Bo
Dai‡
a,
Zichuang
Li‡
a,
Miao
Xu‡
b,
Jiang
Li
c,
Yangfan
Lu
d,
Jiantao
Zai
a,
Liuyin
Fan
e,
Sang-Won
Park
f,
Masato
Sasase
c,
Masaaki
Kitano
*c,
Hideo
Hosono
*c,
Xin-Hao
Li
a,
Tian-Nan
Ye
*a and
Jie-Sheng
Chen
*a
aFrontiers Science Center for Transformative Molecules, School of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China. E-mail: ytn2011@sjtu.edu.cn; chemcj@sjtu.edu.cn
bState Key Laboratory of Space Power-Sources Technology, Shanghai Institute of Space Power-Sources, Shanghai 200245, China
cMDX Research Center for Element Strategy, International Research Frontiers Initiative, Tokyo Institute of Technology, 4259 Nagatsuta, Midori-ku, Yokohama 226-8503, Japan. E-mail: kitano.m.aa@m.titech.ac.jp; hosono@mces.titech.ac.jp
dNational Engineering Research Center for Magnesium Alloys, College of Materials Science and Engineering, Chongqing University, Chongqing 400030, China
eStudent Innovation Center, Shanghai Jiao Tong University, Shanghai 200240, China
fDepartment of Chemical and Materials Engineering, University of Suwon, Hwaseong, Gyeonggi 18323, Republic of Korea
First published on 27th January 2023
The electronic structures of active sites fundamentally determine the catalytic performance in chemical reactions and are also crucial for obtaining a detailed understanding of charge transport and reaction mechanisms. In this study, the regulation of the electronic structure of active metal Pd can be achieved through a multi-step electron transfer process formed by a synergy of [Ca24Al28O64]4+(e−)4 (C12A7:e−) electride and conductive graphene (Gr). The composite catalytic system (Pd/Gr/C12A7:e−) significantly facilitates the transfer of electrons from electron-rich Pd active sites to aryl halides in Suzuki-coupling reactions, which enables superior catalytic performance with TOFs above 20 times higher than well-studied negatively charged Pd catalysts. No catalytic degradation was observed even after impregnating the catalyst in water because of the well-protected C12A7:e− electride by Gr. The present efficient catalyst can further trigger various carbon–carbon cross-coupling reactions with high activities. These results provide significant advantages for expanding the potential applications of electride materials, thereby allowing precise control of the electronic structure of the active sites and aiding in tuning the reaction conditions using a simple method.
10th anniversary statementWe would like to congratulate on the upcoming 10th anniversary celebration of the Journal of Materials Chemistry A. In the discipline of materials and chemistry, the Journal of Materials Chemistry A has a significant influence that cannot be ignored. Many excellent opinions and articles have been published in the journal, playing a decisive role in the profession. As materials chemistry researchers, we feel honoured to publish articles in the Journal of Materials Chemistry A. In the past ten years, measurable progress on the research of materials chemistry in which the future of energy and sustainability is embodied. Electride materials are regarded as one category of promising emergent materials that have more advantages than traditional catalytic materials because of their unique electronic properties that can accelerate or limit specific chemical reactions. At present, there are few cases in which electride materials are used in industrial catalysis because of their chemical instability. Our research goal is to solve this problem and to develop stable electride composite materials that can be successfully put into use in the near future through continuous technological breakthroughs. |
[Ca24Al28O64]4+(e−)4 (C12A7:e−) electride materials, which accommodate anionic electrons in their periodically distributed lattice, have been attracting considerable attention because their low work function feature enables strong electron donation ability.24–30 One of their most successful applications is in ammonia synthesis, in which an electron transfer from C12A7:e− electride to the antibonding orbitals of N2 facilitates N2 dissociation, and the overall activation energy of the reaction is suppressed.31–34 However, C12A7:e− electride is typically prepared by a solid-state method, resulting in a small surface area of a few square meters per gram. Such a low surface area leads to a relatively low dispersion of the loaded active metals, and thus, only the exposure of limited active sites. Moreover, the moisture sensitivity of C12A7:e− severely restricts further applications in aqueous reactions.35,36 Therefore, it is highly desired to achieve C12A7:e− electride with high stability towards moisture while retaining its intrinsic electronic structure. Graphene (Gr)-based materials have been well-studied as water or gas filters,37,38 and thus, we are inspired to combine C12A7:e− electride with Gr to form an encapsulated nano-structure.
Herein, we report the successful preparation of the Pd/Gr/C12A7:e− catalyst and its high catalytic activity towards cross-coupling reactions. In this composite material, the C12A7:e− electride surface was coated with Gr layers, which are highly dispersed with nanometric Pd clusters, to achieve high chemical stability towards moisture. Negatively charged Pd was achieved through a multistep electron transfer process, in which electrons can be donated from interior C12A7:e−via highly conductive Gr to the external Pd active species owing to the work function gap between C12A7:e− (ΦWF = 2.4 eV) and Pd (ΦWF = 5.1 eV).39,40 The electron-rich Pd active sites enable the aryl halide activation, thereby reducing the apparent activation energy of the Suzuki-coupling reaction by ca. 30%; therefore, the reaction could proceed under mild conditions. Impressively, Pd/Gr/C12A7:e− showed no degradation in catalytic activity even after the impregnation of water, which demonstrates the chemical protection of Gr. These results demonstrate the effectiveness of combining an electride and traditional active metal, and they reveal a simple and effective approach to improving catalytic activity and stability.
When electronic conductive materials are in contact, a charge transfer is allowed to balance the chemical potentials of each material. For the Gr/C12A7:e− heterostructure, an inorganic C12A7:e− electride with a low work function of 2.4 eV could be regarded as an electron donor (Fig. S3†) for Gr with a higher work function of 5.0 eV, resulting in an electron transfer from C12A7:e− to Gr (Fig. 1c),39,40 which could be confirmed by Raman spectroscopy (Fig. 1d). In the Raman spectra, the invariant D-band of Gr with and without C12A7:e− suggests a physical interaction between C12A7:e− and Gr. Comparatively, a blue shift of the G-band appeared after the addition of C12A7:e− electride. Regarding the results on carrier-injected carbon nanotubes,41,42 this blue shift should be attributed to the lattice distortion of Gr evoked by the accepted electrons from C12A7:e− and the Fermi level shift induced by electron–phonon coupling. Moreover, the C 1s peak of X-ray photoelectron spectroscopy (XPS) of Pd/Gr/C12A7:e− and Pd/Gr further show that the addition of C12A7:e− up-shifted the binding energies of sp3-hybridized carbon atoms, such as CO, COOH and CO32− over Gr layers, which means that electron transfer is mainly performed on the defect sites and functional groups of Gr (Fig. S4 and Table S1†). The CC peak remains unchanged, indicating no chemisorption process or modification of the graphitic structure. In combination with the Raman shift and up-shifted C 1s peaks in the XPS spectra, it is reasonable to conclude that the electrons can be effectively transferred to the Gr surface.
Given that Gr (ΦWF = 5.0 eV) and Pd (ΦWF = 5.1 eV) exhibit comparable work functions, the electron could easily flow between the Gr and loaded Pd cluster. Negatively charged Pd species are thus achieved, which was confirmed by X-ray absorption near-edge structure (XANES) measurement. The absorption edge for Pd species in Pd/Gr/C12A7:e− was located at a lower energy relative to that of Pd foil (Fig. 1e), implying negatively charged Pd species. However, no energy shift was detected for Pd in Pd/Gr without the presence of C12A7:e− electride. It is noteworthy that Pd on C12A7:e− was also in metallic form, which could be attributed to the uneven charge distribution at large Pd particles on C12A7:e− without the dispersion of Gr (Fig. S5†). As the surface reaction of the catalytic process, we then investigated the surface properties of the catalyst using XPS. Compared with the zero valence state of Pd (335.1 eV), the obviously lower energy shift of the Pd 3d XPS peaks in Pd/Gr/C12A7:e− implies electron-rich Pd species (334.7 eV) (Fig. 1f). These observations combined with the XANES analysis results indicate the key role of the C12A7:e− electride as an electron donor in the modification of the electron density of highly dispersed Pd clusters on Gr.
We initially tested the possibility of using the negatively charged Pd active sites of the electride composite material in organic catalysis to promote the Suzuki coupling of iodobenzene with phenylboronic acid as a model reaction. In Fig. 2a, Pd/Gr/C12A7:e− showed high catalytic performance with a reaction rate of 60.0 mmol g−1 h−1 at room temperature, and its turnover frequency (TOF) value was estimated to be 1413.3 h−1 based on the total amount of Pd, which is in the orders of magnitude larger than the Pd/Gr (5 wt% Pd, 42.4 h−1), Pd/C12A7:e− (1 wt% Pd, 71.0 h−1) and Pd/C12A7:O2− (1 wt% Pd, 43.5 h−1) reference catalysts (Fig. 2a, Table S2†). Impressively, in terms of TOFs, Pd/Gr/C12A7:e− even outperforms our previously reported negatively charged Pd catalysts (Pd/ZrC, ZrPd3 and Y3Pd2) and other benchmarked commercial heterogeneous Pd catalysts (Lindlar catalyst from TCI, 5 wt% Pd; Pd/Al2O3 from Sigma-Aldrich, 5 wt% Pd; Pd/C from Sigma-Aldrich, 10 wt% Pd) 20-fold under the same reaction conditions (Table S3†). The lower reaction rates of Pd/C12A7:e− (1 wt% Pd) and Pd/C12A7:O2− (1 wt% Pd) are attributed to the limited number of exposed active sites owing to the large Pd particle size (Fig. S5 and S6†). It is noteworthy to mention that the activity of Pd/C12A7:e− (0.2 wt% Pd, 1335.6 h−1) is comparable to that of Pd/Gr/C12A7:e− when the Pd particle size was reduced (Fig. S7, Table S2†), indicating an electron donation effect on nanosized Pd particles by C12A7:e−, which could also be confirmed by the same energy location of Pd 3d XPS spectra (Fig. S7b†). Additionally, with a similar Pd particle size (Fig. S8†), the lower activity of the Pd/Gr sample is mainly attributed to the absence of C12A7:e−.
In addition to the high catalytic performance for the Suzuki coupling reactions, the stable recyclability of the Pd/Gr/C12A7:e− catalyst should be noted. Pd/Gr/C12A7:e− catalyst can be recycled at least 8 times under both low (Fig. 2b) and high conversion (Fig. S9†) levels. After the reaction, the crystal structure remained largely unchanged (Fig. S10†). HAADF-STEM observation clearly revealed that the Pd particle size and morphology remained unchanged after recycling (Fig. S11†). XPS measurements showed that the Pd species remained negatively charged, i.e., the surface electronic structure is highly stable (Fig. 2d). The hot filtration experiment indicated that the reaction proceeds only in the presence of Pd/Gr/C12A7:e−, and no more coupling products could be produced after the removal of the Pd/Gr/C12A7:e− catalyst (Fig. S12†). Inductively coupled plasma atomic emission spectroscopy (ICP-AES) measurements also showed that the Pd species in the filtrate was below the detection limit (0.007 ppm). These results for the used Pd/Gr/C12A7:e− catalyst demonstrated the robustness of the Pd active sites and confirmed the excellent stability of this catalyst.
Moisture sensitivity is one of the most serious challenges to the practical application of C12A7:e− electride, which has always undergone a transformation from C12A7:e− to Ca and Al hydrates with the release of the anionic electron, which significantly reduces its carrier density.43 However, Pd/Gr/C12A7:e− was found to be robust against water. A Suzuki coupling reaction using water-impregnated Pd/Gr/C12A7:e− powder as a catalyst was conducted without changing to any other conditions. The obtained coupling rate was almost identical to that of the fresh catalyst, and the color of the samples did not change (Fig. 2c). Decomposition of the catalyst, such as by the generation of Ca and Al hydrates, was not identified from X-ray diffraction (XRD) measurements after the impregnation in water (Fig. S10†). Based on XPS measurements, the partially negatively charged Pd was unaffected by the water impregnation, which demonstrates the robustness of the Pd/Gr/C12A7:e− catalyst (Fig. 2d). In contrast, the catalytic activity of the Pd/C12A7:e− catalyst was obviously degraded after impregnation in water, and the colour of the sample changed from black to light grey, associated with the transformation to the related Ca and Al hydrates (Fig. 2c). This degradation can be attributed to the absence of a protective layer of Gr, which results in the progressive release of anionic electrons from C12A7:e−.
To further investigate the electron donation effect of C12A7:e−, the controlled electron concentration (Ne) of C12A7:e− was investigated. C12A7:e− with different Ne were obtained by treatment with Ti metal at different temperatures (Fig. 3). Fig. 3a shows the UV-vis absorption spectra for the synthesized powders of C12A7:e− or C12A7:O2− with various Ne. There is no adsorption peak of the C12A7:O2− sample (orange line) in the visible region, and the corresponding 3.5 eV absorption edge is attributed to the excitation between the energy level of encaged O2− ions and the cage conduction band (CCB).44 Moreover, a broad absorption peak at around 2.3–2.7 eV (2.3 eV, 2.5 eV and 2.7 eV for green, blue and red lines, respectively) appeared in C12A7:e−, which can be attributed to an intra-cage transition of electrons confined in the cages. The other absorption peak below 2 eV is ascribed to an inter-cage transition as charge transfer occurs from an electron-trapped cage to a neighboring vacant cage.45 Therefore, it can be deduced that O2− ions in the cages of C12A7:O2− are substituted by electrons, thus forming C12A7:e−. Based on our previous studies, the adsorption peak position (Esp) and electron concentration (Ne) are related to the following equation: and one typical case is that represents Ne ≈ 1.0 × 1018 cm−3.46 Thus, the electron concentrations of the above C12A7:e− samples can be calculated to be 1.2 × 1020, 1.1 × 1021, and 2.2 × 1021 cm−3, respectively (Fig. 3b). Accordingly, the sample color changed from white to green to black with increased Ne (Fig. 3b).
Fig. 3c shows the electronic structures of C12A7:e− or C12A7:O2− with various ranges of Ne. For C12A7:O2− with Ne = 0 cm−3, the 2p orbitals of the framework O2− ions and 4s orbitals of the framework Ca2+ ions contributed to the valence band (VB) and conduction band (CB), respectively. In contrast to framework O2− ions, the energy level of the encaged O2− ions is located slightly above the top of the VB of the cage framework. In addition, the electron tunneling among the three-dimensionally connected cages enables the formation of the CCB, which is located at ∼1.0 eV below the bottom of CB.47 In the case of Ne < 1.0 × 1021 cm−3, the cage-encaged electrons with low density formed an F+-like center, and the energy level is located at 0.4 eV below the CCB. When the Ne is higher than 1.0 × 1021 cm−3, the cage-trapped electrons occupy the CCB, raising the Fermi level of C12A7:e− to 0.5 eV above the CCB minimum. Therefore, an intrinsic low work function property was achieved over such a unique electron-occupied CCB electronic structure. The high electron concentration leading to a low work function property enables high electron donation power for C12A7:e−. The catalytic activity of Pd/Gr/C12A7:e− provides information on the critical electron concentration (Ne) of C12A7:e− as a support. Impressively, the reaction rates were observed to increase monotonically with the Ne of C12A7:e− (Fig. 3b). This observation suggests that the catalytic activity of this reaction is directly proportional to the electron donation ability of the support. Accordingly, the measured apparent activation energy (Ea) decreases from 83.0 kJ mol−1 to 58.7 kJ mol−1 with increasing Ne (Fig. S13†), which should be related to the electron transfer from C12A7:e−via surface Pd active sites to aryl halide substrates.
Application of the catalyst was then extended to various aryl halides and boronic acids to verify the general activity of the Pd/Gr/C12A7:e− catalyst towards Suzuki coupling reactions. Table 1, (a) shows that all the tested aryl halides with a range of functional groups could be converted to the corresponding coupled products in high yields, regardless of whether the functional groups were electron-donating or electron-withdrawing. Next, different types of boronic acids were further exploited as coupling partners (Table 1, (b)). The results confirmed the general applicability of the Pd/Gr/C12A7:e− catalyst with high selectivity and functional group tolerance. In addition, the coupling reaction of chlorobenzene with boronic acids was also investigated (Table S4†), and the diphenyl production activity was poor, which may be ascribed to the much higher bonding energy of C–Cl (346 kJ mol−1) than that of C–Br (290 kJ mol−1) and C–I (228 kJ mol−1).48
a Reaction conditions: Pd (0.47 mol% relative to organohalide), 0.5 mmol organohalide, 0.8 mmol arylboronic acid, 1.5 mmol K2CO3, 5 mL ethanol, iodides 30 °C, bromides 60 °C. The yields given below the structure were determined using GC and GC-MS. |
---|
Next, Arrhenius plots were plotted to investigate the apparent activation energies (Ea) of the Suzuki coupling over the Pd/Gr/C12A7:e−, Pd/Gr and Pd/C12A7:O2− catalysts. The Ea of Pd/Gr/C12A7:e− was estimated to be 58.7 kJ mol−1 of the coupling reaction of iodobenzene and phenylboronic acid (Fig. 4a), which shows a ca. 27.4% and 29.3% reduction compared with those of Pd/Gr (80.9 kJ mol−1) and Pd/C12A7:O2− (83.0 kJ mol−1), respectively (Fig. 4c). Moreover, the Ea follows the same trend as the iodobenzene replaced by bromobenzene (Fig. 4b), in which the calculated Ea of Pd/Gr/C12A7:e− (70.8 kJ mol−1) is ca. 31.0% and 28.6% less than those of Pd/Gr (102.6 kJ mol−1) and Pd/C12A7:O2− (99.1 kJ mol−1), respectively (Fig. 4c). These results suggest that the Suzuki coupling reaction is promoted in the presence of C12A7:e−. Kinetic reaction orders were estimated by changing the concentration of aryl halides and phenylboronic acid over Pd/Gr/C12A7:e−. Fig. 4d shows that both reaction rates are sensitive to the concentration of aryl halides but independent of phenylboronic acid. These results imply that the activation of aryl halides controlled the overall reaction process over Pd/Gr/C12A7:e−, which is well accepted for both homogeneous and heterogeneous Pd-based catalysts in Suzuki coupling reactions, i.e., the oxidative addition of the aryl halide is the rate-determining step for the catalytic cycle.49–52 From this kinetic analysis, the smaller Eas and reaction orders reveal that the enhanced activity of Pd/Gr/C12A7:e− over that for the Pd/Gr catalyst exclusively originates from the promoted activation of the aryl halides.
To further understand the activation behaviours of aryl halides, DFT calculations and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) spectra were also performed. As illustrated in Fig. 5a and b, obvious electron-deficient and rich regions are constructed at the interface between C12A7:e− and Pd. The elevated electron density of the Pd sites enables electron transfer (0.15 e−) to the adsorbed iodobenzene substrate, thereby weakening the C–I bond with 9.5% elongation (Fig. 5c). Here, the reaction barriers for the oxidative addition of aryl halide were also calculated over Pd/C12A7:e− and Pd (111). The activation process of the iodobenzene molecule proceeds with an energy barrier of 0.21 eV on Pd/C12A7:e−, much lower than that of 0.58 eV on Pd (111) (Fig. 5d). The tendency of the suppressed activation energy agrees well with the experimental observations, further suggesting that the activation step of aryl halide on Pd/C12A7:e− is remarkably enhanced.
Next, the adsorption behaviour of iodobenzene is experimentally evaluated using DRIFT spectroscopy (Fig. 5e). For Pd (111), the black line shows bands in the range of 1400–1600 cm−1, associated with C–C stretching in aromatic vibration (νC–C).53 In the low-frequency region, the IR vibrational bands of halogen-sensitive vibration (995 cm−1), out-of-plane C–H deformation (β(C–H): 1015 cm−1), and trigonal ring breathing (1062 cm−1) can be observed. Notably, if the catalyst changed to Pd/Gr/C12A7:e−, the IR vibrational peaks of 1577 and 1474 cm−1 associated with the admixture of C–I deformation show a significant redshift compared to the pure Pd case (Fig. 5f). Additionally, a similar redshift was detected for the C–I stretching vibration (1062 cm−1, Fig. 5f). The redshift phenomenon of the C–I bond-related vibration indicates a carbon halogen bond-weakening process. These results agree with the DFT calculation results, in which the low work function C12A7:e− electride generally acts as an electron-donating ligand and improves electron donation to the antibonding orbitals of the aryl halides through the surface Pd active sites, which is responsible for the weakening of the carbon halogen bond, the rate-determining step of the reaction. The results of the kinetic analysis suggest the importance of the electronic effect between C12A7:e− and the highly dispersed active Pd species in terms of electron density modification, which results in the significant enhancement of the catalytic activity of Pd active sites towards the cross-coupling reaction.
Owing to the high activity of the activation of aryl halides over Pd/Gr/C12A7:e−, we became interested in testing the catalyst for various C–C couplings, such as Sonogashira, Stille, Hiyama and Heck coupling reactions. As shown in Table 2, each of the tested iodobenzene and coupling partners, such as phenylacetylene, tributylphenylstannane, phenylethylene, and trimethoxyphenylsilane, could be converted to the corresponding coupled products in a high yield. The results confirmed the versatile applicability of the Pd/Gr/C12A7:e− catalyst for various C–C cross-coupling reactions.
Entry | Aryl halides | Coupling partner | Product | Time (h) | Yield (%) |
---|---|---|---|---|---|
a Reaction conditions: Pd (0.47 mol% relative to organohalide); 0.5 mmol iodobenzene, 0.8 mmol coupling partner, 1.5 mmol K2CO3. The yields given below the structure were determined using GC and GC-MS. b 5 mL ethanol, 60 °C. c 5 mL DMF, 120 °C. | |||||
1 | 12 | 93 | |||
2 | 18 | 91 | |||
3 | 24 | 82 | |||
4 | 24 | 88 |
To check whether the coupling reaction over Pd/Gr/C12A7:e− catalyst is a heterogeneous reaction, the catalyst was removed from the reaction mixture by hot filtration after 1 h of reaction, and the filtrate proceeded under the same reaction condition. For the stability test, each coupling reaction of aryl halides and phenylboronic acid was conducted under the same conditions. After finishing each coupling reaction, the catalyst powder was separated via centrifugation, followed by washing three times with ethanol and water to remove the organic and inorganic residues. Subsequently, the catalyst powder was allowed to dry in a vacuum at room temperature, weighed, and reused in the next run. For the moisture-resistant test, Pd/Gr/C12A7:e− powder was impregnated in water for 1 h at room temperature. The catalyst powder was then separated via centrifugation, followed by the removal of water and drying under a vacuum overnight.
ΔE*x = Esys − Eslab − Ex, |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ta08899a |
‡ These authors have contributed equally. |
This journal is © The Royal Society of Chemistry 2023 |