Chemoselectivity

Last updated

Chemoselectivity is the preferential reaction of a chemical reagent with one of two or more different functional groups. [1]

Contents

In a chemoselective system, a reagent in the presence of an aldehyde and an ester would mostly target the aldehyde, even if it has the option to react with the ester. Chemoselectivity is an area of interest in chemistry because scientists want to recreate complex biological compounds, such as natural products, and make specific modifications to them. [2]

Most chemical reactions bring together atoms that have negative charge character and atoms that have positive charge character. [3] When evaluating possible reaction outcomes, several factors should be considered. The most important being identifying where in the molecule has the most electron density and where has the least. [3] This analysis gives a good prediction of reactivity, but more factors such as connectivity, atomic orbital overlap, solvent effects, and the addition of supporting reagents can affect the reaction outcome.  

Electrophilicity

Main page: electrophile

The carbon-bromine bond is more reactive than the carbon-fluorine bond Electrophilicity comparison.png
The carbon-bromine bond is more reactive than the carbon-fluorine bond

If a molecule has several potential reactive sites, the reaction will occur in the most reactive one. When comparing carbon-halogen bonds, lighter halogens such as fluorine and chlorine have a better orbital overlap with carbon, which makes the bond stronger. [4] Bromine and iodine, on the other hand, are bigger and therefore can undergo chemical reactions more easily.

Carbonyl functional groups

Reactivity ranking of common carbonyl functional groups Functional group reactivity ranking.png
Reactivity ranking of common carbonyl functional groups

The reactivity of carbonyls can be ranked by evaluating how much electron density the neighbouring atoms donate to the carbonyl carbon. [3] Aldehydes are the most reactive because the hydrogen next to the carbon is small and only has one electron, and therefore does not provide steric or electronic shielding. By switching the hydrogen for a carbon group, making a ketone, the carbonyl becomes less reactive since the carbon is bigger and has more electrons. The most stable carbonyls are the ones with atoms with lone pairs next to them, such as amides and esters. [4] Since the electrons are not participating in bonding, they are not as restricted and can readily donate to the carbon. Amides are less reactive than esters because oxygen is more electronegative than nitrogen and therefore it concentrates more of the electron density on itself. [3] Chemists take advantage of the stability of amides by using them as protecting groups to shield sites that they don't want to react. [5]

Metal-assisted selectivity

Selective and unselective reduction of a conjugated ketone Luche corrected.png
Selective and unselective reduction of a conjugated ketone

Some reagents have higher affinity with specific functional groups, which can be used to direct reactivity. A famous example is the Luche Reduction, where an oxophilic metal makes the carbonyl of a conjugated ketone more reactive and directs the reducing agent. [6] On the other hand, copper organometallics have high affinity with carbon-carbon multiple bonds and are used for conjugate addition of nucleophiles into a conjugated ketone. [7]

Reducing and oxidizing reagents

Different hydride reagents have different reactivity towards functional groups so they can be selected according to the desired outcome. [8] Examples include the greater relative chemoselectivity of sodium borohydride versus lithium aluminium hydride for the organic reduction of 4-nitro-2-chlorobenzonitrile to the corresponding aniline, 4-amino-2-chlorobenzonitrile. In another example, the compound 4-methoxyacetophenone is oxidized by bleach at the ketone group at high pH (forming the carboxylic acid) and oxidized by EAS (to the aryl chloride) at low pH. [9]

See also

Related Research Articles

<span class="mw-page-title-main">Carboxylic acid</span> Organic compound containing a –C(=O)OH group

In organic chemistry, a carboxylic acid is an organic acid that contains a carboxyl group attached to an R-group. The general formula of a carboxylic acid is often written as R−COOH or R−CO2H, sometimes as R−C(O)OH with R referring to the alkyl, alkenyl, aryl, or other group. Carboxylic acids occur widely. Important examples include the amino acids and fatty acids. Deprotonation of a carboxylic acid gives a carboxylate anion.

<span class="mw-page-title-main">Grignard reaction</span> Organometallic coupling reaction

The Grignard reaction is an organometallic chemical reaction in which, according to the classical definition, carbon alkyl, allyl, vinyl, or aryl magnesium halides are added to the carbonyl groups of either an aldehyde or ketone under anhydrous conditions. This reaction is important for the formation of carbon-carbon bonds.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

<span class="mw-page-title-main">Sodium borohydride</span> Chemical compound

Sodium borohydride, also known as sodium tetrahydridoborate and sodium tetrahydroborate, is an inorganic compound with the formula NaBH4. It is a white crystalline solid, usually encountered as an aqueous basic solution. Sodium borohydride is a reducing agent that finds application in papermaking and dye industries. It is also used as a reagent in organic synthesis.

<span class="mw-page-title-main">Organoboron chemistry</span> Study of compounds containing a boron-carbon bond

Organoboron chemistry or organoborane chemistry studies organoboron compounds, also called organoboranes. These chemical compounds combine boron and carbon; typically, they are organic derivatives of borane (BH3), as in the trialkyl boranes.

<span class="mw-page-title-main">Organic redox reaction</span> Redox reaction that takes place with organic compounds

Organic reductions or organic oxidations or organic redox reactions are redox reactions that take place with organic compounds. In organic chemistry oxidations and reductions are different from ordinary redox reactions, because many reactions carry the name but do not actually involve electron transfer. Instead the relevant criterion for organic oxidation is gain of oxygen and/or loss of hydrogen. Simple functional groups can be arranged in order of increasing oxidation state. The oxidation numbers are only an approximation:

Reductive amination is a form of amination that involves the conversion of a carbonyl group to an amine via an intermediate imine. The carbonyl group is most commonly a ketone or an aldehyde. It is a common method to make amines and is widely used in green chemistry since it can be done catalytically in one-pot under mild conditions. In biochemistry, dehydrogenase enzymes use reductive amination to produce the amino acid, glutamate. Additionally, there is ongoing research on alternative synthesis mechanisms with various metal catalysts which allow the reaction to be less energy taxing, and require milder reaction conditions. Investigation into biocatalysts, such as imine reductases, have allowed for higher selectivity in the reduction of chiral amines which is an important factor in pharmaceutical synthesis.

<span class="mw-page-title-main">Danishefsky Taxol total synthesis</span>

The Danishefsky Taxol total synthesis in organic chemistry is an important third Taxol synthesis published by the group of Samuel Danishefsky in 1996 two years after the first two efforts described in the Holton Taxol total synthesis and the Nicolaou Taxol total synthesis. Combined they provide a good insight in the application of organic chemistry in total synthesis.

α-Halo ketone

In organic chemistry, an α-halo ketone is a functional group consisting of a ketone group or more generally a carbonyl group with an α-halogen substituent. α-Halo ketones are alkylating agents. Prominent α-halo ketones include phenacyl bromide and chloroacetone.

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.

<span class="mw-page-title-main">Sodium cyanoborohydride</span> Chemical compound

Sodium cyanoborohydride is a chemical compound with the formula Na[BH3(CN)]. It is a colourless salt used in organic synthesis for chemical reduction including that of imines and carbonyls. Sodium cyanoborohydride is a milder reductant than other conventional reducing agents.

The Meerwein–Ponndorf–Verley (MPV) reduction in organic chemistry is the reduction of ketones and aldehydes to their corresponding alcohols utilizing aluminium alkoxide catalysis in the presence of a sacrificial alcohol. The advantages of the MPV reduction lie in its high chemoselectivity, and its use of a cheap environmentally friendly metal catalyst. MPV reductions have been described as "obsolete" owing to the development of sodium borohydride and related reagents.

<span class="mw-page-title-main">Organozinc chemistry</span>

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

<span class="mw-page-title-main">Organocopper chemistry</span> Compound with carbon to copper bonds

Organocopper chemistry is the study of the physical properties, reactions, and synthesis of organocopper compounds, which are organometallic compounds containing a carbon to copper chemical bond. They are reagents in organic chemistry.

Luche reduction is the selective organic reduction of α,β-unsaturated ketones to allylic alcohols. The active reductant is described as "cerium borohydride", which is generated in situ from NaBH4 and CeCl3(H2O)7.

<span class="mw-page-title-main">Takai olefination</span>

Takai olefination in organic chemistry describes the organic reaction of an aldehyde with a diorganochromium compound to form an alkene. It is a name reaction, named for Kazuhiko Takai, who first reported it in 1986. In the original reaction, the organochromium species is generated from iodoform or bromoform and an excess of chromium(II) chloride and the product is a vinyl halide. One main advantage of this reaction is the E-configuration of the double bond that is formed. According to the original report, existing alternatives such as the Wittig reaction only gave mixtures.

