Fact-checked by Grok 2 weeks ago

Macrocycle

A macrocycle is a cyclic consisting of at least 12 atoms arranged in a , distinguishing it from smaller rings and enabling unique conformational flexibility and binding properties. In , macrocycles occupy a critical chemical space between conventional small molecules and biologics, often featuring , peptoid, or heterocyclic backbones that confer rigidity and specificity. They are abundant in natural products, such as antibiotics like and immunosuppressants like rapamycin, and have inspired synthetic analogs for therapeutic development. Key subclasses include crown ethers, cyclodextrins, calixarenes, and porphyrins, each exploiting non-covalent interactions for molecular recognition. Macrocycles hold significant promise in , where their larger size allows disruption of protein-protein interactions intractable to traditional small-molecule inhibitors, with 30–40% demonstrating despite extending beyond . Beyond pharmaceuticals, they enable applications in for , sensing, and materials, including pollutant capture via covalent organic frameworks and stimuli-responsive systems for controlled release. As of 2025, advances such as machine learning-guided and /ML-enhanced have expanded accessible macrocycle diversity, with over 76 FDA-approved macrocycle-based drugs, enhancing their utility in addressing unmet needs in medicine and environmental science.

Definition and Classification

Definition

A macrocycle is defined in as a or containing a ring composed of twelve or more atoms, distinguishing it from smaller cyclic compounds. This ring size criterion is widely adopted in the literature to encompass a diverse class of structures, including both carbocyclic and heterocyclic variants, where the cyclic framework enables unique topological and functional properties. Although the strict IUPAC definition from describes a macrocycle as a cyclic or a cyclic portion of a , in supramolecular and organic contexts, the term is routinely applied to lower molecular weight species meeting the atom count threshold, without implying polymeric scale. The concept of macrocycles gained prominence in the 1960s through the work of Charles J. Pedersen, who coined the term in reference to cyclic polyethers during his discovery of crown ethers. Pedersen's serendipitous synthesis of dibenzo-18-crown-6 in 1961, detailed in his seminal 1967 publication, marked the inception of systematic research into these compounds, highlighting their ability to form complexes with metal cations and laying the foundation for . This historical distinction emphasized macrocycles as larger rings, in contrast to common small cycles like , which exhibit significant angle and torsional strain due to their constrained geometry. Larger ring sizes in macrocycles generally minimize angle and steric repulsion compared to smaller rings, promoting conformational flexibility that allows adoption of multiple stable geometries. This flexibility facilitates the formation of internal cavities capable of hosting guest molecules or ions through non-covalent interactions, a key feature enabling applications in recognition and binding. A representative simple macrocycle is cyclododecane, a 12-membered all-carbon ring that exemplifies these traits without heteroatoms or functional groups, serving as a baseline for understanding unadorned cyclic .

Types of Macrocycles

Macrocycles are broadly classified by the number of atoms in the ring and their elemental composition, with rings containing 12 or more atoms typically qualifying as macrocycles to distinguish them from smaller cyclic compounds. This emphasizes the role of composition in determining properties such as flexibility, binding affinity, and stability. Carbocycles, composed solely of carbon atoms, represent the simplest form and include examples like cyclooctadecane, a saturated 18-membered ring (C18H36) known for its conformational flexibility due to minimal . Heterocycles incorporate one or more heteroatoms such as oxygen, , or into the carbon framework, which often introduces donor sites for coordination or host-guest interactions; crown ethers, featuring multiple oxygen atoms in an all-single-bond ring, exemplify this class and selectively bind cations like K+ through electrostatic interactions. Metallomacrocycles integrate a central metal ion coordinated by the ring's donor atoms, enhancing electronic and catalytic properties; porphyrins, with four units linked by methine bridges and a nitrogen-coordinated metal center (e.g., iron in ), are a prototypical subclass valued for their role in oxygen transport and light harvesting. Key subclasses of macrocycles further diversify their applications based on structural motifs. Crown ethers, first synthesized by Pedersen, are polyether heterocycles named by ring size and oxygen count (e.g., 18-crown-6 for an 18-atom ring with six oxygens), prized for their ionophoric properties in phase-transfer and . Cyclodextrins consist of six to eight glucose units linked by α-1,4-glycosidic bonds, forming toroidal sugar-based heterocycles with hydrophobic interiors that encapsulate guest molecules in aqueous media, as established in foundational enzymatic studies. Calixarenes, derived from p-tert-butylphenol and , adopt a vase-like conformation from typically four to eight phenolic units, enabling cavity-based recognition of cations and neutral guests through hydrogen bonding and π-interactions. Pillararenes feature five para-substituted units linked at para positions, yielding rigid, cylindrical aromatic macrocycles with planar symmetry that excel in and stimuli-responsive materials. macrocycles, formed by cyclizing linear peptides via or bonds, mimic natural constraints to improve proteolytic stability and bioavailability, as seen in therapeutic agents like cyclosporine A. Structural variations within macrocycles influence their dynamics and functionality. Saturated macrocycles, characterized by all single bonds, exhibit high conformational flexibility, facilitating adaptive binding but potentially complicating preorganization; crown ethers and cyclooctadecane illustrate this, with multiple low-energy conformers accessible at . In contrast, unsaturated macrocycles incorporate double bonds, triple bonds, or aromatic systems, imparting rigidity and planarity that stabilize specific geometries; porphyrins and pillararenes demonstrate this through conjugated π-systems that enable delocalization and chromophoric behavior. Bridged macrocycles feature additional covalent connections between non-adjacent ring atoms, restricting flexibility and creating defined cavities compared to unbridged counterparts; examples include - or biaryl-bridged variants that enhance steric control in coordination chemistry. These variations allow tailoring of macrocycle properties for targeted applications, though larger rings generally show increased conformational entropy. Nomenclature for macrocycles follows IUPAC guidelines to reflect composition, size, and substitution systematically. For carbocycles, the von Baeyer system uses the "cyclo-" followed by the total carbon count and suffix, as in cyclooctadecane for an unbranched 18-carbon ring. Heterocyclic macrocycles employ Hantzsch-Widman for heteroatoms (e.g., oxa- for oxygen, aza- for ) combined with cyclo- and a for ring size, though trivial names prevail for established classes like -crown-m (n = total atoms, m = heteroatoms). Metallomacrocycles are named by specifying the (e.g., for the parent) followed by the metal and , such as magnesium in . Functionalized types often use specific descriptors like calixarene for n phenolic units or pillararene for n arene subunits, prioritizing clarity over exhaustive systematic naming in supramolecular contexts.

Structural Properties

Conformational Preferences

Macrocycles containing at least 12 atoms generally exhibit low overall ring strain compared to smaller cycloalkanes, enabling a greater number of accessible conformations due to reduced angle strain and the ability to adopt non-planar geometries that minimize torsional (Pitzer) strain. In these systems, Pitzer strain arises from eclipsed or gauche interactions along the ring backbone, while transannular interactions—steric repulsions between non-adjacent atoms across the ring—play a dominant role in determining energy minima, often leading to flexible, dynamic structures rather than rigid ones. This balance allows macrocycles to sample multiple low-energy conformers, with energy barriers low enough for interconversion at ambient temperatures. Illustrative examples from larger cycloalkanes highlight these principles for macrocyclic systems. Larger rings, such as cyclooctadecane, further reduce and favor near-planar or elliptical shapes, like the or conformations (denoting sequences of trans and gauche bonds), where the ring adopts an elongated, oval profile to minimize both Pitzer and transannular effects, approaching the flexibility of acyclic alkanes. In heteroatom-containing macrocycles like 18-crown-6, multiple low-energy conformations such as , , and twist-boat forms are accessible, influenced by lone-pair repulsions and , enabling selective ion binding. Conformational preferences in macrocycles are probed using NMR spectroscopy, which reveals dynamic equilibria through averaging and coupling constants; for instance, 13C-NMR spectra of cycloalkanes up to C20 show distinct signals for individual conformers that merge upon ensemble averaging, allowing quantification of populated states in solution. Computational modeling complements these experiments by mapping energy landscapes, with the Merck Molecular (MMFF) widely employed for its accuracy in predicting conformer geometries and relative energies in hydrocarbons, often achieving errors below 1 kcal/ when benchmarked against quantum mechanical data. External factors significantly modulate these preferences. Solvent effects arise from differential solvation of polar or hydrophobic surfaces, shifting equilibria toward more extended conformations in polar media like versus compact ones in nonpolar solvents like . Temperature influences conformer populations via , accelerating interconversions and favoring higher-entropy forms at elevated temperatures. Substituent positions introduce additional steric or biases, with groups like methyl at 1,5-positions in medium rings exacerbating transannular and locking preferred geometries.

Stereochemical Aspects

Macrocycles exhibit diverse forms of , broadly categorized into configurational and conformational types. Configurational stereoisomerism arises from fixed geometric arrangements, such as cis-trans (or ) isomerism in unsaturated macrocycles containing double bonds. For instance, ring-closing metathesis reactions often yield mixtures of E and Z isomers in macrocyclic alkenes, as seen in the synthesis of secosteroidal macrocycles where an 8:2 E:Z ratio was obtained with 65% . In contrast, conformational stereoisomerism stems from restricted rotations leading to stable isomers, exemplified by atropisomers in biaryl-containing macrocycles. These atropisomers, such as those in aglycon models, result from hindered rotation around biaryl bonds, enabling isolation of single atropisomers with high atroposelectivity (94% for the R-configured precursor). Helical chirality in macrocycles manifests as a non-superimposable mirror image due to the twisting of the ring structure, often observed in peptide-based and allene-incorporated systems. In peptide macrocycles, stapling techniques can stabilize α-helical conformations with defined handedness, as demonstrated in mixed-stereochemistry macrocyclic N-caps that enhance helix stability through rigid constraints. Allenes contribute to helical chirality when embedded in macrocycles, where the cumulative double bonds impose orthogonal geometry and axial asymmetry, leading to stable helical architectures; for example, alleno-acetylenic macrocycles exhibit pronounced chiroptical properties due to this helical arrangement. Resolution of these helical enantiomers is commonly achieved via chiral high-performance liquid chromatography (HPLC), which separates based on differential interactions with chiral stationary phases, as applied to enantiopure allene-containing macrocycles. Planar chirality in macrocycles, prevalent in calixarenes and substituted pillararenes, results from asymmetric substitution patterns that break the plane of symmetry; for instance, A1/A2-disubstituted pillararenes exhibit planar with racemization barriers around 26 kcal/mol in nonpolar solvents like n-hexane at 298 , enabling enantiomeric stability under certain conditions. These barriers are influenced by solvent and steric bulk, with higher values preventing rapid interconversion. Stereochemical features of macrocycles are characterized using techniques like circular dichroism (CD) spectroscopy and X-ray crystallography. CD spectroscopy detects chiral asymmetries by measuring differential absorption of circularly polarized light, revealing helical or axial handedness; for example, enantiomeric Tröger's base macrocycles show mirror-image CD spectra in the UV region, confirming absolute configurations when correlated with density functional theory simulations. X-ray crystallography provides direct visualization of atomic arrangements to assign absolute configurations, as in the case of chiral macrocyclic dilactams where crystal structures resolved the stereocenters based on known reference configurations. These methods complement each other, with CD offering solution-phase insights and X-ray ensuring solid-state precision.

Synthesis

General Methods

The synthesis of macrocycles often relies on the high-dilution principle, which promotes intramolecular cyclization over intermolecular by conducting reactions at low precursor concentrations, typically below 0.01 M, to statistically favor the formation of the desired ring. This approach, first demonstrated by Paul Ruggli in 1912 for the cyclization of acyclic precursors into large rings and further developed by in 1933 using high-dilution techniques with methods such as acyloin condensation and Dieckmann cyclization, addresses the entropic disadvantage of bringing distant functional groups together in solution. The mathematical basis stems from kinetic considerations: the rate of intramolecular reaction is (proportional to concentration), while intermolecular reactions are second-order (proportional to the square of concentration), making cyclization dominant at low dilutions. For ring size feasibility, provide guidance on preferred geometries for closures, favoring exo-trig modes for rings larger than seven members to minimize strain during nucleophilic or electrophilic attacks. Template-directed synthesis enhances efficiency by preorganizing linear precursors around a central scaffold, such as metal ions or solvents, to position reactive groups for cyclization and reduce the effective dimensionality of the reaction space. A seminal example is J. Pedersen's 1967 synthesis of crown ethers, where cations like served as templates to coordinate precursors during Williamson ether formation, yielding dibenzo-18-crown-6 in high selectivity due to the ion's stabilizing pseudocavity. This method has been foundational for supramolecular macrocycles, as the template enforces proximity and orientation, often increasing yields by orders of magnitude compared to untemplated conditions. Classical reactions for macrocycle construction include the , an intramolecular Claisen-type reaction of diesters to form cyclic β-keto esters, typically under basic conditions and high dilution for rings of 9–16 members, as exemplified in early syntheses of muscone analogs where facilitated closure of long-chain adipate derivatives. Similarly, olefin metathesis emerged in the 1990s as a versatile tool for carbon-carbon bond formation in macrocycles; early applications using molybdenum-based catalysts, such as of diene substrates to afford unsaturated 12–20 membered rings in natural product precursors, demonstrated its tolerance for functional groups and mild conditions, though initial yields were modest (20–50%) without optimization. Despite these advances, challenges in general macrocycle persist, particularly in optimization, where even high-dilution conditions can produce oligomeric byproducts requiring careful control of parameters like and to achieve 30–70% cyclic product formation. Purification often involves to separate cyclic monomers from dimers and polymers based on differences in and , a labor-intensive step that underscores the need for scalable techniques in classical approaches.

Advanced Techniques and Stereocontrol

Ring-closing metathesis (RCM) using catalysts has emerged as a powerful catalytic method for constructing macrocycles, enabling efficient formation of cyclic alkenes with high . Developed by and colleagues, second-generation catalysts facilitate the cyclization of precursors under mild conditions, often achieving E/Z selectivity influenced by and substituents. For instance, in the synthesis of macrocyclic natural products, catalyst-controlled RCM allows stereoselective installation of trans-alkenes in medium to large rings, minimizing oligomerization side products. Post-2015 advances include latent sulfur-chelated iodide benzylidenes, which enhance substrate selectivity for large macrocycles (>14 members) by suppressing intermolecular metathesis, yielding up to 90% isolated products in contexts. Enyne metathesis represents a complementary catalytic approach for synthesizing unsaturated macrocycles, particularly those containing 1,3-diene motifs. Catalyzed by or complexes, this reaction couples alkynes and alkenes intramolecularly, producing cyclic enynes or dienes with defined . Recent progress since 2015 emphasizes relay enyne metathesis in green solvents and under conditions, improving yields for complex polycyclic systems derived from products. These methods have been applied in total syntheses, such as terpenoids, where the metathesis step establishes key stereocenters through subsequent transformations. Stereocontrol in macrocycle often employs chiral auxiliaries or to dictate at remote centers prior to cyclization. Chiral auxiliaries, such as oxazolidinones, direct stereoselective aldol or reactions in linear precursors, ensuring facial selectivity during ring closure. , exemplified by of allylic alcohols, generates epoxy alcohols as versatile building blocks for macrocyclic frameworks, with enantiomeric excesses exceeding 95% in syntheses like polyketides. Dynamic kinetic resolution (DKR) further advances stereocontrol by racemizing substrates while selectively cyclizing one , enabling access to enantioenriched macrocycles. Chemoenzymatic DKR using lipases and racemization catalysts forms macrolactones from secondary alcohols with >99% ee, as demonstrated in depsipeptide analogs. Organocatalytic DKR variants, such as N-heterocyclic carbene-mediated resolutions, target inherently chiral macrocycles by alkylating interconverting conformers, achieving up to 92% ee in 12-16 membered rings. Innovations from 2020 to 2025 have integrated templating and computational tools to enhance macrocycle synthesis. DNA-templated synthesis leverages scaffolds to preorganize building blocks, facilitating high-yield cyclizations via proximity effects; second-generation libraries incorporate diverse macrocyclization linkers like disulfides, yielding cyclic peptides with improved binding affinities for protein targets. Photochemical cyclizations, such as Norrish-Yang reactions, enable mild, stereoselective formation of cyclobutane-fused macrocycles in frameworks, with recent applications achieving >80% diastereoselectivity under visible light. AI-assisted design employs , such as the Macformer model, to predict the success of macrocyclization reactions by analyzing precursor structures and conformational factors, achieving high accuracy in guiding synthesis of macrocyclic candidates. A notable is the synthesis of neopeltolide, a macrolide, where RCM with Grubbs second-generation catalyst closes a 14-membered with trans-stereocontrol at the C12-C13 , guided by chiral in precursor assembly; this step proceeded in 75% , enabling evaluation of antitumor activity. Similarly, the of periplanone B, a , utilized enzymatic kinetic resolution of acetylenic alcohols with lipases to establish at C3 and C11, followed by anionic oxy-Cope rearrangement for ring expansion, affording the germacrane in enantiopure form.

Reactivity

Conformational Effects on Reactivity

In macrocycles, the Curtin-Hammett principle governs reactivity when conformers interconvert rapidly relative to the reaction timescale, such that the product ratio reflects the difference in transition-state free energies rather than ground-state populations. This is particularly relevant in systems with low barriers to or inversion, allowing conformers to contribute disproportionately if their transition states are lower in energy. For instance, in -containing 12-membered cyclic tetrapeptides, the N-acyl exhibits a low rotational barrier of 2.84–3.94 kcal/mol, enabling fast equilibration between and trans conformers; the form, favored under , directs regioselective nucleophilic attack at the α-carbon with >20:1 selectivity, as the for this pathway benefits from reduced steric hindrance. Transannular strain in medium-sized rings (8–11 members) further modulates reactivity by constraining functional group accessibility and promoting intramolecular interactions. This strain arises from non-bonded repulsions across the ring, elevating ground-state energy and lowering activation barriers for reactions that relieve it, such as transannular cyclizations or nucleophilic additions. In crown ethers like 18-crown-6, the flexible yet preorganized conformation optimizes cation coordination (e.g., K⁺ with micromolar affinity in ), which desolvates and activates counteranions, enhancing their nucleophilicity in phase-transfer . Spectroscopic methods, including infrared (IR) and Raman spectroscopy, offer direct evidence for assigning reactive conformers by probing vibrational modes sensitive to ring geometry and strain. Carbonyl stretching frequencies in IR spectra, for instance, shift between cis and trans amide conformers in strained macrocycles, correlating with their differing reactivities toward nucleophiles; Raman spectra complement this by highlighting symmetric ring deformations indicative of transannular interactions. These techniques have identified low-energy conformers in medium-ring diphosphines that predispose them to transannular bond formation. Quantitative assessments reveal that activation energy differences between pathways from preferred versus minor conformers typically span 2–5 kcal/, dictating selectivity under Curtin-Hammett conditions; such modest gaps can shift populations by factors of 10–100 at , amplifying the influence of the more reactive species. In computational studies of macrocyclic peptides, these differences arise from torsional and transannular contributions, directly impacting rates without exceeding 5 kcal/ in most flexible systems.

Peripheral Attack Model

The peripheral attack model, introduced by W. Clark Still and Igor Galynker in the early , serves as a predictive tool for in nucleophilic reactions involving macrocyclic compounds, particularly those with medium-sized rings (8–14 atoms). The model rationalizes that nucleophiles preferentially approach from the less hindered, periphery of the macrocycle rather than the more sterically congested concave face, driven by the ring's preferred low-energy conformations. This framework emphasizes the role of transannular steric interactions in dictating face selectivity during bond formation. Central to the model are conformational analyses that map the macrocycle's structures to distinguish and faces, often facilitated by sp²-hybridized centers adopting quasi-planar geometries to minimize . It has proven particularly useful for forecasting outcomes in aldol additions, where enolates attack carbonyls, and Michael additions to α,β-unsaturated enones embedded in the ring, enabling remote asymmetric with high diastereoselectivity in rigidified systems. Representative examples include stereocontrolled conjugate additions in 10- to 12-membered lactones, yielding products with predictable axial or equatorial substitution patterns. The model's predictions have been substantiated through experimental demonstrations of stereoselective nucleophilic additions. Computational validation, including (DFT) calculations, supports these observations by confirming that low-energy conformers exhibit the anticipated face differentiation; for instance, DFT optimizations of 10-membered conformers align with experimental in peripheral nucleophilic approaches. Despite its utility, the peripheral attack model has limitations, particularly in highly flexible macrocycles exceeding 20 atoms, where rapid conformational interconversions diminish reliable face discrimination and reduce predictive accuracy to moderate levels (typically 60–80% success in literature benchmarks). Post-2010 computational studies have addressed some shortcomings by incorporating effects via implicit solvent models in simulations, enhancing the model's applicability to solution-phase reactions.

Natural Occurrence

In Natural Products

Macrocycles are abundant in natural products, particularly among polyketide-derived compounds isolated from microbial sources, where they often exhibit properties. A prominent class includes antibiotics, such as erythromycin, which features a 14-membered macrocyclic ring and was isolated from the soil bacterium erythreus. Another key example is the depsipeptide valinomycin, a 36-membered macrocycle produced by species that functions as a potassium-selective , highlighting the diversity in ester-amide linkages within these structures. Natural macrocycles are sourced from diverse organisms, including and microorganisms. From marine environments, neopeltolide, a 14-membered with cytotoxic activity, was isolated from deep-water lithistid sponges of the family Neopeltidae collected off the coast of . Microbial sources dominate, with many and depsipeptides derived from actinomycetes such as , underscoring as prolific producers of structurally complex macrocycles. Structural diversity in natural macrocycles encompasses various motifs, including polyene systems in agents. , a 38-membered polyene isolated from nodosus, exemplifies this motif with its conjugated heptaene chain integrated into the ring, contributing to its broad-spectrum activity against fungal pathogens. Recent isolations from 2020 onward have expanded this diversity, including nitrogen-containing macrocycles like the cyclic depsipeptide analogs from uncultured , though natural macrocyclic arenes remain rare and often synthetic mimics of bacterial scaffolds. Biosynthesis of many natural macrocycles relies on polyketide synthases (PKS), multifunctional enzyme assemblies that iteratively build carbon chains before cyclization. In macrolide production, modular type I PKS systems, such as the erythromycin (ery), facilitate ring closure via thioesterase , with analysis revealing conserved modules for chain extension and modification. This pathway, prevalent in bacterial genomes, enables the structural variety observed in isolated macrocycles through variations in PKS architecture and post-cyclization tailoring.

Biological Roles

Macrocycles play critical roles in biological ion transport, particularly through natural ionophores that facilitate selective movement of cations across membranes. Valinomycin, a macrocyclic depsipeptide produced by species, exemplifies this function by forming a rigid cavity that coordinates ions (K⁺) with high specificity, achieving a selectivity ratio of approximately 10,000:1 over sodium ions (Na⁺). This enables valinomycin to shuttle K⁺ across lipid bilayers, disrupting electrochemical gradients in bacterial and mitochondrial membranes, which contributes to its antibiotic activity by uncoupling . Similarly, nonactin, another macrocyclic from , binds K⁺ in a cage-like structure, promoting electroneutral transport and aiding in cellular under stress conditions. In antibiotic mechanisms, macrocycles often target essential cellular processes, including protein synthesis inhibition and membrane disruption. Macrolide antibiotics such as bind to the 50S subunit of the bacterial , specifically at the center, blocking the translocation step of and halting polypeptide chain elongation. This selective inhibition spares eukaryotic ribosomes due to structural differences, making effective against Gram-positive pathogens. For membrane disruption, polyene like interact with in fungal membranes, forming barrel-shaped pores that allow ion leakage and cell lysis, a mechanism that underscores their role in innate defense against fungal infections. Macrocycles also mediate signaling in microbial communities, notably through cyclic peptides involved in quorum sensing. In Gram-positive bacteria, autoinducing peptides (AIPs) such as the thiolactone-based signals in Staphylococcus aureus form cyclic structures that bind receptor histidine kinases, triggering gene expression for virulence factors and biofilm formation at high population densities. These macrocyclic signals enable coordinated behaviors like competence induction in Bacillus subtilis, where cyclic peptides regulate sporulation and DNA uptake. Recent studies have highlighted macrocyclic polyamines in DNA regulation, where synthetic and natural variants like cyclen derivatives modulate chromatin structure and gene expression by binding DNA grooves, influencing epigenetic processes in eukaryotic cells as observed in 2023 investigations on polyamine assembly transitions. From an evolutionary perspective, macrocycles are prevalent in secondary metabolites as defense mechanisms, having diverged to counter biotic pressures like herbivory and microbial competition. Macrocyclic pyrrolizidine alkaloids in plants such as species have evolved through frequent gain and loss events, providing toxicity against herbivores via hepatotoxic effects while deterring pathogens. This prevalence reflects adaptive advantages in ecological niches, where macrocyclic structures enhance and for interspecies defense. Overall, such roles underscore the evolutionary selection for macrocycles in maintaining cellular integrity and community dynamics.

