Acessibilidade / Reportar erro

Properties and structural features of iron doped BABAL glasses

Abstract

The chemical durability, density and structure of the BABAL glasses with batch compositions (100-x)(0.30BaO·0.50B2O3·0.20Al2O3)·xFe 2O3 (1 < x < 10 mol%), were investigated using Mössbauer spectroscopy, electron paramagnetic resonance (EPR), X-ray diffraction, Raman and differential thermal analysis (DTA). The chemical durability for the glass of composition 27BaO·45B2O3·18Al 2O3·10Fe2O3 (mol%) at 90 °C in distilled water was 700 times lower than that of iron phosphate glass 40Fe2O3·60P2O 5 (mol%). The Mössbauer spectra indicate the presence of iron (II) and iron (III) in tetrahedral or octahedral coordination. The results obtained from the g ef = 4.3 EPR line are typical of the occurrence of iron (III) occupying substitutional sites and the line g ef = 2.0 is related to the association of two or more Fe ions found in the interstices (or holes occupied by the glass modifier cations) of the glass network. The paths of X-ray diffraction are typical for glasses based in borate glasses. The Raman spectra showed that the boroxol ring disappears with the increase of iron content, concomitant with the appearance of BO4 and tetraborate structural units. At these conditions, an increase of dissolution rate and clustering of iron ions is observed.

aluminoborate glass; structure; chemical durability


Properties and structural features of iron doped BABAL glasses

Signo Tadeu dos ReisI* * e-mail: reis@umr.edu ; Walter M. PontuschkaII; Jinbo B. YangI; Dalva L A. FariaIII

IGraduate Center for Materials Research, 301 Strautmanis Hall, University of Missouri-Rolla, MO65409, U.S.A.

IIInstitute of Physics, University of São Paulo, C.P. 66318, 053890-970 São Paulo - SP, Brazil

IIIInstitute of Chemistry, University of São Paulo, C.P. 26077, 05389-970 São Paulo - SP, Brazil

ABSTRACT

The chemical durability, density and structure of the BABAL glasses with batch compositions (100-x)(0.30BaO·0.50B2O3·0.20Al2O3)·xFe 2O3 (1 < x < 10 mol%), were investigated using Mössbauer spectroscopy, electron paramagnetic resonance (EPR), X-ray diffraction, Raman and differential thermal analysis (DTA). The chemical durability for the glass of composition 27BaO·45B2O3·18Al 2O3·10Fe2O3 (mol%) at 90 °C in distilled water was 700 times lower than that of iron phosphate glass 40Fe2O3·60P2O 5 (mol%). The Mössbauer spectra indicate the presence of iron (II) and iron (III) in tetrahedral or octahedral coordination. The results obtained from the gef = 4.3 EPR line are typical of the occurrence of iron (III) occupying substitutional sites and the line gef = 2.0 is related to the association of two or more Fe ions found in the interstices (or holes occupied by the glass modifier cations) of the glass network. The paths of X-ray diffraction are typical for glasses based in borate glasses. The Raman spectra showed that the boroxol ring disappears with the increase of iron content, concomitant with the appearance of BO4 and tetraborate structural units. At these conditions, an increase of dissolution rate and clustering of iron ions is observed.

Keywords: aluminoborate glass, structure, and chemical durability

1. Introduction

As long been known1, the considerable practical interest of borate glasses stems mainly from the complex influence of boron oxide on the physical properties2. Some appropriate values of constants such as low melting point, high thermal expansion coefficients and several optical data make these glasses potential candidates for many technological applications such as optical3, medical use4, and solid state electrolytes5.

The local structure of iron doped aluminoborate glasses has been studied using 57Fe Mössbauer, Raman spectroscopy and electron paramagnetic resonance (EPR)6,7. The boroxol ring is the main structural group present in borate glasses and the addition of modifier oxide like CaO gradually converts three-coordinated boron into four-coordinated boron units, resulting in the formation of diborate, tetraborate and pentaborate glass network superstructural units7.

Iron is one of the most pervasive impurities in borate and phosphate glasses. Some specific features of the EPR line width and shape of Fe3+ ions in glass matrices depend on the glass composition and the concentration of paramagnetic ions8. The EPR spectra of most glasses containing iron exhibit the two well known resonances at the effective g values gef» 2 and gef» 4.3 that have been considered as a signature of the presence of Fe3+ ions in a glassy host9. The X-band EPR spectra (~ 9 GHz) of most oxide glasses with low Fe3+ concentrations show a gef = 4.3 line, and in most cases the gef = 2 line is also present. A low field feature (gef ~ 9.7) is also observed in X-band EPR spectra of oxide glasses10. At low iron concentrations (< 1 mol%), only a sharp gef» 4.3 resonance is commonly observed9. As the concentration is increased, a broad line begins to grow at about gef» 2 11. In many oxide glasses the intensity of the line with g = 2 increases much faster than that of g = 4.3 with increasing Fe3+ ions concentration above 5 mol%. This result leads to the hypothesis of Fe3+ clustering, which effectively contributes to the former broad line12. As the linewidth of the gef ~ 4.3 resonance remains roughly constant, in practice the peak-to-peak amplitude of its derivative is proportional to the Fe concentration of the sample13, while the relative concentrations of the Fe2+ and Fe3+ ions are easily determined by Mössbauer spectroscopy.

Bogomolova and Henner14 showed how the X-band the gef» 2 line width can be explained by a clustering of Fe3+ ions. Nevertheless, the Fe3+ concentration determines not only the number of EPR lines and their positions, but the line shapes as well. The theory predicts an increase of line width with spin concentration due to the dipolar broadening. The exchange interaction appears at a sufficiently high concentration of paramagnetic ions (molar contents > 5 mol%)15, when neighboring spins become highly coupled14.

The aim of the present work is to investigate the conditions of the chemical durability of BABAL glasses. Evidence of the presence of diads and microclusters of iron ions and structural studies of the glass matrix were provided by EPR, X-ray diffraction and Mössbauer spectroscopy. The chemical durability of BABAL glasses was determined and compared with those of the iron phosphate glasses.

2. Experimental Procedure

Samples of BABAL glasses with compositions (100-x)(0.30BaO·0.50B2O3 ·0.20Al2O)·xFe2O 3 where x = 0, 1, 2, 4, 8 and 10 mol% (see Table 1) were produced by melting the batches in dense alumina crucibles at 1200 °C for 1 h. The melt was quenched in air by pouring into 1 × 1 × 5 cm3 steel mold. The samples were transferred to an annealing furnace and held at 350 °C for 3 h. The densities of the samples were measured by the Archimedes method using distilled water at room temperature. A portion of the glass samples was crystallized on heating at 750 °C for 8 h in air. Room temperature X-ray powder diffraction (XRD) patterns for both glasses and crystalline counterparts were collected on a X-ray diffractometer (Scintag XDS2000). The chemical durability of the bulk glasses with approximate size of 1 × 1 × 1 cm3 was evaluated by weight loss measurements from glasses exposed to deionized water at 90 °C for 2-8 days. Glasses were polished to 600 grit, finished with SiC paper, cleaned with acetone and suspended in glass flasks containing 100 ml of deionized water at 90 °C. Duplicated measurements were conducted for each glass and the average dissolution rate (DR), normalized to the glass surface area, and the corrosion time was calculated from the weight loss.

The Mössbauer spectra were obtained at room temperature on a spectrometer provided with a 50 mCi rhodium matrix cobalt-57 source. The amount of iron found at the sample holder was 4 mg/cm2. The velocity of the source was calibrated using a pure iron foil that was also used as a reference for the isomer shift value.

The Raman spectra of the glass samples were taken at room temperature using a Renishaw Raman system 3000, coupled to an Olympus metallurgical microscope. The spectra were excited at 6328 nm (He-Ne Spectra Physics mod. 127) and recorded in the 200-1800 cm-1 range. EPR spectra of powdered samples were taken with a Bruker EMX 10/12 homodyne, X-band spectrometer.

3. Results

3.1 Chemical durability

The dissolution rates (DRs) of the BABAL glasses measured from the weight loss experiments conducted in distilled water at 90 °C are given in Table 1 and Fig. 1. Clearly DR varies with Fe2O3 content: it decreases from 4.7 × 10-6 g/cm2min. for glasses containing 0 mol% Fe2O3 containing glass to about 1.0 × 10-7 g/cm2min for glasses containing about 4 mol% of Fe2O3. Even though the samples with Fe2O3 content between 2 and 4 mol% had the lowest chemical durability (see Table 1 and Fig. 1), there is an increase in DR for samples containing more than 4 mol% Fe2O3. As a comparison, the iron phosphate glass containing 40 mol% Fe2O3 16 had a dissolution rate 100 times less than the dissolution rate for window glass and 700 times less than the dissolution rate for the BABAL glass containing 10 mol% of Fe2O3, as shown in the Fig. 2.



The pH of the solution after the corrosion tests for a BABAL glass with 2 mol% Fe2O3 is shown in Fig. 3. The reduction in the pH values for the iron phosphate glass during the first four days is consistent with the observed relatively higher DR values as a result of corrosion in these glasses to be correlated with the amount of phosphoric acid formed from the release of phosphorus from this glass. However for the BABAL glass the higher pH values indicates the formation of considerably higher amount of basic compounds from the release of Ba and Al from interstitial (glass modifier) positions.


3.2 X-ray diffraction and density measurements

All of the compositions listed in Table 1 formed glasses, as no crystalline phases were detected by XRD. The XRD pattern of crystalline counterpart of a sample with x = 0 mol% is given in Fig. 4 which indicates that the major crystalline phase in the crystallized sample is the Ba2Al2B8O1717.


As shown in Fig. 5, the density of the BABAL glasses at room temperature increased linearly from 3.12 to 3.41 g/cm3 with increasing Fe2O3 content (see Table 1).


3.3 EPR

Figure 6 shows the X-band EPR spectra of BABAL glasses containing increasing amounts of 1 to 10 mol% of Fe2O3. The intensities were progressively decreased in order to show the evolution of the relative line shapes and intensities at g = 4.3 from a) isolated ions in local tetrahedral (and eventual octahedral) sites9 and b) at g = 2.0 from clusters of two or more Fe3+ ions which interact among themselves.


No significant changes were observed in the line width and shape of the g = 4.3 EPR spectra, with the Fe2O3 concentrations ranging from 1 to 8 mol%, indicating that the increase of Fe has occurred mainly by substitutions of Fe3+ ions at the sites of the BO4 tetrahedra of the glass forming network.

As the Fe2O3 content reaches about 10 mol% there is observed a significant increase of the resonance at g = 2.0 together with a strong line broadening which is also observed in the resonance at g = 4.3 before it is going to be completely obscured by the former which becomes dominant.

A similar evolution has been observed by H.O. Hooper et al.18 in the EPR spectra of iron containing sodium borate glasses, with the disappearance of the g = 4.3 resonance at about the same relative amount of Fe2O3 (8- 0 mol%). Later on D.W. Moon et al.19 have reported that in the glass system xFe2O3·(100-x)[BaO· 4B2O3] at concentrations greater than 3 mol% Fe2O3, iron ions associated in pairs (diads) and groups of three (triads) became predominant, in addition to the isolated Fe3+ ions. They assigned the shoulder at g = 6.0 to weak crystal field terms and the g = 2.0 resonance to clusters of more than one iron atom with antiferromagnetic interaction.

More recently, L. Cugunov et al.20, E.M. Yahiaoui11, and R. Berger et al.9,21 carried out studies with computer simulations of the Fe3+ resonance in borate glass. The experimental EPR spectra were well described by assuming that the environment of Fe3+ ions are highly distorted and in addition to the orthorhombic distortion related to the g = 4.3 feature, an important part of Fe3+ arises from axial or slightly rhombic distortions.

3.4 Raman

The Raman spectra for the lower concentration of Fe2O3 are dominated by a narrow and intense band at 806 cm-1 (Fig. 7). The 806 cm-1 and 540 cm-1 line intensities decrease with the increasing concentration of Fe2O3 and other bands centered at 580, 780 and 920 cm-1 appear. For higher concentration of Fe2O3 the bands between 580, 780 and 920 cm-1 are predominating in the spectra.


3.5 Mössbauer

The room temperature Mössbauer spectra of the BABAL glasses with iron content > 2 mol% are shown in Fig. 8. The iron valence and hyperfine parameters, isomer shift dand quadrupole splitting DEQ, calculated from the Mössbauer spectra are given in Table 2. It is found that some of the Fe(III) ions in the starting batch are reduced to Fe(II) ions as all of the glasses contain varying amounts of Fe(II) as determined from the Mössbauer spectra. The isomer shift values for Fe(III) and Fe(II) ions range between 0.39 mm/s and 1.08 mm/s while the quadrupole splitting values range between 1.10 mm/s and 2.09, respectively. These values of isomer shifts correspond to tetrahedral coordination for both Fe(II) and Fe(III) in these glasses.


4. Discussion

The chemical durability for the (100x) (0.30BaO·0.50B2O3·0.20Al 2O3)·xFe2O3 glasses with iron content of 2 mol% is 700 times less than that iron phosphate glass containing 40 mol% of Fe2O3. In general the chemical durability for phosphate glasses improves with the increase in iron content16, however the decrease in the chemical durability of BABAL glasses for iron content > 4 mol% could be attributed to the formation of microclusters in substitutional sites in the glass network22.

Figure 6 shows that the EPR lines gef » 2.0 and gef» 4.3 on the spectra these glasses are due to the diluted Fe3+ ions in tetrahedral or orthorhombic coordination9,23 and small clusters of exchange coupled Fe3+ ions in substitutional sites of the glass network. Figure 6 also shows a low field feature, as a shoulder, in the g »10 region. Moreover, this shoulder seems to disappear in the samples with the higher iron concentration.

The Raman band around 806 cm-1 is attributed to boroxol rings, which is the main structural unit of the B2O3 glasses6,7. The addition of Fe2O3 converts tree-coordenated boron into four-coordinated boron units (see Fig. 7), resulting in the formation of tetraborate groups (780 cm-1). The band in 540 cm-1 is attributed to "loose" BO4 groups and "loose" diborate groups24, and the band in the 920 cm-1 can be attributed to the formation of pentaborate groups25. These results are consistent with the investigation of Fe(II) and Fe(III) properties in calcium aluminoborate glasses7.

Mössbauer results indicate that even though there is no Fe(II) in the starting batch, the resulting glasses contain varying amounts of Fe(II) (ranging from 6 to 20%) which shows that some of the Fe(III) ions are reduced to Fe(II) ions during melting. The observed isomer shift values indicate that both Fe(II) and Fe(III) ions have oxygen neighbors in the tetrahedral coordination.

5. Conclusions

The BABAL glass corrosion rate in neutral environment is caused, respectively, by the increase in the pH of the solution (water) due to the attack of OH- and by the reaction between H+ and metallic ions (Fe2+ and Fe3+). EPR lines gef » 2.0 and gef» 4.3 on the spectra of these glasses can be due to the diluted state of Fe3+ ions in tetrahedral or orthorhombic coordination and small clusters of exchange-coupled Fe3+ ions in substitutional sites of the glass network.

Acknowledgements

S.T. Reis thanks Fundação de Amparo a Pesquisa do Estado de São Paulo (FAPESP) for a post-doctoral fellowship. The authors also thank Mr. Hannes Fischer and Mr. José Mário Prison for the analyses of the XRD and EPR.

Received: February 26, 2002

Revised: May 15, 2003

  • 1. Joo, C.; Werner-Zwanziger, U.; Zwanziger, J.W. J. Non-Cryst. Solids, v. 261, p. 282, 2000.
  • 2. Kamitsos, E.I.; Chryssikos, G.D. J. Mol. Struct, v. 247, p. 1, 1991.
  • 3. Guang, H.E.; Przemyslaw, P.; Chau Lin, T. Nature, v. 415, p. 767, 2002.
  • 4. Conzone, S.D.; Brown, F.R.; Day, D.E.; Ehrhardt, G.J. B. Mater Res., v. 60, p. 260, 2002.
  • 5. Friedrich, G.K.; Duffy, J.A. J.Chem.Soc.Faraday Trans., v. 79, p. 661, 1983.
  • 6. Pascoal, H.B.; Pontuschka, W.M.; Rechenberg, H. Appl. Phys, A 70, p. 211, 2000.
  • 7. Pascoal, H.B.; Pontuschka, W.M.; Rechenberg, H. J. of Non-Cryst. Solids, v. 258, p. 92, 1999.
  • 8. Donald, I.W.; Metcalfe, B.L.; Taylor, R.N.J. J. Mater. Sci., v. 32, p. 5851, 1997.
  • 9. Berger, R.; Kliava, J.; Yahiaoui, E.M.; Bissey, J.C.; Zinsou, P.K.; Béziade, P. J. of Non-Cryst. Solids, v. 180, p.151, 1995.
  • 10. Yunfei, L.; Weiwei, H. J. of Non-Cryst. Solids, v. 112, p.136, 1989.
  • 11. Yahiaoui, E.M.; Berger, R.; Servant, Y.; Kliava, J.; Cugunov, L.; Mednis, A. J. Phys.: Condens. Matter, v. 6, p. 9415, 1994.
  • 12. Bogomolova, L.D.; Krasil'nikova, N.A.; Mitrofanov, V.V. Glass Phys. Chem. v. 21, p. 303, 1995.
  • 13. Del Nery, M.; Pontuschka, W.M.; Isotani, S.; Rouse, C.G. Phys. Rev. B49, p. 3760, 1994.
  • 14. Bogomolova, L.D.; Henner, E.K. J. Magn. Reson., v. 41, p. 422, 1980.
  • 15. Stoyanova, R.; Gorova, M. Journal. of Physics and Chemistry of Solids, v. 61, p. 615, 2000.
  • 16. Reis, S.T.; Karabulut, M.; Day, D.E. J. of Non-Cryst. Solids, v. 292, p. 150, 2001.
  • 17. Hubmer, A.K. Neies Jajrb. Mineral, Abh, v. 112, p. 150, 1970.
  • 18. Hooper, H.O.; Beard, G.B.; Catchings, R.M.; Bukrey, R.R.; Forrest, M.; Kenealy, P.F.; Kline, R.W.; Moran Jr., T.J.; O'Keefe, J.G.; Thomas, R.L.; Verhelst, R.A. "Magnetic order in Alkaliborate and Aluminosilicate Glasses Containing Large Concentrations of Iron-Group Ions", in: Amorphous Magnetism, ed. by H.O. Hooper and A.M. de Graaf, Plenum New York, p. 47-63, 1972.
  • 19. Moon, D.W.; Aitken, J.M.; MacCrone, R.K. Phys. Chem. Glasses, v. 16, p. 91, 1975.
  • 20. Cugunov, L.; Mednis, A.; Kliava, J. J. Magn. Resonance, A106,p. 153, 1994.
  • 21. Berger, R.; Bissey, J.C.; Kliava, J.; Soulard, B. J. Magnetism and Magnetic Materials, v. 167, p. 129, 1997.
  • 22. Reis, S.T.; Pontuschka, W.M.; Martinelli, J.R. The existence of iron microsclusters related to the chemical durability of sintered LIP glasses, (submitted for publication).
  • 23. Fuxi, G. "Optical and Spectroscopic properties of glass, Chap. 6, Springer Verlag, 1992.
  • 24. Merra, B.N.; Ramakrishkna, J. J. of Non-Cryst. Solids, v. 159, p. 1, 1993.
  • 25. Maniu, D.; Ardelean, I.; Illiescu, T.; Pantea, C. J. Mater. Sci. Lett., v. 16, p. 19, 1997.
  • *
    e-mail:
  • Publication Dates

    • Publication in this collection
      28 Nov 2003
    • Date of issue
      June 2003

    History

    • Accepted
      15 May 2003
    • Received
      26 Feb 2002
    ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
    E-mail: pessan@ufscar.br