Sobia
Jabeen
a,
Yuanyuan
Cheng
a,
Yaxi
Li
a,
Naiyun
Liu
*a,
Yunliang
Liu
a,
Zhiquan
Lang
a,
Xiuyan
Wang
*b,
Fan
Hu
*c and
Haitao
Li
*a
aSchool of Chemistry and Chemical Engineering, Institute for energy Research, Jiangsu University, Zhenjiang, China. E-mail: liuny@ujs.edu.cn; liht@ujs.edu.cn
bCollege of Materials and Textile Engineering, G60 STI Valley Industry & Innovation Institute, Jiaxing University, Jiaxing, Zhejiang 314001, China. E-mail: xiuyan20@mails.jlu.edu.cn
cHuanggang Center for Disease Control and Prevention, Huanggang, China. E-mail: 17771399439@163.com
First published on 31st October 2025
The addition of metal oxides as impurities to generate an intermediary energy band near the conduction or valence band to reduce the bandgap is the most distinctive approach to improve the photo-absorption characteristics of the material. Herein, we have reported the synthesis of nanohybrid bimetallic heterostructures by linking the interface of CeO2 and CoO, which narrows the electronic band structure of CeO2 from 2.85 eV to 1.5 eV and modulates the distribution of charges at the active sites. The resulting CeO2/CoO hybrid support enhances the dispersion and stability of Pd NPs, resulting in lowering the activation energy (Ea) barrier of the coupling reaction, thereby significantly enhancing its catalytic efficacy. The Ea value of CeO2/CoO/Pd (53.7 kJ mol−1) is much lower compared to that of CeO2/Pd (68.6 kJ mol−1), with excellent catalytic activities (yield: 98%) and exhibiting long-term stability for 5 continuous cycles without any significant loss in activity. Overall, the CeO2/CoO/Pd hybrid system effectively utilized the photothermal effect to facilitate an effective electron transfer, thereby enhancing the rate of the Suzuki–Miyaura coupling reaction. This study offers a feasible and encouraging prospect to use the heterostructured metal oxide-based catalytic system for efficient Suzuki–Miyaura cross-coupling reaction.
The application of rare earth (RE) oxides, as support for metal loading, enables highly effective synergistic catalysis due to their unique electronic structures, crystal topologies, and electromagnetic characteristics.13,14 Collectively, the characteristics of oxide supports, such as their geometric architecture, electrical structure, and defect sites, offer further ways to customize the overall performance of catalysts.15 Among RE oxides, ceria (CeO2) exhibits distinct physical and chemical properties as well as redox features, which can be attributed to its electronic configuration, [Xe] 4f2 5d1 6s2.16 Besides this, it can stabilize metal nanoparticles through its flexible coordination environment and the formation of oxygen vacancies (OVs), which arise due to reversible valence states between Ce4+ and Ce3+ ions,17,18 offering additional possibilities for enhancing the stability and equilibrating the positive charges of metal oxide ions, thereby enhancing the catalytic activity.19 Ceria, being an important semiconductor, has also been explored as a photocatalyst for many chemical processes, i.e., CO2 redox reaction,20–22 organic pollutant degradation,23,24 and air purification.25 The n-type CeO2 is characterized by a broad bandgap (2.8–3.2 eV) and absorbs radiation in the UV region and to a limited extent in the visible range (400 nm).26,27
Regrettably, the wide bandgap of CeO2 hinders its ability to effectively harness abundant solar energy because the visible range comprises ∼50% of the overall solar radiation.28 Additionally, the photo-induced e−/h+ pairs are prone to recombination,29 resulting in the production of unproductive heat and light, which reduces the photocatalytic efficiency of CeO2.30,31 To address the aforementioned problems, various efforts have been made to amend pure CeO2 in order to improve its photocatalytic potential. A highly effective approach to improve the activities of ceria-based catalytic systems is to engineer the interface by integrating it with other metal oxides. Metal-oxide interfaces offer a novel prospect for enhancing catalytic activity through electrical and chemical interactions at the interface.24,32 For instance, the interface engineering strategy has helped to reduce the bandgap (Eg) of CeO2 by raising its valence band (VB) or lowering the conduction band (CB) or executing both processes at once.33,34 Herein, earth-abundant cobalt oxide was utilized to fabricate the CeO2 interface to narrow its bandgap. The addition of CoO to CeO2 can introduce impurity energy levels near the CB and VB within the original bandgap and create lattice defects, which function as recombination centres.35,36 Besides this, cobalt oxide (CoO) has numerous oxygen vacancies, variable oxidation states, and distinct redox properties, which can operate as active sites and promote electron transport in catalytic reactions. Recent research has shown that the efficiency of transition metal oxide-based catalysts is closely linked to their morphologies and variable electronic structures due to the d orbital, which can adjust their adsorption energies to enhance their catalytic performance.37 Therefore, the integration of CoO with ceria will result in the formation of additional OVs to regulate the electronic structure, enhance charge transfer, and offer the most favorable adsorption–desorption energy for the intermediate species during the reaction.38,39 The increased number of active sites and profuse OVs close to the interface have a notable impact on improving the catalytic activity.40 Wang et al. developed a series of self-assembled cobalt–cerium oxides (CoCe-x) with controlled oxygen vacancies to improve the decontamination of air pollution by low-temperature catalytic oxidation. The oxygen vacancies in the catalyst were optimized through the effective integration of Co atoms into the ceria lattice, thereby enhancing the charge transfer by regulating the redox cycles of Co3+/Co2+ (ref. 41) and Ce4+/Ce3+.42
Most reported literature on CoO-based catalysts has focused on applications in water splitting, oxygen evolution, and other oxidation processes. Besides this, there are only a limited number of studies examining CoO-based photocatalysts. To our knowledge, the application of CoO-based photocatalysts in coupling reactions, such as the Suzuki–Miyaura reaction, has been rarely investigated. Consequently, the integration of CoO with nanoceria to create a bimetallic oxide support for Pd NPs presents an innovative approach to broaden the catalytic uses of CoO beyond its traditional functions. In this study, the CeO2/CoO/Pd-based photocatalytic system has been synthesized through the straightforward approach of the calcination-reduction process for catalysing the visible light-stimulated Suzuki coupling. The approach of interface engineering has been utilized to control the electronic band structure of CeO2 at the interface, including the distribution of charges at the active sites. We have shown that CeO2/CoO interfaces with distinct spatial orientations have enhanced the photocatalytic SMC reaction by stabilizing the photogenerated electrons/holes in the conduction/valence band upon exposure to visible light. The photogenerated electrons from the CB of CeO2 are transferred to CoO and then to spatially connected Pd NPs. Mostly, the Pd NPs are disseminated as fine metallic particles (Pd0) or oxidized Pd ions (PdO) on the ceria support. However, Pd2+ ions are found to be much more abundant on the surface of the CeO2 support due to more oxygen vacancies as compared to Pd0 NPs. Therefore, on exposure to light, the photogenerated electron moves to spatially connected PdO, facilitating the surface-adsorbed Pd2+ ion to reduce for the coupling reaction. The bandgap narrowing not only promotes wide light absorption but also improves the usage of photogenerated electrons and holes, thus reducing the energy barrier for the catalytic cycle. The synergistic impact of the bimetallic oxide support (CeO2/CoO) together with photothermal catalysis has enhanced the rate of the SMC reaction.
:
Co molar ratio in the precursor was approximately 3.4
:
1.
:
1 (v/v, 5 × 20 mL) EtOH/water solution. The washing was continued until the filtrate was colourless, ensuring the complete removal of unreacted PdCl2. The resulting catalyst was dried in a vacuum oven at 60 °C and then reduced under an atmosphere of H2/Ar at 300 °C for 3 h to obtain CeO2/CoO/Pd.44 The overall molar ratio of Ce
:
Co
:
Pd in the final composite was roughly 3.3
:
1
:
1.2. The Pd loading was also calculated using inductively coupled plasma mass spectrometry (ICP-MS) and was found to be 0.23 wt% Pd. Other reference catalysts containing CeO2/Pd, CoO/Pd were also prepared without adding cobalt nitrate (Co(NO3)2·6H2O) and CeO2, following the above protocol, to investigate the role of CoO and CeO2 in the enhancement of catalytic activities in the SMC reaction.
:
v = 1
:
1). The reaction temperature was sequentially maintained at 30 °C, 40 °C, 50 °C, 60 °C and 70 °C. A 500 μL aliquot was collected from the reaction mixture every 5 min and poured into a micro vial containing 1 mL of EtOAc and mixed with 1 mL of H2O, shaken well, and the organic layer was extracted with EtOAc (1 mL × 3). The organic layers were further dried using MgSO4 (anhydrous) and filtered. The product yield was quantified using gas chromatography (GC). The addition of more equivalents of phenylboronic acid in the reactants, as compared to the aryl halide, allows us to consider the SMC reaction as a pseudo-first-order kinetic reaction.
XRD and XPS are performed to further characterize the chemical and electronic structure of CeO2/CoO and CeO2/CoO/Pd. The XRD scan is shown in Fig. 1c. CeO2 is identified by eight distinct peaks at 2θ = 28.66° (111), 33.01° (200), 47.38° (220), 56.27° (311), 59.02° (222), 69.57° (400), 76.71° (331), and 79.16° (420) (JCPDS no. 81-0792),50 whereas CoO is identified by three distinct peaks at 36.67° (111), 42.8° (200), and 61.72° (220) (JCPDS no. 43-1004).47,51 The CeO2/CoO/Pd scan shows two distinct peaks of Pd at 40.16° (111) and 46.68° (200) (JCPDS no. 88-2335), which has affirmed the formation of a nanohybrid heterostructure supported on nanoscale bimetallic oxide support (CeO2/CoO). The electronic structure of the hybrid catalyst is investigated by X-ray photoelectron spectroscopy (XPS). XPS spectra are obtained for the Ce 3d, O 1s, Co 3d, and Pd 4d for each of the catalysts. The complete survey scan of precursor (CeO2/CoO) and catalyst (CeO2/CoO/Pd) is demonstrated in Fig. S4, whereas the elemental XPS of precursor and catalyst are presented in Fig. 1(d–g). Fig. 1d shows the XPS scan of Pd. The smaller peaks at the binding energy of 335.8 and 340.8 eV are ascribed to metallic Pd0, whereas the larger peaks at 337.8 and 343.0 eV are ascribed to Pd2+.52 It has been observed that more Pd NPs on the surface of CeO2/CoO are dispersed as oxidized Pd2+ rather than Pd0 NPs.7,53 The Pd2+ ions have also substituted the Ce4+ ions/ Co2+ ions in the composite structure of CeO2/CoO to ensure the even dispersion of Pd NPs because the introduction of Pd caused a significant shift of the Ce 3d and Co 2p XPS peaks, indicating electronic interaction and charge redistribution between the precursor and Pd NPs. This change indicates a potential replacement of Pd into the lattice positions of Ce or Co rather than simply surface deposition.54 The integration of Pd into the CeO2/CoO lattice induces a charge imbalance, which is compensated by the formation of oxygen vacancies, disturbing the local lattice symmetry. Besides this, the partial reduction of Ce4+ to Ce3+ during Pd inclusion also causes lattice expansion because of its larger ionic radius than Ce4+. The observed lattice expansion mostly results from the weakening of local lattice symmetry due to oxygen vacancies and the larger ionic radius of Ce3+ compared to Ce4+. The chemical stability and structural durability of this heterogeneous catalytic system are attributed to the formation of Ce–O–Pd or Co–O–Pd in the composite structure, and this strong interaction of Pd2+ with the precursor has also inhibited the sintering of Pd NPs in the reaction mixture and prevents catalyst deactivation. The ionic Pd2+ formed by Ce–O–Pd/Co–O–Pd has brought much improvement in the catalytic performance of CeO2/CoO/Pd in the SMC reaction.55 In the XPS of Ce, two different forms of Ce are observed: Ce3+ and Ce4+. The peaks at 917.70, 908.45, 902.61, and 899.92. 889.70, and 883.49 eV are attributed to Ce4+ ions, whereas the peaks at 904.46 and 886.43 eV are assigned to Ce3+ ions in the XPS of the precursor material. However, all peaks show slight shifting in the Pd composite structure (CeO2/CoO/Pd); peaks at 916.51, 908.40, 902.42, 899.34, 887.51, 882.8 eV are associated with Ce4+ ions, and 905.51, and 885.10 peaks are attributed to Ce3+ ions56 (Fig. 1e). The shifting of XPS peaks represents the bonding of Pd NPs at the interface of CeO2/CoO during the annealing process and definitive evidence of a change in the Ce oxidation state resulting from strong metal–support interaction (SMSI). The Ce 3d peaks exhibit a shift towards lower binding energy, representing enhanced electron density due to the reduction of Ce4+ to Ce3+ on the surface of the CeO2/CoO support and the formation of oxygen vacancies. A single Ce peak (904.1 → 905.0 eV) exhibits a slight shift toward higher binding energy (+0.9 eV), which is ascribed to Ce atoms at the Pd–support interface undergoing local electron withdrawal as a result of Ce–O–Pd interactions. The XPS of CoO shows two prominent peaks at 781.41 eV and 796.8 eV, confirming Co 2p1/2 and Co 2p3/2, respectively (Fig. 1f).57 Additionally, two shake-up peaks, one at 786.24 eV and another at 802.73 eV, suggest the presence of Co(II) in the CeO2/CoO/Pd sample.56,58 The shifts of Co 2p components along with an increase in FWHM in CeO2/CoO/Pd suggest a more diverse cobalt chemical