Applications

In Supramolecular Chemistry

In , macrocycles serve as versatile hosts for non-covalent assemblies, enabling selective molecular recognition and through specific interactions such as ion-dipole coordination, hydrophobic effects, and π-stacking. Crown ethers exemplify this role in host-guest chemistry, where their cyclic polyether structures coordinate cations via oxygen lone pairs aligning with the guest's electrostatic field. The seminal discovery of crown ethers by Charles J. Pedersen in 1967 introduced compounds like dibenzo-18-crown-6, which demonstrated remarkable selectivity for potassium ions (K⁺) due to a cavity size matching the of approximately 1.33 Å. For instance, 18-crown-6 forms a stable 1:1 complex with K⁺ in , exhibiting an association constant with log K ≈ 6.0, reflecting strong enthalpic contributions from multiple ion-oxygen interactions. Beyond ionic guests, macrocycles like calixarenes facilitate molecular recognition of neutral molecules through their chalice-shaped cavities, which provide hydrophobic environments for encapsulation. Calixarenes and higher homologs, first systematically explored by Gutsche in the 1980s, bind neutral aromatic or aliphatic guests via van der Waals forces and CH-π interactions, often achieving association constants in the range of 10²–10⁴ M⁻¹ in organic solvents. For example, p-tert-butylcalixarene selectively accommodates neutral guests like in its lower rim, demonstrating size and shape complementarity that stabilizes the complex without covalent bonds. Complementing this, pillararenes—rigid, pillar-shaped macrocycles introduced by Ogoshi et al. in 2008—excel in recognizing linear alkanes through their electron-rich π-cavity, promoting CH-π and hydrophobic interactions. Pillararenes, in particular, form pseudorotaxane-like complexes with n-alkanes such as n-hexane, with binding affinities enhanced by the planar arrangement of units facilitating parallel π-stacking alignments. Macrocycles also drive self-assembly into topologically complex structures like rotaxanes and catenanes, where templating directs the interlocked architecture. In rotaxanes, a macrocyclic ring threads onto a linear , often stabilized by hydrogen bonding or donor-acceptor interactions; crown ethers such as 1,5-naphtho-24-crown-8 serve as wheels in J. Fraser Stoddart's templated syntheses, where π-donor-acceptor stacking between the crown and a axle yields yields up to 70% under kinetic control. Catenanes, featuring interlocked rings, emerge similarly, with early examples using crown ethers as one ring and metal coordination for the second, as pioneered by Sauvage, but Stoddart's group advanced organic templates, achieving catenated yields exceeding 50% via clipping strategies. These assemblies rely on the macrocycle's preorganization to lower the entropic penalty of threading, enabling precise control over mechanical bonding. Such host-guest and principles underpin practical applications, including ion sensors and . Crown ether-based sensors exploit binding-induced changes in or electrochemical signals for selective ion detection; for example, 18-crown-6 derivatives in optical sensors achieve micromolar sensitivity for K⁺ via chelation-enhanced quenching, with selectivity ratios over 10³ relative to Na⁺. In , Stoddart's bistable function as switches, where or stimuli shuttle the macrocyclic component along the axle, mimicking muscle-like contraction with energies on the order of 10–20 kJ/mol per cycle, as demonstrated in rotaxane systems incorporating tetrathiafulvalene and units. These developments highlight macrocycles' role in constructing responsive supramolecular devices for and .

In Drug Discovery

Macrocycles have emerged as promising therapeutic agents in drug discovery due to their unique structural properties that address limitations of traditional small molecules and linear peptides. Their rigid scaffolds confer resistance to protease degradation, thereby improving metabolic stability and extending half-life in vivo. This rigidity also enables macrocycles to present a large, conformationally constrained surface area for binding, facilitating interactions with extended or groove-like protein interfaces that are often undruggable by conventional ligands. Notably, certain macrocycles exceed 500 Da yet achieve oral bioavailability rates of 30–40%, as seen in approved drugs, thus broadening their applicability beyond injectable formats. Despite these benefits, macrocycles face significant challenges, particularly in achieving adequate stemming from their size, polarity, and conformational flexibility. Poor membrane permeability often limits systemic exposure, necessitating innovative chemical modifications. One established strategy involves N-methylation of bonds to reduce hydrogen bonding potential and enhance , as demonstrated in cyclosporine A, a inhibitor used for that relies on multiple N-methyl groups for its oral absorption. Such approaches have been pivotal in overcoming these hurdles while preserving . FDA-approved macrocyclic drugs underscore their clinical viability across diverse therapeutic areas. Macrolide antibiotics, such as , feature a 15-membered ring and inhibit bacterial protein synthesis, providing broad-spectrum treatment for infections like and sexually transmitted diseases. Recent approvals from 2020–2025 include (2021), a non-immunosuppressive analog of cyclosporine for that modulates activity with reduced toxicity, and rezfungin (2023), a novel macrocycle for that offers once-weekly dosing due to its long . These examples illustrate macrocycles' role in addressing unmet needs in infectious diseases and autoimmune disorders. Design and optimization of macrocyclic therapeutics rely on advanced tools to navigate their synthetic . Combinatorial libraries, generated via split-pool or DNA-encoded approaches, enable rapid exploration of structural diversity, yielding hits against targets like kinases and proteases. Complementary structure-activity relationship () analyses then refine these leads, often resulting in potency gains of up to 100-fold compared to linear precursors by fine-tuning ring size, substituents, and .

Emerging Applications

Recent advancements in macrocycle research have expanded their utility into innovative areas such as bioimaging, , , and therapeutic design, leveraging their unique and host-guest properties. Fluorescent macrocycles, particularly those incorporating nitrogen-containing arene units, have emerged as promising probes for bioimaging applications due to their tunable photophysical properties and . These compounds enable high-resolution of cellular processes, with often achieved through dynamic exchange reactions that allow for modular assembly and diversification. For instance, a 2025 review highlights examples where nitrogen-rich macrocyclic arenes exhibit enhanced quantum yields (>0.5) and selectivity for biomolecular targets in live-cell imaging. In catalysis, macrocyclic ligands are increasingly employed to facilitate enantioselective reactions, providing precise control over stereochemistry in complex transformations. Recent developments include the use of chiral macrocyclophanes and pillararene derivatives as ligands in asymmetric cross-coupling reactions, achieving enantiomeric excesses exceeding 95% for the synthesis of planar-chiral frameworks. Polyamine-based macrocycles have also shown potential in CO2 capture, where their cyclic architecture enhances adsorption capacity under humid conditions. These applications underscore macrocycles' role in sustainable catalysis and carbon mitigation strategies. Within , pillararene-based polymers have been integrated into nanoparticles for , exploiting their to achieve stimuli-responsive release in tumor microenvironments. These systems demonstrate encapsulation efficiencies of over 80% for hydrophobic therapeutics, improving and reducing off-target effects in cancer models. Complementing this, macrocycle-derived sensors for ion detection have advanced environmental and biomedical monitoring, with fluorescent pillararene probes offering detection limits in the nanomolar range for heavy metal ions like mercury and lead. Such sensors leverage macrocyclic cavities for selective binding, enabling real-time fluorescence-based assays. Emerging trends in macrocycle design incorporate to optimize structures for , accelerating the discovery of eco-friendly variants with reduced synthetic footprints. In , macrocyclic variants have progressed toward , with AI-guided designs yielding potent inhibitors targeting cancer-related proteins such as . These innovations position macrocycles at the forefront of next-generation, multifunctional materials and therapeutics.

References

  1. [1]
    Defining and navigating macrocycle chemical space - PMC - NIH
    Macrocyclic compounds (MCs) – typically defined as organic compounds containing a ring of ≥12 atoms – are a chemotype of particular current interest. Certain ...
  2. [2]
    Stereoselective synthesis of macrocyclic peptides via a dual olefin ...
    Macrocyclic compounds occupy an important chemical space between small molecules and biologics and are prevalent in many natural products and pharmaceuticals.
  3. [3]
    Rapamycin-inspired macrocycles with new target specificity - PMC
    Dec 10, 2018 · Rapamycin and FK506 are macrocyclic natural products with an extraordinary mode of action—they form binary complexes with FKBP through a ...
  4. [4]
    Applications of macrocycle-based solid-state host–guest chemistry
    Oct 2, 2023 · Macrocyclic compounds in the solid state can encapsulate guests with larger affinities than their soluble counterparts. This is crucial for use in applications ...
  5. [5]
    Macrocycles in Drug Discovery—Learning from the Past for the Future
    Macrocycles usually reside in the beyond the Rule of 5 chemical space, but 30–40% of the drugs and clinical candidates are orally bioavailable.
  6. [6]
    Design, synthesis and applications of responsive macrocycles - Nature
    Dec 17, 2020 · This review introduces RMs, covering their design, synthesis and applications. It gives readers insight into the dynamic control of macrocyclic molecules.
  7. [7]
    Macrocyclization of linear molecules by deep learning to facilitate ...
    Jul 28, 2023 · In this study, we propose a computational macrocyclization method based on Transformer architecture (which we name Macformer).
  8. [8]
    macrocycle (M03662) - IUPAC Gold Book
    A cyclic macromolecule or a macromolecule cyclic portion of a macromolecule. Note: In the literature, the term 'macrocycle' is sometimes used for molecules of ...
  9. [9]
    Macrocycles: lessons from the distant past, recent developments ...
    Nov 3, 2014 · A macrocycle is a molecule that contains a cyclic framework of at least twelve atoms. Although the size of naturally occurring macrocycles can ...
  10. [10]
    Cyclic polyethers and their complexes with metal salts
    Note: In lieu of an abstract, this is the article's first page. Free ... Crown Ethers in the Group and Individual Separation of Rare-Earth Elements ...
  11. [11]
    [PDF] Nobel lecture, December 8, 1987 - CHARLES J. PEDERSEN
    I applied the epithet "crown" to the first member of this class of macrocyclic polyethers because its molecular model looked like one and, with it, cations ...Missing: history | Show results with:history
  12. [12]
    Charles J. Pedersen's legacy to chemistry - RSC Publishing
    Apr 11, 2017 · The serendipitous discovery in 1961 of dibenzo-18-crown-6 by Charles J. Pedersen marked the beginning of research on cyclic polyether macrocyclic compounds.
  13. [13]
    Strained Cyclophane Macrocycles: Impact of Progressive Ring Size ...
    The strain energy of these 'small ring' macrocycles results from distortion of standard bond angles and lengths associated with enclosure in a ring. In addition ...
  14. [14]
    Progress in Purely Organic Macrocyclic Emitters: From Molecular ...
    Sep 8, 2025 · Flexible macrocycles, often composed of ether linkages or alkyl chains, combine flexibility with unique cavity structures.,, Being nonconjugated ...
  15. [15]
    Hydrocarbon Macrocycle Conformer Ensembles and 13 C-NMR ...
    Jan 31, 2022 · Except for the small rings, cyclododecane represents the system with the strongest high-field shift of all investigated compounds. From ...
  16. [16]
  17. [17]
    Conformational analysis of cycloalkanes | ChemTexts
    Aug 12, 2015 · Conformational analysis studies the kinetic and thermodynamic properties of molecules, including preferred conformations, energies, and ...Missing: hydrocarbons | Show results with:hydrocarbons
  18. [18]
    The conformation of cyclooctane - AIP Publishing
    Probably the strongest experimental evidence that cyclo- octane in the liqUid phase is predominantly in the boat- chair form comes from the high resolution NMR ...<|separator|>
  19. [19]
    cyclooctadecane, cyclononadecane and cycloicosane - PubMed
    The conformations of cyclooctadecane, cyclononadecane, and cycloicosane were generated by a stochastic program that works in conjunction with MM2.
  20. [20]
    Hydrocarbon Macrocycle Conformer Ensembles and 13 C-NMR ...
    Jan 31, 2022 · Quantum mechanical investigations of hydrocarbon macrocycle ensembles in solution by comparison of experimental and computed 13 C-NMR spectra are described.
  21. [21]
    Benchmark assessment of molecular geometries and energies from ...
    This study provides an extensive test of the performance of different molecular mechanics force fields on a diverse molecule set.
  22. [22]
    Addressing Challenges of Macrocyclic Conformational Sampling in ...
    As the conformational spaces are similar in polar solvents, we investigate conformations of the unprotonated macrocycles in water and chloroform to study ...
  23. [23]
    Effects of solvent polarity and temperature on the conformational ...
    Effects of solvent polarity and temperature on the conformational statistics of a simple macrocyclic polyether ... A Theoretical Case Study of Substituent Effects ...
  24. [24]
    Macrocyclization Reactions: The Importance of Conformational ...
    Aug 6, 2015 · In this work, the term “macrocyclic structure” applies to compounds displaying a ring arrangement of atoms attached through covalent bonds.
  25. [25]
    Atropisomerism in the Pharmaceutically Relevant Realm
    Sep 26, 2022 · While atropisomerism is exemplified by biaryls, it is observed in many other pharmaceutically relevant scaffolds including heterobiaryls, ...
  26. [26]
    Mixed Stereochemistry Macrocycle Acts as a Helix-Stabilizing ...
    Sep 4, 2024 · This mixed stereochemistry macrocyclic N-cap is synthetically accessible, requiring only minor modifications to standard solid-phase peptide synthesis.
  27. [27]
  28. [28]
    Pillar-Shaped Macrocyclic Hosts Pillar[n]arenes: New Key Players ...
    Pillar[n]arenes also have several other characteristic features such as superior host–guest abilities, planar chirality, and the ability to undergo self- ...
  29. [29]
    Resolution and Racemization of a Planar-Chiral A1/A2-Disubstituted ...
    Jun 9, 2019 · The present study has provided, for the first time, thermodynamic parameters of the pillararenes in different solvents that will serve as an ...Missing: barriers | Show results with:barriers
  30. [30]
    Stereochemical Inversion of Rim-Differentiated Pillar[5]arene ...
    Aug 21, 2020 · We synthesized a series of novel rim-differentiated P[5]s with various substituents and examined their rapid rotations by variable-temperature NMR (203–298 K).
  31. [31]
    Chiroptical Signature and Absolute Configuration of Tröger's Base ...
    In this paper, we have addressed the absolute configuration determination of a Tröger's base-based triangular macrocycle, 3MC, by means of vibrational and ...Figure 2 · Figure 4 · Figure 6
  32. [32]
    Design and synthesis of a new chiral macrocyclic dilactam and its ...
    Dec 12, 2024 · The absolute configuration of the four chiral centers in 1 was confirmed via X-ray diffraction, based on the known absolute configuration of the ...Materials And Methods · Nmr · Results And Discussion
  33. [33]
    Macrocyclization Reactions at High Concentration (≥0.2M)
    Jul 1, 2023 · However, most current syntheses still require highly diluted conditions (generally < 0.1M), which makes the synthesis extremely inefficient in ...INTRODUCTION · HIGH CONCENTRATED... · CONCLUSIONS · References
  34. [34]
    Ring Closing Metathesis (RCM) - Organic Chemistry Portal
    The Ring-Closing Metathesis (RCM) allows synthesis of 5- up to 30-membered cyclic alkenes. The E/Z-selectivity depends on the ring strain.
  35. [35]
    Synthesis of macrocyclic natural products by catalyst-controlled ...
    We outline a reliable, practical and general approach for efficient and highly stereoselective synthesis of macrocyclic alkenes by catalytic RCM.
  36. [36]
    Highly Substrate‐Selective Macrocyclic Ring Closing Metathesis
    Feb 22, 2022 · A selective ring-closing metathesis (RCM) reaction for the formation of large macrocycles by using latent sulfur chelated ruthenium iodide benzylidenes.
  37. [37]
    Recent Progress on Enyne Metathesis: Its Application to Syntheses ...
    Olefin metathesis using ruthenium carbene complexes is a useful method in synthetic organic chemistry. Enyne metathesis is also catalyzed by these complexes ...
  38. [38]
    Recent advances in enyne metathesis in non-usual media
    Oct 21, 2025 · This review presents seminal advances in enyne metathesis in non-usual media, including green solvents, supercritical fluid and microwave ...
  39. [39]
    (PDF) Sharpless Asymmetric Epoxidation: Applications in the ...
    Oct 30, 2022 · This review discusses an important synthetic tool proposed by KB Sharpless in 1980, known as the Sharpless asymmetric epoxidation of allylic alcohols.
  40. [40]
    Chemoenzymatic synthesis of macrocycles via dynamic kinetic ...
    Jul 2, 2024 · An ideal strategy for macrolactonization is via dynamic kinetic resolution (DKR), which involves the simultaneous formation of the ester bond and introduction ...
  41. [41]
    Catalytic Enantioselective Synthesis of Inherently Chiral ...
    We report herein DKR of inherently chiral macrocycles through enantioselective alkylation of one of the two rapidly interconverting conformers.Supporting Information · Author Information · References
  42. [42]
    Second-generation DNA-templated macrocycle libraries for the ...
    Here, we have developed and streamlined multiple fundamental aspects of DNA-encoded and DNA-templated library synthesis methodology, including computational ...
  43. [43]
    Recent advances in Norrish-Yang cyclization and dicarbonyl ...
    Oct 30, 2025 · This review summarizes the latest advancements in these reactions for constructing terpenoids, alkaloids, and antibiotics. Through Norrish-Yang ...
  44. [44]
    Accurate de novo design of high-affinity protein-binding macrocycles ...
    Jun 20, 2025 · Here we introduce RFpeptides, a denoising diffusion-based pipeline for designing macrocyclic binders against protein targets of interest.Missing: photochemical | Show results with:photochemical
  45. [45]
    Stereocontrolled synthesis of the macrolactone core of neopeltolide
    A stereoselective synthesis of the macrolactone core of neopeltolide is described. The tetrahydropyran moiety was constructed via the intramolecular ...
  46. [46]
    Organic Synthesis in Pheromone Science - PMC - PubMed Central
    Addition of lithium trimethylsilylacetylide to 10 gave acetylenic alcohol 11, the substrate for the enzymatic kinetic resolution. Similarly, its (3RS,11S)- ...
  47. [47]
  48. [48]
  49. [49]
    Chemical consequences of conformation in macrocyclic compounds
    Journals & Books · View PDF; Download full issue. Search ScienceDirect. Elsevier · Tetrahedron · Volume 37, Issue 23, 1981, Pages 3981-3996. Tetrahedron ...
  50. [50]
    Divergent synthesis of ten-membered lactones: Aspinolides C, F, G ...
    Jun 12, 2023 · On the basis of the peripheral attack model for medium-to-macrocyclic compounds described by Still [6], we carried out a conformational ...
  51. [51]
    [PDF] Medium and Large Rings - Eugene E. Kwan
    “Cyclooctane is unquestionably the conformationally most complex cycloalkane owing to the existence of so many forms of comparable energy…”" -Hendrickson, J. B. ...Missing: hydrocarbons | Show results with:hydrocarbons
  52. [52]
    Macrocyclic Drugs and Synthetic Methodologies toward ... - NIH
    The commonly used synthetic methodologies toward macrocyclization include macrolactonization, macrolactamization, transition metal-catalyzed cross coupling, ...
  53. [53]
    Biosynthesis of depsipeptides, or Depsi: The peptides with varied ...
    Oct 19, 2020 · Important natural product depsipeptides include the piscicide antimycin, the K+ ionophores cereulide and valinomycin, the anticancer agent ...
  54. [54]
    Neopeltolide, a Macrolide from a Lithistid Sponge of the Family ...
    A new marine-derived macrolide designated as neopeltolide (1) has been isolated from a deep-water sponge of the family Neopeltidae.
  55. [55]
    Eucannabinolide, a novel sesquiterpene lactone, suppresses ... - NIH
    Euc elicits the effects of anti-proliferation, anti-metastasis and anti-breast cancer stem cell-like traits in TNBC via targeting STAT3.Missing: plant | Show results with:plant
  56. [56]
    Amphotericin B and Other Polyenes—Discovery, Clinical Use, Mode ...
    Nov 27, 2020 · Polyenes such as amphotericin B have a controversial image. They are the antifungal drug class with the broadest spectrum, resistance ...
  57. [57]
    Cyclic natural product oligomers: diversity and (bio)synthesis of ...
    Nov 25, 2024 · In this review we focus on natural products that involve macrocyclization in their biosynthesis and chemical synthesis, with an emphasis on the function and ...<|control11|><|separator|>
  58. [58]
    Evolution and Diversity of Assembly-Line Polyketide Synthases
    Dec 15, 2019 · In this review, we provide an overview of previous studies that investigated PKS evolution and propose a model that challenges the currently prevailing view.
  59. [59]
    Valinomycin - an overview | ScienceDirect Topics
    Valinomycin is cyclic peptide produced by several species of streptomyces, and it is a natural neutral ionophore that causes permeability of biological ...
  60. [60]
    Ionophore - an overview | ScienceDirect Topics
    Since valinomycin has no net charge it accepts the charge of the K+ but remains lipophilic and freely diffuses through mitochondrial membranes in either the ion ...
  61. [61]
    Understanding the evolution of macrolides resistance: A mini review
    Their mechanism of action involves binding to the bacterial ribosome—the fundamental site of protein synthesis that leads to inhibition of the translation ...
  62. [62]
    How do the polyene macrolide antibiotics affect the ... - PubMed
    It has been shown that the action of polyene antibiotics on cells is not always detrimental: at sub-lethal concentrations these drugs stimulate either the ...
  63. [63]
    Cyclodepsipeptides produced by actinomycetes inhibit ... - PubMed
    Cyclic peptides are commonly used as quorum-sensing autoinducers in Gram-positive Firmicutes bacteria. Well-studied examples of such molecules are thiolactone ...
  64. [64]
    Order–order assembly transition-driven polyamines detection based ...
    Jul 7, 2023 · Here, we display a sensing mode for polyamine detection based on an order–order transition of iron–sulfur complexes upon their assembly.
  65. [65]
    Frequent gain and loss of pyrrolizidine alkaloids in the evolution of ...
    Pyrrolizidine alkaloids (PAs) of the macrocyclic senecionine type are secondary metabolites characteristic for most species of the genus Senecio (Asteraceae).
  66. [66]
    Convergent and divergent evolution of plant chemical defenses
    26 The majority of plant-derived metabolites are constantly evolving to defend against insects and other herbivores, providing impressive chemical diversities.
  67. [67]
    Stability constants of cyclic polyether complexes with univalent cations
    Single-molecule study reveals ion-dependent conformational change and nanomechanical property of crown ether-based polymer.
  68. [68]
    Macrocycles in Drug Discovery Learning from the Past for the Future
    Apr 5, 2023 · We have analyzed FDA-approved macrocyclic drugs, clinical candidates, and the recent literature to understand how macrocycles are used in drug discovery.
  69. [69]
    FDA‐approved drugs featuring macrocycles or medium‐sized rings
    Jan 25, 2025 · This review presents 17 FDA-approved macrocyclic drugs during the past 5 years, highlighting their importance and critical role in modern ...
  70. [70]
    Strategies and Technologies for Drug Discovery - NIH
    Apr 24, 2025 · This review discusses recent innovations in synthetic and computational methodologies that have advanced macrocycle drug discovery over the past five years.Missing: peripheral attack
  71. [71]
    Catalytic Enantioselective Access to Planar‐Chiral Macrocyclophanes
    May 29, 2025 · In this review, it summarizes recent advancements in the catalytic enantioselective synthesis of planar-chiral macrocyclophanes, focusing on various synthetic ...
  72. [72]
    Enantioselective construction of inherently chiral pillar[5]arenes via ...
    Mar 10, 2025 · In this study, we develop an asymmetric extended side-arm Suzuki–Miyaura cross-coupling strategy to construct inherently chiral pillar[5]arenes ...
  73. [73]
    Macrocyclic receptors for anion recognition - RSC Publishing
    Sep 19, 2024 · Macrocyclic receptors have emerged as versatile and efficient molecular tools for the recognition and sensing of anions.
  74. [74]
    Recent Developments of Fluorescence Sensors Constructed ... - MDPI
    This review comprehensively discusses, for the first time, the recent innovations in the synthesis and self-assembly of pillar[n]arene-based supramolecular ...
  75. [75]
    HELM-GPT: de novo macrocyclic peptide design using generative ...
    Jun 12, 2024 · Macrocyclic peptides hold great promise as therapeutics targeting intracellular proteins. This stems from their remarkable ability to bind flat ...