Phase structure and morphology. Figure 1(a) exhibits the XRD pattern of the prepared ZFO, CeO2, and ZFO-CeO2. The high-intensity peak of ZFO is located at 2\({\theta }^\circ\) angle of (3 1 1) plane, verifying the single-phase cubic spinel structure of ZnFe2O4. The cubic crystal structure of the spinel ferrite sample (ZFO) is verified using jade and X\({\prime }\)Pert high score plus software with the JCPD CARDS of #81–0792 and #01-087-1230. The XRD refinement analysis was also carried out to confirm the exact single-phase cubic nature of the synthesized ZFO. From the refinement data, it is verified that ZFO belongs to the Fd-3m space group with lattice parameters a = b = c = 8.4432 Å30. The theoretical parameters of the prepared ZFO sample are in good agreement with the spinel ferrite crystal structure literature 31. While CeO2 phase was also confirmed using X\({\prime }\)Pert high score plus software. It is found that the high intensity peaks are located at 2\({\theta }^\circ\) angle of values of 28.54\(^\circ\) (1 1 1), 33.07\(^\circ\) (2 0 0), 47.47\(^\circ\) (2 2 0), 56.33\(^\circ\) (3 1 1), 59.07\(^\circ\) (2 2 2), 69.40 (4 0 0) with standard JCPDS#01-081-0792. The prepared CeO2 belongs to the fluorite structure with lattice parameter values of a = b = c = 5.412432, 33. In the XRD pattern of ZFO-CeO2 (Supplementary Fig. S1), the diffraction peaks of ZFO and CeO2 are identified without extra peaks, suggesting the successful formation of the heterostructure of ZFO and CeO2.
Morphological analysis of pure ZFO, pure CeO2, and ZFO-CeO2 heterostructure composite were studied using high-resolution transmission electron microscopy (HR-TEM) and energy dispersive spectroscopy (EDS). Figure 1(b) exhibits the randomly distributed spherical-shaped ZFO nanoparticles with a size of 45–50 nm, which identifies a clear interface between ZFO and CeO2 grains (See Fig. b3). Two methods (Line intercept method and Digital Micrograph) were employed to measure the lattice planes and d spacing of the ZFO-CeO2. The d spacing of the ZFO-CeO2 heterostructure agrees with the XRD measured data. The slight difference between lattice spaces, i.e., 2.67 Å and 2.69 Å, as obtained from Digital Micrograph software, verifies the exact formation of ZFO-CeO2 heterostructure. The elemental distribution of ZFO and CeO2 nanoparticles in the heterostructure was further confirmed using Selected Area Electron Diffraction (SAED) and EDS mapping analysis as depicted in Fig. 1c (c1-c5) and 1d (d1 and d2). Zn, Fe, Ce, and O particles are randomly distributed in the sample to indicate the successful heterostructure.
First principle calculations. Density functional theory (DFT) calculation is an appropriate way to theoretically analyze the ions diffusion mechanism and evaluate the semiconducting nature of the prepared samples. The crystal structures and electronic bandgaps of ZFO and CeO2 have also been considered and presented in Fig. 2 (a-d). The structures, electronic bandgap, and density of state (DOS) are constructed and evaluated using Quantum ATK software with LDA and GGA approximations. Further procedure to calculate electronic band structure and DOS has been explained in detail in the experimental section. The obtained electronic bandgap using DFT calculations for ZFO and CeO2 samples is 1.85 eV and 3.5 eV, respectively, which is similar to our experimental values, where the optical bandgaps obtained from UV-Vis spectra are presented in Supplementary Fig. S2. The calculated optical bandgaps of the pure ZFO and CeO2 samples were 1.85 eV and 3.15 eV, respectively. The small error between electronic bandgap and optical bandgap values has been noted and verified with previous literature 34, with the fact that different U values have been used in first principle calculations. The DOS of individual elements and projected DOS indicates the constitution of each element to the valance and conduction zone in the pristine ZFO and CeO2 samples principally, as shown in Fig. 2 (e). In ZFO, it has been observed that O ions principally exist in the valance zone, and Fe ions are in the conduction zone with the smaller contribution of oxygen ions. Steven Jankov et al.35 reported on the existence of each element in its valance and conduction zones, and our work agrees with their study. For CeO2, the fully occupied and unoccupied states are composed of O-2p states and Ce-4 f states, which is in accordance with the report of Rong Han et al.36.
Interfacial disordering and heterojunction. To evaluate the interfacial disorder and conduction mechanism of ZFO-CeO2, DFT was employed with the computational parameters of the constructed structure, as mentioned in the experimental section. Our calculations based on heterostructure showed that the lattice mismatch is less than 5% by indicating valid interface formation between the ZFO and CeO2 using the QUANTUM ATK builder. The interfacial structure without optimization geometry (OG) is presented in the calculations in Fig. 3(a). It can be observed that the constructed heterostructure has an interfacial connection between ZFO and CeO2, and each participating atoms of both samples (Zn, Fe, Ce, and O) are well organized with similar M-O distances as pure ZFO and CeO2 structure. After OG, by applying force tolerance of 1\(\times\)10−5 eV/Å, the disorder crystal structure with different variations in M-O distances are observed, as displayed in Fig. 3 (b). The M-O distances of the constructed heterostructure (before and after) using OG are calculated and inserted in Supplementary Fig. S3 (a and b). At the same time, before OG, the band structure is displayed as I, and after OG, it is inserted as II in Fig. 2(i and ii). Such a kind of disordering induces a disordered oxygen plane at the interface between ZFO and CeO2 by dislocating oxygen atoms. It can be found that the electronic states of oxygen (O) exhibit hybridized nature with those of ZFO and CeO2 species, which play a crucial role in promoting proton migration. The interfacial disorder dominates at the interface grain boundaries of ZFO and CeO2 as both structures break the compatibility and create disorder at the interface during heterostructure formation. The ZFO-CeO2 heterostructure represents to 6 space group and P-2y hall symbol as presented in Supplementary Fig. S4. Interestingly, The ZFO-CeO2 interface atoms are well connected, without crystal structure breakage and presented in Fig. S4 (c). Therefore, we conclude that interface disorder has been successfully created at ZFO/CeO2 hetero-interface because of weak M-O bonding, which can be beneficial for ionic migrations.
Furthermore, the interfacial diagram and electronic states can be verified by the calculated DOS, as illustrated in Fig. 3(c). The crystal orbital overlap population (COOP) suggests a strong relation between ZFO and CeO2 lattices. COOP calculations elaborate the projection of DOS based on specific molecular bonds, which can be constructed as a complementary tool to understand the projected DOS aspects37. COOPs can further reveal the oxygen vacancy concentration and additional bonding information of the built structures. In our case, COOP demonstrates that the oxidation between ZFO and CeO2 is significant in understanding the constructed heterostructure electronic states. The band shifting towards the Fermi energy level is another crucial parameter to understand the DOS of ZFO-CeO2 heterostructure. Naveed et al.38 reported that the band moving near the Fermi level is more favorable to enhance the DOS of neighboring atoms, and the improved DOS at the Fermi level may cause increased efficiency of ionic transport. Our results show that the total DOSs of simple ZFO and CeO2 are far from the fermi level compared to the case of ZFO-CeO2 heterostructure. Moreover, the DOSs of each element in the ZFO-CeO2 heterostructure present a shift towards the fermi level, indicative of an increased transfer rate of the electrons to the adsorbed oxygen species. The higher DOS in ZFO-CeO2 heterostructure, especially in Ce neighboring atoms near the fermi level, can be a reason to increase the charge transfer rate.
When using the ZFO-CeO2 heterostructure to replace conventional electrolytes to fabricate PCFCs, one question may be raised that how to avoid electronic short-circuit problem. Previous studies have proposed to use hetero-junctions to regulate the charge carriers of semiconductors to evade the current leakage risk of fuel cells23, 39. Typically, the p-n bulk-heterojunction based on p-type and n-type semiconductors40, and the metal-semiconductor Schottky formed between the reduced anode (metal or alloy) and the semiconductor electrolyte have been applied in design PCFCs41. In this study, a catalyst Ni0.8Co0.15Al0.05LiO2−δ (NCAL) with high H+/O2−/e− triple conductivity and good catalytic activity was employed as the symmetrical electrodes42. At the anode side, the NCAL can be reduced into Ni/Co alloys in an H2 environment, thus establishing interaction with the semiconductor heterostructure to build a Schottky contact. The Schottky junction involves the propagation of space charge region to block electrons from passing through the anode/electrolyte interface to block electrons crossover and avoid the short-circuiting issue in the fuel cell. Figure 3(d) shows the electron density difference (EDD) of ZFO-CeO2 heterostructure under the equally applied strains to investigate the charger transfer rate. The different positive and negative regions represent the accumulation of electrons in the ZFO-CeO2. The positive region in the EDD indicates electron accumulation, and the negative region reflects the electron depletion of the fabricated device.
Conducting property and electrochemical performance. To understand the oxidation states of the constructed heterostructure, X-ray photoelectron spectra (XPS) measurement was performed to verify the surface chemical states of these samples. Figure 4(a) exhibits the XPS survey spectra of pure ZFO and CeO2 compared to the ZFO-CeO2 heterostructure. The spectra confirm the presence of Zn, Fe, Ce, and O elements on the surface of the ZFO-CeO2 sample. For CeO2, it was observed in work reported by Zhang et al. 43 that the ceria oxidation states play a crucial role in enhancing oxygen vacancies at the constructed sample interface. Figure 4 (b and c) presents the O 1s core-level spectra of pure ZFO and the ZFO-CeO2 heterostructure for comparison. The O 1s peaks of the two samples are deconvoluted into two peaks at 528.7-530.1 eV and 528.4-530.8 eV, respectively, corresponding to adsorbed oxygen specie (Oads) and lattice oxygen species (Olattice). It is calculated that the relative ratio value of (Oads/Olattice) increases from 0.654 for ZFO to 0.817 for ZFO-CeO2, indicative of significantly incremented oxygen vacancies on the surface of the heterostructure sample. Cai et al. 44 attributed the higher relative ratio of from 1.13 for SDC to 1.21 for SDC–STO heterostructure as a result of improved concentration of hetero-interface oxygen vacancies due to lattice mismatch, which would lead to enhanced ionic conductivity.
Figure 4(d and e) displays the measured results of ultraviolet photoelectron spectroscopy (UPS) for pure ZFO and ZFO-CeO2 heterostructure. The optical bandgap for all synthesized samples is calculated using the Kubelka − Munk function. The valance band (VB) maxima in the UPS exhibited an energy cutoff of 21.20 eV, calibrated by He-I light energy. The VB maxima of both ZFO and ZFO-CeO2 heterostructure-based samples are determined by the cutoff of low binding energy and high binding energy level. The UPS measurements show valance band energy levels are 6.17 eV for ZFO and 6.88 eV for ZFO-CeO2 heterostructure. In parallel, the conduction band edges for all synthesized samples are calculated to be 4.33 eV to 4.47 eV. The detailed bandgap analysis of ZFO-CeO2 heterostructure has been carried out using DFT calculations.
Moreover, the fuel cells electrochemical impedance spectroscopy (EIS) was measured at 360\(-\)510 °C under open circuit voltage (OCV) conditions. Typical EIS spectra of ZFO, CeO2, and ZFO-CeO2 based PCFCs measured at 510°C are presented in Fig. 4(f), and the corresponding fitting parameters based on the EIS measured at 485-360°C are displayed in Supplementary Fig. S7 (a, b and c). An equivalent circuit of LRo(R1-Q1)(R2-Q2) was used to fit the experimental data recorded at 510°C by ZSIMPWIN Software, where R represented resistance and Q was the constant phase element representing a non-ideal capacitor. The fitting parameters of ZFO, CeO2 and ZFO-CeO2 devices are listed in Supplementary Tables S1, S2, and S3, respectively. According to the calculated capacitance for each resistance, the R1 can be ascribed to the charge transfer process while the R2 is associated with the mass transfer process. The smaller values of Ro for ZFO-CeO2 PCFC compared to the other two cells indicate higher ionic conduction of ZFO-CeO2 than pure ZFO and CeO2. The ZFO-CeO2 PCFC also presents lower values of R1 and R2, suggesting that the device has higher electrode reaction activity while using the same NCAL electrodes. As reported, the grain boundaries (interface regions) play a crucial role in ion transportation, leading to promoted grain boundary conductance and fast electrode reactions 45, 46, 47. Thus, we conjecture that the high interfacial ionic conductivity of ZFO-CeO2 brought about fast grain-boundary conduction and promoted both electrolyte ionic transport and electrode dynamic processes. The ionic conductivity (σi) was estimated based on the ohmic linear part of the I-V polarization curve and presented in Fig. 5(g). The obtained ionic conductivity of ZFO-CeO2 in the cell exhibits high values of 0.11-0.21 S/cm at 460-510°C, which is remarkably higher than that of pure ZFO and CeO2. The corresponding activation energy (Ea) can be calculated from Arrhenius equation. As shown in the inset of Fig. 5(g), the activation energy values for ZFO-CeO2, CeO2 and ZFO are 0.63 eV, 0.80 eV and 0.88 eV, respectively. The small activation energy of ZFO-CeO2 reflects that proton transport dominates the total conductivity of ZFO-CeO2. It is also noticed that ZFO-CeO2 exhibits lower activation energy as compared to CeO2 and ZFO, which suggests that the ZFO-CeO2 sample probably gained improved proton conduction by forming a heterostructure.
To verify the proton conduction, the proton-conducting property of ZFO-CeO2 was investigated by isotopic effect in 5% H2/95% Ar and 5% D2 95% Ar atmospheres in comparison to air at different operational temperatures. The EIS measurements in the D2, H2 (5%), and H2/air atmosphere at 510°C are demonstrated in Fig S8 (Supplementary Information), while the calculated conductivities from EIS data are presented in Fig. 5(h). The temperature-dependent conductivity of the ZFO-CeO2 sample shows a clear H2-D2 isotopic effect, i.e., the conductivity is significantly higher in H2 (5%) that in D2. We can conclude that proton migration is more favorable in the ZFO-CeO2 heterostructure. Figure 4(i) presents the calculated conductivity of ZFO-CeO2 compared to the reported results of some typical electrolytes 48, 49, 50. As can be seen, the conductivity of ZFO-CeO2 is not only remarkably higher than those of YSZ and La0.9Sr0.1Ga0.8Mg0.2O3 (LSGM) but also superior to that of proton conductors BaZrO3 (BZY) and SmNiO3.
The proton conduction of ZFO-CeO2 heterostructure was further evaluated in PCFCs at 510 − 360°C. The current-voltage (I-V) and current-power (I-P) characteristics of PCFCs were measured in H2/air. The power outputs of the PCFCs based on ZFO, CeO2, and ZFO-CeO2 as electrolytes are presented in Fig. 5(a) for comparison, while I-V & I-P characteristics of the ZFO-CeO2 device at 510 − 360°C is shown in Fig. 5(b), and those for pure ZFO and CeO2 cells are shown in Supplementary Fig. S9 and S10. It is observed that the constructed fuel cells attained high OCVs above 1.05 V at 510 − 360°C. The ZFO, CeO2, and ZFO-CeO2 based devices achieved peak power densities of 427 mW/cm2, 515 mW/cm2, and 1070 mW/cm2 at 510°C, respectively, indicating noticeable performance increment by forming the heterostructure. The obtained higher power outputs can be ascribed to two aspects: i) the interfacial disordering leads to faster ionic transport at the interface, resulting in reduced ohmic resistance of the ZFO-CeO2 as compared to individual CeO2 and ZFO; ii) the hetero-junction provided BIEF at the interface, which accelerated the ions to enhance the mobility thus further reduce the ohmic resistance of ZFO-CeO2. Moreover, the peak power densities (PPDs) of ZFO-CeO2 based cell reach 790 mW/cm2, 732 mW/cm2, 425 mW/cm2, 240 mW/cm2, 201 mW/cm2, 70 mW/cm2 at 485°C, 460°C, 435°C, 410°C, 385°C and 360°C, respectively, manifesting a good capability of low-temperature running, whereas the fuel cells with pure ZFO and CeO2 display much lower outputs and a difficulty to be operated below 460°C. In our measurements, it is also acquired that the synthesized ZFO-CeO2 with mass ratio of 2:8 is the optimal sample with the highest power density, while the performances of the cells with 3:7, 4:6, and 5:5 ZFO-CeO2 present lower power outputs as presented in Supplementary Fig. S11. Additionally, the ZFO-CeO2 fuel cell was assembled with a BaZr0.8Y0.2O3−δ (BZY) filter layer to verify the proton conductivity of ZFO-CeO2 in a configuration of Ni-NCAL/BZY/ZFO-CeO2/BZY/Ni-NCAL, as the BZY layer allows only protons to pass through the electrolyte with negligible oxide ions and electron migration 51. The I-V and I-P characteristics of the ion-filtering cell and the corresponding cross-sectional SEM image are presented in Fig. 5 (c) and Fig. 5(d), respectively. The ion-filtering cell achieved a PPD of 961 mW/cm2 at 510°C, which is slightly lower but comparable to the PPD of the ZFO-CeO2 cell without the filter layer. The slight loss can be attributed to the extra polarization resistance of the device caused by the two additional interfaces of BZY and ZFO-CeO2. For comparison, the ZFO electrolyte cell was also assembled with a BZY filter layer and achieved a PPD of 415 mW/cm2 at 510°C (Supplementary Fig. S12).
Eventually, the durability of the proposed electrolyte and fuel cells was measured. Figure 5(e) exhibits the working voltage of the ZFO-CeO2 cell under a stationary current density of 110 mA/cm2 lasting more than 120 h. The inset figure is an original image directly exported from the I-V data recording software after 72 h. The initial voltage drop is due to the activation process. It subsequently gradually reaches a stable state with a value of around 0.9 V in the following 110 h. Figure 5(f) shows a cross-sectional SEM of the cell after a durability test to assess the compatibility of the ZFO-CeO2 electrolyte with other components, in which the NCAL-Ni cathode, ZFO-CeO2 electrolyte, and NCAL-Ni anode can be distinguished. The electrolyte with a thickness of ~ 650 µm is gas-tight and well adhered to the nether porous electrode but shows partial delamination with the upper electrode, which is probably due to thermal incompatibility between the two materials or the squeezing damage when scissoring fuel cell pellet for SEM characterization. These results reflect the potential of ZFO-CeO2 for practical PCFCs application, while compatibility of the electrode materials with ZFO-CeO2 requires more consideration.