Next Article in Journal
Porous Biomass Carbon Derived from Clivia miniata Leaves via NaOH Activation for Removal of Dye
Next Article in Special Issue
Effect of the P/Al Molar Ratio and Heating Rate on the Composi-Tion of Alumino-Phosphate Binders
Previous Article in Journal
On the Effects of Core Microstructure on Energy Absorbing Capabilities of Sandwich Panels Intended for Additive Manufacturing
Previous Article in Special Issue
Cobalt Minimisation in Violet Co3P2O8 Pigment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Properties and Applications of Metal Phosphates and Pyrophosphates as Proton Conductors

by
Rosario M. P. Colodrero
,
Pascual Olivera-Pastor
,
Aurelio Cabeza
and
Montse Bazaga-García
*
Departamento de Química Inorgánica, Cristalografía y Mineralogía, Facultad de Ciencias, Universidad de Málaga, Campus Teatinos s/n, 29071 Málaga, Spain
*
Author to whom correspondence should be addressed.
Materials 2022, 15(4), 1292; https://doi.org/10.3390/ma15041292
Submission received: 29 December 2021 / Revised: 27 January 2022 / Accepted: 3 February 2022 / Published: 9 February 2022

Abstract

:
We review the progress in metal phosphate structural chemistry focused on proton conductivity properties and applications. Attention is paid to structure–property relationships, which ultimately determine the potential use of metal phosphates and derivatives in devices relying on proton conduction. The origin of their conducting properties, including both intrinsic and extrinsic conductivity, is rationalized in terms of distinctive structural features and the presence of specific proton carriers or the factors involved in the formation of extended hydrogen-bond networks. To make the exposition of this large class of proton conductor materials more comprehensive, we group/combine metal phosphates by their metal oxidation state, starting with metal (IV) phosphates and pyrophosphates, considering historical rationales and taking into account the accumulated body of knowledge of these compounds. We highlight the main characteristics of super protonic CsH2PO4, its applicability, as well as the affordance of its composite derivatives. We finish by discussing relevant structure–conducting property correlations for divalent and trivalent metal phosphates. Overall, emphasis is placed on materials exhibiting outstanding properties for applications as electrolyte components or single electrolytes in Polymer Electrolyte Membrane Fuel Cells and Intermediate Temperature Fuel Cells.

1. Introduction

Metal phosphates (MPs) comprise an ample class of structurally versatile acidic solids, with outstanding performances in a wide variety of applications, such as catalysts [1,2,3], fuel cells [4,5,6], batteries [7], biomedical [8], etc. Depending on the metal/phosphate combinations and the synthetic methodologies, MP solids can be prepared in a vast diversity of crystalline forms, from 3D open-frameworks, through layered networks, to 1D polymeric structures.
The advantages of these solids are, among others, their low cost and easy preparation, hydrophilic and thermally stability and structural designability to a certain extent. Additionally, their structures are amenable to post-synthesis modifications, including the incorporation of ionic or neutral species that significantly affect their functionality and other important properties, such as the formation of hydrogen bonding networks or insertion of proton carriers.
While tetravalent and monovalent metal phosphates are usually prepared by conventional synthetic routes, the synthesis of many divalent and trivalent metal phosphates requires the presence of organic molecules/ions as structure-directing agents, some of which also act as charge-compensating ions [9]. In any case, the metal phosphates contain acidic phosphate groups (H2PO4 and HPO42−) or they can result from specific post-synthesis acid–base reactions, thus, allowing a mostly intrinsic proton conduction.
Additionally, an extrinsic proton conduction, associated with surface proton transport, can also result from specific chemical modification and/or induced morphological changes, e.g., with formation of nanoplatelets or nanorods particles [10,11]. Water molecules act both in the formation H-bonding networks inside the framework and as proton carriers. Moreover, other guest species, including organic moieties, can be inserted in their structure following different strategies, such as intercalation [12,13,14], template [15,16], and post-synthesis adsorption [17,18,19].
Accordingly, proton transport in MPs results from the interplay between water/guest molecules and the hydrophilic network/framework walls in specific ways. Thus, in metal phosphates, such as CsH2PO4 [20] or CaHPO4·2H2O [21] the proton-transfer mechanism involves not only the jump of H+ but also the free rotation of the phosphate groups. In other cases, such as for KH2PO4 the proton-transfer mechanism behavior is by defects [22].
A more complex proton-transfer process arise from the so called “quasi-liquid” state, such as described for H3OUO2PO4·3H2O [23,24], or it may stem from various mobile species moving on the surface, each one using different paths at different rates, e.g., in Ce(HPO4)2·nH2O [25,26]. The proton transport in the quasi-liquid state is often characterized by a “proton jump” or the Grotthuss mechanism, with low activation energy (Ea) [21].
Determining proton conduction mechanisms is key to designing proton conductor materials (Figure 1). Thus, if a vehicle-type mechanism is needed, proton-containing counter ions, such as hydronium ions, NH4+, protonated amine groups, and protonated organic molecules in general, should be incorporated into the structure. Additionally, the immobilization of specific functional groups and the corresponding counter ions onto the framework could be required.
On the other hand, solids exhibiting proton transport by a Grotthuss-type mechanism basically consist of continuous H-bond networks that favour H+ conduction with low activation energy (Ea). A main drawback for the latter materials is that, above certain temperatures, breaking of the H-bond networks can result, accompanied by the release of water molecules, which makes the construction of permanent H-bonding networks with hydrophilic channels or exploring new alternative conductive media necessary [27].
A research area of paramount interest is that relative to fuel-cell (FC) development regarding energy efficiency, environmental issues, and the gradual depletion of conventional fuel reserves [28,29]. The operating temperature regime of these electrochemical devices is determined by the electrolyte. Among the many types of solid electrolytes that have been developed over the past several decades [30], perfluorinated sulfonic acid (PFSA) ionomers, such as Nafion® and Aquivion®, are the most widely used in PEMFCs and DMFCs as polymer electrolyte membranes (PEM) for operating temperatures <120 °C.
This is due to superior conductive properties, ranging from 9 × 10−3 S·cm−1 to 1.2 × 10−1 S·cm−1, and chemical and mechanical stability [31,32]. This PEM acts as a separator for the reactants and, at the same time, provides pathways for proton transport [33,34]. However, Nafion® also shows some important drawbacks, such as methanol permeability, high costs, and complex steps of preparation, working below 100 °C to maintain the hydration of the membrane and other issues related to catalyst efficiency. All this makes necessary the development of new membrane materials [35,36].
Intermediate Temperature fuel cells (ITFCs), which operate between 120 and 300 °C, circumvent some of these problems, achieve less catalyst poisoning and better handling of the hydration conditions, and enable the use of non-precious or low-loading precious metal catalysts [37,38,39]. Nevertheless, for the case of methanol HT-PEMFCs, which must operate above 250 °C and with high pressure, possible degradation of the electrolytes should be a subject of study [40]. Some MP-based electrolytes gather proton conductivity properties to be used as single electrolytes or main components for ITFC applications. Examples are superprotonic CsH2PO4 [41] and M(IV)P2O7-based electrolytes [42].
This review summarizes the state of the art on proton conductive metal phosphates with special attention to the last two decades of developments. A review of publication trends, from the CAS SciFindern database, is given in Figure 2, which confirms the increasing interest in this topic during this period. In this article, our focus is placed on updating recent advances in proton conductors based on tetravalent metal phosphates/pyrophosphates and superprotonic Cs2HPO4. In addition, the design and development of new proton carrier systems relying on divalent and trivalent metal phosphates are also revised.
The potential applications of these materials as proton conductor electrolytes for PEMFCs and ITFCs are also discussed. Other high temperature (above 300 °C) conductive metal phosphates, such as rare earth phosphates and metal polyphosphates, ionic conductors (e.g., NASICON), and electrode materials are out the scope of this revision. The review is not intended to be exhaustive, and therefore certain specific reports and/or applications may not be covered.

2. Tetravalent Metal Phosphates and Pyrophosphates

2.1. Zirconium Phosphates

The prototype of layered metal(IV) phosphates is zirconium hydrogen phosphate, which presents mainly three topologies, known as the forms α: Zr(HPO4)2·H2O (α-ZrP), γ: Zr(PO4)(O2P(OH)2) 2H2O (γ-ZrP), and λ: (Zr(PO4)XY, X = halides, OH, HSO4, and Y = DMTHUS, H2O, etc.) (λ-ZrP). α-ZrP is the most thermal and hydrolytically stable phase and, thus, the most studied compound following the pioneering works by Giulio Alberti and Abraham Clearfield groups [43,44,45].
In the structure of α-ZrP, planar Zr atoms are coordinated by the three oxygen atoms of HO-PO3 groups, while the P-OH bond points toward the interlayer space giving rise to small cavities, where the lattice water is located. Although this water forms hydrogen bonds with P-OH groups, an extended interlayer hydrogen bond network is absent in this structure [44]. On the other hand, the structure of γ-ZrP consists of a biplanar Zr atomic system connected through bridge PO43− groups, while the externally located H2PO4 groups complete the octahedral coordination of the Zr atoms.
Due to the higher stability of the α form, proton conductivity studies have been conducted mainly with this phase. Typical values of 10−6–10−4 S·cm−1 have been determined at RT and 90% RH with activation energy (Ea) ranging between 0.26 (90% RH) and 0.52 eV (5% RH). This conductivity originates from surface transport and is dominated by surface hydration since the long distance between adjacent P-OH groups hinders proton diffusion along the internal layer surface [46].
According to theoretical estimations [47], the interactions between POH and water molecules on the zirconium phosphate surface are relatively large and generate strong local hydrogen-bond networks. The high density of acid groups should produce continuous short O-O distances and generate proton-transfer paths with low activation energy, even under high-temperature and low-humidity conditions.
Studies on microcrystalline α-ZrP revealed that both the morphology and the electric field orientation are important aspects to consider. Thus, pellicular samples of α-ZrP, prepared by intercalation-deintercalation of n-propylamine [48], exhibited a proton conductivity of 10−4 S·cm−1 at 300 °C when it was measured with the electric field parallel to the pellicles. This conductivity was found to be approximately two orders of magnitude higher than that measured on applying the electric field perpendicular to the pellicles [49].
Several other strategies to increase the proton conductivity of α-ZrP have been developed (Table 1). One of the first approaches was the formation of polyhydrated n-propylamine-intercalated compounds (ZrP·xPrNH2·nH2O). For a composition with x = 0.8 and n = 5, a proton conductivity of 1.2 × 10−3 S·cm−1 at 20 °C and 90% RH was determined [50]. Another more sophisticated way of enhancing the proton conductivity of zirconium phosphate was the preparation of mesoporous zirconium phosphate/pyrophosphate, Zr(P2O7)0.81(O3POH)0.38.
This material, obtained from the thermal decomposition of zirconium phosphite benzenediphosphonate, displayed a high specific surface area (215 m2/g) with an enhanced proton conductivity of 1.3 × 10−3 S·cm−1 at 20 °C and 90% RH [51]. One of the most successful ways of obtaining zirconium phosphate derivatives with outstanding proton conductivity properties consisted in preparing mixed zirconium phosphate/phosphonates, an example being [Zr(O3POH)2−x(O3PC6H4SO3H)x], whose proton conductivity was found to increase with x, up to a value of 0.07 S·cm−1 (x = 1.35 at 100 °C and 90% RH) [52].
The great versatility of zirconium phosphate preparation under numerous experimental conditions, has been exploited to synthesize various novel derivatives, which displayed remarkable proton conductivity properties. Two examples are layered (NH4)2[ZrF2(HPO4)2] [53] and the 3D open framework zirconium phosphate (NH4)5[Zr3(OH)3F6(PO4)2(HPO4)] [54]. In both structures, the participation of NH4+ ions in the formation of extended hydrogen-bond networks (Figure 3) resulted in proton conductivities of 1.45 × 10−2 S·cm−1 (at 90 °C and 95% RH) and 4.41 × 10−2 S·cm−1 (at 60 °C and 98% RH), respectively. From the Ea values found, 0.19 and 0.33 eV, respectively, a Grotthuss-type proton-transfer mechanism was inferred for these materials (Table 1).
A one-dimensional zirconium phosphate, (NH4)3Zr(H2/3PO4)3, consisting of anionic [Zr(H2/3PO4)3]n3n− chains bonded to charge-compensating NH4+ ions was reported [55]. In this structure, the phosphates groups are disorderly protonated, while NH4+ ions occupy ordered positions in between adjacent chains (Figure 4). This arrangement generates H-bonding infinite chains of acid–base pairs (N-H···O-P) that lead to a high proton conductivity in anhydrous state of 1.45 × 10−3 S·cm−1, at 180 °C (Table 1). This material also showed a remarkable high performance in PEMFC and DMFC under operation conditions.

2.2. Zirconium Phosphate Composite Membranes

ZrP has been widely studied as a suitable filler for the preparation of Nafion®-composite electrolytes due to its good thermal and chemical stability, proton conductivity, hydrophilic nature, very low toxicity, and stability in hydrogen/oxygen atmospheres. Nafion®, a sulfonated tetrafluoroethylene-based fluoropolymer copolymer is the most widely used proton conductor for polymer electrolyte membrane fuel cells (PEMFCs) [56].
Despite its excellent properties, this polymer is limited to operations at low temperatures of <100 °C, due to dehydration, which drastically reduces the proton conductivity. As operating at higher temperatures is desirable for better efficiency, Nafion® composite membranes have been investigated in which fillers of diverse nature were incorporated to the polymer using different approaches.
Various methodologies of preparing Nafion®/ZrP membranes have been devised [46,57,58,59,60,61,62,63,64,65]. Among them, treatment of the Zr(IV)-exchanged ionomer with phosphoric acid [59], filler/ionomer co-precipitation [61] and casting of ionomer solution/filler gel mixtures [60,61,62]. In general, improvements in mechanical properties and efficiency [61] were highlighted, although high filler loads beyond 25% led to decreased proton conductivity in certain cases.
A derivation of the above methods consisted of using nanosized ZrP particles-containing ionomer Aquivion® as fillers [63]. The resulting composite membranes showed enhanced elastic modulus and lower proton conductivity as compared to Nafion® itself, although they were high enough for fuel cell application [64] with values ranging between 0.16 and 0.23 S·cm−1 at 90% RH [62,65].
Derivatising ZrP with hydrophobic groups resulted in composite membranes mechanically stronger than the pristine Nafion® membrane while displaying high proton conductivities [66,67,68,69,70]. Furthermore, integrating both hydrophobic and hydrophilic ZrP nanofillers showed a synergistic effect with mechanical reinforcement of the composite membrane and better fuel cell performance than that found when single ZrP filler were employed [60]. Other ionomer alternatives to Nafion®, such as sulfonated poly(arylene ether) and poly(ether ether ketone) polymers (SPAEK and SPEEK, respectively), polybenzimidazole (PBI), and vinyl-type (PVA and PVP), have been employed as components of polymer electrolyte membranes [71].
The sulfonated poly(fluorenyl ether ketone) (SPFEK) polymer presents acceptable proton conductivity and stability; however, its performance is lesser in comparison to the benchmark Nafion® ionomer. To overcome this handicap, several approaches have been implemented, one of them being hybridization with α-zirconium phosphate. Recently, SPFEK/ZrP-SO3H composite membranes have been reported with improved proton conductivity, oxidative stability, methanol cross-over barrier and improved H2/O2 fuel cell performance [72,73].
ZrP has been also used as a filler in other polymeric matrices, such as poly(vinylidene fluoride) (PVDF) and (sulfonatedpoly(styrene-block-(ethylene-ran-butylene)-block-styrene) (S40). The resulting composite membranes were tested for potential applications in DMFC, with improved performance especially relative to reducing the methanol permeability [72,74]. PA-PBI/Zr(HPO3) composite membranes (PA-PBI = phosphoric acid-doped polybenzimidazole) exhibited enhanced proton conductivity as compared to a pristine polymer electrolyte membrane at 200 °C and 5% RH [75].
A different approach is based on the substitution of the polymer membranes for others more eco-friendly, inexpensive, and biodegradable materials. Although poly(vinyl alcohol) (PVA) is a suitable candidate for this purpose, it still requires some enhancements related to its ionic conductivity due to its rigid structure. As an example, Gouda et al. [76] doped iotan carrageenan (IC)/sulfonate PVA (SPVA) membranes with different loads of zirconium phosphates. Enhancement of the membrane properties was attributed to abundant hydrogen bond interactions established between the zirconium phosphate particles and the polymers.

2.3. Titanium and Tin(IV) Phosphates

Structurally, titanium(IV) phosphates are more diverse than ZrP, given that structures containing an oxo/hydroxyl ligand or mixed-valence (TiIII/TiIV) titanium ions have been reported. In addition to the zirconium phosphate analogues, α- [77] and γ-TiP [78], two more other layered titanium phosphates, TiO(OH)(H2PO4)·2H2O [79] and Ti2O3(H2PO4)2·2H2O are known [80]. Furthermore, the mixed-valence titanium phosphate TiIIITiIV(HPO4)4·C2N2H9·H2O contains microporous channels [81], where monoprotonated ethylenediamine and lattice water are accommodated, which favours its transformation into a porous phase, Ti2(HPO4)4, at 600 °C with a 3D structure similar to that of τ-Zr(HPO4)2 [82].
Impedance spectroscopy experiments combined with dynamics simulations revealed that this compound was a good proton conductor, with a σ value of 1.2 × 10−3 S·cm−1 (at room temperature and 95% RH) and Ea of 0.13 eV (Table 1). From these data, a water-mediated proton transport originating from the H-bond interaction of water molecules with the bridging oxygen of the porous framework was proposed [83].
The crystal structure of two other open-frameworks Ti2O(PO4)2·2H2O polymorphs (ρ-TiP and π-TiP) has been reported [84,85]. Fibrous ρ-TiP and π-TiP crystallize in triclinic and monoclinic unit cells, respectively, and are composed of TiO6 and TiO4(H2O)2 octahedra bridged through the orthophosphate groups (Figure 5). This type of connectivity creates two types of 1D channels running parallel to the direction of the fibre growth [84,85,86]. 31P MAS-NMR spectroscopy studies revealed that π-TiP has capability of adsorbing superficially protonated phosphate species (H3PO4/H2PO4/HPO42−), which affects the proton conductivity properties as confirmed by AC impedance measurements.
Thus, this solid reached a proton conductivity value of 1.3 × 10−3 S·cm−1, at 90 °C and 95% RH (Table 1). This behaviour was explained as being due to an extrinsic vehicle-type proton transport mechanism. Additionally, H3PO4-impregnated π-TiP solid was used as filler for the preparation of composite membranes of Chitosan (CS) matrices exhibiting a proton conductivity of 4.5 × 10−3 S·cm−1 at 80 °C and 95% RH, which is 1.8-fold higher than that of the pristine CS membrane [86].
Insights in the structure of the isomorphous series of layered α-metal(IV) phosphates [M(IV) = Zr, Ti, Sn) showed appreciable differences in the hydrogen bond interaction of the lattice water with the layers, associated with a variable layer corrugation degree along the series; α-TiP and α-SnP displaying H-bonds stronger than the prototype α-ZrP [87], which, in turn, might have significant implications in making internal surfaces accessible to proton conduction paths.
Nevertheless, studies on structure–property correlations are still lacking, due, in part, to appreciable differences in chemical stability under different conditions, particularly in the case of α-SnP. Recent studies carried out on a nanolayered γ-type tin phosphate, Sn(HPO4)2·3H2O, have revealed that this solid could be obtained in a water-delaminated form, which showed a high proton conductivity of ~1 × 10−2 S·cm−1 at 100 °C and 95% RH (Table 1). This high value was attributed to strong H-bonds between water and the SnP layers, in combination with a high surface area of 223 m2·g−1 [88].

2.4. Other Tetravalent Phosphates

POH groups-containing silicophosphates have received attention as proton-conducting electrolytes [42] but its solid-state structural characterization is elusive, at least, under conditions similar to those used in operating fuel cells [89]. A conductivity value of 2.5 × 10−3 S·cm−1 at 180 °C, under more than 0.4% RH, was reported for a silicophosphate-based composite membrane. However, it was noted that the formation of Si5O(PO4)6 crystals, under dry conditions, was detrimental for the proton conductivity.
NASICON-type compounds with the general formula, AxB2(PO4)3, where A = H+, H3O+, or NH4+ and B = Zr, Ti, Hf, Sn, Ge, etc., are also potential proton conductors [90,91]. HZr2(PO4)3 was firstly prepared by Clearfield et al. [92] and Komorowski et al. [93] described the synthesis of a hydrogen form from Na1 + xZr2SixP3–xO12 by replacing Na+ ions with hydronium ions through ion exchange. This solid showed a bulk proton conductivity of 5.0 × 10−4 S·cm−1 at 100% RH and 25 °C, the proton transport mechanism being a Grotthuss-type one [93].
Other materials, such as HZr2(PO4)3·nH2O, NH4Zr2(PO4)3, or H1 ± xZr2 xMx(PO4)3 · H2O (M = Nb, Y) [94,95], have been reported. Some NASICON-type compounds based on zirconium or titanium hydrogen phosphate exhibit remarkable proton conductivity while showing a high thermal stability, at least for the case of zirconium derivatives [96,97]. Less studied are the NASICON-type hafnium hydrogen phosphates. Among them, (NH4)0.4H0.6Hf2(PO4)3, obtained by thermal decomposition of the cubic NH4Hf2(PO4)3 at 550 °C, showed a proton conductivity of 1.2 × 10−6 S·cm−1 at 500 °C [90].

2.5. Tetravalent Pyrophosphates

In addition to acid metal(IV) phosphates, tetravalent metal pyrophosphates (MP2O7; M = Sn, Ce, Ti, Zr), in particular tin(IV) derivatives, have received considerable attention as proton conductors for intermediate temperature electrolyte applications. These solids combine high thermal stability (T ≥ 400 °C) with conductivities of 10−3–10−2 S·cm−1 in the temperature range of 100–300 °C, under anhydrous or low humidity conditions [98,99,100,101,102]. The origin of this proton conductivity was initially explained because of protons being incorporated into the framework, formed by metal(IV) octahedra and corner-sharing phosphate tetrahedral, after interaction with water vapor.
Defect sites, such as electron holes [103] and oxygen vacancies are believed to favour proton generation [104]. These protons may occupy hydrogen-bonding interstitial sites on either the M−O−P or the P−O−P bonds, thus, generating a hopping proton transport between these sites [105,106]. In support of this, 1H NMR studies revealed two signals corresponding to hydrogen-bonded interstitial protons at phosphate tetrahedral and metal octahedral sites in In3+-doped SnP2O7. Furthermore, indium doping seemed not to affect the proton conduction occurring by a hopping mechanism between octahedral and tetrahedral sites.
Thus, the increased proton conductivity is instead attributable to increased proton concentration [107]. The strategy of dopant insertion into the M(IV) pyrophosphate has been further extended to include other dopant divalent and trivalent metal ions, such as Mg2+, Al3+, Sb3+, Sc3+, and Ga3+, with a maximum conductivity of 0.195 S·cm−1 (Table 1) being reported for In0.1Sn0.9P2O7 [108]. However, recent studies on metal doped and undoped SnP2O7 suggested that it is in co-precipitated phosphorous-rich amorphous phases where the proton conductivity mostly resides, while the crystalline phase exhibit a very low conductivity (~10−8 S·cm−1 at 150−300 °C) [109].
Notwithstanding, the proton conductivity of the high temperature-sintered materials could be recovered by phosphoric acid treatment [110] or by carefully choosing the phosphate precursor. For the latter case, tetrabutylammonium phosphate-based SnP2O7/Nafion® composite membranes showed an increased fuel cell performance, as compared to other analogous SnP2O7/Nafion® composite membranes [111]. A major challenge in using metal(IV) pyrophosphates as electrolytes is forming dense pellets and the associated low open circuit voltages (OCVs) and considerable gas crossover [108].
Several studies to test applicability of composite membranes for ITFCs have been conducted. For this purpose, dispersion of Sb0.2Sn0.8P2O7 nanoparticles into H3PO4-doped PBI membranes were shown to exhibit enhanced conductivities and greater power density relative to the pristine membranes in the single cell tests [112]. Similar claims were made for composite membranes based on Sn0.95Al0.05P2O7 [113]. Other approaches, such as using graphite oxide (GO)/SnP2O7 or biphosphate-containing SnP2O7/Nafion composite membranes have been accomplished.
The former showed improved proton conductivity (~7.6 × 10−3 S·cm−1) peak power density of 18 mW·cm−2 at 220 °C and reduced fuel crossover [114]. The latter, aimed at increasing stability by establishing strong ammonium cation-biphosphate ion interactions, which prevents the loss of biphosphate and, hence, optimises the three-phase interface at the fuel cell electrodes. With these improvements, the H2/O2 fuel cell delivered a power density of 870 mW·cm−2 at 240 °C, with minimal performance loss even under exposure to gas containing 25% carbon monoxide [37].
Mg-doped cerium pyrophosphate, with optimal composition of Ce0.9Mg0.1P2O7, showed a remarkable performance as electrolyte into a fuel cell, generating electricity up to 122 mA·cm−2 at 0.33 V and displaying a peak power value of 40 mW·cm−2 when using 50% H2 at 240 °C. Prepared in disk form and sintered at 450 °C, the Mg-doped material exhibited a significantly high proton conductivity in the presence of moist air (PH2O = 0.114 atm) between 160 and 280 °C. A maximum value of 4.0 × 10−2 S·cm−1 was found at 200 °C, (Table 1), which was even higher than that of CeP2O7 (3.0 × 10−2 S·cm−1, at 180 °C) [115].
Low crystallinity titanium pyrophosphates also exhibit high proton conductivity [116]. Values of 0.0019–0.0044 S·cm−1, at 100 °C and 100% RH were reported for these compounds. The proton conduction is attributable to the presence of H2PO4 and HPO42− species, demonstrated by 31P MAS NMR and FTIR, and it occurs through a water-facilitated Grotthuss-type proton transport mechanism. Titanium phosphates are also prone to be functionalized with phosphonates groups. Thus, mixed titanium phosphate/phosphonates, Ti(HPO4)1.00(O3PC6H4SO3H)0.85(OH)0.30·nH2O [117], displayed an exceptional proton conductivity of 0.1 S·cm−1 at 100 °C (Table 1).
A structurally different proton-conductor metal pyrophosphate is the mixed template-containing vanadium nickel pyrophosphate, (C6H14N2)[NiV2O6H8(P2O7)2]·2H2O (Figure 6). Its crystal structure is composed of octahedrally coordinated V(IV) and Ni(II) interconnected through bridging pyrophosphate groups, thus creating a 3D framework of [NiV2O6H8(P2O7)2]2− unit charged-compensated by protonated DABCO molecules. Hydrogen-bonding networks inside 3D channels facilitate the proton transport, and thus this material exhibits a remarkable proton conductivity of 2.0 × 10−2 S·cm−1 at 60 °C and 100% RH (Table 1) through a Grotthuss-type proton-transfer mechanism (Ea = 0.38 eV) [118].

3. Super Protonic Metal(I) Phosphates

Solid acid proton conductors, with stoichiometry MIHyXO4 (MI = Cs, Rb; X = S, P, Se; y = 1, 2), have received much attention because they exhibit exceptional proton transport properties and can be used as electrolytes in fuel cells operated at intermediate temperatures (120−300 °C). The fundamental characteristics of these materials are the phase transition that occurs in response to heating, cooling or application of pressure [6,119], accompanied by an increase in proton conductivity of several orders of magnitude—referred to as super protonic conductivity (Figure 7).
This property has been associated with the delocalization of hydrogen bonds [120]. For CsH2PO4, a proton conductivity of 6 × 10−2 S·cm−1 (Table 1) was measured above 230 °C corresponding to the super protonic cubic (Pm−3m) phase [121] while it drastically drops in the low temperature phases. Recently, from an ab initio molecular dynamics simulation study of the solid acids CsHSeO4, CsHSO4 and CsH2PO4, it was concluded that efficient long-range proton transfer in the high temperature (HT) phases is enabled by the interplay of high proton-transfer rates and frequent anion reorientation.
In these compounds, proton conduction follows a Grotthuss mechanism with proton transfer being associated with structural reorientation [122]. The super protonic conductor CsH2PO4 is stable under humidified conditions (PH2O = 0.4 atm) [6], but it dehydrates to CsPO3, via the transient phase Cs2H2P2O7, at 230−260 °C, according to the relationship log(PH2O/atm) = 6.11(±0.82) − 3.63(±0.42) × 1000/(Tdehy/K) [123].
Several studies [6,124,125] have shown that CsH2PO4 can be employed as the electrolyte in fuel cells with good long-term stability. Thus, a continuous, stable power generation for both H2/O2 and direct methanol fuel cells operated at ~240 °C was demonstrated for this electrolyte when stabilized with water partial pressures of ~0.3−0.4 atm. In fact, high performances, corresponding to single cell peak power densities of 415 mW·cm−2, was achieved for a humidified H2/O2 system provided with a CsH2PO4 electrolyte membrane of only 25 µm in thickness [126].
Various strategies, including salt modification by cation and anion substitution [127,128,129,130], and mixing with oxide materials [131,132,133,134,135,136,137] or organic additives [121,138,139,140], have been explored to improve the solid acid performance. The preparation of these derivatives and composites involved different synthesis techniques, such as sol-gel, impregnation, thin-casting and electrospinning, depending on the state of the precursor materials and the desired product [6].
Studies on the effects of incorporating different substituents in the structure of CsH2PO4 have shown mixed results. Thus, while incorporation of Rb+, up to 19%, led to a maximum conductivity of 0.03 S·cm−1, at 240 °C, Ba2+-doping only increased the conductivity of the low temperature phase in two orders of magnitude [127]. On the other hand, substitution of phosphorus with Mo or W hardly modified the proton conductivity, as compared to the parent compound. For S-doping, a shift to lower temperatures of the phase transition was observed at low S-contents, while the conductivity increased only for the low temperature phase.
Mixing CsH2PO4 with H3PO4 or CsH5(PO4)2 resulted in composites Cs1−xH2+xPO4 (xmole= 0.01−0.05) with excesses of protons. In comparison with CsH2PO4, these composites exhibited super protonic phase transition at lower temperature (up to ~200 °C) and increased lower-temperature conductivities (σ150 °C~ 2 × 10−3 S·cm−1, x = 0.1) [141]. Addition of Cs2HPO4·2H2O to CsH2PO4 (up to an xmole = 0.03) also resulted in an enhanced low-temperature conductivity and decreased the activation energy to Ea = 0.56 eV. For all these composites no significant changes in the high-temperature conductivity were observed, although the super protonic phase transition became more diffuse and almost disappeared in some case [141,142].
Improvements in the mechanical and proton conductivity properties, as well as in thermal stability of CsH2PO4-based electrolytes have been addressed by mixing with oxide materials, such as zirconia, silica, alumina, and titania (Table 1). Thus, the (1-x)Cs3(HSO4)2(H2PO4)/xSiO2 composite with x = 0.7 increase the proton conductivity up to 10−2 S·cm−1 in the range 60–200 °C. Moreover, the introduction of fine-particle silica reduced the jump in conductivity at the phase-transition temperature and shifted it to lower temperatures. Higher silica contents led to a decrease in the conductivity due to the disruption of conduction paths [132].
The use of acid-modified silica confers high thermal stability at low H2O partial pressure while maintaining high proton conductivity (10−3–10−2 S·cm−1, at 130–250 °C) [134]. A similar trend was found for the composites (1−x)CsH2PO4/xTiO2 and (1−x)CsH2PO4/xZrO2 (contents x = 0.1 and 0.2) [135,136]. High performance was also reported for the composite 8:1:1 CsH2PO4/NaH2PO4/ZrO2, with a stable conductivity of 2.23 × 10−2 S·cm−1 for 42 h being measured for the high temperature phase [137]. Nanodiamonds (ND) are another type of heterogeneous additive giving rise to the (1−x)CsH2PO4-xND (x = 0–0.5) composites exhibiting enhanced low temperature proton conductivities, while maintaining almost unaltered that of high temperature [140].
CsH2PO4 composites based on organic additives have been also intensively investigated. For example, binary mixtures containing N-heterocycles (1,2,4-triazole, benzimidazole and imidazole) displayed enhanced proton conductivity at temperatures below the super protonic phase transition (2 − 8 × 10−4 S·cm−1 at 174–190 °C) [138]. Another approach was reported in which the solid acid CsH2PO4 was combined with fluoroelastomer p(VDF/HFP), producing high conductive composite membranes (1−x)CsH2PO4-xp(VDF/HPF) (x = 0.05–0.25 wt%) with improved mechanical and hydrophobic properties, along with flexibility and reduced thickness.
However, a high concentration of p(VDF/HFP) rendered membranes with a reduced proton conductivity for the HT phase [141]. By using the polymer butvar (polyvinyl butyral) [121], composite membranes (1−x)CsH2PO4-xButvar (x < 0.2 wt%) were obtained showing, in a wide range of composition, a proton conduction behaviour analogous to the pure salt in the high temperature region but with increased low temperature conductivity by three orders of magnitude at x = 0.2.
Table 1. Proton conductivity data for selected monovalent and tetravalent metal phosphates and pyrophosphates.
Table 1. Proton conductivity data for selected monovalent and tetravalent metal phosphates and pyrophosphates.
Compounds/DimensionalityTemperature (°C)/RH(%)Conductivity (S·cm−1)Ea (eV)Ref.
Tetravalent Metal Phosphates
ZrP·0.8PrNH2·5H2O/2D20/901.2 × 10−31.04[50]
Zr(P2O7)0.81(O3POH)0.38/2D20/901.3 × 10−30.19[51]
Zr(O3POH)0.65(O3PC6H4SO3H)1.35/2D100/907.0 × 10−2---[52]
(NH4)2[ZrF2(HPO4)2]/3D90/951.45 × 10−20.19[53]
(NH4)5[Zr3(OH)3F6(PO4)2(HPO4)]/3D60/984.41 × 10−20.33[54]
(NH4)3Zr(H2/3PO4)3/1D90/951.21 × 10−20.30[55]
Ti2(HPO4)4/1D20/951.2 × 10−30.13[83]
Ti2O(PO4)2·2H2O (π-TiP)/3D90/951.3 × 10−30.23[86]
Ti(HPO4)1(O3PC6H4SO3H)0.85(OH)0.30·nH2O/2D100/--0.10.18[117]
Sn(HPO4)2·3H2O/2D100/951.0 × 10−2---[88]
α-ZrP2O7/3D300 1.0 × 10−4---[49]
(NH4)3Zr(H2/3PO4)3/1D180 1.45 × 10−30.26[55]
Tetravalent Pyrophosphates
TiP2O7/3D100/1004.4 × 10−30.14[116]
(C6H14N2)[NiV2O6H8(P2O7)2]·2H2O/3D60/1002.0 × 10−20.38[118]
In0.1Sn0.9P2O7/3D3000.195---[108]
Ce0.9Mg0.1P2O7/3D200 4.0 × 10−2---[115]
CeP2O7/3D180 3.0 × 10−2---[115]
Super protonic Cesium Phosphates
CsH2PO4/3D>230 6.0 × 10−2---[120]
Cs1-xRbxH2PO4/3D240 3.0 × 10−20.92[127]
Cs1−xH2+xPO4/3D150 2.0 × 10−20.70[141]
(1−x)Cs3(HSO4)2(HPO4)/xSiO2200 1.0 × 10−2---[132]
(1-x)CsH2PO4/xTiO2230 2.0 × 10−2---[135]
(1-x)CsH2PO4/xZrO2250 2.6 × 10−2---[136]
CsH2PO4/NaH2PO4/ZrO2230 2.23 × 10−2---[137]

4. Divalent and Trivalent Metal Phosphates

4.1. Divalent Metal Phosphates

Divalent transition metal phosphates show a great structural versatility, from 1D polymeric topologies through layered framework to 3D open-framework structures. Most of these solids are synthetized in the presence of organic molecules, which are retained as protonated guest species (amines, iminazole derivatives, etc.), thus, compensating the anionic charge of the inorganic framework. This is formed by the metal ion, mainly in octahedral or tetrahedral coordination environments, linked to the phosphate groups with different protonated degrees (HxPO4). The presence of the latter makes possible the formation of effective and extensive hydrogen bond networks with participation of water molecules. In addition, protonated guest species and water itself can act as proton carriers, thus, boosting proton conduction (Table 2) [12,13,14].
Several 3D open-framework M(II) phosphates have been reported [14,143], which consist of [CoPO4] or [Zn2(HPO4)2(H2PO4)2]2− anionic frameworks that contain organic charge-compensating ions in their internal cavities. (C2N2H10)0.5CoPO4 exhibited negligible conductivity in anhydrous conditions; however, it displayed a relatively high water-mediated proton conductivity 2.05 × 10−3 S·cm−1 at 56 °C and 98% RH. On the other hand, the solid NMe4·Zn[HPO4][H2PO4] experiences a structural transformation from monoclinic (α) to orthorhombic (β) upon heating at 149 °C. Both polymorphs contain 12-membered rings composed of tetrahedral Zn2+ ions linked to protonated phosphate groups without changing Zn-O-P connectivity (Figure 8).
The α phase transforms into the β phase at high humidity and temperatures above 60 °C, and then reaches a proton conductivity of 1.30 × 10−2 S·cm−1 at 98% RH, a behaviour that might be attributed to the participation of adsorbed water molecules in creating H-bonding networks with effective pathways for proton conduction. The conductivity drastically decreases at 65 °C. In anhydrous conditions, the α phase exhibited a proton conductivity of ~10−4 S·cm−1 at 160 °C, similar values were found for other reported zinc phosphates at temperatures between 130 and 190 °C [12,144,145].
Membrane-electrode assembly prepared with the pelletized solid, gave an observed open circuit voltage (OCV) of 0.92 V at 190 °C measured in a H2/air cell, suggesting that the dominant conductive species are protons, and the proton transport from anode to cathode takes place through H-bond networks in the pellet.
As an example of 2D metal phosphate water-assisted proton conductors, (C2H10N2) [Mn2(HPO4)3](H2O), displayed a proton conductivity of 1.64 × 10−3 S·cm−1 under 99% RH at 20 °C. This proton conductivity was attributed to the formation of dense H-bond networks in the lattice, composed of Mn3O13 units-containing anionic layers [146], which provide efficient proton-transfer pathways for a Grotthuss-type proton transport at high RH.
Another way of improving the proton conductivity in layered divalent metal phosphates is favouring the formation of hydrogen bond networks by ion exchange. For instance [17], partial exchange of Na+ for ethylendiammonium yielded two new crystalline phases, with composition (C2H10N2)xNa1−x[Mn2(PO4)2] (x = 0.37 y 0.54), which influenced formation of extended hydrogen bond networks and concomitant increase in proton conductivity, from 2.22 × 10−5 S·cm−1 for the as-synthesized material (in ethylendiammonium form) to 1.3 × 10−2 S·cm−1 (x = 0.37) and 2.1 × 10−2 S·cm−1 (x = 0.54) at 99% RH and 30 °C.
This strategy of enhancing proton conductivity was further extended successfully to other 2D manganese phosphates [18]. A proton conductivity value as high as 7.72 × 10−2 S·cm−1 was reached at 30 °C and 99% RH for K+-exchanged compounds, which compares well with those of MOF-based open-framework materials [147,148,149]. The 1D solid [Zn3(H2PO4)6(H2O)3](Hbim) (Hbim= benzimidazole) prepared by mechanochemical synthesis is characterized by presenting a dual-function as proton conductor.
It loses the coordinated water by heating transforming into [Zn3(HPO4)6](Hbim), which exhibits an intrinsic proton conductivity higher than the hydrated form, reaching a value of 1.3 × 10−3 S·cm−1 at 120 °C, believed to be due to a rearrangement of the conduction path and the liquid-like behaviour of benzimidazole molecules. In addition, this solid also showed porosity, thus, enabling the adsorption of gaseous methanol that further improved the proton conductivity of the anhydrous phase. This enhancement of proton conductivity in methanol-adsorbed samples was explained by the effective participation of the guest molecule in formation of extended hydrogen-bond interactions [19].
An ordered-to-disordered structural transformation and its implication in proton conduction were investigated for the 1D copper phosphate [ImH2][Cu(H2PO4)2Cl]·H2O (Im = imidazole). In this structure, the protonated imidazole (ImH2) and the water molecule are located in interspaces of the anionic chains [Cu(H2PO4)2Cl]. Upon heating a structural transformation from an ordered crystalline state to a disordered state occurred. Highly mobile and structurally disordered H+ carriers were supposed to be responsible of the high proton conductivity 2 × 10−2 S·cm−1 at 130 °C, under anhydrous conditions [150].
A 1D zinc phosphate-based proton conductor, [Zn3(H2PO4)6(H2O)3](BTA) (BTA = 1,2,3-benzotriazole) has been reported [151] that exhibits high proton conductivity, 8 × 10−3 S·cm−1 in anhydrous glassy-state (120 °C). The glassy-state, developed via melt-quenching, was suggested to induce isotropic disordered domains that enhanced H+ dynamics and conductive interfaces. In fact, the capability of the glassy-state material as an electrolyte was found suitable for the rechargeable all-solid-state H+ battery operated in a wide range of temperatures from 25 to 110 °C.
Focus has been also put on metal phosphate-based solid solutions [152,153]. Vacancies can be generated that introduce extra protons into the structure and increase the proton conduction. This effect was investigated for the 1D rubidium and magnesium polyphosphate compound, RbMg1-xH2x(PO3)3·yH2O, for which system a maximum proton conductivity of 5.5 × 10−3 S·cm−1 was measured at 170 °C with a vehicle-type mechanism of H3O+ conduction.
The proton conductivity results from H-bond interactions between water molecules and corner-sharing PO4 chains that provide formation of sandwiched edge-sharing RbO6-MgO6 chain [152]. Another example is the solid solution with composition Co1-xZnx(H2PO4)2·2H2O (0 < x < 1.0), which showed the highest conductivity value, 2.01 × 10−2 S·cm−1 at 140 °C, for a composition Co0.5Zn0.5(H2PO4)2·2H2O [153].

4.2. Trivalent Metal Phosphates

Zeolite-like open framework metal(III) phosphates consist of metal(III) ions-phosphate species (PO43-, HPO42-, H2PO4) linkages featuring internal cavities, where charge-compensating cations and/or neutral species are located and, thus, dense hydrogen bond networks frequently result. In addition, the robust inorganic framework endues this porous material with better thermal and chemical stability compared with porous coordination polymers/metal organic frameworks (PCPs/MOFs) [13].
Regarding to proton conductivity (Table 2), aluminium phosphate-based solids are by far the most studied compounds [154,155,156,157,158,159,160,161,162,163,164,165,166,167]. The species inside channels affect in different ways to proton conduction. Thus, while water adsorption is key to assist proton transfer in (NH4)2Al4(PO4)4(HPO4)·H2O by a hopping mechanism along H-bond chains [166], densely packed NH4+ ions show negligible contribution because of hampered migration.
By using an organic template-free synthetic methodology, a 3D open-framework aluminophosphate Na6[(AlPO4)8(OH)6]·8H2O (JU103) was prepared [158], which showed a proton conductivity of 3.59 × 10−3 S·cm−1, at 20 °C and 98% RH. It was argued that the enhanced conductivity of the as-synthesized material as compared to its NH4+- or Ag+-exchanged forms is indicative of a beneficial effect of hydrated Na+ ions in generating proton-transfer pathways. The 3D cesium silicoaluminophosphate, Cs2(Al0.875Si0.125)4(P0.875Si0.125O4)4(HPO4), belonging to the structural family of SAPOs, was shown to exhibit a remarkable proton conductivity of 1.70 × 10−4 S·cm−1 at low temperature and RH (20 °C and 30%, respectively) [168,169].
Several 2D aluminophosphates have been reported as proton conductors [154,156]. These compounds (denoted as AlPO-CJ70/2) are structurally characterized by displaying an anionic layer: [Al2P3O12]3-, formed by alternating Al3+ and phosphorus tetrahedra, which is charge-compensated by N,N-dimethylbenzylamine or α-methylbenzylamine ions, respectively. In these layered structures, extended H-bond networks are formed through interactions of the amine N atoms, H2O molecules and protruding phosphate groups of the anionic layer. Consequently, water-mediated proton conduction processes occurred upon immersion in water, with σ values around 10−3 S·cm−1, at 80 °C, and Ea values of 0.16−0.2 eV, typical of a Grotthuss-type proton-transfer mechanism.
By following a synthetic route in which methylimidazolium dihydrogenphosphate was used as a solvent, structure-directing agent, and a phosphorus source, the solid (C4H7N2)(C3H4N2)2·Al3(PO4)4·0.5H2O (SCU−2) was obtained [164]. Its layered structure, built up from corner-sharing Al3+ and PV tetrahedra, features 8-member rings where guest imidazolium ions and water molecules are hosted. At 85 °C and 98% RH, this solid showed a proton conductivity of 5.9 × 10−3 S·cm−1 and a low activation energy (0.20 eV), characteristic of a Grotthuss-type proton-transfer mechanism. An efficient pathway for the proton transfer was attributed to the hydrogen bond network established between the imidazolium ions and water molecules interacting with the host framework.
A few examples of phosphate-based proton conductors of other trivalent metals do exist. Among them, two Fe(III) phosphates, 1D (C4H12N2)1.5[Fe2(OH)(H2PO4)(HPO4)2(PO4)]·0.5H2O [13] and 3D open-framework iron(III) phosphate (NH3(CH2)3NH3)2[Fe4(OH)3(HPO4)2(PO4)3]·4H2O [170], have been reported. Both compounds contain Fe4O20 tetramers as a common structural feature. The 1D solid is composed of chains of tetramers bridged by PO43- groups and having terminal H2PO4 and HPO42− groups [171], while piperazinium cations and water molecules are disorderly situated in between chains. This arrangement gives rise to extended hydrogen bonding interactions and hence proton conducting pathways. The proton conductivity measured at 40 °C and 99% RH was 5.14 × 10−4 S·cm−1, and it was maintained upon dispersion of this solid in PVDF [13]. In the case of the 3D solid, infinite chains of interconnected tetramers are interlinked, in turn by phosphate groups that generate large tunnels (Figure 9). The diprotonated 1,3-diaminopropane and water molecules, localized inside tunnels, form an extended hydrogen bond network with the P–OH groups pointing toward cavities. These interactions favour proton hopping, the measured proton conductivity being of 8.0 × 10−4 S·cm−1, at 44 °C and 99% RH, and with an Ea of 0.32 eV [172]. Furthermore, the proton conductivity of this compound increased up to 5 × 10−2 S·cm−1 at 40 °C upon exposure to aqua-ammonia vapors from 1 M NH3·H2O solution. This result confirms this treatment as an effective way of enhancing proton conductivity, which has been elsewhere demonstrated for the case of coordination polymers [173,174]. The observed variation of Ea with the ammonia concentration also suggested that NH3, as well as H2O, molecules contribute to create proton-transfer pathways, by a Grotthuss mechanism. However, when ammonia concentrations were lower than 0.5 M, the proton conduction mechanism tended to be vehicle-type one [156].
Table 2. Proton conductivity data for selected divalent and trivalent metal phosphates.
Table 2. Proton conductivity data for selected divalent and trivalent metal phosphates.
Compounds/DimensionalityTemperature (°C)/RH(%)Conductivity (S·cm−1)Ea (eV)Ref.
Divalent Metal Phosphates
(C2N2H10)0.5CoPO4/3D56/982.05 × 10−31.01[143]
NMe4·Zn[HPO4][H2PO4] (β phase)/3D60/981.30 × 10−20.92[14]
(C2H10N2) [Mn2(HPO4)3](H2O)/2D20/991.64 × 10−30.22[146]
(C2H10N2)xNa1−x[Mn2(PO4)2]/2D30/992.1 × 10−20.14[17]
(C2H10N2)1-xKx[Mn2(PO4)2]·2H2O/2D30/997.72 × 10−20.18[18]
(C2H10N2)1-xKx[Mn2(HPO4)3] (H2O)/2D30/990.85 × 10−20.081[18]
[Zn3(H2PO4)6(H2O)3](Hbim)/1D120 1.3 × 10−30.50[19]
[ImH2][Cu(H2PO4)2Cl]·H2O/1D130 2.0 × 10−20.1[150]
[Zn3(H2PO4)6(H2O)3](BTA)/1D120 8.0 × 10−30.39[151]
RbMg0.9H0.2(PO3)3·yH2O/1D170 5.5 × 10−3---[152]
Co0.5Zn0.5(H2PO4)2·2H2O/1D140 2.01 × 10−2---[153]
Trivalent Metal Phosphates
Na6[(AlPO4)8(OH)6]·8H2O/3D20/983.59 × 10−30.21[158]
[C9H14N]8[H2O]4·[Al8P12O48H4]/2D80, in water9.25 × 10−40.16[154]
[R-,S-C8H12N]8[H2O]2·[Al8P12O48H4]/2D90/983.01 × 10−30.20[156]
(C4H7N2)(C3H4N2)2·Al3(PO4)4·0.5H2O/2D85/985.94 × 10−30.20[164]
In(HPO4)(H2PO4)(D,L-C3H7NO2)/3D85/982.9 × 10−30.19[16]
(NH3(CH2)3NH3)2[Fe4(OH)3(HPO4)2(PO4)3]·4H2O/1D40/995.0 × 10−2---[170]
Hbim = benzimidazole; Im =imidazole; BTA = 1,2,3-benzotriazole.
For the series of isostructural imidazole cation (ImH2)-templated layered metal phosphates, [ImH2][X-(HPO4)2(H2O)2] (FJU−25-X, X = Al, Ga, and Fe), it was found that the proton conductivity was dependent on mobility of imidazole guests, FJU−25-Fe exhibiting the highest proton conductivity (5.21 × 10−4 S·cm−1 at 90 °C). The determined activation energies (~0.20 eV) were indicative of a Grotthuss-type mechanism of proton conduction.
The amino acid-template indium phosphate, In(HPO4)(H2PO4)(D,L-C3H7NO2) (SCU−12), represents a singular example of 3D metal phosphate-based proton conductors [16].
Its crystal structure is formed by edge-sharing four-ring ladders, with the amino acid molecules attached to the ladders through In–O bonds. Further bridging the indium phosphate ladders by the H2PO4 groups gives rise to a three-dimensional structure. The presence of two kinds of proton carriers, H2PO4 ions and zwitterionic alanine molecules, favours the development of high proton conductivities (2.9 × 10−3 S·cm−1 at 85 °C and 98% RH) through a Grotthuss-type proton-transfer mechanism (Ea = 0.19 eV).
Other trivalent metal phosphates have been reported, e.g., BPOx [175] and CePO4 [176]. The former exhibited a proton conductivity of 7.9 × 10−2 S·cm−1 as self-supported electrolyte and 4.5 × 10−2 S·cm−1 as (PBI)−4BPOx composite membrane, measured at 150 °C and 5% RH, but structure/conductivity correlations were not established because of its amorphous nature. The latter showed a low-temperature (RT) proton conduction < 10−5 S·cm−1 at 100% RH through a structure-independent proton-transport mechanism [176].
There are only a few examples of metal(III) pyrophosphates displaying proton conductivity. Among them is the open framework magnesium aluminophosphate MgAlP2O7(OH)(H2O)2 (JU102) [167]. Its structure is composed of tetrahedral Al3+ and octahedral Mg2+ ions coordinated by pyrophosphate ions. This connectivity results in an open framework with unidirectional 8-ring channels. The proton conduction properties originate from the existence of an H-bond network in which coordinated water molecules participate.
Thus, the proton conductivity measured at 55 °C on water-immersed samples was 3.86 × 10−4 S·cm−1 [167], which raised to 1.19 × 10−3 S·cm−1 when calcined at 250 °C and measured at the same conditions, while the Ea value hardly changed from 0.16 to 0.2 eV. This behaviour was explained as being due to a dehydration–rehydration process that enhances proton conductivity by altering the H-bonding network and the pathway of proton transfer.
Another example of 3D open framework metal(III) pyrophosphate is the compound NH4TiP2O7 [177]. The structure of this solid is composed of negatively charged [TiP4O12] layers, forming one-dimensional six-membered ring channels, where the NH4+ ions are located. Its proton conductivity increased from 10−6 S·cm−1 under anhydrous conditions to 10−3 S·cm−1 at full-hydration conditions and 84 °C. The low Ea value, 0.17 eV, characteristic of a Grotthuss-type proton-transfer mechanism, was associated with the role played by the NH4+ ions in the channels as proton donors and promoters of proton migration. A drop in proton conductivity was observed when the triclinic TiP2O7 phase formed by thermal decomposition of NH4TiP2O7 [177].

5. Outlook

Great efforts have been devoted to the synthesis and characterization of metal phosphate-based proton conductors over more than three decades. Among them, zirconium phosphates are prominent not only because of the feasibility of Zr(IV) and phosphate ions to form a rich variety of crystallographic architectures (from 1D to 3D open frameworks) but also due to their outstanding properties and workability while being environmentally benign and low cost materials.
Although the prototype layered α-zirconium phosphate has been commonly proven as a filler for PEMFCs devices, new synthetic designs of M(IV) phosphates, including pyrophosphate compounds, are promising candidates to broaden their applicability in different electrochemical devices. Other M(IV) phosphates and pyrophosphates (M = Ti, Sn) are less known due, in part, to their amorphous nature or strong tendency to amorphise at working temperatures, though, in some cases, these compounds presented remarkable proton conductivity properties.
CsH2PO4, a super protonic material, was proven as a suitable electrolyte for both H2/O2 and direct methanol fuel cells operated at ~240 °C, and provides excellent performance when controlling its thermal stability. In addition, combinations with other materials are possible to adjust the specific characteristics of the composite CsH2PO4 electrolyte, thus, offering a wide range of compositions with tuning properties.
Recently, research and developments in metal phosphate proton conductors have been addressed to divalent or trivalent metal phosphates, which present a remarkable structural versatility and tunable conductivities; however, more in-depth studies are required to assess their potential use and applicability for low and intermediate temperature fuel cells.
Applications of metal phosphates as electrolytes or as electrolyte components are in continuous progress, although their use for energy storage and conversion remains a challenge. For practical applications in fuel cells at low/intermediate temperatures, phosphate-based proton conducting electrolytes have demonstrated acceptable proton conductivity values; however, other features, such as their mechanical strength, chemical/thermal stabilities, film-forming ability (in the case of composite membranes), durability, and fuel cross-over, are key factors to be improved.

Author Contributions

Writing-original draft preparation, R.M.P.C. and M.B.-G.; writing-review and editing, P.O.-P. and A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by PID2019−110249RB-I00 (MICIU/AEI, Spain) and PY20−00416 (Junta de Andalucia, Spain/FEDER) research projects.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.B.-G. thanks PAIDI2020 research grant (DOC_00272 Junta de Andalucia, Spain) and R.M.P.C. thanks University of Malaga under Plan Propio de Investigación for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shivhare, A.; Kumara, A.; Srivastava, R. Metal phosphate catalysts to upgrade ligno-cellulose biomass into value-added chemicals and biofuels. Green Chem. 2021, 23, 3818–3841. [Google Scholar] [CrossRef]
  2. Zhao, H.; Yuan, Z.-Y. Insights into Transition Metal Phosphate Materials for Efficient Electrocatalysis. ChemCatChem 2020, 12, 3797–3810. [Google Scholar] [CrossRef]
  3. Chen, L.; Zhao, Y.; Yang, J.; Liu, D.; Wei, X.; Wang, X.; Zheng, Y. New Versatile Synthetic Route for the Preparation of Metal Phosphate Decorated Hydrogen Evolution Photocatalysts. Inorg. Chem. 2020, 59, 1566–1575. [Google Scholar] [CrossRef]
  4. Goñi-Urtiaga, A.; Presvytes, D.; Scott, K. Solid acids as electrolyte materials for proton exchange membrane (PEM) electrolysis: Review. Int. J. Hydrog. Energy 2012, 37, 3358–3372. [Google Scholar] [CrossRef]
  5. Paschos, O.; Kunze, J.; Stimming, U.; Maglia, F. A review on phosphate based, solid state, protonic conductors for intermediate temperature fuel cells. J. Phys. Condens. Matter 2011, 23, 234110. [Google Scholar] [CrossRef]
  6. Haile, S.M.; Chisholm, C.R.I.; Sasaki, K.; Boysen, D.A.; Uda, T. Solid acid proton conductors: From laboratory curiosities to fuel cell electrolytes. Faraday Discuss. 2007, 134, 17–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Cheng, Q.; Zhao, X.; Yang, G.; Mao, L.; Liao, F.; Chen, F.; He, P.; Pan, D.; Chen, S. Recent advances of metal phosphates-based electrodes for high-performance metal ion batteries. Energy Storage Mater. 2021, 41, 842–882. [Google Scholar] [CrossRef]
  8. Habraken, W.; Habibovic, P.; Epple, M.; Bohner, M. Calcium phosphates in biomedical applications: Materials for the future? Mater. Today 2015, 19, 69–87. [Google Scholar] [CrossRef]
  9. Mohammad, N.; Mohamad, A.B.; Kadhum, A.A.H.; Loh, K.S. A review on synthesis and characterization of solid acid materials for fuel cell applications. J. Power Sources 2016, 322, 77–92. [Google Scholar] [CrossRef]
  10. Pica, M.; Donnadio, A.; Casciola, M. From microcrystalline to nanosized α-zirconium phosphate: Synthetic approaches and applications of an old material with a bright future. Coord. Chem. Rev. 2018, 374, 218–235. [Google Scholar] [CrossRef]
  11. Wong, N.E.; Ramaswamy, P.; Lee, A.S.; Gelfand, B.S.; Bladek, K.J.; Taylor, J.M.; Spasyuk, D.M.; Shimizu, G.K.H. Tuning Intrinsic and Extrinsic Proton Conduction in Metal−Organic Frameworks by the Lanthanide Contraction. J. Am. Chem. Soc. 2017, 139, 14676–14683. [Google Scholar] [CrossRef] [PubMed]
  12. Horike, S.; Umeyama, D.; Inukai, M.; Itakura, T.; Kitagawa, S. Coordination-network-based ionic plastic crystal for anhydrous proton conductivity. J. Am. Chem. Soc. 2012, 134, 7612–7615. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, K.-M.; Lou, Y.-L.; He, F.-Y.; Duan, H.-B.; Huang, X.-Q.; Fan, Y.; Zhao, H.-R. The water-mediated proton conductivity of a 1D open framework inorganic-organic hybrid iron phosphate and its composite membranes. Inorg. Chem. Commun. 2021, 134, 109032. [Google Scholar] [CrossRef]
  14. Yu, J.-W.; Yu, H.-J.; Ren, Q.; Zhang, J.; Zou, Y.; Luo, H.-B.; Wang, L.; Ren, X.-M. Humidity-sensitive irreversible phase transformation of open-framework zinc phosphate and its water-assisted high proton conduction properties. Dalton Trans. 2021, 50, 8070–8075. [Google Scholar] [CrossRef] [PubMed]
  15. Su, X.; Yao, Z.; Ye, Y.; Zeng, H.; Xu, G.; Wu, L.; Ma, X.; Chen, Q.-H.; Wang, L.; Zhang, Z.; et al. 40-Fold Enhanced Intrinsic Proton Conductivity in Coordination Polymers with the Same Proton-Conducting Pathway by Tuning Metal Cation Nodes. Inorg. Chem. 2016, 55, 983–986. [Google Scholar] [CrossRef]
  16. Shi, J.; Wang, K.; Li, J.; Zeng, H.; Zhang, Q.; Lin, Z. Exploration of new water stable proton-conducting materials in an amino acid-templated metal phosphate system. Dalton Trans. 2018, 47, 654–658. [Google Scholar] [CrossRef]
  17. Zhang, K.-M.; He, F.-Y.; Duan, H.-B.; Zhao, H.-R. An alkali metal ion-exchanged metal-phosphate (C2H10N2)xNa1−x[Mn2(PO4)2] with high proton conductivity of 10−2 S·cm−1. Inorg. Chem. 2019, 58, 6639–6646. [Google Scholar] [CrossRef]
  18. Zhang, K.-M.; Jia, Y.; Gu, Y.; He, F.-Y.; Zhao, H.-R. A facile and efficient method to improve the proton conductivity of open framework metal phosphates under aqueous condition. Inorg. Chem. Commun. 2020, 120, 108128. [Google Scholar] [CrossRef]
  19. Umeyama, D.; Horike, S.; Inukai, M.; Kitagawa, S. Integration of Intrinsic Proton Conduction and Guest-Accessible Nanospace into a Coordination Polymer. J. Am. Chem. Soc. 2013, 135, 11345–11350. [Google Scholar] [CrossRef]
  20. Baranov, A.I.; Khiznichenko, V.P.; Sandler, V.A.; Shuvalov, L.A. Frequency dielectric dispersion in the ferroelectric and superionic phases of CsH2PO4. Ferroelectrics 1988, 81, 183–186. [Google Scholar] [CrossRef]
  21. Colomban, P. Chemistry of Solid-State Materials. In Proton Conductors: Solid, Membranes and Gels-Materials and Devices; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  22. Fragua, D.M.; Castillo, J.; Castillo, R.; Vargas, R.A. New amorphous phase KnH2PnO3n+1(n>>1) in KH2PO4. Rev. Latin Am. Metal. Mat. 2009, 2, 491–497. [Google Scholar]
  23. Skou, E.; Andersen, I.G.K.; Simonsen, K.E.; Andersen, E.K. Is UO2HPO4·4H2O a proton conductor? Solid State Ion. 1983, 9, 1041–1047. [Google Scholar] [CrossRef]
  24. Cabeza, A.; Martinez, M.; Benavente, J.; Bruque, S. Current rectification by H3OUO2PO4 3H2O (HUP) thin films in electrolyte media. Solid State Ion. 1992, 51, 127–131. [Google Scholar] [CrossRef]
  25. Barboux, P.; Morineau, R.; Livage, J. Protonic conductivity in hydrates. Solid State Ion. 1988, 27, 221–225. [Google Scholar] [CrossRef]
  26. Barboux, P.; Livage, J. Ionic conductivity in fibrous Ce(HPO4)2·(3+x)H2O. Solid State Ion. 1989, 34, 47–52. [Google Scholar] [CrossRef]
  27. Li, J.; Yi, M.; Zhang, L.; You, Z.; Liu, X.; Li, B. Energy related ion transports in coordination polymers. Nano Select. 2021, 1–19. [Google Scholar] [CrossRef]
  28. Steele, B.C.H.; Heinzel, A. Materials for fuel-cell technologies. Nature 2001, 414, 245–260. [Google Scholar] [CrossRef]
  29. Wei, J. Proton-Conducting Materials Used as Polymer Electrolyte Membranes in Fuel Cells (ch 9). In Polymer-Based Multifunctional Nanocomposites and Their Applications; Song, K., Liu, C., Guo, J.Z., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; Chapter 9; pp. 345–352. [Google Scholar]
  30. Sazali, N.; Salleh, W.N.W.; Jamaludin, A.S.; Razali, M.N.M. New Perspectives on Fuel Cell Technology: A Brief Review. Membranes 2020, 10, 99. [Google Scholar] [CrossRef] [PubMed]
  31. Dupuis, A.-C. Proton exchange membranes for fuel cells operated at medium temperatures: Materials and experimental techniques. Prog. Mater. Sci. 2011, 56, 289–327. [Google Scholar] [CrossRef]
  32. Peighambardoust, S.J.; Rowshanzamir, S.; Amjadi, M. Review of the proton exchange membranes for fuel cell applications. Int. J. Hydrog. Energy 2010, 35, 9349–9384. [Google Scholar] [CrossRef]
  33. Mauritz, K.A.; Moore, R.B. State of understanding of Nafion. Chem. Rev. 2004, 104, 4535–4585. [Google Scholar] [CrossRef] [PubMed]
  34. Hickner, M.A.; Ghassemi, H.; Kim, Y.S.; Einsla, B.R.; McGrath, J.E. Alternative polymer systems for proton exchange membranes (PEMs). Chem. Rev. 2004, 104, 4587–4612. [Google Scholar] [CrossRef] [PubMed]
  35. Han, X.; Xie, Y.; Liu, D.; Chen, Z.; Zhang, H.; Pang, J.; Jiang, Z. Synthesis and properties of novel poly(arylene ether)s with densely sulfonated units based on carbazole derivative. J. Membr. Sci. 2019, 589, 117230. [Google Scholar] [CrossRef]
  36. Kuzmenko, M.; Poryadchenko, N. Perspective materials for application in fuel-cell technologies. In Fuel Cell Technologies: State and Perspectives; Sammes, N., Ed.; Springer: Dordrecht, The Netherlands, 2005; pp. 253–258. [Google Scholar]
  37. Lee, K.-S.; Maurya, S.; Kim, Y.S.; Kreller, C.R.; Wilson, M.S.; Larsen, D.; Elangovan, S.E.; Mukundan, R. Intermediate temperature fuel cells via an ion-pair coordinated polymer electrolyte. Energy Env. Sci. 2018, 11, 979–987. [Google Scholar] [CrossRef]
  38. Norby, T. Solid-state protonic conductors: Principles, properties, progress and prospects. Solid State Ion. 1999, 125, 1–11. [Google Scholar] [CrossRef]
  39. Li, Q.; He, R.; Jensen, J.O.; Bjerrum, N.J. Approaches and Recent Development of Polymer Electrolyte Membranes for Fuel Cells Operating above 100 °C. Chem. Mater. 2003, 15, 4896–4915. [Google Scholar] [CrossRef] [Green Version]
  40. Loreti, G.; Facci, A.L.; Ubertini, S. High-Efficiency Combined Heat and Power through a High-Temperature Polymer Electrolyte Membrane Fuel Cell and Gas Turbine Hybrid System. Sustainability 2021, 13, 12515. [Google Scholar] [CrossRef]
  41. Boysen, D.A.; Uda, T.; Chisholm, C.R.I.; Haile, S.M. High-Performance Solid Acid Fuel Cells Through Humidity Stabilization. Science 2004, 303, 68–70. [Google Scholar] [CrossRef] [Green Version]
  42. Zhang, J.; Aili, D.; Lu, S.; Li, Q.; Jiang, S.P. Advancement toward Polymer Electrolyte Membrane Fuel Cells at Elevated Temperatures. AAAS Res. 2020, 2020, 9089405. [Google Scholar] [CrossRef]
  43. Clearfield, A.; Smith, S.D. The crystal structure of zirconium phosphate and the mechanism of its ion exchange behavior. J. Colloid Interface Sci. 1968, 28, 325–330. [Google Scholar] [CrossRef]
  44. Clearfield, A.; Smith, G.D. Crystallography and structure of alpha-zirconium bis(monohydrogen orthophosphate) monohydrate. Inorg. Chem. 1969, 8, 431–436. [Google Scholar] [CrossRef]
  45. Troup, J.M.; Clearfield, A. Mechanism of ion exchange in zirconium phosphates. 20. Refinement of the crystal structure of .alpha.-zirconium phosphate. Inorg. Chem. 1977, 16, 3311–3314. [Google Scholar] [CrossRef]
  46. Casciola, M. From layered zirconium phosphates and phosphonates to nanofillers for ionomeric membranes. Solid State Ion. 2019, 336, 1–10. [Google Scholar] [CrossRef]
  47. Ogawa, T.; Ushiyam, H.; Lee, J.-M.; Yamaguchi, T.; Yamashita, K. Theoretical Studies on Proton Transfer among a High Density of Acid Groups: Surface of Zirconium Phosphate with Adsorbed Water Molecules. J. Phys. Chem. C 2011, 115, 5599–5606. [Google Scholar] [CrossRef]
  48. Alberti, G.; Casciola, M.; Costantino, U. Inorganic ion-exchange pellicles obtained by delamination of α-zirconium phosphate crystals. J. Colloid Interface Sci. 1985, 107, 256–263. [Google Scholar] [CrossRef]
  49. Alberti, G.; Casciola, M.; Costantino, U.; Leonardi, M. AC conductivity of anhydrous pellicular zirconium phosphate in hydrogen form. Solid State Ion. 1984, 14, 289–295. [Google Scholar] [CrossRef]
  50. Casciola, M.; Costantino, U.; D’Amico, S. Protonic conduction of intercolation compounds of α-zirconium phosphate with propylamine. Solid State Ion. 1986, 22, 127–133. [Google Scholar] [CrossRef]
  51. Alberti, G.; Casciola, M.; Cavalaglio, S.; Vivani, R. Proton conductivity of mesoporous zirconium phosphate pyrophosphate. Solid State Ion. 1999, 125, 91–97. [Google Scholar] [CrossRef]
  52. Alberti, G.; Casciola, M.; Donnadio, A.; Piaggio, P.; Pica, M.; Sisani, M. Preparation and characterisation of α-layered zirconium phosphate sulfophenylenphosphonates with variable concentration of sulfonic groups. Solid State Ion. 2005, 176, 2893–2898. [Google Scholar] [CrossRef]
  53. Gui, D.; Zheng, T.; Xie, J.; Cai, Y.; Wang, Y.; Chen, L.; Diwu, J.; Chai, Z.; Wang, S. Significantly dense two-dimensional hydrogen-bond network in a layered zirconium phosphate leading to high proton conductivities in both water-assisted low-temperature and anhydrous intermediate-temperature regions. Inorg. Chem. 2016, 55, 12508–12511. [Google Scholar] [CrossRef]
  54. Yu, J.-W.; Yu, H.-J.; Yao, Z.-Y.; Li, Z.-H.; Ren, Q.; Luo, H.-B.; Zou, Y.; Wang, L.; Ren, X.-M. A water-stable open-framework zirconium(IV) phosphate and its water-assisted high proton conductivity. CrystEngComm 2021, 23, 6093–6097. [Google Scholar] [CrossRef]
  55. Gui, D.; Dai, X.; Tao, Z.; Zheng, T.; Wang, X.; Silver, M.A.; Shu, J.; Chen, L.; Wang, Y.; Zhang, T.; et al. Unique proton transportation pathway in a robust inorganic coordination polymer leading to intrinsically high and sustainable anhydrous proton conductivity. J. Am. Chem. Soc. 2018, 140, 6146–6155. [Google Scholar] [CrossRef]
  56. Kusoglu, A.; Weber, A.Z. New Insights into Perfluorinated Sulfonic-Acid Ionomers. Chem. Rev. 2017, 117, 987–1104. [Google Scholar] [CrossRef]
  57. Alberti, G.; Casciola, M.; Capitani, D.; Donnadio, A.; Narducci, R.; Pica, M.; Sganappa, M. Novel Nafion–zirconium phosphate nanocomposite membranes with enhanced stability of proton conductivity at medium temperature and high relative humidity. Electrochim. Acta 2007, 52, 8125–8132. [Google Scholar] [CrossRef]
  58. Grot, W.G.; Rajendran, G.; Hendrickson, J.S. Membranes Containing Inorganic Fillers and Membranes and Electrode Assemblies and Electrochemical Cells Employing Same. International Publication No. WO 96/2975, 26 September 1996. [Google Scholar]
  59. Alberti, G.; Casciola, M.; Pica, M.; Tarpanelli, T.; Sganappa, M. New Preparation Methods for Composite Membranes for Medium Temperature Fuel Cells Based on Precursor Solutions of Insoluble Inorganic Compounds. Fuel Cells 2005, 5, 366–374. [Google Scholar] [CrossRef]
  60. Casciola, M.; Bagnasco, G.; Donnadio, A.; Micoli, L.; Pica, M.; Sganappa, M.; Turco, M. Conductivity and Methanol Permeability of Nafion–Zirconium Phosphate Composite Membranes Containing High Aspect Ratio Filler Particles. Fuel Cells 2009, 9, 394–400. [Google Scholar] [CrossRef]
  61. Arbizzani, C.; Donnadio, A.; Pica, M.; Sganappa, M.; Varzi, A.; Casciola, M.; Mastragostino, M. Methanol permeability and performance of Nafion–zirconium phosphate composite membranes in active and passive direct methanol fuel cells. J. Power Sources 2010, 195, 7751–7756. [Google Scholar] [CrossRef]
  62. Pica, M.; Donnadio, A.; Casciola, M.; Cojocaru, P.; Merlo, L. Short side chain perfluorosulfonic acid membranes and their composites with nanosized zirconium phosphate: Hydration, mechanical properties and proton conductivity. J. Mater. Chem. 2012, 22, 24902–24908. [Google Scholar] [CrossRef]
  63. Pica, M.; Donnadio, A.; Capitani, D.; Vivani, R.; Troni, E.; Casciola, M. Advances in the Chemistry of Nanosized Zirconium Phosphates: A New Mild and Quick Route to the Synthesis of Nanocrystals. Inorg. Chem. 2011, 50, 11623–11630. [Google Scholar] [CrossRef] [PubMed]
  64. Bauer, F.; Willert-Porada, M. Comparison between Nafion and a Nafion zirconium phosphate nanocomposite in fuel cell applications. Fuel Cells 2006, 6, 261–269. [Google Scholar] [CrossRef]
  65. Casciola, M.; Cojocaru, P.; Donnadio, A.; Giancola, S.; Merlo, L.; Nedellec, Y.; Pica, M.; Subianto, S. Zirconium phosphate reinforced short side chain perflurosulfonic acid membranes for medium temperature proton exchange membrane fuel cell application. J. Power Sources 2014, 262, 407–413. [Google Scholar] [CrossRef]
  66. Yang, C.; Srinivasan, S.; Aricò, A.S.; Creti, P.; Baglio, V.; Antonucci, V. Composition Nafion/zirconium phosphate membranes for direct methanol fuel cell operation at high temperature. Electrochem. Solid-State Lett. 2001, 4, A31–A34. [Google Scholar] [CrossRef]
  67. Yang, C.; Costamagna, P.; Srinivasan, S.; Benziger, J.; Bocarsly, A.B. Approaches and technical challenges to high temperature operation of proton exchange membrane fuel cells. J. Power Sources 2001, 103, 1–9. [Google Scholar] [CrossRef] [Green Version]
  68. Costamagna, P.; Yang, C.; Bocarsly, A.B.; Srinivasan, S. Nafion (R) 115/zirconium phosphate composite membranes for operation of PEMFCs above 100 °C. Electrochim. Acta 2002, 47, 1023–1033. [Google Scholar] [CrossRef]
  69. Yang, C.; Srinivasan, S.; Bocarsly, A.B.; Tulyani, S.; Benziger, J.B. A comparison of physical properties and fuel cell performance of Nafion and zirconium phosphate/Nafion composite membranes. J. Membr. Sci. 2004, 237, 145–161. [Google Scholar] [CrossRef]
  70. Lee, H.-K.; Kim, J.-I.; Park, J.-H.; Lee, T.-H. A study on self-humidifying PEMFC using Pt-ZrP-Nafion composite membrane. Electrochim. Acta 2004, 50, 761–768. [Google Scholar] [CrossRef]
  71. Escorihuela, J.; Narducci, R.; Compañ, V.; Costantino, F. Proton Conductivity of Composite Polyelectrolyte Membranes with Metal-Organic Frameworks for Fuel Cell Applications. Adv. Mater. Interfaces 2019, 6, 1801146. [Google Scholar] [CrossRef]
  72. Liu, K.L.; Lee, H.C.; Wang, B.Y.; Lue, S.J.; Lu, C.Y.; Tsai, L.D.; Fang, J.; Chao, C.Y. Sulfonated poly(styrene-block-(ethylene-ran-butylene)-block-styrene (SSEBS)-zirconium phosphate) (ZrP) composite membranes for direct methanol fuel cells. J. Membr. Sci. 2015, 495, 110–120. [Google Scholar] [CrossRef]
  73. Hu, H.; Ding, F.; Ding, H.; Liu, J.; Xiao, M.; Meng, Y.; Sun, L. Sulfonated poly(fluorenyl ether ketone)/Sulfonated α-zirconium phosphate Nanocomposite membranes for proton exchange membrane fuel cells. Adv. Compos. Hybrid Mater. 2020, 3, 498–507. [Google Scholar] [CrossRef]
  74. Pandey, J.; Seepana, M.M.; Shukla, A. Zirconium phosphate based proton conducting membrane for DMFC application. Int. J. Hydrog. 2015, 40, 9410–9421. [Google Scholar] [CrossRef]
  75. He, R.; Li, Q.; Xiao, G.; Bjerrum, N.J. Proton conductivity of phosphoric acid doped polybenzimidazole and its composites with inorganic proton conductors. J. Membr. Sci. 2003, 226, 169–184. [Google Scholar] [CrossRef]
  76. Gouda, M.H.; Tamer, T.M.; Konsowa, A.H.; Farag, H.A.; Mohy Eldin, M.S. Organic-Inorganic Novel Green Cation Exchange Membranes for Direct Methanol Fuel Cells. Energies 2021, 14, 4686. [Google Scholar] [CrossRef]
  77. Alberti, G.; Cardini-Galli, P.; Costantino, U.; Torracca, E. Crystalline insoluble salts of polybasic metals—I Ion-exchange properties of crystalline titanium phosphate. J. Inorg. Nucl. Chem. 1967, 29, 571–578. [Google Scholar] [CrossRef]
  78. Christensen, A.N.; Anderson, E.K.; Andersen, I.G.; Alberti, G.; Nielsen, M.; Lehmann, E.K. X-Ray Powder diffraction study of layer compounds. The crystal structure of α-Ti (HPO4)2·H2O and a proposed structure for γ-Ti (H2PO4)(PO4)·2H2O. Acta Chem. Scand. 1990, 44, 865–872. [Google Scholar] [CrossRef] [Green Version]
  79. Li, Y.J.; Whittingham, M.S. Hydrothermal synthesis of new metastable phases: Preparation and intercalation of a new layered titanium phosphate. Solid State Ion. 1993, 63, 391–395. [Google Scholar] [CrossRef]
  80. Kőrösi, L.; Papp, S.; Dékány, I. A layered titanium phosphate Ti2O3(H2PO4)2·2H2O with rectangular morphology: Synthesis, structure, and cysteamine intercalation. Chem. Mater. 2010, 22, 4356–4363. [Google Scholar] [CrossRef]
  81. Ekambaram, S.; Serre, C.; Férey, G.; Sevov, S.C. Hydrothermal synthesis and characterization of an ethylenediamine-templatedmixed-valence titanium phosphate. Chem. Mater. 2000, 12, 444–449. [Google Scholar] [CrossRef]
  82. Krogh Andersen, A.M.; Norby, P.; Hanson, J.C.; Vogt, T. Preparation and characterization of a new 3-dimensional zirconium hydrogen phosphate, τ-Zr(HPO4)2. Determination of the complete crystal structure combining synchrotron X-ray single-crystal diffraction and neutron powder diffraction. Inorg. Chem. 1998, 37, 876–881. [Google Scholar] [CrossRef]
  83. Mileo, P.G.M.; Kundu, T.; Semino, R.; Benoit, V.; Steunou, N.; Llewellyn, P.L.; Serre, C.; Maurin, G.; Devautour-Vinot, S. Highly Efficient Proton Conduction in a Three-Dimensional Titanium Hydrogen Phosphate. Chem. Mater. 2017, 29, 7263–7271. [Google Scholar] [CrossRef]
  84. Bortun, A.I.; Khainakov, S.A.; Bortun, L.N.; Poojary, D.M.; Rodriguez, J.; Jose, R.; Garcia, J.R.; Clearfield, A. Synthesis and characterization of two novel fibrous titanium phosphates Ti2O(PO4)2·2H2O. Chem. Mater. 1997, 9, 1805–1811. [Google Scholar] [CrossRef]
  85. Salvadó, M.A.; Pertierra, P.; García-Granda, S.; García, J.R.; Fernández-Diaz, M.T.; Dooryhee, E. Crystal structure, including H-atom positions, of Ti2O(PO4)2·(H2O)2 determined from synchrotron X-ray and neutron powder data. Eur. J. Solid State Inorg. Chem. 1997, 34, 1237–1247. [Google Scholar] [CrossRef]
  86. Babaryk, A.A.; Adawy, A.; García, I.; Trobajo, C.; Amghouz, Z.; Colodrero, A.R.M.P.; Olivera-Pastor, P.; Bazaga-García, M.; dos Santos-Gómez, L. Structural and proton conductivity studies of fibrous π-Ti2O(PO4)2·2H2O: Application in chitosan-based composite membranes. Dalton Trans. 2021, 50, 7667–7677. [Google Scholar] [CrossRef] [PubMed]
  87. Bruque, S.; Aranda, M.A.G.; Losilla, E.R.; Olivera-Pastor, P.; Maireles-Torres, P. Synthesis optimization and crystal structures of layered metal(IV) hydrogen phosphates, .alpha.-M(HPO4)2.cntdot.H2O (M = Ti, Sn, Pb). Inorg. Chem. 1995, 34, 893–899. [Google Scholar] [CrossRef]
  88. Huang, W.; Komarneni, S.; Noh, Y.D.; Ma, J.; Chen, K.; Xue, D.; Xuea, X.; Jiang, B. Novel inorganic tin phosphate gel: Multifunctional material. Chem. Commun. 2018, 54, 2682–2685. [Google Scholar] [CrossRef] [PubMed]
  89. Ansari, Y.; Telpriore, G.; Tucker, C.; Angell, A. A novel, easily synthesized, anhydrous derivative of phosphoric acid for use in electrolyte with phosphoric acid-based fuel cells. J. Power Sources 2013, 237, 47–51. [Google Scholar] [CrossRef]
  90. Moshareva, M.A.; Novikova, S.A.; Yaroslavtsev, A.B. Synthesis and ionic conductivity of (NH4)1–x Hx Hf2(PO4)3 (x = 0–1) NASICON-type materials. Inorg. Mater. 2016, 52, 1283–1290. [Google Scholar] [CrossRef]
  91. Clearfield, A. Structural concepts in inorganic proton conductors. Solid State Ion. 1991, 46, 34–43. [Google Scholar] [CrossRef]
  92. Clearfield, A.; Roberts, B.D.; Subramanian, M.A. Preparation of (NH4)Zr2(PO4)3 and HZr2(PO4)3. Mater. Res. Bull. 1984, 19, 219–226. [Google Scholar] [CrossRef]
  93. Komorowski, P.G.; Agryropoulos, S.A.A.; Canaday, J.D.; Kuriakose, A.K.; Wheat, T.A.; Ahmad, A.; Gulens, J. The study of hydronium NASICON conductivity with deuterium. Solid State Ion. 1992, 50, 253–258. [Google Scholar] [CrossRef]
  94. Stenina, I.A.; Kislitsyn, M.N.; Ghuravlev, N.A.; Yaroslavtsev, A.B. Phase transitions and ionic mobility in hydrogen zirconium phosphates with the NASICON structure, H1 ± xZr2–xMx(PO4)3⋅H2O, M = Nb, Y. Mater. Res. Bull. 2008, 43, 377–383. [Google Scholar] [CrossRef]
  95. Stenina, A.; Pinus, I.Y.; Rebrov, A.I.; Yaroslavtsev, A.B. Lithium and hydrogen ions transport in materials with NASICON structure. Solid State Ion. 2004, 175, 445–449. [Google Scholar] [CrossRef]
  96. Stenina, I.A.; Zhizhin, M.G.; Lazoryak, B.I.; Yaroslavtsev, A.B. Phase transitions, structure and ion conductivity of zirconium hydrogen phosphates, H1 ± xZr2–xMx(PO4)3 ⋅ H2O (M = Nb, Y). Mater. Res. Bull. 2009, 44, 1608–1612. [Google Scholar] [CrossRef]
  97. Mieritz, D.; Davidowski, S.K.; Seo, D. Accessing alkali-free NASICON-type compounds through mixed oxoanion sol–gel chemistry: Hydrogen titanium phosphate sulfate, H1−xTi2(PO4)3−x(SO4)x (x=0.5–1). J. Solid State Chem. 2016, 242, 116–125. [Google Scholar] [CrossRef] [Green Version]
  98. Sato, Y.; Shen, Y.; Nishida, M.; Kanematsu, W.; Hibino, T. Proton Conduction in Non-Doped and Acceptor-Doped Metal Pyrophosphate (MP2O7) Composite Ceramics at Intermediate Temperatures. J. Mater. Chem. 2012, 22, 3973–3981. [Google Scholar] [CrossRef]
  99. Singh, B.; Kima, J.-H.; Park, J.-Y.; Song, S.-J. Dense composite electrolytes of Gd3+-doped cerium phosphates for low-temperature proton-conducting ceramic-electrolyte fuel cells. Ceram. Inter. 2015, 41, 4814–4821. [Google Scholar] [CrossRef]
  100. Nagao, M.; Kamiya, T.; Heo, P.; Tomita, A.; Hibino, T.; Sano, M. Proton conduction in In3+-doped SnP2O7 at intermediate temperatures. J. Electrochem. Soc. 2006, 153, 1604–1609. [Google Scholar] [CrossRef]
  101. Nagao, M.; Takeuchi, A.; Heo, P.; Hibino, T.; Sano, M.; Tomita, A. A proton-conducting In3 + -doped SnP2O7 electrolyte for intermediate-temperature fuel cells. Electrochem. Solid-State Lett. 2006, 9, A105. [Google Scholar] [CrossRef]
  102. Jin, Y.; Fujiwara, K.; Hibino, T. High temperature, low humidity proton exchange membrane based on an inorganic–organic hybrid structure. Electrochem. Solid-State Lett. 2010, 13, B8. [Google Scholar] [CrossRef] [Green Version]
  103. Jin, Y.; Shen, Y.; Hibino, T. Proton conduction in metal pyrophosphates (MP2O7) at intermediate temperatures. J. Mater. Chem. 2010, 20, 6214–6217. [Google Scholar] [CrossRef]
  104. Chen, X.; Wang, C.; Payzant, E.; Xia, C.; Chu, D. An oxide ion and proton co-ion conducting Sn0.9In0.1P2O7 electrolyte for intermediate-temperature fuel cells. J. Electrochem. Soc. 2008, 155, B1264. [Google Scholar] [CrossRef]
  105. Kreller, C.R.; Pham, H.H.; Wilson, M.S.; Mukundan, R.; Henson, N.; Sykora, M.; Hartl, M.; Daemen, L.; Garzon, F.H. Intragranular phase proton conduction in crystalline Sn1−xInxP2O7 (x = 0 and 0.1). J. Phys. Chem. C 2017, 121, 23896–23905. [Google Scholar] [CrossRef]
  106. Kreller, C.R.; Wilson, M.S.; Mukundan, R.; Brosha, E.L.; Garzon, F.H. Stability and conductivity of In3+-doped SnP2O7 with varying phosphorous to metal ratios. ECS Electrochem. Lett. 2013, 2, F61–F63. [Google Scholar] [CrossRef]
  107. Foran, G.Y.; Goward, G.R. Site-Specific Proton Dynamics in Indium-Doped Tin Pyrophosphate. J. Phys. Chem. C 2020, 124, 28407–28416. [Google Scholar] [CrossRef]
  108. Scott, K.; Xu, C.; Wu, X. Intermediate temperature proton-conducting membrane electrolytes for fuel cells. Wiley Interdiscip. Rev. Energy Environ. 2014, 3, 24–41. [Google Scholar] [CrossRef]
  109. Anfimova, T.; Lie-Andersena, T.; Pristed Jensen, E.; Brorson Prag, C.; Nielsen, U.G.; Sørensen, D.R.; Skou, E.M.; Christensen, E.; Bjerrum, N.J.; Li, Q. The effect of preparation method on the proton conductivity of indium doped tin pyrophosphates. Solid State Ion. 2015, 278, 209–216. [Google Scholar] [CrossRef]
  110. Li, W.; Bose, A.B.; Rusakova, I.A. An approach for restoring the proton conductivity of sintered tin pyrophosphate membranes for intermediate temperature fuel cells. J. Power Sources 2016, 307, 146–151. [Google Scholar] [CrossRef]
  111. Ramaiyan, K.P.; Herrera, S.; Workman, M.J.; Semelsberger, T.A.; Atanasov, V.; Kerres, J.; Sandip Maurya, S.; Kim, Y.S.; Kreller, C.R.; Mukundan, R. Role of phosphate source in improving the proton conductivity of tin pyrophosphate and its composite electrolytes. J. Mater. Chem. A 2020, 8, 16345–16354. [Google Scholar] [CrossRef]
  112. Wu, X.; Mamlouk, M.; Scott, K. A PBI-Sb0.2Sn0.8P2O7-H3PO4 Composite Membrane. Fuel Cells 2011, 11, 620–625. [Google Scholar] [CrossRef]
  113. Jin, Y.C.; Nishida, M.; Kanematsu, W.; Hibino, T. An H3PO4-doped polybenzimidaz-ole/Sn0.95Al0.05P2O7 composite membrane for high-temperature proton exchange mem-brane fuel cells. J. Power Sources 2011, 196, 6042–6047. [Google Scholar] [CrossRef]
  114. Huang, M.; Huang, X.; Deng, Y.; Fei, M.; Xu, C.; Cheng, J. Graphite oxide-incorporated SnP2O7 solid composite electrolyte for high-temperature proton exchange membrane fuel cells. Int. J. Hydrog. Energy 2017, 42, 1113–1119. [Google Scholar] [CrossRef]
  115. Le, M.-V.; Tsai, D.-S.; Yang, C.-Y.; Chung, W.-H.; Lee, H.-Y. Proton conductors of cerium pyrophosphate for intermediate temperature fuel cell. Electrochim. Acta 2011, 56, 6654–6660. [Google Scholar] [CrossRef]
  116. Hogarth, W.H.J.; Muir, S.S.; Whittaker, A.K.; Diniz da Costa, J.C.; Drennan, J.; Lu, G.Q. Proton conduction mechanism and the stability of sol–gel titanium phosphates. Solid State Ion. 2007, 177, 3389–3394. [Google Scholar] [CrossRef]
  117. Alberti, G.; Costantino, U.; Casciola, M.; Ferroni, S.; Massinelli, L.; Staiti, P. Preparation, characterization and proton conductivity of titanium phosphate sulfophenylphosphonate. Solid State Ion. 2001, 145, 249–255. [Google Scholar] [CrossRef]
  118. Zhang, L.; Liu, X.; Sun, X.; Jian, J.; Li, G.; Yuan, H. Proton conduction in organically templated 3D open-framework vanadium−nickel pyrophosphate. Inorg. Chem. 2019, 58, 4394–4398. [Google Scholar] [CrossRef] [PubMed]
  119. Botez, C.E.; Tackett, R.J.; Hermosillo, J.D.; Zhang, J.; Zhao, Y.; Wang, L. High pressure synchrotron x-ray diffraction studies of superprotonic transitions in phosphate solid acids. Solid State Ion. 2012, 213, 58–62. [Google Scholar] [CrossRef]
  120. Baranov, A.I. Crystals with disordered hydrogen-bond networks and superprotonic conductivity. Review. Crystallogr. Rep. 2003, 48, 1012–1037. [Google Scholar] [CrossRef]
  121. Bagryantseva, I.N.; Gaydamaka, A.A.; Ponomareva, V.G. Intermediate temperature proton electrolytes based on cesium dihydrogen phosphate and Butyral polymer. Ionics 2020, 26, 1813–1818. [Google Scholar] [CrossRef]
  122. Dreßler, C.; Sebastiani, D. Effect of anion reorientation on proton mobility in the solid acids family CsHyXO4 (X = S, P, Se, y = 1, 2) from ab initio molecular dynamics simulations. Phys. Chem. Chem. Phys. 2020, 22, 10738–10752. [Google Scholar] [CrossRef]
  123. Taninouchi, Y.K.; Uda, T.; Awakura, Y.; Ikeda, A.; Haile, S.M. Dehydration behavior of the superprotonic conductor CsH2PO4 at moderate temperatures: 230 to 260 °C. J. Mater. Chem. 2007, 17, 3182–3189. [Google Scholar] [CrossRef]
  124. Mathur, L.; Kim, I.-H.; Bhardwaj, A.; Singh, B.; Park, J.-Y.; Song, S.-J. Structural and electrical properties of novel phosphate based composite electrolyte for low-temperature fuel cells. Composites Part B 2020, 202, 108405. [Google Scholar] [CrossRef]
  125. Otomo, J.; Tamaki, T.; Nishida, S.; Wang, S.; Ogura, M.; Kobayashi, T.; Wen, C.-J.; Nagamoto, H.; Takahashi, H. Effect of water vapor on proton conduction of cesium dihydrogen phosphate and application to intermediate temperature fuel cells. J. Appl. Electrochem. 2005, 35, 865–870. [Google Scholar] [CrossRef]
  126. Uda, T.; Haile, S.M. Thin-membrane solid-acid fuel cell. Electrochem. Solid State Lett. 2005, 8, A245–A246. [Google Scholar] [CrossRef] [Green Version]
  127. Navarrete, L.; Andrio, A.; Escolástico, S.; Moya, S.; Compañ, V.; Serra, J.M. Protonic Conduction of Partially-Substituted CsH2PO4 and the Applicability in Electrochemical Devices. Membranes 2019, 9, 49. [Google Scholar] [CrossRef] [Green Version]
  128. Ponomareva, V.G.; Bagryantseva, I.N.; Gaydamaka, A.A. Study of the Phase Composition and Electrotransport Properties of the Systems Based on Mono- and Disubstituted Phosphates of Cesium and Rubidium. Chem. Sustain. Dev. 2019, 27, 238–245. [Google Scholar]
  129. Martsinkevich, V.V.; Ponomareva, V.G. Double salts Cs1-xMxH2PO4 (M = Na, K, Rb) as proton conductors. Solid State Ion. 2012, 225, 236–240. [Google Scholar] [CrossRef]
  130. Ponomareva, V.G.; Bagryantseva, I.N. Superprotonic CsH2PO4-CsHSO4 solid solutions. Inorg. Mater. 2012, 48, 187–194. [Google Scholar] [CrossRef]
  131. Ponomareva, V.G.; Uvarov, N.F.; Lavrova, G.V.; Hairetdinov, E.F. Composite protonic solid electrolytes in the CsHSO4-SiO2 system. Solid State Ion. 1996, 90, 161–166. [Google Scholar] [CrossRef]
  132. Ponomareva, V.G.; Shutova, E.S.; Matvienko, A.A. Conductivity of Proton Electrolytes Based on Cesium Hydrogen Sulfate Phosphate. Inorg. Mater. 2004, 40, 721–728. [Google Scholar] [CrossRef]
  133. Otomo, J.; Minagawa, N.; Wen, C.-J.; Eguchi, K.; Takahashi, H. Protonic conduction of CsH2PO4 and its composite with silica in dry and humid atmospheres. Solid State Ion. 2003, 156, 357–369. [Google Scholar] [CrossRef]
  134. Ponomareva, V.G.; Shutova, E.S. High-temperature behavior of CsH2PO4 and CsH2PO4-SiO2 composites. Solid State Ion. 2007, 178, 729–734. [Google Scholar] [CrossRef]
  135. Singh, D.; Singh, J.; Kumar, P.; Veer, D.; Kumar, D.; Katiyar, R.S.; Kumar, A.; Kumar, A. The Influence of TiO2 on the Proton Conduction and Thermal Stability of CsH2PO4 Composite Electrolytes. S. Afr. J. Chem. Eng. 2021, 37, 227–236. [Google Scholar] [CrossRef]
  136. Singh, D.; Kumar, P.; Singh, J.; Veer, D.; Kumar, A.; Katiyar, R.S. Structural, thermal and electrical properties of composites electrolytes (1−x) CsH2PO4/x ZrO2 (0 ≤ x ≤ 0.4) for fuel cell with advanced electrode. SN Appl. Sci. 2021, 3, 46. [Google Scholar] [CrossRef]
  137. Veer, D.; Kumar, P.; Singh, D.; Kumar, D.; Katiyar, R.S. A synergistic approach to achieving high conduction and stability of CsH2PO4/NaH2PO4/ZrO2 composites for fuel cells. Mater. Adv. 2021, 3, 409–417. [Google Scholar] [CrossRef]
  138. Aili, D.; Gao, Y.; Han, J.; Li, Q. Acid-base chemistry and proton conductivity of CsHSO4, CsH2PO4 and their mixtures with N-heterocycles. Solid State Ion. 2017, 306, 13–19. [Google Scholar] [CrossRef] [Green Version]
  139. Bagryantseva, I.N.; Ponomareva, V.G.; Khusnutdinov, V.R. Intermediate temperature proton electrolytes based on cesium dihydrogen phosphate and poly(vinylidene fluoride-co-hexafluoropropylene). J. Mater. Sci. 2021, 56, 14196–14206. [Google Scholar] [CrossRef]
  140. Ponomareva, V.G.; Bagryantseva, I.N.; Shutova, E.S. Hybrid systems based on nanodiamond and cesium dihydrogen phosphate. Mater. Today 2020, 25, 521–524. [Google Scholar] [CrossRef]
  141. Ponomareva, V.G.; Bagryantseva, I.N. Effect of the excess protons on the electrotansport, structural and thermodynamic properties of CsH2PO4. Solid State Ion. 2017, 304, 90–95. [Google Scholar] [CrossRef]
  142. Ponomareva, V.G.; Lavrova, G.V. The influence of Cs2HPO4·H2O impurity on the proton conductivity and thermal properties of CsH2PO4. Solid State Ion. 2019, 329, 90–94. [Google Scholar] [CrossRef]
  143. Wang, M.; Luo, H.-B.; Liu, S.-X.; Zou, Y.; Tian, Z.-F.; Li, L.; Liu, J.-L.; Ren, X.-M. Water assisted high proton conductance in a highly thermally stable and superior water-stable open-framework cobalt phosphate. Dalton Trans. 2016, 45, 19466–19472. [Google Scholar] [CrossRef]
  144. Umeyama, D.; Horike, S.; Inukai, M.; Itakura, T.; Kitagawa, S. Inherent Proton Conduction in a 2D Coordination Framework. J. Am. Chem. Soc. 2012, 134, 12780–12785. [Google Scholar] [CrossRef]
  145. Inukai, M.; Horike, S.; Chen, W.; Umeyama, D.; Itakurad, T.; Kitagawa, S. Template-directed proton conduction pathways in a coordination framework. J. Mater. Chem. A. 2014, 2, 10404–10409. [Google Scholar] [CrossRef] [Green Version]
  146. Zhao, H.-R.; Xue, C.; Li, C.-P.; Zhang, K.-M.; Luo, H.-B.; Liu, S.-X.; Ren, X.-M. A Two-Dimensional inorganic−organic hybrid solid of manganese(II) hydrogenophosphate showing high proton conductivity at room temperature. Inorg. Chem. 2016, 55, 8971–8975. [Google Scholar] [CrossRef]
  147. Phang, W.J.; Lee, W.R.; Yoo, K.; Ryu, D.W.; Kim, B.S.; Hong, C.S. PH-dependent proton conducting behavior in a metal-organic framework material. Angew. Chem. Int. Ed. 2014, 53, 8383–8387. [Google Scholar] [CrossRef] [PubMed]
  148. Nagarkar, S.S.; Unni, S.M.; Sharma, A.; Kurungot, S.; Ghosh, S.K. Two-in-one: Inherent anhydrous and water-assisted high proton conduction in a 3D metal-organic framework. Angew. Chem. Int. Ed. 2014, 53, 2638–2642. [Google Scholar] [CrossRef]
  149. Nguyen, N.T.T.; Furukawa, H.; Gándara, F.; Trickett, C.A.; Jeong, H.M.; Cordova, K.E.; Yaghi, O.M. Three-dimensional metal-catecholate frameworks and their ultrahigh proton conductivity. J. Am. Chem. Soc. 2015, 137, 15394–15397. [Google Scholar] [CrossRef]
  150. Horike, S.; Chen, W.; Itakura, T.; Inukai, M.; Umeyama, D.; Asakura, H.; Kitagawa, S. Order-to-disorder structural transformation of a coordination polymer and its influence on proton conduction. Chem. Commun. 2014, 50, 10241–10243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Ma, N.; Kosasang, S.; Yoshida, A.; Horike, S. Proton-conductive coordination polymer glass for solid-state anhydrous proton batteries. Chem. Sci. 2021, 12, 5818–5824. [Google Scholar] [CrossRef] [PubMed]
  152. Matsuda, Y.; Yonemura, M.; Koga, H.; Pitteloud, C.; Nagao, M.; Hirayama, M.; Kanno, R. Synthesis, crystal structure, and ionic conductivity of tunnel structure phosphates, RbMg1-xH2x(PO3)3y(H2O). J. Mater. Chem. A 2013, 1, 15544–15551. [Google Scholar] [CrossRef]
  153. Antraptseva, N.M.; Solod, N.V.; Povshuk, V.A. Electrical conductivity of solid solution of Co(II)-Zn dihydrophosphate and its thermolysis products. Funct. Mater. 2015, 22, 322–326. [Google Scholar] [CrossRef] [Green Version]
  154. Fan, D.; Tian, P.; Xu, S.T.; Wang, D.H.; Yang, Y.; Li, J.Z.; Wang, Q.Y.; Yang, M.; Liu, Z.M. SAPO-34 templated by dipropylamine and diisopropylamine: Synthesis and catalytic performance in the methanol to olefin (MTO) reaction. New J. Chem. 2016, 40, 4236–4244. [Google Scholar] [CrossRef]
  155. Yu, Y.; Zhu, J.; Liu, J.; Yan, Y.; Song, X. Synthesis and characterization of two layered aluminophosphates [R-C8H12N]8[H2O]2·[Al8P12O48H4] and [S-C8H12N]8[H2O]2·[Al8P12O48H4] with a mirror symmetric feature and their proton conductivity. Dalton Trans. 2017, 46, 9157–9162. [Google Scholar] [CrossRef] [PubMed]
  156. Park, S.H.; Choi, W.; Choi, H.J.; Hong, S.B. Organic-Free Synthesis of Silicoaluminophosphate Molecular Sieves. Angew. Chem. Int. Ed. 2018, 57, 9413–9418. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, Y.Y.; Sun, Y.J.; Mu, Y.; Zhang, C.Q.; Li, J.Y.; Yu, J.H. Organotemplate-free hydrothermal synthesis of an aluminophosphate molecular sieve with AEN zeotype topology and properties of its derivatives. Chem. Commun. 2014, 50, 15400–15403. [Google Scholar] [CrossRef] [PubMed]
  158. Sun, Y.J.; Yan, Y.; Wang, Y.Y.; Li, Y.; Li, J.Y.; Yu, J.H. High proton conduction in a new alkali metal-templated open-framework aluminophosphate. Chem. Commun. 2015, 51, 9317–9319. [Google Scholar] [CrossRef]
  159. Parise, J.B. Crystal Structures of Related Novel Aluminophosphate Frameworks: Aipo4-21(py), Aipo4-EN3 (en) and A Structural Model for Alpo4-25. In Studies in Surface Science and Catalysis; Drzâj, S.H.B., Pejovnik, S., Eds.; Elsevier: Amsterdam, The Netherlands, 1985; Volume 24, pp. 271–278. [Google Scholar]
  160. Parnham, E.R.; Morris, R.E. The ionothermal synthesis of cobalt aluminophosphate zeolite frameworks. J. Am. Chem. Soc. 2006, 128, 2204–2205. [Google Scholar] [CrossRef]
  161. Wei, Y.; Tian, Z.; Gies, H.; Xu, R.; Ma, H.; Pei, R.; Zhang, W.; Xu, Y.; Wang, L.; Li, K.; et al. Ionothermal synthesis of an aluminophosphate molecular sieve with 20-ring pore openings. Angew. Chem. Int. Ed. 2010, 49, 5367–5370. [Google Scholar] [CrossRef]
  162. Liu, L.; Yang, J.; Li, J.; Dong, J.; Sîšak, D.; Luzzatto, M.; McCusker, L.B. Ionothermal sythesis and structure analysis of an open-framework zirconium phosphate with a high CO2/CH4 adsorption ratio. Angew. Chem. Int. Ed. 2011, 50, 8139–8142. [Google Scholar] [CrossRef]
  163. Nakayama, M.; Sugiura, Y.; Hayakawa, T.; Nogami, M. A novel proton conductor of imidazole–aluminium phosphate hybrids in the solid state. Phys. Chem. Chem. Phys. 2011, 13, 9439–9444. [Google Scholar] [CrossRef]
  164. Wang, K.; Li, T.; Zeng, H.; Zou, G.; Zhang, Q.; Zhien Lin, Z. Ionothermal synthesis of open-framework metal phosphates using a multifunctional ionic liquid. Inorg. Chem. 2018, 57, 8726–8729. [Google Scholar] [CrossRef]
  165. Zhang, C.; Yan, Y.; Huang, Z.; Shi, H.; Zhang, C.; Cao, X.; Jiang, J. Triclinic AlPO-34 zeolite synthesized with nicotine and its proton conduction properties. Inorg. Chem. Commun. 2018, 96, 165–169. [Google Scholar] [CrossRef]
  166. Xue, C.; Zou, Y.; Liu, S.-X.; Ren, X.-M.; Tian, Z.-F. Two different types of channels exhibiting distinct proton transport behaviour in an open framework aluminophosphate. J. Solid State Chem. 2018, 258, 695–701. [Google Scholar] [CrossRef]
  167. Mu, Y.; Wang, Y.Y.; Li, Y.; Li, J.Y.; Yu, J.H. Organotemplate-free synthesis of an open-framework magnesium aluminophosphate with proton conduction properties. Chem. Commun. 2015, 51, 2149–2151. [Google Scholar] [CrossRef]
  168. Fan, D.; Barrier, N.; Vicente, A.; Gilson, J.-P.; Clevers, S.; Dupray, V.; Coquereld, G.; Valtchev, V. Organic template-free synthesis of an open framework silicoaluminophosphate (SAPO) with high thermal stability and high ionic conductivity. Inorg. Chem. Front. 2020, 7, 542–553. [Google Scholar] [CrossRef]
  169. Zhou, D.; Chen, L.; Yu, J.H.; Li, Y.; Yan, W.F.; Deng, F.; Xu, R.R. Synthesis, crystal structure, and solid-state NMR Spectroscopy of a new open-framework aluminophosphate (NH4)2Al4(PO4)4(HPO4)·H2O. Inorg. Chem. 2005, 44, 4391–4397. [Google Scholar] [CrossRef] [PubMed]
  170. Zhao, H.R.; Jia, Y.; Gu, V.; He, F.Y.; Zhang, K.M.; Tian, Z.F.; Liu, J.L. A 3D open-framework iron hydrogenophosphate showing high proton conductance under water and aqua-ammonia vapor. RSC Adv. 2020, 10, 9046–9051. [Google Scholar] [CrossRef]
  171. Zima, V.; Lii, K.-H. Synthesis and characterization of a novel one-dimensional iron phosphate: [C4H12N2]1.5[Fe2(OH)(H2PO4)(HPO4)2(PO4)]·0.5H2O. J. Chem. Soc. Dalton Trans. 1998, 24, 4109–4112. [Google Scholar] [CrossRef]
  172. Lii, K.-H.; Huang, Y.-F. Large tunnels in the hydrothermally synthesized open-framework iron phosphate (NH3(CH2)3NH3)2[Fe4(OH)3(HPO4)2(PO4)3]·xH2O. Chem. Commun. 1997, 9, 839–840. [Google Scholar] [CrossRef]
  173. Bazaga-García, M.; Colodrero, R.M.P.; Papadaki, M.; Garczarek, P.; Zon, J.; Olivera-Pastor, P.; Losilla, E.R.; Reina, L.L.; Aranda, M.A.; Choquesillo-Lazarte, D.; et al. Guest Molecule-Responsive Functional Calcium Phosphonate Frameworks for Tuned Proton Conductivity. J. Am. Chem. Soc. 2014, 136, 5731–5739. [Google Scholar] [CrossRef]
  174. Lim, D.-W.; Kitagawa, H. Proton Transport in Metal−Organic Frameworks. Chem. Rev. 2020, 120, 8416–8467. [Google Scholar] [CrossRef]
  175. Mamlouk, M.; Scott, K. A boron phosphate-phosphoric acid composite membrane for medium temperature proton exchange membrane fuel cells. J. Power Sources 2015, 286, 290–298. [Google Scholar] [CrossRef] [Green Version]
  176. Pusztai, P.; Haspel, H.; Tóth, I.Y.; Tombácz, E.; László, K.; Kukovecz, Á.; Kónya, Z. Structure-Independent Proton Transport in Cerium(III) Phosphate Nanowires. ACS Appl. Mater. Interfaces 2015, 7, 9947–9956. [Google Scholar] [CrossRef] [PubMed]
  177. Petersen, H.; Stefmann, N.; Fischer, M.; Zibrowius, I.R.; Philippi, W.; Schmidt, W.; Weidenthaler, C. Crystal structures of two titanium phosphate-based proton conductors: Ab initio structure solution and materials properties. Inorg. Chem. 2022, 61, 2379–2390. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Examples of proton transport in phosphate-containing materials: (a) structural reorientation-mediated proton transfer. Zn (magenta), P (orange), O (red) and H (pale pink). (b) hopping through H-bond networks, and (c) carrier-mediated proton conduction. Zn (magenta), Zr (green), Cu (cyan), Cl (light green), P (orange), O (red), N (blue) and H (pale pink) atoms.
Figure 1. Examples of proton transport in phosphate-containing materials: (a) structural reorientation-mediated proton transfer. Zn (magenta), P (orange), O (red) and H (pale pink). (b) hopping through H-bond networks, and (c) carrier-mediated proton conduction. Zn (magenta), Zr (green), Cu (cyan), Cl (light green), P (orange), O (red), N (blue) and H (pale pink) atoms.
Materials 15 01292 g001
Figure 2. Publications on metal phosphates and pyrophosphate proton conductors over time.
Figure 2. Publications on metal phosphates and pyrophosphate proton conductors over time.
Materials 15 01292 g002
Figure 3. Layered structure of compound (NH4)2[ZrF2(HPO4)2] and possible proton-transfer pathways (adapted from [53]).
Figure 3. Layered structure of compound (NH4)2[ZrF2(HPO4)2] and possible proton-transfer pathways (adapted from [53]).
Materials 15 01292 g003
Figure 4. Views of chain packing and H-bond network (dashed grey lines) in compound (NH4)3Zr(H2/3PO4)3 (adapted from [55]).
Figure 4. Views of chain packing and H-bond network (dashed grey lines) in compound (NH4)3Zr(H2/3PO4)3 (adapted from [55]).
Materials 15 01292 g004
Figure 5. Structure of Ti2O(PO4)2·2H2O, a fibrous compound with capability of enhancing surface proton transport by H3PO4 impregnation (adapted from [86]).
Figure 5. Structure of Ti2O(PO4)2·2H2O, a fibrous compound with capability of enhancing surface proton transport by H3PO4 impregnation (adapted from [86]).
Materials 15 01292 g005
Figure 6. (a) View, along b-axis, of the chain packing and (b) 3D-framework of compound (NH4)3Zr(H2/3PO4)3, with H-bond network (dashed blue lines) highlighted (adapted from [118]).
Figure 6. (a) View, along b-axis, of the chain packing and (b) 3D-framework of compound (NH4)3Zr(H2/3PO4)3, with H-bond network (dashed blue lines) highlighted (adapted from [118]).
Materials 15 01292 g006
Figure 7. The characteristic high and low temperature proton conductivity and the corresponding structures for CsH2PO4, adapted from [6,120].
Figure 7. The characteristic high and low temperature proton conductivity and the corresponding structures for CsH2PO4, adapted from [6,120].
Materials 15 01292 g007
Figure 8. Irreversible structural transformation in NMe4Zn[HPO4][H2PO4]4 adapted from [143]. N (sky-blue), O (red), Zn (magenta), P (orange), C (grey) and H (pale pink) atoms.
Figure 8. Irreversible structural transformation in NMe4Zn[HPO4][H2PO4]4 adapted from [143]. N (sky-blue), O (red), Zn (magenta), P (orange), C (grey) and H (pale pink) atoms.
Materials 15 01292 g008
Figure 9. Open-framework structure of (NH3(CH2)3NH3)2[Fe4(OH)3(HPO4)2(PO4)3]·4H2O showing guest species inside channels and H-bond interactions. Fe (green), O (red), P (orange), and C (grey) atoms.
Figure 9. Open-framework structure of (NH3(CH2)3NH3)2[Fe4(OH)3(HPO4)2(PO4)3]·4H2O showing guest species inside channels and H-bond interactions. Fe (green), O (red), P (orange), and C (grey) atoms.
Materials 15 01292 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Colodrero, R.M.P.; Olivera-Pastor, P.; Cabeza, A.; Bazaga-García, M. Properties and Applications of Metal Phosphates and Pyrophosphates as Proton Conductors. Materials 2022, 15, 1292. https://doi.org/10.3390/ma15041292

AMA Style

Colodrero RMP, Olivera-Pastor P, Cabeza A, Bazaga-García M. Properties and Applications of Metal Phosphates and Pyrophosphates as Proton Conductors. Materials. 2022; 15(4):1292. https://doi.org/10.3390/ma15041292

Chicago/Turabian Style

Colodrero, Rosario M. P., Pascual Olivera-Pastor, Aurelio Cabeza, and Montse Bazaga-García. 2022. "Properties and Applications of Metal Phosphates and Pyrophosphates as Proton Conductors" Materials 15, no. 4: 1292. https://doi.org/10.3390/ma15041292

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

Article Metrics

Back to TopTop