Ibrahem O. Baibars‡
a,
Haisen Huang‡b,
Yang Xiaoa,
Shuhao Wanga,
Yan Niea,
Chen Jia
a,
Kamran Dastafkan
a and
Chuan Zhao
*a
aSchool of Chemistry, The University of New South Wales, Sydney, NSW 2052, Australia. E-mail: chuan.zhao@unsw.edu.au
bCAS Key Laboratory of Materials for Energy Conversion, Department of Materials Science and Engineering, University of Science and Technology of China, Hefei, Anhui 230026, P. R. China
First published on 27th May 2025
A macro/mesoporous film with Ni/CeOx interfaces is designed via the dynamic hydrogen bubble template (DHBT) method for ampere-level production of hydrogen in anion exchange membrane water electrolysers (AEMWEs). The AEMWE achieves leading energy efficiencies of 95% and 80%, based on the higher and lower heating values of hydrogen, respectively, at 0.25 A cm−2, producing hydrogen at 42 kW h kg−1 and a cost of $0.84 per kg, thereby meeting the U.S. Department of Energy (DOE) price target ($1 per kg) for 2030. A current density of 5 A cm−2 is achieved at 2.08 V and 60 °C in the AEMWE, with overall cell activation and concentration overpotentials of 594 mV, establishing a leading position in the field. The Ni/CeOx catalyst exhibits superb hydrogen evolution reaction (HER) activity by delivering 1 A cm−2 at an overpotential of 201 mV at 20 °C, far surpassing Ni, CeOx, and benchmark Pt/C catalysts. Electrochemical and theoretical calculations reveal accelerated charge transfer due to the preferential adsorption of intermediates at the tailored defective interfaces during hydrogen evolution. Hydrogen evolving during electro-deposition forms 3D channels for bubble removal in the AEMWE, akin to a hydrogen memory, speeding up mass transfer.
Broader contextOverpotential sources, particularly activation and concentration overpotentials, represent a major barrier to achieving high efficiency in anion exchange membrane water electrolysers (AEMWEs), significantly limiting their real-world performance. A Ni/CeOx heterostructure catalyst with a hierarchical macro/mesoporous morphology is designed to address both overpotentials. The defective interfaces synergistically accelerate charge transfer by promoting water dissociation, hydroxide desorption, and hydrogen adsorption, while the hierarchical pores mitigate mass transfer limitations through improved electrolyte and bubble transport. This design enables the AEMWE to achieve remarkable performance metrics, including a hydrogen generation current density of 5 A cm−2 at 2.08 V and 60 °C with a total overpotential of 594 mV, and energy efficiencies of 95% and 80% (based on higher and lower heating values, respectively) at 0.25 A cm−2. The hydrogen production cost is reduced to $0.84 per kg, surpassing the U.S. Department of Energy's (DoE) target of $1 per kg by 2030. By combining advanced catalyst design with a robust framework for analysing overpotential sources, this work not only meets critical efficiency and cost targets but also establishes a foundation for further optimization and scalability in green hydrogen production. |
Energy efficiencies of water electrolysers are typically governed by three types of overpotentials: activation overpotential (including chemical and charge transfer), concentration overpotential (mass transfer), and ohmic overpotential.7,8 The ohmic overpotential is primarily ionic, influenced by the membrane type and thickness, provided that electronic contributions from catalysts and other cell components are minimized through appropriate design and assembly.4 While AEMWEs benefit from faster oxygen evolution (OER) kinetics in alkaline solutions, the hydrogen evolution reaction (HER) is retarded due to the scarcity of H+ ions, which requires a prerequisite water dissociation step followed by hydroxide removal, resulting in extra energy barriers and a further increase in activation overpotentials.9,10 To overcome these activation overpotentials, various catalyst designs have been developed, including high-entropy alloys,11 single-atom catalysts,12 heteroatom-doped catalysts,13,14 and defect-rich catalysts.15 Among these designs, heterostructured catalysts show great synergy potential at the interfaces formed between a wide range of materials.16,17 Thirdly, concentration overpotential is controlled by mass transfer to (for electrolyte) and from (for gas bubbles) the catalyst surface. At high current densities, concentration overpotential becomes more pronounced due to the increased demand for the electrolyte and the blockage of active sites by numerous bubbles.18 Catalyst morphology is therefore crucial for preventing the accumulation and coalescence of gas bubbles, improving electrolyte accessibility, and minimizing the concentration overpotential.8,19,20
Here, a heterostructured Ni/CeOx catalyst is designed to minimize the activation overpotential via interfacial synergy that facilitates water dissociation and provides suitable sites for the adsorption of H and desorption of OH− and H2. Metal/metal oxide heterostructures have been reported to speed up HER kinetics in alkaline media.21–23 CeO2 is known for non-stoichiometry, O vacancy abundance, facile Ce3+/Ce4+ transition, and effective catalysis features via disrupting chemical bonds in reactants and intermediates at its defective sites.24–27 Theoretical calculations, herein, reveal that CeOx sites help polarise water molecules, host OH− ion desorption, and modulate Ni adsorption of H. Additionally, a highly porous Ni/CeOx catalyst layer is achieved via dynamic hydrogen bubble template (DHBT) electrodeposition.28 During the deposition, catalyst layers grow around the co-produced hydrogen bubbles, forming hierarchical macro/meso-pores with nano-edges along the tracks of the bubbles.28,29 These pores born from gas bubbles during catalyst synthesis are crucial for bubble removal and electrolyte access in MEA.30 Catalyst/membrane interfaces are also enriched by the tortuosity of the catalyst clusters between pores, accelerating electrolyte delivery and minimizing mass transfer resistance. The designed Ni/CeOx film shows minimum mass transfer overpotential and low charge transfer overpotential when tested for the HER in a half-cell, a two-electrode electrolysis cell, and AEMWEs using different membranes.
In this study, various Ni/CeOx HER catalysts are fabricated on Ni foam substrates using a deposition bath with a Ni2+ to Ce3+ molar ratio of 9:
1 by varying the applied current density and deposition times to maintain the same passing charge. The best-performing HER catalyst is found to be deposited at 2 A cm−2 (Fig. 1 and Fig. S15b, ESI†). Ni/CeOx (2 A) shows several hierarchical cone-shaped macropores with various diameters ranging from 16 μm down to 50 nm with tortuous cauliflower-like particles (Fig. 1b and Fig. S1a–c, ESI†). Both dark-field STEM and HRTEM images (Fig. 1c and d) show ∼60 nm-wide particles with mesopores of ∼10 nm diameter. The nitrogen adsorption isotherm from Brunauer–Emmett–Teller (BET) analysis illustrates the distribution of these mesopores, showing an average pore diameter of 7.4 nm (Fig. S2, ESI†). Additionally, the absence of a hysteresis loop indicates the ease of N2 gas desorption and suggests fast H2 bubble removal during the HER (Fig. S2a, ESI†). The wide variation in pore sizes, including macro and mesopores, reflects the diverse sizes of H2 bubbles evolving during deposition. Such variations in pore sizes are crucial to accommodating the range of H2 bubble sizes rising during water electrolysis, akin to a hydrogen memory.30 SEM-EDS mapping analysis (Fig. S3, ESI†) shows the distribution of Ni, Ce, and O elements, with small oxygen content (9.97%), where O sites surround Ce sites in the overlay image, supporting the formation of Ce oxides interfaced with metallic Ni. Despite the quantification limitations of SEM-EDS, the resulting atomic ratio of Ni to Ce is 97
:
3, matching that obtained from ICP-OES (Table S1 and Fig. S3, ESI†). According to the ICP-OES data, deposition faradaic efficiency is calculated to be 7.78%, implying that 92.22% of the passing charge is consumed in producing hydrogen bubble templates (Fig. S4, ESI†). On the other hand, when Ni and CeOx components of the interface are deposited separately using the same method, the Ni film exhibits a macroporous morphology, while the CeOx film does not (Fig. S5–S8, ESI†). Four additional Ni/CeOx films are deposited at different current densities, resulting in two compact non-porous films at low current densities and two macroporous ones at high current densities, with the pore size decreasing as the current density increases (Fig. S5 and S9–S12 and Note S1, ESI†). Among all films, Ni/CeOx (2 A) features the highest electrochemically active surface area (ECSA), as shown in Fig. S20 (ESI†), achieving a balance between the size and number of pores.
High-resolution TEM is employed to determine the chemical structure of Ni/CeOx (2 A). The analysis of d-spacing reveals a metallic Ni facet spaced at 2.03 Å interfacing with a CeO2 facet with a spacing of 3.15 Å, matching those of the (111) facets of cubic Ni36,37 (JCPDS card: 03-065-2865, and d-spacing (111) = 2.035 Å) and cubic CeO2 (JCPDS card: 03-065-5923, and d-spacing (111) = 3.12 Å), respectively (Fig. 1e and Fig. S13a, ESI†). XRD confirms the chemical structure, where peaks at 2θ values of 44.55°, 51.86°, 76.39°, 92.97°, and 98.45° correspond to diffractions from the (111), (200), (220), (311), and (222) crystallographic planes of cubic Ni (JCPDS card: 03-065-2865), respectively (Fig. 1f). Another small peak is also observed at 28.55°, enclosed by a yellow box in Fig. 1f and zoomed in Fig. S13i (ESI†). This peak corresponds to the (111) facet of cubic CeO238,39 (JCPDS card: 03-065-5923) and appears small due to the low Ce content in the film, as reported earlier from ICP-OES and EDS data, in addition to the presence of amorphous regions, Fig. S13b (ESI†). The absence of this peak in the XRD pattern of the Ni film, alongside its appearance in the CeOx pattern, reaffirms the formation of the Ni/CeOx interface (Fig. S13i, ESI†). The patterns of the Ni film and all Ni/CeOx films show the characteristic peaks of cubic metallic Ni, while it is absent for the CeOx film (Fig. 1f and Fig. S13c, ESI†). Slight shifts in the 2θ Bragg angles of cubic metallic Ni peaks are observed for all Ni/CeOx films compared to the Ni film, indicating slight variations in the lattice parameters upon formation of the Ni/CeOx interface, see Fig. S13d–h and Table S2 (ESI†).40 In addition, in all Ni/CeOx films (Fig. S13j, ESI†), the peak corresponding to cubic CeO2 is evident, except for Ni/CeOx (0.1 A), where the presence of that peak cannot be confirmed, likely due to the minimal percentage of Ce in this film as reported by ICP-OES.
Raman spectroscopy confirms the resulting chemical structure, with Ni/CeOx (2 A) showing two bands at 573 and 1065 cm−1, corresponding to the defect-induced mode (D band) and the second-order longitudinal optical mode (2LO band), respectively, as reported for Ni/CeO2.41 Interestingly, the intensity of the F2g symmetry band at 452 cm−1, associated with scattering from symmetric O anions around the cations in the CeO2 structure, is minimal (Fig. 1g).42 The intensity ratio of the D band to the F2g band (ID/IF2g) indicates the amount of O vacancies present in CeO2.41,43,44 Accordingly, the resulting Ni/CeOx (2 A) film is highly defective and non-stoichiometric, and therefore labelled as CeOx instead of CeO2, leading to the appearance of Ce3+ in addition to Ce4+.43 Raman spectra of the other deposited films (Fig. S14a, ESI†) show that the sequence of the (ID/IF2g) ratio of these films is CeOx < Ni/CeOx (0.1 A) < Ni/CeOx (0.5 A) < Ni/CeOx (1 A) < Ni/CeOx (3 A) < Ni/CeOx (2 A). These data refer to structural differences between the designed films, where Ni/CeOx (2 A) exhibits the highest degree of defects.
The oxidation states of the constituent elements in Ni/CeOx (2 A) are investigated through top-surface and depth-profiling XPS analyses. Deconvolution of the Ni 2p spectrum at the top surface (Fig. 2a) indicates that Ni is mostly metallic (Ni0) with the presence of Ni2+ and Ni3+ ions, likely resulting from the oxidation of the top atomic layers by atmospheric oxygen.45,46 This is further supported by depth-profiling XPS, showing that the deeper the layer, the bigger the shift towards the metallic Ni peak (Fig. 2d). The 3d spectrum of Ce at the top surface (Fig. 2b) is deconvoluted into Ce3+ ions, which are related to the formation of O vacancies,47 and Ce4+ ions,48,49 with the former appearing larger in quantity, supporting the formation of defective Ce oxides with multiple Ce oxidation states as deduced from Raman spectra (Fig. 1g). Inspection of depth-profiling XPS of Ce (Fig. 2e) reveals that these oxidation states coexist throughout the tested depth, with the intensity of Ce3+ and Ce4+ peaks getting closer as the depth increases. Supporting these observations, the deconvolution of the O 1s spectrum at the top surface (Fig. 2c) reveals the presence of oxides, adsorbed O, and hydroxides, where hydroxides and adsorbed O appear dominant, likely due to Ni oxidation by atmospheric air, bearing in mind the large atomic percentage of Ni, revealed by ICP-OES and SEM-EDS. Interestingly, substantial changes manifest in the O 1s spectra upon testing deeper layers of the catalyst layer (Fig. 2f). A clear shift towards the oxide peak and a diminution of the hydroxide and adsorbed O peaks are observed at deeper catalyst layers. This is explained by the presence of Ce oxides and the absence of Ni hydroxides. This shift affirms the presence of metallic Ni and Ce oxides at deeper layers and the formation of Ni/CeOx interfaces. On the other hand, top-surface XPS spectra (Fig. S14b and c, ESI†) of the CeOx (2 A) film show a higher Ce4+ content relative to Ce3+, confirming a lower degree of O defects in CeOx (2 A) compared to Ni/CeOx (2 A),47 consistent with Raman data (Fig. S14a, ESI†).
Additionally, the Tafel slope was measured and found to be 41 mV dec−1, referring to the Heyrovsky step as the rate-determining step (RDS) and indicating facilitated charge transfer over Ni/CeOx (2 A) (Fig. 3b).16 This is further supported by impedance spectroscopy with the distribution of relaxation times (DRT) analysis,50 which shows a single, small charge transfer resistance peak (p3, possibly corresponding to the Heyrovsky step), while p1 represents the ohmic resistance in the cell (Fig. 3d). On the other hand, the deposited CeOx film displays a larger Tafel slope of 97 mV dec−1 and a smaller i0 value of 0.17 mA cm−2, suggesting sluggish charge transfer. This value, falling between 120 and 40 mV dec−1, implies a dual control of the HER rate by both Volmer and Heyrovsky steps.16,51 This signifies a notable deceleration occurring at the Volmer step, attributed to the absence of interfaced Ni sites that would help polarise and dissociate water molecules and act as H adsorption sites. Therefore, two much larger charge transfer peaks (p2 and p3, possibly corresponding to Volmer and Heyrovsky steps, respectively) are observed in the DRT analysis. As a result, unobservable HER currents of 0.745 and 2.1 mA cm−2 are driven by overpotentials of 100 and 150 mV. As for the deposited Ni film, it shows a Tafel slope of 64 mV dec−1 and an i0 value of 0.48 mA cm−2, which fall between those of CeOx and Ni/CeOx (2 A), indicating moderate charge transfer kinetics with i0 significantly lower than that of Ni/CeOx (2 A). Accordingly, p2 and p3 DRT peaks appear with the medium area, and overpotentials of 100 and 150 mV drive HER currents of 23.5 and 94.5 mA cm−2 at the Ni catalyst. Such an inferior activity in the absence of CeOx sites highlights again the crucial effect of Ni/CeOx interface formation, where CeOx sites help polarise and dissociate water molecules and provide desorption sites for the hydroxide byproduct.
As for the other Ni/CeOx catalysts deposited at different current densities, the observed activities in the low-current charge-transfer-controlled region follow this order: Ni/CeOx (0.1 A) < Ni/CeOx (0.5 A) < Ni/CeOx (1 A) < Ni/CeOx (3 A) < Ni/CeOx (2 A) (Fig. S15b, ESI†). This order is intriguingly the same as the degree of defects observed among these catalysts from Raman spectra. This observation suggests that the number of O vacancies acts as a descriptor for the charge transfer speed among metal/metal oxide interface catalysts: the more defective the catalyst is, the faster the charge transfer. In agreement, Tafel slopes come in reverse order, approaching the theoretical value of the Volmer RDS as the degree of defects decreases (Fig. S17, ESI†). In conclusion, the higher HER activity of the Ni/CeOx (2 A) film, compared to other Ni/CeOx films in the low-current region, can be attributed to the highest degree of defects in this film, which eases the charge transfer and water polarisation and dissociation processes.
For the concentration overpotential, Ni/CeOx (2 A) exhibits the fastest mass transfer with the smallest low-frequency DRT peak (p4), followed by Ni, whereas CeOx suffers from significant limitations (Fig. 3d). In high-current regions, the non-intrinsic contributions are substantial and correlate with the number of accessible active sites. Accordingly, the ECSA is measured for all deposited films (Fig. S18 and S19, ESI†), revealing that the CeOx film, as expected, has the lowest ECSA due to the absence of macropores, far lower than that of the macroporous Ni film (Fig. S20a, ESI†). Notably, Ni/CeOx (2 A) exhibits the highest ECSA, with 479 cm2ECSA cm−2geo, a high surface area that signifies the non-intrinsic activity component and the accelerated mass transfer. Although all Ni/CeOx catalysts show higher ECSA values than Ni and CeOx films, the gravimetric-specific ECSA, which is more reflective of morphology, refers to a higher value for the macroporous Ni film over the non-porous Ni/CeOx films, as expected.
The gravimetric-specific ECSA of Ni is still inferior to other macroporous Ni/CeOx films, where Ni/CeOx (2 A) keeps exhibiting the highest value (Fig. S20b, ESI†). To reaffirm the intrinsic activity component of Ni/CeOx (2 A), all LSVs are normalised to ECSA values in Fig. S20c (ESI†), where Ni/CeOx (2 A) remains the best-performing catalyst intrinsically. Turnover frequency (TOF) measurements are also conducted to confirm the superiority of the Ni/CeOx (2 A) catalyst, both intrinsically and non-intrinsically (Fig. S21, ESI†). The Ni/CeOx (2 A) catalyst has the highest number of active sites owing to its special balanced macro/mesoporous structure, reflecting enhanced mass transfer and non-intrinsic catalysis superiority. Additionally, after normalising the measured LSVs to obtain the corresponding TOFs, the Ni/CeOx (2 A) catalyst still exhibits the highest TOF, referring to the highest intrinsic activity among the deposited catalysts.
The durability of the Ni/CeOx (2 A) catalyst is evaluated at multiple current densities of up to 1 A cm−2 for 2 weeks (Fig. 3e). The Ni/CeOx film shows great stability even at high current densities like 1 A cm−2, where no decay is noticed, reflecting the ability of the designed interface and macro/mesoporosity to survive under such harsh conditions. Furthermore, post-stability characterization studies (Fig. S22, ESI†) reveal that Ni/CeOx (2 A) maintains its elemental composition, morphology (including the cauliflower-like particles and the macro/mesoporosity), and structure. Nitrogen adsorption BET isotherm data (Fig. S22, ESI†) show nearly the same average mesopore diameter (7.1 nm) and the same type IV(b)52 isotherm with a continuous absence of hysteresis, suggesting continued efficient gas desorption. Reconstruction of the CeOx sites into more ordered crystals is observed during the HER, yielding sharper CeO2 diffraction peaks (Fig. S22g, ESI†). These observations affirm the stability of the designed heterostructured catalyst for practical ampere-level applications.
The performance of Ni/CeOx (2 A) is evaluated for full water electrolysis by coupling with an active OER (Ni–Fe–Zn) catalyst (Fig. S23 and S24 and Note S2, ESI†). In a 2-electrode alkaline electrolysis cell, 500, 1000, and 2000 mA cm−2 are achieved at 60 °C at cell voltages of 1.689, 1.73, and 1.764 V, corresponding to overall cell overpotentials of 489, 530, and 561 mV, respectively (Fig. 3f). Following the overpotential deconvolution procedure outlined in the calculation methods section (ESI†), these overpotentials are found to be mostly activation overpotentials rather than concentration overpotentials. This indicates the efficient mass transfer properties of the catalyst at high current densities. The electrolysis cell exhibits excellent stability for 768 h at 1 A cm−2 (Fig. 3g).
The Ni/CeOx catalyst is further applied for the AEMWE. When the HER and OER catalysts are assembled with a PiperION-20 membrane, current densities of 5, 4, and 3 A cm−2 are achieved at cell voltages of 2.078, 2.001, and 1.918 V at 60 °C, with IR-free overpotentials of 594, 573, and 547 mV, respectively (Fig. 4b), placing the designed catalysts and AEMWEs among the best reported in the literature (Table S4, ESI†). The designed AEMWE produces hydrogen at a high electrical efficiency of 42.0 kW h kg−1, when operating at 0.25 A cm−2 (Fig. 4d). Therefore, the cost of hydrogen production is $0.84 per kg when the electricity cost is $0.02 per kW h,53 meeting the 2030 price target ($1 per kg) of the U.S. Department of Energy (DOE).54 This corresponds to energy efficiencies of 95% and 80%, based on the higher (HHV) and lower (LHV) heating values of hydrogen, respectively (Fig. 4d). Accordingly, the AEMWE device surpasses the International Renewable Energy Agency (IRENA) energy efficiency (LHV) target for 2050 (>75%) when operating at current densities below 0.66 A cm−2.5 On the other hand, with the X37-50 membrane, voltages of 1.754, 1.843, and 1.981 V drive lower current densities of 0.5, 1, and 2 A cm−2, respectively, due to the higher ohmic resistance of the X37-50 membrane (Fig. 4c). Notably, the observed IR-free overpotentials are 503, 542, and 579, respectively, approximately matching those 489, 530, and 561 mV values observed at the IR-free polarization curve at the 2-electrode alkaline electrolysis cell. Accordingly, IR-free testing in 2-electrode alkaline electrolysis cells can sometimes predict the observed overpotentials in AEMWEs free from IR drops. The deconvolution of the observed overpotentials in both AEMWEs shows a similar behaviour to that observed in the 2-electrode cell, with nearly no noticeable concentration overpotentials. This points to the unique macro/mesoporous morphology attained by the DHBT method. The presence of various pore sizes increases the accessible surface area, accommodates different bubble sizes, prevents site blockage, and facilitates electrolyte and bubble (mass) transfer. Reduced activation overpotentials refer to the tailored structures of the designed catalysts to accelerate the charge transfer processes. AEMWE durability is evaluated via chronopotentiometry by applying 0.5 A cm−2 for 96 h. No voltage decay is observed throughout the testing duration, referring to materials’ stability at the operating current and conditions (Fig. 4e). The faradaic efficiency of hydrogen production is measured by gas chromatography (GC) at 0.5 A cm−2, reaching 99.98% (Fig. S25, ESI†).
The Gibbs free energy of hydrogen adsorption (ΔGH) is a main descriptor for the intrinsic activity of catalysts towards the HER. A value proximal to zero indicates superior HER performance, where a balance between H adsorption and desorption is achieved, neither excessively strong nor too weak.55 As shown in Fig. 5a, ΔGH values follow a near-zero-based sequence, Ni/CeO2−2δ (0.22 eV) > Ni/CeO2−δ (0.31 eV) > Ni (0.48 eV) > CeO2−δ (−0.54 eV), mirroring the electrocatalytic sequence towards the HER. This supports the synergistic impact of forming Ni/CeOx interfaces compared to individual Ni or CeOx on boosting HER activity by tuning hydrogen adsorption energies. Meanwhile, it highlights the effect of O vacancy concentration on modulating the catalyst's adsorption capacity for hydrogen, wherein more defects at the interface enhance hydrogen adsorption/desorption energetics.
![]() | ||
Fig. 5 Theoretical calculations. Gibbs free energy of HER processes in alkaline media over Ni, CeO2−δ, Ni/CeO2−δ, and Ni/CeO2−2δ catalysts: (a) hydrogen adsorption process. (b) Water adsorption and dissociation processes. (c) OH* direct desorption process. (d) Water-assisted OH− indirect desorption process. (e) Electron density difference images at the interface of Ni/CeO2 Ni/CeO2−δ, and Ni/CeO2−2δ (blue represents electron rich and red represents electron deficient). (f) Partial density of states (PDOS) of Ni, Ni/CeO2, Ni/CeO2−δ, and Ni/CeO2−2δ (Ni 3d orbits). (g) Partial density of states (PDOS) of CeO2−δ, Ni/CeO2, Ni/CeO2−δ, and Ni/CeO2−2δ (Ce 4f orbits). Atoms colors: Ni: grey, Ce: lime green, O: red, and H: light pink (Table S7, ESI†). |
Water dissociation proceeds through three stages: the initial state (IS), which involves water adsorption, followed by a transition state (TS) of bond disruption, leading to complete dissociation into adsorbed H and OH− in the final state (FS). During these processes, three other Gibbs free energy values are crucial besides ΔGH due to the lack of protons in alkaline media.56 These values are GIS (free energy of water adsorption), ΔGTS–IS (a kinetic parameter referred to as a water dissociation energy barrier), and ΔGFS–IS (a thermodynamic parameter referring to the spontaneity of the dissociation process). The lower these values, the more favourable water dissociation in thermodynamic and kinetic aspects. As illustrated in Fig. 5b, Fig. S35 (ESI†), and Table S6 (ESI†), Ni/CeOx interfaces exhibit rapid (small kinetic energy barriers, ΔGTS–IS) and spontaneous water dissociation (negative ΔGFS–IS values). On the other hand, although CeO2−δ, demonstrates the best water adsorption energy (0.21 eV), it exhibits non-spontaneous (positive ΔGFS–IS value of 0.24 eV) water dissociation with sluggish kinetics and a large energy uphill of 1.36 eV. Separate Ni exhibits the worst water adsorption energy (0.68 eV) along with the largest dissociation energy barrier (1.47 eV). However, it outperforms separate CeOx due to the lower non-spontaneity barrier (ΔGFS–IS) and more optimal hydrogen adsorption energy (ΔGH). For Ni/CeOx interfaces, water adsorption and dissociation processes over Ni/CeO2−2δ are more thermodynamically and kinetically favoured than over Ni/CeO2−δ, reflecting the optimisation impact of the structure defects on these processes besides the hydrogen adsorption process. These observations suggest that at Ni/CeOx interfaces, water adsorption occurs at O vacancies and then disruption of water bonds occurs when hydrogen is pulled towards interfaced Ni, leading to complete water dissociation and leaving dissociated H adsorbed at Ni, while dissociated hydroxide adsorbed at the O vacancy.
The differential charge densities of Ni/CeO2, Ni/CeO2−δ, and Ni/CeO2−2δ are calculated to confirm the proposed role of O vacancies in the HER (Fig. 5e). At the interface of Ni/CeO2, electrons are concentrated around the O atom, resulting in electron enrichment at the interface. This enrichment is advantageous for attracting water H but not for OH, leading to relative deceleration in water adsorption and dissociation. As the concentration of O vacancies increases, a red cloud (indicative of electron deficiency) is observed at the O vacancy site, signifying a localized positive charge in this region, which is beneficial for attracting OH. Meanwhile, electron enrichment occurs around Ni, facilitating the attraction of H. This preferential attraction of OH and H at the O vacancy and Ni sites, respectively, disrupts water bonds and promotes water dissociation, thereby leading to Heyrovsky-controlled kinetics, diverging from the sluggish Volmer RDS as depicted in Fig. 3b.
To verify this concept, the PDOSs of the Ni 3d orbits of Ni, Ni/CeO2, Ni/CeO2−δ, and Ni/CeO2−2δ are calculated (Fig. 5f). The upshift of the d-band centre of Ni/CeO2 compared to Ni indicates a synergistic effect between CeO2 and Ni, with electrons at higher energy levels facilitating H adsorption.57,58 As the concentration of vacancies increases, the d-band centre shifts further upward, indicating that O vacancies can enhance the H adsorption capacity of Ni sites. Similarly, compared to CeO2−δ and Ni/CeO2, Ni/CeO2−δ, and Ni/CeO2−2δ have more bonding orbitals below the Fermi level and fewer antibonding orbitals above the Fermi level (Fig. 5g). Bonding orbitals signify stable bonding in the structure, indicating that O vacancy-rich CeO2 forms a stable interface with Ni.
Hydroxide desorption is a crucial process for the retrieval of active sites through two possible pathways: either direct desorption to the aqueous phase or indirect release from the aqueous phase through a water-assisted proton exchange pathway.59 The latter can accelerate the HER in alkaline media by skipping the water adsorption step. The calculation results (Fig. 5c and d) show that Ni and CeO2−δ prefer the direct hydroxide desorption pathway, while for Ni/CeOx interfaces, the water-assisted hydroxide desorption pathway is preferred, leading to accelerated kinetics. Ni/CeO2−2δ shows the lowest energy barrier for hydroxide desorption, implying the acceleration impact with a higher degree of defects. Based on this discussion and the high conductivity observed from the DOS plots (Fig. S36, ESI†), the intrinsic activity of the Ni/CeOx (2 A) catalyst, with lowered activation overpotential, can be explained.
During the structural optimization process, a 4 × 4 × 1 grid was utilised to sample CeO2, Ni/CeO2, CeO2−δ, Ni/CeO2−δ, and Ni/CeO2−2δ. The CeO2−δ structure was generated by introducing defects onto the CeO2 (111) surface, whereas Ni/CeO2−δ and Ni/CeO2−2δ involved defects created within the oxygen atomic layer at the Ni/CeO2 heterojunction interface. To maximize the exposure of the active oxygen vacancy sites, the surface adsorption configuration of Ni/CeO2−δ was selected in the (100) direction, while Ni/CeO2−2δ was cut on the (010) plane.
For the calculation of the chemical formation energy of defects, taking CeO2−δ as an example, the equation used is:
![]() | (1) |
The Gibbs free energy of hydrogen adsorption (ΔGH), which is a key parameter for determining the HER activity as proposed by Norskov et al.,61 was calculated following the equations below:
![]() | (2) |
ΔGH = ΔEH − ΔZPE − TΔSH | (3) |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ee06113f |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |