Next Article in Journal
Preparation and Elastic Moduli of Germanate Glass Containing Lead and Bismuth
Previous Article in Journal
Characterization of High-Value Bioactives in Some Selected Varieties of Pakistani Rice (Oryza sativa L.)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optical Properties of Bismuth Tellurite Based Glass

by
Hooi Ming Oo
,
Halimah Mohamed-Kamari
* and
Wan Mohd Daud Wan-Yusoff
Physics Department, Faculty of Science, University Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2012, 13(4), 4623-4631; https://doi.org/10.3390/ijms13044623
Submission received: 16 February 2012 / Revised: 26 March 2012 / Accepted: 26 March 2012 / Published: 11 April 2012

Abstract

:
A series of binary tellurite based glasses (Bi2O3)x (TeO2)100−x was prepared by melt quenching method. The density, molar volume and refractive index increase when bismuth ions Bi3+ increase, this is due to the increased polarization of the ions Bi3+ and the enhanced formation of non-bridging oxygen (NBO). The Fourier transform infrared spectroscopy (FTIR) results show the bonding of the glass sample and the optical band gap, Eopt decreases while the refractive index increases when the ion Bi3+ content increases.

1. Introduction

Tellurite glass is a good material used in lasers and nonlinear applications, photonic applications and communication applications [1]. Tellurite glass is often used because it is stable at room temperature, has good thermal, optical and electric properties [2,3]. This glass has a low photon energy, a high linear and non-linear refractive index, and it also can be used in photorefractive materials, nonlinear devices, up-conversion lasers and optical amplifiers [46]. Metal oxides such as Bi2O3, Sb2O3, MoO3 and Nb2O5 have been added to the tellurite glass system in order to enhance the optical behavior [7,8]. Many studies on tertiary tellurite based glass have been done by various authors [916]. For example, Tiefeng et al. (2011) [14] studied the optical non linear properties of TeO2–Bi2O3–BaO glass, and was found that an increase of the ion bismuth content and a decrease of ion barium content increased the non linear properties of the glass as bismuth dissolved in the tellurite glass matrix.
Bismuth (iii) oxide is used in optical fiber and ceramic materials because it can provide a high refractive index and low temperatures [17]. In addition, it is also used in the electronic field due to it having a high valence cation of low field strength and high polarizability [18]. In this paper, the optical and physical properties were determined with different ion bismuth contents of tellurite base glass. Optical properties such as the optical band gap and refractive index of binary bismuth tellurite glass were studied. In addition, the density and molar volume of a tellurite glass system was determined.

2. Results and Discussion

2.1. XRD

The XRD results (Figure 1) show that the bismuth tellurite glass is in an amorphous phase. Sample TB 5 (refer to Table 1) is a transparent glass of a dark orange color, while TB 8, TB10, TB12 and TB15 are transparent glasses of a yellow color.

2.2. Density and Molar Volume

The density of the glass samples increases as the bismuth content increases. The densities of the bismuth tellurite glasses range from 5.43 g/cm3 to 6.26 g/cm3, as shown in Table 1. Figure 2 shows the molar volume and the density increase proportional to the bismuth content. The molar mass of the bismuth (iii) oxide (465.96 g/mol) is heavier than the molar mass of tellurite oxide (159.60 g/mol) and hence, the glass matrix becomes more dense when Bi3+ ions are added into the glass network [19]. In addition, the increase in molar volume is due to the atomic radius of Bi3+ (1.70 Å ) which is higher than that of tellurite (1.60 Å ). Usually, the density of the glass changes in the inverse direction of the molar volume, but in this study, the density and molar volume increase with the bismuth contents, this anomaly was also found by Saddek et al. (2007) and Halimah et al. (2010) [20,21].

2.3. FTIR Results

As shown in Figure 3, the FTIR analysis shows that the TeO2 is the framework former while Bi2O3 is the glass modifier in the bismuth tellurite glasses system. Table 2 shows the ranges of the wave numbers for difference mode. The wave number for the bismuth tellurite glass system is slightly shifted to the lower wave numbers when the ion Bi3+ concentration is increased. There is a diffuse band in the range of 500–600 cm−1, due to the disordered structure; this is considered as the vibration modes of both the TeO3 and the TeO4 entities [22]. The vibration modes of the bonds for TeO3+1 polyhedra is also found at bands around 580 cm−1 where the TeO3+1 polyhedra consists of an intermediate coordination of the tellurium atoms between 3 and 4 [17], this is found in the all glass samples. On the other hand, the shoulder form in the range of 800–900 cm−1 in TB8, TB10, TB12 and TB15 is due to the symmetrical stretching vibration of the Bi–O bonds in BiO3 units. This band is not observed in the TB5 because the Bi3+ ions present in this glass sample contain non-bridging oxygen in the form of [BiO] defects. As the ion Bi3+ content increases, bismuth oxide as the network modifier ion modifies the glass structure and enhances the breaking of axial Te–O–Te linkages in the trigonal bipyramids [TeO4] (tbp) and creates the formation of trigonal pyramid [TeO3] (tp) units and non-bridging oxygen.

2.4. Refractive Index and Optical Absorption

When the bismuth content increases, the refractive index measured at the wavelength of 632.8 nm increases from 1.97 to 2.12. The refractive index increase is due to the ion Bi3+ which has a high polarity that can break the bridging oxygen [BO] to non-bridging oxygen [NBO]. In addition, non-bridging oxygen also has an effect on the refractive index, because of the polarity of non-bridging oxygen (NBO) is higher than that of bridging oxygen [29]. As a result, Bi2O3 will affect the TeO2 based glasses and forms highly polarized non-bridging oxygen ions at the terminals of Te–O bonds.
The optical absorption spectra were taken in the ranges of 350 to 500 nm. For the optical band gap in this bismuth tellurite glass system, the optical band gap energy, Eopt decreases when the ion bismuth content increases as shown in Table 3.
The optical band gap energy is determined by using the following equation [30]
α h v = A ( h v - E opt ) n
where α is the absorption coefficient, hv is the incident photon energy, A is a constant and Eopt is the optical band gap. Values of n are 2 and 1/2 for direct and indirect transitions, respectively. The indirect band gap as a function of photon energy for (Bi2O3)x(TeO2)100−x glass system is plotted and shown in Figure 4.
According to Sayed (2005), amorphous or glassy materials consist of a band tailing in the forbidden energy band gap. The band tailing might arise from random fluctuations of the internal disorder in the amorphous materials. So it can be estimated using the Urbach equation shown as follows [31]
α ( v ) = B e x p ( h v / Δ E g )
where B is a constant and ΔE is the width of the band tail of the electron states. From Figure 5, Urbach energy, Eg is determined from the slope of plot ln (α) versus hv.
The range of the indirect optical band gap decreases from 2.40 eV to 2.60 eV and 3.00 eV to 3.02 eV for the direct band gap as the concentration of ion Bi3+ increases in the bismuth tellurite glass system. However, sample TB5 has a low bismuth content and a low energy band gap because of the formation of TeO2 linkages in Te–O where bridging oxygen is more frequent than non-bridging oxygen. Shifting of the absorption band to a lower energy can be related to the formation of non-bridging oxygen (NBO) which binds exited electrons of non-bridging oxygen less tightly than bridging oxygen [10], and this cause the glass network to become less rigid (Rajendan, 2003) [27]. Another possibility is due to the high polarizability of the Bi3+ ions which can be attributed to the empty d orbitals of the corresponding cations and also their high coordination number towards oxide ions. The coordination number of TeO2 changes from 4 to 3 while the coordination number of Bi2O3 changes from 3 to 6, which then forms a BiO6 octaherdal structural unit with the addition of Bi2O3 into the glass system. According to Yanfei (2008) [16], less tight oxygen anions were allowed when the amount of non-bridging oxygen increased, This was because of an increase in Lewis basicity of oxide ions, which form stronger covalent Te–O bonds in TeO3 units, and so allow less tight oxygen anions in the glass network. Hence the more the Bi2O3 content increases, the fewer tightly bound oxygen anions (valence electrons) are found and the greater the decrease of the optical band gap energies.
The optical band gap energy is opposite to the Urbach energy, which can be attributed to the formation of non-bridging oxygen (Figure 6). The optical band gap energy and Urbach energy values for glass systems are found to lie between 3.10–3.00 eV and 0.25–0.44 eV, respectively. Rajendran (2003) [27] reported that the addition of Bi2O3 glass modifier would make the glass network more loosely packed (breaking the O–Te–O chains to form TeO3 units and non-bridging oxygen) and therefore produce unstable glass. When the concentration of ion Bi3+ is increased, the appearance of TeO3 units and non-bridging oxygen in the glass networks increases. As the concentration of the ion modifier Bi3+ increases, non-bridging oxygen and TeO3 units in the glass structure increase and as a result, the Urbach energy for this glass system also increases.

