A review of the structures of oxide glasses by Raman spectroscopy

Avadhesh Kumar Yadav* and Prabhakar Singh
Department of Physics, Indian Institute of Technology, Banaras Hindu University, Varanasi-221005, India. E-mail: yadav.av11@gmail.com

Received 4th July 2015 , Accepted 13th July 2015

First published on 21st July 2015


Abstract

The family of oxide glasses is very wide and it is continuously developing. The rapid development of advanced and innovative glasses is under progress. Oxide glasses have a variety of applications in articles for daily use as well as in advanced technological fields such as X-ray protection, fibre glasses, optical instruments and lab glassware. Oxide glasses basically consist of network formers, such as borate, silicate, phosphate, borosilicate, borophosphate, and network modifiers such as alkali, alkaline earth and transition metals. In the present review article, Raman spectroscopy results for the structures of borate, silicate, phosphate, borosilicate, borophosphate, aluminosilicate, phosphosilicate, alumino-borosilicate and tellurite glasses are summarized.


1. Introduction

1.1. Brief history of glass

The exact history of glass is not known but there is evidence confirming that it was used around 3500 BCE in Mesopotamia. The available archaeological evidence indicates the presence of glass articles in coastal north Syria, Mesopotamia and Ancient Egypt. The use of glass in South Asia began in 1730 BCE.1 Evidence of the existence of glass has also been found in Hastinapur and an archaeological site in Takshashila, ancient India. Glass objects have been developed in the Romanian Empire and during the Anglo-Saxon period in domestic, industrial and funerary contexts. In China, the development of glass began late but nowadays, glass plays a peripheral role in arts and crafts. The glasses from that period were made of barium oxide (BaO) and lead.2 The tradition of the use of lead–barium glasses disappeared at the end of the Han Dynasty (AD 220). Soda glasses with high alumina content were rarely used in the Mediterranean area or in the Middle East. Soda-lime, borosilicate, glass fibre, lead, alkali-barium silicate, aluminosilicate, vitreous silica and pyrex glasses were investigated from time to time. The large-scale development of glass technology began mainly in the 14th century.3 Recent research on glasses is based on advanced glasses used for safety systems such as X-ray protective glasses, fibre glasses, high-refractive-index glasses for optical instruments, photo-chromatic and high strength glasses.

In the last few decades, oxide glasses have attracted the attention of researchers and scientists due to their excellent properties, which are very useful in many applications. They are easily shaped due to their lack of underlying crystal structure. Glass covers on solar units and photovoltaic units improve their performance. In turbines, glass fiber-reinforced composites are used for the storage of wind energy. Therefore, boron-free glasses are commonly used.4 Oxide glasses may be used in smaller power supplies, including dielectrics for super-capacitors, sealants for high-temperature solid oxide fuel cells (SOFC) and electrolytes for electrochemical devices.5 W. H. Zachariasen investigated silicate glasses and he found that such glasses consist of silicon tetrahedrally coordinated to 4 oxygen atoms.6 Phosphate glasses were investigated by P. B. Price et al.7 Borate glasses have been investigated in various glass systems: B2O3xLi2O3,8 xB2O3–(1 − x)Li2O3,9,10 (30 − x)Li2O–xK2O–10CdO/ZnO–59B2O3,11 60B2O3–(20 − x)Na2O–10PbO–10Al2O3:xTiO2:yNd2O3,12 PbO–Bi2O3–B2O3,13 Bi2O3–B2O3–ZnO–Li2O,14 and Gd2O3–MoO3–B2O3.15 Silicate glasses were investigated in the presence of Na2O, Li2O, K2O and CaO by various research groups.16–25 Phosphate glasses were investigated over a long period of time in many systems. Calcium, sodium, potassium, iron, gadolinium, zinc, lead, chromium, molybdenum, lanthanum, bismuth, tungsten, samarium, copper, cadmium, barium, and silver containing phosphate glasses have been investigated for various applications.122–165 Soon after the investigation of single-network former glasses, double network former glasses were also investigated for the composite network structures of borate, silicate and phosphate, which are named as borosilicate and borophosphate glasses. In recent years, the structures of SnF2–SnO–P2O5 glasses have been studied. In this study, the glass transition temperature was found to be very low.26 CeO2 in cerium iron borophosphate glasses increases the glass transition temperature and also increases the thermal stability.27 The thermal stability of Na2O–FeO–Fe2O3–P2O5 was found to be poorer than sodium-free iron phosphate glasses.28 TiO2 is useful for increasing the thermal stability of BaO–Li2O–diborate glasses.29 Low melting glasses were prepared in the glass system K2O–MgO–Al2O3–SiO2.30 For the purpose of the immobilisation of nuclear waste, a few borosilicate glasses were also investigated in the glass system 55SiO2–15B2O3–5Al2O3–5CaO–(20 − x)Na2O–xCs2O.31

Glasses may be defined as inorganic solids, which are amorphous in nature. They are classified in several groups such as metallic alloys, ionic melts, aqueous solutions, molecular liquids and polymers. They have attracted the attention of researchers and scientists in recent years due to the demand for glasses in many significant applications in the fields of energy science and photonics. There is still a demand for substantial development of glasses.

1.2. Brief chemistry and physics of glasses

Glasses are disordered materials due to the lack of periodicity in their crystal structures.32 They are formed by the cooling of melted inorganic products to a rigid condition without crystallization.33,34 Several materials, such as organic polymers and metal alloys, also have amorphous structures but not all of them are glasses.35 The formation of glass requires critical cooling rates.36–39 Thermodynamically, glasses are non-equilibrium materials because their properties are a function of pressure, temperature and composition. They always approach a nearby metastable state.40–43 Glass formation is possible if the glass contains network formers. Glasses are formed by network formers such as borate, silicate, phosphate, borosilicate and borophosphate. It has been found that the best glass formers have electro-negativity values of 1.7–2.1 on the Pauling scale.44 Network modifiers are used in glass formation to modify the glass properties. The glass network is modified by alkali and alkaline earth metal modifiers such as ZnO, PbO, TeO2, Bi2O3, MgO, CaO, SrO, BaO, CeO2, Cr2O3, La2O3, CdO, CuO, Li2O, Na2O, K2O, Al2O3 and MgF2. Alkali metal ions are easily thermally activated and can move from one site to another within a glass. This type of movement of alkali metal ions within a glass structure enables the replacement of alkali metal ions near the surface of a glass by other ions of the same valence.45,46 Several local packing arrangements are available in glasses, which enable the tuning of their properties through composition and processing. This permits the creation of highly homogeneous materials on a macroscopic scale.47 Zachariasen established a theory that explains the criteria for vitreous structures with highly directed three-dimensional bond arrays. Glass structures comprise microcrystallites of ∼20 Å in size. Glass structures lie between a melt state and a glassy state.48,49 The glassy network structure may be effectively understood by infrared and Raman spectroscopy. Certain viscosity models are used to explain the glass properties on a temperature scale. The five temperatures are named as standard points (associated with the viscous flow of glass), strain point (maximum temperature supported by glass), annealing point (relieves the internal stresses), softening point and working point (most glass forming operations acting at this temperature).50
1.2.1 Classification of the preparation method of glasses. Glasses are mainly synthesized by the melt-quench method, sol–gel method and chemical vapour deposition.51
1.2.1.1 Glass by melt-quench technique. The melt-quench technique is the first discovered technique for the preparation of glasses. In the melt quench method, the inorganic raw materials are accurately weighed according to composition. Then, the raw materials are mixed in a mortar with a pestle or ball mill in acetone media. The well-mixed and dried ingredients are placed into a high-grade alumina or platinum crucible and the crucible is placed in a programmable temperature-controlled furnace for melting. The homogeneous melt is quenched in an aluminium mould. In this way, the glass is formed by the melt quench method. The formed glasses are annealed at relatively low temperatures to the glass casting temperature to remove the residual internal stresses that are produced due to the temperature gradient during the forming and subsequent cooling.52 The melt quench method is advantageous for obtaining materials of a large size compared to single crystal or polycrystalline ceramics. It has another advantage as it provides good flexibility of composition over the chemical vapour deposition or the sol–gel method. This method also has certain disadvantages. For example, this method is not suitable for the preparation of glasses with an ultra-high purity that are used in optical communication systems. Certain impurities from the crucible or furnace materials are also added to the glass during this method. Glasses, in which refractory materials such as SiO2, TiO2, Al2O3 and ZrO2 are used, are difficult to prepare by this method due to the need for extremely high temperatures.
1.2.1.2 Glass by sol–gel method. In the sol–gel method, a colloidal solution (sol) of raw materials is converted into a gelatinous substance (gel). In this method, there is no requirement of melting and quenching. Fine-grained gel bulks consisting of various organic compounds are annealed at a particular temperature and thereafter the annealed gel is sintered at a higher temperature than the annealing temperature. During this process, there are two reactions taking place, namely, hydrolysis and condensation reactions.53
1.2.1.3 Glass by chemical vapour deposition method. The chemical vapour deposition (CVD) method for the preparation of bulk glasses was developed in the early 1940s. The formation of the glass material is followed by thermally activated homogeneous oxidation or hydrolysis of the initial metal halide vapour. The oxidation or hydrolysis reaction is activated by an oxy-hydrogen flame or oxygen plasma. In this method, the initial components are liquid at room temperature and their boiling temperature is very low compared with alkaline halides, alkaline earth, transition metals or rare-earth elements. The purification of raw materials is achieved by repeating distillation below the melting points. This method is advantageous for the preparation of ultra-high purity glasses.54

The structural study of glass networks is important for understanding the composition dependence of oxide glass properties. Various models are proposed for correlating the different properties with the structure and composition of a glassy system. Phillips and Thorpe investigated the compositional dependence of glass properties by analyzing the topology of the glassy network. This approach is based on the comparison of the number of atomic degrees of freedom with the number of interatomic force field constraints. The tendency for glass formation would be maximized if the number of degrees of freedom exactly equalled the number of constraints in the glassy network.55 Gupta and Mauro demonstrated the generalization of the Phillips and Thorpe approach in their study.56,253

This review article is focused on the structural study of various oxide glasses and on how the structure of glassy networks is modified by various dopants such as alkali, alkaline and transition metals. The composition-dependent study of oxide glasses is also summarized in the present review article. This review article is very useful for understanding the formation of glassy networks and their modification by various dopants.

1.3. Brief introduction to glassy networks

A lot of study has been done on glassy materials up to now. Glassy materials consist of various network formers such as borate, silicate, phosphate, borosilicate and borophosphate. The borate network incorporates the vibrations of isolated diborate (B4O72−) groups, breathing of O-atoms, loose diborate vibrations, chain or ring type meta- and pentaborate (B5O8) groups, symmetric breathing vibrations of six-member borate rings with one or two BO3 triangles replaced by a (BO4) tetrahedral, symmetric breathing vibrations of boroxol rings, triborate (B3O5), tetra-borate (B8O132−), ortho-borate (BO33−), pyro-borate (B2O54−) groups and linked [BO3]3− triangles with one free vertex.45,57 The borate glasses have significant importance in several applications in thin amorphous films for battery application, bioactive glasses for tissue engineering, nuclear waste disposal, photonic applications, development of tuneable or short pulse lasers, optical fibre amplifiers and fibre lasers.58–61 The borate networks are shown in Fig. 1.
image file: c5ra13043c-f1.tif
Fig. 1 The network units of borate glasses107 (reproduced with permission from IOP publishing).

The silicate network former is also significant in glass formation; however silica is not easily fusible and does not retain viscosity for a long time. Thus, it cannot easily form glass. Fluxes are used that help in the formation of glassy networks.62 Silicate glasses consist of tetrahedral 6-coordinated 4 oxygen atoms. 6-Coordinated silicon is also found in a few crystalline materials.63 Silicate networks are found in the form of structural units of orthosilicate, Si–O stretching vibrations of tetrahedral silicate units, symmetric stretching vibrations of silicate tetrahedra, inter-tetrahedral Si–O–Si linkages and structural unit QnSi (n = 0 to 4), where QSi represents the tetrahedral unit and n is the number of bridging oxygens (BO) per tetrahedron.64–66 Silicate glasses are generally used in windows, lenses, mirror substrates, crucibles, trays and boats, UV-transmitting optics (synthetic fused silica), IR transmitting optics and metrology components. The network units of silicate glasses are shown in Fig. 2. Similar to silicate networks, phosphate networks are formed from various structural units of phosphate. Phosphate glasses are formed at low temperatures compared to silicate glasses. The vibrational units of phosphate networks are QnP groups of PO4 tetrahedra (n = number of bridging oxygens per PO4 tetrahedron), symmetric stretching of P–O, asymmetric stretching vibrations of P–O, symmetric vibrations of P–O–P, bending vibrations of phosphate polyhedra and O–P–O bending modes. Phosphate glasses have numerous applications in photonics, biomedical applications, solid state electrolytes and fibres.67–69 Phosphate glasses are very sensitive to ionizing particles. This property of phosphate glasses enables their use in the identification of heavy ions.70 The phosphate networks are shown in Fig. 3.


image file: c5ra13043c-f2.tif
Fig. 2 The network units of silicate glasses16 (reproduced with permission of the Mineralogical Society of America).

image file: c5ra13043c-f3.tif
Fig. 3 Network units of phosphate glasses164 (reproduced with permission from Elsevier).

Not only single network former glasses but also double network former glasses were investigated. Borosilicate glasses are one of the double network former glasses. The co-existence of borate and silicate networks is associated with borosilicate glasses. The interlinking of some borate and silicate networks is also possible in such glasses. The high frequency band of >850 cm−1 is related to silicate networks, QnSi, whereas mid frequency bands (400 to 850 cm−1) are associated with ordered superstructures, reedmergnerite [BSi3O8] and danburite [B2Si2O8]2−. Bands ranging from 300 to 500 cm−1 are characteristic of mixed stretching and bending modes of Si–O–Si units, whereas bands in the range 550–850 cm−1 are caused by ring breathing modes. The bands at 670, 770 and 808 cm−1 are the signature of tetraborate groups, four- and three-coordinated boron in diborate and boroxol rings, respectively. Borosilicate glasses are mainly used in lab equipment, high quality medical devices (e.g. ampoules, dental cartridges and veterinary tracking devices), space exploration devices and electronics (microelectromechanical systems).71–74

The properties of borates can be altered by a suitable modification of their chemical compositions within a relatively broad concentration region of their constituents in phosphate glasses. Addition of borate (B2O3) to phosphate glasses improves their mechanical properties and chemical resistance to atmospheric moisture. Borophosphate glasses include networks of both borate and phosphate along with various linking vibrations. These glasses are very useful for potential application in solid electrolytes, glass solders and glass seals.75,76 Phosphosilicate glasses consist of networks of phosphate as well as silicate. Phosphosilicate glasses are used in forming Raman fibre lasers and amplifiers. Phosphosilicate glasses are useful in developing high refractive index gratings.77,78 Aluminosilicate glasses contain units of silicate and aluminate networks. These are very important in geological applications due to their highly refractory nature.79,80 Aluminium-based glasses have great technological importance due to their strong and ductile nature. They play an important role where metallic glasses failed.81 Alumino-borosilicate glasses are formed from networks of aluminate, borate and silicate. These glasses have a wide range of applications in optical, thermal, electrical and mechanical devices. Along with these properties, the addition of Al2O3 to borosilicate matrices improves chemical durability, decreases the thermal expansion coefficient, electrical radiation resistant and widens the scope of the applications of borosilicate glasses.82 Doping of alkali and alkaline earth metals in these glasses modifies the parent network. Rare-earth ion-doped glasses have importance in applications such as optical fibre lasers, amplifiers and scintillating glasses. Sm2+ doped glasses are advantageous in spectral hole burning.83–88

2. Principles of Raman spectroscopy

Raman spectroscopy is a powerful tool for structural analysis and was developed by Sir C. V. Raman and Kirishnan in 1928.89 They pointed out in their study that Raman intensity is 10−6 to 10−9 times less than Rayleigh scattering. Such a low intensity can be produced by lasers.90 When laser light is incident on vibrating molecules, the energy of photons may be changed. The exited molecules or atoms return to different states. The change in energy between the original state and the new state gives the shift of the emitted photon. The excited molecule goes to a higher or lower frequency than the original state and this phenomenon is known as Stoke Raman scattering, (Stoke shift) or anti-Stoke Raman scattering (anti-Stoke shift), respectively.91,92 A detailed description of the working principle is shown in Fig. 4. The chemical composition and structure of molecules influence the modified scattering and no two Raman spectra are exactly the same. Thus, Raman shift may be useful in distinguishing the structures of different constituents and molecules.93,94
image file: c5ra13043c-f4.tif
Fig. 4 Working principle of Raman spectroscopy.

Raman spectroscopy is used for the determination of the structure, environment and dynamics of glassy materials. Furthermore, the portability of the technique allows its use in on-line process monitoring over other techniques like infrared spectroscopy, NMR and X-ray diffraction.95 Raman and infrared (IR) spectroscopy are used as complimentary techniques and both techniques differ slightly according to their selection rules. IR spectra arise from a change in the dipole moment, whereas Raman bands arise from a change in the polarizability. A few transitions are allowed in Raman spectroscopy but are forbidden in IR spectra, so Raman spectroscopy gives particularly useful information in that situation. As with Raman spectroscopy, in FT-IR spectrometry and UV-Vis spectrometry, the characteristics of a sample are analysed using the absorbance (or transmission) spectrum. In IR spectroscopy, molecules like water or acetone are strong IR absorbers, whereas sodium chloride and potassium bromide are very weak absorbers (window materials). Thus, functional groups such as hydroxyl or carbonyl are easily detected, whereas window materials are not easily detected. However, the situation is different for Raman spectroscopy. In Raman spectra, the ordinate axis normally has arbitrary rather than absorption or %T units because it is simply a measure of the number of scattered photons captured by the detector at any particular frequency. If the power of the incident laser on the sample is varied, then the intensity of the Raman spectrum will vary accordingly. Therefore, the peak height in a Raman spectrum is not simply a function of the sample thickness, unlike IR spectra, which depend on sample thickness.

In UV/Visible and IR spectroscopy, the spectral contribution from the instrument is removed by using a reference beam or by subtracting the background spectrum. However, in Raman spectroscopy, any contribution or variation due to the instrument is considered a single-beam. Due to this reason, Raman spectroscopy plays a vital role in quantitative analysis. Unwanted contributions can be minimized by Raman spectroscopy.

