Atropisomer

Last updated
Atropisomers of 6,6'-dinitro-2,2'-diphenic acid were first experimentally described case, by Christie and Kenner (1922). Atropisomer.svg
Atropisomers of 6,6'-dinitro-2,2'-diphenic acid were first experimentally described case, by Christie and Kenner (1922).

Atropisomers are stereoisomers arising because of hindered rotation about a single bond, where energy differences due to steric strain or other contributors create a barrier to rotation that is high enough to allow for isolation of individual conformers. [1] [2] They occur naturally and are important in pharmaceutical design. [3] When the substituents are achiral, these conformers are enantiomers (atropoenantiomers), showing axial chirality; otherwise they are diastereomers (atropodiastereomers). [1]

Contents

Etymology and history

The word atropisomer (Greek : ἄτροπος, atropos , meaning "not to be turned") was coined in application to a theoretical concept by German biochemist Richard Kuhn for Karl Freudenberg's seminal Stereochemie volume in 1933. [4] Atropisomerism was first experimentally detected in a tetra substituted biphenyl, a diacid, by George Christie and James Kenner in 1922. [5] Michinori Ōki further refined the definition of atropisomers taking into account the temperature-dependence associated with the interconversion of conformers, specifying that atropisomers interconvert with a half-life of at least 1000 seconds at a given temperature, corresponding to an energy barrier of 93 kJ mol−1 (22 kcal mol −1) at 300 K (27 °C). [6] [7]

Energetics

The stability of individual atropisomers is conferred by the repulsive interactions that inhibit rotation. Both the steric bulk and, in principle, the length and rigidity of the bond connecting the two subunits contribute. [1] [7] Commonly, atropisomerism is studied by dynamic nuclear magnetic resonance spectroscopy, since atropisomerism is a form of fluxionality. [7] Inferences from theory and results of reaction outcomes and yields also contribute. [8]

Atropisomers exhibit axial chirality (planar chirality). When the barrier to racemization is high, as illustrated by the BINAP ligands, the phenomenon becomes of practical value in asymmetric synthesis. Methaqualone, the anxiolytic and hypnotic-sedative, is a classical example of a drug molecule that exhibits the phenomenon of atropisomerism. [9]

Stereochemical assignment

Determining stereochemistry in atropisomers where substituent A has priority over substituent B Atropisomer stereochem.png
Determining stereochemistry in atropisomers where substituent A has priority over substituent B

Determining the axial stereochemistry of biaryl atropisomers can be accomplished through the use of a Newman projection along the axis of hindered rotation. The ortho, and in some cases meta substituents are first assigned priority based on Cahn–Ingold–Prelog priority rules. One scheme of nomenclature in based on envisioning the helicity defined by these groups. [10] Starting with the substituent of highest priority in the closest ring and moving along the shortest path to the substituent of highest priority in the other ring, the absolute configuration is assigned P or Δ for clockwise and M or Λ for counterclockwise. [1] Alternately, all four groups can be ranked by Cahn–Ingold–Prelog priority rules, with overall priority given to the groups on the "front" atom of the Newman projection. The two configurations are termed Ra and Sa in analogy to the traditional R/S for a traditional tetrahedral stereocenter. [11]

Synthesis

Two examples of atropisomer synthesis SynthesisofAtropisomer.png
Two examples of atropisomer synthesis

Axially chiral biaryl compounds are prepared by coupling reactions, e.g., Ullmann coupling, Suzuki–Miyaura reaction, or palladium-catalyzed arylation of arenes. [12] Subsequent to the synthesis, the racemic biaryl is resolved by classical methods. Diastereoselective coupling can be achieved through the use of a chiral bridge that links the two aryl groups or through the use of a chiral auxiliary at one of the positions proximal to axial bridge. Enantioselective coupling can be achieved through the use of a chiral leaving group on one of the biaryls or under oxidative conditions that utilize chiral amines to set the axial configuration. [1]

