To see the other types of publications on this topic, follow the link: Iron oxalate.

Journal articles on the topic 'Iron oxalate'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Iron oxalate.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Niskanen, Raina. "Extractable aluminium, iron and manganese in mineral soils: II Extractability by oxalate and pyrophosphate." Agricultural and Food Science 61, no. 2 (March 1, 1989): 79–87. http://dx.doi.org/10.23986/afsci.72356.

Full text
Abstract:
The extractability of aluminium, iron and manganese by 0.05 M oxalate and pyrophosphate was studied in samples of 23 mineral soils. Dilute extractants were studied because conventional reagents may cause problems in analytical work. The mean values for Al, Fe and Mn extracted by conventional Tamm’s oxalate were 67, 81 and 1.5 mmol/kg soil, respectively. On the average, 0.05 M oxalate solutions at pH 2.9 and 4.2 extracted Al, Fe and Mn amounts that were 103, 113 and 87 % and 72, 82 and 83 % of the amounts extractable by Tamm’s oxalate, respectively. Eeach metal released by 0.05 M oxalates correlated closely with that dissolved by Tamm’s oxalate; the r values ranged from 0.967*** to 0.997***. The mean values for Al, Fe and Mn extracted by 0.1 M Na4P2O7 and 0.05 M K4P2O7 were 38, 28 and 0.6 and 33, 29 and 0.6 mmol/kg soil, respectively. The amount of each metal extracted by Na4P2O7 correlated closely with that released by K4P2O7; the r values ranged from 0.87*** to 0.97***.
APA, Harvard, Vancouver, ISO, and other styles
2

Diefallah, El-H. M., M. A. Mousa, A. A. El-Bellihi, E. H. El-Mossalamy, G. A. El-Sayed, and M. A. Gabal. "Thermal decomposition of iron(II) oxalate–magnesium oxalate mixtures." Journal of Analytical and Applied Pyrolysis 62, no. 2 (February 2002): 205–14. http://dx.doi.org/10.1016/s0165-2370(01)00117-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Basahel, S. N., A. A. El-Bellihi, M. Gabal, and El-H. M. Diefallah. "Thermal decomposition of iron(III) oxalate-magnesium oxalate mixtures." Thermochimica Acta 256, no. 2 (June 1995): 339–46. http://dx.doi.org/10.1016/0040-6031(94)02158-k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Benhacine, Mohamed Al Amine, Malika Hamadène, Sofiane Bouacida, and Hocine Merazig. "The new one-dimensional coordination polymercatena-poly[[diaquasodium(I)]-μ-oxalato-[diaquairon(III)]-μ-oxalato]." Acta Crystallographica Section C Structural Chemistry 72, no. 3 (February 29, 2016): 243–50. http://dx.doi.org/10.1107/s2053229616002953.

Full text
Abstract:
The oxalate dianion is one of the most studied ligands and is capable of bridging two or more metal centres and creating inorganic polymers based on the assembly of metal polyhedra with a wide variety of one-, two- or three-dimensional extended structures. Yellow single crystals of a new mixed-metal oxalate, namelycatena-poly[[diaquasodium(I)]-μ-oxalato-κ4O1,O2:O1′,O2′-[diaquairon(III)]-μ-oxalato-κ4O1,O2:O1′,O2′], [NaFe(C2O4)2(H2O)4]n, have been synthesized and the crystal structure elucidated by X-ray diffraction analysis. The compound crystallizes in the noncentrosymmetric space groupI41(Z= 4). The asymmetric unit contains one NaIand one FeIIIatom lying on a fourfold symmetry axis, one μ2-bridging oxalate ligand and two aqua ligands. Each metal atom is surrounded by two chelating oxalate ligands and two equivalent water molecules. The structure consists of infinite one-dimensional chains of alternating FeO4(H2OW1)2and NaO4(H2OW2)2octahedra, bridged by oxalate ligands, parallel to the [100] and [010] directions, respectively. Because of thecisconfiguration and the μ2-coordination mode of the oxalate ligands, the chains run in a zigzag manner. This arrangement facilitates the formation of hydrogen bonds between neighbouring chains involving the H2O and oxalate ligands, leading to a two-dimensional framework. The structure of this new one-dimensional coordination polymer is shown to be unique among theAIMIII(C2O4)2(H2O)nseries. In addition, the absorption bands in the IR and UV–Visible regions and their assignments are in good agreement with the local symmetry of the oxalate ligand and the irregular environment of iron(III). The final product of the thermal decomposition of this precursor is the well-known ternary oxide NaFeO2.
APA, Harvard, Vancouver, ISO, and other styles
5

Gabal, M. A., A. A. El-Bellihi, and H. H. El-Bahnasawy. "Non-isothermal decomposition of zinc oxalate–iron(II) oxalate mixture." Materials Chemistry and Physics 81, no. 1 (July 2003): 174–82. http://dx.doi.org/10.1016/s0254-0584(03)00183-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Gismonti, Pedro Rosário, Jéssica Frontino Paulino, and Julio Afonso. "RECOVERY OF METALS FROM ELECTROACTIVE COMPONENTS OF SPENT Ni-MH BATTERIES AFTER LEACHING WITH FORMIC ACID." Detritus, no. 14 (March 31, 2021): 68–77. http://dx.doi.org/10.31025/2611-4135/2021.14063.

Full text
Abstract:
this work describes a route for recovering nickel, cobalt, iron, zinc and lanthanides from spent nickel-metal hydride batteries. Formic acid was used as leachant. Experiments were run at 25-50°C for 1-4 h. Under the best conditions leaching yields surpassed 99 wt.%, except for iron. The insoluble matter contains almost solely iron as iron(III) basic formate. The leachate went through six separation procedures, combining solvent extraction with D2EHPA as extractant, and precipitation reactions. Fe2+ and Zn2+ were extracted together (> 99 wt.%) from the original leachate (pH ~1.5). Yttrium and lanthanides were precipitated as oxalates directly from the raffinate (> 99.9 wt.%) upon addition of sodium oxalate. In the next steps, Mn2+ and Co2+ were extracted with D2EHPA at buffered pH (3 and ~4.8, respectively), after adding NaOHaq. About 10 wt.% of leached Ni2+ was coextracted with Co2+. The remaining Ni2+ was precipitated from the raffinate after addition of aqueous sodium oxalate at pH 6. After precipitation of Al3+ upon addition of NaOHaq. until pH ~8, sodium formate was recovered after slow evaporation of the final aqueous solution at 60oC. It contains ~90 wt.% of the formate present in the leachant.
APA, Harvard, Vancouver, ISO, and other styles
7

Niskanen, Raina. "Release of phosphorus, aluminium and iron in fractionation of inorganic soil phosphorus." Agricultural and Food Science 59, no. 2 (April 1, 1987): 141–45. http://dx.doi.org/10.23986/afsci.72256.

