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 312 毫秒
1.
Two dynamic covalent polymers P1 and P2 were prepared by alternately linking electron‐rich tetrathiafulvalene (TTF) and electron‐deficient bipyridinium (BIPY2+) through hydrazone bonds. In acetonitrile, the polymers were induced by intramolecular donor–acceptor interactions to form pleated foldamers, which unfolded upon oxidation of the TTF units to the radical cation TTF.+. Reduction of the BIPY2+ units to BIPY.+ led to the formation of another kind of pleated secondary structures, which are stabilized by intramolecular dimerization of the BIPY.+ units. The diradical dicationic cyclophane cyclobis(paraquat‐p‐phenylene) (CBPQT2(.+)) could further force the folded structures to unfold by including the BIPY.+ units of the polymers. Upon oxidation of the BIPY.+ units of the cyclophane and polymers to BIPY2+, the first folded state was regenerated. Switching or conversion between the four conformational states was confirmed by UV/Vis spectroscopic experiments.  相似文献   

2.
We report the synthesis of two [2]catenane‐containing struts that are composed of a tetracationic cyclophane (TC4+) encircling a 1,5‐dioxynaphthalene (DNP)‐based crown ether, which bears two terphenylene arms. The TC4+ rings comprise either 1) two bipyridinium (BIPY2+) units or 2) a BIPY2+ and a diazapyrenium (DAP2+) unit. These degenerate and nondegenerate catenanes were reacted in the presence of Cu(NO3)2?2.5 H2O to yield Cu‐paddlewheel‐based MOF‐1050 and MOF‐1051. The solid‐state structures of these MOFs reveal that the metal clusters serve to join the heptaphenylene struts into grid‐like 2D networks. These 2D sheets are then held together by infinite donor–acceptor stacks involving the [2]catenanes to produce interpenetrated 3D architectures. As a consequence of the planar chirality associated with both the DNP and hydroquinone (HQ) units present in the crown ether, each catenane can exist as four stereoisomers. In the case of the nondegenerate (bistable) catenane, the situation is further complicated by the presence of translational isomers. Upon crystallization, however, only two of the four possible stereoisomers—namely, the enantiomeric RR and SS forms—are observed in the crystals. An additional element of co‐conformational selectivity is present in MOF‐1051 as a consequence of the substitution of one of the BIPY2+ units by a DAP2+ unit: only the translational isomer in which the DAP2+ unit is encircled by the crown ether is observed. The overall topologies of MOF‐1050 and MOF‐1051, and the selective formation of stereoisomers and translational isomers during the kinetically driven crystallization, provide evidence that weak noncovalent bonding interactions play a significant role in the assembly of these extended (super)structures.  相似文献   

3.
The ability to control the kinetic barriers governing the relative motions of the components in mechanically interlocked molecules is important for future applications of these compounds in molecular electronic devices. In this Full Paper, we demonstrate that bipyridinium (BIPY2+) dications fulfill the role as effective electrostatic barriers for controlling the shuttling and threading behavior for rotaxanes and pseudorotaxanes in aqueous environments. A degenerate [2]rotaxane, composed of two 1,5‐dioxynaphthalene (DNP) units flanking a central BIPY2+ unit in the dumbbell component and encircled by the cyclobis(paraquat‐p‐phenylene) (CBPQT4+) tetracationic cyclophane, has been synthesized employing a threading‐followed‐by‐stoppering approach. Variable‐temperature 1H NMR spectroscopy reveals that the barrier to shuttling of the CBPQT4+ ring over the central BIPY2+ unit is in excess of 17 kcal mol?1 at 343 K. Further information about the nature of the BIPY2+ unit as an electrostatic barrier was gleaned from related supramolecular systems, utilizing two threads composed of either two DNP units flanking a central BIPY2+ moiety or a central DNP unit flanked by a BIPY2+ moiety. The threading and dethreading processes of the CBPQT4+ ring with these compounds, which were investigated by spectrophotometric techniques, reveal that the BIPY2+ unit is responsible for affecting both the thermodynamics and kinetics of pseudorotaxane formation by means of an intramolecular self‐folding (through donor–acceptor interactions with the DNP unit), in addition to Coulombic repulsion. In particular, the free energy barrier to threading (Δ${G{{{\ne}\hfill \atop {\rm f}\hfill}}}The ability to control the kinetic barriers governing the relative motions of the components in mechanically interlocked molecules is important for future applications of these compounds in molecular electronic devices. In this Full Paper, we demonstrate that bipyridinium (BIPY(2+)) dications fulfill the role as effective electrostatic barriers for controlling the shuttling and threading behavior for rotaxanes and pseudorotaxanes in aqueous environments. A degenerate [2]rotaxane, composed of two 1,5-dioxynaphthalene (DNP) units flanking a central BIPY(2+) unit in the dumbbell component and encircled by the cyclobis(paraquat-p-phenylene) (CBPQT(4+)) tetracationic cyclophane, has been synthesized employing a threading-followed-by-stoppering approach. Variable-temperature (1)H?NMR spectroscopy reveals that the barrier to shuttling of the CBPQT(4+) ring over the central BIPY(2+) unit is in excess of 17 kcal mol(-1) at 343 K. Further information about the nature of the BIPY(2+) unit as an electrostatic barrier was gleaned from related supramolecular systems, utilizing two threads composed of either two DNP units flanking a central BIPY(2+) moiety or a central DNP unit flanked by a BIPY(2+) moiety. The threading and dethreading processes of the CBPQT(4+) ring with these compounds, which were investigated by spectrophotometric techniques, reveal that the BIPY(2+) unit is responsible for affecting both the thermodynamics and kinetics of pseudorotaxane formation by means of an intramolecular self-folding (through donor-acceptor interactions with the DNP unit), in addition to Coulombic repulsion. In particular, the free energy barrier to threading (ΔG(f)(++)) of the CBPQT(4+) for the case of the thread composed of a DNP flanked by two BIPY(2+) units was found to be as high as 21.7 kcal mol(-1) at room temperature. These results demonstrate that we can effectively employ the BIPY(2+) unit to serve as electrostatic barriers in water in order to gain control over the motions of the CBPQT(4+) ring in both mechanically interlocked and supramolecular systems.  相似文献   

4.
The tricyclic azoalkanes, (1α,4α,4aα,7aα)‐4,4a,5,6,7,7a‐hexahydro‐1,4,8,8‐tetramethyl‐1,4‐methano‐1H‐cyclopenta[d]pyridazine ( 1c ), (1α,4α,4aα,6aα)‐4,4a,5,6,6a‐pentahydro‐1,4,7,7‐tetramethyl‐1,4‐methano‐1H‐cyclobuta[d]pyridazine ( 1d ), (1α,4α,4aα,6aα)‐4,4a,6a‐trihydro‐1,4,7,7‐tetramethyl‐1,4‐methano‐1H‐cyclobuta[d]pyridazine ( 1e ), and (1α,4α,4aα,5aα)‐4,4a,5,5a‐tetrahydro‐1,4,6,6‐tetramethyl‐1,4‐methano‐1H‐cyclopropa[d]pyridazine ( 1f ), as well as the corresponding housanes, the 2,3,3,4‐tetramethyl‐substituted tricyclo[3.3.0.02,4]octane ( 2c ), tricyclo[3.2.0.02,4]heptane ( 2d ), and tricyclo[3.2.0.02,4]hept‐6‐ene ( 2e ), were subjected to γ‐irradiation in Freon matrices. The reaction products were identified with the use of ESR and, in part, ENDOR spectroscopy. As expected, the strain on the C‐framework increases on going from the cyclopentane‐annelated azoalkanes and housanes ( 1c and 2c ) to those annelated by cyclobutane ( 1d and 2d ), by cyclobutene ( 1e and 2e ), and by cyclopropane ( 1f ). Accordingly, the products obtained from 1c and 2c in all three Freons used, CFCl3, CF3CCl3, and CF2ClCFCl2, were the radical cations 3c .+ and 2c .+ of 2,3,4,4‐tetramethylbicyclo[3.3.0]oct‐2‐ene and 2,3,3,4‐tetramethylbicyclo[3.3.0]octane‐2,4‐diyl, respectively. In CFCl3 and CF3CCl3 matrices, 1d and 2d yielded analogous products, namely the radical cations 3d .+ and 2d .+ of 2,3,4,4‐tetramethylbicyclo[3.2.0]hept‐2‐ene and 2,3,3,4‐tetramethylbicyclo[3.2.0]heptane‐2,4‐diyl. The radical cations 3c .+ and 3d .+ and 2c .+ and 2d .+ correspond to their non‐annelated counterparts 3a .+ and 3b .+, and 2a .+ and 2b .+ generated previously under the same conditions from 2,3‐diazabicyclo[2.2.1]hept‐2‐ene ( 1a ) and bicyclo[2.1.0]pentane ( 2a ), as well as from their 1,4‐dimethyl derivatives ( 1b and 2b ). However, in a CF2ClCFCl2 matrix, both 1d and 2d gave the radical cation 4d .+ of 2,3,3,4‐tetramethylcyclohepta‐1,4‐diene. Starting from 1e and 2e , the radical cations 4e .+ and 4e′ .+ of the isomeric 1,2,7,7‐ and 1,6,7,7‐tetramethylcyclohepta‐1,3,5‐trienes appeared as the corresponding products, while 1f was converted into the radical cation 4f .+ of 1,5,6,6‐tetramethylcyclohexa‐1,4‐diene which readily lost a proton to yield the corresponding cyclohexadienyl radical 4f .. Reaction mechanisms leading to the pertinent radical cations are discussed.  相似文献   

5.
The reaction of the aromatic distonic peroxyl radical cations N‐methyl pyridinium‐4‐peroxyl (PyrOO.+) and 4‐(N,N,N‐trimethyl ammonium)‐phenyl peroxyl (AnOO.+), with symmetrical dialkyl alkynes 10a – c was studied in the gas phase by mass spectrometry. PyrOO.+ and AnOO.+ were produced through reaction of the respective distonic aryl radical cations Pyr.+ and An.+ with oxygen, O2. For the reaction of Pyr.+ with O2 an absolute rate coefficient of k1=7.1×10?12 cm3 molecule?1 s?1 and a collision efficiency of 1.2 % was determined at 298 K. The strongly electrophilic PyrOO.+ reacts with 3‐hexyne and 4‐octyne with absolute rate coefficients of khexyne=1.5×10?10 cm3 molecule?1 s?1 and koctyne=2.8×10?10 cm3 molecule?1 s?1, respectively, at 298 K. The reaction of both PyrOO.+ and AnOO.+ proceeds by radical addition to the alkyne, whereas propargylic hydrogen abstraction was observed as a very minor pathway only in the reactions involving PyrOO.+. A major reaction pathway of the vinyl radicals 11 formed upon PyrOO.+ addition to the alkynes involves γ‐fragmentation of the peroxy O? O bond and formation of PyrO.+. The PyrO.+ is rapidly trapped by intermolecular hydrogen abstraction, presumably from a propargylic methylene group in the alkyne. The reaction of the less electrophilic AnOO.+ with alkynes is considerably slower and resulted in formation of AnO.+ as the only charged product. These findings suggest that electrophilic aromatic peroxyl radicals act as oxygen atom donors, which can be used to generate α‐oxo carbenes 13 (or isomeric species) from alkynes in a single step. Besides γ‐fragmentation, a number of competing unimolecular dissociative reactions also occur in vinyl radicals 11 . The potential energy diagrams of these reactions were explored with density functional theory and ab initio methods, which enabled identification of the chemical structures of the most important products.  相似文献   

6.
Two water‐soluble para‐xylylene‐connected 4,4′‐bipyridinium (BIPY2+) polymers have been prepared. UV‐Vis absorption, 1H NMR spectroscopy, and cyclic voltammetry experiments support that in water the BIPY2+ units in the polymers form stable 1:1 charge‐transfer complexes with tetrathiafulvalene (TTF) guests that bear two or four carboxylate groups. These charge‐transfer complexes are stabilized by the donor–acceptor interaction between electron‐rich TTF and electron‐deficient BIPY2+ units and electrostatic attraction between the dicationic BIPY2+ units and the anionic carboxylate groups attached to the TTF core. On the basis of UV‐Vis experiments, a lower limit to the apparent association constant of the TTF?BIPY2+ complexes of the mixtures, 1.8×106 m ?1, has been estimated in water. Control experiments reveal substantially reduced binding ability of the neutral TTF di‐ and tetracarboxylic acids to the BIPY2+ molecules and polymers. Moreover, the stability of the charge‐transfer complexes formed by the BIPY2+ units of the polymers are considerably higher than that of the complexes formed between two monomeric BIPY2+ controls and the dicarboxylate‐TTF donor; this has been attributed to the mutually strengthened electron‐deficient nature of the BIPY2+ units of the polymers due to the electron‐withdrawing effect of the BIPY2+ units.  相似文献   

7.
Sterically unprotected thiophene/phenylene co‐oligomer radical cation salts BPnT.+[Al(ORF)4]? (ORF=OC(CF3)3, n=1–3) have been successfully synthesized. These newly synthesized salts have been characterized by UV/Vis‐NIR absorption and EPR spectroscopy, and single‐crystal X‐ray diffraction analysis. Their conductivity increases with chain length. The formed meso‐helical stacking by cross‐overlapping radical cations of BP2T.+ is distinct from previously reported face‐to‐face overlaps of sterically protected (co‐)oligomer radical cations.  相似文献   

8.
We show that the radical cations of adamantane (C10H16.+, 1 H.+) and perdeuteroadamantane (C10D16.+, 1 D.+) are stable species in the gas phase. The radical cation of adamantylideneadamantane (C20H28.+, 2 H.+) is also stable (as in solution). By using the natural 13C abundances of the ions, we determine the rate constants for the reversible isergonic single‐electron transfer (SET) processes involving the dyads 1 H.+/ 1 H, 1 D.+/ 1 D and 2 H.+/ 2 H. Rate constants for the reaction 1 H.++ 1 D? 1 H+ 1 D.+ are also determined and Marcus’ cross‐term equation is shown to hold in this case. The rate constants for the isergonic processes are extremely high, practically collision‐controlled. Ab initio computations of the electronic coupling (HDA) and the reorganization energy (λ) allow rationalization of the mechanism of the process and give insights into the possible role of intermediate complexes in the reaction mechanism.  相似文献   

9.
We have used model tripeptides GXW (with X being one of the amino acid residues glycine (G), alanine (A), leucine (L), phenylalanine (F), glutamic acid (E), histidine (H), lysine (K), or arginine (R)) to study the effects of the basicity of the amino acid residue on the radical migrations and dissociations of odd‐electron molecular peptide radical cations M.+ in the gas phase. Low‐energy collision‐induced dissociation (CID) experiments revealed that the interconvertibility of the isomers [G.XW]+ (radical centered on the N‐terminal α‐carbon atom) and [GXW].+ (radical centered on the π system of the indolyl ring) generally increased upon increasing the proton affinity of residue X. When X was arginine, the most basic amino acid, the two isomers were fully interconvertible and produced almost identical CID spectra despite the different locations of their initial radical sites. The presence of the very basic arginine residue allowed radical migrations to proceed readily among the [G.RW]+ and [GRW].+ isomers prior to their dissociations. Density functional theory calculations revealed that the energy barriers for isomerizations among the α‐carbon‐centered radical [G.RW]+, the π‐centered radical [GRW].+, and the β‐carbon‐centered radical [GRWβ.]+ (ca. 32–36 kcal mol−1) were comparable with those for their dissociations (ca. 32–34 kcal mol−1). The arginine residue in these GRW radical cations tightly sequesters the proton, thereby resulting in minimal changes in the chemical environment during the radical migrations, in contrast to the situation for the analogous GGW system, in which the proton is inefficiently stabilized during the course of radical migration.  相似文献   

10.
Compounds 1 a and 1 b were prepared by appending two tetrathiafulvalene (TTF) units to an aromatic amide segment that is driven by six or two intramolecular N? H???O hydrogen bonds to adopt a folded conformation. UV/Vis absorption experiments revealed that if the TTF units were oxidized to TTF.+ radical cations, the two compounds could form a stable single molecular noncovalent macrocycle in less polar dichloromethane or dichloroethane or a bimolecular noncovalent macrocycle in a binary mixture of dichloromethane with a more polar solvent owing to remarkably enhanced dimerization of the TTF.+ units. The stability of the (TTF.+)2 dimer was evaluated through UV/Vis absorption, electron paramagnetic resonance, and cyclic voltammetry experiments and also by comparing the results with those of control compound 2 . The results showed that introduction of the intramolecular hydrogen bonds played a crucial role in promoting the stability of the (TTF.+)2 dimer and thus the noncovalent macrocyclization of the two backbones in both uni‐ and bimolecular manners.  相似文献   

11.
Organic spin-based molecular materials are considered to be attractive for the generation of functional materials with emergent optoelectronic, magnetic, or magneto-conductive properties. However, the major limitations to the utilization of organic spin-based systems are their high reactivity, instability, and propensity for dimerization. Herein, we report the synthesis, characterization, and magnetic and electronic studies of three ambient stable radical ions ( 1 a.+ , 1 b.+ , and 1 c.+ ). The radical ions 1 b.+ and 1 c.+ with BPh4 and BF4 counter anions, respectively, were synthesized in excellent yields by means of anion metathesis of 1 a.+ with Br as its counter anion. Notably, synthesis of 1 a.+ was achieved in an ecofriendly, solvent-free protocol. The radical ions were characterized by means of single-crystal X-ray diffraction studies, which revealed the discrete nature of the radical ions and extensive hydrogen-bonding interactions within the radical ions and with the counter anions. Thus, radical ions can be organized to form infinite supramolecular arrays using weak noncovalent interactions. In addition, the Br, BF4, and BPh4 anions formed diverse types of anion–π interactions with the naphthalene and imide rings of the radical ions. The radical ions were characterized by means of X-band electron paramagnetic resonance (EPR) spectroscopy in solution and in the solid state. Magnetic studies revealed their paramagnetic nature in the range of 10 to 300 K. The radical ions exhibited high resistivity approaching the gigaohm (GΩ) scale. In addition, the radical ions exhibited panchromism.  相似文献   

12.
Together with high conductivity, high flexibility is an important property required for next generation organic electronic components. Both properties are difficult to achieve together especially when the components are crystalline because of the intrinsic high brittleness of organic molecular crystals. We report an organic radical crystal system that has both high flexibility and high conductivity. The crystal consists of 9,10‐bis(phenylethynyl)anthracene radical cation ( BPEA.+ ) units, and shows flexibility under pressure with high conductivity in ambient condition exhibiting average conductivity of 2.68 S cm?1 when normal linear shape, as well as 2.43 S cm?1 when bent. The structural analysis reveals that both a short π–π distance (3.290 Å) between BPEA.+ units that are aligned along the crystal length direction, and the presence of PF6? counter ions induce flexibility and high electrical conductivity.  相似文献   

13.
The radical anions and the radical cations of dipleiadiene (dicyclohepta[de,ij]naphthalene; 1 ) and its 12b, 12c-homo derivative 2 were characterized by ESR and ENDOR spectroscopy. Their singly occupied orbitals are related to the degenerate nonbonding MOs of a 16-membered π-perimeter. The π-spin distribution over the perimeter is similar in the radical cations 1 .+ and 2 .+, and an analogous statement holds for the radical anions 1 .? and 2 .?. However, deviations of the π-system from planarity lead to a decrease in the absolute values of the negative coupling constants of the perimeter protons in 2 .+ and 2 .? relative to those in 1 .+ and 1 .?. The hyperfine data for the perimeter protons in the radical ions correlate with the changes in 13C chemical shifts on passing from the neutral compounds to the corresponding diions. It is concluded from the coupling constants of the CH2 protons in the radical ions of 2 that the cation 2 .+ exists in the methano-bridged form ( A ) of the neutral 2 (and, presumably, also of the dication 2 2+), whereas the anion 2 .? adopts the bisnorcaradiene form ( B ) of the dianion 2 2?.  相似文献   

14.
The oxidation of elemental sulfur in superacidic solutions and melts is one of the oldest topics in inorganic main group chemistry. Thus far, only three homopolyatomic sulfur cations ([S4]2+, [S8]2+, and [S19]2+) have been characterized crystallographically although ESR investigations have given evidence for the presence of at least two additional homopolyatomic sulfur radical cations in solution. Herein, the crystal structure of the hitherto unknown homopolyatomic sulfur radical cation [S8].+ is presented. The radical cation [S8].+ represents the first step of the oxidation of the S8 molecule present in elemental sulfur. It has a structure similar to the known structure of [S8]2+, but the transannular sulfur⋅⋅⋅sulfur contact is significantly elongated. Quantum-chemical calculations help in understanding its structure and support its presence in solution as a stable compound. The existence of [S8].+ is also in accord with previous ESR investigations.  相似文献   

15.
Radical cations of bis(triarylamine)s, 3 and 4 , in which the triarylamine redox centers are bridged by an ortho ‐phenylene and ortho ‐carborane cluster, respectively, have been prepared to elucidate the difference in intramolecular charge/spin‐transfer (ICT/IST) pathway owing to the two different bridging units affording similar geometrical arrangements between the redox centers. Electrochemistry, absorption spectroscopy, VT‐ESR spectroscopy, and DFT calculations reveal that 3 .+ and 4 .+ are classified into class II and class I mixed‐valence systems, respectively, and therefore, through‐bond and through‐space mechanisms are dominant for the ICT/IST phenomena in 3 .+ and 4 .+, respectively. Moreover, SQUID measurements for dicationic species provide the fact that virtually no spin‐exchange interaction is observed for spins in 4 2+, while the antiferromagnetic interaction for spins in 3 2+, in accordance with the existence of a conjugation pathway for the ortho ‐phenylene bridge.  相似文献   

16.
The addition of 1 and 2 molar equivalents of bromine to a series of 10-alkylphenothiazines, 1a-d (methyl, ethyl, n-propyl, and isopropyl, respectively), yields the corresponding 3-bromo- and 3,7-dibromo-10-alkylphenothiazines ( 11a-d and 12a-d , respectively). Evidence which supports the typical clectrophilic aromatic substitution mechanism is presented. Radical cations ( 12a-d.+ ) arc produced when 12a-d are treated with 1 or 2 molar equivalents of bromine. Upon boiling in acetic acid these radical cations are converted predominantly to 1,3,7,9-lelrabromophenothiazine ( 5 ) and the parent 3,7-dibromo-10-alkylphenothiazine ( 12a-d ) with the evolution of hydrogen bromide. The 10-methyl radical ( 12a ) gives, in addition, 1,3,7-tribromo-10-methylphenothiazine ( 15 ). A mechanism if proposed for these reactions in which initial dealkylution of 12b-d.+ to 3.7-dibromophenothiazine radical cation ( 13 ) occurs followed by reduction of 13.+ by bromide ion to parent 3,7-dibromophenothiazine ( 13 ). Subsequent bromination of 13 by molecular bromine produced in the previous redox reaction yields 1,3,7-tribromo-( 14 ) and 1,3,7,9-tetra-bromo-( 5 ) phenothiazines. The small size of the methyl group allows 12a to be brominated at the 1-position prior to dealkylation. In addition to undergoing bromination at the 3- and 7-position, 10-isopropylphenothiazine ( 1d ) is oxidized to the radical cation 12e.+ when treated with bromine. 10-Benzylphenothiazine ( 1e ), however, undergoes oxidation to radical cation 1e.+ exclusively. This radical cation debenzylates readily at room temperature and is converted finally into phenothiazine.  相似文献   

17.
Thermal transformations of vinylcyclopropane radical cations (VCP.+) in X-ray-irradiated frozen Freon matrices (CFCl2CF2Cl and CFCl3) were studied by ESR; radical processes involving VCP.+ in solid VCP were simulated.Gauche- andanti-VCP .+ were found to be the primary radical cations, however, the former, unlike the latter, is stable only under gas-phase conditions. The thermodynamic equilibrium betweenanti-VCP.+ and its less stable distonic form,dist(90,0)-C 5H8 .+, is established in frozen Freon matrices and the VCP host matrix; the structure of dist(90,0)****-C 5H8 .+ is stabilized by a molecule ofanti-VCP. In CFC3, along with dist(90,0)-C5H8 .+,-dimeric resonance [anti-VCP]2 .+ complex was detected. A general scheme of the transformations of VCP.+ in the solid phase has been proposed.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 1, pp. 11–21, January, 1994.  相似文献   

18.
Although dimer radical ions of aromatic molecules in the liquid-solution phase have been intensely studied, the understanding of charge-localized dimers, in which the extra charge is localized in a single monomer unit instead of being shared between two monomer units, is still elusive. In this study, the formation of a charge-localized dimer radical cation of 2-ethyl-9,10-dimethoxyanthracene (DMA), (DMA)2.+ is investigated by transient absorption (TA) and time-resolved resonance Raman (TR3) spectroscopic methods combined with a pulse radiolysis technique. Visible- and near-IR TA signals in highly concentrated DMA solutions supported the formation of non-covalent (DMA)2.+ by association of DMA and DMA.+. TR3 spectra obtained from 30 ns to 300 μs time delays showed that the major bands are quite similar to those of DMA except for small transient bands, even at 30 ns time delay, suggesting that the positive charge of non-covalent (DMA)2.+ is localized in a single monomer unit. From DFT calculations for (DMA)2.+, our TR3 spectra showed the best agreement with the calculated Raman spectrum of charge-localized edge-to-face T-shaped (DMA)2.+, termed DT.+, although the charge-delocalized asymmetric π-stacked face-to-face (DMA)2.+, termed DF3.+, is the most stable structure of (DMA)2.+ according to the energetics from DFT calculations. The calculated potential energy curves for the association between DMA.+ and DMA showed that DT.+ is likely to be efficiently formed and contribute significantly to the TR3 spectra as a result of the permanent charge-induced Coulombic interactions and a dynamic equilibrium between charge localized and delocalized structures.  相似文献   

19.
One‐electron oxidation of the stibines Aryl3Sb ( 1 , Aryl=2,6‐i Pr2‐4‐OMe‐C6H2; 2 , Aryl=2,4,6‐i Pr3‐C6H2) with AgSbF6 and NaBArylF4 (ArylF=3,5‐(CF3)2C6H3) afforded the first structurally characterized examples of antimony‐centered radical cations 1 .+[BArylF4] and 2 .+[BArylF4]. Their molecular and electronic structures were investigated by single‐crystal X‐ray diffraction, electron paramagnetic resonance spectroscopy (EPR) and UV/Vis absorption spectroscopy, in conjunction with theoretical calculations. Moreover, their reactivity was investigated. The reaction of 2 .+[BArylF4] and p ‐benzoquinone afforded a dinuclear antimony dication salt 3 2+[BArylF4]2, which was characterized by NMR spectroscopy and X‐ray diffraction analysis. The formation of the dication 3 2+ further confirms that the isolated stibine radical cations are antimony‐centered.  相似文献   

20.
One‐electron oxidation of the stibines Aryl3Sb ( 1 , Aryl=2,6‐i Pr2‐4‐OMe‐C6H2; 2 , Aryl=2,4,6‐i Pr3‐C6H2) with AgSbF6 and NaBArylF4 (ArylF=3,5‐(CF3)2C6H3) afforded the first structurally characterized examples of antimony‐centered radical cations 1 .+[BArylF4] and 2 .+[BArylF4]. Their molecular and electronic structures were investigated by single‐crystal X‐ray diffraction, electron paramagnetic resonance spectroscopy (EPR) and UV/Vis absorption spectroscopy, in conjunction with theoretical calculations. Moreover, their reactivity was investigated. The reaction of 2 .+[BArylF4] and p ‐benzoquinone afforded a dinuclear antimony dication salt 3 2+[BArylF4]2, which was characterized by NMR spectroscopy and X‐ray diffraction analysis. The formation of the dication 3 2+ further confirms that the isolated stibine radical cations are antimony‐centered.  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