Individual atropisomers can be isolated by seed-directed crystallization of racemates. Thus, 1,1'-Binaphthyl crystallizes from the melt as individual enantiomers. [13] [14] [15]

Scope

Relaying Asymmetry of Transient Atropisomers RelayAsymmTransientAtropisomer.svg
Relaying Asymmetry of Transient Atropisomers
Structures of BINAP, BINOL, QUINAP BINAP, BINOL, QUINAP.png
Structures of BINAP, BINOL, QUINAP
Example of use of P,N ligand for asymmetric catalysis PNligand3.png
Example of use of P,N ligand for asymmetric catalysis

In one application the asymmetry in an atropisomer is transferred in a chemical reaction to a new stereocenter. [16] The atropisomer is an iodoaryl compound synthesised starting from (S)-valine and exists as the (M,S) isomer and the (P,S) isomer. The interconversion barrier between the two is 24.3 kcal/mol (101.7 kJ/mol). The (M,S) isomer can be obtained exclusively from this mixture by recrystallisation from hexanes. The iodine group is homolytically removed to form an aryl radical by a tributyltin hydride/triethylboron/oxygen mixture as in the Barton–McCombie reaction. Although the hindered rotation is now removed in the aryl radical, the intramolecular reaction with the alkene is so much faster than is rotation of the carbon–nitrogen bond that the stereochemistry is preserved. In this way the (M,S) isomer yields the (S,S) dihydroindolone.

The most important class of atropisomers are biaryls such as diphenic acid, which is a derivative of biphenyl with a complete set of ortho substituents. Heteroaromatic analogues of the biphenyl compounds also exist, where hindered rotation occurs about a carbon-nitrogen or a nitrogen-nitrogen bond. [7] Others are dimers of naphthalene derivatives such as 1,1'-bi-2-naphthol. In a similar way, aliphatic ring systems like cyclohexanes linked through a single bond may display atropisomerism provided that bulky substituents are present. The use of axially chiral biaryl compounds such as BINAP, QUINAP and BINOL, have been found to be useful in the area of asymmetric catalysis as chiral ligands.

Their ability to provide stereoinduction has led to use in metal catalyzed hydrogenation, epoxidation, addition, and allylic alkylation reactions. [1] Other reactions that can be catalyzed by the use of chiral biaryl compounds are the Grignard reaction, Ullmann reaction, and the Suzuki reaction. [17] A recent example in the area of chiral biaryl asymmetric catalysis employs a five-membered imidazole as part of the atropisomer scaffold. This specific phosphorus, nitrogen-ligand has been shown to perform enantioselective A3-coupling. [18]

Natural products, drug design

Many atropisomers occur in nature, and some have applications to drug design. [19] The natural product mastigophorene A has been found to aid in nerve growth. [1] [20] Other examples of naturally occurring atropisomers include vancomycin isolated from an Actinobacterium, and knipholone, which is found in the roots of Kniphofia foliosa of the family Asphodelaceae. The structure complexity in vancomycin is significant because it can bind with peptides due to the complexity of its stereochemistry, which includes multiple stereocenters, two chiral planes in its stereogenic biaryl axis. Knipholone, with its axial chirality, occurs in nature and has been shown to offer good antimalarial and antitumor activities particularly in the M form. [1]

The use of atropisomeric drugs provides an additional way for drugs to have stereochemical variations and specificity in design. [21] One example is (–)-N-acetylallocolchinol, a drug that was discovered to aid in chemotherapy cancer treatment. [21] [22]

Telenzepine is atropisomeric in the conformation of its central thienobenzodiazepine ring. The two enantiomers have been resolved, and it was found that the (+)-isomer which is about 500-fold more active than the (–)-isomer at muscarinic receptors in rat cerebral cortex. [23] However, drug design is not always aided by atropisomerism. In some cases, making drugs from atropisomers is challenging because isomers may interconvert faster than expected. Atropisomers also might interact differently in the body, and as with other types of stereoisomers, it is important to examine these properties before administering drugs to patients. [23]

See also

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">Stereochemistry</span> Subdiscipline of chemistry

Stereochemistry, a subdiscipline of chemistry, involves the study of the relative spatial arrangement of atoms that form the structure of molecules and their manipulation. The study of stereochemistry focuses on the relationships between stereoisomers, which by definition have the same molecular formula and sequence of bonded atoms (constitution), but differ in the geometric positioning of the atoms in space. For this reason, it is also known as 3D chemistry—the prefix "stereo-" means "three-dimensionality".

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

In chemistry, an enantiomer – also called optical isomer, antipode, or optical antipode – is one of two stereoisomers that are nonsuperposable onto their own mirror image. Enantiomers of each other are much like one's right and left hands; without mirroring one of them, hands cannot be superposed onto each other. It is solely a relationship of chirality and the permanent three-dimensional relationships among molecules or other chemical structures: no amount of re-orientiation of a molecule as a whole or conformational change converts one chemical into its enantiomer. Chemical structures with chirality rotate plane-polarized light. A mixture of equal amounts of each enantiomer, a racemic mixture or a racemate, does not rotate light.

<span class="mw-page-title-main">Aldol reaction</span> Chemical reaction

The aldol reaction is a reaction in organic chemistry that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound. Its simplest form might involve the nucleophilic addition of an enolized ketone to another:

<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">BINAP</span> Chemical compound

BINAP (2,2′-bis(diphenylphosphino)-1,1′-binaphthyl) is an organophosphorus compound. This chiral diphosphine ligand is widely used in asymmetric synthesis. It consists of a pair of 2-diphenylphosphinonaphthyl groups linked at the 1 and 1′ positions. This C2-symmetric framework lacks a stereogenic atom, but has axial chirality due to restricted rotation (atropisomerism). The barrier to racemization is high due to steric hindrance, which limits rotation about the bond linking the naphthyl rings. The dihedral angle between the naphthyl groups is approximately 90°. The natural bite angle is 93°.

1,1-Bi-2-naphthol (BINOL) is an organic compound that is often used as a ligand for transition-metal catalysed asymmetric synthesis. BINOL has axial chirality and the two enantiomers can be readily separated and are stable toward racemisation. The specific rotation of the two enantiomers is 35.5° (c = 1 in THF), with the R enantiomer being the dextrorotary one. BINOL is a precursor for another chiral ligand called BINAP. The volumetric mass density of the two enantiomers is 0.62 g cm−3.

<span class="mw-page-title-main">Axial chirality</span> Type of symmetry in molecules

In chemistry, axial chirality is a special case of chirality in which a molecule contains two pairs of chemical groups in a non-planar arrangement about an axis of chirality so that the molecule is not superposable on its mirror image. The axis of chirality is usually determined by a chemical bond that is constrained against free rotation either by steric hindrance of the groups, as in substituted biaryl compounds such as BINAP, or by torsional stiffness of the bonds, as in the C=C double bonds in allenes such as glutinic acid. Axial chirality is most commonly observed in substituted biaryl compounds wherein the rotation about the aryl–aryl bond is restricted so it results in chiral atropisomers, as in various ortho-substituted biphenyls, and in binaphthyls such as BINAP.

<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.

In chemistry, transfer hydrogenation is a chemical reaction involving the addition of hydrogen to a compound from a source other than molecular H2. It is applied in laboratory and industrial organic synthesis to saturate organic compounds and reduce ketones to alcohols, and imines to amines. It avoids the need for high-pressure molecular H2 used in conventional hydrogenation. Transfer hydrogenation usually occurs at mild temperature and pressure conditions using organic or organometallic catalysts, many of which are chiral, allowing efficient asymmetric synthesis. It uses hydrogen donor compounds such as formic acid, isopropanol or dihydroanthracene, dehydrogenating them to CO2, acetone, or anthracene respectively. Often, the donor molecules also function as solvents for the reaction. A large scale application of transfer hydrogenation is coal liquefaction using "donor solvents" such as tetralin.

The Negishi coupling is a widely employed transition metal catalyzed cross-coupling reaction. The reaction couples organic halides or triflates with organozinc compounds, forming carbon-carbon bonds (C-C) in the process. A palladium (0) species is generally utilized as the metal catalyst, though nickel is sometimes used. A variety of nickel catalysts in either Ni0 or NiII oxidation state can be employed in Negishi cross couplings such as Ni(PPh3)4, Ni(acac)2, Ni(COD)2 etc.

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

DIPAMP is an organophosphorus compound that is used as a ligand in homogeneous catalysis. It is a white solid that dissolves in organic solvents. Work on this compound by W. S. Knowles was recognized with the Nobel Prize in Chemistry. DIPAMP was the basis for one of the first industrial scale asymmetric hydrogenation, the synthesis of the drug L-DOPA.

Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target (substrate) molecule with three-dimensional spatial selectivity. Critically, this selectivity does not come from the target molecule itself, but from other reagents or catalysts present in the reaction. This allows spatial information to transfer from one molecule to the target, forming the product as a single enantiomer. The chiral information is most commonly contained in a catalyst and, in this case, the information in a single molecule of catalyst may be transferred to many substrate molecules, amplifying the amount of chiral information present. Similar processes occur in nature, where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer, such as amino acids, that a cell needs to function. By imitating this process, chemists can generate many novel synthetic molecules that interact with biological systems in specific ways, leading to new pharmaceutical agents and agrochemicals. The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers — William Standish Knowles and Ryōji Noyori — being collectively awarded one half of the 2001 Nobel Prize in Chemistry.

Within the area of organocatalysis, (thio)urea organocatalysis describes the use of ureas and thioureas to accelerate and stereochemically alter organic transformations. The effects arise through hydrogen-bonding interactions between the substrate and the (thio)urea. Unlike classical catalysts, these organocatalysts interact by non-covalent interactions, especially hydrogen bonding. The scope of these small-molecule H-bond donors termed (thio)urea organocatalysis covers both non-stereoselective and stereoselective applications.

The Tsuji–Trost reaction is a palladium-catalysed substitution reaction involving a substrate that contains a leaving group in an allylic position. The palladium catalyst first coordinates with the allyl group and then undergoes oxidative addition, forming the π-allyl complex. This allyl complex can then be attacked by a nucleophile, resulting in the substituted product.

In organic chemistry, the Keck asymmetric allylation is a chemical reaction that involves the nucleophilic addition of an allyl group to an aldehyde. The catalyst is a chiral complex that contains titanium as a Lewis acid. The chirality of the catalyst induces a stereoselective addition, so the secondary alcohol of the product has a predictable absolute stereochemistry based on the choice of catalyst. This name reaction is named for Gary Keck.

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

Ugi’s amine is an organometallic compound with the formula (C5H5)Fe(C5H4CH N 2. It is named for the chemist who first reported its synthesis in 1970, Ivar Ugi. It is a ferrocene derivative. Ugi’s amine is a precursor to ligands, most notably, the Josiphos ligands, which have been used in asymmetric catalysis

