Diethyl azodicarboxylate

Last updated
Diethyl azodicarboxylate
DEAD.svg
DEAD-3D-balls.png
Names
IUPAC name
Diethyl diazenedicarboxylate
Systematic IUPAC name
Ethyl N-ethoxycarbonyliminocarbamate
Other names
  • DEAD
  • DEADCAT
  • Diethyl azodiformate
  • Diazenedicarboxylic acid, diethyl ester [1]
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.016.202 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 217-821-7
PubChem CID
UNII
  • InChI=1S/C6H10N2O4/c1-3-11-5(9)7-8-6(10)12-4-2/h3-4H2,1-2H3/b8-7+ Yes check.svgY
    Key: FAMRKDQNMBBFBR-BQYQJAHWSA-N Yes check.svgY
  • InChI=1/C6H10N2O4/c1-3-11-5(9)7-8-6(10)12-4-2/h3-4H2,1-2H3/b8-7+
    Key: FAMRKDQNMBBFBR-BQYQJAHWBW
  • O=C(/N=N/C(=O)OCC)OCC
Properties
C6H10N2O4
Molar mass 174.156 g·mol−1
AppearanceOrange to red to orange liquid [2]
Density 1.11 g/cm3 [3]
Melting point 6 °C (43 °F; 279 K) [4]
Boiling point 104.5 °C (220.1 °F; 377.6 K) at 12 mm Hg [3]
1.420 (20 °C) [3]
Hazards
GHS labelling:
GHS-pictogram-explos.svg GHS-pictogram-exclam.svg
Danger
H240, H302, H312, H315, H319, H332, H335
P210, P220, P234, P261, P264, P270, P271, P280, P301+P312, P302+P352, P304+P312, P304+P340, P305+P351+P338, P312, P321, P322, P330, P332+P313, P337+P313, P362, P363, P370+P378, P370+P380+P375, P403+P233, P403+P235, P405, P411, P420, P501
Flash point 85 °C (185 °F; 358 K) [5]
Safety data sheet (SDS) External MSDS
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

Diethyl azodicarboxylate, conventionally abbreviated as DEAD and sometimes as DEADCAT, [6] [7] 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.

Contents

DEAD is an aza-dienophile and an efficient dehydrogenating agent, converting alcohols to aldehydes, thiols to disulfides and hydrazo groups to azo groups; it is also a good electron acceptor. While DEAD is used in numerous chemical reactions it is mostly known as a key component of the Mitsunobu reaction, a common strategy for the preparation of an amine, azide, ether, thioether, or ester from the corresponding alcohol. [8] It is used in the synthesis of various natural products and pharmaceuticals such as zidovudine, an AIDS drug; FdUMP, a potent antitumor agent; and procarbazine, a chemotherapy drug.

Properties

DEAD is an orange-red liquid which weakens its color to yellow or colorless upon dilution or chemical reaction. This color change is conventionally used for visual monitoring of the synthesis. DEAD dissolves in most common organic solvents, such as toluene, chloroform, ethanol, [9] tetrahydrofuran and dichloromethane [3] [10] but has low solubility in water or carbon tetrachloride; the solubility in water is higher for the related azo compound dimethyl azodicarboxylate. [4]

DEAD is a strong electron acceptor and easily oxidizes a solution of sodium iodide in glacial acetic acid. It also reacts vigorously with hydrazine hydrate producing diethyl hydrazodicarboxylate and evolving nitrogen. Linear combination of atomic orbitals molecular orbital method (LCAO-MO) calculations suggest that the molecule of DEAD is unusual in having a high-lying vacant bonding orbital, and therefore tends to withdraw hydrogen atoms from various hydrogen donors. Photoassisted removal of hydrogen by DEAD was demonstrated for isopropyl alcohol, resulting in pinacol and tetraethyl tetrazanetetracarboxylate, and for acetaldehyde yielding diacetyl and diethyl hydrazodicarboxylate. Similarly, reacting DEAD with ethanol and cyclohexanol abstracts hydrogen producing acetaldehyde and cyclohexanone. Those reactions also proceed without light, although at much lower yields. [9] Thus, in general DEAD is an aza-dienophile and dehydrogenating agent, converting alcohols to aldehydes, thiols to disulfides and hydrazo groups to azo groups. [11] It also undergoes pericyclic reactions with alkenes and dienes via ene and Diels–Alder mechanisms. [12]

Preparation

Although available commercially, diethyl azodicarboxylate can be prepared fresh in the laboratory, especially if required in pure, non-diluted form. A two-step synthesis starts from hydrazine, first by alkylation with ethyl chloroformate, followed by treating the resulting diethyl hydrazodicarboxylate with chlorine (bubbling through the solution), hypochlorous acid, concentrated nitric acid or red fuming nitric acid. The reaction is carried out in an ice bath, and the reagents are added dropwise so that the temperature does not rise above 20 °C. Diethyl hydrazodicarboxylate is a solid with melting temperature of 131–133 °C which is collected as a residue; it is significantly more stable to heating than DEAD and is conventionally dried at a temperature of about 80 °C. [4] [13]

Synthesis of DEAD.png

Applications

Mitsunobu reaction

DEAD is a reagent in the Mitsunobu reaction where it forms an adduct with phosphines (usually triphenylphosphine) and assists the synthesis of esters, ethers, amines and thioethers from alcohols. Reactions normally result in the inversion of molecular symmetry.

Mitsunobu-Reaction V.1.svg

DEAD was used in the original 1967 article by Oyo Mitsunobu, [14] and his 1981 review on the use of diethyl azodicarboxylate is a top-cited chemistry article. [15] [16] The Mitsunobu reaction has several applications in the synthesis of natural products and pharmaceuticals.

Synthesis of AZT from thymidyne.png

In the above reaction, which is assisted either by DEAD or DIAD (diisopropyl azodicarboxylate), thymidine 1 transforms to the derivative 2. The latter easily converts to zidovudine 4 (also known as azidothymidine or AZT), an important antiviral drug, used among others in the treatment of AIDS. [17] [18] [19] Another example of pharmaceutical application of DEAD-assisted Mitsunobu reaction is the synthesis of bis[(pivaloyloxy)methyl [PIVz] derivative of 2'-deoxy-5-fluorouridine 5'-monophosphate (FdUMP), which is a potent antitumor agent. [20]

Michael reaction

The azo group in DEAD is a Michael acceptor. In the presence of a copper(II) catalyst, DEAD assists conversion of β-keto esters to the corresponding hydrazine derivatives. [21]

DEAD1.png

The substitution of boronic acid esters proceeds similarly: [22]

DEAD2.png

Other reactions

DEAD is an efficient component in Diels-Alder reactions and in click chemistry, for example the synthesis of bicyclo[2.1.0]pentane, which originates from Otto Diels. [23] It has also been used to generate aza-Baylis-Hillman adducts with acrylates. [24]

DEAD can be used for synthesis of heterocyclic compounds. Thus, pyrazoline derivatives convert by condensation to α,β-unsaturated ketones: [25]

HuisgenZwitterion.png

Another application is the use of DEAD as an enophile in ene reactions:

DEAD5.png

Safety

DEAD is toxic, shock and light sensitive; it can violently explode when its undiluted form is heated above 100 °C. [2] [3] [26] Shipment by air of pure diethyl azodicarboxylate is prohibited in the United States and is carried out in solution, typically about 40% DEAD in toluene. [27] Alternatively, DEAD is transported and stored on 100–300 mesh polystyrene particles at a concentration of about 1 mmol/g. [3] The time-weighed average threshold limit value for exposure to DEAD over a typical 40-hour working week is 50 parts per million; that is, DEAD is half as toxic as, e.g., carbon monoxide. [28] Safety hazards have resulted in rapid decline of DEAD usage and replacement with DIAD and other similar compounds.

Related Research Articles

<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">Thioester</span> Organosulfur compounds of the form R–SC(=O)–R’

In organic chemistry, thioesters are organosulfur compounds with the molecular structure R−C(=O)−S−R’. They are analogous to carboxylate esters with the sulfur in the thioester replacing oxygen in the carboxylate ester, as implied by the thio- prefix. They are the product of esterification of a carboxylic acid with a thiol. In biochemistry, the best-known thioesters are derivatives of coenzyme A, e.g., acetyl-CoA. The R and R' represent organyl groups, or H in the case of R.

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

Diethyl malonate, also known as DEM, is the diethyl ester of malonic acid. It occurs naturally in grapes and strawberries as a colourless liquid with an apple-like odour, and is used in perfumes. It is also used to synthesize other compounds such as barbiturates, artificial flavourings, vitamin B1, and vitamin B6.

<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">Mitsunobu reaction</span> Chemical reaction

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). 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, cool to 0 °C using an ice-bath, slowly add the DEAD dissolved in THF, then stir at room temperature for several hours. 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 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.

<span class="mw-page-title-main">Organic redox reaction</span> Redox reaction that takes place with organic compounds

Organic reductions or organic oxidations or organic redox reactions are redox reactions that take place with organic compounds. In organic chemistry oxidations and reductions are different from ordinary redox reactions, because many reactions carry the name but do not actually involve electron transfer. Instead the relevant criterion for organic oxidation is gain of oxygen and/or loss of hydrogen. Simple functional groups can be arranged in order of increasing oxidation state. The oxidation numbers are only an approximation:

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

Meldrum's acid or 2,2-dimethyl-1,3-dioxane-4,6-dione is an organic compound with formula C6H8O4. Its molecule has a heterocyclic core with four carbon and two oxygen atoms; the formula can also be written as [−O−(C 2)−O−(C=O)−(CH2)−(C=O)−].

The Barton–McCombie deoxygenation is an organic reaction in which a hydroxy functional group in an organic compound is replaced by a hydrogen to give an alkyl group. It is named after British chemists Sir Derek Harold Richard Barton and Stuart W. McCombie.

<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.

Organophosphorus chemistry is the scientific study of the synthesis and properties of organophosphorus compounds, which are organic compounds containing phosphorus. They are used primarily in pest control as an alternative to chlorinated hydrocarbons that persist in the environment. Some organophosphorus compounds are highly effective insecticides, although some are extremely toxic to humans, including sarin and VX nerve agents.

<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.

The Markó–Lam deoxygenation is an organic chemistry reaction where the hydroxy functional group in an organic compound is replaced by a hydrogen atom to give an alkyl group. The Markó-Lam reaction is a variant of the Bouveault–Blanc reduction and an alternative to the classical Barton–McCombie deoxygenation. It is named for the Belgian chemists István Markó and Kevin Lam.

Electrophilic amination is a chemical process involving the formation of a carbon–nitrogen bond through the reaction of a nucleophilic carbanion with an electrophilic source of nitrogen.

<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">2-Methyl-6-nitrobenzoic anhydride</span> Chemical compound

2-Methyl-6-nitrobenzoic anhydride is an organic acid anhydride also known as the Shiina reagent, having a structure wherein carboxylic acids undergo intermolecular dehydration condensation. It was developed in 2002 by Prof. Isamu Shiina. The compound is often abbreviated MNBA.

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

Ethyl cyanoacetate is an organic compound that contains a carboxylate ester and a nitrile. It is a colourless liquid with a pleasant odor. This material is useful as a starting material for synthesis due to its variety of functional groups and chemical reactivity.

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

Diethyl phosphite is the organophosphorus compound with the formula (C2H5O)2P(O)H. It is a popular reagent for generating other organophosphorus compounds, exploiting the high reactivity of the P-H bond. Diethyl phosphite is a colorless liquid. The molecule is tetrahedral.

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.

<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. Diethyl azodiformate. Webbook.nist.gov (1972-07-28). Retrieved on 2011-03-15.
  2. 1 2 Safety (MSDS) data for diethyl azodicarboxylate Archived 2010-12-09 at the Wayback Machine . Msds.chem.ox.ac.uk. Retrieved on 2011-03-15.
  3. 1 2 3 4 5 6 W. L. F. Armarego; Christina Li Lin Chai (22 May 2009). Purification of laboratory chemicals. Butterworth-Heinemann. pp. 124–. ISBN   978-1-85617-567-8 . Retrieved 12 March 2011.
  4. 1 2 3 Kauer, J. C. "Diethyl azodicarboxylate". Organic Syntheses ; Collected Volumes, vol. 4, p. 411.
  5. L19348 Diethyl azodicarboxylate, 97% – Alfa Aesar – A Johnson Matthey Company. Alfa.com (1972-07-28). Retrieved on 2011-03-15.
  6. Caroline Cooper (23 July 2010). Organic Chemist's Desk Reference. CRC Press. pp. 109–. ISBN   978-1-4398-1164-1 . Retrieved 12 March 2011.
  7. DEAD is sometimes used for another chemical, diethyl acetylenedicarboxylate. P. N. Preston (1980). Benzimidazoles and congeneric tricyclic compounds. John Wiley and Sons. pp. 475–. ISBN   978-0-471-08189-0 . Retrieved 12 March 2011.
  8. Barbara., Czakó (2009). Strategic applications of named reactions in organic synthesis: background and detailed mechanisms ; 250 named reactions. Elsevier. ISBN   978-0-12-369483-6. OCLC   634820219.
  9. 1 2 Yoneda, Fumio; Suzuki, Kunio; Nitta, Yoshihiro (1967). "A new hydrogen-abstracting reaction with diethyl azodicarboxylate". The Journal of Organic Chemistry. 32 (3): 727–729. doi:10.1021/jo01278a049. ISSN   0022-3263.
  10. Kelmara K. Kelly (2009) Novel isotope effects and organic reaction mechanisms, PhD thesis, Texas A&M University, p. 81
  11. Fumio Yoneda; Kunio Suzuki; Yoshihiro Nitta (1966). "A New Hydrogen-Abstracting Reaction with Diethyl Azodicarboxylate". J. Am. Chem. Soc. 88 (10): 2328. doi:10.1021/ja00962a051.
  12. Eric J. Stoner; Amy C. Hart (2010). "Diethyl Azodicarboxylate". Encyclopedia of Reagents for Organic Synthesis. doi:10.1002/047084289X.rd176.pub2. ISBN   978-0-471-93623-7.
  13. Rabjohn, N. (1948). "Diethyl azodicarboxylate". Organic Syntheses. 28: 58. doi:10.15227/orgsyn.028.0058.
  14. Mitsunobu, O.; Yamada, Y. (1967). "Preparation of Esters of Carboxylic and Phosphoric Acid via Quaternary Phosphonium Salts". Bull. Chem. Soc. Jpn. 40 (10): 2380–2382. doi:10.1246/bcsj.40.2380.
  15. Mitsunobu, Oyo (1981). "The Use of Diethyl Azodicarboxylate and Triphenylphosphine in Synthesis and Transformation of Natural Products". Synthesis. 1981: 1–28. doi:10.1055/s-1981-29317.
  16. More than 4,300 citations by March 2011 according to Web of Science
  17. Czernecki, Stanislas and Valery, Jean-marc "Process for preparing AZT (3'-azido-3'-deoxy-thymidine) and related compounds" U.S. patent 5,124,442 issued 23 June 1992
  18. Czernecki, Stanislas; Valéry, Jean-Marc (1991). "An Efficient Synthesis of 3′-Azido-3′-deoxythymidine (AZT)". Synthesis. 1991 (3): 239. doi:10.1055/s-1991-26434.
  19. Dueholm, Kim L.; Aly, Youssef L.; Jørgensen, Per T.; El-Barbary, Ahmed A.; Pedersen, Erik B.; Nielsen, Claus (1993). "Convergent synthesis of 2′,3′-dideoxy-3′-methylthio and 2′,3′-dideoxy-3′-mercapto nucleosides and their disulfide analogues — Potential anti-HIV agents". Monatshefte für Chemie - Chemical Monthly. 124: 37–53. doi:10.1007/BF00808508. S2CID   93871095.
  20. Farquhar, David; Khan, Saeed; Srivastva, Devendra N.; Saunders, Priscilla P. (1994). "Synthesis and Antitumor Evaluation of Bis[(pivaloyloxy)methyl] 2'-Deoxy-5-fluorouridine 5'-Monophosphate (FdUMP): A Strategy To Introduce Nucleotides into Cells". Journal of Medicinal Chemistry. 37 (23): 3902–9. doi:10.1021/jm00049a009. PMID   7966151.
  21. Comelles, C.; et al. (2004). "Ionic and Covalent Copper(II)-Based Catalysts for Michael Additions. The Mechanism". J. Org. Chem. 69 (20): 6834–42. doi:10.1021/jo049373z. PMID   15387609.
  22. Takeshi Uemura; Naoto Chatani (2005). "Copper Salt Catalyzed Addition of Arylboronic Acids to Azodicarboxylates". J. Org. Chem. 70 (21): 8631–8634. doi:10.1021/jo051387x. PMID   16209627.
  23. Gassman PG, Mansfield KT (1969). "BICYCLO[2.1.0]PENTANE". Organic Syntheses. 49: 1. doi:10.15227/orgsyn.049.0001. ISSN   0078-6209.
  24. Shi, Min; Zhao, Gui-Ling (2004). "Aza-Baylis–Hillman reactions of diisopropyl azodicarboxylate or diethyl azodicarboxylate with acrylates and acrylonitrile". Tetrahedron . 60 (9): 2083–2089. doi:10.1016/j.tet.2003.12.059.
  25. Vijay Nair; Smitha C. Mathew; Akkattu T. Biju; Eringathodi Suresh (2007). "A Novel Reaction of the "Huisgen Zwitterion" with Chalcones and Dienones: An Efficient Strategy for the Synthesis of Pyrazoline and Pyrazolopyridazine Derivatives". Angew. Chem. Int. Ed. 46 (12): 2070–2073. doi:10.1002/anie.200604025. PMID   17286329.
  26. G. C. Barrett (1999). Amino acid derivatives: a practical approach. Oxford University Press. pp. 119–. ISBN   978-0-19-855853-8 . Retrieved 12 March 2011.
  27. Diethyl Azodicarboxylate in Chemical Synthesis. Sigmaaldrich.com. Retrieved on 2011-03-15.
  28. Livius Cotarca; Heiner Eckert (2004). Phosgenations—a handbook. Wiley-VCH. p. 42. ISBN   978-3-527-29823-5 . Retrieved 12 March 2011.