Xiaojiao
Hou
ab,
Wenxiu
He
a,
Xu
Zhai
a,
Bingbing
Chen
a,
Yuanlin
Fu
a,
Liying
Zhang
a,
Junyi
Chen
*c and
Yu
Fu
*a
aDepartment of Chemistry, College of Sciences, Northeastern University, Shenyang, 110819, P. R. China. E-mail: fuyu@mail.neu.edu.cn
bExperimental Center, LiaoDong University, Dandong, 118003, P. R. China
cEngineering Laboratory of Chemical Resources Utilization in South Xinjiang, College of Chemistry and Chemical Engineering, Tarim University, Xinjiang Uygur Autonomous Region, Alaer, 843300, P. R. China. E-mail: sln5xn@163.com
First published on 7th June 2024
Metal–organic frameworks (MOFs) permit significant structural diversity and functional design flexibility due to the various combinations of metal centers and different organic linkers. However, their combinations generally adhere to the classic hard and soft acid and base (HSAB) theory. This makes it impossible to directly synthesize desired MOFs with converse Lewis types of metal ions and ligands. Herein, we present a novel Cu(II) ion cleavage reduction strategy that circumvents the limitations of HSAB theory. We demonstrate this approach by utilizing a metastable CuH2DOBDC MOF, whose inherent structural instability facilitates structural transformations and phase transitions in response to external stimuli. By leveraging the unique structure of ascorbic acid to reduce Cu(II) ions and cleave the H2DOBDC2− linkers, we successfully fabricate a new Cu2H2DOBDC structure composed of a soft acid Cu(I) and a hard base (H4DOBDC). The resultant Cu2H2DOBDC integrates the characteristics of mesoporosity and hierarchical nanostructures, resulting in excellent mass transfer and abundant accessible Cu(I) active sites, which endows the material with outstanding catalytic activity in the azide–alkyne cycloaddition reaction. This is the first example of fabricating stable single-phase Cu(I)-carboxylic acid MOFs with nearly all Cu(I) ions, and this work offers a new perspective on the creation of a new category of MOFs with soft-acid and hard-base type coordination bonds.
Functionalization of metal–organic frameworks (MOFs) with Cu(I) sites holds significant promise for various applications including adsorption, separation, catalysis, and sensing.13–18 However, according to the above HSAB theory, Cu(I)-MOFs are usually fabricated using ligands with pyridyl N- or sulfhydryl S-centers. Although carboxylate-based MOFs constitute the overwhelming majority, Cu(I)-MOFs fabricated with carboxylate-based ligands are expected to exhibit thermal stability and have broad catalytic applications.19 Up to now, the direct synthesis of MOFs with Cu(I) metal ions and carboxylate-based ligands has not been reported. Most commonly, Cu(I)-carboxylate MOFs are synthesized using post-treatment approaches. For example, Chen et al. reported that HKUST-1, in the presence of hydroquinone, was reduced to mixed-valence Cu(I)Cu(II)-BTC and could remarkably boost the D2/H2 isotope separation.14 Sun et al. modulated some of the Cu(II) metal nodes in HKUST-1 to Cu(I) by reducing vapor treatment, thereby significantly enhancing the capture activity of aromatic sulfides.20 To our knowledge, only Qiao et al. have recently reported a “molecular scalpel” to chemically cleave conventional CuBDC with a controlled phase transition and synthesize a Cu2BDC MOF, where the Cu(I)/Cu(II) ratio is 1.21 Although the post-treatment approach is an effective strategy for creating new Cu(I)-carboxylate MOFs, achieving a near equal proportion of Cu(I) and Cu(II) sites in the MOF is far from satisfactory. Hence, it is highly desirable to develop an efficient and controllable approach for creating a new MOF with a high proportion of Cu(I) metal ions and carboxylate-based ligands, but this remained a challenge until now.
