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.
Mono‐ and bis‐decylated lumazines have been synthesized and characterized. Namely, mono‐decyl chain [1‐decylpteridine‐2,4(1,3H)‐dione] 6a and bis‐decyl chain [1,3‐didecylpteridine‐2,4(1,3H)‐dione] 7a conjugates were synthesized by nucleophilic substitution (SN2) reactions of lumazine with 1‐iododecane in N,N‐dimethylformamide (DMF) solvent. Decyl chain coupling occurred at the N1 site and then the N3 site in a sequential manner, without DMF condensation. Molecular orbital (MO) calculations show a p‐orbital at N1 but not N3, which along with a nucleophilicity parameter (N) analysis predict alkylation at N1 in lumazine. Only after the alkylation at N1 in 6a , does a p‐orbital on N3 emerge thereby reacting with a second equivalent of 1‐iododecane to reach the dialkylated product 7a . Data from NMR (1H, 13C, HSQC, HMBC), HPLC, TLC, UV‐vis, fluorescence and density functional theory (DFT) provide evidence for the existence of mono‐decyl chain 6a and bis‐decyl chain 7a . These results differ to pterin O‐alkylations (kinetic control), where N‐alkylation of lumazine is preferred and then to dialkylation (thermodynamic control), with an avoidance of DMF solvent condensation. These findings add to the list of alkylation strategies for increasing sensitizer lipophilicity for use in photodynamic therapy.  相似文献   

2.
Bis(trimethylammonium) alkane diiodides dynamically encapsulate dicarboxylic acids through intermolecular hydrogen bonds between the I? anions of the hosts and the carboxylic OH groups of the guests. A selective recognition is realized when the size of the I????HOOC(CH2/CF2)nCOOH???I? superanion matches the dication alkyl chain length. Dynamic recognition is also demonstrated in solution, where the presence of the size‐matching organic salt boosts the acid solubility profile, thus allowing efficient mixture separation.  相似文献   

3.
The paper deals with the synthesis and characterization of a new series of anhydrous conducting acid‐doped complex membranes based on polyimide (PI) and ionic liquid (IL) for high‐temperature fuel cells via a new route. For this purpose, three imidazolium‐based ILs (RIm+BF4?) with different alkyl chain lengths (R=methyl, ethyl, and butyl) are added into polyamic acid (PAA) intermediate prepared from the reaction of benzophenonetetracarboxylic dianhydride and diaminodiphenylsulfone in different –COOH/imidazolium molar ratios (n = 0.5, 1, and 2). Then, the thermally imidized complex membrane was doped with H2SO4. The conductivities of acid‐doped PI/IL complex membranes prepared by taking n of 1 are found to be in the range of 10?4?10?5 S cm?1 at 180°C, whereas the acid‐free PI/IL complex membranes show the conductivity at around 10?9?10?10 S cm?1. Thermogravimetric analysis results reveal that the acid‐doped PI/IL complex membranes are thermally stable up to 250°C. Dynamic mechanical analysis results of the acid‐doped ionically interacted complex membrane show that the mechanical strengths of the PI/IL complex membranes including 1‐methyl imidazolium tetrafluoroborate (MeIm‐BF4) and 1‐ethyl 3‐methyl imidazolium tetrafluoroborate (EtIm‐BF4) are comparable with those of pristine PI until 200°C. Furthermore, it can be clearly emphasized that the ionic interaction between carboxylic acid groups of PAA's and IL's cations offers a positive role in long‐term conductivity stability by preventing the IL migration at high temperatures. On the other hand, preliminary methanol permeability tests of the acid‐doped membranes show that they can also be considered as an alternative for direct methanol fuel cells. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

4.
Multiple bonds between boron and transition metals are known in many borylene (:BR) complexes via metal dπ→BR back‐donation, despite the electron deficiency of boron. An electron‐precise metal–boron triple bond was first observed in BiB2O? [Bi≡B?B≡O]? in which both boron atoms can be viewed as sp‐hybridized and the [B?BO]? fragment is isoelectronic to a carbyne (CR). To search for the first electron‐precise transition‐metal‐boron triple‐bond species, we have produced IrB2O? and ReB2O? and investigated them by photoelectron spectroscopy and quantum‐chemical calculations. The results allow to elucidate the structures and bonding in the two clusters. We find IrB2O? has a closed‐shell bent structure (Cs, 1A′) with BO? coordinated to an Ir≡B unit, (?OB)Ir≡B, whereas ReB2O? is linear (C∞v, 3Σ?) with an electron‐precise Re≡B triple bond, [Re≡B?B≡O]?. The results suggest the intriguing possibility of synthesizing compounds with electron‐precise M≡B triple bonds analogous to classical carbyne systems.  相似文献   

5.
In the title compound, poly­[sodium‐μ4‐3,5‐di­carboxy­benzene­sulfonato‐κ4O:O′:O′′:O′′′‐μ2‐urea‐κ2O:N] monohydrate], {[Na(C8H5O7S)(CH4N2O)]·H2O}n, the organic anions are arranged almost vertically within (001) monolayers, with the sulfonate and carboxylic acid groups pointing into the interlayer region. The inversion‐related aromatic rings of the anions inside the layers are arrayed via offset face‐to‐face interactions into molecular stacks along the crystallographic a axis. The `up' and `down' arrangement of the aromatic portions makes both faces of the layers ionic and hydro­philic, whereas the interiors of the layers are primarily hydro­phobic. The interleaving of the anions is such that the carboxylic acid groups are oriented more toward the interior than are the sulfonate groups. The aromatic rings in neighbouring layers are arranged in a herring‐bone fashion. The coordination sphere of the Na+ ions contains two sulfonate and two carboxylic acid O atoms, from a total of four different acid anions belonging to two neighbouring anionic monolayers. The urea mol­ecules are positioned between translation‐related anionic stacks inside the (001) layers, serving a triple function, viz. they fill in the large meshes (empty cavities) formed within the anionic–cationic network, and they provide additional Na+ coordination and hydrogen‐bond sites.  相似文献   

6.
A new heme–thiolate peroxidase catalyzes the hydroxylation of n‐alkanes at the terminal position—a challenging reaction in organic chemistry—with H2O2 as the only cosubstrate. Besides the primary product, 1‐dodecanol, the conversion of dodecane yielded dodecanoic, 12‐hydroxydodecanoic, and 1,12‐dodecanedioic acids, as identified by GC–MS. Dodecanal could be detected only in trace amounts, and 1,12‐dodecanediol was not observed, thus suggesting that dodecanoic acid is the branch point between mono‐ and diterminal hydroxylation. Simultaneously, oxygenation was observed at other hydrocarbon chain positions (preferentially C2 and C11). Similar results were observed in reactions of tetradecane. The pattern of products formed, together with data on the incorporation of 18O from the cosubstrate H218O2, demonstrate that the enzyme acts as a peroxygenase that is able to catalyze a cascade of mono‐ and diterminal oxidation reactions of long‐chain n‐alkanes to give carboxylic acids.  相似文献   

7.
Amphiphilic copolymers were obtained by grafting arborescent poly(γ‐benzyl l ‐glutamate) (PBG) cores of generations G1–G3 with polyglycidol, poly(ethylene oxide) (PEO), or poly(l ‐glutamic acid) (PGA) chain segments. The PBG substrates were synthesized by two methods: (1) subjecting PBG samples with a dispersity ? = Mw/Mn < 1.1 to partial acidolysis of the benzyl ester groups, to produce randomly distributed carboxylic acid functionalities, and (2) using PBG chains containing a glutamic acid di‐tert‐butyl ester initiator fragment in the last grafting cycle of the PBG core synthesis, and selective acidolysis of the tert‐butyl ester groups to obtain substrates with carboxylic acid termini. Linear polymers with ? < 1.20 and a primary amine terminus were also synthesized to serve as hydrophilic shell materials: Polyglycidol and PEO by anionic polymerization, and PGA by N‐carboxyanhydride ring‐opening polymerization. These polymers, combined with the two different PGB substrate types, allowed the evaluation of the usefulness of random versus chain‐end grafting in producing arborescent copolymers useful as unimolecular micelles in organic and aqueous media. Size exclusion chromatography served to determine the grafting yield, molar mass, dispersity, and branching functionality of the copolymers. Dynamic light scattering measurements provided information on their aggregation behavior in aqueous environments. © 2015 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 1197–1209  相似文献   

