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 31 毫秒
1.
Indole-3-acetaldoxime (IAOx) is a branch point compound of tryptophan (Trp) metabolism in glucosinolate-producing species such as Arabidopsis, serving as a precursor to indole-glucosinolates (IGs), the defense compound camalexin, indole-3-acetonitrile (IAN) and indole-3-acetic acid (IAA). We synthesized [2H5] and [13C1015N2]IAOx and [13C6], [2H5] and [2′,2′-2H2]IAN in order to quantify endogenous IAOx and IAN in Arabidopsis and tobacco, a non-IG producing species. We found that side chain-labeled [2′,2′-2H2]IAN overestimated the amount of IAN by 2-fold compared to when [2H5]IAN was used as internal standard, presumably due to protium-deuterium exchange within the internal standard during extraction of plant tissue. We also determined that [13C1]IAN underestimated the amount of IAN when the ratio of [13C1]IAN standard to endogenous IAN was greater than five to one, whereas either [2H5]IAN or [13C6]IAN showed a linear relationship with endogenous IAN over a broader range of concentrations. Transgenic tobacco vector control lines did not have detectable levels of IAOx or IAN (limit of detection ∼ 100 pg/g fr. wt), while lines expressing either the IAOx-producing CYP79B2 or CYP79B3 genes from Arabidopsis under CaMV 35S promoter control accumulated IAOx in the range of 1-9 μg/g fr. wt. IAN levels in these lines ranged from 0.6 to 6.7 μg/g fr. wt, and IAA levels were ∼9-14-fold above levels in control lines. An Arabidopsis line expressing the same CYP79B2 overexpression construct accumulated IAOx in two of three lines measured (∼200 and 400 ng/g fr. wt) and accumulated IAN in all three lines. IAN is proposed to be a metabolite of IAOx or an enzymatic breakdown product of IGs induced upon tissue damage. Since tobacco does not produce detectable IGs, the tobacco data are consistent with IAN being a metabolite of IAOx. IAOx and IAN were also examined in the Arabidopsis activation tagged yucca mutant, and no accumulation of IAOx was found above the limits of detection but accumulation of IAN (3-fold above wt) occurred. The latter was surprising in light of recent reports that rule out IAOx and IAN as intermediates in YUCCA-mediated IAA synthesis.  相似文献   

2.
The synthesis and crystal structure of four new copper(I) and copper(II) supramolecular amine, and amine phosphonate, complexes is reported. Reaction of copper(I) with 2-,9-dimethyl-1-10-phenanthroline (dmp) produced a stable 4-coordinate Cu(I) species, [Cu(I)(dmp)2]Cl · MeOH · 5H2O (2), i.e., the increased steric hindrance in the ‘bite’ area of dmp did not prevent interaction with the metal and provided protection against oxidation which was not possible for the phen analogue [R. Clarke, K. Latham, C. Rix, M. Hobday, J. White, CrystEngCommun. 7(3) (2005), 28-36]. Subsequent addition of phenylphosphonic acid to (2) produced two structures from alternative synthetic routes. An ‘in situ’ process yielded red block Cu(I) crystals, [Cu(I)(dmp)2] · [C6H5PO3H2 · C6H5PO3H] (4), whilst recrystallisation of (2) prior to addition of the acid (‘stepwise’ process) produced a green, needle-like Cu(II) complex, [Cu(II)(dmp) · (H2O)2 · C6H5PO2(OH)] [C6H5PO2(OH)] (3). However, addition of excess dmp during the ‘stepwise’ process forced the equilibrium towards product (4) and resulted in an optimum yield (99%). The structure of (4) was similar to the phen analogue, [Cu(II)Cl(phen)2] · [C6H5PO2(OH) · C6H5PO(OH)2] (1) [R. Clarke, K. Latham, C. Rix, M. Hobday, J. White, CrystEngCommun. 7(3) (2005), 28-36], but the presence of dmp exerted some influence on global packing, whilst (3) exists as a polymeric layered material. In contrast, reaction of copper(I) with di-2-pyridyl ketone (dpk), followed by phenylphosphonic acid produced purple/blue Cu(II) species, [Cu(II)(dpk · H2O)2] Cl2 · 4H2O (5), and [Cu(II)(dpk · H2O)2] · [C6H5PO2(OH)2 · C6H5PO(OH)2] (6), respectively, i.e., in both cases oxidation of copper occurred. Solid-state luminescence was observed in (2) and (4). The latter showing a 5-fold enhancement in intensity.  相似文献   

