Thermodynamic versus kinetic reaction control

Last updated
Energy profile diagram for kinetic versus thermodynamic product reaction. Thermodyamic versus kinetic control.png
Energy profile diagram for kinetic versus thermodynamic product reaction.

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. [1] [2] [3]

Contents

The conditions of the reaction, such as temperature, pressure, or solvent, affect which reaction pathway may be favored: either the kinetically controlled or the thermodynamically controlled one. Note this is only true if the activation energy of the two pathways differ, with one pathway having a lower Ea (energy of activation) than the other.

Prevalence of thermodynamic or kinetic control determines the final composition of the product when these competing reaction pathways lead to different products. The reaction conditions as mentioned above influence the selectivity of the reaction - i.e., which pathway is taken.

Asymmetric synthesis is a field in which the distinction between kinetic and thermodynamic control is especially important. Because pairs of enantiomers have, for all intents and purposes, the same Gibbs free energy, thermodynamic control will produce a racemic mixture by necessity. Thus, any catalytic reaction that provides product with nonzero enantiomeric excess is under at least partial kinetic control. (In many stoichiometric asymmetric transformations, the enantiomeric products are actually formed as a complex with the chirality source before the workup stage of the reaction, technically making the reaction a diastereoselective one. Although such reactions are still usually kinetically controlled, thermodynamic control is at least possible, in principle.)

Scope

In Diels–Alder reactions

The Diels–Alder reaction of cyclopentadiene with furan can produce two isomeric products. At room temperature, kinetic reaction control prevails and the less stable endo isomer 2 is the main reaction product. At 81 °C and after long reaction times, the chemical equilibrium can assert itself and the thermodynamically more stable exo isomer 1 is formed. [4] The exo product is more stable by virtue of a lower degree of steric congestion, while the endo product is favoured by orbital overlap in the transition state.

Thermodynamic versus kinetic reaction control in reaction of cyclopanetdiene and furan DielsAlderCyclopentadieneFuran.svg
Thermodynamic versus kinetic reaction control in reaction of cyclopanetdiene and furan

An outstanding and very rare example of the full kinetic and thermodynamic reaction control in the process of the tandem inter-/intramolecular Diels–Alder reaction of bis-furyl dienes 3 with hexafluoro-2-butyne or dimethyl acetylenedicarboxylate (DMAD) have been discovered and described in 2018. [5] [6] At low temperature, the reactions occur chemoselectively leading exclusively to adducts of pincer-[4+2] cycloaddition (5). The exclusive formation of domino-adducts (6) is observed at elevated temperatures.

Kinetic and thermodynamic control of the tandem Diels-Alder reaction. Kin therm control diels-alder.svg
Kinetic and thermodynamic control of the tandem Diels–Alder reaction.

Theoretical DFT calculations of the reaction between hexafluoro-2-butyne and dienes 3a-c were performed. The reaction starting with [4+2] cycloaddition of CF3C≡CCF3 at one of the furan moieties occurs in a concerted fashion via TS1 and represents the rate limiting step of the whole process with the activation barrier ΔG≈ 23.1–26.8 kcal/mol.

Gibbs free energy profile for the reaction between bis-dienes 3a-c and hexafluoro-2-butyne. Relative energies are shown in kcal/mol for X = CH2 (plain text), S (italic) and NC(O)CF3 (bold). DHT-calculations.svg
Gibbs free energy profile for the reaction between bis-dienes 3a-c and hexafluoro-2-butyne. Relative energies are shown in kcal/mol for X = CH2 (plain text), S (italic) and NC(O)CF3 (bold).

Further, the reaction could proceed via two competing channels, i.e. either leading to the pincer type products 5viaTS2k or resulting in the formation of the domino product 6viaTS2t. The calculations showed that the first channel is more kinetically favourable (ΔG≈ 5.7–5.9 kcal/mol). Meanwhile, the domino products 6 are more thermodynamically stable than 5G≈ 4.2-4.7 kcal/mol) and this fact may cause isomerization of 5 into 6 at elevated temperature. Indeed, the calculated activation barriers for the 56 isomerization via the retro-Diels–Alder reaction of 5 followed by the intramolecular [4+2]-cycloaddition in the chain intermediate 4 to give 6 are 34.0–34.4 kcal/mol.

In enolate chemistry

In the protonation of an enolate ion, the kinetic product is the enol and the thermodynamic product is a ketone or aldehyde. Carbonyl compounds and their enols interchange rapidly by proton transfers catalyzed by acids or bases, even in trace amounts, in this case mediated by the enolate or the proton source.

In the deprotonation of an unsymmetrical ketone, the kinetic product is the enolate resulting from removal of the most accessible α-H while the thermodynamic product has the more highly substituted enolate moiety. [7] [8] [9] [10] Use of low temperatures and sterically demanding bases increases the kinetic selectivity. Here, the difference in pKb between the base and the enolate is so large that the reaction is essentially irreversible, so the equilibration leading to the thermodynamic product is likely a proton exchange occurring during the addition between the kinetic enolate and as-yet-unreacted ketone. An inverse addition (adding ketone to the base) with rapid mixing would minimize this. The position of the equilibrium will depend on the countercation and solvent.

The kinetic and thermodynamic deprotonation of 2-methylcyclohexanone. Thermodynamic deprotonation methylcyclohexanone.svg
The kinetic and thermodynamic deprotonation of 2-methylcyclohexanone.

If a much weaker base is used, the deprotonation will be incomplete, and there will be an equilibrium between reactants and products. Thermodynamic control is obtained, however the reaction remains incomplete unless the product enolate is trapped, as in the example below. Since H transfers are very fast, the trapping reaction being slower, the ratio of trapped products largely mirrors the deprotonation equilibrium.

The thermodynamic deprotonation of 2-methylcyclohexanone, with trapping of the enolate. Thermodynamic deprotonation methylcyclohexanone enolate trapping.svg
The thermodynamic deprotonation of 2-methylcyclohexanone, with trapping of the enolate.

In electrophilic additions

The electrophilic addition reaction of hydrogen bromide to 1,3-butadiene above room temperature leads predominantly to the thermodynamically more stable 1,4 adduct, 1-bromo-2-butene, but decreasing the reaction temperature to below room temperature favours the kinetic 1,2 adduct, 3-bromo-1-butene. [3]

The addition of HBr to butadiene in ether. Data from Elsheimer (2000). Butadiene hydrogen bromide addition.svg
The addition of HBr to butadiene in ether. Data from Elsheimer (2000).
The rationale for the differing selectivities is as follows: Both products result from Markovnikov protonation at position 1, resulting in a resonance-stabilized allylic cation. The 1,4 adduct places the larger Br atom at a less congested site and includes a more highly substituted alkene moiety, while the 1,2 adduct is the result of the attack by the nucleophile (Br) at the carbon of the allylic cation bearing the greatest positive charge (the more highly substituted carbon is the most likely place for the positive charge).
butadiene hydrobromination mechanism Thermodynamic kinetic control butadiene.svg
butadiene hydrobromination mechanism

Characteristics

    (equation 1)
Unless equilibration is prevented (e.g., by removal of the product from the reaction mixture as soon as it forms), "pure" kinetic control is strictly speaking impossible, because some amount of equilibration will take place before the reactants are entirely consumed. In practice, many systems are well approximated as operating under kinetic control, due to negligibly slow equilibration. For example, many enantioselective catalytic systems provide nearly enantiopure product (> 99% ee), even though the enantiomeric products have the same Gibbs free energy and are equally favored thermodynamically.
    (equation 2)
In principle, "pure" thermodynamic control is also impossible, since equilibrium is only achieved after infinite reaction time. In practice, if A and B interconvert with overall rate constants kf and kr, then for most practical purposes, the change in composition becomes negligible after t ~ 3.5/(kf + kr), or approximately five half-lives, and the system product ratio can be regarded as the result of thermodynamic control.

History

The first to report on the relationship between kinetic and thermodynamic control were R.B. Woodward and Harold Baer in 1944. [18] They were re-investigating a reaction between maleic anhydride and a fulvene first reported in 1929 by Otto Diels and Kurt Alder. [19] They observed that while the endo isomer is formed more rapidly, longer reaction times, as well as relatively elevated temperatures, result in higher exo / endo ratios which had to be considered in the light of the remarkable stability of the exo-compound on the one hand and the very facile dissociation of the endo isomer on the other.

Fulvene maleic anhydride woordward 1944 3.svg

C. K. Ingold with E. D. Hughes and G. Catchpole independently described a thermodynamic and kinetic reaction control model in 1948. [20] They were reinvestigating a certain allylic rearrangement reported in 1930 by Jakob Meisenheimer. [21] Solvolysis of gamma-phenylallyl chloride with AcOK in acetic acid was found to give a mixture of the gamma and the alpha acetate with the latter converting to the first by equilibration. This was interpreted as a case in the field of anionotropy of the phenomenon, familiar in prototropy, of the distinction between kinetic and thermodynamic control in ion-recombination.

Thermo kinetic control ingold 1948.svg

Related Research Articles

In a chemical reaction, chemical equilibrium is the state in which both the reactants and products are present in concentrations which have no further tendency to change with time, so that there is no observable change in the properties of the system. This state results when the forward reaction proceeds at the same rate as the reverse reaction. The reaction rates of the forward and backward reactions are generally not zero, but they are equal. Thus, there are no net changes in the concentrations of the reactants and products. Such a state is known as dynamic equilibrium.

Le Chatelier's principle, also called Chatelier's principle, is a principle of chemistry used to predict the effect of a change in conditions on chemical equilibria. The principle is named after French chemist Henry Louis Le Chatelier, and sometimes also credited to Karl Ferdinand Braun, who discovered it independently. It can be defined as:

If the equilibrium of a system is disturbed by a change in one or more of the determining factors the system tends to adjust itself to a new equilibrium by counteracting as far as possible the effect of the change

In physical chemistry, the Arrhenius equation is a formula for the temperature dependence of reaction rates. The equation was proposed by Svante Arrhenius in 1889, based on the work of Dutch chemist Jacobus Henricus van 't Hoff who had noted in 1884 that the van 't Hoff equation for the temperature dependence of equilibrium constants suggests such a formula for the rates of both forward and reverse reactions. This equation has a vast and important application in determining the rate of chemical reactions and for calculation of energy of activation. Arrhenius provided a physical justification and interpretation for the formula. Currently, it is best seen as an empirical relationship. It can be used to model the temperature variation of diffusion coefficients, population of crystal vacancies, creep rates, and many other thermally-induced processes/reactions. The Eyring equation, developed in 1935, also expresses the relationship between rate and energy.

In electrochemistry, the Nernst equation is a chemical thermodynamical relationship that permits the calculation of the reduction potential of a reaction from the standard electrode potential, absolute temperature, the number of electrons involved in the redox reaction, and activities of the chemical species undergoing reduction and oxidation respectively. It was named after Walther Nernst, a German physical chemist who formulated the equation.

<span class="mw-page-title-main">Gibbs free energy</span> Type of thermodynamic potential; useful for calculating reversible work in certain systems

In thermodynamics, the Gibbs free energy is a thermodynamic potential that can be used to calculate the maximum amount of work, other than pressure-volume work, that may be performed by a thermodynamically closed system at constant temperature and pressure. It also provides a necessary condition for processes such as chemical reactions that may occur under these conditions. The Gibbs free energy is expressed as

In thermodynamics, a spontaneous process is a process which occurs without any external input to the system. A more technical definition is the time-evolution of a system in which it releases free energy and it moves to a lower, more thermodynamically stable energy state. The sign convention for free energy change follows the general convention for thermodynamic measurements, in which a release of free energy from the system corresponds to a negative change in the free energy of the system and a positive change in the free energy of the surroundings.

The equilibrium constant of a chemical reaction is the value of its reaction quotient at chemical equilibrium, a state approached by a dynamic chemical system after sufficient time has elapsed at which its composition has no measurable tendency towards further change. For a given set of reaction conditions, the equilibrium constant is independent of the initial analytical concentrations of the reactant and product species in the mixture. Thus, given the initial composition of a system, known equilibrium constant values can be used to determine the composition of the system at equilibrium. However, reaction parameters like temperature, solvent, and ionic strength may all influence the value of the equilibrium constant.

<span class="mw-page-title-main">Enolate</span> Organic anion formed by deprotonating a carbonyl (>C=O) compound

In organic chemistry, enolates are organic anions derived from the deprotonation of carbonyl compounds. Rarely isolated, they are widely used as reagents in the synthesis of organic compounds.

<span class="mw-page-title-main">Cheletropic reaction</span> Chemical reaction in which a ring is formed/broken by adding/removing a single atom

In organic chemistry, cheletropic reactions, also known as chelotropic reactions, are a type of pericyclic reaction. 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.


Dynamic covalent chemistry (DCvC) is a synthetic strategy employed by chemists to make complex molecular and supramolecular assemblies from discrete molecular building blocks. DCvC has allowed access to complex assemblies such as covalent organic frameworks, molecular knots, polymers, and novel macrocycles. Not to be confused with dynamic combinatorial chemistry, DCvC concerns only covalent bonding interactions. As such, it only encompasses a subset of supramolecular chemistries.

The Curtin–Hammett principle is a principle in chemical kinetics proposed by David Yarrow Curtin and Louis Plack Hammett. It states that, for a reaction that has a pair of reactive intermediates or reactants that interconvert rapidly, each going irreversibly to a different product, the product ratio will depend both on the difference in energy between the two conformers and the energy barriers from each of the rapidly equilibrating isomers to their respective products. Stated another way, the product distribution reflects the difference in energy between the two rate-limiting transition states. As a result, the product distribution will not necessarily reflect the equilibrium distribution of the two intermediates. The Curtin–Hammett principle has been invoked to explain selectivity in a variety of stereo- and regioselective reactions. The relationship between the (apparent) rate constants and equilibrium constant is known as the Winstein-Holness equation.

The Van 't Hoff equation relates the change in the equilibrium constant, Keq, of a chemical reaction to the change in temperature, T, given the standard enthalpy change, ΔrH, for the process. The subscript means "reaction" and the superscript means "standard". It was proposed by Dutch chemist Jacobus Henricus van 't Hoff in 1884 in his book Études de Dynamique chimique.

In biochemistry, equilibrium unfolding is the process of unfolding a protein or RNA molecule by gradually changing its environment, such as by changing the temperature or pressure, pH, adding chemical denaturants, or applying force as with an atomic force microscope tip. If the equilibrium was maintained at all steps, the process theoretically should be reversible during equilibrium folding. Equilibrium unfolding can be used to determine the thermodynamic stability of the protein or RNA structure, i.e. free energy difference between the folded and unfolded states.

<span class="mw-page-title-main">Thermodynamic databases for pure substances</span> Thermodynamic properties list

Thermodynamic databases contain information about thermodynamic properties for substances, the most important being enthalpy, entropy, and Gibbs free energy. Numerical values of these thermodynamic properties are collected as tables or are calculated from thermodynamic datafiles. Data is expressed as temperature-dependent values for one mole of substance at the standard pressure of 101.325 kPa, or 100 kPa. Both of these definitions for the standard condition for pressure are in use.

<span class="mw-page-title-main">Transition state theory</span> Theory describing the reaction rates of elementary chemical reactions

In chemistry, transition state theory (TST) explains the reaction rates of elementary chemical reactions. The theory assumes a special type of chemical equilibrium (quasi-equilibrium) between reactants and activated transition state complexes.

Nucleic acid thermodynamics is the study of how temperature affects the nucleic acid structure of double-stranded DNA (dsDNA). The melting temperature (Tm) is defined as the temperature at which half of the DNA strands are in the random coil or single-stranded (ssDNA) state. Tm depends on the length of the DNA molecule and its specific nucleotide sequence. DNA, when in a state where its two strands are dissociated, is referred to as having been denatured by the high temperature.

In thermodynamics, enthalpy–entropy compensation is a specific example of the compensation effect. The compensation effect refers to the behavior of a series of closely related chemical reactions, which exhibit a linear relationship between one of the following kinetic or thermodynamic parameters for describing the reactions:

  1. Between the logarithm of the pre-exponential factors and the activation energies where the series of closely related reactions are indicated by the index i, Ai are the preexponential factors, Ea,i are the activation energies, R is the gas constant, and α, β are constants.
  2. Between enthalpies and entropies of activation where H
    i
    are the enthalpies of activation and S
    i
    are the entropies of activation.
  3. Between the enthalpy and entropy changes of a series of similar reactions where Hi are the enthalpy changes and Si are the entropy changes.

Equilibrium chemistry is concerned with systems in chemical equilibrium. The unifying principle is that the free energy of a system at equilibrium is the minimum possible, so that the slope of the free energy with respect to the reaction coordinate is zero. This principle, applied to mixtures at equilibrium provides a definition of an equilibrium constant. Applications include acid–base, host–guest, metal–complex, solubility, partition, chromatography and redox equilibria.

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.

In organic chemistry, the Conia-ene reaction is an intramolecular cyclization reaction between an enolizable carbonyl such as an ester or ketone and an alkyne or alkene, giving a cyclic product with a new carbon-carbon bond. As initially reported by J. M. Conia and P. Le Perchec, the Conia-ene reaction is a heteroatom analog of the ene reaction that uses an enol as the ene component. Like other pericyclic reactions, the original Conia-ene reaction required high temperatures to proceed, limiting its wider application. However, subsequent improvements, particularly in metal catalysis, have led to significant expansion of reaction scope. Consequently, various forms of the Conia-ene reaction have been employed in the synthesis of complex molecules and natural products.

References

  1. Organic Chemistry, 3rd ed., M. A. Fox & J. K. Whitesell, Jones & Bartlett, 2004 ISBN   0-7637-2197-2
  2. A Guidebook to Mechanism in Organic Chemistry, 6th Edition, Peter Sykes, Pearson Prentice Hall, 1986. ISBN   0-582-44695-3
  3. 1 2 Introduction to Organic Chemistry I, Seth Robert Elsheimer, Blackwell Publishing, 2000 ISBN   0-632-04417-9
  4. Advanced Organic Chemistry Part A: Structure and Mechanisms, 5th ed., Francis A. Carey, Richard J. Sundberg, 2007 ISBN   978-0-387-44899-2
  5. Kseniya K. Borisova, Elizaveta A. Kvyatkovskaya, Eugeniya V. Nikitina, Rinat R. Aysin, Roman A. Novikov, and Fedor I. Zubkov. “A Classical Example of Total Kinetic and Thermodynamic Control. The Diels-Alder Reaction between DMAD and Bis-furyl Dienes.” J. Org. Chem., 2018, 83 (8), pp 4840-4850. doi:10.1021/acs.joc.8b00336 https://pubs.acs.org/doi/abs/10.1021/acs.joc.8b00336
  6. Kseniya K. Borisova, Eugeniya V. Nikitina, Roman A. Novikov, Victor N. Khrustalev, Pavel V. Dorovatovskii, Yan V. Zubavichus, Maxim L. Kuznetsov, Vladimir P. Zaytsev, Alexey V. Varlamov and Fedor I. Zubkov. “Diels–Alder reactions between hexafluoro-2-butyne and bis-furyl dienes: kinetic versus thermodynamic control.” Chem. Commun., 2018, 54, pp 2850-2853. doi:10.1039/c7cc09466c http://pubs.rsc.org/en/content/articlelanding/2018/cc/c7cc09466c#!divAbstract
  7. Thermodynamic Product vs Kinetic Product
  8. Jean d'Angelo, Tetrahedron report number 25 : Ketone enolates: regiospecific preparation and synthetic uses, Tetrahedron, Volume 32, Issue 24, 1976, Pages 2979-2990, ISSN   0040-4020, doi : 10.1016/0040-4020(76)80156-1
  9. The Chemistry of Carbanions. IX. The Potassium and Lithium Enolates Derived from Cyclic Ketones Herbert O. House, Barry M. Trost J. Org. Chem., 1965, 30 (5), pp 1341–1348 doi : 10.1021/jo01016a001
  10. Chemistry of carbanions. XV. Stereochemistry of alkylation of 4-tert-butylcyclohexanone Herbert O. House, Ben A. Tefertiller, Hugh D. Olmstead J. Org. Chem., 1968, 33 (3), pp 935–942 doi : 10.1021/jo01267a002
  11. Khopade, Tushar; Mete, Trimbak; Arora, Jyotsna; Bhat, Ramakrishna (2018). "An Adverse Effect of Higher Catalyst Loading and Longer Reaction Time on Enantioselectivity in an Organocatalytic Multicomponent Reaction". Chemistry: A European Journal. 24 (23): 6036–6040. doi:10.1002/chem.201800278. PMID   29465758.
  12. Rulli, Giuseppe; Duangdee, Nongnaphat; Baer, Katrin; Hummel, Werner; Berkessel, Albrecht; Gröger, Harald (2011). "Direction of Kinetically versus Thermodynamically Controlled Organocatalysis and Its Application in Chemoenzymatic Synthesis". Angewandte Chemie International Edition. 50 (34): 7944–7947. doi:10.1002/anie.201008042. PMID   21744441. S2CID   42971817.
  13. Only if a subsequent equilibration is as fast or faster is this not true.
  14. Unless one is content with an incomplete reaction, whence a separation of product from unreacted starting material may be necessary.
  15. At worst, Keq will approach 1 as T rises and the proportion of the most stable product will tend toward 50% of the reaction mixture.
  16. will be temperature-independent or nearly so if is small, which would be the case if the rate-determining steps leading to each product were of the same molecularity, for instance if both involved collisions with the same reactant.
  17. will be temperature-independent or nearly so if is small, which would be the case if the overall transformations to each product were of the same molecularity, for instance if both were fragmentations of a molecule to produce a pair of molecules or if both were condensations of two molecules to give a single molecule.
  18. Studies on Diene-addition Reactions. II.1 The Reaction of 6,6-Pentamethylenefulvene with Maleic Anhydride R. B. Woodward, Harold Baer J. Am. Chem. Soc., 1944, 66 (4), pp 645–649 doi : 10.1021/ja01232a042
  19. Diels, O. and Alder, K. (1929), Synthesen in der hydroaromatischen Reihe, IV. Mitteilung: Über die Anlagerung von Maleinsäure-anhydrid an arylierte Diene, Triene und Fulvene (Mitbearbeitet von Paul Pries). Berichte der deutschen chemischen Gesellschaft (A and B Series), 62: 2081–2087. doi : 10.1002/cber.19290620829
  20. Rearrangement and substitution in anionotropic systems. Part III. Mechanism of, and equilibrium in, anionotropic change A. G. Catchpole, E. D. Hughes and C. K. Ingold J. Chem. Soc., 1948, 8-17 doi : 10.1039/JR9480000008
  21. Meisenheimer, J. and Link, J. (1930), Über die Verschiebung in der Allyl-Gruppe. 3. Mitteilung über Substitution und Addition. Justus Liebigs Annalen der Chemie, 479: 211–277. doi : 10.1002/jlac.19304790114