Abstract

We synthesized two enantiomeric three-dimensional chiral coordination polymers, [Cu8((R)-CIA)4(H2O)9]·3H2O (1-D) and [Cu8((S)-CIA)4(H2O)9]·3H2O (1-L), using 5-(1-carboxyethoxy) isophthalic acid as the organic linker. Crystal structure analysis revealed that both 1-D and 1-L exhibit complex and fascinating three-dimensional self-penetrating structures. Further characterization of the properties revealed that 1-D exhibits notable photoluminescence (PL) and magnetic behaviors. Specifically, photoluminescence measurements showed that when excited at 390 nm (λex = 390 nm), 1-D displays maximum emission peaks at 445 and 470 nm, indicating a strong fluorescence emission. The magnetic susceptibility measurements reveal that these coordination polymers exhibit pronounced antiferromagnetic behavior. By integrating the crystal structure with the performance analysis, it is demonstrated that 1-D and 1-L, as chiral coordination polymers, not only show distinct photoluminescent and magnetic characteristics but also provide crucial experimental and theoretical insights for the development and application of chiral functional materials. These research findings provide a solid foundation for the future development of multifunctional materials based on chiral metal coordination polymers, which are expected to play a significant role in areas such as optoelectronic devices, magnetic materials, and catalysis.
1. Introduction
Recently, chiral coordination polymers (CCPs) have attracted considerable interest for their distinct network structures and wide-ranging potential in various applications. Due to their tunable and porous nature, CCPs demonstrate ideal properties in asymmetric catalysis and other applications requiring noncentrosymmetric crystal structures such as enantioselective separation,1−4 chemical sensing,5−7 biomedical imaging,8,9 drug delivery,10,11 conductivity/semiconductivity,12−15 solar energy harvesting,16,17 and nonlinear optical devices.18−20 In particular, CCPs are recognized as a highly adaptable framework for the design of single-site catalysts in various organic transformation processes.21−23 Over the past few decades, asymmetric catalysis of prochiral molecules using chiral metal complexes has become the primary and widely applied synthetic method for producing single enantiomer compounds, gradually evolving into a crucial technology for the pharmaceutical and agrochemical industries.24,25
Despite the advancements in coordination polymers and crystal engineering, designing and rationally constructing CCPs remains one of the significant challenges.26 The current research findings indicate that the method of synthesizing CCPs using chiral ligands is both effective and reliable.27 Therefore, it is particularly important to design and synthesize novel enantiopure ligands.28,29 Lactic acid, as an affordable and straightforward chiral ligand source, can serve as the best enantiopure linker for the construction of CCPs.30 Considering the high flexibility of lactic acid, there are still some limitations in constructing CCPs with 3D and stable architectures directly from lactic acids.31−35
Based on the above considerations, we converted dimethyl 5-hydroxyisophthalate into a chiral linker by introducing a lactic acid moiety, thereby synthesizing a semirigid T-shaped chiral aromatic polycarboxylate ligand, 5-(1-carboxyethoxy) isophthalic acid ((R)-H3CIA and (S)-H3CIA) (Scheme 1). The ligand’s three carboxylate groups can undergo complete or partial deprotonation, allowing them to act as potential donors or acceptors for hydrogen bonds. This property may facilitate the formation of higher-dimensional complexes through intermolecular or intramolecular hydrogen bonding interactions. In this work, two enantiomeric CCPs, [Cu8((R)-CIA)4(H2O)8]·3H2O (1-D) and [Cu8((S)-CIA)4(H2O)8]·3H2O (1-L), were synthesized by using ((R)-H3CIA and (S)-H3CIA) as the starting materials (Table 1). The synthesis, crystal structures, powder X-ray diffraction (PXRD), thermal stability, circular dichroism (CD) spectra, infrared (IR) spectra, as well as luminescent and magnetic properties have been systematically examined.
Scheme 1. Schematic Diagram of the Structure of Enantiopure Linkers.
Table 1. Crystal Structure Parameters of CCPs.
| 1-D | 1-L | |
|---|---|---|
| empirical formula | C44H48Cu8O44 | C44H48Cu8O44 |
| Fw | 1789.14 | 1789.14 |
| crystal system | monoclinic | monoclinic |
| space group | P21 | P21 |
| a (Å) | 10.202(2) | 10.217(2) |
| b (Å) | 26.919(5) | 26.950(5) |
| c (Å) | 10.841(2) | 10.843(2) |
| α (°) | 90 | 90 |
| β (°) | 93.91(3) | 94.21(3) |
| γ (°) | 90 | 90 |
| V (Å3) | 2970.3(10) | 2977.4(10) |
| Z | 2 | 2 |
| Dcalc (g cm–3) | 2.000 | 1.996 |
| μ (mm–1) | 2.923 | 2.916 |
| F (000) | 1792 | 1792 |
| Flack | 0.05(2) | 0.05(2) |
| collected/unique | 23314/10289 | 22707/9806 |
| R(int) | 0.0439 | 0.0762 |
| GOF on F2 | 1.080 | 1.062 |
| Rla[I > 2σ(I)] | 0.0580 | 0.0603 |
| wR2b [I > σ(I)] | 0.1367 | 0.1403 |
| Rla (all) | 0.0695 | 0.0907 |
| wR2b (all) | 0.1423 | 0.1578 |
R1= ∑||Fo| – |Fc||/∑|Fo|;
wR2= ∑ {[w(Fo2– Fc2)2/ ∑w(Fo2)2]}1/2.
