Abstract
20BaO·5ZnO·5Fe2O3·70V2O5 glass annealed at 450°C for 30 min showed a marked decrease in the electric resistivity (ρ) from 4.0×105 to 4.8 Ωcm, while 20BaO·5Cu2O·5Fe2O3·70V2O5 glass from 2.0×105 to 5.0 Ωcm. As for the conduction mechanism, it proved that n-type semiconductor model in conjugation with the small polaron hopping theory was most probable. Since ZnII and CuI have a 3d10-electron configuration in the outer-most orbital, Ga2O3- and GeO2-containing vanadate glasses were explored in this study. 20BaO·5Ga2O3·5Fe2O3·70V2O5 glass showed a less remarkable decrease of ρ from 4.5×105 to 100 Ωcm, and 20BaO·5GeO2·5Fe2O3·70V2O5 glass from 3.3×106 to 400 Ωcm. Activation energies for the conduction (Ea) of GeO2- and Ga2O3-contaning glasses before the annealing were respectively estimated to be 0.42 and 0.41 eV. It proved that barium iron vanadate glass with a smaller Ea value could attain the higher conductivity after the annealing at temperaures higher than the crystalization temperature.
Introduction
Highly conductive vanadate glasses have a lot of industrial applications such as cathode active material for LIB, solid state electrolyte, electric discharge needle, static electricity protecting material, conductive glass paste, hyperfine processing material with laser or FIB, etc. Some conductive vanadate glasses were reviewed in our previous paper [1], together with the results for semiconducting iron silicate glass prepared by recycling a mixture of fly ash (coal ash) and Fe2O3. Small polaron hopping theory has so far been applied to explain the conduction mechanism of semiconducting vanadate glass with a resistivity (ρ) of mega order (MΩcm) [2]. Highly conductive vanadate glass will be utilized as advanced materials because its conductivity (σ) is easily “tunable” in association with the heat treatment at a given temperature higher than glass transition temperature (Tg) or crystallization peak temperature (Tc) determined by differential thermal analysis (DTA) [1], [3], [4], [5], [6], [7], [8].
Isothermal annealing of potassium iron vanadate glass, 25K2O·10Fe2O3·65V2O5 (composition in mol%), at 380°C for only 10 min caused a substantial decrease in ρ from 1.6×107 Ωcm (σ=6.3×10−8Scm−1) to 2.3 kΩcm (4.3×10−4Scm−1) [3], [4]. The annealing was conducted at 340–380°C which was higher than Tg of 217°C, Tc(1) of 284°C and Tc(2) of 344°C. After the heat treatment at 340°C for 10 min, RT-Mössbauer spectra showed a marked decrease in quadrupole splitting (Δ) of FeIII from 0.61 to 0.55 mms−1, which reflected an increased symmetry or a decreased distortion of “distorted” FeO4 and VO4 tetrahedra that were involved with an increased structural relaxation of pseudo-1D network [3], [4]. Nishida et al. concluded that the lowering of ρ in potassium iron vanadate glass, amounting to several orders of magnitude, was ascribed to an increased probability of the small polaron hopping from VIVto VV. In case of potassium iron vanadate glass, 25K2O·10Fe2O3·65V2O5, an isothermal annealing at 380°C for more than 150 min resulted in an increase in ρ due to the precipitation of “insulating” KV3O8 particles. Finally, the ρ became comparable to that of the as-cast potassium iron vanadate glass [3], [4], indicating that the structural relaxation of the network was involved with the high conductivity, as confirmed by a decrease in Δ of FeIII in the Mössbauer spectra.
In case of 15BaO·15Fe2O3·70V2O5 glass (in mol%), further decrease of ρ was observed after heat treatment at temperatures close to Tc(1) of 372°C and Tc(2) of 468°C [5]. In this glass, ρ decreased from 1.0×107 Ωcm (before annealing) to 250 Ωcm (4.0×10−3 Scm−1) and 25 Ωcm (4.0×10−2Scm−1) after the annealing at 370°C for 120 min and at 460°C for 30 min, respectively. It is noted that these ρ values are one- or two-orders of magnitude smaller than that of crystalline V2O5 [9]. It was reported that 15BaO·15Fe2O3·70V2O5 glass has a 3D-network structure in which several “pathways” or “route” were available for the small polaron hopping than in pseudo 1D-network of 25K2O·10Fe2O3·65V2O5 glass [3], [4]. After the heat treatment of 15BaO·15Fe2O3·70V2O5 glass at 460°C for 60 min, Δ of FeIII decreased from 0.67 to 0.60 mms−1, reflecting decreased distortion of “distorted” FeO4 and VO4 tetrahedra or an increased structural relaxation of the 3D-network [5]. Representative Δ values for highly conductive BaO–Fe2O3–V2O5 glasses are summarized in Table 1.
Resistivity (ρ), conductivity (σ), activation energy (Ea) and quadrupole splitting (Δ) of FeIIIobtained for conductive barium iron vanadate glasses containing different metal oxides.
