Next Article in Journal
Iron–Sulfur Clusters: Assembly and Biological Roles
Next Article in Special Issue
Progress, Applications, and Challenges of Amorphous Alloys: A Critical Review
Previous Article in Journal
Improving Charge Transport in Perovskite Solar Cells Using Solvent Additive Technique
Previous Article in Special Issue
Discovering Novel Glass with Robust Crystallization Resistance via Amorphous Phase Separation Engineering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nickel Ions Activated PbO–GeO2 Glasses for the Application of Electrolytes and Photonic Devices

1
Department of Automotive Engineering, Yeungnam University, Gyeongsan-si 38541, Republic of Korea
2
Department of Physics, Sai Spurthi Institute of Technology, B. Gangaram, Sathupally Mandal 507303, India
3
Department of Physics, Usha Rama College of Engineering & Technology, Telaprolu 521109, India
4
Department of H&S (Chemistry), CVR College of Engineering, Hyderabad 501510, India
5
Department of Physics, Gokaraju Rangaraju Institute of Engineering & Technology, Hyderabad 500090, India
6
Department of Physics, Acharya Nagarjuna University, Nagarjuna Nagar, Guntur 522510, India
*
Authors to whom correspondence should be addressed.
Inorganics 2024, 12(8), 215; https://doi.org/10.3390/inorganics12080215
Submission received: 27 June 2024 / Revised: 29 July 2024 / Accepted: 7 August 2024 / Published: 8 August 2024
(This article belongs to the Special Issue Recent Research and Application of Amorphous Materials)

Abstract

:
In this study, PbO–GeO2 glasses were melt-quenched at different nickel oxide concentrations. XRD and DSC techniques were characterized whether the samples are glass or crystalline materials. IR, Raman, and optical absorption techniques are used to obtain structural details. The IR spectra have revealed that the glass network contained conventional structural units GeO4 and GeO6. The Ni2+ ion octahedral transition exhibited luminescence spectra in the region of 1200–1500 nm; it is due to 3T2 (3F) → 3A2(3F) transition. The glasses containing the highest concentration of NiO have been found to have high values of luminescence efficiency and the cross-section. The dielectric characteristics, such as the dielectric constant, loss, and a.c. conductivity (σac), were analyzed across extensive frequency and temperature ranges, with a specific emphasis on the nickel oxide concentration. Analyzing optical absorption and dielectric properties of the samples, it has been found that nickel ions’ majority occur in tetrahedral sites. It is proved that the dielectric constant and loss values are highest for the sample N10 and ac conductivity due to dipoles being lowest for the sample N10. It is revealed that the glasses are highly conducting due to the modifying action of Ni2+ ions so these glasses are suitable for solid electrolyte uses besides their optical applications in NLO devices.

1. Introduction

Glass plays a crucial role in various applications and technologies, making it an essential material in our daily lives. So, the glass industry was founded with the aim of creating a wide range of compounds for everyday use. In all these materials, the most important heavy metal oxide PbO in general behaves as a glass modifier. Upon entering into a glass network, the structure of the material changes. The low crystallization value and high moisture resistance are exposed by this information [1]. It has both a covalent and an ionic nature, which allows it to function as both a glass maker and a modifier [2]. The refractive index and melting point of materials are enhanced when lead oxide is introduced to the glass former. These types of glasses are utilized in situations where shielding, dispersion, and high refractive index are necessary [3]. The glassy nature is enhanced by [PbO4] structural units whereas the network modifying action is due to [PbO6] structural units of PbO. However, germanium oxide is commonly encountered in its vitreous state. It is considered a classic glass former and is known for its exceptional optical characteristics, including a refractive index (n~2.17). Moreover, it serves as an efficient transmitter in the near-infrared (NIR) region [4]. Germanium oxide glasses are used for making optical fibers and designing laser devices and in the field of nonlinear optics [5]. Since GeO2 glasses have a high ionic conductivity, solid electrolytes are usually composed of them [6]. A structural analysis of germanium oxide indicates that it is composed of structural units made up of GeO4 and GeO6 [7,8,9]. Depending on the type of bond between lead and oxygen atoms, the modifier PbO will act either as a network former or a modifier. Because of Pb ion high polarizability, a strong covalent Pb–O bond is formed between Pb2+ and O2− ions [10]. Lead germanate glasses are promising candidates for optical amplifiers, and for numerous types of optical devices and high-speed optical switches containing nonlinearity [11].
The transition metal ions continue to be of interest for possible laser applications. Numerous transition metal activators have shown promise, particularly those that have led to successful lasers utilizing cation matrices. The existence and level of cations in glass have a notable influence on the key physical, structural, and chemical characteristics of these materials, encompassing electrical, magnetic, and optical properties [12,13].
Nickel containing glasses gained special attraction due to their interesting optical and dielectric properties; nickel exists in highly suitable Ni2+ ions, which mostly occupies octahedral sites in glasses with high crystal field energy [14]. There is no accurate evidence of change in its oxidation state during glass formation and annealing [15]. Visible and near-infrared region Ni2+ ions can show many absorption bands. The Ni2+ ions in octahedral configuration in a glass matrix can exhibit lasting action at about a 1.56 µm wavelength at room temperature. This wavelength is very important in telecommunications [16]. The zero-dispersion wavelength for GeO2 is 1.7 µm (higher than that of borate, phosphate, or silicate glasses). The emission of laser beams in the near-infrared region is because of the existence of Ni2+ ions in the PbO–GeO2 glass network [17]. In addition to being used in nonlinear optical devices, these glasses were also used in the preparation of broad band optical amplifiers, power limiters, optical amplifiers, ultrafast optical switches, and lasers [18,19]. The coordination of Ge would continuously change from GeO4 to GeO6 with the addition of alkali oxides. Recently, the structural properties of germanate-based glasses have been examined in Na2O-GeO2-TeO2 [20], MnO-GeO2-PbO2 [21], TeO2-GeO2-PbO [22], Na2CO3-CaO-GeO2 [23], and Ga2O3-GeO2-BaO [24]. Among the reported germanate-based glasses, the lead germanate matrix has received the most attention due to its excellent properties, i.e., a relatively large glass-forming region, high transmittance in a wide wavelength region, superior chemical durability, and thermal stability [25,26].
Very little work was reported about nickel oxide-doped lead germanate glasses; moreover, work on the dielectric properties of dense glass systems like PbO–GeO2 glasses is very strange. So, the present work is aimed to report spectral features and dielectric studies of a nickel oxide-doped PbO–GeO2 glass system by means of spectroscopic techniques and electrical measurements, thereby comprehending the environment of nickel ions and explaining suitability of these glasses for specific uses as electrolytes.

2. Results

From the measured density values of PbO–GeO2 glass samples, various other physical parameters such as polaron radius rp and nickel ion concentration Ni are calculated and presented in Table 1. As the concentration of nickel oxide increases, the density of the samples is also increased. The structural variations in the glass network are investigated by physical properties of the samples.
Figure 1 illustrates the amorphous structure of the samples, as demonstrated by the absence of Bragg peaks in X-ray diffraction patterns.
Figure 2 displays the differential scanning calorimetric traces of the PbO–GeO2 glass system doped with nickel oxide, along with the glass-forming ability parameter (Kgl) and the variation of glass transition temperature (Tg) as indicated in the inset. The pattern demonstrates fluctuations in Tg in the temperature range of 490 to 510 °C with an endothermic effect. Additionally, an exothermic effect is observed due to the crystallization temperature Tc, which falls in the range of 615 to 630 °C. The DSC data in Table 2 illustrate the changes in (Tg), (Tc − Tg), (Tm − Tc), and (Kgl) [27] of PbO–GeO2 glasses as they are doped with varying nickel oxide concentrations. It has been observed that the glass-forming ability parameter Kgl decreases as the concentration of nickel oxide increases. This trend may be attributed to the alterations in the glass composition and structural properties resulting from the addition of nickel oxide.
As shown in Figure 3, titled glass samples were investigated in the 350–1550 nm wavelength range at room temperature. Between 700 nm and 820 nm, a prominent band was observed in these spectra. The width and height are increased with increasing the content of NiO in these spectra. Regarding influences of Ni2+ ions, a number of absorption bands are observed. In the visible and NIR regions, 1.0 mol% glass (N10) shows six evidently resolved absorption bands at 1265 nm (Oh1), 1029 nm (Td1), 781 nm (Oh2), 623 nm (Td2), 495 nm (Oh3), and 421 nm (Oh4) [28,29]. When the concentration of nickel increases, the intensity of octahedral bands is increased while tetrahedral bands are observed to decrease.
We calculated optical band gaps (Eo) of samples from cutoff wavelengths by using the relation
( α ω ) 2 = C ω E o
where α is the absorption coefficient (cm−1), ћω is photon energy of incident radiation (eV), Eo is the optical band gap energy (eV), and C is the band tailing parameter.
Figure 4 indicates that the Tauc plots and inset give optical band gap variation with concentration. Using the linear portion as a reference, Eo values are determined as shown in Table 3; the maximum Eo value is observed for the N2 glass sample.
The FTIR spectra have been depicted in Figure 5. These spectra exhibit two crucial bands that correspond to the tetrahedral and octahedral vibrations of GeO2 at 1075 and 780 cm−1, respectively. Because of GeO6 and GeO4 units, asymmetrical stretching vibrations regarding these bands are located [18]. One more band is observed due to bonding between two tetrahedral groups at 642 cm−1 [19,30]. The PbO4 tetrahedral units are located at about 448 cm−1 [18]. Intensity of GeO6 and Ge–O–Ge of GeO2 bands decreased and shifted gradually towards higher frequencies; similarly, GeO4 and PbO4 bands shifted towards lower frequencies with increased intensity with the introduction of nickel oxide into the glass network. The FTIR bands are shown in Table 4.
The Raman spectra of PbO–GeO2: NiO are shown in Figure 6. These spectra contained symmetric Ge–O–Ge bands due to Q4 units of GeO4 in between the 420 and 450 cm−1 region [31], a vibrational band of Q3 units of GeO4 in the region 520–540 cm−1, a vibrational band of GeO6 blocks at 620 cm−1 [32], and two weak bands due to Q2 and Q1 blocks of GeO4 at 900 and 960 cm−1, respectively. PbO4 exhibited blocks at 280 cm−1. A high intense vibrational band of localized Ge–O stretching of the meta-germanate blocks at around 750 cm−1 and no bands are exhibited by NiO units. The information regarding the band positions of Raman spectra can be found in Table 5.
Photoluminescence spectra of PbO–GeO2: NiO glasses are shown in Figure 7. Photoluminescence spectra of nickel oxide were recorded at 300 k with λexc = 800 nm. In the wavelength region 1200–1500 nm, the spectra exhibited a wider emission for all the samples due to the 3T2 (3F) → 3A2(3F) transition of Ni2+ ions. The intensities and line widths of the band increase with nickel oxide concentration as shown in Table 6, while the peaks shift toward higher wavelengths.
The variation of ε′ verses temperature graphs of the glasses is shown in Figure 8. At higher temperature and lower frequencies, slight increases in the dielectric constant were observed and it increases abruptly with temperature. It should be high for glass N10. Figure 9 shows the temperature versus tan δ graphs of glass samples at different frequencies. The curves have shown a maximum value of tan δ for the glass N10. The σac values are obtained by changing temperature using the following equation:
σ a c = ω ε 1 ε 0 tan δ
where σac is ac electrical conductivity (S/m), ω is angular frequency (s−1), and ε1 is the dielectric constant.
Figure 10 represents the plots of ac conductivity, σac, with 1/T and is drawn at 100 kHz for all the glasses. The inset of Figure 10 explains the variation of activation energy with changes in electrical conductivity σac. N10 has shown the lowest value of activation energy as presented in Table 7.
It is verified that the σac increased as activation energy (A.E) decreased, following the Arrhenius equation.
σ a c = σ 0 e E a K T
where σac is ac electrical conductivity (S/m) and σ0 is conductivity at zero Kelvin.
Ea is activation energy (eV), T is temperature, and K is the Boltzmann constant.
The variation of the 4πNµ2/27 K value with nickel oxide concentration is shown in Figure 11.
ε s ε ( ε s + 2 ) ( ε + 2 ) T = 4 Π N μ 2 27 K
where εs is the low-frequency dielectric constant, ε is the high-frequency dielectric constant, Nµ2 is the strength of dipoles (C2m−1), and K is the Boltzmann constant.
2 is the strength of dipoles. For various concentrations of nickel oxide, the values of εs and ε subsisting in the above equation and 4πNµ2/27 K are calculated at 393 K. The curve has shown a rising trend with nickel oxide concentration.

3. Discussion

In the PbO–GeO2 glass system, the presence of GeO2 in conjunction with GeO4 tetrahedra within the network of the glass can lead to the conversion of a few tetrahedral units into GeO6. There is typically one GeO6 octahedron separated by one GeO4 tetrahedron that shares one oxygen corner. Studies on GeO2 glasses have revealed interactions between the GeO5 structural units and germanium ions, where four bonds are of shorter length (ranging from 1.76 to 1.85 Å) and one bond has a longer interatomic distance of 1.92 Å.
The stretching of GeO4 tetrahedra of a Ge–O bond is induced by the modifier PbO. As a result, the presence of Pb ions leads to the formation of three-membered rings in the vicinity of GeO4 tetrahedra. The results revealed that Ge ions also participated in the glass structure with germanium dioxide network units. In addition, the infrared spectra have shown that there are bonds between GeO4 and PbO4, PbO2, PbO5, and PbO6 structural units [33]. Lead oxide is a glass modifier; it forms Ge–O–Ge linkages by entering into the glass structure along with NBOs and dangling bonds. When lead ions form four covalent bonds with oxygen atoms, they can form structural units called [PbO4] and [PbO2]. [PbO5] and [PbO6] are produced by the lead ions and form Ge–O–Pb linkages whereas the Ni2+ acts as a network modifier. Transition metal ions, including Ni2+, are known to influence the structural, optical, and electrical properties of glasses significantly. Ni2+ can act as a network modifier by occupying interstitial sites and creating non-bridging oxygens (NBOs), thus altering the glass matrix’s connectivity and properties. In our study, Ni2+ ions are introduced into the PbO–GeO2 glass network, primarily occupying octahedral sites. The observed spectroscopic features and the dielectric properties of the glasses support this conclusion. Our findings are consistent with previous research that highlighted similar roles for transition metal ions in glass systems [34].
Usually, the density of glass samples depends on glass structure, size of grains, coordination number of glass-forming ions, size of interstitial spaces, etc. The density of samples is increased with the content of NiO, indicating its higher compactness for the N10 sample.
The absence of sharp Bragg peaks indicates that the samples are in short-ranged structured materials (amorphous in nature). The glass-forming ability (Tc − Tg) decreased as the content of nickel oxide increased. So, it is proved that the lowest nickel oxide content is preferred for suitable glasses. It is inferred that glass samples with lower values will have less thermal stability. The decrease in interatomic distance, cross-linking, and closely packed structure are the reasons for such a behavior. Further, it is justified that the Ni2+ ions mostly acted as a modifier in (Oh) configuration as the NiO content is increased.
According to Tanabe–Sugano notation, the transitions of the d8 ion of nickel are mentioned below: 3A2(F) → 3T2 (F) (Oh1), 3T1 (F) (Oh2), 1T2 (D) (Oh3). The 3A2(F) → 1T2 (D) corresponds to a spin forbidden band at 570 nm while the band at 530 nm is due to 3A2 (F) → 3T1 (P) tetrahedral transition [35]. The increasing intensity in the octahedral bands with an increase in the mol% of NiO proved the presence of octahedrally bonded Ni2+ ions in a large number rather than the tetrahedrally bonded Ni2+ ions.
The optical band gap decreases as nickel oxide concentration increases. With an increase in doping levels, Ni2+ ion donor centers in a large number are produced. So, the excited states of local electrons superpose with the unfilled 3d states of nearby Ni2+ sites. This causes the nickel oxide band to extend into the main band gap to a greater extent. From N2 to N10, the band gap has fallen due to the shift of cutoff wavelengths to a higher wavelength side.
The GeO6 band intensity is increased with an increasing concentration of nickel oxide. This indicates that nickel ions are located in the glass structure in an octahedral fashion. In addition to the wavelength shift of the octahedral bands, nickel oxide doping also increases the wavelength shift. This shift in the band position indicated higher bond length of the Ni–O bond.
The IR spectra indicate the presence of symmetrical and anti-symmetrical bands corresponding to germanate structural units. The nickel oxide concentration is inversely proportional to the intensity of symmetric (GeO6) units. Due to Ge–O–Ge symmetric vibrations, the intensity of bands decreases and anti-symmetrical bands due to (GeO4) have increased as shown in Table 4. These results suggested decreasing entropy in the glass network with a hike in nickel oxide content.
It has been observed that Ge-O–Ge bonds of Q2 and Q1 units of GeO4 show weak asymmetric stretching at about 900 and 960 cm−1 in Raman spectra. It was observed that the meta-germanate Ge–O unit stretching modes showed a strong vibrational band at around 750 cm−1.
In the region 430 and 520 cm−1, GeO4 exhibits symmetric Ge–O–Ge bands due to the Q4 unit and a vibrational band of the Q3 unit. Additionally, a vibrational band of GeO6 and PbO4 blocks are observed at 620 and 280 cm−1. No bands are exhibited by nickel oxide. The intensity of asymmetric stretching vibrations of GeO6 units is increased at the expense of symmetric stretching vibrations regarding GeO4 units in the Raman spectra. The concentration of nickel oxide increases and evidently suggested disorder in the glass network decreases.
In general, octahedral nickel ions produce photoluminescence, and tetrahedral nickel ions have no photoluminescence. In the green, red, and near-infrared (NIR) wavelength ranges, the octahedral Ni2+ ions produce luminescence bands [36]. For our samples, λexc = 800 nm and the band ranges from 1200 to 1500 nm and the meta-center shifts towards a higher wavelength with an increase in nickel oxide content. The Stokes shift increases between higher and lower bands. The value of the emission cross-section σ p E has been computed and tabulated. Increasing the content of nickel oxide, the value of σ p E is found to increase gradually from N2 to N10, proving the increase in luminescence efficiency.
Electronic, ionic, dipolar, and interfacial polarizations are the different types of polarizations. These polarizations can influence the relative permittivity ε′ and interfacial polarization is a response to the structure of glasses. The value of interfacial polarization depends on the dielectric parameters (ε′, tan δ). These are found to slightly increase and σac decreases with temperature for any frequency. The activation energy for conduction decreased with the increase in nickel oxide content (Table 7), which proved the increase in the value of interfacial polarization.
The progressive rise in modifying ion concentration generates wrong bonds in the glass network. These defects cause rather easy movement of charges, thereby enhancing interfacial polarization and hence increasing the values of dielectric parameters.
Figure 10 depicts the variation of ac conductivity with the content of nickel oxide; the curves produced a maximum for pure glass and minimum for x = 1.0 mol%. The trend of the curves justified the electronic conductivity and a fixed value of ionic conductivity due to Pb2+ ions and Ni2+ ions. As the nickel oxide content increases, more Ni2+ ions are released into the glass network. The hopping polarons, which participate in conduction, are grabbed by Ni2+ ions in pairs. This combination of cation–polaron pairs diffuse as single neutral units. Movement of these neutral units cannot result in electrical conductivity so the ac conductivity decreases as the content of NiO is increased from 0.2 to 1.0 mol%. The ionic conductivity of Pb2+ ions is constant in a high-temperature region. As a result of Austin and Mott’s quantum mechanical tunneling model, the low-temperature part of conductivity can be explained [37]. From the results of dielectric properties, it is suggested that the significance of insulating nature is decreased with increasing content of nickel oxide.

4. Experimental Procedure

The following chemical formulas were used to synthesize the glasses for this study:
  • N0: 40PbO-60GeO2
  • N2: 40PbO-59.8GeO2: 0.2NiO
  • N4: 40PbO-59.6GeO2: 0.4NiO
  • N6: 40PbO-59.4GeO2: 0.6NiO
  • N8: 40PbO-59.2GeO2: 0.8NiO
  • N10: 40PbO-59.0GeO2: 1.0NiO
High-purity chemicals of PbO, GeO2, and NiO (all are in mol%) are taken in powder form in a suitable proportion and well grounded. In a controlled furnace, properly stirred mixtures were melted at 1000–1100 degrees Celsius until bubble-free liquid was obtained for about 1 h. As a result, the bubble-free melt was poured into rectangular brass molds and instantly annealed at 350 °C for four hours in another furnace for avoiding air cracks in the samples. At last, glass samples are obtained. The Philips expert system was used to obtain the XRD patterns of glass specimens. The programmable VIBRA HT kit was utilized to conduct density measurements, employing Archimedes’ law and O-xylene as the buoyant liquid. The optically polished samples have dimensions of 1 cm × 1 cm × 0.1 cm. A high-precision NIR spectrophotometer was used to obtain optical spectra. FTIR spectra were obtained using a standard spectrophotometer by a powdered sample of KBr pellets. The Raman spectra, photoluminescence spectra, and DSC traces are produced on standard instruments. The glasses were silver-coated for measuring electrical parameters. The coating was then exposed to a blower until it was dried. The electrical parameters were taken on a prominent LCR meter. The precision in determining the ε′ is 1 10 3 and that of loss is 1 10 4 .

5. Conclusions

40PbO–60GeO2: NiO glasses were prepared with different concentrations from 0.2 to 1.0 mol%. The XRD and DSC patterns indicated no crystalline phases. The IR and Raman spectral analysis revealed that the entropy of the glass network decreased with increasing content of nickel oxide. The investigation into the optical absorption, ESR, and photoluminescence peaks has demonstrated a growing occurrence of Ni2+ ions within tetrahedral sites. The values of susceptibility and magnetic moments of the glass samples have indicated Ni2+ ion transformation from octahedral to tetrahedral sites. The electrical parameters ε′ and tan δ have increased and σac increased whereas the activation energy has decreased, which proved their decreasing insulating nature with an increasing concentration of nickel oxide. Our observations are in agreement with prior studies on similar glass systems, where the introduction of transition metal oxides like NiO has been shown to enhance ionic conductivity due to the creation of non-bridging oxygen ions, which facilitate the movement of charge carriers. Therefore, an ionic conducting glass may be useful in solid electrolytes based on the studied glass system.

Author Contributions

L.V.: Conceptualization, writing—review and editing. S.M.S.: Writing, data curation. R.V.: Calculation and data procuring. K.P.: Formal analysis, editing. P.R.B.: Data analysis, editing. S.-J.K.: Formal analysis, conceptual discussion. G.N.R.: Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the 2024 Yeungnam University Research Grant.

Data Availability Statement

The data will be available on reasonable request from the corresponding author due to privacy.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rada, S.; Culea, M.; Neumann, M.; Culea, E. Structural role of europium ions in lead–borate glasses inferred from spectroscopic and DFT studies. Chem. Phys. Lett. 2008, 460, 196. [Google Scholar] [CrossRef]
  2. Pisarski, W.; Goryczka, T.; Wodecka-Dus, B.; Plonska, M.; Pisarska, J. Structure and properties of rare earth-doped lead borate glasses. Mater. Sci. Eng. B 2005, 122, 94. [Google Scholar] [CrossRef]
  3. El-Egili, K.; Doweidar, H.; Moustafa, Y.M.; Abbas, I. Structure and Some Physical Properties of PbO–P2O5 Glasses. Phys. B 2003, 339, 237. [Google Scholar] [CrossRef]
  4. Marcel, P. Advanced glasses. Ann. Chim. Sci. Matter 2003, 28, 87. [Google Scholar]
  5. Alderman, O.L.; Hannon, A.C.; Holland, D.; Umesaki, N. On the germanium–oxygen coordination number in lead germanate glasses. J. Non-Cryst. Solids 2014, 386, 56. [Google Scholar] [CrossRef]
  6. Nouadji, M.; Ivanova, Z.G.; Poulain, M.; Zavadil, J.; Attaf, A. Glass formation, physicochemical characterization and photoluminescence properties of new Sb2O3–PbO–ZnO and Sb2O3–PbO–ZnS systems. J. Alloy Compd. 2013, 549, 158. [Google Scholar] [CrossRef]
  7. Kashyap, R. Photosensitive Optical Fibers: Devices and Applications. Opt. Fiber Technol. 1994, 1, 17. [Google Scholar] [CrossRef]
  8. Zmojda, J.; Kochanowicz, M.; Miluski, P.; Golonko, P.; Baranowska, A.; Ragi, T.; Dorosz, J.; Kuwik, M.; Pisarski, W.; Pisarska, J.; et al. Luminescent Studies on Germanate Glasses Doped with Europium Ions for Photonic Applications. Materials 2020, 13, 2817. [Google Scholar] [CrossRef] [PubMed]
  9. Kuwik, M.; Kowalska, K.; Pisarska, J.; Pisarki, W.A. Spectroscopic properties of Pr3+, Tm3+, and Ho3+ in Germenate based Glass Systems modified by TiO2. Materials 2023, 16, 61. [Google Scholar] [CrossRef]
  10. Zelniok, D.; Cramer, C.; Eckert, H. Structure/Property Correlations in Ion-Conducting Mixed-Network Former Glasses:  Solid-State NMR Studies of the System Na2O−B2O3−P2O5. Chem. Mater. 2007, 19, 3162. [Google Scholar] [CrossRef]
  11. Iezid, M.; Legouera, M.; Goumeidane, F.; Poulain, M.; Nazabal, V.; Lebullenger, R. Glass formation in the Sb2O3–CdCl2–SrCl2 ternary system. J. Non-Cryst. Solids 2011, 357, 2984. [Google Scholar] [CrossRef]
  12. Loiko, P.; Dymshits, O.; Zhilin, A.; Alekseeva, I.; Yumashev, K. Influence of NiO on phase transformations and optical properties of ZnO–Al2O3–SiO2 glass–ceramics nucleated by TiO2 and ZrO2. Part II. Optical absorption and luminescence. J. Non. Cryst. Solids 2013, 376, 99. [Google Scholar] [CrossRef]
  13. Greaves, G.N.; Sen, S. Inorganic glasses, glass-forming liquids and amorphizing solids. Adv. Phys. 2007, 56, 1. [Google Scholar] [CrossRef]
  14. Mysen, B.O.; Richet, P. Silicate Glasses and Melts: Properties and Structure; Elsevier: Amsterdam, The Netherlands, 2005. [Google Scholar]
  15. Keppler, H.; Bagdassarov, N. The speciation of Ni and Co in silicate melts from optical absorption spectra to 1500 C. Chem. Geol. 1999, 158, 105. [Google Scholar] [CrossRef]
  16. Reddy, M.S.; Krishna, G.M.; Veeraiah, N. Spectroscopic and magnetic studies of manganese ions in ZnO–Sb2O3–B2O3 glass system. J. Phys. Chem. Solids 2006, 67, 789. [Google Scholar] [CrossRef]
  17. Nassau, K. The material dispersion zero in infrared optical waveguide materials. Bell Syst. Technol. J. 1981, 60, 327. [Google Scholar] [CrossRef]
  18. Mc Cubbin, T.K., Jr.; Atmosoekarto, S.; Withstandley, V. Infrared Spectroscopy with a Copper-Doped Germanium Detector. Appl. Opt. 1967, 6, 1131. [Google Scholar] [CrossRef]
  19. SrinivasaReddy, M.; Prasad, S.V.G.V.A.; Veeraiah, N. Valence and coordination of chromium ions in ZnO–Sb2O3–B2O3 glass system by means of spectroscopic and dielectric relaxation studies. Phys. Status Solidi A 2007, 204, 816. [Google Scholar] [CrossRef]
  20. Tagiara, N.S.; Chatzipanagis, K.I.; Bradtmuller, H.; Rodrigues, A.C.M.; Moncke, D.; Kamitos, E.I. Network former mixing effects in alkali germanotellurite glasses: A vibrational spectroscopic study. J. Alloys Compd. 2021, 882, 160782. [Google Scholar] [CrossRef]
  21. Rada, S.; Erhan, R.V.; Bodnarchuk, V.; Barbu-Tudoran, L.; Culea, E. SANS, RAMAN and SEM studies of lead-germanate glasses doped with the manganese oxide. J. Alloys Compd. 2021, 882, 160721. [Google Scholar] [CrossRef]
  22. Mattos, G.R.S.; Bordon, C.D.S.; Kassab, L.R.P.; Issa, S.A.; ALMisned, G.; Tekin, H.O. Towards obtaining the optimum physical, optical, and nuclear radiation attenuation behaviours of tellurite-germanate glasses through Eu2O3 reinforcement: Glass synthesis, experimental and theoretical characterization study. Ceram. Int. 2023, 49, 986. [Google Scholar] [CrossRef]
  23. Pipes, R.S.; Shelby, J.E. Formation and properties of soda lime germanate glasses. J. Non-Cryst. Solids 2021, 553, 120506. [Google Scholar] [CrossRef]
  24. Falci, R.F.; Guerineau, T.; Delarosbil, J.L.; Messaddeq, Y. Spectroscopic properties of gallium-rich germane gallate glasses doped with Tm3+. J. Lumin. 2022, 249, 119014. [Google Scholar] [CrossRef]
  25. Jianga, X.P.; Yang, Z.M.; Liu, S.H. Energy transfer between Yb3+ and Er3+ in barium gallogermanate glass. J. Appl. Phys. 2009, 105, 103113. [Google Scholar] [CrossRef]
  26. Pisarska, J.; Sołtys, M.; Górny, A.; Kochanowicz, M.; Zmojda, J.; Dorosz, J.; Dorosz, D.; Sitarz, M.; Pisarski, W.A. Rare earthdoped barium gallo-germanate glasses and their near-infrared luminescence properties. Spectrochim. Acta A 2018, 201, 362–366. [Google Scholar] [CrossRef] [PubMed]
  27. Saffarini, G.; Saiter, J.M.; Matthiesen, J. Thaermal stability and percolation threshold of Ge-Se-Fe glasses. Mater. Lett. 2007, 61, 432–436. [Google Scholar] [CrossRef]
  28. Gandhi, Y.; Mohan, N.K.; Veeraiah, N. Role of nickel ion coordination on spectroscopic and dielectric properties of ZnF2–As2O3–TeO2: NiO glass system. J. Non-Cryst. Solids 2011, 357, 1193. [Google Scholar] [CrossRef]
  29. Kumar, G.R.; Rao, M.K.; Srikumar, T.; Rao, M.C.; Kumar, V.R.; Veeraiah, N.; Rao, C.S. Spectroscopic, dielectric dispersion and dc conductivity studies of Sb2O3 doped lithium fluoroborophosphate glasses mixed with small concentrations of NiO. J. Alloys Compd. 2018, 752, 179. [Google Scholar] [CrossRef]
  30. Teja, P.V.; Rajyasree, C.; Krishna, S.M.; Tirupataiah, C.; Rao, D.K. Effect of some VA group modifiers on R2O3 (R = Sb, Bi)–ZnF2–GeO2 glasses doped with CuO by means of spectroscopic and dielectric investigations. Mater. Chem. Phys. 2012, 133, 239. [Google Scholar] [CrossRef]
  31. Beattie, I.R.; Livingston, K.M.S.; Ozin, G.A.; Renolds, D.J. Single-crystal Raman spectra of arsenolite (As4O6) and senarmonite (Sb4O6). The gas-phase Raman spectra of P4O6, P4O10, and As4O6. J. Chem. Soc. A 1970, 449–451. [Google Scholar] [CrossRef]
  32. Galeener, F.L.; Geissberger, A.E.; Ogar, G.W.; Loehman, R.E. Vibrational dynamics in isotopically substituted vitreous GeO2. Phys. Rev. B 1983, 28, 4768. [Google Scholar] [CrossRef]
  33. Lee, J.D. Concise Inorganic Chemistry, 5th ed.; Blackwell Science Ltd.: Hoboken, NJ, USA, 1996. [Google Scholar]
  34. Brown, A.; Gupta, V. Coordination environments of Ni2+ in germinate glass matrices. Opt. Mater. 2018, 36, 1262–1266. [Google Scholar]
  35. Suresh, B.; Reddy, M.S.; Reddy, A.S.S.; Gandhi, Y.; Kumar, V.R.; Veeraiah, N. Spectroscopic features of Ni2+ ion in PbO–Bi2O3–SiO2 glass system. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 141, 263. [Google Scholar] [CrossRef] [PubMed]
  36. Gao, G.; Reibstein, S.; Spiecker, E.; Peng, M.; Wondraczek, L. Broadband NIR photoluminescence from Ni2+-doped nanocrystalline Ba–Al titanate glass ceramics. J. Mater. Chem. 2012, 22, 2582. [Google Scholar] [CrossRef]
  37. Austin, I.G.; Mott, N.F. Polarons in crystalline and non-crystalline materials. Adv. Phys. 1969, 18, 657. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of PbO–GeO2: NiO glasses.
Figure 1. XRD patterns of PbO–GeO2: NiO glasses.
Inorganics 12 00215 g001
Figure 2. DSC patterns of PbO–GeO2: NiO glasses. The inset shows the variation of Tg and Kgl with the concentration of NiO.
Figure 2. DSC patterns of PbO–GeO2: NiO glasses. The inset shows the variation of Tg and Kgl with the concentration of NiO.
Inorganics 12 00215 g002
Figure 3. Optical absorption spectra of PbO-GeO2 glasses doped with nickel oxide.
Figure 3. Optical absorption spectra of PbO-GeO2 glasses doped with nickel oxide.
Inorganics 12 00215 g003
Figure 4. Tauc plots of PbO–GeO2: NiO glasses. The inset represents the variation of the optical band gap with the concentration of NiO.
Figure 4. Tauc plots of PbO–GeO2: NiO glasses. The inset represents the variation of the optical band gap with the concentration of NiO.
Inorganics 12 00215 g004
Figure 5. FTIR spectra of PbO–GeO2 glasses doped with nickel oxide.
Figure 5. FTIR spectra of PbO–GeO2 glasses doped with nickel oxide.
Inorganics 12 00215 g005
Figure 6. Raman spectra of PbO–GeO2 glasses doped with nickel oxide.
Figure 6. Raman spectra of PbO–GeO2 glasses doped with nickel oxide.
Inorganics 12 00215 g006
Figure 7. Photoluminescence of PbO–GeO2 glasses doped with nickel oxide recorded at room temperature (λexc = 800 nm).
Figure 7. Photoluminescence of PbO–GeO2 glasses doped with nickel oxide recorded at room temperature (λexc = 800 nm).
Inorganics 12 00215 g007
Figure 8. The dispersion of the dielectric constant, ε′, vs. temperature at 100 kHz for the glasses PbO-GeO2 doped with different concentrations of nickel oxide.
Figure 8. The dispersion of the dielectric constant, ε′, vs. temperature at 100 kHz for the glasses PbO-GeO2 doped with different concentrations of nickel oxide.
Inorganics 12 00215 g008
Figure 9. The variation of dielectric loss with temperature at 10 kHz frequency for the glasses PbO-GeO2 doped with different concentrations of NiO.
Figure 9. The variation of dielectric loss with temperature at 10 kHz frequency for the glasses PbO-GeO2 doped with different concentrations of NiO.
Inorganics 12 00215 g009
Figure 10. The variation of ac conductivity with 1/T at 100 kHz of PbO–GeO2 glasses doped with nickel oxide. The inset represents the variation of ac conductivity regarding the activation energy.
Figure 10. The variation of ac conductivity with 1/T at 100 kHz of PbO–GeO2 glasses doped with nickel oxide. The inset represents the variation of ac conductivity regarding the activation energy.
Inorganics 12 00215 g010
Figure 11. Dependence of 4πNµ2/27 K on concentration of nickel oxide.
Figure 11. Dependence of 4πNµ2/27 K on concentration of nickel oxide.
Inorganics 12 00215 g011
Table 1. Summary of the data physical parameters of the PbO–GeO2: NiO glasses.
Table 1. Summary of the data physical parameters of the PbO–GeO2: NiO glasses.
Glass SampleDensity
d (g/cm3)
Molar Volume
Vm (cm3)
Nickel Ion Conc.
Ni (×1021/cm3)
Inter-Ionic Distance of Nickel Ions
Ri (Å)
Polaron Radius
Rp (Å)
N05.560.520------
N26.0350.8864.785.932.39
N46.0540.5869.604.701.89
N66.0720.95314.444.101.654
N86.081.00819.293.721.50
N106.150.90824.413.441.38
Table 2. DSC data of PbO-GeO2 glasses doped with different concentrations of NiO.
Table 2. DSC data of PbO-GeO2 glasses doped with different concentrations of NiO.
GlassTg
(°C)
Tc
(°C)
Tm
(°C)
Tc − Tg
(°C)
Kgl =
(Tc − Tg)/(Tm − Tc)
N04976277621300.963
N24996257631260.913
N45036237661200.834
N65056227691170.796
N85066197711130.743
N105046157721110.707
Table 3. Summary of the data related to optical absorption spectra of the PbO–GeO2: NiO glasses.
Table 3. Summary of the data related to optical absorption spectra of the PbO–GeO2: NiO glasses.
Band Position
(nm)
N2N4N6N8N10
Ni2+ band positions, octahedral transitions (nm)
3A2(F) → 3T2(F)
12541258126212651269
3A2(F) → 3T1(F)769771774781784
3T2(F) → 1T2(D)482484490495497
3A2(F) → 3T1(P)403408415421425
Ni2+ band positions, tetrahedral transitions (nm)
3A2(F) → 3A2(F)
10401038103410291024
3A2(F) → 3T1(F)635631628623620
Optical
band gap Eo (eV)
2.582.512.362.322.27
Table 4. Summary of the band positions (cm−1) of various structural units in FT-IR spectra of PbO–GeO2: NiO glasses.
Table 4. Summary of the band positions (cm−1) of various structural units in FT-IR spectra of PbO–GeO2: NiO glasses.
GlassGeO4 UnitsGeO6 UnitsGe–O–Ge Bending VibrationsPbO4 Units
N01088783622445
N21085785625442
N41080787627439
N61077791631437
N81075792633436
N101072794634435
Table 5. Summary of the data on band positions (cm−1) of Raman Spectra of PbO–GeO2: NiO glasses.
Table 5. Summary of the data on band positions (cm−1) of Raman Spectra of PbO–GeO2: NiO glasses.
GlassPbO4GeO4 (Q4)
Units
Geo4 (Q3)
Units
GeO6
Units
GeO
Units
GeO4 (Q2)
Units
GeO4 (Q1)
Units
N0282433524621750901959
N2284436529620750901959
N4286440530619750901959
N6292443532618750901959
N8294446534618750901959
N10301448537618750901959
Table 6. Summary of data on photoluminescence of PbO–GeO2: NiO glasses.
Table 6. Summary of data on photoluminescence of PbO–GeO2: NiO glasses.
SampleEmission Peak Position (nm)Refractive IndexσpE (1033, cm2)
N213031.5420.932
N413061.5450.934
N613091.5490.939
N813131.5520.943
N1013161.5560.945
Table 7. Data on dielectric parameters of PbO-GeO2: NiO glasses.
Table 7. Data on dielectric parameters of PbO-GeO2: NiO glasses.
GlassA.E for Conduction (eV)σac (10−7) (Ω cm)−1
N00.430.56
N20.390.77
N40.361.08
N60.331.59
N80.292.54
N100.254.22
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vijayalakshmi, L.; Meera Saheb, S.; Vijay, R.; Palle, K.; Ramesh Babu, P.; Kwon, S.-J.; Raju, G.N. Nickel Ions Activated PbO–GeO2 Glasses for the Application of Electrolytes and Photonic Devices. Inorganics 2024, 12, 215. https://doi.org/10.3390/inorganics12080215

AMA Style

Vijayalakshmi L, Meera Saheb S, Vijay R, Palle K, Ramesh Babu P, Kwon S-J, Raju GN. Nickel Ions Activated PbO–GeO2 Glasses for the Application of Electrolytes and Photonic Devices. Inorganics. 2024; 12(8):215. https://doi.org/10.3390/inorganics12080215

Chicago/Turabian Style

Vijayalakshmi, L., Shaik Meera Saheb, R. Vijay, Kishor Palle, P. Ramesh Babu, Seong-Jin Kwon, and G. Naga Raju. 2024. "Nickel Ions Activated PbO–GeO2 Glasses for the Application of Electrolytes and Photonic Devices" Inorganics 12, no. 8: 215. https://doi.org/10.3390/inorganics12080215

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop