Abstract
Bi2Fe4O9 was synthesized using a polyol-mediated method. X-ray powder diffraction (XRPD) revealed that the as-synthesized sample is nano-crystalline. During heating, the X-ray amorphous powder transformed into a rhombohedral perovskite-type bismuth ferrate followed by a second transformation into mullite-type Bi2Fe4O9 at higher temperatures. This transformation was studied at in-situ conditions by temperature-dependent XRPD and 57Fe Mössbauer spectroscopy. The 57Fe Mössbauer spectra indicate the existence of two Fe3+ species at two different octahedrally coordinated sites leading to the conclusion that the as-synthesized powder of the polyol synthesis possesses a disordered (Bi1–xFex)FeO3 perovskite structure. Rietveld refinements have unambiguously supported this observation and this results suggest that one third of the Bi3+ sites are substituted by Fe3+ representing the initial chemical composition. This study has shown that as-synthesized nano-materials are not always similar to the respective micro-crystalline ones.
1 Introduction
Bismuth ferrates, BiFeO3 and Bi2Fe4O9, have attracted great attention due to their electronic [1], optic [2], catalytic [3], and magnetic [4] properties making them suitable for advanced applications like photocatalysis [5, 6], photovoltaics [7, 8], and spintronics [9]. Bi2Fe4O9 is mainly studied as a gas sensor [10] and as a photocatalytically [11] active material. It also shows multiferroic properties near room temperature [12], interesting crystal chemistry [13–16] and temperature-dependent behavior [17] due to the stereochemically active Bi3+ 6s2 lone electron pairs (LEPs) [18]. Bi2Fe4O9 crystallizes in the orthorhombic space group Pbam with Z = 2 and possesses a mullite-type structure [14] featuring the typical chains of edge-sharing FeO6 octahedra parallel to the c axis. The chains are cross-linked by corner-sharing Fe2O7 double-tetrahedra. The Bi atoms are located between the chains and their LEPs point toward the vacant sites of the channels [17, 18]. In the Bi2O3-Fe2O3 quasi-binary phase field co-crystallization of Bi2Fe4O9 and BiFeO3 is frequently observed [19–21].
BiFeO3 is mostly studied owing to its multiferroic properties at room temperature and photocatalytic activity in the visible range of light [5, 22]. The widely studied BiFeO3 has a Néel temperature of ~640 K and a Curie temperature of ~1100 K [23]. The crystal structure of BiFeO3 as given in Fig. 1 can be described in the space group R3c (Z = 2) as a rhombohedrally distorted perovskite. The tilting of the FeO6 octahedra around the pseudo-cubic threefold <111> axis is a characteristic structural feature. The displacement of the Bi3+ and Fe3+ cations along this axis causes spontaneous polarization [24]. From the distorted 6-fold coordination of the Bi atoms with oxygen a stereochemical activity of the LEP can be assumed.

Crystal structures of perovskite-type BiFeO3 in space group R3c (a) and of mullite-type Bi2Fe4O9 in space group Pbam (b).
Zhang et al. [2] demonstrated that small crystallites of Bi2Fe4O9 exhibit higher photocatalytic activity. Therefore the synthesis and characterization of this compound as nano-material up to ~30 nm is of special interest for improved photocatalytic efficiency. Bi2Fe4O9 was synthesized through a variety of synthesis methods including conventional solid state reaction [25], chemical co-precipitation [1], sol-gel [19] and EDTA routes [20], the glycerine method [13], and hydrothermal processes [26]. In this context, the polyol method [27, 28] provides a cost effective and easy method for the preparation of larger quantities of nano-crystalline samples. In the present study, a precursor was synthesized by the polyol method and characterized by Fourier transform infrared (FTIR) spectroscopy. The transformation of the perovskite-type (ABX3) to the mullite-type (A2B4X9) compound was studied via in-situ temperature-dependent X-ray powder diffraction (XRPD) and 57Fe Mössbauer spectroscopy.
2 Experimental
2.1 Synthesis
Samples were prepared by a polyol-mediated synthesis. In detail, 5 mmol Bi(NO3)3·5H2O (Sigma-Aldrich, ≥98 %), 10 mmol Fe(NO3)3·9H2O (Sigma-Aldrich, ≥98 %) and 40 mmol NaOH (VWR Chemicals, 99.2 %) were dissolved in 42 ml diethylene glycol (DEG) (AppliChem, 99 %). The mixture was heated in a three-necked flask under reflux using a metal bath at 423 K. After 2 h 5 mmol stearic acid was added. After cooling down to ambient conditions, the solid product was separated by centrifugation and washed several times with acetone with intermediate centrifugation steps. For comparison samples were also synthesized hydrothermally at 473 K for 24 h in Teflon coated steel autoclaves. Two series of samples were produced. First the molar ratio of Bi(NO3)3·5H2O to Fe(NO3)3·9H2O was varied from 1:2 to 2:2 in steps of 0.2 using always 10 mmol Fe(NO3)3·9H2O in 20 mL of a 2 mol L–1 NaOH solution. Thereafter a molar 1:1 ratio was used increasing the NaOH concentration up to 12 mol L–1 in steps of 2 mol L–1. After the heat treatment the samples where intensively washed with deionized water. This process was repeated three times before the samples were dried at 393 K for about 12 h.
2.2 Spectroscopy
The FTIR spectra were measured on a Bruker IFS 66v/S spectrometer using the standard KBr method (1 mg sample in 200 mg KBr) between 370 and 4000 cm–1. Background as well as sample spectra were obtained from 128 scans each with a spectral resolution of approx. 1 cm–1. The mode positions were determined by taking the point of the maximum intensity. The temperature-dependent 57Fe Mössbauer spectra were taken between 293 and 973 K in an atmosphere of flowing synthetic air using a standard transmission Mössbauer spectrometer (Halder) in the sinusoidal driving mode employing a 57Co/Rh γ-radiation source with a maximum activity of 1.91 GBq. The velocity scale was calibrated with an α-Fe absorber at room temperature and the isomer shifts (IS) are stated relative to the center of this calibration.
2.3 X-ray diffraction
XRPD data were collected on a X’Pert MPD PRO diffraction system (PANalytical GmbH, Almelo, The Netherlands) equipped with Ni-filtered CuKα1,2 radiation (λα1,2 = 0.15406 nm, 0.15444 nm), a 0.25° divergence slit, a 0.5° antiscatter slit, a 0.04 rad soller slit in the primary beam and a X’Celerator detector system in the secondary beam in Bragg-Brentano geometry. Room-temperature data were measured between 5 and 85° 2θ with a step width of 0.0167° 2θ and a measurement time of 20 s per step. A HTK 1200N heating chamber (Anton Paar, Graz, Austria) was used for temperature-dependent measurements. Data were collected from 5 to 100° 2θ with a step width of 0.0167° 2θ and a measurement time of 75 s per step. The temperature was increased stepwise from 300 to 1120 K in 20 K steps with an equilibration time of 5 min. The obtained data were refined using the Rietveld method (topas V4.2, Bruker AXS, Karlsruhe, Germany). For the profile description the fundamental parameter approach was used, where the fundamental parameters were fitted against a LaB6 standard material.
2.4 Scanning electron microscopy
Scanning electron microscopy (SEM) was carried out on a JSM-6510 (JEOL GmbH, Munich, Germany) equipped with energy dispersive X-ray (EDX) analysis facilities and an XFlash Detector 410-M (Bruker AXS GmbH, Karlsruhe, Germany). To obtain quality data the samples were sputtered with a thin film of gold having a thickness of approx. 10 nm. EDX spectra were collected using an excitation voltage of 20 kV.
3 Results and discussion
3.1 Synthesis
The synthesis of Bi2Fe4O9 using the hydrothermal method was described by Wang et al. [29]; the process strongly depends on the hydroxide concentration during the heat treatment at 373 K. The authors used a multi-step synthesis in which the metal nitrates were first dissolved in nitric acid, which was then brought, by adding slowly dropwise a KOH solution, to a pH = 8 leading to a precipitation of a brown solid. This solid was filtered, washed and transferred into a NaOH solution which was then treated hydrothermally. With a NaOH concentration of 2 mol L–1 they observed the formation of a perovskite phase. Note that the authors [29] took this as the beginning of the crystallization of Bi2Fe4O9, but from the reported diffraction pattern the formation of a perovskite is obvious. With increasing concentration the mullite-type phase was formed. To make the synthesis easier, we simply added the metal nitrates to an aqueous NaOH solution and kept it at 473 K for 24 h. For a first experiment a 2 mol L–1 NaOH solution with a Bi to Fe ratio of 1.2:2 was used to form the Bi2Fe4O9 mullite-type phase. Rietveld refinements of XRPD data confirmed that the product consists of 67 wt% BiFeO3, 24 wt% Fe2O3 and 9 wt% Bi2Fe4O9. However, using a ratio of 1.4:2 as much as 93 wt% Bi2Fe4O9 was formed together with 7 wt% BiFeO3. A further increase of the molar ratio to 1.6:2 and 1.8:2 leads to pure well crystallized Bi2Fe4O9 phases as shown in Fig. 2. The latter sample was used for further investigations. Increasing the cation ratio further to 1:1 and increasing the NaOH concentration at this ratio to 6, 9 or 12 mol L–1 leads to an increase of the average crystallite size, however, along with the formation of a few percent of the sillenite phase Bi25FeO39 [30].

Scanning electron micrographs of Bi2Fe4O9 synthesized by the hydrothermal method (a), by the polyol method (b) and the polyol sample obtained after the XRD heating measurements, showing Bi2Fe4O9 rods and needles (c and d).
The synthesis of nanoscale metal oxides using the polyol method was described by several authors [28, 31, 32]. The metal precursors were heated in a high-boiling alcohol such as most commonly used diethylene glycol (DEG) [31], ethylene glycol [22], triethylene glycol [33] and tetraethylene glycol [33] at elevated temperatures. Using this method various metal oxides can be produced showing particle sizes below ~100 nm [31, 33, 34]. However, to produce a crystalline sample a subsequent heat treatment is often required [19, 31, 35]. In this study, metal nitrates were chosen as metal precursors. They were dissolved in DEG in a three-necked flask followed by addition of 40 mmol NaOH. The temperature of the mixture was then ramped to the target reflux temperature of 423 K, concomitantly releasing NOx gas and forming a brown precipitate at 388 K. Notably, the reflux (423 K) started far below the boiling point of DEG (519 K) owing to the high content of hydrate water in the metal nitrates. The as-synthesized particles are spherical in shape and possess a size distribution ranging from 200 to 700 nm as shown by SEM micrograph analysis (Fig. 2). XRPD gave no clear Bragg reflections. However, well crystalline mullite-type Bi2Fe4O9 could be produced via a subsequent heat treatment. Whereas the hydrothermal synthesis nicely produced crystalline mullite-type material the structure of the polyol synthesized material could hardly be characterized due to the very small average crystallite size leading to very broad diffraction reflections as for an X-ray amorphous material.
3.2 Spectroscopy
3.2.1 Infrared spectroscopy
The infrared spectrum of the hydrothermally synthesized sample corresponds to that reported earlier for Bi2Fe4O9 [36], whereas numerous different absorption bands are visible in the FTIR spectrum of the Bi2Fe4O9 nano-material synthesized by the polyol method as shown in Fig. 3. Most of them are resulting from the (mainly organic) compounds used in the synthesis. The broad absorption bands at 3200 and 1650 cm–1 can be assigned to water in the sample [37]. The NO3– anions are evidenced by the sharp absorption feature at 1385 cm–1. The C–H stretching bands of stearic acid are represented by modes between 2850 and 2920 cm–1 [38], whereas the respective carboxyl groups could be identified by their modes at 1670 cm–1. Remaining DEG (after washing the samples with acetone) leads to the observation of C–O–C, C–OH and C–O vibrations with peak maxima at 1125(3), 1060(3) and 900(3) cm–1, respectively [39]. Octahedral Fe–O stretching vibrations in bismuth ferrates are located in the region between 445 and 548 cm–1 (BiFeO3 [40]) and between 437 and 471 cm–1 (Bi2Fe4O9 [36]). The appearance of absorption peaks around 500 cm–1 of the Bi2Fe4O9 nano-material could therefore be assigned to the Fe–O stretching of FeO6 octahedra. The assignments of the band positions of the vibrational modes of the hydrothermally synthesized mullite-type phase (Fig. 3) agree well with those of Voll et al. [36]. In the spectrum of this sample an additional strong mode with the maximum intensity at 812(1) cm–1 belongs to the Fe–O–Fe stretching mode of the Fe2O7 double tetrahedral unit interconnecting the octahedral chains (as calculated for the isotypic aluminum compound [41]), which is absent in the polyol Bi2Fe4O9 nano-material material.

Infrared spectra of Bi2Fe4O9 synthesized by the hydrothermal method (a) and precursor formed by the polyol method (b), diethylene glycol (DEG) (c), Bi(NO3)3·5H2O (d), and stearic acid (e).
3.2.2 57Fe Mössbauer spectroscopy
To get more insight into the coordination of the iron atoms in the synthesized samples, 57Fe Mössbauer data were collected at different temperatures. The fit of the room-temperature 57Fe Mössbauer spectrum of the polyol-synthesized sample (only polyol from now on) reveals the existence of two Fe3+ species in two different octahedrally coordinated sites, Table 1. The majority component shows exactly the same isomer shift (IS) as the octahedrally coordinated Fe site in Bi2Fe4O9 [15]; thus identical coordination and similar bond lengths are expected. The minority component shows a slightly higher IS, indicating a lower electron charge density at the nucleus, pointing to slightly larger bond lengths. The quadrupole splitting (QS) of the two octahedrally coordinated Fe sites in the polyol material is significantly larger than that of the octahedrally coordinated site in Bi2Fe4O9 [15]. In this context, it is interesting to compare the present results with those of nano-crystalline BiFeO3. Park et al. [4] have studied the Mössbauer spectra of BiFeO3 perovskites with crystallite sizes between <14 nm and about 100 nm. The present Mössbauer spectroscopic parameters exhibit some similarity with those of the sample with <14 nm (Table 1), for example in respect to the relative signal area, indicating a similar distribution of iron over two inequivalent sites in the structure. As seen, also quadrupolar interactions take similarly large values indicating significant coordinative distortion in both materials. However, IS values of the nano-perovskite are significantly lower, indicating a higher electron charge density at the nucleus than in the polyol material. The large room-temperature linewidth (Γ = 0.4 mm s–1, Table 1) of the polyol material provides evidence of some kind of disorder around the two Fe-occupied sites. In conclusion, the polyol material appears to be of disordered nature and to possess an unknown local structure with two non-equivalent octahedral sites for Fe3+ occurring in an approximate 1:3 ratio. According to these signal intensities the material could, for instance, represent a (Bi2Fe)Fe3O9 perovskite.
Hyperfine parameters of room-temperature 57Fe Mössbauer spectra of Bi2Fe4O9 powder prepared by the polyol method, as compared to Bi2Fe4O9 synthesized by the glycerine method [15] and to bulk and nano BiFeO3 [4].
Sample | IS 1/mm s–1 | QS 1/mm s–1 | A 1/% | IS2/mm s–1 | QS 2/mm s–1 | A 2/% | Γ/mm s–1 |
---|---|---|---|---|---|---|---|
polyol | 0.38(1) | 1.09(1) | 27(1) | 0.36(1) | 0.70(1) | 73(1) | 0.40(1) |
hydrothermal | 0.24(1) | 0.95(1) | 29(1) | 0.36(1) | 0.38(1) | 29(1) | 0.25(1) |
Bi2Fe4O9 [15] | 0.23(1) | 0.95(1) | 50(1) | 0.35(1) | 0.37(1) | 50(1) | 0.22(1) |
BiFeO3 [4] bulk | 0.39(3) | –0.10(5)a | 47(2) | 0.38(3) | 0.34(5) | 53(2) | 0.35(2) |
BiFeO3 [4] <14 nm | 0.31(3) | 1.29(5) | 25(2) | 0.33(3) | 0.79(5) | 73(2) | 0.23(2) |
aData from the magnetic bulk material represent the quadrupolar perturbation parameter ε = QS·(3cos2θ–1)/2, where θ is the angle between the direction of the principal component of the electric field gradient and the direction of the local magnetic field.
Temperature-dependent Mössbauer spectra of the hydrothermal and polyol samples are shown in Fig. 4. Whereas the thermal evolution of the 57Fe hyperfine parameters in Bi2Fe4O9 prepared by the hydrothermal method is exactly the same as reported for bulk Bi2Fe4O9 [15], the parameters for the polyol sample show a different behavior. During heating the spectral features change between 573 and 773 K, showing evidence for structural changes leading to the appearance of two differently coordinated iron sites. The two sites of octahedral coordination of Fe3+ in (Bi2Fe)Fe3O9 transform into one tetrahedral and one octahedral site. Finally, at 973 K the spectrum is identical to that of Bi2Fe4O9 prepared by the hydrothermal method. The temperature dependence of isomer shifts (IS) and quadrupole splittings (QS) of Bi2Fe4O9 prepared by the polyol method is presented in Fig. 5.

Temperature-dependent 57Fe Mössbauer spectra of Bi2Fe4O9 powder materials prepared by the polyol (right) and hydrothermal method (left) in flowing synthetic air. The subspectrum shown in red is attributed to a small admixture of Fe2O3.

Temperature-dependent isomer shifts (IS) and quadrupole splittings (QS) of Bi2Fe4O9 prepared by the hydrothermal (left) and polyol method (right).
3.3 Temperature-dependent X-ray powder diffraction
The in-situ heating XRPD experiment of the as-synthesized polyol material showed a transformation from an X-ray amorphous powder into a rhombohedral perovskite-type structure at 680 K followed by a second transformation into Bi2Fe4O9 starting from 800 K on as shown in Fig. 6. Above 920 K the perovskite structure cannot be detected anymore. The sharp reflections which appear at diffraction angles greater than 57° 2θ above 620 K are attributed to the sample holder due to mechanical shrinking of the powder sample.

Temperature-dependent diffraction patterns of the polyol material showing the transformation from as-synthesized amorphous powder into rhombohedral BiFeO3 and orthorhombic Bi2Fe4O9 (different scale range above 900 K for better visibility).
Rietveld refinements of the temperature-dependent data of the polyol material were carried out using the structure model of Moreau et al. [42] for BiFeO3 with R3c symmetry and lattice parameters of a = 558.76(3) pm and c = 1386.7(1) pm in the hexagonal setting. Taking the results of the 57Fe Mössbauer study into account, Fe3+ and Bi3+ were both refined on the same 6a position (0, 0, 0(fixed)) whereas the second 6a position (0, 0, 0.2215(18) at 780 K) was taken as solely occupied by iron ions. The oxygen atoms were found on the 18b position (0.421(11), 0.011(8), 0.952(5) at 780 K). With increasing temperature the (Bi1–xFex)FeO3 crystal structure was observed to be rapidly changing as depicted in Fig. 7. The lattice parameter a increases from 561.4(4) pm (at 740 K) to 562.0(2) pm at 780 K and then decreases to 561.62(6) pm at 860 K as shown in Fig. 8. The lattice parameter c of (Bi1–xFex)FeO3 is significantly smaller than the one reported for BiFeO3 [42], supporting the site co-shared by Fe3+ and Bi3+. Accordingly, the shrinking of the lattice parameter can be explained in terms of the smaller ionic radius of Fe3+ [43]. Starting from 740 K, the lattice parameter c increases linearly with temperature from 1375(2) pm and approaches the value reported for BiFeO3 [42] at about 800 K (c = 1383(1) pm). Notably, this is the temperature at which the lattice parameter a as well as the strain, as shown in Fig. 9, decrease rapidly. The unit cell volume increases linearly with temperature. At 860 K, reflections appear (marked with crosses in Fig. 7) that can be attributed to Fe-bearing sillenite Bi25FeO39 in space group I23 [30]. Above 920 K, neither reflections of sillenite nor of the perovskite structure can be detected.

Left panel: Temperature-dependent XRPD patterns from 680 to 920 K in 20 K steps showing the evolution of (Bi1–xFex)FeO3 and Bi2Fe4O9 phases. Some intense reflections of the BiFeO3 (104) (110) and Bi2Fe4O9 phase (121) (211) are indicated by diamonds and stars, respectively. Reflections marked by a cross can be attributed to Bi25FeO39. Right panel: Magnified feature of a part of the left panel, showing the appearance and disappearance of reflections marked by the cross.

Temperature-dependent lattice parameter a (left), c (middle) and cell volume (right) of (Bi1–xFex)FeO3.

Temperature-dependent strain (left) and occupancy of Bi3+ (right) of the (Bi1–xFex)FeO3 phase.
Substitution of Fe3+ with an effective ionic radius of 64.5 pm [43] on the Bi position causes a high strain in the unit cell due to the much larger ionic radius of Bi3+ 103 pm [43]. The micro-strain at 740 K could be refined to as much as 0.41(1) % giving further evidence that Bi3+ is partially substituted by Fe3+. The micro-strain is reduced to about 0.09(1) % at 860 K. The Rietveld refinement results in an occupancy of 69(2) % for Bi3+ at 740 K, the rest is occupied by Fe3+ according to the constrained occupancy parameter refinements assuming a fully occupied position. This value fits well to the Mössbauer spectroscopic results, which have suggested 1/3 of the occupancy probability. With increasing temperature the amount of Bi3+ on the position is increasing (Fig. 9).
After completion of the temperature-dependent XRD experiments the samples were investigated by scanning electron microscopy. Needle-like rods and plates were identified as shown in the SEM micrograph of Fig. 2. It is assumed that the long chains of the stearic acid present in the precursor caused the pronounced unidirectional growth. The sample contains larger crystals as well as very thin needle-like plates (Fig. 2). The length of the thin needles ranges from 5 to 10 μm, the width takes values of about 250 nm. Due to their transparency in SEM micrographs these needles are assumed to be only a few nm thick. EDX analyses revealed a molar ratio of Bi:Fe of 1:2.1(2) as expected for Bi2Fe4O9. Nevertheless, a remaining content of 3.3(3) mol% Na was found, but we assume, based on the substitution experience with this structure type (e.g. [44]), that sodium is not incorporated in the mullite-type structure.
4 Conclusion
Bi2Fe4O9 was synthesized by a polyol-mediated synthesis. It is assumed that the growth along the chains of stearic acid present in the precursor material leads to needle-like rods and plates during heating experiments. Because smaller dimensions of Bi2Fe4O9 crystals show higher photocatalytical activity [2], such few nm thick plates would be useful as photocatalytically active material. The comparative temperature-dependent 57Fe Mössbauer spectroscopic study demonstrated that the sample synthesized by the polyol method is of different nature than the sample synthesized by the hydrothermal route. During temperature-dependent studies, the polyol material gradually approached the features of the hydrothermal sample. In addition, the transformation of the as-synthesized amorphous polyol sample into rhombohedral (Bi1–xFex)FeO3 perovskite and subsequent transformation into the orthorhombic Bi2Fe4O9 mullite-type structure was confirmed by temperature-dependent X-ray diffraction data analysis showing that the chemical composition of the educts with respect to the metal atoms is conserved during the double reconstructive phase transition process. Mössbauer spectroscopic results as well as Rietveld refinements of temperature-dependent XRPD data suggest that about one third of the octahedral Bi3+ sites in the (Bi1–xFex)FeO3 perovskite is substituted by Fe3+ leading to high strain in the system and a smaller lattice parameter c. Moreover, temperature-dependent Mössbauer spectroscopic and X-ray diffraction investigations have shown that nano-crystalline Bi2Fe4O9 (Bi2O3*2Fe2O3) transforms via the (Bi2Fe)Fe3O9 perovskite [(Bi1–xFex)FeO3] to micro-crystalline mullite-type Bi2Fe4O9, which clearly shows that a nano-crystalline product has to be proven for structural identity with its micro-crystalline appearance before stating the structural relation.
Dedicated to: Professor Wolfgang Jeitschko on the occasion of his 80th birthday.
Acknowledgments
KDB thanks the State of Lower Saxony (Germany) and the Volkswagen Foundation for financial support. TMG gratefully acknowledges the Deutsche Forschungsgemeinschaft (DFG) for the financial support in the Heisenberg program (GE1981/3-1 and GE1981/3-2) and the mullite-LEP project (GE1981/4-1 and GE1981/4-2). We also thank the unknown reviewer for his precise suggestions.
References
[1] Y. Li, Y. Zhang, W. Ye, J. Yu, C. Lu, L. Xia, New J. Chem.2012, 36, 1297.10.1039/c2nj40039aSearch in Google Scholar
[2] Q. Zhang, W. Gong, J. Wang, X. Ning, Z. Wang, X. Zhao, W. Ren, Z. Zhang, J. Phys. Chem. C2011, 115, 25241.10.1021/jp208750nSearch in Google Scholar
[3] N. I. Zakharchenko, Kinet. Catal.2002, 43, 95.10.1023/A:1014209415066Search in Google Scholar
[4] T. Park, G. C. Papaefthymiou, A. J. Viescas, A. R. Moodenbaugh, S. S. Wong, Nano Lett.2007, 7, 766.10.1021/nl063039wSearch in Google Scholar
[5] F. Gao, X. Chen, K. Yin, S. Dong, Z. Ren, F. Yuan, T. Yu, Z. Zou, J. M. Liu, Adv. Mater.2007, 19, 2889.10.1002/adma.200602377Search in Google Scholar
[6] S. Sun, W. Wang, L. Zhang, M. Shang, J. Phys. Chem. C2009, 113, 12826.10.1021/jp9029826Search in Google Scholar
[7] Y. Zang, D. Xie, X. Wu, Y. Chen, Y. Lin, M. Li, H. Tian, X. Li, Z. Li, H. Zhu, T. Ren, D. Plant, Appl. Phys. Lett.2011, 99, 132904.10.1063/1.3644134Search in Google Scholar
[8] Y. Zhang, Y. Guo, H. Duan, H. Li, L. Yang, P. Wang, C. Sun, B. Xu, H. Liu, RSC Adv.2014, 4, 28209.10.1039/C4RA01727GSearch in Google Scholar
[9] G. Catalan, J. F. Scott, Adv. Mater.2009, 21, 2463.10.1002/adma.200802849Search in Google Scholar
[10] A. S. Poghossian, H. V. Abovian, P. B. Avakian, S. H. Mkrtchian, V. M. Haroutunian, Sensors Actuators B1991, 4, 545.10.1016/0925-4005(91)80167-ISearch in Google Scholar
[11] Z. Liu, B. Wu, Y. Zhu, Mater. Chem. Phys.2012, 135, 474.10.1016/j.matchemphys.2012.05.010Search in Google Scholar
[12] A. K. Singh, S. D. Kaushik, B. Kumar, P. K. Mishra, A. Venimadhav, V. Siriguri, S. Patnaik, Appl. Phys. Lett.2008, 92, 132910.10.1063/1.2905815Search in Google Scholar
[13] T. M. Gesing, R. X. Fischer, M. Burianek, M. Mühlberg, T. Debnath, C. H. Rüscher, J. Ottinger, J.-C. Buhl, H. Schneider, J. Eur. Ceram. Soc.2011, 31, 3055.10.1016/j.jeurceramsoc.2011.04.004Search in Google Scholar
[14] H. Schneider, R. X. Fischer, T. M. Gesing, J. Schreuer, M. Mühlberg, Int. J. Mater. Res.2012, 103, 422.10.3139/146.110716Search in Google Scholar
[15] S.-U. Weber, T. M. Gesing, J. Röder, F. J. Litterst, R. X. Fischer, K.-D. Becker, Int. J. Mater. Res.2012, 103, 430.10.3139/146.110701Search in Google Scholar
[16] S. Weber, T. M. Gesing, G. Eckold, R. X. Fischer, F. J. Litterst, K.-D. Becker, J. Phys. Chem. Solids2014, 75, 416.10.1016/j.jpcs.2013.11.013Search in Google Scholar
[17] M. M. Murshed, G. Nénert, M. Burianek, L. Robben, M. Mühlberg, H. Schneider, R. X. Fischer, T. M. Gesing, J. Solid State Chem.2013, 197, 370.10.1016/j.jssc.2012.08.062Search in Google Scholar
[18] M. Curti, T. M. Gesing, M. M. Murshed, T. Bredow, C. B. Mendive, Z. Kristallogr.2013, 228, 629.Search in Google Scholar
[19] X. Wang, M. Zhang, P. Tian, W. S. Chin, C. M. Zhang, Appl. Surf. Sci.2014, 321, 144.10.1016/j.apsusc.2014.09.166Search in Google Scholar
[20] J. Zhao, T. Liu, Y. Xu, Y. He, W. Chen, Mater. Chem. Phys.2011, 128, 388.10.1016/j.matchemphys.2011.03.011Search in Google Scholar
[21] X. H. Wu, J. Miao, Y. Zhao, X. B. Meng, X. G. Xu, S. G. Wang, Y. Jiang, Optoelectron. Adv. Mater. Rapid Commun.2013, 7, 116.Search in Google Scholar
[22] X. Wang, Y. Lin, Z. C. Zhang, J. Y. Bian, J. Sol-Gel Sci. Technol.2011, 60, 1.10.1007/s10971-011-2542-4Search in Google Scholar
[23] P. Fischer, M. Polomska, I. Sosnowska, M. Szymanski, J. Phys. C1980, 13, 1931.10.1088/0022-3719/13/10/012Search in Google Scholar
[24] B. Ruette, S. Zvyagin, A. P. Pyatakov, A. Bush, J. F. Li, V. I. Belotelov, A. K. Zvezdin, D. Viehland, Phys. Rev. B2004, 69, 064114.10.1103/PhysRevB.69.064114Search in Google Scholar
[25] H. Koizumi, N. Niizeki, T. Ikeda, Jpn. J. Appl. Phys.1964, 3, 495.10.1143/JJAP.3.495Search in Google Scholar
[26] Q.-J. Ruan, W.-D. Zhang, J. Phys. Chem. C2009, 113, 4168.10.1021/jp810098fSearch in Google Scholar
[27] F. Fievet, J. P. Lagier, M. Figlarz, MRS Bull.1989, 14, 29.10.1557/S0883769400060930Search in Google Scholar
[28] C. Feldmann, Solid State Sci.2005, 7, 868.10.1016/j.solidstatesciences.2005.01.018Search in Google Scholar
[29] Y. Wang, G. Xu, L. Yang, Z. Ren, X. Wei, W. Weng, P. Du, G. Shen, G. Han, Ceram. Int.2009, 35, 51.10.1016/j.ceramint.2007.09.114Search in Google Scholar
[30] A. M. Lopes, J. P. Araujo, S. Ferdov, Dalt. Trans2014, 43, 18010.10.1039/C4DT01825GSearch in Google Scholar
[31] C. Quievryn, S. Bernard, P. Miele, Nanomater. Nanotechnol.2014, 4, 1.10.5772/58289Search in Google Scholar
[32] C. Feldmann, H. O. Jungk, Angew. Chem. Int. Ed.2001, 40, 359.10.1002/1521-3773(20010119)40:2<359::AID-ANIE359>3.0.CO;2-BSearch in Google Scholar
[33] W. Cai, J. J. Wan, Colloid Interface Sci.2007, 305, 366.10.1016/j.jcis.2006.10.023Search in Google Scholar
[34] H.-O. Jungk, C. Feldmann, J. Mater. Sci.2001, 36, 297.10.1023/A:1004895605613Search in Google Scholar
[35] C. Feldmann, Adv. Funct. Mater.2003, 13, 101.10.1002/adfm.200390014Search in Google Scholar
[36] D. Voll, A. Beran, H. Schneider, Phys. Chem. Miner.2006, 33, 623.10.1007/s00269-006-0108-8Search in Google Scholar
[37] L. Kovács, K. Lengyel, Á. Péter, K. Polgár, A. Beran, Opt. Mater. (Amst).2003, 24, 457.10.1016/S0925-3467(03)00028-4Search in Google Scholar
[38] F. Kimura, J. Umemura, T. Takenaka, Langmuir1986, 2, 96.10.1021/la00067a017Search in Google Scholar
[39] T. T. Nguyen, M. Raupach, L. J. Janik, Clays Clay Miner.1987, 35, 60.10.1346/CCMN.1987.0350108Search in Google Scholar
[40] S. K. Srivastav, N. S. Gajbhiye, J. Am. Ceram. Soc.2012, 95, 3678.10.1111/j.1551-2916.2012.05411.xSearch in Google Scholar
[41] M. M. Murshed, C. B. Mendive, M. Curti, M. Šehović, A. Friedrich, M. Fischer, T. M. Gesing, J. Solid State Chem.2015, 229, 87.10.1016/j.jssc.2015.05.010Search in Google Scholar
[42] J. M. Moreau, C. Michel, R. Gerson, W. J. James, J. Phys. Chem. Solids1971, 32, 1315.10.1016/S0022-3697(71)80189-0Search in Google Scholar
[43] R. D. Shannon, Acta Crystallogr.1976, A32, 751.10.1107/S0567739476001551Search in Google Scholar
[44] Th. M. Gesing, M. Schowalter, C. Weidenthaler, M. M. Murshed, G. Nénert, C. B. Mendive, M. Curti, A. Rosenauer, J.-Ch. Buhl, H. Schneider, R. X. Fischer, J. Mater. Chem. 2012, 22, 18814.10.1039/c2jm33208fSearch in Google Scholar
©2016 by De Gruyter
Articles in the same Issue
- Frontmatter
- In this Issue
- Preface
- Congratulations to Wolfgang Jeitschko
- Review
- Prediction and clarification of structures of (bio)molecules on surfaces
- NaGe6As6: Insertion of sodium into the layered semiconductor germanium arsenide GeAs
- Synthesis, structures, and properties of Mn(II) and Cd(II) thiocyanato coordination compounds with 2,5-dimethylpyrazine as co-ligand
- Synthesis, structure and electronic configuration of [Rh6Te8(PPh3)6]·4C6H6
- Room temperature synthesis, crystal structure and selected properties of the new compound [Mn2(bipy)4SbS4](ClO4)
- The isotypic family of the diarsenates MM′As2O7 (M = Sr, Ba; M′ = Cd, Hg)
- RE3Au5Zn (RE = Y, Sm, Gd–Ho) – A new structure type with five- and six-membered rings as building units in a gold network
- Influence of hydrogenation and mechanical grinding on the structural and ferromagnetic properties of GdFeSi
- Two new ternary chalcogenides Ba2ZnQ3 (Q = Se, Te) with chains of ZnQ4 tetrahedra: syntheses, crystal structure, and optical and electronic properties
- Neutron powder diffraction and theory-aided structure refinement of rubidium and cesium ureate
- High-pressure investigations of lanthanoid oxoarsenates: I. Single crystals of scheelite-type Ln[AsO4] phases with Ln = La–Nd from monazite-type precursors
- Bi2Fe4O9: Structural changes from nano- to micro-crystalline state
- New transition metal oxide fluorides with ReO3-type structure
- Synthesis, photophysical characterization and DFT studies on fluorine-free deep-blue emitting Pt(II) complexes
- Structural studies of CaAl12O19, SrAl12O19, La2/3+δ Al12–δO19, and CaAl10NiTiO19 with the hibonite structure; indications of an unusual type of ferroelectricity
- Synthesis and crystal structure of new K and Rb selenido/tellurido ferrate cluster compounds
- Crystal structures and hydrogenation properties of palladium-rich compounds with elements from groups 12–16
- Hydroalumination and hydrogallation of an aryl-chloro-dialkynylsilane: Si–Cl bond activation by intramolecular Al–Cl and Ga–Cl interactions
- V18P9C2: a complex phosphide carbide
- Gold(III)-mediated cyclization of 2-hydrazinylquinolines
- RE2B8O15 (RE = La, Pr, Nd) – syntheses of three new rare earth borates isotypic to Ce2B8O15
- Substitution of W5+ in monophosphate tungsten bronzes by combinations Mn+/W6+
- On new ternary equiatomic scandium transition metal aluminum compounds ScTAl with T = Cr, Ru, Ag, Re, Pt, and Au
- Triangular Zn3 and Ga3 units in Sr2Au6Zn3, Eu2Au6Zn3, Sr2Au6Ga3, and Eu2Au6Ga3 – structure, magnetism, 151Eu Mössbauer and 69;71Ga solid state NMR spectroscopy
- The crystal structure of Cs2S2O3·H2O
- Trigermanides AEGe3 (AE = Ca, Sr, Ba): chemical bonding and superconductivity
- The coloring problem in the solid-state metal boride carbide ScB2C2: a theoretical analysis
- Synthesis and structure determinantion of the first lead arsenide phosphide Pb2AsxP14–x (x ~ 3.7)
- The new barium compound Ba4Al7+x: formation and crystal structure
Articles in the same Issue
- Frontmatter
- In this Issue
- Preface
- Congratulations to Wolfgang Jeitschko
- Review
- Prediction and clarification of structures of (bio)molecules on surfaces
- NaGe6As6: Insertion of sodium into the layered semiconductor germanium arsenide GeAs
- Synthesis, structures, and properties of Mn(II) and Cd(II) thiocyanato coordination compounds with 2,5-dimethylpyrazine as co-ligand
- Synthesis, structure and electronic configuration of [Rh6Te8(PPh3)6]·4C6H6
- Room temperature synthesis, crystal structure and selected properties of the new compound [Mn2(bipy)4SbS4](ClO4)
- The isotypic family of the diarsenates MM′As2O7 (M = Sr, Ba; M′ = Cd, Hg)
- RE3Au5Zn (RE = Y, Sm, Gd–Ho) – A new structure type with five- and six-membered rings as building units in a gold network
- Influence of hydrogenation and mechanical grinding on the structural and ferromagnetic properties of GdFeSi
- Two new ternary chalcogenides Ba2ZnQ3 (Q = Se, Te) with chains of ZnQ4 tetrahedra: syntheses, crystal structure, and optical and electronic properties
- Neutron powder diffraction and theory-aided structure refinement of rubidium and cesium ureate
- High-pressure investigations of lanthanoid oxoarsenates: I. Single crystals of scheelite-type Ln[AsO4] phases with Ln = La–Nd from monazite-type precursors
- Bi2Fe4O9: Structural changes from nano- to micro-crystalline state
- New transition metal oxide fluorides with ReO3-type structure
- Synthesis, photophysical characterization and DFT studies on fluorine-free deep-blue emitting Pt(II) complexes
- Structural studies of CaAl12O19, SrAl12O19, La2/3+δ Al12–δO19, and CaAl10NiTiO19 with the hibonite structure; indications of an unusual type of ferroelectricity
- Synthesis and crystal structure of new K and Rb selenido/tellurido ferrate cluster compounds
- Crystal structures and hydrogenation properties of palladium-rich compounds with elements from groups 12–16
- Hydroalumination and hydrogallation of an aryl-chloro-dialkynylsilane: Si–Cl bond activation by intramolecular Al–Cl and Ga–Cl interactions
- V18P9C2: a complex phosphide carbide
- Gold(III)-mediated cyclization of 2-hydrazinylquinolines
- RE2B8O15 (RE = La, Pr, Nd) – syntheses of three new rare earth borates isotypic to Ce2B8O15
- Substitution of W5+ in monophosphate tungsten bronzes by combinations Mn+/W6+
- On new ternary equiatomic scandium transition metal aluminum compounds ScTAl with T = Cr, Ru, Ag, Re, Pt, and Au
- Triangular Zn3 and Ga3 units in Sr2Au6Zn3, Eu2Au6Zn3, Sr2Au6Ga3, and Eu2Au6Ga3 – structure, magnetism, 151Eu Mössbauer and 69;71Ga solid state NMR spectroscopy
- The crystal structure of Cs2S2O3·H2O
- Trigermanides AEGe3 (AE = Ca, Sr, Ba): chemical bonding and superconductivity
- The coloring problem in the solid-state metal boride carbide ScB2C2: a theoretical analysis
- Synthesis and structure determinantion of the first lead arsenide phosphide Pb2AsxP14–x (x ~ 3.7)
- The new barium compound Ba4Al7+x: formation and crystal structure