8.
The first part of the catalytic cycle of the pterin‐dependent, dioxygen‐using nonheme‐iron aromatic amino acid hydroxylases, leading to the FeIV?O hydroxylating intermediate, has been investigated by means of density functional theory. The starting structure in the present investigation is the water‐free Fe? O2 complex cluster model that represents the catalytically competent form of the enzymes. A model for this structure was obtained in a previous study of water‐ligand dissociation from the hexacoordinate model complex of the X‐ray crystal structure of the catalytic domain of phenylalanine hydroxylase in complex with the cofactor (6R)‐L ‐erythro‐5,6,7,8‐tetrahydrobiopterin (BH4) (PAH‐FeII‐BH4). The O? O bond rupture and two‐electron oxidation of the cofactor are found to take place via a Fe‐O‐O‐BH4 bridge structure that is formed in consecutive radical reactions involving a superoxide ion, O2?. The overall effective free‐energy barrier to formation of the FeIV?O species is calculated to be 13.9 kcal mol?1, less than 2 kcal mol?1 lower than that derived from experiment. The rate‐limiting step is associated with a one‐electron transfer from the cofactor to dioxygen, whereas the spin inversion needed to arrive at the quintet state in which the O? O bond cleavage is finalized, essentially proceeds without activation.  相似文献   

9.
The gas‐phase elimination kinetics of the ethyl ester of two α‐amino acid type of molecules have been determined over the temperature range of 360–430°C and pressure range of 26–86 Torr. The reactions, in a static reaction system, are homogeneous and unimolecular and obey a first‐order rate law. The rate coefficients are given by the following equations. For N,N‐dimethylglycine ethyl ester: log k1(s?1) = (13.01 ± 3.70) ? (202.3 ± 0.3)kJ mol?1 (2.303 RT)?1 For ethyl 1‐piperidineacetate: log k1(s?1) = (12.91 ± 0.31) ? (204.4 ± 0.1)kJ mol?1 (2.303 RT)?1 The decompositon of these esters leads to the formation of the corresponding α‐amino acid type of compound and ethylene. However, the amino acid intermediate, under the condition of the experiments, undergoes an extremely rapid decarboxylation process. Attempts to pyrolyze pure N,N‐dimethylglycine, which is the intermediate of dimethylglycine ethyl ester pyrolysis, was possible at only two temperatures, 300 and 310°C. The products are trimethylamine and CO2. Assuming log A = 13.0 for a five‐centered cyclic transition‐state type of mechanism in gas‐phase reactions, it gives the following expression: log k1(s?1) = (13.0) ? (176.6)kJ mol?1 (2.303 RT)?1. The mechanism of these α‐amino acids differs from the decarbonylation elimination of 2‐substituted halo, hydroxy, alkoxy, phenoxy, and acetoxy carboxylic acids in the gas phase. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33:465–471, 2001  相似文献   

10.
Carboxylic acid capped γ‐Fe2O3 nanoparticles were prepared by the standard decomposition of Fe(CO)5 in di‐n‐octyl ether and oleic acid. Two methods were employed to introduce surface functionality to the nanoparticles. First, a thermally stable, tert‐butyldiphenylsilyl‐protected hydroxyl group was incorporated into the carboxylic acid surfactant used during the synthesis. Subsequent deprotection and transformation installed a 2‐bromopropionyl ester group on the particle surface (the functional‐group‐interchange method). The resulting nanoparticles were 4.53 nm in average diameter and were characterized with IR spectroscopy, transmission electron microscopy (TEM), selected area electron diffraction, and elemental analysis. Second, a 2‐bromopropionyl ester group was installed on the particle surface after synthesis via the exchange of the surface oleic acid with a carboxylic acid containing the desired 2‐bromopropionyl ester unit (the ligand‐exchange method). The resulting nanoparticles were 4.30 nm in average diameter and were characterized with IR spectroscopy, TEM, and elemental analysis. Monitoring the percentage of bromine incorporated into the nanoparticle sample versus the ligand‐exchange reaction time indicated that the number of initiator‐containing carboxylic acids that could be exchanged onto the surface was limited, presumably by the steric size of the 2‐bromopropionyl ester group. Styrene was then polymerized directly off γ‐Fe2O3 nanoparticles, and this yielded hybrid core–shell structures. The measurements of the magnetic properties of the samples demonstrated that the magnetism of the core γ‐Fe2O3 nanoparticle did not change during the performance of the chemical transformations. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 43: 3675–3688, 2005  相似文献   

11.
A reliable and practical procedure for FeCl3‐promoted ester cleavage has been developed. Lewis acids including TiCl4, ZnO and FeCl3 etc. were investigated as promoters for O‐alkyl cleavage of carboxylic acid ester. Under optimal reaction conditions, FeCl3 (1.5 equiv.) was found to possess the highest activity and efficiently enhanced dealkylation of aryl esters, alkyl esters and aromatic heterocyclic esters to give their corresponding carboxylic acids in 54–98% yield, the method provides a complementary access to dealkylation of ester under neutral condition. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

12.
In solid‐state engineering, cocrystallization is a strategy actively pursued for pharmaceuticals. Two 1:1 cocrystals of 5‐fluorouracil (5FU; systematic name: 5‐fluoro‐1,3‐dihydropyrimidine‐2,4‐dione), namely 5‐fluorouracil–5‐bromothiophene‐2‐carboxylic acid (1/1), C5H3BrO2S·C4H3FN2O2, (I), and 5‐fluorouracil–thiophene‐2‐carboxylic acid (1/1), C4H3FN2O2·C5H4O2S, (II), have been synthesized and characterized by single‐crystal X‐ray diffraction studies. In both cocrystals, carboxylic acid molecules are linked through an acid–acid R 22(8) homosynthon (O—H…O) to form a carboxylic acid dimer and 5FU molecules are connected through two types of base pairs [homosynthon, R 22(8) motif] via a pair of N—H…O hydrogen bonds. The crystal structures are further stabilized by C—H…O interactions in (II) and C—Br…O interactions in (I). In both crystal structures, π–π stacking and C—F…π interactions are also observed.  相似文献   

13.
The title compound, 2‐amino‐5‐carboxy­pyridinium chloride, C6H7N2O2+·Cl?, was isolated from a 1 M HCl aqueous solution containing 2‐amino‐5‐cyano­pyridine. The structure is held together by extensive hydrogen bonding between the chloride ions and the carboxylic acid, amino and pyridinium H atoms. The mol­ecules pack as sheets, with the sheets at a distance of 3.21 (3) Å from one another.  相似文献   

14.
Gas–liquid mass transfer of gaseous reactants is a major limitation for high space–time yields, especially for O2‐dependent (bio)catalytic reactions in aqueous solutions. Herein, oxygenic photosynthesis was used for homogeneous O2 supply via in situ generation in the liquid phase to overcome this limitation. The phototrophic cyanobacterium Synechocystis sp. PCC6803 was engineered to synthesize the alkane monooxygenase AlkBGT from Pseudomonas putida GPo1. With light, but without external addition of O2, the chemo‐ and regioselective hydroxylation of nonanoic acid methyl ester to ω‐hydroxynonanoic acid methyl ester was driven by O2 generated through photosynthetic water oxidation. Photosynthesis also delivered the necessary reduction equivalents to regenerate the Fe2+ center in AlkB for oxygen transfer to the terminal methyl group. The in situ coupling of oxygenic photosynthesis to O2‐transferring enzymes now enables the design of fast hydrocarbon oxyfunctionalization reactions.  相似文献   

15.
2H, 31P, and 1H‐magic‐angle‐spinning (MAS) solid‐state NMR spectroscopic methods were used to elucidate the interaction between sorbic acid, a widely used weak acid food preservative, and 1,2‐dimyristoyl‐sn‐glycero‐3‐phosphocholine (DMPC) bilayers under both acidic and neutral pH conditions. The linewidth broadening observed in the 31P NMR powder pattern spectra and the changes in the 31P longitudinal relaxation time (T1) indicate interaction with the phospholipid headgroup upon titration of sorbic acid or decanoic acid into DMPC bilayers over the pH range from 3.0 to 7.4. The peak intensities of sorbic acid decrease upon addition of paramagnetic Mn2+ ions in DMPC bilayers as recorded in the 1H MAS NMR spectra, suggesting that sorbic acid molecules are in close proximity with the membrane/aqueous surface. No significant 2H quadrupolar splitting (ΔνQ) changes are observed in the 2H NMR spectra of DMPC‐d54 upon titration of sorbic acid, and the change of pH has a slight effect on ΔνQ, indicating that sorbic acid has weak influence on the orientation order of the DMPC acyl chains in the fluid phase over the pH range from 3.0 to 7.4. This finding is in contrast to the results of the decanoic acid/DMPC‐d54 systems, where ΔνQ increases as the concentration of decanoic acid increases. Thus, in the membrane association process, sorbic acids are most likely interacting with the headgroups and shallowly embedded near the top of the phospholipid headgroups, rather than inserting deep into the acyl chains. Thus, antimicrobial mode of action for sorbic acid may be different from that of long‐chain fatty acids. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

16.
The structures of the 1:1 proton‐transfer compounds of 4,5‐dichlorophthalic acid with 8‐hydroxyquinoline, 8‐aminoquinoline and quinoline‐2‐carboxylic acid (quinaldic acid), namely anhydrous 8‐hydroxyquinolinium 2‐carboxy‐4,5‐dichlorobenzoate, C9H8NO+·C8H3Cl2O4, (I), 8‐aminoquinolinium 2‐carboxy‐4,5‐dichlorobenzoate, C9H9N2+·C8H3Cl2O4, (II), and the adduct hydrate 2‐carboxyquinolinium 2‐carboxy‐4,5‐dichlorobenzoate quinolinium‐2‐carboxylate monohydrate, C10H8NO2+·C8H3Cl2O4·C10H7NO2·H2O, (III), have been determined at 130 K. Compounds (I) and (II) are isomorphous and all three compounds have one‐dimensional hydrogen‐bonded chain structures, formed in (I) through O—H...Ocarboxyl extensions and in (II) through N+—H...Ocarboxyl extensions of cation–anion pairs. In (III), a hydrogen‐bonded cyclic R22(10) pseudo‐dimer unit comprising a protonated quinaldic acid cation and a zwitterionic quinaldic acid adduct molecule is found and is propagated through carboxylic acid O—H...Ocarboxyl and water O—H...Ocarboxyl interactions. In both (I) and (II), there are also cation–anion aromatic ring π–π associations. This work further illustrates the utility of both hydrogen phthalate anions and interactive‐group‐substituted quinoline cations in the formation of low‐dimensional hydrogen‐bonded structures.  相似文献   

17.
The crystal structures of 7,7‐dicyclo­but­yl‐5‐hydroxy­meth­yl‐6‐oxabicyclo­[3.2.1]octa­ne‐1‐carboxylic acid, C17H26O4, (I), and 1‐(hydroxy­meth­yl)‐7‐oxaspiro­[bicyclo­[3.2.1]octa­ne‐6,1′‐cyclo­penta­ne]‐5‐carboxylic acid, C13H20O4, (II), determined at 170 K, show that the conformation of the hydroxy­meth­yl group (anti or gauche) affects the dimensionality (one‐ or two‐dimensional) of the supramolecular structures via O—H⋯O hydrogen bonds. In (I), the carbox­yl and hydroxy­meth­yl groups inter­act with themselves, forming a one‐dimensional step‐ladder, while in (II), a two‐dimensional structure is made up of carboxylic acid centrosymmetric R22(8) dimers connected by hydrox­yl‐to‐ether contacts.  相似文献   

18.
The crystal structure determination of the molecular proton‐transfer adduct of Kemp's triacid (ciscis‐1,3,5‐tri­methyl­cyclo­hexane‐1,3,5‐tri­carboxylic acid, KTA) with 2‐amino­pyridine (2‐APY), namely 2‐amino­pyridinium 3,5‐di­carboxy‐1,3,5‐tri­methyl­cyclo­hexane­carboxyl­ate, 2‐APY+·KTA? or C5H7N2+·C12H17O6?, has revealed a centrosymmetric hydrogen‐bonded cyclic KTA homodimer repeating unit [O?O 2.524 (4) Å] linked into a polymer structure through the pyridinium and amino groups of the 2‐APY mol­ecule [O?N 2.736 (4), 2.989 (4) and 2.999 (4) Å].  相似文献   

19.
(?)‐(1S,2R)‐Norbornene‐2‐carboxylic acid alkyl esters (alkyl = Me, Bz, L ‐menthyl, or D ‐menthyl) were successfully prepared by the Diels–Alder reaction of cyclopentadiene with (R)‐(?)‐pantolactone‐O‐yl acrylate followed by epimerization and column chromatography. The enantiomeric excess was 99.9%. These monomers were polymerized by Pd(II)‐based catalysts, and high yields of the polymers were obtained. The methyl ester gave an optically active polymer of high optical rotation (monomer [α]D = ?24.7, polymer [α]D = ?98.5). This high rotation value of the polymer was attributed to the isotactic chain regulation of the polymer. This high rotation was not observed with methyl esters prepared by the transesterification of menthyl esters. The stereoregular polymer exhibited notable resonance peaks at 39 ppm in 13C NMR spectra. No crystallinity was observed. © 2005 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 1263–1270, 2006  相似文献   

20.
Photochemical studies of the reactivity of 6‐(hydroxymethyl)pterin (=2‐amino‐6‐(hydroxymethyl)pteridin‐4(1H)‐one; HPT) in alkaline aqueous solutions (pH 10.2–10.8) at 350 nm and room temperature were performed. The photochemical reactions were followed by UV/VIS spectrophotometry, thin‐layer chromatography (TLC), high‐performance liquid chromatography (HPLC), and an enzymatic method for H2O2 determination. In the presence of O2, 6‐formylpterin (=2‐amino‐3,4‐dihydro‐4‐oxopteridine‐6‐carboxaldehyde; FPT) was the only photoproduct detected. In the absence of O2, we observed a compound with an absorbance maximum at 480 nm, which was oxidized very rapidly by O2 in a dark reaction to yield FPT. The quantum yields of substrates disappearance and of photoproducts formation were determined. The formation of H2O2 during photooxidation was monitored, and the number of mol of H2O2 released per mol of HPT consumed corresponded to a 1 : 1 stoichiometry. HPT was also investigated for efficiency of singlet‐oxygen (1O2) production and quenching in aqueous solution. The quantum yield of 1O2 production (ΦΔ=0.21±0.01) was determined by measurements of the 1O2 luminescence in the near‐IR (1270 nm) upon continuous excitation of the sensitizer. The rate constant of 1O2 total quenching by HPT was determined (kt=3.1?106 M ?1 s?1), indicating that this compound was able to quench 1O2. However, 1O2 did not participate in the photooxidation of HPT to FPT.  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