Crown ethers are a family of macrocyclic polyethers consisting of a cyclic molecular structure with multiple oxygen atoms integrated into the ring, forming a cavity that can selectively encapsulate and bind metal cations via ion-dipole interactions between the cation and the ether oxygen atoms.[1] First synthesized and characterized by Charles J. Pedersen at DuPont in 1967, these compounds were named "crown ethers" due to the crown-like appearance of their molecular models and their capacity to "crown" cations with high specificity based on ring size and ion diameter.[1] The nomenclature for crown ethers follows the format n-crown-m, where n denotes the total number of atoms in the ring and m the number of oxygen atoms; for instance, 18-crown-6 features an 18-membered ring with six oxygen atoms, optimally suited for binding potassium ions.[1]These molecules exhibit remarkable selectivity in complexation, with binding affinity determined by the match between the cation's ionic radius and the size of the crown ether's cavity—smaller rings like 12-crown-4 prefer lithium ions, while larger ones such as 15-crown-5 and 18-crown-6 favor sodium and potassium, respectively.[1] Variations including benzo-substituted or aza-crown ethers expand their structural diversity, enhancing solubility or introducing additional binding sites for applications beyond simple ion solvation.[2] Pedersen's discovery earned him the 1987 Nobel Prize in Chemistry (shared with Jean-Marie Lehn and Donald Cram) for foundational work in supramolecular chemistry, highlighting crown ethers' role in mimicking biological ion channels and enabling controlled ion transport.[3]Crown ethers have found widespread use in chemical analysis, phase-transfer catalysis, and ion extraction processes due to their ability to solubilize inorganic salts in nonpolar organic solvents by forming stable, lipophilic complexes.[4] In analytical chemistry, they serve as selective carriers for alkali and alkaline earth metals in ion-selective electrodes and chromatographic separations, improving detection limits and specificity. More recently, functionalized crown ethers have been incorporated into sensors for environmental monitoring of heavy metals,[5] drug delivery systems to enhance ocular permeability,[6] and solvent extraction processes for rare-earth element separation,[7] underscoring their versatility in modern supramolecular and materials science.
Structure and Nomenclature
Definition and Basic Structure
Crown ethers are macrocyclic polyethers composed of repeating -CH₂-CH₂-O- units that form a closed ring, with the oxygen atoms oriented inward to line a central cavity capable of coordinating cations through electrostatic interactions. These neutral, synthetic compounds were first described as capable of forming stable complexes with alkali metal ions, marking a foundational example of host-guest chemistry in supramolecular systems. Macrocycles, such as crown ethers, are large cyclic molecules (typically exceeding 10 atoms in the ring) that enable non-covalent interactions beyond traditional covalent bonding, a core principle of supramolecular chemistry.The basic structure of crown ethers consists of ethylene oxide oligomers, generally featuring 3 to 10 oxygen atoms separated by two methylene groups each, yielding rings of 12 to 30 atoms in total. This architecture results in a torus-like (doughnut-shaped) form, where the cavity diameter varies with ring size to accommodate different guest ions. The nomenclature follows the convention -crown-n, where m represents the total number of ring atoms and n the number of oxygen atoms; for instance, 12-crown-4 has 12 atoms including four oxygens, while 18-crown-6 is an 18-membered ring with six oxygens.A distinctive feature of crown ethers is their conformational flexibility, often adopting a puckered, crown-like arrangement rather than a fully planar one, which positions the oxygen lone pairs toward the interior for optimal coordination. The exterior surface, formed by the hydrocarbon chains, is hydrophobic, contrasting with the hydrophilic interior created by the polar ether oxygens. This amphiphilic character and the preorganized cavity size facilitate selective inclusion in host-guest assemblies, as originally observed in compounds like dibenzo-18-crown-6 discovered by Charles J. Pedersen in 1967.
