Kai
Zhou
ab,
Yuanyuan
Zhang
ab,
Mingjie
Liu
ab,
Zhenghua
Zhao
ab,
Xiang
Liu
*b,
Zongbi
Bao
ab,
Qiwei
Yang
ab,
Qilong
Ren
ab and
Zhiguo
Zhang
*ab
aKey Laboratory of Biomass Chemical Engineering of Ministry of Education, College of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310058, P. R. China. E-mail: zhiguo.zhang@zju.edu.cn
bInstitute of Zhejiang University–Quzhou, Quzhou 324000, P. R. China. E-mail: liu-xiang@zju.edu.cn
First published on 2nd February 2024
Highly dispersed iron oxide has been rapidly and precisely loaded onto the Zr6 nodes of a stable metal–organic framework (MOF), UiO-66, via an atomic layer deposition (ALD) process to form novel Fe–O–Zr metal cluster sites. By varying the number of ALD cycles, three Fe-decorated UiO-66 materials (denoted as Fe@UiO-66-xc, x = 1, 2, 3) were synthesized. A series of photoelectrochemical measurements, including UV-vis diffuse reflectance spectroscopy, state photoluminescence spectroscopy, and transient photocurrent response, indicate that Fe@UiO-66-1c prepared by a single deposition cycle exhibits improved visible-light absorption ability and enhanced photo-generated charge carrier separation efficiency due to metal-to-metal charge transfer. Moreover, Fe@UiO-66-1c shows excellent photocatalytic activity for aerobic oxidation of N-aryl tetrahydroisoquinolines. This material also exhibits good stability and is capable of cycling the reaction six times and maintaining the crystal structure with a low leaching rate of iron ions. The study explores the application of the atomic layer deposition process in the preparation of advanced photocatalytic materials.
Metal–organic frameworks (MOFs) are crystalline materials consisting of metallic nodes and organic linkers, which have recently emerged as a promising class of heterogeneous photocatalysts for visible-light-induced oxidations.11–15 The organic bridging ligands of MOFs can serve as antennas to harvest light and activate metal nodes, illustrating semiconductor-like behaviour and demonstrating great potential in activating dioxygen and promoting the oxidative process.16–18 However, most functional MOFs still suffer from low efficiency of light absorption, and charge separation and transfer, restricting their further development. A variety of strategies, such as metal doping, ligand exchange, and cavity encapsulation have been explored to circumvent the challenge of charge transfer and hence to improve the photocatalytic performance.19–23 Among them, introducing active metals as photoactive species seems to be the optimal candidate to modify MOFs.24–26 In particular, the introduction of metal clusters or atom-sized metals has been shown to not only maximize the utilization of metals, but also promote the charge separation and transfer for MOFs. For example, many types of metal single atoms (e.g., Cu, Fe, etc.) have been bonded into MOFs by solvothermal or microwave-assisted post-synthetic modification methods, which act as effective electron acceptors for spatial separation and transfer of charge carriers via the ligand–metal charge transfer (LMCT), and metal-to-metal charge transfer (MMCT) occurs upon the connection of two metal centers with different valence states via an oxygen bridge, leading to enhanced photocatalytic performance.27–29 Although there are great achievements on active metal modification of MOFs, more techniques for efficient deposition of the active metals onto MOFs are still expected to be established and utilized.
Atomic layer deposition (ALD) is a chemical vapor deposition technique that uses sequential pulses of precursor gases to react with the surface in a self-limiting manner. ALD is a unique method that offers precise control of atomic-scale thickness, excellent three-dimensional conformality, and large-area uniformity, making it increasingly popular for the direct synthesis and post-synthetic modification of advanced catalysts.30–36 For instance, controlling the precursors and adjusting the number of Fe ALD cycles enables the selective deposition of either Fe single atoms or an ultrathin Fe2O3 film onto the surface of TiO2.37,38 Omar Farha's group has loaded a variety of metals on Zr6-based MOF NU-1000 using ALD and found that the ALD reaction occurs at inorganic nodes.39–41 Herein, we first applied the ALD process to incorporate iron oxide onto the Zr6 nodes of UiO-66. Through controlling the vapor deposition cycle program, we successfully prepared a series of Fe@UiO-66 photocatalysts with different iron loadings. Fe@UiO-66 samples inherit the high stability of UiO-66 and maintain the morphology and porosity. Meanwhile, the Fe–O–Zr sites constructed on the metal nodes extend the visible light absorption due to MMCT effects. Among this series of photocatalysts, the Fe@UiO-66-1c material produced by a single deposition cycle showed the best photocatalytic activity and could achieve aerobic oxidation of N-aryl tetrahydroisoquinolines with excellent yield. Photoelectrochemical measurements indicated that Fe@UiO-66-1c possesses better separation efficiency of photogenerated electron–hole pairs and could activate oxygen efficiently to generate superoxide radicals. In addition, benefiting from the highly dispersed isolated iron sites and the stable Fe–O–Zr bonds, Fe@UiO-66-1c could be reused six times with a high yield of over 80%, and exhibited a low ion leaching rate in the reaction cycles.
Photoelectrochemical measurements were obtained on a CHI 660E electrochemical workstation (Chenhua Instrument, Shanghai, China). The working electrodes were prepared by dropping the sample suspension (10 μL), which was obtained from the mixture of the as-synthesized samples (5 mg), 30 μL Nafion, and 1 mL ethanol under sonication for 30 min, onto the surface of a glassy carbon electrode. After drying at room temperature, electrochemical measurements were performed in a standard three-electrode system with the photocatalyst-coated glassy carbon electrode as the working electrode, Pt plate as the counter electrode, and Ag/AgCl as a reference electrode. A 300 W xenon lamp (HDL-II, Bobei Light Co. Ltd) was used as the light source. A 0.2 M Na2SO4 solution was used as the electrolyte. The photoresponsive signals of the samples were measured at 0.3 V. And the electrochemical impedance spectroscopy (EIS) was performed in the frequency range from 10−1 to 105 Hz with a bias potential of 0.2 V.
Scanning electron microscopy (SEM) images (Fig. 1b and S1†) showed that Fe@UiO-66 samples maintained the morphology of UiO-66. The element mapping images indicate that the Fe element is uniformly distributed on the surface of Fe@UiO-66 samples (Fig. 1c and S2†). Further, the cross-sectional EDS mapping images of Fe@UiO-66-1c indicated a uniform distribution of Fe inside the particles (Fig. S3†). And there were no nanoparticles or thin films observed in the HRTEM images (Fig. S4†). Powder X-ray diffraction (XRD) patterns indicate that the crystal structure of UiO-66 is well maintained after three cycles of the ALD process and no obvious XRD peaks of Fe sites are observed, owing to their high dispersion (Fig. 1d). Thermogravimetric analysis (TGA) showed that the thermal stability of these materials can still be maintained at around 450 °C (Fig. S5†). The surface area was analyzed by the Brunauer–Emmett–Teller (BET) method, which was measured as 1128.5 m2 g−1 for UiO-66, 1082.7 m2 g−1 for Fe@UiO-66-1c, 1072.3 m2 g−1 for Fe@UiO-66-2c, and 849.9 m2 g−1 for Fe@UiO-66-3c (Fig. 1e). This result is consistent with observations in previous studies.43 Furthermore, a significant reduction of surface area after the third deposition cycle was observed, implying the multi-layer deposition of Fe on the Zr6 nodes. In addition, the pore size distributions of the samples, analyzed by nonlinear density functional theory (NL-DFT), remain unchanged. A slight decrease in pore volume is attributed to the occupation of iron oxide (Fig. S6†).
X-ray photoelectron spectroscopy (XPS, Fig. 2a) confirms the coexistence of building elements. The Fe 2p3/2 binding energy of 710.8 eV demonstrates the Fe(III) state (Fig. 2c), and the satellite features of Fe 2p (2p3/2, 715.6 eV; 2p1/2, 729.6 eV) verify the Fe(II) state. Upon introducing Fe species into UiO-66, the peak of Zr 3d5/2 shifts to lower binding energy from 182.95 to 182.81 eV, which could be due to the electron-donating effect of the iron oxide sites (Fig. S7†). Electron paramagnetic resonance (EPR) was used to analyze the dispersion of iron sites.44,45 The EPR peak at g = 4.3 is assigned to high-spin Fe(III), which mainly exists in highly isolated Fe(III) in tetrahedral and distorted tetrahedral coordinations. The signal at g = 2.0 is ascribed to isolated Fe(III) in a high-symmetry octahedral coordination or FexOy oligomers (Fig. 2d). The iron content of the samples was measured by inductively coupled plasma optical emission spectrometry (ICP-OES). The results showed that the iron loading was 1.60%, 3.24%, and 4.94% wt for the samples with 1, 2, and 3 cycles of atomic layer deposition (corresponding to ∼0.58, 1.17, and 1.79 iron atom per Zr6 node), respectively, suggesting that the amount of iron oxide could be well controlled by varying the number of ALD cycles.
![]() | ||
Fig. 2 XPS spectra of (a) a survey scan of Fe@UiO-66-1c, (b) Zr 3d and (c) Fe 2p. (d) EPR spectra of Fe@UiO-66-1c at room temperature. |
Entry | Catalyst | Yieldb [%] |
---|---|---|
a Reaction conditions: 1a (0.1 mmol), catalyst (5 mg), DBN (0.15 mmol), O2 (1 bar), CH3CN (2 mL), white LEDs (10 W), 30 °C, 8 h. b Yields were calculated by 1H NMR with 1,3,5-trimethoxybenzene as the internal standard. | ||
1 | Trace | |
2 | UiO-66 | 35.3 |
3 | Fe@UiO-66-1c | 87.5 |
4 | Fe@UiO-66-2c | 60.3 |
5 | Fe@UiO-66-3c | 48.9 |
6 | Fe2O3 | 7.8 |
Next, we focused on the optimization of reaction parameters, including solvents, base additives, and reaction atmosphere (Table 2). Alcoholic solvents showed extremely poor selectivity towards product 2a. However, changing to alternative polar non-protonic solvents, such as DMF or 1,4-dioxane, could provide competitive yields (entries 1–6, Table 2). Additionally, we examined the influence of other bases, including DBU, DIPEA, and Cs2CO3, but none of them led to an improvement in the yield of 2a (entries 7–10, Table 2). It is worth noting that the reaction proceeded smoothly under an air atmosphere, giving a 75% yield of the desired 2a product (entry 11).
Entry | Base | Solvent | Yieldb [%] |
---|---|---|---|
a Reaction conditions: 1a (0.1 mmol), Fe@UiO-66-1c (5 mg), base (1.5 equiv.), solvent (2 mL), O2 (1 bar), white LEDs (10 W). b Yields were calculated by HPLC with naphthalene as the internal standard. c 1 bar of air. d Under N2 atmosphere. | |||
1 | DBN | MeOH | Trace |
2 | DBN | EtOH | Trace |
3 | DBN | CF3CH2OH | Trace |
4 | DBN | CHCl3 | 6.1 |
5 | DBN | DMF | 80 |
6 | DBN | 1,4-Dioxane | 83 |
7 | DBN | CH3CN | 87 |
8 | DBU | CH3CN | 86 |
9 | DIPEA | CH3CN | Trace |
10 | Cs2CO3 | CH3CN | 37 |
11 | DBN | CH3CN | 75c |
12 | DBN | CH3CN | Traced |
With the optimized reaction conditions in hand, the substrate scope for the aerobic oxidation of amines was next examined (Table 3). Both electron-donating and electron-withdrawing groups on the N-phenyl ring were well-tolerated, affording corresponding products 2a–2f in good to excellent yields (77–94%). In general, electron-donating substituents exhibited superior activity than electron-withdrawing ones. Furthermore, 6,7-dimethoxyl N-aryl tetrahydroisoquinolines are also synthesized efficiently (2g–2h). In contrast, the activation of the α-position C–H of N-alkyl tetrahydroquinoline proved to be more challenging, resulting in a lower yield of 43% (2i). These catalytic results indicated that Fe@UiO-66-1c had comparable performance to the reported organic small molecule photocatalysts, such as rose bengal and eosin Y, in the photocatalytic oxidation of N-aryl tetrahydroisoquinolines (Table S1†).
a Reaction conditions: 1 (0.1 mmol), Fe@UiO-66-1c (5 mg), DBN (0.15 mmol), O2 (1 bar), CH3CN (2 mL), white LEDs (10 W), 30 °C. Yield of isolated product. | ||
---|---|---|
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
Subsequently, we delved into the mechanism of this photocatalytic aerobic oxidation process. UV-vis diffuse-reflectance spectroscopy (UV-DRS) was utilized to verify the optical features of the Fe@UiO-66 samples. A promising photocatalyst is expected to have a broad absorption in the visible light region. As revealed in the obtained spectra (Fig. 3a), UiO-66 showed initial UV absorption at around 330 nm, attributed to the ligand-based absorption affected by the nearby metal clusters. In contrast, Fe@UiO-66 samples exhibited a wide range of visible light absorption, extending to 700 nm. Notably, Fe@UiO-66-1c exhibited a new absorption peak at 646 nm, liking arising from the MMCT process between Fe and Zr sites. The optical band gap of the as-prepared Fe@UiO-66 samples was calculated by Tauc plots based on UV-DRS curves (Fig. 3b). The estimated band gaps were found to be 3.97 and 3.72 eV for UiO-66 and Fe@UiO-66-1c, respectively. The band structure was determined over Mott–Schottky analysis to evaluate its oxidative potential. The positive slopes of the Mott–Schottky plots revealed n-type behavior for UiO-66 and Fe@UiO-66-1c (Fig. S8† and 3c). Typically, in n-type semiconductors, the conduction band (CB) is more negative by about 0.10 V than the flat band potential (Vfb). Therefore, the CB values of UiO-66 and Fe@UiO-66-1c were determined from the intersection with values of −1.02 and −0.86 V vs. NHE, respectively, which are more negative than the reduction potential of O2 to O2˙− (−0.33 V vs. NHE).47 Based on the estimated bandgap values, the valence band (VB) positions were then estimated as 2.95 V (UiO-66) and 2.86 V (Fe@UiO-66-1c) vs. NHE. Moreover, a photocurrent test indicated that Fe@UiO-66-1c had a stronger photocurrent response than UiO-66, suggesting its superior charge transfer efficiency (Fig. 3d). This result was further verified by the electrochemical impedance spectroscopy (EIS) study, which measured the interfacial charge-transfer resistance of the materials. Fe@UiO-66-1c exhibited a smaller radius, indicating lower charge transfer resistance than UiO-66 (Fig. S9a†). Furthermore, the reduced intensity of Fe@UiO-66-1c in photoluminescence (PL) emission spectroscopy corresponded to the higher separation rate of photoinduced charges (Fig. S9b†).
To gain deeper insights into the main active species in the reaction, several radical quenching experiments were carried out under the standard reaction conditions.48,49 The presence of Na2C2O4 (a scavenger for h+) and isopropanol (a scavenger for OH·) did not affect the occurrence of this reaction. However, when 1,4-benzoquinone (BQ, 3 equiv.) was added, the oxidation reaction was markedly inhibited, providing clear evidence for the presence of the superoxide radical (O2˙−).50 Moreover, we employed 5,5-dimethyl-1-pyrroline N-oxide (DMPO) as the radical trapping agent. Under the visible-light irradiation, in the presence of Fe@UiO-66-1c in MeCN and an air atmosphere, in situ EPR spectra revealed a characteristic signal of superoxide radical, confirming that O2˙− was indeed the main active species (Fig. 4b). These findings aligned with the results obtained from the electrochemical analysis of the material potentials. Based on all of the above experimental results, we propose a plausible mechanism in Fig. 4c, where O2˙− serves as reactive oxygen species.
As a heterogeneous photocatalyst, recyclability and reusability are prominent features in industrial and practical processes. After the photocatalytic reactions, Fe@UiO-66-1c was recovered by centrifugation and reused for six consecutive runs under identical conditions. To our delight, the yield of 2a remained almost unchanged throughout these cycles (Fig. S10a†). In addition, the crystal structure of Fe@UiO-66-1c remained intact, as evidenced by the PXRD patterns (Fig. S10b†) and the SEM image (Fig. S11†). Additionally, there was only a slight decrease in the iron loading, from 1.60 to 1.42 wt%, after six runs, indicating the stable anchoring of iron oxide to the Zr6 nodes.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cy01768k |
This journal is © The Royal Society of Chemistry 2024 |