banner
Home / News / Early stage nanocrystallization as a method of enhancement of electrical properties of lead/barium titanate doped vanadium borate glasses
News

Early stage nanocrystallization as a method of enhancement of electrical properties of lead/barium titanate doped vanadium borate glasses

Apr 07, 2024Apr 07, 2024

Scientific Reports volume 13, Article number: 12498 (2023) Cite this article

Metrics details

Glass–ceramic nanocomposites (GCNs) of (10 − x) BaTiO3 (BT)–xPbTiO3 (PT)–60V2O5–30B2O3 with x = 0, 2.5, 5, 7.5 and 10 mol% were formed during heat treatment of conventional melt quenching glasses. X-ray diffraction was used to ensure glass and GCNs formation. Glasses and GCNs densities were measured by Archimedes principle. Fine polar clusters of lead titanate and/or barium titanate incorporation into vanadium borate glass matrix strongly depend on the composition. It was found out that the electrical conductivity of the initial glasses can be considerably improved by proper early stage of nanocrystallization at temperatures approaching the crystallization temperatures determined by DSC method. GCNs show massive increase in electrical conductivity (up to 6 orders of magnitude) as a function of BaTiO3 content. By increasing BaTiO3 content, the activation energy values have been found to increase. The enhancement in electrical conductivity of GCNs can be attributed to the increase in the crystalline phases in the glassy matrix which increases the concentrations of the V ion pairs.

Recently, glass–ceramic nano-composites (GCNs) containing ferroelectric nano-crystallites dispersed in the glass matrix received considerable interest. During the process of glass heat treatment, tiny crystals are formed inside the glass matrix, also, the level of porosity is reduced, which gives a great advantage to GCNs1,2. Glass and glass–ceramic systems embedded with ferroelectric materials have a remarkable applications such as radio frequency filters, actuators, flash memories etc.3. GCNs contaning transition metal oxide (TMO) show semiconducting behavior as a result of above one valency ions. In glass matrix contaning vanadium oxide conduction found due to small polaron hopping (SPH) between V4+ ↔ V5+. The grain size of formed nano-crystalline precepte plays significant role in enhanced electrical conductivity where hopping centers arrangment in that grains minimizing the grain boundary scattering4,5,6.

Lead titanate (PbTiO3) has a Peroviskite structure with good dielectric, thermal stability and ferroelectric properties. In addition, it’s high Curie temperature (490 °C) making it suitable for high temperature device applications, like transducers. The large ionic displacements of lead titanate produce large spontaneous polarization7,8. On the other hand, Barium titanate (BaTiO3) in the last few decades was highly investigated for its technological importance in electronics applications3,9.

For glass forming ability, vanadium pentoxide (V2O5) up to 5 mol%, acts as a network glass modifier while more than 10 mol% it acts as a network former5. Boron trioxide (B2O3) is a basic glass network former in borate glasses with smaller heat of fusion10, while BaTiO3 have a poor glass forming ability3.

Our work aims to study lead/barium titanate doped glass/GCNs prepared by conventional melt quenching technique and early stage nanocryatallization method, respectively. Moreover, the effect of barium titanate substituted by lead titanate, to produce lead free matreial, on the electrical properties was investigated. The formation of nanocrystalline phases were obtained by carful heat treatment at temperatures close to crystallization temperature.

Conventional melt quenching technique was used to prepare (10 − x) BaTiO3 (BT)–xPbTiO3 (PT)–60V2O5–30B2O3 with x = 0, 2.5, 5, 7.5 and 10 mol%. Extra high purity oxides reagent grade of PbTiO3, BaTiO3, B2O3 (Sigma Aldrich, 99%) and V2O5 (Fisher Scientific, 99.99%) were our starting chemicals with a total mixture of 10 g weighed in a stoichiometric ratio. After 10 min. mixing, powder of nominal compositions were placed in a platinum crucible and heated for 90 min in a Muffle furnace at 1250 °C in air. Stainless steel mold was used for casting the melt in plate shape of 2.0 mm thickness. Siemens D5000 X-ray diffractometer with nickel-filtered Cu Kα radiation was used to ensure the amorphicity of the prepared glasses under accelerating voltage of 40 kV and current of 30 mA. Shimatzou DSC 50, differential scanning calorimeter was used for thermal anaysis with heating rate of 10 °C/min under an argon atmosphere. JEOL 2100, high resolution transmission electron microscope (HRTEM) was used to confirm the presence of nano-clusters inside the glass matrix. The as-prepared glass sample was heat treated in air close to its crystallization temperature Tc according to DSC data at 350 °C for 2 h to obtain glass–ceramic nano-composites (GCNs). Glass and GCNs samples were coated by silver pastes for dc conductivity using Pico-ammeter type KEITHLEY 485 in temperatures range 310–450 K. At room temperature, glass and GCNS samples the average densities (ρ) were measured by Archimedes method using Toluene of density 0.866 g/cm3 as an immersion liquid. Density measurements were repeated five times.

XRD patterns of 10 − xBaTiO3 (BT)–xPbTiO3 (PT)–60V2O5–30B2O3 with x = 0, 2.5, 5, 7.5 and 10 mol% glass samples are shown in Fig. 1. A wide broad peak was observed at \(2\uptheta \approx 25\) confirming glassy behavior of our glass system. Figure 2 shows DSC curves of the composition dependence of the glass system. Curves are characterized by an endothermic drift due to glass transition temperature (Tg) and large exothermic peak corresponding to crystallization (Tc). Glass transition temperature shifts to higher values with increasing BaTiO3 content and lies between 280 and 320 °C. Moreover, Tc shifted from 314 to 347 °C according to sample composition. For instance, the glass thermal stability and viscosity depends mainly on the temperature difference between the glass transition and crystallization points11. The temperature difference ΔT = Tc − Tg slightly varied from 36 to 31 °C with increasing BaTiO3 content. Also, the change in the coordination number of the network former and the construction of non-bridging oxygens (NBOs) atoms correlated with Tg. The construction of NBOs causes a decrease into the Tg2,3. In the present work, the continuous increase in Tg suggests continuing increase in the coordination number and destruction of NBOs atoms.

As-quenched glasses X-ray diffraction patterns.

As-quenched glasses DSC curves.

HRTEM, selective area electron diffraction patterns (SAED) and the interplanar spacing of the as-quenched glass samples were presented in Figs. 3 and 4, respectively. Randomly distributed nanoclusters were precipitated in the amorphous glass matrix. The SAED patterns of the as-quenched glass samples confirm amorphous nature. However, spots in this pattern suggest the existence of nanoclusters precipitate that associated with the glass sample. Figure 4 shows the interplanar spacing of the ordered phases of the polar nano-clusters impeded in the glass matrix. The appearance of crystal defects discontinuity and presence of twin boundaries support relaxor ferroelectric-like behavior of the sample. This point was extensively studied in our recent published work12,13.

HRTEM of the as prepared glass samples the inset is the SAED.

The interplanar spacing observed in the present glass samples.

Figure 5 shows XRD patterns of heat-treated glasses at 350 °C for 2 h. Broad peak still associated with partially nanocrystalized matrix confirm GCNs formation. A small (120) peak shown in the XRD result at 25.97° of 10% PT GCNs corresponds to the monoclinic Pb2V2O7. The XRD pattern matches with the JCPDS card no-01-084-1458. For 10% BT GCNs, the (311) peak obtained at 25.83° corresponds to monoclinic Ba2V4O13. Furthermore, Ba2V4O13/Pb2V2O7 coexist in GCNs with mixed PT and BT content. Scherrer equation was used to calculate the crystal size in GCNs which ranged from 7 to 15 nm5. The existence of a very fine crystallites embedded inside the glass matrix was observed also by using HRTEM, Fig. 6. The SAED pattern of 5%BT glass ceramic nano-composite is shown in the inset of Fig. 6. Diffraction spots from nanocrystalline phases are observed with average size larger than observed for the as prepared glass samples. It is clear from the figure that the diffraction spots show high degree of disorder.

X-ray diffraction patterns of GCNs.

HRTEM of 5% BT glass ceramic nanocomposite (a) shows the presence of large particle in the glassy matrix and (b) shows the d spacing of a smal size grains.

The as-quenched glass density decreases from 3.5 up to 2.84 g/cm3 while molar volume increases continuously from 3.15 to 3.7 cm3 with the increase of the BaTiO3 content as shown in Table 1. After heat-treatment at 350 °C for 2 h these values increase. This increase has been attributed to the enhancement of borate tetrahedra formation besides the conversion of bridging oxygens into non-bridging oxygens. Also, the formation of noncrystalline phases compared to corresponding glasses played a significant role in higher density values.

The electrical conduction in glasses consisting of transition metal oxides (TMOs) have been proved to be electronic in nature. The conduction process is believed to occur due to the electron hopping between the ions existing in different valence states in the glass. Figure 7a,b represents the dc conductivity (σ) variations of the glass and corresponding GCNs as a function of reciprocal absolute temperature, respectively. The figure shows a linear temperature dependence up to a certain temperature (θD/2), (θD: Deby temperature), and suddenly the slope changes with deviation from linearity. Along with the activation energy is a temperature dependent. Generally, for glass systems containing TMOs14,15, the electrical conduction at temperatures beyond θD/2 is explained by the Austin-Mott theory of small polaron hopping (SPH) conduction16,17.

Temperature dependance of dc conductivity (ln σ) of different composition for (a) the glass samples (b) GCNs.

High temperature activation energies were obtained from the slope of each curve in the highest range of the measured temperatures. The experimental conductivity data at this situation is well described by activation energy for conduction gives by:

where σo is the pre-exponential factor, W is the activation energy and kB is Boltzmann’s constant and T is the absolute temperature. The diversity of the high temperature conductivities and the high temperature activation energies of glasses and corresponding GCNs are shown in Fig. 8. It is clear from the figure that the conductivity decreases while the activation energy increases by increasing BaTiO3 content.

Effect of BaTiO3 content on activation energy (W) and dc conductivity (ln σ) at fixed temperature (400 K) for the glasses and GCNs samples.

The activation energies for conduction of the glass samples were found to be W = 0.338–0.495 eV at high temperatures. It is known that by adding BaTiO3 to the glass matrix, it decreases the conductivity as a result of decreasing NBO cations18.This lead to decrease the open structure, through which the charge carriers can move with lower mobility. The activation energies for conduction of the GCNs were found to be = 0.179–0.244 eV. As shown from the figure, that the conductivity decreases while the activation energy increases with the increase of the BaTiO3 content similar as observed in the glass samples. However, there is a giant enhancement in the electrical conductivity of the GCNs. This enhancement in electrical conductivity can be attributed to the increase in the Pb2V2O7 nanocrystalline phases in the glassy matrix which increases the concentrations of the V ion pairs4,5. These results are consistent with XRD results.

Furthermore, the improvement of the electrical conductivity of GCNs about 6 order of magnitude under study can be explained in the following way; the key for electronic conduction in the glasses with high amount of V2O5 is the spatial distribution of V4+ and V5+ ions which are centers of hopping for electrons19,20. For initial glass, there is a slight random distribution of such centers. However, the early stage of nanocrystallization at temperatures near to crystallization temperature enhances crystallinity of the formation of nanocrystallites of V2O5 inside the glass matrix. Since the average size of these grains is so small, the interface region between crystalline and amorphous phases is widely ramose and strongly influences over all electrical properties of the nanomaterial. In particular, it may contain the improved concentration of V4+ and V5+ centers dispersed on the surface of V2O5 crystallites19. Figure 9 shows the dependence of dc conductivity (ln σ) as a function of BaTiO3 content for glass and corresponding GCNs at fixed temperature (400 K). It is clear from the figure that there is a massive increase in electrical conductivity of about 6 orders of magnitude after heat treatment.

Effect of BaTiO3 content on dc conductivity (ln σ) at fixed temperature (400 K) for the glasses and corresponding GCNs samples.

Moreover, the decrease in dc conductivities and the increase in activation energies of the studded samples suggest some changes in conduction mechanisms. It has been previously reported21,22 that in glasses which consist of vanadium, the dc conductivity is electronic and depends strongly upon the average distance, R, between the vanadium ions. The average distance, R, was calculated for the studded samples (see Tables 1 and 2) from the relation R = (1/N)1/3, where N is the concentration of vanadium ions per unit volume, calculated from batch composition and the measured density. The density, d, the concentration of vanadium ions per unit volume, N, and average distance, R, are shown in Tables 1 and 2 for all studded systems. The relation between the average distance R, activation energy W, and the dc conductivity, ln σ, for glasses and corresponding GCNs is illustrated in Fig. 10. On the other hand, the theoretical expression for that energy includes a term W = W0(1–rp/R), where W0 is constant and rp denotes a radius of small polaron17,21. Overall, the electronic conductivity increases while activation energy decreases when the distance R between hopping centers decreases.

Effect of the average distance (R) on activation energy (W) and dc conductivity (ln σ) at fixed temperature (400 K) for different glasses and GCNs samples.

The dc conduction process in TMO glasses was investigated by Mott16 in a model in which the conduction process is characterized by phonon assisted hopping of small polarons between localized sates at high temperature T > θD/2. For non-adiabatic hopping regime, where the electron jumping is small during each excitation the dc conductivity can be expressed by

The pre-exponential factor σo in Eq. (2) is given by

where ʋo is the optical phonon frequency which calculated from the electrical conductivity data following the relation (kB θD = hʋo) the values of ʋo are listed in Table 2, N is the transition metal ion density, R the average spacing between transition metal ions (= (1/N)1/3), C is the fraction of reduced transition metal ion(C = V+4/Vtotal), W is the activation energy for hopping conduction and α is the tunneling factor (the ratio of the wave function decay). Suggesting a strong electron–phonon interaction, Austin and Mott showed the following17.

where WH is the hopping energy and WD is the disorder energy which is the difference of electronic energies between two hopping sites.

where ɛs is the static dielectric constant and L is a constant of factor 0.3.

For the adiabatic hopping regime, the electron would make transitions backwards and forwards during excitation between hopping sites. At that case αR in Eqs. (2) and (3) becomes negligible16,17. In this region the conductivity and the pre-exponential factor in Eqs. (2) and (3) were described as the following:

The nature of polaron hopping mechanism (adiabatic or non-adiabatic) can be acquired from a plot of ln σ versus activation energy W at fixed temperature T23. If the temperature estimated Te from the slope of such a plot is close to experimental temperature T, the hopping conduction will be in the adiabatic regime. Nevertheless, if the estimated temperature Te is differed than the experimental temperature T the conduction will be in the non-adiabatic regime. Figure 10 shows the relation between ln σ and activation energy W for glasses and GCNs at T (400 K). The estimated temperatures calculated from the slopes of the plots are Te = 753.5 and Te = 529.9 for glasses and GCNs, respectively which is differ than the experimental temperature. As a consequence, we can confirm the conduction mechanism in the present samples is due to non-adiabatic SPH of electrons17.

The non-adiabatic conduction mechanism is also confirmed from the polaron band width (J) calculation from the following relation24:

The values of J which calculated from the right hand side (R.H.S) of the Eq. (9) or (10) at fixed temperature (400 K) for the glass samples are in the range of (0.015–0.018) eV and for the GCNs are in the range of (0.003–0.014) eV. To clarify whether the polaron is in the adiabatic or in the non-adiabatic regime requires an estimate of the value of J, which can be obtained from25.

where

where ɛs and \({\varepsilon }_{\infty }\) are the static and high frequencies dielectric constants of the samples, respectively. \({\varepsilon }_{p}\) the effective dielectric permittivity.

And \(N({E}_{F})\) is the density of states at Fermi level which estimated from the following expression16.

Table 1 shows the values of \(\mathrm{N}({\mathrm{E}}_{\mathrm{F}})\) of glass and GCNs, then, the calculated values of J given from Eq. (11) is about \(\mathrm{J}\approx 0.0003\) for glasses and about \(\mathrm{J}\approx 0.0001\) for GCNs which are away smaller than those calculated from the R.H.S of Eq. (10) confirming again the non-adiabatic hopping conduction at high temperature for these samples.

In the adiabatic regime the hopping conduction, WH, is given using J value as

where Wp is the polaron binding energy, W/p is the maximum polaron binding energy and WH depends on R26.

Otherwise, for non-adiabatic hopping conduction WH is given by

By using the values of disorder energy WD and activation energies W showed in Table 1, we obtain WH in the range of (0.338–0.493) eV for the glass samples and in the range of (0.179–0.244) eV for the GCNs. These values are much close to W values for the studied systems.

Next, by using the values of mean spacing between vanadium ions, R (Table 1), the polaron radii can be calculated from the relation25.

The values of \({r}_{p}\) and R for glass and corresponding GCNs are listed in Tables 1 and 2.

The values of small polaron coupling constant a measure of the electron–phonon interaction, given bywere also calculated for the studied systems16. The estimated values of are listed in Table 2. The value of  > 4 usually indicates a strong electron–phonon interaction27. The listed values show a decrease in with decreasing BaTiO3 content.

The hopping carrier mobility \(\mu\) in the adiabatic and non-adiabatic hopping regions is described by the following equations28.

Also, the carrier density (Nc) values were calculated from the relation29.

The values of \(\upmu\) and Nc for glass and corresponding GCNs are listed in Table 2. The carrier mobility at 400 K of the studied systems is very small, suggesting that the electrons are highly localized at the V ion sites, corresponding to the strong electron–phonon interaction the large γp27. As the condition of the localized for the conductive electrons is generally for \(\upmu <0.01\) cm2V−1 s−127, confirming the formation of small polaron in our samples. Also, the constant Nc ~ 1018 indicates that the conductivity of such samples is determined by hopping mobility30,31,32,33.

Glass system of (10 − x) BaTiO3 (BT)–xPbTiO3 (PT)–60V2O5–30B2O3 with x = 0, 2.5, 5, 7.5 and 10 mol% were successfully prepared by the conventional melt-quenching technique. The amorphous nature was confirmed by XRD and DSC. HRTEM micrograph along with SAED clarified the existence of polar nanoclusters in the glass matrix and its size increases by heat treating the glass samples. The density values of the glass were found to be increase with the increase in BaTiO3 content while GCNs samples density values decrease. The dc conduction process is believed to occur by electron hopping between the ions existing in different valence states in the glass system (hopping will take place between the V4+ and V5+ ions). The dc conductivity in the glass system decreases while the activation energy increases with the increase of the BaTiO3 content. After heat treating the glass samples, there is a giant enhancement of the electrical conductivity in GCNs. The enhancement in electrical conductivity can be attributed to the increase in the nanocrystalline phases in the glassy matrix which increases the concentrations of the V ion pairs. The non-adiabatic nature of polaron hopping mechanism was confirmed in present samples. The carrier density is almost constant (Nc ~ 1018 cm−3) denoting that the conductivity of these samples is mainly determined by hopping mobility.

The datasets used and/or analyzed during the current study available from the corresponding author on reasonable request.

Al-Assiri, M. S. & El-Desoky, M. M. Synthesis, structural and ferroelectric properties of barium titanate based glass-ceramic nano-composites. J. Non-Cryst. Solids 358(12), 1605–1610 (2012).

Article ADS CAS Google Scholar

Al-Assiri, M. S. & El-Desoky, M. M. Correlation between nanostructural and electrical properties of barium titanate-based glass–ceramic nano-composites. J. Alloy. Compd. 509(36), 8937–8943 (2011).

Article CAS Google Scholar

Thakur, S., Thakur, V., Kaur, A. & Singh, L. Study of the crystallization and structural behavior of bismuth barium titanate glass-ceramics. J. Non-Cryst. Solids 557, 120563 (2021).

Article CAS Google Scholar

Hannora, A. E., Abbas, M. A. & El-Desoky, M. M. Annealing effects on the structural, thermal, and electrical properties of 10PbTiO3–10Fe2O3–30V2O5–50B2O3 glass. J. Mater. Sci. Mater. Electron. 32, 3998–4007 (2021).

Article CAS Google Scholar

Hannora, A. & El-Desoky, M. Correlation between nanostructural and enhanced electrical conductivity of annealed 30V2O5–20Bi2O3–50P2O5 glass. Cryst. Res. Technol. 55(10), 2000128 (2020).

Article CAS Google Scholar

Laorodphan, N. et al. Boron and pentavalent vanadium local environments in binary vanadium borate glasses. J. Non-Cryst. Solids 453, 118–124 (2016).

Article ADS CAS Google Scholar

Rahangdale, V. U. & Deshpande, V. K. Study of physical properties of PbTiO3 based glass-ceramic with variation of heat treatment duration. Ferroelectrics 467(1), 85–98 (2014).

Article ADS CAS Google Scholar

Rahangdale, V. U. & Deshpande, V. K. Lead-zirconate-titanate based glass and glass-ceramics: Preparation, physical, dielectric and ferroelectric properties. Mater. Today Proc. 29, 866–871 (2020).

Article CAS Google Scholar

Thakur, V., Singh, A., Awasthi, A. M. & Singh, L. Temperature dependent electrical transport characteristics of BaTiO3 modified lithium borate glasses. AIP Adv. 5(8), 087110 (2015).

Article ADS Google Scholar

Motke, S. G., Yawale, S. P. & Yawale, S. S. Infrared spectra of zinc doped lead borate glasses. Bull. Mater. Sci. 25(1), 75–78 (2002).

Article CAS Google Scholar

El-Desoky, M. M. Characterization and transport properties of V2O5–Fe2O3–TeO2 glasses. J. Non-Cryst. Solids 351(37), 3139–3146 (2005).

Article ADS CAS Google Scholar

El-Desoky, M. M., Harby, A. E., Hannora, A. E. & Al-Assiri, M. S. Relaxor ferroelectric-like behavior in barium titanate-doped glass via formation of polar clusters. J. Clust. Sci. 28(4), 2147–2156 (2017).

Article CAS Google Scholar

Harby, A. E., Hannora, A. E. & El-Desoky, M. M. Observation of relaxor-like behavior in BT and PT doped glasses for energy storage applications. J. Alloy. Compd. 770, 906–913 (2019).

Article CAS Google Scholar

Ghosh, A. & Chaudhuri, B. K. DC conductivity of V2O5–Bi2O3 glasses. J. Non-Cryst. Solids 83(1), 151–161 (1986).

Article ADS CAS Google Scholar

Mori, H., Kitami, T. & Sakata, H. Electrical conductivity of V2O5–Sb2O3–TeO2 glasses. J. Non-Cryst. Solids 168(1), 157–166 (1994).

Article ADS CAS Google Scholar

Mott, N. F. Conduction in glasses containing transition metal ions. J. Non-Cryst. Solids 1, 1–17 (1968).

Article ADS CAS Google Scholar

Austin, I. G. & Mott, N. F. Polarons in crystalline and non-crystalline materials. Adv. Phys. 18(71), 41–102 (1969).

Article ADS CAS Google Scholar

El-Desoky, M. M. & Al-Assiri, M. S. Structural and polaronic transport properties of semiconducting CuO–V2O5–TeO2 glasses. Mater. Sci. Eng. B 137(1), 237–246 (2007).

Article CAS Google Scholar

Garbarczyk, J., Jozwiak, P., Wasiucionek, M. & Nowinski, J. Enhancement of electrical conductivity in lithium vanadate glasses by nanocrystallization. Solid State Ion. 175, 691–694 (2004).

Article CAS Google Scholar

Al-Assiri, M. & El-Desoky, M. Nanocrystallization as a method of improvement of electrical properties of Fe2O3–PbO2–TeO2 glasses. J. Mater. Sci. Mater. Electron. 25, 3703–3711 (2014).

Article CAS Google Scholar

El-Desoky, M. M. DC conductivity and hopping mechanism in V2O5–B2O3–BaO glasses. Phys. Status Solidi 195(2), 422–428 (2003).

Article ADS CAS Google Scholar

Al-Shahrani, A., Al-Hajry, A. & El-Desoky, M. M. Non-adiabatic small polaron hopping conduction in sodium borate tungstate glasses. Phys. Status Solidi 200(2), 378–387 (2003).

Article ADS CAS Google Scholar

Emin, David & Holstein, T. Studies of small-polaron motion IV. Adiabatic theory of the Hall effect. Ann. Phys. 53(3), 439–520 (1969).

Article ADS Google Scholar

Friedman, L. D. & Holstein, T. Am. Phys. 21, 494–499 (1963).

ADS CAS Google Scholar

Al-Assiri, M. S., Salem, S. A. & El-Desoky, M. M. Effect of iron doping on the characterization and transport properties of calcium phosphate glassy semiconductors. J. Phys. Chem. Solids 67(8), 1873–1881 (2006).

Article ADS CAS Google Scholar

Hirashima, H., Aral, D. & Yoshida, T. Electrical conductivity of PbO–P2O5–V2O5 glasses. J. Am. Ceram. Soc. 68(9), 486–489 (1985).

Article CAS Google Scholar

Mott, N. F. & Davis, E. A. Electronic processes in non-crystalline materials 2nd edn, 32 (Clarendon, 1979).

Google Scholar

Ghosh, A. Transport properties of iron-bismuthate glassy semiconductors. J. Appl. Phys. 66(6), 2425–2429 (1989).

Article ADS CAS Google Scholar

Andelean, I. Solid State Commun. 27, 697–703 (1978).

Article ADS Google Scholar

Fareed, S. A. et al. Structure, seebeck coefficient and DC electrical conductivity of Bi2Mn4O10 prepared by mechanochemical method. J. Mater. Sci. Mater. Electron. 33(19), 15346–15358 (2022).

Article CAS Google Scholar

Szreder, N. A. et al. Electronic and ionic relaxations in strontium–borate glass and glass-ceramics containing bismuth and vanadium oxides. Solid State Ion. 282, 37–48 (2015).

Article CAS Google Scholar

Szreder, N. A. et al. Microstructure and dielectric properties of barium-vanadate glasses. Procedia Eng. 98, 62–70 (2014).

Article CAS Google Scholar

Pietrzak, T. K., Wasiucionek, M. & Garbarczyk, J. E. Towards higher electric conductivity and wider phase stability range via nanostructured glass-ceramics processing. Nanomaterials 11(5), 1321–1344 (2021).

Article CAS PubMed PubMed Central Google Scholar

Download references

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through large group Research Project under Grant Number RGP2/193/44.

Department of Physics, Faculty of Science, Suez University, Suez, Egypt

Amany E. Harby & M. M. El-Desoky

Department of Science and Mathematical Engineering, Faculty of Petroleum and Mining Engineering, Suez University, Suez, 43721, Egypt

Ahmed E. Hannora

Department of Physics, Faculty of Science, King Khalid University, Abha, Saudi Arabia

Atif Mossad Ali

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

All authors equal share contributions, all authors prepared figures, all authors reviewed the manuscript.

Correspondence to M. M. El-Desoky.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Harby, A.E., Hannora, A.E., Ali, A.M. et al. Early stage nanocrystallization as a method of enhancement of electrical properties of lead/barium titanate doped vanadium borate glasses. Sci Rep 13, 12498 (2023). https://doi.org/10.1038/s41598-023-37230-w

Download citation

Received: 14 April 2023

Accepted: 18 June 2023

Published: 02 August 2023

DOI: https://doi.org/10.1038/s41598-023-37230-w

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.