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.
We report the synthesis of novel resorcin[4]arene‐based cavitands featuring two extended bridges consisting of quinoxaline‐fused TTF (tetrathiafulvalene) moieties. In the neutral form, these cavitands were expected to adopt the vase form, whereas, upon oxidation, the open kite geometry should be preferred due to Coulombic repulsion between the two TTF radical cations (Scheme 2). The key step in the preparation of these novel molecular switches was the P(OEt)3‐mediated coupling between a macrocyclic bis(1,3‐dithiol‐2‐thione) and 2 equiv. of a suitable 1,3‐dithiol‐2‐one. Following the successful application of this strategy to the preparation of mono‐TTF‐cavitand 3 (Scheme 3), the synthesis of the bis‐TTF derivatives 2 (Scheme 4) and 19 (Scheme 5) was pursued; however, the target compounds could not be isolated due to their insolubility. Upon decorating both the octol bowl and the TTF cavity rims with long alkyl chains, the soluble bis‐TTF cavitand 23 was finally obtained, besides a minor amount of the novel cage compound 25a featuring a highly distorted TTF bridge (Scheme 6). In contrast to 25a , the deep cavitand 23 undergoes reversible vasekite switching upon lowering the temperature from 293 to 193 K (Fig. 1). Electrochemical studies by cyclic voltammetry (CV) and differential pulse voltammetry (DPV) provided preliminary evidence for successful vasekite switching of 23 induced by the oxidation of the TTF cavity walls.  相似文献   

2.
The synthesis of the quinoxaline‐bridged resorcin[4]arene cavitand 1 was accomplished from 2‐[3,5‐di(tert‐butyl)phenyl]acetaldehyde via formation of the intermediate octol 2 . Such cavitands are known to occur in an open `kite' conformation at low temperature (<213 K) but to adopt a `vase' conformation at elevated temperatures (>318 K). We discovered that protonation of cavitand 1 at room temperature by common acids, such as CF3COOH, also causes reversible switching from `vase' to `kite', and that this conformational change can be conveniently monitored by both 1H‐NMR and UV/VIS spectroscopy.  相似文献   

3.
Conformational changes of amide cavitands A – C were investigated at varied temperatures and in several solvents. While cavitands A and B , with comparatively smaller substituents such as Et and iPr, were always in vase conformation in non‐polar solvents such as CDCl3, CD2Cl2, (D8)THF, and C6D6, their thermoswitching (vase to kite) was observed in polar solvents such as (D7)DMF and (D6)DMSO or in the presence of acid (TFA) and H‐bonding inhibitor (TFE). Intra‐ and interannular H‐bonds of A and B were clearly observed by low‐temperature 1H‐NMR spectra in CDCl3. No conformational change of cavitand C with bigger substituent (tBu) was observed under any tested temperature range and in polar or non‐polar solvents; C was always in the kite conformation.  相似文献   

4.
A number of three‐fold C3v‐symmetrical tribenzotriquinacene (TBTQ) cavitands were synthesized by a “metamorphosis‐to‐half” strategy, employing six‐fold etherification reactions between the hexakis(chloromethyl)‐TBTQ intermediate 2 a and various 5‐functionalized resorcinols. X‐ray structure analyses of single crystals of the cavitands revealed limited rotational flexibility of the resorcinol bridging units, which enables an apical, nearly co‐axial orientation of the three functional groups and, as a consequence, the construction of nanoscale cage‐like molecules via covalent or coordination bonding. On this basis, two TBTQ‐based hemicryptophanes were prepared from the TBTQ cavitands via covalent bond formation in good yields. A dumbbell‐shaped TBTQ‐based metallo‐cryptophane was also synthesized in 34 % yield by a solvothermal reaction between cadmium nitrate and two equivalents of the TBTQ‐cavitand triacid, as confirmed by single‐crystal X‐ray diffraction and MALDI‐ToF mass spectrometry.  相似文献   

5.
We report the detailed investigation of temperature‐ and pH‐triggered conformational switching of resorcin[4]arene cavitands 1 – 10 (Figs. 1, 8, and 9). Depending on the experimental conditions, these macrocycles adopt a vase conformation, featuring a deep cavity for potential guest inclusion, or two kite conformations (kite 1 and kite 2) with flat, extended surfaces (Schemes 1 and 2). The thermodynamic and kinetic parameters for the interconversion between these structures were determined by variable‐temperature NMR (VT‐NMR) spectroscopy (Figs. 27 and 10, and Tables 1 and 2). It was discovered that vasekite switching of cavitands is strongly solvent‐dependent: it is controlled not only by solvent polarity but also by solvent size. Conformational interconversions similar to those of the parent structure 1 with four quinoxaline flaps are also observed when the octol base skeleton is differentially or incompletely bridged. Only octanitro derivative 2 was found to exist exclusively in the kite conformation under all experimental conditions. The detailed insight into the vase?kite conformational equilibrium gained in this investigation provides the basis for the design and construction of new, dynamic resorcin[4]arene cavitands that are switchable between bistable states featuring strongly different structures and functions.  相似文献   

6.
The synthesis, characterization, and photophysical as well as electrochemical properties of the photochromic hybrid systems 11 – 16 and 18 , which contain photoswitchable tetraethynylethene (TEE; 3,4‐diethynylhex‐3‐ene‐1,5‐diyne) and dihydroazulene (DHA) moieties, are presented. The molecular photoswitches were synthesized by a Sonogashira cross‐coupling reaction between an appropriate TEE precursor ( 6 – 10 and 17 ) and an iodinated DHA 1 or its vinylheptafulvene (VHF) isomer ( 4 ) (Schemes 5 – 7). X‐Ray crystal structures of five DHA derivatives ( 1 , trans‐ 11a , cis‐ 11a , 12 , and 13 ) are discussed (Figs. 25). In all compounds, the cyclohexatriene moiety of the DHA chromophore adopts a clear boat conformation (Table 1). Presumably due to crystal‐packing effects, the arylated TEE moieties in the hybrid systems show substantial distortions from planarity, with the dihedral angles between the planes of the central TEE core and the adjacent aryl substituents amounting to 44°. The switching properties were investigated by electronic absorption spectroscopy. Upon light absorption, DHAs 1 , 12 – 16 , and 18 underwent retro‐electrocyclization in solution to give the corresponding VHFs (Figs. 6, 11, and 12). The reaction is thermally reversible, with half‐lives τ1/2 between 3.9 and 5.8 h at 25° in CH2Cl2 (Figs. 7 and 13 and Table 3). A comparatively slower (E)→(Z) isomerization process about the central C=C bond of the TEE moiety was also observed. The N,N‐dimethylanilino‐(DMA) substituted TEE‐DHA hybrid systems trans‐ 11a and cis‐ 11a did not react to the corresponding VHFs upon irradiation (Scheme 9). Instead, only the reversible (E)→(Z) photoisomerization of the TEE core occurred (Fig. 16 and Table 4). This process was further investigated for photofatigue by electronic‐emission spectroscopy (Fig. 17). After protonation of the DMA group, the usual DHA→VHF photoreaction took place. Compound 11 represents a three‐way chromophoric molecular switch with three addressable sub‐units (TEE core, DHA/VHF moiety, and proton sensitive DMA group) that can undergo individual, reversible switching cycles (Scheme 9). A process modeling the function of an `AND' logic gate (Fig. 19) and three write/erase processes could be performed with this system. Cyclic and linear sweep‐voltammetry studies in CH2Cl2 (+Bu4NPF6) revealed the occurrence of characteristic first‐reduction steps in the TEE‐DHA hybrid systems between −1.6 and −1.8 V vs. Fc/Fc+ (ferrocene/ferricinium couple) (Table 5). Oxidations occur at ca. +1.10 V. After photoisomerization to the VHF derivatives, reduction steps at more positive and oxidation steps at more negative potentials were recorded. No DHA→VHF isomerization took place upon electrochemical oxidation or reduction (Fig. 20).  相似文献   

7.
Heptalenecarbaldehydes 1 / 1′ as well as aromatic aldehydes react with 3‐(dicyanomethylidene)‐indan‐1‐one in boiling EtOH and in the presence of secondary amines to yield 3‐(dialkylamino)‐1,2‐dihydro‐9‐oxo‐9H‐indeno[2,1‐c]pyridine‐4‐carbonitriles (Schemes 2 and 4, and Fig. 1). The 1,2‐dihydro forms can be dehydrogenated easily with KMnO4 in acetone at 0° (Scheme 3) or chloranil (=2,3,5,6‐tetrachlorocyclohexa‐2,5‐diene‐1,4‐dione) in a ‘one‐pot’ reaction in dioxane at ambient temperature (Table 1). The structures of the indeno[2,1‐c]pyridine‐4‐carbonitriles 5′ and 6a have been verified by X‐ray crystal‐structure analyses (Fig. 2 and 4). The inherent merocyanine system of the dihydro forms results in a broad absorption band in the range of 515–530 nm in their UV/VIS spectra (Table 2 and Fig. 3). The dehydrogenated compounds 5, 5′ , and 7a – 7f exhibit their longest‐wavelength absorption maximum at ca. 380 nm (Table 2). In contrast to 5 and 5′, 7a – 7f in solution exhibit a blue‐green fluorescence with emission bands at around 460 and 480 nm (Table 4 and Fig. 5).  相似文献   

8.
We report the synthesis of modified Cram‐type cavitands bearing one or two fluorescent labels for single‐molecule spectroscopic studies of vase? kite conformational switching (Scheme 3). Syntheses were performed by stepwise bridging of the four couples of neighboring H‐bonded OH groups of resorcin[4]arene bowls (Schemes 2 and 3). The new substitution patterns enable the construction of a large variety of future functional architectures. 1H‐NMR Investigations showed that the new partially and differentially bridged cavitands feature temperature‐ and pH‐triggered vase? kite conformational isomerism similar to symmetrical cavitands with four identical quinoxaline bridges (Table). It was discovered that vase? kite switching of cavitands is strongly solvent‐dependent.  相似文献   

9.
The majority of reactions currently performed in the chemical industry take place in organic solvents, compounds that are generally derived from petrochemicals. To promote chemical processes in water, we examined the use of synthetic, deep water‐soluble cavitands in the Staudinger reduction of long‐chain aliphatic diazides (C8, C10, and C12). The diazide substrates are taken up by the cavitand in D2O in folded, dynamic conformations. The reduction of one azide group to an amine gives a complex in which the substrate is fixed in an unsymmetrical conformation, with the amine terminal exposed and the azide terminal deep and inaccessible within the cavitand. Accordingly, the reduction of the second azide group is inhibited, even with excess phosphine, and good yields of the monofunctionalized products are obtained. In contrast, the reduction of the free diazides in bulk solution yields diamine products.  相似文献   

10.
Hybrid supramolecular capsules self‐assemble by simultaneously forming hydrogen and metal–ligand coordination bonds on mixing a C2‐symmetrical cavitand (calix[4]resorcinarene‐based cavitands with ureide and terminal 4‐pyridyl units) with platinum or palladium complexes ([Pt(OTf)2] or [Pd(OTf)2] with chelating bisphosphines) in 1:1 ratio. Hemicapsular assemblies formed in the presence of excess amounts of cavitand relative to the platinum or palladium complexes are identified as intermediates in the above self‐assembly process by 2D‐NOESY spectroscopy. External‐anion‐assisted encapsulation of a neutral guest, 4,4′‐diiodobiphenyl, inside the hybrid supramolecular capsules accompanied conformational changes in the hydrogen‐bonding moieties. The in/out exchange ratio of the encapsulated guest depends on the bite angle of the bisphosphine ligand. Addition of DMSO accelerates guest exchange by weakening the hydrogen bonds in the encapsulation complex. Therefore, variations in the structure of the metal complex and amount of polar solvent exert dual control on the dynamics of the guest exchange.  相似文献   

11.
We report the synthesis of the first‐ (G1) and second‐generation (G2) dendritic FeII porphyrins 1?Fe – 4?Fe (G1) and 6?Fe (G2) bearing distal H‐bond donors ideally positioned for stabilization of FeII? O2 adducts by H‐bonding (Fig. 1). A first approach towards the construction of these novel biomimetic systems failed unexpectedly: the Suzuki cross‐coupling between appropriately functionalized ZnII porphyrins and ortho‐ethynylated aryl derivatives, serving as anchors for the distal H‐bond donor moieties, was unsuccessful (Schemes 1, 3, and 5), presumably due to steric hindrance resulting from unfavorable coordination of the ethynyl residue to the Pd species in the catalytic cycle (Scheme 6). The target molecules were finally prepared by a route in which the ortho‐ethynylated meso‐aryl ring is introduced during porphyrin construction in a mixed condensation involving the two dipyrrylmethanes 33 and 34 , and aldehyde 36 (Schemes 7 and 8). Following attachment of the dendrons (Scheme 11), the distal H‐bond donors were introduced by Sonogashira cross‐coupling (Scheme 12), and subsequent metallation afforded the dendritic FeII porphyrins 1?Fe – 6?Fe . 1H‐NMR Spectroscopy proved the location of the H‐bond donor moiety atop the porphyrin surface, and X‐ray crystal‐structure analysis of model system 45 (Fig. 2) revealed that this moiety would not sterically interfere with gas binding. With 1,2‐dimethyl‐1H‐imidazole (DiMeIm) as ligand, the dendritic FeII porphyrins formed five‐coordinate high‐spin complexes (Figs. 3 and 4) and addition of CO led reversibly to the corresponding stable six‐coordinate gas complexes (Fig. 6). Oxygenation, however, did not result in defined FeII? O2 complexes as rapid decomposition to FeIII species took place immediately, even in the case of the G2 dendrimer 6?Fe (DiMeIm) (Fig. 7). In contrast, stable gas adducts are formed between dendritic CoII porphyrins and O2 in the presence of DiMeIm as axial ligand, as revealed by electron paramagnetic resonance (EPR). The possible stabilization of these complexes through H‐bonding involving the distal ligand is currently under investigation in multidimensional and multifrequency pulse EPR experiments.  相似文献   

12.
The 1,4‐cis‐diethynylated α‐D ‐mannopyranose analogue 11 has been prepared from 1,6 : 2,3‐dianhydro‐β‐D ‐allopyranose ( 6 ) by alkynylating epoxide and acetal opening (Scheme 2). Eglinton coupling of 11 gave the cyclodimer 18 (Scheme 3). Crystal‐structure analysis of the corresponding bis(methanesulfonate) 19 revealed substantially bent butadiyne moieties; one mannopyranosyl ring adopts the 4C1 and the other one a slightly distorted OS2 conformation (Fig. 1). Hydrogenation of 18 , followed by deprotection, gave the stable butane‐1,4‐diyl‐bridged cyclodimer 21 (Scheme 3). Crystal‐structure analysis shows the 4C1 conformation of the mannopyranosyl units (Fig. 2). The two butane fragments are characterised by a combination of gauche and antiperiplanar arrangements.  相似文献   

13.
Three bowl‐type cavitand receptors ( 1 – 3 ), consisting of a resorcin[4]arene core with four convergent phenylamidinium groups, were prepared in gram quantities by efficient synthetic routes (Schemes 1 and 2) for the recognition of organic anions in CD3OD and D2O. The key steps in the syntheses are the Suzuki cross‐coupling reactions between the tetraiodo cavitands 12 , 13 , and 23 , respectively, with the m‐cyanophenylboronic ester 14 and subsequent conversion of the nitrile to amidinium groups by the Garigipati reaction. Compounds 1 and 2 displayed limited solubility in protic solvents, and evidence for stoichiometric host‐guest association between 2 and AMP ( 28 ) could only be obtained by FAB‐MS analysis of a complex precipitated from MeOH (Fig. 2). In contrast, receptor 3 with four triethyleneglycol monomethyl ether side chains was readily soluble in the protic environments, and complexation of isophthalates and nucleotides 25 – 37 was extensively studied by 1H‐NMR titrations and Job's method of continuous variation. In CD3OD and pure D2O, isophthalates 25 and 26 formed stable 1 : 2 host‐guest complexes (Table 1 and Fig. 3), whereas upon addition of borate (pH 9.2) or Tris/HCl buffer (pH 8.3) to the D2O solution, the tendency for higher‐order complexation vanishes. Stable 1 : 1 complexes formed in the buffered solutions (Fig. 4) with 5‐methoxyisophthalate being selectively bound over the 5‐NO2 derivative. Complexation‐induced upfield shifts of specific isophthalate 1H‐NMR resonances (Fig. 5) suggest a preferred inclusion of the methoxyphenyl ring into the receptor cavity. Cavitand 3 forms stable 1 : 1 host‐guest complexes with nucleotides in Tris/HCl‐buffered D2O. Association constants increase strongly with increasing guest charge in the series cAMP<AMP<ADP<ATP (Table 2). Association strength is strongly reduced in the presence of high salt (NaCl) concentration (Table 3). Receptor 3 shows a slight preference for the complexation of AMP (Fig. 7) and analogs dAMP or ε‐AMP (Table 4) over nucleotides containing G (guanine), C (cytosine), T (thymine), or U (uracil) as bases. According to the 1H‐NMR analysis, only the nucleobase adenine and derivatives thereof possess the necessary stereoelectronic complementarity for inclusion into the bowl‐type cavity. The major forces stabilizing the complexes of 3 with isophthalates and nucleotides result from ion pairing and ionic H‐bonding between the charged groups of host and guest, and from the desolvation of these groups upon complexation. Apolar interactions and hydrophobic desolvation do not seem to make a large contribution to the measured binding free enthalpies.  相似文献   

14.
The oxidation of 1‐(3,8‐dimethylazulen‐1‐yl)alkan‐1‐ones 1 with 2,3‐dichloro‐5,6‐dicyano‐1,4‐benzoquinone (=4,5‐dichloro‐3,6‐dioxocyclohexa‐1,4‐diene‐1,2‐dicarbonitrile; DDQ) in acetone/H2O mixtures at room temperature does not only lead to the corresponding azulene‐1‐carboxaldehydes 2 but also, in small amounts, to three further products (Tables 1 and 2). The structures of the additional products 3 – 5 were solved spectroscopically, and that of 3a also by an X‐ray crystal‐structure analysis (Fig. 1). It is demonstrated that the bis(azulenylmethyl)‐substituted DDQ derivatives 5 yield on methanolysis or hydrolysis precursors, which in a cascade of reactions rearrange under loss of HCl into the pentacyclic compounds 3 (Schemes 4 and 7). The found 1,1′‐[carbonylbis(8‐methylazulene‐3,1‐diyl)]bis[ethanones] 4 are the result of further oxidation of the azulene‐1‐carboxaldehydes 2 to the corresponding azulene‐1‐carboxylic acids (Schemes 9 and 10).  相似文献   

15.
The cyclic 16‐membered pentadepsipeptide cyclo(Tro‐Aib‐Aib‐Aib‐Aib) ( 1 ) was crystallized from MeOH/AcOEt/CH2Cl2, and its structure was established by X‐ray crystallography (Fig. 1). There are two symmetry‐independent molecules with different conformations in the asymmetric unit. Two intramolecular H‐bonds stabilize two β‐turns in each molecule. On the other hand, two of the four Aib residues are forced to assume a nonfavorable nonhelical conformation in each of the symmetry‐independent molecules (Table 1). The conformational study in CDCl3 solution by NMR spectroscopy and molecular dynamics (MD) simulations indicate that the averaged structure (Fig. 3) is almost the same as in the solid state.  相似文献   

16.
A new family of optically active cyclophane receptors for the complexation of mono‐ and disaccharides in competitive protic solvent mixtures is described. Macrocycles (−)‐(R,R,R,R)‐ 1 – 4 feature preorganized binding cavities formed by four 1,1′‐binaphthalene‐2,2′‐diyl phosphate moieties bridged in the 3,3′‐positions by acetylenic or phenylacetylenic spacers. The four phosphodiester groups converge towards the binding cavity and provide efficient bidentate ionic H‐bond acceptor sites (Fig. 2). Benzyloxy groups in the 7,7′‐positions of the 1,1′‐binaphthalene moieties ensure solubility of the nanometer‐sized receptors and prevent undesirable aggregation. The construction of the macrocyclic framework of the four cyclophanes takes advantage of Pd0‐catalyzed aryl—acetylene cross‐coupling by the Sonogashira protocol, and oxidative acetylenic homo‐coupling methodology (Schemes 2 and 8 – 10). Several cleft‐type receptors featuring one 1,1′‐binaphthalene‐2,2′‐diyl phosphate moiety were also prepared (Schemes 1, 6, and 7). An undesired side reaction encountered during the synthesis of the target compounds was the formation of naptho[b]furan rings from 3‐ethynylnaphthalene‐2‐ol derivatives, proceeding via 5‐endo‐dig cyclization (Schemes 35). Computer‐assisted molecular modeling indicated that the macrocycles prefer nonplanar puckered, cyclobutane‐type conformations (Figs. 7 and 8). According to these calculations, receptor (−)‐(R,R,R,R)‐ 1 has, on average, a square binding site, which is complementary in size to one monosaccharide. The three other cyclophanes (−)‐(R,R,R,R)‐ 2 – 4 feature, on average, wider rectangular cavities, providing a good fit to one disaccharide, while being too large for the complexation of one monosaccharide. This substrate selectivity was fully confirmed in 1H‐NMR binding titrations. The chiroptical properties of the cyclophanes and their nonmacrocyclic precursors were investigated by circular dichroism (CD) spectroscopy. The CD spectra of the acyclic precursors showed a large dependence from the number of 1,1′‐binaphthalene moieties (Fig. 9), and those of the cyclophanes were remarkably influenced by the nature of the functional groups lining the macrocyclic cavity (Fig. 11). Profound differences were also observed between the CD spectra of linear and macrocyclic tetrakis(1,1′‐binaphthalene) scaffolds, which feature very different molecular shapes (Fig. 10). In 1H‐NMR binding titrations with mono‐ and disaccharides (Fig. 13), concentration ranges were chosen to favor 1 : 1 host−guest binding. This stoichiometry was experimentally established by the curve‐fitting analysis of the titration data and by Job plots. The titration data demonstrate conclusively that the strength of carbohydrate recognition is enhanced with an increasing number of bidentate ionic host−guest H‐bonds (Table 1) in the complex formed. As a result of the formation of these highly stable H‐bonds, carbohydrate complexation in competitive protic solvent mixtures becomes more favorable. Thus, cleft‐type receptors (−)‐(R)‐ 7 and (−)‐(R)‐ 38 with one phosphodiester moiety form weak 1 : 1 complexes only in CD3CN. In contrast, macrocycle (−)‐(R,R,R,R)‐ 1 with four phosphodiester groups undergoes stable inclusion complexation with monosaccharides in CD3CN containing 2% CD3OD. With their larger number of H‐bonding sites, disaccharide substrates bind even more strongly to the four phosphodiester groups lining the cavity of (−)‐(R,R,R,R)‐ 2 and complexation becomes efficient in CD3CN containing 12% CD3OD. Finally, the introduction of two additional methyl ester residues further enhances the receptor capacity of (−)‐(R,R,R,R)‐ 3 , and efficient disaccharide complexation occurs already in CD3CN containing 20% CD3OD.  相似文献   

17.
Resorcinarene‐based cavitands 1a – c fold into a deep open‐ended cavity by means of intramolecular hydrogen bonds in both apolar solutions and the solid state. The X‐ray crystal‐structure analysis of cavitand 1a features a seam of secondary amide C=O⋅⋅⋅H−N interactions that bridge adjacent rings and are held in place by intra‐annular hydrogen bonds. This results in a cavity of 9.2×7.0 Å dimensions. The arrangement of the amides in 1a – 1c is cycloenantiomeric, with clock‐ and counterclockwise orientation of the head‐to‐tail amide sequence. Interconversion rates of the two enantiomers are controlled by solvent polarity: the rate is slow on the NMR time‐scale in aromatic solvents and CDCl3, but fast in (D6)acetone. The 1H‐ and 13C‐NMR‐spectral analysis is in agreement with the crystallographic data. Chiral cavitand 1b with eight HN−C(O)−C*HMeEt ((+)‐(S)) groups on its upper rim exists as two cyclodiastereoisomers (in a ca. 3 : 1 ratio) in apolar solution. A `library' of 512 diastereoisomeric cavitands 1c is obtained as a mixture by using the corresponding racemic acid chloride.  相似文献   

18.
The single‐crystal X‐ray data of akuammigine picrate hydrate ( 1 ?Picr?H2O) confirm the relative configuration of the indole alkaloid akuammigine ( 1 ) as epiallo (Fig. 1). With reference to the known (15S)‐configuration due to biosynthesis, the absolute configuration of the other stereogenic centers is thus given by (3R,19S,20S). Four crystallographically independent molecules are observed in the asymmetric unit (Fig. 2). Each of the alkaloid cations forms H‐bonds to a H2O and a picrate anion (Fig. 3). The H2O molecules are further associated by a H‐bond as indicated by the short O???O distance (Table 2). The conformation in the solid state of the picrate hydrate is now firmly established, and a cute H‐bonding motif is observed.  相似文献   

19.
A full account is given of the preparation and use of TADDOLates, which are dendritically incorporated in polystyrene beads (Scheme 1). A series of styryl‐substituted TADDOLs with flexible, rigid, or dendritically branching spacers between the TADDOL core and the styryl groups (2–16 in number) has been prepared ( 5 – 7, 20, 21, 26 in Schemes 24 and Fig. 13). These were used as cross‐linkers in styrene‐suspension polymerization, leading to beads of ca. 400‐μm diameter (Schemes 5 and 6, b). These, in turn, were loaded with titanate and used for the Lewis acid catalyzed addition of Et2Zn to PhCHO as a test reaction (Scheme 6). A comparison of the enantioselectivities and degrees of conversion (both up to 99%), obtained under standard conditions, shows that these polymer‐incorporated Ti‐TADDOLates are highly efficient catalysts for this process (Table 1). In view of the effort necessary to prepare the novel, immobilized catalysts, emphasis was laid upon their multiple use. The performance over 20 cycles of the test reaction was best with the polymer obtained from the TADDOL bearing four first‐generation Fréchet branches with eight peripheral styryl groups ( 6 , p‐ 6 , p‐ 6 ⋅Ti(OiPr)2): the enantioselectivity (Fig. 4), the rate of reaction (Fig. 5), and the swelling factor (Fig. 6) were essentially unchanged after numerous operations carried out with the corresponding beads of 400‐μm diameter and a degree of loading of 0.1 mmol TADDOLate/g polymer, with or without stirring (Fig. 7). The rate with the dendritically polymer‐embedded Ti‐TADDOLate (p‐ 6 ⋅Ti(OiPr)2) was greater than that measured with the corresponding monomer, i.e., 6 ⋅Ti(OiPr)2 (Fig. 8). Possible interpretations of this phenomenon are proposed. A polymer‐bound TADDOL, generated on a solid support (by Grignard addition to an immobilized tartrate ester ketal) did not perform well (Scheme 4 and Table 2). Also, when we prepared polystyrene beads by copolymerization of styrene, a zero‐, first‐, or second‐generation dendritic cross‐linker, and a mono‐styryl‐substituted TADDOL derivative, the performance in the test reaction did not rival that of the dendritically incorporated Ti‐TADDOLate ((p‐ 6 ⋅Ti(OiPr)2) (Scheme 7 and Fig. 10). Finally, we have applied the dendritically immobilized Cl2 and (TsO)2Ti‐TADDOLate as chiral Lewis acid to preferentially prepare one enantiomer of the exo and the endo (3+2) cycloadduct, respectively, of diphenyl nitrone to 3‐crotonoyl‐1,3‐oxazolidinone; in one of these reaction modes, we have observed an interesting conditioning of the catalyst: with an increasing number of application cycles, the amount of polymer‐incorporated Lewis acid required to induce the same degree of enantioselectivity, decreased; the degrees of diastereo‐ and enantioselectivity were, again, comparable to those reported for homogeneous conditions (Fig. 9).  相似文献   

20.
The covalently templated bis‐functionalization of C70, employing bis‐malonate 5 tethered by an anti‐disubstituted dibenzo[18]crown‐6 (DB18C6) ether, proceeds with complete regiospecificity and provides two diastereoisomeric pairs of enantiomeric C70 crown ether conjugates, (±)‐ 7a and (±)‐ 7b , featuring a five o'clock bis‐addition pattern that is disfavored in sequential transformations (Scheme 1). The identity of (±)‐ 7a was revealed by X‐ray crystal‐structure analysis (Fig. 6). With bis‐malonate 6 containing a syn‐disubstituted DB18C6 tether, the regioselectivity of the macrocylization via double Bingel cyclopropanation changed completely, affording two constitutionally isomeric C70 crown ether conjugates in a ca. 1 : 1 ratio featuring the twelve ( 16 ) and two o'clock ((±)‐ 15 ) addition patterns, respectively (Scheme 3). The X‐ray crystal‐structure analysis of the twelve o'clock bis‐adduct 16 revealed that a H2O molecule was included in the crown ether cavity (Figs. 7 and 8). Two sequential Bingel macrocyclizations, first with anti‐DB18C6‐tethered ( 5 ) and subsequently with syn‐DB18C6‐tethered ( 6 ) bis‐malonates, provided access to the first fullerene bis‐crown ether conjugates. The two diastereoisomeric pairs of enantiomers (±)‐ 28a and (±)‐ 28b were formed in high yield and with complete regioselectivity (Scheme 9). The cation‐binding properties of all C70 crown‐ether conjugates were determined with the help of ion‐selective electrodes (ISEs). Mono‐crown ether conjugates form stable 1 : 1 complexes with alkali‐metal ions, whereas the tetrakis‐adducts of C70, featuring two covalently attached crown ethers, form stable 1 : 1 and 1 : 2 host‐guest complexes (Table 2). Comparative studies showed that the conformation of the DB18C6 ionophore imposed by the macrocyclic bridging to the fullerene is not particularly favorable for strong association. Reference compound (±)‐ 22 (Scheme 4), in which the DB18C6 moiety is attached to the C70 sphere by a single bridge only and, therefore, possesses higher conformational flexibility, binds K+ and Na+ ions better by factors of 2 and 20, respectively. Electrochemical studies demonstrate that cation complexation at the crown ether site causes significant anodic shifts of the first reduction potential of the appended fullerene (Table 3). In case of the C70 mono‐crown ether conjugates featuring a five o'clock functionalization pattern, addition of 1 equiv. of KPF6 caused an anodic shift of the first reduction wave in the cyclic voltammogram (CV) by 70 to 80 mV, which is the result of the electrostatic effect of the K+ ion bound closely to the fullerene core (Fig. 14). Addition of 2 equiv. of K+ ions to C70 bis‐crown ether conjugates resulted in the observation of only one redox couple, whose potential is anodically shifted by 170 mV with respect to the corresponding wave in the absence of the salt (Fig. 16). The synthesis and characterization of novel tris‐ and tetrakis‐adducts of C70 are reported (Schemes 5 and 6). Attempts to prepare even more highly functionalized derivatives resulted in the formation of novel pentakis‐ and hexakis‐adducts and a single heptakis‐adduct (Scheme 7), which were characterized by 1H‐ and 13C‐NMR spectroscopy (Fig. 10), as well as matrix‐assisted laser‐desorption‐ionization mass spectrometry (MALDI‐TOF‐MS). Based on predictions from density‐functional‐theory (DFT) calculations (Figs. 12 and 13), structures are proposed for the tris‐, tetrakis‐, and pentakis‐adducts.  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