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 27 毫秒
1.
The N-hydroxymethylation, N-acetylation, and N-acetoxymethylation of mono-, di-, and trialkylglycolurils by reaction with the electrophilic reagents formaldehyde and acetaldehyde have been studied. General methods have been developed for the preparation of mono-, di-, and tri-N-hydroxymethylglycolurils by treatment of differently substituted N-alkylglycolurils with formaldehyde (as hemiformal in methanol) and the synthesis of di-N-and tri-N-acetyl-or N-acetoxymethylglycolurils via the electrophilic substitution of hydrogen atoms for an acetyl group at the nitrogen or oxygen atoms in the hydroxymethyl groups of glycolurils using acetic anhydride. The regioselectivity of the reaction of the 2-t-Bu-and 2-c-C6H11-glycolurils with formaldehyde has been shown to yield a 4,6-di(hydroxymethyl) derivative. It was found that the hydroxymethylation of 2,4-and 2,6-dialkylglycolurils occurs regioselectively with a stoichiometric ratio of glycoluril to hemiformal and permits preparation of their mono-and dihydroxymethyl derivatives. The enantiomeric analysis of the obtained compounds has been carried out for the first time using HPLC on chiral phases. X-ray analysis has been carried out on the previously unreported racemic 2,6-diacetoxymethyl-4,8-dimethylglycoluril. __________ Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 3, pp. 411–423, March, 2006.  相似文献   

2.
Selected synthesis parameters of typical wood adhesive‐type urea–formaldehyde (UF) resins were examined using the 13C NMR spectroscopy. The monomeric hydroxymethylureas and methylene–ether derivatives formed in the initial alkaline reaction polymerize in the subsequent acidic reaction by forming methylene bonds and cleaving some hydroxymethyl groups as formaldehyde. For typical resin syntheses at F/U ratio of 2.10, the resulting UF polymer is found to be a number‐averaged pentamer having 3.25 polymer chain branches with about 80% of chain ends bonded to hydroxymethyl groups and the rest being free amide groups. When the second urea is added during the cooling period, about 67% of hydroxymethyl groups cleave from the UF polymeric components and the freed formaldehyde reacts with second urea to form monomeric hydroxymethylureas. This hydroxymethyl group move is suppressed when the second urea is added at low temperatures, suggesting that wood adhesive‐type UF resins are composed of monomeric and polymeric UF components having hydroxymethyl functional groups in varying proportions. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 995–1007, 1999  相似文献   

3.
Wenhua Huang  Jie Xu 《合成通讯》2013,43(15):1777-1782
The reaction of (hydroxymethyl)triphenylphosphonium with benzylic or allylic halide under basic conditions at room temperature affords terminal alkenes in 61–89% yields. In this reaction, both formaldehyde and triphenylphosphine are in situ generated from (hydroxymethyl)triphenylphosphonium and further undergo Wittig olefination with benzylic or allylic halide.  相似文献   

4.
The gas-phase reaction of the synthesis of tris(hydroxymethyl)phosphine from phosphine and formaldehyde was studied using a calculation scheme based on the density functional theory with hybrid exchange-correlation functional B3LYP in the 6–311++G** basis. The reaction was shown to proceed with the participation of unstable intermediates containing three-membered ring. The transformation into final products includes opening of the three-membered ring and intramolecular proton transfer. The results can be useful at selecting catalysts and for explaining the mechanism of the catalytic reaction of tris(hydroxymethyl)phosphine synthesis.  相似文献   

5.
The hydrolysis of hydroxymethyl cyanoguanidine and methoxymethyl cyanoguianidine in acidic aqueous media were studied by analyzing the eluted formaldehyde and the reaction products. Dihydroxymethyl cyanoguanidine released two formaldehydes with different rates because one of the two hydroxymethyl groups is intramolecularly interactive through the hydrogen bond. Methanol dissociation from methoxymethyl cyanoguanidine was seven to eight times faster than the dissociation of hydroxymethyl group, and the overall hydrolysis kinetics were similar to those of hydroxymethyl cyanoguanidine.  相似文献   

6.
2‐((2,3‐Dihydrothieno[3,4‐b]dioxin‐2‐yl)methoxy)methyl oxirane (EDOT‐MO) was successfully synthesized by the reaction of epichlorohydrin with hydroxymethylated‐3,4‐ethylenedioxylthiophene (EDOT‐MeOH), which was synthesized via a simple four‐step sequence. Poly(hydroxymethylated‐3,4‐ethylenedioxylthiophene) (PEDOT‐MeOH) and poly(2‐((2,3‐dihydrothieno[3,4‐b]dioxin‐2‐yl)methoxy)methyl oxirane) (PEDOT‐MO) were electrosynthesized through electropolymerization of EDOT‐MeOH and EDOT‐MO, respectively. Structural, electrochemical, optical, and thermal properties of as‐formed polymers were investigated by FTIR, cyclic voltammetry, UV–vis, and thermogravimetry. Spectroelectrochemistry studies demonstrated that PEDOT‐MeOH and PEDOT‐MO could be reversibly oxidized and reduced accompany with obvious color changes. Further kinetic studies demonstrated that the introduction of hydroxymethyl or ethylene oxide group significantly improved electrochromic properties of 3,4‐ethylenedioxythiophene (PEDOT) and resulted in high contrast ratios (57.3% at 585 nm) and coloration efficiencies (338.5 cm2 C?1), low switching voltages, and fast response time. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 1989–1999  相似文献   

7.
N,N′-Bis(hydroxymethyl)thiourea reacted with propane-1,3-diamine at a molar ratio of 2 : 1 to give 5,5′-propane-1,3-diylbis(1,3,5-triazinane-2-thione), whereas 1,3,5,7,11,13,15,17-octaazatricyclo[15.3.1.17,11]-docosane-4,14-dithione was obtained in the reaction with equimolar amounts of the reactants. Tricyclic product was also formed in the three-component condensation of thiourea with formaldehyde and propane-1,3-diamine at a ratio of 1 : 3 : 1. The reactions of N,N′-bis(hydroxymethyl)thiourea with ethane-1,2-diamine (2 : 1) and of thiourea with formaldehyde and butane-1,4-diamine (1 : 2 : 1) afforded 5,5′-(ethane-1,2-diyl)bis(1,3,5-triazinane-2-thione) and 5,5′-(butane-1,4-diyl)bis(1,3,5-triazinane-2-thione), respectively.  相似文献   

8.
The results of studies on the kinetics of reaction of 1 mol of parabanic acid (imidazolidine‐2,4,5‐trione) with 1 mol of ethylene oxide (oxirane) or propylene oxide (2‐methyloxirane) carried out in the presence of triethylamine catalyst in dimethylsulfoxide solution are presented. A rate equation describing the reaction is presented and the mechanism proposed. The validity of the proposed mechanism is proved by instrumental analytical methods. The effect of temperature is also presented and the thermodynamic parameters of the reaction evaluated. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 35: 73–80, 2003  相似文献   

9.
Acyclic nucleosides of 4‐nitro‐1H‐imidazole and 4‐nitropyrazole have been synthesized by nucleophilic addition of the appropriate 4‐nitroazole to (?)‐(S)‐(hydroxymethyl)oxirane in the presence of a catalytic amount of potassium carbonate. (+)‐(R)‐3‐(4‐nitro‐1H‐imidazol‐1‐yl)propane‐1,2‐diol and (+)‐(R)‐3‐(2‐methyl‐4‐nitro‐1H‐imidazol‐1‐yl)propane‐1,2‐diol were also obtained in an independent reaction starting from appropriate 1,4‐dinitro‐1H‐imidazole and (+)‐(R)‐3‐aminopropane‐1,2‐diol. (+)‐(R)‐3‐(4‐Nitropyrazol‐1‐yl)propane‐1,2‐diol was also obtained by direct noncatalyzed addition of 4‐nitropyrazole to (?)‐(S)‐(hydroxymethyl)oxirane, whereas the (S)‐enantiomer was obtained by reaction of 4‐nitropyrazole with (+)‐(S)‐1,2‐O‐isopropylideneglycerol under Mitsunobu reaction conditions, followed by a cleavage of the isopropylidene group with 80% AcOH. Racemization during any of these syntheses has not been observed. 3‐(4‐Nitroazol‐1‐yl)propane‐1,2‐diols were incorporated into a 26‐mer oligonucleotide. UV Thermal melting studies of duplexes of the oligonucleotides with 4‐nitropyrazole or 4‐nitro‐1H‐imidazole paired with four natural bases showed moderately decreased stabilities of the duplexes. A narrow range of melting temperatures, typically being within 2° for each acyclic nucleoside, fulfill one of the requirements of using acyclic 4‐nitroazoles as general bases. Single incorporation of 4‐nitroazoles into a 14‐mer triplex forming oligonucleotide resulted in considerably decreased triplex stabilities.  相似文献   

