Sheyi C. Adediwuraa,
Jan K. Wieda,
Christian F. Litterscheid
b and
Jörn Schmedt auf der Günne
*a
aUniversity of Siegen, Faculty IV: School of Science and Technology, Department for Chemistry and Biology, Inorganic Materials Chemistry and Center of Micro- and Nanochemistry and (Bio)Technology (Cμ), Adolf-Reichwein Straße 2, 57076 Siegen, Germany. E-mail: gunnej@chemie.uni-siegen.de
bChemische Fabrik Budenheim KG, Rheinstrasse 27, 55257 Budenheim, Germany
First published on 1st October 2025
Among the fastest hydrogen ion conductors are various hydrogen-rich phosphates, such as phosphoric acid and CsH2PO4. This, in principle, makes Al(H2PO4)3 a promising candidate for proton conduction within its hydrogen bonding network, especially in the medium-temperature range. Here, the phase-pure synthesis and structural details of the ferroelectric crystal C-type Al(H2PO4)3 are reported, where density functional theory (DFT) calculations were used to study the thermodynamic stability of different Al(H2PO4)3 polymorphs. Its ion conductivity, as determined by impedance spectroscopy, is an order of magnitude higher than previously reported. The thermal stability and decomposition route of Al(H2PO4)3 were studied using XRD, thermal analysis, and solid-state NMR, yielding results different from those previously observed in various mixtures containing Al(H2PO4)3. Heating produces an amorphous intermediate phase which converts into phase-pure monoclinic Al2(P6O18), which at even higher temperatures converts into the cubic phase Al4(P4O12)3 and little AlPO4. Rietveld refinements of C-type Al(H2PO4)3 and Al4(P4O12)3 are provided. The results show that while the proton ion conductivity of C-type Al(H2PO4)3 is hampered by the strong hydrogen-bonds, the structure features a Curie temperature significantly higher than room-temperature.
Al(H2PO4)3 is typically synthesised through the reaction of ortho-phosphoric acid H3PO4 with aluminium oxide, aluminium hydroxide, or elemental aluminium.12,13,15–17,18 To date, four polymorphs i.e., A-, B-, C-, and D-type Al(H2PO4)3 have been identified,15,16 although detailed reports exist only for A-, B-, and C-types.12,13,18 The A-type forms when an aged solution of aluminium in concentrated phosphoric acid (Al:
P = 1
:
5.5) is slowly evaporated under constant temperature conditions.13 The phase referred to here as “B-type Al(H2PO4)3” has only been reported in form of the Fe-substituted Al(H2PO4)3 and was prepared under hydrothermal conditions at elevated pressure.18 The C-type is obtained through the evaporation of viscous solutions of aluminium in concentrated phosphoric acid (Al
:
P = 1
:
5).12
The A-type Al(H2PO4)3 adopts a hexagonal crystal structure, with lattice parameters a = 7.858(4) Å and c = 24.956(7) Å (Pearson symbol: hR56, space group Rc).13,15 It consists of AlO6 octahedral that share corners with six PO2(OH)2 tetrahedra, forming a three-dimensional network. The structure of B-type Al(H2PO4)3 was derived by removing the Fe/Al disorder (Pearson symbol: mC88, space group Cc),18 and its simulated XRD-pattern shows good agreement with the XRD pattern of B-type Fe(H2PO4)3. This Fe-substituted form exhibits a distorted sheet-like network linked by M–O–P bridges (M = Fe, Al). The C-type Al(H2PO4)3 also crystallises in a hexagonal structure, with lattice parameters a = 13.727(5) Å and c = 9.152(1) Å (Pearson symbol: hR44, space group R3c)12 but features AlO6 octahedral that are stacked and linked by PO2(OH)2 tetrahedra group in a chain-like arrangement. To the best of our knowledge, its complete crystal structure has not been reported; only lattice parameters, space group and structural images have been provided. Among these Al(H2PO4)3 polymorphs, the thermodynamically most stable phase remains unknown. This can be addressed by comparing ground-state energies computed by first-principle methods at various pressures.19,20
Obtaining phase-pure Al(H2PO4)3 for property characterisation is challenging. This is due to the retention of residual ortho-phosphoric acid in the synthesis mixture, which is difficult to fully separate from the product and affects both thermal stability12 and evolution.2,17,21–24 A convenient tool to study products is NMR, being intrinsically quantitative and element-specific to P, H and Al, while it allows for monitoring all occurring products, including liquid, crystalline and amorphous constituents, which has also been noted in previous studies about the thermal decomposition of samples containing Al(H2PO4)3.17,21–24 Unfortunately, its thermal evolution is not clear, but different condensation and calcination reactions and temperatures have been suggested. Some studies indicate that C-type Al(H2PO4)3 remains stable up to 473 K before transitioning to crystalline AlH2P3O10 below 623 K.17,22 Other reports describe the formation of an amorphous “AlH2P3O10” phase between 473 K and 632 K.2,17,23 At temperatures above 723 K, two aluminium metaphosphate phases, the monoclinic Al2(P6O18) and the cubic Al4(P4O12)3, have been observed.2,17,22–24 Between 773 K and 873 K, a mixture of these phases is typically observed, with Al4(P4O12)3 becoming dominant beyond 873 K,17,22–24 suggesting that Al2(P6O18) may be metastable.24 The cubic phase, Al4(P4O12)3, subsequently decomposes into AlPO4 via the elimination of P4O10 at 1473 K, before transitioning into an amorphous phase beyond 1673 K.2,17,21–24 It should be noted that in those cases where NMR spectra were reported17,22–24 the samples were not single-phase C-type Al(H2PO4)3 but always contained different amounts of ortho-phosphoric acid and its condensation products. In a similar way, the ionic conductivity of pure C-type Al(H2PO4)3 has not been characterised yet, but only mixtures of C-type Al(H2PO4)3 and H3PO4.25 While existing research offers valuable insights into the thermal and conductive properties of C-type Al(H2PO4)3, discrepancies and unresolved questions highlight the need for further investigation into how impurities affect its thermal stability, phase transitions, and ionic conductivity.
The target of this contribution is to get a better understanding of Al(H2PO4)3 starting from phase-pure C-type Al(H2PO4)3, the stability of its different polymorphs, its ionic conductivity and thermal decomposition.
2Al + 6H3PO4 → 2Al(H2PO4)3 + 3H2↑ | (1) |
The second diffractometer used was a Huber Guinier powder camera G621 (Huber, Rimsting, Germany), configured in asymmetric transmission geometry. A curved Ge(111) monochromator was used to focus and monochromatise the X-ray beam, isolating Cu Kα1 radiation. Samples were prepared by evenly spreading the powder onto 1 μm thick Mylar foil affixed to a circular metallic sample holder using a small amount of grease for stabilisation. During measurements, the samples were laterally translated within the X-ray beam to ensure uniform exposure. Diffraction patterns were recorded on BaFX:Eu-based imaging plate (IP) film, with an exposure time of 15 minutes. The exposed imaging plates were subsequently scanned using a Typhoon FLA 7000 scanner (GE Company, USA), and the diffraction images were digitised using IPreader software.26
Structural refinements were performed using the TOPAS academic software package (v 7.24).27 Pawley28 and Rietveld29 refinement techniques were applied. The refinement parameters included: the scale factor and background coefficients, modelled using a Chebyshev function (10–14 free parameters), peak shape determined using the fundamental parameter approach, zero-shift corrections, lattice constants. For Rietveld refinements, hydrogen atom positions were constrained using distance restraints, while other atomic positions were refined considering site symmetry according to the space group. The quality of the structural models was assessed using fit indicators, including Rp (profile R-factor), Rwp (weighted profile R-factor), and Rexp (expected R-factor).
The nutation for 1H, 31P and 27Al were 90.91 kHz, 83.33 kHz and 62.50 kHz, respectively, corresponding to 90° pulse durations of 2.75 μs, 3 μs and 4 μs, respectively. The 1D 1H, 31P and 27Al MAS NMR spectra were recorded using a recycle delay of 5T1 and the dead time delay of 6 μs. 2D 31P–31P double-quantum MAS NMR spectra and 31P double-quantum build-up curves were recorded using the symmetry-based recoupling sequence R1445 (ref. 32) with 560 R elements (i.e., R = (900 − 270180)) and a 16-step phase cycle for the receiver phase to enable 0 → ±2 → 0 → −1 coherence pathway.
The Deconv2Dxy33 program was used for deconvolution of the NMR spectra. The isotropic chemical shift δiso, the anisotropic chemical shift δaniso, and the asymmetry parameter η for the 31P NMR spectra, along with the quadrupole coupling constant CQ and the quadrupolar asymmetry parameter ηQ for the 27Al NMR spectra,31 were extracted from the 31P and 27Al NMR spectra acquired at spinning frequencies of 4 kHz and 6 kHz, respectively. Their values were obtained by fitting the experimental spectra using the SIMPSON34,35 package v4.2.1 together with home-written scripts.
Electrochemical impedance spectroscopy (EIS) measurements were recorded using NEISYS electrochemical impedance analyser (Novocontrol Technologies, Montabaur, Germany) in a home-built cell, which was calibrated before actual measurements based on short/load calibration standards with a 100 Ω resistor as the load. The impedance measurements were recorded in potentiostatic mode, with an amplitude of 7.1 mVrms, in a frequency range from 1 MHz to 50 mHz. The temperature was controlled using a variable temperature and flow controller (NMR Service GmbH) with a constant nitrogen gas flow. Each temperature point was held for 20 minutes to ensure thermal equilibrium with an accuracy of ±0.1 K throughout the EIS measurement. The data analysis was performed using a home-written Python script.36
For each compound, the Monkhorst–Pack45 k-point mesh and energy cutoff were optimised separately. The energy cutoff varied by system, with 120 Ry (1633 eV) applied to specific calculations. The k-point mesh ranged from 4 × 4 × 4 to 10 × 10 × 10 based on convergence tests. The convergence threshold for self-consistency of the electronic wave function was set to 10−14 a.u., while the threshold for the total energy and atomic forces were set to 10−12 a.u., and 10−11 a.u., respectively. All pseudopotentials were taken from pslibrary v1.0,46 while cif2cell v.2.047 was used to convert all the crystallographic files into input files.
The polymorph with the lowest ground-state energy, following full relaxation of both atomic positions and lattice parameters to their most energetically favourable configuration under ambient pressure, is identified as the most stable form of Al(H2PO4)3. As shown in Fig. 1, C-type Al(H2PO4)3 exhibits the highest stability at standard conditions with the lowest formation energy. This observation explains why C-type Al(H2PO4)3 has been the most commonly synthesised polymorph in previous studies.12,17,22,25 The results in Fig. 1 also indicate that C-type Al(H2PO4)3 may transition to A-type Al(H2PO4)3 at approximately 1.43 GPa. Similarly, A-type Al(H2PO4)3 is predicted to transform into B-type Al(H2PO4)3 at around 34.25 GPa. From the volume–pressure curves, the bulk moduli of A-, B-, and C-type were estimated as 69.6 GPa, 37.9 GPa, and 21.7 GPa, respectively (Table S1 and Fig. S1).
Extensive trials showed that dissolving Al powder at 50 °C, then heating the viscous solution to 105 °C for at least 24 hours, followed by precipitation and washing of the product with acetonitrile at least three times, removes excess H3PO4 effectively. This method consistently produced phase-pure C-type Al(H2PO4)3. Purity was confirmed by 31P, 27Al, and 1H magic-angle spinning (MAS) nuclear magnetic resonance (NMR) spectroscopy (Fig. S3). The 31P MAS NMR spectrum displays a single resonance at −16.1 ppm. The 27Al and 1H NMR spectra show single resonances at −15.3 ppm and 10 ppm, respectively. No impurities in the NMR spectra verify phase-pure C-type Al(H2PO4)3 synthesis. Chemical shift tensors for 31P and the quadrupolar coupling constant for 27Al are listed in Tables 1 and 2, respectively.
Phase | δiso/ppm | δaniso/ppm | η | δ11/ppm | δ22/ppm | δ33/ppm |
---|---|---|---|---|---|---|
C-type Al(H2PO4)3 | −16.1 | −64 | 0.81 | 42.1 | −10.2 | −80.2 |
Al2(P6O18) | −36.6 | −105 | 0.61 | 48.0 | −16.1 | −141.6 |
−37.5 | −107 | 0.67 | 51.7 | −19.7 | −144.8 | |
−43.3 | −110 | 0.58 | 43.6 | −20.4 | −153.1 | |
Al4(P4O12)3 | −50.8 | −119 | 0.41 | 33.0 | −15.5 | −169.7 |
α-AlPO4 (cristobalite) | −26.7 | — | — | — | — | — |
The powder X-ray diffraction patterns of synthesised Al(H2PO4)3 were recorded at room temperature using Debye–Scherrer geometry. Pawley refinement was performed on the powder diffractogram in the R3c space group. A surprise were the splittings of the reflections observed in the high-angle region, which are not consistent with the proposed structural model.12 No supercell reflections in the low-angle regime could be observed or anywhere else. According to NMR the samples are phase pure. Structural relaxation of the presented structural model of the B-type Al(H2PO4)3 transformed to the space group P1 with quantum-chemical calculations conserves the high symmetry and indicates the structure is at a true energetic minimum. The splittings can be modelled with two unit cells with slightly different lattice parameters with equal intensity (Fig. S7): a = b = 13.6859(1) Å, c = 9.1314(1) Å and a = b = 13.6961(1) Å, c = 9.1402(1) Å, similar to those previously reported.12,16
The theoretically generated structure was subsequently used for Rietveld refinement analysis, neglecting the splittings in the high-angle regime by using angles up to 65.5°. The analysis yielded lattice parameters of a = b = 13.6862(1) Å and c = 9.1333(1) Å, which align well with the Pawley refinement results and prior studies.12,16 The powder pattern fit (Fig. 2) exhibits strong agreement with the theoretical structure, with minimal deviations observed.
The space group of C-type Al(H2PO4)3 belongs to the non-centrosymmetric pyroelectric point group (3m).49 The structure features a polar axis along which the hydrogen-bonds generate a permanent electric-dipole.49 The presence of a unique polar axis in C-type Al(H2PO4)3, as highlighted in an earlier study,12 suggests that the sample is in a ferroelectric state below its Curie temperature, and a potential transition to the centrosymmetric supergroup Rc in the paraelectric phase may occur at higher temperatures. The observed peak splittings in the high-angle regime (Fig. S7) might be related to this material's property, but NMR (number of peaks and NMR parameters) and quantum-chemical results do not give any evidence of a symmetry reduction or different unit cell.
The proton conductivity σ of C-type Al(H2PO4)3 was determined from the fitted bulk resistance values Rb, at corresponding temperatures, using the expression , where d and A denote the sample thickness and the cross-sectional area, respectively. The activation energy (Fig. 4) of the macroscopic ionic conductivity was obtained using the given expression;
, where a0 is the pre-exponential factor, Ea the activation energy, kB Boltzmann's constant, and T the temperature (Table 4).
The activation energy, conductivity at selected temperatures and corresponding pre-exponential factor for C-type Al(H2PO4)3 are summarised in Table 3. Repeated measurements (Fig. S8 and Table S2) show a minimal variation thereby confirming data reproducibility. The conductivity obtained in this work at 453 K is an order of magnitude higher than the previously reported value at 473 K for C-type Al(H2PO4)3.25 This improvement is attributed to the higher purity of C-type Al(H2PO4)3 synthesised in this work. Despite this improvement, the conductivity is still significantly lower than that of ortho-phosphoric acid,50 even though both materials exhibit similar highly polarisable hydrogen-bonds networks. Second moment analysis of C-type Al(H2PO4)3 from 1H NMR measurements conducted between 300 K to 453 K (Fig. S9) reveals a strong dipolar coupling. The second moment remains nearly constant up to 453 K, the upper limit of impedance spectroscopy measurements. This invariance indicates minimal motional averaging of the dipolar coupling constant as the sample is thermally activated, implying that the proton dynamics are slow on the NMR timescale within this temperature range. These findings align with the high activation energy obtained from impedance spectroscopy, further showing that pure C-type Al(H2PO4)3 is not suitable as an electrolyte for intermediate-temperature fuel cells.9
Predicted structural model | Refined structural model | |
---|---|---|
a/Å | 13.7111 | 13.6862(1) |
b/Å | 13.7111 | 13.6862(1) |
c/Å | 9.1477 | 9.1333(1) |
Crystal system | Hexagonal | Hexagonal |
Space group | R3c (no. 161) | R3c (no. 161) |
Z | 6 | 6 |
Atom | Wyckoff site | x | y | z | Occ. | Beq/Å2 |
---|---|---|---|---|---|---|
Fit residuals Rwp = 5.2%, and Rp = 3.3%. The proton positions were obtained by forcing the H atoms onto the internuclear connectivity vector between the neighbouring O atoms. | ||||||
Al1 | 6a | 0 | 0 | 0 | 1 | 0.61(8) |
P1 | 18b | 0.0033(3) | 0.1623(2) | 0.2656(9) | 1 | 1.24(5) |
O1 | 18b | 0.0340(5) | 0.1271(5) | 0.3998(9) | 1 | 0.83(5) |
O2 | 18b | 0.5811(3) | 0.5354(3) | 0.6204(9) | 1 | 0.83(5) |
O3 | 18b | 0.4512(3) | 0.5045(3) | 0.3784(8) | 1 | 0.83(5) |
O4 | 18b | 0.1251(5) | 0.0307(5) | 0.1256(9) | 1 | 0.83(5) |
H1 | 18b | 0.4967(2) | 0.5153(2) | 0.4631(6) | 1 | 1.24(7) |
H2 | 18b | 0.5506(2) | 0.5401(2) | 0.7123(6) | 1 | 1.24(7) |
C-type Al(H2PO4)3 | |
---|---|
σ373 K/(S cm−1) | 1.68(4) × 10−8 |
σ453 K/(S cm−1) | 1.23(3) × 10−6 |
a0/(S K cm−1) | 558168(1) |
Ea/eV | 0.81(1) |
The first step, up to point b (Fig. 5) is associated with mass loss of 9.1% which corresponds to the condensation of C-type Al(H2PO4)3 to an amorphous “AlH2P3O10·xH2O” (x ∼ 0.4).
Al(H2PO4)3 → “AlH2P3O10·xH2O” + (2 − x)H2O | (2) |
Between points b and c (86%), a gradual mass loss can be observed, which can be interpreted as a continuous loss of water and the formation of amorphous “AlH2P3O10”, theoretically corresponding to a ratio ΔM/M0 of 88.7%, which is in good agreement with the experiment. The observed mass ratio is a little bit too low, which indicates that the consecutive condensation process has partially started before point c, which, for an amorphous material, is not unexpected. Point c (TG) and P2 (DTA) mark the onset at which amorphous “AlH2P3O10” transforms in an exothermic event by condensation and crystallisation into monoclinic Al2(P6O18). The observed ΔM/M0 ratio agrees well with the expected value of 83.0%.
2“AlH2P3O10” → Al2(P6O18) + 2H2O | (3) |
The last thermal event at the highest temperature occurs at point e (onset at 1324 K). Given that water has already been completely released at lower temperatures, this can only be interpreted as a decomposition reaction involving the release of P4O10. The formation of AlPO4 would correspond to a mass ratio ΔM/M0 of 38.4% which does not explain the last step. Cubic Al4(P4O12)3 is expected to form from monoclinic Al2(P6O18) in an exothermic event around 1073 K,17,22 but apparently the effect was below the detection limit.
To confirm this interpretation, NMR and XRD experiments were performed on selected compounds annealed at specific temperatures. In 31P MAS NMR, crystalline and amorphous constituents can be easily identified and quantified: sharp peaks reflect a highly ordered crystal structure, broad peaks reflect the bond-length and bond-angle distribution describing an amorphous material. The 31P and 27Al NMR spectra (Fig. 6) confirm the previous interpretation. An amorphous phase is observed for samples annealed at 573 K and 623 K. The latter already shows a small quantity of the peaks of monoclinic Al2(P6O18), which is also in line with the TG experiments presented above. The peaks of crystalline phases AlPO4, cubic Al4(P4O12)3 and monoclinic Al2(P6O18) were unambiguously identified by their X-ray patterns in agreement with the assignment used in literature.51–53 Typical for 31P NMR is a correlation between the isotropic chemical shift and the number n of bridging O atoms per P atom, which defines the Qn labels.54 Bridging here refers to bridging to another P atom, however, it is known that the correlation is further influenced by other factors, including “bridges” to other types of atoms like Al and H, so that a partial overlap between the chemical shift ranges of the Qn(31P) units results. The correlation of increasing n to lower the isotropic chemical shift values is consistent with the isotropic shift of AlPO4 (Q0), C-type Al(H2PO4)3 (Q0), monoclinic Al2(P6O18) (Q2) and cubic Al4(P4O12)3 (Q2). The broad 31P peak of the amorphous phase overlaps with the peaks in the Q2 range, featuring a significant shoulder to higher ppm values, which could be assigned to the presence of Q1 groups in the amorphous phosphate “AlH2P3O10”. It should be noted that the implicit assumption that the amorphous phosphate corresponds to a triphosphate O3P-O-PO2-O-PO3 is not confirmed by the NMR results. This would require a peak of the Q1 function being twice as high as the Q2 peak, which is not observed. A 31P–31P 2D double-quantum MAS NMR spectrum (Fig. 7) of “AlH2P3O10” shows only a single broad peak and also does not reveal the typical pattern expected for a triphosphate. For this reason, the compound annealed at 573 K most likely should be considered having a composition close to that of the triphosphate AlH2P3O10, but in reality, it contains a mixture of different chain and ring phosphates. 31P NMR unfortunately does not resolve Q0, Q1 and Q2 peaks to gather further insight. This interpretation is corroborated by the observation of three resolved peaks in 27Al MAS NMR, which by a similar argument can be assigned to 4-, 5- and 6-fold coordinated Al-atoms.
The sample annealed at a temperature of 823 K shows that thermal decomposition of phase-pure C-type Al(H2PO4)3 opens a way to the synthesis of phase-pure monoclinic Al2(P6O18). No impurities could be observed in 27Al and 31P NMR nor in XRD, while in previous studies starting from mixtures of C-type Al(H2PO4)3 and H3PO4, only heterogeneous mixtures of monoclinic Al2(P6O18) with additional phases could be observed.17,22–24 This may open a way to samples doped with different luminescent ions in the host structure Al2(P6O18) and their unambiguous characterisation. The XRD pattern of monoclinic Al2(P6O18) (Fig. S10),52 its 31P and 27Al NMR parameters of the different sites determined from fitting NMR spectra (Fig. S5) and a 31P 2D double-quantum MAS NMR (Fig. S11) are shown in the SI and Tables 1 and 2.
NMR and XRD on a sample annealed at a temperature of 1273 K show that monoclinic Al2(P6O18) has converted into cubic Al4(P4O12)3 and a small trace of AlPO4 (Fig. 6). In an attempt to remove the impurity of AlPO4, phase-pure C-type Al(H2PO4)3 was annealed at a temperature of 1273 K in a vacuum of ∼10−2 mbar for 6 and 24 hours, respectively, assuming that the formed small quantity of AlPO4 could be related to a surface coverage by OH functions. Surprisingly, the product of this attempt at 6 hours was mixture of Al2(P6O18), Al4(P4O12)3, and AlPO4 while the product of the attempt at 24 hours produces only AlPO4 (NMR and XRD not shown), which proves that the phase transitions are catalysed by small quantities of the remaining water, which indicates that the transition temperatures of the crystallisation of the amorphous phase observed in thermal analysis may be difficult to reproduce without knowing the partial pressure of water.
For cubic Al4(P4O12)3, the crystal structure published in the year 193753 provides the correct position of reflections but an unsatisfactory match of the intensities (Fig. S13). NMR spectra resolve one P and one Al site, with chemical shift tensors obtained from Fig. S6 spectra listed in Tables 1 and 2. This indicates that the published structure is correct, but the atomic positions could be improved by a Rietveld refinement. For this purpose, the structure for the substituted variant (Al3.39V0.61)(P4O12)3,55 determined by single-crystal X-ray diffraction, was taken as a starting point. The refined structure is in excellent agreement with the powder XRD with cubic lattice parameters of a = 13.735(2) Å (Fig. S12 and Table S3). Furthermore, the structure obtained from Rietveld refinement was converted to P1 space group with respect to the fractional coordinates of the atoms and was relaxed using Quantum Espresso to establish whether the structure corresponds to an energetic minimum. The relaxed structure shows minimal changes, as confirmed by visual inspection and comparison of powder XRD patterns (Fig. S13). The space group also remains I
3d, as confirmed by KPLOT,56 validating the new structure for Al4(P4O12)3, which means that the structure is confirmed through independent evidence by NMR, XRD and quantum-chemical calculations.
CCDC 2485462 and 2485463 contain the supplementary crystallographic data for this paper.57a,b
This journal is © The Royal Society of Chemistry 2025 |