| Composition (in mol%) | Annealing | ρ (Ωcm) | σ (Scm−1) | E a (eV) | Δ (mms−1) |
|---|---|---|---|---|---|
| 15BaO·15Fe2O3·70V2O5 [5] | Unannealed | 1.0×107 | 1.0×10−7 | – | 0.67 |
| 460°C, 30 min | 25 | 4.0×10−2 | – | 0.60 (1 h) | |
| 20BaO·10Fe2O3·70V2O5 [6], [7] | Unannealed | 3.6×105 | 2.8×10−6 | 0.38 | 0.68 |
| 500°C, 60 min | 23 | 4.3×10−2 | 0.13 | 0.50 | |
| 20BaO·10Fe2O3·70V2O5 [8] | Unannealed | 6.3×104 | 1.6×10−5 | – | 0.77 |
| 500°C, 1000 min | 0.91 | 1.1 | – | 0.53 | |
| 20BaO·10Fe2O3·20WO3·50V2O5 [10] | Unannealed | 2.6×105 | 3.9×10−6 | 0.44a | 0.82 |
| 500°C, 240 min | 480 | 2.1×10−3 | 0.24a | 0.76 | |
| 20BaO·10Fe2O3·10WO3·60V2O5 [11] | Unannealed | 5.9×104 | 1.7×10−5 | – | 0.73 |
| 500°C, 1000 min | 10 | 1.0×10−1 | – | 0.59 | |
| 20BaO·10Fe2O3·10MnO2·60V2O5 [12] | Unannealed | 2.2×106 | 4.5×10−7 | 0.33 | 0.76 |
| 500°C, 1000 min | 71 | 1.4×10−2 | 0.11 | 0.49 | |
| 20BaO·5CuO·5Fe2O3·70V2O5 [13] | Unannealed | 2.6×105 | 3.9×10−6 | 0.16 | 0.66 |
| 450°C, 30 min | 3.1 | 3.2×10−1 | 0.10 | 0.54 | |
| 20BaO·5Cu2O·5Fe2O3·70V2O5 [13] | Unannealed | 2.0×105 | 5.1×10−6 | 0.22 | 0.69 |
| 450°C, 30 min | 5.0 | 2.0×10−1 | 0.09 | 0.54 | |
| 20BaO·5ZnO·5Fe2O3·70V2O5 [1] | Unannealed | 4.0×105 | 2.5×10−6 | 0.23b | 0.68 |
| 450°C, 30 min | 4.8 | 2.1×10−1 | 0.14b | 0.61 | |
| 20BaO·5Ga2O3·5Fe2O3·70V2O5 | Unannealed | 4.5×105 | 2.2×10−6 | 0.42c | 0.69 |
| 450°C, 120 min | 104 | 9.6×10−3 | 0.18c | 0.54 | |
| 20BaO·5GeO2·5Fe2O3·70V2O5 | Unannealed | 3.3×106 | 3.0×10−7 | 0.41d | 0.77 |
| 450°C, 120 min | 400 | 2.5×10−3 | 0.21d | 0.60 |
aI. Furumoto, Master thesis Kinki (Kindai) University (2013). bPresent study. cEvaluated on 20BaO·10Ga2O3·70V2O5 glass. dEvaluated on 20BaO·10GeO2·70V2O5 glass.
Heat treatment of 20BaO·10Fe2O3·70V2O5 glass for 60 min at 500°C, which was higher than Tg of 312°C, Tc(1) of 376°C and Tc(2) of 468°C, resulted in a marked decrease in ρ from 3.6×105 Ωcm (2.8×10−6Scm−1) to 23 Ωcm (4.3×10−2 Scm−1) [6], [7]. Mössbauer spectra showed a remarkable decrease in Δ of FeIII from 0.68 to 0.50 (±0.02) mms−1 after the heat treatment. Heat treatment of this glass at 500°C for 1000 min caused a marked decrease in ρ from 6.3×104 Ωcm (1.6×10−5Scm−1) to 0.91 Ωcm (1.1 Scm−1) [8]. X-ray diffraction (XRD) study of the glass ceramic prepared by annealing at 450°C for 2000 min showed the presence of small amounts of “insulating or semiconducting” BaFe2O4 and BaV2O6 particles. Crystal growth became maximal when annealed at temperatures around Tc(1) and Tc(2) [6], [7], at which the formation of BaFe2O4 and BaV2O6 particles were respectively detected, since the FeIII–O bond energy was smaller than VV–O bond energy of 3.9–4.9 eV [14], [15], and hence the precipitation of BaFe2O4 preceded that of BaV2O6. After the annealing at 500°C for 1000 min, precipitation of “semiconducting” FeVO4 particles which intrinsically had ρ of 1.5×106 Ωcm (σ=6.7×10−7 Scm−1) was confirmed in “non-substituted” 20BaO·10Fe2O3·70V2O5 glass and in xR2O·10Fe2O3·(90−x)V2O5 glasses (R=Li, Na, K; x=20, 40) [16]. These results proved that the remarkable decrease of ρ in several vanadate glasses caused by the annealing was not ascribed to the crystalline particles precipitated in the glass matrix, but to the structural relaxation of the 3D-network [1], [3], [4], [5], [6], [7], [8], [10], [11], [12], [13], [16], [17].
Change of the resistivity caused by a heat treatment of barium iron vanadate glasses, in which Fe2O3 was partially replaced by different metal oxides, are summarized in Table 1. Heat treatment of 20BaO·10Fe2O3·20WO3·50V2O5 glass at 500°C for 240 min caused a decrease of ρ from 2.6×105 Ωcm (3.9×10−6 Scm−1) to 480 Ωcm (2.1×10−3 Scm−1) in conjunction with a decrease of Δ from 0.82 to 0.76 mms−1 [10]. In case of 20BaO·10Fe2O3·10WO3·60V2O5 glass [11], an isothermal annealing at 500°C for 1000 min caused a decrease of ρ from 5.9×104 Ωcm (1.7×10−5Scm−1) to 10 Ωcm (1.0×10−1 Scm−1), simultaneously with a decrease in Δ of FeIII from 0.73 to 0.59 mms−1. XRD study of 20BaO·10Fe2O3·xWO3·(70−x)V2O5 glasses revealed a precipitation of crystalline particles such as FeVO4, BaFe2O4, BaFe12O19 and α-Fe2O3 [11]. It was concluded that these crystalline particles were not involved with the marked decrease of ρ, since the intrinsic resistivity of these compounds was much higher than that of WO3-substituted vanadate glasses evaluated at RT after the annealing at 500°C. It proved that cleavage of FeIII–O bonds triggered the crystallization of WO3-substituted glasses [11], as observed in 25K2O·10Fe2O3·65V2O5 glass [3]. Heat treatment of MnO2-substituted vanadate glass, 20BaO·10Fe2O3·10MnO2·60V2O5, at 500°C for 1000 min resulted in a decrease in ρ from 2.2×106 Ωcm (4.5×10−7Scm−1) to 71 Ωcm (1.4×10−2 Scm−1) in conjunction with a decrease in Δ of FeIII from 0.76 to 0.49 mms−1 which evidently showed the structural relaxation [12] (see Table 1).
20BaO·5CuO·5Fe2O3·70V2O5 and 20BaO·5Cu2O·5Fe2O3·70V2O5 glasses showed a higher conductivity after the heat treatment at 450°C for 30 min [13]. Resistivity (ρ) of the former decreased from 2.6×105 Ωcm (3.9×10−6 Scm−1) to only 3.1 Ωcm (3.2×10−1 Scm−1) after the annealing, and the latter glass showed a comparable decrease of ρ from 2.0×105 Ωcm (5.1×10−6 Scm−1) to 5.0 Ωcm (2.0×10−1Scm−1) [13]. In 5 mol% CuO-substituted glass, Δ of FeIII decreased from 0.66 to 0.54 mms−1. In case of 5 mol% Cu2O-substituted glass, Δ decreased from 0.69 to 0.54 mms−1 (see Table 1). It is noted that CuII and CuI in CuO-and Cu2O-substituted glasses respectively have the electron configuration of 3d9 and 3d10 in the outer-most orbital.
