Journal of Materials Science & Technology  2020 , 37 (0): 128-134 https://doi.org/10.1016/j.jmst.2019.06.022

Research Article

(NH4)3PW12O40-H3PO4 composites as efficient proton conductors at intermediate temperatures

Xiaoxiang Xuabcd*, Shunhang Weiad

a Yongxing Special Materials Technology Co., Ltd, Huzhou 313005, China
b Clinical and Central Lab, Putuo People’s Hospital, Tongji University, Shanghai 200060, China;
c School of Engineering, Huzhou University, Huzhou 313000, China
d School of Chemical Science and Engineering, Tongji University, Shanghai 200092, China

Corresponding authors:   *Corresponding author at: Yongxing Special Materials Technology Co., Ltd,Huzhou 313005, China.E-mail address: tongjixxxu@163.com (X. Xu).

Received: 2019-05-29

Revised:  2019-06-17

Accepted:  2019-06-26

Online:  2020-01-15

Copyright:  2020 Editorial board of Journal of Materials Science & Technology Copyright reserved, Editorial board of Journal of Materials Science & Technology

More

Abstract

(NH4)3PW12O40 and (NH4)3PW12O40-H3PO4 composites have been synthesized by precipitation method. Their phase compositions, thermal stability and morphologies have been investigated. The synthesized composites maintain the same structures as pure (NH4)3PW12O40 where phosphoric acid is preserved in residual space of the “spongy crystals” of (NH4)3PW12O40. FT-IR spectra confirm the strong interactions between phosphoric acid and Keggin ions. Pristine (NH4)3PW12O40 shows poor conductivity in air at high temperatures and strongly depends on water molecules for proton transport. The composites exhibit a much higher conductivity compared with pure (NH4)3PW12O40. The highest conductivity achieved is 0.14 S/cm at 170 °C where continuous channels based on phosphoric acid for proton transportation are probably established. Such high conductivity of (NH4)3PW12O40-H3PO4 composites implies promising applications in fuel cells and other electrochemical devices.

Keywords: Proton conductor ; Phosphotungstic acid ; Intermediate temperature ; Keggion ions

0

PDF (2484KB) Metadata Metrics Related articles

Cite this article Export EndNote Ris Bibtex

Xiaoxiang Xu, Shunhang Wei. (NH4)3PW12O40-H3PO4 composites as efficient proton conductors at intermediate temperatures[J]. Journal of Materials Science & Technology, 2020, 37(0): 128-134 https://doi.org/10.1016/j.jmst.2019.06.022

1. Introduction

Heteropolyacids (HPAs) have been recognized as the most proton conductive inorganic solids at near-ambient temperatures [[1],[2]]. Their proton conductivity, such as those from dodecatungstophosphoric acid (H3PW12O40·29H2O) and dodemolybdphosphoric acid (H3PMo12O40·29H2O) can approach > 0.1 S/cm at room temperature and under proper humidity [3]. This high conductivity is mainly attributed to the facile transportation of oxonium (H3O+) or dioxonium cations (H5O2+) in the micro-channels of the crystal structure. One of the major problems that hinder the applications of these promising materials is the dehydration and concomitant loss of proton conductivity, which generally takes place above 40 °C in the normal conditions [3]. Another problem is the solubility of these HPAs whereby the mechanical strength of these materials was lost in the presence of liquid water. How to overcome these problems is a real challenge as proton conductivity at intermediate temperature, i.e. between 100 °C and 600 °C are ideal for fuel cell operations, particularly those used as power plants for automobiles. Two strategies have been proposed. One is to replace water molecules in the HPAs’ structure by other protonic moieties that can sustain high temperatures. This strategy has been realized by replacing water molecules with ionic liquids such as [BMIM][TFSI] [4]. High conductivity above 100 °C and under anhydrous conditions has been achieved. Nevertheless, the cost of ionic liquids is still too high to warrant large scale production. Another strategy is to form heteropolyacid salts which demonstrate reasonable conductivity at high temperatures and insoluble in water. Caesium based heteropolyacid salts have been investigated which shows encouraging results [5]. However, analogues such as ammonium based heteropolyacid salts, e.g. (NH4)3PW12O40 has not been investigated [6]. (NH4)3PW12O40 shows selectivity in catalyzing alkylation of amonia with methanol above 400 °C [7] and reduction of NO2 at a temperature as low as 150 °C [[8],[9]]. It has surface area over 100 m2 g-1 which is remarkably high considering the heavy molecular weight (∼2931). Inumaru and Ito et al. found the “epitaxial self assembly” phenomenon of (NH4)3PW12O40 and categorized (NH4)3PW12O40 as “sponge crystal” since there are large residual space left between the constituent nanocrystallites [[10],[11]]. The high porosity provides ample rooms to accommodate secondary phases such as acids therefore opening possibilities to develop composite proton conductors. Mikhailenko et al. investigated the electrical properties of (NH4)3PW12O40 in dry Ar, 100% RH and in the presence of liquid water [12]. They found that the conductivity of (NH4)3PW12O40 was improved several magnitudes by introducing liquid water (10-5-10-4 S/cm in dry Ar and 3.3-8.0 × 10-2 S/cm mixed with 10% water). Nevertheless, water is easy to evaporate at high temperatures and replacing water with stable acid would be a good choice to further improve the proton conductivity beyond 100 °C. In this paper, we introduce phosphoric acid to replace water and synthesized (NH4)3PW12O40-H3PO4 composites. The proton conductivity of (NH4)3PW12O40-H3PO4 composites at high temperatures was investigated. The high stability and insolubility of (NH4)3PW12O40 as well as its high uptake for H3PO4 signifies promising application as a proton conductor in the intermediate temperatures.

2. Experimental

2.1. Material synthesis

(NH4)3PW12O40-H3PO4 composites were synthesized by precipitation method: calculated amounts of H3PO4 (Aldrich) and NH4HCO3 (Aldrich) were mixed up and dissolved with deionized water simultaneously to obtain a transparent solution. The solution was then dripped into H3PW12O40 solution under vigorously agitations in order to get homogeneous mixtures. White precipitations were produced immediately and the mixtures were heated at 200 °C to remove excess water. The resultant white powders were labeled as NP0, NP3, NP5 and NP10 according to the molar ratios of NH4HCO3/ H3PW12O40/ H3PO4 = 3/1/0, 3/1/3, 3/1/5 and 3/1/10. The powders were stored at 90 °C for 12 h and collected for measurements.

2.2. Analytical procedures

Crystal structure and phase purity were examined by X-ray powder diffraction (XRD) technique on a Bruker D8 Focus diffractometer (Bruker, Germany) using CuKα1 radiation (λ =1.5405 Å) and CuKα2 radiation (λ =1.5444 Å) as incident radiation. 0.01° step size with duration of 0.1 s was adopted for data collection. The microstructures were examined under microscopic conditions using a field emission scanning electron microscope (FESEM, Hitachi S4800, Japan) and a transmission electron microscope (TEM, JEOL JEM-2100, Japan). Fourier transform infrared spectrum (FT-IR) was chosen to inspect the interior bonding conditions for all samples. KBr pellets were made for each sample and the IR spectrum was collected on the PERKIN-ELMER Spectrum GX FTIR systems from 400 cm-1 to 4000 cm-1 with a resolution of 4 cm-1. Thermal analysis was carried out on a Rheometric Scientific TG 1000 M 1000 M + TA instruments SDT 2960 with heating and cooling rate of 5 °C/min under flowing air at a rate of 20 mL/min.

2.3. Conductivity measurements

The conductivity measurements were carried out by the AC impedance method over the frequency range from 1 MHz to 10 mHz with an amplitude of 10 mV on a Zahner electrochemical workstation. Pellets were used for measurements and were obtained by pressing powders under a pressure of 3 × 103 kg/cm2 (diameter 13 mm, thickness 4-5 mm). PTFE bonded carbon (Carbot Vulcan 72R) was dabbed on both sides of the pellets which served as electrodes. Carbon paper (Torory TGPH-090) was subsequently pressed on as current collector. The pellets were mounted into furnaces for impedance measurements. Data were collected from the highest temperature to the lowest on cooling after holding at least one hour at each temperature to reach equilibrium. Air and wet air (humidified at 20 °C with moisture content about 3%) were applied to create different atmosphere and to evaluate conductivity.

3. Results and discussion

3.1. Phase composition, thermal stability and morphology

All samples obtained are white powders. XRD analysis shows similar diffraction patterns (Fig. 1) to standard data (JCPDS: 00-050-0305) with cubic symmetries. There is a clear shift of all diffraction peaks towards low angles implying enlargement of unit cell parameters. The calculated unit cell parameters were listed in Table 1. The lattice parameters and lattice volume were enhanced by introducing phosphoric acid. According to Inumaru [10], the particles of (NH4)3PW12O40 were composed of (NH4)3PW12O40 nanocrystallites which self-assembled with the same crystal orientations and the microporosity of (NH4)3PW12O40 comes from the residual space between these nanocrystallites. Therefore, the increase of unit cell parameters can be rationalized by replacing water molecules with larger H3PO4 molecules in the residual space of (NH4)3PW12O40. However, there seems a saturation of unit cell enlargement at high H3PO4 content as can be seen from Fig. 2. This is probably due to a complete fulfillment of H3PO4 in the residual space of (NH4)3PW12O40.

Fig. 1.   XRD patterns of NP0, NP3, NP5 and NP10, standard pattern for (NH4)3PW12O40 (JCPDS: 00-050-0305) is also included for comparison. Main peak around 26° is enlarged on the right. Dashed line is a guide for the eye.

Table 1   Unit cell parameters calculated from XRD patterns (standard deviation is included in the parenthesis).

SampleSymmetryLattice parameter, a (Å)Volume (Å3)
NP0Cubic11.6729(19)1590.5(5)
NP3Cubic11.7033(18)1603.0(4)
NP5Cubic11.7140(22)1607.4(5)
NP10Cubic11.718(3)1609.0(7)

New window

Fig. 2.   Unite cell parameters as a function of H3PO4 to (NH4)3PW12O40 molar ratio: (a) cell parameter a; (b) cell volume V.

Thermal analysis was carried out to investigate the dehydration and stability of these composites. For NP0, the analysis was continued from room temperature up to 800 °C. The results were shown in Fig. 3(a). According to Belanger et al. [8], ammonium was not released from the solid until the temperature is beyond 400 °C and considerable amount of ammonium was only released at a temperature higher than 550 °C. Therefore, the weight loss observed below 500 °C in Fig. 3(a) could be reasonably attributed to the loss of water. The thermogravimetric curves showed a two-step progress at this temperature range. The first step is a linear loss from room temperature up to around 118 °C, a weight loss of 1.48% was observed corresponding to 2.5 H2O molecules per (NH4)3PW12O40 unit. The weight loss at this region was generally considered as desorption of physically absorbed water due to the hydrophilic properties of (NH4)3PW12O40. The second one is from 118 °C to 500 °C where 1.07% weight loss occurs, corresponding to 1.8 H2O molecules per (NH4)3PW12O40 unit. The thermogravimetric curve at this region presented a smoother slope which could be ascribed to the loss of strongly absorbed water inside the internal structures of (NH4)3PW12O40. As the temperature kept increasing, a sharp loss took place and ended up around 612 °C. If ammonium was completely removed from the solid, a weight loss less than 1.7% should be observed theoretically. The experimental value (2.54%) was significantly larger than the theoretical one, suggesting a decomposition of other constituent elements such as oxygen. XRD patterns of samples after TG analysis confirmed these observations (Fig. 3(c)). The diffraction patterns were attributed to a pure W12PO38.5 (JSPDS card 00-041-0369) and a theoretical weight loss 2.6% can be calculated which is very close to the experimental one (2.54%). From TGA analysis of (NH4)3PW12O40, it is confirmed that the structure of (NH4)3PW12O40 can sustained up to 500 °C and water start to lose quickly even below 50 °C.

Fig. 3.   (a) Thermogravimetric analysis (TGA) of NP0, (b) TGA of NP3, NP5 and NP10, (c) XRD patterns of NP0 after TGA and (d) XRD patterns of NP10 after TGA. Standard patterns for W12PO38.5 (JCPDS: 00-041-0369) and (NH4)3PW12O40 (JCPDS: 00-050-0305) are also included for comparison.

TGA analysis was then performed on (NH4)3PW12O40-H3PO4 composites and the temperature range was set from room temperature to 300 °C (Fig. 3(b)). The corresponding weight loss below 100 °C and from 100 °C to 300 °C was listed in Table 2. It can be seen from the table that weight loss below 100 °C was quite different from NP3 to NP10 due to various extent of hydration during the preparation process. However, the weight loss between 100 °C and 300 °C increased systematically. This weight loss could mainly come from the dehydration of phosphoric acid since only 0.99% was observed for pure (NH4)3PW12O40 whereas more than 2.53% was found in the composites. It has been realized that phosphoric acid could condense into pyrophosphoric acid in a temperature of 175 °C or even lower [13], therefore, the accelerated weight loss above 150 °C for all composites could mainly contributed to the condensation of phosphoric acid. It is worth mentioning that (NH4)3PW12O40 is quite stable in the presence of phosphoric acid at this temperature range. XRD patterns in Fig. 3d revealed that the composite was stable at high acid loadings (NP10) since the diffractions of (NH4)3PW12O40 phase were maintained after TGA analysis.

Table 2   Weight loss of all samples in different temperature regions.

SampleWeight loss below 100 °CWeight loss between 100 °C and 300 °CTotal weight loss
NP00.98%0.99%1.97%
NP32.34%2.53%4.87%
NP50.23%3.24%3.57%
NP101.22%5.62%6.84%

New window

The morphologies of the composites and pure (NH4)3PW12O40 were examined under SEM and TEM conditions. Fig. 4 shows the SEM images of all sample powders. Spherical particles were observed in all cases. The particles sizes were approximately in the range of 100-500 nm which was consistent with literature [6]. According to Inumaru, the particles of (NH4)3PW12O40 were composed of nanocrystallites that self-assembled together in the same orientations and the morphologies of these particles were highly sensitive to the preparation temperatures: with spherical particles in the low temperatures (∼ 20 °C) and symmetric particles in higher temperatures (∼ 90 °C). In addition, they attributed this phenomenon to the various surface energy and/or growth rates between crystal planes. This speculation could be partly true due to the small constituent nanocrystallites (4-6 nm) and exceptionally high surface energy that could be envisaged. However, in our point of view, there could be another factor that might govern the outer shape of (NH4)3PW12O40 which is the hydrogen bonds. According to Okuhara et al. Caesium based heteropolyacid salts like Cs3-xPW12O40 do not have such epitaxial growth phenomenon [14] and the particle dimensions calculated through XRD were significantly smaller (24 nm for Cs3PW12O40) than (NH4)3PW12O40 (> 150 nm) prepared at the same conditions. This observation could not explained by surface energies since the constituent nanocrystallites were in the same dimensions (< 10 nm) and M3PW12O40 (M=Cs, NH4, Rb, K) shared the same crystal structures. Therefore, other factors could trigger the epitaxial growth of (NH4)3PW12O40 and hydrogen bonds are well-known for their direction and strength, it is quite likely that protons in ammonium formed the hydrogen bonds like N-H…O-W and lead to the translation and rotation of Keggin ions during the local dissolution-reprecipitation process. Fig. 5 shows the TEM pictures of NP10 where spherical particles with diameters varied from 150 nm to 500 nm were observed. Inumaru found that the BET surface area and micropore volume of (NH4)3PW12O40 decreased significantly at a high precipitation temperature due to the rearrangements and formation of rigid dodecahedra. It can be seen from Fig. 6 that the particles remained spherical after aging at 100 °C for 2 h and no rigid dodecahedra were observed.

Fig. 4.   Field emission SEM images of freshly prepared sample powders: (a, b) NP0; (c, d) NP3; (e, f) NP5; (g, h) NP10.

Fig. 5.   TEM images of NP10 at different magnifications.

Fig. 6.   FT-IR spectra of all sample powders (85 wt% H3PO4 is also included for comparisons).

3.2. IR spectra

Fig. 6 shows the infra red absorption spectra of pure (NH4)3PW12O40 and all composites. The spectra of 85 wt% phosphoric acid were also shown for comparisons. It can be found that the composites and pure (NH4)3PW12O40 exhibit only minor differences, implying the structure of Keggin ions are preserved in the presence of phosphoric acid. The main differences between pure (NH4)3PW12O40 and the composites lie in the range 1100-1300 cm-1 and 2600-3300 cm-1. Pure (NH4)3PW12O40 shows almost no absorption in these two ranges. The composites show a broad shoulder absorption peak from 1100 cm-1 to 1300 cm-1 corresponding to vibrations of different phosphoric species (HPO42- and H2PO4-) [15] and relatively stronger absorptions at 2600-3300 cm-1 which is due to the ν(O-H) vibrations [16]. These absorptions could come from the phosphoric acid in the structure of (NH4)3PW12O40 if we compare them with 85 wt% phosphoric acid where much stronger and broader absorptions appeared around the same ranges. Park et al. pointed out that hydroxyl groups that have absorption around 2900 cm-1 have higher proton mobility compared with the ones at higher wavelength [17]. Therefore, it is clear that introducing phosphoric acid into (NH4)3PW12O40 significantly enhances the intensities of absorption peaks at this region. Another factor worth noticing is that absorption peaks at 525 cm-1, 795 cm-1, 887 cm-1 and 975 cm-1 which could be assigned to νs(W-O-W), νas(W-Ob-W)and νas(W-Oc-W), and ν(W = Ot) [16] were broadened progressively with increasing phosphoric acid. This phenomenon could be a sign of interactions between phosphoric acid and oxygen ions in Keggin structure such as the formation of hydrogen bonds so that the symmetry of the Keggin ions was consequently lowered. The weak absorption peaks at 596 cm-1, 1400 cm-1 and 1079 cm-1 could be assigned to δ(O-P-O), νas(POa) and deformation vibration of the ammonium ions respectively [[3],[16],[18]].

3.3. Proton conductivity

The conductivity of pure (NH4)3PW12O40 under different atmospheres was shown in Fig. 7. The values are calculated from the Cole-Cole plots obtained by AC impedance spectra at the intersection to the real axis. The pure (NH4)3PW12O40, however, was not a good proton conductor even in wet air and did not obey Arrhius behavior which could be ascribed to the continuous loss of water at high temperatures [[8],[12]]. According to Mikhailenko et al., mechanisms had been proposed for the explanations of proton transport in (NH4)3PW12O40. They attributed the conductivity to the transport of hydrated protons which was developed in the vicinity of NH4+ sites in (NH4)3PW12O40 structures. In our experiment, (NH4)3PW12O40 showed a much better conductivity in wet air than in air which could be the water effects and become an evident sign of proton conductions. More specifically, the differences between wet air and air were enlarged from room temperature to 220 °C due to the thermal activation and subsided above 220 °C, probably due to the dehydrations [8]. The ammonium was not expected to release from (NH4)3PW12O40 below 400 °C, so the better performance in wet atmosphere could come from the strong interactions between ammonium ions and water molecules. This phenomenon gives rise to broad potential applications such as sensors etc. since more than two orders of magnitudes of differences in conductivity was achieved in different atmospheres.

Fig. 7.   (a) Conductivity (σ) of NP0 in air and in wet air (humidified at 293 K) and (b) conductivity of NP3, NP5 and NP10 in air (T: temperature).

The composites, however, exhibited a much better performance on conductivity in air (Fig. 7). The highest conductivity is 0.14 S/cm for NP10 at 170 °C which is almost 4 orders of magnitudes higher than NP3 under the same conditions. It is concerned that phosphoric acid in NP3 was discretely distributed in the so-called “spongy crystal” of (NH4)3PW12O40 where continuous channels for proton conduction could not be fully established. However, in the case of NP10 where a huge amount of phosphoric acid was added, those internal space and channels inside crystals were filled and connected well with each other therefore a high conductivity was observed. The decreased conductivity at high temperature was attributed to condensation of phosphoric acid to polyphosphoric acid where large volume decrease occurs therefore reduce the continuous channels for proton transportations.

4. Conclusion

(NH4)3PW12O40 and (NH4)3PW12O40-H3PO4 composites were synthesized by precipitation method. Their phase compositions, thermal stability and morphologies were investigated. It was found that (NH4)3PW12O40 maintained its structures in the presence of phosphoric acid and the phosphoric acid was believed to be preserved in the micropores of (NH4)3PW12O40. SEM and TEM analysis showed that the particles of all samples were spherically shaped with diameters of 200-500 nm. FTIR spectra confirmed the maintenance of Keggin structures and strong interactions between phosphoric acid and (NH4)3PW12O40. The pure (NH4)3PW12O40 exhibited a poor conductivity in air at high temperatures but the conductivity was improved significantly in wet air due to the water promotion effect. The composites, however, demonstrated a much better conductivity especially at high acid loadings where continuous channels for proton conduction can be established. A conductivity of 0.14 S/cm was achieved at 170 °C for NP10. Such high conductivity gives rise to broad potential applications such as electrolyte in fuel cells and other electrochemical devices.

Acknowledgements

This work was supported financially by the National Natural Science Foundation of China (No. 21401142, 51972233), the Natural Science Foundation of Shanghai (No. 19ZR1459200), the National 1000-Plan Program, the Shanghai Science and Technology Commission (No. 14DZ2261100), the South Taihu Elite Project and the Fundamental Research Funds for the Central Universities.


/