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.
Temperature‐dependent selectivity in nucleophilic additions is affected by the solvent. In this context, we investigated the effect on diastereoselectivity of solvent mixtures with respect to pure solvents. Binary systems of THF/hexane and of four different hydrocarbon mixtures were employed in BuLi addition to 2‐phenylpropanal, (2S)‐2‐[(tert‐butyl)dimethylsilyloxy]‐2‐phenylethanal, and (2S)‐2‐[(tert‐butyl)dimethylsilyloxy]‐N‐(trimethylsilyl)propan‐1‐imine. A 5‐mol‐% of THF in hexane affects the isomer ratio by reducing both ΔΔH and ΔΔS contributions, and suppresses Tinv. On the contrary, in hydrocarbon binary mixtures, the Tinv is still observed and occurs at a higher temperature than in pure solvents. Studying the dependence of Tinv on the hexane/decane mixture composition, we propose the formation of a peculiar solvation cluster that is unaffected by the composition of the bulk reaction solvent.  相似文献   

2.
A series of thermoresponsive polypeptides bearing 1‐butyl, 1‐hexyl, or 1‐dodecyl side‐chains (i.e., 6a ‐ 6c ) were synthesized by copper‐mediated 1,3‐dipolar cycloaddition with high grafting efficiency (>95%) between side‐chain “clickable” polypeptide, namely poly(γ‐4‐(propargoxycarbonyl)benzyl‐L‐glutamate) ( 5 ) and 1‐azidoalkanes. 5 with different degree of polymerization (DP = 48–86) were prepared from triethylamine initiated ring‐opening polymerization of γ‐4‐(propargoxycarbonyl)benzyl‐L‐glutamic acid based N‐carboxyanhydride ( 4 ). 1H NMR, FTIR, and GPC results revealed the successful preparation of the resulting polypeptides. 6a ‐ 6c showed reversible UCST‐type phase behaviors in methanol, ethanol, and ethanol/water solvent mixtures depending on the polymer main‐chain length, alkyl side‐chain length, weight percentage of ethanol (fw) in the binary solvent, and so forth. FTIR analysis revealed the presence of the van der Waals interaction between the alkyl pendants of polypeptides and alkyl groups of alcoholic solvents. Variable‐temperature UV‐vis spectroscopy revealed that the UCST‐type phase transition temperature (Tpt) increased as polymer main‐chain length or concentration increased. In ethanol/water solvent mixtures, polypeptide with short alkyl pendant (i.e., 1‐butyl group) and short main‐chain length (DP = 41) showed the widest fw range and Tpts in the range of 61.0–71.1 °C. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 3425–3435  相似文献   

3.
We have investigated the effect of a series of 18 solvents and mixtures of solvents on the production of singlet molecular oxygen (O2(1Δg), denoted as 1O2) by 9H‐fluoren‐9‐one (FLU). The normalized empirical parameter E derived from ET(30) has been chosen as a measure of solvent polarity using Reichardt's betaine dyes. Quantum yields of 1O2 production (ΦΔ) decrease with increasing solvent polarity and protic character as a consequence of the decrease of the quantum yield of intersystem crossing (ΦISC). Values of ΦΔ of unity have been found in alkanes. In nonprotic solvents of increasing polarity, ΦISC and, therefore, ΦΔ decrease due to solvent‐induced changes in the energy levels of singlet and triplet excited states of FLU. This compound is a poor 1O2 sensitizer in protic solvents, because hydrogen bonding considerably increases the rate of internal conversion from the singlet excited state, thus diminishing ΦΔ to values much lower than those in nonprotic solvents of similar polarity. In mixtures of cyclohexane and alcohols, preferential solvation of FLU by the protic solvent leads to a fast decrease of ΦΔ upon addition of increasing amounts of the latter.  相似文献   

4.
5.
Di(tert‐butyl)diazomethane ( 4 ) is a nucleophilic 1,3‐dipole with strong steric hindrance at one terminus. In its reaction with 2,3‐bis(trifluoromethyl)fumaronitrile ((E)‐ BTE ), a highly electrophilic tetra‐acceptor‐substituted ethene, an imino‐substituted cyclopentene 9 is formed as a 1 : 2 product. The open‐chain zwitterion 10 , assumed as intermediate, adds the second molecule of (E)‐ BTE . The 19F‐ and 13C‐NMR spectra allow the structural assignment of two diastereoisomers, 9A and 9B . The zwitterion 10 can also be intercepted by dimethyl 2,3‐dicyanofumarate ( 11 ) and furnishes diastereoisomeric cyclopentenes 12A and 12B ; an X‐ray‐analysis of 12B confirms the ‘mixed’ 1 : 1 : 1 product. Competing is an (E)‐ BTE ‐catalyzed decomposition of 4 to give 2,3,4,4‐tetramethylpent‐1‐ene ( 7 )+N2; the reaction of (E)‐ BTE with a trace of water appears to be responsible for the chain initiation. The H2SO4‐catalyzed decomposition of diazoalkane 4 , indeed, produced the alkene 7 in high yield. The attack on the hindered diazoalkane 4 by 11 is slower than that by (E)‐ BTE ; the zwitterionic intermediate 21 undergoes cyclization and furnishes the tetrasubstituted furan 22 . In fumaronitrile, electrophilicity and steric demand are diminished, and a 1,3‐cycloaddition produces the 4,5‐dihydro‐1H‐pyrazole derivative 25 . The reaction of 4 with dimethyl acetylenedicarboxylate leads to pyrazole 29 +isobutene.  相似文献   

6.
Zusammenfassung Zur empirischen Beschreibung der Polarität binärer Lösungsmittelgemische wurden mit einem solvatochromen Betain-Farbstoff ET-Werte für folgende Systeme bestimmt: Methanol/Butanol-2, Methanol/2-Methylbutanol-2, Methanol/Pyridin, Methanol/Acetonitril, Methanol/Toluol und Ethanol/ 2,2,2-Trifluorethanol. In allen Systemen treten mehr oder weniger starke Abweichungen vom idealen Mischungsverhalten auf, was im einzelnen diskutiert wird. In den genannten Lösungsmittelgemischen wurde parallel die Kinetik der Ligandensubstitution an einem neutralen Kupfer(II)-bis-chelatkomplex vermessen. Die ET-Werte korrelieren mit den lnkWerten (k = Geschwindigkeitskonstante für den durch das Solvens induzierten Ligandenaustausch). Dies weist darauf hin, daß die Bildung des Übergangszustandes durch H-Brücken erleichtert wird.
Polarity of binary solvent mixtures: Determination of ET(30)-values and correlation with kinetic solvent effects
Summary The polarity of binary solvent mixtures is empirically described by ET-values as determined with a solvatochromic betaine dye. The following mixtures were studied: methanol/butanol-2, methanol/2-methylbutanol-2, methanol/pyridine, methanol/acetonitrile, methanol/toluene and ethanol/2,2,2-trifluoroethanol. All mixtures behave as more or less non-ideal systems which is discussed in detail. Parallel to the determination of ET-values the kinetics of ligand substitution in a neutral copper(II)-bis-chelate complex were studied in the same set of solvent mixtures. The ET-values correlate with the corresponding lnk-values (k = rate constant for solvent induced ligand substitution). This would mean that the formation of the transition state is facilitated by hydrogen bonding.
Für finanzielle Unterstützung bedanken wir uns bei der Deutschen Forschungsgemeinschaft und dem Verband der Chemischen Industrie e.V. Der verwendete Salicylaldehyd wurde uns von der Firma Bayer AG, Leverkusen, zur Verfügung gestellt.  相似文献   

7.
A sensitive and selective method for simultaneous quantitation of 15 benzodiazepines in human whole blood using rapid GC with negative‐ion chemical ionization MS is proposed. A mixed‐mode cation‐exchange polymeric sorbent was used for SPE. Different extraction solvents or mixtures of solvents of different compositions for elution of the adsorbed analytes, and washing steps for eliminating interferences in the column were tested. Analytes were eluted from the column using 5% v/v NH4OH in methanol. A derivatization step using different silylation reagents, time, and temperature was tested. Extracts from SPE were silylated by a mixture of N‐(tert‐butyldimethylsilyl)‐N‐methyltrifluoroacetamide, acetonitrile, ethyl acetate, and subjected to gas chromatographic analysis. The LODs of 15 benzodiazepines in whole blood samples ranged from 0.24–0.62 ng mL?1. The RSDs of samples used for three different quality control concentration levels were lower than 7.0%, and the accuracy ranged from 89.5 to 110.5%. The results show that the developed method is accurate, sensitive, selective, and very fast. Finally, the applicability of this method for determination of trace concentrations of several benzodiazepines in real blood samples has been demonstrated.  相似文献   

8.
Structural and thermodynamic properties of cellulose solutions in the ionic liquid 1‐ethyl‐3‐methylimidazolium acetate (EMIMAc) and its binary mixtures with N,N‐dimethyl formamide (DMF) are studied by small‐angle X‐ray scattering (SAXS). These measurements indicate molecular dissolution of the cellulose chains without any significant aggregation. The power–law relationships of the evaluated correlation length and osmotic modulus to concentration exhibit exponents of ?0.76 and 2.06 for EMIMAc and ?0.80 and 2.14 for DMF/EMIMAc solvent mixture, respectively. Thus, these solvents can be considered to be good solvents for cellulose. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part B: Polym. Phys. 2017 , 55 , 888–894  相似文献   

9.
The effect of solvent on the strength of noncovalent interactions and ionic mobility of the dibenzo‐18‐crown‐6 complex with K+ in water/organic solvents was investigated by using affinity capillary electrophoresis. The proportion of organic solvent (methanol, ethanol, propan‐2‐ol, and acetonitrile) in the mixtures ranged from 0 to 100 vol.%. The stability constant, KKL, and actual ionic mobility of the dibenzo‐18‐crown‐6‐K+ complex were determined by the nonlinear regression analysis of the dependence of the effective electrophoretic mobility of dibenzo‐18‐crown‐6 on the concentration of K+ (added as KCl) in the background electrolyte (25 mM lithium acetate, pH 5.5, in the above mixed hydro–organic solvents). Competitive interaction of the dibenzo‐18‐crown‐6 with Li+ was observed and quantified in mixtures containing more than 60 vol.% of the organic solvent. However, the stability constant of the dibenzo‐18‐crown‐6‐Li+ complex was in all cases lower than 0.5 % of KKL. The log KKL increased approximately linearly in the range 1.62–4.98 with the increasing molar fraction of organic solvent in the above mixed solvents and with similar slopes for all four organic solvents used in this study. The ionic mobilities of the dibenzo‐18‐crown‐6‐K+ complex were in the range (6.1–43.4) × 10?9 m2 V?1 s?1.  相似文献   

10.
This work aims at elucidating the mechanism of solvation of a radical ion pair (RIP) in a micro‐heterogeneous binary solvent mixture using magnetically affected reaction yield (MARY) spectroscopy. For the exciplex‐forming 9,10‐dimethylanthracene/N,N‐dimethylaniline system a comparative, composition‐dependent MARY line‐broadening study is undertaken in a heterogeneous (toluene/dimethylsulfoxide) and a quasi‐homogenous (propyl acetate/butyronitrile) solvent mixture. The half‐saturation field extrapolated to zero‐quencher concentration, B1/2, and the self‐exchange rate constants are analyzed in the light of solvent dynamical properties of the mixtures and a dielectric continuum solvation model. The dependence of B1/2 on the solvent composition is explained by cluster formation giving rise to shortened RIP lifetimes. The results are in qualitative agreement with the continuum solvation model suggesting that it could serve as a theoretical basis for quantitative modeling.  相似文献   