Heat treatment of 20BaO·5ZnO·5Fe2O3·70V2O5 glass at 450°C for 30 min, in which ZnII has an electron configuration of 3d10 in the outer-most orbital, showed a marked decrease in ρ from 4.0×105 Ωcm (2.5×10−6 Scm−1) to 4.8 Ωcm (2.1×10−1 Scm−1), simultaneously with a decrease in Δ from 0.68 to 0.61 mms−1 [1]. It is noted that elevation of the conductivity was most remarkable in case of Cu2O- [13], CuO- [13] and ZnO-substituted glasses [1], and that an electron configuration of 3d10 or 3d9 was preferable to attain the high conductivity. The present study was carried out in order to verify this idea in Ga2O3- and GeO2-substituted vanadate glasses, since GaIII and GeIV of “p-block” elements also have the electron configuration of 3d10 in the outer-most orbital.
Mössbauer spectroscopy has successfully been utilized as a powerful tool for the local structural study of oxide glasses. Mössbauer nuclides like 57Fe and 119Sn could play an effective role as a probe for the local structural study [1], [3], [4], [5], [6], [7], [8], [10], [11], [12], [13], [16], [17], [18], [19], [20], [21]. A “Tg-vs-Δ plot”, applied to several inorganic glasses like silicates, borate, borosilicate, aluminates, tellurite and gallate glasses, is very effective to determine the structural role of FeIII in these glasses, i.e. “substitutional” sites as the network former (NWF) or “interstitial” sites as the network modifier (NWM) [4], [8], [11], [12], [17], [18], [21]. When Tg’s of several oxide glasses, ranging from 180 to 770°C, were plotted against the Δ of FeIII (0.4–1.3 mms−1), one “common” straight line with an identical slope of 680 K(mms−1)−1 was drawn when FeIII atoms occupied substitutional “tetrahedral” NWF sites [4], [8], [11], [12], [17], [18]. For example, “Tg-vs-Δ plot” applied to xBaO·10Fe2O3·(90−x)V2O5 glasses (x=20–40) and 20BaO·10Fe2O3·xWO3·(70-x)V2O5 glasses (x=10–50) yielded comparable slopes of 650 [8] and 680 K(mms−1)−1 [11], respectively. These results proved that FeIIIatoms occupied substitutional “tetrahedral” sites of VIV or VV as NWF. Essentially the same results were obtained in 20BaO·10Fe2O3·xMnO2·(70−x)V2O5 glasses (x=0–30) which yielded a slope of 707 K(mms−1)−1 [12], and in 20R2O·10Fe2O3·xWO3·(70−x)V2O5 glasses (R=Na, K; x=0–50) with a slope of 670 or 680 K(mms−1)−1[17]. In contrast, FeIII atoms in xNa2O·Fe2O3·(99−x)WO3 glasses (30≤x≤42) occupied substitutional “octahedral” sites of WVI as NWF, in which a small slope of 260 K(mms−1)−1 was obtained because of its longer bond length with a weaker bond energy [21].
Distortion or local symmetry at the Mössbauer nuclear sites could be deduced from the Δ of FeIII. Mössbauer spectra of heat-treated vanadate glasses containing FeIII showed a marked decrease in Δ, reflecting a decreased distortion or an increased symmetry of “distorted” FeIIIO4 tetrahedra [1], [3], [4], [5], [6], [7], [8], [10], [11], [12], [13], [16], [17], [18], [19], [20]. This is also the case for “distorted” VO4 tetrahedra in the 3D-network, since one FeO4 unit is randomly surrounded by three or four VO4 units by sharing corner oxygen atoms. It is noted that the electric field gradient (EFG) caused by the valence electrons (eqval) of high-spin FeIII is “zero” since it has an isotropic electron configuration of 3d5 in the outer-most orbital. In this case, Δ of FeIII reflects the EFG brought about by the lattice (eqlat), i.e. by the steric configuration of oxygen atoms constituting the distorted FeO4 and VO4 tetrahedra.
Mössbauer study of heat-treated vanadate glasses indicated that a remarkable decrease of ρ was observed in conjunction with a decrease of Δ, which reflected a decreased distortion of FeIIIO4 tetrahedra and VO4 tetrahedra [1], [3], [4], [5], [6], [7], [8], [12], [13], [14], [15], [16], [17]. It was also considered that FeIII atoms having an isotropic symmetric electron configuration of 3d5 in the outermost orbital was preferable for the remarkable decrease of ρ since half-occupied 3d-orbitals could effectively accept the carriers (polaron) in the remaining unoccupied 3d-orbitals [5], [6], [7], [8].
Experimental
Homogeneous glass samples with compositions of 20BaO·xGa2O3·(10-x)Fe2O3·70V2O5 and 20BaO·xGeO2·(10-x)Fe2O3·70V2O5 (in mol%) were prepared by a melt-quenching method with weighed amounts of BaCO3, Ga2O3, GeO2, Fe2O3 and V2O5 of guaranteed reagent grade. Each reagent mixture placed in an alumina crucible was melted in an electric muffle furnace at 1300°C for 2.5 h in Ga2O3-substituted glass, and at 1100°C for 2 h in GeO2-substituted glass. Glass samples of almost black color were prepared by quenching the melt in the air. Annealing of as-quenched glass sample was carried out at a given temperature using another electric furnace. For determining Tg and Tc values, DTA was conducted at a heating rate of 10 Kmin−1 ranging from RT to 600°C in an N2 atmosphere. A fixed amount of α-Al2O3 powder was used as a reference of the temperature. Electrical resistivity (ρ) of a rectangular glass sample was determined by a conventional dc-four probe method, in which a linear relationship was obtained by plotting the voltage against the electric current applied by the electrometer. Mössbauer measurement was conducted at RT by a constant acceleration method with a source of 57Co(Rh). Small amount of enriched isotope, 57Fe2O3, was used for the sample preparation of Mössbauer measurement. A foil of α-Fe was used as a reference of δ and for calibrating the velocity scale of the spectrometer. A software of Mösswinn 3.0i XP was used for the peak analysis of the Mössbauer spectrum.
Results
20BaO·xGa2O3·(10−x)Fe2O3·70V2O5 glass
DTA of Ga2O3-substituted vanadate glass showed a gradual lowering of Tg from 351°C (x=0) to 328 (x=5) and 324°C (x=10), reflecting lowered heat resistivity. Nishida et al. reported that gallate (Ga2O3-based) glass had higher heat resistivity [4], [18], [19], [20]. When the coordination number (CN) of GaIII was four, GaIII–O bond was reported to have a single bond energy of 2.9 eV [14], [15], which was larger than the FeIII–O bond energy of 2.6 eV [21], [22]. A gradual lowering of Tg observed in the present study will be ascribed to the structural role of GaIII that plays a role of NWM like Ba2+ at the “interstitial” sites of 3D-network of vanadate glass. In such case, six-fold coordinated GaIII atoms have smaller single bond energy of 2.0 eV [14], [15]. Substitution of Ga2O3 for Fe2O3 will cause a lowering of Tg, since the increased fraction of NWM from 20 mol% BaO to 25 mol% (BaO+Ga2O3) is equivalent to a decreased degree of polymerization.
DTA study of Ga2O3-substituted vanadate glass showed Tc values of 410, 422 and 405°C when the Ga2O3 contents were 0, 5 and 10 mol%, respectively. Crystallization of oxide glass is closely related to the bond energy of NWF-oxygen polyhedra [3], [21], [22], [23], [24]. For example, crystallization of FeIII-containing aluminate glass [22] was triggered by a cleavage of FeIII–O bonds, since the Al–O bond energy (3.4–4.4 eV [14], [15]) was much larger than the Fe–O bond energy of 2.6 eV [21], [22]. The same conclusion was drawn in vanadate [3] and tungstate glasses [21]. In the crystallization study of IR-transmitting gallate glass in combination with the heat treatment, Ar+-laser and 60Co-γ ray irradiations, FT-IR spectroscopy was effectively used to explore the kinetics and mechanism of crystallization [19], [23], [24]. Irradiation of Ga2O3-substituted vanadate glass with Ar+-laser and 60Co-γ rays will also show the “memory effect” as observed in the IR-transmitting gallate glass described above.
DTA study of 20BaO·5CuO·5Fe2O3·70V2O5 and 20BaO·5Cu2O·5Fe2O3·70V2O5 glasses showed a small decrease in Tg along with an increasing CuO or Cu2O content, together with a lowering of Tc amounting to ca. 30 K [13]. Similar lowering of Tg and Tc amounting to ca. 30 K was observed in 20BaO·5ZnO·5Fe2O3·70V2O5 glass [1]. In case of 20BaO·10ZnO·70V2O5 glass, a decrease in Tg amounted to ca. 50 K and a decrease in Tc to ca. 30 K [1]. These DTA results proved that substitution of CuO, Cu2O and ZnO for Fe2O3 was effective for the preparation of less heat-resistant vanadate glass with higher conductivity. Such conductive glass might be preferably utilized as sensors, electric discharge needle, conductive glass paste, static electricity-protecting material and hyperfine processing materials combined with FIB, electrons, laser, etc. This will be also the case for Ga2O3- and GeO2-substituted glasses, if the conductivity were highly enough for each purpose.
Figure 1a depicts σ values of Ga2O3-substituted glasses measured at RT after the isothermal annealing at 450°C, which was higher than the Tc’s of 405–422°C. RT-conductivity (σ) of 5 mol% Ga2O3-substituted vanadate glass measured after the heat treatment for 120 min (solid square) increased from 2.2×10−6Scm−1 (ρ=4.5×105 Ωcm) to 9.6×10−3 Scm−1 (104 Ωcm), which was one-order of magnitude lower than that of “non-substituted” 20BaO·10Fe2O3·70V2O5 glass measured in this study (3.5×10−2 Scm−1 or 28 Ωcm). If the heat treatment were made at temperatures higher than 450°C, the conductivity will be enhanced. After the heat treatment at 450°C, σ of 10 mol% Ga2O3-substituted vanadate glass (red solid circle) increased from 6.7× 10−7 Scm−1(1.5×106 Ωcm) to 6.2×10−2 Scm−1(16 Ωcm), which was comparable to that of non-substituted vanadate glass, i.e. 3.5×10−2 Scm−1(28 Ωcm).

(a) Electrical conductivity (σ) of 20BaO·10Fe2O3·70V2O5 (open circle), 20BaO·5Ga2O3·5Fe2O3·70V2O5 (solid square) and 20BaO·10Ga2O3·70V2O5 glasses (red solid circle) measured at RT after isothermal annealing at 450°C. (b) Electrical conductivity (σ) of 20BaO·10Fe2O3·70V2O5 (open circle), 20BaO·5GeO2·5Fe2O3·70V2O5 (solid square) and 20BaO·10GeO2·70V2O5 glasses (red solid circle) measured at RT after isothermal annealing at 450°C.
All the σ values plotted in Fig. 1a are one- or two-orders of magnitude lower than those of 5 mol% CuO-substituted vanadate glass [13], which showed a marked increase of σ from 3.9×10−6 Scm−1 (2.6×105 Ωcm) to 3.2×10−1 Scm−1(3.1 Ωcm) after the annealing at 450°C for 30 min (see Table 1). After the same annealing, 5 mol% ZnO-substituted vanadate glass also showed a remarkable increase in σ from 2.5×10−6 Scm−1 (4.0×105 Ωcm) to 2.1×10−1 Scm−1(4.8 Ωcm) [1]. Conductivities of related barium iron vanadate glasses are summarized in Table 1, together with the activation energy for the electrical conduction (Ea) and Δ values of FeIII obtained from the Mössbauer measurement. As seen from Table 1, CuO-, Cu2O- and ZnO-substituted vanadate glasses, in which CuII atom had an electron configuration of 3d9, and CuI and ZnII atoms 3d10 configuration, were preferable for the preparation of highly conductive vanadate glass.
RT-Mössbauer spectra of 5 mol% Ga2O3-substituted vanadate glass, measured after the isothermal annealing at 450°C, are illustrated in Fig. 2a. A marked decrease in Δ from 0.69 to 0.55 (±0.01) mms−1 was observed after the annealing at 450°C for only 30 min, and Δ showed a slight decrease from 0.55 to 0.52 (±0.01) mms−1after the additional heat treatment up to 300 min. A larger decrease in Δ of 0.12–0.15 (±0.02) mms−1was observed in CuO- and Cu2O-substituted vanadate glasses after the annealing at 450°C for 30 min [13]. In case of ZnO-substituted glass, the decrease in Δ was 0.06–0.07 (±0.01) mms−1 [1]. All the Mössbauer results are summarized in Table 1.

(a) RT-Mössbauer spectra of 20BaO·5Ga2O3·5Fe2O3·70V2O5 glass measured after isothermal annealing at 450°C. (b) RT-Mössbauer spectra of 20BaO·5GeO2·5Fe2O3·70V2O5 glass measured after isothermal annealing at 450°C.
Figure 2a proved that the eqlat at the nuclear sites of 57Fe decreased as a result of heat treatment at 450°C, reflecting a decreased distortion of FeO4 tetrahedra. This is also the case for VO4 tetrahedra since they are directly bonded to FeO4 tetrahedra. In several vanadate glasses, decrease in the resistivity or an increase in the conductivity has been observed in conjunction with a decrease in Δ of FeIII [1], [3], [4], [5], [6], [7], [8], [10], [11], [12], [13], [16], [17]. As shown in Fig. 1a with solid squares, σ values of 5 mol% Ga2O3-substituted glass were smaller than those of non-substituted vanadate glass (open circle) despite that the Mössbauer spectra showed a large decrease from 0.69 to 0.55–0.56 mms−1 after 30–60 min annealing (Fig. 2a). In case of CuO- and Cu2O-substituted vanadate glasses, comparable decrease in Δ [0.12–0.15 (±0.02) mms−1] was observed after the annealing at 450°C for 30 min, and σ showed much more remarkable increase from 3.9×10−6 to 3.2×10−1Scm−1(3.1 Ωcm) [13]. In case of ZnO-substituted glass, a remarkable increase of σ was observed from 2.5×10−6 to 2.1×10−1Scm−1(4.8 Ωcm) in spite that the decrease in Δ was not so large [0.06–0.07 (±0.01) mms−1] [1] (see Table 1). These results prove that elevation of the conductivity is partially related to the decrease in the distortion of FeO4 and VO4 units. The conduction behavior of highly conductive barium iron vanadate glass should be discussed by combining the small polaron hopping theory and the n-type semiconductor model, as discussed below.
20BaO·xGeO2·(10−x)Fe2O3·70V2O5 glasses
DTA study of GeO2-substituted vanadate glass revealed a gradual lowering of Tg from 354°C (x=0) to 330°C (x=5) and 318°C (x=10), as observed in Ga2O3-substituted vanadate glasses. Substitution of 5 mol% GeO2 for Fe2O3 resulted in an increase of Tc from 408 to 424°C. It means that larger thermal energy was required for the rearrangement of each fragment when an equivalent amount of GeO2 and Fe2O3 played a role of NWF in the 3D-network. In case of 10 mol% GeO2-substituted vanadate glass showed a lowering of Tc from 408°C (non-substituted) to 389°C, as also observed in 10 mol% Ga2O3-substituted glass.
RT-conductivities (σ) of 5 mol% GeO2-substituted vanadate glass, measured after an isothermal annealing at 450°C, are plotted with solid squares in Fig. 1b. After the heat treatment at 450°C for 120 min, this glass showed an increase of σ from 3.0×10−7Scm−1 (ρ= 3.3×106 Ωcm) to 2.5×10−3 Scm−1 (400 Ωcm), which was one-order of magnitude smaller than that of non-substituted vanadate glass as plotted with open circles in Fig. 1b, i.e. 3.5×10−2 Scm−1(28 Ωcm).
After the heat treatment at 450°C, σ of 10 mol% GeO2-substituted vanadate glass (red solid circles) increased from 2.0×10−7 Scm−1(5.0×106 Ωcm) to 1.3×10−2 Scm−1(77 Ωcm) which was still smaller than that of non-substituted vanadate glass plotted with open circles, 3.5×10−2 Scm−1(28 Ωcm). Although the σ values of 10 mol% GeO2-substituted vanadate glass plotted with red solid circles (Fig. 1b) were superior to those of 5 mol% GeO2-substituted glass, they were still smaller than those of non-substituted vanadate glass (open circles). It is noted that σ’s of GeO2-substituted glasses were generally one or two-orders of magnitude smaller than those of 5 mol% CuO-substituted glass [13] that showed a marked increase in σ from 3.9×10−6 Scm−1 (2.6×105 Ωcm) to 3.2×10−1 Scm−1(3.1 Ωcm) after the annealing at 450°C for 30 min. Five mol% ZnO-substituted vanadate glass [1] also showed a remarkable increase in σ from 2.5×10−6 Scm−1(4.0×105 Ωcm) to 2.1×10−1 Scm−1(4.8 Ωcm), as summarized in Table 1.
As described above, CuO-, Cu2O- and ZnO-substituted vanadate glasses showed higher conductivity, in which the electron configurations of 3d9 (CuII) and 3d10 (CuI and ZnII) seemed to be preferable to attain the high conductivity [1], [13]. In case of Ga2O3- and GeO2-substituted vanadate glasses, as described above, increase in the σ caused by the “common” heat treatment at 450°C was less than that of Cu2O- and ZnO-substituted glasses, despite GaIIIand GeIV atoms of “p-block elements” had 3d10 configuration. Conduction mechanism will be discussed in the following session by focusing the electron configuration, band gap of the metal oxides and the activation energy for the conduction (Ea).
RT-Mössbauer spectra of 5 mol% GeO2-substituted vanadate glass measured after the isothermal annealing at 450°C are illustrated in Fig. 2b. In several vanadate glasses, the increase in the conductivity has so far been observed in conjunction with a decrease in Δ of FeIII [1], [3], [4], [5], [6], [7], [8], [10], [11], [12], [13], [16], [17]. The isothermal annealing illustrated in Fig. 2b showed a small increase of Δ from 0.74 to 0.81 (±0.01) mms−1 after the annealing for 30 min. This will reflect a temporarily increased local distortion of FeO4 and VO4 tetrahedra of the 3D-network on the way of the structural relaxation. It is considered that the structural relaxation could occur effectively if the heat treatment were conducted at temperatures higher than 450°C. In Fig. 2b, a remarkable decrease in Δ was observed from 0.81 to 0.54 (±0.01) mms−1 after the isothermal annealing for 60–300 min, which evidently demonstrated a decreased distortion or an enhanced structural relaxation of the FeO4 and VO4 units.
Figure 2b reveals that eqlat at the nuclear sites of 57Fe was decreased due to a decreased distortion of FeO4 and VO4 tetrahedra, since eq=eqlat, as observed in 5 mol% Ga2O3-substituted vanadate glass (Fig. 2a). A decrease in Δ amounted to 0.23 (±0.01) mms−1, which was much larger than that of 5 mol% CuO- or Cu2O-substituted vanadate glasses, 0.12–0.15 (±0.02) mms−1, in which σ increased up to 3.2×10−1 Scm−1 (3.1 Ωcm) after the annealing at 450°C for 30 min [12]. As describe above, 5 mol% ZnO-substituted glass showed a smaller decrease of Δ [0.06–0.07 (±0.01) mms−1], despite a remarkable elevation of σ was observed up to 2.1×10−1 Scm−1 (4.8 Ωcm) [1].
Discussion
Activation energy for the conduction (Ea) in the small polaron hopping theory [2] could be calculated with the following equation:
in which σ, T and k are the conductivity, measuring temperature (K) and the Boltzmann constant, respectively. Values of Ea could be calculated from the slope of the straight line obtained by plotting the natural logarithm of σT against the reciprocal of T, as illustrated in Fig. 3.

(a) Natural logarithm of σT for 20BaO·10GeO2·70V2O5 glass plotted against the reciprocal of the measuring temperature (T) before and after the annealing at 450°C for 120 min. (b) Natural logarithm of σT for 20BaO·5ZnO2·5Fe2O3·70V2O5 glass plotted against the reciprocal of the measuring temperature (T) before and after the annealing at 450°C for 30 min.
Non-substituted barium iron vanadate glass, 20BaO·10Fe2O3·70V2O5, showed a marked increase of σ from 2.8×10−6to 4.3×10−2 Scm−1 in conjunction with a decrease of Ea from 0.38 to 0.13 eV when annealed at 500°C for 60 min [7], [8]. Conductivities of this glass showed a linear increase along with a decrease of Ea [8], [11]. The σ values increased significantly when annealed at higher temperatures, since the glass network was partially cleaved to cause the structural relaxation after prolonged annealing. Several Ea’s obtained for a series of barium iron vanadate glasses are summarized in Table 1, in which latest results of Ga2O3-, GeO2- and ZnO-substituted vanadate glasses are summarized together with those for non-substituted [7], [8], WO3- [10], MnO2- [12], CuO- [13] and Cu2O- substituted glasses [13].
Figure 3a depicts the ln σT-vs.-T−1 plot for 10 mol% GeO2-substituted vanadate glass, and Fig. 3b for 5 mol% ZnO2-substituted glass. In these plots all the σT values were normalized with the σT obtained at RT. Ten mol% GeO2-substituted vanadate glass had Ea’s of 0.41 and 0.21 eV before and after the annealing at 450°C for 120 min, respectively (Fig. 3a). Ten mol% Ga2O3-substituted vanadate glass also had comparable Ea’s of 0.42 and 0.18 eV before and after the annealing, respectively (Table 1). In contrast, 5 mol% ZnO2-substituted vanadate glass investigated for comparison (Fig. 3b) had much smaller Ea’s of 0.23 and 0.14 eV before and after the annealing at 450°C for 30 min, respectively. This glass had RT-conductivity of 2.1×10−1 Scm−1 (4.8 Ωcm) which was one-order of magnitude higher than that of non-substituted glass [1]. Smaller Ea’s were also obtained in CuO- (0.16 and 0.10 eV) [13] and Cu2O-substituted glasses (0.22 and 0.09 eV) [13], of which σ’s after the heat treatment were 3.2×10−1 Scm−1 (3.1 Ωcm) and 2.0×10−1 Scm−1 (5.0 Ωcm), respectively.
These Ea values are significantly smaller than the band gap energy (Eg) between the valence band (VB) and the conduction band (CB), reported for typical semiconductors like GaSb (0.23 eV), Ge (0.68 eV) and Si (1.16 eV) [7]. A photoluminescence study of 20BaO·10Fe2O3·70V2O5 glass annealed at 500°C for 60 min yielded an Eg of 2.25 eV [1]. Carrier (electron) density in the CB would increase along with a decrease of Eg and also with the energy gap between the donor level and the CB of n-type semiconductor. It is considered that Ea is equivalent to the energy gap between the donor level and CB. All the experimental results obtained in different vanadate glasses [1], [6], [7], [8], [10], [11], [12], [13], [17] proved that a decrease in Ea caused an increase in σ due to an increased carrier (electron) density in the CB. It is noted that n-type semiconductor model becomes predominant over the small polaron hopping model that has been utilized for the conduction mechanism of “semiconducting” vanadate glass of which σ is of the order of “mega” Ωcm [1], [13].
In the present study, as-quenched Ga2O3- and GeO2-substituted vanadate glasses respectively had Ea’s of 0.42 and 0.41 eV. After the annealing at 450°C for 120 min, Ea’s of Ga2O3- and GeO2-substituted vanadate glasses were 0.18 and 0.21 eV, respectively. Conductivities of these glasses were 9.6×10−3Scm−1 (104 Ωcm) and 2.5× 10−3 Scm−1 (400 Ωcm), as shown in Fig. 1a and b. They were two-orders of magnitude lower than those of ZnO- [1], CuO- [13] and Cu2O-substituted glasses [13]. All the Ea and σ values obtained for non-substituted, ZnO-, CuO, Cu2O-, Ga2O3- and GeO2-substituted vanadate glasses prove that smaller Ea’s obtained before and after the annealing could attain the higher conductivity that was comparable to that of “metal heater” like Ni–Cr alloy.
As described above, electron configuration of 3d10 proved to be preferable for the preparation of highly conductive vanadate glass like ZnO-, CuO- and Cu2O-substituted glasses [1], [13]. It is interesting to discuss the σ value in association with the band gap energy (Eg) of the substituents. The Eg of α-Fe2O3 is reported to be 2.2 eV [25], whereas CuO and Cu2O, respectively have smaller Eg’s of 1.2–1.5 [26], [27] and 2.1 eV [28]. It was reported that ZnO had comparable Eg of 3.3 [29] or 3.4 eV [30]. In contrast, much larger Eg of 5.3 eV was reported for α-Ga2O3 [31], and 4.7–4.9 eV for β-Ga2O3 [31], [32]. The former is known to be stable at relatively lower temperature like 470°C, whereas the latter is stable at relatively higher temperatures of 550–630°C [31]. In case of α-GeO2, a large Eg of 5.95 eV was reported [33]. Larger Ea values obtained for Ga2O3- and GeO2-substituted vanadate glasses before and after the heat treatment will be ascribed to the large Eg of Ga2O3 and GeO2. These results prove that a partial replacement of Fe2O3 by metal oxide with small Eg is effective to attain the high conductivity of vanadate glass.
It was suggested that FeIII atoms with a symmetric electron configuration of 3d5 were preferable for the small polaron hopping from VIV to VV (and FeIII) via oxygen atoms. For example, 25K2O·10Fe2O3·65V2O5 glass annealed at 380°C showed an increase of σ from 6.3×10−8Scm−1(ρ=1.6×107 Ωcm) to 4.3×10−4Scm−1(ρ=2.3 kΩcm) as a result of an increased probability of the small polaron hopping among less-distorted FeO4 and VO4 tetrahedra [3], [4]. Structural relaxation of FeO4 and VO4 tetrahedra was confirmed as a marked decrease of Δ in the Mössbauer spectra [1], [3], [4], [5], [6], [7], [8], [10], [11], [12], [13], [16], [17], and the conductivity was enhanced by the structural relaxation. This is also the case for the vanadate glasses in which Fe2O3 was replaced by metal oxides in which each metal had 3d10 configuration [1], [13].
Small polaron hopping theory [2] has generally been utilized to explore the conduction mechanism of semiconducting vanadate glass of which σ is <10−4Scm−1. This authors’ group suggested that n-type semiconductor model in conjunction with the small polaron hopping theory was most probable to explain the conduction mechanism of highly conductive vanadate glasses with σ higher than ca. 10−4 Scm−1 [1], [13]. In conductive barium iron vanadate glasses, increase in σ was directly proportional to the decrease in Δ of FeIII in the Mössbauer spectra. A marked decrease of Δ was observed in non-substituted [5], [6], [7], [8], WO3- [10], [11], MnO2- [12], CuO- [13], Cu2O- [13] and Ga2O3- or GeO2-substituted barium iron vanadate glasses (this study). In case of ZnO-substituted glass [1], a decrease in Δ was not remarkable (Table 1), but σ of this glass was one order of magnitude larger than that of non-substituted glass. These results evidently prove that n-type semiconductor model in conjunction with the small polaron hopping theory is most probable for the conduction mechanism of highly conductive vanadate glass of which σ value is comparable to that of metal heater like Ni–Cr alloy.
Summary
As for the conduction mechanism of highly conductive vanadate glasses with σ value larger than 10−4 Scm−1, n-type semiconductor model combined with the small polaron hopping theory is most probable.
Heat treatment of barium iron vanadate glass at a given temperature higher than Tc causes a marked increase in the conductivity in conjunction with a decrease in Ea, which could be correlated to the energy gap between the donor level and the CB.
Substitution of Ga2O3 or GeO2 for Fe2O3 was less effective to increase the conductivity, despite that GaIII and GeIV atoms had 3d10 configuration that was preferable to attain the higher conductivity.
This will be ascribed to the intrinsic large Eg’s of Ga2O3 (5.3 eV) and GeO2 (5.95 eV) which are much larger than that of Fe2O3 (2.2 eV).
All the experimental results suggest that replacement of Fe2O3 by the metal oxide with a small Eg is preferable to attain the higher conductivity, as recently discovered in CuO-, Cu2O- and ZnO-substituted vanadate glasses.
Article note
A collection of invited papers based on presentations at the 12th Conference on Solid State Chemistry (SSC-2016), Prague, Czech Republic, 18–23 September 2016.
Acknowledgments
One of the authors (TN) is grateful to Dr. Ken-ichi Kobayashi, Art Beam Co. Ltd. (Tokyo) for his useful discussion. This paper was partially presented at the occasions of ICAME2015 (Hamburg, Sep. 2015) and MECAME2016 (Cavtat, June 2016).
References
[1] T. Nishida, S. Kubuki, K. Matsuda, Y. Otsuka. Croat. Chem. Acta88, 427 (2015).10.5562/cca2760Search in Google Scholar
[2] N. F. Mott. Adv. Phys. 16, 49 (1967).10.1080/00018736700101265Search in Google Scholar
[3] T. Nishida, J. Kubota, Y. Maeda, F. Ichikawa, T. Aomine. J. Mater. Chem.6, 1889 (1996).10.1039/jm9960601889Search in Google Scholar
[4] T. Nishida. “Chemical Applications”, in Introduction to the Mössbauer Spectroscopy-Principles and Applications, F. E. Fujita (Ed.), pp. 169–266. Agne Gijutsu Center, Tokyo (1999), (in Japanese).Search in Google Scholar
[5] K. Fukuda, A. Ikeda, T. Nishida. Solid State Phenom.90/91, 215 (2003).10.4028/www.scientific.net/SSP.90-91.215Search in Google Scholar
[6] T. Nishida. Japan Patent Nos. 3854985 (2006) and 5164072 (2012).Search in Google Scholar
[7] T. Nishida, S. Kubuki. “Mössbauer Study of New Electrically Conductive Glass”, in Mössbauer Spectroscopy: Applications in Chemistry, Biology, Nanotechnology, Industry and Environment, V. K. Sharma, G. Klingelhöfer, T. Nishida (Eds.), pp. 542–551. John Wiley & Sons, Hoboken, NJ (2013).10.1002/9781118714614.ch27Search in Google Scholar
[8] S. Kubuki, H. Sakka, K. Tsuge, Z. Homonnay, K. Sinkó, E. Kuzmann, H. Yasumitsu, T. Nishida. J. Ceram. Soc. Jpn.115, 776 (2007).10.2109/jcersj2.115.776Search in Google Scholar
[9] A. Byström, K. A. Wilhelmi, O. Brotzen. Acta Chem. Scand. 4, 1119 (1950).10.3891/acta.chem.scand.04-1119Search in Google Scholar
[10] I. Furumoto, S. Kubuki, T. Nishida. Radioisotopes61, 463 (2012).10.3769/radioisotopes.61.463Search in Google Scholar
[11] S. Kubuki, H. Masuda, T. Nishida. J. Radioanal. Nucl. Chem. 295, 1123 (2013).10.1007/s10967-012-1887-7Search in Google Scholar
[12] S. Kubuki, H. Masuda, K. Akiyama, I. Furumoto, T. Nishida. Hyperfine Interact.207, 61 (2012).10.1007/s10751-011-0433-2Search in Google Scholar
[13] K. Matsuda, S. Kubuki, T. Nishida. AIP Conf. Proc.1622 (Proc. Mössbauer Spectroscopy in Materials Science-2014, 26–30 May, Czech Rep.), 3 (2014).10.1063/1.4900744Search in Google Scholar
[14] K. H. Sun. J. Am. Ceram. Soc.30, 277 (1947).10.1111/j.1151-2916.1947.tb19654.xSearch in Google Scholar
[15] D. Chakravorty. “Inorganic Glasses”, in Modern Aspect of Solid State Chemistry, C. N. R. Rao (Ed.), pp. 391–423. Plenum Press, New York (1970).10.1007/978-1-4684-1875-0_16Search in Google Scholar
[16] S. Kubuki, K. Matsuda, K. Akiyama, T. Nishida. J. Radioanal. Nucl. Chem. 299, 453 (2014).10.1007/s10967-013-2748-8Search in Google Scholar
[17] S. Kubuki, K. Matsuda, K. Akiyama, Z. Homonnay, K. Sinkó, E. Kuzmann, T. Nishida. J. Non-Cryst. Solids378, 227 (2013).10.1016/j.jnoncrysol.2013.07.012Search in Google Scholar
[18] T. Nishida, J. Iwashita, S. Kubuki. J. Ceram. Soc. Jpn. 107, 408 (1999).10.2109/jcersj.107.408Search in Google Scholar
[19] T. Nishida, S. Saruwatari, Y. Takashima. Bull. Chem. Soc. Jpn.61, 2347 (1988).10.1246/bcsj.61.2347Search in Google Scholar
[20] T. Nishida, T. Ichii, Y. Takashima. J. Mater. Chem.2, 733 (1992).10.1039/jm9920200733Search in Google Scholar
[21] T. Nishida, M. Suzuki, S. Kubuki, M. Katada, Y. Maeda. J. Non-Cryst. Solids194, 23 (1996).10.1016/0022-3093(95)00501-3Search in Google Scholar
[22] T. Nishida, S. Kubuki, M. Shibata, Y. Maeda, T. Tamaki. J. Mater. Chem. 7(9), 1801 (1997).10.1039/a702854gSearch in Google Scholar
[23] T. Nishida, S. Kubuki, Y. Takashima. J. Non-Cryst. Solids177, 193 (1994).10.1016/0022-3093(94)90530-4Search in Google Scholar
[24] S. Kubuki, T. Nishida, P. Kaung, T. Yagi, Y. Maeda. J. Non-Cryst. Solids209, 87 (1997).10.1016/S0022-3093(96)00543-1Search in Google Scholar
[25] N. J. Cherepy, D. B. Liston, J. A. Lovejoy, H. Deng, J. Z. Zhang. J. Phys. Chem. B102, 770 (1998).10.1021/jp973149eSearch in Google Scholar
[26] Y. K. Jeong, G. M. Choi. J. Phys. Chem. Solids57, 81 (1996).10.1016/0022-3697(95)00130-1Search in Google Scholar
[27] Y. S. Chaudhary, A. Agrawal, R. Shrivastav, V. R. Satsangi, S. Dass. Inter. J. Hydro. Ener. 29, 131 (2004).10.1016/S0360-3199(03)00109-5Search in Google Scholar
[28] T. D. Golden, M. G. Shumsky, Y. Zhou, R. A. V. Werf, R. A. Van Leeuwen, J. A. Switzer. Chem. Mater.8, 2499 (1996).10.1021/cm9602095Search in Google Scholar
[29] V. Srikant, D. R. Clarke. J. Appl. Phys. 83, 5447 (1998).10.1063/1.367375Search in Google Scholar
[30] D. C. Look. Mater. Sci. Eng. B80, 383 (2001).10.1016/S0921-5107(00)00604-8Search in Google Scholar
[31] D. Shinohara, S. Fujita. Jpn. J. Appl. Phys. 47, 7311 (2008).10.1143/JJAP.47.7311Search in Google Scholar
[32] H. H. Tippins. Phys. Rev.140, A316 (1965).10.1103/PhysRev.140.A316Search in Google Scholar
[33] T. Lange, W. Njoroge, H. Weis, M. Beckers, M. Wuttig. Thin Sol. Films365, 82 (2000).10.1016/S0040-6090(99)01106-2Search in Google Scholar
©2017 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/
Articles in the same Issue
- Frontmatter
- Graphical abstracts
- In this issue
- Preface
- 12th Conference on Solid State Chemistry (SSC-2016)
- Conference papers
- Structural modifications of metallic glasses followed by techniques of nuclear resonances
- Highly conductive barium iron vanadate glass containing different metal oxides
- Physico-chemical and optical properties of Er3+-doped and Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 chalcogenide glass
- Spectroscopic ellipsometry characterization of spin-coated Ge25S75 chalcogenide thin films
- The challenge of methods of thermal analysis in solid state and materials chemistry
- Mössbauer spectroscopy: epoch-making biological and chemical applications
- Redistribution of iron ions in porous ferrisilicates during redox treatments
- Textural and morphology changes of mesoporous SBA-15 silica due to introduction of guest phase
- Carbon dioxide and methane adsorption over metal modified mesoporous SBA-15 silica
- Titania aerogels with tailored nano and microstructure: comparison of lyophilization and supercritical drying
- Solvent-free, improved synthesis of pure bixbyite phase of iron and manganese mixed oxides as low-cost, potential oxygen carrier for chemical looping with oxygen uncoupling
- Synthesis, structure and thermal expansion of the phosphates M0.5+x M′x Zr2−x (PO4)3 (M, M′–metals in oxidation state +2)
- Visible-light activated photocatalytic effect of glass and glass ceramic prepared by recycling waste slag with hematite
- Structure and properties of nanocrystalline nickel prepared by selective leaching at different temperatures
- Corrosion protection of zirconium surface based on Heusler alloy
- Toward the control of graphenic foams
Articles in the same Issue
- Frontmatter
- Graphical abstracts
- In this issue
- Preface
- 12th Conference on Solid State Chemistry (SSC-2016)
- Conference papers
- Structural modifications of metallic glasses followed by techniques of nuclear resonances
- Highly conductive barium iron vanadate glass containing different metal oxides
- Physico-chemical and optical properties of Er3+-doped and Er3+/Yb3+-co-doped Ge25Ga9.5Sb0.5S65 chalcogenide glass
- Spectroscopic ellipsometry characterization of spin-coated Ge25S75 chalcogenide thin films
- The challenge of methods of thermal analysis in solid state and materials chemistry
- Mössbauer spectroscopy: epoch-making biological and chemical applications
- Redistribution of iron ions in porous ferrisilicates during redox treatments
- Textural and morphology changes of mesoporous SBA-15 silica due to introduction of guest phase
- Carbon dioxide and methane adsorption over metal modified mesoporous SBA-15 silica
- Titania aerogels with tailored nano and microstructure: comparison of lyophilization and supercritical drying
- Solvent-free, improved synthesis of pure bixbyite phase of iron and manganese mixed oxides as low-cost, potential oxygen carrier for chemical looping with oxygen uncoupling
- Synthesis, structure and thermal expansion of the phosphates M0.5+x M′x Zr2−x (PO4)3 (M, M′–metals in oxidation state +2)
- Visible-light activated photocatalytic effect of glass and glass ceramic prepared by recycling waste slag with hematite
- Structure and properties of nanocrystalline nickel prepared by selective leaching at different temperatures
- Corrosion protection of zirconium surface based on Heusler alloy
- Toward the control of graphenic foams