Hammett equation

Last updated

In organic chemistry, the Hammett equation describes a linear free-energy relationship relating reaction rates and equilibrium constants for many reactions involving benzoic acid derivatives with meta- and para- substituents to each other with just two parameters: a substituent constant and a reaction constant. [1] [2] This equation was developed and published by Louis Plack Hammett in 1937 [3] as a follow-up to qualitative observations in his 1935 publication. [4]

Contents

The basic idea is that for any two reactions with two aromatic reactants only differing in the type of substituent, the change in free energy of activation is proportional to the change in Gibbs free energy. [5] This notion does not follow from elemental thermochemistry or chemical kinetics and was introduced by Hammett intuitively. [lower-alpha 1]

The basic equation is:

where

= Reference constant
= Substituent constant
= Reaction rate constant

relating the equilibrium constant, , for a given equilibrium reaction with substituent R and the reference constant when R is a hydrogen atom to the substituent constant σ which depends only on the specific substituent R and the reaction rate constant ρ which depends only on the type of reaction but not on the substituent used. [4] [3]

The equation also holds for reaction rates k of a series of reactions with substituted benzene derivatives:

In this equation is the reference reaction rate of the unsubstituted reactant, and k that of a substituted reactant.

A plot of for a given equilibrium versus for a given reaction rate with many differently substituted reactants will give a straight line.

Substituent constants

The starting point for the collection of the substituent constants is a chemical equilibrium for which the substituent constant is arbitrarily set to 0 and the reaction constant is set to 1: the deprotonation of benzoic acid or benzene carboxylic acid (R and R' both H) in water at 25 °C.

Scheme 1. Dissociation of benzoic acids BenzoicAcidDissociation.svg
Scheme 1. Dissociation of benzoic acids
Substituent constants: para and meta substituted benzene rings. [3] [ needs update ] [6]
Substituentpara- effectmeta- effect
Dimethylamino0.830.211
Amino0.660.161
Butylamino0.510.34
Hydroxy0.37+0.12
Methoxy0.268+0.115
Ethoxy0.25+0.015
Methyl0.1700.069
Trimethylsilyl0.070.04
None0.0000.000
Fluoro+0.062+0.337
Chloro+0.227+0.373
Bromo+0.232+0.393
Iodo+0.276+0.353
Ethoxycarbonyl+0.45+0.37
Trifluoromethyl+0.54+0.43
Cyano+0.66+0.56
Nitro+0.778+0.710

Having obtained a value for K0, a series of equilibrium constants (K) are now determined based on the same process, but now with variation of the para substituent—for instance, p-hydroxybenzoic acid (R=OH, R'=H) or p-aminobenzoic acid (R=NH2, R'=H). These values, combined in the Hammett equation with K0 and remembering that ρ = 1, give the para substituent constants compiled in table 1 for amine, methoxy, ethoxy, dimethylamino, methyl, fluorine, bromine, chlorine, iodine, nitro and cyano substituents. Repeating the process with meta-substituents afford the meta substituent constants. This treatment does not include ortho-substituents, which would introduce steric effects.

The σ values displayed in the Table above reveal certain substituent effects. With ρ = 1, the group of substituents with increasing positive values—notably cyano and nitro—cause the equilibrium constant to increase compared to the hydrogen reference, meaning that the acidity of the carboxylic acid (depicted on the left of the equation) has increased. These substituents stabilize the negative charge on the carboxylate oxygen atom by an electron-withdrawing inductive effect (-I) and also by a negative mesomeric effect (-M).

The next set of substituents are the halogens, for which the substituent effect is still positive but much more modest. The reason for this is that while the inductive effect is still negative, the mesomeric effect is positive, causing partial cancellation. The data also show that for these substituents, the meta effect is much larger than the para effect, due to the fact that the mesomeric effect is greatly reduced in a meta substituent. With meta substituents a carbon atom bearing the negative charge is further away from the carboxylic acid group (structure 2b).

This effect is depicted in scheme 3, where, in a para substituted arene 1a, one resonance structure 1b is a quinoid with positive charge on the X substituent, releasing electrons and thus destabilizing the Y substituent. This destabilizing effect is not possible when X has a meta orientation.

Scheme 3. Hammett Inductive Mesomeric Effects HammettInductiveMesomericEffects.png
Scheme 3. Hammett Inductive Mesomeric Effects

Other substituents, like methoxy and ethoxy, can even have opposite signs for the substituent constant as a result of opposing inductive and mesomeric effect. Only alkyl and aryl substituents like methyl are electron-releasing in both respects.

Of course, when the sign for the reaction constant is negative (next section), only substituents with a likewise negative substituent constant will increase equilibrium constants.

The σp and σp+ constants

Because the carbonyl group is unable to serve a source of electrons for -M groups (in contrast to lone pair donors like OH), for reactions involving phenol and aniline starting materials, the σp values for electron-withdrawing groups will appear too small. For reactions where resonance effects are expected to have a major impact, a modified parameter, and a modified set of σp constants may give a better fit. This parameter is defined using the ionization constants of para substituted phenols, via a scaling factor to match up the values of σp with those of σp for "non-anomalous" substituents, so as to maintain comparable ρ values: for ArOH ⇄ ArO + H+, we define .

Likewise, the carbonyl carbon of a benzoic acid is at a nodal position and unable to serve as a sink for +M groups (in contrast to a carbocation at the benzylic position). Thus for reactions involving carbocations at the α-position, the σp values for electron-donating groups will appear insufficiently negative. Based on similar considerations, a set of σp+ constants give better fit for reactions involving electron-donating groups at the para position and the formation of a carbocation at the benzylic site. The σp+ are based on the rate constants of the SN1 reaction of cumyl chlorides in 90% acetone/water: for ArCMe2Cl + H2O → ArCMe2OH + HCl, we define . Note that the scaling factor is negative, since an electron-donating group speeds up the reaction. For a reaction whose Hammett plot is being constructed, these alternative Hammett constants may need to be tested to see if a better linearity could be obtained.

Rho value

With knowledge of substituent constants it is now possible to obtain reaction constants for a wide range of organic reactions. The archetypal reaction is the alkaline hydrolysis of ethyl benzoate (R=R'=H) in a water/ethanol mixture at 30 °C. Measurement of the reaction rate k0 combined with that of many substituted ethyl benzoates ultimately result in a reaction constant of +2.498. [3] [ needs update ][ non-primary source needed ]

Scheme 2. Hydrolysis of benzoic acid esters BenzoateEsterHydrolysis.png
Scheme 2. Hydrolysis of benzoic acid esters

Reaction constants are known for many other reactions and equilibria. Here is a selection of those provided by Hammett himself (with their values in parentheses):

The reaction constant, or sensitivity constant, ρ, describes the susceptibility of the reaction to substituents, compared to the ionization of benzoic acid. It is equivalent to the slope of the Hammett plot. Information on the reaction and the associated mechanism can be obtained based on the value obtained for ρ. If the value of:

  1. ρ>1, the reaction is more sensitive to substituents than benzoic acid and negative charge is built during the reaction (or positive charge is lost).
  2. 0<ρ<1, the reaction is less sensitive to substituents than benzoic acid and negative charge is built (or positive charge is lost).
  3. ρ=0, no sensitivity to substituents, and no charge is built or lost.
  4. ρ<0, the reaction builds positive charge (or loses negative charge).

These relations can be exploited to elucidate the mechanism of a reaction. As the value of ρ is related to the charge during the rate determining step, mechanisms can be devised based on this information. If the mechanism for the reaction of an aromatic compound is thought to occur through one of two mechanisms, the compound can be modified with substituents with different σ values and kinetic measurements taken. Once these measurements have been made, a Hammett plot can be constructed to determine the value of ρ. If one of these mechanisms involves the formation of charge, this can be verified based on the ρ value. Conversely, if the Hammett plot shows that no charge is developed, i.e. a zero slope, the mechanism involving the building of charge can be discarded.

Hammett plots may not always be perfectly linear. For instance, a curve may show a sudden change in slope, or ρ value. In such a case, it is likely that the mechanism of the reaction changes upon adding a different substituent. Other deviations from linearity may be due to a change in the position of the transition state. In such a situation, certain substituents may cause the transition state to appear earlier (or later) in the reaction mechanism. [7] [ page needed ]

Dominating electronic effects

3 kinds of ground state or static electrical influences predominate:

The latter two influences are often treated together as a composite effect, but are treated here separately. Westheimer demonstrated that the electrical effects of π-substituted dipolar groups on the acidities of benzoic and phenylacetic acids can be quantitatively correlated, by assuming only direct electrostatic action of the substituent on the ionizable proton of the carboxyl group. Westheimer's treatment worked well except for those acids with substituents that have unshared electron pairs such as –OH and –OCH3, as these substituents interact strongly with the benzene ring. [8] [ non-primary source needed ][ non-primary source needed ] [9] [ non-primary source needed ][ needs update ][ non-primary source needed ]

4-substituted bicyclo-2.2.2.-octane-1-carboxylic acid 4-substituted bicyclo(2.2.2)octane-1-carboxylic ester.svg
4-substituted bicyclo-2.2.2.-octane-1-carboxylic acid

Roberts and Moreland studied the reactivities of 4-substituted bicyclo[2.2.2]octane-1-carboxylic acids and esters. In such a molecule, transmission of electrical effects of substituents through the ring by resonance is not possible. Hence, this hints on the role of the π-electrons in the transmission of substituent effects through aromatic systems. [10] [ non-primary source needed ][ non-primary source needed ]

Reactivity of 4-substituted bicyclo[2.2.2]octane-1-carboxylic acids and esters were measured in 3 different processes, each of which had been previously used with the benzoic acid derivatives. A plot of log(k) against log(KA) showed a linear relationship. Such linear relationships correspond to linear free energy relationships, which strongly imply that the effect of the substituents are exerted through changes of potential energy and that the steric and entropy terms remain almost constant through the series. The linear relationship fit well in the Hammett Equation. For the 4-substituted bicyclo[2.2.2.]octane-1-carboxylic acid derivatives, the substituent and reaction constants are designated σ’ and ρ’.

Comparison of ρ and ρ’

Reaction[ citation needed ]ρ'ρDe
Ionization of acids1.4641.46454
Alkaline hydrolysis of ethyl esters2.242.49428
Acids with diphenyldiazomethane0.6980.93724

Reactivity data indicate that the effects of substituent groups in determining the reactivities of substituted benzoic and bicyclo[2.2.2.]-octane-1-carboxylic acids are comparable. This implies that the aromatic π-electrons do not play a dominant role in the transmission of electrical effects of dipolar groups to the ionizable carboxyl group Difference between ρ and ρ’ for the reactions of the acids with diphenylazomethane is probably due to an inverse relation to the solvent dielectric constant De

Comparison of σ and σ’

Substituentσ’ [ citation needed ]σparacσmetacσpara − σ’ [ citation needed ]σmeta − σ’ [ citation needed ]
H00000
OH0.283−0.3410.014−0.624−0.269
CO2C2H50.2970.4020.3340.1050.037
Br0.4540.2320.391−0.222−0.063
CN0.5790.6560.6080.0770.029

For meta-directing groups (electron withdrawing group or EWG), σmeta and σpara are more positive than σ’. (The superscript, c, in table denotes data from Hammett, 1940. [11] [ page needed ]) For ortho-para directing groups (electron donating group or EDG), σ’ more positive than σmeta and σpara. The difference between σpara and σ’ (σpara – σ’) is greater than that between σmeta and σ’(σmeta − σ’). This is expected as electron resonance effects are felt more strongly at the p-positions. The (σ – σ’) values can be taken as a reasonable measurement of the resonance effects.

