Cheletropic reaction

Last updated
Example cheletropic reactions:
Case 1: the single atom is the carbonyl carbon (
.mw-parser-output .template-chem2-su{display:inline-block;font-size:80%;line-height:1;vertical-align:-0.35em}.mw-parser-output .template-chem2-su>span{display:block;text-align:left}.mw-parser-output sub.template-chem2-sub{font-size:80%;vertical-align:-0.35em}.mw-parser-output sup.template-chem2-sup{font-size:80%;vertical-align:0.65em}
C=O) that ends up in carbon monoxide (
C[?]O).
Case 2: the single atom is the nitrogen atom in the diazenyl group (
N=N), which ends up as dinitrogen (
N[?]N).
The above are known as cheletropic eliminations because a small, stable molecule is given off in the reaction.
Case 3 & 4: the single atom is the sulfur in sulfur dioxide (
SO2), which joins the alkene chains to form a ring. Introduction.png
Example cheletropic reactions:
Case 1: the single atom is the carbonyl carbon (C=O) that ends up in carbon monoxide (C≡O).
Case 2: the single atom is the nitrogen atom in the diazenyl group (N=N), which ends up as dinitrogen (N≡N).
The above are known as cheletropic eliminations because a small, stable molecule is given off in the reaction.
Case 3 & 4: the single atom is the sulfur in sulfur dioxide (SO2), which joins the alkene chains to form a ring.

In organic chemistry, cheletropic reactions, also known as chelotropic reactions, [2] are a type of pericyclic reaction (a chemical reaction that involves a transition state with a cyclic array of atoms and an associated cyclic array of interacting orbitals). [1] Specifically, cheletropic reactions are a subclass of cycloadditions. The key distinguishing feature of cheletropic reactions is that on one of the reagents, both new bonds are being made to the same atom. [3]

Contents

Theoretical analysis

In the pericyclic transition state, a small molecule donates two electrons to the ring. The reaction process can be shown using two different geometries, the small molecule can approach in a linear or non-linear fashion. In the linear approach, the electrons in the orbital of the small molecule are pointed directly at the π-system. In the non-linear approach, the orbital approaches at a skew angle. The π-system's ability to rotate as the small molecule approaches is crucial in forming new bonds. The direction of rotation will be different depending on how many π-electrons are in the system. Shown below is a diagram of a two-electron fragment approaching a four-electron π-system using frontier molecular orbitals. The rotation will be disrotatory if the small molecule approaches linearly and conrotatory if the molecule approaches non-linearly. Disrotatory and conrotatory are sophisticated terms expressing how the bonds in the π-system are rotating. Disrotatory means opposite directions while conrotatory means the same direction. This is also depicted in the diagram below.

Using Hückel's rule, one can tell if the π-system is aromatic or antiaromatic. If aromatic, linear approaches use disrotatory motion while non-linear approaches use conrotatory motion. The opposite goes with an anti-aromatic system. Linear approaches will have conrotatory motion while non-linear approaches will have disrotatory motion. [1]

Cheletropic reactions involving SO2

Thermodynamics

In 1995, Suarez and Sordo showed that sulfur dioxide when reacted with butadiene and isoprene gives two different products depending on the mechanism. This was shown experimentally and using ab initio calculations. A kinetic and thermodynamic product are both possible, but the thermodynamic product is more favorable. The kinetic product arises from a Diels–Alder reaction, while a cheletropic reaction gives rise to a more thermodynamically stable product. The cheletropic pathway is favored because it gives rise to a more stable five-membered ring adduct. The scheme below shows the difference between the two products, the path to the right shows the more stable thermodynamic product, while the path to the left shows the kinetic product. [4]

Kinetics

The cheletropic reactions of 1,3-dienes with sulfur dioxide have been extensively investigated in terms of kinetics (see above for general reaction).

In the first quantitative measurement of kinetic parameters for this reaction, a 1976 study by Isaacs and Laila measured the rates of addition of sulfur dioxide to butadiene derivatives. Rates of addition were monitored in benzene at 30 °C with an initial twentyfold excess of sulfur dioxide, allowing for a pseudo first-order approximation. The disappearance of SO2 was followed spectrophotometrically at 320 nm. The reaction showed pseudo first-order kinetics. Some interesting results were that electron-withdrawing groups on the diene decreased the rate of reaction. Also, the reaction rate was affected considerably by steric effects of 2-substituents, with more bulky groups increasing the rate of reaction. The authors attribute this to the tendency of bulky groups to favor the cisoid conformation of the diene which is essential to the reaction (see table below). In addition, the rates at four temperatures were measured for seven of the dienes permitting calculations of the enthalpy of activation (ΔH) and entropy of activation (ΔS) for these reactions through the Arrhenius equation. [5]

-Butadiene104 k /min−1 (30 °C) (± 1-2%) absolute104 k /min−1 (30 °C) (± 1-2%) relativeΔH /kcal mol−1ΔS /cal mol−1 K−1
2-methyl1.831.0014.9-15
2-ethyl4.762.6010.6-20
2-isopropyl13.07.3812.5-17
2-tert-butyl38.220.810.0-19
2-neopentyl17.29.411.6-18
2-chloro0.240.13N/AN/A
2-bromoethyl0.720.39N/AN/A
2-p-tolyl24.713.510.4-19
2-phenyl17.39.45N/AN/A
2-(p-bromophenyl)9.074.96N/AN/A
2,3-dimethyl3.541.9312.3-18
cis-1-methyl0.180.10N/AN/A
trans-1-methyl0.690.38N/AN/A
1,2-dimethylene-cyclohexane24.713.511.4-16
2-methyl-1,1,4,4-d41.96N/AN/AN/A

More recently, a 2002 study by Monnat, Vogel, and Sordo measured the kinetics of addition of sulfur dioxide to 1,2-dimethylidenecycloalkanes. An interesting point presented in this paper is that the reaction of 1,2-dimethylidenecyclohexane with sulfur dioxide can give two different products depending on reaction conditions. The reaction produces the corresponding sulfine through a hetero-Diels–Alder reaction under kinetic control (≤ -60 °C), but, under thermodynamic control (≥ -40 °C), the reaction produces the corresponding sulfolene through a cheletropic reaction. The activation enthalpy for the hetero-Diels–Alder reaction is about 8 kJ/mol smaller than that for the corresponding cheletropic reaction. The sulfolene is about 40 kJ/mol more stable than the isometric sulfine in CH2Cl2/SO2 solution. [6]

Reaction of 1,2-dimethylidenecyclohexane with SO2 gives a sultine through a hetero-Diels-Alder reaction under kinetic control or a sulfolene through a cheletropic reaction under thermodynamic control Sulfolene vs sultine.png
Reaction of 1,2-dimethylidenecyclohexane with SO2 gives a sultine through a hetero-Diels-Alder reaction under kinetic control or a sulfolene through a cheletropic reaction under thermodynamic control

The authors were able to experimentally determine a rate law at 261.2 K for the reaction of 1,2-dimethylidenecyclohexane with sulfur dioxide to give the corresponding sulfolene. The reaction was first order in 1,2-dimethylidenecyclohexane but second order in sulfur dioxide (see below). This confirmed a prediction based on high-level ab initio quantum calculations. Using computational methods, the authors proposed a transition structure for the cheletropic reaction of 1,2-dimethylidenecyclohexane with sulfur dioxide (see figure at right). [6] The reaction is second order in sulfur dioxide because another molecule of sulfur dioxide likely binds to the transition state to help stabilize it. [7] Similar results were found in a 1995 study by Suarez, Sordo, and Sordo which used ab initio calculations to study the kinetic and thermodynamic control of the reaction of sulfur dioxide with 1,3-dienes. [4]

Proposed transition state for reaction of 1,2-dimethylidenecyclohexane with SO2 to give a sulfolene through a cheletropic reaction Sulfolene ts2.png
Proposed transition state for reaction of 1,2-dimethylidenecyclohexane with SO2 to give a sulfolene through a cheletropic reaction

Solvent effects

Cheletropic reaction studied in various solvents Solvent Effects.png
Cheletropic reaction studied in various solvents

The effect of the solvent of the cheletropic reaction of 3,4-dimethyl-2,5-dihydrothiophen-1,1-dioxide (shown at right) was kinetically investigated in 14 solvents. The reaction rate constants of the forward and reverse reaction in addition to the equilibrium constants were found to be linearly correlated with the ET(30) solvent polarity scale.

Reactions were done at 120 °C and were studied by 1H-NMR spectroscopy of the reaction mixture. The forward rate k1 was found to decrease by a factor of 4.5 going from cyclohexane to methanol. The reverse rate k−1 was found to increase by a factor of 53 going from cyclohexane to methanol, while the equilibrium constant Keq decreased by a factor of 140. It is suggested that there is a change of the polarity during the activation process as evidenced by correlations between the equilibrium and kinetic data. The authors remark that the reaction appears to be influenced by the polarity of the solvent, and this can be explained by the change in the dipole moments when going from reactant to transition state to product. The authors also state that the cheletropic reaction doesn’t seem to be influenced by either solvent acidity or basicity.

The results of this study lead the authors to expect the following behaviors:

1. The change in the solvent polarity will influence the rate less than the equilibrium.

2. The rate constants will be characterized by opposite effect on the polarity: k1 will slightly decrease with the increase of ET(30), and k−1 will increase under the same conditions.

3. The effect on k−1 will be larger than on k1. [8]

Carbene additions to alkenes

Addition of a carbene to an alkene to form a cyclopropane Carbene addition to alkene.png
Addition of a carbene to an alkene to form a cyclopropane

One of the most synthetically important cheletropic reactions is the addition of a singlet carbene to an alkene to make a cyclopropane (see figure at left). [1] A carbene is a neutral molecule containing a divalent carbon with six electrons in its valence shell. Due to this, carbenes are highly reactive electrophiles and generated as reaction intermediates. [9] A singlet carbene contains an empty p orbital and a roughly sp2 hybrid orbital that has two electrons. Singlet carbenes add stereospecifically to alkenes, and alkene stereochemistry is retained in the cyclopropane product. [1] The mechanism for addition of a carbene to an alkene is a concerted [2+1] cycloaddition (see figure). Carbenes derived from chloroform or bromoform can be used to add CX2 to an alkene to give a dihalocyclopropane, while the Simmons–Smith reagent adds CH2. [10]

A) The orbitals for singlet carbenes B) Non-linear approach of a) carbene sp orbital and b) carbene p orbital Carbene alkene orbitals.png
A) The orbitals for singlet carbenes B) Non-linear approach of a) carbene sp orbital and b) carbene p orbital

Interaction of the filled carbene orbital with the alkene π system creates a four-electron system and favors a non-linear approach. It is also favorable to mix the carbene empty p orbital with the filled alkene π orbital. Favorable mixing occurs through a non-linear approach (see figure at right). However, while theory clearly favors a non-linear approach, there are no obvious experimental implications for a linear vs. non-linear approach. [1]

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.

In organic chemistry, an electrocyclic reaction is a type of pericyclic rearrangement where the net result is one pi bond being converted into one sigma bond or vice versa. These reactions are usually categorized by the following criteria:

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.

<span class="mw-page-title-main">Ene reaction</span> Reaction in organic chemistry

In organic chemistry, the ene reaction is a chemical reaction between an alkene with an allylic hydrogen and a compound containing a multiple bond, in order to form a new σ-bond with migration of the ene double bond and 1,5 hydrogen shift. The product is a substituted alkene with the double bond shifted to the allylic position.

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. Hence, the reaction is sometimes referred to as the Huisgen cycloaddition. 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.

In organic chemistry, a rearrangement reaction is a broad class of organic reactions where the carbon skeleton of a molecule is rearranged to give a structural isomer of the original molecule. Often a substituent moves from one atom to another atom in the same molecule, hence these reactions are usually intramolecular. In the example below, the substituent R moves from carbon atom 1 to carbon atom 2:

<span class="mw-page-title-main">Thermodynamic versus kinetic reaction control</span>

Thermodynamic reaction control or kinetic reaction control in a chemical reaction can decide the composition in a reaction product mixture when competing pathways lead to different products and the reaction conditions influence the selectivity or stereoselectivity. The distinction is relevant when product A forms faster than product B because the activation energy for product A is lower than that for product B, yet product B is more stable. In such a case A is the kinetic product and is favoured under kinetic control and B is the thermodynamic product and is favoured under thermodynamic control.

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

<span class="mw-page-title-main">Woodward–Hoffmann rules</span> Set of rules pertaining to pericyclic reactions

The Woodward–Hoffmann rules are a set of rules devised by Robert Burns Woodward and Roald Hoffmann to rationalize or predict certain aspects of the stereochemistry and activation energy of pericyclic reactions, an important class of reactions in organic chemistry. The rules originate in certain symmetries of the molecule's orbital structure that any molecular Hamiltonian conserves. Consequently, any symmetry-violating reaction must couple extensively to the environment; this imposes an energy barrier on its occurrence, and such reactions are called symmetry-forbidden. Their opposites are symmetry-allowed.

In organic chemistry, neighbouring group participation has been defined by the International Union of Pure and Applied Chemistry (IUPAC) as the interaction of a reaction centre with a lone pair of electrons in an atom or the electrons present in a sigma or pi bond contained within the parent molecule but not conjugated with the reaction centre. When NGP is in operation it is normal for the reaction rate to be increased. It is also possible for the stereochemistry of the reaction to be abnormal when compared with a normal reaction. While it is possible for neighbouring groups to influence many reactions in organic chemistry this page is limited to neighbouring group effects seen with carbocations and SN2 reactions.

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

Danishefsky's diene is an organosilicon compound and a diene with the formal name trans-1-methoxy-3-trimethylsilyloxy-buta-1,3-diene named after Samuel J. Danishefsky. Because the diene is very electron-rich it is a very reactive reagent in Diels-Alder reactions. This diene reacts rapidly with electrophilic alkenes, such as maleic anhydride. The methoxy group promotes highly regioselective additions. The diene is known to react with amines, aldehydes, alkenes and alkynes. Reactions with imines and nitro-olefins have been reported.

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

Sulfolene, or butadiene sulfone is a cyclic organic chemical with a sulfone functional group. It is a white, odorless, crystalline, indefinitely storable solid, which dissolves in water and many organic solvents. The compound is used as a source of butadiene.

In chemistry, frontier molecular orbital theory is an application of molecular orbital theory describing HOMO–LUMO interactions.

<span class="mw-page-title-main">Germylene</span> Class of germanium (II) compounds

Germylenes are a class of germanium(II) compounds with the general formula :GeR2. They are heavier carbene analogs. However, unlike carbenes, whose ground state can be either singlet or triplet depending on the substituents, germylenes have exclusively a singlet ground state. Unprotected carbene analogs, including germylenes, has a dimerization nature. Free germylenes can be isolated under the stabilization of steric hindrance or electron donation. The synthesis of first stable free dialkyl germylene was reported by Jutzi, et al in 1991.

The retro-Diels–Alder reaction is the reverse of the Diels–Alder (DA) reaction, a [4+2] cycloelimination. It involves the formation of a diene and dienophile from a cyclohexene. It can be accomplished spontaneously with heat, or with acid or base mediation.

The inverse electron demand Diels–Alder reaction, or DAINV or IEDDA is an organic chemical reaction, in which two new chemical bonds and a six-membered ring are formed. It is related to the Diels–Alder reaction, but unlike the Diels–Alder reaction, the DAINV is a cycloaddition between an electron-rich dienophile and an electron-poor diene. During a DAINV reaction, three pi-bonds are broken, and two sigma bonds and one new pi-bond are formed. A prototypical DAINV reaction is shown on the right.

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.

Cycloisomerization is any isomerization in which the cyclic isomer of the substrate is produced in the reaction coordinate. The greatest advantage of cycloisomerization reactions is its atom economical nature, by design nothing is wasted, as every atom in the starting material is present in the product. In most cases these reactions are mediated by a transition metal catalyst, in few cases organocatalysts and rarely do they occur under thermal conditions. These cyclizations are able to be performed with excellent levels of selectivity in numerous cases and have transformed cycloisomerization into a powerful tool for unique and complex molecular construction. Cycloisomerization is a very broad topic in organic synthesis and many reactions that would be categorized as such exist. Two basic classes of these reactions are intramolecular Michael addition and Intramolecular Diels–Alder reactions. Under the umbrella of cycloisomerization, enyne and related olefin cycloisomerizations are the most widely used and studied reactions.

A Fischer carbene is a type of transition metal carbene complex, which is an organometallic compound containing a divalent organic ligand. In a Fischer carbene, the carbene ligand is a σ-donor π-acceptor ligand. Because π-backdonation from the metal centre is generally weak, the carbene carbon is electrophilic.

<span class="mw-page-title-main">Cyclopentadienylcobalt dinitrosyl</span> Organometallic molecule

Cyclopentadienylcobalt dinitrosyl is an organometallic molecule. It is a reactive intermediate in the formation of dinitrosoalkane cobalt complexes. While cyclopentadienylcobalt dinitrosyl has not been isolated and characterized, the preparation of this reactive intermediate in the presence of olefins results in the isolable dinitrosoalkane cobalt complexes. The dinitrosyl intermediate is known for its alkene binding capability. The resulting dinitrosoalkane cobalt complexes are capable of stoichiometric and catalytic C-H bond functionalization.

References

  1. 1 2 3 4 5 6 Eric V. Anslyn and Dennis A. Dougherty Modern Physical Organic Chemistry University Science Books, 2006.
  2. Chelotropic reaction IUPAC GoldBook
  3. Ian Fleming. Frontier Orbitals and Organic Chemistry Reactions. Wiley, 1976.
  4. 1 2 Suarez, D.; Sordo, T. L.; Sordo, J. A. (1995). "A Comparative Analysis of the Mechanisms of Cheletropic and Diels-Alder Reactions of 1,3-Dienes with Sulfur Dioxide: Kinetic and Thermodynamic Controls". J. Org. Chem. 60 (9): 2848–2852. doi:10.1021/jo00114a039.
  5. Isaacs, N. S.; Laila, A. A. R. (1976). "Rates of addition of sulphur dioxide to some 1,3-dienes". Tetrahedron Lett. 17 (9): 715–716. doi:10.1016/S0040-4039(00)74605-3.
  6. 1 2 Monnat, F.; Vogel, P.; Sordo, J. A. (2002). "Hetero-Diels-Alder and Cheletropic Additions of Sulfur Dioxide to 1,2-Dimethylidenecycloalkanes. Determination of Thermochemical and Kinetics Parameters for Reactions in Solution and Comparison with Estimates From Quantum Calculations". Helv. Chim. Acta . 85 (3): 712–732. doi:10.1002/1522-2675(200203)85:3<712::AID-HLCA712>3.0.CO;2-5.
  7. Fernandez, T.; Sordo, J. A.; Monnat, F.; Deguin, B.; Vogel, P. (1998). "Sulfur Dioxide Promotes Its Hetero-Diels−Alder and Cheletropic Additions to 1,2-Dimethylidenecyclohexane". J. Am. Chem. Soc. 120 (50): 13276–13277. doi:10.1021/ja982565p.
  8. Desimoni, G.; Faita, G.; Garau, S.; Righetti, P. (1996). "Solvent effect in pericyclic reactions. X. The cheletropic reaction". Tetrahedron. 52 (17): 6241–6248. doi:10.1016/0040-4020(96)00279-7.
  9. John McMurry Organic Chemistry, 6th Ed. Thomson, 2004.
  10. Robert B. Grossman The Art of Writing Reasonable Organic Reaction Mechanisms Springer, 2003.