11.
The Pd(dba)2‐catalyzed reaction of Z‐1‐aryl‐1‐(tributylstannyl)‐2‐(trimethylsilyl)ethenes with allyl bromide in the presence of copper(I) iodide is reported for the first time. The reaction in the presence of 0.5 mol% Pd(dba)2 and 8 mol% CuI in dimethylformamide takes place at room temperature to give E‐2‐aryl‐1‐(trimethylsilyl)penta‐1,4‐dienes exclusively in isolated yields of 62–99%. A putative reaction mechanism is proposed. Copyright © 2005 John Wiley & Sons, Ltd.  相似文献   

12.
A novel supramolecular fluorescent donor–acceptor type dye molecule, (2E,4E)‐1‐(2‐hydroxyphenyl)‐5‐(pyren‐1‐yl)penta‐2,4‐dien‐1‐one (HPPD) self‐assembles in a mixture of ethanol/chloroform through intermolecular π–π stacking (distance ca. 3.384 Å) to form J‐aggregated single‐crystalline microribbons displaying Fabry–Pèrot (F‐P) type visible‐range optical resonance. The corresponding borondifluoride dye (HPPD‐BF), with a reduced HOMO–LUMO gap, self‐assembles into crystalline microrods acting as an F‐P type resonator in the near‐infrared (NIR) range.  相似文献   

13.
Solvatochromism of the twisted intramolecular charge-transfer (TICT) fluorescence of 4-(N,N-dimethylamino)benzonitrile (DMABN) in pure water, methanol, ethanol, 1-propanol, 2-propanol, acetone, acetonitrile, and in the corresponding aqueous–organic binary mixed solvents was systematically studied and an empirical solvent polarity scale (F B) based on the DMABN solvatochromism was defined. The F B parameters of the explored binary mixed solvents as a function of solvent composition were analyzed by a stepwise solvent-exchange (SSE) model to clarify the preferential solvation (PS) of the probe dye in these binary mixed solvents. Solvation diagrams toward DMABN in the mixed solvents, i.e., the local solvent composition in the solvation shell of DMABN molecules was depicted as a function of bulk solvent composition to visualize the PS in these mixed solvents. For comparison, a similar PS analysis was applied to the solvatochromism of 2,6-diphenyl-4-(2,4,6-triphenylpyridinium-1- yl)phenolate (ET-30) and pyrene (Py) in these mixed solvents; the responsive PS pattern of the mixed solvents toward the specific indicator dye of DMABN, ET-30, and Py was then discussed in terms of the chemical properties of the probe dye, the properties of the mixed solvents, and the solute–solvent and solvent–solvent interactions.  相似文献   

14.
The density functional theory has been used to study the tautomeric equilibrium of 2‐diazo‐4,6‐dinitrophenol(DDNP) in the gas phase and in 14 solvents at the B3LYP/6‐31G* level. The solvent effects on the tautomeric equilibria were investigated by the self‐consistent reaction field theory (SCRF) based on conductor polarized continuum model (CPCM) in apolar and polar solvents and by the hybrid continuum‐discrete model in protic solvent, respectively. Solvent effects on the computed molecular properties, such as molecular geometries, dipole moments, ELUMO, EHOMO, total energies for DDNP tautomers and transition state, tautomerization energies and solvation energies have been found to be evident. The tautomeric equilibrium of DDNP is solvent‐dependent to a certain extent. The tautomer I (cyclic azoxy form) is preferred in the gas phase, while in nonpolar solvents tautomer I and II (quinold form) exist in comparable amounts, and in highly polar solvents, the tautomeric equilibrium is shifted in favor of the more polar tautomer II . © 2010 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

15.
On the line of a previous work on the spectral properties of some of heteroaryl chalcone, the effect of medium acidity and photoreactivity of 3‐(4‐dimethylamino‐phenyl)‐1‐(2,5‐dimethyl‐thiophen‐3‐yl)‐propenone (DDTP) has been investigated in dimethylformamide and in chloromethane solvents such as methylenechloride, chloroform and carbon tetrachloride. The dye solution (ca. 5×10−4 mol·L−1 in DMF) gives a good laser emission in the range 470–560 nm with emission maximum at 515 nm upon pumping by nitrogen laser (λex=337.1 nm). The laser parameters such as gain coefficient (α), emission cross section (δe) and half life energy (E1/2) at maximum laser emission are also determined.  相似文献   

16.
Synthesis, Spectroscopic Characterization, and Molecular Structures of Selected Lewis‐Base Adducts of the Alkali Metal Tri(tert‐butyl)silylphosphanides The metalation of tri(tert‐butyl)silylphosphane with butyllithium and the bis(trimethylsilyl)amides of sodium, potassium, and rubidium yields quantitatively the corresponding alkali metal tri(tert‐butyl)silylphosphanides, which crystallize after addition of appropriate Lewis‐bases as dimeric (DME)LiP(H)SitBu3 ( 1 ), chain‐like (DME)NaP(H)SitBu3 ( 2 ), monomeric ([18]Krone‐6)KP(H)SitBu3 ( 3 ), and dimeric (TMEDA)1.5RbP(H)SitBu3 ( 4 ). The reaction of H2PSitBu3 with cesium bis(trimethylsilyl)amide at room temperature gives monocyclic and tetrameric cesium tri(tert‐butyl)silylphosphanide ( 5 ) with two additional coordinated CsN(SiMe3)2 molecules. At 80 °C this complex reacts with excess of phosphane to the tetrameric toluene adduct (η6‐Toluol)CsP(H)SitBu3 ( 6 ) which contains a central Cs4P4‐heterocubane fragment. The constitution of these compounds was verified by X‐ray structure determinations.  相似文献   

17.
The photophysical properties of 7‐(diethylamino) coumarin‐3‐carboxylic acid (7‐DCCA) were studied in cyclodextrins (α, β, γ,‐CDs), different neat solvents and solvent mixtures by using steady state absorption, emission and time‐resolved fluorescence spectroscopy. We have observed that with gradual increase in concentration of β‐CD the fluorescence quantum yield and lifetime decreased in a regular pattern whereas with gradual increase in concentration of γ‐CD the fluorescence quantum yield and lifetime gradually increased. With addition of urea, the fluorescence quantum yield and lifetime of 7‐DCCA in CDs increased. Binding constant calculation shows that 7‐DDCA forms 1:1 complex with β‐CD and with γ‐CD it forms 1:1 and 1:2 (guest:host) inclusion complex. We proposed that the dye molecule formed capping complex with β‐CD by means of hydrogen bonding and after addition of urea the hydrogen bonding network broke down and part of dye molecule entered inside the cavity of β‐CD. The photophysics of 7‐DCCA was studied in dioxane‐water mixture and ethylene glycol‐acetonitrile mixture to know the effect of polarity and viscosity of the media. The photophysics of 7‐DCCA was also studied in different neat solvents. It was found that the photophysics of 7‐DCCA depended on the structural feature of the solvents and solvent mixtures.  相似文献   

18.
An efficent access to a series of N‐(pyrrol‐2‐yl)amines, namely (E)‐1‐tert‐butyl‐5‐[(4‐chlorobenzylidene)amino]‐1H‐pyrrole‐3‐carbonitrile, C16H16ClN3, (7a), (E)‐1‐tert‐butyl‐5‐[(2,4‐dichlorobenzylidene)amino]‐1H‐pyrrole‐3‐carbonitrile, C16H15Cl2N3, (7b), (E)‐1‐tert‐butyl‐5‐[(pyridin‐4‐ylmethylene)amino]‐1H‐pyrrole‐3‐carbonitrile, C15H16N4, (7c), 1‐tert‐butyl‐5‐[(4‐chlorobenzyl)amino]‐1H‐pyrrole‐3‐carbonitrile, C16H18ClN3, (8a), and 1‐tert‐butyl‐5‐[(2,4‐dichlorobenzyl)amino]‐1H‐pyrrole‐3‐carbonitrile, C16H17Cl2N3, (8b), by a two‐step synthesis sequence (solvent‐free condensation and reduction) starting from 5‐amino‐1‐tert‐butyl‐1H‐pyrrole‐3‐carbonitrile is described. The syntheses proceed via isolated N‐(pyrrol‐2‐yl)imines, which are also key synthetic intermediates of other valuable compounds. The crystal structures of the reduced compounds showed a reduction in the symmetry compared with the corresponding precursors, viz. Pbcm to P from compound (7a) to (8a) and P21/c to P from compound (7b) to (8b), probably due to a severe change in the molecular conformations, resulting in the loss of planarity observed in the nonreduced compounds. In all of the crystals, the supramolecular assembly is controlled mainly by strong (N,C)—H…N hydrogen bonds. However, in the case of (7a)–(7c), C—H…Cl interactions are strong enough to help in the three‐dimensional architecture, as observed in Hirshfeld surface maps.  相似文献   

19.
Amino‐ and halofunctional Siloxititanes Amino‐di‐tert‐butylsilanol reacts with tetrabutoxititane in a molar ratio of 2:1 to give di‐n‐butoxi(bis(di‐tert‐butyl‐n‐butoxi)siloxi)titane, (C4H9OSi(CMe3)2‐O)2Ti(OC4H9)2 ( 1 ), and lithium‐di‐tert‐butylchlorosilanolate in a molar ratio of 3:1 to give n‐butoxi(tris(di‐tert‐butyl‐n‐butoxi)siloxi)titane, (H9C4OSi(CMe3)2‐O)3TiOC4H9 ( 2 ). The amino‐di‐tert‐butylsilanol substitutes the four chloroatoms of TiCl4 in the presence of triethylamine as HCl‐acceptor. The tetrakis(amino‐di‐tert‐butyl)siloxititane ( 3 ) is formed. The lithium salt of di‐tert‐butylfluorosilanol reacts with TiCl4 in a molar ratio of 2:1 to give 1, 1, 3, 3‐tetra‐tert‐butyl‐1‐fluoro‐3‐trichlorotitoxi‐1, 3‐disiloxane, FSi(CMe3)2‐O‐Si(CMe3)2‐O‐TiCl3 ( 4 ). In the reaction of di‐tert‐butyl‐chlorosilanol and TiCl4, the anion [chlorosiloxi‐octa(tri‐μ2‐chlorotitanate)] ( 5 ) with protonated diethylether as counterion is obtained by using diethylether as HCl‐acceptor. The crystal structure determinations of 3 and 5 are reported.  相似文献   

20.
pH‐zone‐refining centrifugal partition chromatography was successively applied in the large‐scale separation of close Rf antipsychotic indole alkaloids directly from CHCl3 fraction of Rauwolfia tetraphylla leaves. Two experiments with increasing mass from 500 mg to 3 g of crude alkaloid extracts ( 1 C) of R. tetraphylla were carried out in normal‐displacement mode using a two‐phase solvent system composed of methyl tert‐butyl ether/ACN/water (4:1:5, v/v/v) where HCl (12 mM) was added to the lower aqueous stationary phase as a retainer and triethylamine (5 mM) to the organic mobile phase as an eluter. The two centrifugal partition chromatography separations afforded a total of 162.6 mg of 10‐methoxytetrahydroalstonine ( 1 ) and 296.5 mg of isoreserpiline ( 2 ) in 97% and 95.5% purity, respectively, along with a 400.9 mg mixture of α‐yohimbine and reserpiline ( 3 and 4 ). Further, this mixture was resolved over medium pressure LC using TLC grade silica gel H (average particle size 10 μm), which afforded 160.4 mg of α‐yohimbine ( 3) and 150.2 mg of reserpiline ( 4) in >95% purities. The purity of the isolated antipsychotic alkaloids was analyzed by high‐performance LC and their structures were characterized on the basis of their 1D, 2D NMR and electrospray ionization‐mass spectroscopic data.  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