fbpx
Wikipedia

Nucleophile

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.

A hydroxide ion acting as a nucleophile in an SN2 reaction, converting a halogenoalkane into an alcohol

Nucleophilic describes the affinity of a nucleophile to bond with positively charged atomic nuclei. Nucleophilicity, sometimes referred to as nucleophile strength, refers to a substance's nucleophilic character and is often used to compare the affinity of atoms. Neutral nucleophilic reactions with solvents such as alcohols and water are named solvolysis. Nucleophiles may take part in nucleophilic substitution, whereby a nucleophile becomes attracted to a full or partial positive charge, and nucleophilic addition. Nucleophilicity is closely related to basicity.

History

The terms nucleophile and electrophile were introduced by Christopher Kelk Ingold in 1933,[1] replacing the terms anionoid and cationoid proposed earlier by A. J. Lapworth in 1925.[2] The word nucleophile is derived from nucleus and the Greek word φιλος, philos, meaning friend.

Properties

In general, in a group across the periodic table, the more basic the ion (the higher the pKa of the conjugate acid) the more reactive it is as a nucleophile. Within a series of nucleophiles with the same attacking element (e.g. oxygen), the order of nucleophilicity will follow basicity. Sulfur is in general a better nucleophile than oxygen.

Nucleophilicity

Many schemes attempting to quantify relative nucleophilic strength have been devised. The following empirical data have been obtained by measuring reaction rates for many reactions involving many nucleophiles and electrophiles. Nucleophiles displaying the so-called alpha effect are usually omitted in this type of treatment.

Swain–Scott equation

The first such attempt is found in the Swain–Scott equation[3][4] derived in 1953:

 

This free-energy relationship relates the pseudo first order reaction rate constant (in water at 25 °C), k, of a reaction, normalized to the reaction rate, k0, of a standard reaction with water as the nucleophile, 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 treatment results in the following values for typical nucleophilic anions: acetate 2.7, chloride 3.0, azide 4.0, hydroxide 4.2, aniline 4.5, iodide 5.0, and thiosulfate 6.4. Typical substrate constants are 0.66 for ethyl tosylate, 0.77 for β-propiolactone, 1.00 for 2,3-epoxypropanol, 0.87 for benzyl chloride, and 1.43 for benzoyl chloride.

The equation predicts that, in a nucleophilic displacement on benzyl chloride, the azide anion reacts 3000 times faster than water.

Ritchie equation

The Ritchie equation, derived in 1972, is another free-energy relationship:[5][6][7]

 

where N+ is the nucleophile dependent parameter and k0 the reaction rate constant for water. In this equation, a substrate-dependent parameter like s in the Swain–Scott equation is absent. The equation states that two nucleophiles react with the same relative reactivity regardless of the nature of the electrophile, which is in violation of the reactivity–selectivity principle. For this reason, this equation is also called the constant selectivity relationship.

In the original publication the data were obtained by reactions of selected nucleophiles with selected electrophilic carbocations such as tropylium or diazonium cations:

 

or (not displayed) ions based on malachite green. Many other reaction types have since been described.

Typical Ritchie N+ values (in methanol) are: 0.5 for methanol, 5.9 for the cyanide anion, 7.5 for the methoxide anion, 8.5 for the azide anion, and 10.7 for the thiophenol anion. The values for the relative cation reactivities are −0.4 for the malachite green cation, +2.6 for the benzenediazonium cation, and +4.5 for the tropylium cation.

Mayr–Patz equation

In the Mayr–Patz equation (1994):[8]

 

The second order reaction rate constant k at 20 °C for a reaction is related to a nucleophilicity parameter N, an electrophilicity parameter E, and a nucleophile-dependent slope parameter s. The constant s is defined as 1 with 2-methyl-1-pentene as the nucleophile.

Many of the constants have been derived from reaction of so-called benzhydrylium ions as the electrophiles:[9]

 

and a diverse collection of π-nucleophiles:

 .

Typical E values are +6.2 for R = chlorine, +5.90 for R = hydrogen, 0 for R = methoxy and −7.02 for R = dimethylamine.

