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.
Soluble guanylyl/guanylate cyclase (sGC), a heme-containing heterodimeric protein of approximately 150 kDa, is the primary receptor for nitric oxide, an endogenous molecule of immense physiological importance to animals. Recent studies have identified compounds such as YC-1 and BAY 41-2272 that stimulate sGC independently of NO binding, properties of importance for the treatment of endothelial dysfunction and other diseases linked to malfunctioning NO signaling pathways. We have developed a novel expression system for sGC from Manduca sexta (the tobacco hornworm) that retains the N-terminal two-thirds of both subunits, including heme, but is missing the catalytic domain. Here, we show that binding of compounds YC-1 or BAY 41-2272 to the truncated protein leads to a change in the heme pocket such that photolyzed CO cannot readily escape from the protein matrix. Geminate recombination of the trapped CO molecules with heme takes place with a measured rate of 6 x 10(7) s(-1). These findings provide strong support for an allosteric regulatory model in which YC-1 and related compounds can alter the sGC heme pocket conformation to retain diatomic ligands and thus activate the enzyme alone or in synergy with either NO or CO.  相似文献   

2.
The heme-regulated eukaryotic initiation factor 2alpha (eIF2alpha) kinase (HRI), which is found primarily in reticulocytes, contains an N-terminal heme-binding domain (NT-HBD). Binding of NO to the heme iron of the NT-HBD of HRI activates its eIF2alpha kinase activity, thus inhibiting the initiation of translation in reticulocyte lysate. The EPR spectrum of the NO-bound NT-HBD showed several derivative-shaped lines around g = 2.00, which is one of the well-documented signature patterns of a six-coordinate NO complex with histidine as the axial ligand. This is in sharp contrast to that of another prototypical NO-sensor protein, soluble guanylate cyclase (sGC), in which the NO binding to the heme iron disrupts the iron-histidyl bond forming a five-coordinate NO. The NO-mediated activation of HRI is, therefore, not triggered by the cleavage of the iron-histidyl bond. As evidenced by the resonance Raman spectra, two inactive forms of HRI, the ferrous ligand-unbound and the CO-bound states of the NT-HBD, contain a six-coordinate complex as found for the NO complex, indicating that the replacement of the sixth ligand of the heme iron is not sufficient to trigger the activation of HRI. Because the configuration of liganded NO is different from that of liganded CO, we propose that specific interactions between liganded NO and surrounding amino acid residues, which would not be formed in the CO complex, are responsible for the NO-induced activation of HRI.  相似文献   

3.
《Polyhedron》1999,18(20):2605-2608
The interaction of iron carbonyls, Fe(CO)5, Fe2(CO)9 and Fe3(CO)12 with Me3NO occurs according to a one-electron redox-disproportionation scheme giving rise to iron carbonyl radical anions: Fe2(CO)8·− (1), Fe3(CO)12·− (2), Fe3(CO)11·− (3) and Fe4(CO)13·− (4). The role of Me3NO, inducing CO-substitution, consists of the generation of reactive 17-electron species with a labile coordination sphere in which the substitution for other ligands occurs, resulting from fast ligand and electron exchange in the confines of the ETC-reaction.  相似文献   

4.
The first examples of the slippage formation of rotaxane‐like structures in the presence of an anion template are reported between a macrocycle, synthesised by exploiting Eglinton coupling, and stoppered pyridinium axle components. The role of the anion template in the slippage process has been explored by kinetic studies. 1H NMR spectroscopic investigations reveal the slippage species formed are not rotaxanes but pseudorotaxanes with some rotaxane character. The anion template significantly influences the amount of rotaxane character and the rate of slippage. Importantly, the fastest slippage rates, kon, are achieved with the non‐coordinating hexafluorophosphate anion, whereas the slowest slippage off rates, koff, are observed in the presence of coordinating anions, such as chloride. Since the koff rates are significantly smaller than the kon rates in the presence of coordinating anions, these anions act as templates favouring formation of the slippage species thermodynamically. Consequently, the resulting pseudorotaxanes with coordinating anions have greater rotaxane character. Two strategies for converting the slippage pseudorotaxanes into rotaxanes using hydrogenation or complexation with cobalt carbonyl are investigated.  相似文献   

5.
Heme reactivity in hemoproteins is governed by the microenvironment near the ligand binding site. In order to quantify polarity effects on heme ligand binding, the kinetics of O2 and CO binding have been measured for a series of synthetic heme models equipped with a range of groups of varying dipole moments positioned near the heme coordination site. For hemes with polar aprotic groups, both O2 on (k′) and off rates (k) are found to be dependent on the dipole moment. For model systems containing protic groups, the O2 off rate is substantially reduced due to hydrogen bonding with the coordinated O2. The hydrogen-bonding stabilization is estimated to be 0.7 and 1.6 kcal/mol for an alcohol and a secondary amide, respectively. CO binding displays little correlation with a polarity effect; instead it seems to depend upon the size and position of the polar group.  相似文献   

6.
The reaction of hemoglobin (Hb), oxyhemoglobin (HbO2), and methemoglobin (metHb) with the tetranitrosyl iron complex of the fu2-S type [Fe2(SC4H3N2)2(NO)4] (1) was studied. The reaction results in the nitrosylation of the free SH group of 93-β-cysteine in these forms of hemoglobin. The change in the Hb, HbO2, and metHb concentrations was monitored by spectrophotometry, recording the difference absorption spectra of the experimental systems with these forms of hemoglobin and the buffer containing complex 1 in the same concentration. The absorption spectra were processed to obtain the components using the MATHCAD method. The nitrosothiol concentration was determined by the Saville reaction. In a protic medium containing 3.3% DMSO, complex 1 spontaneously generates NO due to hydrolysis (k = 3.7 · 10-4 s-1). Oxyhemoglobin reacts with evolved NO to form metHb. Complex 1 reduces metHb with a high rate to yield Hb (k = 6.7 · 10-3 s-1) followed by the formation of HbNO (k = 6.5 · 10-3 s-1). Oxidized complex 1 yields NO with a higher rate than the starting complex does. The reaction of HbO2 and metHb (0.02 mmo1 L-1) with complex 1 affords nitrosothiols in micromolar concentration during 5 min, and no nitrosothiol is formed in the case of Hb.  相似文献   

7.
Flash photolysis of CH3CHO and H2CO in the presence of NO has been investigated by the intracavity laser spectroscopy technique. The decay of HNO formed by the reaction HCO + NO → HNO + CO was studied at NO pressures of 6.8–380 torr. At low NO pressure HNO was found to decay by the reaction HNO + HNO → N2O + H2O. The rate constant of this reaction was determined to be k1 = (1.5 ± 0.8) × 10?15 cm3/s. At high NO pressure the reaction HNO + NO → products was more important, and its rate constant was measured to be k2 = (5 ± 1.5) × 10?19 cm3/s. NO2 was detected as one of the products of this reaction. Alternative mechanisms for this reaction are discussed.  相似文献   

8.
《Polyhedron》2001,20(9-10):1011-1016
The compounds [Fe(ch)(CO)2PPh3] (1) (ch=chalcone) and [Fe(sba)(CO)2PPh3] (2) (sba=sorbic acid) were prepared by irradiating the tetracarbonyltriphenylphosphineiron(0) complex in benzene in the presence of ch or sba. The compounds were characterized by infrared and 31P NMR spectroscopies. Their electrochemical behavior was investigated by cyclic voltammetry and the results suggest that their oxidations occur by more than one electrochemical step, producing free ch and sba, free PPh3 and solvated Fe(III). It was observed that sba ligand contributes more effectively to the stabilization of metal center in these complexes. The X-ray crystal and molecular structures of 1 and 2 were determined; it was shown that the Fe atom adopts a distorted octahedral coordinated geometry in which three of the sites are occupied by the ch or sba ligand. The [Fe(ch)(CO)2PPh3] complex is a monomer and the unit cell of complex 2 contains exist two identical and crystallographically independent molecules of [Fe(sba)(CO)2PPh3] which are linked by short hydrogen bonds OH· · ·O  相似文献   

9.
The H-NOX (Heme-Nitric oxide/OXygen binding) family of diatomic gas sensing hemoproteins has attracted great interest. Soluble guanylate cyclase (sGC), the well-characterized eukaryotic nitric oxide (NO) sensor is an H-NOX family member. When NO binds sGC at the ferrous histidine-ligated protoporphyrin-IX, the proximal histidine ligand dissociates, resulting in a 5-coordinate (5c) complex; formation of this 5c complex is viewed as necessary for activation of sGC. Characterization of other H-NOX family members has revealed that while most also bind NO in a 5c complex, some bind NO in a 6-coordinate (6c) complex or as a 5c/6c mixture. To gain insight into the heme pocket structural differences between 5c and 6c Fe(ii)-NO H-NOX complexes, we investigated the extended X-ray absorption fine structure (EXAFS) of the Fe(II)-unligated and Fe(II)-NO complexes of H-NOX domains from three species, Thermoanaerobacter tengcongensis, Shewanella woodyi, and Pseudoalteromonas atlantica. Although the Fe(II)-NO complex of TtH-NOX is formally 6c, we found the Fe-N(His) bond is substantially lengthened. Furthermore, although NO binds to SwH-NOX and PaH-NOX as a 5c complex, consistent with histidine dissociation, the EXAFS data do not exclude a very weakly associated histidine. Regardless of coordination number, upon NO-binding, the Fe-N(porphyrin) bond lengths in all three H-NOXs contract by ~0.07 ?. This study reveals that the overall heme structure of 5c and 6c Fe(II)-NO H-NOX complexes are substantially similar, suggesting that formal histidine dissociation may not be required to trigger NO/H-NOX signal transduction. The study has refined our understanding of the molecular mechanisms underlying NO/H-NOX signaling.  相似文献   

10.
Reactions of CpMn(CO)(NO)SnCl3 (I) with sodium benzenethiolate and sodium benzenesele-nolate gave orange crystals of the complexes CpMn(CO)(NO)Sn(EPh)3, where E = S (II) or Se (III). Treatment of complex II with photochemically generated W(CO)5(THF) yielded the adduct CpMn(CO)(NO)Sn(SPh)3 · W(CO)5 (IV). A similar treatment of complex III resulted in the formation of the ditungsten complex W2(CO)4(SePh)6 (V) with transfer of all chalcogenate groups from tin to tungsten. In reactions of complexes II and III with a Pt0 complex with phosphine and acetylene, (PPh3)2Pt(Ph2C2), the chalcogenate groups are transferred from tin. Only the known Pt(II) complexes (PPh3)2Pt(EPh)2), where E= S (VI) or Se (VII). Molecular structures IV and V were characterized by X-ray diffraction. It has been found that the Mn-Sn bond in complex IV (2.5479(9) Å) is nearly the same length as that found earlier for complex II (2.5328(17) Å) and is substantially shorter than the sum of the covalent radii of Mn and Sn (2.78 Å). The Sn-S bond is noticeably lengthened (2.5217(11) Å) only for the S atom bound to tungsten (W-S, 2.5696(12) Å), while the other Sn-S bonds (2.4413(12) and 2.4291(12) Å) are virtually the same as in complex II (on average, 2.441 Å). Complex V contains the direct W-W bond (2.8153(16) Å) supplemented with four benzeneselenolate bridges in which the W-Se bonds (on average, 2.642(2) Å) are longer than the two terminal W-SePh bonds (2.571(2) Å). All the W-Se bonds are much shorter than the sum of the covalent radii of W and Se (2.82 Å).  相似文献   

11.
The reaction of [CpMn(CO)(NO)]2 (I) with an equimolar amount of tin dichloride in THF at room temperature gave the product of tin insertion into the Mn-Mn bond, the carbonyl nitrosyl complex [CpMn(CO)(NO)]2SnCl2 (II). The same complex was formed on treatment of CpMn(CO)(NO)SnCl3 with sodium borohydride. Treatment of I with an excess of anhydrous tin dichloride under the same conditions gave the trinitrosyl complex Cp2Mn2(NO)(μ-NO)2SnCl3 (III). According to X-ray diffraction, II contains a Mn-Sn-Mn chain with highly shortened Mn-Sn bonds (2.5570(2) and 2.5754(2) Å). Compound III contains a Mn-Mn-Sn chain (Mn-Mn, 2.5358(10); Mn-Sn, 2.5604(8) Å) with the Mn-Mn bond supplemented by two nitrosyl bridges and one terminal NO group.  相似文献   

12.
Observation of a potentiometric sensor's response behaviour after injection in flow injection analysis at different concentrations allowed studying “on” and “off” kinetics of the analyte's adsorption/diffusion behaviour. The alkaloid metergoline was mostly used as an example. kon and koff rate constant values were measured, and the association constant Kass, and ΔG values of the analyte–surface interaction were calculated with an adsorption-based model which proved to be fully applicable. kon increased by decreasing the sensor dimensions, while koff was unaffected by miniaturization. Increasing acetonitrile concentrations in the running buffer increased koff, while kon was unaffected. The experimentally determined ΔG values of the analyte–surface interaction showed a linear relation to the response of the sensor, in mV. This knowledge was applied to optimize the potentiometric detection of plant alkaloids in (U)HPLC. Sub-micromolar detection limits were obtained with the potentiometric detector/(U)HPLC combination. This is the first time that the response rates and the response itself can be modelled accurately for coated wire potentiometric sensors, and it is the first application of a potentiometric detector in UPLC.  相似文献   

13.
Reaction of complex [(μ-SCH2)2NCH2CO2Me]Fe2(CO)6 (A) with 1,1-bis(diphenylphosphino)ferrocene (dppf) in the presence of the decarbonylating agent Me3NO?2H2O gave complex [(μ-SCH2)2NCH2CO2MeFe2(CO)5]2[(η 5-Ph2PC5H4)2Fe] (1) in 72 % yields, whereas complex [(μ-SCH2)2NPhFe2(CO)5]2[(η 5-Ph2PC5H4)2Fe] (2) was produced by reaction of [(μ-SCH2)2NPh]Fe2(CO)6 (B) with dppf in toluene at reflux in 41 % yield. The new complexes 1 and 2 were characterized by elemental analysis, IR, and 1H (31P, 13C) NMR spectroscopy as well as by single crystal X-ray diffraction analysis. In the crystal structures of 1 and 2, the dppf ligand resides in an apical position of the square-pyramidal geometry of the neighbouring Fe atoms and the crystal structures were stabilized by the intermolecular C–H···O hydrogen bonds.  相似文献   

14.
The heating of the ionic complex [CpMn(CO)2(NO)]+SnCl3-(I) in methylene chloride gives a neutral complex CpMn(CO)(NO)SnCl3 (II). The latter reacts with lithium phenylacetylenide to yield a complex CpMn(CO)(NO)Sn(C≡CPh)3 (III). According to the X-ray diffraction data, complexes II and III contain shortened Mn-Sn bonds (2.5178(5) and 2.5436(12) Å, respectively).  相似文献   

15.
The lack of direct proof in either natural or synthetic systems for trans-dinitrosyl hemes, a key intermediate in the reactions of heme proteins (e.g. soluble guanylate cyclase (sGC), cytochrome c′ and So H-NOX) with nitric oxide (NO), has hampered understanding of the exact reaction mechanisms, such as the formation of the five-coordinate heme complex with NO at the proximal side (5c NOP). Herein, we report the first isolation of a dinitrosyl metalloporphyrin complex, the six-coordinate, low-spin {Mn(NO)2}7 species [Mn(TPP)(NO)2] (TPP2−=meso-tetraphenylporphyrin dianion). The complex shows distinct features, such as an elongated axial bond (1.877(9) vs. 1.641(5) Å), a higher NO stretching bond position (1760 vs. 1735 cm−1) and an isotropic resonance at g = 2.0, in sharp contrast to those of five-coordinate mononitrosyl analogues. In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFT) and EPR studies provided deep insight into the reaction processes, demonstrating different responses of porphyrinates to NO.  相似文献   

16.
《Supramolecular Science》1995,2(3-4):175-182
Steady-state fluorescence and single photon timing have been used to study the effect of the presence of hydrogen bonding on the intermolecular quenching of pyrene covalently linked to a guanine-like receptor I by an aliphatic amine (N,N-dimethylpropylamine) covalently linked to cytosine derivative II. By comparing the fluorescence quenching of I by II with that of 10methylpyrene (1-MP) by triethylamine (TEA), as a model system in which no hydrogen bonding can occur, one could possibly analyze the effect of the hydrogen bonding between receptor and substrate as a quenching as it leads to a higher local concentration of donor and acceptor. While the quenching of I by II was observed with an apparent rate constant kq of (1.78 ± 0.10) × 109 M−1 s−1 and (8.72 ± 0.42) × 108 M−1 s−1 in toluene and acetonitrile, respectively, no quenching could be observed in methanol. Upon excitation of 1-MP, no quenching by II could be detected in the same concentration range as used in the quenching of I. Quenching of I and of 1-MP by TEA (⩾ 10−2 M) in toluene leads to exciplex formation with maxima centred at 540 and 514 nm, respectively. The rate constants of exciplex formation and dissociation of I with TEA were analyzed using a global compartmental analysis. The following values were obtained for the rate constants: k01 = (9.70 ± 0.01) × 106 s−1, k21 = (1.12 ± 0.003) × 109 M−1 s−1, k02 = (5.24 ± 0.01) × 107 s−1 and k12 = (7.74 ± 0.08) × 106 s−1. Quenching of I by TEA in the presence of III, a hydrogen-bonding system without an alkyl amine substituent, leads to exciplex formation centred at 538 nm. The rate constant values for the exciplex formation and dissociation of I with TEA in the presence of III were: k01 = (9.32 ± 0.08) × 106 s−1, k21 = (9.32 ± 0.003) × 108 M−1 s−1, k02 = (6.16 ± 0.03) × 107 s−1 and k12 = (21.90 ± 0.3) × 106 s−1. The apparent rate constants kq for this system was (7.26 ± 0.56) × 106 M−1 s−1. The observed decrease in the rate of exciplex formation of I with TEA in the presence of III could suggest that the guanine-like moiety in I forms hydrogen bonds with the cytosine-like moiety and this could decrease the electron affinity of I. The rate constant of exciplex dissociation increased, indicating that the exciplex is less stable in the presence of III. Because of the single exponential decay of I in the presence and absence of II and of the agreement between steady-state and transient fluorescence measurements, the information available for quantitative analysis of the association between I and II is limited.  相似文献   

17.
Addition of methoxide to either geometric isomer of the benzylidene complex [(η-C5H5)Re(NO)(PPh3)(CHC6H5)]+PF6? (1t, 1k) affords (η-C5H5)Re(NO)(PPh3)(CH(OCH3)C6H5 (2t, 2k) in which a new chiral center has been generated stereospecifically or with high stereoselectivity. Reaction of 2t and 2k with Ph3C+PF6? results in the chemospecific abstraction of a methoxy group and the stereospecific regeneration of 1t and 1k, respectively.  相似文献   

18.
The reaction pattern of the Lewis-acid/base stabilised phosphanylborane [(CO)5W(H2PBH2 · NMe3)] (1) with elemental halogens is comprehensively studied. The reaction with iodine and bromine leads to a selective halogenation at the tungstencarbonyl moiety under formation of [WX2(CO)4(H2PBH2 · NMe3)] (X = I (2), Br (3)). Whereas 2 is a stable product the brominated compound 3 dimerises easily to [WBr2(CO)3(H2PBH2 · NMe3)]2 (4) under lost of CO. The CO elimination reaction of 3 is extensively studied. If 3 is reacted with [Et4N][Br] the ionic compound [Et4N][WBr3(CO)3(H2PBH2 · NMe3)] (5) is formed. Otherwise, if 3 is combined with the donor ligand [H2PBH2 · NMe3], the complex [WBr2(CO)3(H2PBH2 · NMe3)2] (6) is obtained. Compounds 26 are comprehensively characterised by X-ray diffraction analysis, NMR, and IR spectroscopy.  相似文献   

19.
The rate constant k4 has been measured at 268°, 298°, and 334° K for the reaction CH2O + 2OH → CO + 2H2O relative to that for OH + OH (k2) by competition experiments in a discharge flow tube using mass-spectrometric analysis. Based on k2 = 2.24 × 10?12cm3/molec·sec at 298°K and E2 = 4 kJ/mol, k4 = (6.5 ± 1.5) × 10?12cm3/molec·sec at 298°K and E4 = (6 ± 2)kJ/mol.  相似文献   

20.
《Polyhedron》2001,20(15-16):1925-1931
The synthesis and structural characterization of novel organometallic coordination polymers are reported. The reaction of Cd(NO3)2 and 4,4′-bipy in CH3OH/H2O gave a 2D coordination network formulated as {[Cd(4,4′-bpy)2·(H2O)2](NO3)2·4H2O}10, which was used to capture an organic guest species (4-amino-benezopheone, C13H11NO (3)) to obtain {[Cd(4,4′-bpy)2(NO3)(H2O)]·NO3·(C13H11NO)2} (1). Using L (L=4,4′-trimethylenedipyridine) instead of 4,4′-bipy, {[Cd(L)2(H2O)2]·2H2O·2NO3·C13H11NO} (2) was synthesized, which has an interesting configuration.  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