Skip to main content
Nature Communications logoLink to Nature Communications
. 2025 Sep 29;16:8620. doi: 10.1038/s41467-025-63630-9

Electronically tailored metal-ion-chelation strategy promotes ionic liquid catalysis at near-ambient condition

Tianhao Zhang 1,2,#, Yuan Tian 1,2,#, Chong Zhang 1, Tingting Yan 1, Hanwen Yan 1,2, Guoliang Zhang 1, Jie Li 1,2, Zengxi Li 1,2, Gang Wang 1,2,3,, Chunshan Li 1,2,, Suojiang Zhang 1,3,4,
PMCID: PMC12480945  PMID: 41022719

Abstract

Electronic properties of active sites profoundly influence catalytic production of value-added chemicals; however, rational description and modulation is still a significant challenge. Herein, we propose an effective metal-ion-chelation strategy guided by density functional theory prediction and in situ Raman observation to structurally tailor and quantitatively correlate the electronic properties of active sites in ionic liquid. Comprehensive characterizations and theoretical calculations, in combination with electronic properties–performance correlations, reveal the electronic peculiarity of nitrogen and oxygen centers can be controllably restructured for remarkable improvement of catalytic performance at near-ambient condition. The turnover frequency is increased by two folds with deactivation rate suppressed by more than one half, while the kilogram-scaled recycling pilot achieves similar performance for the probe methacrolein synthesis. This strategy further exhibits excellent applicability and tolerance in other substrates with representative functional groups. Our work expresses the significance of electronic properties and provides a valid regulation approach for ionic liquid catalysis.

Subject terms: Chemical engineering, Ionic liquids, Homogeneous catalysis


Accurately describing and tuning active sites remains a major challenge. Here, the authors introduce an effective metal-ion chelation strategy—guided by DFT predictions and in situ Raman measurements—to structurally engineer and quantitatively link the electronic properties of active sites in an ionic liquid.

Introduction

The development of well-designed functional catalysts significantly facilitates the sustainable conversion of petroleum and coal-based downstream products into value-added chemicals13, while the electronic properties of active sites directly affect the catalytic performance4. In contrast to heterogeneous catalytic materials, the description and controllable regulation of electronic properties of homogeneous molecular catalysts is more aligned with intrinsic principles, yet it poses considerable challenges in structural design5,6. Ionic liquids (ILs) as a representative of green medium and catalyst have been extensively utilized in various environmentally friendly and mild catalytic processes, due to the flexible design of their cation and anion structures713. During the transformation in IL catalysis system, the efficient activation of reactants through the nucleophilic or electrophilic interaction is dramatically influenced by the electronic properties of active centers (electron-rich or electron-deficient)1416. Unfortunately, the unbalanced electron density of catalytic sites often leads to a trade-off between the catalytic performance and stability, thereby limiting the efficiency of transformative process17,18. Therefore, the rational description and fine-tuning of the electronic properties of active sites are essential to promote IL catalysis, especially in the aldol reaction of aldehydes19.

In recent years, the synthesis of value-added α, β-unsaturated aldehydes via direct aldol condensation between aldehydes has received considerable attention2022. In particular, the synthesis of methacrolein (MAL) from formaldehyde and propionaldehyde is a crucial step in the production of methyl methacrylate (MMA), which is a significantly important chemical material in the fine chemical, agrochemical, and pharmaceutical industries23,24. Compared to the traditional sec-amine catalysis process at 100–210 °C and 4.5–6.0 MPa25, the ILs show enhanced catalytic performance under near-ambient condition (25–50 °C and 0.1 MPa)2629. Nevertheless, the protonation of sec-amine center in IL leads to the difficulty in dissociation of this catalytically active sites and dramatically reduces the electron density of the nitrogen atom, which decreases the catalytic activity. Additionally, the undesired cyclization generates oxazolidine-type byproducts, which results in catalyst deactivation (Supplementary Fig. 1). Therefore, an effective strategy is urgently required to motivate catalytic activity by increasing the electron density of the nitrogen atom and suppress catalyst deactivation by stabilizing hydroxyl groups as well.

The potential factors influencing the electronic properties of active site in IL are complicated, primarily driven by the ionic bond between the cation and anion, along with polarization, inductive effect, hydrogen bonding, and van der Waals forces30. Over the past decades, great efforts have been dedicated to tune the electronic properties. On one hand, series of multi-scale regulation methods have been developed, including nanoconfinement31,32, multicomponent mixing33, interface enhancement34,35, and electric field assistance36,37. On the other hand, progressive achievements in theoretical calculations3841 and multiple characterization techniques42,43 enable researchers to explore the electronic properties and regulatory mechanisms in microscopic level. These investigations illuminate the critical role of electronic properties in determination of catalytic performance. Nonetheless, the relationship between the structure and electronic properties, as well as their underlying correlations to catalytic performance has not been fully elucidated, due to the inherent complexity of intermolecular interaction and local environment44,45. Thus, it is still challenging to precisely tailor the electronic properties by structure regulation to promote catalytic performance.

Herein, we propose a novel metal-ion-chelation strategy to controllably tailor the electronic peculiarity of catalytically active site in IL with the guidance of theoretical calculations and in situ Raman observations, which significantly prompt catalytic activity and stability at near-ambient condition. The relationship between the electronic properties and performance derived from the descriptors identified the optimal catalyst among the synthesized metal-ion-chelated ILs (MILs). Furthermore, the comprehensive characterizations and density functional theory (DFT) calculations reveal the intramolecular induction and coordination effects resulting from metal-ion chelation restructure the nanoscale morphology and electronic peculiarity, which promotes the dissociation and stability of catalytically active sec-amine center. Consequentially, both catalytic activity and stability in probe aldol condensation can be significantly improved by structurally tailored electronic properties. This strategy also demonstrates remarkable application in other substrates with representative functional groups.

Results

Electronically tailored design of MIL

To modulate the electronic characteristic of the nitrogen atom in protic diethanolamine (DEA)-type IL utilizing the metal-ion-chelation strategy (Fig. 1a), we predicted the Hirshfeld charge of the central nitrogen atom, namely the catalytically active amine site, of various MILs with optimized structure using DFT method (Fig. 1b and Supplementary Tables 12). We first confirmed the negative effect of protonation on the Hirshfeld charge of the central nitrogen atom in protic DEA-type IL using [HDEA]Pc as an example. The [HDEA] cation features symmetrical linear ethanol groups with proton partially delocalized between the cation and anion, resulting in the formation of a N+–H···Pc group. As a result, the Hirshfeld charge on the central nitrogen atom is reduced from −0.1479 to −0.0772 after protonation of DEA. When the [HDEA]Pc IL is chelated with monovalent metal ions of Cs+, K+, Na+, and Li+ through the interaction with hydroxyl groups, a certain degree of bending in the linear structure of cation is caused. While minimal influence is exerted on the anion, thus only slight alteration is discovered on the Hirshfeld charge of the central nitrogen atom. As for the chelation with divalent metal ions of Ba2+ and Mg2+, the Hirshfeld charge of the central nitrogen atom rises to −0.0841 and −0.1031, respectively. Meanwhile, a more significant bending of the alkyl chain is noticed in cation part, suggesting the increasing electronegativity of divalent metal ions correlates the enhanced chelation with hydroxyl groups. Furthermore, a higher coordination number enables them to exhibit interaction with the carbonyl oxygen of anion. Conversely, when we examine the chelation of Mn2+ and Y3+, the Hirshfeld charge of the central nitrogen atom significantly decreases to −0.0789 and −0.0709, even lower than that of [HDEA]Pc IL. As a result, the MgIL with the highest Hirshfeld charge of the central nitrogen atom is identified as the optimal catalyst candidate for the aldol condensation of aldehydes.

Fig. 1. Rational design of MIL inspired by theoretical prediction and in situ observations.