3. Experimental

The raw materials of tellurium dioxide, TeO2 (Aldrich 99.5%), and bismuth (iii) oxide Bi2O3 (Aldrich 99.0%) were used to synthesize the glass sample. The glass sample was prepared by using a melt quenching method [32,33]. The bismuth (iii) oxide, x = 5, 8, 10, 12, 15 in mol% was added into the tellurite oxide and was weighed using a digital weighing machine with an accuracy of ±0.01 g and then mixed together by using a mortar and pestle. The mixture was poured into an alumina crucible and put into an electric furnace set at 100 °C for a period of 30 min. The alumina crucible was used because it can withstand high temperatures and does not react with the raw materials during the melting process as opposed to porcelain crucibles. The temperature was then increased by 10 °C/min until 900 °C and the then maintained for 2 h. When the melting process was completed, the molten liquid was cast into a stainless steel cylindrical shape mould which had been preheated at 400 °C for 30 min. The sample was annealed at 400 °C in a second furnace for 1 h, then the furnace was turned off. The glass sample was cut at a thickness of about 2 mm for the required measurements [33].
The density of the glass samples was measured using the Archimedes principle, and distilled water was used as the immersion liquid. The optical properties such as refractive indices (n) were determined by using EL X-02C high precision ellipsometer. The structure of the bismuth tellurite glass was investigated using X-ray diffraction (XRD) and Fourier transform infrared spectroscopy (FTIR). The optical absorption of the bismuth tellurite glass was measured using UV visible spectroscopy and powder BaSO4 was used as the reference for the UV visible spectroscopy. The wavelength was set from 2600 nm to 220 nm, and the internal spacing was 0.2 nm, using a medium scan speed for glass samples.

4. Conclusions

The density and molar volume of the glass sample increase due to the fact that the atomic mass of bismuth ions is higher than that of tellurite ions, and that the atomic radius of bismuth is also greater than that of tellurite ions. Additionally, the refractive index increases due to the increase of polarity of the Bi3+ ion content in tellurite based glasses. In addition, based on the FTIR results, the TeO2 is the glass former while Bi2O3 is the glass modifier in bismuth tellurite glass systems, and the wave number shifts to a low frequency when the bismuth content increases. The optical band gap shifts to a low energy while the Urbach energy shifts to high energy when non-bridging oxygen (NBO) increases as bismuth content increases.

Acknowledgments

The authors appreciate the financial support for the work from the Ministry of Higher Education of Malaysia through RUGS (9199837).

References

  1. Vijaya Prakash, G.; Narayana Rao, D.; Bhatnagar, A.K. Linear optical properties of niobium based tellurite glasses. Solid State Commun 2001, 119, 39–44. [Google Scholar]
  2. Mohamed, N.B.; Yahya, A.K.; Deni, M.S.M.; Mohamed, S.N.; Halimah, M.K.; Sidek, H.A.A. Effects of concurrent TeO2 reduction and ZnO addition on elastic and structural properties of (90–x)TeO2–10Nb2O5–(x)ZnO glass. J. Non-Cryst. Solids 2010, 356, 1626–1630. [Google Scholar]
  3. Foldvari, I.; Peter, A. Basic spectroscopic properties of bismuth tellurium oxide, Bi2TeO5. J. Mater. Sci 1992, 27, 750–754. [Google Scholar]
  4. Udovic, M.; Valant, M.; Suvorov, D. Phase formation and dielectric characterization of the Bi2O3–TeO2 system prepared in an oxygen atmosphere. J. Am. Ceram. Soc 2004, 87, 591–597. [Google Scholar]
  5. Ozdanova, J.; Ticha, H.; Tichy, L. Optical band gap and Raman spectra in some (Bi2O3)x(WO3)y(TeO2)100−xy and (PbO)x(WO3)y(TeO2)100−xy glasses. J. Non-Cryst. Solids 2009, 355, 2318–2322. [Google Scholar]
  6. Boyer, J.C.; Vetrone, F.; Capobianco, J.A. Optical transitions and upconversion properties of Ho3+ doped ZnO–TeO2 glass. J. Appl. Phys 2003, 93, 9460–9465. [Google Scholar]
  7. Mori, H.; Kitami, T.; Sakata, H. Electrical conductivity of W2O5–Sb2O3–TeO2 glasses. J. Non-Cryst. Solids 1994, 168, 157–166. [Google Scholar]
  8. Szu, S.P.; Chang, F.-S. Impedance study of V2O5–TeO2–BaO glasses. Solid State Ion 2005, 176, 2695–2699. [Google Scholar]
  9. Ayuni, J.N.; Halimah, M.K.; Talib, Z.A.; Sidek, H.A.A.; Daud, W.M.; Zaidan, A.W.; Khamirul, A.M. Optical properties of ternary TeO2–B2O3–ZnO glass system. Mater. Sci. Eng 2011, 17. [Google Scholar] [CrossRef]
  10. Mohamed, N.B.; Yahya, A.K.; Deni, M.S.M.; Mohamed, S.N.; Halimah, M.K.; Sidek, H.A.A. Effects of concurrent TeO2 reduction and ZnO addition on elastic and structural Properties of (90–x)TeO2–10Nb2O5–(x)ZnO glass. J. Non-Cryst. Solids 2010, 356, 1626–1630. [Google Scholar]
  11. Boris, L.S.; Petrov, T.; Yoneda, H.; Rositsa, T.; Boriana, M. Synthesis and nonlinear optical properties of TeO2–Bi2O3–GeO2 glasses. Scr. Mater 2009, 61, 493–496. [Google Scholar]
  12. Durga, D.K.; Veeraiah, N. Dielectric dispersion in ZnF2–Bi2O3–TeO2 glass system. J. Mater. Sci 2001, 36, 5625–5632. [Google Scholar]
  13. Yousef, E.S.S. Characterization of oxyfluoride tellurite glasses through thermal, optical and ultrasonic measurements. J. Phys. D Appl. Phys 2005, 38, 3970–3975. [Google Scholar]
  14. Xu, T.; Chen, F.; Dai, S.; Shen, X.; Wang, X.; Nie, Q.; Liu, C. Glass formation and third-order optical nonlinear properties within TeO2–Bi2O3–BaO pseudo-ternary system. J. Non-Cryst. Solids 2011, 357, 2219–2222. [Google Scholar]
  15. Senthil Murugan, G.; Fargin, E.; Rodriguez, V.; Adamietz, F.; Couzi, M.; Buffeteau, T.; Le Coustumer, P. Temperature-assisted electrical poling of TeO2–Bi2O3–ZnO glasses for non-linear optical application. J. Non-Cryst. Solids 2004, 344, 158–166. [Google Scholar]
  16. Wang, Y.; Dai, S.; Chen, F.; Xu, T.; Nie, Q. Physical properties and optical band gap of new tellurite glasses within the TeO2–Nb2O5–Bi2O3 system. Mater. Chem. Phys 2009, 113, 407–411. [Google Scholar]
  17. El-Mallawany, R.A.H. Tellurite Glasses Handbook; CRC Press Inc: Boca Raton, FL, USA, 2002. [Google Scholar]
  18. El-Mallawanny, R.; Dirar Abdalla, M.; Abbas Ahmed, I. New tellurite glass: Optical properties. Mater. Chem. Phys 2008, 109, 291–296. [Google Scholar]
  19. Sidek, H.A.A.; Rosmawati, S.; Talib, Z.A.; Halimah, M.K.; Daud, W.M. Synthesis and optical properties of ZnO–TeO2 glass system. Am. J. Appl. Sci 2009, 6, 1489–1494. [Google Scholar]
  20. Saddeek, Y.B.; Yasser, B.; Afifi, H.A.; Abd El-Aal, N.S. Interpretation of mechanical properties and structure of TeO2–Li2O–B2O3 glasses. Phys. B 2007, 398, 1–7. [Google Scholar]
  21. Halimah, M.K.; Sidek, H.A.A.; Daud, W.M.; Zainal, A.S.; Mansor, H.; Khamirul, A.M. Physical properties of binary tellurite glass system. Solid State Sci. Tech 2010, 18, 364–370. [Google Scholar]
  22. Munoz-Matin, D.; Villegas, M.A.; Gonzalo, J.; Fernandez-Navarro, J.M. Characterisation of glasses in the TeO2–WO3–PbO system. J. Eur. Ceram. Soc 2009, 29, 2903–2913. [Google Scholar]
  23. Doweidar, H.; Saddeek, Y.B. FTIR and Ultrasonic investigations on modified bismuth borate glasses. J. Non-Cryst. Solids 2009, 355, 348–354. [Google Scholar]
  24. Wanga, G.; Zhanga, J.; Daia, S.; Wena, L.; Yanga, J.; Jianga, Z. Structural investigation on TeO2–BiCl3 glassy system. J. Mol. Struct 2005, 750, 1–6. [Google Scholar]
  25. Ok, K.M.; Bhuvanesh, N.S.P.; Shiv Halasyamani, P. Bi2TeO5: Synthesis, structure, and powder second harmonic generation properties. Inorg. Chem 2001, 40, 1978–1980. [Google Scholar]
  26. Pascuta, P.; Culea, E. FTIR spectroscopic study of some bismuth germinate glasses containing gadolinium ions. Mater. Lett 2008, 62, 4127–4129. [Google Scholar]
  27. Rajendran, V.; Palanivelu, N.; Chaudhuri, B.K.; Goswami, K. Characterisation of semiconducting V2O5–Bi2O3–TeO2 glasses through ultrasonic measurements. J. Non-Cryst. Solids 2003, 320, 195–209. [Google Scholar]
  28. Halimah, M.K.; Daud, W.M.; Sidek, H.A.A.; Zaidan, A.W.; Zainal, A.S. Optical properties of ternary tellurite glasses. Mater. Sci. Pol 2010, 28, 173–180. [Google Scholar]
  29. Chen, Y.; Nie, Q.; Xu, T.; Dai, S.; Wang, X.; Shen, X. A study of nonlinear optical properties in Bi2O3–WO3–TeO2 glasses. J. Non-Cryst. Solids 2008, 354, 3468–3472. [Google Scholar]
  30. Tauc, J. Amorphous and Liqiud Semiconductors; Plenum: New York, NY, USA, 1974. [Google Scholar]
  31. Urbach, F. The long-wavelenght edge of photographic sensitivity and of the electronic absorption of solids. Phys Rev 1953, 92. [Google Scholar] [CrossRef]
  32. Doweidar, H.; Yasser, B.S. FTIR and ultrasonic investigations on modified bismuth borate glasses. J. Non-Cryst. Solids 2009, 355, 348–354. [Google Scholar]
  33. Sidek, H.A.A.; Rosmawati, S.; Talib, Z.A.; Halimah, M.K.; Halim, S.A. Effect of zinc on the elastic behaviour of (TeO2)90(AlF3)10−x(ZnO)x glass system. Int. J. Basic Appl. Sci 2010, 9, 401–407. [Google Scholar]
Figure 1. X-ray diffraction (XRD) analysis of the (Bi2O3)x(TeO2)100−x glass system.
Figure 1. X-ray diffraction (XRD) analysis of the (Bi2O3)x(TeO2)100−x glass system.
Ijms 13 04623f1
Figure 2. Density and molar volume with mole fraction of bismuth (iii) oxide of the (Bi2O3)x(TeO2)100−x glass system.
Figure 2. Density and molar volume with mole fraction of bismuth (iii) oxide of the (Bi2O3)x(TeO2)100−x glass system.
Ijms 13 04623f2
Figure 3. FTIR analysis of (Bi2O3)x(TeO2)100−x glass system.
Figure 3. FTIR analysis of (Bi2O3)x(TeO2)100−x glass system.
Ijms 13 04623f3
Figure 4. Plot of (αhv)1/2 versus hv for indirect band gap of (Bi2O3)x(TeO2)100x glass system.
Figure 4. Plot of (αhv)1/2 versus hv for indirect band gap of (Bi2O3)x(TeO2)100x glass system.
Ijms 13 04623f4
Figure 5. Plot of ln (α) versus hv for Urbach energy of (Bi2O3)x(TeO2)100x glass system.
Figure 5. Plot of ln (α) versus hv for Urbach energy of (Bi2O3)x(TeO2)100x glass system.
Ijms 13 04623f5
Figure 6. The optical band gap energy, Eopt and Urbach energy, Eg of (Bi2O3)x(TeO2)100−x glass system.
Figure 6. The optical band gap energy, Eopt and Urbach energy, Eg of (Bi2O3)x(TeO2)100−x glass system.
Ijms 13 04623f6
Table 1. Molar mass, density and molar volume of the (Bi2O3)x(TeO2)100−x glass system.
Table 1. Molar mass, density and molar volume of the (Bi2O3)x(TeO2)100−x glass system.
Samplesx (mol%)Molar mass (g/mol)Density (g/cm3)Molar Volume (cm3/mol)
TB55174.925.4332.22
TB88184.115.6932.35
TB1010190.235.8532.49
TB1212196.366.0332.56
TB1515205.556.2632.86
Table 2. The ranges of the wave numbers for difference mode.
Table 2. The ranges of the wave numbers for difference mode.
Wavenumber(cm−1)Mode
400–600Bi–O–Bi + Bi–O in BiO6 octahedral [23]
650–660Te–Oax in [TeO4] [24,25]
775Te–Oeq in [TeO4] [26,27]
633, 695(shoulders) Te-O [10]
600–610Bi–O stretch in BiO6 units [28]
860–865Bi–O vibration in distorted BiO6 units [11]
Table 3. Direct optical bandgap, indirect optical bandgap and Urbach energy of (Bi2O3)x(TeO2)100−x glass system.
Table 3. Direct optical bandgap, indirect optical bandgap and Urbach energy of (Bi2O3)x(TeO2)100−x glass system.
Sample glassIndirect bandgap, Eopt (eV)Direct bandgap (eV)Urbach energy, Eg (eV)Refractive index in wavelength 632.8 nm, n
TB 52.403.020.441.97
TB 82.773.100.221.99
TB 102.753.080.232.02
TB 122.703.040.242.07
TB 152.603.000.252.12

Share and Cite

MDPI and ACS Style

Oo, H.M.; Mohamed-Kamari, H.; Wan-Yusoff, W.M.D. Optical Properties of Bismuth Tellurite Based Glass. Int. J. Mol. Sci. 2012, 13, 4623-4631. https://doi.org/10.3390/ijms13044623

AMA Style

Oo HM, Mohamed-Kamari H, Wan-Yusoff WMD. Optical Properties of Bismuth Tellurite Based Glass. International Journal of Molecular Sciences. 2012; 13(4):4623-4631. https://doi.org/10.3390/ijms13044623

Chicago/Turabian Style

Oo, Hooi Ming, Halimah Mohamed-Kamari, and Wan Mohd Daud Wan-Yusoff. 2012. "Optical Properties of Bismuth Tellurite Based Glass" International Journal of Molecular Sciences 13, no. 4: 4623-4631. https://doi.org/10.3390/ijms13044623

Article Metrics

Back to TopTop