Nomenclature and Examples
Crown ethers are named using a trivial convention introduced by Charles J. Pedersen in his seminal work, where the descriptor -crown-n indicates a macrocyclic ring with m total atoms and n oxygen donor atoms arranged at regular intervals.[8] This system simplifies identification, as the systematic International Union of Pure and Applied Chemistry (IUPAC) names employ von Baeyer-style heterocyclic nomenclature, specifying the positions of oxygen atoms in the cycloalkane chain; for instance, 18-crown-6 corresponds to 1,4,7,10,13,16-hexaoxacyclooctadecane.[9]Representative examples illustrate the structural progression in ring size and donor count. The smallest widely studied crown ether, 12-crown-4, features a 12-membered ring with four evenly spaced oxygen atoms, systematically named 1,4,7,10-tetraoxacyclododecane, forming a compact cavity suitable for smaller guests.[10] Larger analogs include 15-crown-5 (1,4,7,10,13-pentaoxacyclopentadecane), with five oxygens in a 15-atom ring, and 18-crown-6 (1,4,7,10,13,16-hexaoxacyclooctadecane), the archetypal member known for its symmetric, nearly planar conformation in solution.[11] A substituted variant, dibenzo-18-crown-6, incorporates two ortho-substituted benzene rings fused to the polyether chain at positions that enhance conformational rigidity, with the IUPAC name 6,7,9,10,17,18,20,21-octahydrodibenzo[b,k][1,4,7,10,13,16]hexaoxacyclooctadecin; this compound was among Pedersen's original syntheses and exemplifies how aromatic substituents modify the basic scaffold.[8][12]Structural diversity extends to larger rings and heteroatom substitutions while retaining the crown nomenclature framework. For instance, 24-crown-8 comprises eight oxygens in a 24-membered ring, named systematically as 1,4,7,10,13,16,19,22-octaoxacyclotetracosane, accommodating bulkier species due to its expanded cavity.[13] Heteroatom variants replace oxygens with sulfur or nitrogen: thiacrown ethers, such as 18-thiacrown-6, incorporate sulfur donors for altered binding properties and follow analogous naming with a "thia-" prefix indicating the number and positions of sulfur atoms; azacrown ethers substitute nitrogen, as in 1-aza-18-crown-6 (systematically 1,4,7,10,13-pentaoxa-16-azacyclooctadecane), where the amine group introduces potential for protonation or further functionalization.[14][15] These modifications, building on Pedersen's designations, allow tailored cavity electronics and geometry without altering the core cyclic polyether motif.[8]
History
Discovery by Charles Pedersen
Charles J. Pedersen, an organic chemist at E. I. du Pont de Nemours and Company, was investigating metal deactivators for petroleum products and lubricants during the early 1960s. His research focused on phenolic antioxidants to prevent oxidative degradation catalyzed by trace transition metals such as copper and vanadium. In this context, Pedersen sought to synthesize bidentate and multidentate phenolic ligands to study their coordination with the vanadyl group and suppress its catalytic activity.[16]The accidental discovery occurred in 1960 when Pedersen attempted to prepare bis[2-(o-hydroxyphenoxy)ethyl] ether using a partially protected catechol contaminated with approximately 10% free catechol, along with bis(2-chloroethyl) ether under basic conditions with sodium hydroxide. This reaction unexpectedly yielded a small amount (about 0.4%) of white, fibrous crystals that were insoluble in most solvents but showed unusual solubility behavior in methanol upon addition of sodium ions. Further investigation revealed these crystals to be a cyclic polyether, later identified as dibenzo-18-crown-6, with a molecular weight twice that of the expected product, confirmed by spectral analysis. To reproducibly synthesize such cyclic compounds, Pedersen employed high-dilution conditions with catechol and diethylene glycol ditosylate, leading to the formation of cyclic byproducts like dibenzo-14-crown-4.[16][8]A key observation highlighting the unique properties of these compounds came from their ability to enhance the solubility of alkali metalsalts in nonpolar solvents. Specifically, dibenzo-18-crown-6 formed stable, colorless crystalline complexes with potassiumpicrate, rendering it soluble in benzene and other nonpolar media, in stark contrast to the insolubility of the salt alone or with linear polyethers. This indicated selective complexation of the potassium cation within the macrocyclic cavity. Dibenzo-18-crown-6 emerged as the first reported crown ether with particular affinity for K⁺ binding.[8]Pedersen's findings were detailed in his seminal 1967 publication in the Journal of the American Chemical Society, marking the initial disclosure of crown ethers and their alkali metal complexes. These early colorless crystals demonstrated remarkable stability in forming host-guest complexes with alkali metals, distinguishing them from non-cyclic polyethers that failed to exhibit such behavior.[8]
