Chiral derivatizing agent

Last updated
(R)-a-methoxy-a-(trifluoromethyl)- phenylacetic acid (Mosher's acid) R-MTPA.png
(R)-α-methoxy-α-(trifluoromethyl)- phenylacetic acid (Mosher's acid)

In analytical chemistry, a chiral derivatizing agent (CDA), also known as a chiral resolving reagent, is a derivatization reagent that is a chiral auxiliary used to convert a mixture of enantiomers into diastereomers in order to analyze the quantities of each enantiomer present and determine the optical purity of a sample. Analysis can be conducted by spectroscopy or by chromatography. Some analytical techniques such as HPLC and NMR, in their most commons forms, cannot distinguish enantiomers within a sample, but can distinguish diastereomers. Therefore, converting a mixture of enantiomers to a corresponding mixture of diastereomers can allow analysis. The use of chiral derivatizing agents has declined with the popularization of chiral HPLC. Besides analysis, chiral derivatization is also used for chiral resolution, the actual physical separation of the enantiomers.

Contents

History

Since NMR spectroscopy has been available to chemists, there have been numerous studies on the applications of this technique. One of these noted the difference in the chemical shift (i.e. the distance between the peaks) of two diastereomers. [1] Conversely, two compounds that are enantiomers have the same NMR spectral properties. It was reasoned that if a mix of enantiomers could be converted into a mix of diastereomers by bonding them to another chemical that was itself chiral, it would be possible to distinguish this new mixture using NMR, and therefore learn about the original enantiomeric mixture. The first popular example of this technique was published in 1969 by Harry S. Mosher. The chiral agent used was a single enantiomer of MTPA (α-methoxy-α-(trifluoromethyl)phenylacetic acid), also known as Mosher's acid. [2] The corresponding acid chloride is also known as Mosher's acid chloride, and the resultant diastereomeric esters are known as Mosher's esters. Another system is Pirkle's Alcohol developed in 1977.

Requirements

The general use and design of CDAs obey the following rules so that the CDA can effectively determine the stereochemistry of an analyte: [3]

  1. The CDA must be enantiomerically pure, or (less satisfactorily) its enantiomeric purity must be accurately known.
  2. The reaction of the CDA with both enantiomers should go to completion under reaction conditions. This acts to avoid enrichment or depletion of one enantiomer of the analyte by kinetic resolution.
  3. CDA must not racemize under derivatization or analysis conditions. Its attachment should be mild enough so that the substrate does not racemize either. If analysis is completed by HPLC, the CDA must contain a chromophore to enhance detectability.
  4. If analysis is completed by NMR, the CDA should have a functional group that gives a singlet in the resultant NMR spectrum, where the singlet must be remote from other peaks.

Mosher's method

Mosher's acid, via its acid chloride derivative, reacts readily with alcohols and amines to give esters and amides, respectively. The lack of an alpha-proton on the acid prevents loss of stereochemical fidelity under the reaction conditions. Thus, using an enantiomerically pure Mosher's acid allows for determination of the configuration of simple chiral amines and alcohols. [4] For example, the (R)- and (S)-enantiomers of 1-phenylethanol react with (S)-Mosher acid chloride to yield (R,S)- and (S,S)-diastereomers, respectively, that are distinguishable in NMR. [5]

CFNA (alternative to Mosher's acid)

A newer chiral derivatizing agent (CDA), α-cyano-α-fluoro (2-naphthyl)-acetic acid (2-CFNA) was prepared in optically pure form by the chiral HPLC separation of a racemic 2-CFNA methyl ester. This ester was obtained by fluorination of methyl α-cyano (2-naphthyl) acetate with FClO3. 2-CFNA has been shown to be a superior CDA than Mosher's agent to determine the enantiomeric excess of a primary alcohol. [6]

Chromatography using CDAs

Amide on silica used as the model compound for Helmchen's Postulates. Amide on silica.png
Amide on silica used as the model compound for Helmchen's Postulates.

