Next Article in Journal
Salivary Proteins Associated with Periodontitis in Patients with Type 2 Diabetes Mellitus
Previous Article in Journal
Optical Properties of Bismuth Tellurite Based Glass
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Elastic Moduli of Germanate Glass Containing Lead and Bismuth

Department of Physics, Faculty of Science, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2012, 13(4), 4632-4641; https://doi.org/10.3390/ijms13044632
Submission received: 31 January 2012 / Revised: 5 March 2012 / Accepted: 30 March 2012 / Published: 11 April 2012
(This article belongs to the Section Materials Science)

Abstract

:
This paper reports the rapid melt quenching technique preparation for the new family of bismuth-lead germanate glass (BPG) systems in the form of (GeO2)60–(PbO)40−x–(½Bi2O3)x where x = 0 to 40 mol%. Their densities with respect of Bi2O3 concentration were determined using Archimedes’ method with acetone as a floatation medium. The current experimental data are compared with those of bismuth lead borate (B2O3)20–(PbO)80−x–(Bi2O3)x. The elastic properties of BPG were studied using the ultrasonic pulse-echo technique where both longitudinal and transverse sound wave velocities have been measured in each glass samples at a frequency of 15 MHz and at room temperature. Experimental data shows that all the physical parameters of BPG including density and molar volume, both longitudinal and transverse velocities increase linearly with increasing of Bi2O3 content in the germanate glass network. Their elastic moduli such as longitudinal, shear and Young’s also increase linearly with addition of Bi2O3 but the bulk modulus did not. The Poisson’s ratio and fractal dimensionality are also found to vary linearly with the Bi2O3 concentration.

1. Introduction

Bismuth germanate-based glasses doped with rare-earth oxides gain much attention from researchers due to their potential applications in non-linear optics devices such as optical communication fibers, solid-state lasers, light converters, waveguide, sensors and scintillation detector in positron camera [13]. Some optical properties of bismuth in borate, phosphate and silicate glasses were also extensively studied [4].
The lead germanate glasses are also of growing interest, because their promising application including new lasing materials, upconverting phosphors and optical waveguides. All are due to their high density, refractive index, thermal expansion, mechanical strength, high chemical durability, temperature stability as well as with excellent transmission in the infrared (IR) region up to 4.5 micron [5].
Recent studies by Rabukhin and Belousova [6] found that the coordination numbers of cation present in the composition study range does not significantly affect the structural grouping character of the bismuth–containing gallate glasses. However, they noticed that the elastic constants of bismuth-containing borate and germinate are strongly dependent on the change of coordination number of B3+ and Ge4+.
The photoelastic constants of germanate glasses containing lead and bismuth oxides were studied by Rabukhin [7,8] where he suggested that the photoelastic and elastooptic data could be used for the light and acoustic lines of acoustooptical devices. He proposed that the high photoelastic constants of lead bismuth germanate glass can be achieved through high refractive index and low modulus of elasticity.
For lead borate glass, at low PbO content the formation of four-coordinated boron proceeds at the rate of two tetrahedral for each added oxygen. The lead enters the glass as modifiers Pb2+ ions. However, at 15–20 mol% PbO, lead enters the network in the form of PbO4 pyramids (with Pb at the apex of the pyramid). These PbO4 units bridge prefentially to BO3 rather than BO4 units [9].
Hamezan et al. [10] studied the elastic constants and thermal properties of lead bismuth borate glasses. They found the density and molar volume of such glasses increase with glass modifier content which attributed to the replacement of Bi2O3 and PbO; both have larger density and molar volume than B2O3 in the glass networks.
So far, the sound velocity in lead and bismuth containing germanate glass and their elastic constants have not been widely studied. In this research the glassy network of GeO2 will be added with PbO and Bi2O3 glass modifiers. Their elastic properties will be investigated by the ultrasonic pulse-echo technique [1114]. The present work reports the preparation and ultrasonic characterization of bismuth-lead germanate (BPG) glass systems.

