Electrocatalytic urea synthesis from NO and CO2 on In1Pd single atom alloys

Tingting Wu*a, Zihan Tiana, Ziyang Zhanga, Ye Tiana and Ke Chu*b
aCollege of Science, Hebei North University, Zhangjiakou 075000, Hebei, China. E-mail: 1329271270@qq.com
bSchool of Materials Science and Engineering, Lanzhou Jiaotong University, Lanzhou 730070, China. E-mail: chuk630@mail.lzjtu.cn

Received 22nd August 2025 , Accepted 5th September 2025

First published on 5th September 2025


Abstract

Electrocatalytic urea synthesis from CO2/NO co-electrolysis (EUCN) has emerged as a promising strategy for sustainable urea production, while simultaneously mitigating greenhouse gas emissions and NO pollutants. Herein, we have developed single-atom In1 alloyed Pd (In1Pd) as a high-performance EUCN catalyst, delivering a remarkable FEurea of 45.9% and a urea yield rate of 55.2 mmol h−1 g−1 in a membrane electrode assembly electrolyzer. The combination of in situ spectroscopic measurements and theoretical calculations reveals the synergy of In1 and Pd, which enables the co-activation of CO2/NO and their C–N coupling while hampering the competing reactions, leading to greatly enhanced EUCN activity and selectivity.


Urea is a vital nitrogen fertilizer that is widely used in global agriculture.1–3 Currently, the Bosch–Meiser process serves as the primary industrial route for urea synthesis, accounting for over 2% of global energy consumption and generating substantial CO2 emissions.4–8 Electrocatalytic urea synthesis from co-reduction of CO2 and NO (EUCN) has emerged as a promising strategy for sustainable urea production, while simultaneously mitigating greenhouse gas emissions and NO pollutants.9 However, the EUCN process involves a complex multi-step reaction pathway requiring the activation of inert CO2 and NO molecules and enhanced C–N coupling kinetics, posing significant challenges for catalyst design.10 Current catalysts still suffer from low urea faradaic efficiency (FEurea) and poor selectivity due possibly to the competitive side reactions (i.e., the hydrogen evolution reaction (HER)) and independent reduction pathways.9 Therefore, the development of high-performance catalysts with high activity, selectivity and durability is crucial to advance the practical application of EUCN.

Single-atom alloys (SAAs), which combine the merits of single-atom catalysts and alloy catalysts, have garnered significant attention in various electrocatalytic reactions involving both carbon and nitrogen cycles,11–13 thus demonstrating their considerable potential for electrocatalytic urea synthesis from CO2/NO co-electrolysis (EUCN). Among them, Pd-based materials have shown particular efficacy attributed to their unfilled d-electron orbitals (4d10) that facilitate optimized adsorption/desorption behavior of key intermediates.14–16 Nevertheless, Pd-based catalysts often suffer from insufficient co-activation capability toward both nitrogen and carbon sources. Additionally, Pd exhibits a strong tendency for H adsorption,17 which competes with the C–N coupling reaction and diminishes both EUCN activity and selectivity for urea synthesis. Notably, these limitations may be effectively mitigated by incorporating p-block metals such as indium (In). By virtue of its partially occupied p-orbitals, In can intrinsically hinder H binding,18 while promoting CO2 reduction to generate and stabilize C-intermediates beneficial for coupling with N-intermediates towards urea generation.19–21 Motivated by these insights, we propose that single-atom In alloyed with Pd (In1Pd) may serve as a highly promising catalyst, enabling enhanced activity and selectivity for efficient EUCN.

In this study, we developed In1Pd as a highly active and selective EUCN catalyst for urea electrosynthesis. Notably, the In1Pd catalyst exhibits exceptional performance in a membrane electrode assembly (MEA) electrolyzer, delivering a remarkable FEurea of 45.9% and a urea yield rate of 55.2 mmol h−1 g−1 at −0.7 V. The catalytic EUCN mechanism of the In1Pd catalyst is further elucidated through combined in situ spectroscopic measurements and theoretical calculations, revealing that the enhanced EUCN efficiency of In1Pd stems from the synergistic effect of In1 and Pd, which promotes the co-activation of CO2 and NO to *CO/*NH2 and their C–N coupling.