Nonlinearity

Rate acceleration EDG Rate acceleration EDG.png
Rate acceleration EDG

The plot of the Hammett equation is typically seen as being linear, with either a positive or negative slope correlating to the value of rho. However, nonlinearity emerges in the Hammett plot when a substituent affects the rate of reaction or changes the rate-determining step or reaction mechanism of the reaction. For the reason of the former case, new sigma constants have been introduced to accommodate the deviation from linearity otherwise seen resulting from the effect of the substituent. σ+ takes into account positive charge buildup occurring in the transition state of the reaction. Therefore, an electron donating group (EDG) will accelerate the rate of the reaction by resonance stabilization and will give the following sigma plot with a negative rho value. [12] [ non-primary source needed ][ non-primary source needed ]

Rate acceleration EWG Rate acceleration EWG.png
Rate acceleration EWG

σ- is designated in the case where negative charge buildup in the transition state occurs, and the rate of the reaction is consequently accelerated by electron withdrawing groups (EWG). The EWG withdraws electron density by resonance and effectively stabilizes the negative charge that is generated. The corresponding plot will show a positive rho value.

In the case of a nucleophilic acyl substitution the effect of the substituent, X, of the non-leaving group can in fact accelerate the rate of the nucleophilic addition reaction when X is an EWG. This is attributed to the resonance contribution of the EWG to withdraw electron density thereby increasing the susceptibility for nucleophilic attack on the carbonyl carbon. A change in rate occurs when X is EDG, as is evidenced when comparing the rates between X = Me and X = OMe, and nonlinearity is observed in the Hammett plot. [13] [ non-primary source needed ][ non-primary source needed ]

Nuc mech.png

The effect of the substituent may change the rate-determining step (rds) in the mechanism of the reaction. A certain electronic effect may accelerate a certain step so that it is no longer the rds. [14] [ non-primary source needed ][ non-primary source needed ]

Change in rds.png
change in rate determining step Change in rds graph.png
change in rate determining step

A change in the mechanism of a reaction also results in nonlinearity in the Hammett plot. Typically, the model used for measuring the changes in rate in this instance is that of the SN2 reaction. [15] [ non-primary source needed ][ non-primary source needed ] However, it has been observed that in some cases of an SN2 reaction that an EWG does not accelerate the reaction as would be expected [16] [ non-primary source needed ][ non-primary source needed ] and that the rate varies with the substituent. In fact, the sign of the charge and degree to which it develops will be affected by the substituent in the case of the benzylic system. [15] [ non-primary source needed ]

Change in emchanism Change in mech graph.png
Change in emchanism

For example, the substituent may determine the mechanism to be an SN1 type reaction over a SN2 type reaction, in which case the resulting Hammett plot will indicate a rate acceleration due to an EDG, thus elucidating the mechanism of the reaction.

Change in mech.png

Another deviation from the regular Hammett equation is explained by the charge of nucleophile. [15] [ non-primary source needed ] Despite nonlinearity in benzylic SN2 reactions, electron withdrawing groups could either accelerate or retard the reaction. If the nucleophile is negatively charged (e.g. cyanide) the electron withdrawing group will increase the rate due to stabilization of the extra charge which is put on the carbon in the transition state. On the other hand, if the nucleophile is not charged (e.g. triphenylphosphine), electron withdrawing group is going to slow down the reaction by decreasing the electron density in the anti bonding orbital of leaving group in the transition state.

Hammett modifications

Other equations now exist that refine the original Hammett equation: the Swain–Lupton equation,[ citation needed ] the Taft equation,[ citation needed ] the Grunwald–Winstein equation,[ citation needed ] and the Yukawa–Tsuno equation.[ citation needed ] An equation that addresses stereochemistry in aliphatic systems has also been developed.[ vague ] [17] [ non-primary source needed ][ non-primary source needed ]

Estimation of Hammett sigma constants

Carbon positions.png

Core-electron binding energy (CEBE) shifts correlate linearly with the Hammett substituent constants (σ) in substituted benzene derivatives. [18] [ non-primary source needed ]

ΔCEBE ≈ κσp

 

 

 

 

(1)

Consider para-disubstituted benzene p-F-C6H4-Z, where Z is a substituent such as NH2, NO2, etc. The fluorine atom is para with respect to the substituent Z in the benzene ring. The image on the right shows four distinguished ring carbon atoms, C1(ipso), C2(ortho), C3(meta), C4(para) in p-F-C6H4-Z molecule. The carbon with Z is defined as C1(ipso) and fluorinated carbon as C4(para). This definition is followed even for Z = H. The left-hand side of ( 1 ) is called CEBE shift or ΔCEBE, and is defined as the difference between the CEBE of the fluorinated carbon atom in p-F-C6H4-Z and that of the fluorinated carbon in the reference molecule FC6H5.

ΔCEBE ≡ CEBE(C4 in p-F-C6H4-Z) – CEBE(C4 in p-F-C6H5)

 

 

 

 

(2)

The right-hand side of Eq. 1 is a product of a parameter κ and a Hammett substituent constant at the para position, σp. The parameter κ is defined by eq. 3 :

κ = 2.3kT(ρ - ρ*)

 

 

 

 

(3)

where ρ and ρ* are the Hammett reaction constants for the reaction of the neutral molecule and core ionized molecule, respectively. ΔCEBEs of ring carbons in p-F-C6H4-Z were calculated with density functional theory to see how they correlate with Hammett σ-constants. Linear plots were obtained when the calculated CEBE shifts at the ortho, meta and para carbon were plotted against Hammett σo, σm and σp constants respectively.

Hence the approximate agreement in numerical value and in sign between the CEBE shifts and their corresponding Hammett σ constant. [19] [ non-primary source needed ][ non-primary source needed ]

See also

Notes

  1. The opening line in his 1935 publication reads: "The idea that there is some sort of relationship between the rate of a reaction and the equilibrium constant is one of the most persistently held and at the same time most emphatically denied concepts in chemical theory". [4]

Related Research Articles

In chemistry, an acid dissociation constant is a quantitative measure of the strength of an acid in solution. It is the equilibrium constant for a chemical reaction

Electrical resistivity is a fundamental specific property of a material that measures its electrical resistance or how strongly it resists electric current. A low resistivity indicates a material that readily allows electric current. Resistivity is commonly represented by the Greek letter ρ (rho). The SI unit of electrical resistivity is the ohm-metre (Ω⋅m). For example, if a 1 m3 solid cube of material has sheet contacts on two opposite faces, and the resistance between these contacts is 1 Ω, then the resistivity of the material is 1 Ω⋅m.

S<sub>N</sub>2 reaction Substitution reaction where bonds are broken and formed simultaneously

Bimolecular nucleophilic substitution (SN2) is a type of reaction mechanism that is common in organic chemistry. In the SN2 reaction, a strong nucleophile forms a new bond to an sp3-hybridised carbon via a backside attack, all while the leaving group detaches from the reaction center in a concerted fashion.

A substitution reaction is a chemical reaction during which one functional group in a chemical compound is replaced by another functional group. Substitution reactions are of prime importance in organic chemistry. Substitution reactions in organic chemistry are classified either as electrophilic or nucleophilic depending upon the reagent involved, whether a reactive intermediate involved in the reaction is a carbocation, a carbanion or a free radical, and whether the substrate is aliphatic or aromatic. Detailed understanding of a reaction type helps to predict the product outcome in a reaction. It also is helpful for optimizing a reaction with regard to variables such as temperature and choice of solvent.

<span class="mw-page-title-main">Nitration</span> Chemical reaction which adds a nitro (–NO₂) group onto a molecule

In organic chemistry, nitration is a general class of chemical processes for the introduction of a nitro group into an organic compound. The term also is applied incorrectly to the different process of forming nitrate esters between alcohols and nitric acid. The difference between the resulting molecular structures of nitro compounds and nitrates is that the nitrogen atom in nitro compounds is directly bonded to a non-oxygen atom, whereas in nitrate esters, the nitrogen is bonded to an oxygen atom that in turn usually is bonded to a carbon atom.

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">Collision theory</span> Chemistry principle

Collision theory is a principle of chemistry used to predict the rates of chemical reactions. It states that when suitable particles of the reactant hit each other with the correct orientation, only a certain amount of collisions result in a perceptible or notable change; these successful changes are called successful collisions. The successful collisions must have enough energy, also known as activation energy, at the moment of impact to break the pre-existing bonds and form all new bonds. This results in the products of the reaction. The activation energy is often predicted using the Transition state theory. Increasing the concentration of the reactant brings about more collisions and hence more successful collisions. Increasing the temperature increases the average kinetic energy of the molecules in a solution, increasing the number of collisions that have enough energy. Collision theory was proposed independently by Max Trautz in 1916 and William Lewis in 1918.

<span class="mw-page-title-main">Hyperconjugation</span> Concept in organic chemistry

In organic chemistry, hyperconjugation refers to the delocalization of electrons with the participation of bonds of primarily σ-character. Usually, hyperconjugation involves the interaction of the electrons in a sigma (σ) orbital with an adjacent unpopulated non-bonding p or antibonding σ* or π* orbitals to give a pair of extended molecular orbitals. However, sometimes, low-lying antibonding σ* orbitals may also interact with filled orbitals of lone pair character (n) in what is termed negative hyperconjugation. Increased electron delocalization associated with hyperconjugation increases the stability of the system. In particular, the new orbital with bonding character is stabilized, resulting in an overall stabilization of the molecule. Only electrons in bonds that are in the β position can have this sort of direct stabilizing effect — donating from a sigma bond on an atom to an orbital in another atom directly attached to it. However, extended versions of hyperconjugation can be important as well. The Baker–Nathan effect, sometimes used synonymously for hyperconjugation, is a specific application of it to certain chemical reactions or types of structures.

In chemistry, the inductive effect in a molecule is a local change in the electron density due to electron-withdrawing or electron-donating groups elsewhere in the molecule, resulting in a permanent dipole in a bond. It is present in a σ (sigma) bond, unlike the electromeric effect which is present in a π (pi) bond.

An electron-withdrawing group (EWG) is a group or atom that has the ability to draw electron density toward itself and away from other adjacent atoms. This electron density transfer is often achieved by resonance or inductive effects. Electron-withdrawing groups have significant impacts on fundamental chemical processes such as acid-base reactions, redox potentials, and substitution reactions.

Physical organic chemistry, a term coined by Louis Hammett in 1940, refers to a discipline of organic chemistry that focuses on the relationship between chemical structures and reactivity, in particular, applying experimental tools of physical chemistry to the study of organic molecules. Specific focal points of study include the rates of organic reactions, the relative chemical stabilities of the starting materials, reactive intermediates, transition states, and products of chemical reactions, and non-covalent aspects of solvation and molecular interactions that influence chemical reactivity. Such studies provide theoretical and practical frameworks to understand how changes in structure in solution or solid-state contexts impact reaction mechanism and rate for each organic reaction of interest.

More O’Ferrall–Jencks plots are two-dimensional representations of multiple reaction coordinate potential energy surfaces for chemical reactions that involve simultaneous changes in two bonds. As such, they are a useful tool to explain or predict how changes in the reactants or reaction conditions can affect the position and geometry of the transition state of a reaction for which there are possible competing pathways.

The Taft equation is a linear free energy relationship (LFER) used in physical organic chemistry in the study of reaction mechanisms and in the development of quantitative structure–activity relationships for organic compounds. It was developed by Robert W. Taft in 1952 as a modification to the Hammett equation. While the Hammett equation accounts for how field, inductive, and resonance effects influence reaction rates, the Taft equation also describes the steric effects of a substituent. The Taft equation is written as:

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.

In physical organic chemistry, the Grunwald–Winstein equation is a linear free energy relationship between relative rate constants and the ionizing power of various solvent systems, describing the effect of solvent as nucleophile on different substrates. The equation, which was developed by Ernest Grunwald and Saul Winstein in 1948, could be written

In physical organic chemistry, the Swain–Lupton equation is a linear free energy relationship (LFER) that is used in the study of reaction mechanisms and in the development of quantitative structure activity relationships for organic compounds. It was developed by C. Gardner Swain and Elmer C. Lupton Jr. in 1968 as a refinement of the Hammett equation to include both field effects and resonance effects.

The Yukawa–Tsuno equation, first developed in 1959, is a linear free-energy relationship in physical organic chemistry. It is a modified version of the Hammett equation that accounts for enhanced resonance effects in electrophilic reactions of para- and meta-substituted organic compounds. This equation does so by introducing a new term to the original Hammett relation that provides a measure of the extent of resonance stabilization for a reactive structure that builds up charge in its transition state. The Yukawa–Tsuno equation can take the following forms:

<span class="mw-page-title-main">Birch reduction</span> Organic reaction used to convert arenes to cyclohexadienes

The Birch reduction is an organic reaction that is used to convert arenes to 1,4-cyclohexadienes. The reaction is named after the Australian chemist Arthur Birch and involves the organic reduction of aromatic rings in an amine solvent with an alkali metal and a proton source. Unlike catalytic hydrogenation, Birch reduction does not reduce the aromatic ring all the way to a cyclohexane.

Electrophilic aromatic substitution is an organic reaction in which an atom that is attached to an aromatic system is replaced by an electrophile. Some of the most important electrophilic aromatic substitutions are aromatic nitration, aromatic halogenation, aromatic sulfonation, alkylation and acylation Friedel–Crafts reaction.

In organic chemistry, the Cieplak effect is a predictive model to rationalize why nucleophiles preferentially add to one face of a carbonyl over another. Proposed by Andrzej Stanislaw Cieplak in 1980, it correctly predicts results that could not be justified by the other standard models at the time, such as the Cram and Felkin–Anh models. In the Cieplak model, electrons from a neighboring bond delocalize into the forming carbon–nucleophile (C–Nuc) bond, lowering the energy of the transition state and accelerating the rate of reaction. Whichever bond can best donate its electrons into the C–Nuc bond determines which face of the carbonyl the nucleophile will add to. The nucleophile may be any of a number of reagents, most commonly organometallic or reducing agents. The Cieplak effect is subtle, and often competes with sterics, solvent effects, counterion complexation of the carbonyl oxygen, and other effects to determine product distribution. Subsequent work has questioned its legitimacy.

References

  1. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " Hammett equation (Hammett relation) ". doi : 10.1351/goldbook.H02732
  2. Keenan, Sheue L.; Peterson, Karl P.; Peterson, Kelly; Jacobson, Kyle (2008). "Determination of Hammett Equation Rho Constant for the Hydrolysis of p-Nitrophenyl Benzoate Esters". J. Chem. Educ. 85 (4): 558. Bibcode:2008JChEd..85..558K. doi:10.1021/ed085p558.
  3. 1 2 3 4 Hammett, Louis P. (1937). "The Effect of Structure upon the Reactions of Organic Compounds. Benzene Derivatives". J. Am. Chem. Soc. 59 (1): 96–103. doi:10.1021/ja01280a022.
  4. 1 2 3 Louis P. Hammett (1935). "Some relations between Reaction Rates and Equilibrium Constants". Chem. Rev. 17 (1): 125–136. doi:10.1021/cr60056a010.
  5. Carey, Francis A.; Sundberg, Richard J. (1983). Advanced Organic Chemistry Part A, 2nd edition. Plenum Press. ISBN   9780306410871. OCLC   1064985344.
  6. Table values are this original 1937 publication, and differ from values appearing in subsequent publications. For more standard values, see: C. Hansch; A. Leo; R. W. Taft (1991). "A survey of Hammett substituent constants and resonance and field parameters". Chem. Rev. 91 (2): 165–195. doi:10.1021/cr00002a004. S2CID   97583278.
  7. E.V. Anslyn & D.A. Dougherty, Modern Physical Organic Chemistry, pp. TBD, Sausalito, CA, US: University Science Books, ISBN   1891389319.[ page needed ]
  8. Westheimer F.H. (1939). "The Electrostatic effect of substituents on the dissociation constants of organic acids. IV. Aromatic acids". J. Am. Chem. Soc. 61 (8): 1977–1980. doi:10.1021/ja01877a012.
  9. Kirkwood J.G.; Westheimer F.H. (1938). "The electrostatic influence of substituents on the dissociation constants of organic acids. I [Missing Subtitle]". J. Chem. Phys. 6 (9): 506. Bibcode:1938JChPh...6..506K. doi:10.1063/1.1750302.[ needs update ]
  10. Roberts J.D.; Moreland Jr. W.T. (1953). "Electrical Effects of Substituent Groups in Saturated Systems. Reactivities of 4-Substituted Bicyclo [2.2.2] octane-1-carboxylic acids". J. Am. Chem. Soc. 75 (9): 2167–2173. doi:10.1021/ja01105a045.
  11. L.P.Hammett, 1940, "Chapter III," "Chapter IV," and "Chapter VII," in Physical Organic Chemistry, New York, NY, US: McGraw-Hill.[ page needed ]
  12. Y. Yukawa & Y. Tsuno, 1959, "Resonance Effect in Hammett Relationship. II. Sigma Constants in Electrophilic Reactions and their Intercorrelation," Bull. Chem. Soc. Jpn.32:965-971, see , accessed 22 June 2015.
  13. Um, Ik-Hwan; Lee, Ji-Youn; Kim, Han-Tae; Bae, Sun-Kun (2004). "Curved Hammett plot in alkaline hydrolysis of O-aryl thionobenzoates: Change in rate-determining step versus ground-state stabilization". J. Org. Chem. 69 (7): 2436–2441. doi:10.1021/jo035854r. PMID   15049643.
  14. Hart, H.; Sedor, Edward A. (1967). "Mechanism of cyclodehydration of 2-phenyltriarylcarbinols". J. Am. Chem. Soc. 89 (10): 2342. doi:10.1021/ja00986a018.
  15. 1 2 3 Stein, Allan R.; Tencer, Michal; Moffatt, Elizabeth A.; Dawe, Robert; Sweet, James (1980). "Nonlinearity of Hammett .sigma..rho. correlations for benzylic systems: activation parameters and their mechanistic implications". J. Org. Chem. 45 (17): 3539–3540. doi:10.1021/jo01305a045.
  16. Young, P. R.; Jencks, W. P. (1979). "Separation of polar and resonance substituent effects in the reactions of acetophenones with bisulfite and of benzyl halides with nucleophiles". J. Am. Chem. Soc. 101 (12): 3288. doi:10.1021/ja00506a025.
  17. Bols, Mikael; Liang, Xifu; Jensen, Henrik H. (2002). "Equatorial contra axial polar substituents. The relation of a chemical reaction to stereochemical substituent constants". J. Org. Chem. 67 (25): 8970–4. doi:10.1021/jo0205356. PMID   12467416.
  18. Linderberg, B.; Svensson, S.; Malmquist, P.A.; Basilier, E.; Gelius, U.; Siegbahn, K. (1976). "Correlation of ESCA shifts and Hammett substituent constants in substituted benzene derivatives". Chem. Phys. Lett. 40 (2): 175. Bibcode:1976CPL....40..175L. doi:10.1016/0009-2614(76)85053-1.
  19. Takahata Y.; Chong D.P. (2005). "Estimation of Hammett sigma constants of substituted benzenes through accurate density-functional calculation of core-electron binding energy shifts". International Journal of Quantum Chemistry . 103 (5): 509–515. Bibcode:2005IJQC..103..509T. doi:10.1002/qua.20533.

Further reading

General

Theory

Surveys of descriptors

History