Mitsunobu reaction

Last updated

Contents

Mitsunobu reaction
Named afterOyo Mitsunobu
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal mitsunobu-reaction
RSC ontology ID RXNO:0000034

The Mitsunobu reaction is an organic reaction that converts an alcohol into a variety of functional groups, such as an ester, using triphenylphosphine and an azodicarboxylate such as diethyl azodicarboxylate (DEAD) or diisopropyl azodicarboxylate (DIAD). [1] Although DEAD and DIAD are most commonly used, there are a variety of other azodicarboxylates available which facilitate an easier workup and/or purification and in some cases, facilitate the use of more basic nucleophiles. It was discovered by Oyo Mitsunobu (1934–2003). In a typical protocol, one dissolves the alcohol, the carboxylic acid, and triphenylphosphine in tetrahydrofuran or other suitable solvent (e.g. diethyl ether), cool to 0 °C using an ice-bath, slowly add the DEAD dissolved in THF, then stir at room temperature for several hours. [2] The alcohol reacts with the phosphine to create a good leaving group then undergoes an inversion of stereochemistry in classic SN2 fashion as the nucleophile displaces it. A common side-product is produced when the azodicarboxylate displaces the leaving group instead of the desired nucleophile. This happens if the nucleophile is not acidic enough (pKa larger than 13) or is not nucleophilic enough due to steric or electronic constraints. A variation of this reaction utilizing a nitrogen nucleophile is known as a Fukuyama–Mitsunobu.

The Mitsunobu reaction Mitsunobu-Reaction V.1.svg
The Mitsunobu reaction

Several reviews have been published. [3] [4] [5] [6] [7]

Reaction mechanism

The reaction mechanism of the Mitsunobu reaction is fairly complex. The identity of intermediates and the roles they play has been the subject of debate.

Initially, the triphenyl phosphine (2) makes a nucleophilic attack upon diethyl azodicarboxylate (1) producing a betaine intermediate 3, which deprotonates the carboxylic acid (4) to form the ion pair 5. The formation of the ion pair 5 is very fast.

The initial stages of the mechanism proposed for the Mitsunobu reaction. Mitsunobu Reaction Mechanism (cropped) Top.png
The initial stages of the mechanism proposed for the Mitsunobu reaction.

The second phase of the mechanism is proposed to be phosphorus-centered, the DEAD having been converted to the hydrazine. The ratio and interconversion of intermediates 811 depend on the carboxylic acid pKa and the solvent polarity. [8] [9] [10] Although several phosphorus intermediates are present, the attack of the carboxylate anion upon intermediate 8 is the only productive pathway forming the desired product 12 and triphenylphosphine oxide (13).

The latter stages of the mechanism proposed for the Mitsunobu reaction. Mitsunobu Reaction Mechanism (cropped)bottom.png
The latter stages of the mechanism proposed for the Mitsunobu reaction.

The formation of the oxyphosphonium intermediate 8 is slow and facilitated by the alkoxide. Therefore, the overall rate of reaction is controlled by carboxylate basicity and solvation. [11]

Order of addition of reagents

The order of addition of the reagents of the Mitsunobu reaction can be important. Typically, one dissolves the alcohol, the carboxylic acid, and triphenylphosphine in tetrahydrofuran or other suitable solvent (e.g. diethyl ether), cool to 0 °C using an ice-bath, slowly add the DEAD dissolved in THF, then stir at room temperature for several hours. If this is unsuccessful, then preforming the betaine may give better results. To preform the betaine, add DEAD to triphenylphosphine in tetrahydrofuran at 0 °C, followed by the addition of the alcohol and finally the acid. [12]

Variations

Other nucleophilic functional groups

Many other functional groups can serve as nucleophiles besides carboxylic acids. For the reaction to be successful, the nucleophile must have a pKa less than 15.

NucleophileProduct
hydrazoic acid alkyl azide
imide substituted imide [13]
phenol alkyl aryl ether (discovered independently [14] [15] )
sulfonamide substituted sulfonamide [16]
arylsulfonylhydrazinealkyldiazene (subject to pericyclic or free radical dediazotization to give allene (Myers allene synthesis) or alkane (Myers deoxygenation), respectively) [17]

Modifications

Several modifications to the original reagent combination have been developed in order to simplify the separation of the product and avoid production of so much chemical waste. One variation of the Mitsunobu reaction uses resin-bound triphenylphosphine and uses di-tert-butylazodicarboxylate instead of DEAD. The oxidized triphenylphosphine resin can be removed by filtration, and the di-tert-butylazodicarboxylate byproduct is removed by treatment with trifluoroacetic acid. [18] Bruce H. Lipshutz has developed an alternative to DEAD, di-(4-chlorobenzyl)azodicarboxylate (DCAD) where the hydrazine by-product can be easily removed by filtration and recycled back to DCAD. [19]

A modification has also been reported in which DEAD can be used in catalytic versus stoichiometric quantities, however this procedure requires the use of stoichiometric (diacetoxyiodo)benzene to oxidise the hydrazine by-product back to DEAD. [20]

Denton and co-workers have reported a redox-neutral variant of the Mitsunobu reaction which employs a phosphorus(III) catalyst to activate the substrate, ensuring inversion in the nucleophilic attack, and uses a Dean-Stark trap to remove the water by-product. [21]

Phosphorane reagents

Tsunoda et al. have shown that one can combine the triphenylphosphine and the diethyl azodicarboxylate into one reagent: a phosphorane ylide. Both (cyanomethylene)trimethylphosphorane (CMMP, R = Me) and (cyanomethylene)tributylphosphorane (CMBP, R = Bu) have proven particularly effective. [22]

The mechanism of the phosphorane variant of the Mitsunobu reaction Phosphorane Mitsunobu Mechanism.png
The mechanism of the phosphorane variant of the Mitsunobu reaction

The ylide acts as both the reducing agent and the base. The byproducts are acetonitrile (6) and the trialkylphosphine oxide (8).

Uses

The Mitsunobu reaction has been applied in the synthesis of aryl ethers: [23]

Mitsunobu reaction Application.png

With these particular reactants the conversion with DEAD fails because the hydroxyl group is only weakly acidic. Instead the related 1,1'-(azodicarbonyl)dipiperidine (ADDP) is used of which the betaine intermediate is a stronger base. The phosphine is a polymer-supported triphenylphosphine (PS-PPh3).

The reaction has been used to synthesize quinine, colchicine, sarain, morphine, stigmatellin, eudistomin, oseltamivir, strychnine, and nupharamine. [24]

Related Research Articles

<span class="mw-page-title-main">Ether</span> Organic compounds made of alkyl/aryl groups bound to oxygen (R–O–R)

In organic chemistry, ethers are a class of compounds that contain an ether group—an oxygen atom connected to two organyl groups. They have the general formula R−O−R′, where R and R′ represent organyl groups. Ethers can again be classified into two varieties: if the organyl groups are the same on both sides of the oxygen atom, then it is a simple or symmetrical ether, whereas if they are different, the ethers are called mixed or unsymmetrical ethers. A typical example of the first group is the solvent and anaesthetic diethyl ether, commonly referred to simply as "ether". Ethers are common in organic chemistry and even more prevalent in biochemistry, as they are common linkages in carbohydrates and lignin.

<span class="mw-page-title-main">Ester</span> Compound derived from an acid

In chemistry, an ester is a compound derived from an acid in which the hydrogen atom (H) of at least one acidic hydroxyl group of that acid is replaced by an organyl group. Analogues derived from oxygen replaced by other chalcogens belong to the ester category as well. According to some authors, organyl derivatives of acidic hydrogen of other acids are esters as well, but not according to the IUPAC.

<span class="mw-page-title-main">Haloalkane</span> Group of chemical compounds derived from alkanes containing one or more halogens

The haloalkanes are alkanes containing one or more halogen substituents. They are a subset of the general class of halocarbons, although the distinction is not often made. Haloalkanes are widely used commercially. They are used as flame retardants, fire extinguishants, refrigerants, propellants, solvents, and pharmaceuticals. Subsequent to the widespread use in commerce, many halocarbons have also been shown to be serious pollutants and toxins. For example, the chlorofluorocarbons have been shown to lead to ozone depletion. Methyl bromide is a controversial fumigant. Only haloalkanes that contain chlorine, bromine, and iodine are a threat to the ozone layer, but fluorinated volatile haloalkanes in theory may have activity as greenhouse gases. Methyl iodide, a naturally occurring substance, however, does not have ozone-depleting properties and the United States Environmental Protection Agency has designated the compound a non-ozone layer depleter. For more information, see Halomethane. Haloalkane or alkyl halides are the compounds which have the general formula "RX" where R is an alkyl or substituted alkyl group and X is a halogen.

<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.

In organic chemistry, an acyl chloride is an organic compound with the functional group −C(=O)Cl. Their formula is usually written R−COCl, where R is a side chain. They are reactive derivatives of carboxylic acids. A specific example of an acyl chloride is acetyl chloride, CH3COCl. Acyl chlorides are the most important subset of acyl halides.

<span class="mw-page-title-main">Appel reaction</span>

The Appel reaction is an organic reaction that converts an alcohol into an alkyl chloride using triphenylphosphine and carbon tetrachloride. The use of carbon tetrabromide or bromine as a halide source will yield alkyl bromides, whereas using carbon tetraiodide, methyl iodide or iodine gives alkyl iodides. The reaction is credited to and named after Rolf Appel, it had however been described earlier. The use of this reaction is becoming less common, due to carbon tetrachloride being restricted under the Montreal protocol.

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

Phosphorus trichloride is an inorganic compound with the chemical formula PCl3. A colorless liquid when pure, it is an important industrial chemical, being used for the manufacture of phosphites and other organophosphorus compounds. It is toxic and reacts readily with water to release hydrogen chloride.

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

Triphenylphosphine (IUPAC name: triphenylphosphane) is a common organophosphorus compound with the formula P(C6H5)3 and often abbreviated to PPh3 or Ph3P. It is versatile compound that is widely used as a reagent in organic synthesis and as a ligand for transition metal complexes, including ones that serve as catalysts in organometallic chemistry. PPh3 exists as relatively air stable, colorless crystals at room temperature. It dissolves in non-polar organic solvents such as benzene and diethyl ether.

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

Diethyl azodicarboxylate, conventionally abbreviated as DEAD and sometimes as DEADCAT, is an organic compound with the structural formula CH3CH2−O−C(=O)−N=N−C(=O)−O−CH2CH3. Its molecular structure consists of a central azo functional group, RN=NR, flanked by two ethyl ester groups. This orange-red liquid is a valuable reagent but also quite dangerous and explodes upon heating. Therefore, commercial shipment of pure diethyl azodicarboxylate is prohibited in the United States and is carried out either in solution or on polystyrene particles.

<i>N</i>,<i>N</i>-Dicyclohexylcarbodiimide Chemical compound

N,N′-Dicyclohexylcarbodiimide (DCC or DCCD) is an organic compound with the chemical formula (C6H11N)2C. It is a waxy white solid with a sweet odor. Its primary use is to couple amino acids during artificial peptide synthesis. The low melting point of this material allows it to be melted for easy handling. It is highly soluble in dichloromethane, tetrahydrofuran, acetonitrile and dimethylformamide, but insoluble in water.

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

Diisopropyl azodicarboxylate (DIAD) is the diisopropyl ester of azodicarboxylic acid. It is used as a reagent in the production of many organic compounds. It is often used in the Mitsunobu reaction, where it serves as an oxidizer of triphenylphosphine to triphenylphosphine oxide. It has also been used to generate aza-Baylis-Hillman adducts with acrylates. It can also serve as a selective deprotectant of N-benzyl groups in the presence of other protecting groups.

Tetrahydropyran (THP) is the organic compound consisting of a saturated six-membered ring containing five carbon atoms and one oxygen atom. It is named by reference to pyran, which contains two double bonds, and may be produced from it by adding four hydrogens. In 2013, its preferred IUPAC name was established as oxane. The compound is a colourless volatile liquid. Derivatives of tetrahydropyran are, however, more common. 2-Tetrahydropyranyl (THP-) ethers derived from the reaction of alcohols and 3,4-dihydropyran are commonly used as protecting groups in organic synthesis. Furthermore, a tetrahydropyran ring system, i.e., five carbon atoms and an oxygen, is the core of pyranose sugars, such as glucose.

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

Triphenylphosphine oxide (often abbreviated TPPO) is the organophosphorus compound with the formula OP(C6H5)3, also written as Ph3PO or PPh3O (Ph = C6H5). This colourless crystalline compound is a common but potentially useful waste product in reactions involving triphenylphosphine. It is a popular reagent to induce the crystallizing of chemical compounds.

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

Phenylmagnesium bromide, with the simplified formula C
6
H
5
MgBr
, is a magnesium-containing organometallic compound. It is commercially available as a solution in diethyl ether or tetrahydrofuran (THF). Phenylmagnesium bromide is a Grignard reagent. It is often used as a synthetic equivalent for the phenyl "Ph" synthon.

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

Oseltamivir total synthesis concerns the total synthesis of the antiinfluenza drug oseltamivir marketed by Hoffmann-La Roche under the trade name Tamiflu. Its commercial production starts from the biomolecule shikimic acid harvested from Chinese star anise and from recombinant E. coli. Control of stereochemistry is important: the molecule has three stereocenters and the sought-after isomer is only 1 of 8 stereoisomers.

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

Triphenylphosphine dichloride, (C6H5)3PCl2, is a chlorinating agent widely used in organic chemistry. Applications include the conversion of alcohols and ethers to alkyl chlorides, the cleavage of epoxides to vicinal dichlorides and the chlorination of carboxylic acids to acyl chlorides.

Shiina esterification is an organic chemical reaction that synthesizes carboxylic esters from nearly equal amounts of carboxylic acids and alcohols by using aromatic carboxylic acid anhydrides as dehydration condensation agents. In 1994, Prof. Isamu Shiina reported an acidic coupling method using Lewis acid, and, in 2002, a basic esterification using nucleophilic catalyst.

In organic chemistry, the Myers allene synthesis is a chemical reaction that converts a propargyl alcohol into an allene by way of an arenesulfonylhydrazine as a key intermediate. This name reaction is one of two discovered by Andrew Myers that are named after him; both this reaction and the Myers deoxygenation reaction involve the same type of intermediate.

In organic chemistry, the Myers deoxygenation reaction is an organic redox reaction that reduces an alcohol into an alkyl position by way of an arenesulfonylhydrazine as a key intermediate. This name reaction is one of four discovered by Andrew Myers that are named after him; this reaction and the Myers allene synthesis reaction involve the same type of intermediate. The other reactions are Myers' asymmetric alkylation and Myers-Saito Cycloaromatization.

<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.

References

  1. Mitsunobu, O.; Yamada, Y. (1967). "Preparation of Esters of Carboxylic and Phosphoric Acid via Quaternary Phosphonium Salts". Bulletin of the Chemical Society of Japan. 40 (10): 2380–2382. doi:10.1246/bcsj.40.2380.
  2. "Organic Syntheses Procedure". orgsyn.org. Retrieved 13 February 2023.
  3. Mitsunobu, O. (1981). "The Use of Diethyl Azodicarboxylate and Triphenylphosphine in Synthesis and Transformation of Natural Products". Synthesis. 1981 (1): 1–28. doi:10.1055/s-1981-29317.
  4. Castro, B. R. (1983). "Replacement of Alcoholic Hydroxyl Groups by Halogens and Other Nucleophiles via Oxyphosphonium Intermediates". Replacement of Alcoholic Hydroxy Groups by Halogens and Other Nucleophiles via Oxyphosphonium Intermediates. Vol. 29. pp. 1–162. doi:10.1002/0471264180.or029.01. ISBN   9780471264187.{{cite book}}: |journal= ignored (help)
  5. Hughes, D. L. (1992). "The Mitsunobu Reaction". Organic Reactions. Vol. 42. pp. 335–656. doi:10.1002/0471264180.or042.02. ISBN   9780471264187.
  6. Hughes, D. L. (1996). "Progress in the Mitsunobu Reaction. A Review". Organic Preparations and Procedures International. 28 (2): 127–164. doi:10.1080/00304949609356516.
  7. Swamy, K. C. K.; Kumar, N. N. B.; Balaraman, E. & Kumar, K. V. P. P. (2009). "Mitsunobu and Related Reactions: Advances and Applications". Chemical Reviews. 109 (6): 2551–2651. doi:10.1021/cr800278z. PMID   19382806.
  8. Grochowski, E.; Hilton, B. D.; Kupper, R. J.; Michejda, C. J. (1982). "Mechanism of the triphenylphosphine and diethyl azodicarboxylate induced dehydration reactions (Mitsunobu reaction). The central role of pentavalent phosphorus intermediates". Journal of the American Chemical Society. 104 (24): 6876–6877. doi:10.1021/ja00388a110.
  9. Camp, D.; Jenkins, I. D. (1989). "The mechanism of the Mitsunobu esterification reaction. Part I. The involvement of phosphoranes and oxyphosphonium salts". The Journal of Organic Chemistry. 54 (13): 3045–3049. doi:10.1021/jo00274a016.
  10. Camp, D.; Jenkins, I. D. (1989). "The mechanism of the Mitsunobu esterification reaction. Part II. The involvement of (acyloxy)alkoxyphosphoranes". The Journal of Organic Chemistry. 54 (13): 3049–3054. doi:10.1021/jo00274a017.
  11. Hughes, D. L.; Reamer, R. A.; Bergan, J. J.; Grabowski, E. J. J. (1988). "A mechanistic study of the Mitsunobu esterification reaction". Journal of the American Chemical Society. 110 (19): 6487–6491. doi:10.1021/ja00227a032.
  12. Volante, R. (1981). "A new, highly efficient method for the conversion of alcohols to thiolesters and thiols". Tetrahedron Letters . 22 (33): 3119–3122. doi:10.1016/S0040-4039(01)81842-6.
  13. Hegedus, L. S.; Holden, M. S.; McKearin, J. M. (1984). "cis-N-TOSYL-3-METHYL-2-AZABICYCLO[3.3.0]OCT-3-ENE". Organic Syntheses . 62: 48; Collected Volumes, vol. 7, p. 501.
  14. Manhas, Maghar S.; Hoffman, W. H.; Lal, Bansi; Bose, Ajay K. (1975). "Steroids. Part X. A convenient synthesis of alkyl aryl ethers". Journal of the Chemical Society, Perkin Transactions 1 (5): 461–463. doi:10.1039/P19750000461.
  15. Bittner, Shmuel; Assaf, Yonit (1975). "Use of activated alcohols in the formation of aryl ethers". Chemistry & Industry (6): 281.
  16. Kurosawa, W.; Kan, T.; Fukuyama, T. (2002). "PREPARATION OF SECONDARY AMINES FROM PRIMARY AMINES VIA 2-NITROBENZENESULFONAMIDES: N-(4-METHOXYBENZYL)-3-PHENYLPROPYLAMINE". Organic Syntheses . 79: 186; Collected Volumes, vol. 10, p. 482..
  17. Myers, Andrew G.; Zheng, Bin (1996). "New and Stereospecific Synthesis of Allenes in a Single Step from Propargylic Alcohols". Journal of the American Chemical Society. 118 (18): 4492–4493. doi:10.1021/ja960443w. ISSN   0002-7863.
  18. Pelletier, J. C.; Kincaid, S. (2000). "Mitsunobu reaction modifications allowing product isolation without chromatography: application to a small parallel library". Tetrahedron Letters. 41 (6): 797–800. doi:10.1016/S0040-4039(99)02214-5.
  19. Lipshutz, B. H.; Chung, D. W.; Rich. B.; Corral, R. (2006). "Simplification of the Mitsunobu Reaction. Di-p-chlorobenzyl Azodicarboxylate: A New Azodicarboxylate". Organic Letters. 8 (22): 5069–5072. doi:10.1021/ol0618757. PMID   17048845.
  20. But, T. Y.; Toy, P. H. (2006). "Organocatalytic Mitsunobu Reactions". Journal of the American Chemical Society. 128 (30): 9636–9637. doi:10.1021/ja063141v. PMID   16866510.
  21. Beddoe, Rhydian H.; Andrews, Keith G.; Magné, Valentin; Cuthbertson, James D.; Saska, Jan; Shannon-Little, Andrew L.; Shanahan, Stephen E.; Sneddon, Helen F.; Denton, Ross M. (30 August 2019). "Redox-neutral organocatalytic Mitsunobu reactions". Science. 365 (6456): 910–914. Bibcode:2019Sci...365..910B. doi: 10.1126/science.aax3353 . ISSN   0036-8075. PMID   31467220. S2CID   201672396.
  22. Tsunoda, T.; Nagino, C.; Oguri, M.; Itô, S. (1996). "Mitsunobu-type alkylation with active methine compounds". Tetrahedron Letters. 37 (14): 2459–2462. doi:10.1016/0040-4039(96)00318-8.
  23. Humphries, P. S.; Do, Q. Q. T.; Wilhite, D. M. (2006). "ADDP and PS-PPh3: an efficient Mitsunobu protocol for the preparation of pyridine ether PPAR agonists". Beilstein Journal of Organic Chemistry. 2 (21): 21. doi: 10.1186/1860-5397-2-21 . PMC   1705810 . PMID   17076898.
  24. Mitsunobu Reaction at SynArchive Accessed 26 April 2014

See also