2. Experimental Section
2.1. General Methods and Materials
The reagents used in this study were purchased from commercial sources and directly used in the experiments. The structure of CCPs was analyzed on a PerkinElmer Spectrum 100 FT-IR spectrometer with a DGTS detector, performing 32 scans within the 4000–500 cm–1 range, and the KBr pellet method was employed. Elemental analysis of carbon and hydrogen was conducted with a PerkinElmer 2400 elemental analyzer. TG analyses were performed on a PerkinElmer STA 6000 thermal analyzer under an air atmosphere, with a heating rate of 10 °C min–1 over the temperature range of 30–800 °C. There are proton (1H) and carbon (13C) nuclear magnetic resonance (NMR) spectra measured using a Bruker Avance (400 MHz) spectrometer with DMSO/CD3Cl as the solvent and TMS as the internal standard. PXRD data were obtained over a 2θ range of 5–50° using a Rigaku D/Max-IIIB X-ray diffractometer with Cu Kα (λ = 1.5406 Å) radiation (40 kV, 200 mA) and a Ni filter. The CD spectra of CCPs were measured at room temperature with a Jasco J-810(S) spectropolarimeter, and the samples were prepared using the KBr pellet method. There are luminescence spectra obtained using a Shimadzu RF-5301 spectrophotometer.
2.2. Synthesis of ((R)-H3CIA and (S)-H3CIA)
Dissolve 4.2 g (20 mmol) of dimethyl 5-hydroxyisophthalate and 2.2 g (21 mmol) of l-(−)-lactic acid methyl ester or d-(−)-lactic acid methyl ester in 40 mL of tetrahydrofuran, and then add 5.8 g (22 mmol) of triphenylphosphine. After stirring the mixture for 10 min, slowly add a 15 mL tetrahydrofuran solution of 3.8 g (22 mol) diethyl azodicarboxylate over 15 min under 0 °C. After the addition is complete, remove the ice bath and allow the reaction to continue stirring at room temperature for 3 h. After the reaction is complete, the reaction mixture is concentrated, the mixture is filtered, and the solvent is removed under reduced pressure. The resulting residue is purified by silica gel column chromatography, with a solvent mixture of petroleum ether and 20% ethyl acetate as the eluent. Dissolve the purified ester (5.9 g, 20 mmol) in 15 mL of methanol, add an equimolar amount of sodium methoxide, and heat the mixture under reflux until the solution becomes clear. After cooling, the pH of the solution was adjusted to 2–3 using concentrated hydrochloric acid and the pH was verified and adjusted to the same range.
The precipitated solid was isolated by filtration, yielding the pure compound (R)-H3CIA or (S)-H3CIA.Yield:4.0 g, 80%. Anal. Calcd for C11H10O7 (254.2): C, 51.93; H, 3.93; found: C, 52.22; H, 4.12. 1H NMR (400 MHz, DMSO-d6), δ (ppm): 13.26 (3H, brs), 8.08 (1H, m), 7.59–7.60 (2H, d, J = 0.8 Hz), 4.96–5.01(1H, q, J = 2.8 Hz), 1.53–1.55 (3H, d, J = 6.8 Hz); 13C NMR (100 MHz, DMSO-d6), δ(ppm): 173.4, 166.7, 167.1, 158.2, 133.0, 123.2, 119.9,72.6, 18.6. As shown in Figures S1–S2, the IR spectrum (solid KBr pellet, cm–1) of (S)-H3CIA exhibits the following peaks: 3393 m, 1715 s, 1314 s, 1279 s, 1136 s, 1048 m, 930 w, 810 w, 759 w, 694 s; (R)-H3CIA IR (solid KBr pellet, cm–1): 3390 m, 1715 s, 1314 s,1279 s, 1136 s, 1048 m, 930 w, 810 w, 759 w, 694 s.
2.3. Synthesis of [Cu8((R)-CIA)4(H2O)8]·3H2O (1-D)
Dissolve 0.2 mmol (50 mg) of CuSO4·5H2O and 0.2 mmol (52 mg) of (R)-H3CIA in 5 mL of deionized water and stir for 15 min. Subsequently, the mixture was transferred into a 15 mL stainless steel reactor with a PTFE lining and allowed to react at 120 °C for 3 days. After the reaction, the mixture was slowly cooled to room temperature at a rate of 10 °C per hour. The green crystals were obtained by filtration, with a yield of 28 mg, corresponding to a 60% yield based on (R)-H3CIA. The elemental analysis results are as follows: for the compound C44 H48 Cu8 O44 (1-D): C, 29.51; H, 2.68; found: C, 29.41; H, 2.24. The IR of [Cu8((R)-CIA)4(H2O)8]·3H2O (1-D) is shown in Figure S3, IR (solid KBr pellet, cm–1): 3445.6 m, 1627.0 s, 1567.3 s, 1456.8 s, 1387.9 s, 1270.2 m, 1233.5 w, 1127.2 w, 1086.9 w, 1041.6 s, 994.3 w, 816.9 m, 749.4 s.