2. Results and Discussion

Glass composition of the ternary bismuth lead germanate glasses (BPG) of the ternary form (GeO2)60–(PbO)40−x–(1/2Bi2O3) together with density, molar weight and molar volume are given in Table 1. The experimental data of bismuth lead borate (BPB) (B2O3)20–(PbO)80−x–(Bi2O3)x [14] is presented for comparison purposes.
It can be seen from Figure 1 that, in contrast to BPB, the density and molar volume of BPG glasses vary monotonically and slowly with the increase of Bi2O3 content. This change in density by the addition of Bi2O3 is much related to the change in the atomic mass and atomic of volume constituent elements as given in Table 1. The atomic mass of the Bi, Pb, Ge and B atoms is 208.98, 207.20, 72.63 and 10.811 and their atomic radii are 1.56, 1.75, 1.22 and 0.90 Å respectively. This explains the increase in density with increasing Bi2O3 content.
Figure 1 shows that densities of all the glass examined are increased with gradual substitution of PbO with Bi2O3. The ternary borate containing lead and bismuth (B) show higher density changes as compared with those of BPG glasses.
Both density and molar volume of the glasses increase with an increasing of Bi2O3. The density and molar volume increases by replacing PbO by Bi2O3 in Bi2O3–PbO–GeO2 glass system. It can be seen from Figure 1 and Figure 2 that the density of BPG glasses varies from 5.90 to 6.05 g cm−3 and the molar volume varies from 25.77 to 25.80 cm3 mol−1. Generally, the density and the molar volume show opposite behavior but, in this study, this was not the case. In this glass, substitution of lead by bismuth causes the expansion of network. Similar trends for densities and molar volumes have already been reported elsewhere for other glass systems [1518].
The molar volume of the ternary BPG glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x (A) is shown in Figure 2 and compared with those of (B2O3)20–(PbO)80−x–(Bi2O3)x (B). It is clear that by increasing Bi2O3, the molar volume increases which is similar with the variation density with increasing Bi2O3 content. For the ternary BPG glasses (A), the change is quite small as compared with the ternary BPB glasses and this might due to the compactness of the glass structure.
The Bi ions may enter the glass network interstitially, hence, some network bonds Ge–O–Ge or Pb–O–Ge are broken and replaced by ionic bonds between Bi ion and singly bonded oxygen atoms. So if one assumed that only effect of adding Bi cations was to break down the network bonds Ge–O–Ge and Pb–O–Ge then an increase in the molar volume with Bi2O3 content would be expected for the entire vitreous range of the studied glass system. Experimentally, this effect increases the molar volume in the glass compositional range from 0 to 40 mol% Bi2O3 content (see Table 1) and as a consequence the values of the density are increased.
The addition of Bi2O3 increased the values of density, which is probably attributable to simultaneous filling up of the vacancies amidst the network by the interstitial Bi ions with atomic mass 208.98. This increase in density indicates a structural change in the glass network which is accompanied by an increase in the molar volume [18,19].
The elastic moduli are proportional to the square of velocity and a plot of sound velocities vs. composition is indicative of relative structure. The compositional dependence of longitudinal (Vl) and transverse (Vs) sound velocities are depicted in Table 2 and Figure 3. Figure 4 shows the variation of elastic moduli as a function of Bi2O3 content. All the elastic moduli show the same trend as the ultrasonic wave velocities.
As can be seen from Figures 3,4, both Vl and Vs, and elastic moduli increase with the addition of Bi2O3 content over the entire composition studied. The results in Table 2 indicate that the elastic moduli increase linearly with Bi2O3 content. Longitudinal modulus changes from 69.10 GPa to 74.53 GPa, shear modulus from 21.57 GPa to 26.51 GPa, Young’s modulus from 54.93 GPa to 64.90 GPa and bulk modulus decrease from 40.34 GPa to 39.18 GPa as depicted in Figure 4. The increase in elastic moduli is due to an increase in the cross-link density [20] and therefore an increase in the rigidity of glass samples.
This variation of ultrasonic wave velocities and elastic moduli can be explained on the basis of the structural consideration of germanate glassy network. In the present glass system, their longitudinal ultrasonic velocity increases from 3422 to 3511 ms−1 and shear velocity increases from 1912 to 2094 ms−1 by increasing of bismuth content. According to Higazy and Bridge [21] the longitudinal strain changes directly with bond stretching force constant. On the other hand, as reported by Reisfeld et al. [4] the shear strain changes with bond bending force constant. The increase in velocities is attributed to the increase in rigidity of the glass network.
Glass is considered as elastic substance and, thus, can be characterized through a modulus of elasticity [11,12]. This modulus increases as the lengthening at a certain applied stress diminishes. That will be the case if the glass structure is rigid and therefore contains the fewest possible non bridging oxygen. When an oxide is introduced to germanate, the strength of the structure depends on the field strength of the cation. With increasing Bi2O3 content in the germanate glass, the structure becomes more rigid and so the density also increase and hence the modulus of elasticity increases [1719]. It may also be noted from Figure 4 that the rate of change of elastic moduli is more pronounced in longitudinal modulus (L) and least in case of transverse or shear modlulus (G). This indicates resistance to deformation and it is most probably due to presence of large number of covalent bonds.
Bulk modulus (K) is the elastic property of material that can be derived most easily from the glass structure [20]. GeO2 posses an open structure characterized by many open spaces. The addition of Bi2O3 will occupy such spaces and this should leads to an increase in bulk modulus. With low Bi2O3 contents, the bulk modulus is small since many open spaces are present which will be filled most quickly by the large Bi3+ ions (ion radius is 1.56 Å), so that the bulk modulus increases. With higher Bi2O3 contents, the ability for deformation of the cations becomes influential. This is clear from Figure 4, which shows the relation between bulk modulus and Bi2O3 content.
The compositional dependence of Poisson’s ratio and fractal dimensionality parameter as a function of Bi2O3 content is given in Figure 5. For BPG glasses their Poisson’s ratio, the ratio of transverse and linear strains for a linear stress, decrease from 0.27 to 0.22. Poisson’s ratio has also been discussed in terms of the dimensionality of glass network and it is observed that the Poisson’s ratio for a three dimensional network is less than that of a two dimensional structure, which in turn is less than that of a one dimensional structure. The decrease in Poisson’s ratio implies to the increase of crosslink density of the glass as proposed by Higazy and Bridge [21].
The fractal dimensionality parameter (d = 4G/B) listed in Table 2 suggest that in all glasses the dimensionality of glass structure lies between 2.14 and 2.71. As it can be seen from Figure 5, the Poisson’s ratio is found to nonlinearly decrease with an increase of Bi2O3 content. Further, the values of Poisson’s ratio are believed to be that of a covalently bonded structure. With an increase of Bi2O3 concentration, an increase in fractal dimensionality is observable. This suggests that, by increasing bismuth cations, dimensionality of BPG glass samples increase.

3. Experimental

The new family of bismuth lead germanate (BPG) glasses in form of (½Bi2O3)x–(PbO)40−x–(GeO2)60 where x = 0 to 40 wt% have been successfully prepared by a rapid melt quenching technique [22]. The reagent grade raw materials (99.9% purity of GeO2, PbO and Bi2O3) were weighed out the ratio of the specified composition (in batches of approximately 15 g for each composition) as presented in Table 1.
The raw materials were mixed carefully in an agate mortar and then the mixtures in alumina crucible were preheated in the first furnace at a temperature of 300 °C for 30 min. The mixtures were then transferred in a second furnace and melted in 1100 °C for 1 h under atmospheric conditions. The mixtures were stirred regularly to obtain a better mixing of the composition and to improve the homogeneity of the glass samples.
The melt was then poured quickly on a preheated split metal mold and transferred to the first furnace for annealing at 420 °C before cooling down to room temperature for 24 h. The glass samples with 12 mm diameter and 30 mm thickness were bubble free and homogeneous with good optical quality. A similar procedure was employed to prepare other glass samples. The color of glass sample varies from light to dark violet as more Bi2O3 were added into germanate glassy networks.
The density (ρ) of the glasses were determined by Archimedes method with acetone as buoyant liquid using the relation
ρ = ( W a W a - W b ) ρ b
where Wa is the glass sample weight in air, Wb the glass sample weight in buoyant and ρb the density of the buoyant. All the glass sample’s weights were measured with a digital balance (±0.0001 g accuracy). Their molar volume was calculated from the molecular weight (M) and density (ρ). The accuracy in the measurement of the density is ±0.01 g cm−3 and the relative error is ±0.05%.
The chemically estimated elemental composition values present in the glass samples with the Atomic Absorption Studies (Perkin–Elmer, Model 1372, USA) were found to be slightly smaller than the corresponding elemental nominal composition values (before glass formation), which we considered to be due to evaporation losses and uncertainties in the chemical analysis.
All the glass samples were checked by X-ray diffraction for their amorphous nature using X’Pert Pro Panalytical PW 3040 MPD X-ray powder diffractometer by employing Cr-Ka radiation. The absence of any crystalline peaks in the XRD patterns of the present glass samples indicates the amorphous nature.
For the measurements of ultrasonic velocity in bismuth lead germanate (BPG) glass samples, the samples were shaped into a circular disc of 12 mm diameter and 10–12 mm thickness. The opposite faces of the disc shaped glass samples were highly polished using very fine lapping papers to achieve a good surface finish with plane parallelism having accuracy of ±5 micron.
Ultrasonic velocity measurements were carried out at a frequency of 10 MHz using x-cut and y-cut quartz transducers. A pulse superposition technique was employed using Ultrasonic Data Acquisition System (MATEC 8020, Matec Instruments, USA) [12,22]. Burnt honey was used as a bonding material between the glass samples and transducers. By measuring the thickness of the sample (d), longitudinal (Vl) and transverse (Vt) wave velocities were calculated using the relation, V = 2d/t [12,13]. The absolute accuracy in the measurement of the velocity is ±5 ms−1 and the relative error is ±0.1%.
Glasses are isotropic and have only two independent elastic constant of L and G which are obtained from longitudinal and shear sound wave velocity and density of the present glass samples. The various elastic properties of the glasses were calculated using the following relations [22]:
Longitudinal modulus : L = ρ V 1 2
Shear modulus : G = ρ V s 2
Bulk modulus : K = ρ ( V l 2 - 4 3 V s 2 )
Young s modulus : E = ρ V s 2 ( 3 V l 2 - 4 V s 2 ) V l 2 - V s 2
Poisson s ratio : σ = ( V l 2 - 2 V s 2 ) 2 ( V l 2 - V s 2 )
Fractal dimensionality : d = 4 G / K

4. Conclusions

New ternary germanate based glass samples added with bismuth and lead oxides were successfully prepared and their glassy natures were confirmed by the XRD method. Based on the result obtained, it demonstrated that the density and molar volume increase with glass modifier content, which more attributed to the replacement of Bi2O3 and PbO; both had larger density and molar volume than germanate glass networks. Such increase in their density and volume with composition are due to the compactness of the structure. The observed higher values in velocity at high modifiers content for this glass types, confirmed a substantial change in glass structure. Their longitudinal, shear and Young’s modulus for bismuth lead germanate glasses are also found to increase with the addition of Bi2O3. Meanwhile, there was a similar pattern in elastic moduli with the increasing of Bi2O3 content, where the values of both ultrasonic wave velocities increased subsequently.

Acknowledgement

The financial support from Universiti Putra Malaysia (UPM), under the Research University Grant Scheme (RUGS), vote no. 91748 is gratefully acknowledged.

References

  1. Reisfeld, R.; Boehm, L. Optical properties of bismuth in germanate, borax and phosphate glasses. J. Non-Cryst. Solids 1974, 16, 83–92. [Google Scholar]
  2. Reisfeld, R.; Bornstein, A.; Boehm, L. Optical characteristics and intensity parameters of Sm3+ in GeO2, ternary germanate, and borate glasses. J. Solid State Chem 1975, 14, 14–19. [Google Scholar]
  3. Cho, Z.H.; Farukhi, M.R. Bismuth germanate as a potential scintillation detector in positron camera. J. Nucl. Med 1977, 18, 840–844. [Google Scholar]
  4. Reisfeld, R.; Boehm, L.; Eckstein, Y.; Lieblich, N. Multiphonon relaxation of rare earth ions in borate, phosphate, germanate and tellurite glasses. J. Lumin 1975, 10, 193–204. [Google Scholar]
  5. Wachtler, M.; Speghini, A.; Gatterer, K.; Fritzer, H.P.; Ajo, D.; Bettinelli, M. Optical properties of rare-earth ions in lead germanate glasses. J. Am. Ceram. Soc 1998, 81, 2045–2052. [Google Scholar]
  6. Rabukhin, A.I.; Belousova, G.V. Science for the Glass Industry; Plenum Publishing Corporation: New York, NY, USA, 1992; pp. 264–266. [Google Scholar]
  7. Rabukhin, A.I. Photoelastic constants of germanate glasses containing lead and bismuth oxides. Glass Ceram 1994, 51, 353–359. [Google Scholar]
  8. Rabukhin, A.I. Acoustooptical efficiency of germanate glasses containing lead and bismuth oxides. Glass Ceram 1995, 52, 203–206. [Google Scholar]
  9. Tawansi, A.; Gohar, I.A.; Holland, D.; El-Shishtawi, N.A. Some physical properties of lead borate glasses. I. Influences of heat treatment and PbO content. J. Phys. D Appl. Phys 1988, 21, 607–613. [Google Scholar]
  10. Hamezan, M.; Sidek, H.A.A.; Zaidan, A.W.; Kaida, K.; Zainal, A.T. Elastic contants and thermal properties of Lead-bismuth Borate glasses. J. Appl. Sci 2006, 6, 943–949. [Google Scholar]
  11. Senin, H.B.; Sidek, H.A.A.; Saunders, G.A. Elastic behaviour of terbium metaphosphate glasses under high pressures. Aust. J. Phys 1994, 47, 795–810. [Google Scholar]
  12. Sidek, H.A.A.; Chow, S.P.; Talib, Z.A.; Halim, S.A. Formation and elastic behaviour of lead-magnesium chlorophosphate glasses. Turk. J. Phys 2004, 28, 67–71. [Google Scholar]
  13. Halimah, M.K.; Daud, W.M.; Sidek, H.A.A. Elastic properties of TeO2–B2O3–Ag2O glasses. Ionics 2010, 16, 807–813. [Google Scholar]
  14. Singh, N.; Singh, K.J.; Singh, K.; Singh, H. Comparative study of lead borate and bismuth lead borate glass systems as gamma-radiation shielding materials. Nucl. Instrum. Methods Phys. Res. B 2004, 225, 305–309. [Google Scholar]
  15. Van Kirk, S.E.; Martin, S.W. Preparation and characterization of high-density PbO–Bi2O3–B2O3 glasses. J. Am. Ceram. Soc 1992, 75, 1028–1031. [Google Scholar]
  16. 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]
  17. Saddek, Y.B. Ultrasonic study and physical properties of some borate glass. Mater. Chem. Phys 2004, 83, 222–228. [Google Scholar]
  18. Saddeek, Y.B. Effect of B2O3 on the structure and properties of tungsten-tellurite glasses. Philos. Mag 2009, 89, 41–54. [Google Scholar]
  19. Saddeek, Y.B. Structural and acoustical studies of lead sodium borate glasses. J. Alloy. Compd 2009, 467, 14–21. [Google Scholar]
  20. Shih, P.Y. Thermal, chemical and structural characteristics of erbium-doped sodium phosphate glasses. Mater. Chem. Phys 2004, 84, 151–156. [Google Scholar]
  21. Higazy, A.A.; Bridge, B. Elastic constants and structure of the vitreous system Co3O4–P2O5. J. Non-Cryst. Solids 1985, 72, 81–108. [Google Scholar]
  22. Halimah, M.K.; Sidek, H.A.A.; Daud, W.M.; Zainul, H.; Talib, Z.A.; Zaidan, A.W.; Zainal, A.S.; Mansor, H. Ultrasonic study and physical properties of borotellurite glasses. Am. J. Appl. Sci 2005, 2, 1541–1546. [Google Scholar]
Figure 1. Density of the ternary lead-bismuth germinate glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x (A) and compared with those of (B2O3)20–(PbO)80-x–(Bi2O3)x (B) [14]. The line is drawn to guide the eye.
Figure 1. Density of the ternary lead-bismuth germinate glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x (A) and compared with those of (B2O3)20–(PbO)80-x–(Bi2O3)x (B) [14]. The line is drawn to guide the eye.
Ijms 13 04632f1
Figure 2. Molar volume of the ternary lead-bismuth germinate glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x (A) and compared with those of (B2O3)20–(PbO)80−x–(Bi2O3)x (B). The line is drawn to guide the eye.
Figure 2. Molar volume of the ternary lead-bismuth germinate glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x (A) and compared with those of (B2O3)20–(PbO)80−x–(Bi2O3)x (B). The line is drawn to guide the eye.
Ijms 13 04632f2
Figure 3. Longitudinal and shear ultrasonic velocities at room temperature in ternary (GeO2)60–(PbO)40−x–(1/2Bi2O3)x glass systems. The line is drawn to guide the eye.
Figure 3. Longitudinal and shear ultrasonic velocities at room temperature in ternary (GeO2)60–(PbO)40−x–(1/2Bi2O3)x glass systems. The line is drawn to guide the eye.
Ijms 13 04632f3
Figure 4. Variation of elastic moduli of ternary (GeO2)60–(PbO)40−x–(1/2Bi2O3)x glass systems with composition. The line is drawn to guide the eye.
Figure 4. Variation of elastic moduli of ternary (GeO2)60–(PbO)40−x–(1/2Bi2O3)x glass systems with composition. The line is drawn to guide the eye.
Ijms 13 04632f4
Figure 5. Variation of fractal dimensionality and Poisson’s ratio of ternary (GeO2)–(PbO)–(1/2Bi2O3) glass systems with composition. The line is drawn to guide the eye.
Figure 5. Variation of fractal dimensionality and Poisson’s ratio of ternary (GeO2)–(PbO)–(1/2Bi2O3) glass systems with composition. The line is drawn to guide the eye.
Ijms 13 04632f5
Table 1. Glass composition of the ternary lead-bismuth germanate glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x together with density, molar weight and molar volume. The experimental data of (B2O3)20–(PbO)80−x–(Bi2O3)x [14] is included for comparison.
Table 1. Glass composition of the ternary lead-bismuth germanate glasses of the form (GeO2)60–(PbO)40−x–(1/2Bi2O3)x together with density, molar weight and molar volume. The experimental data of (B2O3)20–(PbO)80−x–(Bi2O3)x [14] is included for comparison.
SamplesGlass Composition (mol%)Density (g cm−3)Molar Weight (g/mol)Molar Volume (cm3)
GeO2–PbO–1/2Bi2O3
A160–40–05.90152.0625.77
A260–30–105.91153.0425.89
A360–20–205.99154.0125.73
A460–10–306.01154.9925.78
A560–0–406.05155.9725.80
B2O3–PbO–Bi2O3
B120–70–106.69216.7632.40
B220–60–206.90241.0434.93
B320–50–307.41265.3135.80
B420–40–407.26289.5939.89
B520–30–507.39313.8642.47
B620–20–607.36338.1445.94
B720–10–707.24362.4250.06
Table 2. The room temperature ultrasonic wave velocities and elastic moduli of the ternary lead-bismuth germanate glasses together with fractal dimensionality and Poisson’s ratio.
Table 2. The room temperature ultrasonic wave velocities and elastic moduli of the ternary lead-bismuth germanate glasses together with fractal dimensionality and Poisson’s ratio.
Glass samplexVelocity (m/s)Elastic Moduli (GPa)d = 4G/KPoisson’s Ratio

VlVsLEKG
A103422191269.1054.9340.3421.572.140.27
A2103445196370.1657.3939.7922.782.290.26
A3203475202772.2861.1039.4924.592.490.24
A4303482203772.8961.8639.6324.952.520.24
A5403511209474.5364.9039.1826.512.710.22

Share and Cite

MDPI and ACS Style

Sidek, H.A.A.; Bahari, H.R.; Halimah, M.K.; Yunus, W.M.M. Preparation and Elastic Moduli of Germanate Glass Containing Lead and Bismuth. Int. J. Mol. Sci. 2012, 13, 4632-4641. https://doi.org/10.3390/ijms13044632

AMA Style

Sidek HAA, Bahari HR, Halimah MK, Yunus WMM. Preparation and Elastic Moduli of Germanate Glass Containing Lead and Bismuth. International Journal of Molecular Sciences. 2012; 13(4):4632-4641. https://doi.org/10.3390/ijms13044632

Chicago/Turabian Style

Sidek, Hj A. A., Hamid R. Bahari, Mohamed K. Halimah, and Wan M. M. Yunus. 2012. "Preparation and Elastic Moduli of Germanate Glass Containing Lead and Bismuth" International Journal of Molecular Sciences 13, no. 4: 4632-4641. https://doi.org/10.3390/ijms13044632

Article Metrics

Back to TopTop