrocyanide, and ethylenediaminetetraacetic acid and disodium salt (EDTA) to increase the production of pullulan was investigated. Among the above techniques used for the removal of heavy metals, sulphuric acid treatment gave better results regarding polysaccharide concentration, polysaccharide yield, and sugar utilization. Aureobasidium pullulans grown on beet molasses produced a mixture of pullulan and other polysaccharides. The pullulan content of the crude polysaccharide was 30–35%. The addition of nutrients improved the production of polysaccharide. A maximum polysaccharide concentration (32·0±1·0 g litre−1) was achieved in molasses solution (70 g litre 1 initial sugar concentration, pH 6·5–7·5) treated with sulphuric acid and supplemented with K2HPO4 0·5%, -glutamic acid 1%, olive oil 2·5% and Tween 80 0·5%. In this case, the highest values of biomass dry weight (33·8±1·0 g litre−1), polysaccharide yield (63·5±2·5%), and sugar utilization (97·5±1·5%) were obtained at pH 6·5, 3·5, and 4·5–7·5, respectively.  相似文献   

9.
F Pelaez  C de Haro 《FEBS letters》1989,250(2):523-528
The polypeptide chain initiation factor 2 (eIF-2) binds phospholipid (PL) and becomes a potent inhibitor of translation in hemin-supplemented reticulocyte lysates [De Haro et al. (1986) Proc. Natl. Acad. Sci. USA 83, 6711–6715]. This binding is independent of calcium ions and seems to be specific for phosphatidylinositol or phosphatidylserine; phosphatidic and arachidonic acids are inactive. Like -subunit-phosphorylated eIF-2, eIF-2·PL traps GEF in a non-dissociable eIF-2·PL·GEF complex whereby GEF is no longer able to recycle. Initiation is inhibited when no free GEF is available. Translational inhibition by eIF-2·PL is rescued by equimolar amounts of eIF-2·GEF. On the basis of this stoichiometry, we have estimated that reticulocyte lysates contain about 60 pmol of GEF/ml (60 nM) eIF-2·PL also inhibits translation in cell-free mouse liver extracts and this inhibition is prevented by reticulocyte eIF-2·GEF suggesting that GEF also functions in liver. However, the eIF-2·PL complex does not affect translation in such non-mammalian eukaryotic systems as wheat germ and Drosophila embryos.  相似文献   

10.
2-(Diethylphosphonate)-nitrosopropane (DEPNP), prepared by oxidation of the corresponding aminophosphonate, was found to essentially exist as monomer in both water and organic solvents. The mechanisms of its degradation under 80°C heating or visible light exposure were studied by EPR spectroscopy: its decomposition gave rise to paramagnetic by-products, which have been identified as DEPNP / ·C(CH3)2[P(O)(OC2H5)2] and DEPNP / ·P(O)(OC2H5)2 spin adducts. Despite this drawback, DEPNP was successfully used as spin trapping agents to scavenge various carbon — and phosphorus-centred free radicals both in aqueous and organic media, giving rise to intense EPR spectra characteristic of the species trapped.  相似文献   

11.
The reaction of H2[PtCl6] · 6H2O and (H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O (18C6 = 18-crown-6) with 9-methylguanine (MeGua) proceeded with the protonation of MeGua forming 9-methylguaninium hexachloroplatinate(IV) dihydrate (MeGuaH)2[PtCl6] · 2H2O (1).The same compound was obtained from the reaction of Na2[PtCl6] with (MeGuaH)Cl.On the other hand, the reaction of guanosine (Guo) with (H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O in methanol at 60 °C proceeded with the cleavage of the glycosidic linkage and with ligand substitution to give a guaninium complex of platinum(IV), [PtCl5(GuaH)] · 1.5(18C6) · H2O (2).Within several weeks in aqueous solution a slow reduction took place yielding the analogous guaninium platinum(II) complex, [PtCl3(GuaH)] · (18C6) · 2Me2CO (3).H2[PtCl6] · 6H2O and guanosine was found to react in water, yielding (GuoH)2[PtCl6] (4) and in ethanol at 50 °C, yielding [PtCl5(GuoH)] · 3H2O (5).Dissolution of complexes 2 and 5 in DMSO resulted in the substitution of the guaninium and guanosinium ligands, respectively, by DMSO forming [PtCl5(DMSO)].Reactions of 1-methylcytosine (MeCyt) and cytidine (Cyd) with H2[PtCl6] · 6H2O and(H3O)[PtCl5(H2O)] · 2(18C6) · 6H2O resulted in the formation of hexachloroplatinates with N3 protonated pyrimidine bases as cation (MeCytH)2[PtCl6] · 2H2O (6) and (CydH)2[PtCl6] (7), respectively. Identities of all complexes were confirmed by 1H, 13C and 195Pt NMR spectroscopic investigations, revealing coordination of GuoH+ in complex 5 through N7 whereas GuaH+ in complex 3 may be coordinated through N7 or through N9. Solid state structure of hexachloroplatinate 1 exhibited base pairing of the cations yielding (MeGuaH+)2, whereas in complex 6 non-base-paired MeCytH+ cations were found. In both complexes, a network of hydrogen bonds including the water molecules was found. X-ray diffraction analysis of complex 3 exhibited a guaninium ligand that is coordinated through N9 to platinum and protonated at N1, N3 and N7. In the crystal, these NH groups form hydrogen bonds N–HO to oxygen atoms of crown ether molecules.  相似文献   

12.
Reaction of sodium or potassium molybdate and excess malic acid in a wide range of pH values (pH 4.0–7.0) resulted in the isolation of two cis-dioxo-bis(malato)-Mo(VI) complexes, viz. Na3[MoO2H(S-mal)2] and K3[MoO2H(S-mal)2]·H2O (H3mal=malic acid). The sodium complex is also characterized by an X-ray structure analysis, showing that the mononuclear Mo units are linked together via very strong symmetric CO2···H··· O2C-hydrogen bond [2.432(5) Å], forming a polymeric chain. The molybdenum atoms are quasi-octahedrally coordinated by two cis-oxo groups and two bidentate malate ligands via its alkoxy and -carboxyl groups, while the β-carboxylic and carboxylate groups remain uncomplexed, as the coordination of vicinal carboxylate and alkoxide of homocitrate in FeMo cofactor of nitrogenase. The absolute configuration of the metal center in this S-malato complex is assigned as Λ and the homochirality within the chain is established as a homochiral form ···ΛS–ΛS–ΛS–ΛS···. It is proposed that the chiral configuration of the metal center in wild-type FeMo-co biosynthesis might be induced by the early coordination of the chiral R-homocitric acid, while a mixture of raceme might be obtained in the biosynthesis of NifV FeMo-cofactor. The absolute configuration of wild-type FeMo-cofactor is assigned as ΔR.  相似文献   

13.
Water-soluble polysaccharides (12·2% of the algal dry weight) were extracted from marine green seaweed (Ulva spp.) which proliferate along the Brittany shores of France. They were composed of 18·4% rhamnose, 4·4% glucose, 1·9% xylose, 0·9% mannose, 0·9% galactose, 15·2% uronic acid, 15·8% sulphate and 23·7% ash based on the extract dry weight. These polysaccharides, formed a weak gel (about 3 Pa) at a concentration of 1·6% (w/v) in deionized water. The elastic modulus increased to about 160 Pa when boric acid (15–33 m ) was added and reached 250 Pa when both boric acid (7 m ) and calcium chloride (7 m ) were present. Adjusting the pH to 7·5 or higher by sodium tetraborate, phosphate or Tris-HCl buffers was detrimental to the gel. These results demonstrate that the poorly exploited biomass of Ulva spp. is a source of gelling polysaccharides of potential economical value. Mechanisms for gel formation which unusually involve both boron and calcium ions are proposed and will be studied further.  相似文献   