Fig. 1

a Schematic illustration of electronically tailored MIL. X represents CH3CH2COO. b Theoretically calculated Hirshfeld charges of the central nitrogen atom in MIL models (M/IL = 1/1) along with their optimized structures. Atoms are represented in cyan, red, blue, and gray for C, O, N, and H, respectively, with the central metal ions labeled below the structures. c Relationships between the TOF as well as ln rd, and Hirshfeld charge of the central nitrogen and hydroxyl oxygen atoms. The error bars represent the standard deviation of potential inconsistencies in the Hirshfeld charge of two different hydroxyl oxygen atoms. d Relationships between the ΔEM–IL and both TOF and CMAL. CMAL is defined by the MAL concentration measured at the reaction time of 60 min. e Relationships between the ΔEM–IL and both ln rd and stability. The stability is defined as the percentage of remaining catalytically active amine species after 60 min of reaction.

We postulated that catalytic activity is closely related to the Hirshfeld charge of the central nitrogen atom, while catalytic stability depends on the charge of the hydroxyl oxygen atoms in the respective [HDEA]Pc IL and MIL. To confirm this hypothesis, the MIL catalyst series was first synthesized using the fundamental [HDEA]Pc IL. We then employed an in situ Raman spectroscopy system to quantitatively analyze the probe aldol reaction between the formaldehyde and propionaldehyde to produce MAL with the catalysis of [HDEA]Pc IL and MIL (Supplementary Figs. 214). Figure 1c indicates that the turnover frequency (TOF) value is positively correlated with the Hirshfeld charge of the central nitrogen atom in the corresponding MIL. Meanwhile, we found a linear relationship between the catalyst deactivation rate and Hirshfeld charge of the hydroxyl oxygen atoms in MIL. Therefore, it can be concluded that the Hirshfeld charges of the central nitrogen atom and hydroxyl oxygen atoms are strongly relevant to the catalytic activity and stability. To further investigate the effect of chelation by different metal-ions, the relationship between the interaction energy (ΔEM–IL) and catalytic performance of MIL was established (Supplementary Table 3). Specifically, the lower ΔEM–IL reflects the stronger interaction between the metal-ion and [HDEA]Pc IL46. Figure 1d reveals that both TOF and CMAL change in a volcano-shaped trend as the ΔEM–IL increasing progressively from −29.2 to −131.6 kcal·mol−1, with the peak value appearing at the position of Mg2+. The catalytic activity of YIL is even lower than that of pristine [HDEA]Pc IL, attributed to the strongest interaction and lowest Hirshfeld charge of the central nitrogen atom. Furthermore, the catalytic stability of MIL is observably improved in proportion with the decreased ΔEM–IL, namely the catalyst deactivation can be inhibited by modulation of their interactions (Fig. 1e). These observations elucidate that the suitable interaction between the metal-ion and [HDEA]Pc IL via chelation will contribute to the relatively high Hirshfeld charge of the central nitrogen atom and thus electron-rich state of catalytically active amine site through the intramolecular inductive and coordination effects. Besides, the interaction between the metal-ion and hydroxyl groups also hinders the intramolecular cyclization of IMI intermediate. Thereby, the catalytic activity and stability in MAL production from aldol reaction of formaldehyde and propionaldehyde can be promoted. Nevertheless, excessively strong interactions will lead to a decrease in the Hirshfeld charge of the central nitrogen atom, which significantly diminishes catalytic activity despite the catalyst deactivation can be further suppressed. Accordingly, the theoretically predicted MgIL indeed exhibits high efficiency and excellent stability on the probe MAL synthesis via aldol reaction.

Characterization and controllable modulation of MgIL

With the optimal MgIL in hand, which is structurally confirmed using NMR (Supplementary Fig. 15), we further investigated and modulated the underlying physicochemical properties based on the electronic characteristics affected by the metal-ion chelation through changing the Mg2+ content. The cryogenic transmission electron microscope (Cryo-TEM) was employed to directly observe the morphological transformation of [HDEA]Pc IL after chelation with Mg2+. As the Mg2+ content increases, the intrinsic network structure of [HDEA]Pc IL gradually changes into spheroids (MgIL-1.0), with the metastable worm-like structure (MgIL-0.5) as transition state (Fig. 2a and Supplementary Fig. 16). Meanwhile, the average cluster size gradually decreases from 287.5 ± 2.9 ([HDEA]Pc IL) to 140.3 ± 0.7 nm (MgIL-0.5), and then to 93.8 ± 0.7 nm (MgIL-1.0), as determined by DLS (Fig. 2b). These observations at the nanoscale level demonstrate that the chelation of [HDEA]Pc IL with Mg2+ facilitates the formation of more shaped and uniformed microstructural morphology of MgIL. To analyze the electronic characteristics of Mg2+, X-ray photoelectron spectroscopy (XPS) was employed to measure and compare the binding energies of MgIL specimens having different Mg2+ content. As presented in Fig. 2c, the Mg 1s spectra illustrate that the peak located at 1304.0–1303.8 eV is consistent with the Mg2+ species reported in literature47. Moreover, this peak shifts to lower energy compared to that of Mg(Ac)2 (1304.4 eV) as the Mg2+ content increases, indicating that the enhanced chelation effect will give rise to the electron-rich state of Mg2+. Also, the MgIL-1.0 displayed lower white line intensity in X-ray absorption near-edge structure (XANES) spectrum, compared with that of Mg(Ac)2, which further reveals the electron transfer from [HDEA]Pc IL to Mg2+ (Fig. 2d)48,49. We also conducted 13C NMR experiments to test the chelation effect of Mg2+ on the hydroxyl groups of [HDEA], as the chemical shift of α-C atom reflects the changes of chemical environment of the adjacent oxygen atom. As displayed in Fig. 2e, we found an upfield chemical shift for [HDEA]Pc IL compared to DEA, while a series of successive downfield shifts for MgIL with increasing Mg2+ content. When the DEA is protonated by propionic acid (HPc), the decreased inductive effect from the central nitrogen atom increases the electron density of the α-C atom, which contributes to the electronic shielding and decrease in resonating frequency. In contrast, the increased resonating frequency indicates the deshielding effect on the carbon nucleus, explained by the electron donation from the hydroxyl oxygen atoms to Mg2+ during chelation, which reduces the electron density of α-C atom. These results demonstrate that the chelation effect will be enhanced with the increasing Mg2+ content, which is in alignment with the results obtained from the XPS Mg 1s spectra. In addition, the 15N NMR characterization was performed to catch insight into the effect of Mg2+ chelation on the electron density of the central nitrogen atom. It has been known that the isotropic nuclear magnetic deshielding of nitrogen atom will constantly shift downfield when it participates in the formation of an ionic bond50,51. As illustration in Fig. 2f, we first notice a downfield chemical shift after the protonation of DEA by HPc, which is associated with the reduced electron density of the central nitrogen atom by ionization. However, subsequent upfield chemical shifts are found after the chelation of [HDEA]Pc IL with Mg2+, implying the increased electron-rich state of the central nitrogen atom with higher Mg2+ content in MgIL. In addition, these findings are further supported by the observed binding energy shift in XPS N 1s spectra of [HDEA]Pc IL after chelation with different content of Mg2+ (Supplementary Fig. 17). Furthermore, we synthesized a series of MgILs with different anions, including Cl, NO3, SO42− and CF3SO3 to investigate the role of Ac introduced by Mg(Ac)2. Supplementary Fig. 18 demonstrates the largest upfield 15N chemical shift when Ac introduced by Mg salts, suggesting the most effective to facilitate the electron-rich state of the nitrogen atom. Additionally, the combination of [HDEA]Ac and Mg(Ac)2 shows a nearly identical 15N chemical shift to that of [HDEA]Pc and Mg(Ac)2 system. These results suggest that the effect of Ac and Pc on the electronic properties of the nitrogen atom is almost similar when acting as anions in these designed MILs, but Ac shows more beneficial than Cl, NO3, SO42− or CF3SO3.

Fig. 2. Morphological and electronic characterizations of MgIL.

Fig. 2

a Cryo-TEM images of [HDEA]Pc IL, MgIL-0.5 and MgIL-1.0. b Cluster size distribution of [HDEA]Pc IL, MgIL-0.5, and MgIL-1.0. The error bars for the mean values represent the standard deviations calculated from three independent DLS measurements. c XPS Mg 1s spectra of Mg(Ac)2 and MgIL-X (X = 0.25, 0.5, 0.75, 1.0). The samples are prepared using pristine [HDEA]Pc IL without water to avoid volatile components during XPS measurements. d XANES spectra of Mg(Ac)2 and MgIL-1.0. e 13C NMR spectra of DEA, [HDEA]Pc IL, and MgIL-X (X = 0.25, 0.5, 0.75, 1.0). f 15N NMR spectra of DEA, [HDEA]Pc IL, and MgIL-X (X = 0.25, 0.5, 0.75, 1.0).

Aside from the intramolecular inductive and coordination effects derived from the chelation with Mg2+ that will trigger the increased electron density of the central nitrogen atom (namely amine site), the electrostatic interaction between the cation and anion and their local environment also plays an important role in electronic properties5254. We first performed the ionic conductivity measurements and isothermal titration calorimetry (ITC) analysis to explore the bindings between the Mg(Ac)2 and [HDEA]Pc IL. It is held that the conductivity activation energy (Ea) measures the activation barrier of ion migration, while the ionic conductivity (σ) depends on both ion migration and concentration55. As illustrated in Fig. 3a, both saturated aqueous Mg(Ac)2 and [HDEA]Pc IL solutions exhibit relatively high σ but low Ek of 7.1 and 10.0 kJ·mol−1, respectively, due to their intrinsically high ionization and ion migration capabilities. Oppositely, significantly lower σ and higher Ea are observed in the case of MgIL. Specifically, the σ declines substantially while Ek increases from 15.0 to 27.1 kJ·mol−1 when the molar ratio of Mg2+ to [HDEA]Pc IL ranges from 0.25 to 1.0. It suggests that the bindings between the Mg(Ac)2 and [HDEA]Pc IL are continuously strengthened with the rising Mg2+ content, which effectively restricts the migration of free ions. In other words, the chelation of [HDEA]Pc IL with Mg(Ac)2 indeed occurs in their mixtures. The ITC investigations further reveal that the interaction between the Mg2+ and [HDEA]Pc IL is associated with a moderate endothermic process, involving approximate three binding sites (Fig. 3b). These findings signify that these three binding sites include two hydroxyl groups of [HDEA] cation and the other carbonyl group of Pc anion, which thereby support the validity of our proposed chelation structure for DFT calculation.

Fig. 3. Effect of Mg2+ chelation on the local environment of MgIL.

Fig. 3

a Ionic conductivity test of saturated aqueous Mg(Ac)2 solution, [HDEA]Pc IL, and MgIL-X (X = 0.25, 0.5, 0.75, 1.0) samples. b ITC investigations including the thermograms (upper plots) and binding isotherms (lower plots) for the chelation interaction between the Mg2+ and [HDEA]Pc IL. The binding isotherms are calculated using aqueous Mg(Ac)2 solution (20 mmol∙L−1) titrated into water as the blank control and fitted with a first-order binding model. c FT-IR spectra of Mg(Ac)2, [HDEA]Pc IL, and MgIL-X (X = 0.25, 0.5, 0.75, 1.0) samples. d 1H NMR spectra of DEA, [HDEA]Pc IL, and MgIL-X (X = 0.25, 0.5, 0.75, 1.0). e ATR-FIR spectra of [HDEA]Pc IL and MgIL-X (X = 0.25, 0.5, 0.75, 1.0) samples.

To further understand and modulate the impact of Mg2+ chelation on the interaction between the cation and anion, we employed Fourier transform infrared (FT-IR), ¹H NMR, and far-infrared (FIR) spectroscopies to monitor the evolution of molecular-level interactions as a function of Mg2+ concentration. The FT-IR spectroscopic measurements reveal a gradual blue shift from 1562 ([HDEA]Pc IL) to 1572 cm−1 (MgIL-1.0) for the antisymmetric stretching vibration frequency of carbonyl group attributed to the Pc anion, which is distinguished from that in Ac at 1591 cm−1 (Fig. 3c). It demonstrates the successively enhanced coordination effect of Mg2+ on the Pc anion with increasing Mg2+ content in MgIL. While the 1H NMR spectra provide insights into the changes in hydrogen-bonding strength between the cation and anion (Fig. 3d). The downfield 1H chemical shifts of these two pairs of methylene hydrogens in [HDEA]Pc IL suggest that the [HDEA] shows intense hydrogen-bonding interaction with Pc after protonation of DEA by HPc56. However, we subsequently detected a series of upfield 1H chemical shifts for the MgIL samples having different Mg2+ content, indicating that the hydrogen-bonding interaction between the cation and anion will be weakened as the Mg2+ content is raised. According to the reported literature, it is believed that the stronger (or weaker) interaction between the DEA and HPc through protonation will lead to greater shortening of the intermolecular N+–H bond, which is reflected by a blue (or red) shift in FIR stretching57,58. As the content of Mg2+ in MgIL increases, the observed red shift of stretching vibration band identifies that the chelation effect of Mg2+ weakens the N+–H bond, which is benefit for the dissociation of sec-amine site for aldehyde activation (Fig. 3e). Thus, the chelation of Mg2+ also plays a pivotal role in modulating the local environment of cation and anion, ultimately in favor of fine-tuning the electronic properties of MgIL.

Catalytic performance and mechanism

Building on the impressive catalytic efficiency of MgIL, we embarked on a detailed investigation into the effect of Mg2+ content on the performance of MgIL. In situ Raman spectroscopy served as a powerful tool to monitor the dynamic evolution of catalytically active species consumption and formation of MAL from the probe aldol reaction of formaldehyde with propionaldehyde (Fig. 4a, b, and Supplementary Figs. 1922). As the content of Mg2+ incrementally increased from 0 to 3.8 wt.% (corresponding to the molar ratio of Mg2+ to [HDEA]Pc IL from 0 to 1), we can observe a notable enhancement of TOF from 0.67 to 1.36 h−1, while the CMAL rises from 0.25 to 0.54 mmol·g−1. These are accompanied by a dramatic reduction of rd to approximately one-tenth, alongside a 33% enhancement in stability when contrasted with [HDEA]Pc IL (Fig. 4c). Our comparative analysis reveals a striking correlation between the higher Mg2+ content and improved catalytic performance, with MgIL-1.0 (featuring equimolar Mg2+ and [HDEA]Pc IL) emerging as the champion in the view of catalytic activity and stability (Fig. 4d). Additionally, we investigated the situation that Mg2+/[HDEA]Pc IL ratio exceeds 1.0, for instance 1.25, the catalyst becomes unstable, resulting in the precipitation of a white solid. Noting that our attempts to deploy an aqueous solution of Mg(Ac)2 as a catalyst yield no detectable MAL (Fig. 4b). It suggests that Mg(Ac)2 has no catalytic activity in this reaction independently, despite its Lewis acidity, highlighting the indispensable role of sec-amine center as the true catalytically active site for this transformative process. Additional characterizations of these as-synthesized catalysts were conducted using 15N and 13C NMR spectroscopy, with difference in 15N and 13C chemical shifts (Δδ) recorded against [HDEA]Pc IL as a reference. As depicted in Fig. 4e, the incremental rise in the molar ratio of Mg2+ to [HDEA]Pc IL leads to a proportional decrease of Δδ15N from 0 to −0.90 ppm, while the TOF exhibits a corresponding upward trend, spanning 0.67 to 1.36 h−1. Whereas the Δδ13C climbs proportionally from 0 to 0.43 ppm and ln rd exhibits a linear decline (Fig. 4f). These observations consolidate our conclusion that the Mg2+ chelation effect favors the electron-rich state of catalytically amine site, however, electron-deficient state of hydroxyl oxygen, thereby significantly improving catalytic activity and stability.

Fig. 4. Catalytic performance and mechanistic insights.

Fig. 4

a, b Concentration profiles of (a) catalytically active amine species and (b) MAL catalyzed by Mg(Ac)2, [HDEA]Pc IL, and MgIL-X (X = 0.25, 0.5, 0.75, 1.0) as a function of reaction time. c Effect of Mg2+ content in MgIL on the TOF, CMAL, rd and stability. d Comparison of catalytic performance among [HDEA]Pc IL and MgIL-X (X = 0.25, 0.5, 0.75, 1.0). e The relationship between the TOF and 15N chemical shift difference (Δδ15N = δ15NMgIL − δ15N[HDEA]Pc IL); 15N chemical shifts are provided in Fig. 2f. f The relationship between the ln rd and 13C chemical shift difference (Δδ13C = δ13CMgIL − δ13C[HDEA]Pc IL); 13C chemical shifts are illustrated in Fig. 2e. g DFT calculations of the sec-amine site dissociation, formaldehyde (FA) activation, and MB+ enolization steps during the probe aldol condensation of formaldehyde with propionaldehyde catalyzed by [HDEA]Pc IL and MgIL. The embedded structures depict their respective transition states. The atoms in cyan, red, blue, yellow, and gray represent C, O, N, Mg, and H atom, respectively. h Distance of (H)O···C(N+) and corresponding energy for hydroxyl group rotation in IMI intermediate at various angles under the situation of hydrogen-bonding interaction (IMI-Pc) and Mg2+ chelation effect (IMI-Mg). i The comparable relative energies required for rotation of hydroxyl group in IMI intermediate during intramolecular cyclization with hydrogen-bonding (IMI-Pc) and chelation (IMI-Mg) interactions. The blue, green, and red segments in the pie chart respectively represent hydrogen-bonding interaction, van der Waals force, and repulsive interaction. The circular arrows indicate the rotation process. j Kinetic observation of generation of the deactivated byproduct 3-oxazolidineethanol.

DFT calculations subsequently unveil that the chelation of Mg2+ promotes the dissociation of catalytically active sec-amine center compared with pristine [HDEA]Pc IL, which is a crucial step for the following formaldehyde activation (Fig. 4g). It was also observed in other MILs containing various metal-ions that the enhanced dissociation of catalytically active sec-amine center is correlated to the increased Hirshfeld charge on the central nitrogen atom (Supplementary Fig. 23, and Supplementary Table 4). Further analysis of the Gibbs free energy barriers (ΔG) of transition states involved in the critical steps of formaldehyde activation and MB+ enolization reveals that the determining barriers for [HDEA]Pc IL (16.0 kcal·mol−1) and MgIL (15.3 kcal·mol−1) are comparable (Fig. 4g, Supplementary Figs. 2425, and Supplementary Tables 56). Thus, it can be concluded that the enhanced dissociation of sec-amine site predominantly contributes to the improved catalytic activity. Additionally, we also discovered that the hydroxyl group readily combines with the carbocation of IMI intermediate at a rotation angle of 150° to form undesired 3-oxazolidineethanol, primarily leading to catalyst deactivation26, with an energy requirement of only 10.5 kcal·mol−1 when the hydroxyl group is constrained by hydrogen bond with Pc anion (Fig. 4h, i, and Supplementary Table 7). By comparison, the relative energy dramatically rises to 68.1 kcal·mol−1 when the hydroxyl groups are restrained by the Mg2+ chelation. So, we can conclude that the metal-ion chelation effect is significantly stronger than hydrogen-bonding interaction and thereby effectively hinders the undesired cyclization reaction, which is consistent with the in situ kinetic observations of 3-oxazolidineethanol generation (Fig. 4j) and explains the higher stability of MgIL.

Catalytic scale-up and extension

Given the excellent catalytic performance of MgIL-1.0 after controllable modulation, we first conducted the kilogram-scaled aldol condensation for probe MAL synthesis to demonstrate the application of this catalytic system in large-scale production. The scalable preparation of MgIL-1.0 can be easily achieved due to the convenient dissolution and chelation at ambient temperature and pressure in less than an hour, limited only by the size of reactor vessel (Fig. 5a). As illustrated in Fig. 5b and Supplementary Figs. 2628, the MgIL-1.0 catalyst can be recycled for five runs consecutively without a loss of activity, which achieves an average conversion of 98.6% for propionaldehyde and 98.8% selectivity for MAL. Furthermore, we employed FT-IR and 1H NMR to confirm the catalytically active amine component in catalytic system after multiple cycles of reaction, phase separation, evaporation, and recovery. Figure 5c demonstrates that the characteristic peaks of C = O and C–N bonds in the anion and cation of MgIL-1.0 stay consistent after recycling. The 13C NMR spectra indicate that the 13C signals of the catalytically active component are still the most prominent after five cycles, which is in accordance with those of fresh catalyst (Fig. 5d). Additionally, only a few insignificant peaks attributed to the oxazolidine species formed through intramolecular cyclization (catalyst deactivation) were observed, which are in agreement with the characteristic chemical shifts reported in literatures59,60. Aside from the qualitative analysis, the quantitative results reveal that the contents of catalytically active amine component and Mg2+ in MgIL-1.0 remain virtually unchanged, with a high recovery ratio of around 97% (Fig. 5e).

Fig. 5. Reusability and applicability assessment of MgIL-1.0 catalyst.

Fig. 5

a Photograph of catalyst preparation (embedded), kilogram-scaled synthesis, and phase separation systems. b Catalytic performance of MgIL-1.0 during recycling of the probe aldol condensation between the formaldehyde and propionaldehyde. c FT-IR spectra of the fresh and recycled MgIL-1.0 catalyst. d 13C NMR spectra of fresh and recycled MgIL-1.0 catalyst. R represents H–, CH3–, and CH2 = C(CH3)– groups, corresponding to the oxazolidines derived from formaldehyde, propionaldehyde, and MAL, respectively. e Comparison of Mg and N contents in the theoretical, fresh, and recycled MgIL-1.0 catalyst. The error bars for the mean values represent the standard deviations calculated from three independent measurements. f Extensive applications of MgIL-1.0 in the synthesis of representative α, β-unsaturated aldehydes via aldol reaction. The products were identified through GC-MS analysis (Supplementary Figs. 2936), while the stability was evaluated using 1H NMR spectroscopy (Supplementary Figs. 3747).

To further characterize the performance of this MgIL-1.0 catalyst in aldol condensation, we then investigated its applicability with substrates containing various electron-donating and withdrawing groups, such as alkyl, hydroxyl, furan, halogenated and benzoate groups (Fig. 5f). By increasing the alkyl chain length from 1 to 4 carbon atoms and introducing branching, the catalytic performance of MgIL-1.0 still shows a notable enhancement compared to pristine [HDEA]Pc IL, particularly for the catalytic stability improved by 1.4–3.1 times. It also demonstrates a significant improvement in overall performance on high-value fine chemicals applied as commercial fragrances61,62, such as hydroxycitronellal and 3-(5-methyl-2-furyl)propionaldehyde. As for the aldehydes having electron-withdrawing groups like chlorine and phenyl substitution, the stability of MgIL-1.0 is improved by 1.2 to 1.7 times, with selectivity and TOF values increasing by 1.2 to 1.7 and 1.3 to 1.5 times, respectively. Notably, all of these above aldol reactions achieved over 90% selectivity for α, β-unsaturated aldehyde products, a level typically unattainable with conventional IL catalysts. The high catalytic selectivity achieved along with the enhanced TOF values and stability under MgIL-1.0 catalysis exclude the impacts of substituted group and steric hindrance, which suggests the excellent compatibility and overall performance of MgIL on the construction of C = C bonds via aldol reaction.

Discussion

In summary, we theoretically predicted and experimentally synthesized a series of MIL catalysts with tailored electronic characteristics for the efficient and stable synthesis of α, β-unsaturated aldehydes via aldol condensation at near-ambient condition. The electronic modulation mechanism is based on intramolecular inductive and coordination effects, along with the local environment of cation and anion resulting from metal-ion chelation. DFT calculations and in situ observations reveal the catalytic activity and stability of MIL are linearly correlated to the Hirshfeld charge of the central nitrogen (namely catalytically amine site) and hydroxyl oxygen atoms. While the TOF value of MIL shows a volcano-shaped trend with the increasing interaction between the metal-ion and [HDEA]Pc IL, which can be considered as a descriptor for MIL design. Comprehensive characterizations indicate that the electronic properties and local environment of the MgIL can be further controllably regulated by the content of Mg2+, which consequently reconfigures the nanoscale morphology and promotes the dissociation and stability of the catalytically active sec-amine center. As a result, the TOF value can be increased by 2.0-fold with catalytic deactivation rate suppressed by 89.6%, while the kilogram-scaled recycling pilot achieves 98.6% conversion and 98.8% selectivity for probe MAL synthesis. It also demonstrates excellent applicability and tolerance across various substrates with representative functional groups. In light of the impressive catalytic activity, stability, scale-up, and broad extension, this strategy represents a promising approach for sustainable transformation of low-carbon aldehydes into value-added chemicals by constructing new C = C bonds. Our findings also unveil the potential importance of electronic modulation during IL design and provide an avenue for the development of efficient and stable catalytic processes at near-ambient condition.

Methods

Chemicals

The chemical reagents and materials were purchased from commercial suppliers and utilized without additional purification. Diethanolamine (purity, ≥ 99%), propionic acid (purity, ≥ 99%), cesium acetate (purity, ≥ 99.9%), lithium acetate (purity, ≥ 99.99%), sodium acetate (purity, ≥ 99.99%), potassium acetate (purity, ≥ 99%), barium acetate (purity, ≥ 99.99%), magnesium acetate (purity, ≥ 98%), manganese acetate (purity, ≥ 98%), zinc acetate (purity, ≥ 99.99%), yttrium acetate hydrate (purity, ≥ 99.9%), formalin (concentration, 37 wt.% aqueous solution), propionaldehyde (purity, 97%), methacrolein (purity, ≥ 95%), butyraldehyde (purity, 99.5%), valeraldehyde (purity, ≥ 97%), iso-valeraldehyde (purity, 98%), hydroxycitronellal (purity, ≥ 98%), hydrocinnamaldehyde (purity, ≥ 95%), 3-(5-methyl-2-furyl)propionaldehyde (purity, ≥ 98%), ethanol (purity, ≥ 99.5%), 1,4-dioxane (purity, ≥ 99.0%), 4-hydroxy-2,2,6,6-tetramethyl-piperidinooxy (purity, ≥ 98%), 1, 2, 4, 5-tetrachlorobenzene (purity, ≥ 98.0%), deuterated dimethyl sulfoxide (DMSO-d6, purity, d-99.9%) and deuterated chloroform (CDCl3, purity, d-99.8%) were purchased from Aladdin Chemical Co., Ltd (Shanghai, China). Acetaldehyde (purity, 99.5%) and 5-chloropentanal (purity, 98%) were obtained from Macklin Biochemical Co., Ltd. (Shanghai, China). Deionized water was made in the laboratory.

Catalyst preparation

For the synthesis of [HDEA]Pc IL, equimolar amounts of DEA and HPc were mixed in a flask with magnetic stirring under ice bath. Typically, 42.5 g (0.4 mol) of DEA was combined with 127.7 g of water in a 500 mL three-neck flask. Under an argon atmosphere, 29.8 g (0.4 mol) of HPc was slowly added through a pressure-equalizing dropping funnel. The mixture was stirred for 1 h, resulting in a clear and colorless IL. While the MIL was synthesized by gradually adding the required amount of metal acetate to the [HDEA]Pc IL under ultrasonic agitation and vigorous stirring, followed by keeping at room temperature for 30 min. These obtained compounds are designated as MIL-X, where M represents the type of metal-ion, and X indicates the molar ratio of corresponding metal acetate to [HDEA]Pc IL. For the catalysts presented in Fig. 1c–e, the molar ratio of metal ions to [HDEA]Pc IL is 1:2. For the synthesis of MgIL-1.0 sample, 20 g (40 mmol) of the synthesized [HDEA]Pc IL was charged in a 100 mL three-neck flask. Then, 5.8 g (40 mmol) of magnesium acetate was gradually added under ultrasonic agitation and vigorous stirring. The reaction was continued for 30 min to yield a clear and transparent catalyst sample.

Catalyst characterizations

X-ray photoelectron spectroscopy (XPS) was performed on a ThermoFisher ESCALAB 250 Xi spectrometer utilizing Al- radiation, while the analysis was conducted at a power of 200 W. All binding energies are referenced to the C1s peak at 284.8 eV for calibration.

1H/13C/15N NMR spectra were obtained from a Bruker Avance ⅠⅠⅠ HD 600 MHz spectrometer at temperature of 25 °C, with CDCl3 or DMSO-d6 serving as the field frequency lock. The calibration of chemical shift was achieved by referencing the residual solvent peak or tetramethylsilane peak.

Isothermal titration calorimetry (ITC) measurements were carried out employing a Malvern MicroCal Auto-iTC200 calorimeter. A 20 mmol·L−1 aqueous magnesium acetate solution was injected into a reaction cell containing 1 mmol·L−1 [HDEA]Pc IL under the stirring at 800 r·min−1 and 25 °C. The water titration without [HDEA]Pc IL served as the blank control. Titrations employed an initial delay of 60 seconds, followed by a 0.5 µL injection for equilibration but not for fitting. Subsequent 2 µL injections, each lasting 30 s, were administered at 120 s intervals for a total of 20 injections.

Fourier transform infrared (FT-IR) spectroscopy was carried out using a ThermoFisher Nicolet-380 spectrometer, the spectra were collected in the wavenumber ranging from 400 to 4000 cm−1 with 32 scans at a resolution of 2 cm−1. Samples were uniformly applied to potassium bromide tablets for measurements.

X-ray absorption near-edge structure (XANES) spectra of all samples were recorded at the Singapore Synchrotron Light Source. The powdered MgIL-1.0 sample was purified from the aqueous solution using the freeze-drying method. A pair of channel-cut Si (111) crystals was utilized in the monochromator, while the storage ring operated at an energy of 800 MeV, with an average electron current maintained below 200 mA.

Far infrared (FIR) spectroscopic characterizations were conducted on a Thermo Scientific Nicolet IS50 spectrometer equipped with a DLaTGS detector. The catalyst sample was first pretreated under ultrasonic for 10 min to ensure uniformity before introduction into the sample pool to create a liquid film with suitable thickness. Spectra were recorded in the range of 80–450 cm−1 with a resolution of 4 cm−1 and 128 scans.

Ionic conductivity tests were performed using a Mettler Toledo FE-38 conductivity meter with an LE-710 electrode. About 10 mL of catalyst was placed in a 15 mL jacketed kettle equipped with a thermostatic circulation pump. The electrode was positioned below the liquid surface, and conductivity was assessed once the IL attained a stable temperature. For each subsequent measurement, the temperature setting was adjusted accordingly. Ea is defined by the following formula of ln σ(T) = ln σ0 − Ea/RT55. σ(T) represents the conductivity at a specific temperature, σ0 is pre-exponential factor, Ea is conductive activation barrier, R is gas constant of 8.314 J·mol−1·K−1, and T is absolute temperature in Kelvin.

Cryogenic transmission electron microscopy (Cryo-TEM) experiments were conducted using a thin film of MIL solution sample (2 µL) placed onto a supported grid at room temperature (Lacey Formvar/Carbon, 200 mesh, Cu; Ted Pella, Inc.). The preparation of these thin solution films is proceeded under controlled temperature and humidity conditions (97–99%) within a specially designed environmental chamber. After allowing the excess liquid to be blotted with filter paper for 2–3 s, the films were quickly vitrified by plunging them into liquid ethane cooled by liquid nitrogen at freezing point. The grid was then transferred to a Gatan 626 Cryo-holder via a Cryo-transfer device and subsequently to the Thermo Fisher FEI Talos F200C TEM instrument (200 kV). Direct imaging was performed at approximately −175 °C with a 200 kV accelerating voltage, utilizing images captured by a SC 1000 CCD camera (Gatan, Inc., USA). Data analysis was carried out using Digital Micrograph software.

Dynamic light scattering (DLS) analysis was conducted using a ZetaSizer Nano ZS90 equipped with a four-sided optical path sample cell. The sample was sonicated for 10 min before being transferred to the cleaned sample cell. Each sample was tested in triplicate.

The nitrogen content of both fresh and recycled MgIL-1.0 catalysts was determined using an Elementar Vario MACRO cube, while the magnesium content was quantified with an Agilent 5800 ICP-OES.

Experimental setup of in situ Raman spectrometer system

We developed an in situ Raman spectrometer system for real-time monitoring of chemical reactions. The system consists of a laser, an optical path enhancement system, an in situ reaction cell, and a spectrometer (Supplementary Fig. 4). A continuous-wave laser provides up to 0.5 W of power, effectively exciting the samples and generating reliable Raman scattering signals. To improve the Raman signal intensity, we implemented an integrating sphere system based on confocal focusing and right-angle beam enhancement principles, allowing multiple reflections of the pump light within the sample and significantly increasing the efficiency of Raman scattering signal collection. The in situ reaction cell features a jacketed structure for precise temperature control using a circulating temperature-regulating medium. With a volume of 3 mL and sapphire windows for laser transmission and Raman signal acquisition. Data acquisition utilized an Andor IVAC 316 LDC-DD CCD detector paired with an Andor SR500i spectrometer, equipped with a 1200-line grating for a resolution of 1.5 cm−1. The Raman spectroscopic data were processed using Andor Solis software.

In situ Raman observation for aldol condensation

Aldol condensation catalyzed by IL and MIL was conducted on the in situ Raman spectrometer for real-time monitoring of the reaction. First, the Raman spectrometer was activated, and the CCD temperature was lowered to −60 °C. The kinetic acquisition mode was set to cover a wavelength range of 545–587 nm, with an exposure time of 0.2 s and 75 accumulations per acquisition and a total of 260 acquisitions. For example, to evaluate the catalytic performance of MgIL-1.0, 1.8 g of synthesized catalyst was accurately weighed and added to the in situ reaction cell, and the thermostatic circulation pump was activated to maintain a temperature of 0 °C inside the cell. Simultaneously, the pre-prepared aldehyde mixture was kept at a constant temperature in the circulation bath to minimize temperature fluctuations during the initial reaction phase. Following this, 0.8 g of the aldehyde mixture was measured using a syringe. The spectroscopy acquisition program was initiated, and the aldehyde mixture was rapidly injected into the in situ reaction cell, followed by multiple aspirations to ensure thorough mixing of the system.

The concentration of MAL and catalytically active amine species was determined through the quantitative analysis of the characteristic peak at 1693 cm−1 (C = O bond) and 1102 cm−1 (C–N bond), respectively. TOF value was calculated by the following formula: TOF = rPro./CCat., where rPro. represents the formation rate obtained from linear fitting of product concentration at the initial stage of reaction, CCat. represents the initial concentration of catalyst. CMAL is defined by the MAL concentration measured at the reaction time of 60 min. Stability is defined by the formula of Stability (%) = 1 − Q, where Q indicates the catalyst consumption percentage after 60 min of reaction. rd was calculated from the linear fitting of catalytically active amine species concentrations at the initial reaction stage.

Catalytic scale-up and recycling

A total of 540.0 g of MgIL-1.0 was introduced into a 1 L five-necked jacketed reactor using a feed pump, which was equipped with a condenser and an electronic thermometer inserted below the liquid level. Mechanical stirring was initiated at 400 rpm, and a thermostatic circulation bath was activated to maintain the temperature at 40 °C. The feed solution was prepared by thoroughly mixing 203.7 g of formaldehyde, 150.3 g of propionaldehyde, and 0.35 g of 4-hydroxy-2,2,6,6-tetramethyl-piperidinooxy. This mixture was then fed into the reactor by the pump at a drop rate of 17.7 g·min−1 after calibration with the feed solution (Supplementary Fig. 26). After reaction for 10 min, the mixture was transferred to a separatory funnel with a cooling jacket using another pump and then cooled to 10 °C. Once the liquid level stabilized, the aqueous and organic phases were separated by slowly turning on the valve (Supplementary Fig. 27). Both phases were analyzed by gas chromatography using 1, 4-dioxane as an internal standard (Supplementary Fig. 28). The organic phase was subsequently subjected to rotary evaporation to remove excess water before being recycled for the next run.

Catalytic extension assessments

Catalytic extension was performed in a 50 mL three-necked flask equipped with a stirrer. For the evaluation of catalytic selectivity and TOF, the catalyst was added into the reactant mixtures and the reaction was conducted at 25 °C for 3–5 min with a formaldehyde/substrate/catalyst molar ratio of 8:4:1. For the substrates of butyraldehyde, valeraldehyde, and iso-valeraldehyde, the molar ratio was maintained at 4:4:1. Then the mixture was diluted with specified amounts of ethanol and 1, 4-dioxane. The products were confirmed using gas chromatography–mass spectrometry (GC–MS) (QP2020, Shimadzu). Quantitative analysis was performed on an Agilent 7890 A GC equipped with a flame ionization detector and a DB-624 column (60 m × 0.32 mm × 1.8 μm). The relative mass correction factor was determined using a standard mixture, while the effective carbon number method was used for the products without a standard sample. To assess catalytic stability, the reactants and catalyst were mixed in an equimolar ratio of formaldehyde, substrate, and catalyst, and maintained at 25 °C for 1 h to achieve equilibrium. Upon completion, the reaction mixture was diluted with ethanol and then dissolved in DMSO-d6 for 1H NMR analysis.

DFT calculations

All calculations in this study were carried out using the Gaussian 16 software package63. The structural optimization was carried out using B3LYP hybrid density functional with 6-311 + G (d, p) basis set for C, H, O, N, Na, Li, and Mg atoms, and SDD pseudopotential basis set for K, Mn, Y, Cs, and Ba64,65. The dispersion correction was also considered. Vibrational analysis was performed for the optimized structures, and no imaginary frequency was found. In order to obtain more accurate electronic energy, the M06-2X functional, which is suitable for calculation of non-covalent interactions and reasonable description of dispersion effects, was employed along with the def2-TZVP basis set in the single-point energy calculation66,67. In the calculations involving the reaction, water was set as the implicit solvation model. The Integral Equation Formalism of Polarizable Continuum Model was used for structural optimization and vibrational analysis, while the Solvation Model Based on Density was used for single-point energy68. The transition state (TS) was confirmed to have only one imaginary frequency, and the intrinsic reaction coordination analysis was used to obtain the initial state (IS) and final state (FS) structures of the connected TS. The Multifwn software was applied to calculation of electronic properties based on wavefunctions69. The VMD program was used for graphic visualization70. The Gibbs free energy was calculated by Shermo software71. ΔG was calculated by the formula of ΔG =  G(TS)  −  G(IS). Here, G(TS) and G(IS) are the calculated Gibbs free energies of TS and IS in each elementary step, respectively. The temperature and pressure were set as 313.15 K and 1 atm, respectively. The interaction energy (ΔEM–IL) between the metal acetates and [HDEA]Pc IL was calculated using the formula of ΔEM–IL = EMIL − EM − EIL. Here, EMIL refers to the energy of MIL, EM represents the energy of metal acetate, and EIL indicates the energy of [HDEA]Pc IL. In the presence of metal acetate, the interaction energy between HPc and DEA (ΔEM(DEA∙∙∙HPc)) is defined by ΔEM(DEA∙∙∙HPc) = EMIL − EDEA–M − EHPc–M + EM72. Here, EMIL represents the energy of MIL, EDEA–M refers to the energy of the combined DEA and metal acetate, EHPc–M denotes the energy of the combined HPc and metal acetates, and EM is the energy of metal acetate.

Supplementary information

Acknowledgements

For this work, we gratefully acknowledge the financial support of National Natural Science Fund for Distinguished Young Scholars (22025803), Key R&D Program of Henan (241111230500), Natural Science Foundation of Henan (252300421191), Hundred Talents Program of Chinese Academy of Sciences (E3293501) and the Key Technology Team of the Chinese Academy of Sciences (Grant No. GJJSTD20220003).

Author contributions

T.H. Zhang designed the major experiments and analyzed the data. Y. Tian conducted the DFT calculations for the mechanistic investigation. C. Zhang synthesized the catalysts and evaluated their catalytic performance. T.T. Yan analyzed the XANES spectra. H.W. Yan, G.L. Zhang, J.Li and Z.X. Li provided insights into the characterization results. G. Wang, C.S. Li and S.J. Zhang contributed to the project administration, formal analysis and research guidance. The manuscript was written by T.H. Zhang, Y. Tian and G. Wang with input from all authors.

Peer review

Peer review information

Nature Communications thanks Sergio Rossi and the other anonymous reviewer(s) for their contribution to the peer review of this work. A peer review file is available.

Data availability

The source data generated in this study have been deposited in the Figshare database under accession code 10.6084/m9.figshare.29502947.

Competing interests

The authors declare no competing interests.

Footnotes

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

These authors contributed equally: Tianhao Zhang, Yuan Tian.

Contributor Information

Gang Wang, Email: wanggang@ipe.ac.cn.

Chunshan Li, Email: csli@ipe.ac.cn.

Suojiang Zhang, Email: sjzhang@ipe.ac.cn.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-025-63630-9.

References

  • 1.Zhang, X. et al. Regioselective hydroformylation of propene catalysed by rhodium-zeolite. Nature629, 597–602 (2024). [DOI] [PubMed] [Google Scholar]
  • 2.Liu, L. et al. Dealuminated Beta zeolite reverses Ostwald ripening for durable copper nanoparticle catalysts. Science383, 94–101 (2024). [DOI] [PubMed] [Google Scholar]
  • 3.Zhou, H. et al. Cobaltosilicate zeolite beyond platinum catalysts for propane dehydrogenation. Nat. Catal. 8, 357–367 (2025).
  • 4.Li, F. et al. Molecular tuning of CO2-to-ethylene conversion. Nature577, 509–513 (2020). [DOI] [PubMed] [Google Scholar]
  • 5.Witham, C. A. et al. Converting homogeneous to heterogeneous in electrophilic catalysis using monodisperse metal nanoparticles. Nat. Chem.2, 36–41 (2010). [DOI] [PubMed] [Google Scholar]
  • 6.Cui, X. et al. Bridging homogeneous and heterogeneous catalysis by heterogeneous single-metal-site catalysts. Nat. Catal.1, 385–397 (2018). [Google Scholar]
  • 7.Wang, G. & Cai, G. Steric hindrance effect and kinetic investigation for ionic liquid catalyzed synthesis of 4-hydroxy-2-butanone via aldol reaction. Chem. Eng. Sci. 229, 116089 (2021).
  • 8.Dupont, J. et al. Ionic liquids in metal, photo-, electro-, and (bio) catalysis. Chem. Rev.124, 5227–5420 (2024). [DOI] [PubMed] [Google Scholar]
  • 9.Wang, T. et al. Enhancing oxygen reduction electrocatalysis by tuning interfacial hydrogen bonds. Nat. Catal.4, 753–762 (2021). [Google Scholar]
  • 10.Cao, R. et al. Co-upcycling of polyvinyl chloride and polyesters. Nat. Sustain.6, 1685–1692 (2023). [Google Scholar]
  • 11.Coskun, O. K. et al. Synergistic Effects of the Electric Field Induced by Imidazolium Rotation and Hydrogen Bonding in Electrocatalysis of CO2. J. Am. Chem. Soc.146, 23775–23785 (2024). [DOI] [PubMed] [Google Scholar]
  • 12.Piccirilli, L. et al. Versatile CO2 hydrogenation-dehydrogenation catalysis with a Ru-PNP/ionic liquid system. J. Am. Chem. Soc.145, 5655–5663 (2023). [DOI] [PubMed] [Google Scholar]
  • 13.Liu, Y. et al. Improving CO2 photoconversion with ionic liquid and Co single atoms. Nat. Commun.14, 1457 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Wolf, P. et al. Improving the performance of supported ionic liquid phase catalysts for the ultra-low-temperature water gas shift reaction using organic salt additives. ACS Catal.12, 5661–5672 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Qadir, M. I. & Dupont, J. Thermo- and photocatalytic activation of CO2 in ionic liquids nanodomains. Angew. Chem. Int. Ed.62, e202301497 (2023). [DOI] [PubMed] [Google Scholar]
  • 16.Hu, S. et al. Engineering the electronic structure of perovskite oxide surface with ionic liquid for enhanced oxygen reduction reaction. Appl. Catal. B Environ.282, 119593 (2021). [Google Scholar]
  • 17.Jacobsen, E. N., Zhang, W. & Guler, M. L. Electronic tuning of asymmetric catalysts. J. Am. Chem. Soc.113, 6703–6704 (1991). [Google Scholar]
  • 18.Harper, K. C. & Sigman, M. S. Three-dimensional correlation of steric and electronic free energy relationships guides asymmetric propargylation. Science333, 1875–1878 (2011). [DOI] [PubMed] [Google Scholar]
  • 19.Yan, H., Huo, F., Li, P. & Li, C. Mild catalytic mechanism of the mannich reaction for synthesizing methylacrolein by sec-amine short-chain aliphatic acid ionic liquid catalysts. ACS Sustain. Chem. Eng.10, 6687–6698 (2022). [Google Scholar]
  • 20.Xu, J. et al. Synthesis of renewable C8–C10 alkanes with angelica lactone and furfural from carbohydrates. ACS Sustain. Chem. Eng.6, 6126–6134 (2018). [Google Scholar]
  • 21.Zhang, X. et al. Synthesis of high density and low freezing point jet fuels range cycloalkanes with cyclopentanone and lignin-derived vanillins. J. Energy Chem.79, 22–30 (2023). [Google Scholar]
  • 22.Xu, J. et al. Synthesis of high-density aviation fuels with methyl benzaldehyde and cyclohexanone. Green. Chem.20, 3753–3760 (2018). [Google Scholar]
  • 23.Zhang, T. et al. Continuous synthesis of methacrolein over sulfonic acid resin catalysts: unraveling the effect of acid strength and solvent permittivity. Ind. Eng. Chem. Res.62, 12949–12962 (2023). [Google Scholar]
  • 24.Yu, J. et al. Catalytic synthesis of methacrolein via the condensation of formaldehyde and propionaldehyde with L-proline. Green. Chem.22, 4222–4230 (2020). [Google Scholar]
  • 25.Lei, L. et al. Rapid and continuous synthesis of methacrolein with high selectivity by condensation of propanal with formaldehyde in laboratory. Can. J. Chem. Eng.95, 1985–1992 (2017). [Google Scholar]
  • 26.Zhang, T. et al. Proton shuttle mediated by ionic liquid promotes aldol condensation. J. Catal.444, 116001 (2025). [Google Scholar]
  • 27.Zhao, H. et al. Novel continuous process for methacrolein production in numerous droplet reactors. AIChE J.66, e16239 (2020). [Google Scholar]
  • 28.Zhao, Q. et al. Kinetic and mechanistic insights into the roles of protonated morpholine species and weak acidic environment in methylacrolein production catalyzed by morpholine/acid. Chem. Eng. J.460, 140409 (2023). [Google Scholar]
  • 29.Zhao, H. et al. Pickering emulsion stabilized by dual stabilizer: A novel reaction/separation system for methacrolein synthesis. Chem. Eng. Sci.229, 116038 (2021). [Google Scholar]
  • 30.Wang, Y.-L. et al. Microstructural and dynamical heterogeneities in ionic liquids. Chem. Rev.120, 5798–5877 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Comtet, J. et al. Nanoscale capillary freezing of ionic liquids confined between metallic interfaces and the role of electronic screening. Nat. Mater.16, 634–639 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Pham, T. A. et al. Structural anomalies and electronic properties of an ionic liquid under nanoscale confinement. J. Phys. Chem. Lett.11, 6150–6155 (2020). [DOI] [PubMed] [Google Scholar]
  • 33.Villar-Garcia, I. J., Lovelock, K. R. J., Men, S. & Licence, P. Tuning the electronic environment of cations and anions using ionic liquid mixtures. Chem. Sci.5, 2573–2579 (2014). [Google Scholar]
  • 34.Schlaich, A., Jin, D., Bocquet, L. & Coasne, B. Electronic screening using a virtual Thomas-Fermi fluid for predicting wetting and phase transitions of ionic liquids at metal surfaces. Nat. Mater.21, 237–245 (2022). [DOI] [PubMed] [Google Scholar]
  • 35.Sun, T. et al. Water-controlled structural transition and charge transfer of interfacial ionic liquids. J. Phys. Chem. Lett.13, 7113–7120 (2022). [DOI] [PubMed] [Google Scholar]
  • 36.Park, S., Han, D. H., Lee, J. G. & Chung, T. D. Current amplification and ultrafast charge transport in a single microdroplet of bromide/polybromide-based ionic liquid. ACS Appl. Energy Mater.3, 5285–5292 (2020). [Google Scholar]
  • 37.Pudasaini, P. R. et al. Ionic liquid activation of amorphous metal-oxide semiconductors for flexible transparent electronic devices. Adv. Funct. Mater.26, 2820–2825 (2016). [Google Scholar]
  • 38.Xu, C. et al. Dynamics of excess electronic charge in aliphatic ionic liquids containing the bis(trifluoromethylsulfonyl)amide anion. J. Am. Chem. Soc.135, 17528–17536 (2013). [DOI] [PubMed] [Google Scholar]
  • 39.Liu, M. & Yan, T. A unified approach to derive atomic partial charges and polarizabilities of ionic liquids. J. Chem. Theory Comput.18, 4342–4353 (2022). [DOI] [PubMed] [Google Scholar]
  • 40.McDaniel, J. G. & Yethiraj, A. Influence of electronic polarization on the structure of ionic liquids. J. Phys. Chem. Lett.9, 4765–4770 (2018). [DOI] [PubMed] [Google Scholar]
  • 41.Imai, M. et al. Attenuated total reflectance far-ultraviolet and deep-ultraviolet spectroscopy analysis of the electronic structure of a dicyanamide-based ionic liquid with Li+. Phys. Chem. Chem. Phys.22, 21768–21775 (2020). [DOI] [PubMed] [Google Scholar]
  • 42.Cano, I. et al. Blurring the boundary between homogenous and heterogeneous catalysis using palladium nanoclusters with dynamic surfaces. Nat. Commun.12, 4965 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Zou, Y. et al. Ionic liquids tailoring crystal orientation and electronic properties for stable perovskite solar cells. Nano Energy112, 108449 (2023). [Google Scholar]
  • 44.Gebbie, M. A. et al. Ionic liquids behave as dilute electrolyte solutions. Proc. Natl. Acad. Sci. USA110, 9674–9679 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Hunt, P. A., Ashworth, C. R. & Matthews, R. P. Hydrogen bonding in ionic liquids. Chem. Soc. Rev.44, 1257–1288 (2015). [DOI] [PubMed] [Google Scholar]
  • 46.Travis, C. R. et al. WDR5 binding to histone serotonylation is driven by an edge–face aromatic interaction with unexpected electrostatic effects. J. Am. Chem. Soc.146, 27451–27459 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Pérez, O., Odio, O. F. & Reguera, E. XPS as a probe for the bonding nature in metal acetates. N. J. Chem.46, 11255–11265 (2022). [Google Scholar]
  • 48.Wu, R. et al. Intermetallic synergy in platinum–cobalt electrocatalysts for selective C–O bond cleavage. Nat. Catal.7, 702–718 (2024). [Google Scholar]
  • 49.Khamkongkaeo, A. et al. X-ray absorption spectroscopy investigation of relationship between Mg vacancy and magnetic properties of MgO powder. J. Magn. Magn. Mater.460, 327–333 (2018). [Google Scholar]
  • 50.Shmukler, L. E., Fedorova, I. V., Gruzdev, M. S. & Safonova, L. P. Triethylamine-based salts: protic ionic liquids or molecular complexes?. J. Phys. Chem. B123, 10794–10806 (2019). [DOI] [PubMed] [Google Scholar]
  • 51.Wiedemann, C., Fushman, D. & Bordusa, F. 15N NMR studies provide insights into physico-chemical properties of room-temperature ionic liquids. Phys. Chem. Chem. Phys.23, 12395–12407 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Kashin, A. S., Galkin, K. I., Khokhlova, E. A. & Ananikov, V. P. Direct observation of self-organized water-containing structures in the liquid phase and their influence on 5-(hydroxymethyl)furfural formation in ionic liquids. Angew. Chem. Int. Ed.55, 2161–2166 (2016). [DOI] [PubMed] [Google Scholar]
  • 53.Deng, H. et al. Ionosomes: observation of ionic bilayer water clusters. J. Am. Chem. Soc.143, 7671–7680 (2021). [DOI] [PubMed] [Google Scholar]
  • 54.Bian, H. et al. Ion clustering in aqueous solutions probed with vibrational energy transfer. Proc. Natl. Acad. Sci. USA108, 4737–4742 (2011). [Google Scholar]
  • 55.Sugumaran, T. et al. The conductivity and dielectric studies of polymer electrolytes based on iota-carrageenan with sodium iodide and 1-butyl-3-methylimidazolium iodide for the dye-sensitized solar cells. Ionics25, 763–771 (2019). [Google Scholar]
  • 56.Gomes, R. J. et al. Modulating water hydrogen bonding within a non-aqueous environment controls its reactivity in electrochemical transformations. Nat. Catal.7, 689–701 (2024). [Google Scholar]
  • 57.Stange, P., Fumino, K. & Ludwig, R. Ion speciation of protic ionic liquids in water: transition from contact to solvent-separated ion pairs. Angew. Chem. Int. Ed.52, 2990–2994 (2013). [DOI] [PubMed] [Google Scholar]
  • 58.Stange, P., Verevkin, S. P. & Ludwig, R. Combined spectroscopic, thermodynamic, and theoretical approach for detecting and quantifying hydrogen bonding and dispersion interaction in ionic liquids. Acc. Chem. Res.56, 3441–3450 (2023). [DOI] [PubMed] [Google Scholar]
  • 59.Zheng, X. et al. Synthesis of 2-ethyl-3-oxazolidineethanol from propanal and diethanolamine: kinetic and process optimization study. Ind. Eng. Chem. Res.62, 5792–5803 (2023). [Google Scholar]
  • 60.Fenton, D. E. & Papageorgiou, G. Synthesis of unsymmetrical dinucleating ligands bearing nitrogen and oxygen donor atoms. Tetrahedron52, 5913–5928 (1996). [Google Scholar]
  • 61.Gao, Z. & Fletcher, S. P. Asymmetric conjugate addition of alkylzirconium reagents to α, β-unsaturated thioesters: access to fragrances and acyclic stereochemical arrays. Chem. Commun.53, 10216–10219 (2017). [DOI] [PubMed] [Google Scholar]
  • 62.Kouridaki, A., Montagnon, T., Tofi, M. & Vassilikogiannakis, G. Photooxidations of 2-(γ,ε-dihydroxyalkyl) furans in water: synthesis of DE-bicycles of the pectenotoxins. Org. Lett.14, 2374–2377 (2012). [DOI] [PubMed] [Google Scholar]
  • 63.Frisch, M. J. et al. Gaussian 16 Rev. C.02. Gaussian 16 Rev. C.02, https://gaussian.com/relnotes/.
  • 64.Andrae, D. et al. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta. 77, 123–141 (1990). [Google Scholar]
  • 65.Johnson, B. G., Gill, P. M. W. & Pople, J. A. The performance of a family of density functional methods. J. Chem. Phys.98, 5612–5626 (1993). [Google Scholar]
  • 66.Butera, V. Density functional theory methods applied to homogeneous and heterogeneous catalysis: a short review and a practical user guide. Phys. Chem. Chem. Phys.26, 7950–7970 (2024). [DOI] [PubMed] [Google Scholar]
  • 67.Zhao, Y. & Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc.120, 215–241 (2008). [Google Scholar]
  • 68.Marenich, A. V., Cramer, C. J. & Truhlar, D. G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B113, 6378–6396 (2009). [DOI] [PubMed] [Google Scholar]
  • 69.Lu, T. & Chen, F. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem.33, 580–592 (2012). [DOI] [PubMed] [Google Scholar]
  • 70.Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph.14, 33–38 (1996). [DOI] [PubMed] [Google Scholar]
  • 71.Lu, T. & Chen, Q. Shermo: A general code for calculating molecular thermochemistry properties. Comput. Theor. Chem.1200, 113249 (2021). [Google Scholar]
  • 72.Vanommeslaeghe, K. et al. Influence of stacking on the hydrogen bond donating potential of nucleic bases. J. Chem. Theory Comput.2, 1444–1452 (2006). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Data Availability Statement

The source data generated in this study have been deposited in the Figshare database under accession code 10.6084/m9.figshare.29502947.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES