You are on page 1of 36

Ref.1. New intermediate temperature fuel cell with ultra-thin proton conductor electrolyte - J.

Power Sources 152, (2005), 200.

Schematic diagram of hydrogen membrane fuel cell structure.

A new type fuel cell called as hydrogen membrane fuel cell (HMFC). A much thinner electrolyte can be easily realized because it is formed on a solid, nonporous membrane. Another advantage of this approach is an ease of high density stacking because the physical base of this fuel cell is a metal film, not ceramics as in the case of SOFCs. The HMFC can use only proton conductors as an electrolyte because the hydrogen membrane layer only permeates hydrogen. The BaCe0.8Y0.2O3 perovskite was chosen for the electrolyte material. The electrolyte layer, which is 0.7 m in thickness, was coated on Pd film by pulse laser deposition. Test cells were operated from 400 to 600 C. The open circuit voltages were close to theoretical value in all operating temperatures. The power density was 0.9 and 1.4 W cm2 at the operating temperature of 400 and 600 C.

Table . Open circuit voltage of single cell in various conditions IV characteristics of single cell at various temperatures. Anode gas was moist H2 and cathode gas was moist air (both 40 C humidified).
Temperature Anode gas Cathode gas (C) 440 440 440 530 610 H2 = 100% H2 = 50% H2 = 10% H2 = 100% H2 = 100% Air Air Air Air Air Measured OCV (mV) 1103 1082 1036 1081 1051 Theoretical voltage (mV) 1140 1120 1073 1120 1100

Ref.2. Electrochemical analysis of hydrogen membrane fuel cells- J. Power Sources 185, (2008), 922.
Electrochemical analysis was conducted with respect to a hydrogen membrane fuel cell (HMFC) with SrZr0.8In0.2O3 electrolyte. Most of the voltage loss derives from the cathode and the electrolyte, and a small amount of anode polarization was observed only in regions with high current density. The cathode polarization was nearly an order of magnitude lower than that of SOFCs. The conductivity of the film electrolyte was almost identical to that of the sinter at 600 C; however, it was several times as large at 400 C.

VI characteristics of test cells with various electrolyte thickness at 400 C.

AC impedance spectra of test cell with electrolyte thickness of 2 m at 400 C. (a) Effects of the bias current, (b) effects of concentration of H2 in anode gas, and (c) effects of concentration of O2 in cathode gas.

Electrolyte, anode and cathode polarization as a function of current density at 400 C in humidified H2 and air: (a) d = 0.7 m; (b) d = 2.0 m; (c) d = 4.0 m; (d) d = 6.0 m.

Table- Open-circuit voltage of singe cells in various conditions (Anode gas: H2; cathode gas: air.)
Temp. (C) Electrolyte Thickness (m) Measured OCV (mV) Theoretical voltage (mV)

400

Electrolyte resistance as a function of electrolyte thickness at 400 C.

VI characteristics of test cell with electrolyte thickness of 4 m at various temperatures.

500 600

0.7 2.0 4.0 6.0 4.0 4.0

980 1020 1100 1110 1070 1048

1142 1142 1142 1142 1123 1102

Most of the voltage loss of the HMFC is due to cathode activation polarization and electrolyte overpotential, and small anode concentration polarization was observed only in regions with high current density. The cathode polarization was approximately an order of magnitude lower than that of SOFCs, and displayed strong dependence on the temperature. Although the conductivity of the film electrolyte was almost identical to that of the sinter at 600 C, it was several times as large at 400 C. In addition, the conductivity and the thickness of the film electrolyte displayed a non-proportional relation.

TEM cross-section image and schematic model of the film electrolyte. TEM micrograph revealed that the film electrolyte consists mainly of long columnar crystals which can be related to the conductivity enhancement below 600 C.

Ref. 3. Efficient, Anhydrous Proton-Conducting Nanofilms of Y-Doped Zirconium Pyrophosphate at Intermediate Temperatures- Adv. Mater. 20, 2008, 2398. Nanofilms of amorphous Y-doped zirconium phosphate (ZrYP-120 film annealed at 120 0C) and pyrophosphate (ZrYP-400 film annealed at 400 0C) by a novel approach (a combination of the layer-by-layer solgel process and annealing). The obtained defect-free ZrYP-400 nanofilm shows high proton conductivity with ASR as low as 0.085Vcm2 at 340 0C under anhydrous conditions. The conductivity of ZrYP-120 film depends on moisture content (conductive in moist air at 20 0C but non-conductive in dry air) but the conductivity of ZrYP-400 film is independent of moisture content. In ZrYP-120 film conduction may be due to H3O+ or H5O2+ but in ZrYP-400 film it is transport of hydrogen-bonded protons through transformation of intermolecular hydrogen bond

The effective proton concentration is enhanced by Y-doping.

The impedance response of the Zr0.95Y0.05P3400 pyrophosphate nanofilm at different temperature in the flow of dry air.

The conductivity () value of the ZrP3-400 film increases from 1.6x10-7 to 4.2x10-5 S cm-1, from 200 to 400 0C. With the doping of 5 mol % yttrium (Zr0.95Y0.05P3-400) increases from 7.4x10-7 to 1.5x10-4 S cm-1 from 200 to 340 0C. At 340 0C, its proton conductivity is 12.5 times higher than that of the ZrP3-400 nanofilm. Further increase in the doping (Zr0.90Y0.10P3-400), leads to lesser improvement of proton conductivity due to the formation of segregated yttrium phosphate leading to partial disconnection of the P2O7 network for proton conduction.
The turn around 280 0C in the curve of the Zr0.95Y0.05P3-400 film implies that there might be a change of conducting mechanisms or conducting species. The turn around is most probably ascribed to hydrogen-bond sites with higher (OH), which require higher activation energy, that contribute to the proton conduction above 280 0C, while those with lower (OH) are main proton conducting species below 280 0C.

Ref. 4. Thickness-Induced Proton-Conductivity Transition in Amorphous Zirconium Phosphate Thin Films Chem. Mater. 22, (2010), 5528.
amorphous ZrP2.5Ox thin films were a thermally stable, anhydrous proton conductor since these retain the abundant protons bound to metaphosphate glass framework. ZrP2.5Ox films reveal unique conductivity transition as followed by decrease of thickness, which is related to the water content inside film. The proton conductivity in dry air is not dependent on the thickness in d > 60 nm, but it increases as reducing thickness into below 60 nm. As a consequence, the conductivity of 40 nm thick film is 200 times higher than that of 60 nm thick film. Furthermore, the conductivity of films with d of 100 nm increases by 2 orders of magnitude by hydration treatment and becomes as high as that of the as-prepared 40 nm thick film, though the conductivity of 40 nm thick film is kept at high level regardless of hydration. The elevated conductivity of thin films can attribute to the structural modification of metaphosphate glass framework by hydration. The hydrated, high-conductive phase is thermally stable and does not restore to the original unhydrated phase by annealing at 400 C in dry air, but it is stabilized only as the thickness is less than a few tens of nanometers. This findings strongly suggest that the hydration of metaphosphate films is encouraged in the nanometer-thick layer in proximity to the electrode rather than in the surface layer exposed directly to moistures. Consequently, the highly conductive, hydrated phase is confined within the region of a few tens of nanometers thickness from electrode interface, so that the relatively thick films reveal the phase separation between the inner hydrated nanolayer and the outer unhydrated layer. This is the first example of the proton-conductivity transition triggered by the size-confined hydration effect. The origin for this sizeconfinement is still unclear, and is not readily explained by the exponential decay of structure induced by the fixed charge or free energy at the interface. These phenomena reported here are speculated to be operative in various phosphate-based amorphous proton conductors, since the Gibbs energy of amorphous compounds must be larger than that of the crystalline materials under same composition. The current results pose an opportunity to create the high-proton-conductivity electrolyte based on the hydrated metaphosphate glass.

Temperature dependency of dry of ZP film with different thicknesses in dry air.

Plots of dry at 150 C (O) and Eadry () of ZP films and H2O at 150 C () and EaH2O () of ZP-H films.

physically adsorbed and/or chemically bound waters

The amount of desorbed CH4 (m/z = 16), CO (m/z = 28), and CO2 (m/z = 44) is much smaller than that of desorbed H2O and OH in the measured temperature range the water does not evolve by combustion of organic contaminants. There is no clear difference in the intensity of peak at around 400 C between the 40 nm and the 100 nm the concentration of water at elevated temperature in the 40 nm is larger than that in the 100 nm because the volume of the former is much smaller than that of the latter.

proton stable at elevated temperatures

Thermal Decomposition Spectra (TDS) of H2O (m/z = 18) of ZP films with thickness of 100 nm () and 40 nm (- -).

Temperature dependency of proton conductivity of 40 and 100 nm thick -ZrP2.5Ox films treated under various conditions.

The conductivity of 40 nm thick film is not varied by annealing in H2O/air for several hours. On the other hand, H2O of a 100 nm thick ZP-H film is nearly 2 orders of magnitude larger than dry of the corresponding ZP film at temperatures below ca. 200 C and EaH2O of the former is apparently smaller than that of the latter. Above 200 C, ASR of ZR-H films is too small to detect precisely with present setup. dry of 100 nm thick ZP-H/dry film is almost same as H2O of the ZP-H, indicating that the enhanced conductivity of ZP-H is not restored to the original value of ZP even though the ZP-H was annealed again at 400 C in dry air for overnight. These suggest that the relatively thick a-ZrP2.5Ox film reacts with H2O at elevated temperatures to form the hydrated, highconductive phase and the hydrated films do not recover to the original phase, but the high value of of a thin film is maintained regardless of hydration treatment. The D2O of 100 nm thick ZP-H/D is lower than H2O of the ZP-H by a factor of 1.21.4, showing that the conductivity decreases by H/D isotope exchange. This result confirms that the equilibrium between film and moisture actually takes place although of the hydrated ZP-H is not sensitive to the humidity. This ratio of H2O / D2O is closer to the square root of MD/MH 1.4, suggesting that a-ZrP2.5Ox film dominantly conducted protons by thermally activated hopping process.

Ref. 5. Thin Film Fuel Cell Based on Nanometer-Thick Membrane of Amorphous Zirconium Phosphate Electrolyte - J. Electrochem. Soc. 158, (2011), B866.
The hydrogen permeable membrane fuel cell (HMFC) based on amorphous ZrP2.6Ox electrolyte operates at 400C. The electrolyte reveals the protonic transport number of unity without electronic leakage in FC conditions with stable OCV of 1.0 V similar to the theoretical value. The maximum power density is still limited and is about 2 mW cm2 because of the large polarization related to the proton transfer and the oxidative reaction of metal hydride at anode/electrolyte hetero interfaces. Ni interlayer was introduced between the Pd anode and the ZrP2.6Ox electrolyte in order to suppress the deterioration of the electrolyte nanofilm by the deformation of the Pd anode during hydrogen absorption. In the ZrP2.6Ox electrolyte the transport number of proton was unity at 400C as determined by an EMF measurement. The modification of the Ni anode surface by an ultrathin Pt or Pd layer effectively decreased the anode/electrolyte interfacial polarization.

Figure-(a) I-V characteristics of a fuel cell, H2 (pH2 = 0.5 atm), Ni@Pd |ZrP2.6Ox| Pt, air, at 400C. (b) Impedance response of the fuel cell at OCV. The solid line indicates the calculated curve with an equivalent circuit model depicted in (c). The semicircle drawn by red dashed line and blue dashed line is fitted to the high-frequancy semicircle (HS) and low-frequency semicircle (LS), respectively. (c) Equivalent circuit model. R: resistance and Q: pseudo capacitance (constant phase element). SEM images of the surface of Pd anode support (a) before and (b) after exposing to 50%H2/Ar gas at 400 0C for 2 h. (c) Cross-sectional TEM image of HMFC of Ni@Pd/ZrP2.6Ox/Pt. The inset is area-selected electron diffraction pattern from the ZrP2.6Ox film. The scale bar in (a) and (b) is 1 m, and that in (c) is 150 nm.

Figure 3. I-V characteristics of the Ni/Pd |ZrP2.6Ox| Pt cells with various thicknesses of the Ni interlayer. The I-V relationship is not varied by changing the thickness, indicating that the low diffusivity of hydrogen through Ni ((i)) is not responsible for the poralization resistances of the HMFC. Fig. 4. (a) Impedance responses at OCV of H2, Ni@Pd |ZrP2.6Ox| Pt, pO2(Pt) fuel cell at 400C with changing oxygen partial pressure at cathode, pO2(Pt), from 0.1 to 1.0. (b) Impedance responses at OCV of H2, Ni@Pd |ZrP2.6Ox| Pt, air fuel cell at 400C with changing hydrogen partial pressure at anode, pH2(Pd), from 0.2 to 1.0. The resistances of the cell clearly change with the gas concentrations at cathode and anode which indicate that the anodic processes significantly contribute to a large polarization resistances of the cell. Fig. 5. (a) I-V characteristics at 400C of the HMFC formed on anodes of Ni/Pd, Pt@Ni/Pd and Pd@NiPd. (b) impedance responses of the cell at OCV. The HS and LS of the cell with an ultrathin Pt anode surface layer (ca. 5 nm thickness), Pt@Ni/Pd |ZrP2.6Ox| Pt, become smaller than those of the cell without the Pt surface layer and the maximum power density increases by 40% even though OCV is slightly lower than 1.0 V. When an ultrathin Pd layer is implemented, the semicircles at the higher frequency (HS) and lower frequency (LS) are effectively depressed and the maximum power density of the cell with the configuration of Pd@Ni/Pd |ZrP2.6Ox| Pt (1.8 mW cm2) is twice higher than that of Ni/Pd |ZrP2.6Ox| Pt. These results strongly suggest that the electrochemical processes at the anode, which can be assigned to the proton transfer across the hetero interface between the metal anode and the oxide electrolyte and to the hydrogen dissociation reaction of the anode, are responsible for both HS and LS.

Pt@Ni/Pd

Ni/Pd

Pd@NiPd

Fig. 6. Electromotive force of hydrogen concentration cell with a configuration of H2 (pH2 = 1), Ni@Pd |ZrP2.6Ox| Pt, pH2(Pt) = 0.10.8 atm. The black solid line indicates the theoretical EMF given by Eq. (3). The measured (o) and the calculated slope (- - -) of EMF in wet condition. The measured () and the calculated slope ( ) of EMF in dry condition.

These results suggest that the contribution of an oxide ion and an electron to the charge carriers in the ZrP2.6Ox is negligibly small in comparison with that of a proton. The large anodic polarization in our HMFC is not related to the water evolution at the solid-solid hetero interface.

Ref. 6. Proton-Conductivity of Amorphous Aluminum Phosphate Thin Films under Anhydrous Conditions - J. Electrochem. Soc. 158, (2011), P41.
Summary: 100-nm thick films of amorphous aluminum phosphate were prepared in various Al/Si molar ratios by multiple spin-coating with the mixed alkoxide solutions. These were efficient proton conductor at temperatures above 200C under anhydrous conditions. The P-rich films consisting of the aluminum metaphosphate glass phase kept the relatively high conductivity of the order of 105 S cm1 in the temperature region. On the other hand, Al-rich films made of the Al2O3P2O5 mixed glass phase could exhibit such high conductivity only at around 400C due to the large activation energy. The current results strongly suggest that the conductivity of a-AlPnOx glassy films can be improved by forming the optimal glass network structure.
Al1P3

Al1P1

Al3P2

Al1P2

Fig. 2. Colecole plot of 110-nm thick Al1P2 films, in dry Ar (O). Solid line is calculated with equivalent circuit model shown in inset. Rox = resistance and Qox = capacitance of film bulk.

Fig. 3. Temperature dependency of of Al3P2 (d = 125 nm), Al1P1 (d = 105 nm), Al1P2 (d = 110 nm), and Al1P3 (d = 90 nm), measured in dry Ar.

Al-rich films, Al3P2 and Al1P1, reveal the linear Arrhenius relationship in the whole temperature range with an activation energy Ea of 1.0 and 0.9 eV, respectively. The values of Al1P1 are higher than those of Al3P2 by a factor of about 4. On the other hand, P-rich films, Al1P2 and Al1P3, reveal nonlinear temperature dependency and exhibit the apparent change of slope at round 200C. of the P-rich films linearly increases with an activation energy Ea of 0.7 eV below 200C with a break at around 200C and tends to increase with an small Ea of 0.20.3 eV above 220C. The value of Al1P2 is very similar to that of Al1P3 in the whole temperature range.

Furthermore, of P-rich films is at least 2 orders of magnitude higher than that of Al-rich films at temperatures below 200C, but the latter becomes close to the former at around 400C.

Conduction in a-AlPnOx films The resistances in Fig. 4 are normalized with resistance at 0 s in 1%-H2/Ar, R0. When the atmosphere is altered from 1%-H2/Ar to 1%-D2/Ar, both films reveal the abrupt increase of resistivity. The resistivity change exponentially decays and saturates for less than 1 h in both specimens. Consequently, the ratio of resistance in dry 1%-D2/Ar to that in 1%-H2/Ar is about 1.28 for Al1P1 and 1.22 for Al2P2. These values are relatively smaller than the expected by classical diffusion model (MD/MH = 1.42). This may indicate that the protonic conduction in the films obeys nonclassical hopping transport where the dissociation of OH bond is the rate determining step for proton hopping. Hence, it is concluded that a-AlPnOx films predominantly transport the proton even under nonhumidified atmosphere. Fig 4 Time decay curve of the resistance of Al1P1 () and Al1P2 (X) films at 270C as switching the gas flow from 1%-H2/Ar to 1%-D2/Ar. Local environment structure of the aluminum atoms Al-rich films reveal only a strong peak at 1565.5 eV assigned to tetrahedrally coordinated AlO4. P-rich films show the peak at 1565.5 eV with a shoulder at 1567 eV and a clear peak at around 1572 eV, indicating that AlO4 and AlO6 coexist inside these films. The intensity of peaks at 1567 and 1572 eV in Al1P3 is larger than that in Al1P2, indicating that the molar ratio of AlO6 to AlO4 increases with the increasing Al content. The phosphate groups in a-AlPnOx mainly take the form of metaphosphate in every composition, but the Al-rich films include phosphate group belonging to POAl chain built up by alternative linkage of PO4 and AlO4 tetrahedra via vertex sharing. Al K-edge XAS indicates that most of the Al atoms in these films take tetrahedral configuration. The concentration of octahedral AlO6 coordinates clearly increases with the increasing concentration of P, suggesting that the local structure of aAlPnOx films tends to be close to the aluminum metaphosphate Al(PO3)3 where Al3+ ions coordinate to the oxygen of metaphosphate chain in octahedral configuration.

Fig. 5. Al K-edge XAS of a-AlPnOx films of 100 nm thickness. (a) AlPO4 powder, (b) Y-zeolite powder, (c) Al3P2, (d) Al1P1, (e) Al1P2, (f) Al1P3, and (g) -Al2O3 powder.

Fig. 6. FTIR spectra of aAlPnOx films and of references of aluminum orthophosphate AlPO4 and aluminum pyrophosphate Al(PO3)3.

Ref. 7. Proton conductors of cerium pyrophosphate for intermediate temperature fuel cell Electrochem. Acta 56, 2011, 6654.
Abstract The crystal structure & proton conductivity of cerium pyrophosphate are investigated for potential electrolyte applications in intermediate temperature fuel cell. The CeP2O7 thin plate sintered at 450 C exhibits superior proton conductivity (3.0 102 S cm1 at 180 C) under humidified conditions . Doping with 10 mol% Mg on the Ce site of CeP2O7, the conductivity was raised to 4.0 102 S cm1 at 200 C. The Mg doping also shifts and widens its temperature window for electrolyte applications. Ce0.9Mg0.1P2O7 is considered a more appropriate composition, with conductivity >102 S cm1 between 160 and 280 C. A hydrogenair fuel cell with Ce0.9Mg0.1P2O7 electrolyte generates electricity up to 122 mA cm2 at 0.33 V using 50% H2 at 240 C (Peak power 40 mWcm-1 at 240 OC).

Fig. (a) Arrhenius plots for proton conductivity of sintered CeP2O7 at PH2O=0.114atm. (b) Arrhenius plots for proton conductivity of 450 C-sintered CeP2O7 at PH2O=0.0040.114atm.

Fig. (a) Proton conductivities of the Ce1xMgxP2O7, x = 0.0, 0.1, 0.2, and 0.3, calcined at 300 C and sintered at 450 C (measured in moist air of PH2O=0.114atm). (b) XRD pattern of the most conductive sample Ce0.9Mg0.1P2O7.

Fig.- Performance of the FC based on Ce0.9Mg0.1P2O7 The peak power values are much lower than that of Sn0.9In0.1P2O7 cell, 264 mW cm2 at 250 C and Sn0.95Al0.05P2O7 and SEPS polymer cell, 163 mW cm2 at 225 C (Hibino et.al.). The low power of this FC could result from the less-ideal cell potential. The origin of low cell potential can be the porosity of sintered Ce0.9Mg0.1P2O7 disk (12%), which led to certain pores connecting the fuel and air sides, such that hydrogen diffused through the electrolyte membrane.

Ref. 8. Proton conductivity of CeP2O7 for intermediate temperature fuel cells Solid State Ionics 179, 2008, 1138.
Abstract A single phase new proton conductive electrolyte of CeP2O7 was synthesized. The conductivity of CeP2O7 increased with the increasing of temperature and kept above 10 2 S/cm in the temperature range of 150250 C. A single H2/O2 fuel cell with CeP2O7 as electrolyte membrane exhibited a reasonable power density of 25 mW/cm2 at 200 C, showing potential applications.

Fig. 1. XRD patterns of CeP2O7 calcined at different temperature.

Fig. 2. TGA-DSC of CeP2O7 in air atmosphere (phase transformation at 455 C from CeP2O7 to Ce(PO3)4

Fig. 3. Electrical conductivities of CeP2O7 The OCV value is lower than the theoretical value calculated from Nernst's equation, which may be caused by the mechanical leakage of gas through the electrolyte and the electron hole in this material. The proton transfer number calculated base on the EMF value to theoretical value of cell b was about 0.870.9, indicated that CeP2O7 is mainly a proton conductor.

Fig. 4. EMF value of (a) H2/O2 concentration cell (b) H2 (1 atm)/H2 (0.1 atm) concentration cell.

Fig. 5. H2/O2 fuel cell performance with CeP2O7 (1.2 mm) electrolyte.

Ref. 9. Proton Conduction in In3+-Doped SnP2O7 at Intermediate Temperatures- J. Electrochem. Soc. 153(8), (2006), A1604.
Abstract- SnP2O7 -based proton conductors were characterized. Undoped SnP2O7 showed conductivities 102 S cm1 in temperature range of 75300C. The proton transport numbers (tH+) of this material at 250C to be 0.970.99 in humidified H2 and 0.890.92 under fuel cell conditions. Partial substitution of In3+ for Sn4+ led to increase in proton conductivity from 5.56 x102 to 1.95 x101 S cm1 at 250C. FTIR and TPD measurements revealed that the effects of doping on the proton conductivity could be attributed to an increase in the proton concentration in the bulk Sn1xInxP2O7. The deficiency of P2O7 ions in the Sn1xInxP2O7 bulk decreased the proton conductivity by several orders of magnitude. The mechanism of proton incorporation and conduction is examined and discussed in detail. Proton conduction in undoped SnP2O7 -For galvanic cell H2 (1 atm),Pt/C|SnP2O7 |Pt/C, H2 + Ar(0.1 atm) the tH+ of SnP2O7 was in the range 0.970.99, indicating that this material is substantially a pure proton conductor in H2 atmospheres. Cell H2,Pt/C|SnP2O7 |Pt/C, (air) showed EMF values deviating from the theoretical values (tH+ = 0.890.92), although the EMF values were as high as 920 mV. This may be due to electron holes in the SnP2O7 bulk. However, mechanical leakage of gas through the electrolyte may also be responsible for the lower EMF values compared to the theoretical values, because the EMF value was affected by the thickness of the electrolyte used. At pH2O = 0.0008 atm, conductivity increases with increasing pO2 under oxidizing conditions indicating that this material shows mixed electron-hole and proton conduction. This tendency decreased with increasing pH2O & at pH2O = 0.12 atm, the conductivity was almost independent of pO2, indicating the disappearance of electron holes from the bulk. It appears that there is interaction between water vapor and electron holes. The following equilibrium has been proposed as a mechanism of proton incorporation in perovskite oxides such as SrCe0.95Yb0.05O3 H2O + 2h 2H + 1/2 O2 H2O + 2 Oox + 2h 2OHo + 1/2 O2 (3) There is almost no variation of the conductivity with pO2 under reducing conditions pO2 = 10201027 atm was observed, excluding the reduction of Sn4+ and P5+ to lower valences.

Fig.3 pO2 dependence of conductivity of SnP2O7 at 250C and pH2O of 0.0008, 0.0062, and 0.12 atm.

Fig.4 Isotope effect conductivity of SnP2O7.

on

Isotope effect on conductivity of SnP2O7- SnP2O7 yielded a 1.061.44 times higher conductivity and a lower activation energy of 0.03 eV for H2O-containing atmospheres than for D2O-containing atmospheres. This result can be interpreted by a nonclassic H/D isotope effect. When the dissociation of the O-H bond is a rate-determining step for proton conduction, the activation energy for D+ is higher than that for H+ by a difference in zero point energy of 0.05 eV, which is near the difference in activation energy shown above. It is thus proposed that protons migrate via dissociation of O-H bonds hopping mechanism.

XRD patterns of Sn1xInxP2O7

Effect of In3+doping on proton conduction- maximum conductivity at In3+ = 10 mol % which corresponded well with the substitution limit for In3+ from the XRD measurements. IR spectra The absorbance ratios of Sn0.9In0.1P2O7 to SnP2O7 for (OH, 1655 cm1) and (OH, 3410 cm1 ) were 5.2 and 5.7 respectively, which are comparable to their conductivity ratio of 4.6 at 50C Fig. 6. This suggest that the absorption bands are mainly attributable to protons incorporated in the bulk. It is also likely that the protons interact with the lattice oxide ions to form hydrogen bonds.

TPD spectra

The proton concentrations in Sn0.9In0.1P2O7 and SnP2O7 were determined by assuming that all the evolved water vapor and H2 can be attributed to the incorporated protons. The resulting proton concentration values were 10.4 and 2.5 mol % for Sn0.9In0.1P2O7 and SnP2O7, respectively. These values were in good agreement with the proton concentration predicted from the In3+ content of 10 mol %. It thus appears that the protons were fully introduced as point defects by the substitution of In3+ for Sn4+.

Fig. 6 Temperature dependence of conductivity of Sn1xInxP2O7 in unhumidified air pH2O = 0.0075 atm.

Another possible reaction is the following interaction between water vapor and an oxygen vacancy H2O(g) + V 2Hi + Oxo H2O(g) + Oxo + V 2OHo (4) The degree of Reaction (4) cannot be entirely demonstrated at this stage. However, when the proton conductivity of Sn0.9In0.1P2O7 was measured at different pH2O values at 250C, it slightly increased with increasing pH2O. This result is associated with the process of proton incorporation through Reaction 4 rather than 3, because the order of mobility is oxygen vacancy < proton < electron hole. Therefore, Reaction 4 as well as Reaction 3 are possible mechanisms of proton incorporation. Effects of P2O7 deficiency in Sn0.9In0.1P2O7 on proton conduction IR spectra The Sn0.9In0.1(P2O7)1-y IR spectrum showed peaks at almost same wave numbers as those observed for Sn0.9In0.1P2O7. The absorbance ratios of Sn0.9In0.1P2O7 to Sn0.9In0.1(P2O7)1-y were 1.4 and 1.5 for (OH) and (OH), respectively, which are much smaller than their conductivity ratio shown in Fig. 10. A similar behavior was obtained for the TPD spectra. The proton concentration in Sn0.9In0.1(P2O7)0.85 was estimated to be about 8.1 mol % per unit, which is not significantly different from the value of 10.4 mol % for Sn0.9In0.1P2O7. Therefore, the large difference in proton conductivity between Sn0.9In0.1P2O7 and Sn0.9In0.1(P2O7)0.85 can be attributed to the difference in proton mobilities rather than proton concentrations between them. A possible speculation on the proton mobility of Sn0.9In0.1(P2O7)0.85 is that the P2O7 deficiency causes a partial disconnection of the P2O7 network for proton conduction, resulting in a large energy barrier for proton jumps between sites.

Fig.10 Temperature dependence of conductivity of Sn0.9In0.1(P2O7)1-y in unhumidified air pH2O = 0.0075 atm. The proton conductivity of Sn0.9In0.1(P2O7)0.85 was about 2 orders of magnitude lower than that of Sn0.9In0.1P2O7, indicating that the conductivity is strongly affected by the number of P2O74 ions in the lattice.

TPD spectra

Ref. 10. Intermediate-Temperature Proton Conduction in Al3+-Doped SnP2O7 J. Electro. Soc. 154(12), 2007, B1265.
Abstract- Al3+-doped SnP2O7 proton conductors were prepared by controlling the initial composition of the reactants SnO2, AlOH3, and H3PO4. Sn1xAlxPyOz with y<2 displayed conductivities approximately two orders of magnitude lower than Sn1xAlxP2O7, while those of Sn1xAlxPyOz with y 2 exhibited conductivities at a maximum of 1.99 times higher. Partial substitution of Al3+ for Sn4+ in Sn1xAlxP2O7 led to an increase in the conductivity up until x = 0.05. As a result, the conductivity reached 0.045 S cm1 at 100C, 0.15 S cm1 at 200C, and 0.19 S cm1 at 300C when the x and y values were 0.05 and 2, respectively. A hydrogen concentration cell with this material demonstrated that the ionic transport number was 1, and a fuel cell using this material demonstrated that the dc conductivity was comparable to the ac conductivity. Sn0.95Al0.05PyOz contains SnO2 that is formed on the top side of the product due to vaporization of H3PO4. From SEM micrographs of all the samples, the mean size of the primary particles was ~300 nm. However, a highly hygroscopic layer covered the surface of the particles of the sample with y=2.6. From XRD data, it is expected that this layer has an amorphous structure, probably with a composition of PmOn. The small increase in conductivity, with increasing y from 2.0 to 2.6 , is due to additional transport pathways for protons formed by the PmOn layer. Therefore, it is difficult for such a layer to show stable conductivity. From the above results, the optimal value for y in Sn1xAlxPyOz was determined to be 2.

Fig. Conductivity variation of Sn0.95Al0.05PyOz at 200C in dry air. The ordinate is the ratio of measured conductivity to initial value.

Fig. 2. Temperature dependence of conductivity of Sn0.95Al0.05PyOz in dry air.

Effect of using different raw materials as Al sources on the electrical conductivity-

Fig. XRD pattern of Sn0.95Al0.05PxOy

Fig. Variation of conductivity (in dry air) with temperature & XRD Sn0.95Al0.05P2O7 prepared using Al(OH)3, Al2O3, Al(NO3)3, and AlCl3 as Al sources.

pattern

of

Conductivity of the sample prepared from Al2O3 were lower than those of SnP2O7, indicating that the use of Al2O3 had a negative effect on the conductivity. Also, there is a possibility that NO3- and Cl- ions remained in the samples prepared with Al(NO3)3 and AlCl3, causing a deterioration in the conductivity. The sample prepared with Al2O3 showed diffraction peaks assigned to AlPO4, besides SnP2O7 and SnO2 which led to decrease in conductivity as reported by Matsuda.

Effect of Al3+ substitution on the conductivity of Sn1xAlxP2O7In XRD patterns of Sn1xAlxP2O7 with x values of 0.10 and 0.15, some peaks of AlPO4 were present. Therefore, conductivity of these samples was less.

The ionic transport numbers was in the range of 0.97 - 0.99, which was similar to the electrical conduction behavior of Sn0.9In0.1P2O7. This suggests that SnO2 contained in Sn0.95Al0.05P2O7 does not affect ion conduction. The migration of specific ions that contribute to the high ionic transport number of Sn0.95Al0.05P2O7 cannot be identified with this measurement, but it can be assumed that protons are major charge carriers in this material because the ionic size of oxide anions or other cations is too large to achieve high conductivity under the present conditions. The values for dc conductivity were approximately in agreement with the values for ac conductivity at all temperatures tested. The slight difference in conductivity between the dc and ac measurements is probably due to the contact resistance between the electrode and the current collector. The close agreement between the ac and dc conductivity values means that the high proton conductivity of Sn0.95Al0.05P2O7 is also demonstrated under fuel cell conditions.

Ref. 11. Intermediate temperature ionic conduction in Sn1xGaxP2O7 - J. Power Sources 195, (2010), 5596.
Abstract- A novel series of samples Sn1xGaxP2O7 (x = 0.00, 0.01, 0.03, 0.06, 0.09, 0.12, 0.15) are synthesized by solid state reaction. XRD patterns indicate that the samples of x = 0.00 0.09 exhibit a single cubic phase structure, and the doping limit of Ga3+ in Sn1xGaxP2O7 is x = 0.09. The protonic and oxide-ionic conduction in Sn1xGaxP2O7 are investigated using some electrochemical methods at intermediate temperatures (323523 K). The samples exhibit appreciable protonic conduction in hydrogen atmosphere, and a mixed conduction of oxideion and electron hole in dry oxygen-containing atmosphere. The highest conductivities are observed for the sample of x = 0.09 to be 4.6 102 S cm1 in wet H2 and 2.9 102 S cm1 in dry air at 448 K, respectively. The H2/air fuel cell using x = 0.09 as electrolyte (thickness: 1.45 mm) generates a maximum power density of 19.2 mW cm2 at 423 K and 22.1 mW cm2 at 448 K, respectively. The higher conductivities of the Ga3+ doped samples are resulted from the higher oxygen vacancy concentration. While partially substituting Sn4+ with Ga3+ ions, charge compensation is achieved through the formation of oxygen vacancies as indicated by the Eq. (1). Mixed conduction of oxygen ion and electron hole is resulted from Eq. (2) in dry oxygen-containing atmosphere,

Fig. Temperature dependence of electrical conductivity of SnP2O7 and Sn1xGaxP2O7 (x = 0.01 0.12) in (a) dry air and (b) wet H2.

When water vapor is introduced, as shown in Eq. (3) and (4), the conduction of electron hole and oxygen vacancy decreases, at the same time, the protonic conduction appears. In wet H2, the protonic conduction may be further improved according to Eq. (3), (4) and (5), resulting in the prevailing protonic conduction in hydrogen atmosphere.

The influence of the doping level of Ga3+ at Sn4+ sites on the conductivities may be attributed to the effective concentration of oxygen vacancy in the samples. On one hand, the total concentration of oxygen vacancy Vo mainly increases with the increasing of the doping level of Ga3+. On the other hand, the concentrations of point defect pairs, GaSnV0, GaSnV0GaSn, and GaSnOH0, which resulted from the coulombic attraction among the point defects with opposite charges, also increase at the same time. Considering the opposite two factors above, the effective concentrations of oxygen vacancy may reach its largest value at x = 0.09. Moreover, the impurity phase of SnO2 in the samples for x > 0.09 may be also responsible for the decrease in the conductivities. Therefore, these factors result in the highest conductivity at x = 0.09.

Ref. 12. Ionic conduction in Sn1xScxP2O7 for intermediate temperature fuel cells J. Power Sources 196, 2011, 683.
A novel series of samples Sn1xScxP2O7 (x = 0.03, 0.06, 0.09, 0.12, 0.15) were prepared. The doping limit of Sc3+ in SnP2O7 was at least x = 0.09. The conductivities increased with various Sc3+ doping levels in the order: (x = 0.12) < (x = 0.03) < (x = 0.09) < (x = 0.06). The highest conductivity was observed to be 2.76 102 S cm1 for the sample of x = 0.06 under wet H2 atmosphere at 473 K. The ionic transport numbers (tion = 0.950.99) were close to unity, and the relatively low electronic conductivity in wet hydrogen atmosphere. The maximum proton conductivity (2.24 102 S cm1) at 473 K is higher than that in BaCe0.85Y0.15O3 (1.04 102 S cm1) under wet H2 atmosphere at 873 K. The H2/air fuel cells using Sn1xScxP2O7 (x = 0.03, 0.06, 0.09) as electrolytes (thickness: 1.7 mm) generated the maximum power densities of 11.16 mW cm2 for x = 0.03, 25.02 mW cm2for x = 0.06 and 14.34 mW cm2 for x = 0.09 at 423 K, respectively. Sn1xScxP2O7 may be a promising solid electrolyte system for intermediate temperature fuel cells.

Fig. XRD patterns of the Sn1xScxP2O7 (x = 0.03 0.12) samples.

Fig. Temperature dependence of electrical conductivity of Sn1xScxP2O7 in wet H2 atmosphere at 323523 K.

Effect of Sc3+ doping level on the conductivities- The influence of Sc3+ doping on the conductivities may be attributed to the effective concentrtion of oxygen vacancy and impurity phase of Sc(PO3)3. The higher Sc3+ doping level resulted in higher oxygen vacancy concentration, nevertheless, the concentrations of point defect pairs, ScSnVO, ScSnVOScSn, and ScOHO also increase at the same time. Considering above two opposite factors, the effective concentrations of oxygen vacancy may reach its maximum value at x = 0.06. As a result, the highest conductivity 2.76 102 S cm1 was observed for the sample of x = 0.06 under a wet H2atmosphere at 473 K. The conductivities of samples of x > 0.06 decreased with the increasing doping levels due to the formation of the secondary Sc(PO 3 ) 3 phase and lower effective oxygen vacancy concentrations. The effect of the formation of the pseudo-cubic 333 superlattice on the conductivities is still unclear in present stage.

Fig. 5. Partial conductivities of charged species in wet H2 atmosphere.

Fig. 4. EMFs of the H2 concentration cell: H2, Pt |Sn1xScxP2O7 (x = 0.06)| Pt, H2Ar (pH2=1.01410Pa) and water vapor concentration cell: H2 (pH2O=2.34310Pa), Pt| Sn1xScxP2O7 (x = 0.06)| Pt, H2(pH2O=1.23410Pa).

Fig. 6. IVP curves for a hydrogen/air fuel cell using Sn1xScxP2O7 (x = 0.03-0.09) as electrolytes at 423 K. Electrolyte thickness: 1.7 mm.

Ref. 13. Synthesis and characterization of dense SnP2O7SnO2 composite ceramics as intermediate-temperature proton conductors- J.Mater. Chem. 21, (2011), 663.
Abstract: Dense SnP2O7-SnO2 composite ceramics were prepared by reacting a porous SnO2 substrate with an 85% H3PO4 solution at elevated temperatures. Comparison of the observed EMF with the theoretical value in two gas concentration cells demonstrated that this composite ceramic is a pure ion conductor, wherein the predominant ion species are protons. FT-IR and proton magic angle spinning NMR analyses revealed that the protons interacted with lattice oxide ions in the SnP2O7 layer to form hydrogen bonds. An H/D isotope effect suggested that proton conduction in this composite ceramic was based on a proton-hopping mechanism. The proton conductivity is 102 S cm1 in the temperature range of 250600 C.

The electrical conductivity as well as the growth of the SnP2O7 layer increased with increasing carbon content. Unfortunately, a crack-free SnO2 substrate with carbon content above 8 wt% could not be prepared. Thus, the optimal carbon content was concluded to be 8 wt%.

Fig. 1 Influence of temperature on the growth of an SnP2O7 layer, (a) XRD patterns at the ceramic surface; (b) XRD patterns in bulk; (c) relationship between the electrical conductivity of the ceramic at 400 C and the intensity of the main peak (200) for SnP2O7 on ceramic surface and bulk.

Fig. 2 Influence of carbon content in mixtures of carbon and SnO2 powder during the preparation of SnO2 substrate on the growth of SnP2O7 layer. (a) XRD patterns at the ceramic surface; (b) XRD patterns in bulk; (c) relationship between the electrical conductivity of the ceramic at 400 C and the intensity of the main peak (200) for SnP2O7 ceramic surface and bulk.

Denseness and robustness of SnP2O7SnO2 composite ceramic

(Fig. 4)

(Fig. 5)

Fig. 3 Temperature dependence of the electrical conductivity of the SnP2O7SnO2 composite ceramic with various thicknesses. For comparison, data for the SnO2 substrate are also included.

Cross-sectional SEM and EDX images of the SnO2 substrate (Fig. 4) & SnP2O7 SnO2 composite ceramic (Fig 5). (a) SEM; (b) Sn element mapping; (c) P element mapping.

Fig. 6 Hydrogen permeation properties of the SnP2O7SnO2 composite ceramic at various temperatures. For comparison, data for the SnO2 substrate are also included.

Proton conduction in SnP2O7SnO2 composite ceramic 10 vol% H2, AuSnP2O7SnO2Au, 1 vol% H2 E = RT/2F ln(PH2(II)/PH2(I)) H2O concentration in each chamber 3 vol%.

3 vol% H2O, AuSnP2O7SnO2Au, 0.6 vol% H2O E = RT/2F ln(PH2O(II)/PH2O(I))

H2 concentration in each chamber 10 vol%.

Fig. 7 EMF values of (a) hydrogen concentration cell and (b) H2O vapor concentration cell using the SnP2O7SnO2 composite ceramic as an electrolyte at various temperatures.

Fig. 8 Temperature dependence of the conductivity of the SnP2O7SnO2 composite ceramic in various atmospheres. Atmospheric gases used were wet air (PH2O = PD2O = 0.026 atm) and dry air (PH2O = ca. 0.001 atm).

Proton environment in SnP2O7SnO2 composite ceramic

Fig. 9 TPD spectrum of the SnP2O7SnO2 composite ceramic powder.

Fig. 10 FTIR spectrum of the SnP2O7SnO2 composite ceramic powder.

Fig. 11 1H MAS NMR spectrum of the SnP2O7SnO2 composite ceramic powder.

The protons are present in the SnP2O7 layer. The inserted protons can easily jump between adjacent oxide ions by a series of breaking and making of OH bonds.

Ref. 14. Ionic conduction in undoped SnP2O7 at intermediate temperatures- Solid State Ionics 181, (2010), 1521.
Abstract- SnP2O7 is prepared with various initial molar ratios of phosphorus vs. metal ions, Pini/Sn, and different temperatures from 573 to 923 K. The preparation conditions are optimized giving consideration to the influence of H3PO4 concentration, Pini/Sn molar ratio and heattreating temperature. The ionic conduction behaviors indicate that the samples obtained from 85% H3PO4 and SnO2 nanopowders with Pini/Sn 2.4 are a single cubic phase, and that in wet hydrogen atmosphere, the samples are almost pure ionic conductors, the ionic conduction is contributed mainly by proton and partially by oxide ion. An ionic conductivity of 2.17 10 2 S cm 1 is achieved for the sample prepared from Pini/Sn = 2.8 under wet hydrogen atmosphere at 448 K. Single cubic phase of samples is obtained when Pini/Sn = 2.4, 2.6, 2.8 and 3.0, corresponding to Pfin/Sn = 2.02, 2.16, 2.28 and 2.50, respectively. When Pfin/Sn > 2, excess P as amorphous PmOn existed in the grain boundary of SnP2O7 crystal. When Pini/Sn = 2.0, except cubic structure of SnP2O7 as a main phase, a SnO2 impurity phase also is observed. Therefore, in order to obtain a single cubic phase of SnP2O7, Pini/Sn should be controlled to be 2.4 or more. However, moisture absorbability of the samples increases with increasing Pini/Sn molar ratio. Excess P as amorphous PmOn exists in the grain boundary of SnP2O7 crystal. This intergranular PmOn is considered to produce more channels for proton conduction and serves as superionic highways resulting in the increase of the conductivities with increasing Pini/Sn molar ratio. Under the other same conditions, the conductivities increase with the order: (dry air) < (wet air) < (wet H2); (heattreated at 873 K) < (heat-treated at 773 K). The highest conductivity is observed to be 2.17 10 2 S cm 1 for the sample prepared from Pini/Sn = 2.8 under wet H2 atmosphere at 448 K. However, the stability of sample prepared from Pini/Sn of 3.0 became poor, which may be attributed to its higher moisture absorbability though its conductivities are higher.

Fig. Temperature dependence of conductivity of SnP2O7 prepared from different Pini/Sn molar ratios and heat-treating temperatures.

Fig. Transport numbers of the sample prepared from Pini/Sn of 2.4 in wet hydrogen atmosphere.

Fig. Partial conductivities of charged species in the sample prepared from Pini/Sn of 2.4 (in wet hydrogen atmosphere).

Fig. Isotope effect on conductivity of SnP2O7 prepared from different Pini/Sn molar ratios in argon saturated with H2O or D2O vapor at 298 K.

The total conductivity is much higher than that of electronic conductivity, the protonic conductivity dominates ionic conduction, whereas the oxide-ionic conductivity reaches a certain extent in wet hydrogen atmosphere over the entire range of temperatures tested. The oxide-ionic conduction is relevant to oxygen vacancies (VO ) in the sample. One may speculate about the reason for the presence of oxygen vacancies (VO ). H3PO4 and SnO2 as reactants for preparing the sample may be assumed to contain trace of metal ions (Mn+) with lower ion valence than Sn4+ ions. While the sample is prepared, oxygen vacancies (VO ) are formed due to substitution of a small amount of Mn+ on the Sn4+-site, accordingly, resulting in the oxide-ionic conduction to a certain extent.

Ref. 15. Intermediate temperature stable proton conductors based upon SnP2O7, including additional H3PO4 -J. Mater. Chem. 20, (2010), 7827.
Abstract-In order to examine the influence of phospate impurities upon the conduction properties of SnP2O7, SnP2O7-H3PO4 composites were synthesised through different methods with varying starting P:Sn molar ratios. It was found that cubic SnP2O7 is the main crystalline phase and amorphous phases were observed when starting P:Sn ratios exceeded 2:1. Solid State 31P NMR confirmed residual phosphoric acid in samples with high starting P:Sn ratios whilst impedance spectroscopy indicated these to be good proton conductors. The highest conductivity observed was 3.5 102 S/cm at 300 C in air for samples with high starting P:Sn ratios and calcined at higher temperatures. The conductivity stability of the composites was found to be promising.

Table Synthetic history and nomenclatures of all samples


P:Sn molar ratio 2:1 2:1 2.6:1 2.8:1 4:1 2:1 Fig. 4 TEM images of SP02 (a) and (b); SP1 (c); SP2 (d); SP3 (e) and SP4 (f). The face distances marked of 3.22 , 3.51 , 3.98 , 4.59 , 3.55 and 2.81 correspond to (211), (210), (200), (111), (210) and (220) planes respectively for cubic SnP2O7. Raw Material SnO2.nH2O, H3PO4 SnO2.nH2O, H3PO4 SnO2.nH2O, H3PO4 SnO2.nH2O, H3PO4 SnO2.nH2O, H3PO4 SnCl4.5H2O, (NH4)2HPO4 Synthetic history 300 OC, 2.5 h 300 OC, 2.5 h 650 OC, 2.5 h 300 OC, 2.5 h 300 OC, 2.5 h 300 OC, 2.5 h 650 OC, 2.5 h 300 OC, 2.5 h 650 OC, 2.5 h Nomenclatur e SP01 SP02 SP1 SP2 SP3 SP4

The most significant distinctions between different samples are the presence of an amorphous layer when the starting P:Sn ratio exceeds 2:1. The amorphous layer should be rich in phosphorus as the thickness of layer increased from samples SP1 to SP3. Particles with relatively clean edges and grain boundaries were observed for samples with stoichiometric ratio, irrespective of the starting materials. The grid distances marked on Fig. 4 correspond to (211), (210), (200), (111), (210) and (220) planes for cubic SnP2O7.

DP-MAS-NMR spectra show a number of characteristic peaks indicative of free phosphoric acid and tin coordinated groups . Conductivity-The samples with non-stoichiometric ratios, SP1, SP2 and SP3, are highly conductive in both air and wet reducing conditions. A slightly enhanced conductivity in wet conditions implies that these samples are essentially proton conductors. Stoichiometric samples show a much lower conductivity and the conductivity is highly dependent on thermal treatment. A different treatment at 300 and 650 0C can yield a variation in conductivity of almost three magnitudes (sample SP01 and SP02). The samples with poor conductivity are either mainly SnP2O7 (SP02) or even pure cubic SnP2O7(SP4). The slightly higher conductivity in sample SP02 is attributed to the residual acid and the presence of HPO42- groups. The significantly higher conductivity in samples SP01, SP1, SP2 and SP3 is probably due to the presence of considerable amounts of H3PO4 and the SnP2O7 per se is not directly contributing to the high conductivity. This is consistent with the results of our previous report but demonstrates large discrepancies with other groups. This is probably because of the excellent thermal stability and versatility of H3PO4 that is generally difficult to be removed under normal conditions. However, the observed high proton conductivity suggests a new route to develop novel proton conductors since all samples are basically in solid state. The highest conductivity achieved was 3.5 x 10-2 S/cm at 300 0C in air for sample SP3. Conductivity Stability- The conductivity of sample SP1 remains unchanged in both air and wet 5% H2 for more than 50 h, confirming good stability. However, the conductivity of sample SP2 keeps on decreasing over the first 50 h and tends to be stable thereafter which could be due to the slow evaporation of H3PO4 as the initial acid loading in SP2 is relatively high. Another possible explanation is the condensation effect: the slow generation of pyrophosphoric acid could take place at high temperatures and decrease the conductivity to some extent. However, the conductivity of sample SP3 was stable which was probably due to a better thermal stability after the treatment at a higher temperature (650 0C). The SEM images of samples SP1, SP2 and SP3 taken after the conductivity stability testing show no obvious microstructural changes compared to those before the conductivity tests indicating that these samples are quite stable. A TEM picture of sample SP3 after the conductivity stability tests indicates that the core-shell microstructure remains. Fig. Conductivity of SP1, SP2 and SP3 at 250 0C in different atmospheres as a function of time.

Fig. Conductivity versus temperature of SP01; SP1; SP2; SP3 and SP4 in different atmospheres.

Conclusion- The large difference in conductivity between composites and essentially pure cubic SnP2O7 suggested that the conductivity mainly originates from the residual phosphoric acid and cubic SnP2O7 itself is not a good proton conductor.

Ref. 16. Sn0.9In0.1P2O7-Based Organic/Inorganic Composite Membranes: Application to Intermediate-Temperature Fuel Cells- J. Electrochem. Soc. 154(1), 2007, B63.
Abstract- An anhydrous proton conductor, 10 mol % In3+-doped SnP2O7 was composed by 1,8-bis(triethoxysilyl)octane (TES-Oct) and 3(trihydroxysilyl)-1-propanesulfonic acid (THS)Pro-SO3H. The composite membrane with 90 wt % Sn0.9In0.1P2O7 showed high proton conductivities of 0.04 S cm1 or more between 150 and 200C in dry air. The packing of the Sn0.9In0.1P2O7 particles in the matrix was relatively uniform, with no formation of pinholes observed. Fuel cell tests verified that the OCV was maintained at a constant value of ~970 mV regardless of the electrolyte thickness (60200 m), while the Ohmic resistance was decreased to 0.24 cm2 by reducing the electrolyte thickness to 60 m. The peak power densities achieved with dry H2 and air were 109 mW cm2 at 100C, 149 mW cm2 at 150C, and 187 mW cm2 at 200C. Furthermore, fuel cell performance was improved by hotpressing an intermediate layer consisting of Sn0.9In0.1P2O7, Pt/C, TESOct, and (THS)Pro-SO3H between the electrolyte and cathode. The composite membranes with only TES-Oct showed a lower conductivity compared to that in the presence of the (THS)Pro-SO3H component. The observed difference can be ascribed to the -SO3H groups available for proton transfer. The SO3H group likely formed a proton conducting pathway from one Sn0.9In0.1P2O7 cluster to another. The conductivities of all the composite membrane samples decreased with increasing temperature from 200 to 250C, which was irreversible upon heating and cooling. While the former behavior is due to the dehydration of the membrane below 200C, the later behavior is due to the thermal decomposition of components such as PTFE above 200C. Evidence for this reasoning is provided by the TG analysis of the composite membrane and pellet samples.

Measured in dry air

XRD peaks observed for the composite membrane sample were almost identical to those of the pellet sample, suggesting that the polymer components exist as an amorphous phase in the matrix. SEMs of (a) Sn0.9In0.1P2O7 composite membrane and(b) pellet samples. The composite membrane is more compact, suggesting that the polymer functioned as a binder at the interface.

Fig. EMFs of H2 concentration cells with Sn0.9In0.1P2O7 composite membrane and pellet samples, and their proton transport number as a function of temperature in dry atmosphere. It indicates that the polymer does not retard the proton transport number of Sn0.9In0.1P2O7. A decrease in the OCV with decreasing electrolyte thickness was observed for the pellet sample, implying that H2 or O2 crossover through the electrolyte increases with decreasing electrolyte thickness. In contrast, the OCVs for the composite membrane were almost independent of the electrolyte thickness, indicating that both H2 and O2 crossovers through the electrolyte are negligible, which is consistent with the SEM. The lower OCV values for the composite membranes may be due to partial electron-hole conduction in the electrolyte, causing an internal short circuit of the fuel cell. The Ohmic resistances of two samples decreased linearly with decreasing electrolyte thickness. In particular, the composite membrane showed linearity over the thickness range from 60 to 150 m, assuming that the Sn0.9In0.1P2O7 powders were homogeneously distributed at ~10m level in the matrixes. The Ohmic resistance values of the composite membrane samples were always higher than those of the pellet samples at the same electrolyte thickness, reflecting the difference in the proton conductivity for the two samples. However, the composite membrane sample showed a relatively low resistance of 0.24 cm2, while maintaining a high OCV of 986 mV, which could not be achieved for pellet sample. At all of the tested temperatures, OCVs above 950 mV were obtained, and no limiting current behavior was observed at high current densities. In addition, the current-voltage slopes became lower as the operating temperature increased. The peak power density thus reached 109 mW cm2 at 100C, 149 mW cm2 at 150C, and 187 mW cm2 at 200C. However, the power densities were much lower compared to those expected from the Ohmic resistances as an example, 0.24 cm2 at 200C of the electrolyte, which may be ascribed to the large polarization resistance of the fuel cell.

(a)

(b)

Fig. (a) OCVs and (b) Ohmic resistance of fuel cells with Sn0.9In0.1P2O7 composite membrane and pellet samples at 200C as a function of electrolyte thickness. Both H2 and air were dry.

Dry H2 & air flow rate = 30 mL/min electrolyte thickness = 60 m

Dry H2 & air, flow rate = 30 mL/min electrolyte thickness = 60 m

The overpotentials were always dominated by the cathode in the temperature range of 100200C; the cathodic overpotentials account for 8385% of the overall voltage drops during the cell discharge. The anodic overpotentials as well as the IR drops negligibly affected the voltage drops especially at higher temperatures. The water formed electrochemically did not affect the subsequent cathode reaction because the discharge at high current densities gave rise to no limiting currents. It is likely that the charge-transfer reaction of oxygen reduction at the electrolyte/electrode interface proceeded at a very slow rate. Also, only cathode was in physical contact with the electrolyte, probably resulting in low-density three phase boundaries (TPBs). Thus, it is necessary to optimize the electrolyte/electrode interface.

With intermediate layer no intermediate layer

Dry H2 & air, flow rate = 30 mL/min electrolyte thickness = 60 m

An attempt was made to hot-press an intermediate layer constructed from Sn0.9In0.1P2O7, Pt/C, TES-Oct, and (THS)Pro-SO3H between the electrolyte and cathode. As a result of the optimization of the weight ratio of the Sn0.9In0.1P2O7 electrolyte to the Pt/C catalyst, the cathodic overpotential was shown to be the most improved when Sn0.9In0.1P2O7 :Pt/C = 20:1. The voltage drop was reduced by using the intermediate layer, so that the peak power density increased from 149 to 197 mW cm2 at 150C. Similar effects were obtained at other temperatures, although these were smaller at higher temperatures. One can consider that such performance gains are due to an increase in the area of the three-phase boundary for the cathode reaction. Indeed, the total polarization resistance decreased by 1.3 cm2 at 150C by applying the intermediate layer. One may also expect that the fuel cell performance is further enhanced by improving the microstructure, catalytic activity, and proton conductivity of the intermediate layer.

Ref. 17. Proton Conduction in SnP2O7LaP3O9Composite Electrolytes- Electrochem. Solid-State Lett. 12(2),2009, B11.
Abstract-The properties of proton-conducting composite SnP2O7 electrolytes containing LaP3O9 are reported. These electrolytes demonstrate both improved stability and enhanced proton conductivities when compared to their individual constituents. Two different methods were used to prepare samples: sintering or warm-pressing. A conductivity of 1.7104 S cm1 was obtained for warm-pressed SnP2O7LaP3O9 (Sn:La, 82:18) at 350C. Under FC conditions the composite with composition Sn:La, 82:18 exhibited the highest OCV of 0.983 V at 350C. The OCV values confirmed that the electrolytes were predominantly ionically-conducting.

At high La content the structure appears to contain an amorphous component. For the Sn:La 82:18, well-defined crystallites of size 200 nm were observed. EDX showed that the crystallites were mainly SnP2O7. In contrast, SEM image of warm-pressed SnP2O7LaP3O9, Sn:La = 82:18 (Fig. d) shows a nonporous structure with a component that has an amorphous morphology. This may be due to the lower preparation temperature that does not allow the full crystallization of SnP2O7.

Fig. XRD patterns of SnP2O7LaP3O9, (Sn:La = 97:3, 82:18, 70:30, 50:50) composite electrolytes and SnP2O7 and LaP3O9. The arrowed reflections are tentatively associated with LaPO4.

Fig. Cross-sectional SEM images of SnP2O7LaP3O9 composite electrolytes (Sn:La = 50:50 (a), 70:30 (b), and 18:12 (c), and warm-pressed electrolyte with Sn:La = 82:12(d).

The impedance plot shows the typical large electrode dispersion which is attributable to the partially blocking nature of the Pt-paste electrode in a hydrogen containing environment.

Fig. Temperature dependencies of ac bulk conductivities for sintered SnP2O7LaP3O9

Fig. Time dependency of the conductivity of SnP2O7LaP3O9 (Sn:La = 82:18) prepared by warm-pressing at 300C.

The measured OCVs are all lower than the theoretical OCV of 1.18 V vs a standard hydrogen electrode. However, the OCV values are all higher than 0.9 V below 450C, while the cell with electrolyte composition Sn:La = 82:18 exhibited the highest OCV, 0.983 V, at 350C. This confirms that the materials are predominantly ionic conductors. The difference between the theoretical and measured OCVs is likely to be due to some H2 crossover because of a small amount of residual porosity in the sample, or possibly to some electronic conductivity. DC polarization measurements on other rare earth phosphate proton conductors did not indicate, however, a significant electronic contribution to the conductivity. These results show that SnP2O7LaP3O9 composite electrolytes can display promising proton conductivities, where the conductivity/temperature stability can be favorably manipulated by rare-earth additions.

The conductivity decreases with increasing LaP3O9 content but the temperature stability range is significantly extended. The best combination of temperature/conductivity and stability was obtained for a composition Sn:La = 82:18. For the warm-pressed samples, a lower La content resulted in a porous microstructure with a decreased stability under humidified environments. The increased conductivity in humidified, 4% hydrogen supports the presence of proton conductivity. At temperatures below 350C, the warm-pressed sample with Sn:La com position of 82:18 showed significantly higher conductivity, 1.7 x10-4 S cm-1, than the corresponding sintered pellet at that temperature. This higher conductivity might be due to residual phosphoric acid that was not removed by the warm pressing, which would be expected to lead to a rapid conductivity decline when held for a prolonged period at elevated temperatures. To verify this, the sample was held at 300C. Instead, the conductivity remained steady for over 50 h, after an initial drop of less than 5%. This increased stability at 300C compares favorably to the rapid decline of conductivity above 250C for the tin metaphosphates doped with In or Al.

Fig. OCVs of fuel cell with sintered SnP2O7LaP3O9 composite electrolytes. The dashed line indicates the theoretical OCV of a humidified 4% hydrogenair fuel cell.

Superprotonic KH(PO3H)SiO2 composite electrolyte for intermediate temperature fuel cells- Journal of Power Sources 194 (2009) 843846 Novel thin film composite electrolyte membranes, prepared by dispersion of nano-sized SiO2 particles in the solid acid compound KH(PO3H), can be operated under both oxidizing and reducing conditions. Long-term stable proton conductivity is observed at 140 C, i.e., slightly above the superprotonic phase transition temperature of KH(PO3H), under conditions of relatively lowhumidification (pH2O0.02 atm). Unlike sulphate and selenate solid acid electrolytes, KH(PO3H) can be operated in both oxidizing and reducing atmospheres. Slight humidification of the gases in liquid water at room temperature, with an equivalent pH2O of 0.02 atm, is sufficient to prevent impairment of the proton conductivity due to dehydration. Dispersion with nano-sized SiO2 particles leads to improved mechanical properties. The dispersionstrengthened composite electrolyte can be easily made into a thin film in the m range by dip-coating from a suspension. Most important, the elevated temperature of operation and low humidification requirement of the thin film solid acid electrolyte membranes will lead to significant system simplifications in comparison with polymer electrolyte fuel cells. Powders of KH(PO3H) were prepared by slow evaporation of an aqueous solution obtained by mixing of H3PO3 and KOH in molar ratio 1:1. The powders were dried in an oven at ~105 .C for 20 h, ground in an agate mortar and stored in a desiccator due to hygroscopicity of the pure salt. Powders of KH(PO3H).SiO2 composites were prepared by dispersing of SiO2 powder in an aqueous solution of KH(PO3H), followed by drying at ~105 .C for 20 h.
Fig- Dependence of proton conductivity on temperature for (a) pure KH(PO3H) (first heating scan under dry N2), and (b) KH(PO3H)SiO2 composites (air; pH2O 0.02 atm) at different mass fractions, , of SiO2 (particle size 14 nm) in the samples. Also shown in (a) are data for pure KH(PO3H) from differential thermal analysis (DTA) under dry nitrogen. Solid and dashed lines are a guide to the eye.

You might also like