ผู้ใช้:ZeroSixTwo/กระบะทราย8

จากวิกิพีเดีย สารานุกรมเสรี

นิวคลีโอไฟล์ (อังกฤษ: nucleophile) เป็นสปีชีส์ทางเคมีที่ให้คู่อิเล็กตรอนเพื่อสร้างพันธะเคมีในการเกิดปฏิกิริยา โมเลกุลหรือไอออนใด ๆ ก็ตามที่มีคู่อิเล็กตรอนที่ไม่ได้สร้างพันธะอย่างน้อยหนึ่งคู่หรือมีพันธะไพอย่างน้อยหนึ่งตำแหน่งสามารถเป็นนิวคลีโอไฟล์ได้ นิวคลีโอไฟล์จัดเป็นเบสตามทฤษฎีกรด–เบสของลิวอิส

ประวัติ[แก้]

คำว่า นิวคลีโอไฟล์ และ อิเล็กโตรไฟล์ ใช้ครั้งแรกโดยคริสโตเฟอร์ เคลก์ อินโกลด์ใน ค.ศ. 1933[1]เพื่อใช้แทนคำว่า แอนไอออนอยด์ (ไอออนลบเทียม) และ แคตไอออนอยด์ (ไอออนบวกเทียม) ที่เสนอโดยอาร์เธอร์ แลปเวิร์ธใน ค.ศ. 1925[2] คำว่านิวคลีโอไฟล์มาจากคำว่านิวเคลียสรวมกับคำอุปสรรค -phile ซึ่งมาจากภาษากรีก φιλος หรือ philos ซึ่งแปลว่าความรักหรือความชอบ

Properties[แก้]

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[แก้]

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[แก้]

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[แก้]

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:

Ritchie equation diazonium ion reactions

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[แก้]

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]

benzhydrylium ions used in the determination of Mayr–Patz equation

and a diverse collection of π-nucleophiles:

Nucleophiles used in the determination of Mayr–Patz equation, X = tetrafluoroborate anion.

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]

Mayr equation also includes SN2 reactions

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[แก้]

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[แก้]

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

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 reversal 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.

Displacement of bromine by a hydroxide

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 แม่แบบ:Sulfur or the แม่แบบ: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[แก้]

While the halogens aren't 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[แก้]

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[แก้]

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[แก้]

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[แก้]

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

See also[แก้]

References[แก้]

  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". Gold Book. IUPAC. February 24, 2014.
  5. Gold Book definition (Ritchie) Link
  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.
  10. An internet database for reactivity parameters maintained by the Mayr group is available at http://www.cup.uni-muenchen.de/oc/mayr/
  11. 11.0 11.1 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