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.
Summary. Replacing the typical lactam β-alkyl substituents of xanthobilirubinic acid and kryptopyrromethenone, two bilirubin analogs long used as model compounds in studies of its photochemistry and metabolism, leads to increased amphiphilicity. Synthesized by base-catalyzed condensation of 3,4-dimethoxypyrrolin-2-one with the appropriate pyrrole α-aldehyde, the 2,3-dimethoxyl analogs of xanthobilirubinic acid and kryptopyrromethenone are yellow-colored dipyrrinones that form intermolecular hydrogen-bonded dimers in the solid, as determined by X-ray crystallography, and in CHCl3, as revealed by 1H NMR and vapor pressure osmometry. These two new dipyrrinones are approximately ten times more soluble in water than their parent dipyrrinones.  相似文献   

2.
Summary. A new linear tetrapyrrole, with two dipyrrinones connected by a string of 5 carbons, was synthesized from two equivalents of a 9H-dipyrrinone and one of glutaryl dichloride. Unlike typical dipyrrinones, which are intermolecularly hydrogen bonded in the crystal and in CHCl3 solution, 1,3-bis-[2,3,7,8-tetraethyl-(10H)-dipyrrin-9-carbonyl]propane is a monomer in CHCl3, as determined by vapor phase osmometry (VPO) measurements. Its crystal structure determination revealed a folded conformation with a novel type of dipyrrinone to dipyrrinone intramolecular hydrogen bonding. Unexpectedly, the same conformation apparently persists in CHCl3 solution, as shown by 1H NMR spectroscopy.  相似文献   

3.
Sanjeev K. Dey 《Tetrahedron》2009,65(12):2399-2407
Replacing the typical β-alkyl substituents of [6]-semirubin and [6]-oxosemirubin, two intramolecularly hydrogen-bonded bilirubin analogs, with methoxy groups produces amphiphilic dipyrrinones. Synthesized from the respective 9H-dipyrrinones prepared by base-catalyzed condensation of 3,4-dimethoxypyrrolin-2-one with the appropriate pyrrole α-aldehyde, the 2,3-dimethoxy and 2,3,7,8-tetramethoxy analogs of [6]-semirubin are yellow-colored dipyrrinones that form intramolecularly hydrogen-bonded monomers in CDCl3, as deduced from 1H NMR NH chemical shifts. They are monomeric in CHCl3, as determined by vapor pressure osmometry. In contrast, in the solid, X-ray crystallography reveals supramolecular ribbons of intermolecularly hydrogen-bonded (dipyrrinone to dipyrrinone and acid to acid) 2,3,7,8-tetramethoxy-[6]-semirubin. The latter is approximately 20 times more soluble in water than the parent [6]-semirubin with four β-methyl groups.  相似文献   

4.
Summary.  A crystal structure determination of the 9-acyl-dipyrrinone 9-butanoyl-2,3,7,8-tetramethyl-(10H)-dipyrrin-1-one indicates the presence of intermolecularly hydrogen-bonded dimers; however, in CHCl3 solution the pigment is monomeric as determined by vapor pressure osmometry measurements. Lacking an alkyl group at C(8), the 9-acyl-dipyrrinone exhibits only a weak tendency to form dimers in CHCl3 (K A ∼ 60 M −1) as determined by analysis of variable temperature 1H NMR data. In contrast, when the 9-acyl group is replaced by formyl or when the acyl group is fixed in a syn orientation to the pyrrole NH, the dipyrrinone is strongly prone to dimerization in CHCl3. Received August 22, 2000. Accepted September 5, 2000  相似文献   

5.
An indolocarbazole dimer, containing chiral urea appendages, that adopts a helically folded conformation by intramolecular hydrogen bonds as proven by 1H NMR and circular dichroism (CD) spectroscopy has been prepared. Owing to the preferential formation of one helical conformer, strong CD signals appear in relatively non-polar solvents such as chloroform (CHCl3) and dichloromethane (CH2Cl2) but the signal is negligible in dimethyl sulfoxide (DMSO). In addition, the optical rotation of the dimer is highly sensitive to the polarity of solvents. For example, the magnitude of the specific rotation ([α]D) is ? 934° in CH2Cl2 and ? 657° in CHCl3 but it is only ? 75° in DMSO. These observations suggest that the dimer folds to a helical structure by intramolecular hydrogen bonds in relatively non-polar solvents but exists in an unfolded extended conformation in polar solvents such as DMSO. The dimer strongly binds anions such as chloride, acetate and sulfate by multiple hydrogen bonds. In addition, anion binding leads to considerable CD spectral changes with the different pattern and degree of Cotton effects depending on the kind of anions. The dimer may be therefore utilised for the construction of an anion-responsive chiroptical sensor or switch.  相似文献   