Typical N values with s in parenthesis are −4.47 (1.32) for electrophilic aromatic substitution to toluene (1), −0.41 (1.12) for electrophilic addition to 1-phenyl-2-propene (2), and 0.96 (1) for addition to 2-methyl-1-pentene (3), −0.13 (1.21) for reaction with triphenylallylsilane (4), 3.61 (1.11) for reaction with 2-methylfuran (5), +7.48 (0.89) for reaction with isobutenyltributylstannane (6) and +13.36 (0.81) for reaction with the enamine 7.[10]

The range of organic reactions also include SN2 reactions:[11]

 

With E = −9.15 for the S-methyldibenzothiophenium ion, typical nucleophile values N (s) are 15.63 (0.64) for piperidine, 10.49 (0.68) for methoxide, and 5.20 (0.89) for water. In short, nucleophilicities towards sp2 or sp3 centers follow the same pattern.

Unified equation

In an effort to unify the above described equations the Mayr equation is rewritten as:[11]

 

with sE the electrophile-dependent slope parameter and sN the nucleophile-dependent slope parameter. This equation can be rewritten in several ways:

  • with sE = 1 for carbocations this equation is equal to the original Mayr–Patz equation of 1994,
  • with sN = 0.6 for most n nucleophiles the equation becomes
 
or the original Scott–Swain equation written as:
 
  • with sE = 1 for carbocations and sN = 0.6 the equation becomes:
 
or the original Ritchie equation written as:
 

Types

Examples of nucleophiles are anions such as Cl, or a compound with a lone pair of electrons such as NH3 (ammonia) and PR3.

In the example below, the oxygen of the hydroxide ion donates an electron pair to form a new chemical bond with the carbon at the end of the bromopropane molecule. The bond between the carbon and the bromine then undergoes heterolytic fission, with the bromine atom taking the donated electron and becoming the bromide ion (Br), because a SN2 reaction occurs by backside attack. This means that the hydroxide ion attacks the carbon atom from the other side, exactly opposite the bromine ion. Because of this backside attack, SN2 reactions result in a inversion of the configuration of the electrophile. If the electrophile is chiral, it typically maintains its chirality, though the SN2 product's absolute configuration is flipped as compared to that of the original electrophile.

 

An ambident nucleophile is one that can attack from two or more places, resulting in two or more products. For example, the thiocyanate ion (SCN) may attack from either the sulfur or the nitrogen. For this reason, the SN2 reaction of an alkyl halide with SCN often leads to a mixture of an alkyl thiocyanate (R-SCN) and an alkyl isothiocyanate (R-NCS). Similar considerations apply in the Kolbe nitrile synthesis.

Halogens

While the halogens are not nucleophilic in their diatomic form (e.g. I2 is not a nucleophile), their anions are good nucleophiles. In polar, protic solvents, F is the weakest nucleophile, and I the strongest; this order is reversed in polar, aprotic solvents.[12]

Carbon

Carbon nucleophiles are often organometallic reagents such as those found in the Grignard reaction, Blaise reaction, Reformatsky reaction, and Barbier reaction or reactions involving organolithium reagents and acetylides. These reagents are often used to perform nucleophilic additions.

Enols are also carbon nucleophiles. The formation of an enol is catalyzed by acid or base. Enols are ambident nucleophiles, but, in general, nucleophilic at the alpha carbon atom. Enols are commonly used in condensation reactions, including the Claisen condensation and the aldol condensation reactions.

Oxygen

Examples of oxygen nucleophiles are water (H2O), hydroxide anion, alcohols, alkoxide anions, hydrogen peroxide, and carboxylate anions. Nucleophilic attack does not take place during intermolecular hydrogen bonding.

Sulfur

Of sulfur nucleophiles, hydrogen sulfide and its salts, thiols (RSH), thiolate anions (RS), anions of thiolcarboxylic acids (RC(O)-S), and anions of dithiocarbonates (RO-C(S)-S) and dithiocarbamates (R2N-C(S)-S) are used most often.

In general, sulfur is very nucleophilic because of its large size, which makes it readily polarizable, and its lone pairs of electrons are readily accessible.

Nitrogen

Nitrogen nucleophiles include ammonia, azide, amines, nitrites, hydroxylamine, hydrazine, carbazide, phenylhydrazine, semicarbazide, and amide.

Metal centers

Although metal centers (e.g., Li+, Zn2+, Sc3+, etc.) are most commonly cationic and electrophilic (Lewis acidic) in nature, certain metal centers (particularly ones in a low oxidation state and/or carrying a negative charge) are among the strongest recorded nucleophiles and are sometimes referred to as "supernucleophiles." For instance, using methyl iodide as the reference electrophile, Ph3Sn is about 10000 times more nucleophilic than I, while the Co(I) form of vitamin B12 (vitamin B12s) is about 107 times more nucleophilic.[13] Other supernucleophilic metal centers include low oxidation state carbonyl metalate anions (e.g., CpFe(CO)2).[14]

See also

References

  1. ^ Ingold, C. K. (1933). "266. Significance of tautomerism and of the reactions of aromatic compounds in the electronic theory of organic reactions". Journal of the Chemical Society (Resumed): 1120. doi:10.1039/jr9330001120.
  2. ^ Lapworth, A. (1925). "Replaceability of Halogen Atoms by Hydrogen Atoms". Nature. 115: 625.
  3. ^ Quantitative Correlation of Relative Rates. Comparison of Hydroxide Ion with Other Nucleophilic Reagents toward Alkyl Halides, Esters, Epoxides and Acyl Halides C. Gardner Swain, Carleton B. Scott J. Am. Chem. Soc.; 1953; 75(1); 141-147. Abstract
  4. ^ "Swain–Scott equation". The IUPAC Compendium of Chemical Terminology. 2014. doi:10.1351/goldbook.S06201.
  5. ^ "Ritchie equation". The IUPAC Compendium of Chemical Terminology. 2014. doi:10.1351/goldbook.R05402.
  6. ^ Nucleophilic reactivities toward cations Calvin D. Ritchie Acc. Chem. Res.; 1972; 5(10); 348-354. Abstract
  7. ^ Cation–anion combination reactions. XIII. Correlation of the reactions of nucleophiles with esters Calvin D. Ritchie J. Am. Chem. Soc.; 1975; 97(5); 1170–1179. Abstract
  8. ^ Mayr, Herbert; Patz, Matthias (1994). "Scales of Nucleophilicity and Electrophilicity: A System for Ordering Polar Organic and Organometallic Reactions". Angewandte Chemie International Edition in English. 33 (9): 938. doi:10.1002/anie.199409381.
  9. ^ Mayr, Herbert; Bug, Thorsten; Gotta, Matthias F; Hering, Nicole; Irrgang, Bernhard; Janker, Brigitte; Kempf, Bernhard; Loos, Robert; Ofial, Armin R; Remennikov, Grigoriy; Schimmel, Holger (2001). "Reference Scales for the Characterization of Cationic Electrophiles and Neutral Nucleophiles". Journal of the American Chemical Society. 123 (39): 9500–12. doi:10.1021/ja010890y. PMID 11572670. S2CID 8392147.
  10. ^ An internet database for reactivity parameters maintained by the Mayr group is available at http://www.cup.uni-muenchen.de/oc/mayr/
  11. ^ a b Phan, Thanh Binh; Breugst, Martin; Mayr, Herbert (2006). "Towards a General Scale of Nucleophilicity?". Angewandte Chemie International Edition. 45 (23): 3869–74. CiteSeerX 10.1.1.617.3287. doi:10.1002/anie.200600542. PMID 16646102.
  12. ^ Chem 2401 Supplementary Notes. Thompson, Alison and Pincock, James, Dalhousie University Chemistry Department
  13. ^ Schrauzer, G. N.; Deutsch, E.; Windgassen, R. J. (April 1968). "The nucleophilicity of vitamin B(sub 12s)". Journal of the American Chemical Society. 90 (9): 2441–2442. doi:10.1021/ja01011a054. ISSN 0002-7863. PMID 5642073.
  14. ^ Dessy, Raymond E.; Pohl, Rudolph L.; King, R. Bruce (November 1966). "Organometallic Electrochemistry. VII. 1 The Nucleophilicities of Metallic and Metalloidal Anions Derived from Metals of Groups IV, V, VI, VII, and VIII". Journal of the American Chemical Society. 88 (22): 5121–5124. doi:10.1021/ja00974a015. ISSN 0002-7863.

nucleophile, chemistry, nucleophile, chemical, species, that, forms, bonds, donating, electron, pair, molecules, ions, with, free, pair, electrons, least, bond, nucleophiles, because, nucleophiles, donate, electrons, they, lewis, bases, hydroxide, acting, nucl. 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 A hydroxide ion acting as a nucleophile in an SN2 reaction converting a halogenoalkane into an alcohol Nucleophilic describes the affinity of a nucleophile to bond with positively charged atomic nuclei Nucleophilicity sometimes referred to as nucleophile strength refers to a substance s nucleophilic character and is often used to compare the affinity of atoms Neutral nucleophilic reactions with solvents such as alcohols and water are named solvolysis Nucleophiles may take part in nucleophilic substitution whereby a nucleophile becomes attracted to a full or partial positive charge and nucleophilic addition Nucleophilicity is closely related to basicity Contents 1 History 2 Properties 2 1 Nucleophilicity 2 1 1 Swain Scott equation 2 1 2 Ritchie equation 2 1 3 Mayr Patz equation 2 1 4 Unified equation 3 Types 3 1 Halogens 3 2 Carbon 3 3 Oxygen 3 4 Sulfur 3 5 Nitrogen 3 6 Metal centers 4 See also 5 ReferencesHistory EditThe terms nucleophile and electrophile were introduced by Christopher Kelk Ingold in 1933 1 replacing the terms anionoid and cationoid proposed earlier by A J Lapworth in 1925 2 The word nucleophile is derived from nucleus and the Greek word filos philos meaning friend Properties EditIn general in a group across the periodic table the more basic the ion the higher the pKa of the conjugate acid the more reactive it is as a nucleophile Within a series of nucleophiles with the same attacking element e g oxygen the order of nucleophilicity will follow basicity Sulfur is in general a better nucleophile than oxygen Nucleophilicity Edit Many schemes attempting to quantify relative nucleophilic strength have been devised The following empirical data have been obtained by measuring reaction rates for many reactions involving many nucleophiles and electrophiles Nucleophiles displaying the so called alpha effect are usually omitted in this type of treatment Swain Scott equation Edit The first such attempt is found in the Swain Scott equation 3 4 derived in 1953 log 10 k k 0 s n displaystyle log 10 left frac k k 0 right sn This free energy relationship relates the pseudo first order reaction rate constant in water at 25 C k of a reaction normalized to the reaction rate k0 of a standard reaction with water as the nucleophile 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 treatment results in the following values for typical nucleophilic anions acetate 2 7 chloride 3 0 azide 4 0 hydroxide 4 2 aniline 4 5 iodide 5 0 and thiosulfate 6 4 Typical substrate constants are 0 66 for ethyl tosylate 0 77 for b propiolactone 1 00 for 2 3 epoxypropanol 0 87 for benzyl chloride and 1 43 for benzoyl chloride The equation predicts that in a nucleophilic displacement on benzyl chloride the azide anion reacts 3000 times faster than water Ritchie equation Edit The Ritchie equation derived in 1972 is another free energy relationship 5 6 7 log 10 k k 0 N displaystyle log 10 left frac k k 0 right N where N is the nucleophile dependent parameter and k0 the reaction rate constant for water In this equation a substrate dependent parameter like s in the Swain Scott equation is absent The equation states that two nucleophiles react with the same relative reactivity regardless of the nature of the electrophile which is in violation of the reactivity selectivity principle For this reason this equation is also called the constant selectivity relationship In the original publication the data were obtained by reactions of selected nucleophiles with selected electrophilic carbocations such as tropylium or diazonium cations or not displayed ions based on malachite green Many other reaction types have since been described Typical Ritchie N values in methanol are 0 5 for methanol 5 9 for the cyanide anion 7 5 for the methoxide anion 8 5 for the azide anion and 10 7 for the thiophenol anion The values for the relative cation reactivities are 0 4 for the malachite green cation 2 6 for the benzenediazonium cation and 4 5 for the tropylium cation Mayr Patz equation Edit