3.
The thiocarbamates 4-RC6H4NHC(S)NR2′ (R = H, Cl; R′ = Me, Et), 4-ClC6H4NHC(S)NR (NR = 2-pyridylpiperazine) react with cis-[PtCl2(PTA)2] (PTA = 1,3,5-triaza-7-phosphaadamantane) in the presence of base to afford the monocationic platinum(II) complexes cis-[Pt{SC(NR2′) = NC6H4R}(PTA)2]+ (R = H, Cl; R′ = Me, Et), cis-[Pt{SC(NR) = NC6H4Cl}(PTA)2]+ (NR = 2-pyridylpiperazine), which were isolated as their PF6 salts in high yields. The complexes were fully characterised spectroscopically and also by X-ray crystallography. Cytotoxicity of these complexes was studied in vitro in three human cancer cell lines (CH1, A549 and SW480) using the MTT assay.  相似文献   

4.
Reactions of the molybdenum and tungsten precursors [MO2S2]2− with equimolar amounts of benzenedithiol in acetonitrile give the title compounds [M2O2(μ-S)2(bdt)2]2− with M = Mo, W and bdt = benzene-dithiolate. In case a tungsten to ligand ratio of 1:2 is used the dimer forms as well but only as a minor species whereas the monomer [WO(bdt)2]2− is the main product. In both dimeric compounds the syn-isomers are formed referring to the position of the apical oxo ligands with respect to the M2S2 plane. For the molybdenum compound this contrasts with a published crystal structure of the anti-isomer. Both complexes give highly symmetric isomorphous crystals but still show subtle differences in their bond lengths and angles around the central metal. The X-ray crystal structures of both are analyzed in detail and compared with each other and with the isomeric molybdenum compound. Differences and similarities between tungsten and both isomers of molybdenum complexes are shown to be more influenced by the conformation than by the central metal and a reason for the formation of syn- and anti-isomers based on the respective synthetic procedures is proposed.  相似文献   

5.
Bioluminescence resonance energy transfer (BRET) is a powerful tool for the study of protein-protein interactions and conformational changes within proteins. Two common implementations of BRET are BRET1 with Renilla luciferase (RLuc) and coelenterazine h (CLZ, λem ∼ 475 nm) and BRET2 with the substrate coelenterazine 400a (CLZ400A substrate, λem = 395 nm) as the respective donors. For BRET1 the acceptor is yellow fluorescent protein (YFP) (λem ∼ 535 nm), a mutant of green fluorescent protein (GFP), and for BRET2 it is GFP2em ∼ 515 nm). It is not clear from previous studies which of these systems has superior signal-to-background characteristics. Here we directly compared BRET1 and BRET2 by placing two different protease-specific cleavage sequences between the donor and acceptor domains. The intact proteins simulate protein-protein association. Proteolytic cleavage of the peptide linker simulates protein dissociation and can be detected as a change in the BRET ratios. Complete cleavage of its target sequence by thrombin changed the BRET2 ratio by a factor of 28.9 ± 0.2 (relative standard deviation [RSD], n = 3) and changed the BRET1 ratio by a factor of 3.05 ± 0.07. Complete cleavage of a caspase-3 target sequence resulted in the BRET ratio changes by factors of 15.45 ± 0.08 for BRET2 and 2.00 ± 0.04 for BRET1. The BRET2 assay for thrombin was 2.9 times more sensitive compared with the BRET1 version. Calculated detection limits (blank signal + 3σb, where σb = standard deviation [SD] of blank signal) were 53 pM (0.002 U) thrombin with BRET1 and 15 pM (0.0005 U) thrombin with BRET2. The results presented here suggest that BRET2 is a more suitable system than BRET1 for studying protein-protein interactions and as a potential sensor for monitoring protease activity.  相似文献   

6.
A new class of pyrrolo[2,3-d]pyrimidin-4-one corticotropin-releasing factor 1 (CRF1) receptor antagonists has been designed and synthesized. In general, reported CRF1 receptor antagonists possess a sp2-nitrogen atom as hydrogen bonding acceptor (HBA) on their core scaffolds. We proposed to use a carbonyl group of pyrrolo[2,3-d]pyrimidin-4-one derivatives as a replacement for the sp2-nitrogen atom as HBA in classical CRF1 receptor antagonists. As a result, several pyrrolo[2,3-d]pyrimidin-4-one derivatives showed CRF1 receptor binding affinity with IC50 values in the submicromolar range. Ex vivo 125I-sauvagine binding studies showed that 2-(dipropylamino)-3,7-dimethyl-5-(2,4,6-trimethylphenyl)-3,7-dihydro-4H-pyrrolo[2,3-d]pyrimidin-4-one (16b) (30 mg/kg, po) was able to penetrate into the brain and inhibit radioligand binding to CRF1 receptors (frontal cortex, olfactory bulb, and pituitary) in mice. We identified pyrrolo[2,3-d]pyrimidin-4-one derivatives as the first CRF1 antagonists with a carbonyl-based HBA.  相似文献   

7.
A case study on Centaurea gymnocarpa Moris & De Not., a narrow endemic species, was carried out by analyzing its morphological, anatomical, and physiological traits in response to natural habitat stress factors under Mediterranean climate conditions. The results underline that the species is particularly adapted to the environment where it naturally grows. At the plant level, the above-ground/below-ground dry mass (1.73 ± 0.60) shows its investment predominately in the above-ground structure with a resulting total leaf area per plant of 1399 ± 94 cm2. The senescent attached leaves at the base of the plant contribute to limit leaf transpiration by shading soil around the plant. Moreover, the dense C. gymnocarpa leaf pubescence, leaf rolling, the relatively high leaf mass area (LMA = 12.3 ± 1.3 mg cm−2) and leaf tissue density (LTD = 427 ± 44 mg cm−3) contribute to limit leaf transpiration, also postponing leaf death under dry conditions. At the physiological level, a relatively low respiration/photosynthesis ratio (R/PN) in spring results from high R [2.26 ± 0.59 μmol (CO2) m−2 s−1] and PN [12.3 ± 1.5 μmol (CO2) m−2 s−1]. The high photosynthetic nitrogen use efficiency [PNUE = 15.5 ± 0.4 μmol (CO2) g−1 (N) s−1] shows the large amount of nitrogen (N) invested in the photosynthetic machinery of new leaves, associated to a high chlorophyll content (Chl = 35 ± 5 SPAD units). On the contrary, the highest R/PN ratio (1.75 ± 0.19) in summer is due to a significant PN decrease and increase of R in response to drought. The low PNUE [1.5 ± 0.2 μmol (CO2) g−1 (N) s−1] in this season is indicative of a greater N investment in leaf cell walls which may contribute to limit transpiration. On the contrary, the low R/PN ratio (0.05 ± 0.02) in winter is resulting from the limited enzyme activity of the respiratory apparatus [R = 0.23 ± 0.08 μmol (CO2) m−2 s−1] while the low PNUE [3.5 ± 0.2 μmol (CO2) g−1 (N) s−1] suggests that low temperatures additionally limit plant production. The experiment of the imposed water stress confirms that the C. gymnocarpa growth capability is in conformity with the severe conditions of its natural habitat, likewise as it may be the case with others narrow endemic species that have occupied niches with similar extreme conditions.  相似文献   

8.
VDAC1, an outer mitochondrial membrane (OMM) protein, is crucial for regulating mitochondrial metabolic and energetic functions and acts as a convergence point for various cell survival and death signals. VDAC1 is also a key player in apoptosis, involved in cytochrome c (Cyto c) release and interactions with anti-apoptotic proteins. Recently, we demonstrated that various pro-apoptotic agents induce VDAC1 oligomerization and proposed that a channel formed by VDAC1 oligomers mediates cytochrome c release. As VDAC1 transports Ca2 + across the OMM and because Ca2 + has been implicated in apoptosis induction, we addressed the relationship between cytosolic Ca2 + levels ([Ca2 +]i), VDAC1 oligomerization and apoptosis induction. We demonstrate that different apoptosis inducers elevate cytosolic Ca2 + and induce VDAC1 over-expression. Direct elevation of [Ca2 +]i by the Ca2 +-mobilizing agents A23187, ionomycin and thapsigargin also resulted in VDAC1 over-expression, VDAC1 oligomerization and apoptosis. In contrast, decreasing [Ca2 +]i using the cell-permeable Ca2 +-chelating reagent BAPTA-AM inhibited VDAC1 over-expression, VDAC1 oligomerization and apoptosis. Correlation between the increase in VDAC1 levels and oligomerization, [Ca2 +]i levels and apoptosis induction, as induced by H2O2 or As2O3, was also obtained. On the other hand, cells transfected to overexpress VDAC1 presented Ca2 +-independent VDAC1 oligomerization, cytochrome c release and apoptosis, suggesting that [Ca2 +]i elevation is not a pre-requisite for apoptosis induction when VDAC1 is over-expressed. The results suggest that Ca2 + promotes VDAC1 over-expression by an as yet unknown signaling pathway, leading to VDAC1 oligomerization, ultimately resulting in apoptosis. These findings provide a new insight into the mechanism of action of existing anti-cancer drugs involving induction of VDAC1 over-expression as a mechanism for inducing apoptosis. This article is part of a Special Issue entitled: Calcium Signaling in Health and Disease. Guest Editors: Geert Bultynck, Jacques Haiech, Claus W. Heizmann, Joachim Krebs, and Marc Moreau  相似文献   

9.
The kinetics of Ca2+-dependent conformational changes of human cardiac troponin (cTn) were studied on isolated cTn and within the sarcomeric environment of myofibrils. Human cTnC was selectively labeled on cysteine 84 with N-((2-(iodoacetoxy)ethyl)-N-methyl)amino-7-nitrobenz-2-oxa-1,3-diazole and reconstituted with cTnI and cTnT to the cTn complex, which was incorporated into guinea pig cardiac myofibrils. These exchanged myofibrils, or the isolated cTn, were rapidly mixed in a stopped-flow apparatus with different [Ca2+] or the Ca2+-buffer 1,2-Bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid to determine the kinetics of the switch-on or switch-off, respectively, of cTn. Activation of myofibrils with high [Ca2+] (pCa 4.6) induced a biphasic fluorescence increase with rate constants of >2000 s−1 and ∼330 s−1, respectively. At low [Ca2+] (pCa 6.6), the slower rate was reduced to ∼25 s−1, but was still ∼50-fold higher than the rate constant of Ca2+-induced myofibrillar force development measured in a mechanical setup. Decreasing [Ca2+] from pCa 5.0-7.9 induced a fluorescence decay with a rate constant of 39 s−1, which was approximately fivefold faster than force relaxation. Modeling the data indicates two sequentially coupled conformational changes of cTnC in myofibrils: 1), rapid Ca2+-binding (kB ≈ 120 μM−1 s−1) and dissociation (kD ≈ 550 s−1); and 2), slower switch-on (kon = 390s−1) and switch-off (koff = 36s−1) kinetics. At high [Ca2+], ∼90% of cTnC is switched on. Both switch-on and switch-off kinetics of incorporated cTn were around fourfold faster than those of isolated cTn. In conclusion, the switch kinetics of cTn are sensitively changed by its structural integration in the sarcomere and directly rate-limit neither cardiac myofibrillar contraction nor relaxation.  相似文献   

10.
A new synthetic route to the known tripodal tetradentate N3O ligand L1 (HL1 = [N-(3,5-di-tert-butyl-2-hydroxybenzyl)-N,N-di-(2-pyridylmethyl)]amine) is reported. The related compounds HLn (n = 2, 3) were prepared by a similar procedure. Treatment of HLn (n = 1-3) with FeCl3·6H2O in hot methanol led to the mononuclear iron(III) complexes [Fe(Ln)Cl2] (1: n = 1, 2: n = 2, 3: n = 3). The solid-state structures of complexes 1 and 2 were determined by X-ray crystallography. [Fe(L1)Cl2] (1) showed effective nuclease activity in the presence of hydrogen peroxide, converting supercoiled plasmid DNA to its linear form.  相似文献   

11.
By applying the hydrothermal in situ acylation reactions between alkyl-substituted pyridine-2,3-dicarboxylic acids and hydrazine hydrate, six pyridine-monoacylhydrazidate-coordinated transition-metal complexes [Mn(MPDH)2] 1, [M(MPDH)2(H2O)2]·2H2O (M2+ = Co2+2, Zn2+3), [Mn(EPDH)2(H2O)2] 4 and [M(EPDH)2(H2O)2] (M2+ = Zn2+5, Co2+6) (MPDH = 6-methylpyridine-2,3-dicarboxylhydrazidate; EPDH = 5-ethylpyridine-2,3-dicarboxylhydrazidate) were obtained. Although only compound 1 is a 1-D chained coordination polymer and the others are the mononuclear molecular entities, they all further self-assemble into the interesting supramolecular networks via hydrogen-bonded interactions between pyridine-monoacylhydrazidate ligands. Two Zn2+ compounds 3 and 5 possess the fluorescence properties with maximum emissions at 517 nm for 3 and 530 nm for 5 upon excitation, respectively. The magnetic analysis for compound 1 indicates there exists the antiferromagnetic interactions between the Mn(II) ions.  相似文献   

12.
In alkaline aqueous solutions, 3,4-diaminobenzoate (H2(2LPDA)) reacts with PtII to form a 1:2 (Pt:L) complex that intensely absorbs near-infrared (NIR) light at 713 nm (ε = 8.0 × 104 M−1 cm−1). The absorption disappeared at pH < 3 (in DMSO), showing pH-responsive switching of the NIR absorption. By comparing the NIR-absorbing behavior of this complex to that of a complex, [PtII(1LISQ)2]2−, containing the analogous phenylenediamine ligand [(1LISQ)2− = o-diiminobenzosemiquinonate radical], the complex can be formulated as [PtII(2LISQ)2]2−. The assignment of the entity was consistent with the redox and spectroelectrochemical behaviors and electronic spin resonance (ESR) spectroscopy. First, one-electron oxidation of [PtII(2LISQ)2]2− formed an ESR-silent complex assignable to the dimeric complex [{PtII(2LISQ)(2LIBQ)}2]2− [(2LIBQ) = o-iminobenzoquinone form] in which the two radical centers at were antiferromagnetically coupled. Second, the one-electron reduced complex of [PtII(2LISQ)2]2− exhibited an ESR signal attributed to [PtII(2LISQ)(2LPDA)]3−; 34% of the electronic spin was located at the PtII center rather than on the moiety. The pH-responsive switching-off of the NIR absorption was thus rationally explained by oxidation of [PtII(2LISQ)2]2− to [{PtII(2LISQ)(2LIBQ)}2]2− by the increase of the rest potential of the solution in the lower pH region.  相似文献   

13.
1:1 and 1:2 cobalt complexes of bis(benzimidazol-2-ylmethyl)amine (bbma) bis(benzimidazol-2-ylmethyl)sulfide (bbms), bis(benzimidazol-2-ylethyl)sulfide (bbes) and diethylenetriamine (dien) were prepared and their spectral and redox behavior studied. Two geometrical isomers pink-[Co(bbes)2]2+ and blue-[Co(bbes)2]2+ were obtained when the complexes were prepared by using with bbes and they were separated manually and recrystallized. The octahedral structure of pink-[Co(bbes)2]2+ was resolved by X-ray analysis. The electronic spectra show the presence of two geometrical isomers for Co(bbes)22+ in the solid state; for example, the spectral bands of pink-[Co(bbes)2]2+ differs markedly with those of blue-[Co(bbes)2]2+. This is consistent with the results obtained from magnetic measurements (5.10 BM for pink-Co(bbes)22+ and 4.72 BM for blue-[Co(bbes)2]2+). Further, the behavior of the ligands (bbma, bbms, bbes) at different pH conditions was determined on the basis of 13C NMR studies. The redox potentials [Co(II)/Co(I)] of the complexes follow the trend [Co(bbma)2]2+ < [Co(bbms)2]2+ ≈ [Co(bbes)2]2+ which demonstrates the stabilization of the Co(II) ion is more by both weak σ-donor and weak π-acceptor ligands rather than by σ-donor ligand.  相似文献   

14.
Reactions of Au(III)-alkyldiamine complex with l-histidine and imidazole were carried out and monitored time-dependant by 1H and 13C NMR. Kinetics for the [Au(en)Cl2]+ reaction with l-histidine was determined by initial rate method at constant pH and 25 °C using UV-Vis absorption technique, and found to be first order with respect to each component, with a pseudo second order rate constant of 39 ± 3 M−1 s−1. Reaction rates of l-histidine and imidazole reactions with the [Au(en)Cl2]+ complex was found to be strongly dependant on pD. The pD also has profound effect on the stability of the complex. It was observed that concurrent redox reactions also take place in solution in which Au(III) is reduced to metallic Au(0), while l-histidine and imidazole are oxidized to oxy and hydroxyl products. The optimization of the structure of [(His)Au(en)]3+ complex was carried out by gaussian03 at the RB3LYP level that showed a distorted square pyramid with the histidine carboxyl group at the pyramid top.  相似文献   

15.
The ultrafast (< 100 fs) conversion of delocalized exciton into charge-separated state between the primary donor P700 (bleaching at 705 nm) and the primary acceptor A0 (bleaching at 690 nm) in photosystem I (PS I) complexes from Synechocystis sp. PCC 6803 was observed. The data were obtained by application of pump-probe technique with 20-fs low-energy pump pulses centered at 720 nm. The earliest absorbance changes (close to zero delay) with a bleaching at 690 nm are similar to the product of the absorption spectrum of PS I complex and the laser pulse spectrum, which represents the efficiency spectrum of the light absorption by PS I upon femtosecond excitation centered at 720 nm. During the first ∼ 60 fs the energy transfer from the chlorophyll (Chl) species bleaching at 690 nm to the Chl bleaching at 705 nm occurs, resulting in almost equal bleaching of the two forms with the formation of delocalized exciton between 690-nm and 705-nm Chls. Within the next ∼ 40 fs the formation of a new broad band centered at ∼ 660 nm (attributed to the appearance of Chl anion radical) is observed. This band decays with time constant simultaneously with an electron transfer to A1 (phylloquinone). The subtraction of kinetic difference absorption spectra of the closed (state P700+A0A1) PS I reaction center (RC) from that of the open (state P700A0A1) RC reveals the pure spectrum of the P700+A0 ion-radical pair. The experimental data were analyzed using a simple kinetic scheme: An* [(PA0)*A1 P+A0A1] P+A0A1, and a global fitting procedure based on the singular value decomposition analysis. The calculated kinetics of transitions between intermediate states and their spectra were similar to the kinetics recorded at 694 and 705 nm and the experimental spectra obtained by subtraction of the spectra of closed RCs from the spectra of open RCs. As a result, we found that the main events in RCs of PS I under our experimental conditions include very fast (< 100 fs) charge separation with the formation of the P700+A0A1 state in approximately one half of the RCs, the ∼ 5-ps energy transfer from antenna Chl* to P700A0A1 in the remaining RCs, and ∼ 25-ps formation of the secondary radical pair P700+A0A1.  相似文献   

16.
Cobalt(III) and rhodium(III) complexes of the series of [MIIICl3 − n(P)3 + n]n+ (M = Co or Rh; n = 0, 1, 2 or 3) have been prepared with the use of 1,1,1-tris(dimethylphosphinomethyl)ethane (tdmme) and mono- or didentate phosphines. The single-crystal X-ray analyses of both series of complexes revealed that the M-P and M-Cl bond lengths were dependent primarily on the strong trans influence of the phosphines, and secondarily on the steric congestion around the metal center resulting from the coordination of several phosphine groups. In fact, the M-P(tdmme) bonds became longer in the order of [MCl3(tdmme)] < [MCl2(tdmme)(PMe3)]+ < [MCl(tdmme)(dmpe)]2+ (dmpe = 1,2-bis(dimethylphosphino)ethane) < [M(tdmme)2]3+ for both CoIII and RhIII series of complexes, while the M-Cl bond lengths were shortened in this order (except for [M(tdmme)2]3+). Such a steric congestion around the metal center can also account for the structural and spectroscopic characteristics of the series of complexes, [MCl(tdmme)(dmpm, dmpe or dmpp)]2+ (dmpm = bis(dimethylphosphino)methane, dmpp = 1,3-bis(dimethylphosphino)propane). The X-ray analysis for [CoCl(tdmme)(dmpm or dmpe)](BF4)2 showed that all Co-P bonds in the dmpm complex were shorter by 0.03-0.04 Å than those in the dmpe complex. Furthermore, the first d-d transition energy of the CoIII complexes and the 1JRh-P(tdmme) coupling constants observed for the RhIII complexes indicated an unusual order in the coordination bond strengths of the didentate diphosphines, i.e., dmpm > dmpe > dmpp.  相似文献   

17.
Complexes of the type [Pt(amine)4]I2 were synthesized and characterized mainly by multinuclear (195Pt, 1H and 13C) magnetic resonance spectroscopy. The compounds were prepared with different primary amines, but not with bulky amines, due to steric hindrance. In 195Pt NMR, the signals were observed between −2715 and −2769 ppm in D2O. The coupling constant 3J(195Pt-1H) for the MeNH2 complex is 42 Hz. In 13C NMR, the average values of the coupling constants 2J(195Pt-13C) and 3J(195Pt-13C) are 18 and 30 Hz, respectively. The crystal structure of [Pt(EtNH2)4]I2 was determined by X-ray diffraction methods. The Pt atom is located on an inversion center. The structure is stabilized by H-bonding between the amines and the iodide ions. The compound with n-BuNH2 was found by crystallographic methods to be [Pt(n-BuNH2)4]2I3(n-BuNHCOO). The crystal contains two independent [Pt(CH3NH2)4]2+ cations, three iodide ions and a carbamate ion formed from the reaction of butylamine with CO2 from the air. When the compound [Pt(CH3NH2)4]I2 was dissolved in acetone, crystals identified as trans-[Pt(CH3NH2)2(H3CNC(CH3)2)2]I2 were isolated and characterized by crystallographic methods. Two trans bonded MeNH2 ligands had reacted with acetone to produce the two N-bonded Schiff base Pt(II) compound.  相似文献   

18.
Salt marshes near urban, industrial and mining areas are often affected both by heavy metals and by eutrophic water. The aim of this study was to assess and evaluate the main processes involved in the decrease of nitrate concentration in pore water of mine wastes flooded with eutrophic water, considering the presence or absence of plant rhizhosphere. Basic (pH ∼ 7.8) carbonated loam mine wastes and free-carbonated acidic (pH ∼ 6.2) sandy-loam mine wastes were collected from polluted coastal salt marshes of SE Spain which regularly receive nutrient-enriched water. The wastes were put in pots and flooded for 15 weeks with eutrophic water (dissolved organic carbon ∼26 mg L−1, PO43− ∼23 mg L−1, NO3 ∼180 mg L−1). Three treatments were assayed for each type of waste: pots with Sarcocornia fruticosa, pots with Phragmites australis and unvegetated pots. Soluble organic carbon, nitrate, soluble Cd, Pb and Zn, pH and Eh were monitored. But the 2nd day of flooding, nitrate concentrations had decreased between 70% and 90% (equivalent to 1.01-1.12 g N-NO3 m−2 day−1) with respect to the content in the water used for flooding, except in unvegetated pots with acidic wastes. Denitrification was the main mechanism associated with the removal of nitrate. The role of vegetation in improving the rhizospheric environment was relevant in the acidic wastes because higher sand content, lower pH and higher soluble metal concentrations might strongly hinder microbial activity Hence, revegetation of salt marshes polluted by acidic sandy mining wastes might improve the capacity of this type of environment to act as a green filter against excessive nitrate contents flowing through them.  相似文献   

19.
20.
In an unusual reaction of [RuIII(acac)2(CH3CN)2](ClO4) ([1], acac = acetylacetonate) and aniline (Ph-NH2), resulted in the formation of ortho-semidine due to dimerisation of aniline via oxidative ortho-Carom-N bond formation reaction. This oxidation reaction is associated with stepwise chlorination of coordinated acac ligands at the γ-carbon atom resulting in the formation of [RuIII(acac)2L] [2a], [RuIII(Cl-acac)(acac)L] [2b], [RuIII(acac)(Cl-acac)L] [2c] and [RuIII(Cl-acac)2L] [2d] (L = N-phenyl-ortho-semiquinonediimine) complexes, respectively. These have been characterized by 1H NMR, UV-Vis-NIR, ESI-MS and cyclic voltammetry studies. Single crystal X-ray structures of 2c and 2d are reported. Crystallographic structural bond parameters of 2c and 2d revealed bond length equalization of C-C, C-O and M-O bonds. It has been shown that perchlorate () counter anion, present in the starting ruthenium complex, acts as the oxidizing agent in bringing about oxidation of Ph-NH2 to ortho-semidine. The chloronium ions, produced in situ, chlorinate the coordinated acac ligands at the γ-carbon atom. Such electrophilic substitution of coordinated acac ligands indicates that the Ru-acac metallacycles in the reference compounds are aromatic. The complexes showed an intense and featureless band centered near 520 nm, and a structured band near 275 nm. These displayed one reversible cathodic response in the range, −1.1 to −0.8 V and one reversible anodic response between 0.4 and 0.6 V versus the Saturated Calomel reference Electrode, SCE. The response at the anodic potential is due to oxidation of the coordinated ligand L, while the reversible response at cathodic potential is due to reduction of the metal center.  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