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.
Cyclic and square wave voltammetry (−1500 to 1500 mV) of {Mn[μ-(C6H5)2NC4O3]2[H2O]4}n [manganese(II) diphenylaminosquarate] (1) and [Mn(μ-C6H5C4O3)(C6H5C4O3)(H2O)3]n [manganese(II) phenylsquarate] (2) at a gold disk electrode in dimethylsulfoxide (DMSO) and dimethylformamide (DMF), reveal several couples attributable to both ligand and metal-based redox processes. For the manganese(II) phenylsquarate in DMF, the metal-based peaks are more numerous and readily discernible than in DMSO. In either of the solvents, the ligand-based peaks always occur at more positive or more negative potentials than the metal-based ones. In 1 and 2, Mn(II)/Mn(0), Mn(III)/Mn(II), Mn(IV)/Mn(III) and Mn(V)/Mn(IV) couples are observed. However, the manganese redox peaks appear at more negative potentials in 1.  相似文献   

2.
Aspects of manganese metabolism during normal and acidogenic growth of Aspergillus niger were explored. Arginase from this fungus was a Mn[II]-enzyme. The contribution of the arginase protein towards A. niger manganese metabolism was investigated using arginase knockout (D-42) and arginase over-expressing (ΔXCA-29) strains of A. niger NCIM 565. The Mn[II] contents of various mycelial fractions were found in the order: D-42 strain < parent strain < ΔXCA-29 strain. While the soluble fraction forms 60 % of the total mycelial Mn[II] content, arginase accounted for a significant fraction of this soluble Mn[II] pool. Changes in the arginase levels affected the absolute mycelial Mn[II] content but not its distribution in the various mycelial fractions. The A. niger mycelia harvested from acidogenic growth media contain substantially less Mn[II] as compared to those from normal growth media. Nevertheless, acidogenic mycelia harbor considerable Mn[II] levels and a functional arginase. Altered levels of mycelial arginase protein did not significantly influence citric acid production. The relevance of arginase to cellular Mn[II] pool and homeostasis was evaluated and the results suggest that arginase regulation could occur via manganese availability.  相似文献   

3.

Aims

The objectives were to investigate (i) the forms and release pattern of P from an ash-rich biochar-amended sandy soil; (ii) the transformation of biochar P in a soil-plant system.

Methods

Several methodologies (a bioassay test, soluble P extractions, a sequential P fractionation and successive P extractions via resin strips) were used to study the bioavailability and transformation of P in a sandy soil fertilised with either conventional P fertilisers [Ca(H2PO4)2 (CaP) and Sechura phosphate rock (SPR)] or biochars produced from cattle manure (MAe) and alum-treated biosolids (BSe) at four temperatures (250, 350, 450, and 550 °C).

Results

Biochar P mainly contributed to increase soil resin-extractable P- and inorganic NaOH-extractable P-fractions, and thus to plant available P. The decrease in P concentrations of those fractions was caused by the uptake of P by plants rather than their transformations into more stable forms. P release rates diminished following the order: CaP > MAe > BSe > SPR, which indicates a decline in P availability from these P sources.

Conclusions

Phosphorus-rich biochar can be used as a slow-release fertiliser. It is necessary to determine available P (either soil or fertiliser tests) in biochars prior to its application to soil, so that dose, frequency and timing of application are correctly established.  相似文献   

4.
Summary The availability of manganous oxide, manganese dioxide and MnEDTA was compared to that of manganous sulfate at several equivalent rates on a Mn-deficient peaty muck, marl and on Mn-sufficient Scranton sand. Tomato plants grown in pot studies were analyzed for manganese and iron. Mn-deficiency leaf symptoms were in general agreement with the manganese level in the plant and the Fe/Mn ratio. Iron intake and manganese intake showed some evidence of inverse relationships. Manganous sulfate was slightly more available than manganous oxide which, in turn, was more available than MnO2 or MnEDTA.  相似文献   

