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.
New 3,5-bis(2-pyridyl)pyrazolato (bpypz) bridged heterometal dinuclear complexes [(nta)Cr(μ-bpypz)MII(picen)]+ (M = Mn(II), Ni(II)) and [(acac)2Cr(μ-bpypz)NiII(picen)]2+ (nta = nitrilotriacetate, picen = N,N′-bis(2-pyridylmethyl)ethylenediamine, acac = acetylacetate) were synthesized and characterized by the X-ray analysis, ESI-MS and the magnetic measurements, and/or 2H NMR spectra. The molecular structures were compared from a viewpoint of the conformation of the picen depending on MII ionic radii or different modes of hydrogen bonds. The picen in [(nta)Cr(μ-bpypz)MnII(picen)]BF4 takes an abnormal conformation with intramolecular bifurcated three-center hydrogen bonds between two carboxylate oxygens of nta and an amine proton of the picen as found for the previously reported corresponding Fe(II) complex [K. Ni-iya, A. Fuyuhiro, T. Yagi, S. Nasu, K. Kuzushita, S. Morimoto, S. Kaizaki, Bull. Chem. Soc. Jpn. 74 (2001) 1891]. On the other hand, for both Ni(II)-nta and Ni(II)-acac complexes, the picen takes a normal conformation with only a two-center hydrogen bond between non-bridging ligands. The magneto-structural relation is discussed for the Cr(III)–Ni(II) complexes in connection with the orthogonality or orbital overlap arising from the difference in distortion around Cr(III) moiety.  相似文献   

2.
The linkage isomers [Re(NCS)6]2? and [Re(NCS)5(SCN)]2? are obtained by the reaction of [ReBr6]2? with NCS? in dimethylformamide. Some differences in the chemical behavior allowed their separation and structural characterization in the form of (NBu4)2[Re(NCS)6] (1) and [Zn(NO3)(Me2phen)2]2[Re(NCS)5(SCN)] (2), respectively (Bu = n-C4H9 and Me2phen = 2,9-dimethyl-1,10-phenanthroline).  相似文献   

3.
《Inorganica chimica acta》2006,359(9):2879-2887
An investigation of the question of “Is it homogeneous or heterogeneous catalysis?” is reported when using PtII(1,5-COD)X2 (X = halogen, alkyl) precatalysts for the hydrogenation of olefins. Using product studies, kinetic evidence, and Hg0 poisoning experiments, it is shown that PtII(1,5-COD)Cl2 is a precatalyst and must be reduced to Pt0 nanoclusters and bulk metal as the true hydrogenation catalyst. An investigation of the related complex PtII(1,5-COD)(CH3)2 reveals that this complex does not form a hydrogenation catalyst by itself under H2, in agreement with the literature. Kinetic and Hg0 poisoning evidence confirms that PtII(1,5-COD)(CH3)2, too, forms a Pt0 heterogeneous catalyst if other metals (Ir0, Pt0) are used as seeds to initiate the reduction of PtII. A short review of the use of PtII(1,5-COD)Cl2 in hydrosilylation reactions is given, illustrating the continued controversy surrounding the nature of the true catalyst in that literature system.  相似文献   

4.
Two dinuclear metal complexes, [Co2(bhmp)(MeCO2)2]ClO4 · 2H2O (1) and [Ni2(bhmp)(MeCO2)2]ClO4 · 2H2O (2), were synthesized with a dinucleating ligand, 2,6-bis[bis(2-hydroxyethyl)aminomethyl]-4-methylphenol [H(bhmp)]. Both complexes were easily soluble in water as well as in DMF. Electronic spectra for both complexes were measured in both solvents and analyzed using the angular overlap model (AOM). From the electronic spectra and molar conductance, both complexes were determined to exist as [M2(bhmp)(MeCO2)2]+ (M = CoII or NiII) in DMF, dissociating perchlorate ions. On the other hand, in water, it was concluded that the acetate ions were partially dissociated and each complex existed as a mixture of some dissociated species, such as [M2(bhmp)(MeCO2)(H2O)2]2+ and [M2(bhmp)(H2O)4]3+ (M = CoII or NiII). Such dissociation was also confirmed by precipitation of the dissociated species when NaBPh4 was added into an aqueous solution of the nickel complex.  相似文献   

5.
Ionic triads formed by [NiII(bipy)3]2+ (bipy = 2,2′-bipyridyl) and diamagnetic [MIV(CN)8]4? (M = Mo and W) were prepared and structurally characterized. The two compounds are isostructural and their structure consists of a three-dimensional hydrogen-bonded framework where cation–anion interactions occur through short contacts M–CN?H–C(bipy). Before irradiation, the Mo analogue behaves as paramagnet with small intermolecular interactions between the [NiII(bipy)3]2+ cations. Upon irradiation with visible light, it exhibits a reversible photomagnetic effect, which is interpreted in terms of the formation of paramagnetic [MoV(CN)8]3? and [NiII(bipy)2(bipy?)]+ due to the outer-sphere electron transfer.  相似文献   

6.
We review an extensive body of single-crystal high-frequency electron paramagnetic resonance (HFEPR) data in order to determine the transverse spin Hamiltonian parameters that control the tunneling of the direction of magnetization in a variety of integer and half-integer-spin single-molecule magnets (SMMs). The SMMs studied are members of the following families: S = 9/2 [Mn4O3Cl]6+; S = 5 [Mn3NiO4]6+; S = 6 [Mn3ZnO4]6+; and S = 4 [Ni4(OR)4]4+. HFEPR spectra for the half-integer S = 9/2 Mn4 complexes that have C3 symmetry do not provide measurable evidence for transverse spin Hamiltonian terms. This finding is consistent with the relatively large coercive field seen in the magnetization hysteresis loops for these complexes. On the other hand, a low symmetry S = 9/2 complex exhibits a much faster rate of ground-state magnetization tunneling, in agreement with HFEPR spectra for a powder sample that gives a rhombic zero-field splitting (ZFS) parameter of E = 0.140 cm?1. The S = 5 Mn3Ni systems exhibit magnetization tunneling that is much faster than seen for the high-symmetry S = 9/2 Mn4 complexes. This can be attributed to their integer-spin ground states. Like the C3 symmetry Mn4 SMMs, the HFEPR spectra for high-symmetry Mn3Ni complexes do not provide measurable evidence for transverse ZFS terms. However, the spectra exhibit broad peaks, suggesting distributions in the local molecular environments brought about by disordered solvate molecules. This disorder likely explains the fast tunneling in the high-symmetry S = 5 Mn3Ni systems, though one cannot rule out fourth- (and higher-) order interactions that cannot be detected by HFEPR due to the broad resonances. The one S = 6 Mn3Zn complex shows an even faster rate of tunneling compared to the isostructural S = 5 Mn3Ni complex. Finally, the S = 4 [Ni(hmp)(dmb)Cl]4 complex provides unique insights into the origin of fourth- (and higher-) order interactions found for many SMMs on the basis of analysis using a giant spin Hamiltonian (GSH) approximation. We conclude that the fourth-order anisotropy found for the S = 4 ground state of [Ni(hmp)(dmb)Cl]4 originates from the second-order ZFS interactions associated with the individual NiII ions, but only as a result of higher-order processes that occur via S-mixing between the ground state and higher-lying (S < 4) spin-multiplets. The S-mixing is relatively strong in this system because of comparable exchange and anisotropy energy scales. The relatively fast tunneling is a direct consequence of this S-mixing, as opposed to any intrinsic fourth-order (spin–orbit) anisotropy associated with NiII.  相似文献   

7.
《Inorganica chimica acta》2006,359(8):2407-2416
The trimethyltin(IV) polymer [(Me3Sn)2(nip) · EtOH]n (1) of 5-nitroisophthalic acid (H2nip) and its three derivatives with mixed organic N-donor ligands 2,2′-bipy [(Me3Sn)2(nip) · 2H2O] · [(Me3Sn)2(nip) · H2O] · 2(2,2′-bipy) (2) 4,4′-bipy {[(Me3Sn)2(nip)]2(4,4′-bipy)}n (3) or phen [(Me3Sn)2(nip) · H2O] · (phen) (4) have been synthesized by the reaction of trimethyltin(IV) chloride and H2nip when sodium ethoxide was added in the presence of 2,2′-bipy 4,4′-bipy or phen. All complexes 14 were characterized by elemental, IR, 1H, 13C, and 119Sn NMR spectroscopy and X-ray crystallography analyses. Crystal, data collection and structure refinement parameters for complexes 1, 2, 3 and 4 are shown in Table 1, Table 2, respectively. The X-ray data showed the geometries of all the tin atoms in complexes 14 are trigonal bipyramidal. The X-ray analysis of 1 showed that the structure was a polymeric infinite chain with neighboring triorganotin centers being linked by dicarboxylate ligands and hydrogen bonds were found between adjacent chains. For 2, two different monomers were found, in one monomer, Me3Sn were coordinated to both carboxyl groups of the ligand and water molecules were coordinated to the two Sn(IV) centers. In the other monomer, water molecules were coordinated to only one Sn center. Co-crystallized2,2′-bipy was found in 2 and a 2D supermolecular structure was formed via O–H⋯O and O–H⋯N (N atoms derived from 2,2′-bipy) hydrogen bonds. In 3 however, a 2D polymeric block was formed due to the inversion center of the 4,4′-bipy. For 4, due to the O–H⋯O and O–H⋯N (N atoms derived from phen) hydrogen bonds and intermolecular Sn⋯O bonds, a 2D polymeric structure was formed.  相似文献   

8.
9.
The trinuclear cluster (cyclam)Co[(μ-Cl)U(Me2Pz)4]2 (cyclam = 1,4,8,11-tetraazacyclotetradecane, Me2Pz? = 3,5-dimethylpyrazolate) is synthesized through cleavage of the homoleptic dimer [U(Me2Pz)4]2 by (cyclam)CoCl2. A single crystal X-ray diffraction analysis reveals a linear chloride-bridged structure analogous to that previously reported for (cyclam)M[(μ-Cl)U(Me2Pz)4]2 (M = Ni, Cu, Zn). The magnetic exchange coupling of the CoU2 cluster was probed by analyzing the temperature dependence of its magnetic susceptibility. Comparison of χMT versus T between the CoU2 species and the diamagnetic ZnU2 cluster demonstrates the presence of ferromagnetic coupling between the CoII and UIV centers. We present methods for estimating upper and lower bounds for the exchange interaction energy in such systems and find that for CoU2, the exchange constant, J, lies in the range 15–48 cm?1. Application of these methods to the previously reported NiU2 cluster suggests somewhat weaker ferromagnetic exchange, with J lying in the range 2.8–19 cm?1. AC magnetic susceptibility experiments were not indicative of single-molecule magnet behavior for the CoU2 cluster, although qualitative interpretation of the low-temperature magnetization data suggests the presence of significant zero-field splitting in the ground state.  相似文献   

10.
We report the synthesis and magnetic properties of three hexametallic Mn clusters: [Mn6O2(Et-sao)6(O2C-Naphth)2(EtOH)4(H2O)2] (1) (HO2C-Naphth = 1-naphthoic acid, Et-saoH2 = 2-hydroxyphenylpropanone oxime), [Mn6O2(Et-sao)6(O2C-Anthra)2(EtOH)4(H2O)2] · 0.66EtOH · 0.33H2O (HO2C-Anthra = anthracene-9-carboxylic acid) (2 · 0.66EtOH · 0.33H2O) and [Mn6O2(Et-sao)6(O2CPhCCH)2(EtOH)4(H2O)2] · 1.7EtOH · 0.3H2O (HO2CPhC  CH = 4-ethynylbenzoic acid) (3 · 1.7EtOH · 0.3H2O). Clusters 13 exhibit ferromagnetic exchange between all six MnIII centres resulting in S = 12 ground spin states. Ac magnetic susceptibility and single crystal micro-SQUID measurements on 13 confirm SMM behaviour with barriers to magnetisation reversal of 60.12 (1), 60.10 (2) and 66.79 (3) K.  相似文献   

11.
The mononuclear cations of the general formula [(η6-arene)RuCl(pdpt)]+ (pdpt = 5,6-diphenyl-3-(pyridine-2-yl)-1,2,4-triazine; arene = C6H6 (1); C6H5Me (2); p-PriC6H4Me (3); C6Me6 (4)) have been synthesised from 5,6-diphenyl-3-(pyridine-2-yl)-1,2,4-triazine (pdpt) and the corresponding chloro complexes [(η6-C6H6)Ru(μ-Cl)Cl]2, [(η6-C6H5Me)Ru(μ-Cl)Cl]2, [(η6p-PriC6H4Me)Ru(μ-Cl)Cl]2 and [(η6-C6Me6)Ru(μ-Cl)Cl]2, respectively. The X-ray crystal structure analyses of [1][PF6] · (C6H6)2.5 and [2][PF6] · (CH3CN)2 reveal a typical piano-stool geometry around the metal centre and in the crystal packing a complexed networks of intermolecular interactions.  相似文献   

12.
The synthesis and pharmacological characterisation of (1-methyl-1H-imidazol-2-yl)-methanamine and its derivatives in PtII complexes are described. Six out of eleven new PtII complexes showed a significant cytotoxic effect on NCI-H460 lung cancer cell line with EC50 values between 1.1 and 0.115 mM, determined by MTT assay. Compound Pt-4a showed a particularly more potent cytotoxic effect than the previously described PtII complex with 2,2′-bipyridine, [Pt(bpy)Cl2], with an EC50 value equal to 172.7 μM versus 726.5 μM respectively, and similar potency of cisplatin (EC50 = 78.3 μM) in NCI-H460 cell line. The determination of the intracellular and DNA-bound concentrations of 195Pt, as marker of the presence of the complexes, showed that the cytotoxic compound Pt-4a readily diffused into the cells to a similar extent of cisplatin and directly interacted with the nuclear DNA. Pt-4a induced both p53 and p21Waf expression in NCI-H460 cells similar to cisplatin. A direct comparison of the cytotoxic effect between compound Pt-4a and cisplatin on 12 different cancer cell lines demonstrated that compound Pt-4a was in general less potent than cisplatin, but it had a comparable cytotoxic effect on non-small-cell lung cancer NCI-H460 cells, and the colorectal cancer cells HCT-15 and HCT-116. Altogether, these results suggested that the PtII complex with 1-methyl-1H-imidazol-2-yl)-methanamine (compound Pt-4a), displayed a significant cytotoxic activity in cancer cells. Similarly to cisplatin this compound interacts with nuclear DNA and induces both p53 and p21waf, and thus it represents an interesting starting point for future optimisation of new PtII complexes forming DNA adducts.  相似文献   

13.
《Inorganica chimica acta》2006,359(5):1619-1626
The reaction of 1,4-dimethyl-1,4,7-triazacyclononane (L-Me2) with MnCl2 · 4H2O in acetonitrile gives, in the presence of sodium formate, hydrogen peroxide, triethylamine and KPF6, the dinuclear Mn(III)–Mn(IV) complex cation [(L-Me2)2Mn2(O)2(OOCH)]2+ (1) which crystallises as the hexafluorophosphate salt.The analogous reaction with sodium benzoate, however, yields the dinuclear Mn(III)–Mn(III) complex cation [(L-Me2)2Mn2(O)(OOCC6H5)2]2+ (2), isolated also as the hexafluorophosphate salt.In the case of sodium acetate, both cations, the Mn(III)–Mn(IV) complex [(L-Me2)2Mn2(O)2(OOCCH3)]2+ (3) and the known Mn(III)–Mn(III) complex [(L-Me2)2Mn2(O)(OOCCH3)2]2+ (4) are available, depending upon the molar ratio.The single-crystal X-ray structure analyses show for the green crystals of [1][PF6]1.5[Cl]0.5 · 1.5 H2O and [3][PF6]2 · (CH3)2CO, a Mn–Mn distance of 2.620(2) and 2.628(4) Å, respectively, while for the red-violet crystal of [4][PF6]2, a Mn–Mn distance of 3.1416(8) Å is observed.All four compounds show catalytic activity for the oxidation of isopropanol with hydrogen peroxide in water and in acetonitrile to give acetone in the presence of oxalic or ascorbic acid as co-catalysts.  相似文献   

14.
《Inorganica chimica acta》2006,359(4):1275-1281
Two new complexes of composition [Cu(2-NO2bz)2(3-pyme)2(H2O)2] (1) and/or [Cu{3,5-(NO2)2bz}2(3-pyme)2] (2) (3-pyme = 3-pyridylmethanol, ronicol or 3-pyridylcarbinol, 2-NO2bz = 2-nitrobenzoate and 3,5-(NO2)2bz = 3,5-dinitrobenzoate) have been prepared and studied by elemental analysis, electronic, infrared and EPR spectroscopy, magnetic susceptibility measurements and the structure of both complexes has been solved. Complex (1) shows an unusual molecular type of structure consisting of the [Cu(2-NO2bz)2(3-pyme)2(H2O)2] molecules held together by hydrogen bonds and van der Waals interactions. Complex (2) exhibits a polymeric chain-like structure [Cu{3,5-(NO2)2bz}2(3-pyme)2]n with copper atoms doubly bridged by two 3-pyridylmethanol molecules and the polymeric molecules are held together by van der Waals interactions. Complex (1) exhibits a magnetic moment μeff = 1.84 B.M. at 300 K that remains nearly constant within the temperature region (5–300 K). Further cooling results in lowering the magnetic moment to μeff = 1.82 B.M. at 1.8 K. The magnetic susceptibility temperature dependence obeys Curie–Weiss law with Curie constant of 0.423 cm3 K mol−1 and with Weiss constant of −0.06 K. The magnetic moment of (2) exhibits a small increase with a decrease in the temperature (μeff = 1.80 B.M. at 300 K and μeff = 1.85 B.M. at 1.8 K) with Curie constant of 0.409 cm3 K mol−1 and with Weiss constant of +1.1 K, which can indicate a very weak ferromagnetic interaction between the copper atoms within the chain. Applying the molecular field model resulted in obtaining zJ′ values −0.08 cm−1 for complex (1), and −0.07 cm−1 for complex (2), respectively, that could characterize intermolecular and interchain interactions transmitted through π–π stacking.  相似文献   

15.
Bacterial manganese superoxide dismutase (MnSOD) has been shown to localize to the chromosomal portion of the cell and impart protection from ionizing radiation to DNA. The binding affinity of bacterial MnSOD to non-sequence specific double stranded oligomeric DNA has been quantitated previously by nitrocellulose filter binding and gel shift assays. In the current study we have examined the equilibrium binding of Escherichia coli MnSOD to poly(U), poly(A), poly(C), poly(dU) and double-stranded (ds) DNA. Equilibrium association constant, Kobs, was measured by monitoring intrinsic tryptophan fluorescence quenching. Based on the extent of quenching, Kobs was determined as a function of monovalent salt (MX) concentration and type, as well as temperature, from which ΔG°obs and ΔH°obs were determined. It was found that the polynucleotides bind to MnSOD in the following affinity hierarchy, poly(dU) > poly(U) > dsDNA > poly(A) > poly(C). The differences in the hierarchy were not large in magnitude as the poly(dU) bound with less than a 100-fold higher affinity than poly(C) at any given [MX]. For each polynucleotide, Kobs decreases only slightly with increasing [K+], surprising for a relatively non-specific nucleic acid protein. Thus, our finding that MnSOD can bind to RNA leads to the possibility that MnSOD may confer protection to RNA, as well. This is, as of yet, untested. Typically one would expect strong electrostatic interactions to dominate a non-specific binding event like that, but our results show an unexpectedly strong non-electrostatic contribution to the binding.  相似文献   

16.
The bimetallic [M1M2(tren)2(CAn?)]m+ series, where M = GaIII or CrIII and CA is the chloranilate ligand which can take on diamagnetic (CAcat,cat)4? or paramagnetic (CAsq,cat)3? forms, comprises an electronically diverse series of compounds ranging from the closed-shell [Ga2(tren)2(CAcat,cat)]2+ to the S = 5/2 ground state of [Cr2(tren)2(CAsq,cat)]3+. This report deals with the interpretation of the EPR and ENDOR spectra of [Ga2(tren)2(CAsq,cat)](BPh4)2(BF4) (2) and the related derivative [Ga2(tren)2(DHBQ)](BPh4)2(BF4) (2a) (where DHBQ is the fully deprotonated trianionic form of 2,5-dihydroxy-1,4-benzoquinone) in an effort to further characterize the electronic structure of this radical species. The X-band (~9.5 GHz) EPR spectrum of complex 2 acquired in a butyronitrile/propionitrile glass at 4 K reveals a rhombic g-tensor with gxx = 2.0100, gyy = 2.0097, and gzz = 2.0060 with hyperfine interactions due to spin delocalization onto the two Ga nuclei (axx = 4.902 G, ayy = 4.124 G, azz = 3.167 G); the origin of the hyperfine coupling was confirmed by analysis of the room temperature spectra of complexes 2 and 2a. The low-temperature spectrum of complex 2 also indicates the presence of a triplet electronic state characterized by a g-value of 2.009 and axial zero-field splitting of D = 150 G (0.012 cm?1) as determined from measurements carried out at both X- and W-band (~95 GHz) frequencies. This triplet state is believed to arise due to a weak intermolecular Heisenberg exchange interaction between two aggregating complexes. ENDOR measurements on complex 2a at 20 K allowed for a determination of the magnitude of hyperfine coupling to the protons associated with the radical bridge as well as providing a rare example of an ENDOR signal arising from coupling to a gallium nucleus. Finally, these results were combined with literature data on the free semiquinone form of the bridging ligand in order to assess the extent to which density functional theory can predict unpaired spin density distribution in a complex molecule of this type. Although differences between theory and experiment were noted, DFT was able to provide a reasonably accurate picture of the electronic structure of this system as well as provide insight into the spin polarization mechanism(s) responsible for the observed hyperfine interactions.  相似文献   

17.
We present the results of an X-ray magnetic circular dichroism investigation of a reduced version of Mn12 benzoate. At variance with the parent Mn12 benzoate compound, which, analogously to Mn12 acetate, has a ground-state spin equal to ten, the reduced species has a ground-state with total spin S = 19/2. The half-integer spin in the ground-state makes this compound an appealing system where to test parity effects on the efficiency of the quantum tunnelling of the magnetisation. We exploited the sensitivity of X-ray magnetic circular dichroism to the oxidation state of the absorbing metal ion to obtain information about the internal structure of the reduced Mn12 benzoate. In particular, we performed multiplet calculations to analyse the X-ray magnetic circular dichroism spectra at the manganese L2,3 edge and identify the contribution of the MnII ion resulting by the reduction process.  相似文献   

18.
Dinuclear [{M(phen)2}2V4O12] · C6H12O · H2O (M = CoII 1, MnII 2, NiII 3 and CuII 4) and [{Cu(phen)2}2V4O12] · 3.5H2O 5 has been prepared by biphasic and hydrothermal syntheses, respectively. All five structures exhibit the {V4O12}4? cluster in a chair-like configuration, covalently bonded to two [M(phen)2]2+ fragments, producing a super-exchange magnetic phenomenon. The magnetic study of complexes 15 shows that they are very weak antiferromagnetically coupled systems, with J values of ?0.14, 2; ?0.64, 3 and ?0.23, 4 cm?1. Complexes 1 to 3 correspond to isostructural compounds in which the cyclovanadate group acts as a bidentate bridged ligand. In the copper complexes (4 and 5) the {V4O12}4? anion presents the novel monodentate bridging mode, and therefore a more significant distortion from the chair-like configuration. The mentioned complexes, together with that reported in the literature, permit to conclude that it is quite common for a single molecular species to exist in more than one crystalline arrangement. A detailed analysis of the structures of 14 shows that the crystal symmetry cannot be strictly centrosymmetric, due to the presence of the cyclohexanol molecule with a single –OH group in the lattice.  相似文献   

19.
TGase 2 is over-expressed in a variety of inflammatory diseases including allergic asthma. This study aimed to investigate the role of TGase 2 on IgE production and signaling pathways in mast cell activation related to OVA-induced allergic asthma. Bone marrow-derived mast cells (BMMCs) isolated from WT or TGase 2?/? mice were activated with Ag/Ab (refer to act-WT-BMMCs and act-KO-BMMCs, respectively). B cells isolated from splenocytes were activated with anti-mouse IgM (act-B cells), and B cells were co-cultured with BMMCs. WT and TGase 2?/? mice were sensitized and challenged with OVA adsorbed in alum hydroxide. Intracellular Ca2 + ([Ca2 +]i) levels were determined by fluorescence intensity; IgE, mediators and TGase 2 activity by ELISA; the CD138 expression by FACS analyzer; cell surface markers and signal molecules by Western blot; NF-κB by EMSA; co-localization of mast cells and B cells by immunohistochemistry; Fcε RI-mediated mast cell activation by PCA test; expression of cytokines, MMPs, TIMPs, TLR2 and Fc?RI by RT-PCR. In vitro, act-KO-BMMCs reduced the [Ca2 +]i levels, NF-κB activity, expression of CD40/CD40L, plasma cells, total IgE levels and TGase 2 activity in act-B cells co-cultured with act-BMMCs, expression of inflammatory cytokines and MMPs2/9, release of mediators (TNF-α, LTs and cytokines), and activities of signal molecules (PKCs, MAP kinases, I-κB and PLA2), which were all increased in act-WT-BMMCs. TGase 2 siRNA transfected/activated-BMMCs reduced all responses as same as those in act-KO-BMMCs. In allergic asthma model, TGase 2?/? mice protected against PCA reaction, OVA-specific IgE production and AHR, and they reduced co-localization of mast cells and B cells or IgE in lung tissues, expression and co-localization of surface molecules in mast cells (c-kit and CD40L) and B cells (CD23 and CD40), inflammatory cells including mast cells, goblet cells, amounts of collagen and mediator release in BAL fluid and/or lung tissues, which were all increased in WT mice. TLR expression in TGase 2?/? mice did not differ from those in WT mice. Our data suggest that TGase 2 expression and Ca2 + influx required by bidirectional events in mast cell activation facilitate IgE production in B cells via up-regulating mast cell CD40L expression, and induce the expression of numerous signaling molecules associated with airway inflammation and remodeling in allergic asthma.  相似文献   

20.
Dehydrogenases that use ubiquinone as an electron acceptor, including complex I of the respiratory chain, complex II, and glycerol-3-phosphate dehydrogenase, are known to be direct generators of superoxide and/or H2O2. Dihydroorotate dehydrogenase oxidizes dihydroorotate to orotate and reduces ubiquinone to ubiquinol during pyrimidine metabolism, but it is unclear whether it produces superoxide and/or H2O2 directly or does so only indirectly from other sites in the electron transport chain. Using mitochondria isolated from rat skeletal muscle we establish that dihydroorotate oxidation leads to superoxide/H2O2 production at a fairly high rate of about 300 pmol H2O2·min−1·mg protein−1 when oxidation of ubiquinol is prevented and complex II is uninhibited. This H2O2 production is abolished by brequinar or leflunomide, known inhibitors of dihydroorotate dehydrogenase. Eighty percent of this rate is indirect, originating from site IIF of complex II, because it can be prevented by malonate or atpenin A5, inhibitors of complex II. In the presence of inhibitors of all known sites of superoxide/H2O2 production (rotenone to inhibit sites in complex I (site IQ and, indirectly, site IF), myxothiazol to inhibit site IIIQo in complex III, and malonate plus atpenin A5 to inhibit site IIF in complex II), dihydroorotate dehydrogenase generates superoxide/H2O2, at a small but significant rate (23 pmol H2O2·min−1·mg protein−1), from the ubiquinone-binding site. We conclude that dihydroorotate dehydrogenase can generate superoxide and/or H2O2 directly at low rates and is also capable of indirect production at higher rates from other sites through its ability to reduce the ubiquinone pool.  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