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.
The phenol‐tailed porphyrin ligand, H3L was synthesized as a model compound for catalases. H3L and its corresponding iron complex [Fe(L)] were synthesized by using the precursor, 5‐(8‐ethoxycarbonyl‐1‐naphthyl)‐10, 15, 20‐triphenyl porphyrin (ENTPP). They were characterized by 1H NMR spectroscopy, mass spectrometry, X‐ray crystallography, and cyclic voltammetry. All the results have confirmed that the phenol group is covalently attached to the porphyrin. In the iron complex, phenolate oxygen is coordinated to iron(III) as the fifth ligand, leading to the five‐coordinate high‐spin iron(III) species.  相似文献   

2.
The structural parameters, energies, and spectroscopic characteristics of singlet and triplet titanium porphyrin π-complexes Ti(P)(π-L) (P = C20H12N4) with the axial ligands L = C2H2, C2H4, N2H2, HCN, C6H6, N2, and C60 coordinated to the Ti atom through C-C, C-N, and N-N multiple bonds have been calculated by the density functional theory B3LYP method. The changes in the calculated properties of the π-complexes as compared with the properties of the isolated (uncoordinated) Ti(P) and L molecules have been examined. The activation of multiple bonds on coordination to the titanium atom is manifested in (i) their sharp weakening and elongation by 0.10–0.20 Å or more, (ii) a long-wavelength shift of their stretching modes by 300-500 cm-1 or more, (iii) considerable electron density transfer from the porphyrin ring (P ring) to the ligand and the corresponding ligand distortion and polarization, (iv) a strong displacement (0.5-6 Å) of the Ti atom from the P ring plane toward the π-ligand and the dome distortion of the P ring. For the Ti(P)(π-L) systems, the addition of the second axial π-ligand to form six-coordinate ππ-complexes is not typical. In the Ti(P)(π-L)2 with identical ligands, the Ti atom is strongly displaced out of the P ring plane toward one of the ligands and the second ligand is repulsed from the P ring and actually removed from the metal coordination sphere. In the Ti(P)(π-L) (π-L’) complexes with different ligands, according to the relative strength of the Ti–L and Ti-L’ bonds, which decreases in the series N2H2 > C2H2 > HCN > C60 > C2H4 > C6H6 > N2, the weaker ligand is forced out by the stronger ligand (acetylene is pushed out of the coordination sphere by diimine; ethylene, by acetylene and fullerene; fullerene, by hydrogen cyanide; etc.). In mixed πσ-complexes Ti(P)(π-L)(CO) in the singlet state, acetylene pushes out the CO group; conversely, in the triplet state, acetylene is pushed out by carbonyl. There is a trend in the behavior of the activation effects along the series of the above ligands and with a change in the electronic state multiplicity of the complexes.  相似文献   

3.
Carbon-13 NMR studies on a series of high spin iron(III) porphyrins, namely tetraphenylporphyrin iron(III) halides [Fe(TPP) X, X=Cl, Br, I] in CDCl3 solution are reported. As expected the13C shifts are found to be an order of magnitude larger than the corresponding proton shifts. The dipolar contribution, which is quite important for the proton NMR, becomes much less significant for the13C shifts. No systematic variation in the13C shift across the series is observed, except for the meso-carbon which shows a small but gradual decrease in going from the chloro to the iodo complex. The13C shift for the various carbon atoms of the porphyrin ligand shows interesting pattern which is discussed in terms of spin delocalisation mechanisms.  相似文献   

4.
Complex formation in Zn(II) protoporphyrin-IX dimethyl ester has been studied by proton and 13C NMR spectroscopy. The large concentration dependence of the spectra has been studied by the technique of porphyrin/axial ligand titration, which together with selective decoupling and regiospecific deuterium labelling allows the assignment of all the peripheral proton and 13C nuclei in both the monomeric and aggregated species. Titration of the metalloporphyrin with various basic ligands (pyrrolidine, pyridine, lutidine) showed that dissociation of the aggregate was complete for a 1:1 porphyrin/added base ratio. The concentration dependence of the spectra was then analyzed to give the monomer and monomer-dimer shifts for all the assigned nuclei. Analysis of the monomer-dimer shifts in terms of the ring current model gives good agreement with a dimer geometry in which the inter-ring separation is ca. 4.5 Å and there is a smaller lateral displacement of the porphyrin rings. The dimer geometry is such that rings A and B of one porphyrin molecule are situated over rings C and D of the other. These results confirm our earlier suggestions of intermolecular metal-to-porphyrin binding in these metalloporphyrins, and further suggest that charge-transfer interactions may also be present in appropriate cases. The discrepancy between the absolute values of the observed and calculated monomer-dimer shifts, which was formerly attributed to multiple aggregation, is now suggested to be due to ensemble-averaging in the dimer structure.  相似文献   

5.
The 13C and proton NMR spectra of the zinc(II) complexes of the tetramethyl esters of the four coproporphyrin type isomers are reported and assigned. Effects of aggregation phenomena on these shifts are discussed and a method involving addition of a slight excess of pyrrolidine is proposed for measurement of the spectra of the “monomeric” species; spectra obtained under these conditions are capable of simple, straight-forward interpretation and assignment in terms of molecular symmetry. Thus, a facile distinction between the type isomers is obtained.The “monomer” chemical shifts so derived allow consistent SCS parameters to be derived. The Cβ-Me SCS are shown to be related to the bond order of the Cβ-Cβ bond in the porphyrin ring, and are thus quite different from the corresponding SCS in pyrroles.Aggregation shifts in the 13C and proton spectra are shown to be consistent with the presence of “stacked” aggregates with the ring current of one molecule affecting the other, together with an additional effect on the chemical shifts of the meso carbons, which is probably steric in origin.  相似文献   

6.
In the first bis­[ruthenium(II)–porphyrin]–dicarbene complex, μ‐[1,4‐phenyl­ene­bis(phenyl­methyl­idene‐κC)]bis­[(ethanol‐κO)(5,10,15,20‐tetra‐p‐tolyl­porphyrinato‐κ4N)ruthenium(II)] 1,2‐di­chloro­ethane trisolvate, [Ru2(C20H14)(C48H36N4)2(C2H6O)2]·3C2H4Cl2, an inversion center is located at the center of the μ‐phenyl­ene group, leading to a parallel arrangement for the pair of porphyrin ring systems. The bond lengths and angles compare favourably with literature values for ruthenium–porphyrin–monocarbene complexes; the Ru=C(carbene) bond length and the C(phenyl)—C(carbene)—C(phenyl­ene) angle are 1.865 (3) Å and 112.3 (3)°, respectively. The RuII ion is displaced out of the C20N4 porphyrin least‐squares plane (by 0.2373 Å) toward the bridging ligand of the Ci‐symmetry dimer. The porphyrin ring systems of the dimer thus exhibit mildly domed conformations.  相似文献   

7.
The chelate (Cl)AlP was prepared by complexation of porphine (P) with aluminum(III) chloride in refluxing pyridine. Equilibrium coordination of nitrogen-containing ligands (L = 2-methylimidazole, imidazole, pyridine, 3,5-dimethylpyrazole, dimethylformamide) with (Cl)AlP in benzene was studied by spectrophotometric titration and computer simulation. Quantitative and qualitative characteristics of the reaction were obtained. The structure of the mixed-ligand complex formed by intermolecular interaction of the metal porphyrin with a base was determined spectrophotometrically and by quantum-chemical calculations. An effect of additional molecular ligand and of steric strain in the macroring on the stability of the complex was noted. The stability constant (K s) increases with an increase in the basicity (K BH +) of the extra ligand and is proportional to the shift of the main bands (?λ) in the electronic absorption spectra. The geometric and energy characteristics of hexacoordinated aluminum porphyrin were calculated by the PM3 method. Correlations were found between the calculated energy of the interaction of the aluminum atom with the base molecule (E b) and stability of the mixed-ligand complexes (Cl)Al(L)P. The cis and trans effects in the complexes (Cl)Al(L)P were analyzed. The dependence of the strength of the Al-L bond on the nature of the porphyrin and the basicity of the additional molecular ligand was determined from the experimental data and calculation results.  相似文献   

8.
在溶剂热条件下,由2-(4′-羧基苯基)咪唑-4,5-二羧酸(H_4L,C_(12)H_8N_2O_6),合成了4个配位聚合物{[M(H_3L)_2]·2H_2O}_n(M=Zn(1),Cd(2),Co(3)),[Cd(H_2L)(H_2O)]_n(4)。用元素分析、红外光谱、热重分析和单晶X射线衍射对配合物进行了表征和结构分析。结构分析结果表明:1~3是异质同晶。配体失去1个质子以H_3L~-的形式通过单齿和N,O-双齿螯合的配位模式与中心金属离子配位,构成一个略有变形的八面体结构。对于配合物4来说,配体失去2个质子以H_2L~(2-)的形式分别通过单齿和N,O-双齿螯合的配位方式与Cd~(2+)配位,中心离子采取扭曲的七配位五角双锥配位模式,并且通过配体苯环上羧基氧原子的双齿桥联作用连接2个中心离子,形成四元环的双核结构;同时呈现双节点(3,6)-连接的二维拓扑网络(4.4.4)(4.4.4.4.4.4.5.6.6.6.6.6)。测定了产物的固体荧光光谱;用EtBr荧光探针法研究了配体及配合物与ct-DNA的相互作用。  相似文献   

9.
Cobalt(III) tetraphenylporphyrin chloride (TPPCoCl) was experimentally proved to be an active catalyst for poly(propylene carbonate) production. It was chosen as a model catalyst in the present work to investigate the initiation step of propylene oxide (PO)/CO2 copolymerization, which is supposed to be the ring opening of the epoxide. Ring‐opening intermediates ( 1 – 7 ) were detected by using 1H NMR spectroscopy. A first‐order reaction in TPPCoCl was determined. A combination of monometallic and bimetallic ring‐opening pathways is proposed according to kinetics experiments. Addition of onium salts (e.g., bis(triphenylphosphine)iminium chloride, PPNCl) efficiently promoted the PO ring‐opening rate. The existence of axial ligand exchange in the cobalt porphyrin complex in the presence of onium salts was suggested by analyzing collected 1H NMR spectra.  相似文献   

10.
《Polyhedron》2001,20(9-10):995-1003
The formation constants and the isotropic ESR parameters (g-factors, 63Cu, 65Cu, 14N hyperfine coupling constants and relaxation parameters) of the various species were determined by the simultaneous analysis of a series of spectra, taken in a circulating system at various pH and ligand-to-metal concentration ratio. For both systems the new [CuLH]2+ complex was identified in acidic solutions. With the glycyl-l-serine ligand below pH 11.5 the same complexes and coordination modes are formed than with simple dipeptides. The side-chain donor group is bound only over pH 11.5 in the complex [CuLH−2(OH)]2−, where it is deprotonated and substitutes the carboxylate O in the third equatorial site. For the bis complex [CuLH−1(L)] an isomeric equilibrium was shown, where the difference between the isomers was based on which of the donor atoms of the ‘L’ ligand, the peptide O or the amino N, occupies the fourth equatorial position, and which one is coordinated axially. The l-seryl-glycine ligand forms the same species as simple dipeptides and glycyl-l-serine up to pH 8. The only difference is that the axial binding of the alcoholic OH group fairly stabilizes the bidentate equatorial coordination of the ‘L’ ligand through the amino N and peptide O atoms in the [CuL]+ complex as well as in the major isomer of the [CuLH−1(L)] complex. For this system we showed that (1) proton loss and the equatorial coordination of the alcoholic OH group occurs at relatively low pH (over pH 8–9), which results in the [CuL2H−2]2− complex with excess ligand, and also the newly identified species [Cu2L2H−4]2−: (2) this process is in competition with the proton loss of a coordinated water molecule. For both systems, the ESR-inactive species [Cu2L2H−3] was also shown.  相似文献   

11.
A novel nickel(II) complex [Ni(L)2Cl]Cl with a bidentate phosphinopyridyl ligand 6‐((diphenylphosphino)methyl)pyridin‐2‐amine (L) was synthesized as a metal‐complex catalyst for hydrogen production from protons. The ligand can stabilize a low Ni oxidation state and has an amine base as a proton transfer site. The X‐ray structure analysis revealed a distorted square‐pyramidal NiII complex with two bidentate L ligands in a trans arrangement in the equatorial plane and a chloride anion at the apex. Electrochemical measurements with the NiII complex in MeCN indicate a higher rate of hydrogen production under weak acid conditions using acetic acid as the proton source. The catalytic current increases with the stepwise addition of protons, and the turnover frequency is 8400 s?1 in 0.1 m [NBu4][ClO4]/MeCN in the presence of acetic acid (290 equiv) at an overpotential of circa 590 mV.  相似文献   

12.
The preparation, isolation and characterization by proton and 13C NMR of the four possible atropisomers of meso-tetra(2-methoxy-1-naphthyl)porphyrin is described. Chemical shift differences due to atropisomerization effects are observed in the porphyrin and naphthyl rings. Comparison of the naphthyl chemical shifts with those of the model compound 1-isopropenyl-2-methoxynaphthalene allows the shifts due to the porphyrin ring current to be isolated. The observed Δδ values of the naphthyl protons agree well with those calculated from the previously described porphyrin ring current model, and allow both the angle of tilt of the naphthyl ring and the dihedral angle of the 2-methoxy substituent to be estimated. In contrast, the Δδ values for the naphthyl carbons bear no relationship to the calculated ring current shifts. Calculations of the total ring current contribution (porphyrin plus naphthyl rings) at the different naphthyl rings of the unsymmetric type III isomer show that at least part of the observed atropisomerism effects are due to the long-range current shifts of the naphthyl rings. The results also provide a clear demonstration of the identity of the porphyrin ring current in the free base and porphyrin dication.  相似文献   

13.
In the title compound, [RhI(C44H28N4)(C5H5N)]·C7H8, the porphyrin ring experiences significant distortion from planarity (a saddle conformation with a superimposed ruffling), as a result of steric interactions with the 2,6‐H atoms of the axial pyridine ligand. This also leads to a slight lengthening of the Rh–pyridine bond [Rh—N 2.102 (7) Å] relative to those seen in other pyridine adducts of six‐coordinate RhIII. The metric parameters of the porphyrin core are comparable with those of related metalloporphyrin derivatives. No significant intermolecular interactions are observed between the metalloporphyrin and disordered solvate species.  相似文献   

14.
A new coordination mode for the porphyrin ligand is found in [B2O2(BCl3)2(tpClpp)] (tpClpp=dianion of 5,10,15,20-tetra-p-chlorophenylporphyrin; the p-chlorophenyl groups are omitted for clarity in the picture shown on the right). This complex contains a four-membered B2O2 ring in the cavity of the ligand. The two boron atoms are coplanar with the porphyrin molecule, which undergoes an elongation along the B⋅⋅⋅B axis to accomodate the unusual guest.  相似文献   

15.
In order to gain insight into the influence of the H+-accepting terminal ligand in high-valent oxidant mediated proton coupled electron transfer (PCET) reactions, the reactivity of a high valent nickel–fluoride complex [NiIII(F)(L)] ( 2 , L=N,N’-(2,6-dimethylphenyl)-2,6-pyridinecarboxamidate) with substituted phenols was explored. Analysis of kinetic data from these reactions (Evans–Polanyi, Hammett, and Marcus plots, and KIE measurements) and the formed products show that 2 reacted with electron rich phenols through a hydrogen atom transfer (HAT, or concerted PCET) mechanism and with electron poor phenols through a stepwise proton transfer/electron transfer (PT/ET) reaction mechanism. The analogous complexes [NiIII(Z)(L)] (Z=Cl, OCO2H, O2CCH3, ONO2) reacted with all phenols through a HAT mechanism. We explore the reason for a change in mechanism with the highly basic fluoride ligand in 2 . Complex 2 was also found to react one to two orders of magnitude faster than the corresponding analogous [NiIII(Z)(L)] complexes. This was ascribed to a high bond dissociation free energy value associated with H−F (135 kcal mol−1), which is postulated to be the product formed from PCET oxidation by 2 and is believed to be the driving force for the reaction. Our findings show that high-valent metal–fluoride complexes represent a class of highly reactive PCET oxidants.  相似文献   

16.
The reactions of pyrimidine‐phosphine ligand N‐[(diphenylphosphino)methyl]‐2‐pyrimidinamine ( L ) with various metal salts of PtII, PdII and CuI provide three new halide metal complexes, Pt2Cl4(μ‐L)2·2CH2Cl2 ( 1 ), Pd2Cl4(μ‐L)2 ( 2 ), and [Cu2(μ‐I)2L2]n ( 3 ). Single crystal X‐ray diffraction studies show that complexes 1 and 2 display a similar bimetallic twelve‐membered ring structure, while complex 3 consists of one‐dimensional polymeric chains, which are further connected into a 2‐D supramolecular framework through hydrogen bonds. In the binuclear complexes 1 and 2 , the ligand L serves as a bridge with the N and P as coordination atoms, but in the polymeric complex 3 , both bridging and chelating modes are adopted by the ligand. The spectroscopic properties of complexes 1 ‐ 3 as well as L have been investigated, in which complex 3 exhibits intense photoluminescence originating from intraligand charge transfer (ILCT) π→π* and metal‐to‐ligand charge‐transfer (MLCT) excited states both in acetonitrile solution and solid state, respectively.  相似文献   

17.
A new Schiff base 4-[N-hydroxyethyl-N-(methyl)amino]benzaldehyde S-methyl dithiocarbazate (HL, where H is a dissociable proton) and the ruthenium complex [Ru(bpy)2L]PF6 (bpy = 2,2′-bipyridine) have been synthesized. The structural determinations of the ligand and its ruthenium complex, by X-ray crystallography, show that the ligand is coordinated as a monoanionic bidentate N, S-donor, forming a four-member chelate ring with a bite angle of 65.91°. The complex shows intense MLCT transitions in the visible region. Fluorescent and electrochemical properties have been also studied. The complex in DMF solution exhibited a strong two-photon absorption (t.p.a.) at 532 nm nanosecond laser pulses. The t.p.a. coefficient β, t.p.a. cross-section σ and the third-order optical nonlinearity χ(3) of the complex and the ligand have been determined by the Z-scan technique.  相似文献   

18.
This study targets the construction of porphyrin assemblies directed by halogen bonds, by utilizing a series of purposely synthesized Sn(axial ligand)2–(5,10,15,20‐tetraarylporphyrin) [Sn(L)2‐TArP] complexes as building units. The porphyrin moiety and the axial ligands in these compounds contain different combinations of complimentary molecular recognition functions. The former bears p‐iodophenyl, p‐bromophenyl, 4′‐pyridyl, or 3′‐pyridyl substituents at the meso positions of the porphyrin ring. The latter comprises either a carboxylate or hydroxy anchor for attachment to the porphyrin‐inserted tin ion and a pyridyl‐, benzotriazole‐, or halophenyl‐type aromatic residue as the potential binding site. The various complexes were structurally analyzed by single‐crystal X‐ray diffraction, accompanied by computational modeling evaluations. Halogen‐bonding interactions between the lateral aryl substituents of one unit of the porphyrin complex and the axial ligands of neighboring moieties was successfully expressed in several of the resulting samples. Their occurrence is affected by structural (for example, specific geometry of the six‐coordinate complexes) and electronic effects (for example, charge densities and electrostatic potentials). The shortest intermolecular I???N halogen‐bonding distance of 2.991 Å was observed between iodophenyl (porphyrin) and benzotriazole (axial ligand) moieties. Manifestation of halogen bonds in these relatively bulky compounds without further activation of the halophenyl donor groups by electron‐withdrawing substituents is particularly remarkable.  相似文献   

19.
(Di­phenyl phosphite‐κO)(5,10,15,20‐tetra­phenyl­porphyrinato‐κ4N)­manganese(III) hexa­fluoro­antimonate(V), [Mn(C44H28N4)(C12H11O3P)](SbF6), is the first example of a structurally characterized di­aryl or di­alkyl phosphite complex of a metal–porphyrin ion. The axial phosphite ligand binds to the MnIII ion via the P=O O atom, affording a nominally five‐coordinate complex with an Mn—O distance of 2.120 (4) Å. The mean porphyrin Mn—N distance is 2.000 (4) Å and the MnIII ion is displaced from the 24‐atom porphyrin mean plane by 0.1548 (13) Å towards the axial O atom. The porphyrin adopts a marked saddle conformation, with a small domed component. The saddle distortion of the porphyrin ligand reflects the tight back‐to‐back dimers formed in the lattice by pairs of neighboring cations. The `non‐covalent' dimers in the lattice exhibit an unusual (weak) η2‐type coordination of a pyrrole C=C bond from a neighboring mol­ecule, with MnIII⃛C distances of 3.697 (5) and 3.537 (5) Å.  相似文献   

20.
In this paper, the interaction between squarylium cyanine and porphyrin in chloroform is investigated by absorption and fluorescence spectroscopy. Emphasis has been put on the mechanism of intermolecular energy transfer. The overlap integral J between the absorption spectrum of squarylium cyanine and the fluorescence spectrum of porphyrin was calculated, which reveals that the singlet-singlet energy transfer may occur from porphyrin to squarylium cyanine in solution. In comparison of the observed rate constant [kqII=6.1 ×1013 (mol/L)-1·s-1] for fluorescence quenching of porphyrin by squarylium cyanine with the diffusion rate constant in chloroform [kdif=1.1×1010 (mol/L)-1·s-1] and the rate of energy transfer [ket≤6.7×104 (mol/L)-1·s-1 in the experimentally dilute solutions] estimated from Forster formula, the possibility of energy transfer by electron exchange or/and coulombic mechanism could be excluded. So it has been definitely convinced that the intermolecuiar energy transfer between them is  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