Alcohol oxidation is a collection of oxidation reactions in organic chemistry that convert alcohols to aldehydes, ketones, carboxylic acids, and esters where the carbon carries a higher oxidation state. The reaction mainly applies to primary and secondary alcohols. Secondary alcohols form ketones, while primary alcohols form aldehydes or carboxylic acids.

<span class="mw-page-title-main">Carbonyl reduction</span> Organic reduction of any carbonyl group by a reducing agent

In organic chemistry, carbonyl reduction is the conversion of any carbonyl group, usually to an alcohol. It is a common transformation that is practiced in many ways. Ketones, aldehydes, carboxylic acids, esters, amides, and acid halides - some of the most pervasive functional groups, -comprise carbonyl compounds. Carboxylic acids, esters, and acid halides can be reduced to either aldehydes or a step further to primary alcohols, depending on the strength of the reducing agent. Aldehydes and ketones can be reduced respectively to primary and secondary alcohols. In deoxygenation, the alcohol group can be further reduced and removed altogether by replacement with H.

<span class="mw-page-title-main">Reductions with samarium(II) iodide</span>

Reductions with samarium(II) iodide involve the conversion of various classes of organic compounds into reduced products through the action of samarium(II) iodide, a mild one-electron reducing agent.

The Buchner–Curtius–Schlotterbeck reaction is the reaction of aldehydes or ketones with aliphatic diazoalkanes to form homologated ketones. It was first described by Eduard Buchner and Theodor Curtius in 1885 and later by Fritz Schlotterbeck in 1907. Two German chemists also preceded Schlotterbeck in discovery of the reaction, Hans von Pechmann in 1895 and Viktor Meyer in 1905. The reaction has since been extended to the synthesis of β-keto esters from the condensation between aldehydes and diazo esters. The general reaction scheme is as follows:

References

  1. "Chemoselectivity (chemoselective)". IUPAC Compendium of Chemical Terminology (3rd ed.). International Union of Pure and Applied Chemistry. 2006. doi:10.1351/goldbook.C01051 via Online version 3.0.1, 2019.
  2. Shenvi, Ryan A.; O’Malley, Daniel P.; Baran, Phil S. (2009-04-21). "Chemoselectivity: The Mother of Invention in Total Synthesis". Accounts of Chemical Research. 42 (4): 530–541. doi:10.1021/ar800182r. ISSN   0001-4842. PMC   2765532 . PMID   19182997.
  3. 1 2 3 4 Zumdahl, Steven S. (1995). Chemical Principles (2nd ed.). Lexinton, MA, USA: HEATH. ISBN   0669393215.
  4. 1 2 Ouellette, Robert J.; Rawn, J. David (1996). Organic chemistry. Upper Saddle River, N.J: Prentice Hall. ISBN   978-0-02-390171-3.
  5. Greene, Theodora W.; Wuts, Peter G. M. (1999). Protective groups in organic synthesis. A Wiley-Interscience publication (3. ed.). New York Weinheim: Wiley. ISBN   978-0-471-16019-9.
  6. Gemal, Andre L.; Luche, Jean Louis (September 1981). "Lanthanoids in organic synthesis. 6. Reduction of .alpha.-enones by sodium borohydride in the presence of lanthanoid chlorides: synthetic and mechanistic aspects". Journal of the American Chemical Society. 103 (18): 5454–5459. doi:10.1021/ja00408a029. ISSN   0002-7863.
  7. Alexakis, A.; Bäckvall, J. E.; Krause, N.; Pàmies, O.; Diéguez, M. (2008-08-01). "Enantioselective Copper-Catalyzed Conjugate Addition and Allylic Substitution Reactions". Chemical Reviews. 108 (8): 2796–2823. doi:10.1021/cr0683515. ISSN   0009-2665. PMID   18671436.
  8. Walker, E. R. H. (1976). "The functional group selectivity of complex hydride reducing agents". Chemical Society Reviews. 5: 23. doi:10.1039/cs9760500023. ISSN   0306-0012.
  9. Ballard, C. Eric (2010-01-12). "pH-Controlled Oxidation of an Aromatic Ketone: Structural Elucidation of the Products of Two Green Chemical Reactions". Journal of Chemical Education. 87 (2): 190–193. Bibcode:2010JChEd..87..190B. doi:10.1021/ed800054s. ISSN   0021-9584.