In1Pd was synthesized via a one-pot wet chemistry method. The XRD patterns (Fig. 1a) reveal that In1Pd retains the crystalline structure of pristine Pd (JCPDS No. 65-2867), suggesting that In1 incorporation does not change the crystalline phase of pristine Pd. The TEM image (Fig. 1b) shows that In1Pd exhibits a typical graphene-like morphology. Elemental mapping images of In1Pd (Fig. 1c) illustrate a uniform distribution of In1 atoms on Pd. The coordination structure and electronic characteristics of In1Pd are systematically characterized via X-ray absorption spectroscopy. The XANES analysis at the In K-edge (Fig. 1d) reveals that the absorption edge of In1Pd lies between In foil and In2O3 reference samples, indicating an In oxidation state between 0 and +3, which arises from electron transfer from In1 to Pd due to the lower electronegativity of In (1.78) relative to Pd (2.20). The EXAFS spectra (Fig. 1e) show that In1Pd exhibits a prominent In–Pd coordination peak at 2.75 Å, with no observable In–In or In–Pd coordination bonds, confirming the atomic dispersion of In on the Pd substrate.22–24 The corresponding wavelet transform (WT) analysis (Fig. 1f) displays a single In–Pd coordination peak at 6.5 Å−1, further verifying the monoatomic In dispersion within the Pd matrix.


image file: d5dt02018b-f1.tif
Fig. 1 Characterization of In1Pd: (a) XRD patterns, (b) TEM image, (c) elemental mapping images, (d) In K-edge XANES, (e) EXAFS spectra and (f) WT analyses of In1Pd and reference samples.

The electronic structure of In1Pd is systematically investigated using DFT calculations. Electron density difference analysis (Fig. S1a) reveals a distinct electron transfer from In1 to Pd, in line with the XANES result (Fig. 1d). This interfacial In1–Pd electron interaction guarantees the robust bonding between In1 and the Pd substrate, which endows In1Pd with exceptional structural stability. The corresponding PDOS analysis (Fig. S1b) presents favorable orbital overlapping between the 5p orbital of In1 and the 4d orbital of Pd, offering a theoretical foundation for the efficient charge transfer and robust structural stability.25 Additionally, ab initio molecular dynamics (AIMD) simulation results (Fig. S2) indicate that In1Pd maintains stable energy and temperature profiles throughout the simulation, further validating its excellent thermal stability.26–30

The EUCN performance of In1Pd is assessed in a MEA cell containing 0.1 M KHCO3 catholyte saturated with humidified NO and CO2. Gas and liquid products are quantified by gas chromatography and colorimetry, respectively (Fig. S3). LSV curves (Fig. 2a) show a relatively low current density of In1Pd in sole CO2-saturated electrolyte. Strikingly, the current density is significantly enhanced in the presence of both NO and CO2, suggesting the high catalytic EUCN activity of In1Pd toward urea synthesis. The EUCN performance of In1Pd is quantitatively evaluated after 1 h of electrolysis. Remarkably, In1Pd achieves the highest FEurea of 45.9% at −0.7 V, with a corresponding urea yield rate of 55.2 mmol h−1 g−1 (Fig. 2b), surpassing most recently reported catalysts for urea electrosynthesis (Fig. S4 and Table S1). Control experiments are conducted to validate the nitrogen and carbon sources (Fig. S5). No urea formation is observed under conditions lacking NO or CO2, or at the open-circuit potential (OCP), effectively ruling out the possibility of system contamination as C/N sources of urea. Further 13C and 15N isotope tracing via nuclear magnetic resonance (NMR) spectroscopy, using 13CO2 (Fig. 2c) and 15NO (Fig. 2d) as tracers, reveals the characteristic signals of 13CO(NH2)2 and CO(15NH2)2, respectively, further confirming that the synthesized urea originates exclusively from the EUNC process catalyzed by In1Pd.


image file: d5dt02018b-f2.tif
Fig. 2 (a) LSV curves of In1Pd under different conditions. (b) Urea yield rates and FEurea of In1Pd at various potentials. (c) 13C NMR spectra of the 13CO(NH2)2 standard sample and those fed by 13CO2 after electrolysis at −0.7 V. (d) 1H NMR spectra of the CO(15NH2)2 standard sample and those electrolyzed in 15NO electrolyte at −0.7 V. (e) Urea yield rates and FEurea during eight cycling tests at −0.7 V. (f) Comparison of urea yield rates and FEurea between pristine Pd and In1Pd at −0.7 V.

The selectivity of In1Pd toward ECNU is evaluated by quantifying the FEs of other byproducts (CO, H2 and NH4+). At the optimal applied potential of −0.7 V, FEurea remains significantly higher than the FEs of all byproducts (Fig. S6), confirming the exceptional ECNU selectivity of In1Pd for urea synthesis. For stability evaluation, we conducted an eight-cycle test, which shows that both urea yield rate and FEurea exhibit minimal fluctuations (Fig. 2e), confirming the excellent catalysis durability of In1Pd.30–32 Comparative analysis shows that the pristine Pd catalyst (Fig. 2f) exhibits a much inferior ECNU performance relative to In1Pd, highlighting the critical synergistic interaction between In1 and the Pd substrate in boosting the ECNU activity.

To elucidate the fundamental understanding of the significantly improved EUCN performance of In1Pd, we employed in situ FTIR and online differential electrochemical mass spectrometry (DEMS) to identify the reaction intermediates. First, in situ FTIR measurements are conducted on In1Pd over the potential range from the OCP to −0.7 V. As shown in Fig. 3a–c, the characteristic peak at 1396 cm−1 is assigned to the symmetric stretching vibration of the *COOH intermediate, while the C[double bond, length as m-dash]O stretching peak at 2017 cm−1 corresponds to the generated *CO intermediate33 (Fig. 3b). Notably, additional peaks observed at 1695 cm−1 and 1433 cm−1 are attributed to *CONH2 intermediates and C–N bonds,34–36 respectively (Fig. 3c), indicating that EUCN proceeds via *NH2 + *CO → *NH2CO. Meanwhile, the enhanced peak intensity of urea (1610/1195 cm−1) with increasing potentials indicates that the generated *NH2CO intermediates are readily converted into urea.35 In addition, online DEMS measurements (Fig. 3d) show the prominent m/z signals corresponding to key intermediates and products, including *CONH2 (m/z = 44), *CO(NH2)2 (m/z = 60), *NO (m/z = 30), *NH2 (m/z = 16), *CO (m/z = 28) and *COOH (m/z = 45). These DEMS results are in line with in situ FTIR data, collectively providing compelling evidence that In1Pd facilitates the efficient co-reduction of CO2 and NO to urea through a sequential relay catalysis mechanism (Fig. S7), where In1Pd first promotes the co-activation of CO2 and NO, forming critical *CO and *NH2 intermediates. These *CO/*NH2 intermediates then undergo C–N coupling to generate *CONH2, which is ultimately converted into urea.


image file: d5dt02018b-f3.tif
Fig. 3 (a–c) In situ FTIR spectra of In1Pd during the EUNC electrolysis at different potentials from the OCP to −0.7 V. (d) Online DEMS spectra of In1Pd at −0.7 V.

DFT calculations are utilized to unravel the atomic-level EUCN mechanism of In1Pd. Given that the adsorption and activation of NO/CO2 represent the initial step of the catalytic EUCN process, our analysis first focused on NO/CO2 adsorption behaviors on both In1 and Pd sites of In1Pd. As illustrated in Fig. S8, the adsorption free energy calculations reveal that In1 sites are more favorable for CO2 adsorption (−0.26 eV), while Pd sites are more favorable for NO adsorption (−0.48 eV), suggesting that during the EUCN process, In1 sites primarily facilitate CO2 reduction while Pd sites dominate NO reduction.

We then constructed the free energy diagram for the conversion of CO2 to *CO on In1 sites (Fig. 4a), with the corresponding intermediate configurations shown in Fig. S9. Calculation results reveal that the rate-determining step (RDS) for CO2 → *CO reduction is the first CO2 hydrogenation (*CO2 + *H → *COOH), with an energy barrier of 0.38 eV. Notably, the *CO desorption energy barrier on In1 sites is significantly lower (0.19 eV) than that required for further hydrogenation (*CO + *H → *COH, 0.54 eV), indicating that *CO generated on In1 sites is more inclined to desorb and spontaneously migrate closer to Pd sites and participate in the C–N coupling reaction.


image file: d5dt02018b-f4.tif
Fig. 4 (a) Free energy diagram of the CO2 → CO pathway on In1 sites. (b) Free energy diagram of the entire EUNC pathway on Pd sites and possible competing side reactions. (c) PDOS profile of *CONH2 on In1Pd. (d) Adsorption free energies of *H/*CO2 on In1 sites and *H/*NO on Pd sites.

Subsequently, we constructed the free energy diagram for the entire EUCN process on Pd sites, with the corresponding intermediate configurations shown in Fig. S10. As illustrated in Fig. 4b, the RDS of this pathway is also the initial NO hydrogenation (*NO + *H → *NHO, 0.29 eV). Further analysis of the energy barriers for the competing reactions, such as *NO + *CO → *NOCO and *NO + *CO2 → *NOCO2, reveals that compared to direct C–N coupling with *CO or *CO2, *NO tends to undergo hydrogenation reduction via the NHO pathway. Significantly, after stepwise *NOH reduction to *NH2, the generated *NH2 intermediate is more likely to undergo C–N coupling with *CO to form *NH2CO relative to other competing reactions (*NH2 + H → *NH3, *NH2 + CO2 → *NH2CO2). The generated *NH2CO can be spontaneously converted into urea. These findings correlate closely with the above in situ FTIR and DEMS results (Fig. 3).

For the critical *NH2CO intermediate on both pristine Pd and In1Pd surfaces (Fig. S11), the charge density difference map displays that In1Pd provides more electrons to *NH2CO than pristine Pd does, suggesting the significantly enhanced *NH2CO stabilization on In1Pd. Further PDOS analysis (Fig. 4c and Fig. S12) reveals that compared to pristine Pd, the overlap region of electron orbitals between In1Pd and *NH2CO is much expanded, further verifying the stronger *NH2CO activation capability of In1Pd. These results indicate that the introduced In1 not only promotes the reduction of CO2 to *CO but also modulates the electronic structure of the Pd substrate and enhances *NH2CO stabilization and activation towards urea conversion. Given the HER as the major competing reaction for the EUCN,4 we examined the adsorption characteristics of *H on In1Pd. Fig. 4d shows that In1 sites are more favorable for adsorbing *CO2 over *H, while Pd sites exhibit a stronger tendency to adsorb *NO over *H. Molecular dynamics (MD) simulations (Fig. S13) further reveal a more pronounced enrichment effect of *NO on the In1Pd surface, and the corresponding radial distribution function (RDF, Fig. S14) shows that the NO/In1Pd interaction is stronger than the H/In1Pd interaction. These results demonstrate that In1Pd can well hamper the competing HER towards the selective conversion of CO2/NO to urea.

In summary, In1Pd is demonstrated as a high-performance EUCN catalyst for urea electrosynthesis. Combined in situ spectroscopic measurements and theoretical calculations reveal that the enhanced EUCN performance of In1Pd stems from the synergy of In1 and Pd, which enables the co-activation of CO2/NO and their C–N coupling, while hampering the competing reactions. Impressively, In1Pd exhibits an unprecedented urea synthesis performance with urea yield rate up to 55.2 mmol h−1 g−1 and FEurea of 45.9% in a MEA cell. This work provides in-depth insights into the EUCN mechanism and opens up a new avenue for developing efficient and robust catalysts.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

The data supporting this article have been included as part of the SI. Supplementary information is available. See DOI: https://doi.org/10.1039/d5dt02018b.

Acknowledgements

This work is supported by Hebei North University 2025 College Students Innovation Training Program (no. 202510092022).

References

  1. Y. Kohlhaas, Y. S. Tschauder, W. Plischka, U. Simon, R.-A. Eichel, M. Wessling and R. Keller, Joule, 2024, 8, 1579–1600 CrossRef CAS.
  2. C. Mao, J. Byun, H. W. MacLeod, C. T. Maravelias and G. A. Ozin, Joule, 2024, 8, 1224–1238 CrossRef CAS.
  3. R. Ge, J. Huo, P. Lu, Y. Dou, Z. Bai, W. Li, H. Liu, B. Fei and S. Dou, Adv. Mater., 2024, 36, 2412031 CrossRef CAS PubMed.
  4. X. Peng, L. Zeng, D. Wang, Z. Liu, Y. Li, Z. Li, B. Yang, L. Lei, L. Dai and Y. Hou, Chem. Soc. Rev., 2023, 52, 2193–2237 RSC.
  5. J. Liu, S. Zhang, Y. Jiang, W. Li, M. Jin, J. Ding, Y. Zhang, G. Wang and H. Zhang, Chem. Commun., 2024, 60, 11592–11595 RSC.
  6. Y. Wu, H. Lin, Q. Mao, H. Yu, K. Deng, J. Wang, L. Wang, Z. Wang and H. Wang, Small, 2024, 20, 2407679 CrossRef CAS PubMed.
  7. S. Zhang, M. Jin, H. Xu, X. Zhang, T. Shi, Y. Ye, Y. Lin, L. Zheng, G. Wang, Y. Zhang, H. Yin, H. Zhang and H. Zhao, Energy Environ. Sci., 2024, 17, 1950–1960 RSC.
  8. Z. Wang, Y. Wang, S. Xu, K. Deng, H. Yu, Y. Xu, H. Wang and L. Wang, J. Mater. Chem. A, 2025, 13, 305–311 RSC.
  9. Y. Huang, R. Yang, C. Wang, N. Meng, Y. Shi, Y. Yu and B. Zhang, ACS Energy Lett., 2022, 7, 284–291 CrossRef CAS.
  10. D. Chen, J. Liu, J. Shen, Y. Zhang, H. Shao, C. Chen and S. Wang, Adv. Energy Mater., 2024, 14, 2303820 CrossRef CAS.
  11. R. T. Hannagan, G. Giannakakis, M. Flytzani-Stephanopoulos and E. C. H. Sykes, Chem. Rev., 2020, 120, 12044–12088 CrossRef CAS PubMed.
  12. T. Zhang, A. G. Walsh, J. Yu and P. Zhang, Chem. Soc. Rev., 2021, 50, 569–588 RSC.
  13. M. Xu, F. Wu, Y. Zhang, Y. Yao, G. Zhu, X. Li, L. Chen, G. Jia, X. Wu, Y. Huang, P. Gao and W. Ye, Nat. Commun., 2023, 14, 6994 CrossRef CAS PubMed.
  14. S. Zhang, J. Geng, Z. Zhao, M. Jin, W. Li, Y. Ye, K. Li, G. Wang, Y. Zhang, H. Yin, H. Zhang and H. Zhao, EES Catal., 2023, 1, 45–53 RSC.
  15. K. Chen, D. Ma, Y. Zhang, F. Wang, X. Yang, X. Wang, H. Zhang, X. Liu, R. Bao and K. Chu, Adv. Mater., 2024, 36, 2402160 CrossRef CAS PubMed.
  16. K. Chen, J. Xiang, Y. Guo, X. Liu, X. Li and K. Chu, Nano Lett., 2024, 24, 541–548 CrossRef CAS PubMed.
  17. X. Li, P. Shen, X. Li, D. Ma and K. Chu, ACS Nano, 2023, 17, 1081–1090 CrossRef CAS PubMed.
  18. P. Lu, X. Tan, H. Zhao, Q. Xiang, K. Liu, X. Zhao, X. Yin, X. Li, X. Hai and S. Xi, ACS Nano, 2021, 15, 5671–5678 CrossRef CAS PubMed.
  19. W. Guo, X. Tan, J. Bi, L. Xu, D. Yang, C. Chen, Q. Zhu, J. Ma, A. Tayal and J. Ma, J. Am. Chem. Soc., 2021, 143, 6877–6885 CrossRef CAS PubMed.
  20. J. Li, M. Zhu and Y. F. Han, ChemCatChem, 2021, 13, 514–531 CrossRef CAS.
  21. S. Li, X. Lu, S. Zhao, M. Ceccato, X.-M. Hu, A. Roldan, M. Liu and K. Daasbjerg, ACS Catal., 2022, 12, 7386–7395 CrossRef CAS.
  22. R. Lang, X. Du, Y. Huang, X. Jiang, Q. Zhang, Y. Guo, K. Liu, B. Qiao, A. Wang and T. Zhang, Chem. Rev., 2020, 120, 11986–12043 CrossRef CAS PubMed.
  23. S. K. Kaiser, Z. Chen, D. Faust Akl, S. Mitchell and J. Pérez-Ramírez, Chem. Rev., 2020, 120, 11703–11809 CrossRef CAS PubMed.
  24. C. Gao, J. Low, R. Long, T. Kong, J. Zhu and Y. Xiong, Chem. Rev., 2020, 120, 12175–12216 CrossRef CAS PubMed.
  25. D. Wu, K. Chen, P. Lv, Z. Ma, K. Chu and D. Ma, Nano Lett., 2024, 24, 8502–8509 CrossRef CAS PubMed.
  26. J. Min, J. Zhai, T. Dong, D. Xu, Y. Yan, C. S. Garoufalis, S. Baskoutas, Z. Zeng and Y. Jia, Nano Lett., 2023, 23, 4648–4653 CrossRef CAS PubMed.
  27. J. Zhai, T. Dong, Y. Zhou, J. Min, Y. Yan, C. S. Garoufalis, S. Baskoutas, D. Xu and Z. Zeng, Nano Lett., 2023, 23, 3239–3244 CrossRef CAS PubMed.
  28. Z. Zhao, J. Zhang, M. Lei and Y. Lum, Nano Res. Energy, 2023, 2, e9120044 CrossRef.
  29. W. Gou, H. Sun and F. Cheng, Nano Res. Energy, 2024, 3, e9120121 CrossRef.
  30. Q. Wang, H. Wei, P. Liu, Z. Su and X.-Q. Gong, Nano Res. Energy, 2024, 3, e9120112 CrossRef.
  31. Y. Cheng, H. Wang, H. Song, K. Zhang, G. I. Waterhouse, J. Chang, Z. Tang and S. Lu, Nano Res. Energy, 2023, 2, e9120082 CrossRef.
  32. J. Shen and D. Wang, Nano Res. Energy, 2023, 3, e9120096 CrossRef.
  33. Q. Hu, W. Zhou, S. Qi, Q. Huo, X. Li, M. Lv, X. Chen, C. Feng, J. Yu, X. Chai, H. Yang and C. He, Nat. Sustain., 2024, 7, 442–451 CrossRef.
  34. Y. Gao, J. Wang, M. Sun, Y. Jing, L. Chen, Z. Liang, Y. Yang, C. Zhang, J. Yao and X. Wang, Angew. Chem., Int. Ed., 2024, 136, e202402215 CrossRef.
  35. Y. Luo, K. Xie, P. Ou, C. Lavallais, T. Peng, Z. Chen, Z. Zhang, N. Wang, X.-Y. Li, I. Grigioni, B. Liu, D. Sinton, J. B. Dunn and E. H. Sargent, Nat. Catal., 2023, 6, 939–948 CrossRef CAS.
  36. C. Lv, L. Zhong, H. Liu, Z. Fang, C. Yan, M. Chen, Y. Kong, C. Lee, D. Liu, S. Li, J. Liu, L. Song, G. Chen, Q. Yan and G. Yu, Nat. Sustain., 2021, 4, 868–876 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.