1,3-Dipolar cycloaddition

Last updated

Huisgen 1,3-dipolar cycloaddition
Named after Rolf Huisgen
Reaction type Ring forming reaction
Identifiers
Organic Chemistry Portal huisgen-1,3-dipolar-cycloaddition
RSC ontology ID RXNO:0000018

The 1,3-dipolar cycloaddition is a chemical reaction between a 1,3-dipole and a dipolarophile to form a five-membered ring. The earliest 1,3-dipolar cycloadditions were described in the late 19th century to the early 20th century, following the discovery of 1,3-dipoles. Mechanistic investigation and synthetic application were established in the 1960s, primarily through the work of Rolf Huisgen. [1] [2] Hence, the reaction is sometimes referred to as the Huisgen cycloaddition (this term is often used to specifically describe the 1,3-dipolar cycloaddition between an organic azide and an alkyne to generate 1,2,3-triazole). 1,3-dipolar cycloaddition is an important route to the regio- and stereoselective synthesis of five-membered heterocycles and their ring-opened acyclic derivatives. The dipolarophile is typically an alkene or alkyne, but can be other pi systems. When the dipolarophile is an alkyne, aromatic rings are generally produced.

Contents

Example of 1,3-dipolar cycloaddition.tif

Mechanistic overview

Originally two proposed mechanisms describe the 1,3-dipolar cycloaddition: first, the concerted pericyclic cycloaddition mechanism, proposed by Rolf Huisgen; [3] and second, the stepwise mechanism involving a diradical intermediate, proposed by Firestone. [4] After much debate, the former proposal is now generally accepted [5] —the 1,3-dipole reacts with the dipolarophile in a concerted, often asynchronous, and symmetry-allowed π4s + π2s fashion through a thermal six-electron Huckel aromatic transition state. However, a few examples exist of a stepwise mechanism for the catalyst-free 1,3-dipolar cycloaddition reactions of thiocarbonyl ylides, [6] and nitrile oxides [7]

The generic mechanism of a 1,3-dipolar cycloaddition between a dipole and a dipolarophile to give a five-membered heterocycle, through a six-electron transition state. Note that the red curly arrows are conventionally used to denote the reaction process but do not necessarily represent the actual flow of electrons. General mechanism of 1,3-dipolar cycloaddition.png
The generic mechanism of a 1,3-dipolar cycloaddition between a dipole and a dipolarophile to give a five-membered heterocycle, through a six-electron transition state. Note that the red curly arrows are conventionally used to denote the reaction process but do not necessarily represent the actual flow of electrons.

Pericyclic mechanism

Huisgen investigated a series of cycloadditions between the 1,3-dipolar diazo compounds and various dipolarophilic alkenes. [3] The following observations support the concerted pericyclic mechanism, and refute the stepwise diradical or the stepwise polar pathway.

1,3-Dipole

Structure and nomenclature of all second-row 1,3-dipoles consisting of carbon, nitrogen and oxygen centers. The dipoles are categorized as allyl-type or propargyl/allenyl-type based on the geometry of the central atom. 18 species of second-row 1,3-dipoles.tif
Structure and nomenclature of all second-row 1,3-dipoles consisting of carbon, nitrogen and oxygen centers. The dipoles are categorized as allyl-type or propargyl/allenyl-type based on the geometry of the central atom.

A 1,3-dipole is an organic molecule that can be represented as either an allyl-type or a propargyl/allenyl-type zwitterionic octet/sextet structures. Both types of 1,3-dipoles share four electrons in the π-system over three atoms. The allyl-type is bent whereas the propargyl/allenyl-type is linear in geometry. [8] 1,3-Dipoles containing higher-row elements such as sulfur or phosphorus are also known, but are utilized less routinely.

Resonance structures can be drawn to delocalize both negative and positive charges onto any terminus of a 1,3-dipole (see the scheme below). A more accurate method to describe the electronic distribution on a 1,3-dipole is to assign the major resonance contributor based on experimental or theoretical data, such as dipole moment measurements [9] or computations. [10] For example, diazomethane bears the largest negative character at the terminal nitrogen atom, while hydrazoic acid bears the largest negative character at the internal nitrogen atom.

Calculated major resonance structures of diazomethane and hydrazoic acid (doi = 10.1021/ja00475a007) Major resonance structures of diazomethane and hydrazoic acid.tif
Calculated major resonance structures of diazomethane and hydrazoic acid (doi = 10.1021/ja00475a007)

Consequently, this ambivalence means that the ends of a 1,3-dipole can be treated as both nucleophilic and electrophilic at the same time. The extent of nucleophilicity and electrophilicity at each end can be evaluated using the frontier molecular orbitals, which can be obtained computationally. In general, the atom that carries the largest orbital coefficient in the HOMO acts as the nucleophile, whereas that in the LUMO acts as the electrophile. The most nucleophilic atom is usually, but not always, the most electron-rich atom. [11] [12] [13] In 1,3-dipolar cycloadditions, identity of the dipole-dipolarophile pair determines whether the HOMO or the LUMO character of the 1,3-dipole will dominate (see discussion on frontier molecular orbitals below).

Dipolarophile

The most commonly used dipolarophiles are alkenes and alkynes. Heteroatom-containing dipolarophiles such as carbonyls and imines can also undergo 1,3-dipolar cycloaddition. Other examples of dipolarophiles include fullerenes and nanotubes, which can undergo 1,3-dipolar cycloaddition with azomethine ylide in the Prato reaction.

Solvent effects

1,3-Dipolar cycloadditions experience very little solvent effect because both the reactants and the transition states are generally non-polar. For example, the rate of reaction between phenyl diazomethane and ethyl acrylate or norbornene (see scheme below) changes only slightly upon varying solvents from cyclohexane to methanol. [14]

Effect of solvent polarity on 1,3-dipolar cycloaddition reactions(doi:10.3987/S(N)-1978-01-0109.) Solvent effect.tif
Effect of solvent polarity on 1,3-dipolar cycloaddition reactions(doi:10.3987/S(N)-1978-01-0109.)

Lack of solvent effects in 1,3-dipolar cycloaddition is clearly demonstrated in the reaction between enamines and dimethyl diazomalonate (see scheme below). [15] The polar reaction, N-cyclopentenyl pyrrolidine nucleophilic addition to the diazo compound, proceeds 1,500 times faster in polar DMSO than in non-polar decalin. On the other hand, a close analog of this reaction, N-cyclohexenyl pyrrolidine 1,3-dipolar cycloaddition to dimethyl diazomalonate, is sped up only 41-fold in DMSO relative to decalin.

Rate of polar nucleophilic addition reaction versus 1,3-dipolar cycloaddition in decalin and in DMSO (doi:10.1016/S0040-4039(00)70991-9) The solvent effect on the reaction between enamines and diazomalonate.tif
Rate of polar nucleophilic addition reaction versus 1,3-dipolar cycloaddition in decalin and in DMSO (doi:10.1016/S0040-4039(00)70991-9)

Frontier molecular orbital theory

Orbital overlaps in types I, II and III 1,3-dipolar cycloaddition. Frontier molecular orbitals overlap in 1,3-dipolar cycloadditions.tif
Orbital overlaps in types I, II and III 1,3-dipolar cycloaddition.

1,3-Dipolar cycloadditions are pericyclic reactions, which obey the Dewar-Zimmerman rules and the Woodward–Hoffmann rules. In the Dewar-Zimmerman treatment, the reaction proceeds through a 5-center, zero-node, 6-electron Huckel transition state for this particular molecular orbital diagram. However, each orbital can be randomly assigned a sign to arrive at the same result. In the Woodward–Hoffmann treatment, frontier molecular orbitals (FMO) of the 1,3-dipole and the dipolarophile overlap in the symmetry-allowed π4s + π2s manner. Such orbital overlap can be achieved in three ways: type I, II and III. [16] The dominant pathway is the one which possesses the smallest HOMO-LUMO energy gap.

Type I

The dipole has a high-lying HOMO which overlaps with LUMO of the dipolarophile. A dipole of this class is referred to as a HOMO-controlled dipole or a nucleophilic dipole, which includes azomethine ylide, carbonyl ylide, nitrile ylide, azomethine imine, carbonyl imine and diazoalkane. These dipoles add to electrophilic alkenes readily. Electron-withdrawing groups (EWG) on the dipolarophile would accelerate the reaction by lowering the LUMO, while electron-donating groups (EDG) would decelerate the reaction by raising the HOMO. For example, the reactivity scale of diazomethane against a series of dipolarophiles is shown in the scheme below. Diazomethane reacts with the electron-poor ethyl acrylate more than a million times faster than the electron rich butyl vinyl ether. [17]

This type resembles the normal-electron-demand Diels-Alder reaction, in which the diene HOMO combines with the dienophile LUMO.

doi:10.1016/S0040-4039(01)92781-9 The reactivity of Type I 1,3-dipole against dipolarophiles.tif
doi:10.1016/S0040-4039(01)92781-9

Type II

HOMO of the dipole can pair with LUMO of the dipolarophile; alternatively, HOMO of the dipolarophile can pair with LUMO of the dipole. This two-way interaction arises because the energy gap in either direction is similar. A dipole of this class is referred to as a HOMO-LUMO-controlled dipole or an ambiphilic dipole, which includes nitrile imide, nitrone, carbonyl oxide, nitrile oxide, and azide. Any substituent on the dipolarophile would accelerate the reaction by lowering the energy gap between the two interacting orbitals; i.e., an EWG would lower the LUMO while an EDG would raise the HOMO. For example, azides react with various electron-rich and electron-poor dipolarophile with similar reactivities (see reactivity scale below). [18]

doi:10.1021/ja01016a011 The reactivity of Type II 1,3-dipole against dipolarophiles.tif
doi:10.1021/ja01016a011

Type III

The dipole has a low-lying LUMO which overlaps with HOMO of the dipolarophile (indicated by red dashed lines in the diagram). A dipole of this class is referred to as a LUMO-controlled dipole or an electrophilic dipole, which includes nitrous oxide and ozone. EWGs on the dipolarophile decelerate the reaction, while EDGs accelerate the reaction. For example, ozone reacts with the electron-rich 2-methylpropene about 100,000 times faster than the electron-poor tetrachloroethene (see reactivity scale below). [19]

This type resembles the inverse electron-demand Diels-Alder reaction, in which the diene LUMO combines with the dienophile HOMO.

doi:10.1021/ja01016a011 The reactivity of Type III 1,3-dipole against dipolarophiles.tif
doi:10.1021/ja01016a011

Reactivity

Concerted processes such as the 1,3-cycloaddition require a highly ordered transition state (high negative entropy of activation) and only moderate enthalpy requirements. Using competition reaction experiments, relative rates of addition for different cycloaddition reactions have been found to offer general findings on factors in reactivity.

See Huisgen reference
.mw-parser-output cite.citation{font-style:inherit;word-wrap:break-word}.mw-parser-output .citation q{quotes:"\"""\"""'""'"}.mw-parser-output .citation:target{background-color:rgba(0,127,255,0.133)}.mw-parser-output .id-lock-free a,.mw-parser-output .citation .cs1-lock-free a{background:url("//upload.wikimedia.org/wikipedia/commons/6/65/Lock-green.svg")right 0.1em center/9px no-repeat}.mw-parser-output .id-lock-limited a,.mw-parser-output .id-lock-registration a,.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration a{background:url("//upload.wikimedia.org/wikipedia/commons/d/d6/Lock-gray-alt-2.svg")right 0.1em center/9px no-repeat}.mw-parser-output .id-lock-subscription a,.mw-parser-output .citation .cs1-lock-subscription a{background:url("//upload.wikimedia.org/wikipedia/commons/a/aa/Lock-red-alt-2.svg")right 0.1em center/9px no-repeat}.mw-parser-output .cs1-ws-icon a{background:url("//upload.wikimedia.org/wikipedia/commons/4/4c/Wikisource-logo.svg")right 0.1em center/12px no-repeat}.mw-parser-output .cs1-code{color:inherit;background:inherit;border:none;padding:inherit}.mw-parser-output .cs1-hidden-error{display:none;color:#d33}.mw-parser-output .cs1-visible-error{color:#d33}.mw-parser-output .cs1-maint{display:none;color:#3a3;margin-left:0.3em}.mw-parser-output .cs1-format{font-size:95%}.mw-parser-output .cs1-kern-left{padding-left:0.2em}.mw-parser-output .cs1-kern-right{padding-right:0.2em}.mw-parser-output .citation .mw-selflink{font-weight:inherit}
doi:10.1002/anie.196306331. Steric clash of dipolarophile substituents during 1,3-cycloaddition.png
See Huisgen reference doi : 10.1002/anie.196306331.

Stereospecificity

1,3-dipolar cycloadditions usually result in retention of configuration with respect to both the 1,3-dipole and the dipolarophile. Such high degree of stereospecificity is a strong support for the concerted over the stepwise reaction mechanisms. As mentioned before, many examples show that the reactions were stepwise, thus, presenting partial or no stereospecificity.

With respect to dipolarophile

cis-Substituents on the dipolarophilic alkene end up cis, and trans-substituents end up trans in the resulting five-membered cyclic compound (see scheme below). [20]

doi:10.3987/S-1978-01-0147 General scheme describing the stereospecificity of 1,3-dipolar cycloaddition and of the dipolarophile.tif
doi:10.3987/S-1978-01-0147

With respect to dipole

Generally, the stereochemistry of the dipole is not of major concern because only few dipoles could form stereogenic centers, and resonance structures allow bond rotation which scrambles the stereochemistry. However, the study of azomethine ylides has verified that cycloaddition is also stereospecific with respect to the dipole component. Diastereopure azomethine ylides are generated by electrocyclic ring opening of aziridines, and then rapidly trapped with strong dipolarophiles before bond rotation can take place (see scheme below). [21] [22] If weaker dipolarophiles are used, bonds in the dipole have the chance to rotate, resulting in impaired cycloaddition stereospecificity.

These results altogether confirm that 1,3-dipolar cycloaddition is stereospecific, giving retention of both the 1,3-dipole and the dipolarophile.

doi:10.1021/ja00983a052 Stereospecificity of the two termini of the dipole in 1,3-dipolar cycloaddition.tif
doi:10.1021/ja00983a052

Diastereoselectivity

When two or more stereocenters are generated during the reaction, diastereomeric transition states and products can be obtained. In the Diels-Alder cycloaddition, the endo diastereoselectivity due to secondary orbital interactions is usually observed. In 1,3-dipolar cycloadditions, however, two forces influence the diastereoselectivity: the attractive π-interaction (resembling secondary orbital interactions in the Diels-Alder cycloaddition) and the repulsive steric interaction. Unfortunately, these two forces often cancel each other, causing poor diastereoselection in 1,3-dipolar cycloaddition.

Examples of substrate-controlled diastereoselective 1,3-dipolar cycloadditions are shown below. First is the reaction between benzonitrile N-benzylide and methyl acrylate. In the transition state, the phenyl and the methyl ester groups stack to give the cis-substitution as the exclusive final pyrroline product. This favorable π-interaction offsets the steric repulsion between the phenyl and the methyl ester groups. [23] Second is the reaction between nitrone and dihydrofuran. The exo-selectivity is achieved to minimize steric repulsion. [24] Last is the intramolecular azomethine ylide reaction with alkene. The diastereoselectivity is controlled by the formation of a less strained cis-fused ring system. [25]

doi:10.1021/ja00731a056 Substrate-controlled diastereoselectivity of 1,3-dipolar cycloaddition.tif
doi:10.1021/ja00731a056

Directed 1,3-dipolar cycloaddition

Trajectory of the cycloaddition can be controlled to achieve a diastereoselective reaction. For example, metals can chelate to the dipolarophile and the incoming dipole and direct the cycloaddition selectively on one face. The example below shows addition of nitrile oxide to an enantiomerically pure allyl alcohol in the presence of a magnesium ion. The most stable conformation of the alkene places the hydroxyl group above the plane of the alkene. The magnesium then chelates to the hydroxyl group and the oxygen atom of nitrile oxide. The cycloaddition thus comes from the top face selectively. [26]

Directed dipolar cycloaddition.tif

Such diastereodirection has been applied in the synthesis of epothilones. [27]

Use of directed cycloaddition in Epothilones synthesis.tif

Regioselectivity

For asymmetric dipole-dipolarophile pairs, two regioisomeric products are possible. Both electronic/stereoelectronic and steric factors contribute to the regioselectivity of 1,3-dipolar cycloadditions. [28]

Electronic/stereoelectronic effect

The dominant electronic interaction is the combination between the largest HOMO and the largest LUMO. Therefore, regioselectivity is governed by the atoms that bear the largest orbital HOMO and LUMO coefficients. [29] [30]

For example, consider the cycloaddition of diazomethane to three dipolarophiles: methyl acrylate, styrene or methyl cinnamate. The carbon of diazomethane bears the largest HOMO, while the end olefinic carbons of methyl acrylate and styrene bear the largest LUMO. Hence, cycloaddition gives the substitution at the C-3 position regioselectively. For methyl cinnamate, the two substituents (Ph v.s. COOMe) compete at withdrawing electrons from the alkene. The carboxyl is the better electron-withdrawing group, causing the β-carbon to be most electrophilic. Thus, cycloaddition yields the carboxyl group on C-3 and the phenyl group on C-4 regioselectively.

doi:10.1021/ja00444a013 and doi:10.1021/ja00436a062 Electronic effects on the regioselectivity of 1,3-dipolar cycloaddition.tif
doi:10.1021/ja00444a013 and doi:10.1021/ja00436a062

Steric effect

Steric effects can either cooperate or compete with the aforementioned electronic effects. Sometimes steric effects completely outweighs the electronic preference, giving the opposite regioisomer exclusively. [31]

For example, diazomethane generally adds to methyl acrylate to give 3-carboxyl pyrazoline. However, by putting more steric demands into the system, we start to observe the isomeric 4-carboxyl pyrazolines. The ratio of these two regioisomers depends on the steric demands. At the extreme, increasing the size from hydrogen to t-butyl shifts the regioselectivity from 100% 3-carboxyl to 100% 4-carboxyl substitution. [32] [33]

ISBN 0-471-08364-X. and Koszinowski, J. (1980) (PhD Thesis) Steric effects on the regioselectivity of 1,3-dipolar cycloaddition.tif
ISBN   0-471-08364-X. and Koszinowski, J. (1980) (PhD Thesis)

Synthetic applications

1,3-dipolar cycloadditions are important ways toward the synthesis of many important 5-membered heterocycles such as triazoles, furans, isoxazoles, pyrrolidines, and others. Additionally, some cycloadducts can be cleaved to reveal the linear skeleton, providing another route toward the synthesis of aliphatic compounds. These reactions are tremendously useful also because they are stereospecific, diastereoselective and regioselective. Several examples are provided below.

Nitrile oxides

1,3-dipolar cycloaddition with nitrile oxides is a widely used masked-aldol reaction. Cycloaddition between a nitrile oxide and an alkene yields the cyclic isoxazoline product, whereas the reaction with an alkyne yields the isoxazole. Both isoxazolines and isoxazoles can be cleaved by hydrogenation to reveal aldol-type β-hydroxycarbonyl or Claisen-type β-dicarbonyl products, respectively.

Nitrile oxide-alkyne cycloaddition followed by hydrogenation was utilized in the synthesis of Miyakolide as illustrated in the figure below. [34]

Application of nitrile oxide in the synthesis of miyakolide.tif

Carbonyl ylides

1,3-dipolar cycloaddition reactions have emerged as powerful tools in the synthesis of complex cyclic scaffolds and molecules for medicinal, biological, and mechanistic studies. Among them, [3+2] cycloaddition reactions involving carbonyl ylides have extensively been employed to generate oxygen-containing five-membered cyclic molecules. [35]

Preparation of carbonyl ylides for 1,3-dipolar cycloaddition reactions

Ylides are regarded as positively charged heteroatoms connected to negatively charged carbon atoms, which include ylides of sulfonium, thiocarbonyl, oxonium, nitrogen, and carbonyl. [36] Several methods exist for generating carbonyl ylides, which are necessary intermediates for generating oxygen-containing five-membered ring structures, for [3+2] cycloaddition reactions.

Synthesis of carbonyl ylides from diazomethane derivatives by photocatalysis

One of the earliest examples of carbonyl ylide synthesis involves photocatalysis. [37] Photolysis of diazotetrakis(trifluoromethyl)cyclopentadiene* (DTTC) in the presence of tetramethylurea can generate the carbonyl ylide by an intermolecular nucleophilic attack and subsequent aromatization of the DTTC moiety. [37] This was isolated and characterized by X-ray crystallography due to the stability imparted by aromaticity, electron withdrawing trifluoromethyl groups, and the electron donating dimethylamine groups. Stable carbonyl ylide dipoles can then be used in [3+2] cycloaddition reactions with dipolarophiles.

Scheme 1. Photolysis of DTTC in the presence of tetramethylurea. Modified from Janulis, E. P.; Arduengo, A. J. J. Am. Chem. Soc. 1983, 105, 5929. Photolysis of DTTC in the presence of tetramethylurea. Modified from Janulis, E. P.; Arduengo, A. J. J. Am. Chem. Soc. 1983, 105, 5929..png
Scheme 1. Photolysis of DTTC in the presence of tetramethylurea. Modified from Janulis, E. P.; Arduengo, A. J. J. Am. Chem. Soc. 1983, 105, 5929.

Another early example of carbonyl ylide synthesis by photocatalysis was reported by Olah et al. [38] Dideuteriodiazomethane was photolysed in the presence of formaldehyde to generate the dideuterioformaldehyde carbonyl ylide.

Scheme 2. Photolysis of dideuteriodiazomethane with formaldehyde. Modified from Prakash, G. K. S.; Ellis, R. W.; Felberg, J. D.; Olah, G. A. J Am Chem Soc 1986, 108, 1341. Photolysis of Dideuteriodiazomethane with formaldehyde. Modified from Prakash, G. K. S.; Ellis, R. W.; Felberg, J. D.; Olah, G. A. J Am Chem Soc 1986, 108, 1341..png
Scheme 2. Photolysis of dideuteriodiazomethane with formaldehyde. Modified from Prakash, G. K. S.; Ellis, R. W.; Felberg, J. D.; Olah, G. A. J Am Chem Soc 1986, 108, 1341.
Synthesis of carbonyl ylides from hydroxypyrones by proton transfer

Carbonyl ylides can be synthesized by acid catalysis of hydroxy-3-pyrones in the absence of a metal catalyst. [39] An initial tautomerization occurs, followed by elimination of the leaving group to aromatize the pyrone ring and to generate the carbonyl ylide. A cycloaddition reaction with a dipolarophile lastly forms the oxacycle. This approach is less widely employed due to its limited utility and requirement for pyrone skeletons.

Scheme 3. Acid-catalyzed synthesis of carbonyl ylides from hydroxy-3-pyrones. Modified from Sammes, P. G.; Street, L. J. J. Chem. Soc., Chem. Commun. 1982, 1056. Acid-Catalyzed Synthesis of Carbonyl Ylides from Hydroxy-3-Pyrones. Modified from Sammes, P. G.; Street, L. J. J. Chem. Soc., Chem. Commun. 1982, 1056..png
Scheme 3. Acid-catalyzed synthesis of carbonyl ylides from hydroxy-3-pyrones. Modified from Sammes, P. G.; Street, L. J. J. Chem. Soc., Chem. Commun. 1982, 1056.

5-hydroxy-4-pyrones can also be used to synthesize carbonyl ylides by an intramolecular hydrogen transfer. [40] After hydrogen transfer, the carbonyl ylide can then react with dipolarophiles to form oxygen-containing rings.

Scheme 4. Intramolecular hydrogen transfer-mediated synthesis of carbonyl ylides from 5-hydroxy-4-pyrones. Modified from Garst, M. E.; McBride, B. J.; Douglass III, J. G. Tetrahedron Lett. 1983, 24, 1675. Intramolecular Hydrogen Transfer-Mediated Synthesis of Carbonyl Ylides from 5-Hydroxy-4-Pyrones. Modified from Garst, M. E.; McBride, B. J.; Douglass III, J. G. Tetrahedron Lett. 1983, 24, 1675..png
Scheme 4. Intramolecular hydrogen transfer-mediated synthesis of carbonyl ylides from 5-hydroxy-4-pyrones. Modified from Garst, M. E.; McBride, B. J.; Douglass III, J. G. Tetrahedron Lett. 1983, 24, 1675.
Synthesis of α-halocarbonyl ylides from dihalocarbenes

Dihalocarbenes have also been employed to generate carbonyl ylides, exploiting the electron withdrawing nature of dihalocarbenes. [41] [42] [43] Both phenyl(trichloromethyl)mercury and phenyl(tribromomethyl)mercury are sources dichlorocarbenes and dibromocarbenes, respectively. The carbonyl ylide can be generated upon reaction of the dihalocarbenes with ketones or aldehydes. However, the synthesis of α-halocarbonyl ylides can also undesirably lead to the loss of carbon monoxide and the generation of the deoxygenation product.

Scheme 5. a-Halocarbonyl ylide synthesis through dihalocarbene intermediates. Modified from Padwa, A.; Hornbuckle, S. F. Chem Rev 1991, 91, 263. A-Halocarbonyl Ylide Synthesis via Dihalocarbene Intermediates. Modified from Padwa, A.; Hornbuckle, S. F. Chem Rev 1991, 91, 263..png
Scheme 5. α-Halocarbonyl ylide synthesis through dihalocarbene intermediates. Modified from Padwa, A.; Hornbuckle, S. F. Chem Rev 1991, 91, 263.
Synthesis of carbonyl ylides from diazomethane derivatives by metal catalysis

A universal approach for generating carbonyl ylides involves metal catalysis of α-diazocarbonyl compounds, generally in the presence of dicopper or dirhodium catalysts. [44] After release of nitrogen gas and conversion to the metallocarbene, an intermolecular reaction with a carbonyl group can generate the carbonyl ylide. Subsequent cycloaddition reaction with an alkene or alkyne dipolarophile can afford oxygen-containing five-membered rings. Popular catalysts that give modest yields towards synthesizing oxacycles include Rh2(OAc)4 and Cu(acac)2. [45] [46]

Scheme 6. Metal-catalyzed synthesis of carbonyl ylides. Reproduced from Hodgson, D. M.; Bruckl, T.; Glen, R.; Labande, A. H.; Selden, D. A.; Dossetter, A. G.; Redgrave, A. J. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 5450. Metal-Catalyzed Synthesis of Carbonyl Ylides. Reproduced from Hodgson, D. M.; Bruckl, T.; Glen, R.; Labande, A. H.; Selden, D. A.; Dossetter, A. G.; Redgrave, A. J. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 5450..png
Scheme 6. Metal-catalyzed synthesis of carbonyl ylides. Reproduced from Hodgson, D. M.; Bruckl, T.; Glen, R.; Labande, A. H.; Selden, D. A.; Dossetter, A. G.; Redgrave, A. J. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 5450.

Mechanism of the 1,3-dipolar cycloaddition reaction mediated by metal catalysis of diazocarbonyl compounds

The universality and extensive use of 1,3-dipolar cycloaddition reactions mediated by metal catalysis of diazocarbonyl molecules, for synthesizing oxygen-containing five-membered rings, has spurred significant interest into its mechanism. Several groups have investigated the mechanism to expand the scope of synthetic molecules with respect to regio- and stereo-selectivity. However, due to the high turn over frequencies of these reactions, the intermediates and mechanism remains elusive. The generally accepted mechanism, developed by characterization of stable ruthenium-carbenoid complexes [47] and rhodium metallocarbenes, [48] involves an initial formation of a metal-carbenoid complex from the diazo compound. Elimination of nitrogen gas then affords a metallocarbene. An intramolecular nucleophilic attack by the carbonyl oxygen regenerates the metal catalyst and forms the carbonyl ylide. The carbonyl ylide can then react with an alkene or alkyne, such as dimethyl acetylenedicarboxylate (DMAD) to generate the oxacycle.

Scheme 7. Accepted mechanism of the 1,3-dipolar cycloaddition reaction mediated by metal catalysis (example dirhodium catalyst) of diazocarbonyl compounds. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004. Accepted Mechanism of the 1,3-Dipolar Cycloaddition Reaction Mediated by Metal Catalysis of Diazocarbonyl Compounds. Modified from M. Hodgson, D. et al. In Organic Reactions; John Wiley & Sons, Inc.png
Scheme 7. Accepted mechanism of the 1,3-dipolar cycloaddition reaction mediated by metal catalysis (example dirhodium catalyst) of diazocarbonyl compounds. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004.

However, it is uncertain whether the metallocarbene intermediate generates the carbonyl ylide. In some cases, metallocarbenes can also react directly with dipolarophiles. [49] In these cases, the metallocarbene, such as the dirhodium(II)tetracarboxylate carbene, is stabilized through hyperconjugative metal enolate-type interactions. [50] [51] [52] [53] Subsequent 1,3-dipolar cycloaddition reaction occurs through a transient metal-complexed carbonyl ylide. Therefore, a persistent metallocarbene can influence the stereoselectivity and regioselectivity of the 1,3-dipolar cycloaddition reaction based on the stereochemistry and size of the metal ligands.

The dirhodium(II)tetracarboxylate metallocarbene stabilized by pC-Rh-pC=O hyperconjugation. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004. The dirhodium(II)tetracarboxylate metallocarbene stabilized by pC-Rh-pC=O hyperconjugation..png
The dirhodium(II)tetracarboxylate metallocarbene stabilized by πC-Rh→πC=O hyperconjugation. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004.

The mechanism of the 1,3-dipolar cycloaddition reaction between the carbonyl ylide dipole and alkynyl or alkenyl dipolarophiles has been extensively investigated with respect to regioselectivity and stereoselectivity. As symmetric dipolarophiles have one orientation for cycloaddition, only one regioisomer, but multiple stereoisomers can be obtained. [53] On the contrary, unsymmetric dipolarophiles can have multiple regioisomers and stereoisomers. These regioisomers and stereoisomers may be predicted based on frontier molecular orbital (FMO) theory, steric interactions, and stereoelectronic interactions. [54] [55]

Scheme 9. Products of the 1,3-dipolar cycloaddition reaction between carbonyl ylide dipoles and alkenyl or alkynyl dipolarophiles. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004. Products of the 1,3-Dipolar Cycloaddition Reaction Between Carbonyl Ylide Dipoles and Alkenyl or Alkynyl Dipolarophiles. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc. 2004..png
Scheme 9. Products of the 1,3-dipolar cycloaddition reaction between carbonyl ylide dipoles and alkenyl or alkynyl dipolarophiles. Modified from M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004.
Regioselectivity of the 1,3-dipolar cycloaddition reaction mediated by metal catalysis of diazocarbonyl compounds

Regioselectivity of 1,3-dipolar cycloaddition reactions between carbonyl ylide dipoles and alkynyl or alkenyl dipolarophiles is essential for generating molecules with defined regiochemistry. FMO theory and analysis of the HOMO-LUMO energy gaps between the dipole and dipolarophile can rationalize and predict the regioselectivity of experimental outcomes. [56] [57] The HOMOs and LUMOs can belong to either the dipole or dipolarophile, for which HOMOdipole-LUMOdipolarophile or HOMOdipolarophile-LUMOdipole interactions can exist. Overlap of the orbitals with the largest coefficients can ultimately rationalize and predict results.

Scheme 10. diagram of the molecular orbital interactions of HOMOdipole-LUMOdipolarophile or HOMOdipolarophile-LUMOdipole between a carbonyl ylide dipole and alkenyl dipolarophile. Diagram of the Molecular Orbital Interactions of HOMOdipole-LUMOdipolarophile or HOMOdipolarophile-LUMOdipole Between a Carbonyl Ylide Dipole and Alkenyl Dipolarophile..png
Scheme 10. diagram of the molecular orbital interactions of HOMOdipole-LUMOdipolarophile or HOMOdipolarophile-LUMOdipole between a carbonyl ylide dipole and alkenyl dipolarophile.

The archetypal regioselectivity of the 1,3-dipolar cycloaddition reaction mediated by carbonyl ylide dipoles has been examined by Padwa and coworkers. [55] [58] Using a Rh2(OAc)4 catalyst in benzene, diazodione underwent a 1,3-dipolar cycloaddition reaction with methyl propiolate and methyl propargyl ether. The reaction with methyl propiolate affords two regioisomers with the major resulting from the HOMOdipole-LUMOdipolarophile interaction, which has the largest coefficients on the carbon proximal to the carbonyl group of the carbonyl ylide and on the methyl propiolate terminal alkyne carbon. The reaction with methyl propargyl ether affords one regioisomer resulting from the HOMOdipolarophile-LUMOdipole interaction, which has largest coefficients on the carbon distal to the carbonyl group of the carbonyl ylide and on the methyl propargyl ether terminal alkyne carbon.

Scheme 11. Regioselectivity and molecular orbital interactions of the 1,3-dipolar cycloaddition reaction between a diazodione and methyl propiolate or methyl propargyl ether. Modified from Padwa, A.; Weingarten, M. D. Chem Rev 1996, 96, 223. Regioselectivity and Molecular Orbital Interactions of the 1,3-Dipolar Cycloaddition Reaction Between a Diazodione and Methyl Propiolate or Methyl Propargyl Ether. Modified from Padwa, A.; Weingarten, M. D. Chem Rev 1996, 96, 223..png
Scheme 11. Regioselectivity and molecular orbital interactions of the 1,3-dipolar cycloaddition reaction between a diazodione and methyl propiolate or methyl propargyl ether. Modified from Padwa, A.; Weingarten, M. D. Chem Rev 1996, 96, 223.

Regioselectivities of 1,3-dipolar cycloaddition reactions mediated by metal catalysis of diazocarbonyl compounds may also be influenced by the metal through formation of stable metallocarbenes. [49] [59] Stabilization of the metallocarbene, via metal enolate-type interactions, will prevent the formation of carbonyl ylides, resulting in a direct reaction between the metallocarbene dipole and an alkynyl or alkenyl dipolarophile (see image of The dirhodium(II)tetracarboxylate metallocarbene stabilized by πC-Rh→πC=O hyperconjugation.). In this situation, the metal ligands will influence the regioselectivity and stereoselectivity of the 1,3-dipolar cycloaddition reaction.

Stereoselectivity and asymmetric induction of the 1,3-dipolar cycloaddition reaction mediated by metal catalysis of diazocarbonyl compounds

The stereoselectivity of 1,3-dipolar cycloaddition reactions between carbonyl ylide dipoles and alkenyl dipolarophiles has also been closely examined. For alkynyl dipolarophiles, stereoselectivity is not an issue as relatively planar sp2 carbons are formed, while regioselectivity must be considered (see image of the Products of the 1,3-Dipolar Cycloaddition Reaction Between Carbonyl Ylide Dipoles and Alkenyl or Alkynyl Dipolarophiles). However, for alkenyl dipolarophiles, both regioselectivity and stereoselectivity must be considered as sp3 carbons are generated in the product species.

1,3-dipolar cycloaddition reactions between carbonyl ylide dipoles and alkenyl dipolarophiles can generate diastereomeric products. [53] The exo product is characterized with dipolarophile substituents being cis to the ether bridge of the oxacycle. The endo product is characterized with the dipolarophile substituents being trans to the ether bridge of the oxacycle. Both products can be generated through pericyclic transitions states involving concerted synchronous or concerted asynchronous processes.

One early example conferred stereoselectivity in terms of endo and exo products with metal catalysts and Lewis acids. [60] Reactions with just the metal catalyst Rh2(OAc)4 prefer the exo product while reactions with the additional Lewis acid Yb(OTf)3 prefer the endo product. The endo selectivity observed for Lewis acid cycloaddition reactions is attributed to the optimized orbital overlap of the carbonyl π systems between the dipolarophile coordinated by Yb(Otf)3 (LUMO) and the dipole (HOMO). After many investigations, two primary approaches for influencing the stereoselectivity of carbonyl ylide cycloadditions have been developed that exploit the chirality of metal catalysts and Lewis acids. [53]

Facial Selectivity of the 1,3-Dipolar Cycloaddition Reaction using a Metal Catalyst and Lewis Acid Facial Selectivity of the 1,3-Dipolar Cycloaddition Reaction using a Metal Catalyst and Lewis Acid.png
Facial Selectivity of the 1,3-Dipolar Cycloaddition Reaction using a Metal Catalyst and Lewis Acid
Rationale for the Endo Selectivity of the 1,3-Dipolar Cycloaddition Reaction with a Lewis Acid Rationale for the Endo Selectivity of the 1,3-Dipolar Cycloaddition Reaction with a Lewis Acid.png
Rationale for the Endo Selectivity of the 1,3-Dipolar Cycloaddition Reaction with a Lewis Acid

The first approach employs chiral metal catalysts to modulate the endo and exo stereoselectivity. The chiral catalysts, in particular Rh2[(S)-DOSP]4 and Rh2[(S)-BPTV]4 can induce modest asymmetric induction and was used to synthesize the antifungal agent pseudolaric acid A. [61] This is a result of the chiral metal catalyst remaining associated with the carbonyl ylide during the cycloaddition, which confers facial selectivity. However, the exact mechanisms are not yet fully understood.

Asymmetric induction of the 1,3-dipolar cycloaddition reaction with chiral metal catalysts Asymmetric Induction of the 1,3-Dipolar Cycloaddition Reaction with Chiral Metal Catalysts.png
Asymmetric induction of the 1,3-dipolar cycloaddition reaction with chiral metal catalysts

The second approach employs a chiral Lewis acid catalyst to induce facial stereoselectivity after the generation of the carbonyl ylide using an achiral metal catalyst. [62] The chiral Lewis acid catalyst is believed to coordinate to the dipolarophile, which lowers the LUMO of the dipolarophile while also leading to enantioselectivity.

Asymmetric induction of the 1,3-dipolar cycloaddition reaction with chiral Lewis acid catalysts Asymmetric Induction of the 1,3-Dipolar Cycloaddition Reaction with Chiral Lewis Acid Catalysts.png
Asymmetric induction of the 1,3-dipolar cycloaddition reaction with chiral Lewis acid catalysts

Azomethine ylides

1,3-Dipolar cycloaddition between an azomethine ylide and an alkene furnishes an azacyclic structure, such as pyrrolidine. This strategy has been applied to the synthesis of spirotryprostatin A. [63]

Application of azomethine ylide in the synthesis of spirotryprostatin.tif

Ozone

Ozonolysis is a very important organic reaction. Alkenes and alkynes can be cleaved by ozonolysis to give aldehyde, ketone or carboxylic acid products.

Biological applications

The 1,3-dipolar cycloaddition between organic azides and terminal alkynes (i.e., the Huisgen cycloaddition) has been widely utilized for bioconjugation.

Copper catalysis

The Huisgen reaction generally does not proceed readily under mild conditions. Meldal et al. and Sharpless et al. independently developed a copper(I)-catalyzed version of the Huisgen reaction, CuAAC (for Copper-catalyzed Azide-Alkyne Cycloaddition), which proceeds very readily in mild, including physiological, conditions (neutral pH, room temperature and water solution). [64] [65] This reaction is also bioorthogonal: azides and alkynes are both generally absent from biological systems and therefore these functionalities can be chemoselectively reacted even in the cellular context. They also do not react with other functional groups found in nature, so they do not perturb biological systems. The reaction is so versatile that it is termed the "Click" chemistry. Although copper(I) is toxic, many protective ligands have been developed to both reduce cytotoxicity and improve rate of CuAAC, allowing it to be used in in vivo studies. [66]

Copper catalyzed AAC.tif

For example, Bertozzi et al. reported the metabolic incorporation of azide-functionalized saccharides into the glycan of the cell membrane, and subsequent labeling with fluorophore-alkyne conjugate. The result is that the cell membrane is fluorescently labeled, and can therefore be imaged using a fluorescence microscope. [67]

Metabolic labeling with GlcNAz and click chemistry.tif

Strain-promoted cycloaddition

To avoid toxicity of copper(I), Bertozzi et al. developed the strain-promoted azide-alkyne cycloaddition (SPAAC) between organic azide and strained cyclooctyne. The angle distortion of the cyclooctyne helps to speed up the reaction by both reducing the activation strain and enhancing the interactions, thereby enabling it to be used in physiological conditions without the need for the catalyst. [68]

Strained promoted AAC.tif

For instance, Ting et al. introduced an azido functionality onto specific proteins on the cell surface using a ligase enzyme. The azide-tagged protein is then labeled with cyclooctyne-fluorophore conjugate to yield a fluorescently labeled protein. [69]

Enzyme-mediated labeling with azidooctanoic acid and SPAAC.tif

Related Research Articles

<span class="mw-page-title-main">Diels–Alder reaction</span> Chemical reaction

In organic chemistry, the Diels–Alder reaction is a chemical reaction between a conjugated diene and a substituted alkene, commonly termed the dienophile, to form a substituted cyclohexene derivative. It is the prototypical example of a pericyclic reaction with a concerted mechanism. More specifically, it is classified as a thermally-allowed [4+2] cycloaddition with Woodward–Hoffmann symbol [π4s + π2s]. It was first described by Otto Diels and Kurt Alder in 1928. For the discovery of this reaction, they were awarded the Nobel Prize in Chemistry in 1950. Through the simultaneous construction of two new carbon–carbon bonds, the Diels–Alder reaction provides a reliable way to form six-membered rings with good control over the regio- and stereochemical outcomes. Consequently, it has served as a powerful and widely applied tool for the introduction of chemical complexity in the synthesis of natural products and new materials. The underlying concept has also been applied to π-systems involving heteroatoms, such as carbonyls and imines, which furnish the corresponding heterocycles; this variant is known as the hetero-Diels–Alder reaction. The reaction has also been generalized to other ring sizes, although none of these generalizations have matched the formation of six-membered rings in terms of scope or versatility. Because of the negative values of ΔH° and ΔS° for a typical Diels–Alder reaction, the microscopic reverse of a Diels–Alder reaction becomes favorable at high temperatures, although this is of synthetic importance for only a limited range of Diels-Alder adducts, generally with some special structural features; this reverse reaction is known as the retro-Diels–Alder reaction.

An ylide or ylid is a neutral dipolar molecule containing a formally negatively charged atom (usually a carbanion) directly attached to a heteroatom with a formal positive charge (usually nitrogen, phosphorus or sulfur), and in which both atoms have full octets of electrons. The result can be viewed as a structure in which two adjacent atoms are connected by both a covalent and an ionic bond; normally written X+–Y. Ylides are thus 1,2-dipolar compounds, and a subclass of zwitterions. They appear in organic chemistry as reagents or reactive intermediates.

In organic chemistry, a cycloaddition is a chemical reaction in which "two or more unsaturated molecules combine with the formation of a cyclic adduct in which there is a net reduction of the bond multiplicity". The resulting reaction is a cyclization reaction. Many but not all cycloadditions are concerted and thus pericyclic. Nonconcerted cycloadditions are not pericyclic. As a class of addition reaction, cycloadditions permit carbon–carbon bond formation without the use of a nucleophile or electrophile.

In organic chemistry, the diazo group is an organic moiety consisting of two linked nitrogen atoms at the terminal position. Overall charge-neutral organic compounds containing the diazo group bound to a carbon atom are called diazo compounds or diazoalkanes and are described by the general structural formula R2C=N+=N. The simplest example of a diazo compound is diazomethane, CH2N2. Diazo compounds should not be confused with azo compounds or with diazonium compounds.

In chemical synthesis, click chemistry is a class of simple, atom-economy reactions commonly used for joining two molecular entities of choice. Click chemistry is not a single specific reaction, but describes a way of generating products that follow examples in nature, which also generates substances by joining small modular units. In many applications, click reactions join a biomolecule and a reporter molecule. Click chemistry is not limited to biological conditions: the concept of a "click" reaction has been used in chemoproteomic, pharmacological, biomimetic and molecular machinery applications. However, they have been made notably useful in the detection, localization and qualification of biomolecules.

The azide-alkyne Huisgen cycloaddition is a 1,3-dipolar cycloaddition between an azide and a terminal or internal alkyne to give a 1,2,3-triazole. Rolf Huisgen was the first to understand the scope of this organic reaction. American chemist Karl Barry Sharpless has referred to this cycloaddition as "the cream of the crop" of click chemistry and "the premier example of a click reaction".

<span class="mw-page-title-main">Tris(benzyltriazolylmethyl)amine</span> Chemical compound

Tris( methyl)amine (TBTA) is a tertiary amine containing the 1,2,3-triazole moiety. When used as a ligand, complexed to copper(I), it allows for quantitative, regioselective formal Huisgen 1,3-dipolar cycloadditions between alkynes and azides, in a variety of aqueous and organic solvents.

The Diazoalkane 1,3-dipolar cycloaddition is a 1,3-dipolar cycloaddition between a 1,3-dipole diazo compound and a dipolarophile. When the dipolarphile is an alkene, the reaction product is a pyrazoline.

A triazole is a heterocyclic compound featuring a five-membered ring of two carbon atoms and three nitrogen atoms with molecular formula C2H3N3. Triazoles exhibit substantial isomerism, depending on the positioning of the nitrogen atoms within the ring.

<span class="mw-page-title-main">Azomethine ylide</span>

Azomethine ylides are nitrogen-based 1,3-dipoles, consisting of an iminium ion next to a carbanion. They are used in 1,3-dipolar cycloaddition reactions to form five-membered heterocycles, including pyrrolidines and pyrrolines. These reactions are highly stereo- and regioselective, and have the potential to form four new contiguous stereocenters. Azomethine ylides thus have high utility in total synthesis, and formation of chiral ligands and pharmaceuticals. Azomethine ylides can be generated from many sources, including aziridines, imines, and iminiums. They are often generated in situ, and immediately reacted with dipolarophiles.

<span class="mw-page-title-main">1,3-dipole</span> Dipolar compound with electron delocalization and charge separation over 3 atoms

In organic chemistry, a 1,3-dipolar compound or 1,3-dipole is a dipolar compound with delocalized electrons and a separation of charge over three atoms. They are reactants in 1,3-dipolar cycloadditions.

<span class="mw-page-title-main">Nitrile ylide</span>

Nitrile ylides also known as nitrilium ylides or nitrilium methylides, are generally reactive intermediates formally consisting of a carbanion of an alkyl or similar group bonded to the nitrogen atom of a cyanide unit. With a few exceptions, they cannot be isolated. However, a structure has been determined on a particularly stable nitrile ylide by X-ray crystallography. Another nitrile ylide has been captured under cryogenic conditions.

The nitrone-olefin (3+2) cycloaddition reaction is the combination of a nitrone with an alkene or alkyne to generate an isoxazoline or isoxazolidine via a (3+2) cycloaddition process. This reaction is a 1,3-dipolar cycloaddition, in which the nitrone acts as the 1,3-dipole, and the alkene or alkyne as the dipolarophile.

In organic chemistry, a dipolar compound or simply dipole is an electrically neutral molecule carrying a positive and a negative charge in at least one canonical description. In most dipolar compounds the charges are delocalized. Unlike salts, dipolar compounds have charges on separate atoms, not on positive and negative ions that make up the compound. Dipolar compounds exhibit a dipole moment.

The term bioorthogonal chemistry refers to any chemical reaction that can occur inside of living systems without interfering with native biochemical processes. The term was coined by Carolyn R. Bertozzi in 2003. Since its introduction, the concept of the bioorthogonal reaction has enabled the study of biomolecules such as glycans, proteins, and lipids in real time in living systems without cellular toxicity. A number of chemical ligation strategies have been developed that fulfill the requirements of bioorthogonality, including the 1,3-dipolar cycloaddition between azides and cyclooctynes, between nitrones and cyclooctynes, oxime/hydrazone formation from aldehydes and ketones, the tetrazine ligation, the isocyanide-based click reaction, and most recently, the quadricyclane ligation.

Copper-free click chemistry is a bioorthogonal reaction as a variant of an azide-alkyne Huisgen cycloaddition. By eliminating cytotoxic copper catalysts, the reaction proceeds without live-cell toxicity. It was developed as a faster alternative to the Staudinger ligation with the first generation of Cu-free click chemistry, producing rate constants over 63 times faster.

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

Oxanorbornadiene (OND) is a bicyclic organic compound with an oxygen atom bridging the two opposing saturated carbons of 1,4-cyclohexadiene. OND is related to all-carbon bicycle norbornadiene.

A metal-centered cycloaddition is a subtype of the more general class of cycloaddition reactions. In such reactions "two or more unsaturated molecules unite directly to form a ring", incorporating a metal bonded to one or more of the molecules. Cycloadditions involving metal centers are a staple of organic and organometallic chemistry, and are involved in many industrially-valuable synthetic processes.

Montréalone is a mesoionic heterocyclic chemical compound. It is named for the city of Montréal, Canada, which is the location of McGill University, where it was first discovered.

An organic azide is an organic compound that contains an azide functional group. Because of the hazards associated with their use, few azides are used commercially although they exhibit interesting reactivity for researchers. Low molecular weight azides are considered especially hazardous and are avoided. In the research laboratory, azides are precursors to amines. They are also popular for their participation in the "click reaction" between an azide and an alkyne and in Staudinger ligation. These two reactions are generally quite reliable, lending themselves to combinatorial chemistry.

References

  1. Bertrand, Guy; Wentrup, Curt (17 March 1994). "Nitrile Imines: From Matrix Characterization to Stable Compounds". Angewandte Chemie International Edition in English. 33 (5): 527–545. doi:10.1002/anie.199405271.
  2. Huisgen, Rolf (October 1963). "1,3-Dipolar Cycloadditions. Past and Future". Angewandte Chemie International Edition in English. 2 (10): 565–598. doi:10.1002/anie.196305651.
  3. 1 2 Huisgen, Rolf (November 1963). "Kinetics and Mechanism of 1,3-Dipolar Cycloadditions". Angewandte Chemie International Edition . 2 (11): 633–645. doi:10.1002/anie.196306331.
  4. Firestone, R (1968). "Mechanism of 1,3-dipolar cycloadditions". Journal of Organic Chemistry . 33 (6): 2285–2290. doi:10.1021/jo01270a023.
  5. Huisgen, Rolf (1976). "1,3-Dipolar cycloadditions. 76. Concerted nature of 1,3-dipolar cycloadditions and the question of diradical intermediates". Journal of Organic Chemistry . 41 (3): 403–419. doi:10.1021/jo00865a001.
  6. Mloston, G.; Langhals, E.; Huisgen, Rolf (1986). "First Two-Step 1,2-Dipolar Cycloadditons: Nonstereospecificity". J. Am. Chem. Soc. 108 (20): 6401–66402. doi:10.1021/ja00280a053.
  7. Seyyed Amir, Siadati (2015). "An example of a stepwise mechanism for the catalyst-free 1,3-dipolar cycloaddition between a nitrile oxide and an electron rich alkene". Tetrahedron Letters . 56 (34): 4857–4863. doi:10.1016/j.tetlet.2015.06.048.
  8. Huisgen, Rolf (1963). "1,3-Dipolar Cycloadditions. Past and Future". Angewandte Chemie International Edition . 2 (10): 565–598. doi:10.1002/anie.196305651.
  9. Cox, A; Thomas, L; Sheridan, J (1958). "Microwave Spectra of Diazomethane and its Deutero Derivatives". Nature . 181 (4614): 1000–1001. Bibcode:1958Natur.181.1000C. doi:10.1038/1811000a0. S2CID   4245746.
  10. Hilberty, P; Leforestier, C (1978). "Expansion of molecular orbital wave functions into valence bond wave functions. A simplified procedure". Journal of the American Chemical Society . 100 (7): 2012–2017. doi:10.1021/ja00475a007.
  11. McGarrity, J.F.; Patai, Saul (1978). "Basicity, acidity and hydrogen bonding". Diazonium and Diazo Groups. Vol. 1. pp. 179–230. doi:10.1002/9780470771549.ch6. ISBN   9780470771549.
  12. Berner, Daniel; McGarrity, John (1979). "Direct observation of the methyldiazonium ion in fluorosulfuric acid". Journal of the American Chemical Society . 101 (11): 3135–3136. doi:10.1021/ja00505a059.
  13. Muller, Eugen; Rundel, Wolfgans (1956). "Untersuchungen an Diazomethanen, VI. Mitteil.: Umsetzung von Diazoäthan mit Methyllithium". Chemische Berichte . 89 (4): 1065–1071. doi:10.1002/cber.19560890436.
  14. Geittner, Jochen; Huisgen, Rolph; Reissig, Hans-Ulrich (1978). "Solvent Dependence of Cycloaddition Rates of Phenyldiazomethane and Activation Parameters". Heterocycles. 11: 109–120. doi:10.3987/S(N)-1978-01-0109 (inactive 7 March 2024).{{cite journal}}: CS1 maint: DOI inactive as of March 2024 (link)
  15. Huisgen, Rolph; Reissig, Hans-Ulrich; Huber, Helmut; Voss, Sabine (1979). "α-Diazocarbonyl compounds and enamines - a dichotomy of reaction paths". Tetrahedron Letters. 20 (32): 2987–2990. doi:10.1016/S0040-4039(00)70991-9.
  16. Sustmann, R (1974). "Orbital energy control of cycloaddition reactivity". Pure and Applied Chemistry . 40 (4): 569–593. doi: 10.1351/pac197440040569 .
  17. Geittner, Jochen; Huisgen, Rolf (1977). "Kinetics of 1,3-dipolar cycloaddition reactions of diazomethane; A correlation with homo-lumo energies". Tetrahedron Letters. 18 (10): 881–884. doi:10.1016/S0040-4039(01)92781-9.
  18. Huisgen, Rolf; Szeimies, Gunter; Mobius, Leander (1967). "K1.3-Dipolare Cycloadditionen, XXXII. Kinetik der Additionen organischer Azide an CC-Mehrfachbindungen". Chemische Berichte. 100 (8): 2494–2507. doi:10.1002/cber.19671000806.
  19. Williamson, D. G.; Cvetanovic, R. J. (1968). "Rates of ozone-olefin reactions in carbon tetrachloride solutions". Journal of the American Chemical Society. 90 (14): 3668–3672. doi:10.1021/ja01016a011.
  20. Bihlmaier, Werner; Geittner, Jochen; Huisgen, Rolf; ReissigP, Hans-Ulrich (1978). "The Stereospecificity of Diazomethane Cycloadditions". Heterocycles . 10: 147–152. doi:10.3987/S-1978-01-0147 (inactive 7 March 2024).{{cite journal}}: CS1 maint: DOI inactive as of March 2024 (link)
  21. Huisgen, Rolf; Scheer, Wolfgang; Huber, Helmut (1967). "Stereospecific Conversion of cis-trans Isomeric Aziridines to Open-Chain Azomethine Ylides". Journal of the American Chemical Society . 89 (7): 1753–1755. doi:10.1021/ja00983a052.
  22. Dahmen, Alexander; Hamberger, Helmut; Huisgen, Rolf; Markowski, Volker (1971). "Conrotatory ring opening of cyanostilbene oxides to carbonyl ylides". Journal of the Chemical Society D: Chemical Communications (19): 1192–1194. doi:10.1039/C29710001192.
  23. Padwa, Albert; Smolanoff, Joel (1971). "Photocycloaddition of arylazirenes with electron-deficient olefins". Journal of the American Chemical Society. 93 (2): 548–550. doi:10.1021/ja00731a056.
  24. Iwashita, Takashi; Kusumi, Takenori; Kakisawa, Hiroshi (1979). "A Synthesis of dl-isoretronecanol". Chemistry Letters. 8 (11): 1337–1340. doi:10.1246/cl.1979.1337.
  25. Wang, Chia-Lin; Ripka, William; Confalone, Pat (1984). "A short and stereospecific synthesis of (±)-α-lycorane". Tetrahedron Letters. 25 (41): 4613–4616. doi:10.1016/S0040-4039(01)91213-4.
  26. Kanemasa, Shuji (2002). "Metal-Assisted Stereocontrol of 1,3-Dipolar Cycloaddition Reactions". Synlett. 2002 (9): 1371–1387. doi:10.1055/s-2002-33506.
  27. Bode, Jeffrey; Carreira, Erick (2011). "Stereoselective Syntheses of Epothilones A and B via Directed Nitrile Oxide Cycloaddition". Journal of the American Chemical Society. 123 (15): 3611–3612. doi:10.1021/ja0155635. PMID   11472140.
  28. Vsevolod V. Rostovtsev; Luke G. Green; Valery V. Fokin; K. Barry Sharpless (2002). "A Stepwise Huisgen Cycloaddition Process: Copper(I)-Catalyzed Regioselective Ligation of Azides and Terminal Alkynes". Angewandte Chemie International Edition . 41 (14): 2596–22599. doi:10.1002/1521-3773(20020715)41:14<2596::AID-ANIE2596>3.0.CO;2-4. PMID   12203546.
  29. Caramella, Pierluigi; Houk, K.N. (1976). "Geometries of nitrilium betaines. The clarification of apparently anomalous reactions of 1,3-dipoles". Journal of the American Chemical Society. 98 (20): 6397–6399. doi:10.1021/ja00436a062.
  30. Caramella, Pierluigi; Gandour, Ruth W.; Hall, Janet A.; Deville, Cynthia G.; Houk, K. N. (1977). "A derivation of the shapes and energies of the molecular orbitals of 1,3-dipoles. Geometry optimizations of these species by MINDO/2 and MINDO/3". Journal of the American Chemical Society. 99 (2): 385–392. doi:10.1021/ja00444a013.
  31. Huisgen, Rolf (November 1963). "Kinetics and Mechanism of 1,3-Dipolar Cycloadditions". Angewandte Chemie International Edition . 2 (11): 633–645. doi:10.1002/anie.196306331.
  32. Padwa, Albert (1983). 1,3-Dipolar Cycloaddition Chemistry. General Heterocyclic Chemistry Series. Vol. 1. United States of America: Wiley-Interscience. pp. 141–145. ISBN   978-0-471-08364-1.
  33. Koszinowski, J. (1980). thesis (PhD Thesis).
  34. Evans, David; Ripin, David; Halstead, David; Campos, Kevin (1999). "Synthesis and Absolute Stereochemical Assignment of (+)-Miyakolide". Journal of the American Chemical Society . 121 (29): 6816–6826. doi:10.1021/ja990789h.
  35. Synthetic Reactions of M=C and M=N Bonds: Ylide Formation, Rearrangement, and 1,3-Dipolar Cycloaddition; Hiyama, T. W., J., Ed.; Elsevier, 2007; Vol. 11.
  36. Padwa, Albert.; Hornbuckle, Susan F. (1991). "Ylide formation from the reaction of carbenes and carbenoids with heteroatom lone pairs". Chemical Reviews. 91 (3): 263–309. doi:10.1021/cr00003a001.
  37. 1 2 Janulis, Eugene P.; Arduengo, Anthony J. (1983). "Structure of an electronically stabilized carbonyl ylide". Journal of the American Chemical Society. 105 (18): 5929–5930. doi:10.1021/ja00356a044.
  38. Prakash, G. K. S.; Ellis, R. W.; Felberg, J. D.; Olah, G. A. Formaldehyde 0-Methylide, [CH2=O+-CH2: The Parent Carbonyl Ylide] J Am Chem Soc 1986, 108, 1341.
  39. Sammes, P. G.; Street, L. J. Intra molecular Cyclo additions with Oxido pyrylium Ylides J. Chem. Soc., Chem. Commun. 1982, 1056.
  40. Garst, M. E.; McBride, B. J.; Douglass III, J. G. Intramolecular cycloadditions with 2-(ω-alkenyl)-5-hydroxy-4-pyrones Tetrahedron Lett. 1983, 24, 1675.
  41. Gisch, John F.; Landgrebe, John A. (1985). "Dichlorocarbene from flash vacuum pyrolysis of trimethyl(trichloromethyl)silane. Possible observation of 1,1-dichloro-3-phenylcarbonyl ylide". The Journal of Organic Chemistry. 50 (12): 2050–2054. doi:10.1021/jo00212a009.
  42. Huan, Zhenwei; Landgrebe, John A.; Peterson, Kimberly (1983). "Dibromocarbonyl ylides. Deoxygenation of aldehydes and ketones by dibromocarbene". The Journal of Organic Chemistry. 48 (24): 4519–4523. doi:10.1021/jo00172a015.
  43. Martin, Charles W.; Lund, Paul R.; Rapp, Erich; Landgrebe, John A. (1978). "Halogenated carbonyl ylides in the reactions of mercurial dihalocarbene precursors with substituted benzaldehydes". The Journal of Organic Chemistry. 43 (6): 1071–1076. doi:10.1021/jo00400a009.
  44. Hodgson, D. M.; Bruckl, T.; Glen, R.; Labande, A. H.; Selden, D. A.; Dossetter, A. G.; Redgrave, A. J. Catalytic enantioselective intermolecular cycloadditions of 2-diazo-3,6-diketoester-derived carbonyl ylides with alkene dipolarophiles Proceedings of the National Academy of Sciences of the United States of America 2004, 101, 5450.
  45. Padwa, Albert; Hertzog, Donald L.; Nadler, William R. (1994). "Intramolecular Cycloaddition of Isomunchnone Dipoles to Heteroaromatic .pi.-Systems". The Journal of Organic Chemistry. 59 (23): 7072–7084. doi:10.1021/jo00102a037.
  46. Hamaguchi, M.; Ibata, T. New Type of Mesoionic System. 1,3-Dipolar Cycloaddition of Isomunchnon With Ethylenic Compounds Chem Lett 1975, 499.
  47. Park, Soon-Bong; Sakata, Naoya; Nishiyama, Hisao (1996). "Aryloxycarbonylcarbene Complexes of Bis(oxazolinyl)pyridineruthenium as Active Intermediates in Asymmetric Catalytic Cyclopropanations". Chemistry - A European Journal. 2 (3): 303–306. doi:10.1002/chem.19960020311.
  48. Snyder, James P.; Padwa, Albert; Stengel, Thomas; Arduengo, Anthony J.; Jockisch, Alexander; Kim, Hyo-Joong (2001). "A Stable Dirhodium Tetracarboxylate Carbenoid: Crystal Structure, Bonding Analysis, and Catalysis". Journal of the American Chemical Society. 123 (45): 11318–11319. doi:10.1021/ja016928o. PMID   11697986.
  49. 1 2 Hodgson, D. M.; Pierard, F. Y. T. M.; Stupple, P. A. Catalytic enantioselective rearrangements and cycloadditions involving ylides from diazo compounds Chem Soc Rev 2001, 30, 50.
  50. Yoshikai, Naohiko; Nakamura, Eiichi (2003). "Theoretical Studies on Diastereo- and Enantioselective Rhodium-Catalyzed Cyclization of Diazo Compoundvia Intramolecular C—H Bond Insertion". Advanced Synthesis & Catalysis. 345 (910): 1159–1171. doi:10.1002/adsc.200303092.
  51. Nakamura, Eiichi; Yoshikai, Naohiko; Yamanaka, Masahiro (2002). "Mechanism of C−H Bond Activation/C−C Bond Formation Reaction between Diazo Compound and Alkane Catalyzed by Dirhodium Tetracarboxylate". Journal of the American Chemical Society. 124 (24): 7181–7192. doi:10.1021/ja017823o. PMID   12059244.
  52. Costantino, G.; Rovito, R.; Macchiarulo, A.; Pellicciari, R. Structure of metal–carbenoid intermediates derived from the dirhodium(II)tetracarboxylate mediated decomposition of α-diazocarbonyl compounds: a DFT study J Mol Struc-Theochem 2002, 581, 111.
  53. 1 2 3 4 M. Hodgson, D.; H. Labande, A.; Muthusamy, S. In Organic Reactions; John Wiley & Sons, Inc.: 2004.
  54. Suga, Hiroyuki; Ebiura, Yasutaka; Fukushima, Kazuaki; Kakehi, Akikazu; Baba, Toshihide (2005). "Efficient Catalytic Effects of Lewis Acids in the 1,3-Dipolar Cycloaddition Reactions of Carbonyl Ylides with Imines". The Journal of Organic Chemistry. 70 (26): 10782–10791. doi:10.1021/jo051743b. PMID   16356001.
  55. 1 2 Padwa, Albert; Fryxell, Glen E.; Zhi, Lin (1990). "Tandem cyclization-cycloaddition reaction of rhodium carbenoids. Scope and mechanistic details of the process". Journal of the American Chemical Society. 112 (8): 3100–3109. doi:10.1021/ja00164a034.
  56. Houk, K. N.; Sims, Joyner.; Duke, R. E.; Strozier, R. W.; George, John K. (1973). "Frontier molecular orbitals of 1,3 dipoles and dipolarophiles". Journal of the American Chemical Society. 95 (22): 7287–7301. doi:10.1021/ja00803a017.
  57. Houk, K. N.; Rondan, Nelson G.; Santiago, Cielo; Gallo, Catherine J.; Gandour, Ruth Wells; Griffin, Gary W. (1980). "Theoretical studies of the structures and reactions of substituted carbonyl ylides". Journal of the American Chemical Society. 102 (5): 1504–1512. doi:10.1021/ja00525a006.
  58. Padwa, Albert; Weingarten, M. David (1996). "Cascade Processes of Metallo Carbenoids". Chemical Reviews. 96 (1): 223–270. doi:10.1021/cr950022h. PMID   11848752.
  59. Padwa, Albert; Austin, David J. (1996). "Ligand-Induced Selectivity in the Rhodium(II)-Catalyzed Reactions of α-Diazo Carbonyl Compounds†". The Journal of Organic Chemistry. 61: 63–72. doi:10.1021/jo951576n.
  60. Suga, H.; Kakehi, A.; Ito, S.; Inoue, K.; Ishida, H.; Ibata, T. Stereocontrol in a Ytterbium Triflate-Catalyzed 1,3-Dipolar Cycloaddition Reaction of Carbonyl Ylide with N-Substituted Maleimides and Dimethyl Fumarate B Chem Soc Jpn 2001, 74, 1115.
  61. Geng, Zhe; Chen, Bin; Chiu, Pauline (2006). "Total Synthesis of Pseudolaric Acid A". Angewandte Chemie International Edition. 45 (37): 6197–6201. doi:10.1002/anie.200602056. PMID   16906616.
  62. Suga, Hiroyuki; Inoue, Kei; Inoue, Shuichi; Kakehi, Akikazu; Shiro, Motoo (2005). "Chiral 2,6-Bis(oxazolinyl)pyridine−Rare Earth Metal Complexes as Catalysts for Highly Enantioselective 1,3-Dipolar Cycloaddition Reactions of 2-Benzopyrylium-4-olates". The Journal of Organic Chemistry. 70 (1): 47–56. doi:10.1021/jo049007f. PMID   15624905.
  63. Onishi, Tomoyuki; Sebahar, Paul; Williams, Robert (2003). "Concise, Asymmetric Total Synthesis of Spirotryprostatin A". Organic Letters. 5 (17): 3135–3137. doi:10.1021/ol0351910. PMID   12917000.
  64. Tornoe, Christian; Christensen, Caspar; Meldal, Morten (2002). "Peptidotriazoles on Solid Phase: [1,2,3]-Triazoles by Regiospecific Copper(I)-Catalyzed 1,3-Dipolar Cycloadditions of Terminal Alkynes to Azides". Journal of Organic Chemistry. 67 (9): 3057–3064. doi:10.1021/jo011148j. PMID   11975567. S2CID   11957672.
  65. Rostovtsev, Vsevolod; Green, Luke; Fokin, Valery; Sharpless, Barry K. (2002). "A Stepwise Huisgen Cycloaddition Process: Copper(I)-Catalyzed Regioselective Ligation of Azides and Terminal Alkynes". Angewandte Chemie International Edition. 41 (14): 2596–2599. doi:10.1002/1521-3773(20020715)41:14<2596::AID-ANIE2596>3.0.CO;2-4. PMID   12203546.
  66. Besanceney-Webler, Christen; Jiang, Hao; Zheng, Tianqing; Feng, Lei; Soriano del Amo, David; Wang, Wei; Klivansky, Liana M.; Marlow, Florence L.; Liu, Yi; Wu, Peng (2011). "Increasing the Efficacy of Bioorthogonal Click Reactions for Bioconjugation: A Comparative Study". Angewandte Chemie International Edition. 50 (35): 8051–8056. doi:10.1002/anie.201101817. PMC   3465470 . PMID   21761519.
  67. Breidenbach, Mark; Gallagher, Jennifer; King, David; Smart, Brian; Wu, Peng; Bertozzi, Carolyn (2010). "Targeted metabolic labeling of yeast N-glycans with unnatural sugars". Proceedings of the National Academy of Sciences of the United States of America. 107 (9): 3988–3993. Bibcode:2010PNAS..107.3988B. doi: 10.1073/pnas.0911247107 . PMC   2840165 . PMID   20142501.
  68. Agard, Nicholas; Prescher, Jennifer; Bertozzi, Carolyn (2004). "A Strain-Promoted [3 + 2] Azide−Alkyne Cycloaddition for Covalent Modification of Biomolecules in Living Systems". Journal of the American Chemical Society. 126 (46): 15046–15047. doi:10.1021/ja044996f. PMID   15547999.
  69. Fernandez-Suarez, Marta; Baruah, Hemanta; Martinez-Hernandez, Laura; Xie, Kathleen; Baskin, Jeremy; Bertozzi, Carolyn; Ting, Alice (2007). "Redirecting lipoic acid ligase for cell surface protein labeling with small-molecule probes". Nature Biotechnology. 25 (12): 1483–1487. doi:10.1038/nbt1355. PMC   2654346 . PMID   18059260.