Herein, we have successfully fabricated a stable single-phase Cu(I)-carboxylate MOF (denoted as Cu2H2DOBDC), as shown in Fig. 1. We exploited the unique characteristics of metastable MOF structures, which are prone to undergo structural transformation under external stimulation, to achieve full phase transition. Starting with a reported metastable Cu(II)-MOF synthesized from Cu(OAc)2 and 2,5-dihydroxyterephthalic acid (H4DOBDC), we demonstrated that this metastable Cu(II)-MOF could be reduced to construct a novel Cu(I)-carboxylate MOF in the presence of ascorbic acid. A controlled phase transition was achieved by ascorbic acid accurately regulating the chemical state and coordination environment of Cu metal centers. X-ray photoelectron spectrum (XPS) analysis confirmed that nearly all Cu(II) ions were converted to Cu(I) ions. Cu2H2DOBDC, with its high chemical stability, mesoporosity, and uniformly abundant Cu(I) exposure sites, exhibits excellent catalytic capability as a recyclable heterogeneous catalyst for the azide–alkyne cycloaddition (AAC) reaction. To the best of our knowledge, this is the first attempt to fabricate a stable single-phase Cu(I)-carboxylate MOF with near-full Cu(I) ions. It is always fascinating to discover new MOF structures, and this work presents a promising step toward the synthesis of desired MOFs with converse Lewis types of metal ions and ligands, which is impossible to realize by conventional methods.
Fig. 2 (a) SEM images of CuH2DOBDC. (b) SEM images of Cu2H2DOBDC. (c) TEM images of Cu2H2DOBDC. (d) SAED pattern of the selected area in (c). (e) XRD patterns of CuH2DOBDC and Cu2H2DOBDC. |
The XRD pattern showed that the metastable Cu(II)-MOF achieved a full phase transition through the reduction of ascorbic acid (Fig. 2e). As the dosage of the reducing agent increased, the new characteristic peaks at 10.65°, 17.02°, 29.02°, and 42.20° gradually increased in intensity, while the metastable Cu(II)-MOF characteristic peaks gradually faded (Fig. S3†). By adjusting the addition of ascorbic acid, the controlled synthesis of intermediate samples was achieved. The coexistence of two phases is observed in the XRD patterns.
Thermogravimetric (TGA) analysis was employed to analyse the thermostability, as variations in coordination mode may present different mass loss curves. As shown in Fig. 3a, following treatment with ascorbic acid, the experimentally determined mass percentage of CuO in the final product was 44.63%. The final product exhibited a weight loss of about 6.28% in the temperature range of 30–220 °C, which can be attributed to the desorption of physically adsorbed water and solvent. Furthermore, a weight loss of about 49.09% was observed in the temperature range of 220–330 °C, which belonged to ligand decomposition. TGA analysis showed that the ratio of Cu to the H4DOBDC ligand significantly changed from 1:1 to 2:1 following ascorbic acid treatment, indicating a potential structural transformation in the material (see detailed analysis in the ESI†). Consequently, the as-synthesized Cu(I)-carboxylate MOF was denoted as Cu2H2DOBDC. Organic elemental analyzer (EA) results further support our conclusions (see detailed analysis in the ESI†). In addition, the dropping H2DOBDC2− ligand was detected by ultraviolet-visible (UV-vis) spectroscopy in a reducing solution (Fig. S4†). There is a gradual increase in the residual Cu mass percentage observed from CuH2DOBDC to Cu2H2DOBDC, a trend that is further supported by ICP-MS analysis (Table S1†). These results further strengthen the conclusion of TGA. The difference thermogravimetry (DTG) analysis suggested that the thermal stability of Cu2H2DOBDC is weakened due to the formation of Cu(I) species and the variation of coordination mode (Fig. S5†).20 Fourier-transform infrared spectroscopy (FTIR) was employed to investigate the coordination mode between Cu and the carboxylate linker.23 Conclusions about the carboxylate coordination mode are often drawn based on the splitting value between the asymmetric and symmetric stretches of the carboxylic vibration, where Δ = νas(COO−) − νs(COO−) (Fig. S6†).21,24 The metastable CuH2DOBDC MOF contains 2,5-dihydroxyterephthalic acid ligands with two coordination modes: carboxylate groups that connect to a single copper atom through monodentate coordination and those that bridge two copper atoms simultaneously. The FTIR spectra of CuH2DOBDC and Cu2H2DOBDC are shown in Fig. 3b. In CuH2DOBDC, the asymmetric stretch of the carboxylic vibration peak appeared at 1595 cm−1, while the symmetric stretch peak presented two distinct bands at 1425 and 1376 cm−1.25 Consequently, the observed splitting values of 170 cm−1 and 219 cm−1 in CuH2DOBDC correspond to bridging and monodentate characteristics, respectively. In addition, the spectra of CuH2DOBDC exhibited a peak at 1670 cm−1 attributed to the carbonyl stretching vibration of DMF. These results are consistent with the CuH2DOBDC crystal structure. In contrast, for Cu2H2DOBDC, the asymmetric stretch of the carboxylic vibration peak also appeared at 1595 cm−1, but the symmetric stretch peak was only observed at 1332 cm−1. Therefore, a splitting value of 263 cm−1 is consistent with the monodentate mode, indicating that in Cu2H2DOBDC, the Cu ions are coordinated to the carboxylate ligands in a monodentate mode. Notably, the absence of a peak corresponding to the carbonyl stretching vibration of DMF in the spectrum suggests that DMF does not participate in coordination within Cu2H2DOBDC. From the N2 adsorption–desorption measurement, Cu2H2DOBDC exhibited a type IV isotherm with a distinct hysteresis loop (Fig. 3c), indicating the existence of mesopores,26–28 which could be attributed to the ligand loss during the reduction process.20 The pore size distributions shown in Fig. S7† reveal that the size of the mesopores is centered at about 17 nm. The SEM images reveal that the surface of Cu2H2DOBDC exhibits a hierarchical nanostructure assembled by nanosheets, where the intersections and stacking of the surface hierarchical nanostructures also contribute to the formation of mesopores.29
X-ray photoelectron spectroscopy (XPS) was used for the evaluation of the chemical states and quantitative analysis of Cu species in the samples (Fig. 3d). The Cu 2p spectrum of CuH2DOBDC exclusively demonstrated the presence of Cu species in the form of Cu(II) ions. However, in Cu2H2DOBDC, new characteristic peaks appeared at 932.3 eV and 952.0 eV, which can be attributed to the Cu(I) species. Furthermore, a significant decrease in the intensities of the characteristic peaks associated with Cu(II) was observed, along with the near-absence of satellite peaks. This significant transformation indicates a reduction process, where nearly all Cu(II) ions in CuH2DOBDC were reduced to Cu(I) ions following treatment with ascorbic acid. However, two small residual peaks of Cu(II) at 934.8 eV and 954.59 eV were observed, which could be attributed to the unavoidable partial oxidation of the sample due to its exposure to air during the reaction or sample processing.30,31 The chemical state and coordination environment of the Cu atoms in CuH2DOBDC and Cu2H2DOBDC were further investigated by Cu K-edge X-ray absorption near-edge structure (XANES) spectroscopy and extended X-ray absorption fine structure (EXAFS) spectroscopy. The XANES spectra of the Cu K-edge for both CuH2DOBDC and Cu2H2DOBDC are shown in Fig. 3e. In the spectrum of CuH2DOBDC, distinct characteristic peaks were observed, including a pre-edge peak I at 8977 eV corresponding to the 1s → 3d transition, a white line peak II at 8996 eV, and a resonance peak III at approximately 9042 eV, which indicates that the oxidation state of Cu species is +2.32–34 However, in the XANES spectrum of Cu2H2DOBDC, the aforementioned characteristic peaks associated with Cu(II) are absent. Instead, a new peak at 8983 eV emerged, attributed to the 1s → 4p transition of Cu(I) (peak A), indicating that the predominant oxidation state of Cu species is Cu(I). Furthermore, the significant shift of the edge position towards the lower energy direction provides further evidence for the predominance of the Cu(I) state.33,35 This observation is consistent with the XPS results. In the R-space EXAFS spectra, both samples exhibited a dominant peak at 1.5 Å arising from Cu–O bonding (Fig. 3f). However, for Cu2H2DOBDC, the main peak at 1.5 Å possesses a lower intensity, which suggests a different Cu–O coordination environment. Additionally, when compared to CuH2DOBDC, the Cu–C peak also exhibited a decrease in intensity, which could be attributed to the partial removal of the carboxylate ligand.33 The EXAFS fitting curves showed that the Cu–O configuration predominantly exists in CuH2DOBDC and Cu2H2DOBDC, but with a change in the coordination number of O atoms around the Cu atoms (Fig. S8 and S9 and Table S2†).27 In the wavelet transform (WT) contour plots shown in Fig. 3g, Cu2H2DOBDC exhibited an intensity maximum of Cu–O at approximately 5.85 Å−1, which was very close to that of the reference Cu2O at 5.80 Å−1. Compared with the WT plot of CuH2DOBDC (6.0 Å−1), this provides further evidence that the formation of Cu(I) species leads to a change in the Cu–O coordination environment.36 Therefore, the above results indicate that the valence state and coordination environment of Cu atoms changed significantly after ascorbic acid reduction, resulting in the formation of a novel Cu(I)-carboxylate MOF.
Based on the aforementioned characteristics, we have formulated a speculation regarding the local structure of Cu2H2DOBDC. Current understanding of its structure reveals that the coordination mode of Cu(I) and the ligand is in the monodentate form, with DMF not participating in the coordination. Furthermore, the molar ratio of Cu(I) to the ligand is 2:1. Consequently, we speculate that the local structure comprises a symmetric unit with two Cu(I) ions (Cu1 and Cu2) linked by an H2DOBDC2− linker (Fig. S10†). Such a possible formation process is shown in Fig. S11.†
To investigate whether the catalysis was a heterogeneous reaction, the catalysts were separated from the reaction systems by hot filtration after 0.5 h. However, even after continuing the reaction with filtrates, no significant increase in yields was observed (Fig. 4b). The results indicate that Cu2H2DOBDC is an excellent heterogeneous catalyst for the AAC reaction. The recyclability of Cu2H2DOBDC as a heterogeneous catalyst was investigated. Impressively, Cu2H2DOBDC retained satisfactory catalytic activity and could be reused for five consecutive cycles without obvious loss in the catalytic activity, as shown in Fig. 4c. In addition, we further verified the structural stability of the Cu2H2DOBDC catalyst after cycling (Fig. S12†). The SEM results show that the sample morphology remains and the surface still retains hierarchical nanostructures. XRD analysis confirms that the catalyst maintains its crystallinity and structural integrity after cycling. FTIR spectra show no new peaks or shifts in existing peaks, indicating that there is no change in chemical bonds or functional groups. XPS analysis confirms that the predominant oxidation state of Cu species in Cu2H2DOBDC remains as Cu(I) even after undergoing five cycles of reactions. In summary, these results demonstrate the excellent structural and chemical stability of Cu2H2DOBDC, ensuring its potential for long-term and reliable catalytic applications.
To further validate the catalytic generality of Cu2H2DOBDC, we employed various terminal alkynes with different functional groups and azides, including methoxy, nitro, heteroaryl, halogen, or methyl moieties, to investigate the catalytic performance under the optimized conditions. As summarized in Table 1, all the selected substrates could efficiently undergo the Cu2H2DOBDC-catalyzed “click” reaction to deliver their corresponding 1,4-disubstituted 1,2,3-triazoles in high yields (Fig. S13–S28†). Cu2H2DOBDC exhibits superior catalytic performance and regioselectivity, which may be attributed to the high density of uniformly distributed Cu(I) active sites and the hierarchical MOF structure, which enhances accessibility to the active sites and facilitates fast diffusion of the reactants.
Entry | Alkyne | Azide | Product | Yieldb (%) |
---|---|---|---|---|
a Reaction conditions: catalyst (10 mg), azide (1 mmol), alkyne (2 mmol), amyl acetate (1 mmol), acetonitrile (4 ml) and 80 °C, 2.5 h. b Isolated yields were calculated by GC, and amyl acetate was employed as the internal standard. | ||||
1 | 99 | |||
2 | 94 | |||
3 | 99 | |||
4 | 99 | |||
5 | 95 | |||
6 | 99 | |||
7 | 96 | |||
8 | 97 |
Considering previous reports on the CuAAC reaction,39–42 a possible mechanism, as shown in Fig. S29,† has been proposed. Initially, the Cu(I) sites in Cu2H2DOBDC interact with the alkyne to form the Cu(I)-acetylide intermediate A. Subsequently, azides react with the Cu(I)-acetylide intermediate A to generate intermediate B. Then, the nucleophilic nature of the terminal N3 in intermediate B makes it attack the alkyne to form metallacycle C, which further shrinks to yield the Cu–triazole intermediate D. Finally, the protonation of intermediate D yields the target product 1,4-disubstituted 1,2,3-triazole, accompanied by the recovery of the catalyst.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qi00922c |
This journal is © the Partner Organisations 2024 |