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 718 毫秒
1.
Oxidation of As(iii) to As(v) is an effective way to improve the performance of most arsenic removal technologies. In this study, a new alternative biosorbent, TiO2-loaded biochar prepared by waste Chinese traditional medicine dregs (TBC) was applied in remediation for As(iii) from aqueous solution. Compared with unmodified biochar, the specific surface areas and total pore volumes of TBC increased while the average aperture decreased due to the loading of nano-TiO2. The X-ray diffraction (XRD) of TBC confirmed that the precipitated titanium oxide was primarily anatase. pH did not have a significant effect on the adsorption capacity at 10 mg L−1 As(iii) in suspension with a pH ranging from 2 to 10. Adsorption kinetics data were best fitted by the pseudo-second-order model (R2 > 0.999). The Sips maximum adsorption capacity was 58.456 mg g−1 at 25 °C, which is comparable with other adsorbents reported in previous literature. The Gibbs free energy (ΔG) of As(iii) adsorption was negative, indicating the spontaneous nature of adsorption. The results of free radical scavenging and N2 purging experiments indicated that O2 acted as an electron accepter and O2˙ dominated the oxidation of As(iii). The oxidation of As(iii) obviously affected the adsorption capacity for As(iii) by TBC. X-ray photoelectron spectroscopy (XPS) studies showed that As(iii) and As(v) existed on the surface of TBC, suggesting that the oxidation of As(iii) occurred. TBC played multiple roles for As(iii), including direct adsorption and photocatalytic oxidation adsorption. Regeneration and stability experiments showed that TBC was an environment-friendly and efficient adsorbent for As(iii) removal.

TiO2-loaded biochar prepared by waste Chinese traditional medicine dregs (TBC) was applied in remediation for As(iii) from aqueous solution.  相似文献   

2.
Novel Bi2WO6/bentonite (denoted as BWO/BENT) composites were prepared via a typical hydrothermal process and employed for the photocatalytic oxidation of arsenic(iii) (As(iii)). The properties of the prepared samples were characterized through X-ray diffraction, transmission and scanning electron microscopy, UV-visible diffuse reflectance spectroscopy, X-ray photoelectron spectroscopy, and photoluminescence spectroscopy. Effects of the BENT ratio on the As(iii) removal were explored under simulated sunlight, and the best photocatalytic effect was observed for the composite with BWO : BENT = 7 : 3 w/w. Compared with the pure BWO, the BWO/BENT composites exhibited an improved photocatalytic ability in the removal of As(iii), which was mainly ascribed to the enlarged specific surface area and the suppressed electron–hole recombination by the incorporated BENT. Furthermore, photo-generated holes (h+) and superoxide radicals ·O2 were confirmed to be the major contributors to the oxidation of As(iii), and an associated mechanism of photocatalytic oxidation of As(iii) over BWO/BENT composites was proposed.

Novel Bi2WO6/bentonite (denoted as BWO/BENT) composites were prepared via a typical hydrothermal process and employed for the photocatalytic oxidation of arsenic(iii) (As(iii)).  相似文献   

3.
Bifunctional heterogeneous catalytic processes for highly efficient removal of arsenic (As(iii)) are receiving increased attention. However, the agglomerated nature and stability of nanoparticles are major concerns. Herein, we report a new process regarding the anchoring of CuFe2O4 nanoparticles on a substrate material, a kind of Fe–Ni foam, to form porous CuFe2O4 foam (CuFe2O4-foam) by in situ synthesis. The prepared material was then applied to activate peroxymonosulfate (PMS) for fast and efficient removal of As(iii) from water. The results of removal experiments show that the complete removal of arsenic (<10 μg L−1) from 1 mg L−1 As(iii) aqueous solution can be achieved within shorter time (<10 min) using this adsorbent coupled with PMS. The maximum adsorption capability of As(iii) and As(v) on the prepared adsorbent is observed to be about 105.78 mg g−1 and 120.32 mg g−1, respectively. CuFe2O4-foam/PMS couple could work effectively in a wide pH range (3.0–9.0) and temperature range (10–60 °C), which is more beneficial to its application in actual water treatment engineering. The exhausted adsorbents can be refreshed for cyclic runs (at least 7 cycles) with insignificant capacity loss using alkaline solution as a regeneration strategy, suggesting this process has good stability. Investigation of the mechanism reveals that the route to the removal of As(iii) is synchronous oxidation and sequestration in the arsenic removal process. The large As(iii) removal capability and stability of CuFe2O4-foam/PMS show its potential as a promising candidate in real As(iii)-contaminated groundwater treatment.

Bifunctional heterogeneous catalytic processes for highly efficient removal of arsenic (As(iii)) are receiving increased attention.  相似文献   

4.
The impact of calcium on the solubility, redox behavior, and speciation of the An(iii)–EDTA (An = Pu or Cm) system under reducing, anoxic conditions was investigated through batch solubility experiments, X-ray absorption spectroscopy (XAS), density functional theory (DFT), and time-resolved laser fluorescence spectroscopy (TRLFS). Batch solubility experiments were conducted from undersaturation using Pu(OH)3(am) as the solid phase in contact with 0.1 M NaCl–NaOH–HCl–EDTA–CaCl2 solutions at [EDTA] = 1 mM, pHm = 7.5–9.5, and [CaCl2] ≤20 mM. Additional samples targeted brine systems represented by 3.5 M CaCl2 and WIPP simulated brine. Solubility data in the absence of calcium were well-described by Pu(iii)–EDTA thermodynamic models, thus supporting the stabilization of Pu(iii)–EDTA complexes in solution. Cm(iii)–EDTA TRLFS data suggested the stepwise hydrolysis of An(iii)-EDTA complexes with increasing pH, and current Pu(iii)-EDTA solubility models were reassessed to evaluate the possibility of including Pu(iii)–OH–EDTA complexes and to calculate preliminary formation constants. Solubility data in the presence of calcium exhibited nearly constant log m(Pu)tot, as limited by total ligand concentration, with increasing [CaCl2]tot, which supports the formation of calcium-stabilized Pu(iii)–EDTA complexes in solution. XAS spectra without calcium showed partial oxidation of Pu(iii) to Pu(iv) in the aqueous phase, while calcium-containing experiments exhibited only Pu(iii), suggesting that Ca–Pu(iii)–EDTA complexes may stabilize Pu(iii) over short timeframes (t ≤45 days). DFT calculations on the Ca–Pu(iii)–EDTA system and TRLFS studies on the analogous Ca–Cm(iii)–EDTA system show that calcium likely stabilizes An(iii)–EDTA complexes but can also potentially stabilize An(iii)–OH–EDTA species in solution. This hints towards the possible existence of four major complex types within Ca–An(iii)–EDTA systems: An(iii)–EDTA, An(iii)–OH–EDTA, Ca–An(iii)–EDTA, and Ca–An(iii)–OH–EDTA. While the exact stoichiometry and degree of ligand protonation within these complexes remain undefined, their formation must be accounted for to properly assess the fate and transport of plutonium under conditions relevant to nuclear waste disposal.

Combined advanced spectroscopy and solubility studies provide evidence for the formation of novel calcium-containing and hydrolyzed (Cm,Pu)(iii)–EDTA complex(es).  相似文献   

5.
To remove arsenite (As(iii)) from wastewater effectively, the catalytic oxidation of As(iii) to arsenate (As(v)) and As(v) precipitation with iron ions (Fe(iii)) was investigated. The Pt/SiO2 catalyst functioned as a reaction site for As(iii) with oxygen in the atmosphere. The combination of the Pt/SiO2 catalyst and Fe(iii) precipitant improved the removal of As(iii) in the precipitate; Pt/SiO2 worked as both an As(iii) oxidation site and precipitation site with Fe(iii) precipitant.

A Pt/SiO2 catalyst promoted an oxidative reaction of arsenite to arsenate with air, and it also functioned as a nucleation site of its precipitate with iron precipitant, achieving high removal efficiency from water.  相似文献   

6.
Y mainly exists in ionic rare-earth resources. During rare-earth carbonate precipitation, rare-earth ion loss in the precipitated rare-earth mother liquor often occurs due to CO32− coordination and Y(iii) hydration. Microscopic information on the coordination and hydration of CO32− and H2O to Y(iii) has not yet been elucidated. Therefore, in this study, the macroscopic dissolution of Y(iii) in different aqueous solutions of Na2CO3 was studied. The radial distribution function and coordination number of Y(iii) by CO32− and H2O were systematically analyzed using molecular dynamics (MD) simulations to obtain the complex ion form of Y(iii) in carbonate solutions. Density functional theory (DFT) was used to geometrically optimize and calculate the UV spectrum of Y(iii) complex ions. This spectrum was then analyzed and compared with experimentally determined ultraviolet-visible spectra to verify the reliability of the MD simulation results. Results showed that Y(iii) in aqueous solution exists in the form of [Y·3H2O]3+ and that CO32− is present in the bidentate coordination form. In 0–0.8 mol L−1 CO32− solutions, Y(iii) was mainly present as the 5-coordinated complex [YCO3·3H2O]+. When the concentration of CO32− was increased to 1.2 mol L−1, [YCO3·3H2O]+ was converted into a 6-coordinated complex [Y(CO3)2·2H2O]. Further increases in CO32− concentration promoted Y(iii) dissolution in solution in the form of complex ions. These findings can be used to explain the problem of incomplete precipitation of rare earths in carbonate solutions.

Based on MD results, DFT was used to geometrically optimize and calculate the UV spectrum of Y(iii) complex ions. Data validation was further performed using UV-vis experiments to reveal Y(iii) coordination and hydration properties.  相似文献   

7.
In order to effectively destroy the structure of the passive oxidation film that covers zero-valent iron (ZVI), an Fe(iii)-reducing strain, namely Morganella sp., was isolated from anaerobic activated sludge and coated on ZVI, which was distributed in porous ceramsite made of iron dust, kaolin and straw, with a ratio of 7 : 3 : 1. Batch experiments showed that under the optimized conditions, the maximum removal amount of Cr(vi) by ZVI increased from 7.33 mg g−1 to 26.87 mg g−1 in the presence of the Fe(iii)-reducing bacterium. The column experiment was performed with the addition of the agar globules to supply nutrients to the strain. Compared with ZVI, the column penetration time and maximum capture amount of RB-ZVI increased to 17 h and 112.5 mg g−1, respectively, on the 15th day. Furthermore, the service life of RB-ZVI was prolonged in the existence of the strain. Based on X-ray diffraction, Raman spectroscopy and X-ray photoelectron spectroscopy analyses, the key mechanisms for the removal of Cr(vi) by ZVI coated with Fe(iii)-reducing bacterium were determined to be adsorption, reduction, coprecipitation and biomineralization.

To effectively destroy the structure of the passive oxidation film covering zero-valent iron (ZVI), an Fe(iii)-reducing strain, Morganella sp., was isolated from anaerobic activated sludge and coated on the ZVI.  相似文献   

8.
Pd nanoparticles were electrochemically immobilized on a Pt surface in the presence of sodium dodecyl sulfate (SDS) molecules to study the electrokinetics of arsenite oxidation reactions and the corresponding sensing activities. The X-ray photoelectron spectroscopy (XPS) analysis showed that on the Pt surface, Pd atoms exist as adatoms and the contents of Pd(0) and Pd(ii) were 75.72 and 24.28 at%, respectively, and the particle sizes were in the range of 61–145 nm. The experimental results revealed that the catalytic efficiency as well as the charge transfer resistance (at the redox potential of the Fe(ii)/Fe(iii) couple) increased in the order of Pt < Pt–Pd < Pt–Pdsds. A Pt–Pdsds electrode exhibited an open circuit potential (OCP) of 0.65 V in acidic conditions; however, when 50.0 mM NaAsO2 was present, the OCP value shifted to 0.42 V. It has been projected that the As(iii) oxidation proceeds using a sequential pathway: As(iii) → As(iv) → As(v). After optimization of the square wave voltammetric data, the limits of detection of As(iii) were obtained as 1.3 μg L−1 and 0.2 μg L−1 when the surface modification of the Pt surface was executed with Pd particles in the absence and presence of the SDS surfactant, respectively. Finally, real samples were analyzed with excellent recovery performance.

Amplification of true surface area can be improved when Pd particles are deposited on a substrate in the presence of sodium dodecyl sulfate (SDS) surfactant. In acidic medium, As(iii) undergoes a two-step oxidation process.  相似文献   

9.
We report a simple and cost-effective paper-based and colorimetric dual-mode detection of As(iii) and Pb(ii) based on glucose-functionalized gold nanoparticles under optimized conditions. The paper-based detection of As(iii) and Pb(ii) is based on the change in the signal intensity of AuNPs/Glu fabricated on a paper substrate after the deposition of the analyte using a smartphone, followed by processing with the ImageJ software. The colorimetric method is based on the change in the color and the red shift of the localized surface plasmon resonance (LSPR) absorption band of AuNPs/Glu in the region of 200–800 nm. The red shift (Δλ) of the LSPR band observed was from 525 nm to 660 nm for As(iii) and from 525 nm to 670 nm for Pb(ii). The mechanism of dual-mode detection is due to the non-covalent interactions of As(iii) and Pb(ii) ions with glucose molecule present on the surface AuNPs, resulting in the aggregation of novel metal nanoparticles. The calibration curve gave a good linearity range of 20–500 μg L−1 and 20–1000 μg L−1 for the determination of As(iii) and Pb(ii) with the limit of detection of 5.6 μg L−1 and 7.7 μg L−1 for both metal ions, respectively. The possible effects of different metal ions and anions were also investigated but did not cause any significant interference. The employment of AuNPs/Glu is successfully demonstrated for the determination of As(iii) and Pb(ii) using paper-based and colorimetric sensors in environmental water samples.

We report a simple and cost-effective paper-based and colorimetric dual-mode detection of As(iii) and Pb(ii) based on glucose-functionalized gold nanoparticles under optimized conditions.  相似文献   

10.
In this study, a jacobsite–biochar nanocomposite (MnFe2O4–BC) was fabricated and used to simultaneously remove Sb(iii) and Cd(ii) from water via adsorption. The MnFe2O4–BC nanocomposite was prepared via a co-precipitation method and analyzed using various techniques. The results confirm the successful decoration of the biochar surface with MnFe2O4 nanoparticles. The maximum Sb(iii) removal efficiency was found to be higher from bi-solute solutions containing Cd(ii) than from single-solute systems, suggesting that the presence of Cd(ii) enhances the removal of Sb(iii). The Langmuir isotherm model describes well Sb(iii) and Cd(ii) removal via adsorption onto the MnFe2O4–BC nanocomposite. The maximum adsorption capacities are 237.53 and 181.49 mg g−1 for Sb(iii) and Cd(ii), respectively, in a bi-solute system. Thus, the prepared MnFe2O4–BC nanocomposite is demonstrated to be a potential adsorbent for simultaneously removing Sb(iii) and Cd(ii) ions from aqueous solutions.

In this study, a jacobsite–biochar nanocomposite (MnFe2O4–BC) was fabricated and used to simultaneously remove Sb(iii) and Cd(ii) from water via adsorption.  相似文献   

11.
Four different ruthenium(ii) complexes were incorporated into the metal–organic framework (MOF) UiO-67 using three different synthetic strategies: premade linker synthesis, postsynthetic functionalization, and postsynthetic linker exchange. One of these complexes was of the type (N–N)3Ru2+, and three of the complexes were of the type (N–N)2(N–C)Ru+, where N–N is a bipyridine-type ligand and N–C is a cyclometalated phenylpyridine-type ligand. The resulting materials were characterized by PXRD, SC-XRD (the postsynthetic functionalization MOFs), N2 sorption, TGA-DSC, SEM, EDS, and UV-Vis spectroscopy, and were digested in base for subsequent 1H NMR analysis. The absorption profiles of the MOFs that were functionalized with cyclometalated Ru(ii) complexes extend significantly further into the visible region of the spectrum compared to the absorption profiles of the MOFs that were functionalized with the non-cyclometalated reference, (N–N)3Ru2+.

The metal–organic framework (MOF) UiO-67 was functionalized by incorporating different cyclometalated ruthenium(ii) complexes using three different methods: premade linker synthesis, postsynthetic functionalization, and postsynthetic linker exchange.  相似文献   

12.
The antimony(iii) complex of 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetate (DOTA) has been prepared and its exceptionally low stability observed. The Sb(iii) ion in Na[Sb(DOTA)]·4H2O shows an approximately square antiprismatic coordination geometry that is close to superimposable to the Bi(iii) geometry in [Bi(DOTA)] in two phases containing this anion, Na[Bi(DOTA)]·4H2O, [H3O][Bi(DOTA)]·H2O for which structures are also described. Interestingly, DOTA itself in [(H6DOTA)]Cl2·4H2O·DMSO shows the same orientation of the N4O4 metal binding cavity reflecting the limited flexibility of DOTA in an octadentate coordination mode. In 8-coordinate complexes it can however accommodate M(iii) ions with rion spanning a relatively wide range from 87 pm (Sc(iii)) to 117 pm (Bi(iii)). The larger Bi3+ ion appears to be the best metal–ligand size match since [Bi(DOTA)] is associated with greater complex stability. In the solution state, [Sb(DOTA)] is extremely susceptible to transmetallation by trivalent ions (Sc(iii), Y(iii), Bi(iii)) and, significantly, even by biologically important divalent metal ions (Mg(ii), Ca(ii), Zn(ii)). In all cases just one equivalent is enough to displace most of the Sb(iii). [Sb(DOTA)] is resistant to hydrolysis; however, since biologically more abundant metal ions easily substitute the antimony, DOTA complexes will not be suitable for deployment for the delivery of the, so far unexploited, theranostic isotope pair 119Sb and 117Sb.

The antimony(iii) complex of 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetate (DOTA) has been prepared and its exceptionally low stability observed.  相似文献   

13.
A series of metal complexes were prepared from separate reactions of lanthanide nitrate salts (La(iii), Ce(iii), Sm(iii), Gd(iii) and Ho(iii)) with 4-methylbenzoylhydrazide. The structures of the complexes were confirmed by analytical studies, spectral measurements and thermal studies. Complexes were formed with different stoichiometries of 1 : 2 and 1 : 3 (M : L). The ligand chelates by the nitrogen and oxygen atoms of the amino and carbonyl groups of the hydrazide moiety in the neutral keto form. The coordination compounds were converted to metal oxide nanoparticles (MONPs) through solid state thermal decomposition as monocular source precursors. The obtained MONPs were investigated via XRD, TEM and UV-Vis spectra. As a representative, CeO2 was utilized as a nanophotocatalyst to examine the photocatalytic activity of the MONPs for methylene blue (MB) photodegradation. CeO2 showed high removal of MB dye by 90.1% after UV illumination for 220 min. The reported method provides a generalized and systematic method for the preparation of many metal oxide nanoparticles with manageable and reproducible features.

A series of metal complexes were prepared from separate reactions of lanthanide nitrate salts (La(iii), Ce(iii), Sm(iii), Gd(iii) and Ho(iii)) with 4-methylbenzoylhydrazide.  相似文献   

14.
15.
Herein, nanoscale iron (oxyhydr)oxide-coated carbon nanotube (CNT) filters were rationally designed for rapid and effective removal of Sb(iii) from water. These iron (oxyhydr)oxide particles (<5 nm) were uniformly coated onto the CNT sidewalls. The as-fabricated hybrid filter demonstrated improved sorption kinetics and capacity compared with the conventional batch system. At a flow rate of 6 mL min−1, a Sb(iii) pseudo-first-order adsorption rate constant of 0.051 and a removal efficiency of >99% was obtained when operated in the recirculation mode. The improved Sb(iii) sorption performance can be ascribed to the synergistic effects of convection-enhanced mass transport, limited pore size, and more exposed active sorption sites of the filters. The presence of 1–10 mmol L−1 of carbonate, sulfate, and chloride inhibits Sb(iii) removal negligibly. Exhausted hybrid filters can be effectively regenerated by an electrical field-assisted chemical washing method. STEM characterization confirmed that Sb was mainly sequestered by iron (oxyhydr)oxides. XPS, AFS and XAFS results suggest that a certain amount of Sb(iii) was converted to Sb(v) during filtration. DFT calculations further indicate that the bonding energy for Sb(iii) onto the iron (oxyhydr)oxides was 2.27–2.30 eV, and the adsorbed Sb(iii) tends to be oxidized.

Herein, nanoscale iron (oxyhydr)oxide-coated carbon nanotube (CNT) filters were rationally designed for rapid and effective removal of Sb(iii) from water.  相似文献   

16.
In this study, a granular red mud supported zero-valent iron (ZVI@GRM) was successfully prepared and was used to remove Pb(ii) and Cr(vi) from aqueous solution. Zero-valent iron (ZVI) was synthesized by direct reduction of iron oxide in red mud by maize straw as a reductant at 900 °C in an anoxic atmosphere. The technical characterization (SEM, EDS, XRD, FTIR and BET) revealed that ZVI@GRM was loaded with zero-valent iron and contained different size pores. The factors of adsorption experiments include initial concentration, contact time, pH and temperature. The Pb(ii) and Cr(vi) removal by ZVI@GRM well fitted the pseudo-second-order kinetics model and the removal of heavy metals was an endothermic process. Essentially, Pb(ii) was transformed to precipitate forms (Pb0, Pb (OH)2, or 2PbCO3·Pb (OH)2) and Cr(vi) was converted to Cr (OH)3 or Cr3+/Fe3+ hydroxides. The maximum removal capacity for Pb(ii) and Cr(vi) by ZVI@GRM was 149.42 and 37.14 mg g−1. ZVI@GRM was a low-cost material and had outstanding performance and great potential in wastewater treatment.

In this study, a granular red mud supported zero-valent iron (ZVI@GRM) was successfully prepared and was used to remove Pb(ii) and Cr(vi) from aqueous solution.  相似文献   

17.
In this work, an outstanding nanolayered tin phosphate with 15.0 Å interlayer spacing, Sn (HPO4)2·3H2O (SnP–H+), has been synthesized by conventional hydrothermal method and first used in the adsorptive removal of Cr(iii) from aqueous solution. A number of factors such as contact time, initial concentration of Cr(iii), temperature, pH, and ionic strength on adsorption were investigated by batch tests. Moreover, the isothermal adsorption characteristics and kinetic model of Cr(iii) onto SnP–H+ were studied. The results showed that the adsorption of Cr(iii) by SnP–H+ was in accordance with the Langmuir adsorption isotherm model and the pseudo-second-order kinetic model. The adsorption capacity of Cr(iii) onto SnP–H+ at temperature 40.0 °C and pH 3.0 could reach 81.1 mg g−1. And the distribution coefficient Kd was 23.0 g L−1. Overall, experiments certified that SnP–H+ was an excellent adsorbent that can effectively remove Cr(iii) from aqueous solution.

In this work, an outstanding nanolayered tin phosphate with 15.0 Å interlayer spacing, Sn (HPO4)2·3H2O (SnP–H+), has been synthesized by conventional hydrothermal method and first used in the adsorptive removal of Cr(iii) from aqueous solution.  相似文献   

18.
Deep eutectic solvents (DESs) were used as alternatives to the aqueous phase in solvent extraction of iron(iii), zinc(ii) and lead(ii). The selective extraction of iron(iii) and zinc(ii) was studied from a feed of ethaline (1 : 2 molar ratio of choline chloride : ethylene glycol) and lactiline (1 : 2 molar ratio of choline chloride : lactic acid), with the former DES being more selective. A commercial mixture of trialkylphosphine oxides (Cyanex 923, C923) diluted in an aliphatic diluent selectively extracted iron(iii) from a feed containing also zinc(ii) and lead(ii). The subsequent separation of zinc(ii) from lead(ii) was carried out using the basic extractant Aliquat 336 (A336). The equilibration time and the extractant concentration were optimized for both systems. Iron(iii) and zinc(ii) were stripped using 1.2 mol L−1 oxalic acid and 0.5 mol L−1 aqueous ammonia, respectively. An efficient solvometallurgical flowsheet is proposed for the separation and recovery of iron(iii), lead(ii) and zinc(ii) from ethaline using commercial extractants. Moreover, the process was upscaled in a countercurrent mixer-settler set-up resulting in successful separation and purification.

Deep eutectic solvents (DESs) were used as alternatives to the aqueous phase in solvent extraction of iron(iii), zinc(ii) and lead(ii).  相似文献   

19.
Iron(ii) and iron(iii) salts of strong acids form iron glycerolates on heating at 180 °C with glycerol in the presence of an equivalent amount of alkali. Individual iron(iii) glycerolate was obtained for the first time. When Fe3O4 magnetic nanoparticles were heated with glycerol, an iron(iii) glycerolate shell was formed on their surface.

Individual iron(iii) glycerolate was obtained and characterized; a method for the preparation of an iron(iii) glycerolate shell on the surface of Fe3O4 MNPs was proposed.

Currently, glycerolates of various metals (Ti, Co, Fe, Zn, etc.) are used as catalytic systems1,2 or as precursors to obtain nanoparticles, including iron oxide magnetic nanoparticles (MNPs),3 and nanostructure materials for technical and biomedical applications.4–8Glycerolates of biogenic elements (Si, Zn, B and Ti) are of particular interest because of their biological activity. They are used as biocompatible precursors in the sol–gel synthesis of pharmacologically active hydrogels with reparative, regenerative, antioxidant, immunotropic and antimicrobial effects.8–10 In this regard, glycerolate of the biogenic iron element can be considered as an innovative biocompatible precursor in the sol–gel synthesis of composite bioactive hydrogels possessing a haemostatic effect characteristic of various iron compounds.11A promising trend in biomedicine is the core–shell modification of Fe3O4 MNPs for MRI diagnostics or magnetic hyperthermia of tumors.12–14 So, the development of an iron glycerolate shell on the surface of Fe3O4 MNPs15 and studying an opportunity of using modified nanoparticles in magnetic hyperthermia is of particular interest. In addition, the antibacterial activity of Fe3O4 MNPs with glycerol adsorbed on the surface was also demonstrated.16,17In the literature, individual iron(ii) and iron(iii) glycerolates have not so far been described. At the same time, the synthesis of individual forms is extremely important for biomedical purposes in order to determine bioavailability parameters. The available literature data concern only mixed iron(ii,iii) glycerolate that is usually formed as a result of the interaction of di- or trivalent iron oxides, hydroxides or salts (mainly oxalates) with glycerol at elevated temperatures (up to 245 °C).15,18–20 It is worth noting that all attempts to synthesize iron glycerolates from chlorides and sulfates of ferrous or ferric irons proved to be unsuccessful.20 At the same time, iron glycerolate was obtained from iron(iii) nitrate in boiling glycerol under reflux (280 °C);3 however, contents of iron(iii) and iron(ii) were not determined in that product.Regardless of the iron valence state in the starting compound, Fe(ii) and Fe(iii) are present in the resulting glycerolate in all cases. It should be noted that the possible pathways of the redox process for obtaining mixed iron(ii,iii) glycerolate are not discussed in the literature. The quantitative Fe(ii)/Fe(iii) ratio is usually determined by the Mössbauer spectroscopy21 or the colorimetric method.19 The composition of iron(ii,iii) glycerolate is mainly described by the following formulas: Fe2C6H11O6 (powder diffraction file JCPDSD-ICDD PDF 2, card [23-1731])18 and Fe3+2Fe2+3(C3H5O3)4.19–21 It was not possible to obtain single crystals of iron glycerolate and calculate the unit cell parameters.4,18We have found that the reactions of iron(ii) or iron(iii) chlorides and sulfates with glycerol proved to proceed only in the presence of an equivalent amount of alkali to give glycerolates of various chemical compositions. Thus, for the first time, individual iron(iii) glycerolate FeC3H5O3 (1) was obtained in 91% yield on heating iron(iii) chloride hexahydrate FeCl3·6H2O with sodium hydroxide in an excess of glycerol C3H8O3 at 180 °C for 18 h (Scheme 1) (see ESI).Open in a separate windowScheme 1Synthesis of iron(iii) glycerolate 1.The resulting product 1 is a light green powder insoluble in water and organic solvents, thus indicating a probable polymeric structure. It should be noted that the reaction temperature (180 °C) and duration (18 h) appear to be optimal taking into account a high yield of the product and its purity.Heating iron(ii) sulfate heptahydrate FeSO4·7H2O in glycerol in the presence of an equivalent amount of NaOH under the same conditions (180 °C, 18 h) resulted in mixed iron(ii,iii) glycerolate Fe3+2Fe5+3(C3H5O3)7 (2) in 83% yield (see ESI). The resulting product is a dark green powder that is poorly soluble in water and organic solvents.Iron glycerolates 1 and 2 were formed as colored powders; they are storage stable with no change in structure and no noticeable change in color; they do not melt to decomposition temperature. Dilute acids or hot water caused decomposition with the production of glycerol and iron (hydroxy)oxides or salts, as it was noted earlier.18 Plausible pathways for the formation of iron glycerolate 1, as well as iron glycerolate 2 (Scheme 2) and the features of the process are discussed below.Open in a separate windowScheme 2Formation of iron(ii,iii) glycerolate 2.Magnetic materials based on Fe3O4 nanoparticles with a biologically compatible coating are of great interest for biology and medicine.12–14 Previously, we were the first to demonstrate the possibility of forming a shell of iron glycerolate on the surface of Fe3O4 MNPs by a simple and reproducible method, namely, by interacting Fe3O4 MNPs with glycerol at 220 °C for 40 h.15 In this work, we optimized the synthetic procedure and chose the optimum conditions (180 °C, 18 h) (see ESI). The composition of the resulting shell was found to correspond to iron glycerolate 1.To determine the Fe(ii)/Fe(iii) ratio in the obtained products, we used the Mössbauer spectroscopy. Fig. 1 shows the Mössbauer spectra of iron glycerolates 1 (a and c) and 2 (b). The samples were prepared by deposition of the powder onto aluminum foil with a diameter of 22 mm (see ESI).Open in a separate windowFig. 1 57Fe Mössbauer spectra at 295 K of (a) iron(iii) glycerolate 1, (b) iron(ii,iii) glycerolate 2, and (c) iron(iii) glycerolate 1 from Fe3O4 MNPs. The doublets of Fe3+ and Fe2+ ions are marked in red and blue, respectively. The black line represents the sum of these lines. Sodium nitroprusside C5FeN6Na2O was taken as reference.The Mössbauer spectrum of iron glycerolate 1 (Fig. 1a) contains only one doublet (red line) with quadrupole splitting value of 0.48 mm s−1 (iii) positions,21 at the same time there are no signals typical for Fe(ii). The Mössbauer spectrum of iron glycerolate 2 (Fig. 1b) contains two doublets (red and blue lines) with quadrupole splitting Qs values of 0.46 and 2.29 mm s−1 (iii) and Fe(ii) positions, respectively.21 In this case, the content of Fe(ii) was 38%; Fe(iii), 62%.Fitting parameters of 57Fe Mössbauer spectra (Fig. 1) for iron glycerolates
SampleStarting materialSpectral linesIsomer shift, δiso (mm s−1) Q S (mm s−1)Relative content (%)Line width (mm s−1)
Iron(iii) glycerolate 1 (a)FeCl3·6H2OFe3+0.660.481000.31
Iron(iii) glycerolate 1 (c)Fe3O4 MNPsFe3+0.660.511000.33
Iron(ii,iii) glycerolate 2 (b)FeSO4·7H2OFe3+0.660.46620.24
Fe2+1.332.29380.30
Open in a separate windowThe Mössbauer spectrum of a sample obtained from Fe3O4 MNPs (Fig. 1c) also contains a doublet (red line) with Qs = 0.51 mm s−1 (iii). Any signals typical for Fe(ii) are absent, which confirms the presence of a shell of iron glycerolate 1. It should be noted that the signals of Fe(ii) contained in the core of the Fe3O4 MNPs were not recorded under these conditions of spectrum registration.The results of the quantitative determination of Fe(ii) and Fe(iii) by the Mössbauer spectroscopy in the studied products, as well as the data of their elemental analyses (
Iron glycerolateComposition (%)
ExperimentalCalculated
CHFeCHFe
FeC3H5O3 (1)24.753.4438.4024.863.4838.54
Fe2C6H11O6a24.703.2541.0024.783.8138.40
Fe3+2Fe2+3(C3H5O3)4b22.473.3143.6922.683.1743.94
Fe3+2Fe5+3(C3H5O3)7 (2)23.123.2041.8923.663.4841.22
Open in a separate windowaIron(ii,iii) glycerolate synthesized from goethite in boiling glycerole.18bIron(ii,iii) glycerolate synthesized from goethite, lepidocrocite, hematite at 245 °C.19It should be noted that the experimentally determined Fe content (%) in iron(ii,iii) glycerolate18 proved to differ considerably from the calculated composition for the formula Fe2C6H11O6. At the same time, the experimental results for iron(ii,iii) glycerolate19 are in good agreement with the formula Fe3+2Fe2+3(C3H5O3)4. The data of elemental analyses for the mixed iron glycerolate 2 obtained in this study correspond to the formula Fe3+2Fe5+3(C3H5O3)7. However, both mixed iron glycerolates differ significantly in their compositions. Thus, the chemical compositions of iron glycerolates differs significantly depending on the nature of the starting compounds and reaction conditions. Fig. 2 shows the X-ray diffraction (XRD) spectra of the powders of (a) iron glycerolate 1, (b) iron glycerolate 2, and (c) Fe3O4 MNPs with a shell of iron(iii) glycerolate 1. All spectra contain peak at 12.7 deg. 2θ (8.1 Å), which is the main diffraction line of iron glycerolate.18Open in a separate windowFig. 2X-ray diffraction spectra of (a) iron(iii) glycerolate 1, (b) iron(ii,iii) glycerolate 2, (c) Fe3O4 MNPs coated with glycerolate 1. In the insets: photos of the analysed powders. Fig. 3 shows HRTEM images of (a) starting Fe3O4 MNPs and (b) Fe3O4 MNPs coated with iron glycerolate 1. The modified Fe3O4 MNPs had a core–shell structure with an average magnetite core size of 10 nm coated with an iron glycerolate shell 2–4 nm thick.Open in a separate windowFig. 3HRTEM images of (a) starting Fe3O4 MNPs Fe3O4 and (b) MNPs coated with iron glycerolate 1; in inset, an electron diffraction pattern. Fig. 4 shows the IR spectra of (a) iron glycerolate 1, (b) iron glycerolate 2, (c) Fe3O4 MNPs, and (d) MNPS coated with glycerolate 1. The most intensive bands at 2850–2840 cm−1 correspond to the stretching vibration of C–H bonds. The bands in the range 1480–1200 cm−1 are attributed to the deformation vibrations of C–H bonds. The high intensity bands in the ranges 1150–880 and 780–690 cm−1 can be assigned to deformation vibrations of C–O–Fe groups in glycerolate fragments. Band at 537 cm−1 is characteristic of the initial Fe3O4 MNPs. In modified nanoparticles, it is likely to be shifted to 582 cm−1 and superimpose on the bands in the range 600–610 cm−1 corresponding to C–O–Fe vibrations. It should be noted that the IR spectrum of modified Fe3O4 MNPs is similar in position and shape of absorption bands to the spectrum of glycerolate 1. Our measurements are in agreement with the previous IR studies of iron glycerolates.3,15,21Open in a separate windowFig. 4(a) IR spectra of (a) iron(iii) glycerolate 1, (b) iron(ii,iii) glycerolate 2, (c) Fe3O4 MNPs and (d) MNPs coated with glycerolate 1.Thus, the IR and XRD spectra of iron glycerolates 1 and 2 turned out to be similar. Therefore, the quantitative determination of Fe(ii) and Fe(iii) contents in iron glycerolates 1 and 2 proved to be possible only by using the Mössbauer spectroscopy. The results obtained have allowed us to refine the information on the composition of iron glycerolate available in the XRD database [card 23-1731].18Considering the possible pathways for the formation of iron glycerolates 1 and 2, it can be assumed that the process includes the ion exchange reaction between iron salts and alkali with the formation of amorphous iron(iii) and iron(ii) hydroxides, respectively. In this case, iron(ii) hydroxide is partially oxidized by atmospheric oxygen to iron(iii) hydroxide; however, the reduction of iron(iii) to iron(ii) with glycerol does not occur under the reaction conditions. Then iron(ii) and iron(iii) hydroxides enter the reversible condensation reaction with glycerol to form iron glycerolates. As noted above, the resulting products are hydrolysed in hot water, which corresponds to a shift in the equilibrium in the condensation reaction towards the starting materials. However, an excess of glycerol and removal of water when the reaction mixture is heated up to 180 °C leads to a shift in the equilibrium towards the reaction products, as evidenced by their high yields.Thus, iron glycerolate 1 is formed from iron(iii) hydroxide when the latter is reacted with glycerole (Scheme 1); iron glycerolate 2 is formed from iron(ii) and iron(iii) hydroxides (Scheme 2). The graphical formula of iron glycerolate 2 is represented as a conditional notation corresponding to the molecular formula Fe3+2Fe5+3(C3H5O3)7. When m = n = 1, the graphical formula does correspond to the molecular formula Fe3+2Fe2+3(C3H5O3)4 (ref. 21) in iron glycerolate 2, Fe3+2Fe2+3(C3H5O3)4·3FeC3H5O3. It should be noted that all our attempts to synthesize individual iron(ii) glycerolate have failed, even in an inert gas atmosphere.It should be noted that the formation of iron hydroxides and, consequently, iron glycerolates does not occur without alkali, since the hydrolysis of iron salts of strong acids proceeds stepwise and results mainly in the formation of basic salts as the first step of the hydrolysis process. In addition, we cannot exclude that sodium monoglycerolate derived from equilibrium interaction of NaOH with an excess of glycerol is also involved in the formation of iron glycerolates 1 and 2. Sodium monoglycerolate, like NaOH, can enter the ion exchange reaction with iron salts with the formation of glyceroxy iron derivatives.The proposed chemistry of the process, in our opinion, is characteristic of all iron(ii) and iron(iii) salts of strong acids.It is known that iron salts of weak acids, for example, iron(ii) oxalate (as dihydrate FeC2O4·2H2O) or iron(iii) oxalate (as pentahydrate Fe2(C2O4)3·5H2O) are capable of reacting with glycerol without alkali at 240 °C to form mixed iron(ii,iii) glycerolate Fe3+2Fe2+3(C3H5O3)4.21 It is interesting to note that the chemical composition of iron(ii,iii) glycerolate is the same in both cases. It can be assumed that complete hydrolysis of these salts takes place, which is facilitated by an increased temperature, thus resulting in the formation of iron(ii) or iron(iii) hydroxides and oxalic acid. At the same time, in cases of FeC2O4 (as well as FeSO4), the formed iron(ii) hydroxide is partially oxidized by atmospheric oxygen to give iron(iii) hydroxide. Iron(ii) and iron(iii) hydroxides are supposed to react further with glycerol to form a mixed iron(ii,iii) glycerolate. We believe that in the case of Fe2(C2O4)3, the formed iron(iii) hydroxide is reduced by oxalic acid, which is also the product of hydrolysis, to iron(ii) hydroxide (Scheme 3, reaction 1). Further, when reacting with glycerol, a mixture of iron(ii) and iron(iii) hydroxides is expected to form a mixed iron(ii,iii) glycerolate. In our opinion, the reduction with glycerol (Scheme 3, reaction 2), according to the work,19 cannot be considered as the determining process.Open in a separate windowScheme 3Reduction of iron(iii) hydroxide with oxalic acid (1) or glycerol (2).19At the same time, to the best of our knowledge there are scarcely available data on oxidation of glycerol on reacting with iron(iii) salts. It has been established,19 that a carbonyl compound is present in the reaction mixture, as indicated by a low intensity CO stretching vibrational band. However, qualitative tests carried out to determine the character of the CO function proved to be negative for aldehydes and ketones.Thus, the key difference in the pathways for obtaining iron glycerolates from iron(ii) or (iii) salts of strong or weak acids is associated with the step of the formation of iron hydroxides: in case of iron salts of strong acids, iron hydroxides are formed due to the ion exchange reaction of iron salts with alkali, while iron salts of weak acids undergo their complete hydrolysis.In summary, it has first been shown that iron(ii) and iron(iii) salts of strong acids (FeCl3, FeSO4) are able to form iron glycerolates. The previously undescribed individual iron(iii) glycerolate FeC3H5O3 was obtained from FeCl3 as an example, glycerol and NaOH at 180 °C in one-pot synthesis. It has been found that due to the direct interaction of Fe3O4 MNPs with glycerol at 180 °C, an iron(iii) glycerolate shell is formed on the nanoparticle surface. The obtained iron glycerolates were characterized by Mössbauer and IR spectroscopy, XRD and elemental analysis.Individual iron(iii) glycerolate can be considered as a novel biocompatible precursor in the sol–gel synthesis of pharmacologically active nanocomposite materials and for further preparation of advanced composite magnetic nanomaterials with glycerolate shell to be used in magnetic hyperthermia of tumors. In addition, it can be used as a precursor for in the preparation of MNPs and as a catalyst for various chemical processes.  相似文献   

20.
An efficient mercapto-functionalized organic–inorganic hybrid sorbent for selective separation and preconcentration of antimony(iii) in water samples     
Nan You  Tian-Hong Liu  Hong-Tao Fan  Hua Shen 《RSC advances》2018,8(10):5106
This work reported on the application of mercapto-functionalized silica-supported organic–inorganic hybrid sorbent as a solid phase extraction (SPE) extractant for effective separation and preconcentration of Sb(iii) species in real water samples. The influences of pH, sorbent amounts, flow rates and the concentration of eluent on the adsorption and desorption of Sb(iii) species had been evaluated. The recovery of Sb(iii) species at pH 5 with 100 mg mercapto-functionalized hybrid sorbent at the flow rate of 5.0 mL min−1 was greater than 95% without interference from all of metal ions tested. The trapped Sb(iii) species by extractant was then eluted with 5% HCl solution at the flow rate of 5.0 mL min−1. The proposed procedure permitted large enrichment factors of about 200 and higher for 10 μg L−1 of Sb(iii) species. The merits of analytical figures for the determination of Sb(iii) species were as follows: detection limit (3σ, n = 11), 2 ng L−1; precision, 1.6% (n = 11) for 10 μg L−1 of Sb(iii) species; the linear calibration curve presented in the concentration range of 1.0–200.0 μg L−1. The validity of the proposed procedure was checked by the analysis of standard reference materials. Excellent agreement between the analytical results and the certified values (t-test at 95% confidence level) was found. The mercapto-functionalized hybrid sorbent as a SPE extractant was applied to the determination of Sb(iii) species in various water samples with satisfactory results.

This work reported on the application of mercapto-functionalized silica-supported organic–inorganic hybrid sorbent as a solid phase extraction (SPE) extractant for effective separation and preconcentration of Sb(iii) species in real water samples.  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