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.
2,2,6,6-Tetramethylpiperidine-1-oxyl radical (TEMPO) can selectively oxidize primary hydroxyl groups of cellulose to carboxyl groups. However, the depolymerization also occurs during the process. The kinetics and mechanism of carboxyl group formation on the surface of cellulose fiber oxidized by TEMPO/NaClO2/NaClO were discussed. The oxidization and depolymerization of cellulose occurred simultaneously, according to analysis of FTIR and 13C CP/MAS NMR. The glucuronic acid and some small molecular fragments, formed by hydrolysis or β-elimination during the oxidation, are also discussed. The crystallization index increased and crystal size decreased, as shown by X-ray analysis. The degradation steps in the TEMPO/NaClO2/NaClO system was discussed, according to carbon conversion analyzed by 13C CP/MAS NMR. The oxidation of cellulose can be described well by the kinetics model established based on the degradation of cellulose. It was found that temperature is one of the key parameters for controlling the oxdation and degradation level. The possible mechanism for oxidation of cellulose was composed.  相似文献   

2.
沈煜婷  康经武 《色谱》2020,38(10):1238-1242
肝素和低分子量肝素(LMWHs)作为临床上常用的抗凝血药物,其抗凝血活性与硫酸化程度(SD)密切相关。然而,肝素类药物的生产需经历一系列复杂的工艺过程,在制备和储存过程中,肝素的硫酸基团容易水解丢失,影响抗凝血活性,这将直接影响肝素药物的使用安全性。为保证产品质量,需要发展一种快速检测肝素硫酸化程度的技术,以监测原料质量和工艺条件的稳定性。虽然已有一些测定肝素硫酸化程度的报道,但这些方法均有局限性,不适用于肝素生产的质量控制。为此,开发了一种基于毛细管电泳技术(CE)检测肝素和低分子量肝素的平均硫酸化程度的方法。首先,用肝素酶混合液彻底消化未分级肝素(UHF)和低分子量肝素,然后用毛细管电泳分离酶解得到的所有寡糖和二糖构建模块,并对它们进行定性和定量分析。随后,根据每种寡糖和二糖的峰面积及其硫酸酯基团的数量,便可计算出每个构成肝素二糖单元硫酸化程度的平均值。使用该方法对来自两个生产商各4个批次依诺肝素(低分子量肝素)和5个批次肝素原料进行检测,并计算了各批次样品的相对标准偏差(RSD),对不同厂家生产的依诺肝素平均硫酸化程度进行了比较,验证了该方法的实用性。该方法灵敏度高,准确可靠,分析速度快,在肝素类药物生产过程的质量控制中具有良好的应用潜力。  相似文献   

3.
Abstract

Aryl phosphonates can be prepared in good yields from the respective arenes and tri- or dialkylphosphites by either chemical or electrochemical oxidation1. The anodic oxidation proceeds either via phosphonium radical cations which then attack the arenes electrophilically or via arene radical cations which add the trialkyl-phosphite as nucleophile1,2. Aryl phosphonates are also obtained in good yields by chemical oxidation with peroxidisulfate/AgNO3, Iron(III)- or Cerium(IV)-complexes in acetonitrile/water or glacial acetic acid3.  相似文献   

4.
This paper describes a novel strategy for the recycling of polyamide materials and their transformation into functional reactive oligomers with new properties. The method is illustrated by the heating of polyamide-6 (20,000 Da) in the presence of diesters of the phosphonic acid, (RO)2P(O)H, where R could be -CH3, -C2H5 or -C6H5. It is found that the reaction proceeds in several parallel processes: (i) phosphorylation of the amide group by the alkyl esters of the phosphonic acids and (ii) degradation of the main chain through an exchange reaction between the amide and phosphonic acid ester groups. Alternatively the depolymerization could be induced via a radical reaction with the participation of the polyamide moieties and the P-H group. The proceeding of the abovementioned reactions and the structure of the phosphorus-containing oligoamides are confirmed by 31P, 1H and 13C NMR spectroscopy. Their molecular weights are determined by size exclusion chromatography.  相似文献   

5.
The chemical or electrochemical reduction of the trifluoroacetyl complex Co(CO)3(PPh3)(COCF3) involves a single electron transfer yielding trifluoromethyl radical and an anionic cobalt carbonyl complex. The mechanism is proposed to involve electron transfer followed by initial dissociation of either a carbonyl or phosphine ligand from the 19-electron [Co(CO)3(PPh3)(COCF3)] anion. The resulting 17-electron intermediate undergoes subsequent one-electron reductive elimination of trifluoromethyl radical by homolytic cleavage of the carbon-carbon bond of the trifluoroacetyl group. The radical can be trapped by either benzophenone anion, forming the anion of α-(trifluoromethyl)benzhydrol, or Bu3SnH, yielding CF3H. The ultimate organometallic product is an 18-electron anion, either [Co(CO)4] or [Co(CO)3(PPh3)], depending upon which ligand is initially lost. Fluorine-containing products were identified and quantitated by 19F NMR while cobalt-containing products were determined by IR.  相似文献   

6.
The polymerization of methyl methacrylate either by free radical or charge transfer mechanism has been studied in dimethyl sulphoxide at 60° in the presence of oxalic acid and hexakis dimethylsulphoxide iron(III) perchlorate, [Fe(DMSO)6](ClO4)3. Increased rate was noticed for 1:1 mole ratio of oxalic acid to Fe3+ for charge transfer polymerization; a well defined induction period was found for free radical polymerization in the same systems. Mechanisms for the two types of reaction are proposed. The rate constant for the interaction of poly(methyl methacrylate) radical with the iron-oxalate complex was found to be 1.52 × 105 l. mol?1 sec?1 at 60°.  相似文献   

7.
Direct fluoroalkoxylation reactions of (hetero)arenes, carbon-carbon multiple bonds, and substitution reactions at Csp3 carbon centers by CF3O, CHF2O, and (CF3)2CFO groups are discussed. Emphasis on thermal radical, electron transfer, photocatalytic, electrochemical and redox-neutral radical methods are placed to accomplish fluoroalkoxylation reactions. All these methods employ either radical fluoroalkoxylating reagents or some nucleophilic trifluoromethoxylating sources of CF3O. A summary of all these methods is provided in Table 2.  相似文献   

8.
It was shown that trimethylene oxide (oxetane) radical cations were converted at 77 K into either distonic radical cations ·CH2CH2CH=OH+ or 2-oxetanyl radicals, depending on the freonic matrix used, by the action of light at λ = 546 nm and trimethylene sulfide radical cations transformed into distonic radical cations CH2CHSH+CH 2 · under 436-nm irradiation. The quantum yields of the photochemical reactions were determined. Quantum-chemical calculations on the structure and HFC constants of the radical cations and possible paramagnetic products of their transformation were performed. The reasons behind the observed difference in reactivity between the radical cations under the action of light are discussed.  相似文献   

9.
A complete vibrational analysis (IR, Raman) of the methyl viologen radical cation MV+., generated either via thermolysis from the solid MV2+(Cl)2 salt, or via chemical reduction from the dictation intercalated in the layered CdPS3 compound, is reported fpr different isotopic derivatives. The results are discussed with respect to the conformation and the electronic configuration of the radical species, (MV+.) and (MV+.)2.  相似文献   

10.
Low molecular weight heparins (LMWHs) are recognised as the preferred anticoagulants in the prevention and treatment of venous thromboembolism. Anti-Factor Xa (anti-FXa) levels are used to monitor the anticoagulant effect of LMWHs and such assays are routinely employed in hospital diagnostic laboratories. In this study, a fluorogenic anti-FXa assay was developed using a commercially available fluorogenic substrate with an attached 6-amino-1-naphthalene-sulfonamide (ANSN) fluorophore and was used for the determination of two LMWHs, enoxaparin and tinzaparin and the heparinoid, danaparoid. The assay was based on the complexation of heparinised plasma with 100 nM exogenous FXa and 25 μM of the fluorogenic substrate Mes-D-LGR-ANSN (C2H5)2 (SN-7). The assay was tested with pooled plasma samples spiked with anticoagulant concentrations in the range 0–1.6 U mL−1. The statistically sensitive assay range was 0–0.4 U mL−1 for enoxaparin and tinzaparin and 0–0.2 U mL−1 for danaparoid, with assay variation typically below 10.5%. This assay was then compared with a previously published fluorogenic anti-FXa assay developed with the peptide substrate, methylsulfonyl-d-cyclohexylalanyl-glycyl-arginine-7-amino-4-methylcoumarin acetate (Pefafluor FXa). Both assays were compared in terms of fluorescence intensity, lag times and sensitivity to anticoagulants.  相似文献   

11.
The mechanism of depolymerization is one of the most essential issues in chemical engineering and materials science. In this work, we investigate the depolymerization reactions of three typical free‐radical poly(alpha‐methylstyrene) tetramers by using first‐principles density functional theory. The calculated results show that these reactions all need to overcome the energy barriers in the range of 0.58 to 0.77 eV, and that breaking the C?C bond at the chain end leads to the dissociation of alpha‐methylstyrene monomers from the polymers. Electronic‐structure analysis indicates that the reactions occur easily at the CR3 unsaturated end, and that the frontier molecular orbitals that participate in the reactions are mainly localized at the unsaturated ends. Meanwhile, spin population analysis presents the unique net spin‐transfer process in free‐radical depolymerization reactions. We hope the current findings can contribute to understanding the free‐radical depolymerization mechanism and help guide future experiments.  相似文献   

12.
By double or triple depolymerization with sodium hydrogen sulfide in the presence of sodium sulfite, polysulfide polymers of a very low molecular weight can be obtained from disulfide polymers prepared from bis-2-chloroethyl formal. The products are mixtures consisting chiefly of dimers and trimers with monomers. The effect of depolymerization will be greater when three depolymerization processes, instead of two, are completed with the same total amount of NaHS and Na2SO3. When depolymerization is repeated, the degree of depolymerization will be higher with polymers of a higher molecular weight. If depolymerization is made with sodium dithionite, the same effect will be achieved by either single or double depolymerization, provided that the same total amount of dithionite has been used.  相似文献   

13.
We evaluated polyacrylamide gel electrophoresis (PAGE) and size exclusion chromatography coupled with multi-angle laser light scattering (SEC-MALLS) approaches to determine weight-average molecular weight (M w) and polydispersity (PD) of heparins. A set of unfractionated heparin sodium (UFH) and low-molecular-weight heparin (LMWH) samples obtained from nine manufacturers which supply the US market were assessed. For SEC-MALLS, we measured values for water content, refractive index increment (dn/dc), and the second virial coefficient (A 2) for each sample prior to molecular weight assessment. For UFH, a mean ± standard deviation value for M w of 16,773 ± 797 was observed with a range of 15,620 to 18,363 (n = 20, run in triplicate). For LMWHs by SEC-MALLS, we measured mean M w values for dalteparin, tinzaparin, and enoxaparin of 6,717 ± 71 (n = 4), 6,670 ± 417 (n = 3), and 3,959 ± 145 (n = 3), respectively. PAGE analysis of the same UFH, dalteparin, tinzaparin, and enoxaparin samples showed values of 16,135 ± 643 (n = 20), 5,845 ± 45 (n = 4), 6,049 ± 95 (n = 3), and 4,772 ± 69 (n = 3), respectively. These orthogonal measurements are the first M w results obtained with a large heparin sample set on product being marketed after the heparin crisis of 2008 changed the level of scrutiny of this drug class. In this study, we compare our new data set to samples analyzed over 10 years earlier. In addition, we found that the PAGE analysis of heparinase digested UFH and neat LMWH samples yield characteristic patterns that provide a facile approach for identification and assessment of drug quality and uniformity.  相似文献   

14.
2‐Methylideneglutarate mutase is an adenosylcobalamin (coenzyme B12)‐dependent enzyme that catalyses the equilibration of 2‐methylideneglutarate with (R)‐3‐methylitaconate. This reaction is believed to occur via protein‐bound free radicals derived from substrate and product. The stereochemistry of the formation of the methyl group of 3‐methylitaconate has been probed using a `chiral methyl group'. The methyl group in 3‐([2H1,3H]methyl)itaconate derived from either (R)‐ or (S)‐2‐methylidene[3‐2H1,3‐3H1]glutarate was a 50 : 50 mixture of (R)‐ and (S)‐forms. It is concluded that the barrier to rotation about the C−C bond between the methylene radical centre and adjacent C‐atom in the product‐related radical [.CH2CH(O2CC=CH2)CO2] is relatively low, and that the interaction of the radical with cob(II)alamin is minimal. Hence, cob(II)alamin is a spectator of the molecular rearrangement of the substrate radical to product radical.  相似文献   

15.
Monoallyl compounds are not readily homopolymerized by a conventional free‐radical mechanism. However, the polymerization of allylbiguanide hydrochloride was reported to proceed in a concentrated solution of hydrochloric or phosphoric acid in the presence of a radical initiator. Here we have studied the polymerization of allyl alcohol by a radical initiator in the presence of a Lewis acid (ZnCl2, CuCl2 or MgCl2) in an organic solvent (toluene, hexane, methanol or isopropanol). Reactions were performed either at room temperature or 50°C under an atmosphere of nitrogen or in a sealed tube. The same polymerization was also carried out in water and in a concentrated acid solution. The polymer product was purified by dialysis in 0.2–3.7% yield and confirmed by elemental analysis, infrared spectroscopy and 1H NMR. The molecular weight range of poly(allyl alcohol) was 10,000–35,000. The polymerization of allyl acetate by the radical initiator under the above conditions gave poly(allyl acetate) with the molecular weight range of 10,000–13,800 by multi‐angle laser light scattering. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

16.
A summary of the mass spectra of metal dithiocarbamate complex salts (ML2 and ML3) is presented. Only divalent metal dithiocarbamate ions without an electronic configuration containing an inert s-orbital electron pair exhibited both expulsion of a ligand radical (L) and the neutral even electron species (L–H) generated from the ligand via hydrogen transfer to the metal-containing fragment ion. Divalent metal dithiocarbamate ions can be generated either by direct electron impact ionization of gas phase ML2 molecules or ionization of ML3 molecules followed by loss of a ligand radical. A highly stable sp2 hybridized, gas phase ion of a monobidentate lead dithiocarbamate complex is proposed.  相似文献   

17.
Tirapazamine (TPZ) has been tested in clinical trials on radio‐chemotherapy due to its potential highly selective toxicity towards hypoxic tumor cells. It was suggested that either the hydroxyl radical or benzotriazinyl radical may form as bioactive radical after the initial reduction of TPZ in solution. In the present work, we studied low‐energy electron attachment to TPZ in the gas phase and investigated the decomposition of the formed TPZ? anion by mass spectrometry. We observed the formation of the (TPZ–OH)? anion accompanied by the dissociation of the hydroxyl radical as by far the most abundant reaction pathway upon attachment of a low‐energy electron. Quantum chemical calculations suggest that NH2 pyramidalization is the key reaction coordinate for the reaction dynamics upon electron attachment. We propose an OH roaming mechanism for other reaction channels observed, in competition with the OH dissociation.  相似文献   

18.
Preparation of liquid epoxidized natural rubber (ENR) was made by oxidative depolymerization of ENR in latex stage without loss of epoxy group. Epoxidation of fresh natural rubber latex, which was purified by deproteinization with proteolytic enzyme and surfactant, was carried out with freshly prepared peracetic acid. The glass transition temperature (Tg) and gel content of the rubbers increased after the epoxidation, both of which were dependent upon an amount of peracetic acid. The gel content was significantly reduced by oxidative depolymerization of the rubber with (NH4)2S2O8 in the presence of propanal. The resulting liquid epoxidized rubber (Mn≈104) was found to have well-defined terminal groups, i.e. aldehyde groups and α-β unsaturated carbonyl groups. The novel rubber was applied to transport Li+ as an ionic conducting medium, that is, solid polymer electrolyte.  相似文献   

19.
[CuII(Ma)(Mb)]?2+ complexes, where Ma and Mb are dipeptides or tripeptides each containing either a tryptophan (W) or tyrosine (Y) residue, have been examined by means of electrospray tandem mass spectrometry. Collision‐induced dissociations (CIDs) of complexes containing identical peptides having a tryptophan residue generated abundant radical cations of the peptides; by contrast, for complexes containing peptides having a tyrosine residue, the main fragmentation channel is dissociative proton transfer to give [Ma + H]+ and [CuII(Mb – H)]?+. When there are two different peptides in the complex, each containing a tryptophan residue, radical cations are again the major products, with their relative abundances depending on the locations of the tryptophan residue in the peptides. In the CIDs of mixed complexes, where one peptide contains a tryptophan residue and the other a tyrosine residue, the main fragmentation channel is formation of the radical cation of the tryptophan‐containing peptide and not proton transfer from the tyrosine‐containing peptide to give a protonated peptide. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

20.
Lignin modifications resulting from different photochemical pre-treatments were studied using chemiluminescent methods. In the oxidation of lignin in NaOH solutions at 25°C, the intensity increased with increasing temperature and can be described by an Arrheniustype exponential equation with an activation energy of 25.8 ± 2.7 kJ/ mol. The oxidation of lignin model compounds under these conditions indicated 1O2 OH′, and O2 generation. Chemiluminescence of the luminol/H2O2/Fe2+ system was used to study decomposition products of lignin upon irradiation. Unirradiated lignin proved to be an excellent radical trap, an effect initially abolished upon irradiation. At longer irradiation times, however, the radical trapping behavior was restored. The action of the perodidase/H2 O2 system onlignin was also investigated using chemiluminescence. Behavior very similar to that using luninol was observed. The intensity increases with increasing time of irradiation up to an optimum value. More prolonged irradiation results in total quenching of the cheiluminescemce. This is indicative of depolymerization and posterior aggregation.  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