5.
A new family of tridentate ligands PhimpH (2-((2-phenyl-2-(pyridin-2-yl)hydazono)methyl)phenol), N-PhimpH (2-((2-phenyl-2-(pyridin-2-yl)hydrazono)methyl)napthalen-1-ol), Me-PhimpH (2-(1-(2-phenyl-2-(pyridine-2-yl)hydrazono)ethyl)phenol) have been synthesized and characterized. The ligands PhimpH and N-PhimpH after deprotonation react with manganese(II) and manganese(III) starting materials affording [Mn(Phimp)2] (1), [Mn(Phimp)2](ClO4) (2), [Mn(N-Phimp)2] (3), [Mn(N-Phimp)2](ClO4) (4). Complexes [Mn(Phimp)2] (1) and [Mn(N-Phimp)2] (3) convert to [Mn(Phimp)2]+ (cation of 2) and [Mn(N-Phimp)2]+ (cation of 4) respectively upon oxidation. Ligand Me-PhimpH stabilized only manganese(III) centre resulting [Mn(Me-Phimp)2](ClO4) (5). The molecular structures of [Mn(Phimp)2], 1 and [Mn(Phimp)2](ClO4), 2 were determined by single crystal X-ray diffraction. X-ray crystal structures of 1 and 2 have revealed the presence of distorted octahedral MnN4O2 coordination sphere having meridionally spanning ligands. Electrochemical studies for the complexes showed Mn(II)/Mn(III), (E1/2 = 0.14-0.40 V) and Mn(III)/Mn(IV), (E1/2 = 0.80-1.06 V) couples vs. Ag/AgCl. The redox properties were exploited to examine superoxide dismutase (SOD) activity using Mn(II)/Mn(III) couple. The complexes 1, 2, 4 and 5 have been revealed to catalyze effectively the dismutation of superoxide () in xanthine-xanthine oxidase-nitro blue tetrazolium assay and IC50 values were found to be 0.29, 0.39, 1.12 and 0.76 μM respectively. DNA interaction studies with complex 2 showed binding of DNA in a non-intercalative pathway. Complexes 1, 2 and 4 exhibited nuclease activity in presence of H2O2 and inhibition of activity was noted in presence of KI.  相似文献   

6.
Reactions of the structural model hydrolases [M2(OAc)4(H2O)(Im)4]; M=Mn (E); M=Co (D); M=Ni (B) and [M2(OPiv)4(H2O)(tmen)2]; M=Mn (E″); M=Co (D″); M=Ni (B″) with a number of hydroxamic acids, RHA (aceto- (R=CH3), benzo- (R = C6H5) and N-phenylacetohydroxamic acid (NPhAHA)) gave a series of hydroxamate dibridged complexes [M2(OAc)(RA)2(Im)4][OTf] and [M2(OPiv)(RA)2(tmen)2][OTf]; M=Co, Ni, in which the bridging hydroxamates exhibit a novel bonding mode in which the deprotonated hydroxamate hydroxyl bridges the two metal centres only. The formation of this type of structure by NPhAHA is the first example involving a secondary hydroxamic acid. These complexes are good structural models of the acetohydroxamate-inhibited C319A variant of Klebsiella aerogenes urease (KAU) and their structures are close to those previously reported for complexes containing tmen capping ligands. Reaction with glutarodihydroxamic acid leads to hydroxylamine elimination and formation of a dimer containing deprotonated N-hydroxyglutarimide as bridging ligand but in this case the structure contains pentacoordinated Co(II) and only one bridging acetate in contrast to the tmen-based series where the analogous complex contains hexacoordinated Co(II) and two bridging acetates. Reaction of [Mn2(OAc)2(μ-OAc)2(μ-H2O)(tmen)2] with acetohydroxamic acid (AHA) gave the first structurally characterized manganese hydroxamate, [Mn2(OAc)3(AA)(tmen)2] with the same bridging/chelating mode of hydroxamate bonding as in the analogous cobalt and nickel complexes, although only one bridging hydroxamate occurs in the manganese complex in contrast to the two bridging hydroxamates in the cobalt and nickel complexes. The isolation of the dimanganese hydroxamate bridged complex suggests that hydroxamic acids may also inhibit the dimanganese based metallohydrolase, arginase.  相似文献   

7.
The ligand bis(1-methylimidazol-2-yl)ketone (bik) (1) was applied in the synthesis of mononuclear manganese(II) complexes. The complexes [Mn(bik)2Cl2] (2), [Mn(bik)2(OH2)Br]Br × H2O (3b) and [Mn(bik)3](ClO4) (4) were characterised by X-ray crystallography, ESR and UV-Vis methods.  相似文献   

8.
Boris K. Semin  Michael Seibert 《BBA》2006,1757(3):189-197
The role of carboxylic residues at the high-affinity, Mn-binding site in the ligation of iron cations blocking the site [Biochemistry 41 (2000) 5854] was studied, using a method developed to extract the iron cations blocking the site. We found that specifically bound Fe(III) cations can be extracted with citrate buffer at pH 3.0. Furthermore, citrate can also prevent the photooxidation of Fe(II) cations by YZ. Participation of a COOH group(s) in the ligation of Fe(III) at the high-affinity site was investigated using 1-ethyl-3-[(3-dimethylamino)propyl] carbodiimide (EDC), a chemical modifier of carboxylic amino acid residues. Modification of the COOH groups inhibits the light-induced oxidation of exogenous Mn(II) cations by Mn-depleted photosystem II (PSII[−Mn]) membranes. The rate of Mn(II) oxidation saturates at ≥10 μM in PSII(−Mn) membranes and ≥500 μM in EDC-treated PSII (−Mn) samples. Intact PSII(−Mn) membranes have only one site for Mn(II) oxidation via YZ (dissociation constant, Kd = 0.64 μM), while EDC-treated PSII(−Mn) samples have two sites (Kd = 1.52 and 22 μM; the latter is the low-affinity site). When PSII(−Mn) membranes were incubated with Fe(II) before modifier treatment (to block the high-affinity site) and the blocking iron cations were extracted with citrate (pH 3.0) after modification, the membranes contained only one site (Kd = 2.3 μM) for exogenous Mn(II) oxidation by YZ radical. In this case, the rate of electron donation via YZ saturated at a Mn(II) concentration ≥15 μM. These results indicate that the carboxylic residue participating in Mn(II) coordination and the binding of oxidized manganese cations at the HAZ site is protected from the action of the modifier by the iron cations blocking the HAZ site. We concluded that the carboxylic residue (D1 Asp-170) participating in the coordination of the manganese cation at the HAZ site (Mn4 in the tetranuclear manganese cluster [Science 303 (2004) 1831]) is also involved in the ligation of the Fe cation(s) blocking the high-affinity Mn-binding site.  相似文献   

9.
A series of four mononuclear manganese (II) complexes with the N-tridentate neutral ligands 2,2:6,2′′-terpyridine (terpy) and N,N-bis(2-pyridylmethyl)ethylamine (bpea) have been synthesized and crystallographically characterized. The complexes have five- to seven-coordinate manganese(II) ions depending on the additional ligands used. The [Mn(bpea)(Br)2] complex (1) has a five-coordinated manganese atom with a bipyramidal trigonal geometry, while [Mn(terpy)2](I)2 (2) is hexa-coordinated with a distorted octahedral geometry. Otherwise, the reactions of Mn(NO3)2 · 4H2O with terpy or bpea afforded novel seven-coordinate complexes [Mn(terpy)(NO3)2(H2O)] (3) and [Mn(bpea)(NO3)2] (4), respectively. 3 has a coordination polyhedron best described as a distorted pentagonal bipyramid geometry with one nitrate acting as a bidentate chelating ligand and the other nitrate as a monodentate one. 4 possesses a highly distorted polyhedron geometry with two bidentate chelating nitrate ligands. These complexes represent unusual examples of structurally characterized complexes with a coordination number seven for the Mn(II) ion and join a small family of nitrate complexes.  相似文献   

10.
Two six-coordinated manganese(II) complexes [Mn(pydien)Cl](ClO4) · C2H5OH (1), [Mn(pydien)NCS](ClO4) (2) and two seven-coordinated manganese(II) complexes [Mn(pydado)Cl](ClO4) (3), [Mn(pydado)NCS](ClO4) (4) have been obtained using linear penta and hexadentate ligands pydien and pydado (pydien: 1,7-bis(2-pyridylmethyl)-1,4,7-triazaheptane and pydado: 1,10-bis(pyridylmethyl)-1,10-diaza-4,7-dioxadecane). The crystal structures for all compounds have been determined. 1 and 3 crystallize in the triclinic space group , 2 crystallizes in the orthorhombic space group Pbca, whereas 4 crystallizes in the monoclinic space group P21/c. The bound anion (chloro or isothiocyanato) in complexes 1 and 2 has no influence on the geometry of six-coordinate manganese(II) complexes, whereas the geometry and the wrapping of the hexadentate ligand (pydado) around Mn2+ cation depend on the nature of the bound anion. The complex 3 has a capped octahedron geometry with the two pyridyl groups in trans position, while the geometry of complex 4 can be described as pentagonal bipyramid with one pyridyl group and a thiocyanate anion in the axial positions.  相似文献   

11.
1,10-Phenanthroline hydrogen phthalato manganese(II) dimer [Mn2(Hphth)2(phen)4] · 2Hphth · 6H2O (1), monomeric phenanthroline phthalato manganese(II) monomer [Mn(phth)(phen)2(H2O)] · 2.5H2O (2), 2,2′-bipyridine phthalato manganese(II) polymer [Mn(phth)(bpy)(H2O)2]n (3) and 1,10-phenanthroline maleato polymer [Mn(male)(phen)(H2O)2]n · 2nH2O (4) (H2phth = o-phthalic acid, male = maleic acid, phen = 1,10-phenanthroline and bpy = 2,2′-bipyridine) have been synthesized and characterized spectroscopically and structurally. Each Mn(II) atom in dimeric 1 is octahedrally coordinated by two oxygen atoms of phthalate anions and by two cis-phenanthroline ligands. The hydrogen phthalato anion bridges the Mn(II) ions through the deprotonated carboxyl groups, while the carboxylic acid group remains free. In the monomeric 2, the Mn(II) ion is octahedrally surrounded by four nitrogen atoms from two cis-phen ligands, one carboxyl oxygen from a monodentate phth ion, and one coordinated water molecule. The dimeric phthalato complex 1 can be cleaved into monomer 2 under heating with deprotonation, and the course of the reaction can be qualitatively traced by IR spectra. The phthalate group in the complex 3 binds to two manganese atoms through the vicinal carboxyl-oxygen atoms in syn-syn bridging mode. The Mn(II) atoms are linked by the phthalate group to yield a one-dimensional chain running along the a-axis. The coordination polymer 3 can be obtained from the reaction of dichloro dibipyridine manganese with phthalate under heating. In polymer 4, the manganese atom is six-coordinated by two nitrogen atoms from phen, two oxygen atoms from the coordinated water molecules and two oxygen atoms from two different maleate dianions. Each maleato unit links two neighboring manganese atoms to yield one-dimensional chain along b-axis in bis-monodentate mode. The single-chain polymer 4 prepared at low temperature can be converted to double-chain coordination polymer [Mn(male)(phen)]n · nH2O (5) with dehydration in warm solution.  相似文献   

12.
Successful revegetation of bauxite residue sand (BRS) requires large inputs of nutrients such as manganese (Mn), yet Mn deficiency is still encountered, raising doubts about sustainable revegetation of BRS disposal areas. The application of deep placement of Mn, a measure common in agriculture, was examined as a method for improving productivity and sustainability when lucerne (Medicago sativa L.) is used as a species for BRS revegetation. In pots containing BRS, Mn was banded at 2.5-, 10- and 20-cm depths at rates of 10, 20 and 50 g g–1 BRS. Two lucerne genotypes used were Salado, a Mn-deficiency-tolerant variety, and Sirosal, a Mn-deficiency-sensitive variety. Banding at 10-cm depth produced the best shoot growth of Sirosal at each Mn rate. Greatest shoot growth in Salado was found at 2.5-, 10- and 20-cm depths for 10, 20 and 50 g Mn g–1 BRS, respectively. Deep banding 20 g Mn g–1 BRS at 10-cm depth significantly increased lucerne growth compared with mixing through the profile. Banding at 20 cm produced Mn deficiency symptoms in lucerne during early growth, but symptoms were alleviated when sufficient amounts of roots proliferated in the banding zone. Dissolution and movement of Mn away from the fertiliser band were also investigated. In pots without plants, water throughput from watering twice weekly to 110% field capacity had no effect on the amount of extractable Mn at distances more than 1 cm away from the original Mn band position. Whilst not only providing a more effective supply of Mn for BRS revegetation over one growth period, deep-banding of adequate rates of Mn may also result in a longer residual value, reducing the need for frequent broadcast applications.  相似文献   

13.

Background and aims

We sought to describe the species and functional composition of Brazilian campos rupestres plant communities on severely nutrient-impoverished white sands, to test hypotheses relating plant communities and physiological adaptations to infertile soils. Based on recently-published information on a south-western Australian dune chronosequence, we hypothesised that campos rupestres plant communities would similarly contain a relatively large proportion of non-mycorrhizal species, because of the phosphorus-(P) impoverished nature of the soils. We also sought to test the hypothesis that many of these non-mycorrhizal species have high leaf manganese (Mn) concentrations as a consequence of carboxylate exudation to mobilise soil P.

Methods

We conducted flora surveys and quantified mycorrhizal status and foliar Mn concentrations in field sites with strongly-weathered sandy soils. Rhizosphere carboxylates were collected from glasshouse-grown plants to assess a potential correlation of carboxylates and leaf Mn concentrations.

Results

Soils were depleted of all major plant nutrients. Non-mycorrhizal plants were abundant in most field sites (mean relative cover = 48%). Vellozia species were dominant aboveground; belowground, roots were colonised more by dark septate endophytic fungi than by mycorrhizal fungi. From the field sites, foliar Mn concentrations in non-mycorrhizal species increased with decreasing soil P concentrations, but only when soil Mn concentrations were above a minimum threshold (exchangeable [Mn] above detection limit). Across all species, however, there was no relationship of foliar Mn concentrations with soil P concentrations.

Conclusions

Our hypothesis that white-sand campos rupestres communities contain a relatively large proportion of non-mycorrhizal plants was supported. Comparison with similar ecosystems in south-western Australia suggests that plant communities on severely P-impoverished sandy soils, despite differing evolutionary histories and little overlap in plant families, follow convergent evolutionary paths towards increasing abundance of non-mycorrhizal species.
  相似文献   

14.
Manganese (Mn) is an essential nutrient that can be toxic in excess concentrations, especially during early development stages. The mechanisms of Mn toxicity is still unclear, and little information is available regarding the role of Mn speciation and fractionation in toxicology. We aimed to investigate the toxic effects of several chemical forms of Mn in embryos of Danio rerio exposed during different development stages, between 2 and 122 h post fertilization. We found a stage-specific increase of lethality associated with hatching and removal of the chorion. Mn(II), ([Mn(H2O)6]2+) appeared to be the most toxic species to embryos exposed for 48 h, and Mn(II) citrate was most toxic to embryos exposed for 72 and/or 120 h. Manganese toxicity was associated with calcium disruption, manganese speciation and metal fractionation, including bioaccumulation in tissue, granule fractions, organelles and denaturated proteins.  相似文献   

15.
The complex [Mn(mesalim)2Cl] (1), (Hmesalim = methyl salicylimidate) has been synthesized and fully characterized. The manganese(III) complex is formed by the reaction of the ligand Hmesalim with manganese(II) chloride. Complex 1 is mononuclear and crystallizes in the space group . Electrochemical studies were performed for complex 1, as well as for the related complexes [Mn(mesalim)2(OAc)(MeOH] · MeOH (2) and [Mn2(etsalim)4(Hetsalim)2](ClO4)2 (3), (Hetsalim = ethyl salicylimidate). The complexes display intricate oxidation-reduction behaviour, and coulometric analyses in combination with electrochemical analyses have been used to understand the electron transfer mechanisms occurring at the electrodes.  相似文献   

16.

Aims

The thiosulphate induced accumulation of mercury by the three plants Brassica juncea var.LDZY, Brassica juncea var.ASKYC and Brassica napus var. ZYYC and the transformation of mercury fractionation in the rhizosphere of each plant was investigated in the field.

Methods

Experimental farmland was divided into control and thiosulphate plots. Each plot was divided into three subplots with each planted with one of the plants. After harvesting, the mercury concentration in plants, mercury fractionation in rhizosphere soil before and after phytoextraction, and the vertical distribution of bioavailable mercury in bulk soil profiles was analyzed.

Results

The cultivar B. juncea var.LDZY accumulated a higher amount of mercury in shoots than the other two plants. Thiosulphate treatment promoted an increase in the concentration of metal in plants and a transformation of Fe/Mn oxide-bound and organic-bound mercury (potential bioavailable fractions) into soluble and exchangeable and specifically-sorbed fractions in the rhizosphere. The observed increase in bioavailable rhizosphere mercury concentration was restricted to the root zone; mercury did not move down the soil profile as a function of thiosulphate application to soil.

Conclusions

Thiosulphate-induced phytoextraction has the potential to manage environmental risk of mercury in soil by decreasing the concentration of mercury associated with potential bioavailable fraction that can be accumulated by crop plants.  相似文献   

17.
Iwasaki  Kōozō  Maier  Peter  Fecht  Marion  Horst  Walter J. 《Plant and Soil》2002,238(2):281-288
The effects of silicon (Si) supply on manganese (Mn) toxicity symptoms and Mn and Si concentrations in the leaf apoplast in a Mn-sensitive cowpea cultivar (Vigna unguiculata (L.) Walp. cv. TVu 91) were investigated in solution culture experiments. When 1.44 mM Si was supplied concurrently with 50 M Mn, the Mn toxicity symptoms were clearly avoided without decreasing the total Mn concentration. On the other hand, the symptoms were not completely alleviated when the plants were pretreated with 1.44 mM Si and then exposed to 50 M Mn without concurrent Si supply. Plants of both of these treatments exhibited lower Mn concentrations in the apoplastic washing fluids but higher amounts of adsorbed Mn on the cell walls than the plants treated with 50 M Mn without Si supply. However, the difference in Mn concentration between plants with continuous and interrupted Si supply was not significant. Moreover, the Mn concentration in the apoplastic washing fluids of the plants with continuous supply of 1.44 mM Si and 50 M Mn and not showing Mn toxicity symptoms was higher than that of the plants grown at 10 M Mn without Si supply which showed distinct Mn toxicity symptoms. These results show that Si supply alleviates Mn toxicity not only by decreasing the concentration of soluble apoplastic Mn through the enhanced adsorption of Mn on the cell walls. A role of the soluble Si in the apoplast in the detoxicification of apoplastic Mn is indicated.  相似文献   

18.
Two new complexes, [Cu(mamba)2] and [Mn(mamba)2] (mamba, N-(2-methylpyridine)-2-aminomethyl benzoate) were synthesized and characterized by X-ray crystallography. Whereas the [Cu(mamba)2] complex crystallizes in a monoclinic P21/c space group, the [Mn(mamba)2] complex crystallizes in a triclinic space group. The nature of the metal ion greatly influences the lattices and the molecular structures of the compounds. In the crystal lattice of the copper complex are four cocrystallized methanol solvent, which are all involved in building six strong H-bonds with the complex. However, the lattice for the manganese complex contain only one cocrystallized methanol, along with one NaClO4, that is also involved in making one H-bond with the [Mn(mamba)2] unit. Nevertheless, the sodium ion is coordinated to the ClO4, the methanol and two [Mn(mamba)2] to form a stable extended chain metal complex. Electrochemical studies indicated that both complexes undergo quasi reversible one electron reduction in acetonitrile.  相似文献   

19.
Two hexacoordinated dinuclear compounds [Mn(L1)(dca)]2(ClO4/PF6)2·CH3OH (1/2) and two heptacoordinated coordination polymers [Mn(L2)(dca)]n(ClO4/PF6)n (3/4) [L1 = N,N′-(bis-(pyridin-2-yl)benzylidene)-1,3-propanediamine; L2 = N,N′-(bis-(pyridin-2-yl)benzylidene)diethylenetriamine; dca = dicyanamide] are synthesized and characterized. Structures of 1-3 have been solved by X-ray diffraction measurements. Each manganese(II) center in 1/2 is located in a distorted octahedral environment with an MnN6 chromophore coordinated by the four N atoms of L1 and two nitrile N atoms of bibridged μ1,5 dca. Interestingly, the coordination polymer 3 forms a 1D chain through single Mn-(NCNCN)-Mn units in which each manganese(II) center adopts a pentagonal bipyramidal geometry with an MnN7 chromophore occupied with five N atoms of L2 and two nitrile N atoms of monobridged μ1,5 dca. Magnetic susceptibility measurements of 1-3 in the 2-300 K temperature range reveal weak antiferromagnetic interactions.  相似文献   

20.
A family of four new phenoxo-bridged binuclear manganese(III) complexes of the general formula, [Mn(L)(X)]2 where L = [N,N′-bis(salicylidene)]propane-1,2-diamine and X = salicylaldehyde anion (sal) (1); NCS (2); NCO (3) and [Mn(L′)(N3)]2·2C2H5OH (4) where L′ = [N,N′-bis(2-hydroxyacetophenylidene)]propane-1,2-diamine has been prepared. The syntheses have been achieved by reacting manganese perchlorate with 1,2-diaminopropane and salicylaldehyde (or 2-hydroxyacetophenone for 4) or along with the respective pseudohalides so that the tetradentate Schiff base H2L or H2L′ is obtained in situ to bind the Mn(III) ion. The complexes have been characterized by IR spectroscopy, elemental analysis, crystal structure analysis and variable-temperature magnetic susceptibility measurements. The single crystal X-ray diffraction studies show that the compounds are isostructural containing dimeric Mn(III) units with bridging phenolate oxygen atoms. Low temperature magnetic studies indicate that the complexes 1-3 exhibit intradimer ferromagnetic exchange as well as single-molecule magnet (SMM) behavior while complex 4 is found to undergo an intradimer antiferromagnetic coupling.  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