Qi
Wang
ab,
Jie
Hou
a,
Yun
Fan
a,
Xiu-an
Xi
a,
Jun
Li
ab,
Ying
Lu
ab,
Ge
Huo
ab,
Lin
Shao
ab,
Xian-Zhu
Fu
*a and
Jing-Li
Luo
*ac
aCollege of Materials Science and Engineering, Shenzhen University, Shenzhen 518060, China. E-mail: xz.fu@szu.edu.cn; Jingli.Luo@ualberta.ca
bKey Laboratory of Optoelectronic Devices and Systems of Ministry of Education and Guangdong Province, College of Optoelectronic Engineering, Shenzhen University, Shenzhen 518060, China
cDepartment of Chemical and Materials Engineering, University of Alberta, Edmonton, Alberta T6G 2G6, Canada
First published on 25th December 2019
The performance of low-temperature solid-oxide fuel cells (LT-SOFCs) is heavily dependent on the electrocatalytic activity of the cathode toward the oxygen reduction reaction (ORR). To overcome the obstacles of the poor activity and stability of traditional cathode materials, a Pr2BaNiMnO7−δ double-layered Ruddlesden–Popper structural oxide was developed that exhibits high ORR activity, exceptional low-temperature cell performance, long-term stability, and excellent chemical compatability with a BaZr0.1Ce0.7Y0.1Yb0.1O3−δ (BZCYYb) proton-conducting electrolyte. When Pr2BaNiMnO7−δ is used as the cathode electrocatalyst for BZCYYb-based SOFCs, it demonstrates a peak power density of 1070 mW cm−2 at 700 °C with excellent stability under 0.7 V after 100 h of discharging. Even at 500 and 400 °C, the peak power densities still reach 259 and 135 mW cm−2, respectively. The area-specific polarization resistance of this cell is 0.084 Ω cm2 at 700 °C under open-circuit voltage (OCV) conditions. Interestingly, the activation energy (Ea) of the polarization resistance derived from the assembled single cell is lower than that in the reported literature, corresponding to a value of 0.96 eV. The excellent performance is higher than those of the state-of-the-art Ruddlesden–Popper structure materials used as cathode electrocatalysts for LT-SOFCs. Thus, these results suggest that the double-layered Ruddlesden–Popper perovskite oxide Pr2BaNiMnO7−δ is a preeminent highly active low-temperature cathode material.
To achieve a feasible, alternative cathode, research efforts have concentrated on developing a novel cathode material with good properties. Traditional cathode materials, such as La0.6Sr0.4MnO3 (LSM), show a relatively high area specific resistance (ASR) compared to the available anode materials.13 A layered LnBaCo2O5+δ perovskite oxide cathode suffers from various stability issues during long-term operation, including A-site segregation, CO2 poisoning to form carbonates and unmatched TEC compared to those of typical electrolytes.14,15 As an alternative cathode material system, Ruddlesden–Popper (R–P) perovskite oxides with layered structures have attracted attention, in consideration of their extremely high oxygen ion conductivities and electronic conductivity inside the rock salt layer.11,16 R–P series materials have promising ORR applications for high-temperature SOFC electrodes. As reported in the literature, one layered R–P oxides have been studied extensively, such as La2NiO4 and Pr2NiO4,17–26 and Pr2NiO4 has better ORR activity than La2NiO4 or Nd2NiO4.27,28 In addition, according to the calculated and experimental results, the double-layered R–P phase structure La3Ni2O7 exhibits better oxygen ion and proton transport behaviour than single-layered La2NiO4.29 In the double-layered Ln3Ni2O7, the transition dopant metals (Mn, Fe, Co and Cu) are favourable for enhancing oxygen diffusion and proton diffusivity to some extent at the same time. Oxygen ion conduction is found to be closely related to the microelectronic structure and charge density gradients generated from the chemical bonding between O and the B-site atoms along the oxygen migration pathway. While, for proton diffusion, the dopants can weaken the proton association and enlarge the capacity of the ‘electron pocket’ around the Fermi level, which is beneficial to fast proton diffusivity in the R–P phase structure.29 Intuitively, the typological transfer mode for the R–P cathode is shown in Fig. 1, in which the proton electrolyte SOFC cathode allows simultaneous oxygen ion, electron and proton transportation. The ORR routes are similar to those reported in the literature,30 and a recent report31 demonstrated that oxygen vacancy is not the key point for R–P oxides to realize high ORR activity at high temperature. Instead, a high concentration and fast migration of the interstitial oxygen
and lattice oxygen with high activity are favourable for high-temperature catalytic activity. High-valence ion doping can enhance the
concentration and the lattice oxygen activity.
![]() | ||
| Fig. 1 Schematic diagram of R–P cathode with different transportation routes (including e−, O2−, H+). | ||
Inspired by previous research on the R–P phase cathode, in this study we have developed a novel double-layered R–P structure Pr2BaNiMnO7−δ cathode material. Normally, the structural stability of ABO3 perovskite is related to the ionic radii of the A- and B-sites and the stability is usually evaluated by the Goldschmidt tolerance factor ‘t’ (eqn (1)).32–34 Consideration is given to the ionic radii of Ba2+ (12-coordination with oxygen, 1.61 Å) and Pr3+/Pr4+ (8-coordination, 1.126/0.96 Å) compared with Ni3+ (6-coordination with oxygen, high spin, 0.6 Å) and Ni2+ (6-coordination with oxygen, 0.69 Å). If the A-site is Pr3+/Pr4+, the ‘t’ value is about 0.80–0.86, which is lower than 1. If the A-site is changed to Ba2+, the ‘t’ value is very close to 1 and Ba2+ can easily form a 12-coordination structure with oxygen. In this case, the stability of the perovskite structure layer is significantly improved when the Pr3+/Pr4+ are partially replaced with Ba2+. In the B-site, the ionic radii and valence state of doped Mn3+/Mn4+ (6-coordination, 0.645/0.53 Å) is lower than that of Ni2+/Ni3+. The Mn substitution in the B-site lattice in the perovskite structure layer could cause some distortion to the BO6 octahedra. Further, the changed microelectronic structures of perovskite may improve the oxygen ion conduction.29 Hence, in the selected Pr3Ni2O7 system, the barium element is doped into the A-site for considerations of structural stability, and the aliovalent Mn (Mn3+/Mn4+) cation doping in the B-site is an effective strategy for modifying the catalytic activity. Pr2BaNiMnO7−δ (PrBaNiMn) was synthesized via a facile citric-acid–nitrate combustion process. The physicochemical and electrical properties of the powders and the anode-supporting single-cell NiO–BZCYYb/AFL/BZCYYb/PrBaNiMn were also investigated using corresponding techniques. Finally, the electrochemical performance in terms of the power output and the polarization impedance of the SOFC with the PrNaNiMn cathode was evaluated to demonstrate the feasibility of the materials as promising ORR catalysts for LT-SOFCs.
![]() | (1) |
:
1
:
1
:
1 were dissolved and well mixed, with citric acid as the chelating agent n(citric acid)/n(metals) = 1.5 in the raw material solution. Then, the solution was heated to evaporate under continuous stirring until it changed into a sol–gel and finally ignited into a flame. The electrolyte powder BZCYYb was synthesized using BaCO3, Zr(NO3)4·5H2O, Ce(NO3)3·6H2O, Y(NO3)3·6H2O, and Yb(NO3)3·6H2O. After combustion, the two as-prepared ash-like precursors were calcined at 1250 and 1000 °C for 5 h in air to obtain PrBaNiMn and BZCYYb powders, respectively.
:
35 weight ratio) was also prepared by the one-step gel combustion process and calcined at 1000 °C for 5 h. The 80 wt% composite powder NiO–BZCYYb and pore-forming material (20 wt% starch) were well mixed to prepare a porous anode powder. The anode supported half cells (NiO–BZCYYb/AFL/BZCYYb) with a tri-layered structure were fabricated by a co-pressing method36 and then co-sintered at 1350 °C for 5 h at a 3 °C min−1 heating rate. The cathode paste was obtained by mixing the PrBaNiMn powder and a 10 wt% ethylcellulose–terpineol binder to get PrBaNiMn ink. Then, the ink was painted onto a dense BZCYYb electrolyte membrane surface with a 0.28 cm2 effective area and about 20 μm thickness, and sintered at 1100 °C for 3 h in air to get a porous cathode layer. Ag paste and wire were applied as a current collector and the conducting wire, respectively. Then, the button anode supported single cells were installed on an alumina supporting tube for further measurements. A scheme of the fabrication process for anode-supported single cells is shown in Fig. S1.†
:
1 calcined at 1000 °C for 5 h. Then, the mixture was studied to characterize the crystalline structure via XRD.
The valence states and the composition of the elements on the PrBaNiMn surface were investigated using an X-ray photon spectroscopy (XPS, Thermo ESCALAB 250) method. The thermal expansion coefficient (TEC) of the PrBaNiMn sample was measured using a thermal expansion instrument (DIL402C). For this measurement, 2 g of PrBaNiMn powder were dry-pressed into a rectangular bar of dimensions 2.50 × 0.45 × 0.30 cm (length × width × height) after sintering at 1450 °C for 10 h in air. The performance of the button single fuel cells installed on an alumina supporting tube was measured at 400–700 °C with wet hydrogen (∼3% H2O) as fuel and stationary air as oxidant. The linear sweep voltammetry (LSV) and electrochemical impedance spectroscopy (EIS) characteristics of the cells were measured using an electrochemistry workstation (Solartron 1287 and 1260). The EIS measurements were carried out under open circuit conditions at 5 mV with AC amplitude 0.1–100 kHz. The cell ohmic and polarization resistances were clearly obtained from the EIS under open circuit conditions. The morphology, microstructure, and mapping of the tested single cell were examined using a scanning electron microscope (SEM, Hitachi, SU-70). The HRTEM and lattice fringes of the cathode material PrBaNiMn were measured by JEM-F200 at 200 kV accelerating voltage.
along the (a–b) plane. And the perovskite-like sheets have rich oxygen vacancies and excellent electrical conductivity which are beneficial for the ORR and electron transportation. In this case, the obtained powder PrBaNiMn should be a good cathode material for SOFCs. In addition, the XRD patterns of the powder mixture of PrBaNiMn and BZCYYb (50
:
50 wt%) after annealing at 1000 °C for 5 h (Fig. 2c), with no detectable impurity peaks, suggests good chemical compatibility between the cathode and the electrolyte phases.
Fig. 3a displays a TEM image of PrBaNiMn powder with an average particle size of ∼300 nm. Corresponding energy dispersive X-ray spectroscopy (EDS) mapping images of PrBaNiMn for Pr, Ba, Ni, Mn and O elements exhibit a homogeneous distribution, and no observable elemental segregation could be detected (Fig. S2†). All this evidence suggests that PrBaNiMn double-layered R–P perovskite oxides have been successfully prepared without a secondary phase or elemental segregation, which is essential for the R–P structural cathodes to take part in an ORR in the fuel cell operating process. In addition, from the high-resolution TEM image (Fig. 3b and c) of the powder, the interplanar crystal spacing is calculated to be 3.804 Å, which corresponds to the (101) plane for the R–P structure (space group I4/mmm). The SEM image of the powder indicates the porous structure and good dispersibility of the prepared PrBaNiMn, which is beneficial for oxygen adsorption and transportation (Fig. 3d).
The elemental compositions and valence states of PrBaNiMn were investigated by XPS. Table 1 shows the elemental compositions (Pr, Ba, Ni, Mn, O), valence states and content of the elements with multiple valence states (Pr, Ni, Mn) which are calculated from the peak fitting by XPSPEAK4.1 software. The elemental compositions are close to the stoichiometric ratio of the designed material. In A-sites, the Pr 3d spectrum of PrBaNiMn (Fig. 4a) comprises four intense peaks at 928.5 (Pr3+ 3d5/2), 948.4 (Pr3+ 3d3/2), 932.8 (Pr4+ 3d5/2) and 952.8 (Pr4+ 3d3/2), respectively, with a splitting energy of about 20 eV between 3d5/2 and 3d3/2 core-levels, in agreement with the reported literature.38 The spectrum of the satellite located at 956.5 eV corresponds to Pr6O11 (Fig. 4a) which is associated with the multiple valence effect of the Pr and oxygen O KLL Auger peak at 971.5 eV.38–40 The tetravalent Pr atoms are the charge compensation for the Ba atoms with bivalence. For the B-site atoms Ni and Mn, both have two oxidation states: these are 854.2 eV (Ni2+ 2p3/2) and 855.5 eV (Ni3+ 2p3/2) for Ni atoms (Fig. 4b), and 642.1 eV (Mn3+ 2p3/2), 653.8 (Mn3+ 2p1/2), 645.6 eV (Mn4+ 2p3/2) and 658.2 eV (Mn4+ 2p1/2) for Mn atoms (Fig. 4c).41,42 The discrepancy between the Ni2+/Ni3+ and Mn3+/Mn4+ atomic ratios at the B-sites in the surface of the sample reflect the charge and structural adjustments. Under these conditions, the percentage of Oads (adsorbed oxygen) at 531.2 eV is higher than the Olatt (lattice oxygen) at 529.0 eV, which is beneficial for oxygen ion transportation and ORR at the surface of the samples (Fig. 4d). What is more, the rich Oads can increase the proton (H+/OH−) conduction, which can enhance the cathode activity.29 The 7.34% oxygen species at 532.8 eV for O 1s indicates the porous surface in the contaminants (hydroxyl/carbonyl groups).43 The Ba 3d5/2 peaks show that two chemical states are present (Fig. S3†). The lower binding energy (BE) peak at 778.2 eV can be related to PrBaNiMn while the higher BE peak at 779.8 eV is typical of BaCO3 as a result of the adsorption of CO2 by the sample in an ambient environment.44
| The sample PrBaNiMn surface elemental composition (at%) | ||||
|---|---|---|---|---|
| Pr 3d | Ba 3d | Ni 2p | Mn 2p | O 1s |
| 14.42 | 7.10 | 5.44 | 6.22 | 66.82 |
| Pr3+ | Pr4+ | Ni2+ | Ni3+ | Mn3+ | Mn4+ | Olatt | Oads | H2O |
|---|---|---|---|---|---|---|---|---|
| 30.59 | 69.41 | 30.09 | 69.91 | 81.12 | 18.88 | 44.11 | 48.55 | 7.34 |
Along the ab-plane in the R–P structure, the rock-salt A2O2 layers leave favorable transportation pathways for
and the electron transportation occurs predominantly along the perovskite-like layers. Because of the variable oxidation state in both the A- and B-site cations of the R–P structure, the O2− can transport via the 3D channels in the perovskite-like layers by the hopping mechanism and in the rock-salt layers by the
migration mechanism. The formation of oxygen vacancies
can be expressed by eqn (2)–(4). Although the perovskite-like layers do not allow for the existence of interstices, the O2− ions can diffuse via the BO6 octahedra with a corner-shared structure between perovskite-like layers and the rock-salt layers (shown in Fig. 2b). As mentioned above, the obtained PrBaNiMn should be a good cathode material in SOFC with nice ORR activity.
![]() | (2) |
![]() | (3) |
![]() | (4) |
These three redox couples accelerate the ORR reaction, which could be described in Kröger–Vink notation as follows:45,46
![]() | (5) |
It is generally accepted that the ORR at the cathode of H–SOFC could be described with the most plausible rate-limiting steps 1–4, as follows:
Step 1: O2 → 2Oad (dissociative adsorption).
Step 2: Oad + e− → Oad− (charge transfer).
Step 3: Oad− → OTPB− (surface diffusion, TPB, three-phase boundary).
Step 4: OTBP− + e− → OTPB2− (charge transfer at TPB).
The adsorbed Oad, Oad− and Oad2− could be incorporated into the PrBaNiMn phase, combining with the oxygen vacancies directly in the double-layered R–P structure cathode PrBaNiMn, which could be written as steps 5–7.
Furthermore, the OTPB2− at the TPB of PrBaNiMn–BZCYYb-gas could be injected into the PrBaNiMn phase (step 8) and react with HTPB+ conducted from the electrolyte (steps 9–12).
Step 9: Helectrolyte+ → HTPB+.
Step 10: OTPB2− + HTPB+ → OHTPB−.
Step 11: OHTPB− + HTPB+ → H2OTPB.
Step 12: H2OTPB → H2O(g).
![]() | ||
| Fig. 5 Cross-sectional SEM images of the single cell. (a) The four-layer single cell with an anode functional layer (AFL) and PrBaNiMn cathode after testing; (b) PrBaNiMn cathode layer; (c) the cathode/BZCYYb; (d) BZCYYb electrolyte; (e) AFL/BZCYYb; (f) the porous anode NiO–BZCYYb and anode functional layer (AFL) AFL provides abundant triple-phase boundaries (TPBs) with the electrolyte allowing fast dissociation of the fuel gas and the rapid occurrence of charge transfer after the reduction of NiO (Fig. 5e). The larger pores in the anode support obtained from the pore-forming materials offer rapid diffusion of fuel gas to the AFL layer and then to the TPBs (Fig. 5f). The pore size distribution of the anode support plates is measured by the mercury porosimeter method described in theESI.†47 As shown in Fig. S5 and S6,† the cumulative intrusion of mercury has three stages as the pressure increases and the corresponding pore size distributions are about 100–200 μm, 1–0.4 μm and 50–250 nm. The rational distributions of pore sizes for the anode support accelerate the fuel gas diffusion, transmission and reaction. On the basis of the superior single-cell components and microstructures, infusive electrochemical performance in an intermediate temperature range should be achieved when a PrBaNiMn cathode is applied to BZCYYb-based solid-oxide fuel cells. | ||
| Year | Cathode | Electrolyte (thickness, μm) | 700 | 650 | 600 | 550 | 500 | 450 | 400 | Ref. |
|---|---|---|---|---|---|---|---|---|---|---|
| 2009 | Pr2NiO4+δ | BaCe0.9Y0.1O3−δ (40) | 132 | 96 | 53 | 55 | ||||
| 2014 | La2NiO4+δ | BaZr0.1Ce0.7Y0.2O3−δ (20) | 398 | 283 | 196 | 120 | 56 | |||
| 2017 | La3Ni1.6Co0.4O7−δ | BaZr0.3Ce0.5Y0.2O3−δ (20) | 398 | 293 | 215 | 50 | ||||
| 2018 | SrEu2Fe1.8Co0.2O7−δ | BaZr0.1Ce0.7Y0.2O3−δ (15) | 562 | 365 | 210 | 51 | ||||
| 2018 | La1.2Sr0.8NiO4 | BaZr0.1Ce0.7Y0.2O3−δ (30) | 461 | 330 | 223 | 49 | ||||
| 2018 | Pr1.2Sr0.8NiO4 | BaZr0.1Ce0.7Y0.2O3−δ (30) | 352 | 212 | 127 | 49 | ||||
| 2019 | La2NiO4+δ nanofiber | BaZr0.1Ce0.7Y0.2O3−δ (15) | 508 | 349 | 246 | 175 | 57 | |||
| 2019 | La1.2Sr0.8Ni0.6Fe0.4O4+δ | BaZr0.1Ce0.7Y0.2O3−δ (15) | 782 | 658 | 421 | 257 | 139 | 58 | ||
| This work | PrBaNiMn | BaZr0.1Ce0.7Y0.1Yb0.1O3−δ (12) | 1070 | 776 | 570 | 407 | 259 | 245 | 135 |
In detail, the activation energy Ea of the assembled single cell in this work was calculated with the Arrhenius equation and compared with various R–P single-phase cathodes for BaCeO3-based proton electrolytes under the same conditions (hydrogen with about 3% H2O as fuel gas).
As shown in Fig. 6e, the Ea of the RP of the PrBaNiMn cathode is lower than that in the reported literature, corresponding to the value of 0.96 eV, being comparable with the singe-phase cathode with a one-layer R–P structure La1.2Sr0.8NiO4+δ and Pr1.2Sr0.8NiO4+δ with Ea values of 1.27 and 1.38 eV, respectively.49 In addition, the PrBaNiMn cathode also has a relatively lower Ea than double-layered R–P structure cathode materials, such as La3Ni1.6Co0.4O7−δ or SrEu2Fe1.8Co0.2O7−δ.50,51 Although the La2NiO4+δ nano-fiber presents relatively lower Ea at 0.98 eV, which is also little higher than the PrBaNiMn cathode in this study. Hence, it is clear that PrBaNiMn is a novel single-phase cathode material for H–SOFC combined with the compatibility analysis by XRD and EDS mapping mentioned above. And the thermal expansion coefficient (TEC) of the prepared PrBaNiMn is 10.35 × 10−6 K−1 (Fig. 7), which is close to that of the BaCeO3-based electrolyte.52 In this case, it possesses better compatibility with the electrolyte than conventional perovskite materials. From the XPS study, the variable valence of the cations in the R–P structure can accelerate the oxygen transfer ability and oxygen reduction reaction catalytic activity. The lower Ea of the PrBaNiMn cathode can mitigate the performance degradation of the cell with a reduction in operation temperature. As expected, the performance of the PrBaNiMn cathode may be improved further when it is prepared by special synthesis methods or the particle size is decreased by reducing the roasting temperature with the guarantee of a single R–P phase structure. The molten salt,53 electrostatic spinning7 and microwave sintering54 methods may provide a new approach to improve the properties of the cathode further. In the stability test, the single cell with the PrBaNiMn cathode remains stable for 100 hours with no degradation and a power output of around 350 mW cm−2 under a cell voltage of 0.7 V at 600 °C (Fig. 6f). This demonstrates that the PrBaNiMn cathode possesses fine chemical stability and compatibility with the electrolyte layer against water vapor and CO2 under working condition. Combined with the higher peak power density and lower RP compared with other reported cathodes under similar testing conditions, PrBaNiMn could be an outstanding alternative cathode material for proton-conducting SOFCs (H–SOFCs).
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ta11212j |
| This journal is © The Royal Society of Chemistry 2020 |