10.
Abstract

Novel P-H-diphosphinoalkanes with pyridyl substituents at the phosphorus atoms have been obtained in good or moderate yields by the interaction of primary 2-pyridyl-phosphine with dichloroalkanes in superbasic medium. The formation of novel 1-(2-pyridyl)-1-phosphacyclopentane together with bis(2-pyridylphosphino)butane was observed as the result of the interaction of 2-pyridylphosphine and 1,4-dichlorobutane under the same reaction conditions. The reaction of the bis(2-pyridylphosphino)alkanes thus obtained with formaldehyde leads to the formation of the corresponding hydroxymethyl derivatives.  相似文献   

11.
In this three-component reaction, alkynes undergo a copper(I)-catalyzed cycloaddition with sodium azide and formaldehyde to yield 2-hydroxymethyl-2 H-1,2,3-triazoles, which are useful intermediates that can be readily converted to polyfunctional molecules. The hydroxymethyl group can also be removed, providing convenient access to N H-1,2,3-triazoles. The reaction is experimentally simple and readily scalable.  相似文献   

12.
Bare vanadium oxide and hydroxide cluster cations, V(m)O(n)+ and V(m)O(n-1) (OH)+ (m = 1-4, n = 1-10), generated by electrospray ionization, were investigated with respect to their reactivity toward methanol using mass spectrometric techniques. Several reaction channels were observed, such as abstraction of a hydrogen atom, a methyl radical, or a hydroxymethyl radical, elimination of methane, and adduct formation. Moreover, dehydrogenation of methanol to generate formaldehyde was found to occur via four different pathways. Formaldehyde was released as a free molecule either upon transfer of two hydrogen atoms to the cluster or upon transfer of an oxygen atom from the cluster to the neutral alcohol concomitant with elimination of water. Further, formaldehyde was attached to V(m)O(n)+ upon loss of H2 or neutral water to produce the cation V(m)O(n)(OCH(2))+ or V(m)O(n-1) (OCH(2))+, respectively. A reactivity screening revealed that only high-valent vanadium oxide clusters are reactive with respect to H2 uptake, oxygen transfer, and elimination of H2O, whereas smaller and low-valent cluster cations are capable of dehydrogenating methanol via elimination of H2. For comparison, the reactivity of methanol with the corresponding hydroxide cluster ions, V(m)O(n-1) (OH)+, was studied also, for which dominant pathways lead to both condensation and association products, i.e., generation of the ions V(m)O(n-1) (OCH(3))+ and V(m)O(n-1) (OH)(CH(3)OH)+, respectively.  相似文献   

13.
A series of (hydroxymethyl)hydroxyphenyldimethylsiloxanes of linear and branched structures were synthesized by the reaction with formaldehyde in aqueous alkaline medium of organosilicon phenols of general formula R′[Si(CH3)2O] n [Si(CH3)RO] m SiMe2R′, where R′ was 4-hydroxy-3-methoxyphenylpropyl, R was methyl or 4-hydroxy-3-methoxyphenylpropyl. The structure and composition of the siloxanes were confirmed by elemental analysis, NMR and IR spectroscopy. The homocondensation of (hydroxymethyl)hydroxyphenyldimethylsiloxanes at the hydroxymethyl groups was investigated. A possibility of reaction of (hydroxymethyl) hydroxyphenyldimethylsiloxanes with phenol-formaldehyde resin to form copolymers was demonstrated.  相似文献   

14.
(2-Chloroethyl)oxirane, (3-chloropropyl)oxirane, and (4-chlorobutyl)oxirane were prepared from the corresponding alkenols and polymerized using a triethylaluminum-water-acetylacetone initiator system. Copolymerizations with propylene oxide and epichlorohydrin were also accomplished; the copolymerization activities of the (chloroalkyl)oxiranes were similar to the activity of epichlorohydrin. Poly[(2-chloroethyl)oxirane], poly[(3-chloropropyl)oxirane], and poly[(4-chlorobutyl)oxirane] exhibit elastomeric properties and are highly reactive in nucleophilic substitution reactions.  相似文献   

15.
Summary In the reaction of formaldehyde with 1,1-dichloro-l-alkenes in concentrated sulfuric acid, the corresponding -(hydroxymethyl) carboxylic acids were formed.  相似文献   

