Yifan
Yan‡
a,
Xi
Cai‡
a,
Jiangrong
Yang‡
a,
Yu
Fu
a,
Qiwei
Shi
a,
Pengjie
Hao
a,
Hua
Zhou
a,
Zhenhua
Li
*ac,
Mingfei
Shao
*ac and
Haohong
Duan
*bde
aState Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, P. R. China. E-mail: LZH0307@mail.buct.edu.cn; shaomf@mail.buct.edu.cn
bDepartment of Chemistry, Tsinghua University, 30 Shuangqing Rd, Haidian Qu, Beijing 100084, P. R. China. E-mail: hhduan@mail.tsinghua.edu.cn
cQuzhou Institute for Innovation in Resource Chemical Engineering, Quzhou 324000, P. R. China
dHaihe Laboratory of Sustainable Chemical Transformations, Tianjin 300192, China
eEngineering Research Center of Advanced Rare Earth Materials (Ministry of Education), Department of Chemistry, Tsinghua University, Beijing 100084, China
First published on 2nd May 2025
α,β-Unsaturated ketones, crucial in organic synthesis and life sciences, are conventionally produced through aldol condensation of ketones and aldehydes. However, traditional synthesis methods involve high temperature, pressure, and the use of environmentally harmful solvents, hindering sustainable development. Herein, we present one-step electrosynthesis of benzylidene acetones and 2-methylenephenyl cyclohexanone via tandem reactions, by coupling electrooxidation of benzylic alcohols to the corresponding aldehyde, followed by aldol condensation between the aldehyde and the ketone. Selective formation of benzaldehydes is key to the tandem reaction and was achieved over a cubic oxide-supported gold catalyst (Au/CuO) as the anode, showing the ability to adsorb benzylic alcohols and generate the active adsorbed oxygen species (OH*) for selective oxidation. The tandem reaction strategy demonstrates its versatility in the synthesis of α,β-unsaturated ketones from benzyl alcohols with different substituents and acetone/cyclohexanone. As proof of concept, we constructed a flow electrolyzer and achieved continuous electrosynthesis of benzylidene acetone coupled with H2 production at ampere-level current, delivering a benzylidene acetone productivity of 9.5 mmol h−1 and a H2 productivity of 0.4 L h−1. This study demonstrates the potential of coupling electrocatalysis and thermocatalysis in tandem, with implications for synthesis of more value-added chemicals.
Green foundation1. This work describes a sustainable tandem electrocatalytic strategy for one-step synthesis of α,β-unsaturated ketones, eliminating the need for high temperatures, high pressures, and hazardous organic solvents traditionally required for aldol condensation. The use of renewable electricity further enhances its green chemistry potential.2. The electrocatalytic oxidation of aromatic alcohols over an Au/CuO catalyst achieves high benzylidene acetone productivity (9.5 mmol h−1) with 84% selectivity, while simultaneously producing 0.4 L h−1 of H2 as a valuable byproduct. This approach demonstrates a high AE (89%) and a low E-factor (0.71), making it a more sustainable alternative. 3. Further research could focus on renewable feedstock integration, optimizing catalyst durability, and scaling up flow electrolyzers to industrial levels, which may improve efficiency and overall sustainability. |
![]() | ||
Scheme 1 Tandem coupling of alcohol electrooxidation with aldol condensation. Ar: aromatic substituent. |
Electrocatalytic synthesis is regarded as a promising approach for producing high-value-added compounds, thanks to its mild electrochemical reaction conditions and the possibility to integrate renewable electricity.12–16 Very recently, tandem electrocatalysis has been reported for the production of value-added chemicals, by combining electro- and thermo-catalytic processes.17–21 For example, Zhang and coworkers reported electrosynthesis of cyclohexanone oxime via tandem electroreduction of NO2− to NH2OH* (as an intermediate) and the cyclohexanone-hydroxylamine reaction.21 Zou and coworkers, in a tandem electrochemical-chemical-electrochemical reaction, coupled the electroreduction of NO3− to NH2OH with its chemical reaction with pyruvic acid to produce pyruvic oxime, which was electroreduced to alanine.18 In addition, our group reported the electrosynthesis of lactic acid from glycerol, a byproduct of biodiesel, using tandem reactions. These reactions include the electrooxidation of glycerol to glyceraldehyde or dihydroxyacetone intermediates, followed by base-catalyzed dehydration and the Cannizzaro rearrangement, ultimately producing the biodegradable plastic monomer lactic acid.17
Inspired by previous achievements in tandem reactions, we envisaged that high selectivity for electrooxidation of alcohols to aldehydes and efficient aldol condensation are two key points for one-step synthesis of α,β-unsaturated ketones.10,22 Herein, we report the selective electrooxidation of benzylic alcohols to the corresponding benzaldehydes using a cooperative Au/CuO catalyst, followed by spontaneous aldol condensation between the benzaldehyde and a ketone (acetone and cyclohexanone) in an alkaline environment, efficiently producing the corresponding α,β-unsaturated ketones in one step (Scheme 1b). To maximize electron economy and energy efficiency, we further integrated the electrosynthesis of benzylidene acetone at the anode with its electroreduction at the cathode to a saturated ketone, 4-phenyl-butane-2-one. This study underscores the great potential of coupling electro- and thermo-catalysis in a tandem manner, paving the way for enhanced synthesis of valuable chemicals.
For the preparation of the Au/CuO catalyst, CuO nanosheets were first grown on Ni foam via a hydrothermal method, followed by a calcination process at 453 K. Then, Au nanoparticles (NPs) were electrodeposited on CuO to obtain Au/CuO (Fig. 1a). Pure CuO nanosheets and Au NPs on Ni foam were also synthesized with the same methods and act as reference samples (Fig. S1, ESI†). Fig. S2 (ESI)† shows the X-ray diffraction (XRD) patterns of CuO and Au/CuO. The diffraction peaks at 35.3° and 38.5° are attributed to the (002) and (111) planes of CuO, respectively (JCPDS #44-0706). The peaks at 44.5° and 51.8° are attributed to the Ni foam substrate (JCPDS #04-0850). For Au/CuO, the peak at 38.2° is broader than that of CuO, which can be attributed to the (111) plane of face-centered cubic Au (JCPDS #04-0784). Scanning electron microscope (SEM) images (Fig. 1b) show that the CuO nanosheets are vertically aligned on the Ni foam, with an average thickness of ∼20 nm and a diameter of 300–400 nm. The energy-dispersive spectroscopy (EDS) mapping results show that Au NPs are uniformly distributed on the surface of the CuO nanosheets with an average diameter of 15 nm (Fig. 1c). High-resolution transmission electron microscopy (HRTEM) images show that both Au and CuO are well crystallized and exposed to the (111) plane (Fig. S3, ESI†).
We then investigated the effect of oxidation potential on benzyl alcohol oxidation over Au/CuO via chronoamperometric (CA) tests. The reaction products were quantitatively analyzed by high-performance liquid chromatography (HPLC). As shown in Fig. S5 (ESI),† the conversion rate of benzyl alcohol over Au/CuO increases gradually with the increase of the oxidation potential from 0.96 to 1.56 V vs. RHE, reaching its highest at 1.56 V vs. RHE with a rate of 0.92 mmol cm−2 h−1. In contrast, benzaldehyde selectivity decreases with the increase of reaction potentials (from 73% to 47%) due to the formation of benzoic acid via overoxidation (Fig. S5(c) and S6, ESI†). By collectively considering the reaction rate and benzaldehyde selectivity, 1.16 V vs. RHE was selected as the optimal reaction potential. Additionally, we observed that reducing the KOH concentration enhances benzaldehyde selectivity (Fig. S7, ESI†); however, it markedly decreases the reaction rate. Considering both factors, we selected 0.1 M KOH as the electrolyte.
We then compared the performance of electrooxidation of benzyl alcohol over the Au, CuO and Au/CuO catalysts. As shown in Fig. 1e, CuO exhibits a high benzaldehyde selectivity of 95% at 1.16 V vs. RHE, while its benzyl alcohol conversion rate is only 0.0024 mmol cm−2 h−1. The benzyl alcohol conversion rate over Au reaches 0.066 mmol cm−2 h−1, together with a benzaldehyde selectivity of 81%. On Au/CuO, the conversion rate of benzyl alcohol achieves 0.44 mmol cm−2 h−1, which is 6.7-fold and 183-fold higher than that on Au and CuO, respectively, while maintaining a relatively good benzaldehyde selectivity of 67%. We also examined the role of nickel foam, as shown in Fig. S8 (ESI),† confirming that Ni foam functions solely as a substrate in this study.