environment, characterized by the coexistence of Co2+ and Co3+ states, interface disorder, and defect formation (oxygen vacancies). This behavior aligns with the redox interaction between Co and Pd (Co2+ ↔ Co3+). The existence of Co2+ indicates that CoO is strongly interacting with the reducible ceria support, rather than creating bigger cobalt oxide aggregates. The O 1s XPS scan of the precursor (Fig. 1g) shows peaks at 529.6 and 532.3 eV ascribed to the binding energy of lattice oxygen (O) and defective or surface-adsorbed oxygen (Oads).59 The surface-adsorbed oxygen includes superoxide (O2−) and peroxide (O22−) and is believed to perform a critical role in defining the catalytic activity and reaction rate. The shifting of oxygen species at high binding energies (531.7, 532.5, and 533.6 eV) in the O 1s XPS scan of CeO2/CoO/Pd signifies the development of increasingly electron-deficient oxygen environments. The 533.3 eV peak probably refers to oxygen atoms interacting with several metal centers (Ce–O–Pd or Co–O–Pd bridges). Besides this, the incorporation of palladium has resulted in the formation of PdO on the catalyst surface and has enhanced the oxygen vacancies, leading to the displacement of the O 1s peak towards higher binding energies (532–533 eV).47,60 PdO may also interact with CeO2/CoO, inducing the integration of Pd NPs into the lattices of CeO2 or CoO, as discussed before in the discussion of Pd XPS. Besides this, the substitution of Pd2+ ions into the Ce4+ ions/Co2+ ions during the annealing process under an inert atmosphere would result in the formation of OVs, which also cause the shifting of the O 1s peak to high binding energy. This endorses that the interaction of PdO and CeO2/CoO is accountable for the formation of abundant Oads, resulting in better catalytic activity.
The FTIR spectra of CeO2/CoO, CeO2/Pd, and CeO2/CoO/Pd are shown in Fig. 2a. The CeO2/CoO FTIR spectra display two weak signals at 570 cm−1 and 660 cm−1 attributed to Co2+ in the composite structure of CeO2/CoO, thereby confirming the formation of CoO.61 The small signal at 557 cm−1 represents the stretching vibration of Ce–O in CeO2. The other stretching bands of Ce–O are visible in the spectra at 663 and 863 cm−1, further revealing the metal–oxygen bond. There are noticeable absorption peaks at 3421 cm−1 and 1630 cm−1 related to –OH stretching vibrations resulting from entrapped water molecules from the environment due to the high surface-to-volume ratio of nanomaterials.62Fig. 2b illustrates the Raman spectra of the CeO2/CoO composite and CeO2/CoO/Pd composite. Both samples show intense peaks at 462 and 456 cm−1 corresponding to the F2g phase of the CeO2 fluorite structure. The vibrational mode F2g is usually allied with the elongation of the Ce–O bond in the composite structure, and a slight change in the Ce–O bond position on the addition of Pd has facilitated the shifting of band location.63 Besides this, the peak shifting (F2g band) to lower wavenumber in the CeO2/CoO/Pd sample is attributed to various factors such as defect formation, variation in phonon relaxation, and phonon confinement relative to particle size. Other bands in the CeO2/CoO composite are 620 and 689 cm−1, corresponding to defect-induced (D) mode in the CeO2 structure due to oxygen vacancies and the valence state of the dopant material (Co). Bands in the bimetallic oxide support of CeO2/CoO at 480 cm−1 (Eg), 521 cm−1 (F22g), 622 cm−1 (F32g), and 690 cm−1 (A1g) are attributed to CoO.62 In particular, the band at 521 cm−1 in the CeO2/CoO composite corresponds to the Co–O stretching frequency, signifying the presence of CoO in the matrix.64 However, the weak signal at 1169 cm−1 is associated with the 2nd-order longitudinal optical (2LO) mode. Following the deposition of Pd, new bands at 640 cm−1, attributable to the vibrational mode of the Pd–O bond, emerged in the CeO2/CoO/Pd catalyst. The literature survey indicates that certain defects may influence the peak intensity of the defect-induced mode, whereas oxygen vacancies in the ceria lattice may enhance the peak intensity. However, the intensity of the CoO band has significantly reduced in the composite catalyst structure of CeO2/CoO/Pd, representing the addition of Pd NPs.65 The textural properties, such as surface area, porosity, and pore volume of the bimetallic oxide support and its catalyst are determined using nitrogen adsorption isotherms (Table 1). Fig. 2c displays each catalyst's nitrogen adsorption and desorption isotherms and pore size distribution. The impregnated samples (CeO2/CoO and CeO2/CoO/Pd) exhibited a type IV physisorption isotherm with an H1 hysteresis loop signifying the microporous nature of the catalyst, and nanoclusters/spherical particles are distributed in a homogenous manner.66 The pore size distribution indicates uniformity and facile pore connectivity. As indicated in Table 1, the BET surface area of the CeO2/CoO/Pd has shrunk, while the average pore diameter of the CeO2/CoO/Pd has enlarged due to Pd NPs. Pd NPs cover the active sites on the CeO2/CoO surface, reducing the surface area overall. Besides this, the detonation or clogging of the micropores by deposited Pd NPs during the impregnation and calcination process is the cause of the significant drop in the BET surface area of CeO2/CoO/Pd.56,67
| Catalyst | BET surface area (m2 g−1) | Pore volume (cm3 g−1) | Average pore diameter (Å) |
|---|---|---|---|
| CeO2/CoO | 37.6038 | 7.6 × 10−3 | 8.09 |
| CeO2/CoO/Pd | 9.9052 | 2.7 × 10−3 | 11.07 |
UV-vis diffuse reflectance spectroscopy is performed to examine the light absorption characteristics of the catalyst (Fig. 2d and S5a). The strong absorption band is seen in the UV range of 300–350 nm. The catalyst bandgap is calculated by using eqn (1):
| αhv = A(hv − Egap)n/2 | (1) |
The efficacy of this ternary nanohybrid catalytic system (CeO2/CoO/Pd) is investigated by catalyzing the SMC reaction of 4-bromotoluene and phenylboronic acid. The reaction conditions for SMC reaction are optimized by testing different solvents and bases under light at 50 °C (Table S1). The electrons are not energetic enough below 50 °C to be injected into adsorbate states, therefore contributing to the augmentation of the reaction rate by means of photothermal effects. The synergistic impact of photothermal catalysis has been operated, which has accelerated the rate of reaction and altered the selectivity pattern, resulting in superior catalytic efficacy. The effect of temperature on the reaction kinetics is discussed in detail in the next section. Following the optimization of reaction conditions, the activities of different catalysts (CeO2/CoO/Pd, CeO2/CoO, CeO2/Pd, CoO/Pd) are evaluated using the standard reaction of 4-bromotoluene and phenylboronic acid under optimized reaction conditions. The results are summarized in Fig. 3a (Table S2). The catalyst CeO2/CoO/Pd has shown the highest catalytic activity in the SMC reaction compared to the other catalysts (CeO2/CoO, CeO2/Pd). The CeO2/CoO/Pd catalyst has achieved a 98% yield, roughly 3.3 times greater than CeO2/Pd (30%) and 4.3 times greater than CoO/Pd (23%), whereas the reaction scarcely proceeds without Pd, suggesting that the CeO2/CoO system lacks catalytic activity. This comparison underscores the synergistic interaction between CeO2 and CoO in promoting electron transport, stabilizing Pd NPs, and enhancing catalytic efficiency. The CeO2/CoO/Pd catalyst demonstrates exceptional reactivity and selectivity (99%) in the SMC reaction. The SMC reaction of 4-bromotoluene gives 98% yield of cross-coupled product with 99% selectivity and TOF of 2675 h−1 in only 30 min.70 The increased reaction yield is attributed to the decreased bandgap of CeO2 (2.85 eV to 2.38 eV) as Co is incorporated into the CeO2 structure, which is further reduced to 1.99 eV with the addition of Pd NPs. Light exposure excites the electrons from the valence shell of CeO2 and transfers them to the CoO in the composite structure and finally to the surface Pd2+ ions to form Pd0.69 The bimetallic CeO2/CoO support has enhanced the rate of charge transfer, making CoO a highly active co-catalyst for the SMC reaction.69,71 The photocatalytic efficacy of CeO2/CoO/Pd for the SMC process is further examined in the dark (solely thermal reaction) and under light illumination. Control experiments under dark conditions at 50 °C clearly indicate that the catalytic performance is significantly enhanced under visible light irradiation, confirming that the activity arises predominantly from photocatalysis rather than purely thermal contributions (Fig. S8). Furthermore, the correlation between activity and light intensity (Fig. 3b) demonstrates that the contribution of photonic processes to the total reaction efficiency escalates from 43% in the dark (thermal reaction) to 84% at 30 mW cm−2 to 98% at 50 mW cm−2. The conversion of 4-bromotoluene in the absence of light irradiation is attributed to the thermal effect; however, only about 43% yield of the coupled product was obtained under thermal conditions in 30 min. The reaction was conducted again at 50 °C under light exposure, yielding about 98% of the coupled product. However, as light intensity increases, the contribution of light becomes more significant. Increasing the irradiation intensity from 10 to 50 mW cm−2 enhanced the yield of the cross-coupled product from 65% to 98%. The number of photoexcited electrons at elevated energy levels has been augmented by increasing the light intensity. This linear correlation between reaction conversion and light intensity indicates a photoexcited electron-driven chemical process. Furthermore, all light within the wavelength range of 420–600 nm (intensity 50 mW cm−2) enhances the yield of the coupled product; however, the yield varies with different wavelengths, and the contribution of the photonic process to reaction efficiency slowly diminishes as the wavelength of the incident light increases (Fig. 3c). Light on/off experiments further confirm that the catalyst activity is induced by light rather than thermal factor (Fig. S9), whereas wavelength-dependent studies indicate that catalytic activity has extended into the visible-light spectrum, aligning with the decreased bandgap energy. In pure CeO2 (2.85 eV), activity is mostly limited to UV excitation; however, after the inclusion of CoO, the initiation of photocatalytic reaction shifts to longer wavelengths, corresponding to the reduced bandgap of 2.38 eV. The relationship between the absorption edge and the catalytic performance demonstrates that bandgap narrowing facilitates the use of lower-energy photons, therefore improving visible-light-driven activity. The trend of the apparent quantum efficiency (AQE) curve corresponds with the UV-vis absorption spectra of CeO2/CoO/Pd (Fig. 3d). The action spectrum, representing AQE as a function of excitation wavelength, exhibited elevated AQEs at shorter wavelengths, with an estimated AQE of 6.3% attained under 420 nm light illumination. The AQEs corresponded closely with the absorption spectra of CeO2/CoO/Pd, therefore validating the role of light absorption amplification in the enhancement of yield in the coupling process.46 The significant increase in activity with shorter-wavelength photons suggests that photo enhancement at 420 nm is not only attributable to a basic photothermal effect. Photons of shorter wavelengths are capable of exciting electrons to elevated energy levels, making these electrons more efficient in facilitating coupling processes compared to those stimulated by photons with a larger wavelength of 600 nm. The excited electrons surpassing the energy barrier are introduced into the LUMO of the aryl halide adsorbed on Pd NPs. The contribution from this photoexcitation mechanism intensifies at a low wavelength (420 nm). Meanwhile, electrons stimulated by light of longer wavelengths (>500 nm) lack enough energy for excitation and they may participate in the process via the photothermal effect. Therefore, the nanohybrid heterostructure of the composite catalyst (CeO2/CoO/Pd) exhibited remarkable photocatalytic activity at low wavelengths due to both augmented light absorption and increased separation of e−/h+ pairs. The enhanced light absorption is attributed to the narrowing of the bandgap structure of nanoceria by incorporating CoO and Pd, which has kinetically facilitated the promotion of surface charge carriers, whereas thermal energy promotes the rapid formation of the oxidative addition complex, thereby enhancing the reaction rate. The data suggest that photogenerated charge carriers (including electrons and holes) play significant roles in the photocatalytic SMC reaction and photo–thermo catalysis processes function synergistically to augment the total photocatalytic rate. Temperature-dependent reaction kinetics are studied under optimized reaction conditions between phenylboronic acid and 4-bromotoluene in the temperature range of 30–70 °C. A sample from the reaction mixture is taken every 5 min and monitored by GC using nitrobenzene as an internal standard. The product yield of the SMC reaction linearly accelerates with the increase in temperature. According to Fig. 3e, the yield of the cross-coupled product is lowered at temperatures below 50 °C, but it has significantly increased to 98% when the temperature is raised to 50 °C within 30 min. The increase in temperature has resulted in the upgrading of electrons to higher energy levels via channeling through CeO2 → Co → Pd, which are further elevated in energy upon light absorption when irradiated. These excited electrons under the photothermal effect enhance the probability of rapid activation of aryl halide molecules adsorbed on the Pd NPs' surface, enabling them to overcome the activation barrier and thus initiate the reaction. However, when we further increased the temperature to 70 °C, the yield reached 99% in only 15 min, but the reaction selectivity dropped from 97% to 87% as a result of the temperature rise. Thus, it may be proposed that the rise in temperature has accelerated the rate of the SMC reaction, but this elevated temperature is not conducive to achieving the desired selectivity of the SMC coupled product because more heat energy has stimulated the self-coupling reaction in the solution, indicating that more molecules have crossed the activation barrier, thus increasing the rate of side reaction. To differentiate photocatalytic effects from thermal effects, the activation energy (Ea) for the CeO2/CoO/Pd catalyst under visible-light irradiation and in the dark at the same reaction temperature is investigated. Then the curves of 4-bromotoluene conversion with respect to time at each temperature are drawn (Fig. S10a–c). The time-dependent ln(Co/Ct) curve exhibits pseudo-first-order reaction kinetics with respect to the concentration of phenylboronic acid (Fig. S10d–f).68 Reaction rate constant (k) values are computed from the slope of ln(Co/Ct)–time curves. The activation energy (Ea) of the reaction is determined from the Arrhenius plot by drawing ln
k versus 1000/T (Fig. 3f). The Ea of CeO2/CoO/Pd under visible light irradiation (53.7 kJ mol−1) is markedly lower than that under thermal circumstances (76.3 kJ mol−1), indicating that photoexcited charge carriers efficiently reduce the reaction energy barrier accelerating the rate of the coupling process. This observation suggests that the primary factor contributing to the increased reaction rate is photocatalytic rather than solely thermal. In contrast, CeO2/Pd under illumination also exhibits higher Ea (68.6 kJ mol−1), highlighting the critical role of CoO in facilitating charge separation and reducing the reaction energy barrier.11 The smaller bandgap of CeO2/CoO/Pd, as shown by UV-vis DRS measurement, has improved the availability of charge carriers, necessitating less activation energy for the photocatalytic reaction, as shown in Fig. 3g. This noticeable relationship between Eg and Ea indicates that bandgap narrowing not only promotes wide light absorption but also improves utilization of photogenerated electrons and holes, thus reducing the energy barrier for the catalytic cycle. The surface-active sites act as entrapping centers for the e−/h+ pairs, thereby reducing their recombination rate and facilitating their effectual participation in the reaction. This represents that the bimetallic CeO2/CoO active sites have significantly lowered the bandgap and Ea for the reaction, subsequently promoting the rate of the SMC coupling reaction.71 Furthermore, the activation energy (Ea) under illumination is considerably lower than that observed in darkness, showing that visible-light photoexcitation dramatically reduces the energy barrier for the reaction and is the primary factor in the improved catalytic performance. Thus, the ability to proficiently integrate thermal and light energy sources to enhance the efficiency of chemical transformations renders this catalytic system superior to traditional photocatalytic systems and conventional thermal catalysts. Furthermore, the rate constant values are used to investigate the initial rate enhancement using CeO2/CoO/Pd and CeO2/Pd with the increase in temperature (Table 2). We have observed that the rate enhancement (13.61) is more pronounced at low temperatures as compared to 70 °C (7.66), which signifies the catalyst efficiency at low temperatures. The rate enhancement of CeO2/CoO/Pd at low temperature is more significant, representing the amplifying role of CoO in the catalytic system as the reaction with CeO2/Pd is markedly sluggish or insignificant at low temperatures.
| Temperature (°C) | k (CeO2/CoO/Pd) (μM min−1) | k (CeO2/Pd) (μM min−1) | Rate enhancement |
|---|---|---|---|
| 30 | 0.26 | 3.54 | 13.61 |
| 40 | 0.53 | 6.54 | 12.34 |
| 50 | 1.51 | 12.08 | 7.99 |
| 60 | 4.01 | 28.61 | 7.13 |
| 70 | 4.95 | 37.95 | 7.66 |
To understand the light-stimulated activation of Pd2+ ions on the catalyst surface, the coupling reaction has been conducted in the presence of electron and hole scavengers.72 The results are described in Table 3. Fig. 3h shows the effect of the electron and hole scavengers on the reaction rate. The SMC reaction is performed under optimized conditions between 4-bromotoluene and phenylboronic acid with benzoquinone (electron scavenger). The conversion of 4-bromotoluene to the cross-coupled product is virtually impeded in the presence of benzoquinone under light-stimulated conditions, demonstrating that the photoexcited electrons initiated the oxidative addition of 4-bromotoluene (R1–X) to Pd0, forming the R1–Pd–X bond. Fig. 3i illustrates the mechanistic role of photoexcited e−/h+ in the SMC reaction. Meanwhile, the conversion of 4-bromotoluene to the coupled product is nearly the same when the reaction is conducted at high temperature (100 °C) even in the presence of benzoquinone. The catalyst activity at high temperature in the presence of benzoquinone shows that Pd2+ ions capture the more thermally excited electrons moving across the CeO2 to CoO on the surface to form Pd0. Nonetheless, the selectivity of the cross-coupled product decreases from 99% to 67% in the presence of benzoquinone. The decrease in selectivity is attributed to the role of benzoquinone in obstructing the thermally excited electrons from being grasped by the surface of Pd2+ ions, and hence the pathway of Pd0 activation by electrons is hindered. Therefore, the yield of the self-coupled product of phenylboronic acid is high, which has lowered the selectivity of the reaction.72 A control experiment was performed at 100 °C without adding any scavenger to differentiate between photocatalytic and merely thermal effects. The reaction achieved a conversion of almost 99%, representing the thermal effect on the rate of reaction. However, the selectivity decreased to 78%. The decrease in selectivity for the cross-coupled product is attributed to the increased kinetic energy of the reactants at higher temperature, encouraging the formation of the self-coupled product. This outcome verifies that while thermal activation can accelerate the rate of reaction, it challenges the selectivity of the cross-coupling process. This contrast verifies that the light-induced amplification is mostly facilitated by photogenerated electrons rather than being only attributable to localized heating or thermal effects. This result indicates that even under thermal conditions, surface-mediated electron transfer to Pd2+ contributes to maintaining catalytic reactivity, whereas the partial suppression of this process by benzoquinone highlights the significance of surface electron dynamics in photothermal contexts.
| Entry | Reaction condition | Benzoquinone | TEA | Conversion (%) | Selectivity (%) |
|---|---|---|---|---|---|
Reaction conditions: 4-bromotoluene (0.5 mmol), phenylboronic acid (0.55 mmol), K2CO3 (1.5 mmol), catalyst (20 mg), EtOH/H2O (1 : 1, 4 mL), time 30 min, Xe lamp (150 W) equipped with an optical filter (cut off <420 nm). |
|||||
| 1. | Light | 50 mg | — | 17 | 84 |
| 2. | 100 °C | 50 mg | — | 91 | 67 |
| 3. | Light | — | 50 mg | 59 | 47 |
| 4. | 100 °C | — | 50 mg | 87 | 79 |
| 5. | 100 °C | — | — | 99 | 78 |
The next experiment is performed to investigate the function of photogenerated holes in the activation of phenylboronic acid. Phenylboronic acid in the reaction solution reacts with potassium carbonate (base) to produce anionic species, which are adsorbed by electrostatic interaction onto the catalyst's surface (CeO2/CoO/Pd). On exposure to light, the photogenerated holes diffuse to the adsorption site of aryl boronic acid on the catalyst surface and trigger the oxidation of the adsorbed phenylboronic acid molecules, resulting in the cleavage of the C–B bond. Therefore, in the photocatalytic SMC reaction, phenylboronic acid is activated by holes, and this phenomenon has been experimentally verified by performing the SMC reaction in the presence of triethanolamine (TEA) as a hole scavenger.45 The conversion of 4-bromotoluene to the SMC cross-coupled product decreased from 99% to 59%, whereas the appearance of the self-coupled product of 4-bromotoluene shows that phenylboronic acid is unable to couple with the aryl halide in the presence of a hole scavenger (TEA) under light conditions. The impact of holes on the activation of phenylboronic acid in SMC reactions is also investigated by conducting the reaction in the presence of a hole scavenger (TEA) at 100 °C in the absence of light. The comparable yield (87%) and selectivity (79%) of the cross-coupled product in this reaction suggest that TEA has no impact on the reactivity or selectivity of the coupled product in the dark. These results have affirmed that phenylboronic acid undergoes activation by photo-induced holes on the surface of CeO2/CoO/Pd upon exposure to light irradiation. Fig. 3i also shows the mechanistic proposal of the flow of electrons in the composite catalytic system CeO2/CoO/Pd. O2−(ads.) participates in the redox cycle. Surface oxygen donates electrons to reduce Co3+ → Co2+, generating oxygen vacancies and releasing O2 on exposure to light. The photogenerated electrons in the CB of Co2+ readily move to the surface-adsorbed Pd2+, thereby reducing it into Pd0 and accelerating it for the oxidative addition step of aryl halide (R1–X), forming the R1–Pd–X complex. In addition to activity, recyclability is a crucial element for assessing the performance of catalysts. Following the reaction completion, the catalyst is separated from the concoction using the centrifugal method, washed extensively with water and ethanol to remove impurities, dried in a vacuum oven at 60 °C, and used for the next reaction with a fresh batch of the reaction mixture. Even after five cycles, the yield of the coupled product in the SMC standard reaction is still above 90%. This demonstrates the catalyst's exceptional cycle stability (Fig. S11a). Furthermore, the hot filtration test investigates the heterogeneity of the catalyst. The catalyst is removed from the reaction concoction after 15 min by centrifugation and filtration. The supernatant is collected, poured into the reaction vessel again, and stirred for another 1 h. The reaction aliquots are collected every 15 min, diluted with EtOAc and brine, extracted in EtOAc, and investigated by GC by adding nitrobenzene as an internal standard. It is important to note that no more coupled product is produced even after 1 h of reaction, ensuring that no leached Pd contents are present in the solution after filtration, which can act as a homogenous catalyst during the reaction (Fig. S11b). This experiment has affirmed the catalyst's heterogeneous nature and also shows that the designed catalytic system is extremely stable, in which Pd NPs strongly adhered to the CeO2/CoO hybrid support without forming palladium black in the solution or palladium agglomerate during the reaction process.73 Furthermore, ICP-MS analysis of the hot filtrate indicated a Pd concentration of 0.004 mg L−1, representing approximately 0.4% of the total Pd initially contained in 20 mg catalyst, thereby confirming that Pd leaching under the reaction conditions is nearly negligible. The morphological characteristics of the recycled CeO2/CoO/Pd catalyst are examined by XRD and HRTEM. The XRD analysis of CeO2/CoO/Pd catalyst, both pre- and post-recycled catalyst, showed comparable results, indicating that the structure and oxidation state of the catalyst are essentially the same after the reaction (Fig. S12a). HRTEM is subsequently conducted on the recycled catalyst (after five consecutive cycles), showing the diminutive change in the particle size of Pd NPs that is ascribed to a degree of agglomeration occurring during the reaction process, signifying the slight loss in catalytic activity after every run. Fig. S12c presents the HRTEM image of the CeO2/CoO/Pd, where lattice fringes of each element are distinctly visible, and the interplanar distance measured 0.317 nm, 0.25 nm, and 0.225 nm, corresponding accurately to the (111) lattice plane of Ce, (111) lattice plane of Co and (111) lattice plane of Pd (Fig. S12b and c). Elemental mapping is conducted on a recycled CeO2/CoO/Pd sample, revealing a homogeneous distribution of Ce, Co, O, and Pd even after recycling 5 times (Fig. S12d–h). The findings demonstrated that the CeO2/CoO/Pd catalyst maintained stability following the aforementioned process, exhibiting a slight loss of activity.
The SMC reaction with several other activated and non-activated aryl halide substrates is carried out under optimized reaction conditions (Table 4). Good to excellent yields are achieved using the designed catalyst under light-stimulated conditions. In the activated aryls (having electron-withdrawing groups), entries 1, 2, 4, 9 and 13–15 (Table 4) have given 100% yield under light irradiation conditions with TOF values of 2703, 1802, 5404, and 4095 h−1, respectively.74 However, with deactivated aryl halides (having electron-donating groups), i.e., entries 3, 7, 11 and 12, the Pd complex has given good quantitative conversion with 99–96% yield of the coupled product, having a TOF of 1311, 1730, 2675, 2648 h−1 (Table 4). Table S3 shows a detailed comparative analysis of the current work with the literature. To summarize, CeO2/CoO/Pd has demonstrated excellent catalytic activities in a short reaction time (30 min) and can tolerate a wide range of functional groups.75,76
| Entry | X | R | Temp (°C) | Light condition | |||
|---|---|---|---|---|---|---|---|
| Time (min) | Conversion (%) | Selectivity (%) | TOF (h−1) | ||||
| Reaction conditions: aryl halide (0.5 mmol), phenylboronic acid (0.55 mmol), K2CO3 (1.5 mmol), catalyst (20 mg), EtOH/H2O, Xe lamp (150 W) equipped with an optical filter (cut off <420 nm). | |||||||
| 1 | I | –H | 25 | 30 | 100 | 99 | 2703 |
| 2 | I | –CH3 | 25 | 45 | 100 | 99 | 1802 |
| 3 | I | –OCH3 | 25 | 60 | 97 | 98 | 1311 |
| 4 | I | –CN | 25 | 15 | 100 | 99 | 5405 |
| 5 | Br | –H | 50 | 20 | 99 | 99 | 4054 |
| 6 | Br | –CH3 | 50 | 30 | 98 | 99 | 2675 |
| 7 | Br | –OCH3 | 50 | 45 | 96 | 98 | 1730 |
| 8 | Br | –CHO | 50 | 15 | 99 | 98 | 5351 |
| 9 | Br | –NO2 | 50 | 15 | 100 | 99 | 5405 |
| 10 | Br | –COCH3 | 50 | 15 | 98 | 98 | 5297 |
| 11 | Br | –NH2 | 50 | 30 | 99 | 99 | 2675 |
| 12 | Br | –OH | 50 | 30 | 98 | 99 | 2648 |
| 13 | Br | –COOH | 50 | 15 | 100 | 99 | 5405 |
| 14 | Br | –CN | 50 | 20 | 100 | 99 | 4095 |
| 15 | Br | –CONH2 | 50 | 20 | 100 | 99 | 4095 |
| 16 | Br | –COOCH3 | 50 | 20 | 99 | 99 | 4054 |
| 17 | Cl | –H | 50 | 120 | 51 | 82 | 270 |
| 18 | Cl | –CH3 | 50 | 120 | 39 | 88 | 203 |
| 19 | Cl | –COOH | 50 | 120 | 63 | 90 | 338 |
| 20 | Cl | –CN | 50 | 120 | 70 | 92 | 365 |
In situ diffuse reflectance infrared Fourier-transform spectroscopy (DRIFTS) analysis has been further performed to analyze the reaction mechanism. In situ FTIR gives real-time information about molecular dynamics on the catalyst surface and enables continuous monitoring of the reaction progress over time. In situ FTIR has verified the successful coupling of phenylboronic acid and 4-bromotoluene. The reaction intermediates appeared progressively as the time of light exposure increased. The distinct IR absorption signals undergo a shift or weaken the reactant peaks, and the emergence of new signals has verified the formation of the coupled product under light exposure. Fig. 4a illustrates that the peak at 2900 cm−1 along with its shoulder peak at 2980 cm−1 becomes more obvious as the reaction proceeds, indicating the stretching vibration of the C–H bond in –CH3 in the coupled product. The prominent peak at 1470 cm−1 signifies C–C stretching in the phenyl ring that has shifted to 1450 cm−1 due to the coupling of two phenyl rings (Fig. 4b). Notably, the C–H peak intensity progressively enhanced as the time of light exposure increased, certifying a photocatalytic SMC reaction. The IR signals in the range of 1650–1750 cm−1 demonstrate the stretching vibration of aromatic C
C bonds, which further affirms the formation of the biphenyl product from the SMC reaction. Another noticeable signal at 1050 cm−1 is attributed to the stretching vibration of C–O due to solvent (ethanol) interaction with the coupled product, indicating the deformation vibration of the –CH3 group in the product due to the solvation effect. Fig. 4c illustrates the plausible pathway for the SMC reaction catalyzed by the CeO2/CoO/Pd catalyst. The CeO2/CoO bimetallic hybrid system absorbed photons of light following the electron excitation from VB to CB, generating an electron/hole pair for conducting the SMC reaction. Indeed, the catalytic efficacies of active centers in the SMC reaction are significantly influenced by their electronic states. Since the catalyst must act as a nucleophile during the oxidative addition phase and as an electrophile during both the transmetalation and the reductive elimination phases of the coupling reaction.71,77 The integration of cobalt oxide into the CeO2 structure has been proposed to reduce the bandgap and promote the formation of OVs, thus augmenting the concentration of surface-adsorbed oxygen (Oads) and enhancing the mobility of oxygen, as evidenced by XPS analysis of O 1s spectra. These oxygen vacancies are stabilized by the reversible redox process of Co2+ + Ce4+ ↔ Co3+ + Ce3+. In this hybrid system, CoO has also enabled the more effective absorption of light due to the narrowing of the bandgap of the bimetallic composite support (CeO2/CoO) compared to pristine CeO2, and also acts as an electron shuttle for the transfer of excited electrons between CeO2 and Pd. The addition of cobalt oxide to CeO2 has introduced impurity energy levels in between the CB and the VB of the original bandgap of CeO2, which function as intermediate energy levels to capture photoexcited electrons, thus stabilizing the redox cycle between Co2+ and Pd2+. The e− emigrated from the O2−(abs) of CeO2 to the empty 3d orbital of Co, which is lower in position than the CB of Ce 4f, signifying less energy requirement for excitation. The photogenerated electrons in the CB of Co readily move to the surface-adsorbed Pd2+, thereby reducing it into Pd0 and accelerating it for the oxidative addition step of aryl halide (R1–X), forming the R1–Pd–X complex. Conversely, the defective surface structure of CeO2 hosts several oxidative species (O2−/O2−), which are readily transported from the CeO2 surface to the neighboring interface of CoO, facilitating the redox cycle of the Co3+/Co2+ (highly active sites) surface. The kinetic energy is provided by the heating effect of the reaction system to the photoexcited electrons, further averting the recombination rate of e−/h+ pairs. The holes formed in the valence band of ceria (CeO2) are retained by the VB of O 2p and activate the aryl boronic acid for the coupling cycle. In fact, the aryl boronic acid in the basic media generates anionic species and is adsorbed on the catalyst surface through electrostatic interaction. However, on exposure to light, electron/hole pairs are generated, diffusing the holes to the adsorption site of aryl boronic acid. The migration of holes to the adsorption site of aryl boronic acid results in the release of the oxidized boronic acid complex that reacts with R1–Pd–X to exchange the halide group with another aryl group from boronic acid and forms the R1–Pd–R2 complex. Finally, reductive elimination releases the coupled (R1–R2) product. The Pd2+ ions on the catalyst surface function as a strong electron acceptor, promoting the rate of transmetalation and reductive elimination step. In this context, the concurrent utilization of photo and thermal energy enhances the pace of the coupling reaction. In brief, the synergistic impact of CeO2/CoO/Pd in the heterojunction significantly amplifies the photocatalytic activity of the catalytic system, where CeO2 provides stability and keeps the e−/h+ pairs separated, and CoO has improved light absorption and provides shuttle service to the electron transfer process between CeO2 and Pd NPs.
![]() | ||
| Fig. 4 (a and b) In situ FTIR spectrum of light-stimulated SMC reaction using the CeO2/CoO/Pd catalyst. (c) Mechanism for the Suzuki coupling reaction catalyzed by the CeO2/CoO/Pd catalyst. | ||
Supplementary information is available. See DOI: https://doi.org/10.1039/d5cy00994d.
| This journal is © The Royal Society of Chemistry 2025 |