16.
以D-(+)-葡萄糖酸内酯为原料,经三甲硅基保护羟基后与5-溴-2-氯-4′-乙氧基二苯甲烷偶联制得(2S,3R,4S,5S,6R)-2-[4-氯-3-(4-乙氧苄基)苯基]-6-(羟甲基)-2-甲氧基四氢-2H-吡喃-3,4,5-三醇(2); 2经羟基保护、氧化和羟醛缩合等5步反应制得(3S,4S,5R,6S)-3,4,5-三(苄氧基)-6-[4-氯-3-(4-乙氧苄基)苯基]-2-(羟甲基)-6-甲氧基四氢-2H-吡喃-2-甲醛(7); 7经还原、脱苄同时关环制得埃格列净(1S,2S,3S,4R,5S)-5-[4-氯-3-(4-乙氧苄基)苯基]-1-(羟甲基)-6,8-二氧杂二环[3.2.1]辛烷-2,3,4-三醇,其结构经1H NMR和LC-MS表征。  相似文献   

17.
Absolute configurations of enantiomeric methylenecyclopropanecarbinols obtained by reaction of (R)- and (S)-epichlorohydrin 5 with methylenetriphenylphosphorane or resolution of the corresponding oxaphospholane 6 via a salt with L-(+)-tartaric acid and subsequent Wittig transformation with formaldehyde were revised. The (-)-oxaphospholane 6 has the S,S and (-)-(methylenecyclopropyl)carbinol (4) the R configuration. The configurations of (+)-6 and (+)-4 are then R,R and S, respectively. These assignments are in accord with an initial attack of phosphorane at the oxirane ring of epichlorohydrin. An improved preparation of key enantiomeric intermediates (R)-1a and (S)-1a, important for synthesis of antiviral purine methylenecyclopropane analogues of nucleosides, is also described.  相似文献   

18.
The reaction mechanisms of phenol with formaldehyde in the first and second addition at the ortho- and para-position in acid solution were theoretically investigated at the PW91/DNP level with solvent effects included. The reaction of phenol with protonated methanediol firstly forms an adduct intermediate, via a SN2 mechanism with a water molecule as the leaving group. From the adduct intermediate, there are two reaction channels involving a proton transfer to form the addition products. One is that a proton directly transfers via a four-membered ring transition state with a notable energy barrier (Four-member mechanism). Another mechanism involving a water molecule as catalyst to mediate the proton transfer (WCP mechanism), is a barrierless process, indicating that the formation of the adduct intermediate, the first reaction step, is rate-limiting. The reaction products are free hydroxymethyl phenols and/or hydroxybenzy carbocation (HOC6H4CH2+) which plays an important role in the following formation of methylene and methylene ether linkages. The second addition reactions between formaldehyde and hydroxymethyl phenol at all possible reaction sites of the phenol ring in acid solution were also investigated and discussed.  相似文献   

19.
The condensation reaction between α-keto-β-aroyl (or acyl) -γ-butyrolactones, 4a-4e and o-phenylenediamine or 2, 3-diaminonaphthalene leads under retrograde aldol condensation involving loss of formaldehyde to formation of 3-substituted-3, 4-dihydro-2 (1H) quinoxalinones or benzo [g] quinoxalinones, 7a-7g , respectively as a new convenient synthesis of this type of heterocyclic systems. The reaction of type 4 compound with 4, 5-diaminopyromidine, 8 , was found to proceed differently. 2-[(4-Amino-5-pyrimidinyl)amine]-4-oxo-3-(hydroxymethyl)-4-phenyl-2-butenoic acid 9 was the only product formed when the reaction between 4a and 8 was run in ethanol. The same reaction in glacial acetic acid proceeds with loss of formaldehyde, to afford 7-phenacylidene-7,8-dihydro-6 (1H)-pteridione 10 . The reaction between type 4 compounds and ethylenediamine or 1, 4-phenylenediamine leads to the formation of the bis-condensation products 13–15 , respectively.  相似文献   

20.
The mechanism of the reaction of phenols with formaldehyde was studied by computer simulation. At first starting molecules were stored in a computer and the hypothetical reaction yielded conditions like the reactivity ratio of hydroxymethyl group to formaldehyde and that of orthohydrogen to parahydrogen. The molecular weight distribution of the hypothetical product in the computer was compared with that of the prepared resin determined from GPC measurement. Reaction mechanisms were discussed. We also confirmed, by computer simulation, that the rate of methylenation is larger than that of hydroxymethylation in an acid-catalyzed system and that the reactivity ratio of hydroxymethyl group to formaldehyde is 5–12 for the reaction of phenols such as o-cresol, p-cresol, and phenol with formaldehyde. The opposite results were obtained in a base-catalyzed system. It also became apparent that information regarding molecular structures, such as the number of branches, the number of phenolic nuclei in the longest chain, and the number of o,o′-,o,p′- and p,p′-methylene linkages, can be obtained by computer simulation. The most probable values of these factors for 10 mer of a phenol–formaldehyde condensation molecule are 2, 7, 2, 5, and 2.  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