Upon reaction of a CDA with the target analyte, chromatography can be used to separate the resulting products. In general, chromatography can be used to separate chiral compounds to bypass difficult crystallizations and/or to collect all diastereomer pairs in solution. Chromatography also has many variations (e.g. HPLC, Gas Chromatography, flash chromatography) with a wide array of applicability to diverse categories of molecules. The ability for CDAs to separate chiral molecules is dependent on two major mechanisms of chromatography: [7]

  1. Differential solvation in the mobile phase
  2. Differential adsorption to the stationary phase

Helmchen's postulates

Helmchen's Postulates [8] [9] are the theoretical models used to predict the elution order and extent of separation of diastereomers (including those formed from CDAs) that are adsorbed onto a surface. Although Helmchen's postulates are specific for amides on silica gel using liquid chromatography, the postulates provide fundamental guidelines for other molecules. Helmchen's Postulates are:

  1. Conformations are the same in solution and when adsorbed.
  2. Diastereomers bind to surfaces (silica gel in normal phase chromatography) mainly with hydrogen bonding.
  3. Significant resolution of diastereomers is only expected when molecules can adsorb to silica through two contact points (two hydrogen bonds). This interaction can be perturbed by substituents.
  4. Diastereomers with bulky substituents on the alpha carbon (R2) and on the nitrogen (R1) can shield the hydrogen bonding with the surface, thus the molecule will be eluted before similar molecules with smaller substituents.

Helmchen's postulates have been proven to be applicable to other functional groups such as: carbamates, [7] esters, [10] and epoxides. [11]

Chiral stationary phases

Stationary phases can react with CDAs to form chiral stationary phases which can resolve chiral molecules. [12] By reacting with alcohols on a silicate stationary phase, CDAs add a chiral center to the stationary phase, which allows for the separation of chiral molecules.

CDAs in NMR spectroscopy

CDAs are used with NMR spectroscopic analysis to determine enantiomeric excess and the absolute configuration of a substrate. Chiral discriminating agents are sometimes difficult to distinguish from chiral solvating agents (CSA) and some agents can be used as both. The speed of the exchange between the substrate and the metal center is the most important determining factor to differentiate between the use of a compound as a CDA or CSA. Generally, a CDA has a slow exchange whereas a CSA has a fast exchange. [13] CDAs are more widely used than CSAs to determine absolute configurations because the covalent bonding to the substrate and auxiliary reagent produce species with greater conformational rigidity which creates greater differences in the NMR spectra. [14] CDAs and CSAs can be used together to improve chiral recognition, although this is not a common.

NMR shift reagents such as EuFOD, Pirkle's alcohol, and TRISPHAT take advantage of the formation of diastereomeric complexes between the shift reagent and the analytical sample. [15]

Primary concerns when using CDAs

The primary concerns to take into consideration when using a CDA in NMR spectroscopy are kinetic resolution, racemization during the derivatization reaction and that the reagent should have 100% optical purity. Kinetic resolution is especially significant when determining optical purity, but it is somewhat negligible when the CDA is being used to assign the absolute configuration of an optically pure substrate. [13] Kinetic resolution can be overcome using excess of the CDA. [16] Racemization can occur to either the CDA or the substrate and in both cases it has the potential to significantly affect the results.

Strategies for NMR analysis

The two basic methods of NMR analysis are single- and double-derivatization. Double-derivatization is generally considered more accurate, but single-derivatization usually requires less reagents and, thus, is more cost effective.

Single-derivatization methods

The NMR spectrum of the product formed from the reaction of the substrate with a CDA at room temperature is compared with one of the following: [14]

  1. the spectrum for the same derivative when registered at lower temperature
  2. the spectrum of the same derivative after forming a complex with a metal salt
  3. the spectrum of the substrate without derivatization

Double-derivatization methods

Either the enantiomer of the substrate is derivatized with two enantiomers of the CDA or both enantiomers of the substrate are derivatized with one enantiomer of the CDA. Two diastereomers form in both cases and the chemical shifts of their nuclei are evaluated to assign the configuration of the substrate. [16]

NMR techniques

The most common NMR techniques used when discriminating chiral compounds are 1H-NMR, 19F-NMR and 13C-NMR. 1H-NMR is the primary technique used to assign absolute configuration. 19F-NMR is almost exclusive applied to optical purity studies, and 13C-NMR is primarily used to characterize substrates that do not have protons that are directly bonded to an asymmetrical carbon atom. [14]

Related Research Articles

<span class="mw-page-title-main">Stereoisomerism</span> When molecules have the same atoms and bond structure but differ in 3D orientation

In stereochemistry, stereoisomerism, or spatial isomerism, is a form of isomerism in which molecules have the same molecular formula and sequence of bonded atoms (constitution), but differ in the three-dimensional orientations of their atoms in space. This contrasts with structural isomers, which share the same molecular formula, but the bond connections or their order differs. By definition, molecules that are stereoisomers of each other represent the same structural isomer.

<span class="mw-page-title-main">Enantiomer</span> Stereoisomers which are non-superposable mirror images of each other

In chemistry, an enantiomer – also called optical isomer, antipode, or optical antipode – is one of two stereoisomers that are non-superposable onto their own mirror image. Enantiomers are much like one's right and left hands; without mirroring one of them, hands cannot be superposed onto each other. No amount of reorientation in three spatial dimensions will allow the four unique groups on the chiral carbon to line up exactly. The number of stereoisomers a molecule has can be determined by the number of chiral carbons it has. Stereoisomers include both enantiomers and diastereomers.

<span class="mw-page-title-main">Diastereomer</span> Molecules which are non-mirror image, non-identical stereoisomers

In stereochemistry, diastereomers are a type of stereoisomer. Diastereomers are defined as non-mirror image, non-identical stereoisomers. Hence, they occur when two or more stereoisomers of a compound have different configurations at one or more of the equivalent (related) stereocenters and are not mirror images of each other. When two diastereoisomers differ from each other at only one stereocenter, they are epimers. Each stereocenter gives rise to two different configurations and thus typically increases the number of stereoisomers by a factor of two.

<span class="mw-page-title-main">Chirality (chemistry)</span> Geometric property of some molecules and ions

In chemistry, a molecule or ion is called chiral if it cannot be superposed on its mirror image by any combination of rotations, translations, and some conformational changes. This geometric property is called chirality. The terms are derived from Ancient Greek χείρ (cheir) 'hand'; which is the canonical example of an object with this property.

<span class="mw-page-title-main">Enantioselective synthesis</span> Chemical reaction(s) which favor one chiral isomer over another

Enantioselective synthesis, also called asymmetric synthesis, is a form of chemical synthesis. It is defined by IUPAC as "a chemical reaction in which one or more new elements of chirality are formed in a substrate molecule and which produces the stereoisomeric products in unequal amounts."

In chemistry, stereoselectivity is the property of a chemical reaction in which a single reactant forms an unequal mixture of stereoisomers during a non-stereospecific creation of a new stereocenter or during a non-stereospecific transformation of a pre-existing one. The selectivity arises from differences in steric and electronic effects in the mechanistic pathways leading to the different products. Stereoselectivity can vary in degree but it can never be total since the activation energy difference between the two pathways is finite: both products are at least possible and merely differ in amount. However, in favorable cases, the minor stereoisomer may not be detectable by the analytic methods used.

In stereochemistry, enantiomeric excess (ee) is a measurement of purity used for chiral substances. It reflects the degree to which a sample contains one enantiomer in greater amounts than the other. A racemic mixture has an ee of 0%, while a single completely pure enantiomer has an ee of 100%. A sample with 70% of one enantiomer and 30% of the other has an ee of 40%.

Derivatization is a technique used in chemistry which converts a chemical compound into a product of similar chemical structure, called a derivative.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

Chiral column chromatography is a variant of column chromatography that is employed for the separation of chiral compounds, i.e. enantiomers, in mixtures such as racemates or related compounds. The chiral stationary phase (CSP) is made of a support, usually silica based, on which a chiral reagent or a macromolecule with numerous chiral centers is bonded or immobilized.

In organic chemistry, kinetic resolution is a means of differentiating two enantiomers in a racemic mixture. In kinetic resolution, two enantiomers react with different reaction rates in a chemical reaction with a chiral catalyst or reagent, resulting in an enantioenriched sample of the less reactive enantiomer. As opposed to chiral resolution, kinetic resolution does not rely on different physical properties of diastereomeric products, but rather on the different chemical properties of the racemic starting materials. The enantiomeric excess (ee) of the unreacted starting material continually rises as more product is formed, reaching 100% just before full completion of the reaction. Kinetic resolution relies upon differences in reactivity between enantiomers or enantiomeric complexes.

Chiral resolution, or enantiomeric resolution, is a process in stereochemistry for the separation of racemic mixture into their enantiomers. It is an important tool in the production of optically active compounds, including drugs. Another term with the same meaning is optical resolution.

Mosher's acid, or α-methoxy-α-trifluoromethylphenylacetic acid (MTPA) is a carboxylic acid which was first used by Harry Stone Mosher as a chiral derivatizing agent. It is a chiral molecule, consisting of R and S enantiomers.

<span class="mw-page-title-main">Pirkle's alcohol</span> Chemical compound

Pirkle's alcohol is an off-white, crystalline solid that is stable at room temperature when protected from light and oxygen. This chiral molecule is typically used, in nonracemic form, as a chiral shift reagent in nuclear magnetic resonance spectroscopy, in order to simultaneously determine absolute configuration and enantiomeric purity of other chiral molecules. The molecule is named after William H. Pirkle, Professor of Chemistry at the University of Illinois whose group reported its synthesis and its application as a chiral shift reagent.

<span class="mw-page-title-main">Diastereomeric recrystallization</span>

Diastereomeric recrystallisation is a method of chiral resolution of enantiomers from a racemic mixture. It differs from asymmetric synthesis, which aims to produce a single enantiomer from the beginning, in that diastereomeric recrystallisation separates two enantiomers that have already mixed into a single solution. The strategy of diastereomeric recrystallisation involves two steps. The first step is to convert the enantiomers into diastereomers by way of a chemical reaction. A mixture of enantiomers may contain two isomers of a molecule with one chiral center. After adding a second chiral center in a determined location, the two isomers are still different, but they are no longer mirror images of each other; rather, they become diastereomers.

Chiral Lewis acids (CLAs) are a type of Lewis acid catalyst. These acids affect the chirality of the substrate as they react with it. In such reactions, synthesis favors the formation of a specific enantiomer or diastereomer. The method is an enantioselective asymmetric synthesis reaction. Since they affect chirality, they produce optically active products from optically inactive or mixed starting materials. This type of preferential formation of one enantiomer or diastereomer over the other is formally known as asymmetric induction. In this kind of Lewis acid, the electron-accepting atom is typically a metal, such as indium, zinc, lithium, aluminium, titanium, or boron. The chiral-altering ligands employed for synthesizing these acids often have multiple Lewis basic sites that allow the formation of a ring structure involving the metal atom.

Nuclear magnetic resonance spectroscopy of stereoisomers most commonly known as NMR spectroscopy of stereoisomers is a chemical analysis method that uses NMR spectroscopy to determine the absolute configuration of stereoisomers. For example, the cis or trans alkenes, R or S enantiomers, and R,R or R,S diastereomers.

<span class="mw-page-title-main">Emanuel Gil-Av</span> Russian-Israeli chemist

Emanuel Gil-Av (Zimkin) was an Israeli chemist. The main emphasis of his work constituted chiral chromatography for the analytical separation of enantiomers.

Chemical compounds that come as mirror-image pairs are referred to by chemists as chiral or handed molecules. Each twin is called an enantiomer. Drugs that exhibit handedness are referred to as chiral drugs. Chiral drugs that are equimolar (1:1) mixture of enantiomers are called racemic drugs and these are obviously devoid of optical rotation. The most commonly encountered stereogenic unit, that confers chirality to drug molecules are stereogenic center. Stereogenic center can be due to the presence of tetrahedral tetra coordinate atoms (C,N,P) and pyramidal tricoordinate atoms (N,S). The word chiral describes the three-dimensional architecture of the molecule and does not reveal the stereochemical composition. Hence "chiral drug" does not say whether the drug is racemic, single enantiomer or some other combination of stereoisomers. To resolve this issue Joseph Gal introduced a new term called unichiral. Unichiral indicates that the stereochemical composition of a chiral drug is homogenous consisting of a single enantiomer.

Chiral analysis refers to the quantification of component enantiomers of racemic drug substances or pharmaceutical compounds. Other synonyms commonly used include enantiomer analysis, enantiomeric analysis, and enantioselective analysis. Chiral analysis includes all analytical procedures focused on the characterization of the properties of chiral drugs. Chiral analysis is usually performed with chiral separation methods where the enantiomers are separated on an analytical scale and simultaneously assayed for each enantiomer.

References

  1. J. L. Mateos and D. J. Cram (1959). "Studies in Stereochemistry. XXXI. Conformation, Configuration and Physical Properties of Open-chain Diastereomers". J. Am. Chem. Soc. 81 (11): 2756–2762. doi:10.1021/ja01520a037.
  2. J. A. Dale, D. L. Dull and H. S. Mosher (1969). "α-Methoxy-α-trifluoromethylphenylacetic acid, a versatile reagent for the determination of enantiomeric composition of alcohols and amines". J. Org. Chem. 34 (9): 2543–2549. doi:10.1021/jo01261a013.
  3. Gawley, Robert E.; Aubé, Jeffrey (2012). Principles of Asymmetric Synthesis.
  4. D. Parker (1991). "NMR determination of enantiomeric purity". Chem. Rev. 91 (7): 1441–1457. doi:10.1021/cr00007a009.
  5. Stereochemistry and Chiral Derivatizing Agents (PDF)
  6. New efficient derivatizing agent, alpha-cyano-alpha-fluoro(2-naphthyl)acetic acid (2-CFNA). application to the EE determination of (−)-3-acetoxy-2-fluoro-2-(hexadecyloxymethyl)propan-1-ol., Toyama, Japan: Toyama Medical & Pharmaceutical University, 2000, archived from the original on July 5, 2013
  7. 1 2 Pirkle, W. H.; J. R. Hauske (1977). "Broad spectrum methods for the resolution of optical isomers. A discussion of the reasons underlying the chromatographic separability of some diastereomeric carbamates". J. Org. Chem. 42 (11): 1839. doi:10.1021/jo00431a004.
  8. Helmchen, G.; K. Sauber; R. Ott (1972). "Gezielte Trennung und absolute Konfiguration von enantiomeren Carbonsäuren und Aminen (1. Mitteilung) Günter Helmchen". Tetrahedron Letters. 13 (37): 3873. doi:10.1016/s0040-4039(01)94184-x.
  9. Helmchen, G.; G. Nill; D. Flockerzi; W. Schuhle; M.S.K. Youssef (1979). "Preparative Scale Directed Resolution of Enantiomeric Aminesvia Liquid Chromatography of Diastereomeric 4-Hydroxybutyramides". Angew. Chem. Int. Ed. Engl. 18 (1): 62. doi:10.1002/anie.197900651.
  10. Pirkle, W. H.; J. R. Hauske (1977). "Design of chiral derivatizing agents for the chromatographic resolution of optical isomers. Asymmetric synthesis of some chiral fluoroalkylated amines". J. Org. Chem. 42 (14): 2436. doi:10.1021/jo00434a019.
  11. Pirkle, W.H.; P.L. Rinaldi (1979). "Synthesis and enantiomeric purity determination of the optically active epoxide disparlure, sex pheromone of the gypsy moth". J. Org. Chem. 44 (7): 1025. doi:10.1021/jo01321a001.
  12. Blaschke, G. (1980). "Chromatographic Resolution of Racemates. New analytical methods (17)". Angew. Chem. Int. Ed. Engl. 19 (1): 13. doi:10.1002/anie.198000131.
  13. 1 2 Wenzel, Thomas J. Discrimination of Chiral Compounds Using NMR Spectroscopy. John Wiley & Sons, Inc. pp. 1–7.
  14. 1 2 3 J. M. Seco; E. Quiñoá; R. Riguera* (June 2012). "Assignment of the Absolute Configuration of Polyfunctional Compounds by NMR Using Chiral Derivatizing Agents". Chemical Reviews. 112 (8): 4603–4641. doi:10.1021/cr2003344. PMID   22658125.
  15. Sastri, V.S; Bünzli, Jean-Claude; Rao, V. Ramachandra; Rayudu, G.V.S; Perumareddi, J.R (2003). "Lanthanide Nmr Shift Reagents". Modern Aspects of Rare Earths and Their Complexes. pp. 779–843. doi:10.1016/B978-044451010-5/50024-9. ISBN   9780444510105.
  16. 1 2 Katarzyna M. Błażewskaa; Tadeusz Gajda (July 2009). "Assignment of the absolute configuration of hydroxy- and aminophosphonates by NMR spectroscopy". Tetrahedron: Asymmetry. 20 (12): 1337–1361. doi:10.1016/j.tetasy.2009.05.021.