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.
A spectroscopic investigation of the products formed in the reaction of emeraldine base (EB-PANI) with copper(II) ions in dimethylacetamide (DMA) is presented. It is well known that metal cations can dope emeraldine base polyaniline (EB-PANI) through a pseudo-protonation reaction. Resonance Raman, UV–vis-NIR, and EPR data, obtained for Cu2+/EB-PANI solutions prepared using CuCl2·2 H2O, Cu(NO3)2· 3 H2O or Cu(CH3COO)2·H2O as Cu2+ sources, showed that the species formed in reactions of EB-PANI and Cu2+ ions are dependent on the anions of the copper salt employed. EPR spectra pointed out that the environments of Cu2+ ions with acetate, chloride or nitrate as anions in DMA solution are distinct. Resonance Raman and UV–vis-NIR data demonstrated that the main reactions are the oxidation of EB-PANI to pernigraniline base (PB-PANI) and doping of EB-PANI to ES-PANI (emeraldine salt) when a direct coordination of Cu2+ ions to PANI exists. With nitrate as very weak coordinating anion, ES-PANI is formed preferentially. When copper chloride is used, both oxidation and doping of EB-PANI are verified. Conversely with acetate, the dimeric cage structure of this copper salt is preserved in solution, and oxidation of EB-PANI to PB-PANI is the only observed reaction. These results demonstrate the possibility of modulating the products of reaction between Cu2+ ions and EB-PANI in DMA solution by changing the counter ion of the Cu2+ source.  相似文献   

2.
Herein, we demonstrated miniature solid-contact ion-selective electrodes (ISEs) using a commercial mesoporous carbon black (MCB) as ion-to-electron transducer. MCB is attractive in its high surface area, good conductivity, relative low cost and availability. ISEs for potassium (K+) and nitrate (NO3) ions were prepared by subsequently coating the sealed glass capillaries (1.5 mm) with MCB and ion-selective membranes. Addition of MCB substantially stabilized electrode response by providing adequate double-layer capacitance and lowering resistance by more than 100× compared to the coated-wire electrodes. The electrodes exhibited near-Nernstian slopes of 59.6 (K+ ionophore), 57.8 (K+ ion-exchanger) and −54.8 (NO3 ion-exchanger) with standard solutions in the range of 10−5 to 10−1 M. Fast response (∼10 s) and reproducible sensitivities were also obtained in a mixed electrolyte containing interfering ions, although with a baseline drift of 2–10 mV/day in the long term. Importantly, the electrodes can be simply stored in air between measurements and used directly without conditioning in solutions. With simple fabrication and free maintenance, these sensors offer a low cost and convenient alternative to bulk ISEs, especially when sample volumes or space are limited.  相似文献   

3.
The solutions containing one of the copper salts (CuCl2, Cu(ClO4)2, Cu(NO3)2, and CuSO4) and one of the non-steroidal anti-inflammatory drugs (NSAIDs, ibuprofen, ketoprofen or naproxen) were analyzed by electrospray ionization mass spectrometry. Three of the salts, namely CuCl2, Cu(ClO4)2 and Cu(NO3)2, yielded binuclear complexes of drug:metal stoichiometry 1:2. Existence of the complexes of such stoichiometry has not been earlier observed. For copper(II) chloride the complexes (ions of the type [M-HCOOH+Cu2Cl]+ and [M+Cu2Cl]+, M stands for the drug molecule) were formed in the gas phase. When copper(II) perchlorate or copper(II) nitrate was used, the observed binuclear copper complexes (ions of the type [M-H+Cu2(ClO4)2+CH3OH]+, [M-H+Cu2(ClO4)2]+ and [M-H+Cu2(NO3)2+CH3OH]+, [M-H+Cu2(NO3)2]+) were observed at low cone voltage, thus these complexes must have already existed in the solution analysed. Therefore, such complexes may also exist under physiological conditions.   相似文献   

4.
The electrical conductivity of polyacetylene, prepared as a free-standing film, changes drastically when treated with aqueous solutions of nitrate in the presence of sulphuric acid (5.7 M). Linear calibration plots are obtained for 2–40 × 10?3 M nitrate. Common non-oxidating ions do not interfere.  相似文献   

5.
The interaction between lanthanide ions LnIII (Ln = La, Nd, Sm–Dy, Er, Yb) and nitrate ions is investigated by FT-IR spectroscopy in dilute anhydrous MeCN solution. The work is performed for ratios R = [NO]t/[LnIII]t ranging from 0 to 8 and for solutions generally 0.05M in LnIII, prepared from anhydrous lanthanide perchlorates Ln(ClO4)3. When nitrate is progressively added to the Ln(ClO4)3 solutions, the formation of [Ln(NO3)n](3?n)+ species is clearly evidenced by the FT-IR spectra. All the NO3? ions are coordinated and bidentate. A quantitative study was performed using the v1 and v6 vibrational modes for coordinated NO ions. The average coordination numbers estimated for Nd, Eu, Tb, and Er in solutions of trinitrates are 9.0, 9.1, 8.3 and 8.2, respectively (±0.3 unit). In presence of an excess NO, these numbers become 9.8, 10.2, 10.0, 9.8, 9.9, and 9.9 (±0.3 unit) for La, Nd, Eu, Tb, Er, and Yb, respectively. No hexanitrato species forms under the experimental conditions used (R up to 8). The structural aspect of the various nitrato species is also investigated. In the pentanitrato species, all the ligands appear to be equivalent, while large inequivalences are observed for Ln(NO3)3 solutions. Since for the latter most of the absorption bands assigned to nitrate vibrations contain several components, a curve-fitting procedure has been used for decomposing the v2, v4 and v6 vibrations. There is a considerable difference between LnIII ions, the nitrate inequivalences being larger in the middle of the series.  相似文献   

6.
《中国化学会会志》2018,65(8):982-988
CuAg nanoparticles (CuAgNPs) were electrochemically formed in situ on pre‐anodized, screen‐printed carbon electrodes (SPCEs) that possessed many oxygen‐containing functional groups capable of adsorbing metal ions, namely Cu2+ and Ag+. Pre‐anodization was achieved using continuous cyclic voltammetry in the range of potential 0.3–2.0 V under a scan rate of 50 mV/s. Cu2+ and Ag+ ions were adsorbed on the pre‐anodized SPCE by immersing the electrode in solutions containing both metal ions, and then CuAgNPs were formed in situ via electrochemical reduction in a deaerated, neat NaClO4 solution after the electrode was ultrasonicated to remove physically adsorbed metal ions. Although CuNPs showed higher activity than AgNPs toward both nitrate (NO3) and nitrite (NO2) ions, the instability of CuNPs hindered the application, so CuAgNPs were employed to achieve a compromise between sensitivity and stability. The SPCE/anodized/CuAgNP electrodes showed activity toward the electrochemical reduction of NO3 and NO2, respectively, with the limit of detection (LOD) of 15.6 μM (0.97 ppm) and 11.1 μM (0.51 ppm), which is sufficient to fit the allowed values (50 and 3 ppm, respectively) in drinking water as suggested by the World Health Organization (WHO).  相似文献   

7.
Abstract

The hydrated metal nitrates (M(NO3)3.6H2O, M[dbnd]Co, Ni, Cu, Zn and Cd) have been crystallised from water in the presence of 18-crown-6 and their structures determined by X-ray crystallography. In the case of copper, a pseudo four-coordinate square planar complex resides in an extended six-coordinate octahedral array which is further bound in a single-stranded one-dimensional hydrogen bonded polymeric mode. For M[dbnd]Co,Ni,Zn and Cd isomorphous complexes are isolated where the octahedral [M(H2O)5(NO3)+ cation resides in a two-dimensional polymeric network through hydrogen bonds between the water ligands and either the crown ether oxygens or unbound nitrate ions or water molecules.  相似文献   

8.
Carrier-free90Y isotope was used to study the distribution between the aqueous solution of nitric acid and sodium nitrate and HDOP in cyclohexane. The distribution coefficients (D) were measured as a function of the acidity and the nitrate concentration (MNO 3) in the aqueous phase. The acid dependence was studied in a HNO3 concentration range of 1–14M. The curve IgD vs. lgMHNO 3 in the range of 1–4 gives a slope of 3. The anion effect of nitrate ions shows a complexing influence on the yttrium ions changing its distribution between the two phases. The results can be useful in hydrometallurgical processes in which the solutions are highly acidic and contain a large amount of the dissolved materials as nitrate salts.  相似文献   

9.
Mass-analysed ion kinetic energy spectrometry (MIKES) with collision-induced dissociation (CID) has been used to study the fragmentation processes of a series of deuterated 2,4,6-trinitrotoluene (TNT) and deuterated 2,4,6-trinitrobenzylchloride (TNTCI) derivatives. Typical fragment ions observed in both groups were due to loss of OR′ (R′ = H or D) and NO. In TNT, additional fragment ibns are due to the loss of R2′O and 3NO2, whilst in TNTCI fragment ions are formed by the loss of OCI and R2′OCI. The TNTCI derivatives did not produce molecular ions. In chemical ionization (Cl) of both groups. MH+ ions were observed, with [M – OR′]+ fragments in TNT and [M – OCI]+ fragments in TNTCI. In negative chemical ionization (NCI) TNT derivatives produced M?′, [M–R′]?, [M–OR′]? and [M–NO]? ions, while TNTCI derivatives produced [M–R]?, [M–Cl]? and [M – NO2]? fragment ions without a molecular ion.  相似文献   

10.
Three mononuclear copper(II) complexes of copper nitrate with 2, 6‐bis(pyrazol‐1‐yl)pyridine ( bPzPy ) and 2, 6‐bis(3′,5′‐dimethylpyrazol‐1‐yl)pyridine ( bdmPzPy ), [Cu(bPzPy)(NO3)2] ( 1 ), [Cu(bPzPy)(H2O)(NO3)2] ( 2 ) and [Cu(bdmPzPy)(NO3)2] ( 3 ) were synthesized by the reaction of copper nitrate with the ligand in ethanol solution. The complexes have been characterized through analytical, spectroscopic and EPR measurements. Single crystal X‐ray structure analysis of complexes 1 and 2 revealed a five‐coordinate copper atom in 1 , whereas 2 contains a six‐coordinate (4+2) CuII ion with molecular units acting as supramolecular nodes. These neutral nodes are connected through O–H ··· O(nitrate) hydrogen bonds to give couples of parallel linear strips assembled in 1D‐chains in a zipper‐like motif.  相似文献   

11.
The mechanism of pit growth of pure aluminum (Al) in sulfate ion (SO4 2–)- or nitrate ion (NO3 )-containing 0.1 M sodium chloride solutions has been studied in terms of the morphological changes of artificial pits using potentiodynamic polarization experiment, potentiostatic current transient technique and optical microscopy. The increase in SO4 2– and NO3 ion concentrations in NaCl solution raised the pitting potential E pit of pure Al, which is ascribed to the impediment to pit initiation on pure Al by the addition of SO4 2– or NO3 ions. From the potentiostatic current transients of artificial pits in aqueous 0.1 M NaCl solution, the average value of the pit current was observed to increase with increasing SO4 2– ion concentration, whereas that value of the pit current in the presence of NO3 ions increased up to ca. 0.4 M NO3 ion concentration and then decreased abruptly with increasing NO3 ion concentration. From observations of the morphologies of the pits, it appears that the pit grows preferentially in the lateral direction or in the downward direction by adding SO4 2– or NO3 ions to aqueous 0.1 M NaCl solution, respectively. Based upon the experimental results, two different pit growth mechanisms by anion additives can be proposed: (1) pit growth by the preferential attack of both SO4 2– and Cl to the pit wall in SO4 2–-containing solutions; (2) pit growth by the creation of an aggressive environment at the pit bottom up to 0.4 M NO3 ion concentration due to the lower mobility of NO3 than Cl in NO3 -containing solutions. Electronic Publication  相似文献   

12.
1,2,4-Triazole-3-one (TO) is anticipated to have applications as a high performance alternative gas generating agent, while basic copper nitrate (BCN) is typically used as the oxidizing agent in air bag systems. In order to obtain a better understanding of the thermal properties of TO/BCN mixtures, thermal behavior was investigated using the differential scanning calorimetry. Mixtures of TO with copper, copper oxide, and trihydrated copper nitrate (Cu(NO3)2·3H2O) were also examined for comparison purposes. Samples were prepared at TO/BCN ratios (on a per mass basis) of 10/0, 7/3, 5/5, 3/7, 2/8, 1.6/8.4, 1/9, and 0/10. The endothermic onset temperatures for TO/BCN mixtures were lower than those for either pure TO or pure BCN. TO/BCN mixtures exhibited an initial exothermic peak immediately after an endothermic peak, in the range of 219–234 °C. TO/BCN mixtures with ratios of 3/7, 2/8, 1.6/8.4, and 1/9 displayed a second series of exothermic peaks in the range of 260–300 °C, which appear to result from the oxidation–reduction reaction of previously formed intermediate species with NO2 and NO generated by unreacted BCN. The TO/CuO mixtures are believed to undergo reaction between molten TO and CuO at approximately 230 °C. In general, the presence of copper was shown to be effective at promoting the decomposition of TO. The reaction between TO and Cu(NO3)2·3H2O seems to be initiated by the melting of Cu(NO3)2·3H2O, following which TO reacts with nitric acid resulting from the dissociation of Cu(NO3)2·3H2O. Overall, the triggering event for the reaction between TO and each of the copper nitrate species is a phase change of one of the two mixture components.  相似文献   

13.
Ternary Lithium Rare Earth Nitrates with Lonesome Nitrate Ions: Li3[M(NO3)5](NO3) (M = Gd? Lu, Y). The Crystal Structure of Li3Er(NO3)6 Single crystals of the ternary nitrate Li3Er(NO3)6 are obtained from a solution of “Er(NO3)3” in the melt of LiNO3. In Li3Er(NO3)6 (monoclinic, P21/n, Z = 4; a = 776.0(1); b = 748.86(8); c = 2 396(1) pm; β = 90.76(3)°; R1 = 0.0490; wR2 = 0.0792), Er3+ is surrounded by five bidentate nitrate ligands yielding the anionic units [Er(NO3)5]2?. These are arranged in the direction of the 21 screw axis. Two lonesome NO3? ions are in the middle of such a “helix” and are connected by Li+ with the anions [Er(NO3)5]2?. The helices are moved against each other by about half of the lattice constant a and are connected by further Li+ ions.  相似文献   

14.
Collision-induced dissociation (CID) experiments were performed on atmospheric ion adducts [M + R] formed between various types of organic compounds M and atmospheric negative ions R- [such as O2 , HCO3 , COO(COOH), NO2 , NO3 , and NO3 (HNO3)] in negative-ion mode atmospheric pressure corona discharge ionization (APCDI) mass spectrometry. All of the [M + R] adducts were fragmented to form deprotonated analytes [M – H] and/or atmospheric ions R, whose intensities in the CID spectra were dependent on the proton affinities of the [M – H] and R fragments. Precursor ions [M + R] for which R- have higher proton affinities than [M – H] formed [M – H] as the dominant product. Furthermore, the CID of the adducts with HCO3 and NO3 -(HNO3) led to other product ions such as [M + HO] and NO3 , respectively. The fragmentation behavior of [M + R] for each R observed was independent of analyte type (e.g., whether the analyte was aliphatic or aromatic, or possessed certain functional groups).   相似文献   

15.
Synthesis and Structure of the First Anhydrous Ternary Lithium Nitrates of the Lanthanides, Li2[M(NO3)5] (M = La, Pr? Eu). Single crystals of the ternary lithium nitrates of the lanthanides, Li2[M(NO3)5] (M = La, Pr? Eu), are obtained by dissolving the respective anhydrous nitrate, previously obtained by dehydration of M(NO3)3 · 6 H2O at 180°C under vacuum, in a melt of LiNO3. In the crystal structure of Li2[Pr(NO3)5] (orthorhombic, Pnnm, Z = 4, a = 899.6(2), b = 1 052.7(2), c = 1 178.6(2)pm; R = 0.072, Rw = 0.034) there are two crystallographically different Pr3+ ions, each surrounded by six bidentate nitrate ligands. One nitrate group is bridging between Pr1 and Pr2 resulting in a winded chain, [(O2/2NO2/1)Pr1(NO3)4(O2/2NO2/1)Pr2(O2/2NO2/1)(NO3)4]5?, running along [010]. The chains are packed hexagonally and held together by lithium ions. The coordination polyhydron of Li+ may be described as a bicapped trigonal prism.  相似文献   

16.
A synthetic graft copolymer of cross-linked starch/acrylonitrile was used as an adsorbent for the removal of Cu(II) ions from an aqueous solution of copper nitrate hexahydrate Cu(NO3)2 · 6H2O at different temperatures and fixed pH. The amount adsorbed increased with increasing concentration of Cu(II) ions and decreasing temperature. The length of time taken to reach equilibrium of the adsorption of Cu(II) ions was the same at all temperatures tested. Kinetics studies showed that the adsorption process obeyed first-order reversible kinetics and the adsorption isotherms followed the Freundlich model. Furthermore, the thermodynamic parameters, i.e. standard free energy (ΔG), standard enthalpy (ΔH), and standard entropy (ΔS), of the adsorption process were calculated and the results are discussed in detail.  相似文献   

17.
Displacement-extraction of tracer concentrations of210Po in 1.0M (H, Na)NO3 solutions has been studied by using copper dithizonate–CCl4 solutions. Furthermore, based on the results of the displacement-extraction of polonium, a mixture of210Po,210Bi, and210Pb of tracer concentrations in 1.0M (H, Na)NO3 solutions could be satisfactorily separated with successive extractions by copper dithizonate–CCl4 and dithizone–CCl4 solutions in acidic conditions.  相似文献   

18.
Reaction of 2,2′‐bi­pyridine (bpy) and copper(II) nitrate in methanol results in two complexes, namely light‐blue bis(2,2′‐bi­pyridine)­nitrato­copper(II) nitrate methanol solvate, [Cu(NO3)(C10H8N2)2]NO3·CH3OH, (I), which is unstable in air, and the product of its decomposition, catena‐poly­[[[bis(2,2′‐bi­pyridine)copper(II)]‐μ‐nitrato‐O:O′] nitrate], {[Cu(NO3)(C10H8N2)2]NO3}n, (II). The crystal structures of both compounds were determined from one crystal at room temperature. Later, the structure of (I) was redetermined at low temperature. In (I) and (II), the Cu atom is coordinated by two bpy and one or two nitrate ions, respectively. The second nitrate ion in (I), along with the methanol solvent mol­ecule, is found in the outer coordination sphere, not bonded to Cu. The nitrate in (I) is chelating, while in (II), it bridges (bpy)2Cu complexes, forming a one‐dimensional chain structure. The Cu cation in (II) lies on a twofold axis and the uncoordinated NO3? ion is located close to a twofold axis and is therefore disordered. Compound (I) converts into (II) upon loss of solvent.  相似文献   

19.
An amperometric detector unit equipped with a Cu(II)-containing poly(3-methylthiophene) working electrode is described for the single-column ion chromatographic detection of electroinactive inorganic anions, such as F?, Cl?, Br?, NO2? and NO3?. Chromatograms obtained with this unit and with a commercial conductivity detector are almost identical with regard to peak height. Thus, an amperometric unit employing this modified electrode can be used as a conductance monitor in ion chromatographic analysis. Although the responses of this electrode seem to be conductivity related, the detection principle is probably based on a dual mechanism involving equilibria between copper ions and various anions of the system in addition to simple conductivity changes associated with the passage of analyte plugs. This explains the difference in responses observed with platinum and stainless-steel electrodes used in the same cell configuration. The detector displays a linear range of at least two orders of magnitude on a logarithmic scale.  相似文献   

20.
This paper describes three new methods: the first may be used for the determination of nitrite; the second is applicable to determination of nitrate; and the third permits sequential determination of both nitrite and nitrate in mixtures with no prior separation. For the determination of nitrite and nitrate in synthetic mixtures containing 1:5 to 5:1 ratios of the ions, in tap water, and in river water, mean recoveries (for 3 to 22 μg of added NO3and NO2) are 96.1 and 98.1% (n= 15) and coefficients of variation are 2.2 and 2.5% for NO3and NO2(n= 5), respectively.  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