14.
Dimethylammonium 2,4-dichlorophenoxyacetate (2,4-D · DMA) induced strand breaks in PM2 DNA when incubated with CuCl2, whereas 2,4-D · DMA alone or CuCl2 alone did not show any or only a negligible effect. The formation of single strand breaks increased linearly with time and concentration of 2,4-D · DMA. Neocuproine, a specific Cu(I) chelator totally prevented strand break formation. So did catalase (up to 100mM 2,4-D · DMA), but DMSO had only a small protective effect. 2,4-Dichlorophenol, CO2 and formaldehyde were detected as reaction products of 2,4-D and CuCl2. From these results a redox reaction of Cu(II) and 2,4-D is proposed, which could explain the DNA damaging properties of CuCl2/2,4-D · DMA.  相似文献   

15.
Medium improvement for the production of cholesterol oxidase (CO, EC 1.1.3.6) by Rhodococcus equi No. 23 was investigated using an orthogonal array design in two steps. Results revealed that yeast extract, Tween 80 and zinc sulphate had positive effects on CO production, but magnesium sulphate had an inhibitory effect. In addition, interaction between cholesterol and sodium chloride also had a significant effect on enzyme production. The improved medium consisted of 2·0 g/litre cholesterol, 8·0 g/litre yeast extract, 1·0 g/litre NH4Cl, 1·0 g/litre NaCl, 0·50 g/litre KH2PO4, 0·25 g/litre Na2HPO4, 0·10 g/litre -valine, 0·15 g/litre -tyrosine, 0·15 g/litre MgSO4·7H2O, 0·01 g/litre ZnSO4·7H2O, 0·10 g/litre FeSO4·7H2O and 4·0 ml/litre Tween 80. CO production at 60 h (about 0·24 units/ml) was about four-fold greater than with the control medium.  相似文献   

16.
Five heterometallic compounds with formulae [Ba(H2O)4Cr2(μ-OH)2(nta)2] · 3H2O (I), [M(bpy)2(H2O)2] [Cr2(OH)2(nta)2] · 7H2O, where M2+ = Zn, (II); Ni, (III); Co, (IV) and [Mn(H2O)3(bpy)Cr2(OH)2(nta)2] · (bpy) · 5H2O (V); bpy = 2,2′-bipyridine, (nta = nitrilotriacetate ion) have been prepared by reaction of I with the corresponding MII-sulfates in the presence of 2,2′-bipyridine. Substances I–V have been characterized by magnetic susceptibility measurements, EPR and X-ray determinations. I represents a 2D coordination polymer formed by coordination of centrosymmetrical dimeric chromium(III) units and Barium cations. The 10-coordinate Ba polyhedron is completed by four water molecules. Compounds II–IV are isostructural and consist of non-centrosymmetric dimeric anions [Cr2(μ-OH)2(nta)2]2−, complex cations [MII(bpy)2(H2O)2]2+ and solvate water molecules. The octahedral coordination of chromium atoms implies four donor atoms of the nta3− ligands and two bridging OH groups. Multiple hydrogen bonds of coordinated and solvate water molecules link anions and cations in a 3D network. A similar [Cr2(μ-OH)2(nta)2]2− unit is found in V. The bridging function is performed by a carboxylate oxygen atom of the nta ligand that leads to the formation of a trinuclear complex [Mn(bpy)(H2O)2Cr2(μ-OH)2(nta)2]. Experimental and calculated frequency and temperature dependences of EPR spectra of these compounds are presented. The fine structure appearing on the EPR spectra of compound V is analyzed in detail at different temperatures. It is established that the main part of the EPR signals is due to the transitions in the spin states of a spin multiplet with S = 2. Analyses of experimental and calculated spectra confirm the absence of interaction between metal ions (MII) and Cr-dimers in complexes III and IV and the presence of weak Mn–Cr interactions in V. The temperature dependence of magnetic susceptibilities for I–V was fitted on the basis of the expression derived from isotropic Hamiltonian including a bi-quadratic exchange term.  相似文献   

17.
A selection programme to increase the cellular eicosapentaenoic acid (EPA) content has been carried out with the microalga Isochrysis galbana. The selection process involved two stages of single selection. EPA content continuously increased from 2·4% dry weight (d.w.) of the ‘parent’ culture to an average value of 5·3% d.w. in the final stage. The proportion of total EPA variation attributable to the genetic variation (heritability in a broad sense) was 0·99 showing the importance of the genome in the determination of this fatty acid. The growth and fatty acid profile of an EPA-rich isolate grown as a chemostat in a cylindrical photobioreactor have been studied. A decrease in EPA content was observed (5·21% w/w to 2·8% w/w) at the lowest dilution rate D = 0·024 h−1, up close to the maximum growth rate, D = 0·038 h−1. At the same time, the biomass concentration also decreased from 1015 mg/litre to 202 mg/litre over the abovementioned range of dilution rate (D). Nonetheless, the EPA productivity increases with D, with a maximum of 15·26 mg/litre/day at D = 0·0208 h−1. Furthermore, steady-state dilution rates may be related to average internal light intensity. Reverse-phase, high-pressure liquid chromatography (HPLC) on octadecylsilyl semi-preparative columns was used to separate stearidonic acid (SA), EPA and docosohexaenoic acid (DHA) in polyunsaturated fatty acid concentrate obtained by the urea complexation method from a fatty acid solution previously obtained by direct saponification of biomass. Isolate SA, EPA and DHA fraction purity was 94·8, 96·0 and 94·9%, respectively, with yields of 100·0, 99·6 and 94·0%.  相似文献   

18.
The effect of the molecular weight of the gum arabic sample on droplet-size distributions of n-hexadecane-in-water emulsions (1% wt gum, 10% vol. oil) has been investigated at neutral pH. A high-molecular-weight fraction (0·87% nitrogen) corresponding to 10% of a natural gum (0·38% N) gives initially slightly larger droplets but better emulsion stability than the low-molecular-weight fraction (0·35% N) corresponding to the residual 90% of the original gum. Samples of a different gum arabic (0·35% N) subjected to different degrees of controlled degradation give decreasing emulsion stability with reduction in weight-average molecular weight from 3·1 × 105 to 2·2 × 105Da.  相似文献   

19.
Quaternized xylans: synthesis and structural characterization   总被引:3,自引:0,他引:3  
《Carbohydrate polymers》1994,24(4):301-308
Quaternization xylans were prepared by treating xylans differing in structural and molecular properties with 3-chloro-2-hydroxypropyltrimethylammonium chloride (CHMAC) in aqueous NaOH under mild reaction conditions. The degree of substitution with quaternary ammonium groups can be monitored by altering the amount of CHMAC at optimum molar ratios of NaOH/CHMAC in the range 1·2-1·6, without changing the structure of the xylan backbones. Hot water pretreatment of the xylans in the alkaline activation step significantly enhances their reactivity. The presence of alkylammonium substituents was revealed by IR and NMR spectroscopy. The applicability of quaternized xylans as a retention aid in papermaking is demonstrated on thermomechanical pulp. Xylans, therefore, represent an important renewable resource for the production of novel cationic biopolymers.  相似文献   

20.
The Piedmont System is a collection of equipment for efficiently removing the juice from sweet sorghum stalks for the production of ethanol. The concept is to separate the whole stalks into pith and rind-leaf fractions, pass only the pith fraction through a screw press, and thus achieve an improvement in juice-expression efficiency and press capacity. An energy analysis was done for two options of this proposed harvesting/processing system: (Option 1) The juice is evaporated to syrup and used throughout the year to produce ethanol, and the by-products are used as cattle feed. (Option 2) The juice is fermented as it is harvested, and the by-products (along with other cellulosic materials) are used as feedstock for the remainder of the year. Energy ratios (energy output/energy input) of 0·9, 1·1 and 0·8 were found for sweet sorghum Option 1, sweet sorghum Option 2, and corn, respectively, as feedstocks for ethanol. If only liquid fuels are considered, the ratios are increased to 3·5, 7·9 and 4·5.  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