In the Mayr Patz equation 1994 8 log k s N E displaystyle log k s N E The second order reaction rate constant k at 20 C for a reaction is related to a nucleophilicity parameter N an electrophilicity parameter E and a nucleophile dependent slope parameter s The constant s is defined as 1 with 2 methyl 1 pentene as the nucleophile Many of the constants have been derived from reaction of so called benzhydrylium ions as the electrophiles 9 and a diverse collection of p nucleophiles Typical E values are 6 2 for R chlorine 5 90 for R hydrogen 0 for R methoxy and 7 02 for R dimethylamine Typical N values with s in parenthesis are 4 47 1 32 for electrophilic aromatic substitution to toluene 1 0 41 1 12 for electrophilic addition to 1 phenyl 2 propene 2 and 0 96 1 for addition to 2 methyl 1 pentene 3 0 13 1 21 for reaction with triphenylallylsilane 4 3 61 1 11 for reaction with 2 methylfuran 5 7 48 0 89 for reaction with isobutenyltributylstannane 6 and 13 36 0 81 for reaction with the enamine 7 10 The range of organic reactions also include SN2 reactions 11 With E 9 15 for the S methyldibenzothiophenium ion typical nucleophile values N s are 15 63 0 64 for piperidine 10 49 0 68 for methoxide and 5 20 0 89 for water In short nucleophilicities towards sp2 or sp3 centers follow the same pattern Unified equation Edit In an effort to unify the above described equations the Mayr equation is rewritten as 11 log k s E s N N E displaystyle log k s E s N N E with sE the electrophile dependent slope parameter and sN the nucleophile dependent slope parameter This equation can be rewritten in several ways with sE 1 for carbocations this equation is equal to the original Mayr Patz equation of 1994 with sN 0 6 for most n nucleophiles the equation becomeslog k 0 6 s E N 0 6 s E E displaystyle log k 0 6s E N 0 6s E E dd or the original Scott Swain equation written as log k log k 0 s E N displaystyle log k log k 0 s E N dd with sE 1 for carbocations and sN 0 6 the equation becomes log k 0 6 N 0 6 E displaystyle log k 0 6N 0 6E dd or the original Ritchie equation written as log k log k 0 N displaystyle log k log k 0 N dd Types EditExamples of nucleophiles are anions such as Cl or a compound with a lone pair of electrons such as NH3 ammonia and PR3 In the example below the oxygen of the hydroxide ion donates an electron pair to form a new chemical bond with the carbon at the end of the bromopropane molecule The bond between the carbon and the bromine then undergoes heterolytic fission with the bromine atom taking the donated electron and becoming the bromide ion Br because a SN2 reaction occurs by backside attack This means that the hydroxide ion attacks the carbon atom from the other side exactly opposite the bromine ion Because of this backside attack SN2 reactions result in a inversion of the configuration of the electrophile If the electrophile is chiral it typically maintains its chirality though the SN2 product s absolute configuration is flipped as compared to that of the original electrophile An ambident nucleophile is one that can attack from two or more places resulting in two or more products For example the thiocyanate ion SCN may attack from either the sulfur or the nitrogen For this reason the SN2 reaction of an alkyl halide with SCN often leads to a mixture of an alkyl thiocyanate R SCN and an alkyl isothiocyanate R NCS Similar considerations apply in the Kolbe nitrile synthesis Halogens Edit While the halogens are not nucleophilic in their diatomic form e g I2 is not a nucleophile their anions are good nucleophiles In polar protic solvents F is the weakest nucleophile and I the strongest this order is reversed in polar aprotic solvents 12 Carbon Edit Carbon nucleophiles are often organometallic reagents such as those found in the Grignard reaction Blaise reaction Reformatsky reaction and Barbier reaction or reactions involving organolithium reagents and acetylides These reagents are often used to perform nucleophilic additions Enols are also carbon nucleophiles The formation of an enol is catalyzed by acid or base Enols are ambident nucleophiles but in general nucleophilic at the alpha carbon atom Enols are commonly used in condensation reactions including the Claisen condensation and the aldol condensation reactions Oxygen Edit Examples of oxygen nucleophiles are water H2O hydroxide anion alcohols alkoxide anions hydrogen peroxide and carboxylate anions Nucleophilic attack does not take place during intermolecular hydrogen bonding Sulfur Edit Of sulfur nucleophiles hydrogen sulfide and its salts thiols RSH thiolate anions RS anions of thiolcarboxylic acids RC O S and anions of dithiocarbonates RO C S S and dithiocarbamates R2N C S S are used most often In general sulfur is very nucleophilic because of its large size which makes it readily polarizable and its lone pairs of electrons are readily accessible Nitrogen Edit Nitrogen nucleophiles include ammonia azide amines nitrites hydroxylamine hydrazine carbazide phenylhydrazine semicarbazide and amide Metal centers Edit Although metal centers e g Li Zn2 Sc3 etc are most commonly cationic and electrophilic Lewis acidic in nature certain metal centers particularly ones in a low oxidation state and or carrying a negative charge are among the strongest recorded nucleophiles and are sometimes referred to as supernucleophiles For instance using methyl iodide as the reference electrophile Ph3Sn is about 10000 times more nucleophilic than I while the Co I form of vitamin B12 vitamin B12s is about 107 times more nucleophilic 13 Other supernucleophilic metal centers include low oxidation state carbonyl metalate anions e g CpFe CO 2 14 See also EditElectrophile A chemical species that accepts an electron pair from a nucleophile Lewis acids and bases Chemical bond theory Nucleophilic abstraction Type of organometallic reaction Addition to pi ligands Organometallic chemistry rulePages displaying short descriptions of redirect targetsReferences Edit Ingold C K 1933 266 Significance of tautomerism and of the reactions of aromatic compounds in the electronic theory of organic reactions Journal of the Chemical Society Resumed 1120 doi 10 1039 jr9330001120 Lapworth A 1925 Replaceability of Halogen Atoms by Hydrogen Atoms Nature 115 625 Quantitative Correlation of Relative Rates Comparison of Hydroxide Ion with Other Nucleophilic Reagents toward Alkyl Halides Esters Epoxides and Acyl Halides C Gardner Swain Carleton B Scott J Am Chem Soc 1953 75 1 141 147 Abstract Swain Scott equation The IUPAC Compendium of Chemical Terminology 2014 doi 10 1351 goldbook S06201 Ritchie equation The IUPAC Compendium of Chemical Terminology 2014 doi 10 1351 goldbook R05402 Nucleophilic reactivities toward cations Calvin D Ritchie Acc Chem Res 1972 5 10 348 354 Abstract Cation anion combination reactions XIII Correlation of the reactions of nucleophiles with esters Calvin D Ritchie J Am Chem Soc 1975 97 5 1170 1179 Abstract Mayr Herbert Patz Matthias 1994 Scales of Nucleophilicity and Electrophilicity A System for Ordering Polar Organic and Organometallic Reactions Angewandte Chemie International Edition in English 33 9 938 doi 10 1002 anie 199409381 Mayr Herbert Bug Thorsten Gotta Matthias F Hering Nicole Irrgang Bernhard Janker Brigitte Kempf Bernhard Loos Robert Ofial Armin R Remennikov Grigoriy Schimmel Holger 2001 Reference Scales for the Characterization of Cationic Electrophiles and Neutral Nucleophiles Journal of the American Chemical Society 123 39 9500 12 doi 10 1021 ja010890y PMID 11572670 S2CID 8392147 An internet database for reactivity parameters maintained by the Mayr group is available at http www cup uni muenchen de oc mayr a b Phan Thanh Binh Breugst Martin Mayr Herbert 2006 Towards a General Scale of Nucleophilicity Angewandte Chemie International Edition 45 23 3869 74 CiteSeerX 10 1 1 617 3287 doi 10 1002 anie 200600542 PMID 16646102 Chem 2401 Supplementary Notes Thompson Alison and Pincock James Dalhousie University Chemistry Department Schrauzer G N Deutsch E Windgassen R J April 1968 The nucleophilicity of vitamin B sub 12s Journal of the American Chemical Society 90 9 2441 2442 doi 10 1021 ja01011a054 ISSN 0002 7863 PMID 5642073 Dessy Raymond E Pohl Rudolph L King R Bruce November 1966 Organometallic Electrochemistry VII 1 The Nucleophilicities of Metallic and Metalloidal Anions Derived from Metals of Groups IV V VI VII and VIII Journal of the American Chemical Society 88 22 5121 5124 doi 10 1021 ja00974a015 ISSN 0002 7863 Retrieved from https en wikipedia org w index php title Nucleophile amp oldid 1124630623, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.