2.4. Synthesis of [Cu8((S)-CIA)4(H2O)8]·3H2O (1-L)
An experimental protocol identical to that of 1-D was followed, with the sole modification being the substitution of (S)-H3CIA for its R enantiomer. After filtration, green crystals were isolated, yielding 24 mg (52%, calculated relative to that of (S)-H3CIA). Elemental analysis (%) Calc for 1-L (C44 H48 Cu8 O44): C, 29.51; H, 2.68; found: C, 29.41; H, 2.24. The IR of [Cu8((S)-CIA)4(H2O)8]·3H2O (1-L) is shown in Figure S4, IR (solid KBr pellet, cm–1): 3449.9 m, 1626.1 s, 1566.3 s, 1456.3 s, 1391.7 s, 1273.1 m, 1233.5 w, 1129.4 w, 1095.6 w, 1044.5 s, 992.4 w, 816.9 m, 748.4 s.
2.5. X-ray Crystallography
The crystal structures of the CCPs were analyzed using single-crystal X-ray diffraction with a Rigaku R-AXIS RAPID diffractometer, employing Mo Kα radiation (λ = 0.71073 Å) and a graphite monochromator. Data collection was carried out at a temperature of 291 K. During the data processing, absorption corrections were applied empirically using equivalent reflections. The crystal structure was initially solved using direct methods and subsequently optimized through full-matrix least-squares calculations based on F2. The refinement process was carried out using the SHELXS-97 software package for crystallographic analysis.36 In the analysis, isolated oxygen atoms were assumed to be water molecules. The structural parameters of CCPs 1-D and 1-L are listed in Table 1, while Tables S1 and S2 provide the bond lengths and bond angles.
3. Results and Discussion
3.1. Structure
Given that CCPs 1-D and 1-L are enantiomers, only the structural features of 1-D are discussed here. Single-crystal X-ray diffraction analysis indicates that CCP 1-D adopts an intriguing three-dimensional structure, crystallizing in the noncentrosymmetric space group P21, with a Flack parameter of 0.05(2). The asymmetric unit of CCP 1-D comprises eight individual Cu(II) ions and four deprotonated (R)-CIA4– ligands, eight water molecules coordinated to the metal centers, and an additional four lattice water molecules (Figure 1). It is noteworthy that a hydroxylation reaction occurred to the benzene ring of the ligand during the coordination process under hydrothermal conditions.
Figure 1.

Asymmetric unit structure of 1-D, with H and lattice H2O removed to ensure a clear representation. Symmetry operations include: I: x + 1, y, z; II: –x + 1, y + 1/2, –z + 1; III: x + 1, y, z.
The presence of Cu2+ ion and SO42– in the reaction vessel might enhance the hydroxylation, leading to the production of phenol.37,38 On the one hand, Cu2+ ions can promote the hydroxylation reaction by coordinating with the reactants or by acting as electron transfer catalysts in the redox process. On the other hand, SO42– ions may stabilize the copper ions, modify the polarity of the reaction environment, or regulate solubility, thus optimizing the reaction conditions and facilitating the production of phenol.37,38 The distorted square-pyramid coordination geometries of Cu1, Cu2, and Cu5 involve the binding of three carboxylate oxygen atoms from three (R)-H3CIA ligands and two hydroxyl oxygen atoms from two (R)-H3CIA ligands. For Cu3 and Cu7, their coordination environments comprise three O from the carboxylate groups of the (R)-H3CIA ligand and two from coordinated H2O, forming an overall square-pyramidal geometry. Cu4 and Cu8 adopt distorted pentagonal bipyramidal geometries, where three carboxylate O from (R)-H3CIA ligands coordinate to each metal center, and the two oxygen atoms originate from the two water molecules that are coordinated, while an additional oxygen atom is derived from the ether group present in the ligand. Cu6 shows a distorted square-pyramid geometry, coordinated by two carboxylate oxygen atoms from two (R)-H3CIA ligands, two hydroxyl oxygen atoms from two (R)-H3CIA ligands, and one oxygen atom from one coordinated water molecule. The coordination geometry of Cu6 is a distorted square pyramid, consisting of two carboxylate O from two (R)-H3CIA ligands, two hydroxyl O, and one O donor from bound water.
The distances between Cu and the O atoms vary from 1.886(8) to 2.884(2) Å, which are all in reasonable range. The ligand (R)-H3CIA shows versatile coordination sites: (A) the carboxylate group from the lactic unit; (B) the two carboxylate groups attached to the benzene ring; (C) the ether group; (D) the hydroxyl group introduced by hydroxylation reaction (Figure 2). Among the reported (R)-H3CIA complexes in the literature, no similar structures have been found to exhibit such a diverse and intricate coordination mode.39,40
Figure 2.

