Ruthenium tetroxide

Last updated
Ruthenium(VIII) oxide
Ruthenium tetroxide.svg
Ruthenium-tetroxide-3D-balls.png
Names
IUPAC name
Ruthenium(VIII) oxide
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.039.815 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 243-813-8
PubChem CID
UNII
  • InChI=1S/4O.Ru
    Key: GJFMDWMEOCWXGJ-UHFFFAOYSA-N
  • O=[Ru](=O)(=O)=O
Properties
RuO4
Molar mass 165.07 g/mol
Appearanceyellow easily melting solid
Odor pungent
Density 3.29 g/cm3
Boiling point 129.6 [1]  °C (265.3 °F; 402.8 K)
2% w/v at 20 °C
Solubility in other solventsSoluble in
Carbon tetrachloride
Chloroform
Structure
tetrahedral
zero
Hazards
NFPA 704 (fire diamond)
NFPA 704.svgHealth 3: Short exposure could cause serious temporary or residual injury. E.g. chlorine gasFlammability 0: Will not burn. E.g. waterInstability 1: Normally stable, but can become unstable at elevated temperatures and pressures. E.g. calciumSpecial hazards (white): no code
3
0
1
Safety data sheet (SDS) external MSDS sheet
Related compounds
Related compounds
Ruthenium dioxide
Ruthenium trichloride
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Yes check.svgY  verify  (what is  Yes check.svgYX mark.svgN ?)

Ruthenium tetroxide is the inorganic compound with the formula RuO4. It is a yellow volatile solid that melts near room temperature. [2] It has the odor of ozone. [3] Samples are typically black due to impurities. The analogous OsO4 is more widely used and better known. It is also the anhydride of hyperruthenic acid (H2RuO5). One of the few solvents in which RuO4 forms stable solutions is CCl4. [4]

Contents

Preparation

RuO4 is prepared by oxidation of ruthenium(III) chloride with NaIO4. [2] The reaction initially produces sodium diperiodo­dihydroxo­ruthenate(VI), which then decomposes in acid solution to the tetroxide: [5]

8 Ru3+(aq) + 5 IO4(aq) + 12 H2O(l) → 8 RuO4(s) + 5 I(aq) + 24 H+(aq) [6]

Due to its challenging reactivity, RuO4 it is always generated in situ and used in catalytic quantities, at least in organic reactions. [4]

Structure

RuO4 forms two crystal structures, one with cubic symmetry and another with monoclinic symmetry, isotypic to OsO4. The molecule adopts a tetrahedral geometry, with the Ru–O distances ranging from 169 to 170 pm. [7]

Uses

Isolation of ruthenium from ores

The main commercial value of RuO4 is as an intermediate in the production of ruthenium compounds and metal from ores. Like other platinum group metals (PGMs), ruthenium occurs at low concentrations and often mixed with other PGMs. Together with OsO4, it is separated from other PGMs by distillation of a chlorine-oxidized extract. Ruthenium is separated from OsO4 by reducing RuO4 with hydrochloric acid, a process that exploits the highly positive reduction potential for the [RuO4]0/- couple. [8] [9]

Organic chemistry

RuO4 is of specialized value in organic chemistry because it oxidizes virtually any hydrocarbon. For example, it will oxidize adamantane to 1-adamantanol. Because it is such an aggressive oxidant, reaction conditions must be mild, generally room temperature. Although a strong oxidant, RuO4 oxidations do not perturb stereocenters that are not oxidized. Illustrative is the oxidation of the following diol to a carboxylic acid:

RuO4oxidation.png

Oxidation of epoxy alcohols also occurs without degradation of the epoxide ring:

RuO4epoxy.png

Under milder conditions, oxidative reaction yields aldehydes instead. RuO4 readily converts secondary alcohols into ketones. Although similar results can be achieved with other cheaper oxidants such as PCC- or DMSO-based oxidants, RuO4 is ideal when a very vigorous oxidant is needed, but mild conditions must be maintained. It is used in organic synthesis to oxidize internal alkynes to 1,2-diketones, and terminal alkynes along with primary alcohols to carboxylic acids. When used in this fashion, the ruthenium(VIII) oxide is used in catalytic amounts and regenerated by the addition of sodium periodate to ruthenium(III) chloride and a solvent mixture of acetonitrile, water and carbon tetrachloride. RuO4 readily cleaves double bonds to yield carbonyl products, in a manner similar to ozonolysis. OsO4, a more familiar oxidant that is structurally similar to RuO4, does not cleave double bonds, instead producing vicinal diol products. However, with short reaction times and carefully controlled conditions, RuO4 can also be used for dihydroxylation. [10]

Because RuO4 degrades the "double bonds" of arenes (especially electron-rich ones) by dihydroxylation and cleavage of the C-C bond in a way few other reagents can, it is useful as a "deprotection" reagent for carboxylic acids that are masked as aryl groups (typically phenyl or p-methoxyphenyl). Because the fragments formed are themselves readily oxidizable by RuO4, a substantial fraction of the arene carbon atoms undergo exhaustive oxidation to form carbon dioxide. Consequently, multiple equivalents of the terminal oxidant (often in excess of 10 equivalents per aryl ring) are required to achieve complete conversion to the carboxylic acid, limiting the practicality of the transformation. [11] [12] [13]

RuO4-degradation-rev.png

Although used as a direct oxidant, due to the relatively high cost of RuO4 it is also used catalytically with a cooxidant. For an oxidation of cyclic alcohols with RuO4 as a catalyst and bromate as oxidant under basic conditions, RuO4 is first activated by hydroxide, turning into the hyperruthenate anion:

RuO4 + OH → HRuO5

The reaction proceeds via a glycolate complex.

Other uses

Ruthenium tetroxide is a potential staining agent. It is used to expose latent fingerprints by turning to the brown/black ruthenium dioxide when in contact with fatty oils or fats contained in sebaceous contaminants of the print. [14]

Gaseous release by nuclear accidents

Because of the very high volatility of ruthenium tetroxide (RuO
4
) ruthenium radioactive isotopes with their relative short half-life are considered as the second most hazardous gaseous isotopes after iodine-131 in case of release by a nuclear accident. [15] [3] [16] The two most important radioactive isotopes of ruthenium are 103Ru and 106Ru. They have half-lives of 39.6 days and 373.6 days, respectively. [3]

Related Research Articles

<span class="mw-page-title-main">Osmium</span> Chemical element, symbol Os and atomic number 76

Osmium is a chemical element; it has symbol Os and atomic number 76. It is a hard, brittle, bluish-white transition metal in the platinum group that is found as a trace element in alloys, mostly in platinum ores. Osmium is the densest naturally occurring element. When experimentally measured using X-ray crystallography, it has a density of 22.59 g/cm3. Manufacturers use its alloys with platinum, iridium, and other platinum-group metals to make fountain pen nib tipping, electrical contacts, and in other applications that require extreme durability and hardness.

<span class="mw-page-title-main">Ruthenium</span> Chemical element, symbol Ru and atomic number 44

Ruthenium is a chemical element; it has symbol Ru and atomic number 44. It is a rare transition metal belonging to the platinum group of the periodic table. Like the other metals of the platinum group, ruthenium is inert to most other chemicals. Karl Ernst Claus, a Russian-born scientist of Baltic-German ancestry, discovered the element in 1844 at Kazan State University and named ruthenium in honor of Russia. Ruthenium is usually found as a minor component of platinum ores; the annual production has risen from about 19 tonnes in 2009 to some 35.5 tonnes in 2017. Most ruthenium produced is used in wear-resistant electrical contacts and thick-film resistors. A minor application for ruthenium is in platinum alloys and as a chemistry catalyst. A new application of ruthenium is as the capping layer for extreme ultraviolet photomasks. Ruthenium is generally found in ores with the other platinum group metals in the Ural Mountains and in North and South America. Small but commercially important quantities are also found in pentlandite extracted from Sudbury, Ontario, and in pyroxenite deposits in South Africa.

<span class="mw-page-title-main">Osmium tetroxide</span> Chemical compound

Osmium tetroxide (also osmium(VIII) oxide) is the chemical compound with the formula OsO4. The compound is noteworthy for its many uses, despite its toxicity and the rarity of osmium. It also has a number of unusual properties, one being that the solid is volatile. The compound is colourless, but most samples appear yellow. This is most likely due to the presence of the impurity OsO2, which is yellow-brown in colour. In biology, its property of binding to lipids has made it a widely-used stain in electron microscopy.

Sharpless asymmetric dihydroxylation is the chemical reaction of an alkene with osmium tetroxide in the presence of a chiral quinine ligand to form a vicinal diol. The reaction has been applied to alkenes of virtually every substitution, often high enantioselectivities are realized, with the chiral outcome controlled by the choice of dihydroquinidine (DHQD) vs dihydroquinine (DHQ) as the ligand. Asymmetric dihydroxylation reactions are also highly site selective, providing products derived from reaction of the most electron-rich double bond in the substrate.

Chromic acid is jargon for a solution formed by the addition of sulfuric acid to aqueous solutions of dichromate. It consists at least in part of chromium trioxide.

<span class="mw-page-title-main">Benzyl group</span> Chemical group (–CH₂–C₆H₅)

In organic chemistry, benzyl is the substituent or molecular fragment possessing the structure R−CH2−C6H5. Benzyl features a benzene ring attached to a methylene group group.

<span class="mw-page-title-main">Permanganate</span> Chemical compound

A permanganate is a chemical compound with the manganate(VII) ion, MnO
4
, the conjugate base of permanganic acid. Because the manganese atom has a +7 oxidation state, the permanganate(VII) ion is a strong oxidising agent. The ion is a transition metal ion with a tetrahedral structure. Permanganate solutions are purple in colour and are stable in neutral or slightly alkaline media. The exact chemical reaction depends on the carbon-containing reactants present and the oxidant used. For example, trichloroethane (C2H3Cl3) is oxidised by permanganate ions to form carbon dioxide (CO2), manganese dioxide (MnO2), hydrogen ions (H+), and chloride ions (Cl).

<span class="mw-page-title-main">Wacker process</span> Chemical reaction

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride and copper(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

<span class="mw-page-title-main">Periodate</span> Negatively-charged molecule made of oxygen and iodine

Periodate is an anion composed of iodine and oxygen. It is one of a number of oxyanions of iodine and is the highest in the series, with iodine existing in oxidation state +7. Unlike other perhalogenates, such as perchlorate, it can exist in two forms: metaperiodateIO
4
and orthoperiodateIO5−
6
. In this regard it is comparable to the tellurate ion from the adjacent group. It can combine with a number of counter ions to form periodates, which may also be regarded as the salts of periodic acid.

<span class="mw-page-title-main">Tetrapropylammonium perruthenate</span> Chemical compound

Tetrapropylammonium perruthenate (TPAP or TPAPR) is the chemical compound described by the formula N(C3H7)4RuO4. Sometimes known as the Ley–Griffith reagent, this ruthenium compound is used as a reagent in organic synthesis. This salt consists of the tetrapropylammonium cation and the perruthenate anion, RuO−4.

Dihydroxylation is the process by which an alkene is converted into a vicinal diol. Although there are many routes to accomplish this oxidation, the most common and direct processes use a high-oxidation-state transition metal. The metal is often used as a catalyst, with some other stoichiometric oxidant present. In addition, other transition metals and non-transition metal methods have been developed and used to catalyze the reaction.

Glycol cleavage is a specific type of organic chemistry oxidation. The carbon–carbon bond in a vicinal diol (glycol) is cleaved and instead the two oxygen atoms become double-bonded to their respective carbon atoms. Depending on the substitution pattern in the diol, these carbonyls will be ketones and/or aldehydes.

Osmium compounds are compounds containing the element osmium (Os). Osmium forms compounds with oxidation states ranging from −2 to +8. The most common oxidation states are +2, +3, +4, and +8. The +8 oxidation state is notable for being the highest attained by any chemical element aside from iridium's +9 and is encountered only in xenon, ruthenium, hassium, iridium, and plutonium. The oxidation states −1 and −2 represented by the two reactive compounds Na
2
[Os
4
(CO)
13
]
and Na
2
[Os(CO)
4
]
are used in the synthesis of osmium cluster compounds.

Ruthenium compounds are compounds containing the element ruthenium (Ru). Ruthenium compounds can have oxidation states ranging from 0 to +8, and −2. The properties of ruthenium and osmium compounds are often similar. The +2, +3, and +4 states are the most common. The most prevalent precursor is ruthenium trichloride, a red solid that is poorly defined chemically but versatile synthetically.

The Milas hydroxylation is an organic reaction converting an alkene to a vicinal diol, and was developed by Nicholas A. Milas in the 1930s. The cis-diol is formed by reaction of alkenes with hydrogen peroxide and either ultraviolet light or a catalytic osmium tetroxide, vanadium pentoxide, or chromium trioxide.

Alcohol oxidation is a collection of oxidation reactions in organic chemistry that convert alcohols to aldehydes, ketones, carboxylic acids, and esters where the carbon carries a higher oxidation state. The reaction mainly applies to primary and secondary alcohols. Secondary alcohols form ketones, while primary alcohols form aldehydes or carboxylic acids.

<span class="mw-page-title-main">Dichlorotris(triphenylphosphine)ruthenium(II)</span> Chemical compound

Dichlorotris(triphenylphosphine)ruthenium(II) is a coordination complex of ruthenium. It is a chocolate brown solid that is soluble in organic solvents such as benzene. The compound is used as a precursor to other complexes including those used in homogeneous catalysis.

<span class="mw-page-title-main">Oxoammonium-catalyzed oxidation</span>

Oxoammonium-catalyzed oxidation reactions involve the conversion of organic substrates to more highly oxidized materials through the action of an N-oxoammonium species. Nitroxides may also be used in catalytic amounts in the presence of a stoichiometric amount of a terminal oxidant. Nitroxide radical species used are either 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) or derivatives thereof.

<span class="mw-page-title-main">Jones oxidation</span> Oxidation of alcohol

The Jones oxidation is an organic reaction for the oxidation of primary and secondary alcohols to carboxylic acids and ketones, respectively. It is named after its discoverer, Sir Ewart Jones. The reaction was an early method for the oxidation of alcohols. Its use has subsided because milder, more selective reagents have been developed, e.g. Collins reagent.

<span class="mw-page-title-main">Barium manganate</span> Chemical compound

Barium manganate is an inorganic compound with the formula BaMnO4. It is used as an oxidant in organic chemistry. It belongs to a class of compounds known as manganates in which the manganese resides in a +6 oxidation state. Manganate should not be confused with permanganate which contains manganese(VII). Barium manganate is a powerful oxidant, popular in organic synthesis and can be used in a wide variety of oxidation reactions.

References

  1. Koda, Yoshio (1986). "Boiling Points and Ideal Solutions of Ruthenium and Osmium Tetraoxides". Journal of the Chemical Society, Chemical Communications. 1986 (17): 1347–1348. doi:10.1039/C39860001347.
  2. 1 2 H. L. Grube (1963). "Ruthenium (VIII) Oxide". In G. Brauer (ed.). Handbook of Preparative Inorganic Chemistry, 2nd Ed. Vol. 1. NY: Academic Press. pp. 1599–1600.
  3. 1 2 3 Backman, U., Lipponen, M., Auvinen, A., Jokiniemi, J., & Zilliacus, R. (2004). Ruthenium behaviour in severe nuclear accident conditions. Final report (No. NKS–100). Nordisk Kernesikkerhedsforskning.
  4. 1 2 Martín, V. S.; Palazón, J. M.; Rodríguez, C. M.; Nevill, C. R. (2006). "Ruthenium(VIII) Oxide". Encyclopedia of Reagents for Organic Synthesis. doi:10.1002/047084289X.rr009.pub2. ISBN   978-0471936237.
  5. Mercer, E. E.; Meyer, S. M. (1972) [1 July 1971]. "A periodate complex of ruthenium(VI)". J. Inorg. Nucl. Chem. 34 (2). Great Britain: Pergamon: 777–778. doi:10.1016/0022-1902(72)80466-4.
  6. Carlsen, Per H. J.; Katsuki, Tsutomu; Martin, Victor S.; Sharpless, K. Barry (September 1981). "A greatly improved procedure for ruthenium tetroxide catalyzed oxidations of organic compounds". The Journal of Organic Chemistry. 46 (19): 3936–3938. doi:10.1021/jo00332a045. ISSN   0022-3263.
  7. Pley, M.; Wickleder, M. S. (2005). "Two Crystalline Modifications of RuO4". Journal of Solid State Chemistry. 178 (10): 3206–3209. Bibcode:2005JSSCh.178.3206P. doi:10.1016/j.jssc.2005.07.021.
  8. Bernardis, Francesco L.; Grant, Richard A.; Sherrington, David C. (2005). "A review of methods of separation of the platinum-group metals through their chloro-complexes". Reactive and Functional Polymers. 65 (3): 205–217. doi:10.1016/j.reactfunctpolym.2005.05.011.
  9. Swain, P.; Mallika, C.; Srinivasan, R.; Mudali, U. K.; Natarajan, R. (2013). "Separation and recovery of ruthenium: a review". Journal of Radioanalytical and Nuclear Chemistry. 298 (2): 781–796. doi:10.1007/s10967-013-2536-5. S2CID   95804621.
  10. Plietker, Bernd (2005). "Selectivity versus reactivity - recent advances in RuO4-catalyzed oxidations". Synthesis. 5 (15): 2453–2472. doi:10.1055/s-2005-872172.
  11. Nunez, M. Teresa; Martin, Victor S. (March 1990). "Efficient oxidation of phenyl groups to carboxylic acids with ruthenium tetraoxide. A simple synthesis of (R)-.gamma.-caprolactone, the pheromone of Trogoderma granarium". The Journal of Organic Chemistry. 55 (6): 1928–1932. doi:10.1021/jo00293a044. ISSN   0022-3263.
  12. Nasr, Khaled; Pannier, Nadine; Frangioni, John V.; Maison, Wolfgang (February 2008). "Rigid Multivalent Scaffolds Based on Adamantane". The Journal of Organic Chemistry. 73 (3): 1056–1060. doi:10.1021/jo702310g. ISSN   0022-3263. PMC   2505186 . PMID   18179237.
  13. Mander, Lewis N.; Williams, Craig M. (2003-02-17). "Oxidative degradation of benzene rings". Tetrahedron. 59 (8): 1105–1136. doi:10.1016/S0040-4020(02)01492-8. ISSN   0040-4020.
  14. Mashiko, K.; Miyamoto, T. (1998). "Latent Fingerprint Processing by the Ruthenium Tetroxide Method". Journal of Forensic Identification. 48 (3): 279–290. doi: 10.3408/jasti.2.21 .
  15. Ronneau, C.; Cara, J.; Rimski-Korsakov, A. (1995). "Oxidation-enhanced emission of ruthenium from nuclear fuel". Journal of Environmental Radioactivity. 26: 63–70. doi:10.1016/0265-931X(95)91633-F.
  16. Beuzet, Emilie; Lamy, Jean-Sylvestre; Perron, Hadrien; Simoni, Eric; Ducros, Gérard (2012). "Ruthenium release modelling in air and steam atmospheres under severe accident conditions using the MAAP4 code". Nuclear Engineering and Design. 246: 157–162. doi:10.1016/j.nucengdes.2011.08.025.

Further reading