Lanthanide trifluoromethanesulfonates

Last updated

Lanthanide triflates are triflate salts of the lanthanides. These salts have been investigated for application in organic synthesis as Lewis acid catalysts. These catalysts function similarly to aluminium chloride or ferric chloride, but they are water-tolerant (stable in water). Commonly written as Ln(OTf)3·(H2O)9 the nine waters are bound to the lanthanide, and the triflates are counteranions, so more accurately lanthanide triflate nonahydrate is written as [Ln(H2O)9](OTf)3. [1]

Contents

Synthesis

Lanthanide triflates are synthesized from lanthanide oxide and aqueous triflic acid: [2]

Ln2O3 + 6HOTf + 18H2O → 2[Ln(H2O)9](OTf)3 + 3H2O

Anhydrous lanthanide triflates can be produced by dehydrating their hydrated counterparts by heating between 180 and 200 °C under reduced pressure:[ citation needed ]

[Ln(H2O)9](OTf)3 → Ln(OTf)3 + 9H2O

Example reactions

Friedel-Crafts reactions

Lanthanide triflates are proposed for Friedel-Crafts acylations and alkylations, which are often carried out with AlCl3 as the catalyst in an organic solvent. The nature of the Friedel-Craft reaction, especially the acylation, forces the AlCl3 to irreversibly complex with any oxygen-containing group in the product, with the only way of decomplexing it being to destroy the AlCl3 part with water altogether. An estimated 0.9 kg of AlCl3 is wasted per kilogram of typical product- it is hydrolysed into Al2O3 and the extremely corrosive HCl. [3]

In contrast, lanthanide triflates' complexes with the product are easily separated by water, and the lanthanide triflate hydrate thus formed can be simply heated to boil the water away (This does not work for aluminium chloride due to loss of HCl; same goes for the lanthanide chlorides, hence the necessity of the triflate counterion). This avoids the need to use organic solvents- one can just use water as the solvent.

Ln(OTf)3 catalysts are used for esterifications. [4]

Other C-C bond-forming reactions

La(OTf)3 catalysts have been used for Diels-Alder, aldol, and allylation reactions. [5] Some reactions require a mixed solvent, such as aqueous formaldehyde, although Kobayashi et al. have developed alternative surfactant-water systems. [6]

Michael additions are another very important industrial method for creating new carbon-carbon bonds, often with particular functional groups attached. Addition reactions are inherently atom efficient, so are preferred synthesis pathways. La(OTf)3 catalysts not only enable these reactions to be carried out in water, but can also achieve asymmetric catalysis, yielding a desired enantio-specific or diastereo-specific product. [5]

C-N bond-forming reactions

Lanthanide triflates can be used to synthesize pyridine by catalysing either the condensation of aldehydes and amines, or the aza Diels-Alder reaction catalytic synthesis. Again, water can be used as a solvent, and high yields can be achieved under mild conditions. [7]

Nitro compounds are common in pharmaceuticals, explosives, dyes, and plastics. As for carbon compounds, catalysed Michael additions and aldol reactions can be used. For aromatic nitro compounds, synthesis is via a substitution reaction. The standard synthesis is carried out in a solution of nitric acid, mixed with excess sulfuric acid to create nitronium ions. These are then substituted on to the aromatic species. Often, the para-isomer is the desired product, but standard systems have poor selectivity. As for acylation, the reaction is normally quenched with water, and creates copious acidic waste. Using a La(OTf)3 catalyst in place of sulfuric acid reduces this waste considerably. Clark et al. report 90% conversion using just 1 mol% of ytterbium triflate in weak nitric acid, generating only a small volume of acidic waste. [3]

Green catalysts

Lanthanide triflates are stable in water, so avoid the need for organic solvents, and can be recovered for reuse. [8] on their catalytic effect in water, the range of researched applications for La(OTf)3 catalysts has exploded. [6] [9] [10]

See also

Related Research Articles

The Friedel–Crafts reactions are a set of reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring. Friedel–Crafts reactions are of two main types: alkylation reactions and acylation reactions. Both proceed by electrophilic aromatic substitution.

The aldol reaction is a reaction that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound.

<span class="mw-page-title-main">Aldol condensation</span> Type of chemical reaction

An aldol condensation is a condensation reaction in organic chemistry in which two carbonyl moieties react to form a β-hydroxyaldehyde or β-hydroxyketone, and this is then followed by dehydration to give a conjugated enone.

<span class="mw-page-title-main">Fischer–Speier esterification</span>

Fischer esterification or Fischer–Speier esterification is a special type of esterification by refluxing a carboxylic acid and an alcohol in the presence of an acid catalyst. The reaction was first described by Emil Fischer and Arthur Speier in 1895. Most carboxylic acids are suitable for the reaction, but the alcohol should generally be primary or secondary. Tertiary alcohols are prone to elimination. Contrary to common misconception found in organic chemistry textbooks, phenols can also be esterified to give good to near quantitative yield of products. Commonly used catalysts for a Fischer esterification include sulfuric acid, p-toluenesulfonic acid, and Lewis acids such as scandium(III) triflate. For more valuable or sensitive substrates other, milder procedures such as Steglich esterification are used. The reaction is often carried out without a solvent or in a non-polar solvent to facilitate the Dean-Stark method. Typical reaction times vary from 1–10 hours at temperatures of 60-110 °C.

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

Cerium(III) chloride (CeCl3), also known as cerous chloride or cerium trichloride, is a compound of cerium and chlorine. It is a white hygroscopic salt; it rapidly absorbs water on exposure to moist air to form a hydrate, which appears to be of variable composition, though the heptahydrate CeCl3·7H2O is known. It is highly soluble in water, and (when anhydrous) it is soluble in ethanol and acetone.

<span class="mw-page-title-main">Triflate</span> Chemical group (–OSO2CF3) or anion (charge –1)

In organic chemistry, triflate, is a functional group with the formula R−OSO2CF3 and structure R−O−S(=O)2−CF3. The triflate group is often represented by −OTf, as opposed to −Tf, which is the triflyl group, R−SO2CF3. For example, n-butyl triflate can be written as CH3CH2CH2CH2OTf.

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

Aluminium chloride, also known as aluminium trichloride, is an inorganic compound with the formula AlCl3. It forms a hexahydrate with the formula [Al(H2O)6]Cl3, containing six water molecules of hydration. Both the anhydrous form and the hexahydrate are colourless crystals, but samples are often contaminated with iron(III) chloride, giving them a yellow colour.

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

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">Henry reaction</span>

The Henry reaction is a classic carbon–carbon bond formation reaction in organic chemistry. Discovered in 1895 by the Belgian chemist Louis Henry (1834–1913), it is the combination of a nitroalkane and an aldehyde or ketone in the presence of a base to form β-nitro alcohols. This type of reaction is also referred to as a nitroaldol reaction. It is nearly analogous to the aldol reaction that had been discovered 23 years prior that couples two carbonyl compounds to form β-hydroxy carbonyl compounds known as "aldols". The Henry reaction is a useful technique in the area of organic chemistry due to the synthetic utility of its corresponding products, as they can be easily converted to other useful synthetic intermediates. These conversions include subsequent dehydration to yield nitroalkenes, oxidation of the secondary alcohol to yield α-nitro ketones, or reduction of the nitro group to yield β-amino alcohols.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

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

Scandium trifluoromethanesulfonate, commonly called scandium triflate, is a chemical compound with formula Sc(SO3CF3)3, a salt consisting of scandium cations Sc3+ and triflate SO
3
CF
3
anions.

<span class="mw-page-title-main">Grignard reagent</span> Organometallic compounds used in organic synthesis

A Grignard reagent or Grignard compound is a chemical compound with the general formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

Silyl enol ethers in organic chemistry are a class of organic compounds that share a common functional group composed of an enolate bonded through its oxygen end to an organosilicon group. They are important intermediates in organic synthesis.

<span class="mw-page-title-main">Mukaiyama aldol addition</span>

The Mukaiyama aldol addition is an organic reaction and a type of aldol reaction between a silyl enol ether and an aldehyde or formate. The reaction was discovered by Teruaki Mukaiyama (1927–2018) in 1973. His choice of reactants allows for a crossed aldol reaction between an aldehyde and a ketone or a different aldehyde without self-condensation of the aldehyde. For this reason the reaction is used extensively in organic synthesis.

<span class="mw-page-title-main">Mukaiyama Taxol total synthesis</span>

The Mukaiyama taxol total synthesis published by the group of Teruaki Mukaiyama of the Tokyo University of Science between 1997 and 1999 was the 6th successful taxol total synthesis. The total synthesis of Taxol is considered a hallmark in organic synthesis.

In organic chemistry, the Baylis–Hillman, Morita–Baylis–Hillman, or MBH reaction is a carbon-carbon bond-forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst, such as a tertiary amine or phosphine. The product is densely functionalized, joining the alkene at the α-position to a reduced form of the electrophile.

In Lewis acid catalysis of organic reactions, a metal-based Lewis acid acts as an electron pair acceptor to increase the reactivity of a substrate. Common Lewis acid catalysts are based on main group metals such as aluminum, boron, silicon, and tin, as well as many early and late d-block metals. The metal atom forms an adduct with a lone-pair bearing electronegative atom in the substrate, such as oxygen, nitrogen, sulfur, and halogens. The complexation has partial charge-transfer character and makes the lone-pair donor effectively more electronegative, activating the substrate toward nucleophilic attack, heterolytic bond cleavage, or cycloaddition with 1,3-dienes and 1,3-dipoles.

<span class="mw-page-title-main">1-Tetralone</span> Chemical compound

1-Tetralone is a bicyclic aromatic hydrocarbon and a ketone. In terms of its structure, it can also be regarded as benzo-fused cyclohexanone. It is a colorless oil with a faint odor. It is used as starting material for agricultural and pharmaceutical agents. The carbon skeleton of 1-tetralone is found in natural products such as Aristelegone A (4,7-dimethyl-6-methoxy-1-tetralone) from the family of Aristolochiaceae used in traditional Chinese medicine.

<span class="mw-page-title-main">Teruaki Mukaiyama</span> Japanese chemist (1927–2018)

Teruaki Mukaiyama was a Japanese organic chemist. One of the most prolific chemists of the 20th century in the field of organic synthesis, Mukaiyama helped establish the field of organic chemistry in Japan after World War II.

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

Hafnium(IV) triflate or hafnium trifluoromethansulfonate is an inorganic substance with the idealized formula Hf(OSO2CF3)4, also written as Hf(OTf)4. Hafnium triflate is used as an impure mixture as a catalyst. Hafnium (IV) has an ionic radius of intermediate range (Al < Ti < Hf < Zr < Sc < Ln) and has an oxophilic hard character typical of group IV metals. This solid is a stronger Lewis acid than its typical precursor hafnium tetrachloride, HfCl4, because of the strong electron-withdrawing nature of the four triflate groups, which makes it a great Lewis acid and has many uses including as a great catalyst at low Lewis acid loadings for electrophilic aromatic substitution and nucleophilic substitution reactions.

References

  1. Harrowfield, J. M.; Keppert, D. L.; Patrick, J. M.; White, A. H. (1983). "Structure and stereochemistry in "f-block" complexes of high coordination number. VIII. The [M(unidentate)9] system. Crystal structures of [M(OH2)9] [CF3SO3]3, M = lanthanum, gadolinium, lutetium, or yttrium". Australian Journal of Chemistry . 36 (3): 483–492. doi:10.1071/CH9830483.
  2. Kobayashi, S.; Hachiya, I. (1994). "Lanthanide Triflates as Water-Tolerant Lewis Acids. Activation of Commercial Formaldehyde Solution and Use in the Aldol Reaction of Silyl Enol Ethers with Aldehydes in Aqueous Media". J. Org. Chem. 59 (13): 3590–6. doi:10.1021/jo00092a017.
  3. 1 2 Clark, J.; Macquarie, D. (2002). Handbook of Green Chemistry & Technology. Oxford, UK: Blackwell Science. ISBN   978-0-632-05715-3.
  4. Barrett, A.; Braddock, D. (1997). "Scandium(III) or Lanthanide(III) Triflates as Recyclable Catalysts for the Direct Acetylation of Alcohols with Acetic Acid". Chem. Commun. 1997 (4): 351–352. doi:10.1039/a606484a.
  5. 1 2 Engberts, J., Feringa, B., Keller, E. & Otto, S. 1996, “Lewis-acid Catalysis of Carbon Carbon Bond Forming Reactions in Water”, Recuil des Travaux Chimiques des Pays-Bas 115(11-12), 457-464
  6. 1 2 Kobayashi, S.; Manabe, K. (2000). "Green Lewis Acid Catalysts in Organic Synthesis". Pure Appl. Chem. 72 (7): 1373–1380. doi:10.1351/pac200072071373. S2CID   16770637.
  7. Wenhua Xie; Yafei Jin; Peng George Wang (1999). "Lanthanide triflates as unique Lewis acids". Chemtech . 29 (2): 23–29.
  8. Kobayashi, S.; Uchiro, H.; Fujishita, Y.; Shiina, I.; Mukaiyama, T. (1991). "Asymmetric Aldol Reaction between Achiral Silyl Enol Ethers and Achiral Aldehydes by use of a Chiral Promoter System". J. Am. Chem. Soc. 113 (11): 4247–4252. doi:10.1021/ja00011a030.
  9. Shu Kobayashi† and Kei Manabe. Green Lewis acid catalysis in organic synthesis. Pure Appl. Chem., Vol. 72, No. 7, pp. 1373–1380, 2000.
  10. Waller, F.J.; Barrett. A.G.M.; Braddock, D.C.; Ramprasad, D. "Lanthanide (III) Triflates as Recyclable Catalysts for Atom Economic Aromatic Nitration." Chem. Commun. 1997, 613-614.