Skip to content
Publicly Available Published by De Gruyter April 20, 2016

Substitution of W5+ in monophosphate tungsten bronzes by combinations Mn+/W6+

  • Subrata Chandra Roy , Wilfried Assenmacher , Thomas Linden , Lars Esser , Werner Mader and Robert Glaum EMAIL logo

Abstract

A series of hitherto unknown mixed-metal phosphates of the monophosphate tungsten bronze structure family [MPTB; (WO3)2m(PO2)4] have been obtained by solution combustion synthesis followed by annealing (ϑ = 850°C) at appropriate oxygen pressures. These new phosphates show substitution of W5+ by either M3+1/3W6+2/3 (M: V, Cr, Fe, Mo) or Ti4+1/2W6+1/2. Members of the MPTB structural series with m = 2 [e.g. CrIII4/3WVI8/3O12(PO2)4; TiIV6/3WVI6/3O12(PO2)4] and m = 4 [e.g. Cr4/3W20/3O24(PO2)4] have been obtained. In the course of our investigation the crystal structure of WOPO4 (MPTB with m = 2: W4O12(PO2)4) has been re-determined from X-ray single-crystal data, showing monoclinic instead of the orthorhombic symmetry reported in literature (P21/m, Z = 1, 80 parameters, 1832 independent reflections R1 = 0.027, wR2 = 0.063). The crystal structures of MoIII4/3WVI8/3O12(PO2)4 and CrIII4/3WVI8/3O12(PO2)4 (MPTBs with m = 2) were also refined from single-crystal data {(Mo/W (Cr/W): P21/m, Z = 1, 80(86) parameters, 1782(1769) independent reflections, R1 = 0.035(0.059), wR2 = 0.081(0.146)}. These refinements indicate statistical distribution of MIII and WVI over the metal sites. By selected area electron diffraction the unit cell dimensions of CrIII4/3WVI8/3O12(PO2)4 and CrIII4/3WVI20/3O24(PO2)4 derived from XRPD and SXRD are confirmed. HRTEM images of Cr4/3W20/3O24(PO2)4 are in agreement with its assumed close structural relation to W8O24(PO2)4 and show an highly ordered atomic arrangement.

1 Introduction

Monophosphate tungsten bronzes (MPTBs) form a family of compounds with general composition (WO3)2m(PO2)4 (0 ≤ m ≤ 12) [1, 2]). Their crystal structures are built from slabs of WO3 with ReO3-type structure which are separated by phosphate groups, as it is emphasized by the general formula. The alternative formulations WOPO4 (W4O4(PO4)4, m = 2) and W8O16(PO4)4 (m = 4) put special emphasis on the distinction between oxygen from phosphate groups and oxide ions. In some structures the WO3 slabs are separated by pyrophosphate groups (W12O28(P2O7)4 [3], W2O3(P2O7) [4]). Since their discovery in the early 1980s by Raveau and co-workers [2] a plethora of papers have reported on synthesis, crystal structures and physical properties of this class of mixed-valent tungsten (V,VI) phosphates [1, 5, 6]. In particular characterization of electrical conductivity and in recent years charge localization (“charge-density waves”) made these phosphates highly interesting research subjects [79]. As for classical tungsten bronzes AxWO3 [1014], incorporation of large cations into voids in the ReO3-type slabs is possible in MPTBs and has been studied extensively [15, 16].

It is rather surprising that despite all these investigations on monophosphate tungsten bronzes yet no reports were available on the substitution of tungsten(V) in the ReO3-type slabs by other transition metals. In our contribution we close this gap by reporting the first examples of mixed-metal monophosphate tungsten bronzoids (mm-MPTBs) were W5+ has been completely substituted either by M3+1/3W6+2/3 (M: V, Cr, Fe, Mo) or Ti4+1/2W6+1/2. In addition, we report on the re-determination of the crystal structure of WOPO4 (MPTB with m = 2), which resolves some inconsistencies in the literature [17, 18].

2 Experimental

2.1 Synthesis and crystallization

2.1.1 Starting material

Monoclinic WVI2O3(PO4)2 as microcrystalline, off-white powder was prepared from an equimolar mixture of WO3 (Sigma Aldrich, p. A.) and (NH4)2HPO4 (Merck, Darmstadt, p. A.) by solid state reaction according to [19]. FeIIIPO4 was also synthesized by solid state reaction from an equimolar ratio of Fe(NO3)3·9 H2O (Acros, NJ, USA, p. A.) and (NH4)2HPO4 at ϑ ≈ 600°C [20]. MoO2 was obtained by CVT according to [21].

WOPO4 as microcrystalline, dark brown powder was prepared from m-(W2O3)(PO4)2, WO3 (Sigma Aldrich, p. A.) and red phosphorus (Knapsack Electronic, 6N) at 1000°C in sealed silica tubes according to eq. 1.

(1)2W2O3(PO4)2+WO3+P5WOPO4(s) (1)

2.1.2 Solution combustion synthesis (SCS)

The mixed metal oxide phosphates TiIV6/3WVI6/3O12(PO2)4), MIII4/3WVI8/3O12(PO2)4, and MIII4/3W20/3O24(PO2)4 (MIII: V, Cr, Fe) were synthesized via SCS [22, 23] followed by heating of the reaction intermediates (see Table S1 of the Supporting Information). For SCS the starting materials TiO(acac)2 (abcr, Karlsruhe), NH4VO3 (Chempur, Karlsruhe), Cr(NO3)3·9H2O (Acros, NJ, USA), FePO4, Fe(NO3)3·9H2O (Acros, NJ, USA), (NH4)6W12O39·4.8H2O (Alfa Aesar, Karlsruhe), and (NH4)2HPO4 (Merck, Darmstadt), were used as source of TiO2, V2O5, Cr2O3, Fe2O3, WO3, and P4O10, respectively. Glycine (Grüssing, Filsum) was used as chelator and fuel in the SCS experiments. The starting materials were dissolved at appropriate molar ratios in a minimum amount of water together with conc. HNO3 as oxidant. The ratio of n(metal):n(glycine):n(HNO3) = 1:3:8 was kept in all experiments. The nitric acid solutions were carefully (ϑmax ≤ 100°C) evaporated to dryness to prevent premature ignition. The dry residues were then placed in a furnace preheated to 400°C. After ignition the reaction intermediates were ground in an agate mortar and subsequently heated with stepwise rising of the temperature until the mm-MPTBs were single-phase. Progress of the reactions was monitored by XRPD (IP Guinier technique). For the mm-MPTBs with M: Ti, Cr and Fe heat treatment was carried out in air whereas for (VIII1/3WVI2/3)OPO4 and VIII4/3WVI20/3O24(PO2)4 the reaction was carried out in an argon flow (p(O2) ≈ 20 ppm). For the synthesis of (VIII1−xWV,VIx)OPO4 (x = 0.8) which contains strongly reducing tungsten(V) subsequent annealing of the precursor from SCS was done in moist hydrogen (p(O2) ≈ 10−29 bar). For details see Table S1.

Crystals of WOPO4, (MoIII1/3WVI2/3)OPO4, and (CrIII1/3WVI2/3)OPO4 suitable for single-crystal structure analyses were grown by chemical vapor transport [24] in sealed silica tubes (l = 13 cm, d = 1.5 cm). For WOPO4 and (MoIII1/3WVI2/3)OPO4 iodine was used as transport agent and the temperature gradient 1000 → 900°C was applied. Thus, crystals of WOPO4 with edge length up to 4 mm were obtained by CVT at a transport rate of m˙= 3 mgh1. For (CrIII1/3WVI2/3)OPO4 chlorine as transport agent (from in situ decomposition of PtCl2 [25]), rather high temperatures (1050 → 900°C) and a long reaction period of 30 days had to be employed to achieve crystal growth.

2.2 X-ray powder diffraction (IP Guinier technique)

Powder diffraction patterns were recorded at ambient temperature using an imaging plate Guinier camera with an integrated read out system (HUBER G670, CuKα1 radiation, λ = 1.54059 Å for details see [26]). Lattice parameters for the mm-MPTBs (see Table 1) were determined using the program SOS [27] and SiO2 (Merck, Darmstadt, p. A.) as internal standard. The assignment of the reflections was checked by comparing the observed XRPD pattern to simulations based on the single crystal structure refinements.

Table 1:

Lattice parameters of mm-MPTBs derived from WOPO4 (MPTB with m = 2, P21/m, Z = 4) and W8O24(PO2)4 (MPTB with m = 4, P21, Z = 1), respectively.

Phosphatesa (Å)b (Å)c (Å)β (deg)
WOPO46.5538(4)5.2237(8)11.1866(8)90.332(7)a
TiIV1/2WVI1/2OPO46.48(4)5.21(5)11.08(4)90.4(6)
VIII1/3WVI2/3OPO46.513(2)5.202(1)11.054(2)90.12(1)
VIII0.20WV,VI0.80OPO46.531(1)5.206(7)11.124(2)90.12(2)
CrIII1/3WVI2/3OPO46.474(1)5.1569(4)11.049(1)90.11(1)
FeIII1/3WVI2/3OPO46.516(1)5.1771(6)11.054(1)90.20(1)
MoIII1/3WVI2/3OPO46.5417(6)5.276(5)11.1730(7)90.307(6)
W8O24(PO2)46.569(1)5.285(2)17.351(3)90b
V4/3W20/3O24(PO2)46.4929(8)5.2217(5)17.322(3)90.486(7)
Cr4/3W20/3O24(PO2)46.483(2)5.215(1)17.22(1)90.32(3)
Fe4/3W20/3O24(PO2)46.504(3)5.222(1)17.334(3)90.32(4)

aLattice parameters of WOPO4 [this work], brefinement of (WO3)8(PO2)4 in spacegroup P212121 [2], transformation to P21: 0 1̅ 0 1̅ 0 0 0 0 1.

2.3 Single-crystal X-ray diffraction

Crystal structure analyses were carried out for WVOPO4, (CrIII1/3WVI2/3)OPO4, and (MoIII1/3WVI2/3)OPO4 from single-crystal data. Details on data collections and reductions are summarized in Table 2. The starting parameters for the least-squares refinements were obtained by Direct Methods with Shelx-97 [31]. Full-matrix least-squares refinement in space group P21/m with anisotropic displacement parameters was carried out by using Shelx-97 in the WinGX [32] framework of programs. Mixed occupancy of the metal sites by tungsten and molybdenum or chromium was accounted for. The final atomic coordinates and anisotropic displacement parameters as well as selected inter-atomic distances are provided as supporting information in Tables S3 to S9.

Table 2:

WOPO4, (Mo1/3W2/3)OPO4, and (Cr1/3W2/3)OPO4. Summary of crystallographic data, single crystal data collection and structure refinements.

Empirical formulaWPO5(Mo1/3W2/3)PO5(Cr1/3W2/3)PO5
Structural formula(WO3)4(PO2)4(Mo1/3W2/3O3)4(PO2)4(Cr1/3W2/3O3)4(PO2)4
Formula weight294.81265.81251.30
Crystal systemMonoclinicMonoclinicMonoclinic
Space groupP21/m (11)P21/m (11)P21/m (11)
T, K293293293
λ, Å0.710730.710730.71073
a, Åa6.5538(4)6.5417(6)6.474(1)
b, Åa5.2237(8)5.276(5)5.1569(4)
c, Åa11.1866(8)11.1730(7)11.049(1)
β, dega90.33(2)90.307(6)90.11(1)
V, Å3a382.97(7)381.36(5)368.87(9)
Z444
Dcalcd., g·cm−35.125.104.48
μ, mm−130.525.322.1
Crystal size, mm30.02 × 0.06 × 0.100.2 × 0.1 × 0.080.06 × 0.06 × 0.06
ColorBronzeDark brownGreen
F(000), e516474449
DiffractometerX8-KappaApex II (Bruker)κ-CCD (Enraf-Nonius Inc.)κ-CCD (Enraf-Nonius Inc.)
SoftwareApex2, Saint [28]HKL Denzo, Scalepack [29]HKL Denzo, Scalepack [29]
Radiation/λ, ÅMoKα/0.71073MoKα/0.71073MoKα/0.71073
MonochromatorGraphiteGraphiteGraphite
Measured refls.56221627613046
Independent refls.183218321782
Rint0.0390.0710.130
Absorption corr.Multiscan [30]MultiscanMultiscan
No. of parameters808286
θ range, deg1.82–35.092.91–35.011.8–35.12
Index ranges hkl–10:+6, ±8, –13:+18–10:+6, ±8, –13:+18±10, ±8, ±17
R1/wR2b [I > 2 σ(I)]0.027/0.0630.035/0.0810.058/0.146
GooFb1.0501.311.278
Δρmax, e Å−32.2 (0.41 Å from O6)3.5 (0.77 Å from M2)3.8 (0.61 Å from M2)
Δρmin, e Å−32.2 (0.11 Å from O5)–2.2 (0.07 Å from O5)–2.9 (1.30 Å from O1)

aCell parameters from XRPD; bR1 = ∑||Fo| – |Fc||/∑|Fo|; wR2 = [∑w(Fo2Fc2)2/∑w(Fo2)2]1/2, w = [σ2(Fo2)+(AP)2+BP]−1, where P = (Max(Fo2, 0)+2Fc2)/3; GooF = S = [∑w(Fo2Fc2)2/(nobsnparam)]1/2.

Further details of the three crystal structure investigations may be obtained from Fachinformationszentrum Karlsruhe, 76344 Eggenstein-Leopoldshafen, Germany (fax: +49-7247-808-666; e-mail: , http://www.fiz-karlsruhe.de/request_for_deposited_data.html) on quoting the deposition numbers CSD-430823 for WOPO4, CSD-430824 for (Mo1/3W2/3)OPO4, and CSD-430825 for (Cr1/3W2/3)OPO4.

2.4 Magnetic measurements

The magnetic susceptibility of crushed crystals of WOPO4 and of microcrystalline powders of (Cr1/3W2/3)OPO4, and (Fe1/3W2/3)OPO4 was measured as a function of temperature (2 ≤ T ≤ 300 K, Figure 11) using a vibrating sample magnetometer (Quantum Design QD-PPMS-VSM). The measurement on WOPO4 was performed by first cooling the sample from room temperature to 2 K under applied magnetic field and then rising the temperature using the same temperature sweep rate. A diamagnetic correction was applied to the observed susceptibilities.

2.5 Electron microscopy

TEM studies were conducted on transmission electron microscopes (i) FEI-Philips CM300 UT/FEG operated at 300 kV and (ii) FEI-Philips CM30 T/LaB6, operated at 300 kV, both equipped with a Gatan CCD for image recording and with a Thermo NSS system for EDS analysis using HPGe and Si(Li) Nanotrace detectors, respectively. The powder samples were prepared both dispersed in cyclo-hexane by ultrasonication and without dispersion in a solvent on holey carbon films supported on a copper grid. Electron diffraction pattern and HRTEM image simulations were carried out using the Jems software package [33].

SEM EDS [34, 35] analyses of CrIII4/3WVI8/3O12(PO2)4 and CrIII4/3WVI20/3O24(PO2)4 were carried out by REM, DSM-940, Zeiss (operated in 20 kV, equipped with PV 9800 energy dispersive detector [36]). Prior to the EDS analyses crystals were soaked in 5% HF for 2 days, washed with distilled water and sputtered with gold. No evidence for impurity elements was found. The observed ratio n(Cr): n(W): n(P) was always close to the expected one.

3 Results and discussion

3.1 Synthesis, crystallization, and thermal decomposition

Several hitherto unknown mixed transition metal tungsten phosphates of the MPTB structure family have been synthesized (Tables S1 and S2) and characterized by their XRPD pattern (Table 1, and Figs. 1, 2 and 4). Phosphates (TiIV1/2WVI1/2)OPO4, (MIII1/3WVI2/3)OPO4 (M: V, Cr, Fe, Mo) and (VIII1−xWV,VIx)OPO4 (x ≤ 0.8) are isotypic to WOPO4 (MPTB with m = 2 [17, 23]). The phosphates MIII4/3WVI20/3O24(PO2)4 (M: V, Cr, Fe) with higher content of WO3 are isotypic to W8O24(PO2)4, the MPTB with m = 4 [2].

Fig. 1: Comparison of the XRPD pattern (IP Guinier technique, CuKα1 radiation) of phosphates (MIII1/3WVI2/3)OPO4 with M: V, Cr, Fe, Mo. Asterisks indicate a small impurity of V4/3W20/3O24(PO2)4 [23, 37]. The Guinier diagram of (MoIII1/3WVI2/3)OPO4 is compared to the simulated diffraction pattern based on the crystal structure refinements [this work].
Fig. 1:

Comparison of the XRPD pattern (IP Guinier technique, CuKα1 radiation) of phosphates (MIII1/3WVI2/3)OPO4 with M: V, Cr, Fe, Mo. Asterisks indicate a small impurity of V4/3W20/3O24(PO2)4 [23, 37]. The Guinier diagram of (MoIII1/3WVI2/3)OPO4 is compared to the simulated diffraction pattern based on the crystal structure refinements [this work].

Fig. 2: Comparison of the XRPD pattern (IP Guinier technique, CuKα1 radiation) of phosphates MIII4/3W20/3O24(PO2)4 with M: V, Cr, Fe. Asterisks indicate a small impurity of monoclinic WO3 [38, 39]. The pattern of FeIII4/3W20/3O24(PO2)4 is compared to the simulated diffraction pattern based on the crystal structure of W8O24(PO2)4 [2].
Fig. 2:

Comparison of the XRPD pattern (IP Guinier technique, CuKα1 radiation) of phosphates MIII4/3W20/3O24(PO2)4 with M: V, Cr, Fe. Asterisks indicate a small impurity of monoclinic WO3 [38, 39]. The pattern of FeIII4/3W20/3O24(PO2)4 is compared to the simulated diffraction pattern based on the crystal structure of W8O24(PO2)4 [2].

Synthesis of (TiIV1/2WVI1/2)OPO4, (MIII1/3WVI2/3)OPO4, and MIII4/3WVI20/3O24(PO2)4 (M: Cr, Fe) was achieved by SCS followed by annealing in air with stepwise rising of the temperature to 900°C. Synthesis of the “reduced” phases (VIII1/3WVI2/3)OPO4 and VIII4/3WVI20/3O24(PO2)4 containing vanadium(III) besides tungsten(VI) started also from precursors obtained by SCS (Tables S1, S2). Yet, subsequent heat treatment (550°C) of the reaction intermediate after ignition was carried out in an argon flow (p(O2) ≈ 2·10−5 bar) to set the oxidation state of vanadium to +3. Formation of the solid solution (VIII1−xWV,VIx)OPO4 (e.g. x = 0.8) containing vanadium(III) and tungsten(V) besides tungsten(VI) required reduction of the precursors from SCS in moist hydrogen. Hydrogen saturated at ambient temperature with water vapor (∑p = 1 bar, p(H2O) = 0.023 bar [40]) corresponds at ϑ = 550°C to an oxygen pressure p(O2) ≈ 10−29 bar, which is sufficient to reduce W6+ to W5+. Subsequent heat treatment of thus reduced precursors in sealed silica tubes at ϑ = 800°C for several days improved the crystallinity of the products. Annealing in air of precursors from SCS for synthesis of (VIII0.333WVI0.667)OPO4 led to oxidation and (according to XRPD) the formation of WVI2O3(PO4)2 (monoclinic modification [19]), a second phase very similar to VIII4/3WVI20/3O24(PO2)4, yet probably with a higher oxidation state of vanadium and slightly lower tungsten content. In addition, the XRPD pattern showed an increased background, possibly due to the formation of vitreous vanadium(V) phosphate.

The general procedures for the synthesis of the mm-MPTBs derived from W8O24(PO2)4 (MPTB with m = 4) from precursors obtained via SCS are very similar to those leading to the phases with m = 2. Therefore, it is quite remarkable that phases with higher WO3 content (m = 4) could also be synthesized via vapor phase moderated solid state reactions in sealed silica tubes using small amounts of NH4Cl or chlorine (from in situ decomposition of PtCl2) as mineralizer. Surprisingly, this procedure never led to the formation of the mm-MPTBs derived from WOPO4. Instead, formation of the mixed-metal tungsten ortho-pyrophosphates MIII(WVIO2)2(P2O7)(PO4) (M: V, Cr, Fe, Mo) [41] with the same overall chemical composition was observed. These phases with M: V, Cr, and Fe are also formed upon heating (MIII1/3WVI2/3)OPO4 (from SCS and subsequent annealing) to temperatures above 950°C, while the reverse reaction, the transformation of phases MIII(WVIO2)2(P2O7)(PO4) into the corresponding phase (MIII1/3WVI2/3)OPO4, did never occur in our experiments. All these observations suggest that mm-MPTBs (MIII1/3WVI2/3)OPO4 are thermodynamically metastable with respect to the ortho-pyrophosphates MIII(WVIO2)2(P2O7)(PO4), while the mm-MPTBs MIII4/3WVI20/3O24(PO2)4 (M: V, Cr, Fe) with a higher content of WO3 are indeed thermodynamically stable phases. Attempts to synthesize mm-MPTBs with even higher content of WO3 (m > 4) have failed so far. Such experiments led always to WO3 besides the mm-MPTBs with m = 2 and 4.

Chemical vapor transport of WOPO4 using water and iodine as transport agents is expected to proceed according to equations (2) and (3) in agreement with literature on CVT of WO2 [42], WO3 [43, 44], WP2O7 [45] and further anhydrous phosphates [24]. By analogy to anhydrous molybdenum phophates [46] and MoO2 [47], for the mixed phosphate (MoIII1/3WVI2/3)OPO4 heterogeneous equilibria similar to (2) and (3) are expected.

(2)8WOPO4(s) + 8H2O(g) 8WO2(OH)2(g) + P4O10(g) + P4O6(g) (2)
(3)8WOPO4(s) + 8I2(g) + P4O6(g)  8WO2I2(g) + 3P4O10(g) (3)

In agreement with literature on chemical vapor transport of Cr2O3 [48], WO3 and the phosphates CrPO4 and W2O3(PO4)2 [24] in a temperature gradient 1050 → 900°C using chlorine as transport agent, CVT of (Cr1/3W2/3)OPO4 is expected to proceed via equilibrium (4).

(4)12(Cr1/3W2/3)OPO4(s)+ 12Cl2(g)  4CrO2Cl2(g) + 8WO2Cl2(g) + 3P4O10(g) + 3O2(g) (4)

(TiIV1/2WVI1/2)OPO4 was never obtained as single-phase product, which hampered the determination of its lattice parameters (Table 1). Monoclinic W2O3(PO4)2 [19], monoclinic WO3 [38, 39] and TiP2O7 [49] were observed as by-products. Upon heating at temperatures above 850°C transformation into bi-phasic equilibrium mixtures consisting of WO3 and TiP2O7 was observed.

Upon heating at temperatures above 900°C (CrIII1/3WVI2/3)OPO4 transforms to CrIII(WVIO2)2(P2O7)(PO4) [41] (eq. 5a), which eventually decomposes at ϑ ≥ 1000°C to W2O3(PO4)2 [19] and α-CrPO4 [50] (eq. 5b). Decomposition of CrIII4/3WVI20/3O24(PO2)4 into W2O3(PO4)2 [19], WO3 [38, 39], and α-CrPO4 is observed at above 1000°C (eq. 6).

(5a)3(CrIII1/3WVI2/3)OPO4 CrIII(WVIO2)2(P2O7)(PO4) (5a)
(5b)CrIII(WVIO2)2(P2O7)(PO4)  WVI2O3(PO4)2+ CrPO4 (5b)
(6)3/4(CrIII4/3WVI20/3)O24(PO2)4 W2O3(PO4)2+ 3WO3+ CrPO4 (6)

3.2 Crystal structures

For a survey of the crystal structures of monophosphate tungsten bronzes (WO3)2m(PO2)4 we refer to [1, 5]. WOPO4 can be considered a second member (m = 2) of this series (Fig. 3a). Its crystal structure has been solved more than two decades ago (Pna21, Z = 4, a = 11.174(3) Å, b = 6.550(2) Å, c = 5.228(1) Å [17]). It is closely related to those of monoclinic and orthorhombic NbOPO4 [5153] and to TaOPO4 [54, 55].

Fig. 3: Crystal structures of WOPO4 [this work] (a) and of W8O24(PO2)4 [2] (b). Mixed occupancy (MIII1/3WVI2/3) for metal sites in the isotypic mm-MPTBs (MIII1/3WVI2/3)OPO4 and MIII4/3WVI20/3O24(PO2)4 is indicated by light green octahedra.
Fig. 3:

Crystal structures of WOPO4 [this work] (a) and of W8O24(PO2)4 [2] (b). Mixed occupancy (MIII1/3WVI2/3) for metal sites in the isotypic mm-MPTBs (MIII1/3WVI2/3)OPO4 and MIII4/3WVI20/3O24(PO2)4 is indicated by light green octahedra.

The XRPD pattern (IP Guinier photograph) of WOPO4 observed in the course of our study (Fig. 4) shows splitting of several reflections compared to the simulation based on the reported orthorhombic structure model WOPO4 [17]. Closer inspection of this discrepancy led to a monoclinic unit cell with β = 90.33°. The monoclinic symmetry is confirmed by the re-determination of the crystal structure in space group P21/m from single-crystal data (Tables 2, S4, S7, and S8). Despite the rather low residuals slightly enlarged anisotropic displacement parameters for oxygen atoms O5 and O6 (Fig. 5) suggested even lower symmetry. However, refinements in space groups P21 and Pm (including assumed pseudomerohedral twinning) did not yield an improvement. Evaluation of the XRPD pattern of the series of mm-MPTB given in Table 1 led for these phases to monoclinic unit cells similar to WOPO4. Single-crystal structure refinements of (MoIII1/3WVI2/3)OPO4 and (CrIII1/3WVI2/3)OPO4 (Tables 2, S4, S5, S6, S8, S9) confirmed the statistical distribution of Mo/W and Cr/W over the two chemically very similar metal sites in the structures. Selected interatomic distances for the mm-MPTBs are given in Table S5 and Fig. 5 and show no peculiarities. As for WOPO4 the structure refinement for (MoIII1/3WVI2/3)OPO4 led to enlarged ADPs for O5 and O6 (Fig. 5). For (CrIII1/3WVI2/3)OPO4 these anisotropies became even larger and led to the introduction of only half occupied split sites (O5a, O5b, O6a, O6b; Fig. 5). Selected area electron diffraction pattern of (CrIII1/3WVI2/3)OPO4 (Figs. 6 and 7) are in agreement with the chosen monoclinic unit cell and provide no hint on an enlarged cell as it is observed for NbOPO4 and TaOPO4 [51, 55].

Fig. 4: WOPO4. Experimental (IP Guinier technique, CuKα1 radiation, middle) and simulated diffraction pattern based on the revised single crystal structure refinement (bottom). Magnification of characteristically split reflections (top).
Fig. 4:

WOPO4. Experimental (IP Guinier technique, CuKα1 radiation, middle) and simulated diffraction pattern based on the revised single crystal structure refinement (bottom). Magnification of characteristically split reflections (top).

Fig. 5: ORTEP representation of [MO6] and [PO4] polyhedra in WOPO4 (a), (Mo1/3W2/3)OPO4 (b), and (Cr1/3W2/3)OPO4 with split positions (c). Displacement ellipsoids are drawn at the 50% probability level, software Diamond 4.0 [56].
Fig. 5:

ORTEP representation of [MO6] and [PO4] polyhedra in WOPO4 (a), (Mo1/3W2/3)OPO4 (b), and (Cr1/3W2/3)OPO4 with split positions (c). Displacement ellipsoids are drawn at the 50% probability level, software Diamond 4.0 [56].

Fig. 6: (CrIII1/3WVI2/3)OPO4. Contrast-inverted selected area electron diffraction (SAED) pattern in [100] zone axis orientation; dobs(0 0 1) = 11.18 Å, dobs(0 1 0) = 5.10 Å (see also Fig. S1).
Fig. 6:

(CrIII1/3WVI2/3)OPO4. Contrast-inverted selected area electron diffraction (SAED) pattern in [100] zone axis orientation; dobs(0 0 1) = 11.18 Å, dobs(0 1 0) = 5.10 Å (see also Fig. S1).

Fig. 7: (CrIII1/3WVI2/3)OPO4. Contrast-inverted SAED pattern in [001] zone axis orientation. dobs(1 0 0) = 6.472 Å, dobs(0 1 0) = 5.194 Å (see also Fig. S2).
Fig. 7:

(CrIII1/3WVI2/3)OPO4. Contrast-inverted SAED pattern in [001] zone axis orientation. dobs(1 0 0) = 6.472 Å, dobs(0 1 0) = 5.194 Å (see also Fig. S2).

Substitution of WV by MIII1/3WVI2/3 is also possible in the MPTB with m = 4, W8O24(PO2)4, which is described in literature [2] by an orthorhombic structure model (P212121, Z = 1, a = 5.285(2) Å, b = 6.569(1) Å, c = 17.351(3) Å). The powder diffraction pattern of the mm-MPTBs (MIII4/3WVI20/3)O24(PO2)4 (MIII: V, Cr, Fe) clearly exhibit a small but significant monoclinic distortion (Table 1) with the a axis of the orthorhombic structure model becoming the monoclinic b axis. Despite several attemps we did not succeed in growing single crystals of the mm-MPTBs suitable for single-crystal structure analysis. The simulated powder diffraction pattern for (FeIII4/3WVI20/3)O24(PO2)4 (Fig. 2) is based on the structure model of W8O24(PO2)4 (Fig. 3b) after its transformation to space group P21 (online tool TRANSTRU at the Bilbao Crystallographic Server [5760]). For the simulation refined lattice parameters were used. The orthorhombic structure model of W8O24(PO2)4 consists of two chemically different metal sites W1 and W2. Octahedra [W1O6] are sharing five vertices with [WO6] octahedra and only one with an adjacent [PO4] tetrahedron. In contrast, octahedra [W2O6] are sharing three vertices with [PO4] tetrahedra and three with neighboring [WO6] octahedra. Based on valence sum considerations [61] we assumed in the simulation preferred occupancy of the sites originating from W2 by trivalent cations. Intensities calculated for random distribution of iron and tungsten over the metal sites are, however, not significantly different. The unit cell dimensions obtained from the XRPD pattern of mm-MPTBs (MIII4/3WVI20/3)O24(PO2)4 are in agreement with the SAED pattern of (CrIII4/3WVI20/3)O24(PO2)4 (Figs. 8 and 9). Further support for the structure model of (CrIII4/3WVI20/3)O24(PO2)4 is drawn from the HRTEM image overlayed in Fig. 10 with a structural image based on coordination polyhedra. The characteristic herring bone pattern of the HRTEM image is nicely matched.

Fig. 8: (CrIII4/3WVI20/3)O24(PO2)4. Contrast-inverted SAED pattern in [100] zone axis orientation; dobs(0 1 0) = 6.64 Å, dobs(0 0 1) = 17.23 Å (see also Fig. S5).
Fig. 8:

(CrIII4/3WVI20/3)O24(PO2)4. Contrast-inverted SAED pattern in [100] zone axis orientation; dobs(0 1 0) = 6.64 Å, dobs(0 0 1) = 17.23 Å (see also Fig. S5).

Fig. 9: (CrIII4/3WVI20/3)O24(PO2)4. Contrast-inverted SAED pattern in [010] zone axis orientation; dobs(1 0 0) = 5.30 Å, dobs(0 0 1) = 17.27 Å (see also Fig. S6).
Fig. 9:

(CrIII4/3WVI20/3)O24(PO2)4. Contrast-inverted SAED pattern in [010] zone axis orientation; dobs(1 0 0) = 5.30 Å, dobs(0 0 1) = 17.27 Å (see also Fig. S6).

Fig. 10: (CrIII4/3WVI20/3)O24(PO2)4. Fourier-filtered HRTEM image (CM30 T) with overlaid polyhedral representation of the structure; view along [100].
Fig. 10:

(CrIII4/3WVI20/3)O24(PO2)4. Fourier-filtered HRTEM image (CM30 T) with overlaid polyhedral representation of the structure; view along [100].

3.3 Magnetic behavior

The temperature dependent magnetic behavior of WVOPO4, (CrIII1/3WVI2/3O)PO4 and (FeIII1/3WVI2/3O)PO4 is shown in Fig. 11. For the chromium and iron compounds over the whole temperature range Curie–Weiss behavior [62] with constant magnetic moments of μexp = 3.85 μB per Cr3+ (θp = –6.3 K) and 5.85 μB per Fe3+ (θp = –14.8 K) is observed. These moments are close to the spin-only values of 3.87 μB (d3) and 5.92 μB (d5, high spin) expected for Cr3+ and Fe3+, respectively. This agreement can be taken as further confirmation for the assumed chemical composition of these mm-MPTBs. WVOPO4 shows Curie–Weiss behavior (μexp = 1.06 μB, θP = –78 K) above 100 K and antiferromagnetic ordering (TN = 15.6 K) at low temperatures. These observations are in full agreement with information in the literature [9].

Fig. 11: Temperature-dependent molar susceptibility, χmol and its reciprocal χmol−1 of WVOPO4, and phosphates (MIII1/3WVI2/3)OPO4 (M: Cr, Fe) in the range 2 ≤ T ≤ 300 K.
Fig. 11:

Temperature-dependent molar susceptibility, χmol and its reciprocal χmol−1 of WVOPO4, and phosphates (MIII1/3WVI2/3)OPO4 (M: Cr, Fe) in the range 2 ≤ T ≤ 300 K.

4 Conclusions and outlook

It has been shown that complete substitution of tungsten(V) by either (TiIV1/2WVI1/2) or (MIII1/3WVI2/3) with MIII: V, Cr, Mo, Fe is possible in monophosphate tungsten bronzes (WO3)2m(PO2)4 with m = 2 and 4. These mixed-metal monophosphate tungsten bronzoids (mm-MPTBs) do not show the typical bronze color of MPTBs. Instead, their colors, which are typical for the corresponding M3+ ions, suggest localized electrons and insulating behavior. In contrast to the parent MPTBs, where the strongly reducing W5+ basically prevents incorporation of redox-active transition metal ions, the mm-MPTBs show much wider chemical flexibility. It is part of our ongoing research on this new class of phosphates to establish their crystal chemical limits. Of particular interest will be, whether catalytically active cations like V4+, V5+, Mo6+, or Cu2+ can be incorporated into mm-MPTBs. This should allow combining the structural features of MPTBs with the redox chemistry of these cations.

4.1 Supporting information

Tables S1 to S9 and Figures S1 to S6 are given as Supporting Information available online (DOI: 10.1515/znb-2016-0036).


Dedicated to: Professor Wolfgang Jeitschko on the occasion of his 80th birthday


Acknowledgments:

We thank Dr. Gregor Schnakenburg and Dr. Jörg Daniels for collection of X-ray single crystal diffraction data. We also thank Mr. Norbert Wagner for the magnetic measurements (all at University of Bonn).

References

[1] P. Roussel, O. Perez, P. H. Labbe, Acta Crystallogr.2001, B57, 603.10.1107/S0108768101009685Search in Google Scholar

[2] J. P. Giroult, M. Goreaud, P. H. Labbe, B. Raveau, Acta Crystallogr.1981, B37, 2139.10.1107/S0567740881008248Search in Google Scholar

[3] B. Domenges, M. Goreaud, P. H. Labbe, B. Raveau, Acta Crystallogr.1982, B38, 1724.10.1107/S0567740882007031Search in Google Scholar

[4] A. Leclaire, J. Chardon, B. Raveau, J. Solid State Chem.2000, 155, 112.10.1006/jssc.2000.8910Search in Google Scholar

[5] P. Roussel, P. H. Labbe, D. Giroult, Acta Crystallogr.2000, B56, 377.10.1107/S0108768199016195Search in Google Scholar

[6] A. Leclaire, M. M. Borel, J. C. Chardon, B. Raveau, J. Solid State Chem.1997, 28, 191.10.1006/jssc.1996.7182Search in Google Scholar

[7] E. Wang, M. Greenblatt, I. El-I. Rachidi, E. Canadell, M. H. Whangbo, S. Vadlamannati, Phys. Rev. B1989, 39, 12969.10.1103/PhysRevB.39.12969Search in Google Scholar

[8] Z.-T. Zhu, J. L. Musfeldt, H.-J. Koo, M.-H. Whangboo, Z. S. Teweldemedhin, M. Greenblatt, Chem. Mater.2002, 14, 2607.10.1021/cm011675jSearch in Google Scholar

[9] Z. S. Teweldemedhin, K. V. Ramanujachary, M. Greenblatt, J. Solid State Chem.1991, 95, 21.10.1016/0022-4596(91)90372-OSearch in Google Scholar

[10] A. Magnéli, Key Eng. Mater.1992, 68, 293.10.4028/www.scientific.net/KEM.68.293Search in Google Scholar

[11] C. G. Granqvist, Sol. Energy Mater. Sol. Cell2000, 60, 201.10.1016/S0927-0248(99)00088-4Search in Google Scholar

[12] A. Magnéli, Ark. Foer Kemi.1949, 1, 213.Search in Google Scholar

[13] A. Magnéli, Acta Chem.Scand.1953, 7, 315.10.3891/acta.chem.scand.07-0315Search in Google Scholar

[14] A. Hussain, L. Kihlborg, Acta Crystallogr.1976, A32, 551.10.1107/S0567739476001216Search in Google Scholar

[15] D. J. M. Bevan, P. Hagenmuller, Non-stoichiometric Compounds: Tungsten Bronzes, Vanadium Bronzes and Related Compounds, Pergamon Press, Oxford, 1973.10.1016/B978-0-08-018776-1.50006-8Search in Google Scholar

[16] B. Ingham, S. C. Hendy, S. V. Chong, J. L. Tallon, Phys. Rev.2005, B72, 075109.10.1103/PhysRevB.72.075109Search in Google Scholar

[17] S. L. Wang, C. C. Wang, K. H. Lii, J. Solid State Chem.1989, 44, 298.10.1016/0022-4596(89)90295-8Search in Google Scholar

[18] N. Kinomura, M. Hirose, N. Kumada, F. Muto, T. Ashida, J. Solid State Chem.1988, 77, 156.10.1016/0022-4596(88)90103-XSearch in Google Scholar

[19] C. Litterscheid, Ph. D. Thesis, Bonn University, Bonn, 2009; http://hss.ulb.uni-bonn.de/2009/1928/1928.htm (accessed March 2016).Search in Google Scholar

[20] H. N. Ng, C. Calvo, Canad. J. Chem.1975, 53, 2064.10.1139/v75-287Search in Google Scholar

[21] H. Oppermann, Z. Anorg. Allg. Chem.1971, 383, 285.10.1002/zaac.19713830309Search in Google Scholar

[22] M. S. Hedge, G. Madras, K. C. Patil, Acc. Chem. Res.2009, 42, 704.10.1021/ar800209sSearch in Google Scholar PubMed

[23] S. C. Roy, Ph.D. Thesis, Bonn University, Bonn, 2015.Search in Google Scholar

[24] M. Binnewies, R. Glaum, M. Schmidt, P. Schmidt, Chemical Vapor Transport Reactions, W. de Gruyter, Berlin, Boston, 2012.10.1515/9783110254655Search in Google Scholar

[25] G. Brauer (Hrsg.), Handbuch der Präparativen Anorganischen Chemie, Ferdinand Enke Verlag, Stuttgart, 1975.Search in Google Scholar

[26] I. Tanaka, M. Yao, M. Suzuki, K. Hikichi, T. Matsumoto, M. Kozasa, C. Katayama, J. App. Crystallogr.1990, 23, 334.10.1107/S0021889890004009Search in Google Scholar

[27] G. Meyer, J. Soose, SOS, Staatsexamensarbeit, Justus-Liebig Universität Giessen, Giessen (Germany) 1990.Search in Google Scholar

[28] G. M. Sheldrick, Shelx-97 (includes Shelxs-97, Shelxl-97, Ciftab), Programs for Crystal Structure Analysis (release 97-2), University of Göttingen, Göttingen (Germany) 1998.Search in Google Scholar

[29] L. J. Farrugia, J. Appl. Cryst.1990, 32, 837.10.1107/S0021889899006020Search in Google Scholar

[30] Apex2 (version 2009/2), Saint (version 2009/2), Area Detector Control and Integration Software, Bruker Analytical X-ray Instruments Inc., Madison, WI (USA) 2009.Search in Google Scholar

[31] Z. Otwinowski, W. Minor, in Methods in Enzymology, Vol. 276, Macromolecular Crystallography, Part A (Eds.: C. W. Carter Jr, R. M. Sweet), Academic Press, New York, 1997, pp. 307.10.1016/S0076-6879(97)76066-XSearch in Google Scholar

[32] R. H. Blessing, Acta Crystallogr.1995, A51, 33.10.1107/S0108767394005726Search in Google Scholar PubMed

[33] P. A. Stadelmann, Jems, EMS java version, 2004.Search in Google Scholar

[34] L. Reimer, Scanning electron microscopy, Springer-Verlag, Berlin, Heidelberg, 1985.10.1007/978-3-662-13562-4Search in Google Scholar

[35] D. E. Newbury, C. J. David, P. Echlin, Advanced scanning electron microscopy and x-ray microanalysis, Plenum Press, New York, London, 1987.10.1007/978-1-4757-9027-6Search in Google Scholar

[36] New Products, Powder Diffr.1987, 2, 212. doi:10.1017/S0885715600012781 (accessed March 2016).10.1017/S0885715600012781Search in Google Scholar

[37] D. R. Lide (Ed.), CRC Handbook of Chemistry and Physics, CRC Press, Boca Raton, FL, 2005.Search in Google Scholar

[38] S. C. Roy, B. Raguž, W. Assenmacher, R. Glaum, Solid State Sci.2015, 49, 18.10.1016/j.solidstatesciences.2015.09.006Search in Google Scholar

[39] H. Schornstein, R. Gruehn, Z. Anorg. Allg. Chem.1990, 582, 51.10.1002/zaac.19905820108Search in Google Scholar

[40] G. Krabbes, U. Gerlach, E. Wolf, W. Reichelt, H. Oppermann, J. C. Launay, Proc. 6th Symposium on Material Science under Microgravity Condition, Bordeaux, 1986.Search in Google Scholar

[41] H. Schäfer, T. Grofe, M. Trenkel, J. Solid State Chem.1973, 8, 14.10.1016/0022-4596(73)90014-5Search in Google Scholar

[42] H. Mathis, R. Glaum, R. Gruehn, Acta Chem. Scand.1991, 45, 781.10.3891/acta.chem.scand.45-0781Search in Google Scholar

[43] M. Saiful Islam, Ph. D. Thesis, Bonn University, Bonn, 2012; http://hss.ulb.uni-bonn.de/2012/2760/2760.htm (accessed March 2016).Search in Google Scholar

[44] H. Schäfer, J. Cryst. Growth1971, 9, 17.10.1016/0022-0248(71)90204-1Search in Google Scholar

[45] K. Nocker, R. Gruehn, Z. Anorg. Allg. Chem.1993, 619, 1530.10.1002/zaac.19936190907Search in Google Scholar

[46] S. C. Roy, R. Glaum, D. Abdullin, O. Schiemann, N. Q. Bac, K.-H. Lii, Z. Anorg. Allg. Chem.2014, 640, 1876.10.1002/zaac.201400160Search in Google Scholar

[47] P. M. Woodward, A. W. Sleight, T. Vogt, J. Phys. Chem. Solids1995, 56, 1305.10.1016/0022-3697(95)00063-1Search in Google Scholar

[48] T. Vogt, P. M. Woodward, B. A. Hunter, J. Solid State Chem.1999, 144, 209.10.1006/jssc.1999.8173Search in Google Scholar

[49] S. T. Norberg, G. Svensson, J. Albertsson, Acta Crystallogr.2001, C57, 225.Search in Google Scholar

[50] R. Glaum, R. Gruehn, M. Möller, Z. Anorg. Allg. Chem.1986, 543, 111.10.1002/zaac.19865431214Search in Google Scholar

[51] T. G. Amos, A. W. Sleight, J. Solid State Chem.2001, 160, 230.10.1006/jssc.2001.9227Search in Google Scholar

[52] D. L. Serra, S.-J. Hwu, Acta Crystallogr.1992, C48, 733.Search in Google Scholar

[53] E. M. Levin, R. S. Roth, J. Solid State Chem.1970, 2, 250.10.1016/0022-4596(70)90077-0Search in Google Scholar

[54] J. M. Longo, J. W. Pierce, J. A. Kafalas, Mater. Res. Bull.1971, 6, 1157.10.1016/0025-5408(71)90051-1Search in Google Scholar

[55] H. Chahboun, D. Groult, M. Hervieu, B. Raveau, J. Solid State Chem.1986, 65, 331.10.1016/0022-4596(86)90105-2Search in Google Scholar

[56] K. Brandenburg, Diamond, Crystal and Molecular Structure Visualization, Crystal Impact-K. Brandenburg & H. Putz GbR, Bonn (Germany) 2004; http://www.crystalimpact.com/diamond/ (accessed March 2016).Search in Google Scholar

[57] http://www.cryst.ehu.es/ (accessed March 2016).Search in Google Scholar

[58] M. I. Aroyo, J. M. Perez-Mato, D. Orobengoa, E. Tasci, G. de la Flor, A. Kirov, Bulg. Chem. Commun.2011, 43, 183.Search in Google Scholar

[59] M. I. Aroyo, J. M. Perez-Mato, C. Capillas, E. Kroumova, S. Ivantchev, G. Madariaga, A. Kirov, H. Wondratschek, Z. Kristallogr.2006, 221, 15.10.1524/zkri.2006.221.1.15Search in Google Scholar

[60] M. I. Aroyo, A. Kirov, C. Capillas, J. M. Perez-Mato, H. Wondratschek, Acta Crystallog.2006, A62, 115.10.1107/S0108767305040286Search in Google Scholar PubMed

[61] I. D. Brown, The chemical bond in inorganic chemistry: the bond valence model, IUCr Monographs in Crystallography 12, Oxford Science Publications, Oxford, 2002.Search in Google Scholar

[62] A. Weiss, H. Witte, Magnetochemie, Verlag Chemie, Weinheim, 1973.Search in Google Scholar


Supplemental Material:

The online version of this article (DOI: 10.1515/znb-2016-0036) offers supplementary material, available to authorized users.


Received: 2016-2-5
Accepted: 2016-2-19
Published Online: 2016-4-20
Published in Print: 2016-5-1

©2016 by De Gruyter

Downloaded on 10.12.2023 from https://www.degruyter.com/document/doi/10.1515/znb-2016-0036/html
Scroll to top button