X-ray fluorescence (XRF) is also useful for elemental analysis of glassy networks, whereas vibrational networks cannot be significantly understood by XRF. Moreover, a very high energy is required for its operation. X-ray powder diffraction is also used for understanding the nature of glassy materials but this method is particularly useful for the network analysis of glassy materials.96–98

3. Discussion

3.1. Borate glasses

A typical Raman spectrum of lithium borate glass is shown in Fig. 5.99 A Raman study of lithium borate glasses depicts networks of B(4)–O vibrations of tetracoordinated boron at 520.5, 769 and 784 cm−1, vibrations of B5O8(OH)4 units at 555.5 cm−1 and vibrations of tricoordinated boron near 919 cm−1. The characteristic intense peaks at 500.5 and 884.5 cm−1 correspond to the vibrations of B(3)–O in B(OH3). The sol-derived borate glasses show the presence of boroxol rings as a band found at 807 cm−1.99
image file: c5ra13043c-f5.tif
Fig. 5 Raman spectrum of lithium borate glass99 (reproduced with permission from Springer).

Li cations are found in two distinct modes as non-bridging oxygen (NBO) type and bridging oxygen (BO) type sites. The primary network is formed by NBO and BO atoms in tetrahedral borate units, which reside close to NBO atoms. They modify the network by forming clusters around NBO atoms.9 The Raman spectra of xLi2O·(l − x)B2O3 glasses consist of one component at 1030 cm−1 corresponding to B–O bond stretching of BO4 units in diborate polycrystals and others near 900, 940 and 1040 cm−1 corresponding to the same vibration but in tetraborate units. Increasing the concentration of Li2O leads to the formation of pentaborate, tetraborate, diborate, metaborate and pyroborate.10 The spectra of Li2O·2B2O3 fluxures show features of a crystalline nature with strong characteristic bands at 783, 1034 and 1174 cm−1 and some weaker bands between 167 and 723 cm−1. Spectra of lithium metaborate (Li2O·B2O3) show clear bands at 545–582, 765, 955, 1108–128 and 1467 cm−1. When the sample is heat treated above the melting temperature, then bands around 640, 671, 724 cm−1 and two less resolved bands in the range 1475–1495 cm−1 were also observed.100 Raman spectra of Li2O·4B2O3 at room temperature indicate the formation of boroxol rings containing one BO4 tetrahedron. With increasing temperature, the stability of the BO3 tetrahedron decreases and it is finally destroyed.101 The networking structure of Li2O–PbO–B2O3 is characterised by [BO3/2]0, [B2O4/2] and the B–O–B bending mode at 1485, 967 and 765 cm−1, respectively. The band due to diborate units at 535 cm−1 is completely eradicated by the addition of PbO, whereas a few additional bands in the region of 280–300 cm−1 are incorporated. The emergence of peaks in the region of 1200 cm−1 is associated with the symmetric stretching of B3 units and is generated during the incorporation of Pb–O into the structure. The band at 640 cm−1 was attributed to a bending mode of the Pb–O–B links in the new structure.102 Raman spectra in the range 150–1600 cm−1 of (30 − x)Li2O − xK2O − 10CdO/ZnO − 59B2O3 (x = 0, 10, 15, 20, and 30) doped with 1 MnO2/1 Fe2O3 showed four regions of bands as follows: (i) 190–600 cm−1; (ii) 600–820 cm−1; (iii) 900–1200 cm−1 and (iv) 1200–1600 cm−1. Strong peaks around 775, 650, 500 cm−1 and a weak peak at ∼805 cm−1 are present, which are attributed to the localized breathing motions of oxygen atoms in the boroxol ring. The peaks around 950 and 1110 cm−1 are due to diborate groups in the structure. The peaks in the high frequency region are attributed to BO2O triangles linked to BO4 units and BO2O triangles linked to other triangular units.103

The polarized Raman spectra of sodium borate glasses showed the short range structure of BO3 and BO2O units. In HH (parallel to the inherent polarisation of the excitation laser) spectra, some strong Raman bands were found in the frequency region 700–850 cm−1. The strongest band due to the boroxol ring was observed at 805 cm−1 in the spectrum of glass. A broader and asymmetric band at the lower frequency side was attributed to the increase of Na2O. A band near 1500 cm−1 was found in the spectra and this shifts towards the lower frequency side with increasing Na2O concentration, whereas VH (perpendicular to the inherent polarisation of the excitation laser) spectra showed significant differences compared to HH spectra. VH spectra showed some more bands at 773, 730 and 670 cm−1.104 The spectra of B2O3–Na2O glasses depict five major peaks at 480, 660, 770, 806 and 920 cm−1. The peaks due to the isolated diborate groups and the ring-type metaborate groups polymerized by BO3 and BO4 units remain unaffected by the addition of Al2O3, whereas the peak at 770 cm−1 was shifted towards the low frequency side in the spectrum. This meant that BO4 units were consumed and the number of boroxol rings (B3O6)3− gradually increased. Another new vibrational peak at about 920 cm−1 appeared in 6 mol% Al2O3. This peak is characteristic of an orthoborate-type structure containing BO3 units, which can be transformed by the linking of pentaborate and tetraborate groups. The presence of orthoborate groups and the absence of pentaborate groups proves the feasibility of the transformation of pentaborate to orthoborate groups as a result of addition of Al2O3 to the structure. This suggests that the addition of Al2O3 transformed high polymer borate units like pentaborate into low polymer groups such as boroxol rings, triborate and orthoborate groups.105 There are four Raman bands in the glassy matrix of 60B2O3·(20 − x)Na2O·10PbO·10Al2O3:xTiO2:yNd2O3. The bands at 755 and 772 cm−1 are assigned to the chain-type metaborate groups and the symmetric breathing vibration of six membered rings with one BO4 tetrahedron, respectively. The position of the boroxol ring was found at 797 cm−1 in this system. When TiO2 concentration was increased, back conversion of BO4–BO3 took place. The incorporation of Ti4+ led to ∼40% smooth reduction of the BO4 groups due to the BO4–BO3 back conversion effect.12

The Raman spectra of lead borate glasses in glass compositions, 80% PbO, 20% Bi2O3, 90% PbO, and 1% B2O3 showed a very sharp band at 130 cm−1 in the first composition and shifted to 139 cm−1 in the second composition. A few broad bands centred at 710, 905, 1024, 1239 and 1308 cm−1 are also observed in the pattern. In 1 mol% MoO3 doped lead borate glass, two other peaks at 1927 and 3341 cm−1 are introduced. On increasing the doping concentration of MoO3, the band near 1255 cm−1 becomes weaker. A sharp band at 133 cm−1 in the 2 mol% Mo-containing lead borate glasses was found. The Raman peak in un-doped lead borate glasses appears at 139 cm−1 and this band was located at 144 cm−1 in crystalline PbO. This suggests the presence of Pb2+ in high lead borate glasses in the form of PbO4 groups. The sharp band observed at 305 cm−1 in lead borate glass, containing 5 mol% of MoO3, is not observed in un-doped lead borate glasses, whereas it appeared at 296 cm−1 in crystalline lead oxide and 282 cm−1 in crystalline molybdenum trioxide MoO3. This can be predicted as the Raman band at 305 cm−1 originates from a different site of Pb2+. The weak broad Raman bands in MoO3–lead borate glasses at 3311–3341 cm−1 are assigned to molecular water.106 Sodium-doped lead borate glasses in the glass system xNa2O·(100 − x) (10PbO·90B2O3) are made from structural units of boroxol rings and growth of triborate.107 The spectra of bismuth-containing lead borate glasses showed Raman peaks at 400, 550, 710, 920 and 1220 cm−1. The peaks for heavy metal oxides appear in the range 380–580 and 650–950 cm−1 due to the bridging anion modes and the non-bridging anion modes, respectively. The polarization behaviour of bismuth and lead cations is similar due to these having similar atomic weights. The first two peaks correspond to the bridge-anion motion due to symmetric stretching of Bi–O–Bi and Pb–O–Pb combined with Bi–O–Pb.108 The Raman spectra for xFe2O3·(100 − x)[3B2O3·0.7PbO·0.3Ag2O] glasses (Fig. 6) confirm the presence of boroxol rings [B3O4.5], pyro-B2O54−, ditri-[B3O8] and dipenta-borate [B5O11] groups as Raman bands are present at ∼770, ∼800, ∼1040 and ∼1340 cm−1. A wide envelope appears around ∼465 cm−1 due to isolated diborate groups as well as Pb–O linking vibrations and the envelope at ∼700 cm−1 is a characteristic of symmetric breathing vibrations of metaborate rings. The silver ions peak shifted from 806 to 800 cm−1. Increasing the Fe2O3 content in lead borate glasses containing silver decreases the intensity of peaks at 770 and 800 cm−1 and increases the intensity of the bands at 1040 and 1340 cm−1.109 The Raman bands in PbO–BaO–B2O3 glasses appear at 770, 806, 1230, 1280 and 1450 cm−1.110 The substitution of cadmium by lead in lead borate glasses was represented by Raman bands in the range of 200–300 cm−1. The band near 840 cm−1 is not found and this suggests that CdO4 is not formed in the glass network. Raman bands were found at 3691, 1275, 927, 720, 623 and 219 cm−1 in the spectra of glass with composition 30PbO2·20CdO·50B2O3.111


image file: c5ra13043c-f6.tif
Fig. 6 Raman spectra for xFe2O3·(100 − x)[3B2O3·0.7PbO·0.3Ag2O] glasses.109

Barium borate glasses are mainly characterized by two bands near 460 and 805 cm−1. With the addition of 0.5 mol% Mn into the glassy matrix, one more band at about 990 cm−1 appeared and the intensity of this band increases with increasing concentration of Mn, i.e. MnO is helpful for the formation of the orthoborate group. The addition of up to 3 mol% Mn ions causes the intensity of the band at ∼460 cm−1 to increase and thereafter it decreases. Two new bands at ∼630 and 700 cm−1 are also introduced with the addition. The band at ∼630 cm−1 was assigned to the vibration of the ring and chain types of meta- and penta-borate groups, whereas the band at 700 cm−1 is attributed to vibrations of chain or ring-type metaborate groups.112 There are five Raman bands at ∼490, ∼690, ∼800, ∼875 and ∼1250 cm−1 that are present in the 3B2O3–As2O3 glass matrix. The band at ∼490 cm−1 is caused by the vibrations of isolated diborate groups and/or vibrations of As–O bonds. The band near 875 cm−1 decreases progressively with the addition of silver ions. The Raman spectra of xAg2O·(1 − x)[zB2O3·As2O3] (z = 1, 2, 3) glasses showed bands at 490, 685, 803, 880, 960 and 1250 cm−1. The addition of manganese ions to an xAg2O·(1 − x)[2B2O3·As2O3] vitreous matrix leads to structural modifications and to an increase in the degree of disorder, particularly for higher manganese ions concentrations (y ≥ 20 mol%) in the glass system x[(1 − y)Ag2yMnO]·(100 − x)[2B2O3·As2O3].113,114

Raman studies of xSrO·(1 − x)B2O3 (0.2 ≤ x ≤ 0.7) indicate the absence of a band near 806 cm−1, whereas bands are present at 989, 677 and 552 cm−1 for x = 0.2. The Raman spectrum of 0.3SrO:0.7B2O3 is dominated by the band at 671 cm−1, which is assigned to di-pentaborate groups, whereas for 0.4SrO·0.6B2O3, a band at 799 cm−1 is observed. In a glass sample with composition 0.5SrO·0.5B2O3, the band near 744 cm−1 is attributed to six-membered rings with one BO4 tetrahedron. The intensity of this band decreases with the addition of SrO. Addition of metal cations to these glasses introduced a new band near 423 cm−1. The peak near 1413 cm−1 in the undoped sample was shifted towards a higher wavenumber with the addition of Fe2+, Mn2+ and Zn2+ ions.115 The structure of glasses in the glass system 20MO·55Bi2O3·25B2O3 (M = Sr, Ba) was determined by FT-Raman spectroscopy. A broad Raman band at 124 cm−1 was observed in both barium bismuth borate and strontium bismuth borate glasses. A few resolved Raman bands are also present in the region 50–400 cm−1. These vibrations are assigned to Bi atoms in crystalline Bi24B2O39. The external vibration due to α-Bi2O3 appeared in the region 0–150 cm−1, whereas the internal vibrations are in the range 150–500 cm−1. The peak centred at ∼128 cm−1 in the Raman spectrum indicates the presence of [BiO3] and [BiO6] units in the structure of glass and glass ceramics. Raman bands at 696 and 723 cm−1 are associated with the symmetric bending vibration of the [BO3]3− anion in Bi24B2O39 and indicate the violation of the planarity of the [BO3]3− anion. The band due to the boroxol ring at 804 cm−1 disappeared during heat treatment of glasses. The band at 773 cm−1 is predominant with associated bands at 960, 663 and 487 cm−1. These bands are comparatively weak in barium bismuth borate glass ceramics compared to those for strontium bismuth borate glass ceramics. The overtones of the [BO3]3− anion are observed in the range of 1100–1500 cm−1. The strong band centred at ∼1412 cm−1 was assigned to linked [BO3]3− triangles with one free vertex as in BiB3O6 heat treated samples. The heat treatment of glasses increases diborate units as indicated by the increased intensity of the bands at 1117, 1031 and 930 cm−1 (strontium containing) and 1107, 1026 and 925 cm−1 (barium containing).116 The Raman study of Bi2O3–B2O3–ZnO–Li–O showed that Bi can form [BiO3] pyramidal or [BiO6] octahedral units. The symmetric stretching anion motion in an angularly constrained Bi–O–Bi configuration was found in the region 300–600 cm−1. The vibrations of Bi–O–Bi and [BiO6] octahedral units were present near 390 cm−1. The position of this band is shifted towards a lower wavenumber with increasing Li2O. The shoulder around 555 cm−1 was assigned to stretching vibrations of Bi–O–/Bi–O–Zn and its intensity increases with increasing ZnO content. The Raman peak at 220 cm−1 indicates the presence of Zn–O tetrahedral bending vibrations of ZnO4 units in the present glass system.14

The polarized Raman spectra of zinc borate glasses showed features of binary metaborate glasses. The polarized bands appeared at 950, 840 and 1280 cm−1. The two weak bands at ∼770 and 800 cm−1 were also an observed Raman pattern. The unpolarized bands were found at 690 and 1420 cm−1. The strongly polarized shoulder localized at about 250 cm−1 is ascribed to bending modes of ZnO4 units. Doping of Eu3+ in zinc borate increases the intensity of the Raman band. The band at 440 cm−1 is assigned to a Eu3+–O stretching/BO33− vibrational mode.117 The Raman spectra of glasses in the glass system 60B2O3–10TeO2–5TiO2–24R2O:1CuO (where R = Li, Na, K) depict the bands of boroxol rings, vibrations of ring/chain type metaborate units and stretching vibrations of B–O bonds with non-bridging oxygen. The networking modifying behaviour of Na2O and K2O is stronger than Li2O.118

Raman spectra of (1 − x)[3B2O3·K2O]·xTiO2 glasses consist of bands centred at 420, 475, 600, 670, 770, 800, 850, 930, 1230 and 1450 cm−1. The intensity of the band at 770 cm−1 is higher than that of the band near 800 cm−1. The band at 770 cm−1 is significant up to x = 0.2. At higher concentrations of titanium oxide, the intensity of bands at 420, 475 and 670 cm−1 was found to increase. For x = 0.5, two shoulders appear at 850 and 600 cm−1. Further addition of TiO2 introduces the appearance of ring type metaborate groups and loose BO4 tetrahedra. Therefore, the number of non-bridging oxygens increases with increasing titanium concentration and the glass structure becomes more randomized. The addition of K2O in borate glasses changes the boron coordination number from 3 to 4.119 Calcium oxide modified the structure of borate glasses. The structure of borate glasses consists of boroxol groups, a smaller number of pentaborate groups, diborate groups, chain type metaborate groups and pyroborate groups at a lower concentration of calcium oxide, whereas at a higher concentration of calcium oxide, pentaborate, orthoborate and metaborate groups also appear. With an increasing content of calcium oxide, the number of non-bridging oxygen ions increases.120 Yttrium oxide in calcium borate glasses further modified the structure.121 Glasses with compositions of 22.5Gd2O3·xWO3·(47.5 − x)MoO3·30B2O3 (x = 0–40) consist of structural units centred at 344, 840 and 944 cm−1, as shown in Fig. 7. When these glasses were heat treated, a few more sharp Raman bands were also introduced. The Raman peak at ∼994 cm−1 was assigned to symmetric stretching vibrations in (WO4)2− tetrahedral units in the glass composition with Sm2O3 for x = 10.15 Peak assignments of various vibrational units in borate glasses are listed in Table 1.


image file: c5ra13043c-f7.tif
Fig. 7 Raman spectra of glasses in system 22.5Gd2O3·xWO3·(47.5 − x)MoO3·30B2O3 (x = 0–40)15 (reproduced with permission from Elsevier).
Table 1 Assignment of main Raman bands in the spectra of borate glasses
Wavenumber (cm−1) Raman assignments Reference
465–500 Isolated diborate groups 105 and 113–115
535 Diborate units 102
600–650 Symmetric breathing vibrations of metaborate rings 100,103,111,115 and 119
650–660 Pentaborate groups 103 and 105
700–735 Symmetric breathing vibrations of metaborate chains 10,12,100,104,106,108,111 and 116
740–775 Symmetric breathing vibrations of six-membered rings with one BO4 tetrahedron unit 10,12,100,104,105,109,115 and 119
765 B–O–B bending mode 102
800–808 Boroxol ring 12,99,103,105,109,112,112–115,117 and 119
835–840 Pyroborate vibrations 115 and 126
875–1000 Ortho-borate groups 10,105,106,108,111–114,117 and 119
1000–1110 Diborate groups 10,100,102,103,106 and 112–115
1200 Symmetric stretching of B3 units 102
1216–1260 Pyro-borate groups 102,106,108,109 and 119
1300–1600 B–O stretching in metaborate rings and chains 102,106,109,115 and 116
3300–3500 Molecular water 102
0–150 External vibration of α-Bi2O3 116
128 Presence of [BiO3] and [BiO6] units 116
150–500 Internal vibrations of α-Bi2O3 116
220 Zn–O tetrahedral bending vibrations of ZnO4 units 117
250 Bending modes of ZnO4 units 118
390 [BiO6] octahedral units 117
300–600 Angularly constrained Bi–O–Bi configuration 117
555 Stretching vibrations of Bi–O–/Bi–O–Zn 117
440 Eu3+–O stretching/BO33− vibrational mode 118


3.2. Silicate glasses

The networks of silicate glass are shown in Fig. 8. Raman spectroscopy studies of silicate glasses showed the formation of structural units of orthosilicate, silicon–oxygen stretching vibrations of tetrahedral silicate units, symmetric stretching vibrations of silicate tetrahedra, inter-tetrahedral Si–O–Si linkages and the structural unit QnSi, where Q represents the tetrahedral unit and n the number of bridging oxygens (BO) per tetrahedron. For silicon compounds, n varies between 0 and 4, where Si is a central tetrahedral atom ranging from Q0, which represents orthosilicates SiO44−, Q4Si (tectosilicates), Q3Si, Q2Si and Q1Si representing intermediate silicate structures.
image file: c5ra13043c-f8.tif
Fig. 8 The selected networks of silicate glasses16 (reproduced with permission of the Mineralogical Society of America).

These networks can be modified by adding certain network modifiers such as alkali and alkali earth atoms. The high frequency bands are described in mainly polarized Raman bands centred in the ranges 1050–1100, 950–1000, 900 and 850 cm−1; they are designated as disilicate, metasilicate, pyrosilicate and orthosilicate, respectively. The low frequency bands are mainly associated with Si–O–Si linkages.16 The presence of the water group in SiO2 glasses was confirmed by a Raman band at 3598 cm−1. The weak band near 2350 cm−1 arose due to Si–OH groups involved in intratetrahedral hydrogen bonding across an edge of the SiO4 tetrahedron. The Si–OH stretching mode at 970 cm−1 is more prominent than other bands and shows that more silanol groups are present in comparison to Suprasil. The bands at 430, 800, 1060 and 1200 cm−1 arise due to fundamental vibrations of the dry SiO2 glass. The sharp peaks at 490 and 600 cm−1 are characteristic of defect bands. Raman bands at 541 and 1100 cm−1 in the spectra of dry and wet sodium silicate glasses indicate the presence of SiO units (QSi, species: SiO4 units with one non-bridging oxygen) and vibrations of the bridging oxygen in the Si–O–Si linkage, respectively. The shoulder peaks near 850, 960 and 1000 cm−1 suggest that three components are present in the high-frequency envelope, whereas four components near 970, 1060, 1100, and 1150 cm−1 are similar to hydrous sodium silicate glasses.17

The effect of cations on the symmetric vibrational wavenumber of NBO of Si–O–T in the high frequency range is small, whereas the intensity of the Raman peak increases as the radius of the cation increases. The scattering cross section increases in the order of Li, Na, K, Rb and Cs.18 The Raman spectrum of sulphur-induced sodium calcium silicate glasses indicates that the band due to sulphur ions is located at 990 cm−1. Not only this band, but some other bands due to silicate networks are also present in the pattern.19 The presence of rare earth oxides in soda-lime-silicate glasses led to a shift of the peak positions at 1100, 790 and 550 cm−1, which are attributed to Si–O–Si asymmetric stretching, Si–O–Si symmetric stretching and bending vibrations, respectively. It was observed that the peak positions of bands at 1100 and 790 cm−1 are shifted towards low frequency and the peak at 550 cm−1 was shifted towards high frequency with doping of lanthanide elements (La2O3, CeO2, Nd2O3 and Gd2O3) into soda-lime-silicate glass. Y2O3 in soda-lime-silicate glass led the frequencies of peaks at 1100, 790 and 550 cm−1 to shift towards the high frequency side. The QnSi species in soda-lime-silicate glasses changes as Si2O76− + 2Si2O52− = 5SiO32− + SiO2 with addition of lanthanide elements.20 The band near 980 cm−1 was observed in both the bulk (Suprasil) and sol–gel silica glass. The defect mode bands at 493 and 606 cm−1 are also present in the matrix. Raman bands at ∼356, ∼264, ∼207 and ∼128 cm−1 in quartz crystal are attributed to lattice modes, whereas the strong peak at ∼465 cm−1 is characteristic of a symmetric stretching vibration. At higher frequencies, the bands at ∼807 and ∼1083 cm−1 are assigned to Si–O–Si bending and SiO4 asymmetric stretching vibrations, respectively. The addition of CaO to silica glass changes the relative intensities of bands, as indicated in Fig. 9. The intensity of the band at 1050 cm−1 increased due to addition of CaO. With the addition of a modifier containing MgO, the peak position of 805 cm−1 shifted towards a lower wavenumber. Comparatively, CaO changes the structure more significantly than MgO.21,22


image file: c5ra13043c-f9.tif
Fig. 9 Raman spectra of xCaO·(1 − x)SiO2 (x = 0, 0.1, 0.2, 0.4) glasses21 (reproduced with permission from Springer).

The main silicate structure does not change much for smaller contents of metal oxide, but two additional peaks appear at 500 and 600 cm−1. The modifying nature of Cs, Rb, K, Na and Li in silica glass is in the increasing order because the intensity of the peak at 950 cm−1 increases in the same order. The silicate bands broaden in the order Cs < Rb < K < Na < Li < Ca < Mg, which may suggest that the perturbation of silicate units increases with increasing cation strength.23

Raman spectra of potash-lime-silica glasses depict bands near 300 and 450 cm−1. The addition of sulphur to a glass matrix increases the Raman cross-section of vibrational modes due to S–O bands in comparison to less polarized Si–O bonds. The intense Raman features in such glasses at 1000–1030 cm−1 indicates a mixture of three calcium sulphates: gypsum (CaSO4·2H2O), (1008 cm−1), basanite (CaSO4·1/2H2O), (1016 cm−1) and anhydrite (CaSO4), (1026 cm−1). The bands in Raman spectra in the ranges of 608–679 cm−1 and 1110–1167 cm−1 are characteristic of silicate networks, whereas the peak in the range 412–494 cm−1 is characteristic of cation-oxygen vibrational modes.24 In Raman spectra of lead silicate glasses, a band near 1000 cm−1 was shifted towards the longer wavelength and becomes broader with increasing content of lead. The intensity of the Raman peak at 140 cm−1 was increased by increasing the content of lead. Other peaks at 450, 950 and 1060 cm−1 are also modified by PbO. After UV irradiation with an energy density of 150 mJ cm−2, significant changes in the spectra were observed. The intensity of a Pb–O band at 140 cm−1 decreased after UV irradiation and no new band appeared in the Raman spectra. The decrease in intensity after UV irradiation is caused by the broken Pb–O bond in lead silicate glasses and the broken Pb–O bond is related to the energy density of the UV beam.25 Lead silicate glasses provide a simple absorption spectrum with gamma irradiation.122 Band assignments of various vibrational units in silicate glasses are listed in Table 2.

Table 2 Assignment of main Raman bands in the spectra of silicate glasses
Wavenumber (cm−1) Raman assignments Reference
580 Si–O0 rocking motions in fully polymerized SiO2 (Q4) units 20
600 Si–O–Si bending vibration in depolymerized structural units 20
700–850 Si–O–Si symmetric stretching of bridging oxygen between tetrahedra 20
1050–1100 Disilicate 16
950–1000 Metasilicate 16
790 Si–O–Si symmetric stretching 20
807 Si–O–Si bending 21 and 22
850 Orthosilicate, SiO44− 16
900 Pyrosilicate 16
970 Si–OH stretching mode 17
1083 SiO4 asymmetric stretching vibration 21 and 22
1100 Si–O–Si asymmetric stretching 20
2350 Si–OH groups involved in intratetrahedral hydrogen bonding across an edge of the SiO4 tetrahedron 17


3.3. Phosphate glasses

The typical Raman spectra of phosphate glass are shown in Fig. 10. The structure of phosphate glasses mainly consists of QnP groups of PO4 tetrahedra (n = number of bridging oxygens per PO4 tetrahedron). The symmetric stretching of P–O is assigned at 1260 cm−1 of an asymmetric profile (620 cm−1), symmetric stretching vibration of P–O (1380 cm−1), the symmetric stretching of a non-bridging oxygen on a Q2P tetrahedron at 1170 cm−1, P–O–P symmetric vibrations at 690 cm−1, the symmetric stretching of the orthophosphate groups PO43− near 960 cm−1. These bands are modified by dopants such as earth and alkaline-earth atoms. The addition of TiO2 up to 10 mol% causes the depolymerization of the phosphate glass network by systematic conversion of Q2P structural units into Q1P and finally into Q0P structural units. Even though Q2P to Q1P conversion takes place due to the breaking of P–O–P linkages, formation of P–O–Ti and P–O–Al linkages provides cross linking between short P-structural units. Above 10 mol% TiO2 content, the network is highly depolymerized due to the formation of Q0P structural units and cross-linking becomes poor. The band near 700 cm−1 is assigned to the symmetric stretching of P–O–P linkages in Q2P and Q1P structural units. On addition of TiO2 to phosphate glass, the intensity of the bands near 1280, 1170 and 700 cm−1 decreased and a few new bands at 1210, 1090, 900, 750 and 625 cm−1 are introduced. The bands at 1210 and 1090 cm−1 are assigned to asymmetric and symmetric stretching vibrations of PO3 groups (Q1 structural units), respectively. The coordination of Ti4+ ions with oxygen atoms is found in Raman spectra and a band near 625 cm−1 is attributed to the vibration of octahedral titanate units TiO6, as its relative intensity increases with increasing TiO2. The band centred at 930 cm−1 is associated with a shorter Ti–O bond present in octahedral titanium and is attributed to the titanyl bonds (Ti–O or –Ti–O–Ti–) associated with five-coordinated Ti. The bands at 900 and 525 cm−1 are related to asymmetric stretching of P–O–P bridges and bending vibrations of P–O bonds, respectively.123 The bands in Raman spectra of a 39AlF3⋅11NaF⋅10LiF·(40 − x)CaF2·MgF2·SrF2·BaF2·xNaPO3 system at 515, 570, and 620 cm−1 are assigned to octahedral [ALF6], five-coordinated [AlF5] and tetrahedral [AlF4] structural units, respectively. The intensity of a band at 1330 cm−1 decreases, whereas the intensity of bands at 1070, 860 and 775 cm−1 increases with increasing phosphate content. The low frequency band at 70 cm−1 is a boson band. Raman peaks at 250 and near the range 350–420 cm−1 are associated with symmetric stretching vibrations of fluorine species bonded by the modifying cations.124 The bands at 1194, 1294 and 1349 cm−1 are prominent in the Raman spectra of Na2O–Al2O3–P2O5. With an increase in the content of Al2O3, the intensity of bands near 1200 cm−1 is enhanced, whereas the band at 1161 cm−1 disappears. The band at 676 cm−1 is shifted towards a higher wavenumber due to the formation of an aluminium metaphosphate structure. Introduction of Al2O3 causes the band located at 1349 cm−1 to disappear. The weak vibrations due to PO2 bending and bending motions in chains of O–P–O appear at 300 and 360 cm−1, respectively. With addition of Fe2O3 to this glass system, the intensity of the bands at 1257, 1185 and 706 cm−1 is increased, whereas the position of the band at 1349 cm−1 shifted to 1309 cm−1 due to depolymerization. The ferric oxide in sodium phosphate glasses forms a metaphosphate structure connected by P–O–Fe bonds.125 The bands in the Raman spectrum of zinc–alumino–sodium–phosphate glasses with added Pr3+ and Nd3+ are located at 545, 698, 1042 and 1162 cm−1.126 Raman spectra of a 55P2O5⋅30CaO·(25 − x)Na2xTiO2 (0 ≤ x ≤ 5) system have bands near 695, 705, 930, 1180, 1280 and 1370 cm−1. The band at 930 cm−1 is attributed to Ti–O and the intensity of the band at 1370 cm−1 increased by the addition of TiO2 of up to 5 mol%.127 Raman spectra of Li2O–BaO–Al2O3–La2O3–P2O5 glasses showed bands at 1850, 1700, 1180 and 700 cm−1.128 The spectra of phosphate glasses containing TiO2 consist of bands at 340–540, 693–697, 1010–1020, 1160–1173 and 1248–1271 cm−1. The band in the range 1010–1020 cm−1 disappears at a higher content of TiO2 and a peak at 905 cm−1 is assigned to asymmetric TiO5.129
image file: c5ra13043c-f10.tif
Fig. 10 Raman spectra of 40Na2O–10Al2O3xTiO2–(50 − x)P2O5 glasses.123

Raman investigations of a xMoO3·(100 − x)[P2O5·CaO], 0 ≤ x ≤ 50 mol% glass system showed that Raman bands are centred at 316, 392, 536, 693, 1175 and 1274 cm−1. Following addition of MoO3 at about 0.3 ≤ x ≤ 1 mol%, a new band at 1032 cm−1 is introduced. The intensity of a band at 683 cm−1 decreased at 5 ≤ x ≤ 10 mol% of MoO3, whereas a peak at 436 shifted towards a higher wavenumber. The band located at 700 cm−1 becomes weaker and the peak at 1270 cm−1 disappears. MoO3 is a strong modifier and forms a band at 600 cm−1 due to asymmetric vibration of Mo–O–Mo bond. The band at 894 cm−1 is caused by orthomolybdate, (MoO4)2− units. The band at 963 cm−1 is assigned to the stretching vibrations of partially isolated Mo–O bonds in deformed (MoO6) units.130 The structure of xCaO·(100 − x) (0.4Fe2O3·0.6P2O5) (x = 0, 10, 20, 30, 40, 50 mol%) glasses is made up of structural units of phosphate and is modified by calcium as well as iron. Their Raman bands are observed at 626, 772, 950, 1071 and 1241 cm−1.131 Raman spectra of calcium coacervate show peaks at 1168 and 698 cm−1. The Raman spectrum of the croconate ion consists of bands located at 564, 649, 1605 and 1722 cm−1.132 In Raman spectra of glasses in the glass system xGa2O3·(1 − x)P2O5, there are five Raman bands located at 1320, 1210, 715, 620 and 350 cm−1. Depolarization by Na, Ca, Zn, Mg, Al and Be is found at 350 cm−1. The low frequency band is assigned to Ga–O–M bridges. The band near 640 cm−1 is assigned to a gallate system with four-fold coordinated Ga atoms.133,134

The structure of PbO–MoO3–P2O5 glasses consists of networks of phosphate units and is modified by MoO3. The incorporation of MoO3 results in the transformation of octahedral MoO6 units into tetrahedral MoO4 units. The band at 930 cm−1 is shifted towards the high frequency side with increasing content of MoO3. The intense band near 1158 cm−1 is shifted towards lower frequency with increasing content of MoO3. The bands in the region 850–970 cm−1 are attributed to Mo–O and Mo–O bonds in MoO6 tetrahedra.135 The structure of PbO–Ga2O3–P2O5 glasses doped with Cr2O3 is modified by the formation of tetrahedral CrO42− structural units. The phosphate network is modified by Pb, Ga and Cr ions. The band at 398 cm−1 is assigned to Ga–O–P linkages and the peak at 368 cm−1 is attributed to GaO6 vibrational groups. The band at 270 cm−1 is associated with Cr–O vibrations of CrO42− units.136 A lead phosphate network can also be modified by Zn and Cu ions. The phosphate network after modification is assigned to positions 690, 747, 904, 1026 and 1155 cm−1. The intensity of a band at 1155 cm−1 decreases with addition of BaO.137 A high ZnO content in lead phosphate glasses forms a network of P–O–Zn bonds.138 The network of lead–gallium phosphate glasses can be modified by rare earth ions (Eu3+, Dy3+, Tb3+, and Er3+). The spectrum contains two shoulders centred at 1050 and 1120 cm−1, corresponding to symmetric stretching vibrations of diphosphates in Q1 units. The other two shoulders near 900 and 1250 cm−1 are attributed to symmetric stretching vibrations of PO4 in the Q0P units and asymmetric stretching vibrations of PO2 in Q2P units, respectively. The high frequency band at about 1120 cm−1 is related to the phonon energy of the lead phosphate glass and its intensity decreases when PbO is substituted by PbF2 components.139 The bands near 1172, 1096, 1032, 973, 930, 742, 627 and 474 cm−1 are found in glasses in the system PbO–Fe2O3–P2O5. The bands at 1080 and 966 cm−1 are related to [P2O7]2− and terminal [PO3]2− symmetric groups in Q1P pyrophosphates.140 In2O3 in PbO–P2O5–As2O3 plays an important role in network modification and forms the structural unit of InO6. The band centred at 200 cm−1 is related to In–O vibrations of InO6 structural units. The bands near 110 and 250–350 cm−1 are attributed to Pb–O ionic bond vibrations and PbO4 structural vibrations, respectively. With an increase in the concentration of In2O3, the intensity of the bands due to symmetric vibrations of the phosphate groups is increased, whereas the intensity of bands due to torsional vibrations of PO4 structural units and As2O3 structural vibrations is decreased.141 The fraction of orthophosphate Q0P units in lead iron phosphate glasses increases with increasing Cr2O3 content. The band at 600 cm−1 is assigned to different P–O, Fe–O and Pb–O stretching and bending modes. An intense peak appears at 1062 cm−1 in Cr-free glass. The intensity of a band at 979 cm−1 is enhanced with Cr addition. Ferric oxide also forms a network of Fe–O stretching vibrations in FeO6 octahedra in α-Fe2O3 and this is indicated by a band near 614 cm−1.142,143

The Raman spectra of iron phosphate glasses investigated by Milankovic et al. showed that networks of phosphate glasses are modified by the incorporation of MoO3. Addition of MoO3 to a glass matrix replaced the stronger P–O–P bonds with weaker Mo–O–P bonds as the bond covalence decreased from P–O to Mo–O and the bands located at 850 and 980 cm−1 became stronger. Iron works as a network modifier for phosphate networks and a band at 400 cm−1 assigned to PO4 polyhedra was modified by it. The Raman band near 864 cm−1 is associated with the stretching mode of isolated orthomolybdate (MoO4)2−. Due to an increase in NBO and (PO4)3−, some bands show splitting in the spectra. The asymmetric vibration of Mo–O–Mo is assigned at 620 cm−1, whereas the symmetric Mo–O–Mo vibrational mode appears at 520 cm−1.144 The addition of La2O3 in iron phosphate glasses lowers the peak position of a band at 1072 cm−1 and increases the relative area of a band at 1231 cm−1, whereas this decreases for a band at 1072 cm−1. This result is related to the depolymerization of the phosphate network 2Q1P = Q2P + Q0P.145 The most intense peak in Raman spectra of cesium-doped iron phosphate glasses is at 1085 cm−1 and is assigned to Q1P. This band is shifted towards a lower wavenumber with increasing Cs content. The intensity of a band at 555 cm−1 is enhanced due to the disproportion of the pyrophosphate linkage. The intensity of this band increases with increasing cesium loading and indicates the increasing concentration of orthophosphate units. The bands at 270 and 200 cm−1 are related to bending vibrations of P–O–P and Cs–O vibrations, respectively.146 The P–O–P network is depolymerized with increasing CeO2 content in glass compositions and results in the conversion of Q1P to Q2P structural units. The addition of CeO2 in iron phosphate glasses introduced some new bands near 1000, 1150 and 1279 cm−1. The band at 1037 cm−1 is shifted towards a higher wavenumber with the substitution of FeO3 by CeO2. The band at 522 cm−1 is related to FeO bonds and a new peak at 460 cm−1 is attributed to cerium oxygen polyhedron vibration of the [CeO8] unit symmetrical stretching mode at higher contents of CO2.147 Depolymerization of the phosphate network, in the form of conversion of Q1P to Q0P groups (2QnP + Na2O = 2Qn−1P), takes place with the addition of Na2SO4 and results in the conversion of P–O–P bridging oxygens to P–O–Na+ non-bridging oxygens. The band at 1039 cm−1 is shifted to the higher frequency region with the addition of the modifier Na2O (Na2SO4 = Na2O + SO3). The intensity of the band at 931 cm−1 is also increased with this addition.148 The vibrations at 170, 306 and 420 cm−1 arise due to network vibrations and vibrations of Fe–O polyhedral and phosphate network units, respectively. The peak position of 1090 cm−1 is shifted towards a lower wavenumber with heat treatment at the transition temperature of the glass.149

The MoO3 plays an important role in the modification of structural units of the glass system ZnF2–Bi2O3–P2O5. With addition of MoO3 to the glass matrix, the structural units of MoO4 and MoO6 are formed. The intensity of the band due to asymmetric vibrations of Mo–O–Mo linkages is observed to increase with an increase in dopant concentration with a slight shift in the position towards a lower wavenumber. In the Raman spectra of the abovementioned system, bands due to BiO6 and BiO3 pyramidal structural units are assigned at 547 and 230 cm−1, respectively. The modifier Zn takes the interstitial positions in the form of dangling bonds along with non-bridging oxygens as Zn–O–P linkages and fluorine breaks the P–O network.150,151 The doping of WO3 in the place of MoO3 forms a network of bending vibrations of O–W–O with corner shared octahedra. A noticeable decrease in the intensity of W–O stretching was also observed, which is associated with WO4 tetrahedral units, as shown in Fig. 11.152 Trivalent samarium doped K–Mg–Al zinc fluorophosphate glasses are formed by networks of phosphate and modified by K, Mg, Al, Zn and F. The stretching vibration of Al–O linkages in AlO4 structural units was observed.153,154 The structure of a 2P2O5–CaO–0.05ZnO glass matrix contains Raman bands near 345, 460, 903, 959, 1011, 1137, 1168, 1297 and 1400 cm−1. The content of Ag2O in the glass system modified the network of ring structures. Some depolymerization had also taken place in the glass matrix. Raman spectra of xAg2O·(100 − x)[2P2O5·CaO·0.05ZnO] confirm the modifying nature of Ca, Zn and Ag. With increasing content of Ag2O, the intensity of a band near 554 cm−1 is increased and the peak position of a band at 690 cm−1 is shifted towards the lower frequency side due to changes in the chain P–O–P bond angles. Silver oxide favours the presence of orthophosphate, pyrophosphate and small metaphosphate chains; this information is revealed by the increase in intensity of the bands at 554, 1100 and 670 cm−1.155


image file: c5ra13043c-f11.tif
Fig. 11 Raman spectra of ZnF2·Bi2O3·P2O5 glass-ceramics doped with different concentrations of WO3 (ref. 152) (reproduced with permission from Elsevier).

Raman spectroscopy investigations of TiO2–P2O5–CaO glasses showed that the structure consists of a distorted Ti octahedral linked to a pyrophosphate unit through P–O–Ti bonds at higher contents of TiO2 and thus depolymerization took place in the P–O–P network. In Raman spectra of TiO2-free calcium phosphate glasses, the strong bands near 680, 1175 and 1290 cm−1 are assigned to the glassy network, whereas with addition of TiO2, the existing bands disappeared. New bands at 765 and 920 cm−1 are also introduced with TiO2 addition.156 Li2O further modified the network of TiO2–P2O5–CaO glasses.157 Raman spectra of (P2O5–K2O–Al2O3–CaF2) glasses depict bands at 335, 543, 625, 728, 1075, 1135 and 1284 cm−1. The band centred at 728 cm−1 is related to vibrations of P–F chains, whereas a less intense band at 625 cm−1 is assigned to Al2O vibrations. The band at 335 cm−1 is attributed to K2O.158,159 CuO and V2O5 work as network modifiers for calcium phosphate glasses. At higher concentrations of CuO and V2O5, the glass matrix forms a network of V–O–P and Cu–O–P. The band at 1175 cm−1 indicates depolymerization of the phosphate network. The band at 980 cm−1 is related to V–O vibration in VO5 tetragonal pyramids, and the stretching vibrations of O–V–O in metavanadate chains appear at 970–905 cm−1.160 Raman spectra of (50 − y)Na2yCuO·10Bi2O3·40P2O5 (0 ≤ y ≤ 25 mol%) glasses depict the formation of P–O–Bi and P–O–Cu bonds. Bi2O3 acts as a network modifier and is incorporated in the form of BiO6 units. No depolymerization took place with the replacement of Na2O by CuO.161 As CuO content increases in xCuO·(1 − x)P2O5 glasses, the intensity of the P–O band associated with the Q3 Cu tetrahedra increases to x < 0.33 and decreases with a concomitant increase of the intensity of the band at 1265 cm−1 due to the asymmetric vibration of the PO2 groups on Q2 tetrahedra. When x > 0.33, the isolated Cu-octahedra begin to share common oxygens to form a sub-network in the phosphate matrix. The intensity of a band at 640 cm−1 increases with the addition of CuO. CuO in phosphate glasses promotes depolymerization.162

Raman spectra of cadmium phosphate glasses reveal the formation of P–O–V and V–O–V bonds instead of P–O–P vibrations after the addition of V2O5. The presence of a band at 760 cm−1 is attributed to V–O–V vibration. The intensity of bands at 615 and 974 cm−1 is increased with addition of V2O5 at about 20 mol%. These bands are related to P–O–V and V–O vibration of VO5 trigonal bipyramids.163 The Li+ and Cd2+ cations in the phosphate glass network are responsible for shifting the bands as well as for intensity modification. The band at 1120 cm−1 is shifted towards the high frequency side with increase of the Li2O content.164 Increasing the content of BaO in lithium barium phosphate glasses causes the intensity of all bands to be enhanced. The bands at 712 and 1163 cm−1 are shifted slightly towards a lower wavenumber with an increase in BaO. The decrease in wavenumber is related to an increase in the average length of the P–O bond resulting from π-bond delocalization on the Q3P tetrahedra.165 Raman spectra of phosphate glasses with addition of Er3+/Yb3+ ions depict bands at 312, 380, 685, 1165 and 1275 cm−1.166 The assignments of the main Raman bands in the spectra of phosphate glasses are listed in Table 3.

Table 3 Assignment of main Raman bands in the spectra of phosphate glasses
Wavenumber (cm−1) Raman assignments Reference
1300–1350 P[double bond, length as m-dash]O stretching of terminal oxygen 123
1267–1294 PO2 asymmetric stretching of non-bridging oxygen in Q2 units 123
1260, 1380 Symmetric stretching of P–O 123
1210 Asymmetric stretching vibrations of PO3 groups 123
1170 Symmetric stretching of a non-bridging oxygen on a Q2P tetrahedron 123
1161 PO2 symmetric stretching of non-bridging oxygen in Q2 units 123
1065–1097 PO2 symmetric stretching of non-bridging oxygen in Q1 units 123
1090 Symmetric stretching vibrations of PO3 groups 123
1080 [P2O7]2− symmetric groups in Q1P pyrophosphates 140
966 Terminal [PO3]2− symmetric groups in Q1P pyrophosphates 140
960 Symmetric stretching of the orthophosphate groups PO43− 123
940 PO2 symmetric stretching of non-bridging oxygen in Q0 units 123
900 Asymmetric stretching of P–O–P bridges 123
700 The symmetric stretching of P–O–P linkages in Q2P and Q1P structural units 123
690 P–O–P symmetric vibrations 123
676–708 P–O–P symmetric stretching of bridging oxygen in Q2 units 123
525 Bending vibrations of P–O bonds 123
390–410 O–P–O bending modes 123
398 Ga–O–P linkages 136
360 Chains of O–P–O bending motions 125
300 PO2 bending O–P–O bending motions 125


3.4. Borosilicate glasses

Typical Raman spectra of alkali borosilicate glasses are shown in Fig. 12. The network structure of borosilicate glasses is formed by borate as well as silicate structural units. The structure of lithium silicate glasses produces Raman bands at 850 cm−1 (Q2Si), 950 cm−1 (Q2Si), 1000 cm−1 (Q3Si), 1050 cm−1 (Q3Si), 1080 cm−1 (Q3Si) and 1150 cm−1 (Q4Si). The borosilicate structure is associated with the vibrational modes of the two medium-range (400 to 850 cm−1) order superstructures, reedmergnerite [BSi3O8] and danburite [B2Si2O8]2− and high frequency region (850–1250 cm−1) vibrations are related to silicon QnSi vibrational modes.167
image file: c5ra13043c-f12.tif
Fig. 12 Raman spectra of alkali borosilicate glasses167 (reproduced with permission from Elsevier).

Raman spectra of sodium borosilicate glasses depict bands centred at 350, 500, 630, 670, 770 and 808 cm−1. The bands between 300 and 500 cm−1 are attributed to mixed stretching and bending modes of Si–O–Si units. The bands in the 550–850 cm−1 range are caused by ring breathing modes. The bands at 670, 770 and 808 cm−1 are the signature of tetraborate groups, four- and three-coordinated boron in diborate and boroxol rings, respectively. The intensity of the peak at 808 cm−1 decreases with increasing sodium content, whereas the band near 770 cm−1 shows the opposite behaviour. This means that sodium ions favour the formation of BO4 units at the expense of BO3 units. Polymerization of the silicate network first increases with Na2O content up to Na2O/B2O3 = 0.6 and thereafter decreases with increasing Na2O content.168 Sulphate in borosilicate glass showed its presence in the form of vibrational modes of SO4 tetrahedra. The major bands of sulphate at 1100, 1000, 990, 620 and 460 cm−1 are related to asymmetric S–O stretch modes, symmetric S–O stretching, symmetric S–O stretch vibrations of tetrahedral SO42−, asymmetric O–S–O bend modes and symmetric O–S–O bend modes, respectively.169 When the B2O3 content is increased, broad bands near 1500 and 800 cm−1 are developed. The shoulder peak at 590 cm−1 is assigned to the symmetric oxygen breathing vibrations of three-membered siloxane rings. The network modifier ions Na+ and Ba2+ are used for the conversion of BO3 units into BO4 units before forming non-bridging oxygen (NBO) in the network. As the silica content increases in the glassy matrix, the ratio Na/Si (or Ba/Si) decreases, which could depolymerise the silicate network after the formation of BO4 units. It can be concluded that the silicate network in SiO2-rich borosilicate glasses remains more polymerized than in boron-rich glasses.170 The bands in the range 1200–1100 cm−1 disappear as the ratio B2O3/Na2O tends to zero. The intensity of the band at 1077 cm−1 is also increased. With an increasing amount of Na2O in the glass, the intensity of the band around 630 cm−1 increased and the band in the range 550–450 cm−1 began to dominate, due to formation of [BO4] tetrahedra with a small degree of polymerization. The intensities of bands at 385, 360, 287, 1850 and 1975 cm−1 are increased as the ratio B2O3/(Na2O + 3La2F6) increases.171 When barium is substituted for the sodium cation, the position of the peak at 540 cm−1 is shifted to the low frequency region and the intensity of a band at 635 cm−1 was decreased. A high frequency band around 1100 cm−1 becomes stronger with the introduction of sodium. The substitution of Ba and Ca cations enhances the intensity of a band at 960 cm−1 and shifts it towards the low frequency region.172 Raman spectra of Na2O–B2O3–SiO2 exhibit bands at ∼430, 485, 600, 800, 1060 and 1200 cm−1. Bands around 485 and 600 cm−1 (known as defect lines D1 and D2) are assigned to the ring breathing mode of four and three-membered silicate rings, respectively. Raman peaks at 770 and 805 cm−1 become more significant for glasses with higher borate content.173 In Raman spectra of 30Na2O·5SiO2·65[(1 − x)P2O5·xB2O3] glasses, the intensity of bands at 680 and 770 cm−1 decreases with an increase in B2O3; these are attributed to (P–O–P)sym stretching vibrations and BO4 stretching in borate species, respectively. Bands at 1200 and 1350 cm−1 disappear for x = 0.250. The intensity of a band at 620 cm−1 increases with an increase in the B2O3 content. Bands appear at 500, 780 and 1100 cm−1 in B2O3 rich glasses. The Raman band at 1160 cm−1 is attributed to the symmetric stretching mode of the terminal oxygen on each tetrahedron. The intensity of a band at 780 cm−1 increases and is assigned to a BO3 stretching vibration.174 The intensity of Raman bands at 810 cm−1 is increased, whereas it is decreased for a band at 795 cm−1 as a result of heat treatment of glasses.175 The S–O symmetrical stretching modes near 1000 cm−1 from tetrahedral SO4 environments were observed. The peak near 300 cm−1 is assigned to S–S stretching in the dithionate anion S2O62−.176,177

The wettability of diamond by borosilicate glass melt at temperatures above 700 °C was enhanced due to oxidation of the diamonds. The irregular pores become distinct in diamond–borosilicate glass composites sintered above 800 °C.178 Depolarization of the silicate network takes place with an increase in zirconia and causes reduction of the intensity of the band due to Si–O–Si links. Some of the highly polymerized Q3Si and Q4Si units are replaced by the less polymerized Q2Si units in the glass structure. Most of the borate units within the glass structure are probably linked to silicate tetrahedra in the range 650–800 cm−1 and this is indicated by the lack of any prominent peaks in this range. The band near 703 cm−1 is related to the zirconia network (Zr–O stretch).179 The spectrum of Eu3+ ion doped MgO–PbO–B2O3–SiO2–Nd2O3 glasses consists of bands centred at 1510, 1385, 955, 842, 732, 645, 487 and 298 cm−1 (Fig. 13). The latter band near 298 cm−1 was attributed to Ln–O–Ln (Nd and Eu) stretching vibrations. The position of the first band was shifted towards a higher wavenumber with an increase of the doping concentration of Eu. The bands at 605 and 527 cm−1 are related to B–O–Si linkages along with PbO4 units and Si–O–Pb, respectively. Polymerization of the glass network takes place as a result of replacement of B–O–B and Si–O–Si bonds with more resistant B–O–Si and Si–O–Pb bonds with addition of Eu3+ ions.180 Structural units of TiO4 exist in CaO–B2O3–SiO2–TiO2 glasses and the degree of depolarization decreases with an increase in the CaO/SiO2 ratio. An increase in B2O3 leads to an increase of BO4 tetrahedral units in the glass matrix. The addition of TiO2 to this glass system introduces two peaks near 840 and 726 cm−1, whereas the existing bands at 500–800 cm−1 disappear. Bands near 948, 1040 and 1300–1500 cm−1 shifted towards low frequency. The bands at 840 and 726 cm−1 are characteristic of the vibrations of Ti–O–Si or Ti–O–Ti and deformation of O–Ti–O or O–(Si,Ti)–O in chain or sheet units or both.181


image file: c5ra13043c-f13.tif
Fig. 13 Raman spectra of MgO–PbO–B2O3–SiO2–Nd2O3–Eu2O3 glasses180 (reproduced with permission from Elsevier).

The structure of borosilicate glass in the presence of the modifier BaO also contains a network of borate and silicate. The addition of BaO decreases the intensity of a band near 816 cm−1. A new band at 780 cm−1 is incorporated at 42 mol% of BaO. Bands at ∼843, 877, 934, 1000 and 1040 cm−1 are shifted to the lower frequency region and are caused by low network polymerization after the formation of BO4 units due to the increase in BaO content instead of SiO2.182 The position of a peak at 675 cm−1 is shifted towards a higher wavenumber with an increasing ratio of BaO/SrO. This is associated with variation in the ionic radii of Ba and Sr. La2O3 and Fe2O3 in barium strontium titanate (BST) borosilicate glasses shift the peak positions of 675 and 820 cm−1 towards a higher wavenumber.183–185 A Raman band at about 735 cm−1 occurs due to metaborate groups and symmetric breathing vibrations of BO3 triangles replaced by boron tetrahedra (BO4). The Raman shift has been shifted towards a higher wavenumber with an increase in the doping concentration of Fe2O3.52 Raman spectra of AO–SiO2–B2O3–La2O3, (A = Mg, Ca, Sr, Ba) glasses showed bands near 770–779 and 1274–1280 cm−1, which are related to borate networks. The bands near 796–800 and 1060–1070 cm−1 are caused by silicate networks. Tetrahedral coordination occurs in both silicon and boron in this glass system. Lanthanum occupies interstitial octahedral co-ordination in the local network. Modifiers, such as Ba, Ca, Mg, Sr and La, break the B–O or Si–O bonds and occupy interstitial or substitutional positions in the glass network. As the ionic radii of the modifier ion increases in the order Mg2+ > Ca2+ > Sr2+ > Ba2+, polarization follows Fajan's rule, i.e. weak polarization by Mg and strong polarization by Ba.186 Raman spectra of lead bismuth borosilicate glasses show bands at 1582 and 1382 cm−1 in the high frequency region.187 After irradiation, the peak at 480 cm−1 is shifted towards higher wavenumber. This suggests a decrease of the Si–O–Si bond angle. A new weak peak at 605 cm−1 is introduced by irradiation in glasses and is assigned to the D2 defect peak involving the three-membered rings of SiO4 tetrahedra. The weak band at 630 cm−1 is associated either with the hindered bending vibration of ring-type metaborate groups or the breathing mode of danburite-like rings. The QnSi species of silicon remains almost unchanged after Ar-irradiation.188 The Raman band assignments of borosilicate glasses are listed in Table 4.

Table 4 Assignment of main Raman bands in the spectra of borosilicate glasses
Wavenumber (cm−1) Raman assignments Reference
1510–1570 B–O stretching modes involving one NBO of [BO3] triangles and molecular oxygen stretching vibration modes 183–185
1385–1397 Stretching vibrations of B–O bond in BO4 units from different borate groups 183–185
1160 Symmetric stretching mode of the terminal oxygen on each tetrahedron 174
947–955 Stretching vibrations of Si–O bond due to the presence of BO4 units 183–185
840–842 Pyroborate unit along with ortho-silicate (SiO4)4− composition or vibrations of Ti–O–Si or Ti–O–Ti structural units 181
808 Boroxol rings 168
722–732 Chain-type metaborate groups containing NBO or deformation of O–Ti–O or O–(Si,Ti)–O in chain or sheet units or both 181
550–850 Ring breathing modes 167 and 168
400–850 Reedmergnerite [BSi3O8] and danburite [B2Si2O8]2− 167
605 B–O–Si linkages along with PbO4 units 180
590 Symmetric oxygen breathing vibrations of three-membered siloxane rings 170
300–500 Mixed stretching and bending modes of Si–O–Si units 167
475–487 Si–O–Si, Si–O–B isolated vibrations 180


3.5. Borophosphate glasses

Typical Raman spectra of tin borophosphate glasses are shown in Fig. 14. In this study, it was found that the Raman bands were centred at 1330, 1270, 1050, 1000, 950, 760, 700 cm−1 and some bands appeared near 500 cm−1. The vibrations at 1330 and 700 cm−1 are attributed to metaborate groups, whereas the band near 1000 cm−1 is attributed to pyrophosphate or orthophosphate units. The presence of metaborate chains in the glass matrix was indicated by the vibration at 760 cm−1. Evidence of pyrophosphate units was also confirmed by the peak located at 1050 cm−1. The vibration at 1270 cm−1 is caused by orthoborate/pyroborate units.189 A network of lithium borophosphate glasses, in the presence of germanium oxide, consists of stretching vibrations of phosphate structural units in high (1000–1300 cm−1) and mid frequency regions (500–700 cm−1) and these units are modified by Ge ions. Ge-free glasses showed a band at 1159 cm−1, which is caused by a metaphosphate structure. The vibration at 1260 cm−1 is related to an asymmetric, νas(PO2), stretching mode in Q2P units, whereas the weaker band centred at 1093 cm−1 is attributed to νs(PO3) in Q1P units. The bands at 703 and 665 cm−1 are assigned to the symmetric stretching mode of the P–O–P bridging bond in Q1P and Q2P structures, respectively. A broad band around 320 cm−1 is attributed to the cage vibrational frequencies of Li+ ions. The addition of 5 mol% GeO2 shifted the band around 1159 to 1106 cm−1 and also caused depolymerization of the Q2P structure by the formation of more Q1P phosphate units. The intensity of the band at 665 cm−1 is reduced with increasing GeO2 and it is completely eradicated by the addition of 15 to 25 mol% of GeO2, whereas the intensity of bands at 700 and 751 cm−1 is enhanced due to more depolymerization of Q1P units by this addition. A new band at 600 cm−1 is introduced due to Ge–O–Ge bending modes or Ge–O–P stretching modes.190 Four intense bands at 650, 696, 880 and 1112 cm−1 are found in the Raman spectrum of 75P2O5–20B2O3–4.9Na2O–0.1Er2O3 glass.191
image file: c5ra13043c-f14.tif
Fig. 14 Raman spectra of various tin borophosphate glasses189 (reproduced with permission from Elsevier).

Raman bands at 636 and 350 cm−1 in Na2O–Ga2O3–P2O5 glasses are attributed to GaO4 and GaO6 groups, respectively. The bands centred at 515 and 540 cm−1 are caused by Ga–O–Ga bonds between GaO4 tetrahedra, and the band at 675 cm−1 is associated with vibrations involved with non-bridging oxygens from the GaO4 tetrahedra. The band at ∼540 cm−1 was attributed to [GaO4]. In Ti-containing glasses, an intense band near 900 cm−1 is caused by [TiO4] groups.192 The Raman spectra of calcium borophosphate consisted of bands attributed to networks of ortho Q0P and pyro Q1P phosphate units. With the addition of 20 mol% of B2O3 to calcium phosphate glass, a band at 1045 cm−1 is shifted to 1038 cm−1 and new bands are introduced at 1000 and 675 cm−1.193 In the Raman spectra of (1 − x)[0.5K2O·0.1B2O3·0.4P2O5xNb2O5 glasses, a strong band in the region 907–911 cm−1 was observed and the bands assigned to phosphate groups appear at 1070–1148 cm−1 and 1212–1244 cm−1. The new bands at 811–817 and 638 cm−1 are introduced in glasses for more than 16.7 mol% Nb2O5. The vibrations centred at 907 cm−1 are characteristic of symmetric stretching vibrations of isolated NbO6 octahedra. The dominating bands at 817 and 638–648 cm−1 are assigned to Nb–O–Nb bonds in glass with 37.5 mol% of Nb2O5. The bands at 820–840 cm−1 in barium–potassium niobato–phosphate glasses are attributed to the deformation vibrational modes of Nb–O–Nb bridges between NbO6 octahedra.194

Raman spectroscopy studies of 50PbO·10B2O3·40P2O5·xTiO2 glasses showed that major bands are located at 329, 729, 936, 1100 and 1210 cm−1 (Fig. 15). The strong peaks at 1100 and 1210 cm−1 are shifted towards a lower wavenumber, whereas the peak at 729 and 936 cm−1 are shifted towards a higher wavenumber with increasing content of TiO2.195 TeO2 in lead borophosphate glasses forms the structural units of TeO3 and TeO4 at low and high concentrations, respectively. The intensity of a band at 1092 cm−1 decreases with TeO2 content and also shifts towards the low frequency side. This change is associated with the depolymerisation of the phosphate network. The bands at 550 and 850 cm−1 are attributed to the vibration of oxygen atoms in P–O–P bridges and are gradually replaced by a band that is a characteristic of different Te–O vibrations. New bands in the 453–458 cm−1 region are ascribed to bending vibrations of O–Te–O and Te–O–Te linkages. The bands lying at 620–630 cm−1 and 660–670 cm−1 are related to the stretching vibrations of TeO4 trigonal bipyramids, whereas the vibrations at 720–750 cm−1 and 760–800 cm−1 are associated with stretching vibrations of TeO3 trigonal pyramids.196


image file: c5ra13043c-f15.tif
Fig. 15 Raman spectra of the glass series 50PbO–10B2O3–40P2O5 + xTiO2 (ref. 196) (reproduced with permission from Elsevier).

Zinc oxide plays an important role in the modification of borate and phosphate structural units. In the spectra, a sharp band at 808 cm−1 is characteristic of B2O3 and the band at 968 cm−1 is caused by the vibrations of isolated PO4 units in the structural network of borophosphate glasses.197 If TiO2 is added to zinc borophosphate glasses, the intense bands at 1162 and 666 cm−1 are shifted towards the lower frequency side, whereas the band at 747 cm−1 is shifted to 774 cm−1. The band at 506 cm−1 disappeared above 4 mol% of TiO2. The shifts of bands are associated with the distortion of borate and phosphate units. The shift in sodium containing glasses is less than that in Zn modified glasses due to the difference in the ionic field strength.198 All the Raman bands are shifted towards a lower wavenumber with the substitution of Sb2O3 instead of TiO2 in the abovementioned glass system. The vibrations of structural units containing Sb–O are assigned in the range 350–700 cm−1. Sb2O3 modifies the network in the form of SbO3 units with a single atom of Sb. The low concentration of Sb2O3 forms the isolated SbO3 unit. The existence of three bands in place of two is associated with the splitting of the symmetry of SbO3 units in the glass network. The Sb2O3 content replicates the depolymerisation of phosphate chains. Raman bands in the 520–690 cm−1 range are attributed to SbO3 pyramids. SbO3 units are linked into chains and clusters with Sb–O–Sb bridges at higher Sb2O3 contents and are characterised by bands in the 380–520 cm−1 regions.199 MoO3 weakened the bond strength for zinc borophosphate glasses. The presence of the polarized vibrational band at 976 cm−1 is associated with MOx symmetric stretching vibrations and the depolarized band at 878 cm−1 is ascribed to the Mo–O–Mo stretching vibration. The incorporation of molybdate units in the glass matrix is responsible for the depolymerization of phosphate chains and the formation of P–O–Mo bonds.200 Tellurium oxide in the abovementioned glass system forms structural units of TeO3, TeO3+1 and TeO4. The phosphate units are depolymerized by the incorporation of TeO2 and also boron coordination is modified. When the ratio of B2O3/P2O5 increases, TeO4 units are replaced by TeO3 units as the number of oxygen atoms in the glass decreases.201–204 The peak assignments of Raman spectra of borophosphate glasses are listed in Table 5.

Table 5 Assignment of Raman bands in the spectra of borophosphate glasses
Wavenumber (cm−1) Raman assignments Reference
1260, 1380 Symmetric stretching of P–O 123
1210 Asymmetric stretching vibrations of PO3 groups 123
1170 Symmetric stretching of a non-bridging oxygen on a Q2P tetrahedron 123
1080 [P2O7]2− symmetric groups in Q1P pyrophosphates 140
966 Terminal [PO3]2− symmetric groups in Q1P pyrophosphates 140
960 Symmetric stretching of the orthophosphate groups PO43− 123
600 Ge–O–Ge bending modes or Ge–O–P stretching modes 190
550, 850 Vibration of oxygen atoms in P–O–P bridges 196
968 Vibrations of isolated PO4 units 199


3.6. Aluminosilicate glasses

Raman spectra of glasses for the glass system 5Na2O·20CaO·5Al2O3·(60 − x)SiO2·xZnO (where x = 0, 4, 7 and 10 mol%) showed the bands by deconvolution, which are centred at 868, 956, 1020 and 1075 cm−1 and are attributed to the Si\NBO stretching modes of mainly silicate units, QnSi (Fig. 16). The band near 868 cm−1 is characteristic of Q1Si. The peak at 956 cm−1 is assigned to Q2Si, whereas the peak at 1019 cm−1 is caused by the asymmetric stretching of BO and is related to Q1Si–Q3Si units. The band at 1020 cm−1 is associated with Q3Si. Raman bands in the wavenumber range of 400–800 cm−1 are at 412, 544, 622 and 779 cm−1. The vibrations in the range 400–650 cm−1 are caused by the bending vibrations of the bridging oxygen (BO) bonds of SiO4 and the band near 779 cm−1 is caused by the Si–O–Si network, ZnO4 tetrahedra and AlO4 units with three BOs and one NBO. The low frequency bands located at 222, 279 and 338 cm−1 are related to the cations of modifying networks. The vibration of cations Na+ and Ca2+ are centred at 222 and 279 cm−1, respectively. The coupled vibration of the Ca–O and SiO4 networks is found at 338 cm−1. The addition of ZnO to the glass matrix shifts the bands at 279 and 338 cm−1 to 274 and 346 cm−1, respectively. With an increasing ZnO/SiO2 molar ratio, the band at 622 cm−1 is shifted towards 645 cm−1 and finally, splits into two components for 10 mol% ZnO content glasses.203
image file: c5ra13043c-f16.tif
Fig. 16 Raman spectra of glasses for glass system 5Na2O·20CaO·5Al2O3·(60 − x)SiO2·xZnO203 (reproduced with permission from Elsevier).

Sodium silicate and aluminosilicate can be analyzed by Raman spectroscopy in the low frequency region of 20–250 cm−1. In this region, the peak at 69 cm−1 is the signature of Raman scattering involving rotational motions of interconnected tetrahedral units. The position of this peak is shifted to 78 cm−1 with the addition of Al2O3 up to 6 mol%. The Raman peak at 170 cm−1 is assigned to a narrower distribution of T–O–T angles or to a structural localization of those vibrations in specific clusters of molecules in the glass network. The band centred at 490 cm−1 is shifted to 482 cm−1. The band at 597 cm−1 becomes broader with an increase in the content of Al/(Al + Na). This result suggests the presence of three-membered rings in the glass network. The Raman band located at 781 cm−1 is shifted to 799 cm−1 with replacement of Na2O by Al2O3. A shoulder near 810–820 cm−1 is a feature of Si–O stretching involving oxygen motions in the Si–O–Si plane or of the motion of the Si atom in its oxygen cage or a threefold-degenerate rigid cage vibrational mode of TO2 units. The high frequency bands at 960, 1070, 1100, 1150 and 1200 cm−1 are characteristic of silicate networks and vibrational modes of TO4 tetrahedra.204 A Raman band at 495 cm−1 was observed in albite glass and showed the presence of polymeric structures in such glasses. The band at 595 cm−1 is characteristic of structural defects in the glass involving T–O (non-bridging) vibrations (Fig. 17). The defect line, D2, of SiO2 is at 606 cm−1. Al in borosilicate glass shifted the band positioned at 950 cm−1 to 900 cm−1, while peaks at 980 and 1110 cm−1 showed the opposite trend to the previous result. The polymerization becomes more significant with an increasing Al/Na ratio.205


image file: c5ra13043c-f17.tif
Fig. 17 Raman spectra of 75SiO2·(25 − x)Na2xAl2O3 glasses204 (reproduced with permission from Elsevier).

The structure of aluminosilicate glasses consists of silicate and aluminate networks. The bands centred at 1060 and 1200 cm−1 are attributed to Si–O stretching vibrations of SiO4 tetrahedra. The intensities of bands at 1000 and 1100 cm−1 are increased with an increase in Al2O3. This increase in the intensity of these bands is due to the interlinking of the Si–O–Al network. The intensity of the band at 435 cm−1 was decreased with an increase in Al2O3 and this was believed to be the result of a decrease of Si–O–Si linkages in tetrahedral units.206 Aluminosilicate glasses along with silica–calcium, silica sodium aluminate and silica–potassium aluminate showed the appearance of bands centred at 1120, 1100, 930 and 890 cm−1. The degree of metal cation induced network clustering increases in the order K < Na < Li < Ca < Mg within the SiO2–MAlO2 glasses (M = K, Na, Li, Ca and Mg).207,208 The low frequency bands at 450, 500 and 600 cm−1 are associated with the motions of bridged oxygen in T–O–T linkages and the band at 560 cm−1 in the Al-rich sample is caused by the presence of Al–O–Al bridges in the CaO–Al2O3–SiO2 system. The intensity of a band at 800 cm−1 decreased with a decrease in SiO2 and shifted towards a lower wavenumber. The high frequency Raman spectra consists of peaks at 1000 and 1080 cm−1. Al coordination exists mainly in four-fold type and varies between Q2Si and Q4Si species as a function of the MO/Al2O3 ratio (M = Ca, Sr, Ba). At much higher concentrations of CaO (more than 62 mol%), a Raman band between 300 and 400 cm−1 was also present. A Raman band at 170 cm−1 and a shoulder around 300 cm−1 was observed in Ba containing glasses. The intensity of the peak at 170 cm−1 increases with Ba content. The peak position of Raman bands due to metal cations decreases with the size of the cation.209–212 A Raman band located at 450 cm−1 is characteristic of the motion of bridging oxygens in the plane perpendicular to the Si–O–Si (A1) bond. The vibration at 710 cm−1 provides evidence of AlO4 tetrahedra and the band at 760 cm−1 corresponds to the stretching mode of the Al–O bond with aluminium in four-fold coordination. The band at 960 cm−1 in the spectra is caused by stretching vibrations of silicon–oxygen tetrahedra with two corners shared with aluminium–oxygen or calcium–oxygen polyhedra.213 The depolarization ratios in calcium alumino-silicate glasses are 0.007 and 0.485 for intense bands at 540 and 96 cm−1, respectively. This result showed that the vibration at 540 cm−1 is highly symmetric in nature. The polarized spectrum for this glass system showed peaks centred at 96, 540, 767 and 867 cm−1. Similar bands are observed in other glass system with a slight variation in their peak positions.214 The main peak assignments in the Raman spectra of aluminosilicate glasses are listed in Table 6.

Table 6 Assignment of Raman bands in the spectra of Aluminosilicate glasses
Wavenumber (cm−1) Raman assignments Reference
1060, 1200 Si–O stretching vibrations of SiO4 tetrahedra 206
960 Stretching vibration of silicon–oxygen tetrahedra with two corners shared with aluminium–oxygen or calcium–oxygen polyhedra 213
810–820 Si–O stretching involving oxygen motions in the Si–O–Si plane or the motion of the Si atom in its oxygen cage or the threefold-degenerate rigid cage vibrational mode of TO2 units 204
779 Si–O–Si network, ZnO4 tetrahedra and AlO4 units with three BOs and one NBO 203
450, 500, 600 Motions of bridged oxygen in T–O–T linkages 209–212
400–650 Bending vibrations of the bridging oxygen (BO) bonds of SiO 203


3.7. Phosphosilicate glasses

A typical Raman spectrum of phosphosilicate glass is shown in Fig. 18.215 Raman spectra of phosphosilicate glasses consist of networks of phosphate as well as silicate. A Raman peak at 1320 cm−1 is attributed to the stretching vibrations of O–P double bonds and the bands below 860 cm−1 are assigned to Si–O and P–O bonds in Si–O–Si, P–O–Si and P–O–P linkages. The peaks at 1200 and 1020 cm−1 are caused by P–O–P linkages and the band near 1145 cm−1 is due to P–O–Si linkages.215,216 The bands are located at 340, 426, 580, 648, 720, 780, 882, 956, 1040 and 1450 cm−1 in the Raman spectra of various compositions of a CaO–MgO–SiO2 system doped with B2O3, P2O5, Na2O and CaF2 (Fig. 19).
image file: c5ra13043c-f18.tif
Fig. 18 Raman spectrum of phosphosilicate glass215 (reproduced with permission from Elsevier).

image file: c5ra13043c-f19.tif
Fig. 19 Raman spectra of calcium magnesium silicate glasses doped with B2O3, P2O5, Na2O and CaF2 (ref. 217) (reproduced with permission from Elsevier).

The vibrations in the region 800–1300 cm−1 are characteristic of the asymmetric vibration of SiO4 tetrahedra, whereas the bands at 952, 590 and 425 cm−1 are assigned to the symmetric stretching of the P–O bonds and O–P–O bending modes of the orthophosphate PO43− unit (Q2P), respectively.217 The silicate networks Si–O–Si and the intensity of the Si–O–NBO stretching mode are modified by cations Na, Mg and Ca.218,219 The addition of silver to calcium phosphosilicate glasses changes some networks, for example the contribution of Q0P, Q3P and Q4P progressively vanishes as a result of increasing the concentration of silver and a weak band around 1025 cm−1 also appears. The position of a band at 956 cm−1 was shifted to 967 cm−1 with this addition. The spectra also show the occurrence of a depolymerization process of the SiO4 network.220 The average distance between silicates, phosphates, inter silicate–phosphate and aluminium–phosphate chains was increased by increasing the concentrations of AlO6 structural units and this leads to an increase in the average bond lengths of Tm–O and Er–O due to weaker fields around Tm–O and Er–O ions.221 The main peak assignments in the Raman spectra of phosphosilicate glasses are listed in Table 7.