Full text
Abstract:
Release of phosphorus, aluminium and iron by a modified Chang and Jackson procedure was studied in five mineral soils. Quantities of aluminium and iron released during the procedure and extracted by acid ammonium oxalate were compared. The extractability of P, Al and Fe by 1 M NH4CI and that of Al and Fe by alkaline 0.5 M NH4F was poor. Proportions of P extracted by 0.5 M NH4F (0.2—10.4 mmol/kg soil) and 0.1 M NaOH (0.1— 9.8 mmol/kg soil) were related to the molar ratio of oxalate-extractable iron and aluminium. P extracted by 0.25 M H2SO4 amounted to 2.1—12.2 mmol/kg soil. Al extracted by 0.1 M NaOH (7—174 mmol/kg soil) and 0.25 M H2SO4 (17—112 mmol/kg soil) amounted to 55—94 % and 16—245 % of oxalate-extractable Al, respectively. Fe released by 0.1 M NaOH (1—10 mmol/kg soil) and 0.25 M H2SO4 (30—196 mmol/kg soil) amounted to 1—13 % and 62—272 % of oxalate-extractable Fe, respectively. In total, 91—309 % of oxalate-extractable Al and 70—285 % of oxalate-extractable Fe were released by NaOH and H2SO4.
APA, Harvard, Vancouver, ISO, and other styles
8

Paris, R., and K. V. Desboeufs. "Effect of atmospheric organic complexation on iron-bearing dust solubility." Atmospheric Chemistry and Physics Discussions 13, no. 2 (February 4, 2013): 3179–202. http://dx.doi.org/10.5194/acpd-13-3179-2013.

Full text
Abstract:
Abstract. Recent studies reported that the effect of organic complexation may be a potentially important process to be considered in models to estimate atmospheric iron flux to the ocean. In this study, we investigated this effect by a series of dissolution experiments on iron-bearing dust in presence or absence of various organic compounds typically found in the atmospheric waters (acetate, formate, oxalate, malonate, succinate, glutarate, glycolate, lactate, tartrate and humic acid as an analogue of humic like substances (HULIS)). Only 4 of tested organic ligands (oxalate, malonate, tartrate and humic acid) caused an enhancement of iron solubility which was associated with an increase of dissolved Fe(II) concentrations. For all of these organic ligands, a positive linear dependence of iron solubility to organic concentrations was observed and showed that the extent of organic complexation on iron solubility decreased in order oxalate > malonate = tartrate > humic acid. This was attributed to the ability of electron donors of organic ligands and implied a reductive ligand-promoted dissolution. This study confirmed that oxalate is the most effective ligand playing on dust iron solubility and showed, for the first time, the potential effect of HULIS on iron dissolution in atmospheric conditions.
APA, Harvard, Vancouver, ISO, and other styles
9

Piro, O. E., G. A. Echeverría, and E. J. Baran. "Spontaneous enantiomorphism in poly-phased alkaline salts of tris(oxalato)ferrate(III): crystal structure of cubic NaRb5[Fe(C2O4)3]2." Acta Crystallographica Section E Crystallographic Communications 74, no. 7 (June 8, 2018): 905–9. http://dx.doi.org/10.1107/s2056989018008022.

Full text
Abstract:
We show here that the phenomenon of spontaneous resolution of enantiomers occurs during the crystallization of the sodium and rubidium double salts of the transition metal complex tris(oxalato)ferrate(III), namely sodium pentarubidium bis[tris(oxalato)ferrate(III)], NaRb5[Fe(C2O4)3]2. One enantiomer of the salt crystallizes in the cubic space groupP4332 withZ= 4 and a Flack absolute structure parameterx= −0.01 (1) and its chiral counterpart in the space groupP4132 withx= −0.00 (1). All metal ions are at crystallographic special positions: the iron(III) ion is on a threefold axis, coordinated by three oxalate dianions in a propeller-like conformation. One of the two independent rubidium ions is on a twofold axis in an eightfold coordination with neighbouring oxalate oxygen atoms, and the other one on a threefold axis in a sixfold RbO6coordination. The sodium ion is at a site ofD3point group symmetry in a trigonal–antiprismatic NaO6coordination.
APA, Harvard, Vancouver, ISO, and other styles
10

Luo, Chao, and Yuan Gao. "Aeolian iron mobilisation by dust - acid interactions and their implications for soluble iron deposition to the ocean: a test involving potential anthropogenic organic acidic species." Environmental Chemistry 7, no. 2 (2010): 153. http://dx.doi.org/10.1071/en09116.

Full text
Abstract:
Environmental context. Studying the input of atmospheric soluble iron to the ocean is important as the soluble form of iron is bioavailable for phytoplankton uptake in the surface ocean to support photosynthesis. In this paper, the effect of organic acidic species on atmospheric iron dissolution is addressed through a global model for the first time. The new results contribute to a better understanding of iron dissolution processes in the atmosphere and the role of atmospheric iron in ocean biogeochemical cycles. Abstract. Dust deposition is a major source of iron in certain oceanic regions. Many atmospheric processes, such as heterogeneous reactions with acidic species, may convert insoluble iron in dust to soluble forms that become bioavailable for phytoplankton uptake in the surface ocean. Here we report for the first time the effects of organic acidic species on iron dissolution using laboratory-measured conversion rates by oxalate, simulated in a global model to estimate soluble iron fluxes to the ocean. With the complexity and limited data from measurements relating to different sources for oxalate, we focus on the effect of oxalate of anthropogenic origin in this work as a first-step testing, and we apply a scaling factor for oxalate based on its relationship with aerosol sulfate observed by in situ measurements in the continental sites. The results show better correlation with the observations than the work including inorganic acids alone, suggesting the contribution of organic acids to Fe dissolution. However, the simulated iron solubility is lower than that derived from measurements, suggesting additional processes may contribute to Fe dissolution that should be included in the model. Total deposition of soluble iron to the global ocean including the effect by anthropogenic oxalate is ~0.34 Tg year–1.
APA, Harvard, Vancouver, ISO, and other styles
11

Paris, R., and K. V. Desboeufs. "Effect of atmospheric organic complexation on iron-bearing dust solubility." Atmospheric Chemistry and Physics 13, no. 9 (May 14, 2013): 4895–905. http://dx.doi.org/10.5194/acp-13-4895-2013.

Full text
Abstract:
Abstract. Recent studies reported that the effect of organic complexation may be a potentially important process to be considered by models estimating atmospheric iron flux to the ocean. In this study, we investigated this process effect by a series of dissolution experiments on iron-bearing dust in the presence or the absence of various organic compounds (acetate, formate, oxalate, malonate, succinate, glutarate, glycolate, lactate, tartrate and humic acid as an analogue of humic like substances, HULIS) typically found in atmospheric waters. Only 4 of tested organic ligands (oxalate, malonate, tartrate and humic acid) caused an enhancement of iron solubility which was associated with an increase of dissolved Fe(II) concentrations. For all of these organic ligands, a positive linear dependence of iron solubility to organic concentrations was observed and showed that the extent of organic complexation on iron solubility decreased in the following order: oxalate >malonate = tartrate > humic acid. This was attributed to the ability of electron donors of organic ligands and implies a reductive ligand-promoted dissolution. This study confirms that among the known atmospheric organic binding ligands of Fe, oxalate is the most effective ligand promoting dust iron solubility and showed, for the first time, the potential effect of HULIS on iron dissolution under atmospheric conditions.
APA, Harvard, Vancouver, ISO, and other styles
12

Balmer, Marianne E., and Barbara Sulzberger. "Atrazine Degradation in Irradiated Iron/Oxalate Systems: Effects of pH and Oxalate." Environmental Science & Technology 33, no. 14 (July 1999): 2418–24. http://dx.doi.org/10.1021/es9808705.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Goldberg, Margarita A., Marat R. Gafurov, Fadis F. Murzakhanov, Alexander S. Fomin, Olga S. Antonova, Dinara R. Khairutdinova, Andrew V. Pyataev, et al. "Mesoporous Iron(III)-Doped Hydroxyapatite Nanopowders Obtained via Iron Oxalate." Nanomaterials 11, no. 3 (March 22, 2021): 811. http://dx.doi.org/10.3390/nano11030811.

Full text
Abstract:
Mesoporous hydroxyapatite (HA) and iron(III)-doped HA (Fe-HA) are attractive materials for biomedical, catalytic, and environmental applications. In the present study, the nanopowders of HA and Fe-HA with a specific surface area up to 194.5 m2/g were synthesized by a simple precipitation route using iron oxalate as a source of Fe3+ cations. The influence of Fe3+ amount on the phase composition, powders morphology, Brunauer–Emmett–Teller (BET) specific surface area (S), and pore size distribution were investigated, as well as electron paramagnetic resonance and Mössbauer spectroscopy analysis were performed. According to obtained data, the Fe3+ ions were incorporated in the HA lattice, and also amorphous Fe oxides were formed contributed to the gradual increase in the S and pore volume of the powders. The Density Functional Theory calculations supported these findings and revealed Fe3+ inclusion in the crystalline region with the hybridization among Fe-3d and O-2p orbitals and a partly covalent bond formation, whilst the inclusion of Fe oxides assumed crystallinity damage and rather occurred in amorphous regions of HA nanomaterial. In vitro tests based on the MG-63 cell line demonstrated that the introduction of Fe3+ does not cause cytotoxicity and led to the enhanced cytocompatibility of HA.
APA, Harvard, Vancouver, ISO, and other styles
14

Kashyap, Rajarshi, Dhruba Joyti Talukdar, and Sanjay Pratihar. "Iron oxalate capped iron–copper nanomaterial for oxidative transformation of aldehydes." New Journal of Chemistry 39, no. 2 (2015): 1430–37. http://dx.doi.org/10.1039/c4nj01966k.

Full text
Abstract:
An efficient, sustainable and green procedure for the synthesis of selective orthorhombic iron(oxalate) capped Fe–Cu bimetallic oxide nanomaterial [Fe(ox)Fe–CuOx] was developed using a sodium borohydride reduction of iron(ii) salt in the presence of oxalic acid at room temperature followed by addition of copper sulfate in water.
APA, Harvard, Vancouver, ISO, and other styles
15

Teer, J. E., D. J. Leak, A. W. L. Dudeney, A. Narayanan, and D. C. Stuckey. "The treatment of iron oxalate leach liquors in a UASB with sulfate reduction." Water Science and Technology 36, no. 6-7 (September 1, 1997): 383–90. http://dx.doi.org/10.2166/wst.1997.0614.

Full text
Abstract:
The presence of small amounts of iron (>0.013% Fe) in sand creates problems in the manufacture of high quality glass. Removal by hot sulphuric acid is possible, but creates environmental problems, and is costly. Hence organic acids such as oxalic have been investigated since they are effective in removing iron, and can be degraded anaerobically. The aim of this work was to identify key intermediates in the anaerobic degradation of oxalate in an upflow anaerobic sludge blanket reactor (UASB) which was removing iron from solution in the sulphide form, and to determine the bacterial species involved. 2-bromoethanesulfonic acid (BES) and molybdenum were selected as suitable inhibitors for methanogenic and sulphate reducing bacteria (SRB) respectively. 40mM molybdenum was used to inhibit the SRB in a reactor with a 12hr HRT. Total SRB inhibition took place in 20 hrs, with a complete breakthrough of influent sulphate. The lack of an immediate oxalate breakthrough confirmed Desulfovibrio vulgaris subspecies oxamicus was not the predominant oxalate utilising species. Nevertheless, high concentrations of molybdenum were found to inhibit oxalate utilising bacteria in granular reactors but not in suspended population reactors; this observation was puzzling, and at present cannot be explained. Based on the intermediates identified, it was postulated that oxalate was degraded to formate by an oxalate utilising bacteria such as Oxalobacter formigenes, and the formate used by the SRBs to reduce sulphate. Acetate, as a minor intermediate, existed primarily as a source of cell carbon for oxalate utilising bacteria. Methanogenic inhibition identified that 62% of the CH4 in the reactor operated at 37°C originated from hydrogenotrophic methanogenesis, whilst this figure was 80% at 20°C. Possible irreversible effects were recorded with hydrogenotrophic methanogens.
APA, Harvard, Vancouver, ISO, and other styles
16

Lin, Qinhao, Xinhui Bi, Guohua Zhang, Yuxiang Yang, Long Peng, Xiufeng Lian, Yuzhen Fu, et al. "In-cloud formation of secondary species in iron-containing particles." Atmospheric Chemistry and Physics 19, no. 2 (January 30, 2019): 1195–206. http://dx.doi.org/10.5194/acp-19-1195-2019.

Full text
Abstract:
Abstract. The increase in secondary species through cloud processing potentially increases aerosol iron (Fe) bioavailability. In this study, a ground-based counterflow virtual impactor coupled with a real-time single-particle aerosol mass spectrometer was used to characterize the formation of secondary species in Fe-containing cloud residues (dried cloud droplets) at a mountain site in southern China for nearly 1 month during the autumn of 2016. Fe-rich, Fe-dust, Fe-elemental carbon (Fe-EC), and Fe-vanadium (Fe-V) cloud residual types were obtained in this study. The Fe-rich particles, related to combustion sources, contributed 84 % (by number) to the Fe-containing cloud residues, and the Fe-dust particles represented 12 %. The remaining 4 % consisted of the Fe-EC and Fe-V particles. It was found that above 90 % (by number) of Fe-containing particles had already contained sulfate before cloud events, leading to no distinct change in number fraction (NF) of sulfate during cloud events. Cloud processing contributed to the enhanced NFs of nitrate, chloride, and oxalate in the Fe-containing cloud residues. However, the in-cloud formation of nitrate and chloride in the Fe-rich type was less obvious relative to the Fe-dust type. The increased NF of oxalate in the Fe-rich cloud residues was produced via aqueous oxidation of oxalate precursors (e.g., glyoxylate). Moreover, Fe-driven Fenton reactions likely increase the formation rate of aqueous-phase OH, improving the conversion of the precursors to oxalate in the Fe-rich cloud residues. During daytime, the decreased NF of oxalate in the Fe-rich cloud residues was supposed to be due to the photolysis of Fe-oxalate complexes. This work emphasizes the role of combustion Fe sources in participating in cloud processing and has important implications for evaluating Fe bioavailability from combustion sources during cloud processing.
APA, Harvard, Vancouver, ISO, and other styles
17

Straub, Steffen, and Peter Vöhringer. "Ultrafast “end-on”-to-“side-on” binding-mode isomerization of an iron–carbon dioxide complex." Physical Chemistry Chemical Physics 23, no. 33 (2021): 17826–35. http://dx.doi.org/10.1039/d1cp02300d.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Varela, Elisa, and Ming Tien. "Effect of pH and Oxalate on Hydroquinone-Derived Hydroxyl Radical Formation during Brown Rot Wood Degradation." Applied and Environmental Microbiology 69, no. 10 (October 2003): 6025–31. http://dx.doi.org/10.1128/aem.69.10.6025-6031.2003.

Full text
Abstract:
ABSTRACT The redox cycle of 2,5-dimethoxybenzoquinone (2,5-DMBQ) is proposed as a source of reducing equivalent for the regeneration of Fe2+ and H2O2 in brown rot fungal decay of wood. Oxalate has also been proposed to be the physiological iron reductant. We characterized the effect of pH and oxalate on the 2,5-DMBQ-driven Fenton chemistry and on Fe3+ reduction and oxidation. Hydroxyl radical formation was assessed by lipid peroxidation. We found that hydroquinone (2,5-DMHQ) is very stable in the absence of iron at pH 2 to 4, the pH of degraded wood. 2,5-DMHQ readily reduces Fe3+ at a rate constant of 4.5 × 103 M−1s−1 at pH 4.0. Fe2+ is also very stable at a low pH. H2O2 generation results from the autoxidation of the semiquinone radical and was observed only when 2,5-DMHQ was incubated with Fe3+. Consistent with this conclusion, lipid peroxidation occurred only in incubation mixtures containing both 2,5-DMHQ and Fe3+. Catalase and hydroxyl radical scavengers were effective inhibitors of lipid peroxidation, whereas superoxide dismutase caused no inhibition. At a low concentration of oxalate (50 μM), ferric ion reduction and lipid peroxidation are enhanced. Thus, the enhancement of both ferric ion reduction and lipid peroxidation may be due to oxalate increasing the solubility of the ferric ion. Increasing the oxalate concentration such that the oxalate/ferric ion ratio favored formation of the 2:1 and 3:1 complexes resulted in inhibition of iron reduction and lipid peroxidation. Our results confirm that hydroxyl radical formation occurs via the 2,5-DMBQ redox cycle.
APA, Harvard, Vancouver, ISO, and other styles
19

Yang, Zhang, Liu, Ning, Han, Liu, and Huo. "Preparation of Iron Carbides Formed by Iron Oxalate Carburization for Fischer–Tropsch Synthesis." Catalysts 9, no. 4 (April 9, 2019): 347. http://dx.doi.org/10.3390/catal9040347.

Full text
Abstract:
Different iron carbides were synthesized from the iron oxalate precursor by varying the CO carburization temperature between 320 and 450 °C. These iron carbides were applied to the high-temperature Fischer–Tropsch synthesis (FTS) without in situ activation treatment directly. The iron oxalate as a precursor was prepared using a solid-state reaction treatment at room temperature. Pure Fe5C2 was formed at a carburization temperature of 320 C, whereas pure Fe3C was formed at 450 °C. Interestingly, at intermediate carburization temperatures (350–375 °C), these two phases coexisted at the same time although in different proportions, and 360 °C was the transition temperature at which the iron carbide phase transformed from the Fe5C2 phase to the Fe3C phase. The results showed that CO conversions and products selectivity were affected by both the iron carbide phases and the surface carbon layer. CO conversion was higher (75–96%) when Fe5C2 was the dominant iron carbide. The selectivity to C5+ products was higher when Fe3C was alone, while the light olefins selectivity was higher when the two components (Fe5C2 and Fe3C phases) co-existed, but the quantity of Fe3C was small.
APA, Harvard, Vancouver, ISO, and other styles
20

Choudhury, Amitava, Srinivasan Natarajan, and C. N. R. Rao. "Hybrid Open-Framework Iron Phosphate-Oxalates Demonstrating a Dual Role of the Oxalate Unit." Chemistry – A European Journal 6, no. 7 (April 3, 2000): 1168–75. http://dx.doi.org/10.1002/(sici)1521-3765(20000403)6:7<1168::aid-chem1168>3.0.co;2-s.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Choudhury, Amitava, Srinivasan Natarajan, and C. N. R. Rao. "Hybrid Open-Framework Iron Phosphate-Oxalates Demonstrating a Dual Role of the Oxalate Unit." Chemistry - A European Journal 6, no. 7 (April 3, 2000): 1168–75. http://dx.doi.org/10.1002/(sici)1521-3765(20000403)6:7<1168::aid-chem1168>3.3.co;2-j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Skjemstad, JO, HVA Bushby, and RW Hansen. "Extractable Fe in the surface horizons of a range of soils from Queensland." Soil Research 28, no. 2 (1990): 259. http://dx.doi.org/10.1071/sr9900259.

Full text
Abstract:
The levels of iron and aluminium extracted from 36 surface soils by pyrophosphate, oxalate and dithionite are compared with a number of other soil properties. Correlations suggest that aluminium released by these extraction procedures is largely associated with organic matter while only a small fraction of the iron released is in this form. Significant correlations between soil pH and the negative logarithms of both oxalate (r = 0.715) and pyrophosphate (r = 0.959) extractable iron in soils with >20% clay content indicate that pH is the most significant factor in determining the level of ferrihydrite and iron/organic matter complexes in surface soils. The significance of these relationships in terms of soil weathering processes is discussed. Further, the data suggest that pyrophosphate extractable iron is a useful indicator of the most active, mobile component of iron in surface soils.
APA, Harvard, Vancouver, ISO, and other styles
23

González, Ernesto, F. González, J. A. Muñoz, M. Luisa Blázquez, and Antonio Ballester. "Reductive Dissolution of Iron Oxides and Manganese Bioleaching by Acidiphilium cryptum JF-5." Advanced Materials Research 1130 (November 2015): 347–50. http://dx.doi.org/10.4028/www.scientific.net/amr.1130.347.

Full text
Abstract:
In the development of new processes to use the potential of iron reducing bacteria,Acidiphilium cryptum, the main bacteria involved in the reduction of Fe (III) compounds in acidic environments, could play an important biohydrometallurgical role. Thus, the bioleaching of hematite, goethite and a low-grade manganese ore was assayed, in vials and columns, using three different media; two of which included a ligand, oxalate, or a redox mediator, thionine.Although the presence ofA. cryptumwas essential for promoting the dissolution of both iron oxides and the bioleaching of manganese ore, the addition of oxalate to the media tripled and quadrupled the microbial dissolution of hematite and goethite, respectively. Oxalate also had a positive effect in assays performed in columns, however, the addition of thionine to the medium allowed to reach significant hematite dissolution.
APA, Harvard, Vancouver, ISO, and other styles
24

Gabal, M. A., A. A. El-Bellihi, and S. S. Ata-Allah. "Effect of calcination temperature on Co(II) oxalate dihydrate–iron(II) oxalate dihydrate mixture." Materials Chemistry and Physics 81, no. 1 (July 2003): 84–92. http://dx.doi.org/10.1016/s0254-0584(03)00137-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Kutalkova, E., T. Plachy, J. Osicka, M. Cvek, M. Mrlik, and M. Sedlacik. "Electrorheological behavior of iron(ii) oxalate micro-rods." RSC Advances 8, no. 44 (2018): 24773–79. http://dx.doi.org/10.1039/c8ra03409e.

Full text
Abstract:
The application of rod-like iron(ii) oxalates particles led to significant electrorheological effect as proved e.g. via the creep-recovery experiments under the application of an external electric field.
APA, Harvard, Vancouver, ISO, and other styles
26

Zelinka, Samuel L., Carl J. Houtman, Kolby Hirth, Steven Lacher, Linda Lorenz, Emil Engelund Thybring, and Christopher G. Hunt. "The Effect of Acetylation on Iron Uptake and Diffusion in Water Saturated Wood Cell Walls and Implications for Decay." Forests 11, no. 10 (October 21, 2020): 1121. http://dx.doi.org/10.3390/f11101121.

Full text
Abstract:
Acetylation is widely used as a wood modification process that protects wood from fungal decay. The mechanisms by which acetylation protects wood are not fully understood. With these experiments, we expand upon the literature and test whether previously observed differences in iron uptake by wood were a result of decreased iron binding capacity or slower diffusion. We measured the concentration of iron in 2 mm thick wood sections at 0, 10, and 20% acetylation as a function of time after exposure to iron solutions. The iron was introduced either strongly chelated with oxalate or weakly chelated with acetate. The concentrations of iron and oxalate in solution were chosen to be similar to those found during brown rot decay, while the concentration of iron and acetate matched previous work. The iron content of oxalate-exposed wood increased only slightly and was complete within an hour, suggesting little absorption and fast diffusion, or only slight surface adsorption. The increase in iron concentration from acetate solutions with time was consistent with Fickian diffusion, with a diffusion coefficient on the order of 10−16 m2 s−1. The rather slow diffusion rate was likely due to significant binding of iron within the wood cell wall. The diffusion coefficient did not depend on the acetylation level; however, the capacity for iron absorption from acetate solution was greatly reduced in the acetylated wood, likely due to the loss of OH groups. We explored several hypotheses that might explain why the diffusion rate appears to be independent of the acetylation level and found none of them convincing. Implications for brown rot decay mechanisms and future research are discussed.
APA, Harvard, Vancouver, ISO, and other styles
27

Krause, M., C. Michalk, B. Lippold, M. Benedix, and J. Suwalski. "Slow paramagnetic relaxation in oxalate (2,2′-bipyridine) iron." physica status solidi (a) 97, no. 2 (October 16, 1986): 521–26. http://dx.doi.org/10.1002/pssa.2210970225.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Rudenkov, A. S., M. A. Yarmolenko, A. A. Rogachev, A. P. Surzhikov, A. P. Luchnikov, and O. A. Frolova. "Phase composition and morphology of nanostructured coatings deposited by laser dispersion of a mixture of polyethylene with iron oxalate." Bulletin of the Karaganda University. "Physics" Series 99, no. 3 (September 30, 2020): 22–30. http://dx.doi.org/10.31489/2020ph3/22-30.

Full text
Abstract:
Peculiarities of forming of iron oxide coatings with reinforced carbon nanostructures from gas phase generated by laser dispersion of composite target were explored. Influence of technological modes of heat treatment on morphology and phase composition of nanostructured film layers was determined. It was found that on a substrate highly dispersed layers containing carbon nanostructures are formed. Using Raman spectroscopy it was shown that in oxide matrix carbon structures, which are mainly in the form of planar located nanotubes, appear. It was found that with a mass ratio of polyethylene and iron oxalate equal to 1:1, the distribution of the formed nanostructures in size is unimodal with a maximum near 20 nm. At dispersing of polyethylene and iron oxalate mixture with mass ratio 1:2 in deposited layers nanotubes have the least defectiveness. Patterns of influence on morphology and coatings phase composition of relative component abundance in being dispersed by laser radiation composite target were determined. It was shown that with the growing of iron oxalate concentration in the target coating structural heterogeneity increases, subroughness and average size of separate nanostructures in the deposited condensate grow. The obtained polymer matrix nanocomposite films can be used in sensors.
APA, Harvard, Vancouver, ISO, and other styles
29

Milardović, Stjepan, Zorana Grabarić, Vlatko Rumenjak, Nenad Blau, and Dražen Milošević. "Use of a Ruthenium(III), Iron(II), and Nickel(II) Hexacyanometallate-Modified Graphite Electrode with Immobilized Oxalate Oxidase for the Determination of Urinary Oxalate." Journal of AOAC INTERNATIONAL 84, no. 6 (November 1, 2001): 1927–33. http://dx.doi.org/10.1093/jaoac/84.6.1927.

Full text
Abstract:
Abstract This paper describes the performance of a biosensor with an Ru(III), Ni(II), and Fe(II) hexacyanometallate-modified graphite electrode and immobilized oxalate oxidase for the determination of urinary oxalate. The addition of ruthenium enhances the electrochemical reversibility and chemical stability of the electrocrystallized layer and improves the sensitivity of the biosensor. Hydrogen peroxide, produced by the enzyme-catalyzed oxidation of oxalate, was measured at −50 mV vs an Hg|Hg2Cl2|3M KCl electrode in a solution of pH 3.6 succinic buffer, 0.1M KCl, and 5.4mM ethylenediaminetetraacetic acid. The linear concentration range for the determination of oxalate was 0.18–280 μM. The recoveries of added oxalate (10–35 μM) from aqueous solution ranged from 99.5 to 101.7%, whereas from urine samples without oxalate (or with a concentration of oxalate below the detection limit) the recoveries of added oxalate ranged from 91.4 to 106.6%. The oxalate in 24 h urine samples, taken during their daily routine from 35 infants and children, was measured and found to range from 0.6 to 121.7 mg/L. There were no interferences from uric acid, acetylsalicylic acid, and urea in the concentration range investigated, but paracetamol and ascorbic acid did interfere. A good correlation (R2 = 0.9242) was found between values obtained for oxalate in real urine samples by 2 laboratories, with the proposed biosensor and ion chromatography, respectively.
APA, Harvard, Vancouver, ISO, and other styles
30

Pawlowski, Jake W., Noelle Kellicker, Cedric E. Bobst, and Igor A. Kaltashov. "Assessing the iron delivery efficacy of transferrin in clinical samples by native electrospray ionization mass spectrometry." Analyst 141, no. 3 (2016): 853–61. http://dx.doi.org/10.1039/c5an02159f.

Full text
Abstract:
Serum transferrin is a key player in iron homeostasis, and its ability to deliver iron to cellsviathe endosomal pathway critically depends on the nature of anion (carbonate or oxalate) that binds this protein synergistically with ferric ion.
APA, Harvard, Vancouver, ISO, and other styles
31

Chandorkar, Suneeta S., and Kafila Jaipuri. "Enhanced In Vitro Iron Availability from Traditional Foods of Western India: Effect of Soaking, Germination and Fermentation." South Asian Journal of Experimental Biology 2, no. 4 (September 26, 2012): 177–83. http://dx.doi.org/10.38150/sajeb.2(4).p177-183.

Full text
Abstract:
Soaking, germination and fermentation are simple techniques which are widely practiced at household level in India. These are also reported to improve the nutritive value of foods. However, regional data on composition and bioavailability of nutrients from cooked food are scanty. Therefore, selected foods commonly consumed in the western region of India were analysed for phytates, tannates, oxalates, calcium, phosphorus, ascorbic acid, total, soluble and ionisable iron. Subsequently, the percent bioavailable iron was calculated from the foods under study. The foods selected were, soaked and cooked field beans, red gram, Bengal gram, kabuli chana, cow pea, peas, soyabean; germinated and soyabean, green gram, lentil, moth beans and fermented foods prepared using rice and split legume combination viz. idli, khaman, dhokla and handwa. All the processing treatments brought about a significant reduction in the phytate, tannate and oxalate content. A concomitant increase was observed in the soluble and ionisable iron content. Ascorbic acid showed a negligible increase in fermented foods only. The calcium: phosphorus ratio improved on processing.
APA, Harvard, Vancouver, ISO, and other styles
32

Dehner, Carolyn A., Jonathan D. Awaya, Patricia A. Maurice, and Jennifer L. DuBois. "Roles of Siderophores, Oxalate, and Ascorbate in Mobilization of Iron from Hematite by the Aerobic Bacterium Pseudomonas mendocina." Applied and Environmental Microbiology 76, no. 7 (January 29, 2010): 2041–48. http://dx.doi.org/10.1128/aem.02349-09.

Full text
Abstract:
ABSTRACT In aerobic, circumneutral environments, the essential element Fe occurs primarily in scarcely soluble mineral forms. We examined the independent and combined effects of a siderophore, a reductant (ascorbate), and a low-molecular-weight carboxylic acid (oxalate) on acquisition of Fe from the mineral hematite (α-Fe2O3) by the obligate aerobe Pseudomonas mendocina ymp. A site-directed ΔpmhA mutant that was not capable of producing functional siderophores (i.e., siderophore− phenotype) did not grow on hematite as the only Fe source. The concentration of an added exogenous siderophore (1 μM desferrioxamine B [DFO-B]) needed to restore wild-type (WT)-like growth kinetics to the siderophore− strain was ∼50-fold less than the concentration of the siderophore secreted by the WT organism grown under the same conditions. The roles of a reductant (ascorbate) and a simple carboxylic acid (oxalate) in the Fe acquisition process were examined in the presence and absence of the siderophore. Addition of ascorbate (50 μM) alone restored the growth of the siderophore− culture to the WT levels. A higher concentration of oxalate (100 μM) had little effect on the growth of a siderophore− culture; however, addition of 0.1 μM DFO-B and 100 μM oxalate restored the growth of the mutant to WT levels when the oxalate was prereacted with the hematite, demonstrating that a metabolizing culture benefits from a synergistic effect of DFO-B and oxalate.
APA, Harvard, Vancouver, ISO, and other styles
33

Hong, Seok Yong, Dong Hyun Chun, Jung-Il Yang, Heon Jung, Ho-Tae Lee, Sungjun Hong, Sanha Jang, Jung Tae Lim, Chul Sung Kim, and Ji Chan Park. "A new synthesis of carbon encapsulated Fe5C2 nanoparticles for high-temperature Fischer–Tropsch synthesis." Nanoscale 7, no. 40 (2015): 16616–20. http://dx.doi.org/10.1039/c5nr04546k.

Full text
Abstract:
A novel Fe5C2@C catalyst bearing small iron carbide particles ∼10 nm in diameter was prepared using a simple thermal treatment of iron oxalate dihydrate cubes, employed in high-temperature Fischer–Tropsch synthesis.
APA, Harvard, Vancouver, ISO, and other styles
34

Bolland, MDA, RJ Gilkes, RF Brennan, and DG Allen. "Comparison of seven phosphorus sorption indices." Soil Research 34, no. 1 (1996): 81. http://dx.doi.org/10.1071/sr9960081.

Full text
Abstract:
Seven methods of estimating or predicting phosphorus (P) sorption capacity were compared for 47 neutral to acid Western Australian soils. Two methods, the P buffer capacity (PBC) and the Fox and Kamprath procedure, provided reliable indices of P sorption from well defined P sorption isotherms, but they are not quick routine methods because several levels of P addition are required. The other five routine procedures included two versions of the P retention index (PRI), determined by adding one level of P, and three soil properties, oxalate extractable iron (oxalate Fe), oxalate extractable aluminium (oxalate Al), and pH measured in sodium fluoride [pH (F)], that are known to indicate P sorption capacity. All the indices were well related to one another. The oxalate Fe index was the least well related to PBC whereas oxalate Al, one of the PRI indices, and pH (F) were closely related to PBC and could be used as quick, economical procedures to assess the P sorption capacity of soils.
APA, Harvard, Vancouver, ISO, and other styles
35

Capitan, F., A. Arrebola Ramirez, and C. Jimenez Linares. "Spectrophotometric determination of trace amounts of iron(III) in oxalates by extraction of the mixed-ligand iron-oxalate-purpurin complex." Analyst 110, no. 7 (1985): 819. http://dx.doi.org/10.1039/an9851000819.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

Kumar, Harish, and A. K. Shukla. "Fabrication Fe/Fe3O4/Graphene Nanocomposite Electrode Material for Rechargeable Ni/Fe Batteries in Hybrid Electric Vehicles." International Letters of Chemistry, Physics and Astronomy 19 (October 2013): 15–25. http://dx.doi.org/10.18052/www.scipress.com/ilcpa.19.15.

Full text
Abstract:
Fe/Fe3O4/Graphene composite electrode material was synthesized by a thermal reduction method and then used as anode material along with Nickel cathode in rechargeable Ni/Fe alkaline batteries in hybrid electric vehicles. Reduced graphene /Fe/Fe3O4 composite electrode material was prepared using a facile three step synthesis involving synthesis of iron oxalate and subsequent reduction of exfoliated graphene oxide and iron oxalate by thermal decomposition method. The synthesis approach presents a promising route for a large-scale production of reduced graphene /Fe/Fe3O4 composite as electrode material for Ni/Fe rechargeable batteries. The particle size and structure of the samples were characterized by SEM and XRD.
APA, Harvard, Vancouver, ISO, and other styles
37

Pegu, Rupa, Krishna Joyti Majumdar, Dhruba Joyti Talukdar, and Sanjay Pratihar. "Oxalate capped iron nanomaterial: from methylene blue degradation to bis(indolyl)methane synthesis." RSC Adv. 4, no. 63 (2014): 33446–56. http://dx.doi.org/10.1039/c4ra04214j.

Full text
Abstract:
An efficient, sustainable and green procedure for the synthesis of selective orthorhombic iron(oxalate) capped Fe(0) [Fe(ox)–Fe(0)] nanomaterial is developed using sodium borohydride (NaBH4) reduction of iron(ii) salt in the presence of oxalic acid at room temperature in water.
APA, Harvard, Vancouver, ISO, and other styles
38

Molinier, Michel, Daniel J. Price, Paul T. Wood, and Annie K. Powell. "Biomimetic control of iron oxide and hydroxide phases in the iron oxalate system *." Journal of the Chemical Society, Dalton Transactions, no. 21 (1997): 4061–68. http://dx.doi.org/10.1039/a704400c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Kizewski, Fiona R., Paul Boyle, Dean Hesterberg, and James D. Martin. "Mixed Anion (Phosphate/Oxalate) Bonding to Iron(III) Materials." Journal of the American Chemical Society 132, no. 7 (February 24, 2010): 2301–8. http://dx.doi.org/10.1021/ja908807b.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Tyapkin, P. Yu, S. A. Petrov, A. P. Chernyshev, A. I. Ancharov, L. A. Sheludyakova, and N. F. Uvarov. "Structural features of hydrate forms of iron(III) oxalate." Journal of Structural Chemistry 57, no. 6 (November 2016): 1134–40. http://dx.doi.org/10.1134/s0022476616060111.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Reddy, S. Lakshmi, Kenichi Uehara, and Tamio Endo. "Synthesis of Nano Iron Oxalate – Structures and Optical Transitions." MRS Proceedings 1454 (2012): 273–77. http://dx.doi.org/10.1557/opl.2012.1231.

Full text
Abstract:
AbstractSynthesis of FeC2O42H2O nano particles was carried out by thermal double decomposition of solutions of oxalic acid dihydrate (C2H2O4 2H2O) and FeSO4 7H2O employing CATA -2R microwave reactor. Structural elucidation was carried out by employing X-ray diffraction, particle size and shape were studied by transmission electron microscopy and nature of bonding was investigated by Optical absorption and near-infrared spectral studies. The powder resulting from this method is possesses distorted rhombic octahedral structure. The particle grain size is about 50 nm. Details of optical transitions are mentioned in terms of energy states.
APA, Harvard, Vancouver, ISO, and other styles
42

Mangani, Stefano, and Luigi Messori. "EXAFS studies on the oxalate adduct of iron transferrin." Journal of Inorganic Biochemistry 46, no. 1 (April 1992): 1–6. http://dx.doi.org/10.1016/0162-0134(92)80057-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Shekhanov, R. F., and S. N. Gridchin. "Electrodeposition of zinc-iron coatings from ammonium oxalate baths." Гальванотехника и обработка поверхности 29, no. 2 (2021): 19–24. http://dx.doi.org/10.47188/0869-5326_2021_29_2_19.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Gabal, M. A., S. S. Ata-Allah, A. O. Al-Youbi, S. N. Basahel, and S. A. Al-Thabaiti. "Formation of LaFeO3 and thermal decomposition reactions in lanthanum(III) oxalate–iron(II) oxalate crystalline mixture." Journal of Materials Science 41, no. 22 (October 11, 2006): 7597–603. http://dx.doi.org/10.1007/s10853-006-0848-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

PRISSANAROON, W., N. BRACK, P. J. PIGRAM, and J. LIESEGANG. "CO-DOPED POLYPYRROLE COATINGS FOR STAINLESS STEEL PROTECTION." Surface Review and Letters 13, no. 02n03 (April 2006): 319–27. http://dx.doi.org/10.1142/s0218625x06008268.

Full text
Abstract:
Polypyrrole (PPy) films have been successfully electrodeposited on stainless steel substrates in aqueous solution. In this work, three systems of electrolytes were studied: oxalic acid, dodecylbenzenesulfonic acid (DBSA) and a mixture of oxalic acid and DBSA. A combination of XPS and TOF–SIMS revealed the formation of an iron oxalate layer at the interface between the oxalic acid-doped PPy (PPy(Ox)) and stainless steel and a thin layer of DBSA was observed at the interface between DBSA-doped PPy (PPy(DBSA)) and stainless steel. Similar to the PPy(Ox) system, an iron oxalate was also present at the co-doped PPy/stainless steel interface. Cyclic voltammetry indicated that an iron oxalate layer initially formed at the surface of the stainless steel when the co-doping system was used. The adhesion strength and corrosion performance of the PPy coating on stainless steel were evaluated by lap shear tests and an anodic potentiodynamic polarization technique, respectively. The co-doped PPy-coated stainless steel exhibited the best adhesion and a significant shift of corrosion potential to the positive direction. This finding opens the possibility for the co-doped PPy coating to be deployed as a strongly adherent corrosion inhibitor by using a simple one-step electropolymerization process.
APA, Harvard, Vancouver, ISO, and other styles
46

Ocampo-López, C., M. E. Ramírez-Carmona, and E. Vélez-Ortiz. "Thermodynamic analysis of stability in iron removal from kaolin by using oxalic acid." Cerâmica 59, no. 350 (June 2013): 326–30. http://dx.doi.org/10.1590/s0366-69132013000200019.

Full text
Abstract:
The graphical representation of global stability for a system, or Pourbaix diagram, was constructed to perform a thermodynamic study of iron removal from kaolin using oxalic acid as an oxidant. To do this the free energies of formation of the oxalate complex of the system were calculated, and it was found that the more stable specie is Fe(C2O4)3-3, with a calculated free energy of formation of -3753.88 kcal/mol. Thermodynamic stability functions were estimated for the system as a function of pH and Eh known as potential of oxide reduction. It was built a global stability diagram for the removal system; it showed that the specie trioxalate Fe(C2O4)3-3 is the only oxalate in equilibrium with other compounds associated with the removal of iron in kaolin.
APA, Harvard, Vancouver, ISO, and other styles
47

Lewis, DG, and CM Cardile. "Hydrolysis of FeIII solution to hydrous iron oxides." Soil Research 27, no. 1 (1989): 103. http://dx.doi.org/10.1071/sr9890103.

Full text
Abstract:
A range of hydrous iron oxides was prepared from ferric nitrate solutions by varying the rate and manner of hydrolysis. The properties of the resultant products were determined by using various techniques including size fractionation with Amicon Diafilters, rate and extent of dissolution in acid ammonium oxalate solution, X-ray diffraction and Mossbauer spectroscopy. The variation in properties is interpreted in terms of the conditions under which the various materials had been formed. The products of spontaneous hydrolysis at low pH and low FeIII concentration had properties distinctly different from those samples in which hydrolysis was induced rapidly by addition of alkali. The low-pH products were far more resistant to dissolution in acid ammonium oxalate and were relatively well ordered as determined by X-ray diffraction and Mossbauer spectroscopy at 77 K. The crystallographic and magnetic ordering of the rapidly formed phase is normally very limited, but these properties were increased significantly if formation occurred in the presence of low-pH polymeric material which may act as a template for freshly precipitating oxides.
APA, Harvard, Vancouver, ISO, and other styles
48

Krause, M., C. Michalk, B. Lippold, M. Benedix, and J. Suwalski. "Identification of three different iron sites in a polymeric oxalate (2,2-bipyridine) iron compound." Inorganica Chimica Acta 126, no. 1 (January 1987): 45–47. http://dx.doi.org/10.1016/s0020-1693(00)81238-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Gnanamani, Muthu Kumaran, Hussein H. Hamdeh, Wilson D. Shafer, Shelley D. Hopps, and Burtron H. Davis. "Hydrogenation of carbon dioxide over iron carbide prepared from alkali metal promoted iron oxalate." Applied Catalysis A: General 564 (August 2018): 243–49. http://dx.doi.org/10.1016/j.apcata.2018.07.034.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Fine, P., and M. J. Singer. "Contribution of Ferrimagnetic Minerals to Oxalate- and Dithionite-Extractable Iron." Soil Science Society of America Journal 53, no. 1 (January 1989): 191–96. http://dx.doi.org/10.2136/sssaj1989.03615995005300010035x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography