Home Synthesis of schafarzikite-type (PbBi)(Fe1−xMn x )O4: a study on structural, spectroscopic and thermogravimetric properties
Article Open Access

Synthesis of schafarzikite-type (PbBi)(Fe1−xMn x )O4: a study on structural, spectroscopic and thermogravimetric properties

  • Carla M. Uribe-Rincón ORCID logo , Thorsten M. Gesing ORCID logo and Mohammad Mangir Murshed ORCID logo EMAIL logo
Published/Copyright: June 23, 2025
Become an author with De Gruyter Brill

Abstract

In the mineral schafarzikite (FeSb2O4) the stereochemical active 5s2 lone electron pairs (LEPs) play pivotal roles to diverse crystal-physico-chemical properties. The complete substitution of Sb(III) with an equivalent amount of Pb(II) and Bi(III) leads to higher oxidation state of the transition element (e.g., Fe3+) and offers a mixed stereochemical activity of the 6s2 LEPs. Here we report the synthesis and characterization of a solid solution (PbBi)(Fe1−xMn x )O4 (x = 0–1.0; Δx = 0.1). The polycrystalline samples are synthesized in sealed quartz tubes at 923 K and a pressure of about 1.0 Pa. The crystal structure of each member of the solid solution has been analyzed by X-ray power diffraction data Rietveld refinements. The metric and the geometric parameters (bond lengths, angles, distortion of the polyhedra) have been followed with respect to the chemical composition. The Wang-Liebau eccentricity (WLE) parameter, which measures the strength of the stereochemical activity, has been found to be associated with the over-bonding nature of the LEP cations. The vibrational properties are characterized using Raman and infrared spectroscopy, supporting the structural features. The optical band gap is evaluated from the UV/Vis diffuse reflectance spectra using both conventional Tauc’s and derivation of absorption spectrum fitting (DASF) methods. This study particularly highlights the miscibility between Fe(III) and Mn(III) cations on a single Wyckoff site within the schafarzikite-type structures, where the LEP driven void channels seem to be more suitable for a small cation (e.g., Li+) insertion than the mineral schafarzikite.

1 Introduction

While the mineral schafarzikite (FeSb2O4) has been known since 1921, 1 the isostructural Mn-containing mineral manganoschafarzikite (MnSb2O4) 2 was discovered very recently. For a general description, the chemical composition can be written either as ML2O4 with M2+ and L3+ cations, where L represents elements with stereochemically active lone electron pairs (LEPs), or as L2MO4 with L2+, L3+ and M3+. The crystal structure can be described as mullite-type, 3 where the edge-sharing MO6 octahedra chains run along the c-axis, as shown in Figure 1.

Figure 1: 
Crystal structure of FeSb2O4, showing the edge-sharing FeO6 octahedra chains and the distorted tetrahedral SbO3E coordination (E = 5s2 active lone electron pair). The E lobes (yellow extension of the Sb atoms) form a quasi-1D channel along the c-axis.
Figure 1:

Crystal structure of FeSb2O4, showing the edge-sharing FeO6 octahedra chains and the distorted tetrahedral SbO3E coordination (E = 5s2 active lone electron pair). The E lobes (yellow extension of the Sb atoms) form a quasi-1D channel along the c-axis.

The L2MO4 mullite-type O8 phases (L4M2O8), 4 , 5 , 6 , 7 , 8 where the L-cation (e.g., Pb2+, Sb3+, Bi3+) keeps a stereochemically active ns2 LEP located at the apex of a distorted LO3E nido-like tetrahedron (LEP = E) that connects the MO6 octahedra of the respective chains. The E lobes expose as pseudo-ligands, producing a negatively charged void channel along the c-axis. This LEP-driven quasi-1D channel draws particular attention in designing functional energy materials. For instance, the LEP induced void space in MnSb2O4, 9 CoSb2O4 and NiSb2O4 10 are demonstrated as potential anode candidates for Li-ion batteries. Moreover, FeSb2O4 11 has been shown to be a suitable host for fluoride (F) inclusion. The reversible topochemical fluoride was inserted into the quasi-1D tunnels (Figure 1) followed by Fe2+/Fe3+ redox process along with the consequent lattice volume expansion/contraction. During the Fe2+→Fe3+ oxidation, the 5s2 LEPs interact with the F-2p states, forming a larger SbO3EF coordination environment, however, with relatively weak Sb–F bonds. 11

The strength of the stereochemical activity of the LEPs essentially stabilizes the channel, which can be measured by the Wang-Liebau eccentricity (WLE) parameter. This scalar quantity is found to be a function of the structural bond valence sum (BVS) of the LEP cations. 12 , 13 , 14 , 15

Earlier attempts to study a solid solution of (M1−xM x )L2O4 included partial or full substitution of the M cation to tailor the endmember properties. Using 57Fe Mössbauer spectroscopy Berry et al. 16 demonstrated that when Mg2+ substantially replaces Fe2+ the non-magnetic MgO6 subunit disrupts the magnetic exchange interactions between the Fe2+ cations along the c-axis, and a composition of (Fe0.25Co0.75)Sb2O4 shows a different antiferromagnetic (AFM) exchange between Fe2+ and Co2+ than those of the endmembers. Cumby et al. 17 reported (Co1−xMn x )Sb2O4 and (Co1−xFe x )Sb2O4 solid solutions, where the canted AFM structure gradually changes from C-type to A-type with increasing Co content due to rotation of the magnetic moment vectors from c- to a-direction. Within a MO6 octahedra chains the direct exchange promotes an AFM ground state while a 90° super-exchange favors a ferromagnetic (FM) ordering between two neighboring M cations. 17 For co-sharing the 4d Wyckoff site by two magnetic cations, the relative magnitude of the exchange interactions plays pivotal roles for the magnetic orderings. Within the solid solution (Mn1−xV x )Sb2O4 a FM ordering is found for 0 < x ≤ 0.6. Clearly, the chemical tailoring of the L and M sites can significantly influence the physical properties. 18 , 19 , 20 , 21 , 22 Introducing Pb2+ into FeSb2O4 results in partial oxidation of Fe2+ to Fe3+ in Fe(Sb1−xPb x )2O4. 23 As such, the Néel temperature (T N ) increases from 45 K (FeSb2O4) to 69 K (FeSb1.4Pb0.6O4) and the parental A-type AFM turns into C-type AFM ordering. Complete substitution of Sb3+ by mixed valence L-type cations has recently been published 24 for two new members of the schafarzikite-type family: (PbBi)FeO4 keeps the parental A-type AFM ordering at T N of 24 K and (PbBi)MnO4 a G-type AFM at T N of 43 K. Therefore, it is worth testing whether a complete miscibility exists between Fe3+ and Mn3+ within the solid solution (PbBi)(Fe1−xMn x )O4, which eventually offers a systematic study to follow the crystal chemical properties.

2 Experimental

2.1 Synthesis

Polycrystalline samples of (PbBi)(Fe1−xMn x )O4 (x = 0–1) were synthesized by the conventional solid-state method, grinding respective stoichiometric amounts of PbO (99.9 %, Alfa Aesar), Bi2O3 (99 %, Alfa Aesar), Fe2O3 (99 %) and Mn2O3 (99 %) in a agate mortar. Afterward, the mixture was pressed into pellets and transferred into dried quartz tubes. The tubes were sealed with an internal pressure of about 1.0 Pa followed by heating in a muffle furnace at 923 K for 72 h.

2.2 X-ray powder diffraction

X-ray powder diffraction (XRPD) data were collected on a Bruker D8 Advance diffractometer (Bruker AXS GmbH, Karlsruhe Germany) using a Bragg-Brentano geometry with CuKα1 radiation and a Johansson monochromator. Measurements were taken at ambient conditions from 10° to 130° 2θ with a step width of 0.0145° and a data collection time of 1.25 s per step. The XRPD data were refined using the Rietveld method (TOPAS V4.2, Bruker AXS). The respective fundamental parameters were calibrated using a LaB6 standard and used for the profile line-shape. During the refinements general parameters such as sample displacement, scale factor and background parameter (Chebychev polynomial) were optimized. The lattice parameters, atomic coordinates and atomic displacement parameters were also refined. The isotropic atomic displacement parameters (ADP) were constrained between atoms located at the same Wyckoff site (Bi/Pb, Fe/Mn) to avoid unphysical outcomes such as negative ADPs. Besides the heavy elements Pb and Bi, the ADPs of the light oxygen atoms were as well refined constrained to each other. The starting atomic coordinates were taken from the crystal structure data of FeSb2O4, where the 8h Wyckoff site was set to co-share between Pb2+ and Bi3+ with an occupancy of 0.5 each. 25 Since the X-ray scattering factors of Pb2+ and Bi3+ are similar, the Rietveld refinement cannot distinguish the occupancy parameters between the respective atoms. As such, the nominal compositional occupancy was fixed, which was supported by the EDX results. For the same reason, the occupancy parameter between Fe and Mn were set to their nominal compositions.

2.3 UV/Vis diffuse reflectance spectroscopy

A Shimadzu UV-2600 (Shimadzu, Kyoto, Japan) spectrophotometer, equipped with an ISR-2600 plus two-detector integrating sphere (Pike Technologies, USA), was used for the collection of optical diffuse reflectance spectra at room temperature. Data were recorded in the wavelength range between 190 nm and 950 nm with a step of 1 nm. A BaSO4 powder sample was used as reference standard for the baseline correction. The Tauc-type 26 , 27 band gaps (direct or indirect) were compared with those determined by the DASF method 28 , 29 after converting the reflection data to absorption ones using the Kubelka-Munk transformation. 30

2.4 Vibrational spectroscopy

Raman spectra were recorded on a LabRam ARAMIS (Horiba Jobin–Yvon) Micro-Raman spectrometer. A laser of 525 nm providing power of less than 20 mW was focused on the sample surface through a 50× objective (Olympus) with a numerical aperture of 0.75. The focus spot was estimated to be about 1 μm when closing the confocal hole to 200 μm. Using a backscattering geometry each spectrum was recorded between 40 cm−1 and 850 cm−1 with a spectral resolution of approx. 3.2 cm−1 using a grating of 1800 grooves/mm and a thermoelectrically cooled CCD detector (Synapse, 1024 × 256 pixels). Fourier transform infrared (FTIR) spectra were collected in the mid-infrared range from 400 cm−1 to 4000 cm−1 on a Bruker VERTEX 80v spectrometer equipped with a DTGS Detector. Standard KBr pellet method was used for data collection.

2.5 Thermal analysis

Simultaneous thermogravimetric analysis and differential scanning calorimetry (TGA/DSC) measurements were performed on a simultaneous TGA/DSC 3+ STAR e system (Mettler Toledo, Schwerzenbach, Switzerland). Approximately 10 mg of each polycrystalline sample was measured with a heating rate of 10 K/min and a continuous N2 flow of 20 mL/min from 300 K to 1500 K. An empty corundum crucible was used as reference. Data were normalized to the respective mass and a drift correction was applied based on empty crucible versus empty crucible data.

3 Results and discussion

3.1 Crystal structure

XRPD data Rietveld refinements confirm that each member of the (PbBi)(Fe1−xMn x )O4 solid solution is found to be phase pure and crystallizes in the space group P42/mbc. A representative Rietveld plot of (PbBi)(Fe0.5Mn0.5)O4 is shown in Figure 2 and the refined crystal structural parameters of the solid solution (PbBi)(Fe1−xMn x )O4 are given in Table S1 (Supplementary information). The metric parameters (a = b, c and V) are given in Table S2, and the interatomic bond distances and angles in Table S3.

Figure 2: 
Rietveld plot of (PbBi)(Fe0.5Mn0.5)O4 schafarzikite-type sample.
Figure 2:

Rietveld plot of (PbBi)(Fe0.5Mn0.5)O4 schafarzikite-type sample.

Complete substitution of Fe3+ with Mn3+ in (PbBi)(Fe1−xMn x )O4, leads to a slight elongation (0.3(1) %) of the (a, b)- and a contraction (−0.7(2) %) of the c-lattice parameters. The latter parameter plays the dominant role for almost a negligible cell volume contraction of −0.1(4) % between the endmembers, as depicted in Figure 3. The observed cell volumes of the endmembers (PbBi)FeO4 (V = 441.35(4)·106 pm3) and (PbBi)MnO4 (V = 440.76(6)·106 pm3) are in excellent agreement with the reported ones. 24 , 31 The slight deviation of each lattice parameter from the Vegard’s rule line can be understood due to these small changes of the metric parameters with respect to the chemical composition. The interplay between the geometric distortion of the MO6 and LO3E polyhedra and the stereochemical activities of the associated LEPs should also be considered to explain the slight deviation from the Vegard’s lines. Moreover, the adoption of the high-spin (HS) and low-spin (LS) configurations of the M cations directly influence the respective ionic radii, and the metric parameters as well. Due to Fe3+/Mn3 mixed occupancy the average ionic radius of the M cations should consider four different spin configurations: Fe3+/Mn3+ (LS–LS), Fe3+/Mn3+ (LS–HS), Fe3+/Mn3+ (HS–LS) and Fe3+/Mn3+ (LS–LS). The trends of the metric parameters fit well to the Fe3+/Mn3+ (LS–HS) configuration (Figure S2).

Figure 3: 
Changes of the metric parameters in (PbBi)(Fe1−xMn
x
)O4 with respect to the x-value of the chemical composition.
Figure 3:

Changes of the metric parameters in (PbBi)(Fe1−xMn x )O4 with respect to the x-value of the chemical composition.

The axial Fe–O1 (202.8(11) pm) and equatorial Fe–O2 (203.8(7) pm) bond lengths for the endmember (PbBi)FeO4 are found to be close (Table S3) to the respective Mn–O1 (200.0(7) pm) and Mn–O2 (200.2(7) pm) bond lengths in (PbBi)MnO4. Due to mixed occupancy the axial (Fe/Mn)–O1 and the equatorial (Fe/Mn)–O2 bond distances lie close to each other irrespective to the chemical composition, as depicted in Figure 4.

Figure 4: 
Fe/Mn–O interatomic bond distances (left), polyhedral volumes (middle) and M–M distances with respect to chemical composition in (PbBi)(Fe1−xMn
x
)O4. The solid line is for eye guideline.
Figure 4:

Fe/Mn–O interatomic bond distances (left), polyhedral volumes (middle) and MM distances with respect to chemical composition in (PbBi)(Fe1−xMn x )O4. The solid line is for eye guideline.

The gradual decrease of the MM distance from 305.3(8) pm (Fe–Fe) to 303.1(6) pm (Mn–Mn) in the octahedral chains can be understood from the deviation between the Fe–O and Mn–O distances, which leads to a contraction of the c-lattice parameter. The MO6 octahedral volume remains almost constant at 10.94(4)·106 pm3 followed by a sharp contraction for x = 1.0(10.42(6)·106 pm3) (Figure 4), indicating a dominant influence of the Fe3+ cations in the structure. The underlying geometric consequences should influence the polyhedral distortions. The mixed valence LEP cations (Pb2+ and Bi3+) on the same atomic position imposes additional distortion to the LO3E tetrahedra 32 due to the asymmetric distribution of the O–Pb–O–Bi–O distances. Indeed, the bond distortion in the LO4E polyhedra shows a slight increase, with an insignificant change of the MO6 polyhedra, for an increasing Mn content in the solid solution (Figure S1).

The bond valence sum of the M- and L-cations are calculated from the M–O and L–O bond distances (Table S3) in the MO6 and LO3 polyhedra, respectively, for any given chemical composition. 33 , 34 The structural BVS of the L-cation exhibits an over-bonding nature above the average 2.5 v.u., as shown in Figure 5, which is a common finding for stereochemically active LEP containing cations. 14 , 15 On the other hand, except for the endmember (PbBi)MnO4 where Mn shows a BVS of 2.65(4) v.u., the BVS of the M3+ cation shows an under-bonding nature below the formal value 3 v.u. with almost a constant deviation of about 0.2 v.u. throughout the solid solution. Such a constant deviation may not be explained due to experimental uncertainty. Since all oxygen atoms of the LO3E polyhedra are linked to the MO6 octahedral apexes, the respective BVS deviations in opposite directions can be understood from geometric distortions. The observed absolute value of the WLE parameters (|Φi|) of 3.72(2)·10−5 and 3.43(2)·10−5 for (PbBi)FeO4 and (PbBi)MnO4, respectively, are within the typical range of a stereochemically active Pb2+/Bi3+ cations in the crystal structures. 14 Most of the |Φi| values lie close to 3.75·10−5, indicating an asymmetric distribution of the LEP 35 followed by a withdrawal of the positive charge of the nucleus toward the bonding oxygen ligands. Such redistribution leads to an increase in the bonding capacity of an L-cation, which may extend the first coordination sphere 36 along with a higher coordination number. This phenomenological concept can further explain the inclusion of Li+ cations into the schafarzikite void channels. 11 According to the Hamani LEP localization model 37 where each L-cation is decomposed into cationic cores (M*) and LEP spatial spheres, as given in Figure 6, allowing for the calculation of the distance between M* and center of the LEP sphere (d M*-LP ). Irrespective of the chemical composition throughout the solid solution the LEP sphere radius of the Pb2+/Bi3+ cations is found to be 120(1) pm and the center of the LEP sphere locates 68(1) pm away from the Pb2+/Bi3+ core which itself possesses a radius of 77(1) pm. Since the shortest interatomic distance between two LEP cations is 375 pm, the estimated corresponding LEP-LEP distance of about 239(1) pm (375 pm − 2*68(1) pm = 239(1) pm) well allows free space in the center of a pseudo-tetrahedron formed by four LEP cations, for a sphere with a radius of ∼80 pm. Such empty space well fits, for instance, to the cationic diameter of lithium of 118 pm. 38 , 39

Figure 5: 
Deviations of bond valence sum (BVS) from formal valence (forV) of the M and L-cations in (PbBi)(Fe1−xMn
x
)O4 (left), and the correlation between the absolute value of the Wang-Liebau eccentricity parameter (|
Φ


i

|) and the BVS for Pb/Bi (right).
Figure 5:

Deviations of bond valence sum (BVS) from formal valence (forV) of the M and L-cations in (PbBi)(Fe1−xMn x )O4 (left), and the correlation between the absolute value of the Wang-Liebau eccentricity parameter (| Φ i |) and the BVS for Pb/Bi (right).

Figure 6: 
Schematic representation for the estimation of effective free space in the (PbBi)FeO4 schafarzikite structure using Hamani
37
 model, where the L-cation is decomposed into a cationic core (M*) and a LEP spatial sphere with radii of rM* and r
LEP
, respectively, and dM*-LEP is the distance between their centers. The free space sphere with a radius of rFS is located at the center of a pseudo-tetrahedron formed by four LEP cations, which fits well for small cations such as lithium.
Figure 6:

Schematic representation for the estimation of effective free space in the (PbBi)FeO4 schafarzikite structure using Hamani 37 model, where the L-cation is decomposed into a cationic core (M*) and a LEP spatial sphere with radii of rM* and r LEP , respectively, and dM*-LEP is the distance between their centers. The free space sphere with a radius of rFS is located at the center of a pseudo-tetrahedron formed by four LEP cations, which fits well for small cations such as lithium.

3.2 Vibrational properties

For the room-temperature Raman and FTIR spectra of (PbBi)(Fe1−xMn x )O4, which are given in Figure 7, the factor group analysis predicts 62 vibrational modes (5A1g + 3A1u + 7A2g + 5A2u + 7B1g + 5B1u + 5B2g + 3B2u + 13Eu + 9Eg). As listed in Table S4, 26 of them are Raman (5A1g + 7B1g + 5B2g + 9Eg), 16 are infrared (4A2u + 12Eu) active, 2 are acoustic (A2u + Eu) and a substantial number of 18 of them are optically silent (3A1u + 7A2g+ 5B1u + 3B2u). Fitting the resolved Raman spectra requires 19 and 17 pseudo-Voigt peaks for the endmembers (PbBi)FeO4 and (PbBi)MnO4, respectively. The vibrational features of each Raman spectrum can be categorized into three spectral regions of low- (<140 cm−1), mid- (140–600 cm−1) and high-frequency (>600 cm−1).

Figure 7: 
Stack plots of FTIR (left) and Raman (right) spectra of (PbBi)(Fe1−xMn
x
)O4.
Figure 7:

Stack plots of FTIR (left) and Raman (right) spectra of (PbBi)(Fe1−xMn x )O4.

The low-frequency region is mainly dominated by the L–O–L and O–L–O bending contributions of the LO4E nido-tetrahedra. The mid-range covers the M–O–M and O–M–O bending of the MO6 octahedra, and the L–O stretching modes of LO4E polyhedra. The high-frequency region is purely associated with the M–O symmetric and asymmetric vibrations in the MO6 octahedra. More details of the vibrational description of the endmember (PbBi)MnO4 are available elsewhere. 24 The low-frequency Raman modes at 106(1) cm−1 and 134(1) cm−1 of (PbBi)FeO4 gradually shift to 95(2) cm−1 and 130(1) cm−1, respectively, as shown in Figure S3, which is mainly due to weaking of the associated bending force constant as the average L-O bond distances between the endmembers (Table S3) does not significantly change. Some high-frequency modes show slight blue shifts, indicating to the observed shortening of the M–O bond lengths in the MO6 octahedra. To better understand the vibrational spectra of the centrosymmetric structures the FTIR spectra of the investigated phases provide complementary information (Figure S4). The compact nature of the FTIR spectra due to merging and the instrumental limitation (>400 cm−1), the shift of the mode frequencies could hardly be followed with changing chemical composition of the solid solution. However, an apparent quasiharmonic trend of each FTIR peak maximum complements the mid- and high-frequency Raman modes.

3.3 Optical band gap

The UV/Vis diffuse reflectance spectra of the endmembers of the (PbBi)(Fe1−xMn x )O4 solid solution are shown in Figure S5 and Figure S6. To determine the optical band-gap transition energy (E g ) from a given absorption spectra by using the Tauc method, 27 the reflectance data are transformed into the corresponding absorption spectra using the Kubelka-Munk function. 40 Absorption data are then recalculated for the use of the Tauc method to calculate a band-gap transition energy assuming a direct (E d ) or indirect (E i ) electron transition. Each analysis is performed by plotting a tangent to the baseline at low energies and a second tangent to the slope in the linear region of the absorption onset. 41 , 42 , 43 The energy of the band gap can then be estimated by the intersection of these two lines (Figure S5). Since the Tauc method cannot assert the nature of the transition type (E d or phonon assisted E i ), the derivative absorption spectra fitting (DASF) method 28 is demonstrated 29 , 44 to compare the optical band-gap types. With gradual increase of Mn content in the solid solution the reflectivity becomes noticeably lower, which hampers calculating the band-gap energies using Tauc’s approach. However, data treated with the DASF method appears as peak shape (Figure S5), allowing for calculation of the band-gap energy (E g ) from the fitted peak-maximum. A comparison between E d and E i with the DASF E g value guides to ascertain the type of transition. In the present, since each band-gap energy obtained by the DASF method (Table S5) agrees well with E d , each member of the solid solution presumably follows direct transition. Beside the band-gap energy, the FWHM (F g ) of the DASF line shape essentially describes the edge of the valence and the conduction bands split into different quantized levels based on the distribution of crystallite sizes and available defects. If one assumes the slope of the absorption edge corresponds directly to the sharpness of the transition (the shaper the energy level separation the flatter the valence and/or the conduction bands in the transition region), the F g of the DASF peak is a direct measure for this behavior. Although the optical band gaps of the schafarzikite family members are not well explored, the experimental E g of minium Pb3O4 (2.1 eV) 45 and MnSb2O4 (2.65(2) eV) 9 are comparable to those of the present compounds. Considering the direct transition (PbBi)FeO4 holds a band gap of 2.24(3) eV which drops to 1.77(3) eV for x = 0.3. Afterward, a slight linear decrease results in 1.72(3) eV for the endmember (PbBi)MnO4, as shown in Figure 8.

Figure 8: 
Changes of direct (E
d
) and indirect (E
i
) electron transfer to the conduction band and band-gap energy (E
g
) with respect to the chemical composition x in (PbBi)(Fe1−xMn
x
)O4.
Figure 8:

Changes of direct (E d ) and indirect (E i ) electron transfer to the conduction band and band-gap energy (E g ) with respect to the chemical composition x in (PbBi)(Fe1−xMn x )O4.

3.4 Thermal properties

The TG and the DSC stack plots of the compounds (PbBi)(Fe1−xMn x )O4, as shown in Figure 9, suggest that samples with higher Mn content exhibit ∼1 % weight loss at around 1080(3) K, 1084(1) K, 1075(3) K, and 1062(1) K, respectively, for x = 0.7, 0.8, 0.9 and 1.0. Such tiny weight loss is most likely associated with the evaporation of Bi-containing amorphous contents. Exothermic signals are observed at 981(1) K and 1000(1) K, respectively, for the endmembers (PbBi)FeO4 and (PbBi)MnO4. The other members (0 < x < 0.7) of the solid solution exhibit an exothermic signal at slightly lower temperatures. The onset temperature (first point of inflection) has been taken as the decomposition temperature, which fluctuates around 1360 K within 20 K for any given composition of the solid solution, as shown in Figure S7.

Figure 9: 
Thermoanalytic evaluation of (PbBi)(Fe1−xMn
x
)O4.
Figure 9:

Thermoanalytic evaluation of (PbBi)(Fe1−xMn x )O4.

4 Conclusions

The solid solution of schafarzikite-type (PbBi)(Fe1−xMn x )O4 has been successfully synthesized. X-ray powder data Rietveld refinements confirm each member to be phase pure and to crystallizes in space group P42/mbc. According to Hume-Rothery rules 46 stricto sensu for intermetallics, the atomic size, the electro-chemical effect and the relative valency between Fe(III) and Mn(III) favor a complete miscibility for a given chemical composition. While the X-ray scattering factors can hardly distinguish between the cations in the same site (8h: Pb(II) and Bi(III); 4d: Fe(III) and Mn(III)), EDX based atomic content unambiguously supports the chemical composition. Both Raman and FTIR spectra complement the structural subunit arrangements. Since the spin-only magnetic moment differs between Fe3+ and Mn3+ along with their spin configurations (high-spin or low-spin), how the magnetic property changes with respect to the Fe3+/Mn3+ in the solid solution leaves scopes for a future study. While incorporation of fluoride anion into the FeSb2O4 structure 11 leads to Fe2+/Fe3+ redox process, the void space formed by the stereochemical active lone electron pairs of these tetragonal phases can be suitable for Li+ insertion for an opposite M3+/M2+ reversible conversions for designing energy materials.


Corresponding author: Mohammad Mangir Murshed, University of Bremen, Institute of Inorganic Chemistry and Crystallography, Leobener Straße 7, D-28359 Bremen, Germany; and University of Bremen, MAPEX Center for Materials and Processes, Bibliothekstraße 1, Universität Bremen, D-28359 Bremen, Germany, E-mail:

Funding source: Deutsche Forschungsgemeinschaft (DFG)

Award Identifier / Grant number: MU4415/3-1

Award Identifier / Grant number: INST144/521-1FUGG

Acknowledgments

The University of Bremen is cordially acknowledged for technical and administrative supports.

  1. Research ethics: The local Institutional Review Board deemed the study exempt from review.

  2. Informed consent: Informed consent was obtained from all individuals included in this study.

  3. Author contributions: Carla M. Uribe-Rincón: drafting the article, acquisition of data, analysis and interpretation of data. Thorsten M. Gesing: conception and design of the study, analysis and interpretation of data, revising the manuscript critically for important intellectual contents. M. Mangir Murshed: conception and design of the study, analysis and interpretation of data, revising the manuscript critically for important intellectual contents.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: The authors hereby state no conflict of interest.

  6. Research funding: Deutsche Forschungsgemeinschaft (DFG) supporting the project within grant MU4415/3-1 (project-ID: 492161439) and DFG the funding of the IR spectrometer under grant number INST144/521-1FUGG.

  7. Data availability: Data can be obtained upon request to the corresponding author.

References

1. Krenner, J. A. Z. Kristallogr. 1921, 56, 198–200.10.1016/S0140-6736(01)38702-0Search in Google Scholar

2. Holtstam, D.; Langhof, J.; Friis, H.; Karlsson, A.; Erambert, M. Eur. J. Mineral. 2024, 36, 311–322; https://doi.org/10.5194/ejm-36-311-2024.Search in Google Scholar

3. Fischer, R. X.; Schneider, H.; Rahman, S.; Freimann, S. Mullite; Wiley VCH: Weinheim, 2005; pp. 1–70.10.1002/3527607358.ch1aSearch in Google Scholar

4. Murshed, M. M.; Gesing, T. M. Acta Crystallogr. A 2012, 68, s195; https://doi.org/10.1107/s0108767312096249.Search in Google Scholar

5. Murshed, M. M.; Mendive, C. B.; Curti, M.; Nenert, G.; Kalita, P. E.; Lipinska, K.; Cornelius, A. L.; Huq, A.; Gesing, T. M. Mater. Res. Bull. 2014, 59, 170–178; https://doi.org/10.1016/j.materresbull.2014.07.005.Search in Google Scholar

6. Gogolin, M.; Murshed, M. M.; Ende, M.; Miletich, R.; Gesing, T. M. J. Mater. Sci. 2020, 55, 177–190; https://doi.org/10.1007/s10853-019-04013-6.Search in Google Scholar

7. Gesing, T. M.; Mendive, C. B.; Curti, M.; Hansmann, D.; Nenert, G.; Kalita, P. E.; Lipinska, K. E.; Huq, A.; Cornelius, A. L.; Murshed, M. M. Z. Kristallogr. 2013, 228, 532–543; https://doi.org/10.1524/zkri.2013.1640.Search in Google Scholar

8. Curti, M.; Mendive, C. B.; Bredow, T.; Mangir Murshed, M.; Gesing, T. M. J. Phys.: Condens. Matter. 2019, 31, 345701; https://doi.org/10.1088/1361-648x/ab20a1.Search in Google Scholar PubMed

9. Roelsgaard, M.; Norby, P.; Eikeland, E.; Sondergaard, M.; Iversen, B. B. Dalton Trans. 2016, 45, 18994–19001; https://doi.org/10.1039/c6dt03459d.Search in Google Scholar PubMed

10. Jibin, A. K.; Reddy, M. V.; Subba Rao, G. V.; Varadaraju, U. V.; Chowdari, B. V. R. Electrochim. Acta 2012, 71, 227–232; https://doi.org/10.1016/j.electacta.2012.03.145.Search in Google Scholar

11. Zaheer, W.; Andrews, J. L.; Parija, A.; Hyler, F. P.; Jaye, C.; Weiland, C.; Yu, Y.-S.; Shapiro, D. A.; Fischer, D. A.; Guo, J.; Velázquez, J. M.; Banerjee, S. ACS Energy Lett. 2020, 5, 2520–2526; https://doi.org/10.1021/acsenergylett.0c01328.Search in Google Scholar

12. Liebau, F.; Wang, X. Z. Kristallogr. 2005, 220, 589–591; https://doi.org/10.1524/zkri.220.7.589.67103.Search in Google Scholar

13. Liebau, F.; Wang, X.; Liebau, W. Chem. – Eur. J. 2009, 15, 2728–2737; https://doi.org/10.1002/chem.200802098.Search in Google Scholar PubMed

14. Wang, X.; Liebau, F. Acta Crystallogr. B 1996, 52, 7–15; https://doi.org/10.1107/s0108768195004472.Search in Google Scholar

15. Wang, F. L. X. Z. Kristallogr. 1996, 211, 437–439.10.1524/zkri.1996.211.7.437Search in Google Scholar

16. Berry, F. J.; de Laune, B. P.; Greaves, C.; Hah, H. Y.; Johnson, C. E.; Johnson, J. A.; Kamali, S.; Marco, J. F.; Thomas, M. F.; Whitaker, M. J. Hyper. Interact. 2018, 239, 8.10.1007/s10751-018-1501-7Search in Google Scholar

17. Cumby, J.; de Laune, B. P.; Greaves, C. J. Mater. Chem. C 2016, 4, 201–208; https://doi.org/10.1039/c5tc03333k.Search in Google Scholar

18. Tiwari, A.; Kakarla, D. C.; Chen, W.-L.; Dhanasekhar, C.; Hu, Y.-J.; Lin, J.-Y.; Wang, C.-W.; Kuo, T.-W.; Pal, A.; Chou, M. M.-C.; Yang, H.-D. Adv. Phys. Res. 2024, 3, 2400056; https://doi.org/10.1002/apxr.202400056.Search in Google Scholar

19. Fjellvåg, H.; Kjekshus, A.; Leskelä, T.; Leskelä, M.; Hoyer, E. Acta Chem. Scand. 1985, 39, 389–395.10.3891/acta.chem.scand.39a-0389Search in Google Scholar

20. de Laune, B. P.; Greaves, C. J. Solid State Chem. 2012, 187, 225–230; https://doi.org/10.1016/j.jssc.2012.01.032.Search in Google Scholar

21. de Laune, B. P.; Rees, G. J.; Whitaker, M. J.; Hah, H. Y.; Johnson, C. E.; Johnson, J. A.; Brown, D. E.; Tucker, M. G.; Hansen, T. C.; Berry, F. J.; Hanna, J. V.; Greaves, C. Inorg. Chem. 2017, 56, 594–607; https://doi.org/10.1021/acs.inorgchem.6b02466.Search in Google Scholar PubMed

22. de Laune, B. P.; Rees, G. J.; Marco, J. F.; Hah, H. Y.; Johnson, C. E.; Johnson, J. A.; Berry, F. J.; Hanna, J. V.; Greaves, C. Inorg. Chem. 2017, 56, 10078–10089; https://doi.org/10.1021/acs.inorgchem.7b01613.Search in Google Scholar PubMed

23. Whitaker, M. J.; Bayliss, R. D.; Berry, F. J.; Greaves, C. J. Mater. Chem. 2011, 21, 14523–14529; https://doi.org/10.1039/c1jm12645h.Search in Google Scholar

24. Uribe-Rincón, C. M.; Ghosh, K.; Frederichs, T.; Gesing, T. M.; Murshed, M. M. J. Solid State Chem. 2025, 125215.10.1016/j.jssc.2025.125215Search in Google Scholar

25. Zemann, J. Tschermaks Mineral. Petro. Mitt. 1951, 2, 166–175; https://doi.org/10.1007/bf01127892.Search in Google Scholar

26. Tauc, J. Mater. Res. Bull. 1968, 3, 37–46; https://doi.org/10.1016/0025-5408(68)90023-8.Search in Google Scholar

27. Tauc, J.; Grigorovici, R.; Vancu, A. Phys. Stat. Solidi (b) 1966, 15, 627–637; https://doi.org/10.1002/pssb.19660150224.Search in Google Scholar

28. Souri, D.; Tahan, Z. E. App. Phys. B 2015, 119, 273–279; https://doi.org/10.1007/s00340-015-6053-9.Search in Google Scholar

29. Kirsch, A.; Murshed, M. M.; Schowalter, M.; Rosenauer, A.; Gesing, T. M. J. Phys. Chem. C 2016, 120, 18831–18840; https://doi.org/10.1021/acs.jpcc.6b04773.Search in Google Scholar

30. Kubelka, F. P. A. M. Z. Technische Phys. 1931, 12, 593–601.10.1007/BF01618784Search in Google Scholar

31. Wu, J.; Yan, X.; Lei, X.; He, L.; Guo, W.; Yan, S.; Kuang, X.; Yin, C. J. Solid State Chem. 2023, 318; https://doi.org/10.1016/j.jssc.2022.123770.Search in Google Scholar

32. Halasyamani, P. S. Chem. Mater. 2004, 16, 3586–3592; https://doi.org/10.1021/cm049297g.Search in Google Scholar

33. Brese, N. E.; O’Keeffe, M. Acta Crystallogr. B 1991, 47, 192–197.10.1107/S0108768190011041Search in Google Scholar

34. Brown, I. D.; Altermatt, D. Acta Crystallogr. B 1985, 41, 244–247; https://doi.org/10.1107/s0108768185002063.Search in Google Scholar

35. Murshed, M. M.; Rusen, A.; Fischer, R. X.; Gesing, T. M. Mater. Res. Bull. 2012, 47, 1323–1330; https://doi.org/10.1016/j.materresbull.2012.03.014.Search in Google Scholar

36. Brandes, R.; Hoppe, R. Z. Anorg. Allg. Chem. 1994, 620, 1549–1558; https://doi.org/10.1002/zaac.19946200910.Search in Google Scholar

37. Hamani, D.; Masson, O.; Thomas, P. J. App. Crystallogr. 2020, 53, 1243–1251; https://doi.org/10.1107/s1600576720010031.Search in Google Scholar

38. Shannon, R. D. Acta Crystallogr. A 1976, 32, 751–767; https://doi.org/10.1107/s0567739476001551.Search in Google Scholar

39. Shannon, R. D.; Prewitt, C. T. Acta Crystallogr. B 1969, 25, 925–946; https://doi.org/10.1107/s0567740869003220.Search in Google Scholar

40. Kubelka, P. J. Opt. Soc. Am. 1948, 38, 448–457; https://doi.org/10.1364/josa.38.000448.Search in Google Scholar PubMed

41. M. M. R. Center; Department of Chemical Engineering, Stanford University, 2017. Available from: http://mmrc.caltech.edu.Search in Google Scholar

42. Makuła, P.; Pacia, M.; Macyk, W. J. Phys. Chem. Lett. 2018, 9, 6814–6817; https://doi.org/10.1021/acs.jpclett.8b02892.Search in Google Scholar PubMed

43. Escobedo-Morales, R.-L. A.; Ruiz-Peralta, M. de L.; Tepech-Carrillo, L.; Sánchez-Cantú, M.; Moreno-Orea, J. E. Heliyon 2019, 5, e01505.10.1016/j.heliyon.2019.e01505Search in Google Scholar PubMed PubMed Central

44. Teck, M.; Murshed, M. M.; Schowalter, M.; Lefeld, N.; Grossmann, H. K.; Grieb, T.; Hartmann, T.; Robben, L.; Rosenauer, A.; Madler, L.; Gesing, T. M. J. Solid State Chem. 2017, 254, 82–89; https://doi.org/10.1016/j.jssc.2017.07.013.Search in Google Scholar

45. Terpstra, H. J.; De Groot, R. A.; Haas, C. J. Phys. Chem. Solids 1997, 58, 561–566; https://doi.org/10.1016/s0022-3697(96)00165-5.Search in Google Scholar

46. Hume-Rothery, W.; Smallman, R. E.; Haworth, C. W. The Structure of Metals and Alloys; Metals & Metallurgy Trust: London, 1969.Search in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/zkri-2025-0023).


Received: 2025-05-24
Accepted: 2025-06-04
Published Online: 2025-06-23

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 9.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/zkri-2025-0023/html?licenseType=open-access
Scroll to top button