Open access peer-reviewed chapter

Analysis of the Electrochemical Transport Properties of Doped Barium Cerate for Proton Conductivity in Low Humidity Conditions: A Review

Written By

Laura I.V. Holz, Vanessa C.D. Graça, Francisco J.A. Loureiro and Duncan P. Fagg

Submitted: 15 March 2020 Reviewed: 10 September 2020 Published: 04 November 2020

DOI: 10.5772/intechopen.93970

From the Edited Volume

Analytical Chemistry - Advancement, Perspectives and Applications

Edited by Abhay Nanda Srivastva

Chapter metrics overview

509 Chapter Downloads

View Full Metrics

Abstract

Proton-conducting perovskites are among the most promising electrolytes for Proton Ceramic Fuel Cells (PCFCs), electrolysers and separation membranes. Particularly, yttrium-doped barium cerate, BaCe1-xYxO3-δ (BCY), shows one of the highest protonic conductivities at intermediate temperatures (σ ∼ 10−3 S cm−1 at 400°C); values that are typically achieved under humidified atmospheres (pH2O ∼ 10−2 atm). However, BCY has commonly been discarded for such applications due to its instability in the presence of water vapour and carbonaceous atmospheres. A recent discovery has shown that BCY10 exhibits pure protonic conductivity under very low humidity contents (∼10−5–10−4 atm), owing to its very high equilibrium constant for hydration. This peculiar characteristic allows this material to retain its functionally as a proton conductor in such conditions, while preventing its decomposition. Hence, this chapter explores the electrochemical properties of the BaCe0.9Y0.1O3-δ (BCY10) composition, comprehensively establishing its limiting operation conditions through defect chemistry and thermodynamic analyses. Moreover, the importance of such conditions is highlighted with respect to potential industrially relevant hydrogenation/de-hydrogenation reactions at low temperatures under low humidity.

Keywords

  • perovskite
  • barium cerate
  • protonic conductivity
  • transport number
  • nominally dry conditions

1. Introduction

Ceramic proton conductors have been highlighted for electrochemical synthesis, as potential membranes in hydrogenation and dehydrogenation reactions [1]. One of the best compositions for this role is that of the doped barium cerate, e.g. BaCe1-xMxO3-δ (M = Y3+, In3+, Gd3+, etc.), which can show very high levels of proton conductivity at intermediate temperatures (i.e. σ ∼ 10−3 S cm −1 at 400°C) [2, 3, 4, 5, 6, 7, 8, 9]. This material belongs to the perovskite family with ABO3 ceramic oxide structure, including a divalent alkaline earth element, such as Ba2+ (also, Sr2+ or Ca2+), in the A-cation site, while a tetravalent rare-earth element, Ce4+, is present in the B-cation site. The introduction of dopants in the B-site with suitable acceptor elements, such as Y3+, In+3 or Gd3+ trivalent cations, leads to the formation of charge compensating oxygen vacancies [9]:

BaO+1yCeO2+y2M2O3BaBax+1yCeCex+yMCe+3y2OOx+y2VOE1

In addition to potential oxide-ion conductivity, these acceptor-substituted materials are also capable of offering both protonic and electronic conductivity, depending on the temperature and atmospheric conditions. The protonic conductivity is the most significant characteristic of these materials that is usually associated with the existence of protonic defects (OHO), upon filling of these oxygen vacancies in the presence of water vapour, as expressed by Eq. (2) [10, 11, 12, 13]:

H2O+VO+OOx2OHOE2

Accordingly, the equilibrium constant for hydration, Kw, is given by the following equation:

KwOHO2pH2OVOOOxE3

Due to the significant importance of humidity to promote protonic conductivity, most of the reported studies of barium cerate based materials have focused on highly wetted atmospheres with typical water vapour partial pressure pH2O ∼ 3 × 10−2 atm [14, 15, 16, 17, 18]. Unfortunately, these works also underline the tendency of this material for reacting with acidic gases, viz. carbon dioxide (CO2) and water vapour (H2O), leading to the formation of insulating carbonate or hydroxide phases, respectively, on the surface of the material. This complication impedes the ability of this material to be used in highly humidified and carbon-based fuels, thus, limiting its potential application range [3, 14, 15, 16, 17, 18, 19, 20]. The typical degradation reactions in such atmospheres include:

BaCeO3s+CO2gBaCO3s+CeO2sE4
BaCeO3s+H2OgBaOH2g+CeO2sE5

The chemical stability of doped barium cerates is well documented in the literature and huge efforts have been made to explore the reasons behind its chemical instability, using both conventional and non-conventional techniques [21, 22, 23, 24, 25]. For instance, Matsumoto et al. [22] studied the effect of dopant M in BaCe0.9M0.1O3-δ (M = Y, Tm, Yb, Lu, In, or Sc) on the electrical conductivity in the temperature range 400–900°C and on the chemical stability with respect to CO2 by thermogravimetry (TG). Both the electrical conductivity (moistened H2 or O2, pH2O = 1.9 × 10−2 atm) and the stability against carbonate formation were shown to decrease with increasing ionic radius (Figure 1), corresponding to an increase in basicity. Nonetheless, all compounds were found to interact with pure CO2 at temperatures below 900°C, failing to succeed in the mitigation of the chemical instability in the doped barium cerate.

Figure 1.

Carbonate formation temperature (blue) and the conductivity isotherm at 400°C of BaCe0.9M0.1O3-δ (M = Y, Tm, Yb, Lu, In or Sc) in moist H2 as a function of the ionic radius of the dopant. Adapted from [22].

Against this scenario, one common alternative is the use of the barium zirconates or compounds containing both Ce and Zr elements, where the introduction of Zr can significantly increase their chemical stability. Nonetheless, it has also been demonstrated that increased amounts of Zr negatively impact the total conductivity of these materials, due to an increase in their refractive nature and in their grain growth, which aggravate the problem of resistive grain boundaries. As such, much lower values of total conductivity are, typically, reported for the zirconate materials than for their cerate analogues, even though their bulk protonic conductivities are actually greater [9, 26, 27, 28, 29, 30, 31, 32, 33].

More recently, the work of Kim et al. [34] reported that the chemical instability of the barium cerates is due to the presence of a nanometre-thick amorphous phase found at the grain boundaries in proton-conducting BaCeO3 polycrystals, which not only leads to a reduced proton mobility, but also can act as a penetration path for H2O and CO2 gas molecules, facilitating chemical decomposition and collapse of the microstructure (Figure 2a). Furthermore, this effect could be minimised by controlling the composition to obtain Ba-deficient samples in which the intergranular amorphous layer could be minimised, leading to a mitigation of the reactivity with such gases (Figure 2b). The presence of an amorphous layer on the interfaces between grains has also been documented in barium zirconate-based compositions [26, 35], where this feature can exert significant complications during fabrication of complete electrochemical cells [19, 36].

Figure 2.

Schematic representation of microstructural changes upon reaction with water and carbon dioxide: (a) Ba-stoichiometric compositions (thick amorphous intergranular phase); (b) Ba-deficient compositions (thin amorphous intergranular phase).

In summary, the high electrical conductivity and the facile processing of the doped barium cerates demands further investigation to succeed to overcome their limited stabilities. In fact, it is only very recently that research in these materials has moved towards a more fundamental and, yet, critical aspect, concerning a deeper understanding of the limiting atmospheric conditions that are necessary to retain their functionality. Taking this into account, Loureiro et al. [37] reanalysed the barium cerate stability limits by thermodynamic calculations, considering its decomposition products in the presence of water vapour and CO2 (Figure 3). According to this theoretical study, no degradation would be expected for humidity values of ∼3 × 10−2 atm and temperatures higher than ∼500°C. However, when considering the formation of barium carbonate (Figure 3), the thermodynamics predict that much stricter conditions need to be applied, where only very low partial pressures of CO2 (e.g. pCO2 < ∼10−8 atm at 400°C) are able to avoid barium cerate degradation.

Figure 3.

Thermodynamic stability of carbon dioxide partial pressure (pCO2) and water vapour partial pressure (pH2O) as function of temperature considering the equilibrium of BaCeO3 and its decomposition products (i.e. BaCO3 and Ba(OH)2) [38] (reproduced by permission of The Royal Society of Chemistry).

For this reason, only very few reports can be found on successful applications of BCY membranes for chemical reactions. Most of these have concerned, ammonia synthesis [39, 40, 41], or the conversion of propane to propylene [42]. In these cases, no chemical instability has been reported and the survival of the BCY material is likely to be related to the effective absence of CO2 or significant water vapour in these operations. To understand this further, Figure 4 presents the maximum water vapour partial pressure (pH2O) that could be tolerated in different carbonaceous atmospheres to provide an equilibrium partial pressure of CO2 that remains below that of the BCY stability limit. These results demonstrate that, for example, at 400°C, these values should range between the values of 10−3 < pH2O < ∼ 10−4 atm in order to avoid decomposition of the perovskite phase, for the potential hydrocarbon atmospheres of CH4, C3H8 or C6H6 [43].

Figure 4.

Thermodynamic equilibrium for the formation of carbon dioxide from a hydrocarbon-based mixture and water at 400°C [38] (reproduced by permission of The Royal Society of Chemistry).

Nonetheless, one of the requirements for operating in such low water vapour partial pressures is that the protonic conductivity must be maintained in order to ensure the functionality of the electrolyte membrane in these applications. In this respect, protonic conductors are complex materials as they are capable to offer mixed conductivity (protonic, oxide-ion and electronic), depending on the temperature and on the nature of the surrounding atmosphere [37, 38]. One of the most promising compositions for this type of application is that of the yttrium-doped barium cerate, BaCe1-xYxO3-δ (BCY), which has very high protonic conductivity at lower temperatures under humidified atmospheres (e.g. ∼10−3 S cm−1 at 400°C, pH2O ∼ 10−2 atm) [1, 38].

Therefore, the current chapter will be focus on the electrochemical transport properties of the BaCe0.9Y0.1O3-d (BCY10) in reducing and oxidising conditions when operating in very low humidity levels. The aim of this chapter is to comprehensively explain the working limits of BCY10 and to assess its applicability as an electrolyte membrane for fuel cell, electrolysers and other electrochemical-based applications, with special focus on operation under low water vapour partial pressures.

Advertisement

2. Electrochemical properties of BCY10 in nominally dry reducing conditions

Figure 5 depicts the total conductivity of BCY10 analysed by impedance spectroscopy between 100 and 500°C in H2, 10%H2-N2 and N2, highlighting that no significant differences can be observed in the conductivity measured under these atmospheres. In addition, at the higher temperature range, a notable decrease of the activation energy is observed in all cases, as a result of the exsolution of protons from the structure of BCY10, and the concomitant decrease of the protonic contribution to the electrical transport [37]. Interestingly, and also surprisingly at first sight, current results of total conductivity in nominally dry H2 are close to those corresponding to the available data in literature for humidified H2 (Table 1).

Figure 5.

(a) Temperature dependency of the total conductivity of BCY10 obtained in the temperature range 100–500°C in nominally dry conditions for H2, 10% H2-N2 and N2; (b) BCY10 total conductivity as function of hydrogen partial pressure (pH2) under nominally dry conditions in the temperature range of 350–600°C [37] (reproduced by permission of The Royal Society of Chemistry).

Conductivity (S cm−1)pH2O (atm)Reference
3.59 × 10−3∼10−5 atm (dry H2)[37]
2.67 × 10−3∼10−2 atm (wet H2)[14]
1.85 × 10−3[17]
1.96 × 10−3[15]
2.60 × 10−3[16]
8.48 × 10−4[18]

Table 1.

Comparison of literature studies of total conductivity of BCY10 in nominally dry and wet H2 at 400°C [37] (reproduced by permission of The Royal Society of Chemistry).

To analyse the contribution of electronic conductivity to this material in nominally dry conditions, the total conductivity was also analysed as a function of hydrogen partial pressure (pH2) [37], as shown in Figure 5b. A slight increase in total conductivity can be observed towards higher pH2 values.

To be able to understand this behaviour, firstly the potential for an electronic component to conductivity must be assessed. In reducing conditions (e.g. H2-containing atmospheres), the cerium cations from the B-site of the perovskite structure of BCY10 can reduce from a higher oxidation state, Ce4+, to a lower one, Ce3+, altering the contribution of the concentration of the electronic charge carriers. This phenomenon is well documented in the literature for various cerium-based compositions [37, 38, 44, 45, 46, 47], being described as small-polaron electronic conductivity (i.e., a localised, mobile electron, CeCe). Due to the high mobility of electronic conductors, such electronic contribution can exceed that of the ionic, under very reducing conditions and high temperatures [37, 45, 46, 47]. However, in the case of BCY10, the extent of cerium reduction has been assessed by Loureiro et al. [37], who performed coulometric titration measurements to study the potential role of electronic contribution in BaCe0.9Y0.1O3-δ in reducing conditions as a function of temperature. This technique has been widely adopted to quantify the changes in the oxygen non-stoichiometry (Δδ), which can be associated with the reduction of Ce4+ to Ce3+, following the equation:

2CeCex+OOxVO+2CeCe+12O2E6

with the equilibrium constant for reduction reaction given by:

KR=VOCeCe2pO212OOxCeCex2E7

The results of coulometric titration (Figure 6) [37] show considerable variations of Δδ with oxygen partial pressure only at very high temperature, with a lower impact as temperature decreases. Thus, Figure 6 demonstrates that very extreme reducing conditions and very high temperatures are required to produce appreciable increase in the oxygen-vacancy and electronic concentrations in BCY [47, 48]. These results contrast with those of fluorite-ceria-based materials which usually show high reducibility under milder conditions [46, 49].

Figure 6.

Oxygen non-stoichiometry as function of oxygen partial pressure (pO2) [37] (reproduced by permission of The Royal Society of Chemistry).

Thus, to take the possibility of reduction into account, the methodology applied by Loureiro et al. [37] for the determination of reduction equilibrium follows the method reported elsewhere [50], as described below.

The corresponding mass action constant (Eq. (7)) can be combined with the electroneutrality condition:

2VO+OHOYCe+CeCeE8

and other mass and lattice position restrictions, on neglecting defect interactions and assuming nearly ideal behaviour, with the following relations between the concentrations of relevant species, stoichiometric changes (Δδ), and fraction of trivalent additive (x):

CeCe=Zv02∆δE9
VO=Zv0∆δ+x2E10
CeCex=Zv01x2∆δE11
OOx=Zv03x2∆δE12

where Z is the number of atoms per unit cell and n0, the unit cell volume. Substitution in Eq. (7) leads to the values of the equilibrium constant for reduction (KR) from the entire range of values of Δδ versus pO2 at a given temperature T:

KRT=4δ2∆δ+x2pO21/23x2∆δ1x2∆δ2E13

The following equation was then determined to describe the temperature dependence of KR, from the results of oxygen-nonstoichiometry shown in Figure 6:

KRT=4.47·1014exp7.85·104/Tatm1/2E14

with an enthalpy for reduction, ΔHR = 804.99 kJ mol−1. This value is significantly higher than those obtained by other authors for fluorite ceria-based materials (Table 2) [46, 49], underscoring the low reducibility of BCY10 in such conditions from intermediate to low temperatures.

Compound δhr (kJ mol−1)Reference
BaCe0.9Y0.1O3-x/2-Δδ805[37]
Ce0.9Gd0.1O2-x/2-Δδ410-420[46]
438[51]
Ce0.8Gd0.2O2-x/2-Δδ430[46]
385[51]
Ce0.9Sm0.1O2-x/2-Δδ400[52]
Ce0.8Sm0.2O2-x/2-Δδ385[52]
375[49]

Table 2.

Enthalpy (ΔHR) for reduction of different ceria-based based solid solutions materials.

On the basis of these results, the potential rehydration of the BCY10 material was then assessed by thermogravimetric experiments [37]. Figure 7 depicts the concentration of protonic charge carriers as a function of temperature, calculated from the following methodology.

Figure 7.

Concentration protonic defects obtained from TG in N2 and from the simulation performed in [37] (reproduced by permission of The Royal Society of Chemistry).

By expressing the equilibrium constant for water incorporation reaction (Eq. (3)) in terms of entropy, ΔSw, and enthalpy, ΔHw:

Kw=expΔSwR.expΔHwRTE15

where T and R have usual meanings. Given Eq. (1) and knowing that the number of oxygen sites per formula unit of barium cerate is restricted to 3, implying the site restriction relationship:

2VO+OHO+OOx=3E16

with Eqs. (3), (15), (16), Kw can be reformulated as

Kw=exp4OHO2pH2OSOHO6SOHOE17

and then, the concentration of protonic defects is given by

OHO=3.K9K6K.S+K.S2+24S4S2K4E18

where K′ = Kw.pH2O and S = [YCe’]. Because the formation of protonic defects is usually accompanied by a significant weight increase, the concentration of protonic defects as a function of temperature and water vapour partial pressure is generally measured by thermogravimetric analysis (TG). From Figure 7, one can observe an increase in the concentration of protonic species as a function of decreasing temperature, even in nominally dry 10%H2/N2. This factor is most likely related to the intrinsic formation of water vapour under the presence of hydrogen and oxygen impurities in the feed stream:

H2g+12O2gH2OgE19

This result emphasises the existence of protonic conductivity in nominally dry hydrogen-containing atmospheres, as even trace amounts of oxygen can form water vapour, potentially contributing to the hydration of the BCY10 material. Hence, the partial conductivities can be obtained by combining the results from both coulombic titration and TG experiments using a defect chemistry methodology [37].

Figure 8 shows the partial conductivities of all species (protons, oxide-ions and electrons) obtained at the temperature range (350–600°C) in nominally dry H2 (pH2O ∼ 10−4 atm). One can observe a dominance of the ionic charge carriers over the electronic carriers in the whole temperature range, corroborating the negligible reducibility of cerium cations measured by coulombic titration (Figure 6).

Figure 8.

Total (experimental and calculated) and partial conductivities vs. temperature. Data obtained in the temperature range 350–600°C in nominally dry conditions [37] (reproduced by permission of The Royal Society of Chemistry).

Furthermore, at the low temperature range (350–400°C), the dominance of protonic conductivity is related to the high equilibrium constant for water incorporation in BCY10, allowing a significant hydration even at pH2O values as low as ∼10−4 atm [53], as confirmed by TG (Figure 7). This behaviour also explains the slight pH2 dependence of conductivity shown in Figure 5b that is due, not to electronic behaviour, but to changes in the effective water vapour partial pressure arising from Eq. (19) and subsequent slight increase in ionic conductivity due to a higher level of hydration Eq. (18). In contrast at higher temperatures in the (550–600°C) range, oxide-ion conductivity starts to become dominant at due to the loss of protons from the structure (Figure 7).

Advertisement

3. Electrochemical properties of BCY10 in low humidity oxidising conditions

The transport numbers of BCY10 in oxidising atmospheres were firstly studied by Oishi et al. [54] and by Grimaud et al. [55]. Later, Lim et al. [56] determined the concentration of charge carriers in BCY10 by thermogravimetric analysis (TGA) under two different humidity conditions (dry and wet, pH2O ∼ 10−5 and 10−3 atm, respectively). More recently, Loureiro et al. [38] focused on the determination of the transport properties of this composition at temperatures below 600°C and under very low humidity levels (pH2O ≤ 10−4 atm).

In oxidising conditions, the absence of hydrogen species, shifts the water formation reaction, Eq. (19), away from the water product, leading to a lower intrinsic water vapour partial pressure that can, in turn, decrease the protonic transport number [38]. Therefore, at the intermediate temperature range, 350–600°C, it is necessary to externally add humidity to guarantee a sufficient level of protonic conductivity. Moreover, BCY10 is known to possess p-type electronic conductivity in oxidising atmospheres, which can importantly impact the total conductivity in these conditions [38], as expressed by

12O2g+VOOOx+2hE20

with the following mass action constant

KOh2VO.pO21/2E21

Figure 9 shows the total conductivity of BCY10 measured in the temperature range 350–600°C in wet and low humidity O2 and N2. From Figure 9, this expected decrease in the concentration of protonic species is corroborated, as in both, N2 and O2, total conductivity is shown to be higher in wet conditions (pH2O ∼ 10−3 atm) than in low humidity conditions (pH2O ∼ 10−7 atm). It is also possible to observe that low humidity N2 (pH2O ∼ 10−7 atm) the total conductivity is lower in the whole measured temperature range in comparison to wet N2 (pH2O ∼ 10−3 atm), as a result of dehydration of the sample according to Eq. (22). In contrast, in O2, the total conductivity in low humidity and wet conditions are similar, particularly at higher temperatures, a factor that can be explained due to the presence and dominance of p-type electronic conductivity [57, 58] (see Eq. (20)):

Figure 9.

Total conductivity of BCY10 in wet (pH2O ∼ 10−3 atm) and low humidity (pH2O ∼ 10−7 atm) N2 and O2. Reproduced from [38] with permission from Elsevier.

2OHOH2O+VO+OOxE22

In agreement, the presence of p-type electronic conductivity can explain the slightly higher activation energy registered in low humidity O2, 0.49 eV, in comparison to the other studied atmospheres.

Figure 10 illustrates the partial conductivities obtained in wet (pH2O ∼ 10−3 atm) and low humidity (pH2O ∼ 10−7 atm) conditions in N2 and O2. Figure 10a and b show that in moderate wet conditions (pH2O ∼ 10−3 atm) the protonic conductivity is dominating in both atmospheres with activation energies similar to that obtained for the protonic conduction (∼0.4–0.5 eV) [16, 17]. In contrast, in low humidity conditions (Figure 10c and d) a drop on protonic conductivity with increasing temperature is observed, due to predominant oxide-ion conductivity in both atmospheres. In the case of hole conductivity, the activation energies obtained were found to be lower at low humidity conditions (0.61–1.03 eV, T = 350–500°C) in comparison with those obtained in wet conditions (1.29–1.75 eV, T = 350–500°C). This can be explained by the creation of electronic defects (Eq. (20)), upon filling the oxygen vacancies.

Figure 10.

Partial conductivities obtained in wet and low humidity conditions in (a) and (b) N2, and (c) and (d) O2. The activation energy values, Ea, were calculated in the temperature range 350–500°C. Reproduced from [38] with permission from Elsevier.

Advertisement

4. Comparison between reducing and oxidising conditions under low humidity

As discussed previously, to maximise the protonic conductivity is necessary to maintain a minimum level of humidity in the order of 10−4–10−5 atm. It is also important to emphasise that, while this level of humidity is intrinsically formed in nominally dry hydrogen-containing atmospheres, in the case of oxidising atmospheres this level must be externally supplied. A comparison of the partial conductivities in all cases is shown in Figure 11, for pH2O ∼ 10−4 atm. At temperatures below 450°C, the total conductivity is dominated by protonic conductivity, with the oxide-ion conductivity taking a negligible role. In contrast, at higher temperatures (T > 450°C), the oxide- ion conductivity dominates the total conductivity with a simultaneous decrease of protonic conductivity. With respect to the electronic conductivity, this term increases as pO2 increases, being only relevant in oxidising conditions and/or high temperatures. This can be explained due to the creation of electronic holes, which become more relevant with increasing pO2 and temperature (Eq. (20)).

Figure 11.

Temperature dependence of partial conductivities in at pH2O ∼ 10−4 atm: (a) H2, (b) N2 and (c) O2. Activation energy values, Ea, calculated in the temperature range 350-500°C. Reproduced from [38] with permission from Elsevier.

Overall, BCY10 is shown to be a predominant protonic conductor in both reducing and oxidising atmospheres at sufficiently low temperatures ≤500°C, even under relatively low water vapour partial pressures (pH2O ∼ 10−4–10−5 atm). Moreover, the level of conductivity measured at 400°C in these conditions is high, e.g. ∼10−3 S cm−1. The origin of protonic conductivity is due to a high equilibrium constant for water absorption that allows this material to offer high bulk protonic conductivity at intermediate temperatures in these very low humidity conditions. From Figure 12, one can immediately envisage that this is a particular behaviour of BCY10 that cannot be obtained in other competing proton-conducting perovskites, due to their much lower values of Kw.

Figure 12.

Equilibrium constant for hydration of several perovskite proton conductors. Adapted from [9].

This is a very exciting result since it opens a wide range of possibilities for using the BCY material, in particular, in different applications that involve very low humidity levels and low temperatures of operation. The most well-known is that of ammonia electrochemical synthesis [39, 40, 41], although many other processes concerning hydrogenation and de-hydrogenation reactions can also be considered that agree with these operating conditions (Table 3).

TypeReactionH2980/kJmol1
Dehydrogenations2CH4gC2H4g+2H2g202
6CH4gC6H6g+9H2g89
C3H8gC3H6g+H2g124.3
iC4H10giC4H8g+H2g122
C8H10gC8H8g+H2g117.6
HydrogenationsC10H8g+2H2gC10H12g−134
C6H10g+H2gC6H12g−120
C6H6g+3H2gC6H12g−207
N2g+3H2g2NH3g−109

Table 3.

Examples of dehydrogenation/hydrogenation reactions that can occur at very low humidity conditions.

Reproduced from [38] with permission from Elsevier.

Advertisement

5. Conclusions

The current chapter highlights that the transport properties of BaCe0.9Y0.1O3-δ (BCY10) in very low humidity conditions are dependent on the nature of the surrounding atmosphere and on the temperature, being significantly different in reducing and oxidising conditions and at high and low temperatures. In reducing conditions, BCY10 shows a very high protonic conductivity (e.g. ∼ 10−3 S cm−1) at low temperatures i.e. < 400°C, even in nominally dry atmospheres with negligible oxide-ion/electronic influence.

In the other hand, in oxidising conditions, the same behaviour can only be obtained by externally supplying humidity in the range (pH2O ∼ 10−4–10−5 atm) at low temperatures ≤500°C. At higher temperatures, at this low humidity, the onset of hole conductivity can be noted at higher oxygen partial pressures due partial material dehydration.

The present discussion shows the importance of controlling the humidity levels in order to maximise the protonic conductivity of BCY under operation. The very low levels of humidity required (pH2O ∼ 10−4–10−5 atm), to ensure predominant proton conductivity in both reducing and oxidising atmospheres at low temperatures ≤500°C, are highly interesting as they highlight the possibility of using this composition in applications where low humidity levels and temperatures are required, such as the suggested de-hydrogenation/hydrogenation chemical reactions, while maintaining its stability against decomposition.

Advertisement

Acknowledgments

The authors acknowledge Fundação para a Ciência e Tecnologia (FCT) for the PhD grants – PD/BDE/142837/2018, SFRH/BD/130218/2017, and PD/BDE/114353/2016. The authors also acknowledge the projects UID/EMS/00481/2019-FCT and CENTRO-01-0145-FEDER-022083 - Centro Portugal Regional Operational Programme (Centro2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund (ERDF).

References

  1. 1. Morejudo SH, Zanón R, Escolástico S, Yuste-Tirados I, Malerød-Fjeld H, Vestre PK, et al. Direct conversion of methane to aromatics in a catalytic co-ionic membrane reactor. Science (80- ) [Internet]. 2016 Aug 5;353(6299):563 LP – 566. Available from: http://science.sciencemag.org/content/353/6299/563.abstract
  2. 2. Bi L, Boulfrad S, Traversa E. Steam electrolysis by solid oxide electrolysis cells (SOECs) with proton-conducting oxides. Chem Soc Rev [Internet]. 2014;43(24):8255–70. Available from: http://dx.doi.org/10.1039/C4CS00194J
  3. 3. Fabbri E, Bi L, Pergolesi D, Traversa E. Towards the Next Generation of Solid Oxide Fuel Cells Operating Below 600 °C with Chemically Stable Proton-Conducting Electrolytes. Adv Mater [Internet]. 2012;24(2):195–208. Available from: http://dx.doi.org/10.1002/adma.201103102
  4. 4. Bannykh A V, Kuzin BL. Electrical conductivity of BaCe0.9Nd0.1O3−α in H2+H2O+Ar gas mixture. Ionics (Kiel) [Internet]. 2003;9(1):134–9. Available from: https://doi.org/10.1007/BF02376550
  5. 5. Kuzin BL, Beresnev SM, Bannykh A V, Perfil’yev M V. Transport numbers of H+ and O2- in the electrochemical system (H2 + H2O),Me/BaCe0.9Nd0.1O3-α/Me,(H2 + H2O). Russ J Electrochem [Internet]. 2000;36(4):424–30. Available from: https://doi.org/10.1007/BF02756951
  6. 6. Qiu L-G, Ma G-L, Wen D-J. Properties and Application of Ceramic BaCe0.8Ho0.2O3−α. Chinese J Chem [Internet]. 2005 Dec 22;23(12):1641–5. Available from: https://doi.org/10.1002/cjoc.200591641
  7. 7. Virkar AN, Maiti HS. Oxygen ion conduction in pure and yttria-doped barium cerate. J Power Sources [Internet]. 1985;14(4):295–303. Available from: http://www.sciencedirect.com/science/article/pii/0378775385800458
  8. 8. Suksamai W, Metcalfe IS. Measurement of proton and oxide ion fluxes in a working Y-doped BaCeO3 SOFC. Solid State Ionics [Internet]. 2007;178(7):627–34. Available from: http://www.sciencedirect.com/science/article/pii/S0167273807000549
  9. 9. Kreuer KD. Proton conducting oxides. Annu Rev Mater Res. 2003 Aug;33(1):333–59
  10. 10. Malavasi L, Fisher CAJ, Islam MS. Oxide-ion and proton conducting electrolyte materials for clean energy applications: structural and mechanistic features. Chem Soc Rev [Internet]. 2010;39(11):4370–87. Available from: http://dx.doi.org/10.1039/B915141A
  11. 11. Fabbri E, D’Epifanio A, Di Bartolomeo E, Licoccia S, Traversa E. Tailoring the chemical stability of Ba(Ce0.8-xZrx)Y0.2O3-d protonic conductors for Intermediate Temperature Solid Oxide Fuel Cells (IT-SOFCs). Solid State Ionics. 2008;179(15–16):558–64
  12. 12. Brett DJL, Atkinson A, Brandon NP, Skinner SJ. Intermediate temperature solid oxide fuel cells. Chem Soc Rev [Internet]. 2008;37(8):1568–78. Available from: http://dx.doi.org/10.1039/B612060C
  13. 13. Liu M, Lynch ME, Blinn K, Alamgir FM, Choi Y. Rational SOFC material design: new advances and tools. Mater Today [Internet]. 2011;14(11):534–46. Available from: http://www.sciencedirect.com/science/article/pii/S1369702111702796
  14. 14. Coors WG, Readey DW. Proton Conductivity Measurements in Yttrium Barium Cerate by Impedance Spectroscopy. J Am Ceram Soc [Internet]. 2004 Dec 20;85(11):2637–40. Available from: http://www3.interscience.wiley.com/journal/118935979/abstract
  15. 15. Ma G, Shimura T, Iwahara H. Simultaneous doping with La3+ and Y3+ for Ba2+ − and Ce4+ −sites in BaCeO3 and the ionic conduction. Solid State Ionics. 1999;120(1):51–60
  16. 16. Kreuer KD, Dippel T, Baikov YM, Maier J. Water solubility, proton and oxygen diffusion in acceptor doped BaCeO3: A single crystal analysis. Solid State Ionics [Internet]. 1996;86–88, Par:613–20. Available from: http://www.sciencedirect.com/science/article/pii/0167273896002214
  17. 17. Bonanos N, Ellis B, Knight KS, Mahmood MN. Ionic conductivity of gadolinium-doped barium cerate perovskites. Solid State Ionics [Internet]. 1989;35(1–2):179–88. Available from: http://www.sciencedirect.com/science/article/pii/0167273889900283
  18. 18. Slade RCT, Singh N. The perovskite-type proton-conducting solid electrolyte BaCe0.90Y0.10O3−α in high temperature electrochemical cells. Solid State Ionics [Internet]. 1993;61(1–3):111–4. Available from: http://www.sciencedirect.com/science/article/pii/016727389390342Z
  19. 19. Fabbri E, Pergolesi D, Traversa E. Materials challenges toward proton-conducting oxide fuel cells: a critical review. Chem Soc Rev [Internet]. 2010;39(11):4355–69. Available from: http://dx.doi.org/10.1039/B902343G
  20. 20. Fabbri E, Pergolesi D, Traversa E. Electrode materials: a challenge for the exploitation of protonic solid oxide fuel cells. Sci Technol Adv Mater [Internet]. 2010 Aug 10;11(4):44301. Available from: http://www.ncbi.nlm.nih.gov/pmc/articles/PMC5090333/
  21. 21. Kruth A, Irvine JTS. Water incorporation studies on doped barium cerate perovskites. Solid State Ionics. 2003;162–163:83–91
  22. 22. Matsumoto H, Kawasaki Y, Ito N, Enoki M, Ishihara T. Relation Between Electrical Conductivity and Chemical Stability of BaCeO3-Based Proton Conductors with Different Trivalent Dopants. Electrochem Solid-State Lett [Internet]. 2007 Apr 1;10(4):B77–80. Available from: http://esl.ecsdl.org/content/10/4/B77.abstract
  23. 23. Eriksson Andersson AK, Selbach SM, Grande T, Knee CS. Thermal evolution of the crystal structure of proton conducting BaCe0.8Y0.2O3-[small delta] from high-resolution neutron diffraction in dry and humid atmosphere. Dalt Trans [Internet]. 2015;44(23):10834–46. Available from: http://dx.doi.org/10.1039/C4DT03948C
  24. 24. Lacz A. Effect of microstructure on chemical stability and electrical properties of BaCe0.9Y0.1O3{\thinspace}−{\thinspace}$δ$. Ionics (Kiel) [Internet]. 2016;22(8):1405–14. Available from: http://dx.doi.org/10.1007/s11581-016-1665-6
  25. 25. Li Y, Su P-C, Wong LM, Wang S. Chemical stability study of nanoscale thin film yttria-doped barium cerate electrolyte for micro solid oxide fuel cells. J Power Sources [Internet]. 2014;268:804–9. Available from: http://www.sciencedirect.com/science/article/pii/S0378775314009926
  26. 26. Sun Z, Fabbri E, Bi L, Traversa E. Lowering grain boundary resistance of BaZr0.8Y0.2O3-[small delta] with LiNO3 sintering-aid improves proton conductivity for fuel cell operation. Phys Chem Chem Phys [Internet]. 2011;13(17):7692–700. Available from: http://dx.doi.org/10.1039/C0CP01470B
  27. 27. Donglin H, Naoyuki H, Tetsuya U. Chemical Expansion of Yttrium-Doped Barium Zirconate and Correlation with Proton Concentration and Conductivity. J Am Ceram Soc [Internet]. 2016 Jun 30;99(11):3745–53. Available from: https://doi.org/10.1111/jace.14377
  28. 28. Muccillo R, Muccillo ENS, Andrade TF, Oliveira OR. Thermal analyses of yttrium-doped barium zirconate with phosphor pentoxide, boron oxide and zinc oxide addition. J Therm Anal Calorim [Internet]. 2017;130(3):1791–9. Available from: https://doi.org/10.1007/s10973-017-6523-x
  29. 29. Narendar N, Mather GC, Dias PAN, Fagg DP. The importance of phase purity in Ni–BaZr0.85Y0.15O3−δ cermet anodes – novel nitrate-free combustion route and electrochemical study. RSC Adv [Internet]. 2013;3(3):859–69. Available from: http://dx.doi.org/10.1039/C2RA22301E
  30. 30. Soares HS, Zhang X, Antunes I, Frade JR, Mather GC, Fagg DP. Effect of phosphorus additions on the sintering and transport properties of proton conducting BaZr0.85Y0.15O3−δ. J Solid State Chem [Internet]. 2012;191:27–32. Available from: http://www.sciencedirect.com/science/article/pii/S0022459612001557
  31. 31. Yang T, Loureiro FJAFJA, Queirós RP, Pukazhselvan D, Antunes I, Saraiva JAJA, et al. A detailed study of hydrostatic press, sintering aids and temperature on the densification behavior of Ba(Zr,Y)O3−delectrolyte. Int J Hydrogen Energy. 2016;41(27):1–10
  32. 32. Bozza F, Bator K, Kubiak WW, Graule T. Effects of Ni doping on the sintering and electrical properties of BaZr0.8Y0.2O3−δ proton conducting electrolyte prepared by Flame Spray Synthesis. J Eur Ceram Soc [Internet]. 2016;36(1):101–7. Available from: http://www.sciencedirect.com/science/article/pii/S0955221915301291
  33. 33. Grant H, Anthony M, G. CW, Sandrine R. Chemical expansion in BaZr0.9−xCexY0.1O3−δ (x = 0 and 0.2) upon hydration determined by high-temperature X-ray diffraction. J Am Ceram Soc [Internet]. 2017 Oct 7;101(3):1298–309. Available from: https://doi.org/10.1111/jace.15275
  34. 34. Kim H-S, Bae H Bin, Jung W, Chung S-Y. Manipulation of Nanoscale Intergranular Phases for High Proton Conduction and Decomposition Tolerance in BaCeO3 Polycrystals. Nano Lett [Internet]. 2018 Feb 14;18(2):1110–7. Available from: https://doi.org/10.1021/acs.nanolett.7b04655
  35. 35. Park J-S, Lee J-H, Lee H-W, Kim B-K. Low temperature sintering of BaZrO3-based proton conductors for intermediate temperature solid oxide fuel cells. Solid State Ionics [Internet]. 2010;181(3):163–7. Available from: http://www.sciencedirect.com/science/article/pii/S0167273809002549
  36. 36. Loureiro FJA, Nasani N, Reddy GS, Munirathnam NR, Fagg DP. A review on sintering technology of proton conducting BaCeO3-BaZrO3 perovskite oxide materials for Protonic Ceramic Fuel Cells. J Power Sources [Internet]. 2019;438:226991. Available from: http://www.sciencedirect.com/science/article/pii/S037877531930984X
  37. 37. Loureiro FJA, Pérez-Coll D, Graça VCD, Mikhalev SM, Ribeiro AFG, Mendes A, et al. Proton conductivity in yttrium-doped barium cerate in nominally dry reducing conditions for application in chemical synthesis. J Mater Chem A [Internet]. 2019;7:18135–42. Available from: submitted
  38. 38. Loureiro FJA, Ramasamy D, Ribeiro AFG, Mendes A, Fagg DP. Underscoring the transport properties of yttrium-doped barium cerate in nominally dry oxidising conditions. Electrochim Acta [Internet]. 2019;334:135625. Available from: https://doi.org/10.1016/j.electacta.2020.135625
  39. 39. Otomo J, Noda N, Kosaka F. Electrochemical Synthesis of Ammonia with Proton Conducting Solid Electrolyte Fuel Cells at Intermediate Temperatures. ECS Trans [Internet]. 2015 Jun 2;68(1):2663–70. Available from: http://ecst.ecsdl.org/content/68/1/2663.abstract
  40. 40. Li Z, Liu R, Wang J, Xu Z, Xie Y, Wang B. Preparation of double-doped BaCeO3 and its application in the synthesis of ammonia at atmospheric pressure. Sci Technol Adv Mater [Internet]. 2007;8(7–8):566–70. Available from: http://www.sciencedirect.com/science/article/pii/S1468699607001404
  41. 41. Marnellos G, Stoukides M. Ammonia Synthesis at Atmospheric Pressure. Science (80- ) [Internet]. 1998 Oct 2;282(5386):98 LP – 100. Available from: http://science.sciencemag.org/content/282/5386/98.abstract
  42. 42. Feng Y, Luo J, Chuang KT. Propane Dehydrogenation in a Proton-conducting Fuel Cell. J Phys Chem C [Internet]. 2008 Jul 1;112(26):9943–9. Available from: http://dx.doi.org/10.1021/jp710141c
  43. 43. Bale CW, Bélisle E, Chartrand P, Decterov SA, Eriksson G, Gheribi AE, et al. FactSage thermochemical software and databases, 2010–2016. Calphad [Internet]. 2016;54:35–53. Available from: http://www.sciencedirect.com/science/article/pii/S0364591616300694
  44. 44. Bonanos N, Willy Poulsen F. Considerations of defect equilibria in high temperature proton-conducting cerates. J Mater Chem [Internet]. 1999;9(2):431–4. Available from: http://dx.doi.org/10.1039/A805150J
  45. 45. Pérez-Coll D, Aguadero A, Núñez P, Frade JR. Mixed transport properties of Ce1−xSmxO2−x/2 system under fuel cell operating conditions. Int J Hydrogen Energy [Internet]. 2010;35(20):11448–55. Available from: http://www.sciencedirect.com/science/article/pii/S0360319910009900
  46. 46. Pérez-Coll D, Marrero-López D, Ruiz-Morales JC, Núñez P, Abrantes JCC, Frade JR. Reducibility of Ce1−xGdxO2−δ in prospective working conditions. J Power Sources [Internet]. 2007;173(1):291–7. Available from: http://www.sciencedirect.com/science/article/pii/S0378775307009044
  47. 47. Bonanos N. Transport study of the solid electrolyte BaCe0.9Gd0.1O2.95 at high temperatures. J Phys Chem Solids [Internet]. 1993;54(7):867–70. Available from: http://www.sciencedirect.com/science/article/pii/002236979390258S
  48. 48. Chen W, Nijmeijer A, Winnubst L. Oxygen non-stoichiometry determination of perovskite materials by a carbonation process. Solid State Ionics [Internet]. 2012 Dec 14;229:54–8. Available from: http://www.sciencedirect.com/science/article/pii/S0167273812005723
  49. 49. Abrantes JCC, Pérez-Coll D, Núñez P, Frade JR. Electronic transport in Ce0.8Sm0.2O1.9−δ ceramics under reducing conditions. Electrochim Acta [Internet]. 2003 Aug 15;48(19):2761–6. Available from: http://www.sciencedirect.com/science/article/pii/S0013468603003955
  50. 50. Pérez-Coll D, Marrero-López D, Ruiz-Morales JC, Núñez P, Abrantes JCC, Frade JR. Reducibility of Ce1-xGdxO2-δ in prospective working conditions. J Power Sources. 2007;173(1):291–7
  51. 51. Wang S, Inaba H, Tagawa H, Hashimoto T. Nonstoichiometry of Ce0.8Gd0.2 O 1.9 − x . J Electrochem Soc [Internet]. 1997 Nov 1;144(11):4076–80. Available from: http://jes.ecsdl.org/content/144/11/4076.abstract
  52. 52. Kobayashi T, Wang S, Dokiya M, Tagawa H, Hashimoto T. Oxygen nonstoichiometry of Ce1−ySmyO2−0.5y−x (y=0.1, 0.2). Solid State Ionics [Internet]. 1999;126(3):349–57. Available from: http://www.sciencedirect.com/science/article/pii/S0167273899002593
  53. 53. Kreuer KD. Aspects of the formation and mobility of protonic charge carriers and the stability of perovskite-type oxides. Solid State Ionics [Internet]. 1999 Oct;125(1–4):285–302. Available from: http://www.sciencedirect.com/science/article/pii/S0167273899001885
  54. 54. Oishi M, Akoshima S, Yashiro K, Sato K, Mizusaki J, Kawada T. Defect structure analysis of B-site doped perovskite-type proton conducting oxide BaCeO3: Part 2: The electrical conductivity and diffusion coefficient of BaCe0.9Y0.1O3−δ. Solid State Ionics [Internet]. 2008;179(39):2240–7. Available from: http://www.sciencedirect.com/science/article/pii/S016727380800564X
  55. 55. Grimaud A, Bassat JM, Mauvy F, Simon P, Canizares A, Rousseau B, et al. Transport properties and in-situ Raman spectroscopy study of BaCe0.9Y0.1O3−δ as a function of water partial pressures. Solid State Ionics [Internet]. 2011;191(1):24–31. Available from: http://www.sciencedirect.com/science/article/pii/S0167273811001767
  56. 56. Lim D-K, Im H-N, Song S-J, Yoo H-I. Hydration of Proton-conducting BaCe0.9Y0.1O3−δ by Decoupled Mass Transport. Sci Rep [Internet]. 2017;7(1):486. Available from: http:https://dx.doi.org/10.1038/s41598-017-00595-w
  57. 57. Heras-Juaristi G, Pérez-Coll D, Mather GC. Temperature dependence of partial conductivities of the BaZr0.7Ce0.2Y0.1O3-δ proton conductor. J Power Sources [Internet]. 2017;364:52–60. Available from: http://www.sciencedirect.com/science/article/pii/S0378775317310352
  58. 58. Triviño-Peláez Á, Pérez-Coll D, Mather GC. Electrical properties of proton-conducting BaCe0.8Y0.2O3-δ and the effects of bromine addition. Acta Mater [Internet]. 2019;167:12–22. Available from: http://www.sciencedirect.com/science/article/pii/S1359645419300436

Written By

Laura I.V. Holz, Vanessa C.D. Graça, Francisco J.A. Loureiro and Duncan P. Fagg

Submitted: 15 March 2020 Reviewed: 10 September 2020 Published: 04 November 2020