Atul D.
Sontakke
and
Kalyandurg
Annapurna
*
Glass Science and Technology Section, CSIR-Central Glass and Ceramic Research Institute, 196, Raja S. C. Mullick Road, Kolkata, 700 032, India. E-mail: annapurnak@cgcri.res.in; Fax: +91-33 2473 0957; Tel: +91-33 2473 3469
First published on 9th November 2012
Glass formation in a newly formulated low germanium alkaline earth gallate (LGMGa, M = Ca/Mg/Zn/Sr/Ba) system by a high temperature melt quenching route is reported here. The proportionate ratios of tetrahedral to octahedral coordination of both Ga3+ and Ge4+ network cations are found to be crucial for vitrification. The enhancement in octahedral coordination led to surface crystallization and phase separation in cast melts. This behavior has been attributed to the overall reduction in tetrahedrally coordinated network units due to the failure of Ca2+ ions in charge compensation of the [GaO4]− tetrahedron. Among different network modifying oxides, BaO in combination with CaO (mixed alkaline earth) containing compositions have proved to yield clear glass formation with improved tetrahedral networking. This has been explained though the basicity equalization concept and also due to efficient charge compensation of the [GaO4]− tetrahedron requiring lower field strength cations like Ba2+. The effective phonon energy of this low germanium calcium barium gallate (LGCBGa) glass is found to be around 650 cm−1, with an extended infrared transparency up to around 7 μm, which is superior to most of the low phonon energy oxide glass systems, making it promising for photonic and mid-infrared luminescence applications. Furthermore, the observed dark yellow to brown coloration in the glass has been attributed to the precipitation of metallic nanoparticles, and this mechanism has been discussed in detail.
Al2O3, being a glass network intermediate oxide, forms stable glasses with the aid of comparatively lower field strength cations such as alkali metal (group IA) or alkaline earth metal (group IIA) ions.2 The structural chemistry of Al3+ ions prefers the [AlO6] octahedron as the most stable polyhedral unit,3 but its moderate field strength gives the possibility for squeezing the oxygen coordination sphere down to a four-fold [AlO4]− tetrahedron,3,4 as evidenced in the excess alkali containing silicate glasses and calcium-aluminate glasses.2,5–7 The isovalent gallium (Ga3+) also follows the footsteps of Al3+ in a similar manner and forms a [GaO4]− tetrahedron to participate in the glass network structure.8,9
The substitution of trivalent Al3+ over tetravalent Si4+ in the alkali silicate network gives rise to the polymerization of the anionic framework, where mono-valent alkali metal ions serve as charge compensators, leading to reduction of the overall density of non-bridging oxygens (NBOs) in the network. The inclusion of Al3+ in the alkali silicate network enhances both the viscosity and glass transition temperature (Tg) of glass along with added strength, chemical durability and thermal stability.2 Furthermore, the lower bond strength of the AlIII–O bond in the [AlO4]− tetrahedron compared to the SiIV–O bond reduces the lattice vibrational energy in low silica alkali/alkaline earth aluminate or binary alkali/alkaline earth aluminate glass.10,11 The reduced phonon energy of aluminate glasses (∼800 cm−1) extends the infrared window with the cut-off lying at around 6–6.5 μm, which is comparable to the heavy metal oxide glasses, besides exhibiting significantly low optical loss.12 The substitution of Al3+ by Ga3+ in these networks can further extend the infrared transparency up to 7–8 μm, owing to the fairly low lattice vibrational energy of the GaIII–O bond in the [GaO4] tetrahedron (∼650 cm−1).13
But, unlike Al3+, the glass forming ability of Ga3+ is comparatively poor. Shelby observed that gallosilicate glasses are more prone to phase separation than aluminosilicate glasses.8 The larger ionic radius of Ga3+ compared to Al3+ reduces the cationic field strength and makes it less favorable to form tetrahedral coordination with oxygen.3 In ideal conditions, one alkali ion per [MIIIO4]− tetrahedron is sufficient for the charge compensation. But it is observed that a fraction of M3+ cations tends to occupy the octahedral coordination, which demands excess alkali ions to enhance the overall tetrahedral/octahedral ratio in the network.14 The basicity equalization concept of Duffy and Ingram gives a satisfactory explanation of this behavior.15 Accordingly, the structural units that have a group basicity close to the glass basicity are more abundant in the network. Miura et al.16 and Tanaka et al.17 could successfully predict the coordination chemistry of [BO4] to [BO3] proportionality in borosilicate glasses based on the basicity equalization concept. This concept provides clear guidelines for the structural chemistry of such glasses involving multi-coordinated network units.
The present work has been initiated to develop stable glass compositions based on Ga2O3 as the main network former, in association with alkaline earth cations and a small amount of GeO2, in order to achieve low phonon characteristics for photonic applications. The glasses with low phonon energy characteristics are preferred hosts for luminescent ions in achieving better quantum efficiency, owing to their reduced non-radiative losses, apart from their extended infra-red transparency. Furthermore, some specific luminescence transitions, which are otherwise impossible in high phonon energy glass, could be achieved in such glasses. Thus, these gallate glass systems could be useful for the development of mid-infrared lasers and waveguides with low optical losses. Nishida et al.,18,19 have also proposed the possibility of calcium-gallate glasses in optical memory devices. However, they have prepared such glasses by the rapid quenching of melt in ice-cold water and thus obtained samples with a transmission of about 40% only in the infrared region. Our initial experiments also revealed that the glass formation of the binary calcium-gallate system is quite poor. So, in this work, a series of compositional variations have been examined to achieve good glass formation in the gallate glass system. The low silica calcium aluminate (LSCA) system has been selected as a starting glass, owing to its superior properties among low phonon oxide glass systems.10 Since Ga3+ and Al3+ are iso-structural in the network, we have tried to maintain the similar structural network to get the advantages of LSCA glasses in the gallate glass system.
This work provides a detailed insight into the structural modifications due to the substitution of lower field strength Ga3+ and Ge4+ in the LSCA network, along with the effect of different alkaline earth metal ions on the glass formation ability. Infrared spectroscopy, differential thermal analysis and optical absorption spectroscopy have been adopted for thorough analysis of the structural and optical properties of prepared glasses. The results have been explained based on the glass basicity and the role of alkaline earth metal ions in the network, as well as their ability for charge compensation.
Glass | Composition (mol%) | |||||||||
---|---|---|---|---|---|---|---|---|---|---|
Al2O3 | Ga2O3 | CaO | ZnO | SrO | BaO | MgO | SiO2 | GeO2 | La2O3 | |
Series 1 | ||||||||||
G-0a | 28 | — | 56 | — | — | — | 4 | 10 | — | 2 |
G-1a | 21 | 7 | 56 | — | — | — | 4 | 10 | — | 2 |
G-2a | 14 | 14 | 56 | — | — | — | 4 | 10 | — | 2 |
G-3a | 7 | 21 | 56 | — | — | — | 4 | 10 | — | 2 |
G-4a | — | 28 | 56 | — | — | — | 4 | 10 | — | 2 |
G-5a | — | 28 | 58 | — | — | — | 4 | 8 | — | 2 |
G-6a | — | 28 | 60 | — | — | — | 4 | 6 | — | 2 |
Series 2 | ||||||||||
G-4b | — | 28 | 56 | — | — | — | 4 | 8 | 2 | 2 |
G-4c | — | 28 | 56 | — | — | — | 4 | 6 | 4 | 2 |
G-4d | — | 28 | 56 | — | — | — | 4 | 3 | 7 | 2 |
G-4e | — | 28 | 56 | — | — | — | 4 | — | 10 | 2 |
Series 3 | ||||||||||
G-3e | 7 | 21 | 56 | — | — | — | 4 | — | 10 | 2 |
G-4eMg | — | 28 | 50 | — | — | — | 10 | — | 10 | 2 |
G-4eZn | — | 28 | 28 | 28 | — | — | 4 | — | 10 | 2 |
G-4eSr | — | 28 | 28 | — | 28 | — | 4 | — | 10 | 2 |
G-4eBa | — | 28 | 28 | — | — | 28 | 4 | — | 10 | 2 |
Fig. 1a shows the infrared reflectance spectra of the glasses in Series 1. It can be seen that the G-0a sample exhibits an intense band at 790 cm−1 with two shoulders on both sides, one at 650 cm−1 and the other peaking at 890 cm−1, respectively. Table 2 presents the assignments of these bands, which are ascribed to the vibrational transitions of the Al–O bond in tetrahedral coordination, Al–O in octahedral units, and the Si–O vibrations in the non-bridging Q(4−x) state as well as in the ‘isolated’ tetrahedron, respectively.10,20 This 790 cm−1 band was systematically reduced upon Ga2O3 substitution and two new bands appeared at around 636 cm−1 and 503 cm−1 due to the Ga–O bond vibration in the [GaO4] tetrahedron and [GaO6] octahedron, respectively.13 Furthermore, the vibrational band at ∼440 cm−1 showed a red shift to 360 cm−1, due to the fading effect in Al–O bending modes on Ga2O3 inclusion, which exhibits bending vibrations at around 360 cm−1.13 The results suggest a systematic replacement of Al3+ in both the tetrahedral and octahedral coordination by Ga3+ cations. However, a close observation reveals that the relative intensities of tetrahedral to octahedral units in aluminate (G-0a) and gallate (G-4a) glasses exhibit a comparatively higher fraction of octahedral coordination in G-4a glass. Interestingly, this trend was reversed upon reducing the SiO2 content, as can be seen from the spectra of the G-5a and G-6a glasses respectively, where the 636 cm−1 band showed an enhancement and the 503 cm−1 band went down compared to G-4a glass, implying improvement of the tetrahedral [GaO4] coordination at the expense of the [GaO6] units in the glass network.
![]() | ||
Fig. 1 (a) Infrared reflectance spectra of Series 1, and (b) Series 2 glasses. |
IR band | Position (±10 cm−1) | Description |
---|---|---|
1 | 890 cm−1 | Si–O stretching vibration in [SiO4]x− tetrahedron |
2 | 790 cm−1 | Al–O stretching vibration in [AlO4] tetrahedron |
3 | 650 cm−1 | Al–O stretching vibration in [AlO6] octahedron |
4 | 636 cm−1 | Ga–O stretching vibration in [GaO4] tetrahedron |
5 | 503 cm−1 | Ga–O stretching vibration in [GaO6] octahedron |
6 | 440 cm−1 | Al–O and Si–O bending vibrations |
7 | 360 cm−1 | Ga–O bending vibrations |
8 | 745 cm−1 | Ge–O stretching vibration in [GeO4]x− tetrahedron |
9 | 610 cm−1 | Ge–O stretching vibration in [GeO6] octahedron |
10 | 407 cm−1 | Ge–O bending vibrations |
11 | >300 cm−1 | MII–O vibrations in multi-polyhedral units |
This strange cation coordination behavior has its explanation in the concept of glass basicity. Table 3 depicts the basicity of the glasses under study, along with some important physical properties. The basicity has been estimated by three different methods, based on (i) Pauling's electronegativity scale,3,15 (ii) the oxide ‘optical’ basicity concept of Duffy and Ingram,15,21 and (iii) the experimental method using measured refractive indices22 (Appendix I). The basicity values calculated from these methods show significant deviations from each other, which may be due to the different approaches used for their estimation.15 The theoretical basicity based on Duffy's method follows a similar trend to the experimental basicity values, increasing with the inclusion of Ga2O3 in the glass; but the basicity values estimated using Pauling's electronegativity scale show a decrease in glass basicity up to G-4a, and then a small increase for G-5a and G-6a glasses. Pauling defined electronegativity as the relative tendency towards electron attraction.3 Gallium, being more electronegative attracts a larger electron cloud in Ga–O bonding and thus reduces the total charge density of the anionic ligand, which in turn affects the overall electron donating power of the network and reduces the glass basicity. Furthermore, the reduction in acidic SiO2 content in the G-5a and G-6a glasses again enhances the theoretical basicity of the glass. Since the metal ion favors higher coordination in an acidic environment, based on the above argument the proportionality of octahedral coordination increases from G-0a to G-4a glass due to the reduction in glass basicity, and then decreases on SiO2 reduction, favoring the formation of tetrahedral coordination of Ga3+ cations. Further reduction in the theoretical basicity (Pauling) of the glass with the GeO2 substitution for SiO2 content in Series 2 compositions hints at the formation of more fractions of octahedrally coordinated units in the network, which might have resulted in phase separation of the glass melt. The infrared reflectance spectra shown in Fig. 1b give more insight into the structural modifications induced by GeO2 inclusion. The Si–O stretching band at 890 cm−1 diminished upon GeO2 substitution, accompanied by a broadening of the 636 cm−1 band with a slight blue shift in its peak position to 665 cm−1.
Glass | Nature of glass | Physical properties | Glass basicity (Λ) | Thermal properties | ||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
d (g cm−3) | M avg (g mol−1) | n d | Pauling | Duffy | Expa | T g (±3 °C) | T p1 (±1 °C) | T p2 (±1 °C) | T m (±2 °C) | T x − Tg (±5 °C) | ||
a Derived from experimental refractive index and oxide ion polarizability.(Glass + SC) refers to quick devitrification of upper surface layer of poured glass. (Glass + C) refers to partially crystallized glass – phase separated. | ||||||||||||
G-0a | Glass | 3.068 | 74.1 | 1.676 | 0.697 | 0.747 | 0.799 | 806 | 941 | 970 | 1343 | 113 |
G-1a | Glass | 3.361 | 80.1 | 1.690 | 0.689 | 0.760 | 0.795 | 786 | 918 | 943 | 1298 | 104 |
G-2a | Glass | 3.558 | 86.1 | 1.708 | 0.679 | 0.772 | 0.824 | 761 | 880 | 915 | 1274 | 96 |
G-3a | Glass + SC | 3.766 | 92.0 | 1.728 | 0.671 | 0.784 | 0.849 | 740 | 846 | 891 | 1250 | 91 |
G-4a | Glass + SC | 3.973 | 98.0 | 1.751 | 0.662 | 0.797 | 0.874 | 738 | 829 | 865 | 1257 | 90 |
G-5a | Glass + SC | 4.013 | 96.8 | 1.766 | 0.667 | 0.804 | 0.886 | 701 | 824 | 844 | 1228 | 93 |
G-6a | Glass + SC | 4.067 | 95.6 | 1.774 | 0.672 | 0.812 | 0.889 | 670 | 778 | 821 | 1213 | 95 |
G-4b | Glass + C | 4.038 | 98.9 | 1.772 | 0.661 | 0.800 | 0.881 | 736 | 839 | — | 1261 | 86 |
G-4c | Glass + C | 4.083 | 99.8 | 1.774 | 0.660 | 0.803 | 0.881 | 738 | 834 | — | 1260 | 81 |
G-4d | Glass + C | 4.093 | 101.1 | 1.780 | 0.659 | 0.807 | 0.894 | 739 | 827 | — | 1244 | 72 |
G-4e | Glass + C | 4.106 | 102.5 | 1.783 | 0.658 | 0.812 | 0.905 | 730 | 816 | 960 | 1257 | 72 |
G-3e | Glass + C | — | 96.5 | — | 0.666 | 0.800 | — | 736 | 860 | — | 1238 | 85 |
G-4eMg | Crystallized | — | 101.5 | — | 0.647 | 0.804 | — | 690 | 722 | 801 | 1244 | <50 |
G-4eZn | Crystallized | — | 109.6 | — | 0.581 | 0.804 | — | 658 | 715 | 750 | 1238 | <30 |
G-4eSr | Glass + C | — | 115.8 | — | 0.670 | 0.828 | — | 726 | 814 | — | 1228 | 77 |
G-4eBa | Glass | 4.546 | 129.7 | 1.789 | 0.686 | 0.837 | 0.964 | 710 | 839 | 861 | 1178 | 106 |
This is due to the superposition of the Ga–O bond vibration with the Ge–O stretching vibrational bands in the [GeO4] tetrahedron peaking at around 745 cm−1.23 On critically observing the shift as well as the enhancement of the 636 cm−1 band in Series 2 glasses, it is observed that both peak shifting and intensity enhancement is slowly retarded with the increase in GeO2 content up to G-4d glass and then this trend is reversed for the G-4e glass sample. Usually, the structure of germinate glasses is complicated, since unlike SiO2, GeO2 exhibits two kinds of structural units, [GeO4] and [GeO6].15,17 The observed changes in infrared spectra on GeO2 inclusion can be correlated with the relative enhancement in octahedral [GeO6] units in higher GeO2 containing glasses (Series 2). This combined effect of increased proportions of both [GaO6] and [GeO6] units in Series 2 glasses reduces the network connectivity and results in partial devitrification of glasses, as observed in the present case.
This observation suggests that to obtain a good glass in the present Ga2O3 based composition, more Ga3+ should be in a tetrahedral coordination, which needs a more basic environment to facilitate an effective charge compensation of the [GaO4]− tetrahedron. The above assumption is strengthened by the observation of complete crystallization of the melts having MgO (G-4eMg) and ZnO (G-4eZn) as network modifiers in Series 3, which gives a more acidic nature to the network, and thus poor charge compensation. In this series, the addition of SrO to the glass composition eases this problem to some extent, but still the cast glass exhibits a small amount of partial crystallization. However, the BaO containing composition showed an excellent glass forming ability, resulting in clear glass without any crystallization, as can be noticed from the image of G-4eBa glass presented in the inset of Fig. 2a. Fig. 2a shows the infrared reflectance spectra of G-4eBa glass in comparison with the G-4e glass sample. It can be realized that the intensity of the 665 cm−1 vibrational band shows significant enhancement on BaO inclusion in the network. Furthermore, the shoulder band at 745 cm−1 is more pronounced in G-4eBa glass compared to G-4e. These changes indicate a relative enhancement in the tetrahedral coordination of both the Ga3+ and Ge4+ cations, which is confirmed by the deconvoluted components of vibrational bands depicted in Fig. 2b and 2c, respectively. The increased fraction of tetrahedral coordination units in G-4eBa glasses allows for better network linking and thus yields good glass formation.
![]() | ||
Fig. 2 (a) Infrared reflectance spectra of G-4e and G-4eBa glasses along with their deconvoluted components (b) and (c), respectively (inset: image of the G-4eBa glass sample). |
From all these observations, it can be understood that the theoretical basicity based on Pauling's electronegativity exhibits some definite correlation with the glass formation that has been achieved in the worked out compositions in the present investigation. The thermal analysis of the prepared glass samples also supports the same trend, as can be noticed from the glass stability factor (Tx − Tg) listed in Table 3 along with the glass transition temperature (Tg), crystallization peak temperatures (Tp1, Tp2) and melting temperature (Tm). The data clearly indicates that the glass stability factor is highest for the G-0a and G-4eBa samples, which practically could form clear glasses by an ordinary melt quenching technique. Another interesting observation of thermal analysis is that there is a decreasing trend in the glass transition and melting temperatures in the studied series of glasses, which can be attributed to the decreased bond strength with the modifications in chemical composition. The melting temperature is almost 165 °C lower in G-4eBa glass compared to the G-0a glass, which is one of the advantages of the gallate glass system, requiring a lower melting temperature despite maintaining similar thermal stability over crystallization. The differential thermal analysis (DTA) curves of some selected glasses are depicted in Fig. 3. It is evident that the crystallization peak profile undergoes striking changes with the composition modifications. The broad crystallization peak observed in G-0a glass becomes sharper in G-4a and G-4e glasses and then again becomes broad in G-4eBa glass. Such sharpening of the crystallization peak may arise due to the reduced viscosity of glass on Ga2O3 inclusion. But, on addition of Ba2+ into the network, the viscosity of the glass may again increase, due to the rise in network forming tetrahedral units in G-4eBa glass, accompanied by the presence of larger cations (Ba2+) in the network creating obstacles in the process of crystallization.24 This effect has been experienced practically in the present glass preparation.
![]() | ||
Fig. 3 DTA profiles of G-0a, G-4a, G-4e and G-4eBa glasses. |
![]() | ||
Fig. 4 Infrared absorption spectra of G-0a, G-4a, G-4e and G-4eBa glasses (inset: infrared transmission of glasses). |
Similarly to the infrared band edge, the UV band edge also exhibits a red shift on Ga2O3 inclusion in the glass, as can be seen from the UV-Vis-NIR absorption spectra of the glasses shown in Fig. 5. This shift in UV cut-off is expected, owing to the comparatively lower bond strength of Ga–O bonding compared to Al–O bonding. The optical transparency of these glasses is observed to be more than 80%. Apart from the UV band edge shift, a small absorption band could be detected in the absorption spectra of gallate glasses peaking at around 450 nm. The inset shows an enlarged view of absorption spectra, giving a clearer profile of the absorption band. It can be seen that the intensity of this absorption band increases from G-4a to G-4eBa glass, imparting a slightly brown colored tint to the glass. A similar observation has also been reported by Shaw and Shelby14 in barium gallosilicate glass, where they observed the appearance of a broad absorption spectrum in the visible region on increasing the BaO content in the glass.
![]() | ||
Fig. 5 UV-Vis-NIR optical absorption spectra of G-0a, G-4a, G-4e and G-4eBa glasses (inset: enlarged view of the spectra in the UV-Vis region). |
Since the raw materials used in the glass preparation were highly pure and the subsequent processing was carried out by taking sufficient precautions to avoid any contamination of transition or rare earth ion impurities, the only culprit behind this absorption band can be the surface plasmon resonance (SPR) of metal nanoparticles generated in the glass matrix.25 So, to confirm the presence of metal nanoparticles in the glass matrix, X-ray diffraction patterns of glass samples were recorded, but it did not show any distinct diffraction pattern, suggesting the glasses are X-ray amorphous. However, TEM analysis of the glass sample displayed the presence of nanoparticles scattered throughout the glass matrix, as seen from the TEM image of G-4eBa glasses presented in Fig. 6a. From the TEM image, it is clear that the volume percentage of nanoparticles in the matrix is smaller, which might be a reason behind the absence of distinct diffraction peaks in the XRD pattern. However, the selected area electron diffraction (SAED) analysis under TEM revealed a lattice diffraction pattern, confirming the presence of nanoparticles, as shown in Fig. 6b. This pattern has been identified as being due to the metallic Ga0 nanoparticles (JCPDS file No. 02 0480) exhibiting different diffraction planes, as shown in Fig. 6b. The high resolution transmission electron microscopy (HR-TEM) image of the nanoparticle is shown in Fig. 6c along with the corresponding FFT pattern (Fig. 6d). The HR-TEM image shows a lattice plane with inter-planar spacing of 3.2 Å, further confirming the presence of Ga0 nanoparticles in glass. Germanium, Ge0, also exhibits the same inter-planar spacing of 3.2 Å in the cubic lattice. Also, the diffraction planes of metallic Ge0 match well with the SAED pattern shown in Fig. 6b (JCPDS file No. 03 478). However, Ga0 exhibits more diffraction planes compared to Ge0 and all of the diffraction planes of Ga0 could be matched, confirming its presence in the glass. However, the diffraction pattern and HR-TEM still also hint at the formation of Ge0 nanoparticles and their presence in the glass, accompanied by Ga0 nanoparticles.
![]() | ||
Fig. 6 (a) Bright-field TEM image of G-4eBa glass, showing nanoparticles in the glass; (b) selected area electron diffraction (SAED) pattern; (c) HR-TEM image of nanoparticles; and (d) fast Fourier Transform (FFT) pattern. |
In fact, the reducing environment of basic glasses is ideal for the formation of metal nanoparticles. The high basicity (experimental) of present glasses can spontaneously reduce the metal cations to metallic atoms, which then agglomerate to form nanoparticles. The tendency of such reduction of metal cations depends on the standard reduction potential (E0) of the reaction. Actually, the reduction potential is higher for germanium than for gallium ions, as illustrated in the following reactions:26
Ge4+ + 4e− → Ge0 (E0 = +0.124 V) |
Ga3+ + 3e− → Ga0 (E0 = −0.549 V) |
So, this indicates that Ge4+ is more prone to form nanoparticles over Ga3+, and thus the coexistence of Ge0 nanoparticles along with Ga0 in the glass can be anticipated.
The deconvoluted components of the absorption spectrum of G-4eBa glass have been plotted in Fig. 7, representing the SPR bands of metal nanoparticles as well as the base glass spectrum. The spectrum indicates the presence of two distinct SPR peaks centered at 3.41 eV and 2.87 eV, respectively.27,28 The combined SPR spectrum of the metal nanoparticles is presented in the inset, showing strong absorption in the UV-visible region of the spectrum. Although the presence of metal nanoparticles imparts brown colored tint in the glasses, their large surface area to volume ratio gives rise to an enhanced Coulombic field, advantageous for superior luminescent properties of dopant rare earth ions. It is also possible to minimize this SPR absorption band by introducing some suitable oxidizing agents like Sb2O3 into glasses, which oxidizes the metal nanoparticles and thus gives clear glass. This effect of oxidizing agents on metal nanoparticles in gallate glasses will be discussed in our next report.
![]() | ||
Fig. 7 UV-Vis absorption spectra of G-4eBa glass along with its deconvoluted components (inset: SPR spectrum of nanoparticles). |
![]() | (1) |
![]() | (2) |
Another approach of direct experimental estimation of optical basicity is based on the determination of the refractive index (n), which in turn represents the electronic polarizability (αO2−).21
![]() | (3) |
![]() | (3a) |
This journal is © The Royal Society of Chemistry 2012 |