To demonstrate this assumption, we investigated the effect of acetone addition to the electrolyte on the electrocatalytic performance of Au/CuO. As shown in Fig. 2a, the current density for the OER and benzyl alcohol oxidation over Au/CuO remained nearly unchanged after acetone addition, indicating that acetone is stable and does not participate in the electrocatalytic process. The slight decrease in current may be due to increased electrolyte resistance. Previous studies have also shown that acetone is stable under alkaline conditions and does not polymerize in a short time.32
We subsequently verified the spontaneous aldol condensation between benzaldehyde and acetone in 0.1 M KOH. As shown in Fig. S9 (ESI),† 0.15 M benzaldehyde spontaneously converts entirely to benzylidene acetone with acetone within 20 minutes, significantly faster than benzyl alcohol electrooxidation (0.44 mmol cm−2 h−1 over Au/CuO), suggesting that the in situ produced benzaldehyde can be quickly consumed by reacting with acetone.
We then carried out CA measurements at 1.16 V vs. RHE to verify the feasibility of electrooxidation of benzyl alcohol to produce benzylidene acetone. As shown in Fig. 2b and S10 (ESI),† benzylidene acetone was successfully synthesized during the benzyl alcohol oxidation process in 0.1 M KOH, and the selectivities of benzylidene acetone over Au, CuO and Au/CuO are all higher than 80% (the byproduct is benzoic acid). Notably, for Au/CuO, the selectivity of benzylidene acetone in the presence of acetone (84%) is significantly higher than that of benzaldehyde without acetone (67%), which may be due to the facile aldol condensation reaction between acetone and benzaldehyde, lowering the amount of benzaldehyde being oxidised to benzoic acid. In terms of reaction rate, Au/CuO shows a benzylidene acetone productivity of 0.35 mmol cm−2 h−1, which is 9-fold and 159-fold higher than that of Au and CuO, respectively. Moreover, the yield of benzylidene acetone reached 84% when the conversion of benzyl alcohol was higher than 98% (Fig. 2c), suggesting that the reaction strategy has potential for practical application that requires high conversion.
According to the above experimental results, we propose the following reaction path for the benzylidene acetone electrosynthesis. As shown in Fig. 2d, benzyl alcohol is first oxidized to the benzaldehyde intermediate via an electrocatalytic process. The generated benzaldehyde intermediate rapidly reacts with the excess acetone in the solution, producing benzylidene acetone through aldol condensation while preventing the overoxidation of benzaldehyde. For the stability of the Au/CuO catalyst, the preliminary results show that the conversion rate of benzyl alcohol and the selectivity of benzylidene acetone were largely maintained after 10 batches (overall 10 h; Fig. S11, ESI†), together with the preservation of the original nanosheet structure (Fig. S12, ESI†), demonstrating high stability.
To demonstrate the generality of this one-step electrosynthesis of α,β-unsaturated ketones from alcohols, benzyl alcohols with diverse substituents (–Cl, –F, –Br, –CH3, and –OCH3) were examined. The electrooxidation reactions were performed at 60 °C to guarantee the full dissolution of alcohols in 0.1 M KOH (Fig. S13, ESI†). As shown in Fig. 2e, benzylidene acetones with diverse substituents (–Cl, –F, –Br, –CH3, and –OCH3) were successfully synthesized in 0.1 M KOH with 0.3 M acetone (Fig. S14, ESI†), together with high selectivity (>60%). Additionally, Au/CuO exhibits a 1.8 to 4.8-fold enhancement in productivity compared with Au for these α,β-unsaturated ketones, indicating the excellent performance of Au/CuO to produce α,β-unsaturated ketones. Moreover, we replaced acetone with cyclohexanone and successfully obtained 2-phenylmethylene cyclohexanone from benzyl alcohol in one step with a selectivity of 91% (Fig. S15, ESI†), further demonstrating the generality of this strategy.
To further elucidate the mechanism behind the high performance of Au/CuO, we calculated the turnover frequency (TOF) value to investigate the intrinsic activity of Au.14 As shown in Fig. S16 (ESI),† the TOF value of Au/CuO is calculated to be 0.34 s−1 at 1.16 V vs. RHE, which is 4.8-fold higher than that of pure Au, indicating that the excellent activity of Au/CuO stems from the synergistic effect of Au and CuO. Due to the much higher electrochemically active surface area (ECSA) of Au/CuO compared to Au (Fig. S17, ESI†), we normalized the conversion rate of benzyl alcohol by the ECSA to further exclude the possibility that the excellent performance of Au/CuO arises from its higher ECSA. As shown in Fig. S18 (ESI),† the normalized conversion rate of benzyl alcohol over Au/CuO is 1.9 times higher than that of Au, ruling out the possibility that the high activity of Au/CuO is from its nanosheet structure.
Previous studies have shown that in alcohol electrooxidation, Au adsorbs OH− to generate OH*, thereby oxidizing the alcohol molecule for dehydrogenation.15 We first compared the OH− adsorption ability of Au and Au/CuO via zeta potential tests.33 As shown in Fig. 3b, Au/CuO shows a lower zeta potential value than Au, showing that Au/CuO can adsorb more OH− in the inner Helmholtz layer. Then we investigated the ability of Au and Au/CuO to produce OH*. To ensure accurate detection of the OH* signal, Au and Au/CuO were deposited onto a glass carbon electrode for the electrochemical tests. As shown in Fig. 3c, during the anodic scan in 0.1 M KOH, Au exhibits two oxidation peaks, which are attributed to the adsorption of OH− to form Au–OH* (from ∼0.85 V vs. RHE) and its further oxidation to AuOx.34 For Au/CuO, three oxidation peaks were observed. The first and second anodic peaks are attributed to the oxidation of Cu0 to Cu1+ (from 0.5 to 0.7 V vs. RHE) and its further oxidation to Cu2+ (from 0.7 to 0.9 V vs. RHE).35–37 The intensity of the third oxidation peak, attributed to Au–OH*, is significantly higher than that of pure Au, indicating that the ability to generate OH* on Au/CuO is stronger than on Au.
After adding 0.15 M benzyl alcohol, the oxidation current of both Au and Au/CuO begins to increase from ∼0.8 V vs. RHE (Fig. 3c inset), which is due to the oxidation of benzyl alcohol by Au–OH*.38 Note that the current density with Au/CuO is significantly higher than that with Au, demonstrating the better performance of Au/CuO for benzyl alcohol oxidation. The oxidation current of Au and Au/CuO diminishes at ∼1.45 V vs. RHE because of the formation of AuOx.
According to previous reports, the electron deficiency of Au is theoretically more conducive to adsorbing OH−, which may be more beneficial for generating OH*.14,39,40 Therefore, we investigated the electronic structure of Au/CuO using X-ray photoelectron spectroscopy (XPS). As shown in Fig. S19 (ESI),† the peaks of Au 4f over Au/CuO shift ∼0.2 eV towards higher energy compared to Au. Conversely, the peaks of Cu 2p in Au/CuO shift negatively by ∼0.2 eV relative to Cu. Previous studies have shown that loading Au onto other oxyhydroxides would also transfer electrons from Au to oxyhydroxides, thereby forming electron-deficient Au, which could promote OH* production and thus improve alcohol oxidation activity.28 To verify the specificity of Au/CuO, we constructed Au/NiOOH and Au/CoOOH catalysts and investigated their performance. XPS results show that there is an electron transfer at Au/NiOOH and Au/CoOOH interfaces from Au to NiOOH/CoOOH, forming electron-deficient Au (Fig. S20, ESI†). However, the performance of Au/CuO is still significantly better than that of the other catalysts (Fig. 3d and S21, ESI†), implying that the CuO support also plays a key role in the superior performance of Au/CuO.
We subsequently investigated the benzyl alcohol adsorption ability of Au/CuO and pure Au. Open circuit potential (OCP) measurements were carried out to detect the change in organic adsorbent content in the inner Helmholtz layer of the electrode.44 As shown in Fig. S22 (ESI),† the OCP changes of Au/CuO are more obvious than those of Au (0.44 vs. 0.37 V) when 0.05 M benzyl alcohol was injected, indicating that more benzyl alcohol molecules are adsorbed in the inner Helmholtz layer of Au/CuO. The enhanced adsorption of benzyl alcohol over Au/CuO was also examined by electrochemical adsorbate-stripping measurements (see the ESI† for experimental details), and the oxidation charges of benzyl alcohol (QOx(BA)) can be used to quantify the amount of adsorbed benzyl alcohol (Fig. S23, ESI†).45 As shown in Fig. 3f, the QOx(BA) of Au/CuO is 2.4-fold higher than that of pure Au, suggesting that Au/CuO has a stronger ability to adsorb benzyl alcohol. Therefore, the significantly enhanced reaction rate of benzyl alcohol oxidation on Au/CuO can be attributed to the synergistic effect of Au and CuO to enhance the generation of OH* and the adsorption of benzyl alcohol (Fig. 3f inset).
The Gibbs free energy changes (ΔGOH*) for the generation of the active oxygen species (OH*) on Au and Au/CuO were calculated for different geometric sites (Fig. 4a); the corresponding models are shown in Fig. S25 (ESI).† The ΔGOH* values of OH* at the vertex site on Au/CuO and Au are lower than that at the edge site, which is attributed to the smaller coordination number of the vertex site. Moreover, the ΔGOH* of OH* at the vertex site on Au/CuO is 0.67 eV, which is lower than that on Au (1.29 eV). The Hirshfeld charge analysis indicated that electrons transferred from Au atoms to CuO, resulting in Au atoms at the Au/CuO interface being more electron-deficient than those in pure Au (Fig. S26, ESI†), consistent with the XPS results. As a result, the nucleophilic molecule H2O and the ions OH− are adsorbed preferentially on electron-deficient Au clusters in Au/CuO to generate OH*.
Moreover, the energies of benzyl alcohol adsorption (Eads) on Au, CuO, and Au/CuO were calculated and are displayed in Fig. 4b. The Eads of benzyl alcohol on Au/CuO is −1.03 eV, which is higher than that on Au (−0.63 eV) and CuO (−0.71 eV), demonstrating that the Au/CuO interface is more favorable for adsorbing benzyl alcohol. The analysis of the density of states (DOS) (Fig. S27, ESI†) and charge density difference (Fig. S28, ESI†) revealed that benzyl alcohol is chemically adsorbed on Au/CuO via a covalent σ bond between the 2p orbital of the O atom in the hydroxyl group of benzyl alcohol and the 3d orbital of Cu in Au/CuO and a π bond between the unoccupied π* orbital of benzyl alcohol and the occupied Au 5d orbital of the Au cluster, thereby strengthening the adsorption of benzyl alcohol at the Au/CuO interface and thus improving the reaction rate (Fig. 4c).28
In addition, we calculated the Gibbs free energy diagrams of the whole reaction process, including the electrooxidation process of benzyl alcohol to benzaldehyde on the catalyst surface and aldol condensation to generate benzylidene acetone in solvent (Fig. 4d). As shown in Fig. 4e, the benzyl alcohol is first adsorbed at the Au/CuO interface with a ΔG of −0.96 eV, which is higher than that on Au (−0.49 eV). Then, the nucleophilic OH* is generated on Au clusters for Au/CuO and Au, which attacks Ph-CH2OH* to generate one molecule of H2O and the adsorbed radical Ph-CH2O*. Afterwards, another nucleophilic OH* is also generated on the Au clusters and reacts with Ph-CH2O* to generate Ph-CHO* and H2O. Then, the benzaldehyde is desorbed from the catalyst into the solvent, followed by aldol condensation with acetone to produce benzylidene acetone. For the electrooxidation process over Au/CuO and Au, the potential-determining steps (PDS) are the benzaldehyde desorption and the generation of OH*, respectively. According to the Brønsted–Evans–Polanyi relation,47 the ΔG for the benzaldehyde desorption from Au/CuO is lower than the ΔGOH* of OH* on Au (0.88 vs. 1.05 eV), which means that the activation energy barrier for this elementary reaction is lower on Au/CuO. Therefore, the theoretical reaction rate is higher on Au/CuO than on Au at the same potential. Note that although the desorption energy of benzaldehyde on Au/CuO is slightly higher than that on Au (0.88 vs. 0.39 eV), the desorption can be overcome by the kinetic energy of benzaldehyde at room temperature, especially under strong stirring, to participate in the subsequent aldol condensation process.
Subsequently, we assembled a stacked membrane-free flow electrolyzer consisting of three units with a total working area of 150 cm2 (Fig. 5b). CA measurements were performed to evaluate the electrosynthesis of benzylidene acetone at industrially relevant current. The electrochemical tests were performed at a constant cell voltage of 1.4 V to exclude the influence of the OER. Moreover, an intermittent potential (IP) strategy, proposed in our previous work,17 was employed to stabilize the catalyst that keeps the absolute current at a high level during long-term testing. As depicted in Fig. 5c and d, the stacked electrolyzer exhibited an absolute current of ∼1 A at 1.4 V, achieving a benzylidene acetone productivity of 103.4 mmol in 10.9 h (0.06 mmol cm−2 h−1) with relatively high selectivity (75%) and FE (60%) (Fig. S30†). Concurrently, cathodic H2 productivity reached 0.4 L h−1 with an energy consumption of 3.2 kWh Nm−3 H2, significantly lower than the advanced water electrolysis system (∼4.2 kWh Nm−3 H2).48
We then evaluated the green chemistry metrics and profitability of the electrochemical synthesis of benzylidene acetone from benzyl alcohol, as presented in this study. Specifically, we calculated key green chemistry indicators, including atom economy (AE), environmental factor (E-factor), and carbon efficiency (CE). Detailed calculation procedures for these metrics are provided in Note S2 (ESI).† The results from the electrochemical synthesis of benzylidene acetone show a high AE of 89%, a low E-factor of 0.71, and a high CE of 74%, suggesting that this synthesis method is potentially green and sustainable. To evaluate the economic market potential of sustainable benzylidene acetone electrosynthesis, a techno-economic analysis (TEA) was conducted (Fig. S31, ESI†).16,49 The results indicate that the profit per ton of benzylidene acetone from this electrocatalytic technology can reach $936 (marked as a white star in Fig. 5e) and further reach $1009 when including the benefit of hydrogen (Note S3, ESI†), demonstrating the economic potential of this technique.
Leveraging the excellent catalytic performance of Au/CuO for the BOR and CoP for the benzylidene acetone reduction reaction (denoted as BRR, Fig. S32, ESI†), we established an electrolytic system for co-producing benzylidene acetone (anode) and 4-phenylbutan-2-one (cathode) in a separate H-type quartz cell. Note that benzylidene acetone at the cathode is externally added, and the product at the anode does not cross the membrane to the cathode. The three coupling systems are denoted as HER//OER, HER//BOR and BRR//BOR, respectively. As shown in Fig. 6b, the current density of BRR//BOR is higher than that of HER//BOR and HER//OER. We then compared the product distribution and productivities in the HER//BOR and BRR//BOR coupling systems with the same electricity input (a constant potential of 1.5 V). As shown in Fig. 6c, for the HER//BOR system, the H2 productivity was 0.13 mmol cm−2 h−1. Additionally, benzylidene acetone was obtained at the anode with a productivity of 0.08 mmol cm−2 h−1 and an FE of 63% (Fig. S33, ESI†), showcasing the advantage of co-producing a high value-added product and H2 fuel compared with traditional water electrolysis. For the BRR//BOR system, benzylidene acetone productivity reaches 0.1 mmol cm−2 h−1, which is 1.3-fold higher than that of the HER//BOR system. Moreover, the productivity of 4-phenylbutan-2-one is 0.09 mmol cm−2 h−1, and the FE is 70% (Fig. S34, ESI†), indicating that the co-production of benzylidene acetone and benzylacetone by electrocatalysis was realized.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5gc01155h |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |