Praseodymium(III,IV) oxide

Last updated
Praseodymium(III,IV) oxide
Names
IUPAC name
Praseodymium(III,IV) oxide
Identifiers
3D model (JSmol)
ECHA InfoCard 100.031.676 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 234-857-9
PubChem CID
  • InChI=1S/11O.6Pr
    Key: AICMLAQRDYRMRP-UHFFFAOYSA-N
  • O=[Pr]=O.O=[Pr]=O.O=[Pr]=O.O=[Pr]=O.O=[Pr]O[Pr]=O
Properties
Pr6O11
Molar mass 1021.44 g/mol
Appearancedark brown powder
Density 6.5 g/mL
Melting point 2,183 °C (3,961 °F; 2,456 K). [1]
Boiling point 3,760 °C (6,800 °F; 4,030 K) [1]
Hazards
GHS labelling:
GHS-pictogram-exclam.svg
Warning
H315, H319, H335
P261, P305+P351+P338
Lethal dose or concentration (LD, LC):
5000 mg·kg−1 Rat oral
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

Praseodymium(III,IV) oxide is the inorganic compound with the formula Pr6O11 that is insoluble in water. [2] It has a cubic fluorite structure. [3] It is the most stable form of praseodymium oxide at ambient temperature and pressure. [4]

Contents

Properties and structure

Pr6O11 adopts a cubic fluorite crystal structure, measured by XRD, TEM and SEM methods. [3] [5] It can be considered an oxygen deficient form of praseodymium(IV) oxide (PrO2), with the Pr ions being in a mixed valency state Pr(III) and Pr(IV). [5] This characteristic is what gives the oxide its many useful properties for its catalytic activity.

Synthesis

Praseodymium oxide nanoparticles are generally produced via solid-state methods such as thermolysis, molten salt method, calcination or precipitation. [3] [4] [6] Practically all processes, however, contain a calcination step in order to obtain a crystalline Pr6O11 nanoparticles.

Calcination

Typically, praseodymium nitrate Pr(NO3)3·6H2O [3] [5] or praseodymium hydroxide Pr(OH)3 [7] is heated at high temperatures (usually above 500 °C) under air to give praseodymium(III,IV) oxide. While less common, synthesis from other organic precursors such as praseodymium acetate, oxalate [8] and malonate [9] have also been reported in chemical literature.

The physical properties of the prepared nanoparticles such as particle shape or lattice parameter depend strongly on the conditions of calcination, such as the temperature or duration, as well as the different preparation methods (calcination, sol-gel, precipitation, for example). As a result, many synthesis routes have been explored to obtain the precise morphology desired. [3] [4] [5]

Uses

Praseodymium(III,IV) oxide has a number of potential applications in chemical catalysis, and is often used in conjunction with a promoter such as sodium or gold to improve its catalytic performance. It has a high-K dielectric constant of around 30 and very low leakage currents [10] which have also made it a promising material for many potential applications in nanodevices and microelectronics. [6]

Oxidative coupling of methane

Sodium or lithium promoted praseodymium(III,IV) oxide displays good conversion rate of methane with a good selectivity towards ethane and ethene as opposed to unwanted byproducts such as carbon dioxide. [11] [12] While the precise mechanism for this reaction is still under debate, it has been proposed that typically, methane is activated to a methyl radical by oxygen on the surface of the catalyst which combines to form ethane. Ethene is then formed by reduction of ethane either by the catalyst or spontaneously. The multiple oxidation states of Pr(III) and Pr(IV) allows rapid regeneration of the active catalyst species involving a peroxide anion O2−2. [11]

This reaction is of particular interest as it enables the conversion of abundant methane gas (composing up to 60% of natural gas) [11] [12] into higher order hydrocarbons, which provide more applications. As a result, the oxidative coupling of methane is an economically desirable process.

CO oxidation

In the proposed mechanism for Pr6O11catalysed oxidation of CO to CO2, CO first binds to the catalyst surface to create a bidentate carbonate then converted to a monodentate carbonate species which can decompose as CO2, completing the catalyst cycle. The conversion of a bidentate carbonate to a monodentate species leaves an oxygen vacancy on the catalyst surface which can quickly be filled due to the high oxygen mobility deriving from the mixed oxidation states of Pr centres. This proposed mechanism is presented schematically below, adapted from Borchert, et al. [5]

Praseodymium oxide-catalyzed CO oxidation mechanism CO oxidation mechanism.png
Praseodymium oxide-catalyzed CO oxidation mechanism

Addition of gold promoters to the catalyst may significantly lower the reaction temperature from 550 °C to 140 °C, but the mechanism is yet to be discovered. It is believed that there is a certain synergistic effect between gold and praseodymium(III,IV) oxide species. [13]

The interest in CO oxidation lies in its ability to convert toxic CO gas to non-toxic CO2 and has applications in car exhaust, for example, which emits CO. [14]

Pr6O11 is also used in conjunction with other additives such as silica or zircon to produce pigments for use in ceramics and glass [15]

Related Research Articles

<span class="mw-page-title-main">Catalysis</span> Process of increasing the rate of a chemical reaction

Catalysis is the increase in rate of a chemical reaction due to an added substance known as a catalyst. Catalysts are not consumed by the reaction and remain unchanged after it. If the reaction is rapid and the catalyst recycles quickly, very small amounts of catalyst often suffice; mixing, surface area, and temperature are important factors in reaction rate. Catalysts generally react with one or more reactants to form intermediates that subsequently give the final reaction product, in the process of regenerating the catalyst.

<span class="mw-page-title-main">Praseodymium</span> Chemical element, symbol Pr and atomic number 59

Praseodymium is a chemical element; it has symbol Pr and the atomic number 59. It is the third member of the lanthanide series and is considered one of the rare-earth metals. It is a soft, silvery, malleable and ductile metal, valued for its magnetic, electrical, chemical, and optical properties. It is too reactive to be found in native form, and pure praseodymium metal slowly develops a green oxide coating when exposed to air.

In chemistry, dehydrogenation is a chemical reaction that involves the removal of hydrogen, usually from an organic molecule. It is the reverse of hydrogenation. Dehydrogenation is important, both as a useful reaction and a serious problem. At its simplest, it's a useful way of converting alkanes, which are relatively inert and thus low-valued, to olefins, which are reactive and thus more valuable. Alkenes are precursors to aldehydes, alcohols, polymers, and aromatics. As a problematic reaction, the fouling and inactivation of many catalysts arises via coking, which is the dehydrogenative polymerization of organic substrates.

<span class="mw-page-title-main">Hopcalite</span> Catalyst to oxidise carbon monoxide at room temperature

Hopcalite is the trade name for a number of mixtures that mainly consist of oxides of copper and manganese, which are used as catalysts for the conversion of carbon monoxide to carbon dioxide when exposed to the oxygen in the air at room temperature.

The water–gas shift reaction (WGSR) describes the reaction of carbon monoxide and water vapor to form carbon dioxide and hydrogen:

<span class="mw-page-title-main">Photocatalysis</span> Acceleration of a photoreaction in the presence of a catalyst

In chemistry, photocatalysis is the acceleration of a photoreaction in the presence of a photocatalyst, the excited state of which "repeatedly interacts with the reaction partners forming reaction intermediates and regenerates itself after each cycle of such interactions." In many cases, the catalyst is a solid that upon irradiation with UV- or visible light generates electron–hole pairs that generate free radicals. Photocatalysts belong to three main groups; heterogeneous, homogeneous, and plasmonic antenna-reactor catalysts. The use of each catalysts depends on the preferred application and required catalysis reaction.

<span class="mw-page-title-main">Cerium(IV) oxide</span> Chemical compound

Cerium(IV) oxide, also known as ceric oxide, ceric dioxide, ceria, cerium oxide or cerium dioxide, is an oxide of the rare-earth metal cerium. It is a pale yellow-white powder with the chemical formula CeO2. It is an important commercial product and an intermediate in the purification of the element from the ores. The distinctive property of this material is its reversible conversion to a non-stoichiometric oxide.

<span class="mw-page-title-main">Catalytic reforming</span> Chemical process used in oil refining

Catalytic reforming is a chemical process used to convert petroleum refinery naphthas distilled from crude oil into high-octane liquid products called reformates, which are premium blending stocks for high-octane gasoline. The process converts low-octane linear hydrocarbons (paraffins) into branched alkanes (isoparaffins) and cyclic naphthenes, which are then partially dehydrogenated to produce high-octane aromatic hydrocarbons. The dehydrogenation also produces significant amounts of byproduct hydrogen gas, which is fed into other refinery processes such as hydrocracking. A side reaction is hydrogenolysis, which produces light hydrocarbons of lower value, such as methane, ethane, propane and butanes.

<span class="mw-page-title-main">Rhodium(III) oxide</span> Chemical compound

Rhodium(III) oxide (or Rhodium sesquioxide) is the inorganic compound with the formula Rh2O3. It is a gray solid that is insoluble in ordinary solvents.

<span class="mw-page-title-main">Manganese(II,III) oxide</span> Chemical compound

Manganese(II,III) oxide is the chemical compound with formula Mn3O4. Manganese is present in two oxidation states +2 and +3 and the formula is sometimes written as MnO·Mn2O3. Mn3O4 is found in nature as the mineral hausmannite.

Methanation is the conversion of carbon monoxide and carbon dioxide (COx) to methane (CH4) through hydrogenation. The methanation reactions of COx were first discovered by Sabatier and Senderens in 1902.

<span class="mw-page-title-main">Catalyst support</span> Porous material with a high specific surface area supporting a catalyst

In chemistry, a catalyst support is a material, usually a solid with a high surface area, to which a catalyst is affixed. The activity of heterogeneous catalysts is mainly promoted by atoms present at the accessible surface of the material. Consequently, great effort is made to maximize the specific surface area of a catalyst. One popular method for increasing surface area involves distributing the catalyst over the surface of the support. The support may be inert or participate in the catalytic reactions. Typical supports include various kinds of activated carbon, alumina, and silica.

The oxidative coupling of methane (OCM) is a potential chemical reaction studied in the 1980s for the direct conversion of natural gas, primarily consisting of methane, into value-added chemicals. Although the reaction would have strong economics if practicable, no effective catalysts are known, and thermodynamic arguments suggest none can exist.

Cuprospinel is a mineral. Cuprospinel is an inverse spinel with the chemical formula CuFe2O4, where copper substitutes some of the iron cations in the structure. Its structure is similar to that of magnetite, Fe3O4, yet with slightly different chemical and physical properties due to the presence of copper.

<span class="mw-page-title-main">Carbon nanotube supported catalyst</span> Novel catalyst using carbon nanotubes as the support instead of the conventional alumina

Carbon nanotube supported catalyst is a novel supported catalyst, using carbon nanotubes as the support instead of the conventional alumina or silicon support. The exceptional physical properties of carbon nanotubes (CNTs) such as large specific surface areas, excellent electron conductivity incorporated with the good chemical inertness, and relatively high oxidation stability makes it a promising support material for heterogeneous catalysis.

Gadolinium-doped ceria (GDC) (known alternatively as gadolinia-doped ceria, gadolinium-doped cerium oxide (GCO), cerium-gadolinium oxide (CGO), or cerium(IV) oxide, gadolinium-doped, formula Gd:CeO2) is a ceramic electrolyte used in solid oxide fuel cells (SOFCs). It has a cubic structure and a density of around 7.2 g/cm3 in its oxidised form. It is one of a class of ceria-doped electrolytes with higher ionic conductivity and lower operating temperatures (<700 °C) than those of yttria-stabilized zirconia, the material most commonly used in SOFCs. Because YSZ requires operating temperatures of 800–1000 °C to achieve maximal ionic conductivity, the associated energy and costs make GDC a more optimal (even "irreplaceable", according to researchers from the Fraunhofer Society) material for commercially viable SOFCs.

<span class="mw-page-title-main">Heterogeneous gold catalysis</span>

Heterogeneous gold catalysis refers to the use of elemental gold as a heterogeneous catalyst. As in most heterogeneous catalysis, the metal is typically supported on metal oxide. Furthermore, as seen in other heterogeneous catalysts, activity increases with a decreasing diameter of supported gold clusters. Several industrially relevant processes are also observed such as H2 activation, Water-gas shift reaction, and hydrogenation. One or two gold-catalyzed reactions may have been commercialized.

Praseodymium(III) fluoride is an inorganic compound with the formula PrF3, being the most stable fluoride of praseodymium.

Steven L. Suib is an American inorganic chemist, academic and researcher. He is a Board of Trustees Distinguished Professor of Chemistry at University of Connecticut. He is a director of the Institute of Materials Science and of the Center for Advanced Microscopy and Materials Analysis.

Praseodymium compounds are compounds formed by the lanthanide metal praseodymium (Pr). In these compounds, praseodymium generally exhibits the +3 oxidation state, such as PrCl3, Pr(NO3)3 and Pr(CH3COO)3. However, compounds with praseodymium in the +2 and +4 oxidation states, and unlike other lanthanides, the +5 oxidation state, are also known.

References

  1. 1 2 "Praseodymium Oxide Nanoparticles (Pr6O11) – Properties, Applications". AZoNano.com. 2013-04-17. Retrieved 2018-03-15.
  2. "Praseodymium Oxide (Pr6O11)". www.reade.com. Retrieved 2018-03-15.
  3. 1 2 3 4 5 Matović, Branko; Pantić, Jelena; Prekajski, Marija; Stanković, Nadežda; Bučevac, Dušan; Minović, Tamara; Čebela, Maria (2013). "Synthesis and characterization of Pr6O11 nanopowders". Ceramics International. 39 (3): 3151–3155. doi:10.1016/j.ceramint.2012.09.098.
  4. 1 2 3 Zinatloo-Ajabshir, Sahar; Salavati-Niasari, Masoud (2015). "Novel poly(ethyleneglycol)-assisted synthesis of praseodymium oxide nanostructures via a facile precipitation route". Ceramics International. 41 (1): 567–575. doi:10.1016/j.ceramint.2014.08.105.
  5. 1 2 3 4 5 Borchert, Yulia; Sonström, Patrick; Wilhelm, Michaela; Borchert, Holger; Bäumer, Marcus (2008). "Nanostructured Praseodymium Oxide: Preparation, Structure, and Catalytic Properties". The Journal of Physical Chemistry C. 112 (8): 3054–3063. doi:10.1021/jp0768524. ISSN   1932-7447.
  6. 1 2 Shamshi Hassan, M., Shaheer Akhtar, M., Shim, KB. et al. Morphological and Electrochemical Properties of Crystalline Praseodymium Oxide Nanorods. Nanoscale Res Lett 5, 735 (2010). https://doi.org/10.1007/s11671-010-9547-8
  7. Ma, Lin; Chen, Weixiang; Zhao, Jie; Zheng, Yifeng; Li, Xiang; Xu, Zhude (2007). "Microwave-assisted synthesis of praseodymium hydroxide nanorods and thermal conversion to oxide nanorod". Materials Letters. 61 (8–9): 1711–1714. doi:10.1016/j.matlet.2006.07.116.
  8. Hussein, Gamal A.M. (1994). "Formation of praseodymium oxide from the thermal decomposition of hydrated praseodymium acetate and oxalate". Journal of Analytical and Applied Pyrolysis. 29 (1): 89–102. doi:10.1016/0165-2370(93)00782-i.
  9. Muraishi, Kazuo; Yokobayashi, Hiroko; Nagase, Kenzo (1991). "Systematics on the thermal reactions of lanthanide malonates Ln2(C3H2O4)3*nH2O in the solid state". Thermochimica Acta. 182 (2): 209–217. doi:10.1016/0040-6031(91)80006-5.
  10. Osten HJ, Liu JP, P Gaworzewski, E Bugiel, Zaumseil P: IEDM Technical Digest 653. 2000.
  11. 1 2 3 GAFFNEY, A (1988). "Oxidative coupling of methane over sodium promoted praseodymium oxide". Journal of Catalysis. 114 (2): 422–432. doi:10.1016/0021-9517(88)90045-0.
  12. 1 2 Poirier, Michel G.; Breault, Raymond; Kaliaguine, Serge; Adnot, Alain (1991). "Oxidative coupling of methane over praseodymium oxide catalysts". Applied Catalysis. 71 (1): 103–122. doi:10.1016/0166-9834(91)85009-k.
  13. Huang, P. X.; Wu, F.; Zhu, B. L.; Li, G. R.; Wang, Y. L.; Gao, X. P.; Zhu, H. Y.; Yan, T. Y.; Huang, W. P. (2006-02-01). "Praseodymium Hydroxide and Oxide Nanorods and Au/Pr6O11 Nanorod Catalysts for CO Oxidation". The Journal of Physical Chemistry B. 110 (4): 1614–1620. doi:10.1021/jp055622r. ISSN   1520-6106. PMID   16471724.
  14. Kim, Il Hee; Seo, Hyun Ook; Park, Eun Ji; Han, Sang Wook; Kim, Young Dok (2017-01-16). "Low Temperature CO oxidation over Iron Oxide Nanoparticles Decorating Internal Structures of a Mesoporous Alumina". Scientific Reports. 7: 40497. Bibcode:2017NatSR...740497K. doi:10.1038/srep40497. ISSN   2045-2322. PMC   5238452 . PMID   28091561.
  15. Kar, J. K; Stevens, R.; Bowen, C. R (2005). "Processing and characterisation of Pr–zircon pigment powder". Advances in Applied Ceramics. 104 (5): 233–238. Bibcode:2005AdApC.104..233K. doi:10.1179/174367605X16699. S2CID   55057492.