Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Raman, green and infrared emission cross-sectionsof Er3+ doped TZPPN tellurite glass

Open Access Open Access

Abstract

The spectroscopic properties of Er3‏+ ion in tellurite glass of molar composition 76TeO2∙10ZnO∙9.0PbO∙1.0PbF2∙3.0Na2O∙1.0Er2O3 was investigated experimentally. The three phenomenological intensity parameters Ωk (k = 2, 4, 6) were determined from the absorption spectral intensities using the Judd-Ofelt (J-O) theory. Several radiative properties such as spontaneous transition probabilities, fluorescence branching ratios and radiative life times were determined by using these intensity parameters. The special attention was attributed to the visible emissions that could be obtained by pumping using a blue laser diode. The stimulated emission cross-section and CIE chromaticity coordinates were calculated. The latter were used to evaluate green light emitting by Er:TZPPN glass. Subsequently, the stimulated emission cross-section, around 1.5 μm, was calculated from McCumber theory. Gain cross-section for laser transition 4I13/24I15/2 of Er3+-ions was obtained. In comparison with other Er-doped laser glasses, the calculated parameters show that Er:TZPPN glass satisfies the fundamental spectral condition for laser emission around 1.5 μm. Moreover the Raman gain coefficient of the present glass was obtained from Raman scattering experiments using 532 nm excitation [(532 nm Laser type Diode-pumped, solid state (DPSS)]. The developed glass showed the widest bandwidths of gain cross section from 249 to 1,106 cm−1.

© 2014 Optical Society of America

1. Introduction

Infrared solid state and fiber lasers/optical amplifiers are of great interest for numerous applications, such as telecommunication, Raman laser amplifiers, optical parametric oscillators, vibrancies lasers, chemical sensors and medicine and atmosphere transmission. Particularly, tellurite glasses are attractive in optical fiber lasers, amplifier applications and frequency up-converters [17].Transparent tellurite glasses doped withTm3+ [46] and Er3+ [5,7] -doped have shown a great potential for optical amplifiers in the second and third telecommunications windows (at 1.3 and 1.5 µm, respectively).

Generally, tellurite glasses have a wider transmission range than silica glass. They also have much lower phonon energy, and their glass stability and corrosion resistance is superior to that of fluoride glass. Because the rheological behavior of tellurite glass is Newtonian, its viscosity does not depend on the shear rate. Consequently, the fiber drawing speed is not likely to affect the fiber quality and fiber fabrication will not be a technical challenge [1, 6].

In previously work [8], the results of differential thermal analysis (DTA) indicated that the composition 76TeO2∙10ZnO∙9.0PbO∙1.0PbF2∙3.0Na2O doped with 1% Er2O3 (denoted as Er:TZPPN glass) has a high thermal stability and a low tendency towards crystallization. Especially the thermal stability factor is ΔT = 152 °C (the difference between crystallization and glass transition temperature) [8]. In this work, we investigate the optical transitions for this glass. The Judd-Ofelt intensity parameters of Er:TZZPN glass are calculated from the absorption spectra. The measured absorption, around 1,500nm is analyzed by McCumber theory [11, 12] in order to obtain stimulated emission cross sections and gain coefficient ofI413/2I415/2 transition. We also present a study of photoluminescence, by excitation at 490nm, in Er:TZPPN glass. Moreover the Raman gain coefficient of the present glass is obtained from Raman scattering experiments using 532 nm excitation [(532 nm Laser type Diode-pumped, solid state (DPSS)]

2. Experimental procedure

A glass with the composition76TeO2∙10ZnO∙9.0PbO∙1.0PbF2∙3.0Na2O∙1.0Er2O3 was prepared by mixing specified weights of raw materials. The powder mixture was given in a covered gold crucible and heated in a melting furnace to a temperature of 900 °C for 30 min; the melt was stirred from time to time. The highly viscous melt which was cast at 850 °C on a graphite mould. Subsequently, the sample was transferred to an annealing furnace and kept for 2h at 270°C (below Tg-15K). Then the furnace was switched off and the glass sample was allowed to cool.

The vertical (VV) polarized spontaneous Raman spectra of the prepared glass were acquired using a Thermo Scientific DXR Raman Microscope spectroscopy setup with 532 nm excitation [(532 nm Laser type Diode-pumped, solid state (DPSS)]. An incoming vertically surface of the bulk sample, and V-polarized Raman scattered signal collected in the back scattering geometry with a 100x microscope objective.

3. Results and discussion

3.1 Absorption spectrum and Judd-Ofelt analysis

The Judd-Ofelt theory has mostly been used to evaluate the probability of forced electric dipole transitions of rare-earth ions in various environments as well as in calculating spectroscopic parameters [810,13]. It has been shown that for glasses the Judd-Ofelt parameters are related to local structures in the vicinity of rare-earth ion sites, which is useful information in order to estimate the emission properties of rare-earth-doped glasses [8]. The transitions between states that meet the transition-selective rulesΔS=ΔL=0,ΔJ=0,±1 comprise magnetic dipole (Smd) as well as electric dipole (Sed) transitions, which can be calculated by:

Sedcalc(ψJ,ψ'J')=k=2,4,6Ωk|ψJU(k)ψ'J'|2
Smd(ψJ,ψ'J')=[h4πmc]2|ψJL+2SψJ'|2
whereΩkrepresent the Judd–Ofelt parameters to be determined. We have calculated the reduced matrix elements|ψJU(k)ψ'J'|based on the free-ion wave functions of Er3+ determined by Maalej et al [14] (Table 1).

Tables Icon

Table 1. The reduced matrix elementsof Er3+.

So the measured electric dipole line strengths Sedmeas(ψJ,ψ'J')from the absorption spectrum can be given by [8, 13]:

Sedmeas(ψJ,ψ'J')=14πε0[9n(n2+2)2][3ch8π3e2(2J+1)Nλ¯×2.303JJ'OD(λ)dλnSmd]
where λ¯ is the mean wavelength of the absorption band, n is the refractive index of the host with respect to λ¯, ( = 10 mm) is the thickness of the studied sample, N (=4.420×1020ion/cm3)is the Er3+ ions concentration, e is the electron charge andJJ'OD(λ)dλrepresents the experimental integrated optical density in the wavelength range and can be obtained by calculating the total area under the absorbed band. S, L, and J are spin angular momentum, orbit angular momentum, and total angular momentum of the initial state, respectively.

The linear refractive index for TZPPN:Er was calculated using Wemple relation:

1n2(E)1=EsEdE2EsEd
The Ed and Es, equal to 20.52 eV and 6.59 eV respectively, obtained in previous work [8].

Figure 1 shows the absorption spectrum of Er3+ doped TZPPN glass. Er3+ doped TZPPN glass has numerous absorption bands located at 1530, 975, 803, 654, 545, 523 and 489nm, which correspond to the transitions from I415/2to I413/2, I411/2, I49/2, F49/2, S43/2, H2(21)11/2 and F47/2, respectively.

 figure: Fig. 1

Fig. 1 Optical density for the TZPPN glass doped 1%Er2O3 from I415/2 level.

Download Full Size | PDF

The three intensity parameters Ωk(k=2,4,6) of Er3+ ions in TZPPN glass were calculated by using Eqs. (1-3). The Judd-Ofelt parameters are found to be Ω2=4.28×1020cm2, Ω4=1.68×1020cm2, Ω6=1.38×1020cm2 by using the least squares method.

Such calculated and measure dielectric dipole line strengths Sedcal and Sedmeas values of different transitions are given in Table 2. An estimation of the accuracy of the calculations of Ωt is given by the rms deviation:

δrms=((SedmeasSedcal)2(p3))1/2
Where p is the number of spectral bands analyzed. The value of the rms deviation calculated in our case was 0.20 ×1020cm2, which denotes a good agreement between the calculated and the experimental data and, consequently, a good precision in the determination of the intensity parameters.

Tables Icon

Table 2. Average wavelengths, refractive indexes and electric and magnetic dipole line strengths for Er3+ doped TZPPNglass.

The calculated spontaneous emission probabilities for electricAed and magneticAmddipole transitions, the predicted radiative lifetime τrfor any specific emitting state (which is an important parameter in consideration of the pumping requirement for laser action threshold) and the intermultiplet luminescence branching ratios βwere estimated using the calculated intensity parameters and correcting for the refractive index. The values of all these parameters are listed in Table 3.

Tables Icon

Table 3. Electric and magnetic dipole line strengths (Sedcal)and (Sedmeas), electric dipole transition probabilities (Aed), magnetic dipole transition probabilities (Amd), radiative branching ratios (β) and radiative lifetimes (τr)of the energy levels of Er3+ doped TZPPN glass.

The branching ratios obtained for the transitions from the I413/2, I411/2, I49/2, F49/2, S43/2, and F47/2 levels to the I415/2 ground state are larger than 0.74, which predicts efficient emissions from such levels under suitable excitation conditions.

3.2 Fluorescence properties

Figure 2 shows the fluorescence spectrum of the Er3+ ion which is excited with 490nm to the 4F7/2 multiplet at room temperature. The band with a maximum at 553nm is attributed to the transition from the 4S3/2 level to the 4I15/2 ground level. The next one with a maximum at 666nmrepresents the emission from 4F9/2 level to the 4I15/2 ground level. The band with a maximum at 844 nm is attributed to the transition from 4S3/2 level to the 4I13/2 first excited state of Er3+. The band with a maximum at 532 nm has been attributed to the 2H11/24I15/2 transition.

 figure: Fig. 2

Fig. 2 Fluorescence spectrum of the Er:TZPPN glass excited by 490 nm.

Download Full Size | PDF

For the calculation of the emission cross-section [15] the Füchtbauer–Ladenburg formula is applied:

σem(λ)=βλ5I(λ)8πn2cτRλI(λ)dλ
whereI(λ) is the intensity of the emission spectrum, n is the refractive index of the material,τRandβare the radiative lifetimes and luminescence branching ratios, determined from the J-O theory. The highest value of the emission cross-section,σem, equal to 6.61×1021cm2at 553 nm is obtained for theS43/2 I415/2transition of Er3+.

3.3 CIE Chromaticity Coordinates

The assessment and quantification of color is referred to colorimetry or the ‘science of color’. The CIE (Commission International de l’Eclairage) system is the most common method to describe the compositions of any color in terms of three primariesx¯(λ), y¯(λ)and z¯(λ), which are called color matching functions [16,17]. Artificial “colors,” denoted by X, Y and Z, also called tristimulus values, can be added to produce real spectral colors [16,17]. The degree of simulation required to match the color of given power spectral density (P(λ)) can be expressed as:

X=λx¯(λ)P(λ)dλ
Y=λy¯(λ)P(λ)dλ
Z=λz¯(λ)P(λ)dλ

The chromaticity coordinates, x, y and z are calculated from the tristimulus values according [17] to the equations:

x=XX+Y+Z
y=YX+Y+Z
z=ZX+Y+Z

Generally (x,y) coordinates are used to represent the color. The locus of all monochromatic color coordinates makes the perimeter of CIE1931 chromaticity diagram. All the multi-chromatic wavelengths will lie within the area of the chromaticity diagram.

The CIE coordinate by using suitable software is calculated for Er3+doped tellurite glass upon excitation at 490nm and is found to be (x = 0.308, y = 0.684)which lies within the green region (Fig. 3). Because of the above reason, present glass gives emission in the green region with appreciable intensity for display applications, light emitting diodes and laser action.

 figure: Fig. 3

Fig. 3 The CIE coordinates for the Er:TZPPNglass upon excitation at490 nm

Download Full Size | PDF

3.4 Stimulated emission cross-section and gain coefficient at 1.5μm

According to the measured absorption spectra shown in Fig. 1, the absorption cross-sections of Er3+ ion for theI413/2I415/2transition can be calculated. The relation between the absorption cross-sectionsσabs(λ) and wavelength can be expressed by using the Beer-Lambert equation [13]:

σabs(λ)=ln(I0(λ)/I(λ))N=2.303OD(λ)N
whereI0(λ) is the incident optical intensity,I(λ) is the optical intensity throughout the sample, OD(λ)is the optical density, is the thickness of the sample, and N is the concentration of Er3+.

The emission cross section σem(λ) of Er3+ for I413/2I415/2can be calculated from the absorption cross-section by using the McCumber formula to evaluate the possibility of laser effect [11,12]:

σem(λ)=σabs(λ)ZlZuexp[hckT(1λZL1λ)]
Where Zl and Zu are the partition functions, respectively, for the lower and the upper levels involved in the considered optical transition, T is the temperature (here is the room temperature), k is the Boltzmann constant and λZL is the wavelength for the transition between the lower Stark sublevels of the emitting multiplets and the lower Stark sublevels of the receiving multiplets (zero-phonon line). Generally, the room temperature value of the Zl/Zu ratio is typically in the range of 0.757–1.045 for the most Er-doped crystals [12] and the λZLis approximately between 1,514 and 1,547nm [12].

The true value of Zl/Zu is not known exactly for glasses, but in the high-temperature limit, the ratio of the partition functions of the lower and upper states Zl/Zu simply becomes the degeneracy weighting of the two states [12,15]. In the following calculation, it will be assumed that the ratio Zl/Zu is equal to 16/14 [12,15]whereas the zero-phonon line is assumed to beλZL=1,531nm [7].

Figure 4 shows the calculated absorption and emission cross-sections for the glass. The emission cross-sections are very similar to those calculated for other Er doped glasses [5,7]. The peak of stimulated emission cross-section (σempeak) is about 1.03×1020cm2.This larger value for the emission cross-section is related to the larger value of the line strength of the I413/2I415/2 transition,S(I413/2,I415/2)=0.019Ω2+0.117Ω4+1.432Ω6, and, more specifically, on the large value of the Ω6 parameter found in glass under study. The value of σempeak for the studied glass is much higher than that for silicate, phosphate, germanate [18,19] and other tellurite glasses [7, 2022].

 figure: Fig. 4

Fig. 4 Absorption cross-sections σa(λ)and stimulated emission cross-sectionσe(λ)for the Er:TZPPNglass.

Download Full Size | PDF

The full width at half maximum (FWHM) is also a critical parameter that is used to evaluate the gain bandwidth properties of the optical amplifiers. The full width at half maxima (FWHM) of the emission peak is 46nm for Er3+-doped TZPPN glass.

Due to the large overlap of the absorption and emission spectrum of Er3+ ions at 1.5 µm, reabsorption will occur and cause a change in the fluorescence spectrum. Thus, due to the asymmetric profile of the emission line, it is more reasonable to calculate an effective bandwidth, instead of the FWHM. The effective bandwidth (Δλ) can be expressed as Δλ=σem(λ)dλ/σempeak. The effective bandwidth is 72.5 nm. This value is similar to those of other tellurite glasses [10] and it is very large with respect to those of silicate, phosphate, germanium glasses and boro-tellurite glasses [2224].

In order to understand the band profile of the I413/2I415/2emission of the Er3+ ions and to estimate the Stark splitting for the I413/2 emitting and the I415/2 ground levels in the tellurite glass under study, a Gaussian deconvolution of the 1.5 µm band developed assuming a simplified model of 4 Stark levels system for the first two levels of the Er3+ ions in the tellurite glass.Fig. 5 shows the emission spectra due to the I413/2I415/2transition of Er3+ ions and the deconvoluted Gaussian amplitude peaks obtained from the fitting of the emission spectra of the Er:TZPPN glass (dotted lines). Peak positions and the widths of these subcomponent peaks are labeled as A, B, C and D and tabulated in Table 4. In order to explain 1.5 μm emissions of the Er3+ ions, an equivalent model of four levels system is shown in Fig. 6 [19,2527].

 figure: Fig. 5

Fig. 5 Emission spectra of the Er:TZPPN glass and deconvolution into Gaussian peaks.

Download Full Size | PDF

Tables Icon

Table 4. Peak positions (λ) and the half maximum(W) of the A-D subcomponents.

 figure: Fig. 6

Fig. 6 An equivalent model of four level system for describing 1.5 µm emission of Er3+.

Download Full Size | PDF

Which show the ground I415/2 level splits into two sublevels at around 0cm1and227cm1. The excited I413/2 level also splits into two sublevels (Starks levels) at around 6,530cm1and 6,650cm1as seen in Fig. 6 together with all of the transitions possible between these subcomponents. Thus the energy differences ΔE1=2270=227cm1and ΔE2=6,6506,530=120cm1are the values of the energy range of the Stark splitting of the I415/2 and the I413/2multiplets, respectively. The ground state presents a larger Stark splitting than the emitting level for the tellurite glass under study, in a similar way to what has been found in Er3+ doped phosphate glasses [27], silicate glasses [28] and tellurite glasses [29]. The results also indicate that the bandwidth is strongly dependent on the overall extent of the Stark splitting.

Optical gain coefficient is an important factor for evaluating the performance of a laser media. If the absorption and emission cross sections for the transitions between two laser operating levels are obtained, the optical gain coefficient g(λ) can be calculated from the following formula:

g(λ)=N[P.σem(λ)(1P).σabs(λ)]
whereP is the population inversion between the upper and the lower levels. Figure 7 shows the gain cross-section as a function of the wavelength under different population inversions. It can be seen that the peak gain cross-section increases and the gain band extends to the short-wavelength side as P increases for both transitions. Higher P gives rise to both broader bandwidth and higher peak value of the gain cross-section. In the case of total inversion (P=1) at 1,532 nm, we obtain a gain coefficient equal to4.54cm1 for Er:TZPPN glass. This value is very large than those of other tellurite glasses [30].

 figure: Fig. 7

Fig. 7 The gain coefficient for theI413/2I415/2 transition of the Er:TZPPNglass.

Download Full Size | PDF

3.5 Raman spectra

Raman bands are obtained in present glasses in Fig. 8.These bands are deconvoluted into five symmetrical Gaussian peaks at about 382, 484, 560, 710 and 803 cm−1 in the following are denoted as peak A, B, C, D and E, respectively.

 figure: Fig. 8

Fig. 8 Deconvolution of the Raman spectra of the Er:TZPPN glass. Experimental spectra: symbol; fitted curves: dashed lines.

Download Full Size | PDF

The assignment of the deconvoluted peaks is performed based on the literature on the tellurite glasses [3134]. The phase structure α-TeO2 is similar to those tellurite glasses estimated by Sekiya et al. [35]; which consists of three dimensional network of TeO4trigonalbipyramid (tbp) connected with asymmetric Te-eqOax-Te bond. The Raman bands in the low frequency region at 123 and 152 cm−1 correspond to the intra-molecular asymmetric motion of the Te-O bonds. These bands are not appeared in the present glass. When added of metal oxide to TeO2 leads to a break of the axial Te-O bonds because of the strong polarizabilty of the tellurium lone pair electrons and the formation of non-bridging Te-O bonds. Moreover the TeO4 units are transformed into TeO3+1 and TeO3 polyhedra. In the present glass,the band at 382 cm−1 (labeled as A) can be contributed to the axial bending vibration mode (Oax- Te-Oax). A strong band labeled B at 482 cm−1can be attributed stretching vibrations of Te-O- and Te = O bonds containing non-bridging oxygen in TeO3tps and TeO3+1 polyhedral. Furthermore this band is not observed in pure TeO2 glass, when the addition of network modifiers results in a cleavage of Te-O-Te linkages of the initially polymerized structure and in the transformation of TeO4tbp into TeO3+1polyhedrawithnon bridging oxygen (NBO) or into TeO3tp with even more NBO atoms [36]. The weak band centered around 560 cm−1(labeled C) and 803 cm−1 (labeled E)observed in the present glass are attribute able to the (Teeq-O)s and the (Teeq-O)as vibrational modes of TeO3+1polyhedra and/or TeO3trigonal pyramids. Furthermore, a strong and broad band with the peak at 710 cm−1 (peak D)is observed. We suggest that this band is due to the symmetric stretching vibrations of Te-O-Te in which both the Te-O bonds have lengths of about 2.0 Å. Hoppe et al [37] determined Te-O and Zn-O coordination in zinc tellurite glasses by X-ray and neutron scattering measurements. They concluded from radial distribution function analyses of neutron scattering data that increasing the ZnO concentration to 10 mol% ZnO in the TZPPN glass, leads to a decrease in the mean Te-O coordination number due to a conversion of TeO4 into TeO3+1 and TeO3 structural units.

The intensities of various peaks in the Raman spectra of the prepared Er2O3 doped glass are higher than of other tellurite glasses reported in the literature [3134]. This indicates the transformation of TeO4 tbp into TeO3+1/TeO3 tp in the studied glasses seen from the high intensity of the band at 710 cm−1.

3.6 Stimulated Raman gain coefficient

We can calculate the Raman gain coefficient, G, of the present glass using the equation [38]:

G=σTλS3c2hn2[N(w,T)+1]

Where σT is the corrected scattering cross section at temperature T(K), λs is the Stokes wavelength, c is velocity and n is the refractive index at the excitation wavelength. N(w,T) is the Bose-Einstein factor:

N(w,T)=1exp(wKBT)1
the gain coefficient of silica glass at 440 cm−1for a pump wavelength of 532 nm is 1.861013m/W [39].The shape of the Raman gain spectrum of76TeO2∙10ZnO∙9.0PbO∙1.0PbF2∙3.0Na2O∙1.0Er2O3glass is shown in Fig. 9.The Raman gain peak for the prepared glass at about 705 cm−1 is equal to9.3×1011m/W.This value was ~500 times larger compared to that silica glass at 532 nm. For selecting a material for Raman amplifiers, the gain bandwidth is an additional important parameter which can be obtained from the wave number dependency of the Raman gain coefficient. The FWHM bandwidth for the band centered at 705 cm−1 is equal to 161 cm−1.

 figure: Fig. 9

Fig. 9 Raman gain spectra for the Er:TZPPN glass calculated from absolute spontaneous Raman cross-section.

Download Full Size | PDF

Figure 10 shows the Raman cross section spectra of the prepared glass and its deconvolution within the wave number range from 249 to 1106 cm−1. The shows a decrease in the cross section and Raman gain coefficient at around 434 and 880 cm−1 related to vibration of Te-O-Te bridges and TeO4polyhedra. Otherwise, the decrease of the number of TeO4 units leads to the formation TeO3+1 or TeO3 structural units and hence to depolymerization of the tellurite glass matrix and results in increasing Raman gain coefficient and cross sections at 705 cm−1. In conclusion, the introduction of Er3+ ions has a major effect by the depolymerization of the tellurite network leads to local polarizability/hyperpolarizability and highest Raman gain compared with other oxide tellurite glass without doping by Er3+ [4042]. Therefore, the prepared material could be a promising candidate for ultrabroadband Raman amplifier.

 figure: Fig. 10

Fig. 10 Deconvolution of the corrected Raman cross section spectra of the Er:TZPPN glass.

Download Full Size | PDF

4. Conclusion

The optical transitions of Er3+ in tellurite glass were investigated. The CIE coordinate was calculated for the glass upon excitation at 490 nm and was found about (x = 0.3080,y = 0.6840) that indicated the purity of the emission spectra for green band. Using the Füchtbauer-Ladenburg formula, the highest value of the emission cross-section in visible region, equal to 6.61×1021cm2at 553nm, was obtained for theS43/2 I415/2transition of Er3+.Using the Judd-Ofelt parameters of Er3+in TZPPN glass, the spontaneous radiative lifetime (τr) of I413/2I415/2was calculated to be2.723ms. Stimulated emission cross section in the 1.5 µm region was obtained by using McCumber theory and the optical gain coefficient to the population inversion of the I413/2 level was analyzed. We obtain a gain coefficient of4.54cm1, an effective bandwidth of 72.5nm for Er:TZPPN glass. Therefore, the spectroscopy investigations suggest that the TZPPN glass doped with 1mol% Er3+ions might be a promising material for broadband amplification in the third telecommunications window as well as to generate green light in color display devices.

In addition, the Raman gain coefficient for a pump wavelength of 532 nm at about 705cm1is 9.3×1011m/W.The FWHM bandwidth for the band centered at 705cm1is equal to161cm1. Thus, the prepared tellurite glass could be a candidate material to realize highly efficient ultra broadband fiber Raman amplifier with higher Raman gain.

Acknowledgments

This research was supported by a Grant of King Abdulaziz City for Science and Technology (Code Number: 10-ADV1160-07) from King of Saudi Arabia.

References

1. J. S. Wang, E. M. Vogel, and E. Snitzer, “Tellurite glass: a new candidate for fiber devices,” Opt. Mater. 3(3), 187–203 (1994). [CrossRef]  

2. N. Jaba, A. Kanoun, H. Mejri, A. Selmi, S. Alaya, and H. Maaref, “Infrared to visible up-conversion study for erbium-doped zinc tellurite glasses,” J. Phys. Condens. Matter 12(20), 4523–4534 (2000). [CrossRef]  

3. E. Yousef, M. Hotzel, and C. Rüssel, “Effect of ZnO and Bi 2O 3 addition on linear and non-linear optical properties of tellurite glasses,” J. Non-Cryst. Solids 353(4), 333–338 (2007). [CrossRef]  

4. D. Zhou, R. Wang, Z. Yang, Z. Song, Z. Yin, and J. Qiu, “Spectroscopic properties of Tm3+ doped TeO2-R 2O-La2O3 glasses for 1.47 μm optical amplifiers,” J. Non-Cryst. Solids 357(11-13), 2409–2412 (2011). [CrossRef]  

5. R. Balda, J. Fernández, S. García-Revilla, and J. M. Fernández Navarro, “Spectroscopy and concentration quenching of the infrared emissions in Tm3+-doped TeO2-TiO2-Nb2O5 glass,” Opt. Express 15(11), 6750–6761 (2007). [CrossRef]   [PubMed]  

6. E. R. Taylor, L. N. Ng, and N. P. Sessions, “Spectroscopy of Tm3+-doped tellurite glasses for 1,470 nm fiber amplifier,” J. Appl. Phys. 92, 112–117 (2002).

7. N. Jaba, H. BenMansour, A. Kanoun, A. Brenier, and B. Champagnon, “Spectral broadening and luminescence quenching of 1.53 μm emission in Er3+-doped zinc tellurite glass,” J. Lumin. 129(3), 270–276 (2009). [CrossRef]  

8. E. S. Yousef, K. Damak, R. Maalej, and C. Rüssel, “Thermal stability and UV–Vis-NIR spectroscopy of a new erbium-doped fluorotellurite glass,” Philos. Mag. 92(7), 899–911 (2012). [CrossRef]  

9. B. R. Judd, “Optical absorption intensities of rare-earth ions,” Phys. Rev. 127(3), 750–761 (1962). [CrossRef]  

10. G. S. Ofelt, “Intensities of crystal spectra of rare-earth ions,” J. Chem. Phys. 37(3), 511–520 (1962). [CrossRef]  

11. D. E. McCumber, “Einstein relations connecting broadband emission and absorption spectra,” Phys. Rev. A 136(4), 954–957 (1964). [CrossRef]  

12. P. A. Stephen, A. Payne, L. L. Chase, L. K. Smith, W. L. Kway, and W. F. Krupke, “Infrared cross-section measurements for crystals doped with Er3+, Tm3+, and Ho3+,” IEEE J. Quantum Electron. 28, 2619–2630 (1992).

13. K. Damak, R. Maalej, E. S. Yousef, A. H. Qusti, and C. Rüssel, “Thermal and spectroscopic properties of Tm3+ doped TZPPN transparent glass laser material,” J. Non-Cryst. Solids 358(22), 2974–2980 (2012). [CrossRef]  

14. R. Maalej, A. Chehaidar, and M. Kamoun, “Crystal-field analysis of Er3+ emission spectrum in epitaxial Ca(1-x)ErxF2+x thin films,” Phys. Status Solidi (B) Basin Res. 221, 657–666 (2000).

15. B. Zhou, L. Tao, C. Y.-Y. Chan, Y. H. Tsang, and W. Jin, “Intense near-infrared emission of 1.23 µm in erbium-doped low-phonon-energy fluorotellurite glass,” Spectrochim. Acta [A] 111, 49–53 (2013). [CrossRef]  

16. A. C. Harris and I. L. Weatherall, “Objective evaluation of colour variation in the sand-burrowing beetle Chaerodestrachyscelides White (Coleoptera: Tenebrionidae) by instrumental determination of CIELAB values,” J. R. Soc. N. Z. 20(3), 253–259 (1990). [CrossRef]  

17. H. S. Fairman, M. H. Brill, and H. Hemmendinger, “How the CIE 1931 color-matching functions were derived from wright-guild data,” Color Res. Appl. 22(1), 11–23 (1997). [CrossRef]  

18. R. R. Xu, Y. Tian, M. Wang, L. L. Hu, and J. J. Zhang, “Spectroscopic properties of 1.8 μm emission of thulium ions in germanate glass,” Appl. Phys. B 102(1), 109–116 (2011). [CrossRef]  

19. Z. Ling, Z. Ya-Xun, D. Shi-Xun, X. Tie-Feng, N. Qiu-Hua, and S. Xiang, “Effect of Ga2O3 on the spectroscopic properties of erbium-doped boro-bismuth glasses,” Spectrochim. Acta [A] 68(3), 548–553 (2007). [CrossRef]  

20. G. Bilir, G. Ozen, D. Tatar, and M. L. Öveçoğlu, “Judd–Ofelt analysis and near infrared emission properties of the Er3+ ions in tellurite glasses containing WO3 and CdO,” Opt. Commun. 284(3), 863–868 (2011). [CrossRef]  

21. U. R. Rodríguez-Mendoza, E. A. Lalla, J. M. Caceres, F. Rivera-Lopez, S. F. Leon-Luís, and V. Lavín, “Optical characterization, 1.5μm emission and IR-to-visible energy upconversion in Er3+-doped fluorotellurite glasses,” J. Lumin. 131(6), 1239–1248 (2011). [CrossRef]  

22. R. Rolli, M. Montagna, S. Chaussedent, A. Monteil, V. K. Tikhomirov, and M. Ferrari, “Erbium-doped tellurite glasses with high quantum efficiency and broadband stimulated emission cross section at 1.5 μm,” Opt. Mater. 21(4), 743–748 (2003). [CrossRef]  

23. S. Jianga, B.-C. Hwang, T. Luo, K. Seneschal, F. Smektala, S. Honkanen, J. Lucas, and N. Peyghambarian, “Net gain of 15.5 dB from a 5.1 cm long Er3+-doped phosphate glass fiber,” Optical Fiber Comm. 4, 181–183 (2000).

24. T. Xu, X. Shen, Q. Nie, and Y. Gao, “Spectral properties and thermal stability of Er3+/Yb3+codoped tungsten–tellurite glasses,” Opt. Mater. 28(3), 241–245 (2006). [CrossRef]  

25. W. Deng, J. Zhang, J. Sun, Y. Luo, J. Lin, X. Wang, and W. Xu, “Analysis of spectral components in the 1.5 μm emission band of Er3+ doped borosilicate glass,” J. Non-Cryst. Solids 336(1), 44–48 (2004). [CrossRef]  

26. G. Bilir and G. Ozen, “Optical absorption and emission properties of Nd3+ in TeO2–WO3 and TeO2–WO3–CdO glasses,” Physica B 406(21), 4007–4013 (2011). [CrossRef]  

27. F. Rivera-López, P. Babu, L. Jyothi, U. R. Rodríguez-Mendoza, I. R. Martín, C. K. Jayasankar, and V. Lavín, “Er3+–Yb3+codoped phosphate glasses used for an efficient 1.5 μm broadband gain medium,” Opt. Mater. 34(8), 1235–1240 (2012). [CrossRef]  

28. E. Desurvire and J. R. Simpson, “Evaluation of 4I15/2 and 4I13/2 Stark-level energies in erbium-doped aluminosilicate glass fibers,” Opt. Lett. 15(10), 547–549 (1990). [CrossRef]   [PubMed]  

29. X. Li and W. Zhang, “Temperature-dependent fluorescence characteristics of an ytterbium-sensitized erbium-doped tellurite glass,” Physica B 403(18), 3286–3288 (2008). [CrossRef]  

30. X. Shen, Q. Nie, and X. Wang, “Effect of Bi2O3 on spectroscopic properties of Er3+ -doped tellurite bismuth glasses for broadband optical amplifiers,” IEEE International Conference on Industrial Informatics, 1233 (2006). [CrossRef]  

31. I. Shaltout and Y. Badr, “Manifestation of Nd ions on the structure, Raman and IR spectra of (TeO2 -MoO-Nd2O3) glasses,” J. Mater. Sci. 40(13), 3367–3373 (2005). [CrossRef]  

32. T. Hayakawa, M. Koduka, M. Nogami, J. R. Duclere, A. P. Mirgorodsky, and P. Thomas, “Metal oxide doping effects on Raman spectra and third-order nonlinear susceptibilities of thallium–tellurite glasses,” Scr. Mater. 62(10), 806–809 (2010). [CrossRef]  

33. A. Mogusmilankovic, A. Šantić, S. T. Reis, K. Furic, and D. E. Day, “Studies of lead–iron phosphate glasses by Raman, Mössbauer and impedance spectroscopy,” J. Non-Cryst. Solids 351(40-42), 3246–3258 (2005). [CrossRef]  

34. V. K. Malinovesky and A. P. Sokolov, “The nature of boson peak in Raman scattering in glasses,” Solid State Commun. 57(9), 757–761 (1986).

35. T. Sekiya, N. Mochida, and A. Ohtsuka, “Raman spectra of MO- TeO2 (M = Mg, Sr, Ba and Zn) glasses,” J. Non-Cryst. Solids 168(1-2), 106–114 (1994). [CrossRef]  

36. G. Guery, A. Farges, T. Cardinal, M. Dussauze, F. Adamietz, V. Rodriguez, J. D. Musgraves, K. Richardson, and P. Thomas, “Impact of tellurite-based glass structure on Raman gain,” Chem. Phys. Lett. 554, 123127 (2012).

37. U. Hoppe, E. Yousef, C. Rüssel, J. Neuefeind, and A. C. Hannon, “Structure of zinc and niobium tellurite glasses by neutron and x-ray diffraction,” J. Phys. Condens. Matter 16(9), 1645–1663 (2004). [CrossRef]  

38. K. Damak, E. S. Yousef, A. S. Al-Shihri, H. J. Seo, C. Rüssel, and R. Maâlej, “Quantifying Raman and emission gain coefficients of Ho3+ doped TeO2·ZnO·PbO·PbF2·Na2O (TZPPN) tellurite glass,” Solid State Sci. 28, 74–80 (2014). [CrossRef]  

39. R. H. Stolen, C. Lee, and R. K. Jain, “Development of simulated Raman spectrum in single mode silica fibers,” J. Opt. Soc. Am. B 1(4), 652–662 (1984). [CrossRef]  

40. C. Rivero, K. Richardson, R. Stegeman, G. Stegeman, T. Cardinal, E. Fargin, M. Couzi, and V. Rodriguez, “Quantifying Raman gain coefficients in tellurite glasses,” J. Non-Cryst. Solids 345-346(346), 396–401 (2004). [CrossRef]  

41. R. Jose and Y. Ohishi, “Higher Raman scattering cross-sections, bandwidths, and nonlinear indices in the TeO2- ZnO- Nb2O5- Mo2O3 quaternary glass system,” Appl. Phys. Lett. 89, 121122 (2006).

42. R. Jose, G. Qin, Y. Arai, and Y. Ohishi, “Tailoring of Raman gain bandwidth of tellurite glasses for designing gain-flattened fiber Raman amplifiers,” J. Opt. Soc. Am. B 25(3), 373–382 (2008). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1
Fig. 1 Optical density for the TZPPN glass doped 1%Er2O3 from I 4 15/2 level.
Fig. 2
Fig. 2 Fluorescence spectrum of the Er:TZPPN glass excited by 490 nm.
Fig. 3
Fig. 3 The CIE coordinates for the Er:TZPPNglass upon excitation at490 nm
Fig. 4
Fig. 4 Absorption cross-sections σ a ( λ ) and stimulated emission cross-section σ e ( λ ) for the Er:TZPPNglass.
Fig. 5
Fig. 5 Emission spectra of the Er:TZPPN glass and deconvolution into Gaussian peaks.
Fig. 6
Fig. 6 An equivalent model of four level system for describing 1.5 µm emission of Er3+.
Fig. 7
Fig. 7 The gain coefficient for the I 4 13 / 2 I 4 15 / 2 transition of the Er:TZPPNglass.
Fig. 8
Fig. 8 Deconvolution of the Raman spectra of the Er:TZPPN glass. Experimental spectra: symbol; fitted curves: dashed lines.
Fig. 9
Fig. 9 Raman gain spectra for the Er:TZPPN glass calculated from absolute spontaneous Raman cross-section.
Fig. 10
Fig. 10 Deconvolution of the corrected Raman cross section spectra of the Er:TZPPN glass.

Tables (4)

Tables Icon

Table 1 The reduced matrix elementsof Er3+.

Tables Icon

Table 2 Average wavelengths, refractive indexes and electric and magnetic dipole line strengths for Er3+ doped TZPPNglass.

Tables Icon

Table 3 Electric and magnetic dipole line strengths ( S ed cal ) and ( S ed meas ) , electric dipole transition probabilities ( A ed ) , magnetic dipole transition probabilities ( A md ) , radiative branching ratios ( β ) and radiative lifetimes ( τ r ) of the energy levels of Er3+ doped TZPPN glass.

Tables Icon

Table 4 Peak positions (λ) and the half maximum(W) of the A-D subcomponents.

Equations (17)

Equations on this page are rendered with MathJax. Learn more.

S e d c a l c ( ψ J , ψ ' J ' ) = k = 2 , 4 , 6 Ω k | ψ J U ( k ) ψ ' J ' | 2
S m d ( ψ J , ψ ' J ' ) = [ h 4 π m c ] 2 | ψ J L + 2 S ψ J ' | 2
S ed meas ( ψJ,ψ'J' )= 1 4π ε 0 [ 9n ( n 2 +2 ) 2 ][ 3ch 8 π 3 e 2 ( 2J+1 ) N λ ¯ × 2.303 JJ' OD( λ )dλ n S md ]
1 n 2 (E)1 = E s E d E 2 E s E d
δ rms = ( ( S ed meas S ed cal ) 2 ( p3 ) ) 1/2
σ em ( λ )= β λ 5 I( λ ) 8π n 2 c τ R λI( λ )dλ
X= λ x ¯ ( λ )P( λ )dλ
Y= λ y ¯ ( λ )P( λ )dλ
Z= λ z ¯ ( λ )P( λ )dλ
x= X X+Y+Z
y= Y X+Y+Z
z= Z X+Y+Z
σ abs ( λ )= ln( I 0 (λ)/I(λ) ) N = 2.303OD( λ ) N
σ em ( λ )= σ abs ( λ ) Z l Z u exp[ hc kT ( 1 λ ZL 1 λ ) ]
g ( λ ) = N [ P . σ e m ( λ ) ( 1 P ) . σ a b s ( λ ) ]
G= σ T λ S 3 c 2 h n 2 [ N( w,T )+1 ]
N ( w , T ) = 1 exp ( w K B T ) 1
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.