Next Article in Journal
Analogous Anti-Ferroelectricity in Y2O3-Coated (Pb0.92Sr0.05La0.02)(Zr0.7Sn0.25Ti0.05)O3 Ceramics and Their Energy-Storage Performance
Next Article in Special Issue
Effect of A-Site Nonstoichiometry on Defect Chemistry and Electrical Conductivity of Undoped and Y-Doped SrZrO3
Previous Article in Journal
Effect of Recrystallization Annealing on Corrosion Behavior of Ta-4%W Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Reversible Protonic Ceramic Cell with Symmetrically Designed Pr2NiO4+δ-Based Electrodes: Fabrication and Electrochemical Features

1
Laboratory of Electrochemical Devices Based on Solid Oxide Proton Electrolytes, Institute of High Temperature Electrochemistry, Yekaterinburg 620137, Russia
2
Institute of New Materials and Technologies, Ural Federal University, Yekaterinburg 620002, Russia
3
Institute of Chemical Engineering, Ural Federal University, Yekaterinburg 620002, Russia
4
Laboratory of Solid State Oxide Fuel Cells, Institute of High Temperature Electrochemistry, Yekaterinburg 620137, Russia
5
Graduate School of Economics and Management, Ural Federal University, Yekaterinburg 620002, Russia
*
Author to whom correspondence should be addressed.
Materials 2019, 12(1), 118; https://doi.org/10.3390/ma12010118
Submission received: 23 November 2018 / Revised: 12 December 2018 / Accepted: 26 December 2018 / Published: 31 December 2018

Abstract

:
Reversible protonic ceramic cells (rPCCs) combine two different operation regimes, fuel cell and electrolysis cell modes, which allow reversible chemical-to-electrical energy conversion at reduced temperatures with high efficiency and performance. Here we present novel technological and materials science approaches, enabling a rPCC with symmetrical functional electrodes to be prepared using a single sintering step. The response of the cell fabricated on the basis of P–N–BCZD|BCZD|PBN–BCZD (where BCZD = BaCe0.5Zr0.3Dy0.2O3−δ, PBN = Pr1.9Ba0.1NiO4+δ, P = Pr2O3, N = Ni) is studied at different temperatures and water vapor partial pressures (pH2O) by means of volt-ampere measurements, electrochemical impedance spectroscopy and distribution of relaxation times analyses. The obtained results demonstrate that symmetrical electrodes exhibit classical mixed-ionic/electronic conducting behavior with no hydration capability at 750 °C; therefore, increasing the pH2O values in both reducing and oxidizing atmospheres leads to some deterioration of their electrochemical activity. At the same time, the electrolytic properties of the BCZD membrane are improved, positively affecting the rPCC’s efficiency. The electrolysis cell mode of the rPCC is found to be more appropriate than the fuel cell mode under highly humidified atmospheres, since its improved performance is determined by the ohmic resistance, which decreases with pH2O increasing.

Graphical Abstract

1. Introduction

Solid oxide systems with predominant protonic transport are considered as advanced applied materials for the energy sector. A particular interest is associated with their utilization as proton-conducting electrolytes in protonic ceramic cells (rPCCs), reversible solid oxide cells that are used to convert different types of energy with high efficiency and no harmful impact [1,2,3,4,5]. Compared with conventional solid oxide cells with oxygen-ionic electrolytes, rPCCs are able to operate at reduced temperatures due to their high proton mobility and low activation energy [6,7,8]. These advantages may feasibly lead to their commercialization in the near future [9]. Therefore, carrying out the necessary applied and fundamental research in this field, as well as developing new approaches for the optimization of the relevant technological and electrochemical processes, are among highly relevant current trends.
Recently, many efforts have been made to simplify technological processes involved in the fabrication of solid oxide devices based on oxygen-ionic or proton-conducting electrolytes [10,11]. One of these efforts consists in designing electrochemical cells having symmetrical electrodes [12] as an efficient strategy for reducing fabrication costs. Here, the economic benefit lies in minimizing the number of functional materials used. Moreover, this strategy can help to resolve problems associated with thermal incompatibility and electrochemical degradation if the latter has a reversible nature [13,14]. However, materials for symmetrical electrodes preparation must satisfy a number of requirements, such as excellent electrochemical activity in both reduced (Red) and oxidized (Ox) conditions, as well as appropriate phase and thermal behavior under RedOx cycling. As a rule, only a limited series of materials can be used for this purpose, including Ti-, Fe-, Cr- and Mn-based oxides with simple or double perovskite structures [12].
Although the utilization of symmetrical solid oxide fuel cells based on oxygen-ionic electrolytes has been intensively studied in the past few years [13,14,15,16,17], the application of this strategy to PCCs is only at the beginning of its development. In the present work, we propose to use a Pr2NiO4+δ-based oxide in symmetrical functional layers for a rPCC. In contrast to the previously mentioned simple or double perovskite, a layered structure of praseodymium nickelate is substituted under reducing atmospheres to complete Ni reduction and formation of a Ni–Pr2O3 cermet (Figure 1) with good electrocatalytic properties [18,19]. Along with the symmetrical electrode application and reversible operation mode, the close thermal expansion coefficients (TECs) of Pr2NiO4+δ- and Ba(Ce,Zr)O3-based materials [20,21] allow a one-step sintering procedure to be used. According to the literature analysis, a single temperature processing step is a highly attractive approach for reducing the fabrication costs; in particular, this strategy has recently been adopted for the production of PCCs [8,22,23]. However, in these works the anode and cathode layers consisted of different functional materials, which can cause a mechanical misbalance leading to the deformation of whole cells following their sintering. Utilizing the same component for both functional electrode layers minimizes the possible strain, representing significant benefits in technological aspects, as well as the in terms of the quality of the target product.

2. Materials and Methods

2.1. Preparation of Materials

The protonic ceramic cell (PCC) was fabricated from three functional materials, including BaCe0.5Zr0.3Dy0.2O3−δ (BCZD) as a proton-conducting electrolyte layer (EL), its mixture with nickel oxide as a substrate for supporting fuel electrode layer (SFEL) and Pr1.9Ba0.1NiO4+δ (PBN) as a basis for functional oxygen (FOEL) and functional fuel (FFEL) electrode layers.
The BCZD and BPN powders were obtained using the citrate-nitrate synthesis method described in detail in our previous works [24,25].

2.2. Characterization of Materials

The phase structure of the individual materials (BCZD, PBN) and their mixture (1:1 at ratio) calcined at 1350 °C for 3 h was studied by X-ray diffraction analysis (Rigaku D/MAX-2200VL [26]). The scans were performed under CuKα1 radiation between 20° and 80° with a scan step of 0.02° and a scan rate of 3° min−1.
The morphology of the ceramic samples and multilayered cell was studied using scanning electron microscopy (SEM) and energy dispersive X-ray (EDX) analyses on a Tescan Mira 3 LMU microscope with an Oxford Instruments INCA Energy X-MAX 80 spectrometer [26].
Thermogravimetric (TG) technique (Netzsch STA 449 F3 Jupiter) was employed to study reduction behavior of the PBN powder.
Conductivity measurements were carried out in air and wet H2 atmospheres using a convenient 4-probe DC current method (Zirconia-318 measurement station).

2.3. Fabrication of the PCC

The single-phase BCZD was mixed with NiO and starch (pore former) in a weight ratio of 3:2:1 to prepare the powder for the SFEL, while BCZD was mixed with PBN and pore former in a weight ratio of 4:1:1 to be used as FOEL and FFEL. The mixing stages were performed using a Fritsch Pulverisette 7 planetary ball mill with the following conditions: zirconia milling balls, acetone media, 500 rpm for 0.5 h. The corresponding powders were mixed with an organic binder (butadiene rubber in acetone/benzene solvent) and rolled used a Durtson rolling mill to fabricate the functional films having the required thicknesses. These films were then co-rolled with each other (tape-calendering method), adjusted by ~30 µm for the raw EL, ~20 µm for the raw FOEL and FFEL and ~800 µm for the raw SFEL. The green multilayered NiO–BCZD|PBN–BCZD|BCZD|PBN–BCZD structure was slowly (1 °C min−1) heated up to 900 °C to decompose and remove organic residue, then heated further to 1350 °C at a heating rate of 5 °C min−1, sintered at this temperature for 3 h and cooled to room temperature at a cooling rate of 5 °C min−1.

2.4. Characterization of the PCC

The fabricated PCC with an effective electrode area of 0.21 cm2 was characterized at a temperature range of 600–750 °C under reversible operation (bias changes from 0.1 to 1.5 V with a step of 25 mV). The volt-ampere dependences and impedance spectra were obtained using a complex of Amel 2550 potentiostat/galvanostat and Materials M520 frequency response analyzer. To evaluate the electrode response, hydrogen as a fuel and air as an oxidant were humidified to varying degrees. The target water vapor partial pressure values were set by passing the corresponding atmospheres through a water bubbler heated to certain temperatures. The impedance spectra were obtained across a frequency range of 10−2–105 Hz at a perturbation voltage of 25 mV and were then analyzed utilizing the methods of equivalent circuits (Zview software) and distribution of relaxation times (DRT, DRTtools core of the Matlab software [27]).

3. Results and Discussion

3.1. Pr1.9Ba0.1NiO4+δ Functionality

To assess the applied prospect of PBN as electrodes for PCCs, its properties were comprehensively studied (Figure 2).
This material is characterized by a high chemical compatibility with the proton-conducting electrolytes based on Ba(Ce,Zr)O3 (Figure 2a), since the high-temperature treatment results in maintaining the basic structures for both phases, although a small amount of NiO exists at the same time (this impurity phase is also visible at the interface regions, see the Ni-element distribution map on Figure 3c). According to our best knowledge, we have used the highest temperature (1350 °C), at which chemical interactions of cerate-zirconates and nickelates were studied. Such excellent results in phase stability can be explained by the fact that Pr2NiO4+δ is doped with barium, that compensates the Ba-concentration difference between two components and, therefore, diminishes a degree of chemical interaction [28,29].
As it is shown in Figure 2b, the PBN material in a powder state starts to reduce at 300 °C, when interstitial (and a part of lattice-site) oxygen is gradually removed [30]; this nickelate is almost completely decomposed at temperatures above 600 °C, which is in line with the following simplified reaction: Pr2NiO4+δ→Pr2O3 + Ni, see Figure 1.
Conductivity of the as-prepared ceramic sample in air atmosphere differs from one of the reduced samples in wet hydrogen by more than four orders of magnitude (Figure 2d). It can be concluded that the reduced sample represents a mixture of (Pr,Ba)2O3 and Ni with a mole ratio of 2:1. In other words, the total conductivity is determined by oxide phases, since no continuous high conductive metallic framework is formed under such a reduction. Nevertheless, Ni-particles appear during the exsolution procedure (Figure 2c), which is found to be a remarkable factor in electrode processes improvement [31]. According to Figure 2c, a certain part of Pr2O3 is transformed to Pr(OH)3, but this hydroxide is formed at relatively low temperatures only (under cooling of the sample in static air), when Pr2O3 chemisorbs steam of air. This is confirmed by the fact that the total weight change of PBN under full reduction is equal to 95.2% (δ@RT = 0.17, 4 + δ@1000 °C = 2.95; see Figure 2b) corresponding with the formation of 0.95 mole of Pr2O3 and 0.1 mole of BaO at 1000 °C. Otherwise, if Pr(OH)3 was formed during the reduction procedure, the overall weight change should amount to 96.1%. This level corresponds to δ@RT = −0.07 (4 + δ@1000 °C = 2.95), which is in disagreement with that reached for Pr2NiO4+δ, which has a close composition, δ = 4.23–4.25 [32,33].

3.2. Microstructural Features

The PCC fabricated using the one-step sintering procedure shows a well-organized, multi-layered structure without any visible deformation, material delamination or cracks (Figure 3a). The obtained results can be explained by the excellent thermal compatibility of the Ln2NiO4+δ-based (Ln = La, Pr, Nd) materials with Ba(Ce,Zr)O3 proton-conducting electrolytes, especially the low chemical expansion in contrast to that of many electrode materials having perovskite-related structures [34,35].
The resulting thickness of the FOEL, EL and FFEL are estimated to be about 15, 25 and 17 µm, respectively (Figure 3b), while the total PCC thickness is equal to 700 µm. The porosity of SFEL evaluated using ImageJ software amounts to 40 ± 5 vol.%; the porosity of the functional electrodes does not exceed 20% (measured on the individually prepared pellets with the same composition, BCZD:PBN:starch = 4:1:1), indicating their strong sintering behavior despite 20 wt.% of pore former used.
According to EDX analysis (Figure 3c), elements are evenly distributed and do not show significant interdiffusion, forming clearly separated interphase boundaries. This also supports the conclusion regarding the chemical compatibility of materials used, at least at a sintering temperature of 1350 °C.

3.3. Volt-Ampere Dependences and Related Properties

The operability of the reversible PCC is shown in Figure 4a. In the current-free mode, this cell generates open circuit voltages (OCV) of 1.076, 1.051, 1.006 and 0.979 V at 600, 650, 700 and 750 °C, respectively. These values are somewhat lower than those theoretically predicted, which amount to 1.135, 1.126, 1.118 and 1.109 V, respectively. Since this cell maintains gas-impenetrability (measured at room temperature under 10−3 atm/1 atm of total gas pressures gradient), the most likely reason for the observed differences is non-ionic conduction of the BCZD electrolyte. Indeed, both BaCeO3- and BaZrO3-based materials demonstrate meaningful electron conductivity in oxidizing conditions at high temperatures [36,37]. Regarding the BCZD electrolyte, its ionic transference numbers (ti) estimated as a ratio of its ionic conductivity measured in wet H2 to the total conductivity measured in wet air only reached 0.78, 0.62, 0.50 and 0.44 at 600, 650, 700 and 750 °C [38]. It should be noted that these ti values were determined for the samples in system with unseparated gas space, when all the samples’ sides were in contact with the same atmosphere. In the case of PCC, the electrolyte membrane separates two gas spaces. Therefore, one subsurface of the membrane exhibits predominant proton transport, while another one features mixed ionic-electronic transport. Therefore, the resulting (or average) ionic transference numbers, ti,av, should be much higher than ti. The exact ti,av values under OCV conditions can be calculated as follows [39]:
t i , av = 1 R O R O + R p ( 1 E meas E ) ,
where, RO and Rp are the ohmic and polarization resistances of the PCC, while Emeas and E are the measured and theoretically predicted potentials. As can be seen, this equation includes the parameters related to the resistances of the functional materials, which can be separated using the impedance spectroscopy method. The separation procedure will be described in Section 3.4 and the calculation of ti,av values shown in Section 3.6.
The PCC yields such maximal power densities (Pmax) as 305, 360, 395 and 430 mW cm−2 at 600, 650, 700 and 750 °C, respectively (Figure 4b,d). In electrolysis cell mode of operation, the maximal current densities reach about 640, 780, 950 and 1070 mA cm−2 (U = 1.5 V) at the corresponding temperatures and about 295, 405, 535 and 690 mA cm−2 under conditions close to the thermoneutral mode (UTN ≈ 1.3 V, Figure 4c,d). This mode is determined by the thermodynamic parameter of the resulting reaction occurring in the rPCC (Equations (2) and (3)) and corresponds to the conditions when the cell is in thermal equilibrium. More precisely, the rPCC consumes heat when the bias is lower than UTN and, conversely, produces heat when the bias exceeds UTN.
H 2 + 1 / 2 O 2 H 2 O ,
U T N = Δ H z F .
Here, ΔH is the molar enthalpy of reaction (2), z is the number of participating electrons (z = 2) and F is the Faraday constant.

3.4. Analysis of Impedance Data

First, the PCC’s functionality was characterized under OCV mode using EIS analysis (Figure 5). As can been seen from these data, all the obtained impedance spectra consist of two clearly separated arcs, corresponding to low- and high-frequency processes. In order to correlate these partial processes with the corresponding resistances, the experimental results were fitted by the model lines originated from the used L—RO—(R1Q1)—(R2Q2) equivalent circuit scheme. Here, L is the inductance associated with cables, wires and their junctions, RO is the ohmic resistance of the electrolyte membrane, R1 and R2 are the resistances of low-(I) and high-(II) processes, Q1 and Q2 are the corresponding constant phase elements. Two fitting results (presented in Figure 5 as examples) confirm a good agreement between experimental and model data, implying the success of the used scheme.
According to the fitting procedure, the RO, R1, R2 parameters, along with the polarization resistance of the electrodes (Rp = R1 + R2) and total resistance of the PCC (Rtotal = RO + Rp), were successfully determined, as shown in Figure 6. With increasing temperature, the total polarization resistance of the PCC decreases from 0.90 to 0.52 Ω cm2 (Figure 6a); at the same time, this resistance is determined by the ohmic component (Figure 6b), the contribution of which varies between 57 and 86%. Two factors contribute to this result: the rather high thickness of the electrolyte used (25 µm) and the excellent electrochemical properties of the electrodes, despite their fairly low porosity. The total polarization resistance of the electrodes decreases from 0.39 Ω cm2 at 600 °C to 0.07 Ω cm2 at 750 °C and is regulated by the R2 level, which contributes to amounts to 91% and 71%, respectively.
All the constituent resistances have a thermo-activated nature with different activation energies, Ea (see Appendix A, Figure A1). The Process II exhibits the highest Ea value, indicating a strong correlation with temperature; conversely, the Ea level of Process I is lower than that of Process II by ~3 times. The Ea level of the PCC’s total resistance is quite low (0.37 eV), since it is regulated by the predominant influence of the ohmic resistance with the lowest Ea. Two important conclusions are supported by the obtained data:
(1)
The low Ea for the ohmic resistance is an indirect evidence of proton behavior, since protons migrate much more easily than massive oxygen-ions [40,41];
(2)
The slight temperature behavior of the total resistance of the PCC is a characteristic feature of the cells, corresponding to the condition of Rp < RO. Therefore, their output parameters (as shown in Figure 4d) also change slightly with temperature variation.
In order to evaluate the prospect of utilizing the Pr2NiO4+δ-based material as symmetrical electrodes, Processes I and II were thoroughly analyzed by calculating the capacitance/frequency values and utilizing the DRT method.
The values for the capacitance and frequency characteristics (at the top of the arcs) were estimated from the impedance spectra fitting as follows:
C j = ( R j Q j ) 1 / n j R j 1 ,
f j = ( R j Q j ) 1 / n j ( 2 π ) 1 ,
where j = 1 or 2, nj is the exponent factor [42,43].
Process I semicircles are characterized by the characteristic capacitances of 3.6·× 10−3–6.4·× 10−3 F cm−2 and characteristic frequencies of 4 × 102–6 × 102 Hz; these values reach 8.5 × 10−1–9.0 × 10−1 F cm−2 and 4 × 10−1–4 × 101 Hz, respectively, for the semicircles of Process II. These results are also supported by the DRT data (Figure 7). As can be seen, Processes I and II correspond to the medium- and low-frequency stages, respectively. In detail, the first rate-determining step can be attributed either to the surface charge-transfer phenomenon [44] or to ionic diffusion at the electrode [45] with direct participation of proton charge carriers due to having a low Ea value (Figure A1). Although low values are achieved for this stage at the estimated frequencies, it should be noted that the electrodes (at least, the oxygen one) represent quite a dense structure for an electrolytic component, which might provide proton transportation. Due to high Ea and C2 values, the second rate-determining (and dominating) step corresponds to sluggish gas-diffusion and adsorption of electrochemically-active components [46,47], which correlates with the mentioned low porosity of the electrodes.
Along with these two steps, the high-resolution DRT method [48,49] gives an additional peak around tens of Hz (see inset in Figure 7). The partial polarization resistance corresponding to this peak (R1’) does not exceed 0.005 and 0.002 Ω cm2 (below 1.5% of Rp) at 600 and 650 °C, respectively; therefore, it cannot be resolved under convenient impedance data analysis by equivalent circuits. Considering its mediate frequencies and strong dependence on temperature, such a peak can likely be associated with the dissociation of adsorbed molecules [46,47].
Summarizing, the symmetrically-formed electrodes yield a promising performance of the PCC regardless of their low-porous microstructure. Moreover, the obtained results are in line with characteristics for similar PCC electrolyte and oxygen electrode materials (Table 1, [50,51,52,53]).

3.5. Effect of Air and Hydrogen Humidification

It is well-known that water vapor partial pressure is a parameter determining proton transport in proton-conducting electrolyte membranes [54]. From the viewpoint of the bulk structure of such membranes, humidification of both atmospheres is favorable, since it results in a decrease both of oxygen vacancies and hole concentrations:
V O + O O x + H 2 O 2 O H O ,
V O + 1 / 2 O 2 2 h + O O x ,
and, correspondingly, improved proton transport. At the same time, air humidification is considered to be a more effective and easy way of suppressing some of the undesirable electronic conductivity of cerates and zirconates [55,56].
From the perspective of thermodynamic features, the maximal achievable electrical potential difference of a PCC (E or OCV) decreases with gas humidification, which follows from the corresponding decrease in partial pressure gradients [57,58]:
E = t i , a v R T 4 F ln ( p O 2 p O 2 ) + t H , a v R T 2 F ln ( p H 2 O p H 2 O ) = t i , a v E O + t H , a v E H 2 O ,
where p’O2 and p’’O2 are the oxygen partial pressures in reducing and oxidizing atmospheres, p’H2O, p’’H2O are the water vapor partial pressures, EO and E H 2 O are the electrical potential differences of oxygen- and steam-concentration cells, tH,av is the average proton transference number. Such a decrease negatively affects the performance of the cells operated in fuel cell mode because, in order to obtain high power densities, high OCVs are necessary [59]:
P max = E 2 4 R t o t a l .
Conversely, the lowest possible OCVs are needed for cells operated in electrolysis cell mode if these OCVs are not associated with significant electron transport of the electrolytes or imperfect system gas-tightness. In detail, a current density will be higher at a certain voltage value (U), while a difference of U—E will be higher (or E will be lower).
In order to check the pH2O effect, both atmospheres were consequentially humidified: first, air atmosphere and then hydrogen atmosphere. Moreover, this allows the response of each electrode to be revealed and even their contributions to the total polarization resistance, Rp, to be estimated.
Figure 8 shows the main electrochemical characteristics of the PCC obtained under isothermal conditions with gradual increase of pH2O. As can be seen from these data, the OCVs drop from 0.979 to 0.899 V, the maximal power density decreases from 430 to 290 mW cm−2, while the maximal achievable hydrogen flux density increases from 4.4 to 5.1 ml min−1 cm−2, respectively, when p’’H2O increases from 0.03 to 0.5 atm. This is in complete agreement with the abovementioned thermodynamic predictions.
EIS and DRT analyses were further utilized to reveal the main tendencies in Ro and Rp changes and their effects on the PCC’s performance.
Air humidification virtually does not change the spectra’s shape (Figure A2); they, as well as the original spectra, can accurately be described by an equivalent circuit scheme with two RQ-elements. On the base of DRT data (Figure A3), the distribution function consists of three well-separated peaks at high p’’H2O values, two of which merge at lower p’’H2O values. The total polarization resistance of the electrodes amounts 0.07, 0.08, 0.10 and 0.12 Ω cm2 at 0.03, 0.10, 0.30 and 0.50 atm of p’’H2O. At the same time, the partial resistance components vary differently (Figure A4):
(1)
The absolute value of R1 is equal to 0.02 Ω cm2, but its contribution as part of Rp decreases from 30 to 17%;
(2)
The contribution of R1’ in Rp does not exceed 4.5% or 0.004 Ω cm2 in absolute units;
(3)
The contribution of R2 increases from 70 to 80%, remaining the dominant parameter in the electrode performance.
All the mentioned components are sensitive towards air humidification. Therefore, these stages are primarily associated with oxygen electrode behavior.
Returning to the data of Figure 8, it can be stated that performance of the PCC is regulated by the electrode activity, which dominates under OCV and fuel cell modes of operation. When the bias exceeds the OCV level, the electrode resistance drops rapidly [60]; in this case, an improvement in PCC’s performance is related with a lower RO as a result of achieving excellent proton conductivity in highly humid conditions. Making a preliminary conclusion, it can be noted that the Pr2NiO4+δ-based electrodes operate as a classical mixed oxygen-ionic/electronic conductor with no evidence of proton transportation revealed in previously published works [51,61]. However, this might be explained by the relatively high measured temperature (750 °C) leading to the insignificant water uptake capability of nickelates.
Finally, the PCC was tested depending on temperature under 50%H2O/H2—50%H2O/air conditions, corresponding to both highly-moisturized gases (Figure 9). As listed in Table A1, further hydrogen humidification from 0.03 to 0.5 atm has little effect on both maximal power density (decreases by ~3%) and the hydrogen evolution rate (increases by ~6%). Therefore, differences in output parameters obtained at p’H2O = p’’H2O = 0.03 atm (condition 1) and p’H2O = p’’H2O = 0.5 atm (condition 2) are mainly attributed with the pH2O variation in air atmosphere. Such differences amount to −37% of Pmax and +13% of jH2 at the same comparison temperature as shown in Figure 9d. The obtained data are also in agreement with thermodynamic predictions, in particular with OCVs, which reach 0.991, 0.951, 0.912 and 0.874 V at 600, 650, 700 and 750 °C, respectively.
With increasing pH2O in hydrogen atmosphere, the impedance spectra cannot be described by the used equivalent circuit schemes, implying the appearance of additional rate-determining steps. These steps might be attributed either to fuel electrode processes or even to those taking place at the oxygen electrode. The latter is realized due to the fact that a change in a potential-determined parameter from the one side of an electrolyte membrane results in a redistribution of the overall potential and internal (ionic and electronic) currents, which can in turn affect the electrode process at the other side of the same membrane. As indicated in Figure A5, hydrogen humidification (when the oxidizing composition is unchanged) leads to the formation of a new distribution function, consisting in different number of peaks, as well as their intensity and displacement. Such a distribution function becomes more complicated with decreasing temperature (Figure 10): here, at least five independent peaks and corresponding processes can be distinguished.
Due to the already mentioned peculiarities, the RO values were determined as a high-frequency intercept of spectra with the x-axis, while the Rp values were separately estimated as the total area bounded by a γ(τ) function. For 50%H2O/H2—50%H2O/air and OCV conditions, RO is equal to 0.51, 0.48, 0.45 and 0.42 Ω cm2, whereas Rp is equal to 0.99, 0.46, 0.21 and 0.14 Ω cm2 at 600, 650, 700 and 750 °C, respectively. Comparing 50%H2O/H2—50%H2O/air and 3%H2O/H2—50%H2O/air conditions at 750 °C, the RO parameter is virtually unchanged, showing the possibility of full hydration of the proton-conducting electrolyte. At the same time, Rp increases from 0.12 to 0.14 Ω cm2. Such an increment of Rp reaches 0.99 Ω cm2 at 600 °C and 0.21 Ω cm2 at 700 °C for 50%H2O/H2—50%H2O/air conditions as against 0.39 Ω cm2 at 600 °C and 0.12 Ω cm2 at 700 °C for 3%H2O/H2—3%H2O/air conditions. Therefore, not only air humidification, but also hydrogen moisturization leads to higher Rp values and correspondingly lower performance characteristics. It should be noted that neither RO nor Rp indicates the efficiency of the PCC, which can, however, be estimated on the basis of the electrolytic domain boundaries (absolute level and contribution of ionic conductivity).

3.6. Electrolytic Properties of the BCZD Membrane

In order to estimate the electrolytic properties of BCZD, which determines rPCCs’ efficiency [62,63,64], the ti,av values were calculated (Equation (1)) using data on RO and Rp. According to Figure 11, the ionic transference numbers decrease naturally with increasing temperature due to an increase in the hole conductivity contribution. Nevertheless, the ti,av parameter achieves quite a high level (0.90 at 750 °C) and, at the same time, can be further increased (up to 0.95 at the same temperature) via humidification of both gases. As a result of high saturation, oxygen vacancies are almost fully filled with steam, leading to a higher concentration of proton charge carriers (Equation (6)) and inhibition of hole charge carrier formation due to the lower concentration of free oxygen vacancies (Equation (7)).
The average ionic conductivity of the electrolyte membranes, determined as
σ i , a v = h R O t i , a v ,
should be considered as the most appropriate parameter (instead of total conductivity) related simultaneously with the performance (h/RO) as well as the efficiency (ti,av) of the electrochemical devices. For the present case, the average ionic conductivity of BCZD reaches 4.7 and 4.9 mS cm−1 at 600 and 700 °C under condition 1 and 4.9 and 5.4 mS cm−1, respectively, under condition 2. These belong to a range of the highest values reached for proton-conducting electrolytes (Table 2, [37,52,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79]), which agrees with the independently measured comparison of ionic conductivities of Y- and Dy-doped Ba(Ce,Zr)O3 [80]. The abovementioned results allow the following important conclusions to be formulated:
(1)
The BCZD electrolyte forms the basis for the design of novel electrochemical cells with improved output parameters due to its higher ionic conductivity compared with those for the most-studied Y-containing cerate-zirconates;
(2)
Despite the negative electrochemical response of the electrodes to gas humidification, the average ionic transference and ionic conductivity values take the opposite direction, resulting in improved PCC efficiency.

4. Conclusions

This work presents the results of fabrication and characterization of a reversible protonic ceramic cell having symmetrically-organized electrodes made of Pr1.9Ba0.1NiO4+δ (PBN). Utilization of the same material as the functional fuel and oxygen electrode allows minimization of the thermo-chemical stress between the functional materials during high-temperature steps and even a reduction of these steps to one sintering stage, promising significant techno-economic benefits. The fabricated cell based on a 25 µm-thick BaCe0.5Zr0.3Dy0.2O3−δ (BCZD) proton-conducting electrolyte demonstrated output characteristics as high as ~300 mW cm−2 at 600 °C in fuel cell mode and ~300 mA cm−2 in electrolysis cell mode at thermoneutral conditions. The cell was tested under conditions of varying humidity in order to evaluate electrode and electrolyte performance. It was found that the PBN–BCZD oxygen electrode determined the overall electrode performance at 750 °C, operating as a dual conducting (O2−/h) system due to the negative electrochemical response to gas humidification. As a consequence of the high contribution of polarization resistance to the total resistance, the response of the rPCC’s performance was the same under open circuit voltage and fuel cell modes. On the other hand, the ohmic resistance and the average ionic transference numbers of the electrolyte membrane increased with increasing humidification, demonstrating its proton-conducting nature. Along with increased efficiency (as a result of improved electrolytic properties), the performance of the PCC was higher under humidification due to the total cell resistance being determined by the ohmic component. Although only medium performance was reached in this work (due to the rather thick electrolyte, 25 µm), the proposed strategies can effectively be used in future to resolve a number of technological issues.

Author Contributions

Conceptualization, D.M.; Methodology, J.L. and G.V.; Validation, A.D.; Formal analysis, A.F. and S.P.; Investigation, A.T., J.L. and G.V.; Resources, A.T. and J.L.; Writing—original draft preparation, D.M.; Writing—review & editing, D.M.; Visualization, A.T.; Supervision, A.D.; Project administration, D.M.; Funding acquisition, D.M.

Funding

The majority of this work was carried out under the budgetary plans of Institute of High Temperature Electrochemistry. The design of new electrode materials and their characterization was also funded by the Russian Foundation for Basic Research, grant number 18-38-20063. Dr. Dmitry Medvedev is also grateful to the Council of the President of the Russian Federation (scholarship СП-161.2018.1) for supporting the studies devoted to search of new Co-free electrode materials.

Acknowledgments

The characterization of powder and ceramic materials was carried out at the Shared Access Centre “Composition of Compounds” of the Institute of High Temperature Electrochemistry [26].

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Figure A1. Temperature dependences of constitute resistances (Rj) of the fabricated PCC in Arrhenius coordinates. The corresponding activation energies calculated in Frenkel coordinates (log(T/Rj) = ƒ(1/T)) are also presented.
Figure A1. Temperature dependences of constitute resistances (Rj) of the fabricated PCC in Arrhenius coordinates. The corresponding activation energies calculated in Frenkel coordinates (log(T/Rj) = ƒ(1/T)) are also presented.
Materials 12 00118 g0a1
Figure A2. Impedance spectra of the PCC at 750 °C and OCV conditions depending on different pH2O in air atmosphere (fuel gas is 3%H2O/H2): original spectra (a) and those obtained after subtracting the ohmic resistance (b).
Figure A2. Impedance spectra of the PCC at 750 °C and OCV conditions depending on different pH2O in air atmosphere (fuel gas is 3%H2O/H2): original spectra (a) and those obtained after subtracting the ohmic resistance (b).
Materials 12 00118 g0a2
Figure A3. DRT results for the impedance spectra presented in Figure A2.
Figure A3. DRT results for the impedance spectra presented in Figure A2.
Materials 12 00118 g0a3
Table A1. Effect of hydrogen humidification on the PCC performance at 700 °C, when p’’H2O = 0.50 atm. RO and Rp are presented for OCV mode, while jH2 is presented at U = 1.3 V.
Table A1. Effect of hydrogen humidification on the PCC performance at 700 °C, when p’’H2O = 0.50 atm. RO and Rp are presented for OCV mode, while jH2 is presented at U = 1.3 V.
p’H2O, atmRO, Ω cm2Rp, Ω cm2Pmax, mW cm−2jH2, ml min−1 cm−2
0.030.470.122935.1
0.500.450.212845.6
Figure A4. Contribution of partial resistances in the total resistance of the electrodes (Rp). These data were obtained from the DRT results (Figure A3).
Figure A4. Contribution of partial resistances in the total resistance of the electrodes (Rp). These data were obtained from the DRT results (Figure A3).
Materials 12 00118 g0a4
Figure A5. DRT results for the impedance spectra measured for the PCC at 750 °C under OCV conditions at different pH2O values in hydrogen atmosphere (p’’H2O in air is equal 0.5 atm).
Figure A5. DRT results for the impedance spectra measured for the PCC at 750 °C under OCV conditions at different pH2O values in hydrogen atmosphere (p’’H2O in air is equal 0.5 atm).
Materials 12 00118 g0a5

References

  1. Medvedev, D.; Murashkina, A.; Pikalova, E.; Demin, A.; Podias, A.; Tsiakaras, P. BaCeO3: Materials development, properties and application. Prog. Mater. Sci. 2014, 60, 72–129. [Google Scholar] [CrossRef]
  2. Bi, L.; Boulfrada, S.; Traversa, E. Steam electrolysis by solid oxide electrolysis cells (SOECs) with proton-conducting oxides. Chem. Soc. Rev. 2014, 43, 8255–8270. [Google Scholar] [CrossRef] [PubMed]
  3. Choi, S.M.; An, H.; Yoon, K.J.; Kim, B.-K.; Lee, H.-W.; Son, J.-W.; Kim, H.; Shin, D.; Ji, H.-L. Electrochemical analysis of high-performance protonic ceramic fuel cells based on a columnar-structured thin electrolyte. Appl. Energy 2019, 233–234, 29–36. [Google Scholar] [CrossRef]
  4. Mohato, N.; Banerjee, A.; Gupta, A.; Omar, S.; Balani, K. Progress in material selection for solid oxide fuel cell technology: A review. Prog. Mater. Sci. 2015, 72, 141–337. [Google Scholar] [CrossRef]
  5. Gómez, S.Y.; Hotza, D. Current developments in reversible solid oxide fuel cells. Renew. Sust. Energy Rev. 2016, 61, 155–174. [Google Scholar] [CrossRef]
  6. Choi, S.; Kucharczyk, C.J.; Liang, Y.; Zhang, X.; Takeuchi, I.; Ji, H.-I.; Haile, S.M. Exceptional power density and stability at intermediate temperatures in protonic ceramic fuel cells. Nat. Energy 2018, 3, 202–210. [Google Scholar] [CrossRef] [Green Version]
  7. Bae, K.; Jang, D.Y.; Choi, H.J.; Kim, D.; Hong, J.; Kim, B.-K.; Lee, J.-H.; Son, J.-W.; Shim, J.H. Demonstrating the potential of yttrium-doped barium zirconate electrolyte for high-performance fuel cells. Nat. Commun. 2017, 8, 14553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Duan, C.; Tong, J.; Shang, M.; Nikodemski, S.; Sanders, M.; Ricote, S.; Almansoori, A.; O’Hayre, R. Readily processed protonic ceramic fuel cells with high performance at low temperatures. Science 2015, 349, 1321–1326. [Google Scholar] [CrossRef]
  9. Shim, J.H. Ceramics breakthrough. Nat. Energy 2018, 3, 168–169. [Google Scholar] [CrossRef]
  10. Ding, D.; Li, X.; Lai, S.Y.; Gerdes, K.; Liu, M. Enhancing SOFC cathode performance by surface modification through infiltration. Energy Environ. Sci. 2014, 7, 552–575. [Google Scholar] [CrossRef]
  11. Zheng, Y.; Wang, J.; Yu, B.; Zhang, W.; Chen, J.; Qiao, J.; Zhang, J. A review of high temperature co-electrolysis of H2O and CO2 to produce sustainable fuels using solid oxide electrolysis cells (SOECs): Advanced materials and technology. Chem. Soc. Rev. 2017, 46, 1427–1463. [Google Scholar] [CrossRef] [PubMed]
  12. Su, C.; Wang, W.; Liu, M.; Tadé, M.O.; Shao, Z. Progress and prospects in symmetrical solid oxide fuel cells with two identical electrodes. Adv. Energy Mater. 2015, 5, 1500188. [Google Scholar] [CrossRef]
  13. Choi, S.; Sengodan, S.; Park, S.; Ju, Y.-W.; Kim, J.; Hyodo, J.; Jeong, H.Y.; Ishihara, T.; Shin, J.; Kim, G. A robust symmetrical electrode with layered perovskite structure for direct hydrocarbon solid oxide fuel cells: PrBa0.8Ca0.2Mn2O5+δ. J. Mater. Chem. A 2016, 4, 1747–1753. [Google Scholar] [CrossRef]
  14. Shen, J.; Chen, Y.; Yang, G.; Zhou, W.; Tadé, M.O.; Shao, Z. Impregnated LaCo0.3Fe0.67Pd0.03O3−δ as a promising electrocatalyst for “symmetrical” intermediate-temperature solid oxide fuel cells. J. Power Sources 2016, 306, 92–99. [Google Scholar] [CrossRef]
  15. Tao, H.; Xie, J.; Wu, Y.; Wang, S. Evaluation of PrNi0.4Fe0.6O3−δ as a symmetrical SOFC electrode material. Int. J. Hydrogen Energy 2018, 43, 15423–15432. [Google Scholar] [CrossRef]
  16. Niu, B.; Jin, F.; Zhang, L.; Shen, P.; He, T. Performance of double perovskite symmetrical electrode materials Sr2TiFe1−xMoxO6−δ (x = 0.1, 0.2) for solid oxide fuel cells. Electrochim. Acta 2018, 263, 217–227. [Google Scholar] [CrossRef]
  17. Lan, R.; Cowin, P.I.; Sengodan, S.; Tao, S. A perovskite oxide with high conductivities in both air and reducing atmosphere for use as electrode for solid oxide fuel cells. Sci. Rep. 2016, 6. [Google Scholar] [CrossRef]
  18. Gao, Z.; Ding, X.; Ding, D.; Ding, L.; Zhang, S.; Yuan, G. Infiltrated Pr2NiO4 as promising bi-electrode for symmetrical solid oxide fuel cells. Int. J. Hydrogen Energy 2018, 43, 8953–8961. [Google Scholar] [CrossRef]
  19. Nicollet, C.; Flura, A.; Vibhu, V.; Rougier, A.; Bassat, J.-M.; Grenier, J.-C. An innovative efficient oxygen electrode for SOFC: Pr6O11 infiltrated into Gd-doped ceria backbone. Int. J. Hydrogen Energy 2016, 41, 15538–15544. [Google Scholar] [CrossRef]
  20. Lyagaeva, J.; Medvedev, D.; Pikalova, E.; Plaksin, S.; Brouzgou, A.; Demin, A.; Tsiakaras, P. A detailed analysis of thermal and chemical compatibility of cathode materials suitable for BaCe0.8Y0.2O3−δ and BaZr0.8Y0.2O3−δ proton electrolytes for solid oxide fuel cell application. Int. J. Hydrogen Energy 2017, 42, 1715–1723. [Google Scholar] [CrossRef]
  21. Pikalova, E.; Kolchugin, A.; Bogdanovich, N.; Medvedev, D.; Lyagaeva, J.; Vedmid’, L.; Anayev, M.; Plaksin, S.; Farlenkov, A. Suitability of Pr2−xCaxNiO4+δ as cathode materials for electrochemical devices based on oxygen ion and proton conducting solid state electrolytes. Int. J. Hydrogen Energy 2018. [Google Scholar] [CrossRef]
  22. Sun, W.; Wang, Y.; Fang, S.; Zhu, Z.; Yan, L.; Liu, W. Evaluation of BaZr0.1Ce0.7Y0.2O3−δ-based proton-conducting solid oxide fuel cells fabricated by a one-step co-firing process. Electrochim. Acta 2011, 56, 1447–1454. [Google Scholar] [CrossRef]
  23. Dai, H.; Da’as, E.H.; Shafi, S.P.; Wang, H.; Bi, L. Tailoring cathode composite boosts the performance of proton-conducting SOFCs fabricated by a one-step co-firing method. J. Eur. Ceram. Soc. 2018, 38, 2903–2908. [Google Scholar] [CrossRef]
  24. Lyagaeva, J.; Danilov, N.; Vdovin, G.; Bu, J.; Medvedev, D.; Demin, A.; Tsiakaras, P. A new Dy-doped BaCeO3–BaZrO3 proton-conducting material as a promising electrolyte for reversible solid oxide fuel cells. J. Mater. Chem. A 2016, 4, 15390–15399. [Google Scholar] [CrossRef]
  25. Lyagaeva, Y.G.; Danilov, N.A.; Gorshkov, M.Y.; Vdovin, G.K.; Antonov, B.D.; Demin, A.K.; Medvedev, D.A. Functionality of lanthanum, neodymium, and praseodymium nickelates as promising electrode systems for proton-conducting electrolytes. Russ. J. Appl. Chem 2018, 91, 583–590. [Google Scholar] [CrossRef]
  26. Equipment. Shared Access Center. Available online: http://www.ihte.uran.ru/?page_id=3154 (accessed on 11 December 2018).
  27. DRTTOOLS. Available online: https://sites.google.com/site/drttools/ (accessed on 11 December 2018).
  28. Pikalova, E.; Kolchugin, A.; Filonova, E.; Bogdanovich, N.; Pikalov, S.; Ananyev, M.; Molchanova, N.; Farlenkov, A. Validation of calcium-doped neodymium nickelates as SOFC air electrode materials. Solid State Ionics 2018, 319, 130–140. [Google Scholar] [CrossRef]
  29. Akbari-Fakhrabadi, A.; Toledo, E.G.; Canales, J.I.; Meruane, V.; Chan, S.H.; Gracia-Pinilla, M.A. Effect of Sr2+ and Ba2+ doping on structural stability and mechanical properties of La2NiO4+δ. Ceram. Int. 2018, 44, 10551–10557. [Google Scholar] [CrossRef]
  30. Kravchenko, E.; Khalyavin, D.; Zakharchuk, K.; Grins, J.; Svensson, G.; Pankov, V.; Yaremchenko, A. High-temperature characterization of oxygen-deficient K2NiF4-type Nd2−xSrxNiO4−δ (x = 1.0–1.6) for potential SOFC/SOEC applications. J. Mater. Chem. A 2015, 3, 23852–23863. [Google Scholar] [CrossRef]
  31. Hua, B.; Li, M.; Sun, Y.-F.; Li, J.-H.; Luo, J.-L. Enhancing perovskite electrocatalysis of solid oxide cells through controlled exsolution of nanoparticles. ChemSusChem 2017, 10, 3333–3341. [Google Scholar] [CrossRef]
  32. Pikalova, E.Y.; Medvedev, D.A.; Khasanov, A.F. Structure, stability, and thermomechanical properties of Ca-substituted Pr2NiO4+δ. Phys. Solid State 2017, 59, 694–702. [Google Scholar] [CrossRef]
  33. Ceretti, M.; Wahyudi, O.; André, G.; Meven, M.; Villesuzanne, A.; Paulus, W. (Nd/Pr)2NiO4+δ: Reaction intermediates and redox behavior explored by in situ neutron powder diffraction during electrochemical oxygen intercalation. Chem. Mater. 2018, 57, 4657–4666. [Google Scholar] [CrossRef] [PubMed]
  34. Løken, A.; Ricote, S.; Wachowski, S. Thermal and chemical expansion in proton ceramic electrolytes and compatible electrodes. Crystals 2018, 8, 365. [Google Scholar] [CrossRef]
  35. Kharton, V.V.; Kovalevsky, A.V.; Avdeev, M.; Tsipis, E.V.; Patrakeev, M.V.; Yaremchenko, A.A.; Naumovich, E.N.; Frade, J.R. Chemically induced expansion of La2NiO4+δ-based materials. Chem. Mater. 2007, 19, 2027–2033. [Google Scholar] [CrossRef]
  36. Kochetova, N.; Animitsa, I.; Medvedev, D.; Demin, A.; Tsiakaras, P. Recent activity in the development of proton-conducting oxides for high-temperature applications. RSC Adv. 2016, 6, 73222–73268. [Google Scholar] [CrossRef]
  37. Jeong, S.; Kobayashi, T.; Kuroda, K.; Kwon, H.; Zhu, C.; Habazaki, H.; Aoki, Y. Evaluation of thin film fuel cells with Zr-rich BaZrxCe0.8−xY0.2O3−δ electrolytes (x ≥ 0.4) fabricated by a single-step reactive sintering method. RSC Adv. 2018, 8, 26309–26317. [Google Scholar] [CrossRef]
  38. Danilov, N.A.; Lyagaeva, J.G.; Medvedev, D.A.; Demin, A.K.; Tsiakaras, P. Transport properties of highly dense proton-conducting BaCe0.8−xZrxDy0.2O3−δ materials in low- and high-temperature ranges. Electrochim. Acta 2018, 284, 551–559. [Google Scholar] [CrossRef]
  39. Liu, M.; Hu, H. Effect of interfacial resistance on determination of transport properties of mixed-conducting electrolytes. J. Electrochem. Soc. 1996, 143, L109–L112. [Google Scholar] [CrossRef]
  40. Kreuer, K.D. Proton-conducting oxides. Annu. Rev. Mater. Res. 2003, 33, 333–359. [Google Scholar] [CrossRef]
  41. Malavasi, L.; Fisher, C.A.J.; Islam, M.S. Oxide-ion and proton conducting electrolyte materials for clean energy applications: Structural and mechanistic features. Chem. Soc. Rev. 2010, 39, 4370–4387. [Google Scholar] [CrossRef]
  42. Gregori, G.; Merkle, R.; Maier, J. Ion conduction and redistribution at grain boundaries in oxide systems. Prog. Mater. Sci. 2017, 89, 252–305. [Google Scholar] [CrossRef]
  43. Nechache, A.; Cassir, M.; Ringuedé, A. Solid oxide electrolysis cell analysis by means of electrochemical impedance spectroscopy: A review. J. Power Sources 2014, 258, 164–181. [Google Scholar] [CrossRef]
  44. Zheng, M.; Wang, S.; Yang, Y.; Xia, C. Barium carbonate as a synergistic catalyst for the H2O/CO2 reduction reaction at Ni–yttria stabilized zirconia cathodes for solid oxide electrolysis cells. J. Mater. Chem. A 2018, 6, 2721–2729. [Google Scholar] [CrossRef]
  45. Sun, S.; Cheng, Z. Electrochemical behaviors for Ag, LSCF and BSCF as oxygen electrodes for proton conducting IT-SOFC. J. Electrochem. Soc. 2017, 164, F3104–F3113. [Google Scholar] [CrossRef]
  46. Shi, N.; Su, F.; Huan, D.; Xie, Y.; Lin, J.; Tan, W.; Peng, R.; Xia, C.; Chen, C.; Lu, Y. Performance and DRT analysis of P-SOFCs fabricated using new phase inversion combined tape casting technology. J. Mater. Chem. A 2017, 5, 19664–19671. [Google Scholar] [CrossRef]
  47. Wang, X.; Ma, Z.; Zhang, T.; Kang, J.; Ou, X.; Feng, P.; Wang, S.; Zhou, F.; Ling, Y. Charge transfer modeling and polarization DRT analysis of proton ceramics fuel cells based on mixed conductive electrolyte with the modified anode-electrolyte interface. ACS Appl. Mater. Interfaces 2018, 10, 35047–35059. [Google Scholar] [CrossRef] [PubMed]
  48. Boukamp, B.A. Fourier transform distribution function of relaxation times; application and limitations. Electrochim. Acta 2015, 154, 35–46. [Google Scholar] [CrossRef]
  49. Ivers-Tiffée, E.; Weber, A. Evaluation of electrochemical impedance spectra by the distribution of relaxation times. J. Ceram. Soc. Jpn. 2017, 125, 193–201. [Google Scholar] [CrossRef] [Green Version]
  50. Yang, S.; Wen, Y.; Zhang, J.; Lu, Y.; Ye, X.; Wen, Z. Electrochemical performance and stability of cobalt-free Ln1.2Sr0.8NiO4 (Ln=La and Pr) air electrodes for proton-conducting reversible solid oxide cells. Electrochim. Acta 2018, 267, 269–277. [Google Scholar] [CrossRef]
  51. Li, W.; Guan, B.; Ma, L.; Hu, S.; Zhang, N.; Liu, X. High performing triple-conductive Pr2NiO4+δ anode for proton-conducting steam solid oxide electrolysis cell. J. Mater. Chem. A 2018, 6, 18057–18066. [Google Scholar] [CrossRef]
  52. Li, G.; Jin, H.; Cui, Y.; Gui, L.; He, B.; Zhao, L. Application of a novel (Pr0.9La0.1)2(Ni0.74Cu0.21Nb0.05)O4+δ-infiltrated BaZr0.1Ce0.7Y0.2O3-δ cathode for high performance protonic ceramic fuel cells. J. Power Sources 2017, 341, 192–198. [Google Scholar] [CrossRef]
  53. Dailly, J.; Marrony, M.; Taillades, G.; Taillades-Jacquin, M.; Grimaud, A.; Mauvy, F.; Louradour, E.; Salmi, J. Evaluation of proton conducting BCY10-based anode supported cells by co-pressing method: Up-scaling, performances and durability. J. Power Sources 2014, 255, 302–307. [Google Scholar] [CrossRef]
  54. Wang, W.; Medvedev, D.; Shao, Z. Gas Humidification impact on the properties and performance of perovskite-type functional materials in proton-conducting solid oxide cells. Adv. Funct. Mater. 2018, 28, 1802592. [Google Scholar] [CrossRef]
  55. Heras-Juaristi, G.; Pérez-Coll, D.; Mather, G.C. Temperature dependence of partial conductivities of the BaZr0.7Ce0.2Y0.1O3−δ proton conductor. J. Power Sources 2017, 364, 52–60. [Google Scholar] [CrossRef]
  56. Lyagaeva, J.; Danilov, N.; Korona, D.; Farlenkov, A.; Medvedev, D.; Demin, A.; Animitsa, I.; Tsiakaras, P. Improved ceramic and electrical properties of CaZrO3-based proton-conducting materials prepared by a new convenient combustion synthesis method. Ceram. Int. 2017, 43, 7184–7192. [Google Scholar] [CrossRef]
  57. Norby, T. EMF method determination of conductivity contributions from protons and other foreign ions in oxides. Solid State Ionics 1988, 28–30, 1586–1591. [Google Scholar] [CrossRef]
  58. Sutija, D. Transport number determination by the concentration-cell/open-circuit voltage method for oxides with mixed electronic, ionic and protonic conductivity. Solid State Ionics 1995, 77, 167–174. [Google Scholar] [CrossRef]
  59. Gong, Z.; Sun, W.; Cao, J.; Shan, D.; Wu, Y.; Liu, W. Ce0.8Sm0.2O1.9 decorated with electron-blocking acceptor-doped BaCeO3 as electrolyte for low-temperature solid oxide fuel cells. Electrochim. Acta 2017, 228, 226–232. [Google Scholar] [CrossRef]
  60. Huan, D.; Shi, N.; Zhang, L.; Tan, W.; Xie, Y.; Wang, W.; Xia, C.; Lu, Y. New, Efficient, and reliable air electrode material for proton-conducting reversible solid oxide cells. ACS Appl. Mater. Interfaces 2018, 10, 1761–1770. [Google Scholar] [CrossRef]
  61. Philippeau, B.; Mauvy, F.; Mazataud, C.; Fourcade, S.; Grenier, J.-C. Comparative study of electrochemical properties of mixed conducting Ln2NiO4+δ (Ln = La, Pr and Nd) and La0.6Sr0.4Fe0.8Co0.2O3−δ as SOFC cathodes associated to Ce0.9Gd0.1O2−δ, La0.8Sr0.2Ga0.8Mg0.2O3−δ and La9Sr1Si6O26.5 electrolytes. Solid State Ionics 2013, 249–250, 17–25. [Google Scholar] [CrossRef]
  62. Nakamura, T.; Mizunuma, S.; Kimura, Y.; Mikami, Y.; Yamauchi, K.; Kuroha, T.; Taniguchi, N.; Tsuji, Y.; Okuyama, Y.; Amezawa, K. Energy efficiency of ionic transport through proton conducting ceramic electrolytes for energy conversion applications. J. Mater. Chem. A 2018, 6, 15771–15780. [Google Scholar] [CrossRef]
  63. Ji, H.-I.; Kim, H.; Lee, H.-W.; Kim, B.-K.; Son, J.-W.; Yoon, K.J.; Lee, J.-H. Open-cell voltage and electrical conductivity of a protonic ceramic electrolyte under two chemical potential gradients. Phys. Chem. Chem. Phys. 2018, 20, 14997–15001. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, J.-H.; Lei, L.-B.; Liu, D.; Zhao, F.-Y.; Ni, M.; Chen, F. Mathematical modeling of a proton-conducting solid oxide fuel cell with current leakage. J. Power Sources 2018, 400, 333–340. [Google Scholar] [CrossRef]
  65. Wachowski, S.; Li, Z.; Polfus, J.M.; Norby, T. Performance and stability in H2S of SrFe0.75Mo0.25O3-δ as electrode in proton ceramic fuel cells. J. Eur. Ceram. Soc. 2018, 38, 163–171. [Google Scholar] [CrossRef]
  66. Nasani, N.; Ramasamy, D.; Mikhalev, S.; Kovalevsky, A.V.; Fagg, D.P. Fabrication and electrochemical performance of a stable, anode supported thin BaCe0.4Zr0.4Y0.2O3−δ electrolyte protonic ceramic fuel cell. J. Power Sources 2015, 278, 582–589. [Google Scholar] [CrossRef]
  67. Wan, Y.; He, B.; Wang, R.; Ling, Y.; Zhao, L. Effect of Co doping on sinterability and protonic conductivity of BaZr0.1Ce0.7Y0.1Yb0.1O3−δ for protonic ceramic fuel cells. J. Power Sources 2017, 347, 14–20. [Google Scholar] [CrossRef]
  68. Dailly, J.; Taillades, G.; Ancelin, M.; Pers, P.; Marrony, M. High performing BaCe0.8Zr0.1Y0.1O3−δ−Sm0.5Sr0.5CoO3−δ based protonic ceramic fuel cell. J. Power Sources 2017, 361, 221–226. [Google Scholar] [CrossRef]
  69. Ding, H.; Xue, X. GdBa0.5Sr0.5Co2O5+δ layered perovskite as promising cathode for proton conducting solid oxide fuel cells. J Alloys Compd. 2010, 496, 683–686. [Google Scholar] [CrossRef]
  70. Shi, H.; Ding, Z.; Ma, G. Electrochemical performance of cobalt-free Nd0.5Ba0.5Fe1−xNixO3−δ cathode materials for intermediate temperature solid oxide fuel cells. Fuel Cells 2016, 16, 258–262. [Google Scholar] [CrossRef]
  71. Dailly, J.; Ancelin, M.; Marrony, M. Long term testing of BCZY-based protonic ceramic fuel cell PCFC: Micro-generation profile and reversible production of hydrogen and electricity. Solid State Ionics 2017, 306, 69–75. [Google Scholar] [CrossRef]
  72. Lee, S.; Park, S.; Wee, S.; Baek, H.; Shin, D. One-dimensional structured La0.6Sr0.4Co0.2Fe0.8O3−δ-BaCe0.5Zr0.35Y0.15O3−δ composite cathode for protonic ceramic fuel cells. Solid State Ionics 2018, 320, 347–352. [Google Scholar] [CrossRef]
  73. Amiri, T.; Singh, K.; Sandhu, N.K.; Hanifi, A.R.; Etsell, T.H.; Luo, J.-L.; Thangadurai, V.; Sarkar, P. High performance tubular solid oxide fuel cell based on Ba0.5Sr0.5Ce0.6Zr0.2Gd0.1Y0.1O3-δ proton conducting electrolyte. J. Electrochem. Soc. 2018, 165, F764–F769. [Google Scholar] [CrossRef]
  74. Li, F.; Tao, Z.; Dai, H.; Xi, X.; Ding, H. A high-performing proton-conducting solid oxide fuel cell with layered perovskite cathode in intermediate temperatures. Int. J. Hydrogen Energy 2018, 43, 19757–19762. [Google Scholar] [CrossRef]
  75. Lyagaeva, J.; Danilov, N.; Tarutin, A.; Vdovin, G.; Medvedev, D.; Demin, A.; Tsiakaras, P. Designing a protonic ceramic fuel cell with novel electrochemically active oxygen electrodes based on doped Nd0.5Ba0.5FeO3−δ. Dalton Trans. 2018, 47, 5149–8157. [Google Scholar] [CrossRef] [PubMed]
  76. Lyagaeva, J.; Vdovin, G.; Hakimova, L.; Medvedev, D.; Demin, A.; Tsiakaras, P. BaCe0.5Zr0.3Y0.2−xYbxO3−δ proton-conducting electrolytes for intermediate-temperature solid oxide fuel cells. Electrochim. Acta 2017, 251, 554–561. [Google Scholar] [CrossRef]
  77. Gui, L.; Ling, Y.; Li, G.; Wang, Z.; Wan, Y.; Wang, R.; He, B.; Zhao, L. Enhanced sinterability and conductivity of BaZr0.3Ce0.5Y0.2O3−δ by addition of bismuth oxide for proton conducting solid oxide fuel cells. J. Power Sources 2016, 301, 369–375. [Google Scholar] [CrossRef]
  78. Zhang, L.; Li, C. BaFe0.6Co0.3Ce0.1O3−δ as cathode materials for proton-conducting solid oxide fuel cells. Ceram. Int. 2016, 42, 10511–10515. [Google Scholar] [CrossRef]
  79. Danilov, N.A.; Tarutin, A.P.; Lyagaeva, J.G.; Pikalova, E.Y.; Murashkina, A.A.; Medvedev, D.A.; Patrakeev, M.V.; Demin, A.K. Affinity of YBaCo4O7+δ-based layered cobaltites with protonic conductors of cerate-zirconate family. Ceram. Int. 2017, 43, 15418–15423. [Google Scholar] [CrossRef]
  80. Danilov, N.; Pikalova, E.; Lyagaeva, J.; Antonov, B.; Medvedev, D.; Demin, A.; Tsiakaras, P. Grain and grain boundary transport in BaCe0.5Zr0.3Ln0.2O3−δ (Ln-Y or lanthanide) electrolytes attractive for protonic ceramic fuel cells application. J. Power Sources 2017, 366, 161–168. [Google Scholar] [CrossRef]
Figure 1. Principal scheme of a symmetrically designed PCC and Pr2NiO4+δ reduction with the formation of a Ni-based cermet.
Figure 1. Principal scheme of a symmetrically designed PCC and Pr2NiO4+δ reduction with the formation of a Ni-based cermet.
Materials 12 00118 g001
Figure 2. (a) XRD pattern of the BCZD – PBN mixture calcined at 1350 °C for 5 h; (b) TG curve obtained under reduction of the PBN powder in 50%H2/N2 atmosphere; (c) XRD pattern of the reduced product of the PBN material (after TG analysis); (d) Conductivity of the PBN samples in oxidizing and reducing atmospheres.
Figure 2. (a) XRD pattern of the BCZD – PBN mixture calcined at 1350 °C for 5 h; (b) TG curve obtained under reduction of the PBN powder in 50%H2/N2 atmosphere; (c) XRD pattern of the reduced product of the PBN material (after TG analysis); (d) Conductivity of the PBN samples in oxidizing and reducing atmospheres.
Materials 12 00118 g002
Figure 3. Cross-section images of the fabricated rPCC at different magnification (a,b) and maps of the elements distribution (c).
Figure 3. Cross-section images of the fabricated rPCC at different magnification (a,b) and maps of the elements distribution (c).
Materials 12 00118 g003
Figure 4. Reversible operation of the PCC at different temperatures under 3%H2O/H2–3%H2O/air conditions: volt-ampere curves (a), power density characteristics (b), maximal achievable hydrogen flux density (c), maximal power density and hydrogen flux density at U = 1.3 V depending on temperature (d).
Figure 4. Reversible operation of the PCC at different temperatures under 3%H2O/H2–3%H2O/air conditions: volt-ampere curves (a), power density characteristics (b), maximal achievable hydrogen flux density (c), maximal power density and hydrogen flux density at U = 1.3 V depending on temperature (d).
Materials 12 00118 g004
Figure 5. Impedance spectra of the PCC at different temperatures under 3%H2O/H2—3%H2O/air and OCV conditions: original spectra (a) and ones obtained after subtracting the ohmic resistance (b).
Figure 5. Impedance spectra of the PCC at different temperatures under 3%H2O/H2—3%H2O/air and OCV conditions: original spectra (a) and ones obtained after subtracting the ohmic resistance (b).
Materials 12 00118 g005
Figure 6. (a) Constituent resistances (Rj) of the total resistance of the PCC depending on temperature; (b) Contributions of Process II in the total polarization resistance and the ohmic resistance in the total resistance of the PCC at different temperatures.
Figure 6. (a) Constituent resistances (Rj) of the total resistance of the PCC depending on temperature; (b) Contributions of Process II in the total polarization resistance and the ohmic resistance in the total resistance of the PCC at different temperatures.
Materials 12 00118 g006
Figure 7. DRT results for the obtained impedance spectra (see details in Figure 5a).
Figure 7. DRT results for the obtained impedance spectra (see details in Figure 5a).
Materials 12 00118 g007
Figure 8. Reversible operation of the PCC at 750 °C depending on different pH2O in wet air with the constant fuel gas composition (3%H2O/H2): volt-ampere curves (a), power density characteristics (b), maximal achievable hydrogen flux density (c), maximal power density and hydrogen flux density at U = 1.3 V depending on pH2O (d).
Figure 8. Reversible operation of the PCC at 750 °C depending on different pH2O in wet air with the constant fuel gas composition (3%H2O/H2): volt-ampere curves (a), power density characteristics (b), maximal achievable hydrogen flux density (c), maximal power density and hydrogen flux density at U = 1.3 V depending on pH2O (d).
Materials 12 00118 g008
Figure 9. Reversible operation of the PCC at different temperatures under 50%H2O/H2—50%H2O/air conditions: volt-ampere curves (a), power density characteristics (b), maximal achievable hydrogen flux density (c), maximal power density and hydrogen flux density at U = 1.3 V depending on temperature compared with those (dashed columns) obtained under 3%H2O/H2—3%H2O/air conditions (d).
Figure 9. Reversible operation of the PCC at different temperatures under 50%H2O/H2—50%H2O/air conditions: volt-ampere curves (a), power density characteristics (b), maximal achievable hydrogen flux density (c), maximal power density and hydrogen flux density at U = 1.3 V depending on temperature compared with those (dashed columns) obtained under 3%H2O/H2—3%H2O/air conditions (d).
Materials 12 00118 g009
Figure 10. DRT results for the impedance spectra measured for the PCC at different temperatures under 50%H2O/H2—50%H2O/air and OCV conditions: general view (a) and its parts at different magnifications (be).
Figure 10. DRT results for the impedance spectra measured for the PCC at different temperatures under 50%H2O/H2—50%H2O/air and OCV conditions: general view (a) and its parts at different magnifications (be).
Materials 12 00118 g010
Figure 11. Temperature dependences of the average ionic transference numbers of the BCZD electrolyte membrane in the current-free mode of the rPCC under condition 1 (p’H2O = p’’H2O = 0.03 atm) and condition 2 (p’H2O = p’’H2O = 0.5 atm).
Figure 11. Temperature dependences of the average ionic transference numbers of the BCZD electrolyte membrane in the current-free mode of the rPCC under condition 1 (p’H2O = p’’H2O = 0.03 atm) and condition 2 (p’H2O = p’’H2O = 0.5 atm).
Materials 12 00118 g011
Table 1. Polarization behavior of the Pr2NiO4+δ-based electrodes of PCCs under OCV mode of operation 1: T is the temperature, Rp is the total polarization of the electrodes.
Table 1. Polarization behavior of the Pr2NiO4+δ-based electrodes of PCCs under OCV mode of operation 1: T is the temperature, Rp is the total polarization of the electrodes.
ElectrolyteElectrodeT, °CRp, Ω cm2Ref.
BaCe0.5Zr0.3Dy0.2O3−δ
(BCZD)
Pr1.9Ba0.1NiO4+δ–BCZD6000.39This work
7000.12
BaCe0.7Zr0.1Y0.2O3−δ (BCZY1)Pr1.8Sr0.2NiO4+δ6002.17[50]
7000.33
BaCe0.6Zr0.2Y0.2O3−δ (BCZY2)Pr2NiO4+δ–BCZY26000.21[51]
7000.06
BCZY1(Pr0.9La0.1)2Ni0.74Cu0.21Nb0.05O4+δ
BCZY1 (infiltration)
6000.32[52]
7000.13
BaCe0.9Y0.1O3−δPr2NiO4+δ6000.80[53]
1 Polarization resistance of the Ni-cermets is assumed to be much lower than that of Pr2NiO4+δ-based electrodes.
Table 2. Electrolytic properties of proton-conducting membranes of PCCs under OCV conditions 1: h is the thickness, T is the temperature, E is the OCV, RO and Rp are the ohmic and polarization resistances, ti,av is the average ionic transference number, σav and σi,av are the average values of total and ionic conductivities of the electrolyte membranes.
Table 2. Electrolytic properties of proton-conducting membranes of PCCs under OCV conditions 1: h is the thickness, T is the temperature, E is the OCV, RO and Rp are the ohmic and polarization resistances, ti,av is the average ionic transference number, σav and σi,av are the average values of total and ionic conductivities of the electrolyte membranes.
Oxygen Electrode 2Electrolyte 3h, μmT, °CE, VRO, Ω cm2Rp, Ω cm2ti,avσav, mS cm−1σi,av, mS cm−1Ref.
PBNBCZD256001.0760.520.390.974.84.7This work
7001.0060.470.120.925.34.9
SFMBZCY812007000.9638.093.590.873.22.7[65]
SFMBZCY8206001.030.801.470.972.52.4[66]
7000.960.570.330.913.53.2
PLNCNBCZY1126000.990.330.320.943.63.4[52]
7000.950.210.130.915.85.2
PBCBCZYYC106001.000.370.340.942.72.5[67]
7000.990.260.120.923.83.5
SSC–BCZY1/BCZY1/96001.130.470.421.001.91.9[68]
7001.070.310.100.972.92.8
GBSCBCZY1206001.010.650.390.933.12.8[69]
7001.020.520.080.923.83.6
NBFNBCZY1406001.110.840.710.994.84.7[70]
LSCFBCZY4306001.070.720.240.964.23.9[37]
7001.010.580.070.915.24.7
BCZY6306001.060.580.130.955.24.9
7000.990.460.040.906.55.8
BCZY7306001.041.370.650.942.22.1
7000.960.970.210.883.12.7
BZY306000.931.340.820.892.21.9
7000.841.130.290.802.72.1
BSCFBCZY1/66001.040.370.850.981.61.6[71]
LSCF–BCZY3.5BCZY3.586001.050.311.250.992.62.6[72]
7001.020.220.300.963.63.5
LSCF–BSCZGYBSCZGY106001.150.413.461.002.42.4[73]
7001.130.191.821.005.45.4
PBFM–SSCBCZY1256001.010.410.540.956.15.8[74]
NBFCBCZD306001.050.680.660.964.44.2[75]
7001.010.420.240.947.16.7
NBFC’BCZYY256001.041.040.830.952.42.3[76]
7001.010.650.220.923.83.6
PBC–BCZYBCZY0.3176001.010.320.640.965.35.1[77]
BFCCBCZYY’306001.060.490.280.966.15.9[78]
YBCZBCZD206001.030.770.510.942.62.5[79]
7000.950.490.180.894.43.6
1 In the most cases for the listed PCCs, the gas compositions represent wet (3%H2O) H2 and static (or wet) air. 2 Abbreviations of oxygen electrodes: PBN = Pr1.9Ba0.1NiO4+δ, SFM = SrFe0.75Mo0.25O3−δ, PLNCN = (Pr0.9La0.1)2Ni0.74Cu0.21Nb0.05O4+δ, PBC = PrBaCo2O5+δ, SSC = Sm0.5Sr0.5CoO3−δ, GBSC = GdBa0.5Sr0.5Co2O5+δ, NBFN = Nd0.5Ba0.5Fe0.9Ni0.1O3−δ, LSCF = La0.6Sr0.4Co0.2Fe0.8O3−δ, BSCF = Ba0.5Sr0.5Co0.8Fe0.2O3−δ, PBFM = (PrBa)0.95(Fe0.9Mo0.1)2O5−δ, NBFC = Nd0.5Ba0.5Fe0.9Co0.1O3−δ, NBFC’ = Nd0.5Ba0.5Fe0.9Cu0.1O3−δ, PBC = PrBaCo2O5+δ, BFCC = BaFe0.6Co0.3Ce0.1O3−δ, YBCZ = YBaCo3.5Zn0.5O7+δ. 3 Abbreviations of electrolytes: BCZD = BaCe0.5Zr0.3Dy0.2O3−δ, BCZY1 = BaCe0.7Zr0.1Y0.2O3−δ, BCZY1/ = BaCe0.8Zr0.1Y0.1O3−δ, BCZY3 = BaCe0.5Zr0.3Y0.2O3−δ, BCZY3.5 = BaCe0.5Zr0.35Y0.15O3−δ, BCZY4 = BaZr0.4Ce0.4Y0.2O3−δ, BCZY6 = BaZr0.6Ce0.2Y0.2O3−δ, BCZY7 = BaZr0.7Ce0.1Y0.2O3−δ, BZCY8 = BaCe0.1Zr0.8Y0.1O3−δ, BZY = BaZr0.8Y0.2O3−δ, BSCZGY = Ba0.5Sr0.5Ce0.6Zr0.2Gd0.1Y0.1O3−δ, BCZYY = BaCe0.5Zr0.3Y0.1Yb0.1O3−δ, BCZYY’ = BaCe0.7Zr0.1Y0.1Yb0.1O3−δ, BCZYYC = BaCe0.68Zr0.1Y0.1Yb0.1Co0.02O3−δ.

Share and Cite

MDPI and ACS Style

Tarutin, A.; Lyagaeva, J.; Farlenkov, A.; Plaksin, S.; Vdovin, G.; Demin, A.; Medvedev, D. A Reversible Protonic Ceramic Cell with Symmetrically Designed Pr2NiO4+δ-Based Electrodes: Fabrication and Electrochemical Features. Materials 2019, 12, 118. https://doi.org/10.3390/ma12010118

AMA Style

Tarutin A, Lyagaeva J, Farlenkov A, Plaksin S, Vdovin G, Demin A, Medvedev D. A Reversible Protonic Ceramic Cell with Symmetrically Designed Pr2NiO4+δ-Based Electrodes: Fabrication and Electrochemical Features. Materials. 2019; 12(1):118. https://doi.org/10.3390/ma12010118

Chicago/Turabian Style

Tarutin, Artem, Julia Lyagaeva, Andrey Farlenkov, Sergey Plaksin, Gennady Vdovin, Anatoly Demin, and Dmitry Medvedev. 2019. "A Reversible Protonic Ceramic Cell with Symmetrically Designed Pr2NiO4+δ-Based Electrodes: Fabrication and Electrochemical Features" Materials 12, no. 1: 118. https://doi.org/10.3390/ma12010118

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop