Fact-checked by Grok 2 weeks ago

Stereoselectivity

Stereoselectivity is the preferential formation in a of one stereoisomer over another. This phenomenon arises due to differences in the transition states leading to different stereoisomeric products, often influenced by steric, electronic, or conformational factors. When the stereoisomers are enantiomers, the selectivity is termed enantioselectivity and is quantified by enantiomeric excess (ee); for diastereomers, it is diastereoselectivity, measured by diastereomeric excess (de). In , stereoselectivity is distinguished from , where the latter refers to reactions in which stereoisomeric starting materials yield stereoisomerically related products, such that the product's stereochemistry is directly determined by the reactant's . All stereospecific reactions are inherently stereoselective, but stereoselective reactions may produce mixtures of stereoisomers with one predominating, rather than a single product exclusively. Common examples include syn or anti addition to alkenes, such as the syn using , which favors one face of the over the other. Stereoselectivity plays a pivotal role in , enabling the controlled construction of complex molecules with defined three-dimensional architectures essential for . In pharmaceutical development, stereoisomers of drugs can exhibit dramatically different pharmacological profiles—one enantiomer may be therapeutic while the other is inactive or toxic—necessitating stereoselective methods to produce enantiomerically pure compounds. Advances in asymmetric , such as those using chiral ligands or enzymes, have revolutionized stereoselective , improving efficiency and sustainability in producing and natural products.

Definitions and Principles

Definition

Stereoselectivity refers to the preferential formation in a of one stereoisomer over another when multiple stereoisomers are possible from the same reactant and reaction pathway. This property arises because reactions can favor certain spatial arrangements of atoms due to differences in energies, leading to an unequal mixture where the preferred stereoisomer predominates. Stereoisomers are molecules with identical molecular formulas and connectivity of atoms but differing in the three-dimensional arrangement of those atoms. Key types include enantiomers, which are nonsuperimposable mirror images of each other and exhibit identical physical properties except for , and diastereomers, which are stereoisomers that are not mirror images and often differ in physical properties such as melting points and solubilities. The importance of stereoselectivity in chemistry lies in its role in efficiently producing chiral molecules, which are critical for applications in and synthesis, as the specific often dictates and pharmacological efficacy. For instance, in pharmaceuticals, the wrong stereoisomer can lead to reduced potency or adverse effects, underscoring the need for selective reactions to isolate bioactive forms. Quantitative assessment of stereoselectivity uses metrics like enantiomeric excess (ee) for reactions involving and diastereomeric excess (de) for . Enantiomeric excess is defined as ee = |(% major ) - (% minor )|, providing a measure of chiral purity; for example, a mixture with 85% (R)- and 15% (S)- yields ee = 70%. Diastereomeric excess follows an analogous calculation, de = |(% major ) - (% minor )|, to evaluate selectivity between .

Distinction from Stereospecificity

Stereospecificity describes a in which stereoisomerically pure reactants yield correspondingly distinct stereoisomeric products, such that different stereoisomers of the starting material produce different stereoisomers of the product. This contrasts with stereoselectivity, which, as defined earlier, involves the preferential but not necessarily exclusive formation of one stereoisomer from a given reactant. In essence, stereospecific reactions exhibit complete control over product stereochemistry based on reactant configuration, rendering them a subset of stereoselective processes; however, many stereoselective reactions lack this strict dependency and may produce mixtures of stereoisomers. A classic illustration of stereospecificity is the bimolecular nucleophilic substitution () reaction, where the reactant undergoes complete inversion of configuration at the reaction center, as established through kinetic and stereochemical studies. Similarly, in bimolecular elimination (E2) reactions, the anti-periplanar requirement for the departing groups ensures that the geometry of the resulting directly reflects the relative of the starting material, producing Z or E products predictably from diastereomeric precursors. The foundational concepts distinguishing these reaction behaviors emerged from Christopher Ingold's mechanistic investigations into and elimination processes during , particularly his elucidation of concerted pathways that impose rigid constraints. In non- additions, such as certain electrophilic additions to alkenes, the process may favor one stereoisomer due to steric or electronic factors but does not mandate a unique outcome tied to any inherent reactant stereochemistry, often yielding mixtures rather than pure stereoisomers. This distinction is crucial for understanding reaction design, as stereospecificity guarantees predictability in stereochemical transfer, whereas stereoselectivity offers tunable preference without absolute certainty.

Types

Enantioselectivity

Enantioselectivity refers to the preferential formation in a of one over the other when the stereoisomers involved are enantiomers. This phenomenon is central to asymmetric synthesis, where reactions starting from achiral substrates produce a chiral product with a bias toward one mirror-image form, typically requiring the use of chiral catalysts, , or to induce the selectivity. A key concept in enantioselectivity is chiral induction, whereby a chiral element in the environment differentiates the two enantiotopic faces of a prochiral , favoring the formation of one over the other. The outcome is measured in terms of , assigned using the Cahn-Ingold-Prelog () priority rules to designate the as or ; for instance, a highly enantioselective might yield predominantly the ()-enantiomer if the clockwise arrangement of substituents (with the lowest priority group pointing away) predominates. The degree of enantioselectivity is quantified by enantiomeric excess (), defined as ee = \frac{|[] - [minor]|}{[major] + [minor]} \times 100%, where [major] and [minor] represent the concentrations of the predominant and subordinate enantiomers, respectively. This metric directly reflects the purity of the chiral product; for example, an ee of 90% corresponds to a 95:5 of the major to minor enantiomer, indicating strong but incomplete selectivity. Theoretically, enantioselectivity arises from the difference (ΔΔG‡) between the transition states leading to each ; a larger ΔΔG‡ results in greater differentiation, as the lower-energy pathway dominates kinetically, while even small differences (on the order of 1-2 kcal/) can yield useful levels of selectivity in practice.

Diastereoselectivity

Diastereoselectivity describes the preferential formation of one over another in chemical reactions where the already contains one or more chiral centers, resulting in products that are stereoisomers but not mirror images of each other. This type of selectivity arises because possess different physical and chemical properties due to their non-superimposable, non-mirror-image configurations at multiple stereocenters. In reactions involving facial selectivity, such as nucleophilic additions to chiral carbonyl compounds, qualitative models guide the prediction of preferred diastereomer formation. Cram's rule posits that the nucleophile approaches the carbonyl from the less hindered face in a conformation where the largest substituent adjacent to the chiral center is positioned anti to the incoming group, minimizing steric interactions. Complementing this non-chelation model, chelation control occurs when a metal coordinates simultaneously to the carbonyl oxygen and a nearby heteroatom (such as oxygen or ) in the substrate, rigidifying the conformation and directing the nucleophile to the opposite face for enhanced selectivity. Diastereoselectivity is measured using the diastereomeric excess (de), defined as the percentage difference in the amounts of the two diastereomers formed: \text{de} = \frac{|[\text{D}_1] - [\text{D}_2]|}{[\text{D}_1] + [\text{D}_2]} \times 100\% where [\text{D}_1] and [\text{D}_2] represent the concentrations or mole fractions of the major and minor diastereomers, respectively. This metric is analogous to enantiomeric excess but applies to diastereomers; for instance, a 10:1 ratio corresponds to a de of approximately 82%, indicating the major diastereomer constitutes 91% of the product mixture. Compared to enantioselectivity, which distinguishes between enantiomers in a chiral environment and often yields moderate levels due to their energetic equivalence, diastereoselectivity is typically higher because the existing chiral centers in the create distinct energy differences between transition states leading to diastereomeric products. Enantioselectivity represents a related but distinct aspect of stereoselectivity focused on mirror-image products.

Mechanisms and Control

Kinetic vs. Thermodynamic Control

In stereoselective reactions, kinetic control determines the outcome when the reaction proceeds irreversibly, favoring the stereoisomer formed through the transition state with the lower activation energy. The selectivity arises from the difference in activation free energies, denoted as \Delta \Delta G^\ddagger, between competing pathways, where a smaller \Delta \Delta G^\ddagger leads to higher stereoselectivity by exponentially favoring the lower-energy route according to the Eyring equation. Low temperatures enhance this regime by minimizing thermal energy available for alternative pathways, ensuring the reaction captures the kinetically preferred product before significant side reactions occur. Under thermodynamic , stereoselectivity is governed by the relative stabilities of the stereoisomeric products, allowing the most stable to predominate at . This occurs in reversible reactions where products and intermediates interconvert freely, driven by differences in their Gibbs free energies (\Delta G), with the lower-energy stereoisomer accumulating over time. Higher temperatures promote this control by increasing the of equilibration, enabling the system to overcome barriers and reach the global energy minimum. Catalysts can influence this regime by lowering reversal barriers without altering product stabilities, thus facilitating thermodynamic resolution. Qualitatively, energy diagrams for these regimes differ markedly: in kinetic control, the profile emphasizes the heights of transition state barriers from reactants, where the path with the shallowest barrier dictates selectivity, often leading to metastable products. In contrast, thermodynamic control focuses on the depths of product energy wells relative to each other, with transition states between products being surmountable under equilibrating conditions, resulting in the deepest well (most stable stereoisomer) prevailing. The switch between regimes is primarily triggered by reversibility—irreversible for kinetic, reversible for thermodynamic—modulated by and catalyst design to tune barrier accessibility.

Factors Influencing Selectivity

Steric factors significantly dictate the stereochemical outcome of reactions by imposing spatial constraints that favor transition states with minimal non-bonded repulsions. In nucleophilic additions to aldehydes or ketones bearing an α-, bulkier substituents on the stereocenter rotate to an position relative to the incoming , as described by the Felkin-Anh model, thereby directing approach from the less hindered face. This model, supported by computational analyses, highlights how increasing steric bulk enhances selectivity by widening the energy gap between competing conformers. Electronic factors complement steric influences by modulating energies through polarization and orbital interactions. Polar substituents can stabilize specific geometries via electrostatic effects, while stereoelectronic interactions, such as from adjacent σ-bonds, align to lower the barrier for the preferred pathway in additions or eliminations. For example, electron-withdrawing groups may polarize the substrate to favor nucleophilic attack from one face, thereby controlling diastereoselectivity in conjugate additions. Chiral auxiliaries and catalysts impose on otherwise achiral substrates or , amplifying differences in free energies (ΔΔG‡) to drive high selectivity. Chiral auxiliaries, like the Evans oxazolidin-2-one, are covalently attached to create a rigid, shielded environment that directs geometry in aldol reactions, ensuring facial discrimination through or steric blocking. Similarly, chiral catalysts, such as BINAP-rhodium complexes in , form enantioselective binding pockets that differentiate enantiotopic faces via non-covalent interactions, enabling efficient enantiocontrol across diverse substrates. Solvent and temperature further tune selectivity by affecting solvation and population dynamics. Polar solvents stabilize charged or polar s, often enhancing electronic control and inverting selectivity compared to nonpolar media, as seen in Diels-Alder reactions where protic solvents promote endo products through hydrogen bonding. Temperature modulates the kinetic regime, with lower values favoring the thermodynamically less stable but kinetically preferred stereoisomer by limiting access to higher-energy pathways, a principle applied in cryogenic conditions for aldol additions to maximize diastereoselectivity.

Applications and Examples

In Synthetic Organic Chemistry

Stereoselectivity plays a pivotal role in synthetic by enabling the efficient construction of chiral molecules with precise three-dimensional architectures, which is essential for the of pharmaceuticals and products. Through the design of chiral catalysts and , chemists can control the stereochemical outcome of reactions, often achieving high enantiomeric excess () or diastereomeric excess () to minimize the need for laborious separations. This control is particularly valuable in multi-step syntheses where a single stereogenic center can dictate the and safety of the final compound. One landmark example of stereoselectivity in synthesis is the Sharpless asymmetric epoxidation, which provides enantioselective access to epoxy alcohols from allylic alcohols. This reaction employs a chiral titanium catalyst formed from titanium tetraisopropoxide and , along with tert-butyl hydroperoxide as the oxidant, to deliver the with predictable facial selectivity based on the tartrate . For kinetic resolution of racemic secondary allylic alcohols, the method selectively epoxidizes one , allowing isolation of the unreacted alcohol in high (often >95%) and the epoxy alcohol product in complementary high , thus enabling the preparation of both s from a racemic starting material. This enantioselective process exemplifies kinetic control in asymmetric synthesis, with applications in building complex chiral scaffolds. In chemistry, the Diels-Alder reaction demonstrates diastereoselectivity when using chiral dienophiles, guided by the rule that favors the where the dienophile's substituents align endo to the for enhanced orbital overlap. Seminal work using chiral N-acyloxazolidinone auxiliaries attached to α,β-unsaturated acyl groups as dienophiles with yields cycloadducts with high diastereoselectivity (de >90%), where the auxiliary dictates facial approach and the endo geometry predominates. This approach allows for the stereocontrolled formation of up to four new stereocenters in a single step, facilitating the synthesis of polycyclic frameworks with defined relative . Enantioselective hydrogenation of alkenes represents another cornerstone, where variants of (RhCl(PPh₃)₃) modified with chiral bisphosphine ligands enable precise stereocontrol. The use of (R,R)-DIPAMP as a ligand in complexes achieves high enantioselectivity in the reduction of enamides, such as those derived from α-acetamidoacrylic acids, yielding precursors with ee up to 94%. This method's has made it industrially viable for producing enantiopure building blocks. Advances in stereoselective synthesis have profoundly impacted pharmaceutical development, particularly in the of statins like and , where stereocontrol at key hydroxyl-bearing centers enhances potency and reduces side effects. Highly stereoselective hydrogenations serve as critical steps in these routes, ensuring the correct configuration at multiple chiral centers to mimic the natural substrate for inhibition.

In Biosynthesis

Enzymes serve as chiral catalysts in biosynthesis, enforcing stereoselectivity through their three-dimensional active sites that provide pocket-like environments for substrate binding, akin to the lock-and-key model proposed by Emil Fischer, where the enzyme's structure complements the substrate's geometry to favor one stereoisomer over others. This selectivity arises from precise steric and electronic interactions within the enzyme's chiral pocket, ensuring high fidelity in the formation of stereocenters during natural product assembly. In biological systems, such enzymatic control minimizes the production of inactive or deleterious stereoisomers, directing metabolic pathways toward bioactive molecules. A prominent example of enzymatic stereoselectivity occurs in biosynthesis, particularly the cyclization of to form steroids. epoxide undergoes a stereoselective polycyclization catalyzed by oxidosqualene cyclase (OSC), which guides the substrate through a series of rearrangements to yield with precise at multiple chiral centers, essential for subsequent production. This process exemplifies how enzymes like OSC achieve near-perfect stereocontrol by stabilizing specific transition states within their active sites, preventing alternative cyclization pathways. In synthesis, diastereoselective s are mediated by ketoreductase () domains within modular polyketide synthases (PKSs), which convert β-ketoacyl intermediates to hydroxyacyl units with defined . For instance, in the of actinorhodin, the actKR domain performs a regio- and stereoselective at the C9 carbonyl, producing the (R)-hydroxy through hydride transfer from NADPH, controlled by the KR's Rossmann fold and residues. These KR domains classify into subtypes (A, B) that dictate syn or anti formation, enabling the diversity of polyketide natural products like antibiotics. From an evolutionary perspective, stereoselectivity in biosynthesis ensures the bioactivity of biomolecules by favoring configurations compatible with physiological targets. The universal adoption of L-amino acids in proteins, rather than D-enantiomers, likely originated from early chiral biases in prebiotic chemistry and RNA-world scenarios, where L-amino acids enhanced folding stability and receptor interactions, conferring survival advantages. This homochirality extends to other biopolymers, such as D-sugars in DNA/RNA, underscoring how stereoselective enzymatic machinery evolved to maintain functional specificity and avoid metabolic interference from mirror-image molecules. Recent advances in , post-2020, have leveraged computational design and to enhance enzymatic stereoselectivity for biotech applications. For example, learning-guided of amidases has improved enantioselectivity ratios from 10:1 to over 100:1 in , enabling scalable production of chiral intermediates for pharmaceuticals. Similarly, semi-rational of ketoreductases has broadened scope while preserving diastereoselectivity, facilitating greener synthesis routes in industrial biocatalysis. In 2025, biocatalytic dynamic kinetic has been applied to synthesize enantioenriched atropisomers from achiral precursors, achieving high enantioselectivity in the formation of axially chiral biaryls.

References

  1. [1]
    stereoselectivity (S05991) - IUPAC Gold Book
    The preferential formation in a chemical reaction of one stereoisomer over another. When the stereoisomers are enantiomers, the phenomenon is called ...
  2. [2]
    Stereoselectivity in Addition Reactions to Double Bonds
    Jan 22, 2023 · Addition reactions to this function might occur in three different ways, depending on the relative orientation of the atoms or groups that add to the carbons ...
  3. [3]
  4. [4]
    Significance and challenges of stereoselectivity assessing methods ...
    Stereoselectivity in drug metabolism can not only influence the pharmacological activities, tolerability, safety, and bioavailability of drugs directly, but ...
  5. [5]
    Multiple stereoselectivity and its application in organic synthesis
    Stereoselectivity and stereoselective methods in organic synthesis are a problem of fundamental importance and will be even more important in the future as ...
  6. [6]
    Assymetric Induction - MSU chemistry
    ... diastereomeric excess (%de), defined as the mole fraction difference in diastereomers times 100. Thus, a mixture of 80% diastereomer A and 20% diastereomer ...<|control11|><|separator|>
  7. [7]
    stereospecificity (S05994) - IUPAC Gold Book
    A reaction is termed stereospecific if starting materials differing only in their configuration are converted into stereoisomeric products.Missing: origin Ingold
  8. [8]
    257. Reaction kinetics and the Walden inversion. Part VI. Relation of ...
    Relation of steric orientation to mechanism in substitutions involving halogen atoms and simple or substituted hydroxyl groups. W. A. Cowdrey, E. D. Hughes, ...Missing: SN2 | Show results with:SN2
  9. [9]
    Introduction: Enantioselective Catalysis | Chemical Reviews
    Enantioselective catalysis entails the catalytic, selective, and reproducible generation of a given enantiomer of a chiral product from achiral reactants.
  10. [10]
    Metal Stereogenicity in Asymmetric Transition Metal Catalysis
    Mar 29, 2023 · This review provides a comprehensive survey of reported chiral transition metal catalysts in which the metal formally constitutes a stereocenter.<|separator|>
  11. [11]
    The R and S Convention for Absolute Configuration
    The absolute stereochemistry of chiral molecules is defined by the R/S convention developed in 1956 by RS Cahn, C. Ingold and V. Prelog.
  12. [12]
    Do the Terms “%ee” and “%de” Make Sense as Expressions of ... - NIH
    It is recommended that the terms ee and de be abandoned in favor of er, dr and q as descriptors of stereoisomer composition and stereoselectivity.
  13. [13]
    Illustrated Glossary of Organic Chemistry - Term
    Diastereoselective: A process such as a chemical reaction or total synthesis in which a mixture of diastereomers is possible, but one of these diastereomers ...Missing: diastereoselectivity | Show results with:diastereoselectivity
  14. [14]
    [PDF] Stereospecificity and Stereoselectivity
    A stereoselective reaction is one in which a single reactant can give two or more stereoisomeric products, and one or more of these products is preferred over ...
  15. [15]
    [PDF] Chemistry 425b: Organic Reactions
    MO Description: The new occupied orbital is lower in energy. When you stabilize the electrons in a system, you stabilize the system as a whole.
  16. [16]
    [PDF] Diastereoselective synthesis of tertiary alcohols by nucleophilic ...
    Chelation-control of stereochemistry is one of the most effective means by which diastereoselective carbonyl addition reactions can be accomplished.9 Two ...
  17. [17]
    [PDF] Diastereoselection in Lewis-Acid-Mediated Aldol Additions
    This review covers the evolution of stereoselective. Lewis-acid-mediated aldol-type addition up to the recent development of chiral Lewis acids. Mukaiyama et al ...
  18. [18]
    Assymetric Induction - MSU chemistry
    The ee percentage given in the last equation stands for enantiomeric excess ... (%de), defined as the mole fraction difference in diastereomers times 100.Missing: calculation | Show results with:calculation
  19. [19]
    8.4. Stereoselectivity – Introduction to Organic Chemistry - Saskoer.ca
    Diastereoselectivity can be achieved if there is a moderate to large energy difference between transition states leading to the products. This is usually ...
  20. [20]
    [PDF] Kinetic and Thermodynamic Control | Dalal Institute
    The selectivity arising from kinetic vs thermodynamic control is extremely important in the case of asymmetric synthesis. This can be attributed to the fact ...
  21. [21]
    [PDF] Kinetic vs Thermodynamic Control
    Reaction of 1,2-Diols with a bis-enol ether to give dispiroketals. – The dispiroketal forms as a single diastereomer as the result of its formation being ...Missing: stereoselective | Show results with:stereoselective
  22. [22]
    [PDF] Remote Control of Stereochemistry: Communicating Information via ...
    Nov 5, 2007 · Kinetic and thermodynamic control in strategies for stereochemical relay. A and B represent two alternative stereo- or conformational ...
  23. [23]
    [PDF] of 1-2 asymmetric induction. the importance of antiperiplanarity
    A qualitative interpretation of Felkin's model, based on perturbational arguments, is suggested which stresses the ... In their original papers 7. Felkin ...
  24. [24]
    Stereoelectronic Effects. A Bridge between Structure and Reactivity ...
    Feb 1, 2017 · Transition states are similarly stabilized by stereoelectronic effects, what has an impact on reactivities and selectivities. The ninth chapter ...
  25. [25]
    [PDF] Chiral Auxiliaries in Asymmetric Synthesis - ResearchGate
    More recent studies by the Evans group have demonstrated an extension of the aldol process, which employs the same oxazolidinone 1 or the thiazolidine thione. 2 ...
  26. [26]
    [PDF] Key Concepts in Stereoselective Synthesis
    In stereospecific reactions, the stereochemistry of the starting material determines the stereochemistry of the product. For a reaction to be stereospecific, ...
  27. [27]
    Solvent effects on stereoselectivity: more than just an environment
    Feb 12, 2009 · Intensive investigations into the role of solvents on diastereo- and enantioselective reactions, as well as temperature-dependent measurements ...
  28. [28]
    The first practical method for asymmetric epoxidation
    Mechanism of the Sharpless Epoxidation Reaction: A DFT Study. The Journal of Physical Chemistry A 2024, 128 (11) , 2072-2091. https://doi.org/10.1021/acs ...
  29. [29]
    Kinetic resolution of racemic allylic alcohols by enantioselective ...
    Kinetic resolution of racemic allylic alcohols by enantioselective epoxidation. A route to substances of absolute enantiomeric purity?Missing: original | Show results with:original
  30. [30]
    Asymmetric Diels-Alder cycloaddition reactions with chiral .alpha ...
    ... Evans Auxiliary Induced by Mg(ClO4)2. The Journal of Organic Chemistry 2012, 77 (24) , 11091-11095. https://doi.org/10.1021/jo302160s. Mangilal Chouhan ...Missing: original | Show results with:original
  31. [31]
    Asymmetric hydrogenation with a complex of rhodium and a chiral ...
    Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1975, 97, 9, 2567–2568 ... Plus‐Size DiPAMP Hybrids for Advanced Rhodium(I)‐Catalyzed ...
  32. [32]
    Highly stereoselective hydrogenations—As key‐steps in the total ...
    Nov 6, 2009 · We summarize here two new approaches for the total synthesis of the most important representatives, atorvastatin, and rosuvastatin, based on highly ...
  33. [33]
    Rationalization of stereoselectivity in enzyme reactions - Chan - 2019
    Dec 5, 2018 · As early as 1894, Fischer proposed the “key and lock” theory which refers to the special relationships between an enzyme and its substrate.22 ...
  34. [34]
    Enzymatic strategies for asymmetric synthesis - RSC Publishing
    Jun 1, 2021 · In this review, the major enzymatic strategies broadly applicable in the asymmetric synthesis of optically pure chiral compounds are presented1. Stereoselective Reactions... · 1.1. C · Notes And References
  35. [35]
    A Case Study in Biomimetic Total Synthesis: Polyolefin ...
    Steroid biosynthesis from squalene and its derivatives is generally treated in two segments. The first is an enzyme-catalyzed series of carbocyclizations, ...
  36. [36]
    Biocatalytic stereocontrolled head-to-tail cyclizations of unbiased ...
    Jun 10, 2024 · In this study, we showcase the remarkable capabilities of squalene-hopene cyclases (SHCs) in the chemoenzymatic synthesis of head-to-tail-fused terpenes.
  37. [37]
    Path to Actinorhodin: Regio- and Stereoselective Ketone Reduction ...
    In the actinorhodin type II polyketide synthase, the first polyketide modification is a regiospecific C9-carbonyl redn., catalyzed by the ketoreductase (actKR).
  38. [38]
    The Epimerase and Reductase Activities of Polyketide Synthase ...
    Stereochemistry of reductions catalyzed by methyl-epimerizing ketoreductase domains of polyketide synthases. ... Essential role of the donor acyl carrier protein ...Missing: diastereoselective | Show results with:diastereoselective
  39. [39]
    Amino Acid Chirality: Stereospecific Conversion and Physiological ...
    This mini-review summarizes the overall mechanistic insights into the interconversion of l- and d-amino acids by the amino acid racemases.
  40. [40]
    Chirality and the Origin of Life - MDPI
    Nov 30, 2021 · It has long been believed that mammalian life is based exclusively on L-amino acids, and their enantiomers, i.e., D-amino acids, are unsuitable ...
  41. [41]
    Machine learning-assisted amidase-catalytic enantioselectivity ...
    Oct 10, 2024 · Biocatalysis is an attractive approach for the synthesis of chiral pharmaceuticals and fine chemicals, but assessing and/or improving the ...
  42. [42]
    Redesigning Enzymes for Biocatalysis: Exploiting Structural ... - NIH
    Here, we discuss rational or semi-rational protein engineering approaches to alter aspects of enzyme chemistry with a focus on improving stereo- and/or ...