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 89 毫秒
1.
2.
The azoles 2-(2′-hydroxyphenyl)oxazole (HPO) and 2-(2′-hydroxyphenyl)-4-methylthiazole (HPT) have been synthesised and studied in order to compare their photophysical characteristics. Their absorption and emission properties are reported in non-polar, alcoholic and aqueous media. Ground and excited state pK data were determined by spectroscopy and a model is proposed to explain the behaviour of HPT and HPO as a function of the pH. Excitation spectra and quantum chemical calculations suggest an equilibrium of ground state conformers. The calculations also predict a small energy barrier for rotation in the first excited singlet state for the proton transferred tautomers. The resulting twisted structure of the tautomer form possesses a biradicaloid nature, and is near-degenerate in energy with the first excited triplet state.  相似文献   

3.
The 2,2′-dimethyl-4,4′-bithiazole ligand (1), (dm4bt), and its Zn and Hg complexes have been prepared. A conformational property calculation at the DFT level for the ligand shows the anti conformation is energetically more stable by about 22.83 kJ/mol and the rotational barrier is about 32.01 kJ/mol for the anti → syn conversion, a phenomena happening during complex formation. The complexes [Zn(dm4bt)Cl2] (2) and [Hg(dm4bt)Cl2] (3) have spectral properties typical for d10 metal diimine systems. The structures of the ligand and the two complexes have been determined by the single crystal diffraction method. The X-ray structure determinations show that both complexes are four coordinated by two chloride atoms and one bidentate dm4bt. In the Hg complex one of the two chlorides is set at a semi-bridging position.  相似文献   

4.
2,2′-Bis[(4,7-dimethyl-inden-1-yl)methyl]-1,1′-binaphthyl and [2,2′-bis[(4,5,6,7-tetrahydroinden-1-yl)methyl]-1,1′-binaphthyl]titanium and -zirconium dichlorides have been synthesized from 2,2′-bis(bromomethyl)-1,1′-binaphthylene. 2,2′-Bis(bromomethyl)-1,1′-binaphthylene was alkylated with the lithium salt of 4,7-dimethylindene to yield 2,2′-bis[1-(4,7-dimethyl-indenylmethyl)]-1,1′-binaphthylene (S)-(−)-9. The lithium salt of 9 was metalated with either titanium trichloride followed by oxidation or zirconium tetrachloride to give titanocene dichloride (S)-(+)-10 and zirconocene dichloride 11. The known complexes ansa-[2,2′-bis[(1-indenyl)methyl]-1,1′-binaphthyl]titanium and -zirconium dichlorides were formed and hydrogenated to ansa-[2,2′-bis[(4,5,6,7-tetrahydroinden-1-yl)methyl]-1,1′-binaphthyl]titanium and -zirconium dichlorides 12 and 14 or to ansa-[2,2′-bis[(4,5,6,7-tetrahydroinden-1-yl)methyl]-5,5′,6,6′,7,7′,8,8′-octahydro-1,1′-binaphthyl]titanium dichloride 13 whose solid state structure was determined by X-ray crystallography. Complex 13 adopts a C1-symmetrical conformation in the solid state, but is conformationally mobile in solution, exhibiting C2-symmetry in its room temperature NMR spectra.  相似文献   

5.
Experiments are described which show that a substance isolated from the root bark of Casimiroa edulis Llave et Lex. is 2′,5,6-trimethoxyflavone (IIa). It is demonstrated that zapotin and zapotinin, isolated previously from the same plant, are very probably 2′,5,6,7-tetramethoxyflavone (IIIa) and 5-hydroxy-2′,5,6-trimethoxyflavone (IIIb), respectively. The occurrence in nature of flavonoids of type IIa, IIIa and IIIb bearing a single oxygen substituent in ring B at the 2′-position is unusual.  相似文献   

6.
The coordination of 1,5-bis-(1′-phenyl-3′-methyl-5′-pyrazolone-4′)-1,5-pentanedione (BPMPPD) and 2,2′-bipyridine (bipy) with lanthanide ions in water-alcohol solution has been studied. Binuclear complexes of the types : Ln2(BPMPPD)3(bipy)2·nH2O (n = 2 for Y, n = 4 for Eu, Gd, Dy, Ho, Er, Tm and Yb); Ln2(BPMPPD)3bipy·nH2O (n = 10 for La, n = 3 for Pr, Nd, Sm and Tb) were formed. The compounds were characterized by elemental analysis, molar conductance, IR, UV, 1H NMR spectroscopy, thermogravimetric analysis and fluorescence spectra.  相似文献   

7.
The complexes of 3,3′,5,5′-tetrabromo-2,2′-biphenol (TBBPh) with 7-methyl-1,5,7-triazabicyclo[4.4.0]dec-5-ene (MTBD) and triethylamine (TEA) were studied by FTIR spectroscopy. In chloroform and in acetonitrile a proton transfer from TBBPh to N-bases (MTBD, TEA) occurs. In chloroform solution the protonated N-base molecules are hydrogen-bonded to the deprotonated TBBPh molecules, whereas in acetonitrile the complexes dissociate. The intra- as well as intermolecular hydrogen bonds within the chains show large proton polarizability.  相似文献   

8.
7,8,2′,3′,4′,5′-Hexachloro-11-methoxy-6H-cyclopenta[a]pyrene-6-spiro-1′-cyclopenta-2′,4′-diene 5 has been obtained by the reaction of phenalenone and 1,2,3,4-tetrachlorocyclopentadiene. The gross structure of 5 has been determined by the X-ray analysis. The ground-state properties of 5were discussed from the spectroscopic data, dipole moment and bond lengths.  相似文献   

9.
Conformational properties of 2,2′‐bithiazole and 4,4′‐dimethyl‐2,2′‐ bithiazole have been studied by using AM1 and PM3 semiemperical methods and ab initio HF/6‐311+G* and B3LYP/6‐311+G* calculations. All methods agree that the planar s‐trans conformation is the global minimum and the perpendicular conformation is the transition state. Additional local minima were found using the Hartree–Fock (HF) and B3LYP levels for 2,2′‐bithiazole while for 4,4′‐dimethyl derivative the minima was located only at the MP2//B3LYP level. The barrier heights for rotation are 1.72, 7.69, and 7.88 kcal/mol at the PM3, HF, and B3LYP levels, respectively, and methyl substitution did not affect appreciably this value. Fourier expansion terms and bond orders were used to explain the origins of the rotational barrier in terms of π conjugation, electrostatic interaction, and steric effects, which represent the main factors in the shape of the rotational barrier. © 2000 John Wiley & Sons, Inc. Int J Quant Chem 79: 367–377, 2000  相似文献   

10.
3′,5′-di-tert-butyldimethylsilyl-2,2′-anhydrouridine 3 was hydrolyzed under basic conditions to yield selectively 3′-tert-butyldimethylsilylarabinouridine 5 in 85 % yield. 3′,5′-disilyl-arabinouridine derivatives 4 and 6 also led selectively to 5 under the same conditions. These reactions suggest an intramolecular participation of the 2′-hydroxyl group in the desilylation and can be used to prepare rapidly and in high yield 3′-silylaraU derivatives.  相似文献   

11.
The room temperature photophysical properties of several sulphonated and unsulphonated 6-(2′-hydroxy-4′-methoxyphenyl)-s-triazines were investigated in a range of solvents by means of steady state and picosecond fluorescence spectroscopy. Compounds possessing phenyl or p-tolyl groups in the s-triazinyl ring exhibit only a very weak normal Stokes-shifted fluorescence, arising from the initially excited chromophore. Substitution of phenoxy groups into the s-triazinyl ring results in the appearance of an additional longer-wavelength fluorescence which is assigned to the keto tautomer, formed following excited state intramolecular proton transfer (ESIPT). The rate constant for the (ESIPT) process that occurs in sodium 3-(3′,5′-diphenoxy-2′,4′,6′-triazinyl)-4-hydroxy-2-methoxybenzene sulphonate in water is estimated to be greater than 1011 s−1.  相似文献   

12.
3-exo,3′-exo-(1R,1′R)-bicamphor (12) is obtained from 3-exo,3′-exo-(1R,1′R)-bithtiocamphor (3) by condensation with hydrazine hydrate followed by hydrolysis of the resulting dihydropyridazine 11. Deprotonation of 12 with NaH and subsequent treatment with potassium hexacyanoferrate (III) furnishes the 2,2′-dioxo-3,3′-bibornanylidene 13, whilst reduction of 12 with L1AlH4 affords the 3,3′-biisoborneol 16. Further related transformations to various 2,2′-difunctional 3,3′-bibornane derivatives are described, which are could be of interest as chiral ligands  相似文献   