Table 7 Assignment of Raman bands in the spectra of phosphosilicate glasses
Wavenumber (cm−1) Raman assignments Reference
1320 Stretching vibrations of O–P double bond 215 and 216
1200, 1020 P–O–P linkages 215 and 216
1145 P–O–Si linkages 215 and 216
860 Si–O and P–O bonds in Si–O–Si, P–O–Si and P–O–P linkages 215 and 216


3.8. Alumino-borosilicate glasses

Raman spectra of NaF alumino-borosilicate glasses consist of vibrations located at 1075, 947, 802, 723 and 480 cm−1 (Fig. 20). The Raman band located at 802 cm−1 is characteristic of Si–O–Si vibrations and symmetric breathing vibrations of the boroxol rings. Sodium fluoride (NaF) in alumino-borosilicate glass increases the stability of the boroxol ring. The band at 723 cm−1 is assigned to the metaborate group, whereas lower intensity bands centered at 1075, 947 and 480 cm−1 are attributed to the stretching vibrations of Si–O–Si and B–O–Si. Instead of the abovementioned bands, a few more bands at 1114, 1000, 885, 770, 657, 598, 520, 491 and 447 cm−1 also appear in the Raman spectrum. The band located at 1114 cm−1 is attributed to diborate groups. The band at 885 cm−1 indicates the presence of pyroborate groups. The existence of symmetric breathing vibrations of six-membered rings in the BO4 tetrahedron was confirmed by a band located at 770 cm−1. The bands at 1000 and 520 cm−1 indicate the availability of the vibrations of Si–O, B–O–B, B–O–Si and Si–O–Si bending. The rocking vibration is also found in a network of glass at 475 cm−1. A Raman band at 951 cm−1 is attributed to the B–O–Si stretching vibration. The shoulder peak at 447 cm−1 is assigned to the bending or rocking vibrations of B–O–Si linkages, B–O–B, B–O–Si and Si–O–Si.222
image file: c5ra13043c-f20.tif
Fig. 20 Raman spectra of NaF alumino-borosilicate glasses222 (reproduced with permission from Elsevier).

Raman spectroscopy of sodium-alumino-borosilicate glasses in the presence of Gd2O3 showed that the band at 1050 cm−1 is shifted towards a lower wavenumber with the addition of Gd2O3 and the intensity of the band centred at 450 cm−1 was increased. One more band at 300 cm−1 appeared in glasses with 7.06 and 13.58 mol% of Gd2O3. The borate network was much influenced by network modifying cations as compared to the silicate network. Gd cations are more effective than Na cations with respect to the modification of networks.223 Raman spectra of strontium alumino-borosilicate glasses exhibit vibrations of borate as well as silicate networks along with some linking networks. Asymmetric ring breathing vibrations are characterised by a band at 430 cm−1, symmetric stretching vibrations of Si–O–Si by a band at 800 cm−1, stretching vibrations associated with SiO4 composed of sites with three and four-fold rings ranging from 1000 to 1200 cm−1 and symmetric stretching vibrations are characterised by a band at 1200 cm−1. The borate networks in pure B2O3 exhibit a strong band at 805 cm−1 due to the symmetric beating vibrations of boroxol ring oxygens. The boroxol ring structure is split into different ring structures (i.e. triborate, tetraborate or pentaborate) in the presence of suitable modifiers in the glass network (Fig. 21). After irradiation with γ-rays, the intensity of bands due to asymmetric vibrations increased and shifted to a higher wavenumber. The effect of radiation was less in Al2O3 containing glasses.224 Not only γ-ray irradiation but also β-ray irradiation changes the network of the glasses. The bending vibration of Si–O–Si at 460 cm−1 is shifted in glasses irradiated with 109 Gy. β-Ray radiation decreases the ratio of Q2Si species to Q3Si ones.225,226 β-Ray irradiation of rare earth alumino-borosilicate glasses reduces the number of Cr and Mn ions in the matrix. The Raman band at 460 cm−1 is shifted to 480 cm−1 with addition of Cr instead of Sm and Gd. Silicate networks in the range 900–1200 cm−1 are not significantly affected by Cr, whereas new bands centered at 845 and 900 cm−1 are introduced in the spectrum. The behaviour of the Raman band at ∼1550 cm−1 due to molecular oxygen could not be analyzed under irradiation due to polishing of the glass surface.226–229 As with β-ray irradiation, γ-irradiation is effective in reducing Sm3+ to Sm2+ ions with a 10 kGy dose.230 The Raman band assignments in alumino-borosilicate glasses are listed in Table 8.


image file: c5ra13043c-f21.tif
Fig. 21 Raman spectra of glass 40SrO–5Al2O3–15B2O3–40SiO2 irradiated with doses of (1) 0 kGy; (2) 10 kGy; and (3) 30 kGy (ref. 224) (reproduced with permission from Elsevier).
Table 8 Assignment of Raman bands in the spectra of alumino-borosilicate glasses
Wavenumber (cm−1) Raman assignments Reference
1114 Diborate groups 222
800–1300 Asymmetric vibration of SiO4 tetrahedra 222
1000, 520 Vibration of Si–O, B–O–B, B–O–Si and Si–O–Si bending 222
952 Symmetric stretching of the P–O bonds 222
951 B–O–Si stretching vibration 222
885 Pyroborate groups 222
770 Symmetric breathing vibration of six-membered rings, one BO4 tetrahedron 222
590, 425 O–P–O bending modes of the orthophosphate PO43− unit 222
447 Bending or rocking vibrations of the B–O–Si linkages, B–O–B, B–O–Si and Si–O–Si 222


3.9. Tellurite glasses

The Raman spectra of various tellurite glasses are shown in Fig. 22. In tellurite glasses, tellurium oxide works as a glass former. Tellurite oxide has three basic structures, namely, TeO4 (trigonal bipyramid, tbp), TeO3 (pyramid) and an intermediate with a TeO3+δ polyhedron. The four oxygen atoms are covalently bonded with a central tellurium atom in TeO4. The bipyramid structure is formed by two equatorial and two apex oxygen sites. The trigonal pyramid structure consists of two bridging oxygen sites and one non-bridging oxygen atom.231,232 The Raman bands near 667 cm−1 are related to the combined vibrations of asymmetric stretching of Te–eqOax–Te bonds and symmetric stretching of TeO4 tbps. The addition of Na2O to the glassy matrix changes the intensity of bands due to stretching vibrations of non-bridging Te–O bonds in TeO3 tps (753 and 792 cm−1). The position of a band assigned to the symmetric bending vibration of TeO4 tbps is shifted towards a lower wavenumber. The intense band at 40 cm−1 is associated with a Boson peak. An increase in the Na2O content of tellurite glasses results in the transformation of TeO4 tbps to TeO3 tps due to an increased number of NBOs.233,234 Assignments of the main Raman bands in the spectra of tellurite glasses are listed in Table 9.
image file: c5ra13043c-f22.tif
Fig. 22 Raman spectra of various tellurite glasses231 (reproduced with permission from Elsevier).
Table 9 Assignment of main Raman bands in the spectra of tellurite glasses
Wavenumber (cm−1) Raman assignments Reference
40 Boson peak 233 and 234
60–80 β-TeO2 239
110–143 γ-TeO2 239
300 TeO3 236
440–478 Stretching and bending vibration of Te–O–Te linkages in TeO4 (tbps), TeO3+δ polyhedra and TeO3 (tps) 231,235,236,243 and 246
581–623 Vibration of the continuous network comprised of TeO4 tbps 231
648–700 Antisymmetric vibrations of Te–O–Te in TeO3+δ, TeO4, TeO3 and TeO4 networks 231,233,234,239 and 250
716–753 Stretching vibrations between tellurium and non-bridging oxygen (NBOs) atoms 231,233,234,239 and 250
773–793 Continuous network vibration of TeO4 tbps and a TeO stretching vibration of TeO3+δ polyhedra or TeO3 231,233,234,239 and 250


The study of xNa2O·(35 − x)V2O5·65TeO2 glass showed an increase in the intensities of TeO4 tbp (443–478 cm−1 and 671–675 cm−1) and TeO3 tp (783–793 cm−1) with an increase in the Na2O content. This result is also consistent with the abovementioned result of the transformation of TeO4 tbps to TeO3 tps.235 The network Te–O–Te bridges are broken by additional network modifiers, Na2O, ZnO and ZnF2, in tellurite and form the non-bridging oxygens. The intensity of vibrations of TeO3 (300 cm−1) and Te–O–Te (470 cm−1)/TeO4 (675 cm−1) are reduced by the replacement of ZnO with ZnF2 in the glassy matrix. The tellurium chain was broken by the introduction of P2O5 to tellurite glasses, which suggests the formation of Te–O–P bonds.236 Raman peaks of crystallized sodium tellurite glasses are very sharp and intense. The peak positions were shifted towards a lower wavenumber with an increase in the crystallization temperature.237 The network of TeO4 bipyramids are broken to form networks of TeO3+1 and TeO3 polyhedra at high pressures.238 Raman spectra of zinc tellurite glasses confirm the presence of β-TeO2 (60–80 cm−1) and γ-TeO2 (110–143 cm−1). The three dimensional network of asymmetrical vibrations of TeO4 units with one lone pair of electrons at an equilateral position is linked to Teax–Oeq–Te linkages. The formation of Zn2Te3O8 units changes the bands positioned at 295, 343 and 700–800 cm−1.239–241 The peak positions of bands at 423 cm−1 is found to shift towards a lower wavenumber with increasing ZnO content, whereas bands at 661 and 720 cm−1 are shifted towards a higher wavenumber with increasing ZnO concentration. The intensities of bands at 661 cm−1 (decrease) and 720 cm−1 (increase) are reversed in order with an increase of ZnO content from 18 to 35 mol%. This indicates a decrease in Te–O coordination due to the conversion of TeO4 into TeO3 structural units with an increase in ZnO content.242 A Raman study of TeO2–La2O3–TiO2 glasses shows the chain formation of Te–O–Ti–O–Te– at a lower content of TiO2, whereas the chain formation becomes saturated and some TiO4 polyhedra are also formed in the glassy matrix.243 The structural change from TeO4 trigonal bipyramids (tbps) to TeO3 trigonal pyramids (tps) via [TeO3+1] is also observed with an increasing Ta2O5 content in glass.244 The Raman spectra of (89 − x) TeO2–10TiO2–1Nd2O3xWO3 indicates the depolymerisation of tellurite glass as the Te–O–Te inter-chain linkages are progressively substituted by stronger Te–O–W bridges as a result of the addition of WO3.245 The addition of Nb2O5 in tellurite glasses decreases the connectivity by the deformation of Te–O–Te linkages.246 The network modifying nature of K+ ions is greater than that of Li+ ions in the alkali metal tellurite glasses due to the larger ionic radius of potassium compared to that of lithium.247 The thallium ions in tellurite glasses induce the island-like structure of [TeO3]2−.248 The alkali ion (Li and Na) transports drive the secondary event as NBO migration through BO–NBO switching.249 Khanna et al. studied the Raman spectra of boro-tellurite and aluminoboro-tellurite glasses. A Raman spectrum of boro-tellurite glass depicts bands at 450, 503, 615, 665, 718, 762, and 902 cm−1, whereas alumino-boro-tellurite glasses have Raman bands located at 337, 450, 505, 567, 606, 666, 724, 754 and 864 cm−1.250 The structural units TeO4 tbps and TeO3+1 polyhedra change into TeO3 tps with an increase in the B2O3 content.251 The diamond in tellurite glasses is a good medium for quantum information. The first order diamond Raman band is at 1350 cm−1 in nano-diamond induced tellurite glasses.252

4. Conclusions

A detailed study of Raman spectroscopy data on various alkali and alkaline earth metal doped oxide glasses is reviewed. The various models were helpful in the accurate correlation of data. A general agreement was reached that the structure of glassy networks is composition dependent. The vibrational properties of structural units are better understood with respect to their structure and bonding as vibrational spectroscopy cannot be used as a tool for analyzing their detailed molecular structure. Raman spectroscopy studies clarify the modification of vibrational networks of borate by alkali and alkaline earth metals. The basic borate networks are formed by NBO and BO atoms in a tetrahedral borate unit.9 The formation of pentaborate, tetraborate, diborate, metaborate and pyroborate units10 and the disappearance of diborate units,102 boroxol rings116 and high polymeric units such as pentaborate105 caused by certain dopants was clearly noted by Raman spectroscopy studies of various borate glasses. Raman spectroscopy of oxide glasses showed that the modifying nature of alkali and alkaline earth metals are in the order Cs < K < Na < Li < Ca < Mg.23 The alkaline metals are effective in the order Ba < Sr < Ca < Mg, whereas the alkali metals are in the order K < Na < Li. Significant information regarding the glass forming ability of various glassy networks was obtained from Raman spectroscopy studies in the present review article. On the basis of the abovementioned investigations of oxide glasses, dopants MoO3, CeO2, Bi2O3, Sb2O3 and TeO2 form mainly MoO6 or MoO4, CeO8, BiO3 or BiO6, SbO3 and TeO3 or TeO4 units, respectively. Raman studies also provided significant information about UV, β and γ-ray irradiation. It was found that these radiations decrease the Q2Si species compared to Q3Si species and reduce the oxidation state of cations.230 Tellurium oxide works as a network former as well as a network modifier. With the addition of alkali metal oxides, TeO4 tbps units transform into TeO3 tps due to an increased number of NBOs.

Great advances have been seen in the structural studies of glasses in the past few decades; however, a few questions remain open to study. The presence or absence of super-structural units in the glass network is the most important question. The mesounits in an intermediate- or medium-range order in the oxide structure affect a variety of the properties of the glass and are constituted by 7–20 or more atoms. The primary information about mesounits and their possible arrangements can be derived from the study of crystal structures that are isocompositional with the glasses. Such types of information could be obtained from Raman spectroscopy. Quantitative information on structural units in glasses may also be obtained by Raman spectroscopy analysis. It is always a question whether information obtained by Raman spectroscopy is complete or not. The vibrations, which are antisymmetrical with respect to the centre, are Raman inactive. In this case, researchers might question whether anti-symmetric vibrations can be analysed by Raman spectroscopy. Such vibrations are analysed by infrared spectroscopy. What do the analogies with organic high polymers teach us with respect to the glass? In what respect can a given glass be called a polymer?

The absence of a change of P-coordination in phosphate glasses as found in borate is not explained by the structural study of phosphate glasses by Raman spectroscopy. No particular relationship is seen between the distribution of tetrahedral linkages and the glass properties. A question also arises whether there are any differences between the non-bridging oxygens on individual Q2P and Q1P tetrahedra.

Acknowledgements

The authors are thankful to the publishers Elsevier, Mineralogical Society of America, Springer, IOP science and EBSCO Publishing for permitting the reproduction of cited contributions for academic purposes in this review article. The authors acknowledge the authors of cited articles. The funding agency CSIR, New Delhi, (India) is gratefully acknowledged for financial support under project No. 03(1300)/13/EMR-II.

References

  1. L. Dussubieux, B. Gratuze and M. Blet-Lemarquand, J. Archaeol. Sci., 2010, 37, 1646 CrossRef .
  2. A. Macfarlane and G. Martin, Glass: A World History, University of Chicago Press, 2002 Search PubMed .
  3. R. K. Brow and M. L. Schmitt, J. Eur. Ceram. Soc., 2009, 29, 1193 CrossRef CAS .
  4. B. Wei, H. Cao and S. Song, Mater. Des., 2010, 31, 4244 CrossRef CAS .
  5. A. Goel, M. J. Pascual and J. M. F. Ferreira, Int. J. Hydrogen Energy, 2010, 35, 6911 CrossRef CAS .
  6. W. H. Zachariasen, J. Am. Chem. Soc., 1932, 54, 3841 CrossRef CAS .
  7. P. B. Price, L. M. Cook and A. Markar, Nature, 1987, 325, 137 CrossRef CAS .
  8. T. Lopez, E. Haro-Poniatowski, P. Bosh, M. Asomoza, R. Gomez, M. Massot and M. Balkanski, J. Sol-Gel Sci. Technol., 1994, 2, 891 CrossRef CAS .
  9. C.-P. E. Varsamis, A. Vegiri and E. I. Kamitsos, Condens. Matter Phys., 2001, 4, 119 CrossRef .
  10. K. S. Oo, M. Lwin, P. Kaung and S. Htoon, Journal of the Myanmar Academy of Arts and Science, 2006, 4, 277 Search PubMed .
  11. G. Padmaja and P. Kistaiah, J. Phys. Chem. A, 2009, 113, 2397 CrossRef CAS PubMed .
  12. N. C. A. de Souza, C. C. Santos, I. Guedes, N. O. Dantas and M. V. D. Vermelho, Opt. Mater., 2013, 35, 2544 CrossRef CAS .
  13. Q. Chen, Q. Chen and S. Wang, Open J. Inorg. Non-Met. Mater., 2011, 1, 1 CAS .
  14. S. Bale and S. Rahman, ISRN Spectrosc., 2012, 2012, 1 Search PubMed .
  15. Y. Wang, T. Honma and T. Komatsu, J. Non-Cryst. Solids, 2014, 383, 86 CrossRef CAS .
  16. P. McMillan, Am. Mineral., 1984, 69, 622 CAS .
  17. P. F. McMillan and R. L. Remmele Jr., Am. Mineral., 1986, 71, 772 CAS .
  18. Y. Jinglin, J. Guochang, C. Hui and X. Kuangdi, Rare Met., 2006, 25, 431 Search PubMed .
  19. R. P. Hapanowicz and R. A. Condrate Sr., J. Solid State Chem., 1996, 123, 183 CrossRef CAS .
  20. M. Wang, J. Cheng, M. Li and F. He, Phys. B, 2011, 406, 3865 CrossRef CAS .
  21. A. G. Kalampounias, Bull. Mater. Sci., 2011, 34, 299 CrossRef CAS .
  22. B. O. Mysen, L. W. Finger, D. Vigro and F. A. Seifert, Am. Mineral., 1982, 67, 686 CAS .
  23. P. McMillan, Am. Mineral., 1984, 69, 645459 Search PubMed .
  24. L. D. Ferri, P. P. Lottici and G. Vezzalini, Corros. Sci., 2014, 80, 434 CrossRef .
  25. H. Jia, G. Chen and W. Wang, Opt. Mater., 2006, 29, 445 CrossRef CAS .
  26. H. Liu, J. Ma, J. Gong and J. Xu, J. Non-Cryst. Solids, 2015, 419, 92 CrossRef CAS .
  27. M. Karabulut, H. Ertap and M. Yüksek, J. Non-Cryst. Solids, 2015, 417–418, 39 CrossRef CAS .
  28. L. Ma, R. K. Brow, L. Ghussn and M. E. Schlesinger, J. Non-Cryst. Solids, 2015, 409, 131 CrossRef CAS .
  29. M. Farouk, J. Non-Cryst. Solids, 2014, 402, 74 CrossRef CAS .
  30. L. Song, J. Wu, Z. Li, X. Hao and Y. Yu, J. Non-Cryst. Solids, 2015, 419, 16 CrossRef CAS .
  31. P. Stoch and A. Stoch, J. Non-Cryst. Solids, 2015, 411, 106 CrossRef CAS .
  32. P. G. Debenedetti and F. H. Stillinger, Nature, 2001, 410, 259 CrossRef CAS PubMed .
  33. M. Yamane and Y. Asahara, Glasses for Photonics, Cambridge University Press, Cambridge, 2000 Search PubMed .
  34. A. L. Greer, Nature, 1999, 402, 132 CrossRef CAS .
  35. E. Axinte, Mater. Des., 2011, 32, 1717 CrossRef CAS .
  36. J. Schroers, Nature, 2014, 512, 142 CrossRef CAS PubMed .
  37. N. Karpukhina, R. G. Hilla and R. V. Law, Chem. Soc. Rev., 2014, 43, 2174 RSC .
  38. G. W. Morey, J. Ind. Eng. Chem., 1922, 823 CrossRef CAS .
  39. J. Kieffer, J. Phys. Chem. B, 1999, 103, 4153 CrossRef CAS .
  40. A. K. Varshneya and J. C. Mauro, Glass Technol.: Eur. J. Glass Sci. Technol., Part A, 2010, 51, 28 CAS .
  41. L. Wondraczek, S. Sen, H. Behrens and R. E. Youngman, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 014202 CrossRef .
  42. J. C. Mauro, D. C. Allan and M. Potuzak, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 094204 CrossRef .
  43. J. C. Mauro and M. M. Smedskjaer, J. Non-Cryst. Solids, 2014, 396–397, 41 CrossRef CAS .
  44. J. E. Stanworth, Nature, 1952, 169, 581 CrossRef CAS .
  45. C. R. Gautam and A. K. Yadav, ISRN Ceram., 2012, 2012, 1 CrossRef .
  46. M. C. Eckersley, P. H. Gaskell, A. C. Barnes and P. Chieux, Nature, 1988, 335, 525 CrossRef CAS .
  47. S. Singh, M. D. Ediger and J. J. de Pablo, Nat. Mater., 2013, 12, 139 CrossRef CAS PubMed .
  48. Z.-H. Jiang and Q.-Y. Zhang, Prog. Mater. Sci., 2014, 61, 144 CrossRef CAS .
  49. M. Poulain, Nature, 1981, 293, 279 CrossRef CAS .
  50. A. Fluegel, Glass Technol.: Eur. J. Glass Sci. Technol., Part A, 2007, 48, 13 CAS .
  51. A. K. Yadav and C. R. Gautam, Adv. Appl. Ceram., 2014, 113, 193 CrossRef .
  52. A. K. Yadav and C. R. Gautam, Spectrosc. Lett., 2015, 48, 514 CrossRef CAS .
  53. A. K. Yadav and C. R. Gautam, Adv. Appl. Ceram., 2014, 113, 193 CrossRef .
  54. M. Yamane and Y. Asahara, Glasses for photonics, Cambridge University, Cambridge, UK, 2000 Search PubMed .
  55. J. C. Phillips and M. F. Thorpe, Solid State Commun., 1985, 53, 699 CrossRef CAS .
  56. P. K. Gupta and J. C. Mauro, J. Chem. Phys., 2009, 130, 094503 CrossRef PubMed .
  57. P. G. Pavani, K. Sadhana and V. C. Mouli, Phys. B, 2011, 406, 1242 CrossRef .
  58. K. Cho, J. Oh, T. Lee and D. Shin, J. Power Sources, 2008, 183, 431 CrossRef CAS .
  59. S. Baccaro, N. Catallo, A. Cemmi and G. Sharma, Nucl. Instrum. Methods Phys. Res., Sect. B, 2011, 269, 167 CrossRef CAS .
  60. Y. J. Chen, X. H. Gong, Y. F. Lin, Z. D. Luo and Y. D. Huang, Opt. Mater., 2010, 33, 71 CrossRef CAS .
  61. S. P. Singh, R. P. S. Chakradhara, J. L. Rao and B. Karmakar, J. Alloys Compd., 2010, 493, 256 CrossRef CAS .
  62. R. Gehring, The chemistry of glass, Scientific American Inc, 1890, vol. 737, p. 11776 Search PubMed .
  63. R. Dupree, D. Holland and M. G. Mortuza, Nature, 1987, 328, 416 CrossRef CAS .
  64. L. Cormier, G. Calas and B. Beuneu, J. Non-Cryst. Solids, 2011, 357, 926 CrossRef CAS .
  65. A. M. Efimov, J. Non-Cryst. Solids, 1999, 253, 95 CrossRef CAS .
  66. E. M. Rabinovich, J. Mater. Sci., 1976, 11, 925 CrossRef CAS .
  67. M. Karabulut, E. Melnik, R. Stefan, G. K. Marasinghe, C. S. Ray, C. R. Kurkjin and D. E. Day, J. Non-Cryst. Solids, 2001, 288, 8 CrossRef CAS .
  68. K. El-Egili, H. Doweidar, Y. M. Moustafa and I. Abbas, Phys. B, 2003, 339, 237 CrossRef CAS .
  69. G. Guimbretière, D. Bégué, M. Dussauze and V. Rodriguez, Vib. Spectrosc., 2012, 63, 426 CrossRef .
  70. P. B. Price, L. M. Cook and A. Marker, Nature, 1987, 325, 137 CrossRef CAS .
  71. A. A. Osipov, L. M. Osipova and V. E. Eremyashev, Glass Phys. Chem., 2013, 39, 105 CrossRef CAS .
  72. O. N. Koroleva, L. A. Shabunina and V. N. Bykov, Glass Ceram., 2011, 67, 340 CrossRef CAS .
  73. A. Goel, J. S. McCloy, C. F. Windisch Jr, B. J. Riley, M. J. Schweiger, C. P. Rodriguez and J. M. F. Ferreira, Int. J. Appl. Glass Sci., 2012, 1 CAS .
  74. J. Blizard, Ind. Eng. Chem., 1947, 39, 1215 CrossRef CAS .
  75. L. Koudelka, P. Mosner and J. Šubčík, IOP Conf. Ser.: Mater. Sci. Eng., 2009, 2, 012015 CrossRef .
  76. D. Kim, C. Hwang, D. Gwoo, T. Kim, N. Kim and B.-K. Ryu, Electron. Mater. Lett., 2011, 7, 343 CrossRef CAS .
  77. A. Naumaan and J. T. Boyd, J. Electrochem. Soc., 1980, 127, 1414 CrossRef CAS .
  78. A. Tilocca, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 224202 CrossRef .
  79. S. W. Lee, K. S. Ryoo, J. E. Kim, J. H. Lee, C. D. Kim and K. S. Hong, J. Mater. Sci., 1994, 29, 4577 CrossRef CAS .
  80. J. L. Rygel, Y. Chen, C. G. Pantano, T. Shibata, J. Du, L. Kokou, R. Woodman and J. Belcher, J. Am. Ceram. Soc., 2011, 94, 2442 CrossRef CAS .
  81. R. W. Cahn, Nature, 1989, 341, 183 CrossRef .
  82. M. Handke, M. Sitarz, M. Rokita and E. Galuskin, J. Mol. Struct., 2003, 651, 39 CrossRef .
  83. L. Cormier and D. R. Neuville, Chem. Geol., 2004, 213, 103 CrossRef CAS .
  84. S. L. S. Rao, G. Ramadevudu, M. Shareefuddin, A. Hameed, M. N. Chary and M. L. Rao, Int. J. Eng. Sci. Tech., 2012, 4, 25 Search PubMed .
  85. D. P. Singh and G. P. Singh, J. Alloys Compd., 2013, 546, 224 CrossRef CAS .
  86. G. Calas, L. Cormier, L. Galoisy and P. Jollivet, C. R. Chim., 2002, 5, 831 CrossRef CAS .
  87. J. Qiu, X. Jiang, C. Zhu, H. Inouye, J. Si and K. Hirao, Opt. Lett., 2004, 29, 370 CrossRef CAS PubMed .
  88. Q. Yu, F. Chen, T. Xu, S. Dai and Q. Zhang, J. Non-Cryst. Solids, 2013, 378, 254 CrossRef CAS .
  89. Introduction to Modern Vibrational Spectroscopy, ed. M. Diem, Wiley-Interscience, New York, 1993 Search PubMed .
  90. C. V. Raman and K. S. Kirishnan, Nature, 1928, 121, 501 CrossRef CAS .
  91. C. V. Raman, Indian J. Phys., 1928, 2, 387–398 CAS .
  92. C. C. Lin, M. T. Kuo and H. C. Chang, J. Med. Biol. Eng., 2010, 30, 343–354 CrossRef .
  93. C. V. Raman and K. S. Krishnan, Indian J. Phys., 1928, 2, 399 CAS .
  94. C. V. Raman and K. S. Krishnan, Nature, 1928, 121, 619 CrossRef CAS .
  95. L. A. Lyon, C. D. Keating, A. P. Fox, B. E. Baker, H. Lin, S. R. Nicewarner, S. P. Mulvaney and M. J. Natan, Anal. Chem., 1998, 70, 341R CrossRef CAS PubMed .
  96. H. H. Willard Jr., L. L. Merritt and J. A. Dean, Raman Spectroscopy, in Instrumental Methods of Analysis, D. Van Nostrand Co., New York, 5th edn, 1974, p. 189 Search PubMed .
  97. B. H. Stuart, Infrared Spectroscopy, Fundamentals and Applications, J. Wiley, Chichester, West Essex, England, 2004, p. 2 Search PubMed .
  98. M. S. Shackley, X-Ray Fluorescence Spectrometry (XRF), in Geoarchaeology, Springer, New York, 2010, p. 16 Search PubMed .
  99. T. Lopez, E. H. Poniatowski, P. Bosh, M. Asomoza, R. Gomez, M. Massot and M. Balkanski, J. Sol-Gel Sci. Technol., 1994, 2, 891 CrossRef CAS .
  100. M. Loubser, Chemical and Physical Aspects of Lithium Borate Fusion, Distertation M. Sc., University of Pretoria, Pretoria, 2009 .
  101. W. Di, W. Song-ming, Z. Qing-li, S. Dun-lu, Y. Shao-tang, Y. Jing-lin and W. Yuan-yuan, Chin. J. Quantum Electron., 2011, 28, 241 Search PubMed .
  102. M. Ganguli and K. J. Rao, J. Solid State Chem., 1999, 145, 65 CrossRef CAS .
  103. G. Padmaja and P. Kistaiah, J. Phys. Chem. A, 2009, 113, 2397 CrossRef CAS PubMed .
  104. T. Yano, N. Kunimine, S. Shibata and M. Yamane, J. Non-Cryst. Solids, 2003, 321, 137 CrossRef CAS .
  105. Q. Jiao, X. Yu, X. Xu, D. Zhou and J. Qiu, J. Solid State Chem., 2013, 202, 65 CrossRef CAS .
  106. A. A. A. Shama and F. H. El-Batal, Egypt. J. Solid., 2006, 29, 49 Search PubMed .
  107. M. Affatigato, S. Feller, A. K. Schue, S. Blair, D. Stentz, G. B. Smith, D. Liss, M. J. Kelley, C. Goater and R. Leelesagar, J. Phys.: Condens. Matter, 2003, 15, S2323 CrossRef CAS .
  108. Q. Chen, Q. Chen and S. Wang, Open J. Inorg. Non-Met. Mater., 2011, 1, 1 CAS .
  109. V. Timar, R. Lucăcel-Ciceo and I. Ardelean, Semicond. Phys., Quantum Electron. Optoelectron., 2008, 11, 221 CAS .
  110. J. Schwarz and H. Tichá, J. Optoelectron. Adv. Mater., 2003, 5, 69 CAS .
  111. A. A. Alemi, H. Sedghi, A. R. Mirmohseni and V. Golsanamlu, Bull. Mater. Sci., 2006, 29, 55 CrossRef CAS .
  112. A. Toderas, M. Toderas, S. Filip and I. Ardelean, Optoelectron. Adv. Mater., Rapid Commun., 2012, 6, 331 CAS .
  113. B. S. Cristian, Study of some structural and physical properties of some B2O3 based glasses containing Ag2O and 3d ions, Ph.D. thesis, Babeş-Bolyai University, 2011 .
  114. S. C. Baidoc and I. Ardelean, J. Optoelectron. Adv. Mater., 2008, 10, 3205 CAS .
  115. R. Uning, W. M. Hua and R. Hussin, Structural studies of strontium borate glass doped with Zn2+, Mn2+ and Fe2+ ions, UMTAS, 2011 Search PubMed .
  116. S. R. Rejisha, P. S. Anjana, N. Gopakumar and N. Santha, J. Non-Cryst. Solids, 2014, 388, 68 CrossRef CAS .
  117. M. Bettinelli, A. Speghini, M. Ferrari and M. Montagna, J. Non-Cryst. Solids, 1996, 201, 211 CrossRef CAS .
  118. S. Suresh, M. Prasad, G. Upender, V. Kamalaker and V. C. Mouli, Indian J. Pure Appl. Phys., 2009, 47, 163 CAS .
  119. D. Maniu, I. Ardelean, T. Iliescu, S. Cınta, V. Nagel and W. Kiefer, J. Mol. Struct., 1999, 480, 657 CrossRef .
  120. D. Maniu, T. Iliescu, I. Ardelean, S. Cinta-Pinzaru, N. Tarcea and W. Kiefer, J. Mol. Struct., 2003, 651, 485 CrossRef .
  121. C. N. Santos, D. D. S. Meneses, P. Echegut, D. R. Neuville, A. C. Hernandes and A. Ibanez, Appl. Phys. Lett., 2009, 94, 151901 CrossRef .
  122. R. S. Barker, D. A. Richardson, E. A. G. Mcconkey and R. Rimmer, Nature, 1960, 187, 135 CrossRef CAS .
  123. B. Tiwari, A. Dixit, G. P. Kothiyal, M. Pandey and S. K. Deb, Preparation and characterization of phosphate glasses containing titanium, National Symposium on Science & Technology of Glass and Glass-Ceramics (NSGC-06) held during September 15–16, at BARC, Mumbai, 2006.
  124. L. F. Santos, R. M. Aldeida, V. K. Tikhomirov and A. Jha, J. Non-Cryst. Solids, 2001, 284, 43 CrossRef CAS .
  125. A. Mogus-Milankovic, A. Gajovic, A. Santic and D. E. Day, J. Non-Cryst. Solids, 2001, 289, 204 CrossRef CAS .
  126. K. Brahmachary, D. Rajesh, S. Babu and Y. C. Ratnakaram, J. Mol. Struct., 2014, 1064, 6 CrossRef CAS .
  127. A. Kiani, J. V. Hanna, S. P. King, G. J. Rees, M. E. Smith, N. Roohpour, V. Salih and J. C. Knowles, Acta Biomater., 2012, 8, 333 CrossRef CAS PubMed .
  128. M. Elisa, I. C. Vasiliu, C. E. A. Grigorescu, B. Grigoras, H. Niciu, D. Niciu, A. Meghea, N. Iftimie, M. Giurginca, H. J. Trodahl and M. Dalley, Opt. Mater., 2006, 28, 621 CrossRef CAS .
  129. F. H. El-Batal, Indian J. Pure Appl. Phys., 2009, 47, 631 CAS .
  130. D. Toloman, R. Ciceo-Lucacel, D. A. Magdas, A. Regos, A. R. Biris, C. Leostean and I. Ardelean, J. Alloys Compd., 2013, 556, 67 CrossRef CAS .
  131. B. Qian, X. Liang, C. Wang and S. Yang, J. Nucl. Mater., 2013, 443, 140 CrossRef CAS .
  132. C. I. R. D. Oliveira, L. F. C. D. Oliveira, F. A. D. Filho, Y. Messaddeq and S. J. L. Ribeiro, Spectrochim. Acta, Part A, 2005, 61, 2023 CrossRef PubMed .
  133. D. Ilieva, B. Jivov, G. Bogachev, C. Petkov, I. Penkov and Y. Dimitriev, J. Non-Cryst. Solids, 2001, 283, 195 CrossRef CAS .
  134. D. Ilieva, B. Jivov, D. Kovacheva, T. Tsacheva, Y. Dimitriev, G. Bogachev and C. Petkov, J. Non-Cryst. Solids, 2001, 293–295, 562 CrossRef CAS .