Subsequent Developments and Recognition
Following Pedersen's initial publication in 1967, research on crown ethers expanded significantly during the 1970s, with chemists developing efficient syntheses for unsubstituted crown compounds, such as 18-crown-6, through methods like the reaction of polyethylene glycols with sulfonyl chlorides, achieving good yields without high-dilution techniques.[17] Researchers including Reed M. Izatt contributed to this progress by exploring macrocyclic polyethers and their ion-binding properties in comprehensive reviews and synthetic studies.[18] Concurrently, Jean-Marie Lehn and Donald J. Cram advanced the field by integrating crown ethers into supramolecular chemistry, with Lehn synthesizing bicyclic cryptands in 1969 for enhanced ion selectivity and Cram designing chiral crown derivatives for stereospecific binding.[19]Pedersen himself extended his work through a series of publications from 1967 to 1979, detailing the synthesis of approximately 60 macrocyclic polyether derivatives containing 12 to 60 ring atoms and 4 to 10 heteroatoms, including variations with nitrogen or sulfur.[16] Key papers included reports on their preparation and complexation behavior in Journal of the American Chemical Society (1967), Journal of Organic Chemistry (1971), and Angewandte Chemie (1972).[16] In parallel, George W. Gokel pioneered the synthesis of azacrown ethers in the 1970s, replacing oxygen atoms with nitrogen to create ligands with improved solubility and binding for transition metals, as detailed in early communications and subsequent reviews.[20]The foundational impact of these efforts was recognized in 1987 when Charles J. Pedersen, Donald J. Cram, and Jean-Marie Lehn shared the Nobel Prize in Chemistry "for their development and use of molecules with structure-specific interactions of high selectivity," highlighting crown ethers as pioneers in host-guest chemistry.[19] This recognition underscored the molecules' role in mimicking enzymatic selectivity and laid the groundwork for supramolecular systems.[19]By the 1970s, crown ethers saw early commercialization, particularly in ion-selective electrodes for analytical applications like potassium detection in clinical settings, leveraging their carrier properties in neutralmembrane systems introduced around that time.[19][21] Their influence extended to host-guest chemistry, becoming a cornerstone of modern supramolecular science by enabling selective molecular recognition and transport.[19]
Synthesis
Classical Methods
The classical synthesis of crown ethers, pioneered by Charles J. Pedersen in the 1960s, primarily relies on high-dilution cyclization reactions to favor intramolecular ring formation over undesirable polymerization. In this approach, diols such as ethylene glycol or catechol are reacted with dihalides (e.g., 1,2-dibromoethane) or ditosylates under basic conditions, typically involving alkali metal hydroxides like sodium or potassium hydroxide. The high-dilution technique involves slow addition of reactants to a large volume of solvent, maintaining concentrations around 0.001 M to enhance cyclization yields, which can reach 40-50% for key compounds. For instance, the first crown ether, dibenzo-18-crown-6, was synthesized by treating catechol with bis(2-chloroethyl) ether in the presence of sodium hydroxide, yielding up to 45% of the cyclic product as white fibrous crystals after purification.[8][16]Template synthesis represents a refinement of these methods, where metal ions such as K⁺ serve to preorganize the linear precursors around the cation, thereby directing efficient ring closure and improving yields. This templating effect is particularly effective for larger rings, as the ion coordinates with oxygen atoms in the oligoethylene glycol chains, stabilizing the transition state for cyclization. A seminal example is the preparation of 18-crown-6, achieved by refluxing triethylene glycol with 1,2-bis(2-chloroethoxy)ethane and potassium hydroxide in tetrahydrofuran for 18-24 hours, followed by evaporation, extraction with dichloromethane, distillation under reduced pressure (100-167°C at 0.2 mm Hg), and recrystallization from acetonitrile to afford colorless crystals with a yield of 38-44% crude and 56-66% overall after purification.[8]This scheme, employed by Pedersen for synthesizing over 50 crown ethers, underscores the versatility of cyclization in basic media, with the choice of alkali metal influencing selectivity and efficiency due to templating.[8]
Modern and Functionalized Syntheses
Direct functionalization of crown ethers typically occurs post-cyclization to introduce targeted substituents, simplifying the preparation of derivatives with enhanced properties. Alkylation via radical-mediated cross-dehydrogenative coupling (CDC) allows attachment of alkyl, cyano, or alkynyl groups to aliphatic crowns; for example, photochemical activation of 15-crown-5 with cyanoarenes in the presence of an iridium photocatalyst yields functionalized products in 71-84% efficiency under mild conditions. Incorporation of chromophores, such as BODIPY units on formylated benzo-15-crown-5 via the Duff reaction followed by condensation, enables optical signaling upon ion binding, with yields exceeding 70% in key steps. These modifications expand crown ether utility in sensing and catalysis without disrupting the macrocyclic core.Thia-crown ethers, featuring sulfur atoms substituted for oxygen, are prepared to enable selective coordination of soft ions like transition metals due to sulfur's polarizability. Synthesis often involves reacting oligoethylene glycol dichlorides with dimercaptans or sodium sulfide under high-dilution conditions, yielding unsubstituted thia-18-crown-6 in 20-40% isolated amounts after purification. Further post-cyclization thiol functionalization, such as converting sulfonated dibenzo-crowns to thiols via reduction, facilitates grafting onto surfaces or incorporation into hybrid materials.Advanced methods leverage templating and plasma techniques for complex architectures. The crown ether active template synthesis (CEATS), introduced in 2023, employs kinetic control to assemble rotaxanes by stabilizing transition states within the crown cavity; for instance, 24-crown-8 threads onto an axle via SN2 reaction of a primary amine with an alkyl bromide in chloroform, affording the rotaxane in 48% yield with >100:1 selectivity over non-interlocked products. In 2024, the solution plasma process (SPP) embedded 18-crown-6 into graphene sheets through C-H activation and polymerization of aromatic precursors like benzene in a DMF-ethanol mixture under nonequilibrium plasma, producing 2D materials with ordered cavities verified by TEM and XPS for ion-selective membrane applications.One-step condensations have streamlined aza-crown ether production. Rh(II)-catalyzed [3+6+3+6] cycloadditions of morpholines with α-diazo-β-ketoesters generate diaza macrocycles in a single pot, proceeding via metal carbene intermediates to form the ring with high regioselectivity, as yields reach 50-70% for symmetric derivatives under reflux in benzene. Molecular layer deposition (MLD) in 2025 enabled polymeric crown ethers by alternating exposures to 4,10-diaza-15-crown-5 and malonyl chloride vapors, yielding conformal films with a growth rate of 0.8-1.2 Å per cycle for ion-conducting coatings.Recent innovations include crown ether-cycloparaphenylene (CPP) hybrids synthesized in 2025 for supramolecular applications. These multimacrocycles, such as [22]CPP fused with 18-crown-6, were obtained via Suzuki-Miyaura coupling of iodinated dibenzo-crown ethers with a diboronate precursor using [DPPF Pd G4] catalyst and K3PO4 in 1,4-dioxane-water at 85°C for 24 hours (70-85% for intermediates), followed by reductive aromatization with sodium naphthalenide in THF at -78°C, isolating the target in 20% overall yield. The trimacrocycle analog with 24-crown-8 achieved 10% yield under similar conditions, demonstrating scalability through standard Pd-catalyzed protocols adaptable to larger scales with optimized purification.
Properties
Cation Affinity and Selectivity
Crown ethers bind metal cations primarily through coordination of the lone pairs on their oxygen atoms to the positively charged ion, forming a host-guestcomplex stabilized by electrostatic and ion-dipole interactions.[23][24] The macrocyclic ring adopts a roughly circular conformation, with the cation positioned in the central cavity, where the oxygen donors point inward to encircle the guest. This binding is most effective when the cavity size of the crown ether matches the ionic radius of the cation, allowing for optimal orbital overlap and minimal strain. For instance, 18-crown-6 has a cavity radius of 1.34–1.55 Å, which complements the potassiumion (K⁺) with an ionic radius of 1.33 Å.[16][25]Selectivity in crown ether-cation binding arises mainly from principles of size complementarity, where the ring's cavity diameter aligns with the guest ion's dimensions to maximize enthalpic contributions from coordination while minimizing entropic penalties from conformational restriction. Smaller crowns like 12-crown-4, with a cavity suited to ions around 0.74 Å, preferentially bind lithium (Li⁺), whereas 15-crown-5 favors sodium (Na⁺) at approximately 0.98 Å.[26][24] This size-based discrimination follows the hard-soft acid-base (HSAB) theory to some extent, with neutral oxygen donors in crown ethers acting as hard bases that prefer hard cations like alkali metals over softer ones. Additionally, the Hofmeister series can influence selectivity in aqueous environments by modulating solvation shells around ions, affecting competition between crown binding and hydration.[27] The stability of these complexes is quantified by the stability constant K, where \log K = -\Delta G^\circ / (RT \ln 10), and the binding free energy \Delta G^\circ incorporates enthalpic terms from ion-dipole interactions and solvation changes, as well as entropic contributions from desolvation and ligand wrapping.[28]Representative examples illustrate these affinities: 18-crown-6 forms a stable 1:1 complex with K⁺ in water, with \log K \approx 6 at 25°C, reflecting strong selectivity over Na⁺ (\log K \approx 4.3) due to better size fit. Crown ethers also bind ammonium ions (NH₄⁺) through hydrogen bonding between the ion's N-H protons and ether oxygens, often with affinities comparable to K⁺ in non-aqueous solvents, enabling applications in recognition of protonated amines. Substituents on the crown ring, such as benzo groups or alkyl chains, enhance selectivity by increasing rigidity to preorganize the cavity or modulating donor strength, as seen in dibenzo-18-crown-6, which shows heightened affinity for K⁺ over Rb⁺ compared to the parent compound.[16]Several factors modulate crown ether selectivity beyond intrinsic structure. Solvent polarity plays a key role, as protic solvents like water stabilize hydrated ions and reduce binding constants (e.g., \log K for 18-crown-6–K⁺ drops from ~10 in nonpolar solvents like 1,2-dichloroethane to ~6 in water), while aprotic media enhance complexation by weakening ionsolvation.[29]Temperature influences stability through van't Hoff effects, with endothermic desolvation often offset by exothermic coordination, leading to decreased K at higher temperatures. Comparisons with variants reveal that neutral oxygen-based crowns excel for hard s-block cations, whereas charged crowns or those with softer sulfur/nitrogen donors (e.g., thia- or aza-crowns) improve selectivity for borderline or soft ions like transition metals.[30][31][32]
Spectroscopic and Thermodynamic Properties
Crown ether complexation with cations is characterized by thermodynamic parameters that reflect enthalpic contributions from ion-dipole interactions between the metal ion and ether oxygen atoms, as well as entropic effects arising from desolvation of the cation and ligand. For the 1:1 complex of 18-crown-6 with K⁺ in methanol at 25 °C, the Gibbs free energy change (ΔG) is approximately -7 kcal/mol, driven primarily by a favorable enthalpy (ΔH ≈ -11 kcal/mol) with an unfavorable entropy term (TΔS ≈ -4 kcal/mol). Similar trends hold for related systems, such as N-benzylaza-18-crown-6 with K⁺, where log K_s = 5.00, confirming the dominance of enthalpic stabilization in polar solvents.[33]Spectroscopic techniques provide insights into conformational dynamics and binding interactions. Nuclear magnetic resonance (NMR) spectroscopy reveals shifts in the CH₂ proton signals of the ethylene units upon cation binding, indicative of conformational adjustments in the macrocycle. For instance, in 18-crown-6 complexes with alkali cations like K⁺ and Ba²⁺, the CH₂-O-CH₂ and CH₂-N-CH₂ protons shift upfield by 0.13–0.23 ppm due to the deshielding effect of the encapsulated ion, as observed in electrophoretic NMR studies. Infrared (IR) spectroscopy detects changes in the asymmetric C-O-C stretching vibrations, typically in the 950–1100 cm⁻¹ region; upon complexation, these bands shift to lower frequencies, such as from 1128 cm⁻¹ to 1116 cm⁻¹ in dibenzo-18-crown-6 with metal ions, reflecting altered ether bond polarities.[34][35][36]Ultraviolet-visible (UV-Vis) spectroscopy is particularly useful for chromophore-substituted crown ethers, where ionbinding induces spectral shifts due to changes in electronic conjugation or charge transfer. In benzothiazole-appended crown ethers, alkali metal complexation causes minimal absorption changes but can lead to red shifts in related protonated forms, enhancing sensitivity for detection.[37]Transport properties of crown ether-cation complexes highlight their role in ion mobility and diffusion. In aqueous solutions, the self-diffusion coefficient of Na⁺ decreases significantly upon binding to 18-crown-6, forming stable aggregates that reduce mobility compared to free ions, as measured by pulsed-field gradient NMR. In polymer membranes functionalized with 12-crown-4, the diffusion coefficient of Li⁺ exceeds that of Na⁺ by factors influencing selectivity, with values on the order of 10⁻⁶ cm²/s for monovalent ions.[38][39]Luminescent derivatives of crown ethers exhibit fluorescence modulation upon ion binding, enabling probe design. BODIPY-equipped benzo-18-crown-6 shows a fluorescence enhancement factor of 4.7 with K⁺ (K_d = 18 mM), attributed to inhibition of photoinduced electron transfer (PET), while the benzo-15-crown-5 analog yields a factor of 7.3 for Na⁺ (K_d = 276 mM); these responses are pH-independent over 3.5–9.5. Conversely, some systems display quenching, such as in benzoxazolyl-alanine crown ethers with Pd²⁺, via chelation-enhanced quenching (CHEQ).[40][41]
Applications
Phase Transfer Catalysis and Extraction
Crown ethers serve as effective phase-transfer catalysts by forming lipophilic complexes with metal cations, which solubilize associated anions in organic phases and enhance their reactivity in two-phase systems.[42] In these processes, the crown ether encapsulates the cation, shielding it from solvation and allowing the "naked" anion to participate in reactions that would otherwise be hindered by poor solubility in nonpolar solvents.[43] A classic example is the use of 18-crown-6 to facilitate reactions of potassium cyanide (KCN) with organic substrates, often accelerating the reaction rate by orders of magnitude compared to uncatalyzed conditions.[43] This mechanism relies on the dynamic equilibrium of complex formation and ion transport across the interface, with catalytic efficiency depending on the crown's lipophilicity and cation-binding affinity.[44]In extraction applications, crown ethers enable the selective partitioning of metal ions from aqueous to organic solvents by forming neutral, stable complexes that enhance solubility in the receiving phase.[45] The lipophilicity of the crown-metal complex drives this transport, with extraction efficiency quantified by distribution coefficients (D), which measure the ratio of metal concentration in the organic versus aqueous phase.[46] For instance, dicyclohexano-18-crown-6 (DC18C6) is widely employed in nuclear waste processing to selectively extract strontium (Sr²⁺) from acidic aqueous streams into organic diluents like kerosene or octanol, due to its strong affinity for divalent cations and resistance to radiolysis under harsh conditions. This application supports the remediation of high-level radioactive wastes by isolating fission products like ⁹⁰Sr, minimizing secondary contamination.[47]Recent industrial advancements leverage crown ethers for lithium recovery from brines, addressing the growing demand for lithium in battery production.[48] Functionalized crowns, such as dibenzo-14-crown-4 derivatives, are integrated into membranes or sorbents to selectively bind Li⁺ over competing ions like Na⁺ and K⁺, with extraction efficiencies exceeding 90% in pilot-scale tests on salt lake brines since the early 2020s.[49] These methods, often combined with supercritical fluid extraction or nanofiltration, offer environmentally friendly alternatives to traditional evaporation ponds, reducing water usage and processing time.[50] The selectivity stems from the crown's cavity size matching Li⁺'s ionic radius, enabling high separation factors (up to 1000 for Li⁺/Na⁺) in complex aqueous matrices.[51]
Pharmaceutical and Biological Applications
Crown ethers play a significant role in pharmaceutical applications by enhancing the solubility of poorly water-soluble drugs through complexation with their macrocyclic cavities, thereby improving bioavailability. For instance, 18-crown-6 has been shown to increase the aqueous solubility of riboflavin, an ocular drug, by 42% at concentrations of 30 mg/mL, facilitating better formulation for topical delivery.[6] This complexation mechanism also extends to antifungal agents, where dicyclohexano-18-crown-6 forms stable 1:1 complexes with sulfanilamide, a sulfa-based antifungaldrug, primarily stabilized by hydrogen bonding, which can enhance drug stability and release profiles in formulations.[52]In terms of permeability enhancement, crown ethers like 12-crown-4, 15-crown-5, and 18-crown-6 improve drug transport across biological barriers by sequestering divalent cations such as Ca²⁺, which loosens tight junctions in corneal tissue. In vitro studies using excised bovine corneas demonstrated that 12-crown-4 at 30 mg/mL increased riboflavin permeation by up to 7.75-fold compared to controls, reaching steady-state fluxes of 62 ng/min·cm², although in vivo rabbit models showed more modest improvements due to ocular clearance.[6] These properties position crown ethers as excipients in ocular and potentially transdermal formulations to boost drugabsorption without significant irritation at low concentrations.[53]Certain derivatives, such as thia-crown ethers (sulfur-containing analogs), exhibit direct antibacterial activity by disrupting bacterial cell membranes through ionophoric action, targeting Gram-positive and Gram-negative pathogens like Staphylococcus aureus and Pseudomonas aeruginosa. However, their clinical translation is limited by high cytotoxicity to mammalian cells, often arising from non-selective ion transport that perturbs cellular homeostasis, with IC₅₀ values as low as 0.5 μM in cancer cell lines for some adamantyl-substituted variants.[54]Crown ethers mimic natural ion carriers like valinomycin, a potassium-selective ionophore, by facilitating selective transmembrane transport of alkali metal cations, which can be harnessed for targeted drug delivery systems. Adamantyl-substituted diaza-crown ethers, for instance, act as ionophores to reverse P-glycoprotein-mediated multidrug resistance in cancer cells by inhibiting drug efflux, thereby enhancing intracellular accumulation of chemotherapeutics like paclitaxel and inducing apoptosis at concentrations comparable to verapamil (IC₅₀ ~5-10 μM).[54] This ion transport capability supports applications in liposomal or nanoparticle-based carriers, where crown ethers enable controlled release across lipid bilayers in response to physiological ion gradients.[53]Recent developments in the 2020s have focused on optimizing crown ethers for enhanced drugpermeation in biological models, including skin and ocular tissues, while addressing cytotoxicity challenges. A 2025 review highlighted crown ethers' role as permeation enhancers in topical ocular delivery, though concentrations above 10 mg/mL risk mild epithelial toxicity.[55] Similarly, silacrown ether analogs, introduced in 2024 studies, exhibit reduced cytotoxicity compared to traditional crowns while maintaining iontransport efficacy in cardiomyocyte models, suggesting potential for safer transdermal applications; however, ongoing challenges include optimizing selectivity to avoid off-target effects in vivo.[56]
Sensors and Supramolecular Devices
Crown ethers have been integral to the development of ion-selective sensors, particularly fluorescent and electrochemical probes that enable precise detection of alkali metal ions in complex environments. For instance, BODIPY-equipped benzo-crown-6 ethers serve as fluorescent sensors for potassium (K⁺) ions, exhibiting enhanced fluorescence upon binding with sensitivity in the 1-10 mM range suitable for monitoring K⁺ levels in biological fluids like blood.[40] Similarly, electrochemical sensors incorporating 18-crown-6 derivatives, such as 4-aminobenzo-18-crown-6 self-assembled monolayers, demonstrate high selectivity for K⁺ over other cations like Na⁺ and Ca²⁺, with Nernstian responses in the micromolar range for clinical diagnostics.[57] These probes leverage the size-complementary binding cavities of crown ethers to achieve specificity, often integrated with nanomaterials like graphene oxide for amplified sensitivity.[58]In supramolecular chemistry, crown ethers form the basis of advanced assemblies such as rotaxanes and catenanes, which facilitate controlled cargo release in molecular devices. A notable example is a [59]rotaxane carrier where a crown ether ring threads an axle, enabling autonomous cargo release driven by a piperidine-based chemical oscillator; this system periodically cleaves the rotaxane structure, releasing the crown ether-bound cargo in a self-sustained manner, as demonstrated in 2025 studies.[60]Catenane architectures incorporating crown ethers similarly exhibit interlocked dynamics for ion-responsive shuttling, contributing to switchable supramolecular machines.[61] For detection applications, crown ether conjugation to silicon photonic platforms has enabled highly selective sensing of toxic ions; a 2024 integration via Fischer esterification on silicon waveguides achieved ppb-level detection of Pb²⁺ (1–262,000 ppb range) with selectivity over competing ions, adaptable for monitoring environmental toxins.[62]Recent innovations highlight the versatility of crown ether derivatives in sensing and assembly. Aza-crown ethers, with nitrogen-substituted rings, have been theoretically and experimentally validated for fluorescent biosensors targeting alkali metal ions, enabling selective imaging in biological systems through enhanced chelation and emission changes, as explored in 2025 computational studies.[63]Hybrid multimacrocycles combining crown ethers with cycloparaphenylenes offer tunable host–guest binding for diverse guests, showing promise in biological evaluations for targeted delivery and sensing due to their improved photophysical properties and selectivity.[64] As of November 2025, crown ethers have also seen emerging applications in sustainable energy, such as ion-selective membranes for lithium-sulfur batteries and CO₂ capture sorbents, enhancing efficiency in energy storage and environmental remediation.[65]Crown ethers also underpin practical devices like selective membranes and adsorbents. Crown-ether-embedded crystal channel membranes mimic biological ion channels, providing subnanometer pores for Na⁺ transport with high selectivity (e.g., Na⁺/K⁺ ratio >100) and flux rates suitable for desalination.[65] In lithium recovery, redox-active crown ether copolymers and magnetic core–shell adsorbents enable reversible, selective Li⁺ extraction from brines, achieving uptake capacities up to 50 mg/g with minimal interference from Na⁺ or Mg²⁺, as reported in 2025 advancements.[66][67]