<span class="mw-page-title-main">Jayaraman Sivaguru</span>

Jayaraman Sivaguru (Siva) is the Antonia and Marshall Wilson Professor of Chemistry and the Associate Director, Center for Photochemical Sciences at the Department of Chemistry, Bowling Green State University, Bowling Green, Ohio. He is a recipient of 2008 National Science Foundation CAREER Award, 2010 Grammaticakis-Neumann Prize from the Swiss Chemical Society, 2011 young-investigator award from the Inter-American Photochemical Society (I-APS), and 2012-young investigator award from Sigma Xi. His honors also include Excellence in Research award, 2011 Excellence in Teaching award, and the 2012 PeltierAward for Innovation in Teaching. Prof. Siaguru was a visiting young professor at the Global Centre for Excellence at Osaka University, Japan and was a visiting fellow for the Chinese Academy of Sciences President's International Fellowship Initiative in 2018. He is an editor for the Journal of Photochemistry and Photobiology A: Chemistry and from 2020 serves as the co-Editor-in-Chief of Journal of Photochemistry and Photobiology published by Elsevier. He is an international board member of the International Union of Pure and Applied Chemistry (IUPAC) photochemistry symposium.

In organic chemistry, carboboration describes an addition of both a carbon and a boron moiety to certain carbon-containing double and triple bonds, such as alkenes, alkynes, and allenes.

References

  1. 1 2 3 4 5 6 7 8 Bringmann, Gerhard; Mortimer, Anne J. Price; Keller, Paul A.; Gresser, Mary J.; Garner, James; Breuning, Matthias (2005). "Atroposelective Synthesis of Axially Chiral Biaryl Compounds". Angewandte Chemie International Edition. 44 (34): 5384–5427. doi:10.1002/anie.200462661. PMID   16116589.
  2. Anslyn, Eric V.; Dougherty, Dennis A. (2006). Modern physical organic chemistry. Mill Valley, CA: University Science Books. ISBN   1-891389-31-9. OCLC   55600610.
  3. Clayden, Jonathan; Moran, Wesley J.; Edwards, Paul J.; LaPlante, Steven R. (2009). "The Challenge of Atropisomerism in Drug Discovery". Angew. Chem. Int. Ed. 48 (35): 6398–6401. doi:10.1002/anie.200901719. ISSN   1433-7851. PMID   19637174.
  4. Kuhn Richard (1933). "Molekulare asymmetrie". Stereochemie (Karl Freudenberg, Ed.). Leipzig-Wien:Franz-Deutike. pp. 803–824.
  5. Christie, George Hallatt; Kenner, James (1922-01-01). "LXXI.—The molecular configurations of polynuclear aromatic compounds. Part I. The resolution of γ-6 : 6′-dinitro- and 4 : 6 : 4′ : 6′-tetranitro-diphenic acids into optically active components". Journal of the Chemical Society, Transactions. 121: 614–620. doi:10.1039/CT9222100614.
  6. Ōki, Michinori (2007). "Recent Advances in Atropisomerism". Topics in Stereochemistry. pp. 1–81. doi:10.1002/9780470147238.ch1. ISBN   9780470147238.
  7. 1 2 3 4 Alkorta, Ibon; Elguero, José; Roussel, Christian; Vanthuyne, Nicolas; Piras, Patrick (2012). Atropisomerism and Axial Chirality in Heteroaromatic Compounds. Advances in Heterocyclic Chemistry. Vol. 105. pp. 1–188. doi:10.1016/b978-0-12-396530-1.00001-2. ISBN   9780123965301. OCLC   781540796.
  8. LaPlante, Steven R.; Edwards, Paul J.; Fader, Lee D.; Jakalian, Araz; Hucke, Oliver (2011). "Revealing Atropisomer Axial Chirality in Drug Discovery". ChemMedChem. 6 (3): 505–513. doi:10.1002/cmdc.201000485. PMID   21360821. S2CID   27354841.
  9. Prost, Francine; Thormann, Wolfgang (2003). "Assessment of the stereoselective metabolism of methaqualone in man by capillary electrophoresis". Electrophoresis. 24 (15): 2598–2607. doi:10.1002/elps.200305512. PMID   12900872. S2CID   27581783.
  10. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " helicity ". doi : 10.1351/goldbook.H02763
  11. IUPAC , Compendium of Chemical Terminology , 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006) " axial chirality ". doi : 10.1351/goldbook.A00547
  12. Cepanec, Ivica (2004). Synthesis of biaryls (1st ed.). Amsterdam: Elsevier. ISBN   978-0-08-044412-3. OCLC   162567758.
  13. Wilson, Keith R.; Pincock, Richard E. (1975-03-01). "Thermally induced resolution of racemic 1,1'-binaphthyl in the solid state". Journal of the American Chemical Society. 97 (6): 1474–1478. doi:10.1021/ja00839a033.
  14. Einhorn, Cathy; Durif, André; Averbuch, Marie-Thérèse; Einhorn, Jacques (2001). "Solid-State Isomerization of Atropodiastereomers: Effective Diastereoselection through Polymorphic Transformations". Angewandte Chemie International Edition. 40 (10): 1926–1929. doi:10.1002/1521-3773(20010518)40:10<1926::AID-ANIE1926>3.0.CO;2-3. PMID   11385675.
  15. Pu, Lin (1998-11-05). "1,1'-Binaphthyl Dimers, Oligomers, and Polymers: Molecular Recognition, Asymmetric Catalysis, and New Materials". Chemical Reviews. 98 (7): 2405–2494. doi:10.1021/cr970463w. PMID   11848968.
  16. Petit, Marc; Lapierre, Andre J. B.; Curran, Dennis P. (2005-11-01). "Relaying Asymmetry of Transient Atropisomers of o-Iodoanilides by Radical Cyclizations". Journal of the American Chemical Society. 127 (43): 14994–14995. doi:10.1021/ja055666d. PMID   16248616.
  17. Cozzi, Pier Giorgio; Emer, Enrico; Gualandi, Andrea (2011). "Atroposelective Organocatalysis". Angewandte Chemie International Edition. 50 (17): 3847–3849. doi:10.1002/anie.201008031. PMID   21448867.
  18. Cardoso, Flavio S. P.; Abboud, Khalil A.; Aponick, Aaron (2013-10-02). "Design, Preparation, and Implementation of an Imidazole-Based Chiral Biaryl P,N-Ligand for Asymmetric Catalysis". Journal of the American Chemical Society. 135 (39): 14548–14551. doi:10.1021/ja407689a. PMID   24044433.
  19. Smyth, Jamie E.; Butler, Nicholas M.; Keller, Paul A. (2015). "A twist of nature – the significance of atropisomers in biological systems". Natural Product Reports. 32 (11): 1562–1583. doi:10.1039/c4np00121d. PMID   26282828.
  20. Fukuyama, Yoshiyasu; Asakawa, Yoshinori (1991-01-01). "Novel neurotrophic isocuparane-type sesquiterpene dimers, mastigophorenes A, B, C and D, isolated from the liverwort Mastigophora diclados". Journal of the Chemical Society, Perkin Transactions 1 (11): 2737–2741. doi:10.1039/P19910002737.
  21. 1 2 Zask, Arie; Murphy, John; Ellestad, George A. (2013). "Biological Stereoselectivity of Atropisomeric Natural Products and Drugs". Chirality. 25 (5): 265–274. doi:10.1002/chir.22145. PMID   23620262.
  22. Joncour, Agnès; Décor, Anne; Thoret, Sylviane; Chiaroni, Angèle; Baudoin, Olivier (2006). "Biaryl Axis as a Stereochemical Relay for the Enantioselective Synthesis of Antimicrotubule Agents". Angewandte Chemie International Edition. 45 (25): 4149–4152. doi:10.1002/anie.200600451. PMID   16688690.
  23. 1 2 Clayden, Jonathan; Moran, Wesley J.; Edwards, Paul J.; LaPlante, Steven R. (2009). "The Challenge of Atropisomerism in Drug Discovery". Angewandte Chemie International Edition. 48 (35): 6398–6401. doi:10.1002/anie.200901719. PMID   19637174.

Further reading