Yuhui Liuab,
Xiaoxu Deng*b,
Chuangyun Guoa,
Shuang-Feng Yin*cd and
Peng Chen*a
aSchool of Chemistry and Chemical Engineering, Guizhou University, Guiyang 550025, Guizhou, China. E-mail: pchen3@gzu.edu.cn
bKey Laboratory of Electronic Composites of Guizhou Province, College of Big Data and Information Engineering, Guizhou University, Guiyang, 550025, Guizhou, China. E-mail: dxjdeng7@163.com
cCollege of Chemistry and Chemical Engineering, Central South University of Forestry and Technology, Changsha 410004, P. R. China. E-mail: sf_yin@hnu.edu.cn
dAdvanced Catalytic Engineering Research Center of the Ministry of Education, State Key Laboratory of Chemo/Biosensing and Chemometrics, College of Chemistry and Chemical Engineering, Hunan University, Changsha 410082, P. R. China
First published on 25th August 2025
Efficiently activating inert C–H bonds while maintaining control over the selective pathways of complex chemical reactions involving high-energy species remains a highly challenging and as-yet unattained objective. Herein, we propose a novel concept called ‘dynamic local free radical confinement-mediated mechanism’ to efficiently achieve synergetic selective oxidation of toluene and hydrogen generation for the first time via a CdS–CdIn2S4 semi-coherent heterojunction (CCS) under two-phase conditions. Surprisingly, the optimized CCS-2 exhibited amazing catalytic efficiency and long-term stability in gram-scale experiments and automatically separated the catalyst from the product. The mechanistic study indicates that the unstable semi-coherent interface in CCS-2 establishes a channel for directed carrier migration. Furthermore, the unstable semi-coherent interface facilitates the assembly of low-coordinate cadmium sites with surface hydroxyl groups, resulting in the formation of a first-layer hydrogen bonding framework, which effectively cleaves the C–H bond and dynamically inhibits the adsorption of aldehydes, thereby inducing spatial separation between the two phases. Our study highlights a new insight into the selective regulation mediated by surface radicals and introduces a novel and universal approach for achieving environmentally friendly chemical synthesis and energy conversion.
Considering the aforementioned factors, if selective oxidation of benzylic C–H bonds can be achieved while simultaneously producing clean and renewable H2 energy, it would offer significant advantages in terms of dynamic and thermodynamic benefits as well as carrier utilization rate compared to conventional methods involving multiple steps for interconversion with electron–hole or active species.11–15 However, the low oxidation kinetics of hole in the photocatalytic C–H bond coupling process provide dynamic advantages, enabling simultaneous hydrogen generation instead of selective C–H bond oxidation (Fig. 1a).16 Furthermore, despite water-induced species (·OH) can activate C–H bond bonds to form oxygenated products, the high-energy species are prone to excessive oxidation in reactions and capturing active proton hydrogen results in the inability to generate hydrogen.4 Moreover, due to their limited solubility in water, C–H compounds are frequently misused with inappropriate solvents, posing challenges in the separation process flow related to solvent usage. Therefore, efficiently activating inert C–H bonds while maintaining control over the selective pathways of complex chemical reactions involving high-energy species remains a highly challenging and yet unattained objective.
To address these challenges and enhance the efficiency of photocatalysis, we propose a novel concept called ‘dynamic local free radical confinement mediated mechanism’ for the first time. This method exploits the unstable semi-coherent interface of CdS–CdIn2S4 amphipathic heterojunction with structural disorder and a decrease in coordination number. Notably, the unstable semi-coherent interface facilitates the assembly of low-coordinate cadmium sites with surface hydroxyl groups, resulting in the formation of a first-layer hydrogen bonding framework that confines highly active radicals of ·OH and toluene and benzyl alcohol within a localized area but dynamically inhibits the adsorption of aldehydes, thereby inducing spatial separation between the two phases. This overcomes the long-standing challenge of low selectivity for C–H activation and excessive oxidation. Moreover, the unstable semi-coherent interface in CCS-2 establishes a channel for directed carrier migration. Remarkably, the optimized CCS-2 exhibited large-scale production of H2 and oxidized compounds while maintaining long-term selectivity retention and automatically separated the catalyst from the product (see the SI experiment section for details, Fig. S1, SI). The dynamic biphasic demonstrate practical benefits in terms of retrieving the product, expenses, efficiency, and expandability for solar hydrogen production and selective oxidation of organic compounds.
This phenomenon can be attributed to the lattice transformation of the (311) face of CdIn2S4 into the (002) face of CdS by tensile extrusion. Moreover, the geometric phase analysis (GPA) reveals that CdS (002) experiences compression while most of the CdIn2S4 undergoes tension on CCS-2 (Fig. 1f). This has been confirmed by observing an increasing trend and a shift to higher angles in the (002) peak of CdS, while the (311) peak of CdIn2S4 demonstrates a decreasing tendency with an increase in sulfur content (Fig. 1g and S7, SI). In addition, X-ray diffraction (XRD) Rietveld refinement results show that on CCS-2, CdS is compressed while CdIn2S4 is stretched (Fig. S8 and Table S1, SI), further suggesting that CdS (002) originates from the lattice transition of stretched CdIn2S4 (311).20 Besides, the molar ratio of the Cd/In in CCS-2 was confirmed to be 1.08 by the plasma mass spectrometry (ICP-MS) (Table S2, SI), demonstrating the unstable surfaces of CdS–CdIn2S4 heterojunction are successfully prepared. Those metastable crystal faces are prone to causing structural disorder and a decrease in coordination number, which is conducive to the enhancement of material surface activity.
Synchrotron X-ray absorption fine structure (XAFS) spectroscopy was used to further identify the structural alterations in CCS-2. As shown in Fig. 1h, S9 and Table S3 (SI), CCS-2 manifestly favors a lower energy state compared to CdS/CdIn2S4, which corresponds to an Cd oxidation state of +3.64, which is lower than that of CdS/CdIn2S4 at +3.97, suggesting that less coordination of metastable surface. In addition, electron paramagnetic resonance (EPR) analysis verified sulfur vacancies (Fig. S10, SI), demonstrating the low-coordination structure. The Fourier transform (FT) extended X-ray absorption fine structure (EXAFS) spectra, compared to CdS/CdIn2S4, show that the phase-shift corrected Cd–S bond peak is displaced from 2.07 to 2.04 Å, indicating the presence of low-coordinated Cd sites and lattice contraction in CCS-2 (Fig. 1i). As shown in Fig. 1j–l, the wavelet transform visualizes the changes occurring in the lattice of the Cd–S within the samples.
Furthermore, the material's surface electronic changes were observed via X-ray photoelectron spectroscopy (XPS). The S 2p pattern may be deconvoluted into two different peaks at 161.5 and 162.7 eV (Fig. S11a, SI), corresponding to S 2p3/2 and S 2p1/2 signals, corresponding the presence of S2−.21,22 The two In 3d peaks at 444.8 and 452.4 eV correspond to In 3d5/2 and In 3d3/2 (Fig. S11b, SI), respectively, which can be attributed to typical In3+ in CCS-2.23 As shown in Fig. S11c (SI), two peaks at 405.3 and 412.0 eV in the Cd 3d spectrum of CCS-2, corresponding to Cd 3d5/2 and Cd 3d3/2 signals, indicating the Cd2+ oxidation state of Cd, which consistent with the synchrotron radiation data.24,25 The results indicated that the binding energy shifts of S 2p, In 3d, and Cd 3d in CCS-2 were lower than in CdS/CdIn2S4 heterojunctions, indicating a very low degree of coordination that may aid in the adsorption of the reactants.26 In conclusion, the stretching and extrusion of the lattice lead to the formation of sub-stable transition surfaces. The defects on these surfaces eventually exhibit a low-coordination state, which enhances substrate adsorption and influences the PCH reaction.
For the most reported works, the over-oxidation product of benzoic acid or CO2 could be obtained by prolong reaction time.27 Therefore, the time-dependent toluene conversion over CCS-2 sample was also tested. With the increase of irradiation time from 3 to 48 h, the toluene conversion rate increased (conversion rate: 0.08%, 0.17%, 0.36%, 0.73%, 1.46%, Fig. 2c and Table S4, SI), while benzoic acid was not detected in the product (Fig. S16, SI). Importantly, CCS-2 will sink into the water the material after operated for 48 h, indicating that the prepared catalyst can effectively suppress excessive oxidation of toluene to benzoic acid (Fig. S17, SI). Moreover, the experiment of hydrogen production from D2O proved that the hydrogen in the reaction system came from water rather than toluene (Fig. S18, SI). In the isotope experiment using H218O, the 16O in the product was replaced by 18O, which suggesting that the oxygen in the products originated from water (Fig. 2d and e). Furthermore, pH-dependent studies revealed optimal catalytic performance under neutral conditions, whereas acidic media induced catalyst deactivation and competitive hydrogen evolution, suppressing organic transformations, while alkaline conditions diminished activity due to hydroxyl-induced active-site passivation (Fig. S19).
The toluene substrates extension has been investigated to broaden the reaction applicability. The observed differences in reactivity can indeed be attributed to electronic effects, where electron-donating groups (e.g., –CH3 or –C(CH3)3) enhance the electron density of the aromatic ring, facilitating C–H bond activation, while electron-withdrawing groups (e.g., –Cl or –Br) exhibit slightly lower reactivity due to reduced electron density. This trend aligns with our mechanistic studies, which highlight the role of surface hydroxyl groups in modulating substrate adsorption and activation (Table S5, SI). For industrial production, the recyclability and stability of the catalyst is an important role and are verified by cyclic experiments. As shown in Fig. 2f, S20 and S21 (SI), there is no obvious deactivation after five cycles of photocatalytic test, indicating that the prepared catalyst has high stability and recyclability. Above all, CCS-2 exhibited excellent photocatalytic activity, selectivity and durability, which means that the considerable prospects in photocatalytic hydrogen production synergistic with toluene oxidation. Additionally, our reaction system was implemented delicately without using solvents and co-catalyst, this implies easier separation of products and results in an efficient and economically viable reaction system, with promising industrial application prospects.
While the drift trend of charge carriers is easily understood, predicting their intrinsic drift behavior remains challenging. There is reported that the (002) facet of CdS is a reductive crystal plane,28 which originates from a lattice transition induced by tensile strain in CdIn2S4 and offers significant advantages for carrier transport in the heterojunction. To confirm the functionality of intermediate coherent interfaces, density functional theory (DFT) calculation has been carried out. As shown in Fig. S27 (SI), the differential charge density suggests that the coherent interface of CdS acts as an internal carrier convergence center, which drives the excited electrons from CdS into CdIn2S4. To further illustrate the transmission of photogenerated electrons at the interface of CCS heterojunction, light-assisted Kelvin Probe Force Microscopy (KPFM) has been carried out (Fig. 3a). The higher contact potential difference value of 4.6 mV for CCS-2, compared to 2.8 mV for CdS/CdIn2S4, underscores the enhanced charge separation efficiency facilitated by the semi-coherent interface in CCS-2. This difference arises from the unique electron screening effect at the unstable interface, which reduces carrier loss. These results collectively highlight the critical role of the semi-coherent interface in optimizing charge transfer dynamics.
Femtosecond time-resolved transient absorption (fs-TA) spectroscopy was conducted to confirm the generation and transfer mechanism of photogenerated carriers at the intermediate coherent interface (Fig. 3b–e and S28, SI). CCS-2 and CdS/CdIn2S4 exhibited a four-step electron relaxation process.29 Since the electron diffusion lifetime on the lattice is very short (a few picoseconds) attributing τ1 (3.23 ps) to electron leaps,30 attributing τ2 (50.42 ps) to sulfur defect trapping,31 and the slower quenching process of τ3 (1124.99 ps) describes the recombination of the electrons in the CB with the holes in the VB.32 Therefore, attributing the new fitted lifetime of 60.13 ps (τ4) attributed to interfacial electron transport is reasonable,33 suggesting that additional steps are required for electron transport from CdS to CdIn2S4 at the CCS-2 interface. In addition, the interface transport time is long compared to CdS/CdIn2S4, which is notably slower than in CdS/CdIn2S4, confirming the semi-coherent interface's role as an electron trap that mitigates recombination.
By implementing a logical regulation of charge carriers, one can anticipate a remarkable trend in the migration of carriers. The charge effective mass calculations showed that CCS has the least charge effective mass, indicating the maximum charge transfer rate in the heterojunction, as shown in Fig. S29 (SI). Furthermore, CCS-2 exhibited weakest photoluminescence (PL) intensity and the longest charge average lifetimes (Fig. 3f and g), confirming the highest carrier separation efficiency in the heterojunctions. Additional evidence demonstrates that CCS-2 has the largest photocurrent density and the smallest electrochemical impedance (Fig. 3h and i), indicating a significant enhancement of photogenerated charge separation at the CCS, thereby leading to heightened photocatalytic activity. In summary, the coherent interface of CdS serves as an internal carrier convergence center. This facilitates the formation of the built-in electric field, leading to the directional combination of charges instead of random recombination at the heterojunction interfaces.
In our system, the key is to control the simultaneous oxidation of hydrogen and toluene while maintaining the selectivity of toluene oxidation. Fig. S35 (SI) illustrates that CCS-2 exhibits a low contact angle and similar affinity for water and toluene, indicating its potential as a Pickering emulsion catalyst for simultaneous oxidation of toluene and hydrogen production through regulated catalytic reactions within distinct interfacial environments.42 However, the catalyst has a smaller contact angle with the toluene–benzyl alcohol solution compared to toluene and water, but a larger contact angle with the toluene–benzyl alcohol-benzaldehyde solution, suggesting CCS-2 is able to suppress the oxidation of toluene to benzoic acid or CO2 (Fig. S36, SI). Therefore, it can be concluded that CCS-2 exhibits unique surface properties at the interfaces of two liquid media, enabling simultaneous activation of toluene and intermediates as well as water.
To further understand the unique two-phase interface relationship, in situ FTIR measurements using various toluene-based substances and water were conducted. As shown in Fig. 4b, compared with CdS, CdIn2S4, and CdS/CdIn2S4, the characteristic ν(C–H) band of toluene of CCS-2 has a red shift in the concomitant toluene and water, indicating that CCS-2 has a preference for adsorbing toluene.43,44 Furthermore, the characteristic ν(C–O) band of benzyl alcohol also exhibited a red shift on CCS-2 in the concomitant benzyl alcohol and water (Fig. 4c), while benzaldehyde (Fig. 4d), benzoic acid (Fig. S37, SI) and 1,2-diphenylethane (Fig. S38, SI) adsorption exhibited a blue shift,45,46 suggesting the CCS-2 support the existence of benzyl alcohol and detest benzaldehyde, benzoic acid and 1,2-diphenylethane in a two-phase aqueous solution. In addition, fitting the broad peaks of H2O in the in situ FTIR spectra yielded a ratio of the linear structure (DA) to the hydrogen-bonding network (DDAA) of CCS-2 of 2.01, which is higher than that of CdS (1.48), CdIn2S4 (1.38), and CdS/CdIn2S4 (1.69) (Fig. S39, SI).47 The DA structure facilitates water adsorption on the surface of the material and further formation of surface hydroxyl groups.
In order to analyze the effect of surface hydroxyl groups on the reaction kinetics, the density distributions, free diffusion coefficients and coordination numbers of CCS (without surface hydroxyl groups) and CCS-OH (with surface hydroxyl groups) were simulated using density distribution curves, the radial distribution function (RDF) and the mean square displacement (MSD) method. Firstly, the theoretical models of CCS-OH and CCS are analyzed, and it is found that organic matter and water on CCS-OH are closer to the interface (Fig. S40 and S41, SI). Second, in the benzyl alcohol–water system, water is closer to the interface than in the other systems, corresponding to the contact angle results (Fig. 4e). As shown in Fig. S42 and S43 (SI), the distances of both organic matter and water from the active site Cd are shorter in CCS-OH than on CCS. In addition, the RDFs of organic matter and water with the active site Cd in CCS-OH and CCS were compared and the coordination number was calculated. The high peak value and coordination number of CCS-OH structure indicate that the surface hydroxyl group promotes the adsorption of H2O on the catalyst, which is conducive to the formation of hydroxyl radical (Fig. 4f and Table S6, SI). Interestingly, the coordination number of cadmium active sites on CCS-OH with toluene and benzyl alcohol is higher than that of CCS, while the opposite is true with benzaldehyde and benzoic acid, suggesting that the adsorption of organic substances by surface hydroxyl groups is selective (Fig. S44 and Table S7, SI). Finally, MSD analysis showed that the surface hydroxyl group reduced the diffusion coefficient, again proving that the surface ·OH was beneficial to the adsorption of organic matter and water (Fig. 4g).
The Gibbs free energy have been computed in order to better understand the internal reaction process. Remarkably, the ΔGH* for CCS is significantly smaller (−0.29 eV) than that of CdS/CdIn2S4 (−0.13 eV), CdS (−0.08 eV) and CdIn2S4 (−0.04 eV), suggesting that CCS demonstrates superior activity as an HER catalyst compared to other materials (Fig. S45, SI). On the other hand, compare to CdS, CdIn2S4, and CdS/CdIn2S4, CCS exhibit the lowest free energies in the toluene oxidation (Fig. 4h, S46 and S47, SI). Notably, the reaction from toluene to benzyl alcohol becomes increasingly exergonic, while the reactions from benzyl alcohol to benzaldehyde and then to benzoic acid become increasingly endergonic, suggesting CCS can fundamentally control their selectivity. Furthermore, the adsorption energy of the toluene coupling process is notably positive than that of the oxidation process (Fig. S48, SI), suggesting the competitive oxidation process of toluene yields benzyl alcohol and benzaldehyde as the primary products. To further confirm the impact of surface hydroxyl groups, DFT calculations are being carried out (Fig. S49, SI). As shown in Fig. S50 (SI), CCS exhibit the related adsorption energies for toluene (−4.77 eV) and H2O (−4.69 eV), indicating even under initial conditions (without hydroxyl radicals), the catalyst could theoretically react with water and toluene at the same time (consistent with contact angle results). Notably, the adsorption energy for benzyl alcohol is −5.12 eV, while benzaldehyde, benzoic acid, and 1,2-diphenylethane are −1.57, −1.47 and −1.16 eV, demonstrating stronger adsorption of benzyl alcohol than others. It is noteworthy that the adsorption energies of toluene and benzyl alcohol are more negative and the adsorption energy of benzaldehyde is more positive in the paths with surface hydroxyl groups (Fig. 4i). Therefore, the low-coordinate CdS–CdIn2S4 heterojunction assembles hydroxyl groups to dynamically regulate substrate adsorption, enhancing both the selectivity of toluene oxidation and the rate of hydrogen production.
Consequently, the final path for toluene oxidation is obtained (Fig. S51, SI). The catalyst initially adsorbs both water and toluene during the reaction. With the formation of surface hydroxyl groups, the catalyst enhances the adsorption of toluene and oxidizes it to benzyl alcohol, with the concomitant production of hydrogen. Benzyl alcohol is more susceptible to the surface hydroxyl groups of the catalyst, and its adsorption capacity to the toluene–benzyl alcohol solution increases, leading to a gradual transfer of the catalyst to the toluene solution, which in turn oxidizes benzyl alcohol to benzaldehyde. However, when benzaldehyde is present in the toluene–benzyl alcohol solution, the catalyst repels benzaldehyde, leading to a gradual transfer of the catalyst to water again, at which point further oxidation of benzaldehyde to benzoic acid is no longer possible, and ultimately, the selectivity of benzyl alcohol and benzaldehyde is greatly improved. Most importantly, we explored a way for surface hydroxyl groups to modulate interfacial adsorption, which is conducive to improving the selectivity of organic oxidation.
This journal is © The Royal Society of Chemistry 2025 |