13.
R. Kikumoto  T. Kobayashi 《Tetrahedron》1966,22(10):3337-3343
Oxindole reacts with p-nitrobenzyl chloride ot give 3-(4′-nitrobenzyl) oxindole, but with o-nitrobenzyl chloride abnormal product, 2′-hydroxy-spiro[2H-indole-2,3′-3′H-indole] (Vb) is produced. The structure of Vb has been elucidated on the basis of the IR, UV and mass spectra, and confirmed by the analogous reactions of 3-methyl-, 4-methyl- and 3,3-dimethyloxindoles with o-nitrobenzyl chloride. Isatin reacts with o-nitrobenzyl chloride to give o-nitrobenzyloxireno[,3]-oxindole (X).  相似文献   

14.
1,1′-Methylene-3,3′-dialkyldiimidazolium salts have been deprotonated with n-butylithium in the presence of palladium(II) iodide to give the percarbene complexes 1 (alkyl=Me) and 2 (alkyl=Et), each containing two bidentate 1,1′-methylene-3,3′-dialkyldiimidazolin-2,2′-diylidene chelate ligands. The X-ray structure analysis of 1 reveals a stereochemistry in which the two spiro-linked six-membered metallacycles adopt boat-like conformations strongly bending out of the PdC4 coordination plane in opposite directions. The carbenoid imidazole rings, which are rotated by +42 and −43°, respectively, relative to this plane, break down into two tightly bound π-systems (N=4C=4N,= C=C) connected by long C---N bonds.  相似文献   

15.
The template condensation of 6,6″-bis(-methylhydrazino)-2,2′: 6′,2″-terpyridines L2 and L3 with 2,6-pyridinedialdehyde may give a number of different products depending upon the metal ion which is used. In the presence of nickel(II) the products are either the nickel(II) complexes of the 18-membered ring macrocycles L4 or L5 or the free macrocycles. The metal ion acts as a transient template and is removed in a chloride ion specific demetallation. The use of dimethyltin(IV) as a template results in the formation of complexes of the ring contracted macrocycles L6 or L7.  相似文献   

16.
Substituted thymidine derivatives are of interest because of their potential antiviral properties. We demonstrate a general strategy for synthesis of 3′-substituted thymidine derivatives, consisting of activation via N-3 alkylation of 2,3′-cyclothymidine followed by nucleophilic opening at the 3′-position. Examples include demonstration of carbon-carbon bond formation at the 3′-position.  相似文献   

17.
Enantiopure 1,1′-binaphthyl-2,2′-dicarboxylic acids (R)-1 and (S)-1 have been synthesized through the lipase-catalyzed kinetic resolution of the racemic 2,2-bis(hydroxymethyl)-1,1′-binaphthyl (±)-2 and subsequent oxidation of the hydroxymethyl groups.  相似文献   

18.
Novel Oligodeoxynucleotide analogues containing 3′-C-threo-methylene phosphodiester internucleoside linkages were synthesized on automated DNA-synthesizers using the phosphoramidite approach. The sugar modified phosphoramidite building block 5 was obtained by phosphitylation of 1-(2,3-dideoxy-5-O-(4,4′-dimethoxytrityl)-3-C-hydroxymethyl-β-D-threo-pentofuranosyl)thymine (4) which was synthesized in only three steps from 5′-O-(4,4′-dimethoxytrityl)thymidine (1). The hybridization properties and enzymatic stability of the oligonucleotide analogues were studied by UV experiments. 17-Mers having one or three modifications in the middle or two modifications in each end hybridized to DNA with moderate lowered affinity compared to unmodified 17-mers (ΔTm 1–3°C per modification). Furthermore, the end-modified and all-modified oligonucleotides were stable towards snake venom phosphodiesterase.  相似文献   

19.
The enantiomers of the title compound, the important photochromic material (RS)-1b, have been enriched semipreparatively by liquid chromatography. As a consequence, we were able to determine the barrier of the thermal interconversion (R)-1b(S)-1b via time-dependent polarimetry, amounting to ΔG=85.9 kJ/mol at 22.0°C in d6-DMSO (Table 2). The thermal equilibration of the corresponding merocyanine 2b was monitored in d6-DMSO by time-dependent 1H NMR, resulting in ΔG1=102.8 and ΔG2=92.0 kJ/mol at 22°C (Table 1). This means that, starting from (RS)-1b, the opened isomer 2b is attained by a slow reaction (ΔG1=102.8 kJ/mol, Fig. 4). Therefore, the merocyanine 2b cannot be identified with the intermediate required for the fast process of C(sp3)–O bond cleavage (ΔG=85.9 kJ/mol) upon the above enantiomerization of (RS)-1b. Apparently, these two thermal isomerizations (Fig. 4) are independent of each other. The structure of the unknown intermediate of the interconversion (R)-1b(S)-1b must therefore differ from the known one of merocyanine 2b.
Table 1. Equilibration between spiro compounds (RS)-1 and merocyanines 2 at 22°C, measured by time-dependent UV absorptions[3] for (RS)-1a2a and by time-dependent 1H NMR intensities for the other compounds

Article Outline

1. Introduction
2. Equilibration of the merocyanine 2b with the spiro compound (RS)-1b
3. Preparative separation and characterization of the enantiomers of the spiro compound (RS)-1b
4. Enantiomerization of the spiro compounds (R)- and (S)-1b
5. Discussion of the two different isomerizations investigated
6. Experimental
6.1. General methods
6.2. (±)-6-Nitro-1′,3′,3′-trimethylspiro[2H-1-benzopyran-2,2′-indoline] 1b[43]
6.3. (+)436-6-Nitro-1′,3′,3′-trimethylspiro[2H-1-benzopyran-2,2′-indoline] 1b
6.4. (−)436-6-Nitro-1′,3′,3′-trimethylspiro[2H-1-benzopyran-2,2′-indoline] 1b
6.5. 4-Nitro-2-[(E)-2′-(1′′,3′′,3′′-trimethyl-3H′′′-2′′-indoliumyl)-1′-ethenyl]-1-phenolate 2b[19]
Acknowledgements
References

1. Introduction

Many derivatives of 1′,3′,3′-trimethylspiro[2H-1-benzopyran-2,2′-indoline] 1a (Scheme 1) are of interest because of their photochromism.[2] The parent molecule 1a can be transformed photochemically into the merocyanine 2a which isomerizes thermally with a very high rate back to 1a.[3] Therefore, unsubstituted 1a has no practical value with respect to photochromism. This situation changes upon the introduction of a nitro group into the 6-position: the title compound 1b has probably been cited in the literature most often among all photochromic materials. The corresponding merocyanine 2b is obtained by irradiation and reverts to the equilibrium mixture (Scheme 1) consisting predominantly of the spiro compound 1b. The rate of isomerization of 2b is much lower than that of the 2a1a reversal.[3, 4, 5, 6, 7 and 8] Although analogs have now been found which are more stable to light than 1b, the latter has been significant for the development of practical applications of photochromism and continues to be significant for basic research,[2, 9 and 10] e.g. with respect to 1b chemically bonded to another molecule. A further nitro group in the 8-position again changes the properties: only a very small amount of the spiro compound 1c appears in the thermal equilibrium[11 and 12] ( Scheme 1) in dipolar aprotic solvents, which means that the observed photochromism is a reversible one with limited applicability.  相似文献   

20.
The molecular structure of isolated canonical 2′-deoxyrinobucleosides was calculated using the density functional theory. It was demonstrated that the geometry of the base unit (BU) is almost unchanged compared to free nucleobases. Only slight out-of-plane deformation of the pyrimidine ring in deoxy-cytidine is observed. The conformation of the furanose ring strongly depends on the nature and orientation of the nucleobase. All nucleosides possess different conformations of this ring. Significant influence of the steric repulsion between the nucleobase and the sugar unit (SU) on puckering of the furanose ring and variation of the C–O and glycosyl bond lengths was demonstrated. The C(3′)-endo conformer of the furanose ring is more stable at the anti-orientation BU with respect to SU. An opposite trend is observed for the syn-orientation which is additionally stabilized by an intramolecular hydrogen bond with participation of the C(5′)OH group.  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