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.
The data on the uranium metal corrosion rate in the solutions of nitric acid (0,1 – 4 M) and effect of complex forming agents on uranium corrosion properties are presented. The increase of HNO3 concentration caused the shift of corrosion potential from 38 mV to 446 mV and the increase of the corrosion rate from 0,02 to 0,62 mg.cm-2h-1. Transpassivation potential of U metal was found weakly effected by HNO3 concentration varying from 448 to 470 mV/Ag/AgCl. The addition of HCOOH to the electrolytes containing less than 3 M HNO3 found to shift the values of corrosion potentials about 500 mV towards negative direction reducing the passivation of U metal. The data on the kinetics of oxidative dissolution of PuO2 using Ag(II) and Am(VI,V) as mediators and the effect of the mediator generation techniques are discussed. The electrochemical properties of UC in the solutions 2 – 4 M HNO3, results of the quantitative determination of “oxidizable carbon” in dissolver solutions are presented. The results of corrosion and dissolution studies of Tc metal and Tc - Ru alloys containing from 19 to 70 at.% Ru in 0.5 0– 6 M HNO3 indicate the formation of passive films of Tc(IV) – Ru(III,IV) hydroxides at the electrode surface in the solutions containing less than 2 M HNO3 at the potentials less than 650 mV/Ag/AgCl. The increase of HNO3 concentration to values exceeding 3 M and the shift of the electrode potential towards positive direction causes the transition of the Tc and Tc-Ru alloys to transpassive state. The values of transpassivation potentials increase with the increasing with HNO3 concentration. Quantitative dissolution of Tc metal without application of oxidation potential becomes possible in the electrolytes, containing more than 4 M HNO3. The rate of Tc – Ru alloys dissolution is noticed to slow down with the increase of Ru content in the alloy.  相似文献   

2.
The extraction of Tc(VII) by the mixture of tri-n-butyl phosphate (TBP) and 2-nitrophenyl octyl ether (NPOE) has been studied. 0.2M NPOE-TBP can extract Tc(VII) effectively from 1M HNO3 and 1M NaOH solutions with distribution ratios of 57.1 and 12.3, respectively. The distribution ratio of Tc(VII) decreases with increasing (>0.5M) HNO3 concentration but increases with the increase of NaOH concentration. A pH 9 NaOH solution has proven to be suitable for Tc(VII) stripping. A simple extraction-stripping cycle can remove Tc(VII) from a sodium hydroxide solution. A more sophisticated extraction process is proposed to remove Tc(VII) from nitric acid solution because the co-extracted HNO3 prevents the direct stripping of Tc(VII) by NaOH solution of pH 9.  相似文献   

3.
In order to further test the actinides/lanthanides separation performance by Cyanex 301 extraction, a countercurrent extraction experiment was carried out in the present work. The separation process consisted of 7-stage extraction, 3-stage scrubbing and 4-stage stripping. 14 miniature centrifugal contactors were installed in glove box for the experiment. The feed solution from TRPO process, mainly consisting of 4.4 M HNO3, ∼3.7×108Bq Am-241, 0.02 M lanthanides, ∼120 ppm Mo and 100 ppm Fe, was pretreated by the following procedures: denitrating to 0.2 M HNO3 by formic acid, adjusting pH to 2 by 8 M NaOH, removing most Fe3+ by 0.2 M Cyanex 301-kerosene cross-flow extraction, and then adjusting pH to 3.5 by 1 M NaOH. About 5.6×108Bq Pm-147 was added into the feed solution to trace lanthanides during the experiment. The experiment lasted for 9 hours with a feed flow rate of 30 ml/h. The results show that 99.95% Am was separated from lanthanides and only 0.1% lanthanides were extracted together with Am. 99.3% Am and 96.6% Pm were stripped from load Cyanex 301 by 1.0 M HNO3, respectively.  相似文献   

4.
Technetium-99 is one of the most abundant, long-lived radiotoxic isotopes in used nuclear fuel (UNF). As such, it is targeted in UNF separation strategies such as UREX+, for isolation and encapsulation in solid waste forms for storage in a nuclear repository. We report here results regarding the incorporation of Tc-99 into ternary oxides of different structure types: pyrochlore (Nd2Tc2O7), perovskite (SrTcO3), and layered perovskite (Sr2TcO4). The goal was to determine synthesis conditions of these potential waste forms to immobilize Tc-99 and to harvest crystallographic, thermophysical and hydrodynamic data. Within these studies, Rietveld structure refinement was applied to allow for crystallographic characterization, while a physical property measurement system (PPMS) was used to determine thermophysical properties. The ternary oxides exhibited good crystallinity and their lattice parameters and atomic coordinates could be refined to high accuracy. Low refinement residuals (RBragg) of 2.0, 2.4, and 3.9% were achieved for Nd2Tc2O7, SrTcO3, and Sr2TcO4, respectively. The strontium technetates, SrTcO3 and Sr2TcO4, show superconductivity at rather high critical temperatures of Tc = 7.8 K and 7 K, respectively. Nd2Tc2O7 did not show any changes in magnetic properties above 3 K.  相似文献   

5.
This work introduces the feasibility of using sugar cane bagasse (SCB) – a sugar cane industry waste – as a selective solid phase extractor for Fe(III). The order of metal uptake capacities in μmol g?1 for the extraction of six tested metal ions from aqueous solution using static technique is Fe(III) > Cu(II) > Pb(II) > Zn(II) > Cd(II) > Co(II). Since SCB exhibits remarkable binding characteristics for Fe(III), special interest was devoted for optimizing its uptake and studying its selectivity properties under static and dynamic conditions. In this respect, batch experiments were carried out at the pH range 1.0–4.0, initial concentration of metal ion (10–100 μmol), weight of phase (25, 50, 75, 100, 125 and 150 mg) and shaking time (10, 30, 45, 60, 90, 120 and 150 min). FT-IR spectra of SCB before and after uptake of Fe(III) were recorded to explore the nature of the functional groups responsible for binding of Fe(III) onto the studied natural biosorbent. The equilibrium data were better fitted with Langmuir model (r2 = 0.985) than Freundlich model (r2 = 0.934). Moreover, Fe(III) sorption was fast and completed within 60 min. The adsorption kinetics data were best fitted with the pseudo-second-order type. As a view to find a suitable application of SCB based on its unique property as a benign sorbent, it was found that, Fe(III) spiked natural water samples such as doubly distilled water (DDW), drinking tap water (DTW), natural drinking water (NDW), ground water (GW) and Nile River water (NRW) was quantitatively recovered (>95.0%) using batch and column experiments, with no matrix interferences.  相似文献   

6.
The cross-sections for formation of metastable state of 99Tc (99mTc, 140.511 keV, 6.01 h) through natRu(n,x)99mTc reaction induced by 13.5 MeV and 14.8 MeV neutrons were measured. Fast neutrons were produced via the 3H(d,n)4He reaction on the K-400 neutron generator. Induced gamma activities were measured by a high-resolution gamma-ray spectrometer with a high-purity germanium (HpGe) detector. Measurements were corrected for gamma-ray attenuations, dead time and fluctuation of neutron flux. Data for natRu(n,x)99mTc reaction cross sections are reported to be 9.6±1.5 and 9.2±1.1 mb at 13.5±0.2 and 14.8±0.2 MeV incident neutron energies, respectively. Results were compared with the data by other anthors.  相似文献   

7.
Gum kondagogu (Cochlospermum gossypium), a naturally occurring tree biopolymer, is exploited as a biosorbent to remove metal ions from aqueous solutions. The removal efficiency of toxic metals by gum kondagogu was determined quantitatively in the order Cd2+ > Cu2+ > Fe2+ > Se2+ > Pb2+ > total Cr > Ni2+ > Zn2+ > Co2+ > As2+ at pH 5.0 ± 0.1 and temperature 25 ± 2 °C by inductively coupled plasma-mass spectrometry (ICP-MS). The biosorption (%) of various metal ions tested was found to be in the range of 97.3–16.7%, at pH 5.0. The morphological and mechanisms of interaction of toxic metal ions with gum kondagogu were assessed by scanning electron microscopy coupled with energy dispersive X-ray analysis (SEM-EDXA) and X-ray diffraction (XRD) spectrum. The analysis indicated that biosorption process included morphological changes, precipitation, complexation and ion exchange mechanism for the removal of metal ions by the gum. XRD analysis indicated the amorphous nature of gum kondagogu, which facilitate metal biosorption. The metal ions adsorption leads to its deposition on the gum kondagogu matrix in a crystalline state.  相似文献   

8.
Both TcO(2)F(3) and ReO(2)F(3) are infinite chain, fluorine-bridged polymers in the solid state. Their solution structures have been studied by (19)F and (99)Tc NMR spectroscopy in SO(2)ClF solution and shown to exhibit cyclic (MO(2)F(3))(3) (M = Tc, Re) and (ReO(2)F(3))(4) structures that have been confirmed by simulation of the (19)F NMR spectra. The trimers dominate in both the technetium and rhenium systems, with both the tetramer and trimer existing in equilibrium in the rhenium system. A low concentration of a higher, possibly pentameric, cyclic rhenium polymorph is also present in equilibrium with the trimer and tetramer.  相似文献   

9.
A new PVC membrane based strontium(II) ion-selective electrode has been constructed using acetophenone semicarbazone as a neutral carrier. The sensor exhibits a Nerstian response for strontium(II) ion over a wide concentration range 1.0 × 10−2–1.0 × 10−7 M with the slope of 29.4 mV/per decade. The limit of detection was 2.7 × 10−8 M. It was relatively fast response time (<10 s for concentration ⩾1.0 × 10−3 and <15 s for concentration of ⩾1.0 × 10−6 M) and can be used for 8 months without any considerable divergence in potentials. The proposed sensor revealed relatively good selectivity and high sensitivity for strontium(II) over a mono, di, trivalent cation and can be used in a pH range of 2.5–10.5. It was also successfully used as an indicator electrode in potentiometer titration and in the analysis of concentration in various real samples.  相似文献   

10.
The metal–metal bond in [M2(CO)9{C(OEt)R}] (M = Mn (1), Re (2), R = 2-thienyl (a), 2-bithienyl (b)) is readily cleaved with halogens to afford cis-[M(CO)4(X){C(OEt)R}] (M = Mn (3), X = I; M = Re (4), X = Br). In the binuclear manganese complex, the carbene ligand is found in an axial position due to steric reasons, whereas the electronically favoured equatorial position is found for the carbene ligands in the corresponding rhenium complexes and in [Mn2(CO)9{C(NH2)thienyl}] (5a), containing a sterically less demanding NH2-substituent.  相似文献   

11.
Mg–Al layered double hydroxides (Mg–Al LDHs) intercalated with 1,3,6-naphthalenetrisulfonate (NTS3?) and 3-amino-2,7-naphthalenedisulfonate (ANDS2?) ions were prepared by coprecipitation and were characterized by X-ray diffraction and chemical analyses. Based on X-ray diffraction patterns, the naphthalene rings of NTS3? and ANDS2? were most likely oriented parallel to the brucite-like host layers of the Mg–Al LDH, midway between layers. The prepared Mg–Al LDHs were able to selectively take up aromatics from aqueous solutions, and the order of percentage uptake was as follows: 1,3-dinitrobenzene > nitrobenzene > benzaldehyde > N,N-dimethylaniline > anisole > 1,2-dimethoxybenzene. The differences in the extent of π–π stacking interactions occurring between the benzene rings of the aromatics and the naphthalene ring of the intercalated NTS3? and ANDS2? probably resulted in these differences among the absorbed quantities of the various aromatics.  相似文献   

12.
The photostabilization of poly(methyl methacrylate) (PMMA) films by Schiff bases of 2,5-dimercapto-1,3,4-thiadiazole compounds was investigated. The PMMA films containing concentration of complexes 0.5% by weight were produced by the casting method from chloroform solvent. The photostabilization activities of these compounds were determined by monitoring the hydroxyl index with irradiation time. The changes in viscosity average molecular weight of PMMA with irradiation time were also tracked (using benzene as a solvent). The quantum yield of the chain scission (Φcs) of these complexes in PMMA films was evaluated and found to range between 4.19 × 10?5 and 8.75 × 10?5. Results obtained showed that the rate of photostabilization of PMMA in the presence of the additive followed the trend:[1] > [2] > [3] > [4] > [5].According to the experimental results obtained, several mechanisms were suggested depending on the structure of the additive. Among them, UV absorption, peroxide decomposer, and radical scavenger for photostabilizer mechanisms were suggested.  相似文献   

13.
《Fluid Phase Equilibria》2005,235(2):196-200
This work contributes to the development of an enrichment process for antioxidant compounds in aqueous alcoholic extracts of boldo (Peumus boldus M.) leaves by using high-pressure CO2 as the solvent. Specifically we measured the high-pressure solubility (y2, molar fraction) of a selected bioactive compound in boldo leaves (boldine) in CO2 as a function of system temperature (298 K  T  333 K) and pressure (8 MPa  P  40 MPa). Experimental data was correlated by using a density-based model which is valid for solvent densities >607 kg/m3. Predicted solubility values are low (4 × 10−7  y2  6 × 10−5) but comparable with those of nitrogen-containing organic compounds with similar molecular weight (327.4 Da) and solubility parameter (28.3 MPa0.5 at 313 K) as boldine.  相似文献   

14.
The reactions of [ReOX3(AsPh3)2] (X = Cl, Br) with 4,7-diphenyl-1,10-phenanthroline (dpphen) have been examined and the complexes [ReO(OMe)X2(dpphen)] and [Re2O3X4(dpphen)2]·2/3CH2Cl2 (X = Cl, Br) have been obtained. They were characterized by IR, UV–Vis spectroscopy, and single crystal X-ray analysis has been performed for [ReO(OMe)Cl2(dpphen)] and [Re2O3Br4(dpphen)2]·2/3CH2Cl2. The nature of the frontier orbitals of [ReO(OMe)Cl2(dpphen)] and [Re2O3Br4(dpphen)2] and the electronic transitions involved in the absorption spectra have been studied by means of the density functional and time-dependent density functional methods.  相似文献   

15.
The Freundlich and Langmuir isotherms were used to describe the biosorption of Cu(II), Pb(II), and Zn(II) onto the saltbush leaves biomass at 297 K and pH 5.0. The correlation coefficients (R2) obtained from the Freundlich model were 0.9798, 0.9575, and 0.9963 for Cu, Pb, and Zn, respectively, while for the Langmuir model the R2 values for the same metals were 0.0001, 0.1380, and 0.0088, respectively. This suggests that saltbush leaves biomass sorbed the three metals following the Freundlich model (R2 > 0.9575). The KF values obtained from the Freundlich model (175.5 · 10−2, 10.5 · 10−2, and 6.32 · 10−2 mol · g−1 for Pb, Zn, and Cu, respectively), suggest that the metal binding affinity was in the order Pb > Zn > Cu. The experimental values of the maximal adsorption capacities of saltbush leaves biomass were 0.13 · 10−2, 0.05 · 10−2, and 0.107 · 10−2 mol · g−1 for Pb, Zn, and Cu, respectively. The negative ΔG values for Pb and the positive values for Cu and Zn indicate that the Pb biosorption by saltbush biomass was a spontaneous process.  相似文献   

16.
Diels-Alder reaction of (η5-cyclopentadienyl)M(CO)x1-N-maleimidato) complexes (M = Fe, Mo, W, x = 2 or 3) with cyclopentadiene has been studied. The observed order of reactivity was: N-ethylmaleimide > W complex > Mo complex > Fe complex. The X-ray structures of the adducts have been determined for M = W and Fe. DFT calculations on the starting complexes have been performed to explain the observed reactivity order.  相似文献   

17.
The effect of gel composition, absorbed dose and pH of the solution on the uranyl ion uptake capacity of N-isopropylacrylamide/maleic acid copolymeric hydrogels containing 0–3 mol% of maleic acid at 48 kGy have been investigated. Uranyl uptake capacity of hydrogels are found to increase from 18.5 to 94.8 mg [UO22+]/g dry gel as the mole % of maleic acid content in the gel structure increased from 0 to 3. The percent swelling, equilibrium swelling and diffusion coefficient values have been evaluated for poly(N-isopropylacrylamide/maleic acid) hydrogels at 500 ppm of uranyl nitrate solution.  相似文献   

18.
This paper describes a novel strategy for actinide separation by extraction chromatography with Np(III) valence adjustment. Neptunium(IV) was reduced to Np(III) using Cr(II) and then selectively separated from uranium (IV) on a TEVA resin. After elution, Np(III) was retained on a DGA resin in order to remove any detrimental chromium impurities. Neptunium(III) formation was demonstrated by the complete and selective elution of Np from TEVA resin (99 ± 7%) in less than 12 mL of 9 M HCl from U(IV) (0.7 ± 0.7%). It was determined by UV–visible and kinetic studies that Cr(II) was the only species responsible for the elution of Np(IV) as Np(III) and that the Cr(II) solution could be prepared from 2 to 30 min before its use without the need of complex degassing systems to prevent the oxidation of Np(III) by oxygen. The methodology proposed here with TEVA/DGA resins provides removal of Cr(III) impurities produced at high decontamination factors (2.8 × 103 and 7.3 × 104 respectively).  相似文献   

19.
《Polyhedron》2002,21(12-13):1163-1175
The results of density functional theory (DFT) calculations on a set of binuclear nonachloride complexes M′M″Cl9 4− (M′=V, Nb, Ta; M″=Cr, Mo, W) and M′M″Cl9 2− (M′=Cr, Mo, W; M″=Mn, Tc, Re), in which each metal possesses a nominal d3 valence electronic configuration, are reported. When compared with previous studies on same-group dimers (typified by the M′M″Cl9 3− complexes of the chromium triad), the present results display an increased tendency for electron donation from M′ to M″. Structural trends evident for the M′M″Cl9 4− and M′M″Cl9 2− series of dimers are remarkably consistent: weak ferromagnetic coupling between M′ and M″ is the most favorable intermetallic interaction when M″ is a first-row transition metal, antiferromagnetic coupling dominates when M′ is first-row but M″ is second- or third-row, and metalmetal triple bond formation generally yields the lowest-energy structure when neither metal is first-row. These structural trends, and other characteristics of the dimers described here, can be satisfactorily rationalized in terms of the tendency for electron transfer from M′ to M″, coupled with effects due to spin polarization and ligand field splitting of the valence d orbitals on M′ and M″.  相似文献   

20.
A high-resolution continuum source atomic absorption spectrometric method was developed and validated for the determination of NiII(3-OMe-salophene) (a complex with anticancer activity in vitro) in MCF7 and HT29 cancer cell lines. The primarily most sensitive line 232.003 nm was selected for analysis. Compared to the standard nickel, the absorbance values obtained for NiII(3-OMe-salophene) complex was at least 93% at the upper end of linear range of the calibration curve. The use of common matrix modifiers including magnesium nitrate, palladium nitrate, ammonium hydrogen phosphate, lanthanum chloride and calcium nitrate brought no significant improvement in the GF AAS measurement. The dynamic linear working range of the calibration curve was found to be between 2.16 and 12.0 μg L? 1 (ppb). This covers a concentration range of the complex from 0.036 μM to 0.204 μM. Typical coefficients of variation (n = 6) ranged from 0.2% to 6.7% for Ni in NiII(3-OMe-salophene). Detection and quantitation limits were 0.65 and 2.16 μg L? 1 (ppb), respectively. The proposed method has been successfully applied to the analysis of NiII(3-OMe-salophene) complex in cell lines of breast cancer (MCF7) and colon cancer (HT29). However, being based on the determination of nickel in the salophene complex, the method was unsuitable for the quantitation of NiII(3-OMe-salophene) in serum albumin, which originally contains significant amount of nickel. For this purpose, a high performance liquid chromatographic method with a monolithic silica RP-18e column has been developed to quantitate the complex in serum albumin. The developed chromatographic method depends on detecting the whole complex in serum rather than the bounded nickel. A mobile phase consisting of 25 mM phosphate buffer pH 3/methanol (30:70, v/v) was pumped at a flow rate of 1 mL min? 1. The eluted complex was monitored at a wavelength of 250 nm. The dynamic linear working range of the calibration curve for the developed LC method was found to be between 100 and 20,000 μg L? 1 (0.23–46.18 μM). Detection and quantitation limits were 30 and 100 μg L? 1 (ppb), 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