Coordination site of (R)-H3CIA in 1-D Color code: C gray, O red, Cu turquoise.
First of all, two (R)-H3CIA ligands coordinate to different central metals (Cu1 and Cu2/Cu5 and Cu6) with different angles to form a plane. Two planes are connected to each other by the coordination of two chiral carboxylate groups (O31–C43–O30 and O7–C10–O6) and Cu1, 2/Cu5, 6, forming a right-handed helical double-layer secondary structure [Cu2(CIA)2]22– (Figure 3a). Such secondary structures [Cu2(CIA)2]22– are further linked with each other by another two chiral carboxylate group (O14–C21–O15 and O22–C32–O23) coordinating with Cu3, 4/Cu7, and 8, eventually generating a 2D framework perpendicular to the c axis (Figure 3b).
Figure 3.
(a) Right-handed helical double-layer secondary structure, (b) double-layer 2D framework, (c) 3D framework of 1-D, (d) 3D interpenetration frameworks in 1-D, and (e) diagram illustrating the self-penetrating topological structure of CCP 1-D.
On the other hand, the two neighboring [Cu2(CIA)2]22– are connected by two achiral carboxylate groups in [Cu2(CIA)2]22– (O11–C19–O12,O9–C18–O10, O1–C7–O2, O3–C8–O4 and O17–C29–O18,O19–C30–O20, O25–C40–O26, O27–C41–O28) coordinating with metal Cu3/Cu4/Cu7/Cu8 to get a 3D framework {[Cu2(CIA)2]2}n perpendicular to c axis (Figure 3c). The 3D framework {[Cu2(CIA)2]2}n could also be generated by the chiral carboxylate groups coordinating with Cu3/Cu4/Cu7/Cu8. (Figure 3d).
Employing a topological method that simplifies complex multidimensional architectures into basic node-link frameworks allows for a clearer understanding of the structural characteristics of CCP 1-D. According to the above analysis, we can simplify the whole framework by ignoring the chelated chiral carboxylate groups and simply every two Cu atoms (Cu1 and Cu2, Cu3 and Cu4, Cu5 and Cu6, Cu7 and Cu8) to one point, thus obtaining a polycatenated 2-nodal 2,6-connected topology with a point symbol of {2} {44.2.6.44.65} (Figure 3e).
3.2. PXRD Patterns and Thermogravimetric Analysis
PXRD analyses were conducted to confirm the phase purity of the solid-state samples. In the case of CCPs 1-D and 1-L, characteristic diffraction peaks appeared at 6.66 and 8.79, which are associated with the [020] and [011] planes, respectively, and match well with those predicted in the simulated powder diffraction pattern, confirming that the obtained structures accurately represent the bulk crystals. Additionally, variations in peak intensity may arise from the preferred orientation of the crystals, as demonstrated in Figure 4.
Figure 4.

PXRD patterns of 1-D and 1-L.
A comparison of the infrared spectra between the ligand and CCPs (Figures S1–S4) reveals several noticeable differences. Several noticeable variations are observed, indicating a relationship between the ligand and metal ions. For example, the strong absorption peak at 1715 cm–1 in the ligand’s infrared spectrum is typically attributed to the C = O stretching vibration of the carboxyl group. In the metal complex, this peak undergoes a significant blue shift (from 1715 to 1626.1 cm–1). This shift indicates coordination between the metal ion and the ligand’s carboxyl group, which affects the electron density of the C = O bond and, in turn, alters the vibration frequency.
The stability of the complexes was assessed through thermogravimetric analysis (TGA) (Figure 5). The TGA was conducted on single-crystal samples of the CCPs, with the temperature increase set at 10 °C per minute. The structures of the CCPs exhibit an initial weight reduction of 12.2% between 30 and 175 °C; this results from the solvent molecules evaporating. The subsequent weight loss of 9.6% takes place between 205 and 250 °C; there is likely a loss of coordinated water molecules. In the third stage, the ligand framework begins to decompose at temperatures above 250 °C, with the remaining mass being attributed to the formation of CuO.
Figure 5.

TGA diagram of CCPs.
3.3. CD Spectra Analysis
The enantiomeric properties of CCPs were further confirmed through solid-state CD spectroscopy (Figure 6). The spectrum of 1-D exhibits a prominent positive Cotton effect (CE) peak around 251 nm, indicating its single chirality. Additionally, the presence of a mirror-image signal in 1-L confirms the enantiomeric relationship between CCPs 1-D and 1-L.
Figure 6.
(a) Solid-state CD spectra of (R)-H3CIA and (S)-H3CIA and (b) solid-state CD spectra of 1-D and 1-L.
Through the CD measurements of both the ligand and the CCPs 1-D and 1-L, it is evident that the two ligands are a pair of corresponding isomers, and thus, the resulting CCPs are also a pair of corresponding isomers. When comparing the CD spectra of the free ligand with those of the CCPs, distinct shifts in the peak positions are observed. The observed differences stem from the fact that the binding of the metal ion to the free ligand can induce a change in the ligand’s spatial structure due to the coordination, leading to shifts in the peak positions and changes in the peak intensity in the CD spectra.
3.4. Photophysical Properties
This study focuses on the solid-state excitation and emission spectra of CCP 1-D, measured at room temperature (Figure 7). The emission spectra of 1-D show peaks at 445 and 470 nm, with maxima observed at these wavelengths when excited at 390 nm (λex = 390 nm). To further investigate the source of these emission bands, the free ligand’s luminescence was also measured at room temperature, showing maximum emission peaks at 415 and 462 nm when excited at 395 nm (λex = 395 nm). Therefore, the emission bands observed in 1-D can be ascribed to the intralinker π–π* transition. The fluorescence emission variation between CCP 1-D and the free ligand can be attributed to changes in the ligand’s configuration and the π–π* stacking interactions. Molecular interactions in the complex bring the luminophores closer, which enhances electronic coupling between them and leads to spectral shifts.41,42 Based on these observations, CCP 1-D appears to be a potential candidate for use in photoactive material applications.
Figure 7.

Room-temperature solid-state fluorescence emission spectrum of CCP 1-D.
3.5. Magnetic Properties Study
Magnetic susceptibility measurements of 1-D were carried out in a 1000 Oe field across a temperature range of 2 to 300 K, yielding a χMT value of 1.715 cm3 K mol–1 at 300 K. (Figure 8a). When the temperature reaches 2 K, the χMT value shows a continuous decline, eventually reaching 0.069 cm3 K mol–1. n 1-D, the Cu1–Cu2 and Cu5–Cu6 ion pairs are each bridged by two (R)-H3CIA ligands, to form a plane with a very short Cu1···Cu2 distance of 2.8841 and Cu5···Cu6 distance of 2.9468 Å, which can be regarded as the direct overlap of the magnetic orbitals. This result indicates overall antiferromagnetic coupling among the paramagnetic Cu(II) ion centers in 1-D.
Figure 8.
(a) χMT versus temperature (T) plots for 1-D and the temperature variation of 1/χM, along with the red curve representing the best fit to the Curie–Weiss law (color online). (b) Magnetization versus magnetic field (M–H) curve of 1-D at 2 K. (c) AC magnetic susceptibility measured at frequencies ranging from 1 to 1000 Hz.
Additionally, we observed that the low χMT value of 1-D at 300 K may be attributed to two factors: The bridging angle serves as a crucial factor that significantly impacts the magnetic properties. Thompson and his team discovered that in planar macrocyclic dinuclear copper(II) complexes, the strength of antiferromagnetic coupling increases linearly as the Cu–O–Cu bridging angle expands.43 This observed trend is in agreement with the research findings of Ruiz et al., who reported similar behavior in hydroxo- and alkoxo-bridged coordination complexes.44 According to our experimental data, the bridging angle of 1-D is relatively large, and this leads to pronounced antiferromagnetic coupling between the copper ions under ambient conditions, thus explaining the observed low χMT value. Moreover, the magnetic coupling between the Cu ions could be influenced by the π–π* interactions present in the coordination environment, further decreasing the observed χMT value. In summary, the magnetic behavior of 1-D is significantly influenced by its coordination environment and intermolecular interactions.
Magnetization measurements were performed at 2 K under varying applied magnetic fields to gain a more detailed understanding of the magnetic behavior. The M(H) curve of 1 consistently approaches a value of 0.08Nβ as the magnetic field increases, but it does not reach saturation even at the maximum field strength of 70 kOe. The maximum magnetization (M) observed deviates from the anticipated net spin value for Cu(II) ions. It is also evidence of an antiferromagnetic interaction existing in this complex (Figure 8b). No out-of-phase signals were observed in the ac magnetic susceptibility measurements across the oscillating frequency range of 1–1000 Hz, suggesting the absence of superparamagnetic behavior or slow magnetic relaxation. As a result, this complex does not exhibit the characteristics of a single-molecule magnet (SMM) (Figure 8c).
4. Conclusions
To conclude, a pair of homochiral CCPs have been successfully synthesized, 1-D and 1-L, using two enantiopure organic linkers, namely, ((R)-H3CIA and (S)-H3CIA). As a pair of enantiomers, CCPs 1-D and 1-L demonstrate a very interesting and complicated 3D polycatenated framework with eight independent Cu(II) centers and versatile coordination sites in one asymmetric unit. Remarkably, a hydroxylation reaction occurred during the coordination process under hydrothermal conditions, introducing one hydroxyl group to the benzene ring of the ligand. The fascinating fluorescence and magnetic behaviors indicate that CCPs 1-D and 1-L may have promising applications as materials with both photoluminescent and magnetic properties. We plan to systematically study the impact of CCP morphology on optical properties (such as CD and fluorescence) in future research.
Acknowledgments
The authors acknowledge a Project Supported by Scientific Research Fund of Zhejiang Provincial Education Department (Y202455509), Zhejiang Public Welfare Research Project (LGJ20E020001), Ningbo Polytechnic National Research Project Cultivation Topic (NZ22GJ004), the National Natural Science Foundations of China (Grant Nos. 52250005, 21875271, U20B2021, 21707147, 51372046, 51479037, 91226202, and 91426304), the support of the Key R & D Projects of Zhejiang Province (Nos. 2022C01236, 2019C01060), Zhejiang Key Laboratory of Data-Driven High-Safety Energy Materials and Applications, and the Entrepreneurship Program of Foshan National Hi-tech Industrial Development Zone.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.4c10465.
Crystallographic data, IR data, and CCDC reference numbers: 1525408 (1-D) and 1525409 (1-L) (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Chen B.; Liang C.; Yang J.; Contreras D. S.; Clancy Y. L.; Lobkovsky E. B.; Yaghi O. M.; Dai S. A microporous metal–organic framework for gas-chromatographic separation of alkanes. Angew. Chem., Int. Ed. 2006, 45 (9), 1390–1393. 10.1002/anie.200502844. [DOI] [PubMed] [Google Scholar]
- Li J.-R.; Kuppler R. J.; Zhou H.-C. Selective gas adsorption and separation in metal–organic frameworks. Chem. Soc. Rev. 2009, 38 (5), 1477–1504. 10.1039/b802426j. [DOI] [PubMed] [Google Scholar]
- Sumida K.; Rogow D. L.; Mason J. A.; McDonald T. M.; Bloch E. D.; Herm Z. R.; Bae T.-H.; Long J. R. Carbon dioxide capture in metal–organic frameworks. Chem. Rev. 2012, 112 (2), 724–781. 10.1021/cr2003272. [DOI] [PubMed] [Google Scholar]
- Zhang Z.; Zhao Y.; Gong Q.; Li Z.; Li J. MOFs for CO2 capture and separation from flue gas mixtures: the effect of multifunctional sites on their adsorption capacity and selectivity. Chem. Commun. 2013, 49 (7), 653–661. 10.1039/C2CC35561B. [DOI] [PubMed] [Google Scholar]
- Allendorf M. D.; Houk R. J.; Andruszkiewicz L.; Talin A. A.; Pikarsky J.; Choudhury A.; Gall K. A.; Hesketh P. J. Stress-induced chemical detection using flexible metal-organic frameworks. J. Am. Chem. Soc. 2008, 130 (44), 14404–14405. 10.1021/ja805235k. [DOI] [PubMed] [Google Scholar]
- Kreno L. E.; Leong K.; Farha O. K.; Allendorf M.; Van Duyne R. P.; Hupp J. T. Metal–Organic Framework Materials as Chemical Sensors. Chem. Rev. 2012, 112 (2), 1105–1125. 10.1021/cr200324t. [DOI] [PubMed] [Google Scholar]
- Hu Z.; Deibert B. J.; Li J. Luminescent metal–organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43 (16), 5815–5840. 10.1039/C4CS00010B. [DOI] [PubMed] [Google Scholar]
- Della Rocca J.; Lin W. Nanoscale metal–organic frameworks: magnetic resonance imaging contrast agents and beyond. Eur. J. Inorg. Chem. 2010, 2010 (24), 3725–3734. 10.1002/ejic.201000496. [DOI] [Google Scholar]
- Ay B.; Takano R.; Ishida T.; Yildiz E. Tricopper (II) bis (2-((hydrogen phosphonato) methyl) benzylphosphonate) as a layered oxo-bridged copper (II) coordination polymer: Synthesis, structure, magnetic property, and catalytic activity. Polyhedron 2022, 225, 116038 10.1016/j.poly.2022.116038. [DOI] [Google Scholar]
- Horcajada P.; Serre C.; Maurin G.; Ramsahye N. A.; Balas F.; Vallet-Regí M.; Sebban M.; Taulelle F.; Férey G. Flexible porous metal-organic frameworks for a controlled drug delivery. J. Am. Chem. Soc. 2008, 130 (21), 6774–6780. 10.1021/ja710973k. [DOI] [PubMed] [Google Scholar]
- Horcajada P.; Chalati T.; Serre C.; Gillet B.; Sebrie C.; Baati T.; Eubank J. F.; Heurtaux D.; Clayette P.; Kreuz C.; Chang J. S.; Hwang Y. K.; Marsaud V.; Bories P. N.; Cynober L.; Gil S.; Férey G.; Couvreur P.; Gref R. Porous metal-organic-framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2010, 9 (2), 172–178. 10.1038/nmat2608. [DOI] [PubMed] [Google Scholar]
- Kobayashi Y.; Jacobs B.; Allendorf M. D.; Long J. R. Conductivity, Doping, and Redox Chemistry of a Microporous Dithiolene-Based Metal–Organic Framework. Chem. Mater. 2010, 22 (14), 4120–4122. 10.1021/cm101238m. [DOI] [Google Scholar]
- Givaja G.; Amo-Ochoa P.; Gómez-García C. J.; Zamora F. Electrical conductive coordination polymers. Chem. Soc. Rev. 2012, 41 (1), 115–147. 10.1039/C1CS15092H. [DOI] [PubMed] [Google Scholar]
- Narayan T. C.; Miyakai T.; Seki S.; Dincă M. High charge mobility in a tetrathiafulvalene-based microporous metal-organic framework. J. Am. Chem. Soc. 2012, 134 (31), 12932–12935. 10.1021/ja3059827. [DOI] [PubMed] [Google Scholar]
- Manna F.; Oggianu M.; Auban-Senzier P.; Novitchi G.; Canadell E.; Mercuri M. L.; Avarvari N. A highly conducting tetrathiafulvalene-tetracarboxylate based dysprosium(iii) 2D metal–organic framework with single molecule magnet behaviour. Chem. Sci. 2024, 15 (46), 19247–19263. 10.1039/D4SC05763E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kent C. A.; Mehl B. P.; Ma L.; Papanikolas J. M.; Meyer T. J.; Lin W. Energy transfer dynamics in metal-organic frameworks. J. Am. Chem. Soc. 2010, 132 (37), 12767–12769. 10.1021/ja102804s. [DOI] [PubMed] [Google Scholar]
- Kent C. A.; Liu D.; Ma L.; Papanikolas J. M.; Meyer T. J.; Lin W. Light Harvesting in Microscale Metal–Organic Frameworks by Energy Migration and Interfacial Electron Transfer Quenching. J. Am. Chem. Soc. 2011, 133 (33), 12940–12943. 10.1021/ja204214t. [DOI] [PubMed] [Google Scholar]
- Férey G. Hybrid porous solids: past, present, future. Chem. Soc. Rev. 2008, 37 (1), 191–214. 10.1039/B618320B. [DOI] [PubMed] [Google Scholar]
- Farrusseng D.; Aguado S.; Pinel C. Metal-organic frameworks: opportunities for catalysis. Angew. Chem., Int. Ed. 2009, 48 (41), 7502–7513. 10.1002/anie.200806063. [DOI] [PubMed] [Google Scholar]
- Manna K.; Zhang T.; Greene F. X.; Lin W. Bipyridine- and phenanthroline-based metal-organic frameworks for highly efficient and tandem catalytic organic transformations via directed C-H activation. J. Am. Chem. Soc. 2015, 137 (7), 2665–2673. 10.1021/ja512478y. [DOI] [PubMed] [Google Scholar]
- Ma L.; Falkowski J. M.; Abney C.; Lin W. A series of isoreticular chiral metal-organic frameworks as a tunable platform for asymmetric catalysis. Nat. Chem. 2010, 2 (10), 838–846. 10.1038/nchem.738. [DOI] [PubMed] [Google Scholar]
- Cho S. H.; Ma B.; Nguyen S. T.; Hupp J. T.; Albrecht-Schmitt T. E. A metal-organic framework material that functions as an enantioselective catalyst for olefin epoxidation. Chem. Commun. 2006, (24), 2563–2565. 10.1039/B600408C. [DOI] [PubMed] [Google Scholar]
- Falkowski J. M.; Sawano T.; Zhang T.; Tsun G.; Chen Y.; Lockard J. V.; Lin W. Privileged phosphine-based metal–organic frameworks for broad-scope asymmetric catalysis. J. Am. Chem. Soc. 2014, 136 (14), 5213–5216. 10.1021/ja500090y. [DOI] [PubMed] [Google Scholar]
- Noyori R. Asymmetric catalysis: science and opportunities (Nobel lecture). Angew. Chem., Int. Ed. 2002, 41 (12), 2008–2022. . [DOI] [PubMed] [Google Scholar]
- Solano F.; Auban-Senzier P.; Olejniczak I.; Barszcz B.; Runka T.; Alemany P. Bis (Vinylenedithio)-Tetrathiafulvalene-Based Coordination Networks. Chem. - Eur. J.. 2023, 29 (8), e202203138 10.1002/chem.202203138. [DOI] [PubMed] [Google Scholar]
- Niu X.; Zhao R.; Yan S.; Li H.; Yang J.; Cao K.; Liu X.; Wang K. Chiral MOFs encapsulated by polymers with poly-metallic coordination as chiral biosensors. Microchimica Acta 2023, 190 (6), 230 10.1007/s00604-023-05807-x. [DOI] [PubMed] [Google Scholar]
- Huang T.-T.; Li Z.-X.; Shi X.-N.; Yue Q.; Gao E.-Q. Homochiral coordination polymers based on proline-derivative: structures, magnetic properties, and selective detection of Cr2O72– anion. J. Solid State Chem. 2022, 308, 122894 10.1016/j.jssc.2022.122894. [DOI] [Google Scholar]
- Yang W.; Lin X.; Jia J.; Blake A. J.; Wilson C.; Hubberstey P.; Champness N. R.; Schröder M. A biporous coordination framework with high H 2 storage density. Chem. Commun. 2008, (3), 359–361. 10.1039/B712201B. [DOI] [PubMed] [Google Scholar]
- Chen Y.-B.; Kang Y.; Zhang J. New mimic of zeolite: heterometallic organic host framework accommodating inorganic cations. Chem. Commun. 2010, 46 (18), 3182–3184. 10.1039/b927101e. [DOI] [PubMed] [Google Scholar]
- Cong Y.; Zhou Y.; Bao J.; Zhang X. Crystallization, recrystallization and melting behaviors in supramolecular poly (l-lactic acid) bonded by metal-ligand coordination. Polymer 2024, 294, 126702 10.1016/j.polymer.2024.126702. [DOI] [Google Scholar]
- Chen Z.; Liu X.; Zhang C.; Zhang Z.; Liang F. Structure, adsorption and magnetic properties of chiral metal–organic frameworks bearing linear trinuclear secondary building blocks. Dalton Trans. 2011, 40 (9), 1911–1918. 10.1039/c0dt01278e. [DOI] [PubMed] [Google Scholar]
- Reger D. L.; Horger J. J.; Smith M. D.; Long G. J.; Grandjean F. Homochiral, Helical Supramolecular Metal–Organic Frameworks Organized by Strong π···π Stacking Interactions: Single-Crystal to Single-Crystal Transformations in Closely Packed Solids. Inorg. Chem. 2011, 50 (2), 686–704. 10.1021/ic102256t. [DOI] [PubMed] [Google Scholar]
- Tan Y.-X.; He Y.-P.; Zhang J. Serine-Based Homochiral Nanoporous Frameworks for Selective CO2 Uptake. Inorg. Chem. 2011, 50 (22), 11527–11531. 10.1021/ic201442u. [DOI] [PubMed] [Google Scholar]
- Reger D. L.; Leitner A. P.; Smith M. D. Homochiral Helical Metal–Organic Frameworks of Potassium. Inorg. Chem. 2012, 51 (19), 10071–10073. 10.1021/ic301228j. [DOI] [PubMed] [Google Scholar]
- Reger D. L.; Leitner A.; Smith M. D.; Tran T. T.; Halasyamani P. S. Homochiral Helical Metal–Organic Frameworks of Group 1 Metals. Inorg. Chem. 2013, 52 (17), 10041–10051. 10.1021/ic401327h. [DOI] [PubMed] [Google Scholar]
- Kurtz S. K.; Perry T. A powder technique for the evaluation of nonlinear optical materials. J. Appl. Phys. 1968, 39 (8), 3798–3813. 10.1063/1.1656857. [DOI] [Google Scholar]
- Mahapatro S. N.; Panigrahi A. K.; Panda R.; Patro D. M. Mechanism of oxidation of mandelic acid by Fenton’s reagent. Inorg. Chem. 1984, 23 (24), 4119–4120. 10.1021/ic00192a060. [DOI] [Google Scholar]
- Walling C.; Camaioni D. M.; Kim S. S. Aromatic hydroxylation by peroxydisulfate. J. Am. Chem. Soc. 1978, 100 (15), 4814–4818. 10.1021/ja00483a030. [DOI] [Google Scholar]
- Hu D.-H.; Sun C.-Y.; Liu F.-H.; Qin C.; Wang X.-L.; Su Z.-M. A series of coordination complexes based on unsymmetrical multicarboxylate ligands: syntheses, structures and properties. CrystEngComm 2013, 15 (34), 6769–6775. 10.1039/c3ce40782a. [DOI] [Google Scholar]
- Xu Z.-X.; Xiao Y.; Kang Y.; Zhang L.; Zhang J. Homochiral Cluster-Organic Frameworks Constructed from Enantiopure Lactate Derivatives. Cryst. Growth Des. 2015, 15 (9), 4676–4686. 10.1021/acs.cgd.5b00958. [DOI] [Google Scholar]
- Tang Y.-Z.; Xiong J.-B.; Gao J.-X.; Tan Y.-H.; Xu Q.; Wen H.-R. Spontaneous Resolution, Asymmetric Catalysis, and Fluorescence Properties of Δ- and Λ-[Cu(Tzmp)]n Enantiomers from in Situ [2 + 3] Cycloaddition Synthesis. Inorg. Chem. 2015, 54 (11), 5462–5466. 10.1021/acs.inorgchem.5b00478. [DOI] [PubMed] [Google Scholar]
- Han M.-L.; Chang X.-H.; Feng X.; Ma L.-F.; Wang L.-Y. Temperature and pH driven self-assembly of Zn (II) coordination polymers: crystal structures, supramolecular isomerism, and photoluminescence. CrystEngComm 2014, 16 (9), 1687–1695. 10.1039/c3ce41968a. [DOI] [Google Scholar]
- Thompson L. K.; Mandal S. K.; Tandon S. S.; Bridson J. N.; Park M. K. Magnetostructural Correlations in Bis(μ2-phenoxide)-Bridged Macrocyclic Dinuclear Copper(II) Complexes. Influence of Electron-Withdrawing Substituents on Exchange Coupling. Inorg. Chem. 1996, 35, 3117–3125. 10.1021/ic9514197. [DOI] [PubMed] [Google Scholar]
- Ruiz E.; Alemany P.; Alvarez S.; Cano J. Toward the Prediction of Magnetic Coupling in Molecular Systems: Hydroxo-and Alkoxo-Bridged Cu(II) Binuclear Complexes. J. Am. Chem. Soc. 1997, 119, 1297–1303. 10.1021/ja961199b. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




