Hydroformylation

Last updated

Contents

Hydroformylation
Reaction type Addition reaction
Identifiers
RSC ontology ID RXNO:0000272
Hydroformylation of an alkene (R to R organyl groups (i. e. alkyl- or aryl group) or hydrogen) Hydroformylation V.2 en.png
Hydroformylation of an alkene (R to R organyl groups (i. e. alkyl- or aryl group) or hydrogen)

In organic chemistry, hydroformylation, also known as oxo synthesis or oxo process, is an industrial process for the production of aldehydes (R−CH=O) from alkenes (R2C=CR2). [1] [2] This chemical reaction entails the net addition of a formyl group (−CHO) and a hydrogen atom to a carbon-carbon double bond. This process has undergone continuous growth since its invention: production capacity reached 6.6×106 tons in 1995. It is important because aldehydes are easily converted into many secondary products. For example, the resultant aldehydes are hydrogenated to alcohols that are converted to detergents. Hydroformylation is also used in speciality chemicals, relevant to the organic synthesis of fragrances and pharmaceuticals. The development of hydroformylation is one of the premier achievements of 20th-century industrial chemistry.

The process entails treatment of an alkene typically with high pressures (between 10 and 100 atmospheres) of carbon monoxide and hydrogen at temperatures between 40 and 200 °C. [3] In one variation, formaldehyde is used in place of synthesis gas. [4] Transition metal catalysts are required. Invariably, the catalyst dissolves in the reaction medium, i.e. hydroformylation is an example of homogeneous catalysis.

History

The process was discovered by the German chemist Otto Roelen in 1938 in the course of investigations of the Fischer–Tropsch process. Aldehydes and diethylketone were obtained when ethylene was added to an F-T reactor. Through these studies, Roelen discovered the utility of cobalt catalysts. HCo(CO)4, which had been isolated only a few years prior to Roelen's work, was shown to be an excellent catalyst. [5] [6] The term oxo synthesis was coined by the Ruhrchemie patent department, who expected the process to be applicable to the preparation of both aldehydes and ketones. Subsequent work demonstrated that the ligand tributylphosphine (PBu3) improved the selectivity of the cobalt-catalysed process. The mechanism of Co-catalyzed hydroformylation was elucidated by Richard F. Heck and David Breslow in the 1960s. [7]

A generic rhodium catalyst, where PAr3 = triphenylphosphine or its sulfonated analogue Tppts. See tris(triphenylphosphine)rhodium carbonyl hydride. HRhP3CO.png
A generic rhodium catalyst, where PAr3 = triphenylphosphine or its sulfonated analogue Tppts. See tris(triphenylphosphine)rhodium carbonyl hydride.

In 1968, highly active rhodium-based catalysts were reported. [8] Since the 1970s, most hydroformylation relies on catalysts based on rhodium. [9] Water-soluble catalysts have been developed. They facilitate the separation of the products from the catalyst. [10]

Mechanism

Mechanism of cobalt-catalyzed hydroformylation. The process begins with dissociation of CO from cobalt tetracarbonyl hydride to give the 16-electron species (step 1). Subsequent binding of alkene gives an 18e species (step 2). In step 3, the olefin inserts to give the 16e alkyl tricarbonyl. Coordination of another equivalent of CO give alkyl tetracarbonyl (step 4). Migratory insertion of CO gives the 16e acyl in step 5. In step 6, oxidative addition of hydrogen gives a dihydrido complex, which in step 7 releases aldehyde by reductive elimination. Step 8 is unproductive and reversible. Hydroformylation Mechanism V.1.svg
Mechanism of cobalt-catalyzed hydroformylation. The process begins with dissociation of CO from cobalt tetracarbonyl hydride to give the 16-electron species (step 1). Subsequent binding of alkene gives an 18e species (step 2). In step 3, the olefin inserts to give the 16e alkyl tricarbonyl. Coordination of another equivalent of CO give alkyl tetracarbonyl (step 4). Migratory insertion of CO gives the 16e acyl in step 5. In step 6, oxidative addition of hydrogen gives a dihydrido complex, which in step 7 releases aldehyde by reductive elimination. Step 8 is unproductive and reversible.

Selectivity

A key consideration of hydroformylation is the "normal" vs. "iso" selectivity. For example, the hydroformylation of propylene can afford two isomeric products, butyraldehyde or isobutyraldehyde:

These isomers reflect the regiochemistry of the insertion of the alkene into the M–H bond. Since both products are not equally desirable (normal is more stable than iso), much research was dedicated to the quest for catalyst that favored the normal isomer.

Steric effects

Markovnikov's rule addition of the cobalt hydride to primary alkenes is disfavored by steric hindrance between the cobalt centre and the secondary alkyl ligand. Bulky ligands exacerbate this steric hindrance. Hence, the mixed carbonyl/phosphine complexes offer a greater selectivity for anti-Markovnikov addition, thus favoring straight chain products (n-) aldehydes. Modern catalysts rely increasingly on chelating ligands, especially diphosphites. [12]

n (top) vs. iso (bottom)-selectivity. Hydroformylation steric effect Markovnikov.svg
n (top) vs. iso (bottom)-selectivity.

Electronic effects

Additionally, electron-rich the hydride complex are less proton-like. Thus, as a result, the electronic effects that normally favour the Markovnikov addition to an alkene are less applicable. Thus, electron-rich hydrides are more selective.

Acyl formation

To suppress competing isomerization of the alkene, the rate of migratory insertion of the carbonyl into the carbon-metal bond of the alkyl must be relatively fast. The rate of insertion of the carbonyl carbon into the C-M bond is likely to be greater than the rate of beta-hydride elimination. [13]

Hydroformylation of oct-2-ene Hydroformulation 2-octene.png
Hydroformylation of oct-2-ene

Asymmetric hydroformylation

Hydroformylation of prochiral alkenes creates new stereocenters. Using chiral phosphine ligands, the hydroformylation can be tailored to favor one enantiomer. [14] Thus, for example, dexibuprofen, the (+)−(S)-enantiomer of ibuprofen, can be produced by enantioselective hydroformylation followed by oxidation.

Processes

The industrial processes vary depending on the chain length of the olefin to be hydroformylated, the catalyst metal and ligands, and the recovery of the catalyst. The original Ruhrchemie process produced propanal from ethene and syngas using cobalt tetracarbonyl hydride. Today, industrial processes based on cobalt catalysts are mainly used for the production of medium- to long-chain olefins, whereas the rhodium-based catalysts are usually used for the hydroformylation of propene. The rhodium catalysts are significantly more expensive than cobalt catalysts. In the hydroformylation of higher molecular weight olefins the separation of the catalyst from the produced aldehydes is difficult.

BASF-oxo process

The BASF-oxo process starts mostly with higher olefins and relies on cobalt carbonyl-based catalyst. [15] By conducting the reaction at low temperatures, one observes increased selectivity favoring the linear product. The process is carried out at a pressure of about 30 MPa and in a temperature range of 150 to 170 °C. The cobalt is recovered from the liquid product by oxidation to water-soluble Co2 +, followed by the addition of aqueous formic or acetic acids. This process gives an aqueous phase of cobalt, which can then be recycled. Losses are compensated by the addition of cobalt salts. [16]

Exxon process

The Exxon process, also Kuhlmann- or PCUK – oxo process, is used for the hydroformylation of C6–C12 olefins. The process relies on cobalt catalysts. In order to recover the catalyst, an aqueous sodium hydroxide solution or sodium carbonate is added to the organic phase. By extraction with olefin and neutralization by addition of sulfuric acid solution under carbon monoxide pressure the metal carbonyl hydride can recovered. This is stripped out with syngas, absorbed by the olefin, and returned to the reactor. Similar to the BASF process, the Exxon process is carried out at a pressure of about 30 MPa and at a temperature of about 160 to 180 °C. [16]

Shell process

The Shell process uses cobalt complexes modified with phosphine ligands for the hydroformylation of C7–C14 olefins. The resulting aldehydes are directly hydrogenated to the fatty alcohols, which are separated by distillation, which allows the catalyst to be recycled. The process has good selectivity to linear products, which find use as feedstock for detergents. The process is carried out at a pressure of about 4 to 8 MPa and at a temperature range of about 150–190 °C. [16]

Union Carbide process

The Union Carbide (UCC) process, also known as low-pressure oxo process (LPO), relies on a rhodium catalyst dissolved in high-boiling thick oil, a higher molecular weight condensation product of the primary aldehydes, for the hydroformylation of propene. The reaction mixture is separated in a falling film evaporator from volatile components. The liquid phase is distilled and butyraldehyde is removed as head product while the catalyst containing bottom product is recycled to the process. The process is carried out at about 1.8 MPa and 95–100 °C. [16]

BiPhePhos is representative diphosphite ligand popularized by workers at Union Carbide. BiPhePhos.svg
BiPhePhos is representative diphosphite ligand popularized by workers at Union Carbide.

Ruhrchemie/Rhone–Poulenc process

Process scheme of the Ruhrchemie/Rhone-Poulenc process Idroformilazione LPO.svg
Process scheme of the Ruhrchemie/Rhône–Poulenc process

The Ruhrchemie/Rhone–Poulenc process (RCRPP) relies on a rhodium catalyst with water-soluble TPPTS as ligand (Kuntz Cornils catalyst) for the hydroformylation of propene. [19] The tri-sulfonation of triphenylphosphane ligand provides hydrophilic properties to the organometallic complex. The catalyst complex carries nine sulfonate-groups and is highly soluble in water (about 1 kg L−1), but not in the emerging product phase. [20] The water-soluble TPPTS is used in about 50-fold excess, whereby the leaching of the catalyst is effectively suppressed. Reactants are propene and syngas consisting of hydrogen and carbon monoxide in a ratio of 1.1:1. A mixture of butyraldehyde and isobutyraldehyde in the ratio 96:4 is generated with few by-products such as alcohols, esters and higher boiling fractions. [20] The Ruhrchemie/Rhone-Poulenc-process is the first commercially available two-phase system in which the catalyst is present in the aqueous phase. In the progress of the reaction an organic product phase is formed which is separated continuously by means of phase separation, wherein the aqueous catalyst phase remains in the reactor. [20]

The process is carried out in a stirred tank reactor where the olefin and the syngas are entained[ clarification needed ] from the bottom of the reactor through the catalyst phase under intensive stirring. The resulting crude aldehyde phase is separated at the top from the aqueous phase. The aqueous catalyst-containing solution is re-heated via a heat exchanger and pumped back into the reactor. [20] The excess olefin and syngas is separated from the aldehyde phase in a stripper and fed back to the reactor. The generated heat is used for the generation of process steam, which is used for subsequent distillation of the organic phase to separate into butyraldehyde and isobutyraldehyde. [20] Potential catalyst poisons coming from the synthesis gas migrate into the organic phase and removed from the reaction with the aldehyde. Thus there is no accumulation of catalyst poisons, and the elaborate fine purification of the syngas can be omitted. [20]

A plant was built in Oberhausen in 1984, which was debottlenecked in 1988 and again in 1998 up to a production capacity of 500,000 t/a butanal. The conversion rate of propene is 98% and the selectivity to n-butanal is high. During the life time of a catalyst batch in the process less than 1 ppb rhodium is lost. [21]

Laboratory process

Recipes have been developed for the hydroformylation on a laboratory scale, e.g. of cyclohexene. [3] [14]

Substrates other than alkenes

Cobalt carbonyl and rhodium complexes catalyse the hydroformylation of formaldehyde and ethylene oxide to give hydroxyacetaldehyde and 3-hydroxypropanal, which can then be hydrogenated to ethylene glycol and propane-1,3-diol, respectively. The reactions work best when the solvent is basic (such as pyridine). [22] [23]

In the case of dicobalt octacarbonyl or Co2(CO)8 as a catalyst, pentan-3-one can arise from ethene and CO, in the absence of hydrogen. A proposed intermediate is the ethylene-propionyl species [CH3C(O)Co(CO)3(ethene)] which undergoes a migratory insertion to form [CH3COCH2CH2Co(CO)3]. The required hydrogen arises from the water shift reaction. For details, see [24]

If the water shift reaction is not operative, the reaction affords a polymer containing alternating carbon monoxide and ethylene units. Such aliphatic polyketones are more conventionally prepared using palladium catalysts. [25]

Functionalized olefins such as allyl alcohol can be hydroformylated. The target product 1,4-butanediol and its isomer is obtained with isomerization free catalysts such as rhodium-triphenylphosphine complexes. The use of the cobalt complex leads by isomerization of the double bond to n-propanal. [26] The hydroformylation of alkenyl ethers and alkenyl esters occurs usually in the α-position to the ether or ester function.

The hydroformylation of acrylic acid and methacrylic acid in the rhodium-catalyzed process leads to the Markovnikov product in the first step. [27] By variation of the reaction conditions the reaction can be directed to different products. A high reaction temperature and low carbon monoxide pressure favors the isomerization of the Markovnikov product to the thermodynamically more stable β-isomer, which leads to the n-aldehyde. Low temperatures and high carbon monoxide pressure and an excess of phosphine, which blocks free coordination sites, can lead to faster hydroformylation in the α-position to the ester group and suppress the isomerization. [27]

Side- and consecutive reactions

Alkenes

Side reactions of the alkenes are the isomerization and hydrogenation of the double bond. While the alkanes resulting from hydrogenation of the double bond do not participate further in the reaction, the isomerization of the double bond with subsequent formation of the n-alkyl complexes is a desired reaction. The hydrogenation is usually of minor importance; However, cobalt-phosphine-modified catalysts can have an increased hydrogenation activity, where up to 15% of the alkene is hydrogenated.

Aldehydes

A usually desired consecutive reaction is the hydrogenation of the aldehydes to alcohols. Higher temperatures and hydrogen partial pressures favor the hydrogenation of the resulting aldehyde to the alcohol. For the reaction mechanism it is believed that the aldehyde initially forms a CO-π-complex with the catalyst. This is rearranged to the alkoxide complex and by subsequent oxidative addition of hydrogen the alcohol is eliminated and the starting complex is formed.

The aldehydic carbon-oxygen double bond can also be subject to hydroformylation, which leads to formic acid and its esters. The reaction requires the carbon monoxide insertion into the oxygen-metal bond of the alkoxide complex. The resulting formyl complex can converted into the formic acid esters and the starting complex by oxidative addition of hydrogen. The initially produced aldehydes can react further by aldol condensation to either target product precursors like 2-ethylhexenal or higher molecular weight condensation products, so-called thick oil.

Catalyst and ligands

Conditions for hydroformylation catalysis can induce degradation of supporting organophosphorus ligands. Triphenylphosphine is subject to hydrogenolysis, releasing benzene and diphenylphosphine. The insertion of carbon monoxide in an intermediate metal-phenyl bond can lead to the formation of benzaldehyde or by subsequent hydrogenation to benzyl alcohol. [28] One of the ligands phenyl-groups can be replaced by propene, and the resulting diphenylpropylphosphine ligand can inhibit the hydroformylation reaction due to its increased basicity. [28]

See also

Related Research Articles

<span class="mw-page-title-main">Organometallic chemistry</span> Study of organic compounds containing metal(s)

Organometallic chemistry is the study of organometallic compounds, chemical compounds containing at least one chemical bond between a carbon atom of an organic molecule and a metal, including alkali, alkaline earth, and transition metals, and sometimes broadened to include metalloids like boron, silicon, and selenium, as well. Aside from bonds to organyl fragments or molecules, bonds to 'inorganic' carbon, like carbon monoxide, cyanide, or carbide, are generally considered to be organometallic as well. Some related compounds such as transition metal hydrides and metal phosphine complexes are often included in discussions of organometallic compounds, though strictly speaking, they are not necessarily organometallic. The related but distinct term "metalorganic compound" refers to metal-containing compounds lacking direct metal-carbon bonds but which contain organic ligands. Metal β-diketonates, alkoxides, dialkylamides, and metal phosphine complexes are representative members of this class. The field of organometallic chemistry combines aspects of traditional inorganic and organic chemistry.

<span class="mw-page-title-main">Hydrogenation</span> Chemical reaction between molecular hydrogen and another compound or element

Hydrogenation is a chemical reaction between molecular hydrogen (H2) and another compound or element, usually in the presence of a catalyst such as nickel, palladium or platinum. The process is commonly employed to reduce or saturate organic compounds. Hydrogenation typically constitutes the addition of pairs of hydrogen atoms to a molecule, often an alkene. Catalysts are required for the reaction to be usable; non-catalytic hydrogenation takes place only at very high temperatures. Hydrogenation reduces double and triple bonds in hydrocarbons.

In chemistry, homogeneous catalysis is catalysis where the catalyst is in same phase as reactants, principally by a soluble catalyst a in solution. In contrast, heterogeneous catalysis describes processes where the catalysts and substrate are in distinct phases, typically solid-gas, respectively. The term is used almost exclusively to describe solutions and implies catalysis by organometallic compounds. Homogeneous catalysis is an established technology that continues to evolve. An illustrative major application is the production of acetic acid. Enzymes are examples of homogeneous catalysts.

<span class="mw-page-title-main">Wilkinson's catalyst</span> Chemical compound

Wilkinson's catalyst is the common name for chlorido­tris(triphenylphosphine)­rhodium(I), a coordination complex of rhodium with the formula [RhCl(PPh3)3], where 'Ph' denotes a phenyl group). It is a red-brown colored solid that is soluble in hydrocarbon solvents such as benzene, and more so in tetrahydrofuran or chlorinated solvents such as dichloromethane. The compound is widely used as a catalyst for hydrogenation of alkenes. It is named after chemist and Nobel laureate Sir Geoffrey Wilkinson, who first popularized its use.

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

Rhodium(III) chloride refers to inorganic compounds with the formula RhCl3(H2O)n, where n varies from 0 to 3. These are diamagnetic solids featuring octahedral Rh(III) centres. Depending on the value of n, the material is either a dense brown solid or a soluble reddish salt. The soluble trihydrated (n = 3) salt is widely used to prepare compounds used in homogeneous catalysis, notably for the industrial production of acetic acid and hydroformylation.

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

Dicobalt octacarbonyl is an organocobalt compound with composition Co2(CO)8. This metal carbonyl is used as a reagent and catalyst in organometallic chemistry and organic synthesis, and is central to much known organocobalt chemistry. It is the parent member of a family of hydroformylation catalysts. Each molecule consists of two cobalt atoms bound to eight carbon monoxide ligands, although multiple structural isomers are known. Some of the carbonyl ligands are labile.

<span class="mw-page-title-main">Bite angle</span>

In coordination chemistry, the bite angle is the angle on a central atom between two bonds to a bidentate ligand. This ligand–metal–ligand geometric parameter is used to classify chelating ligands, including those in organometallic complexes. It is most often discussed in terms of catalysis, as changes in bite angle can affect not just the activity and selectivity of a catalytic reaction but even allow alternative reaction pathways to become accessible.

In chemistry, carbonylation refers to reactions that introduce carbon monoxide (CO) into organic and inorganic substrates. Carbon monoxide is abundantly available and conveniently reactive, so it is widely used as a reactant in industrial chemistry. The term carbonylation also refers to oxidation of protein side chains.

In organometallic chemistry, a migratory insertion is a type of reaction wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

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

3,3′,3′′-Phosphanetriyltris(benzenesulfonic acid) trisodium salt (abbreviated TPPTS), is an organic compound that is also known as sodium triphenylphosphine trisulfonate. The compound has the formula P(C6H4SO3Na)3. This white solid is an unusual example of a water-soluble phosphine. Its complexes are also water-soluble. Its complex with rhodium is used in the industrial production of butyraldehyde.

<span class="mw-page-title-main">Concurrent tandem catalysis</span>

Concurrent tandem catalysis (CTC) is a technique in chemistry where multiple catalysts produce a product otherwise not accessible by a single catalyst. It is usually practiced as homogeneous catalysis. Scheme 1 illustrates this process. Molecule A enters this catalytic system to produce the comonomer, B, which along with A enters the next catalytic process to produce the final product, P. This one-pot approach can decrease product loss from isolation or purification of intermediates. Reactions with relatively unstable products can be generated as intermediates because they are only transient species and are immediately used in a consecutive reaction.

<span class="mw-page-title-main">Organocobalt chemistry</span> Chemistry of compounds with a carbon to cobalt bond

Organocobalt chemistry is the chemistry of organometallic compounds containing a carbon to cobalt chemical bond. Organocobalt compounds are involved in several organic reactions and the important biomolecule vitamin B12 has a cobalt-carbon bond. Many organocobalt compounds exhibit useful catalytic properties, the preeminent example being dicobalt octacarbonyl.

<span class="mw-page-title-main">Organorhodium chemistry</span> Field of study

Organorhodium chemistry is the chemistry of organometallic compounds containing a rhodium-carbon chemical bond, and the study of rhodium and rhodium compounds as catalysts in organic reactions.

<span class="mw-page-title-main">Cobalt tetracarbonyl hydride</span> Chemical compound

Cobalt tetracarbonyl hydride is an organometallic compound with the formula HCo(CO)4. It is a volatile, yellow liquid that forms a colorless vapor and has an intolerable odor. The compound readily decomposes upon melt and in absentia of high CO partial pressures forms Co2(CO)8. Despite operational challenges associated with its handling, the compound has received considerable attention for its ability to function as a catalyst in hydroformylation. In this respect, HCo(CO)4 and related derivatives have received significant academic interest for their ability to mediate a variety of carbonylation (introduction of CO into inorganic compounds) reactions.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

<span class="mw-page-title-main">Metal carbonyl hydride</span>

Metal carbonyl hydrides are complexes of transition metals with carbon monoxide and hydride as ligands. These complexes are useful in organic synthesis as catalysts in homogeneous catalysis, such as hydroformylation.

<span class="mw-page-title-main">Tris(triphenylphosphine)rhodium carbonyl hydride</span> Chemical compound

Carbonyl hydrido tris(triphenylphosphine)rhodium(I) [Carbonyl(hydrido)tris(triphenylphosphane)rhodium(I)] is an organorhodium compound with the formula [RhH(CO)(PPh3)3] (Ph = C6H5). It is a yellow, benzene-soluble solid, which is used industrially for hydroformylation.

Clark Landis is an American chemist, whose research focuses on organic and inorganic chemistry. He is currently a Professor of Chemistry at the University of Wisconsin–Madison. He was awarded the ACS Award in Organometallic Chemistry in 2010, and is a fellow of the American Chemical Society and the American Association for the Advancement of Science.

In organic chemistry, hydrovinylation is the formal insertion of an alkene into the C-H bond of ethylene. The more general reaction, hydroalkenylation, is the formal insertion of an alkene into the C-H bond of any terminal alkene. The reaction is catalyzed by metal complexes. A representative reaction is the conversion of styrene and ethylene to 3-phenybutene:

Shuttle catalysis is used to describe catalytic reactions where a chemical entity of a donor molecule is transferred to an acceptor molecule. In these reactions, while the number of chemical bonds of each reactant changes, the types and total number of chemical bonds remain constant over the course of the reaction. In contrast to many organic reactions which exothermicity practically renders them irreversible, reactions operated under shuttle catalysis are often reversible. However, the position of the equilibrium can be driven to the product side through Le Chatelier’s principle. The driving forces for this equilibrium shift are typically the formation of a gas/precipitation, the use of high ground-state energy reactants or the formation of stabilized products or the excess equivalents of a reactant.

References

  1. Robert Franke, Detlef Selent, Armin Börner (2012). "Applied Hydroformylation". Chem. Rev. 112 (11): 5675–5732. doi:10.1021/cr3001803. PMID   22937803.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  2. Ojima I., Tsai C.-Y., Tzamarioudaki M., Bonafoux D. (2000). "The Hydroformylation Reaction". Org. React. 56: 1. doi:10.1002/0471264180.or056.01. ISBN   0-471-26418-0.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  3. 1 2 Pino, P.; Botteghi, C. (1977). "Aldehydes from olefins: cyclohexanecarboxaldehyde". Organic Syntheses. 57: 11. doi:10.15227/orgsyn.057.0011.
  4. Makado, Gouki; Morimoto, Tsumoru; Sugimoto, Yasuko; Tsutsumi, Ken; Kagawa, Natsuko; Kakiuchi, Kiyomi (15 February 2010). "Highly Linear-Selective Hydroformylation of 1-Alkenes using Formaldehyde as a Syngas Substitute". Advanced Synthesis & Catalysis. 352 (2–3): 299–304. doi:10.1002/adsc.200900713. hdl: 10061/11759 . ISSN   1615-4150.
  5. Boy Cornils; Wolfgang A. Herrmann; Manfred Rasch (1994). "Otto Roelen, Pioneer in Industrial Homogeneous Catalysis". Angewandte Chemie International Edition in English. 33 (21): 2144–2163. doi:10.1002/anie.199421441.
  6. "Archived copy" (PDF). Archived from the original (PDF) on 28 September 2007. Retrieved 7 January 2007.{{cite web}}: CS1 maint: archived copy as title (link)
  7. 1 2 Richard F. Heck; David S. Breslow (1961). "The Reaction of Cobalt Hydrotetracarbonyl with Olefins". Journal of the American Chemical Society. 83 (19): 4023–4027. doi:10.1021/ja01480a017..
  8. Evans, D., Osborn, J. A., Wilkinson, G. (1968). "Hydroformylation of Alkenes by Use of Rhodium Complex Catalyst". Journal of the Chemical Society. 33 (21): 3133–3142. doi:10.1039/J19680003133.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. J. F. Hartwig; Organotransition metal chemistry – from bonding to catalysis. University Science Books. 2009. 753, 757–578. ISBN   978-1-891389-53-5.
  10. Cornils, B.; Herrmann, W. A. (eds.) "Aqueous-Phase Organometallic Catalysis" VCH, Weinheim: 1998
  11. Jack Halpern (2001). "'Organometallic chemistry at the threshold of a new millennium. Retrospect and prospect". Pure and Applied Chemistry. 73 (2): 209–220. doi: 10.1351/pac200173020209 .
  12. Gual, Aitor; Godard, Cyril; de la Fuente, Verónica; Castillón, Sergio (2012). "Design and Synthesis of Phosphite Ligands for Homogeneous Catalysis". Phosphorus(III) Ligands in Homogeneous Catalysis: Design and Synthesis. pp. 81–131. doi:10.1002/9781118299715.ch3. ISBN   978-1-118-29971-5.
  13. Kuil, M.; Soltner, T.; van Leeuwen, P. W. N. M.; Reek, J. N. H. (2006). "High-Precision Catalysts: Regioselective Hydroformylation of Internal Alkenes by Encapsulated Rhodium Complexes". Journal of the American Chemical Society. 128 (35): 11344–45. doi:10.1021/ja063294i. PMID   16939244.
  14. 1 2 Gene W. Wong; Tyler T. Adint; Clark R. Landis (2012). "Synthesis of (2R)-3-[[(1,1-Dimethylethyl)dimethylsilyl]oxy]-2-methylpropanal by Rhodium-Catalyzed Asymmetric Hydroformylation". Org. Synth. 89: 243. doi: 10.15227/orgsyn.089.0243 .
  15. Duembgen G., Neubauer D. (1969). "Grosstechnische Herstellung von Oxo-Alkoholen aus Propylen in der BASF.". Chemie Ingenieur Technik. 41 (17): 974–80. doi:10.1002/cite.330411708.
  16. 1 2 3 4 Boy Cornils, Wolfgang A. Herrmann, Chi-Huey Wong, Horst Werner Zanthoff: Catalysis from A to Z: A Concise Encyclopedia, 2408 Seiten, Verlag Wiley-VCH Verlag GmbH & Co. KGaA, (2012), ISBN   3-527-33307-X.
  17. Cuny, Gregory D.; Buchwald, Stephen L. (1993). "Practical, High-Yield, Regioselective, Rhodium-Catalyzed Hydroformylation of Functionalized α-olefins". Journal of the American Chemical Society. 115 (5): 2066–2068. doi:10.1021/ja00058a079.
  18. Van Rooy, Annemiek; Kamer, Paul C. J.; Van Leeuwen, Piet W. N. M.; Goubitz, Kees; Fraanje, Jan; Veldman, Nora; Spek, Anthony L. (1996). "Bulky Diphosphite-Modified Rhodium Catalysts: Hydroformylation and Characterization". Organometallics. 15 (2): 835–847. doi:10.1021/OM950549K.
  19. W. A. Herrmann, C. W. Kohlpaintner, Angew. Chem. 1993, 105, 1588.
  20. 1 2 3 4 5 6 Wiebus Ernst, Cornils Boy (1994). "Die großtechnische Oxosynthese mit immobilisiertem Katalysator.". Chemie Ingenieur Technik. 66 (7): 916–923. doi:10.1002/cite.330660704.
  21. Manfred Baerns, Arno Behr, Axel Brehm, Jürgen Gmehling, Hanns Hofmann, Ulfert Onken: Technische Chemie Lehrbuch. 480 Abbildungen, 190 Tabellen. Wiley VCH Verlag GmbH, September 2006, ISBN   3-527-31000-2.
  22. Chan A.S.C.; Shieh H-S. (1994). "A mechanistic study of the homogeneous catalytic hydroformylation of formaldehyde: synthesis and characterization of model intermediates". Inorganica Chimica Acta . 218 (1–2): 89–95. doi:10.1016/0020-1693(94)03800-7.
  23. A. Spencer (1980). "Hydroformylation of formaldehyde catalysed by rhodium complexes". Journal of Organometallic Chemistry . 194 (1–2): 113–123. doi:10.1016/S0022-328X(00)90343-7.
  24. Murata K.; Matsuda A. (1981). "Application of Homogeneous Water-Gas Shift Reaction III Further Study of the Hydrocarbonylation – A highly Selective Formation of Diethyl Keton from Ethene, CO and H2O". Bulletin of the Chemical Society of Japan . 54 (7): 2089–2092. doi: 10.1246/bcsj.54.2089 .
  25. J. Liu; B.T. Heaton; J.A. Iggo; R. Whyman (2004). "The Complete Delineation of the Initiation, Propagation, and Termination Steps of the Carbomethoxy Cycle for the Carboalkoxylation of Ethene by Pd–Diphosphane Catalysts". Angew. Chem. Int. Ed. 43 (1): 90–94. doi:10.1002/anie.200352369. PMID   14694480.
  26. Fell Bernhard, Rupilius Wolfgang, Asinger Friedrich (1968). "Zur Frage der Isomerenbildung bei der Hydroformylierung höhermolekularer Olefine mit komplexen Kobalt- und Rhodiumkatalysatoren". Tetrahedron Letters. 9 (29): 3261–3266. doi:10.1016/S0040-4039(00)89542-8.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  27. 1 2 Jürgen Falbe, Ch. R. Adams: Carbon Monoxide in Organic Synthesis, Springer Verlag, 1970, ISBN   3-540-04814-6
  28. 1 2 Arno Behr: Angewandte homogene Katalyse, Wiley-VCH. Weinheim, ISBN   3-527-31666-3

Further reading