  135. L. Koudelka, I. Rösslerová, J. Holubová, P. Mošner, L. Montagne and B. Revel, J. Non-Cryst. Solids, 2011, 357, 2816 CrossRef CAS .
  136. G. L. Flower, M. S. Reddy, G. S. Baskaran and N. Veeraiah, Opt. Mater., 2007, 30, 357 CrossRef CAS .
  137. S. S. Sastry and B. R. V. Rao, Phys. B, 2014, 434, 159 CrossRef CAS .
  138. G. le Saout, F. Fayon, C. Bessada, P. Simon, A. Blin and Y. Vaills, J. Non-Cryst. Solids, 2001, 293–295, 657 CrossRef CAS .
  139. W. A. Pisarski, L. Zur, T. Goryczka, M. Sołtys and J. Pisarska, J. Alloys Compd., 2014, 587, 90 CrossRef CAS .
  140. A. M. Milankovic, A. Santic, S. T. Reis, K. Furic and D. E. Day, J. Non-Cryst. Solids, 2005, 351, 3246 CrossRef .
  141. G. S. Baskaran, G. L. Flower, D. K. Rao and N. Veeraiah, J. Alloys Compd., 2007, 431, 303 CrossRef CAS .
  142. A. Šantić, Ž. Skoko, A. Gajović, S. T. Reis, D. E. Day and A. M. Milanković, J. Non-Cryst. Solids, 2011, 357, 3578 CrossRef .
  143. A. Santic, A. M. Milankovic, K. Furic, V. Bermanec, C. W. Kim and D. E. Day, J. Non-Cryst. Solids, 2007, 353, 1070 CrossRef CAS .
  144. A. M. Milankovic, A. Santic, A. Gajovic and D. E. Day, J. Non-Cryst. Solids, 2003, 325, 76 CrossRef .
  145. B. Qian, X. Liang, S. Yang, S. He and L. Gao, J. Mol. Struct., 2012, 1027, 31 CrossRef CAS .
  146. K. Joseph, M. Premila, G. Amarendra, K. V. G. Kutty, C. S. Sundar and P. R. V. Rao, J. Nucl. Mater., 2012, 420, 49 CrossRef CAS .
  147. Y. M. Lai, X. F. Liang, S. Y. Yang, J. X. Wang, L. H. Cao and B. Dai, J. Mol. Struct., 2011, 992, 84 CrossRef CAS .
  148. Y. M. Lai, X. F. Liang, S. Y. Yang, J. X. Wang and B. T. Zhang, J. Mol. Struct., 2012, 1013, 134 CrossRef CAS .
  149. S. Chakraborty and A. K. Arora, Vib. Spectrosc., 2012, 61, 99 CrossRef CAS .
  150. P. S. Rao, P. M. V. Teja, A. R. Babu, C. Rajyasree and D. K. Rao, J. Non-Cryst. Solids, 2012, 358, 3372 CrossRef CAS .
  151. N. Kitamura, K. Fukumi, J. Nakamura, T. Hidaka, T. Ikeda, H. Hashima and J. Nishii, J. Non-Cryst. Solids, 2011, 357, 1188 CrossRef CAS .
  152. P. S. Rao, S. B. M. Krishna, S. Yusub, P. R. Babu, C. Tirupataiah and D. K. Rao, J. Mol. Struct., 2013, 1036, 452 CrossRef .
  153. S. Thomas, R. George, M. Rathaiah, V. Venkatramu, S. N. Rasool and N. V. Unnikrishnan, Phys. B, 2013, 431, 69 CrossRef CAS .
  154. S. H. Santagneli, G. Poirier, M. T. Rinke, S. J. L. Ribeiro, Y. Messaddeq and H. Eckert, J. Non-Cryst. Solids, 2011, 357, 2126 CrossRef CAS .
  155. R. C. Lucacel, A. O. Hulpus, V. Simonn and I. Ardelean, J. Non-Cryst. Solids, 2009, 355, 425 CrossRef .
  156. A. M. B. Silva, R. N. Correia, J. M. M. Oliveira and M. H. V. Fernandes, J. Eur. Ceram. Soc., 2010, 30, 1253 CrossRef CAS .
  157. I. D. F. Gimenez, I. O. Mazali and O. L. Alves, J. Phys. Chem. Solids, 2001, 62, 1251 CrossRef .
  158. S. N. Rasool, L. R. Moorthy and C. K. Jayasankar, Opt. Commun., 2013, 311, 156 CrossRef .
  159. M. Elisa, B. A. Sava, I. C. Vasiliu, R. C. C. Monteiro, J. P. Veiga, L. Ghervase, I. Feraru and R. Iordanescu, J. Non-Cryst. Solids, 2013, 369, 55 CrossRef CAS .
  160. N. Vedeanu, O. Cozar, I. Ardelean, B. Lendl and D. A. Magdas, Vib. Spectrosc., 2008, 48, 259 CrossRef CAS .
  161. A. Chahine, M. Et-tabirou and J. L. Pascal, Mater. Lett., 2004, 58, 2776 CrossRef CAS .
  162. E. Metwalli, M. Karabulut, D. L. Sidebottom, M. M. Morsi and R. K. Brow, J. Non-Cryst. Solids, 2004, 344, 128 CrossRef CAS .
  163. N. Kerkouri, M. Haddad, M. Et-Tabirou, A. Chahine and L. Laanab, Phys. B, 2011, 406, 3142 CrossRef CAS .
  164. C. Ivascu, I. B. Cozar, L. Daraban and G. Damian, J. Non-Cryst. Solids, 2013, 359, 60 CrossRef CAS .
  165. C. Ivascu, A. Timar Gabor, O. Cozar, L. Daraban and I. Ardelean, J. Mol. Struct., 2011, 993, 249 CrossRef CAS .
  166. S. Hraiech, M. Ferid, Y. Guyot and G. Boulon, J. Rare Earths, 2013, 31, 685 CrossRef CAS .
  167. B. G. Parkinson, D. Holland, M. E. Smith, C. Larson, J. Doerr, M. Affatigato, S. A. Feller, A. P. Howes and C. R. Scales, J. Non-Cryst. Solids, 2008, 354, 1936 CrossRef CAS .
  168. D. Manara, A. Grandjean and D. R. Neuville, Am. Mineral., 2009, 94, 777 CrossRef CAS .
  169. M. Lenoir, A. Grandjean, S. Poissonnet and D. R. Neuville, J. Non-Cryst. Solids, 2009, 355, 1468 CrossRef CAS .
  170. M. Lenoir, A. Grandjean, Y. Linard, B. Cochain and D. R. Neuville, Chem. Geol., 2008, 256, 316 CrossRef CAS .
  171. M. Jroda and C. Paluszkiewicz, J. Mol. Struct., 2007, 834–836, 302 Search PubMed .
  172. V. E. Eremyashev, A. A. Osipov and L. M. Osipov, Glass Ceram., 2011, 68, 205 CrossRef CAS .
  173. A. Winterstein-Beckmann, D. Möncke, D. Palles, E. I. Kamitsos and L. Wondraczek, J. Non-Cryst. Solids, 2014, 401, 110 CrossRef CAS .
  174. H. Yamashita, K. Nagata, H. Yoshino, K. Ono and T. Maekawa, J. Non-Cryst. Solids, 1999, 248, 115 CrossRef CAS .
  175. D. Chen, H. Miyoshi, H. Masui, T. Akai and T. Yazawa, J. Non-Cryst. Solids, 2004, 345–346, 104 CrossRef .
  176. D. A. McKeown, I. S. Muller, H. Gan, I. L. Pegg and C. A. Kendziora, J. Non-Cryst. Solids, 2001, 288, 191 CrossRef CAS .
  177. D. A. McKeown, I. S. Muller, H. Gan, I. L. Pegg and W. C. Stolte, J. Non-Cryst. Solids, 2004, 333, 74 CrossRef CAS .
  178. X. Zhang, Y. Wang, J. Lu, J. Zang, J. Zhang and E. Ge, Int. J. Refract. Met. Hard Mater., 2010, 28, 260 CrossRef CAS .
  179. D. A. McKeown, I. S. Muller, A. C. Buechele, I. L. Pegg and C. A. Kendziora, J. Non-Cryst. Solids, 2000, 262, 126 CrossRef CAS .
  180. T. G. V. M. Rao, A. R. Kumar, K. Neeraja, N. Veeraiah and M. R. Reddy, J. Alloys Compd., 2013, 557, 209 CrossRef CAS .
  181. Z. Wang, Q. Shu and K. Chou, ISIJ Int., 2011, 51, 1021 CrossRef CAS .
  182. L. Rezazadeh, S. Baghshahi, A. Nozad Golikand and Z. Hamnabard, Ionics, 2013, 1 Search PubMed .
  183. A. K. Yadav, C. R. Gautam, A. Gautam and V. K. Mishra, Phase Transitions, 2013, 86, 1000 CrossRef CAS .
  184. A. K. Yadav and C. R. Gautam, Opt. Photonics J., 2013, 3, 1 Search PubMed .
  185. C. R. Gautam, A. K. Yadav, V. K. Mishra and K. Vikram, Open J. Inorg. Non-Met. Mater., 2012, 2, 47 CAS .
  186. G. Kaur, O. P. Pandey and K. Singh, J. Non-Cryst. Solids, 2012, 358, 2589 CrossRef CAS .
  187. C. R. Gautam, A. K. Singh and A. K. Yadav, Int. J. Appl. Nat. Sci., 2012, 1, 69 Search PubMed .
  188. G. F. Zhang, T. S. Wang, K. J. Yang, L. Chen, L. M. Zhang, H. B. Peng, W. Yuan and F. Tian, Nucl. Instrum. Methods Phys. Res., Sect. B, 2013, 316, 218 CrossRef CAS .
  189. C. Gejke, E. Zanghellini, J. Swenson and L. Borjesson, J. Power Sources, 2003, 119–121, 576 CrossRef CAS .
  190. P. Mošner, M. Vorokhta, L. Koudelka, L. Montagne, B. Revel, K. Sklepić and A. M. Milanković, J. Non-Cryst. Solids, 2013, 375, 1 CrossRef .
  191. N. Sdiri, H. Elhouichet and M. Ferid, J. Non-Cryst. Solids, 2014, 389, 38 CrossRef CAS .
  192. K. J. Rao, K. C. Sobha and S. Kumar, Proc.–Indian Acad. Sci., Chem. Sci., 2001, 113, 497 CrossRef CAS .
  193. A. Saranti, I. Koutselas and M. A. Karakassides, J. Non-Cryst. Solids, 2006, 352, 390 CrossRef CAS .
  194. L. Koudelka, J. Pospisil, P. Mosner, L. Montagne and L. Delevoye, J. Non-Cryst. Solids, 2008, 354, 129 CrossRef CAS .
  195. L. Koudelka, P. Mosner, M. Zeyer and C. Jager, J. Non-Cryst. Solids, 2003, 326–327, 72 CrossRef .
  196. L. Koudelka, I. Rösslerová, Z. Černošek, P. Mošner, L. Montagne and B. Revel, J. Non-Cryst. Solids, 2014, 401, 124 CrossRef CAS .
  197. L. Koudelka and P. Mosner, Mater. Lett., 2000, 42, 194 CrossRef CAS .
  198. L. Koudelka, P. Mosnera, J. Pospisila, L. Montagne and G. Palavit, J. Solid State Chem., 2005, 178, 1837 CrossRef CAS .
  199. L. Koudelka, J. Subcik, P. Mosner, L. Montagne and L. Delevoye, J. Non-Cryst. Solids, 2007, 353, 1828 CrossRef CAS .
  200. J. Šubcík, L. Koudelka, P. Mošner, L. Montagne, B. Revel and I. Gregora, J. Non-Cryst. Solids, 2009, 355, 970 CrossRef .
  201. P. Mosner, K. Vosejpková, L. Koudelka, L. Montagne and B. Revel, Mater. Chem. Phys., 2010, 124, 732 CrossRef CAS .
  202. K. Vosejpkova, L. Koudelka, Z. Cernosek, P. Mosner, L. Montagne and B. Revel, J. Phys. Chem. Solids, 2012, 73, 324 CrossRef CAS .
  203. S. Petrescu, M. Constantinescu, E. M. Anghel, I. Atkinson, M. Olteanu and M. Zaharescu, J. Non-Cryst. Solids, 2012, 358, 3280 CrossRef CAS .
  204. C. L. Losq, D. R. Neuville, P. Florian, G. S. Henderson and D. Massiot, Geochim. Cosmochim. Acta, 2014, 126, 495 CrossRef .
  205. D. A. McKeown, F. L. Galeener and G. E. Brown Jr, J. Non-Cryst. Solids, 1984, 68, 361 CrossRef CAS .
  206. M. Okuno, N. Zotov, M. Schmucker and H. Schneider, J. Non-Cryst. Solids, 2005, 351, 1032 CrossRef CAS .
  207. P. McMillan, B. Piriou and A. Navrotsky, Geochim. Cosmochim. Acta, 1982, 46, 2021 CrossRef CAS .
  208. P. McMillan and B. Piriou, J. Non-Cryst. Solids, 1983, 55, 221 CrossRef CAS .
  209. D. R. Neuville, L. Cormier and D. Massiot, Geochim. Cosmochim. Acta, 2004, 68, 5071 CrossRef CAS .
  210. D. R. Neuville, L. Cormier and D. Massiot, Chem. Geol., 2006, 229, 173 CrossRef CAS .
  211. D. R. Neuville, L. Cormier, A. M. Flank, V. Briois and D. Massiot, Chem. Geol., 2004, 213, 153 CrossRef CAS .
  212. M. Licheron, V. Montouillout, F. Millot and D. R. Neuville, J. Non-Cryst. Solids, 2011, 357, 2796 CrossRef CAS .
  213. C. Huang and E. C. Behrman, J. Non-Cryst. Solids, 1991, 128, 310 CrossRef CAS .
  214. J. F. Stebbins, E. V. Dubinsky, K. Kanehashi and K. E. Kelsey, Geochim. Cosmochim. Acta, 2008, 72, 910 CrossRef CAS .
  215. V. G. Plotnichenko, V. O. Sokolov, V. V. Koltashev and E. M. Dianov, J. Non-Cryst. Solids, 2002, 306, 209 CrossRef CAS .
  216. A. Trukhin and B. Capoen, J. Non-Cryst. Solids, 2005, 351, 3640 CrossRef CAS .
  217. S. Agathopoulos, D. U. Tulyaganov, J. M. G. Ventura, S. Kannan, A. Saranti, M. A. Karakassides and J. M. F. Ferreira, J. Non-Cryst. Solids, 2006, 352, 322 CrossRef CAS .
  218. H. Aguiar, E. L. Solla, J. Serra, P. González, B. León, F. Malz and C. Jäger, J. Non-Cryst. Solids, 2008, 354, 5004 CrossRef CAS .
  219. H. Aguiar, J. Serra, P. González and B. León, J. Non-Cryst. Solids, 2009, 355, 475 CrossRef CAS .
  220. A. Vulpoi, L. Baia, S. Simon and V. Simon, Mater. Sci. Eng., C, 2012, 32, 178 CrossRef CAS .
  221. Y. Gandhi, M. V. R. C. Rao, C. S. Rao, I. V. Kityk and N. Veeraiah, J. Lumin., 2011, 131, 1443 CrossRef CAS .
  222. K. Cheng, J. Wan and K. Liang, Mater. Lett., 2001, 47, 1 CrossRef CAS .
  223. H. Li, Y. Su, L. Li and D. M. Strachan, J. Non-Cryst. Solids, 2001, 292, 167 CrossRef CAS .
  224. A. R. Kumar, T. G. V. M. Rao, K. Neeraja, M. R. Reddy and N. Veeraiah, Vib. Spectrosc., 2013, 69, 49 CrossRef .
  225. E. Malchukova, B. Boizot, G. Petite and D. Ghaleb, J. Non-Cryst. Solids, 2008, 354, 3592 CrossRef CAS .
  226. E. Malchukova and B. Boizot, Phys. Solid State, 2008, 50, 1687 CrossRef CAS .
  227. E. Malchukova, B. Boizot, G. Petite and D. Ghaleb, Eur. Phys. J.: Appl. Phys., 2009, 45, 10701 CrossRef .
  228. E. Malchukova, B. Boizot, D. Ghaleb and G. Petite, J. Non-Cryst. Solids, 2006, 352, 297 CrossRef CAS .
  229. E. Malchukova, B. Boizot, G. Petite and D. Ghaleb, J. Non-Cryst. Solids, 2007, 353, 2397 CrossRef CAS .
  230. E. Malchukovaa, B. Boizot, D. Ghaleb and G. Petite, Nucl. Instrum. Methods Phys. Res., Sect. A, 2005, 537, 411 CrossRef .
  231. A. jha, B. Richards, G. Jose, T. T. Fernandez, P. Joshi, X. Jiang and J. Lousteau, Prog. Mater. Sci., 2012, 57, 1426 CrossRef CAS .
  232. F. Pietrucci, S. Caravati and M. Bernasconi, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 064203 CrossRef .
  233. A. Santic, A. M. Milankovic, K. Furic, M. R. Linaric, C. S. Ray and D. E. Dayd, Croat. Chem. Acta, 2008, 81, 559 CAS .
  234. J. Heo, D. Lam, G. H. Sigel, Jr., E. A. Mendoza and D. A. Hensley, J. Am. Ceram. Soc., 1992, 75, 277 CrossRef CAS .
  235. M. M. Umair and A. K. Yahya, Mater. Chem. Phys., 2013, 142, 549 CrossRef CAS .
  236. M. Irannejad, G. Jose, A. Jha and P. Steenson, Opt. Commun., 2012, 285, 2646 CrossRef CAS .
  237. A. Jha, P. Joshi and S. Shen, Opt. Express, 2008, 16, 13526 CrossRef CAS PubMed .
  238. E. Stavrou, C. Raptis and K. Syassen, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 174202 CrossRef .
  239. D. Dutta, M. P. F. Graca, M. A. Valente and S. K. Mendiratta, Solid State Ionics, 2013, 230, 66 CrossRef CAS .
  240. M. Ceriotti, F. Pietrucci and M. Bernasconi, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 104304 CrossRef .
  241. S. Manning, H. Ebendorff-Heidepriem and T. M. Monro, Opt. Mater. Express, 2012, 2, 140 CrossRef CAS .
  242. A. Kaur, A. Khanna, C. Pesquera, F. González and V. Sathe, J. Non-Cryst. Solids, 2010, 356, 864 CrossRef CAS .
  243. W. Stambouli, H. Elhouichet and M. Ferid, J. Mol. Struct., 2012, 1028, 39 CrossRef CAS .
  244. I. Yakine, A. Chagraoui, A. Moussaoui and A. Tairi, J. Mater. Environ. Sci., 2012, 3, 776 CAS .
  245. H. Fares, I. Jlassi, H. Elhouichet and M. Férid, J. Non-Cryst. Solids, 2014, 396–397, 1 CrossRef CAS .
  246. A. Kaur, A. Khanna, V. G. Sathe, F. Gonzalez and B. Ortiz, Phase Transitions, 2013, 86, 598 CrossRef CAS .
  247. R. Akagi, K. Handa, N. Ohtori, A. C. Hannon, M. Tatsumisago and N. Umesaki, J. Non-Cryst. Solids, 1999, 256 & 257, 111 Search PubMed .
  248. T. Hayakawa, M. Koduka, M. Nogami, J. R. Duclere, A. P. Mirgorodsky and P. Thomas, Scr. Mater., 2010, 62, 806 CrossRef CAS .
  249. M. H. Bhat, M. Ganguli and K. J. Rao, Curr. Sci., 2004, 86, 676 CAS .
  250. N. Kaur and A. Khanna, J. Non-Cryst. Solids, 2014, 404, 116 CrossRef CAS .
  251. P. Damas, J. Coelho, G. Hungerford and N. S. Hussain, Mater. Res. Bull., 2012, 47, 3489 CrossRef CAS .
  252. M. R. Henderson, B. C. Gibson, H. Ebendorff-Heidepriem, K. Kuan, S. V. Afshar, J. O. Orwa, I. Aharonovich, S. Tomljenovic-Hanic, A. D. Greentree, S. Prawer and T. M. Monro, Adv. Mater., 2011, 23, 2806 CrossRef CAS PubMed .
  253. M. M. Smedskjaer, J. C. Mauro, S. Sen and Y. Yue, Chem. Mater., 2010, 22, 5358 CrossRef CAS .

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.