Edwards equation

Last updated

The Edwards equation in organic chemistry is a two-parameter equation for correlating nucleophilic reactivity, as defined by relative rate constants, with the basicity of the nucleophile (relative to protons) and its polarizability. This equation was first developed by John O. Edwards in 1954 [1] and later revised based on additional work in 1956. [2]

Contents

The general idea is that most nucleophiles are also good bases because the concentration of negatively charged electron density that defines a nucleophile will strongly attract positively charged protons, which is the definition of a base according to Brønsted–Lowry acid-base theory. Additionally, highly polarizable nucleophiles will have greater nucleophilic character than suggested by their basicity because their electron density can be shifted with relative ease to concentrate in one area.

History

Prior to Edwards developing his equation, other scientists were also working to define nucleophilicity quantitatively. Brønsted and Pederson first discovered the relationship between basicity, with respect to protons, and nucleophilicity in 1924: [3]

where

where kb is the rate constant for nitramide decomposition by a base (B) and βN is a parameter of the equation.

Swain and Scott later tried to define a more specific and quantitative relationship by correlating nucleophilic data with a single-parameter equation [4] [5] derived in 1953:

This equation relates the rate constant k, of a reaction, normalized to that of a standard reaction with water as the nucleophile (k0), to a nucleophilic constant n for a given nucleophile and a substrate constant s that depends on the sensitivity of a substrate to nucleophilic attack (defined as 1 for methyl bromide). This equation was modeled after the Hammett equation.

However, both the Swain–Scott equation and the Brønsted relationship make the rather inaccurate assumption that all nucleophiles have the same reactivity with respect to a specific reaction site. There are several different categories of nucleophiles with different attacking atoms (e.g. oxygen, carbon, nitrogen) and each of these atoms has different nucleophilic characteristics. [3] The Edwards equation attempts to account for this additional parameter by introducing a polarizability term.

Edwards equations

The first generation of the Edwards equation [1] was

where k and k0 are the rate constants for a nucleophile and a standard (H2O). H is a measure of the basicity of the nucleophile relative to protons, as defined by the equation:

where the pKa is that of the conjugate acid of the nucleophile and the constant 1.74 is the correction for the pKa of H3O+.

En is the term Edwards introduced to account for the polarizability of the nucleophile. It is related to the oxidation potential (E0) of the reaction (oxidative dimerization of the nucleophile) by the equation:

where 2.60 is the correction for the oxidative dimerization of water, obtained from a least-squares correlation of data in Edwards’ first paper on the subject. [1] α and β are then parameters unique to specific nucleophiles that relate the sensitivity of the substrate to the basicity and polarizability factors. [6] However, because some β's appeared to be negative as defined by the first generation of the Edwards equation, which theoretically should not occur, Edwards adjusted his equation. The term En was determined to have some dependence on the basicity relative to protons (H) due to some factors that affect basicity also influencing the electrochemical properties of the nucleophile. To account for this, En was redefined in terms of basicity and polarizability (given as molar refractivity, RN):

where

The values of a and b, obtained by the method of least squares, are 3.60 and 0.0624 respectively. [2] With this new definition of En, the Edwards equation can be rearranged:

where A= αa and B = β + αb. However, because the second generation of the equation was also the final one, the equation is sometimes written as , especially since it was republished in that form in a later paper of Edwards’, [7] leading to confusion over which parameters are being defined.

Significance

A later paper by Edwards and Pearson, following research done by Jencks and Carriuolo in 1960 [8] [9] led to the discovery of an additional factor in nucleophilic reactivity, which Edwards and Pearson called the alpha effect, [7] where nucleophiles with a lone pair of electrons on an atom adjacent to the nucleophilic center have enhanced reactivity. The alpha effect, basicity, and polarizability are still accepted as the main factors in determining nucleophilic reactivity. As such, the Edwards equation is applied in a qualitative sense much more frequently than in a quantitative one. [10] In studying nucleophilic reactions, Edwards and Pearson noticed that for certain classes of nucleophiles most of the contribution of nucleophilic character originated from their basicity, resulting in large β values. For other nucleophiles, most of the nucleophilic character came from their high polarizability, with little contribution from basicity, resulting in large α values. This observation led Pearson to develop his hard-soft acid-base theory, which is arguably the most important contribution that the Edwards equation has made to current understanding of organic and inorganic chemistry. [11] Nucleophiles, or bases, that were polarizable, with large α values, were categorized as “soft”, and nucleophiles that were non-polarizable, with large β and small α values, were categorized as “hard”.

The Edwards equation parameters have since been used to help categorize acids and bases as hard or soft, due to the approach's simplicity. [12]

See also

Related Research Articles

<span class="mw-page-title-main">Acid–base reaction</span> Chemical reaction

An acid–base reaction is a chemical reaction that occurs between an acid and a base. It can be used to determine pH via titration. Several theoretical frameworks provide alternative conceptions of the reaction mechanisms and their application in solving related problems; these are called the acid–base theories, for example, Brønsted–Lowry acid–base theory.

In chemistry, a nucleophile is a chemical species that forms bonds by donating an electron pair. All molecules and ions with a free pair of electrons or at least one pi bond can act as nucleophiles. Because nucleophiles donate electrons, they are Lewis bases.

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

<span class="mw-page-title-main">Lewis acids and bases</span> Chemical bond theory

A Lewis acid (named for the American physical chemist Gilbert N. Lewis) is a chemical species that contains an empty orbital which is capable of accepting an electron pair from a Lewis base to form a Lewis adduct. A Lewis base, then, is any species that has a filled orbital containing an electron pair which is not involved in bonding but may form a dative bond with a Lewis acid to form a Lewis adduct. For example, NH3 is a Lewis base, because it can donate its lone pair of electrons. Trimethylborane (Me3B) is a Lewis acid as it is capable of accepting a lone pair. In a Lewis adduct, the Lewis acid and base share an electron pair furnished by the Lewis base, forming a dative bond. In the context of a specific chemical reaction between NH3 and Me3B, a lone pair from NH3 will form a dative bond with the empty orbital of Me3B to form an adduct NH3•BMe3. The terminology refers to the contributions of Gilbert N. Lewis.

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. Because electrophiles accept electrons, they are Lewis acids. Most electrophiles are positively charged, have an atom that carries a partial positive charge, or have an atom that does not have an octet of electrons.

HSAB concept is a jargon for "hard and soft (Lewis) acids and bases". HSAB is widely used in chemistry for explaining stability of compounds, reaction mechanisms and pathways. It assigns the terms 'hard' or 'soft', and 'acid' or 'base' to chemical species. 'Hard' applies to species which are small, have high charge states, and are weakly polarizable. 'Soft' applies to species which are big, have low charge states and are strongly polarizable.

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

In organic chemistry, the Michael reaction or Michael 1,4 addition is a reaction between a Michael donor and a Michael acceptor to produce a Michael adduct by creating a carbon-carbon bond at the acceptor's β-carbon. It belongs to the larger class of conjugate additions and is widely used for the mild formation of carbon-carbon bonds.

<span class="mw-page-title-main">Nucleophilic conjugate addition</span> Organic reaction

Nucleophilic conjugate addition is a type of organic reaction. Ordinary nucleophilic additions or 1,2-nucleophilic additions deal mostly with additions to carbonyl compounds. Simple alkene compounds do not show 1,2 reactivity due to lack of polarity, unless the alkene is activated with special substituents. With α,β-unsaturated carbonyl compounds such as cyclohexenone it can be deduced from resonance structures that the β position is an electrophilic site which can react with a nucleophile. The negative charge in these structures is stored as an alkoxide anion. Such a nucleophilic addition is called a nucleophilic conjugate addition or 1,4-nucleophilic addition. The most important active alkenes are the aforementioned conjugated carbonyls and acrylonitriles.

In organic chemistry, umpolung or polarity inversion is the chemical modification of a functional group with the aim of the reversal of polarity of that group. This modification allows secondary reactions of this functional group that would otherwise not be possible. The concept was introduced by D. Seebach and E.J. Corey. Polarity analysis during retrosynthetic analysis tells a chemist when umpolung tactics are required to synthesize a target molecule.

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. This equation was developed and published by Louis Plack Hammett in 1937 as a follow-up to qualitative observations in his 1935 publication.

The alpha effect refers to the increased nucleophilicity of an atom due to the presence of an adjacent (alpha) atom with lone pair electrons. This first atom does not necessarily exhibit increased basicity compared with a similar atom without an adjacent electron-donating atom, resulting in a deviation from the classical Brønsted-type reactivity-basicity relationship. In other words, the alpha effect refers to nucleophiles presenting higher nucleophilicity than the predicted value obtained from the Brønsted basicity. The representative examples would be high nucleophilicities of hydroperoxide (HO2) and hydrazine (N2H4). The effect is now well established with numerous examples and became an important concept in mechanistic chemistry and biochemistry. However, the origin of the effect is still controversial without a clear winner.

In chemistry, a reaction intermediate or an intermediate is a molecular entity that is formed from the reactants but is consumed in further reactions in stepwise chemical reactions that contain multiple elementary steps. Intermediates are the reaction product of one elementary step, but do not appear in the chemical equation for an overall chemical equation.

An acidity function is a measure of the acidity of a medium or solvent system, usually expressed in terms of its ability to donate protons to a solute. The pH scale is by far the most commonly used acidity function, and is ideal for dilute aqueous solutions. Other acidity functions have been proposed for different environments, most notably the Hammett acidity function, H0, for superacid media and its modified version H for superbasic media. The term acidity function is also used for measurements made on basic systems, and the term basicity function is uncommon.

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:

<span class="mw-page-title-main">Flippin–Lodge angle</span>

The Flippin–Lodge angle is one of two angles used by organic and biological chemists studying the relationship between a molecule's chemical structure and ways that it reacts, for reactions involving "attack" of an electron-rich reacting species, the nucleophile, on an electron-poor reacting species, the electrophile. Specifically, the angles—the Bürgi–Dunitz, , and the Flippin–Lodge, —describe the "trajectory" or "angle of attack" of the nucleophile as it approaches the electrophile, in particular when the latter is planar in shape. This is called a nucleophilic addition reaction and it plays a central role in the biological chemistry taking place in many biosyntheses in nature, and is a central "tool" in the reaction toolkit of modern organic chemistry, e.g., to construct new molecules such as pharmaceuticals. Theory and use of these angles falls into the areas of synthetic and physical organic chemistry, which deals with chemical structure and reaction mechanism, and within a sub-specialty called structure correlation.

In theoretical chemistry, Specific ion Interaction Theory is a theory used to estimate single-ion activity coefficients in electrolyte solutions at relatively high concentrations. It does so by taking into consideration interaction coefficients between the various ions present in solution. Interaction coefficients are determined from equilibrium constant values obtained with solutions at various ionic strengths. The determination of SIT interaction coefficients also yields the value of the equilibrium constant at infinite dilution.

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

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

The White catalyst is a transition metal coordination complex named after the chemist by whom it was first synthesized, M. Christina White, a professor at the University of Illinois. The catalyst has been used in a variety of allylic C-H functionalization reactions of α-olefins. In addition, it has been shown to catalyze oxidative Heck reactions.

<span class="mw-page-title-main">Alpha-silicon effect</span>

Generally speaking, second-row elements such as silicon (Si) are known to stabilize α-carbanions with greater effectiveness than a first-row element, which also means Si could destabilize the α-carbocations. This effect is known as silicon alpha effect. Another term that always associates with silicon alpha effect is the so-called silicon beta effect, which means Si at the β position could support formation of carbocations.

References

  1. 1 2 3 Edwards, J.O. (1954). "Correlation of Relative Rates and Equilibria with a Double Basicity Scale". Journal of the American Chemical Society. 76 (6): 1540–1547. doi:10.1021/ja01635a021.
  2. 1 2 Edwards, J.O. (1956). "Polarizability, Basicity and Nucleophilic Character". Journal of the American Chemical Society. 78 (9): 1819–1820. doi:10.1021/ja01590a012.
  3. 1 2 Harris, J. Milton; McManus, Samuel P. (1987). Nucleophilicity . Vol. 215. American Chemical Society. pp.  7–8. doi:10.1021/ba-1987-0215. ISBN   9780841209527.
  4. Swain, C. Gardner; Scott, Carleton B. (January 1953). "Quantitative Correlation of Relative Rates. Comparison of Hydroxide Ion with Other Nucleophilic Reagents toward Alkyl Halides, Esters, Epoxides and Acyl Halides". Journal of the American Chemical Society. 75 (1): 141–147. doi:10.1021/ja01097a041.
  5. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " Swain–Scott equation ". doi : 10.1351/goldbook.S06201
  6. Carroll, Felix (2010). Perspectives on Structure and Mechanism in Organic Chemistry (2 ed.). New Jersey: Wiley. pp. 506–507. ISBN   978-0470276105.
  7. 1 2 Edwards, J.O.; Pearson, Ralph G. (1962). "The Factors Determining Nucleophilic Reactivities". Journal of the American Chemical Society. 84 (1): 16–24. doi:10.1021/ja00860a005.
  8. Jencks, William P.; Carriuolo, Joan (April 1960). "Reactivity of Nucleophilic Reagents toward Esters". Journal of the American Chemical Society. 82 (7): 1778–1786. doi:10.1021/ja01492a058.
  9. Jencks, William P.; Carriuolo, Joan (February 1960). "General Base Catalysis of the Aminolysis of Phenyl Acetate". Journal of the American Chemical Society. 82 (3): 675–681. doi:10.1021/ja01488a044.
  10. Hudson, Michael J.; Laurence M. Harwood; Dominic M. Laventine; Frank W. Lewis (2013). "Use of Soft Heterocyclic N‑Donor Ligands To Separate Actinides and Lanthanides". Inorganic Chemistry. 52 (7): 3414–3428. doi:10.1021/ic3008848. PMID   22867058.
  11. Pearson, Ralph G. (1963). "Hard and Soft Acids and Bases". J. Am. Chem. Soc. 85 (22): 3533–3539. doi:10.1021/ja00905a001.
  12. Yingst, Austin; MacDaniel, Darl H. (1967). "Use of the Edwards Equation to Determine Hardness of Acids". Inorganic Chemistry. 6 (5): 1067–1068. doi:10.1021/ic50051a051.