6.
Using their amide (and pyrrole) groups, dipyrrinones act as hydrogen bonding receptors for carboxylic acids, as found in a large number of 10-oxo-semirubins (1-6). The latter can be synthesized readily by Friedel-Crafts coupling of 9-H dipyrrinones with half-ester acid chlorides or diacid dichlorides of α,ω-dicarboxylic acids, ranging from C2 to C10. With ω-oxo-alkanoic acid chains of C5 or ≥C5, intramolecular hydrogen bonding is observed. With acid chains <C5 hydrogen bonding is not observed. Uncharacteristically (for dipyrrinones), 10-oxo-dipyrrinone acids (1-6) and their corresponding esters (1e-6e) remain monomeric in hydrogen bond promoting solvents.  相似文献   

7.
A high-yield straightforward conversion of lactams to lactim ethers is shown by the conversion of (10H)-dipyrrin-1-ones to (11H)-dipyrrin-1-ol methyl and ethyl ethers in 90% yield from heating in neat trimethyl or triethyl phosphite at 160°C. Unlike the parent dipyrrinones, which form intermolecularly hydrogen-bonded dimers in CHCl3, their lactim ethers are shown to be monomeric by vapor pressure osmometry. The latter react with boron trifluoride etherate to N,N′-bridged BF2 derivatives that exhibit strong fluorescence (φF 0.6–0.8) near 535 nm. X-Ray crystal structures were obtained of the lactim ethyl ether of kryptopyrromethenone and the BF2 derivative of the lactim ethyl ether 2,3-diethyl-7,8-dimethyl-(10H)-dipyrrin-1-one.  相似文献   

8.
Summary. Bilirubin congeners with dipyrrinones conjoined to a diaceteylene unit (–CC–CC–) rather than to –CH2– were synthesized and examined spectroscopically. This new class of rubrified linear tetrapyrroles cannot easily fold or bend in the middle, but the dipyrrinones can rotate independently about the diacetylene unit. Thus, unlike bilirubin, which is bent in the middle and has a ridge-tile shape, the diacetylene unit orients the attached dipyrrinones along a linear path, and intramolecular hydrogen bonding between the dipyrrinones and opposing carboxylic acids preserves a twisted linear molecular shape when the usual propionic acids are replaced by hexanoic. In a bis-hexanoic acid rubin, the extended planes of the dipyrrinones intersect along the –(CC)2– axis at an angle of 102° for the conformation stabilized by intramolecular hydrogen bonding. With propionic acid chains, however, neither CO2H can engage an opposing dipyrrinone in intramolecular hydrogen bonding, and the energy-minimum conformation of this linear pigment, shows nearly co-planar dipyrrinones, with an intersection of an angle of 180° of the extended planes of the dipyrrinones. Spectroscopic evidence for such linearized and twisted (bis-hexanoic) and planar (bis-propionic) structures comes from the pigments NMR spectral data and their exciton UV-Vis and induced circular dichroism spectra.  相似文献   

9.
A high-yield straightforward conversion of lactams to lactim ethers is shown by the conversion of (10H)-dipyrrin-1-ones to (11H)-dipyrrin-1-ol methyl and ethyl ethers in 90% yield from heating in neat trimethyl or triethyl phosphite at 160°C. Unlike the parent dipyrrinones, which form intermolecularly hydrogen-bonded dimers in CHCl3, their lactim ethers are shown to be monomeric by vapor pressure osmometry. The latter react with boron trifluoride etherate to N,N′-bridged BF2 derivatives that exhibit strong fluorescence (φF 0.6–0.8) near 535 nm. X-Ray crystal structures were obtained of the lactim ethyl ether of kryptopyrromethenone and the BF2 derivative of the lactim ethyl ether 2,3-diethyl-7,8-dimethyl-(10H)-dipyrrin-1-one. Correspondence: David A. Lightner, Department of Chemistry, University of Nevada, Reno, Nevada 89557-0020 USA.  相似文献   

10.
Roth SD  Shkindel T  Lightner DA 《Tetrahedron》2007,63(45):11030-11039
A rare and unusual class of tripyrrolic compounds, violet-colored tripyrrin-1,14-diones, can be prepared easily and in moderately high yields from base (piperidine)-catalyzed condensation of 3-pyrrolin-2-ones with 2,5-diformylpyrroles. Dipyrrinones adopt the all-syn-Z conformation leading to helical, lock-washer like structures, which form dimers that are held together by intermolecular hydrogen bonds in nonpolar solvents and in the crystal. Strong bathochromic spectral shifts of the tripyrrindione ∼480 nm long wavelength UV-visible absorption band are seen with added base: DBU, 615 nm; TFA, 573 nm; and Zn(OAc)2, 586 nm.  相似文献   

11.
Bromophilic attack by the transition metal carbonyl anion, [Re(CO)5]Na (pKa = 21.1), on 2-(1-bromoalkylidene)thiazolidin-4-ones is significantly faster than abstraction of an acidic lactam hydrogen (pKa ∼17-18), when the generated carbanion is stabilized by an α-CN or α-PhCO group. The bromophilic reaction of 2-(1-bromoalkylidene)thiazolidin-4-one, having an α-CN electron-withdrawing group, resulted in formation of a new metallacyclic anionic complex. With less reactive vinyl bromides, containing an α-CONHPh or α-CO2Et group, only deprotonation is observed. The role of the metal carbonyl anion is highlighted by a comparison with the 9-methylfluorenide carbanion (pKa of 9-methylfluorene is 22.3), which reacts exclusively via a deprotonation pathway.  相似文献   

12.
In this study the synthesis of novel chiral calix[4]azacrown derivatives has been reported. The enantioselectivity of chiral receptors was investigated by using UV-vis spectroscopy. All the chiral calix[4]arene derivatives exhibited certain chiral recognition toward the enantiomers of phenylalanine (Phe-OMe·HCl) and alanine methyl ester hydrochlorides (Ala-OMe·HCl). As a chiral receptor, the furfuryl-armed calix[4]azacrown ether 7 has the best enantiomeric discriminating ability for α-amino acid ester hydrochlorides (up to KL/KD=2.08, ΔΔG0=−1.82 kJ mol−1) in CHCl3. The enantiomeric recognition abilities for guests are also discussed from a thermodynamic point of view.  相似文献   

13.
A new method based on negligible depletion hollow fiber-protected liquid-phase microextraction coupled with high-performance liquid chromatography (HPLC) was developed for the simultaneous determination of partitioning coefficients (KOW) and acid dissociation constants (pKa), by using phenol, 4-chlorophenol and 2,4-dichlorophenol as model compounds. A 37-mm length polypropylene hollow fiber membranes (600 μm inner diameter, 200 μm wall-thickness, 0.2 μm pore size, ∼70% porosity) with two-end sealed were filled with 1-octanol by ultrasonic agitation to prepare the extraction device. The extraction device was deployed in sample solutions, prepared by spiking target analytes in 1-octanol saturated aqueous solutions (500 mL), for negligible depletion extraction. After equilibrium was reached (∼5 h), the 1-octanol in the lumen of the hollow fiber membrane was collected for HPLC determination of the target analytes. As the depletion of the analytes in aqueous samples was negligible, the distribution coefficient (DOW) could be calculated based on the measured equilibrium concentration in 1-octanol (CO) and the initial concentration (CW) in the aqueous sample of the target analyte (DOW = CO/CW). The DOW values measured at various pH values were nonlinearly regressed with pH to obtain the KOW and pKa values of a compound. Results showed that the measured values of the KOW and pKa of these model compounds agreed well with literature data.  相似文献   

14.
The first representatives of chiral boron atom-containing amine-cyanomethoxycarbonyl boranes (A · BH(CN)COOMe) have been synthesized either from the corresponding amine-bromocyanomethoxycarbonylborane complexes with [Bu4N]CN or from Me3N · BH(CN)COOMe and an amine in a base-exchange reaction. Acid hydrolyses of methyl esters generated the free acids (A · BH(CN)COOH), which are isoelectronic to the α-cyano carboxylic acids. Their pKa values and hydrolysis half-lives in acidic medium (that is rate of proton reduction) have been determined. Similarly to the alpha cyano carboxylic acids, the cyano group attached to the boron (in alpha position to the COOH group) increased the acid strength of carboxy boranes with 2.0-2.5 orders of magnitude. Independently from the type of the amine, pKa values of the amine-cyanocarboxyboranes (6.34-5.82) decrease consistently with the increase of pKb values of the amines. Hydrolytic decomposition rate of the alkylamine complexes increases with increasing pKb values of the amines while the opposite was found for pyridine base complexes. When considering both types of the amines, hydrolysis half-lives of the complexes range over several orders of magnitude from 0.005 to 400 h. Based on these observations protonation of the amine nitrogen atom appears to be the rate determining step in the hydrolysis process. With loss of methanol, 2-NH2-py · BH(CN)COOMe transformed into a five membered lactam derivative. X-ray diffraction revealed that the pyridine ring is coplanar with the five membered lactam ring. In the crystal two molecules are connected in a head to tail arrangement by strong intermolecular H-bonds between N(2)-H and the carbonyl oxygen (O1) with a donor and acceptor distance of 2.867(3) Å. Three new cyanomethoxycarbonylborates having the composition of K[BHn(CN)3−nCOOMe] (n = 1, 2) and K[B(OH)(CN)2COOMe] have also been synthesized and their properties examined.  相似文献   

15.
For the continuing study of the molecular recognition of the π-electron-poor host comprised of two pyromellitic diimides and two dialkoxynaphthalenes, its inclusion with π-electron-rich polymethoxybenzenes has been examined. The UV-vis titration studies indicated following order of the association constants (Ka's) as 1:1 complexes in CHCl3: 1,3,5-trimethoxybenzene (31.3 M−1)>1,3-dimethoxybenzene (9.2 M−1)>1,2-dimethoxybenzene (6.5 M−1)>1,4-dimethoxybenzenes (2.8 M−1). The X-ray structural analysis of the complexes between the host and dimethoxybenzenes proved the intracavity 1:1 complexes and provided useful information on the structures of the complexes. Not only charge transfer interactions but also other weak interactions such as electrostatic, van der Waals, and the unique hydrogen bonds between the α-hydrogen atoms of the naphthalene and the methoxy oxygen atoms were considered to be responsible for the magnitude of the Ka's. Thus molecular recognition of polymethoxybenzenes has been accomplished by the neutral host using multipoint weak interactions in an organic solvent.  相似文献   

16.
A novel type of ionophore ligands, 3′-(2,3-dihydroxypropylthio)-phthalonitrile and 4′(2,3-dihydroxypropylthio)-phthalonitrile, and their α- and β-tetrasubstituted metallo phthalocyanines, (MPc), (M = ZnII, CoII, MnIIICl, FeIIIAc, CuII) have been prepared and fully characterized by elemental analysis, FT-IR, 1H and 13C NMR, and MS (ESI and Maldi-TOF). The complexes are soluble in both polar and non-polar solvents, such as MeOH and EtOH, THF, CHCl3 and CH2Cl2. The spectroscopic properties of the complexes are affected strongly by the electron-donating sulfanyl units on the periphery of the phthalocyanines. The cation binding properties of the complexes, for example using AgI and PdII, were evaluated by UV-Vis spectroscopy and the results show the formation of polynuclear phthalocyanine complexes. Functional donors on the periphery of the zinc and copper complexes coordinate to AgI and PdII to give ca. a 2:1 metal-phthalocyanine complex binding ratio for the concentration of 2.5 × 10−5 M (Pc) and 1.0 × 10−3 M (Metal ions). Voltammetric and in-situ spectroelectrochemical studies were performed to characterize the redox behavior of the complexes. An in-situ electrocolorimetric method was applied to investigate the colors of the electro-generated anionic and cationic forms of the complexes.  相似文献   

17.
The kinetics of oxidative addition of CH3I to [Rh(FcCOCHCOCF3)(CO)(PPh3)], where Fc = ferrocenyl and (FcCOCHCOCF3) = fctfa = ferrocenoylacetonato, have been studied utilizing UV/Vis, IR, 1H and 31P NMR techniques. Three definite sets of reactions involving isomers of at least two distinctly different classes of RhIII-alkyl and two different classes of RhIII-acyl species were observed. Rate constants for this reaction in CHCl3 at 25 °C, applicable to the reaction sequence below, were determined as k1 = 0.00611(1) dm3 mol−1 s−1, k−1 = 0.0005(1) s−1, k3 = 0.00017(2) s−1 and k4 = 0.0000044(1) s−1 while k−3 ? k3 and k−4 ? k4 but both ≠0. The indeterminable equilibrium K2 was fast enough to be maintained during RhI depletion in the first set of reactions and during the RhIIIalkyl2 formation in the second set of reactions. From a 1H and 31P NMR study in CDCl3, Kc1 was found to be 0.68, Kc2 = 2.57, Kc3 = 1.00, Kc4 = 4.56 and Kc5 = 1.65.  相似文献   

18.
An efficient polycondensation reaction between α,ω-diformyl functional aromatics, namely 2,5-diheptyloxy-1,4-diformylbenzene and side-chain substituted α,ω-diformyl oligo(p-phenylene vinylene) (OPV) with hydrazine afforded novel soluble and processable conjugated polymers, P1 and P2, respectively. The reaction conditions were investigated and structural analysis of the polymers was carried out by means of 1H, 13C, 15N NMR, indicating all-trans configured CN-NC linkages. The molecular weights Mn were observed at ∼1100-1400 g/mol for P1 and ∼8700-10,500 g/mol for P2. The optical properties showed absorption maxima at ∼455 nm and ∼487 nm for P1 and P2 (CHCl3 solutions), respectively, red shifted by 31-60 nm relative to the monomer aromatics due to conjugation through the azine linkage. The emission maxima are observed at ∼515 nm and ∼560 nm for P1 and P2 (CHCl3 solutions), respectively. Thin films of P2 readily undergo n-doping.  相似文献   

19.
Reaction of AlEt3 with 2-methyl-8-quinolinol gave ethylbis(2-methyl-8-quinolinolato)aluminum complex [Al(Et)(q′)2] 1. The complex 1 provided photoluminescent Al complexes by reactions with phenols, carboxylic acid, and H2O. The α-CH2 hydrogens in the Et group of 1 was diastereotropic as revealed by 1H NMR spectroscopy because of the presence of a chiral center at Al. The chirality at Al was dynamically lost at elevated temperature in CDCl2CDCl2 and DMSO-d6, as indicated by temperature dependent 1H NMR spectroscopy. This dynamic or fluxional behavior of 1 is explained by rotation of the 2-methyl-8-quinolinolato ligand. The kinetic parameters of the dynamic process were estimated at ΔH = 135 kJ mol−1 and ΔS = 159 J K−1 mol−1 in CDCl2CDCl2 and at ΔH = 124 kJ mol−1 and ΔS = 151 J K−1 mol−1 in DMSO-d6, respectively, at 350 K. Structures of some of the obtained Al complexes were confirmed by single-crystal X-ray crystallography. These Al complexes showed photoluminescence peaks at 492-507 nm in CHCl3 with quantum yields of 7-23%.  相似文献   

20.
In solid triphenylmethanol, the molecules are arranged in hydrogen-bonded tetramers, and it is already well established that the hydrogen bonding in this material undergoes a dynamic switching process between different hydrogen bonding arrangements. In addition to this motion, we show here, from solid-state 2H NMR studies of the deuterated material (C6D5)3COH, that each phenyl ring in this material undergoes a 180°-jump reorientation about the C6D5-C(OH) bond, with an activation energy of ca. 50 kJ mol−1. The timescale for the phenyl ring dynamics is several orders of magnitude longer than the timescale for the hydrogen bond dynamics in this material, and is uncorrelated with the dynamics of the hydrogen bonding arrangement.  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