Next Article in Journal
Minkowski Island and Crossbar Fractal Microstrip Antennas for Broadband Applications
Next Article in Special Issue
Nanocomposite Antimony-Germanate-Borate Glass Fibers Doped with Eu3+ Ions with Self-Assembling Silver Nanoparticles for Photonic Applications
Previous Article in Journal
Real Time Quantum Dynamics of Spontaneous Translational Symmetry Breakage in the Early Stage of Photo-Induced Structural Phase Transitions
Previous Article in Special Issue
Effect of ZnO Addition and of Alpha Particle Irradiation on Various Properties of Er3+, Yb3+ Doped Phosphate Glasses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Spectroscopic Investigation of Tm3+:Tellurite Glasses for 2-μm Lasing Applications

1
Laser Research Laboratory, Departments of Physics and Electrical-Electronics Engineering, Koç University Rumelifeneri, Sariyer, Istanbul 34450, Turkey
2
Center for Free-Electron Laser Science, Deutsches Elektronen Synchrotron (DESY), Notkestrasse 85, 22607 Hamburg, Germany
3
Physics Department, University of Hamburg, Luruper Chaussee 149, 22761 Hamburg, Germany
4
Teknofil. Inc., Zekeriyakoy, Istanbul 34450, Turkey
5
Dipartimento di Biotecnologie, University of Verona and INSTM UdR Verona, Ca’Vignal, Strada Le Grazie 15, 37134 Verona, Italy
6
Koç University Surface Science and Technology Center (KUYTAM), Koç University Rumelifeneri, Sariyer, Istanbul 34450, Turkey
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2018, 8(3), 333; https://doi.org/10.3390/app8030333
Submission received: 10 January 2018 / Revised: 13 February 2018 / Accepted: 21 February 2018 / Published: 27 February 2018
(This article belongs to the Special Issue Rare-Earth Doping for Optical Applications)

Abstract

:
We performed a comparative spectroscopic analysis on three novel Tm3+:tellurite-based glasses with the following compositions Tm2O3:TeO2-ZnO (TeZnTm), Tm2O3:TeO2-Nb2O5 (TeNbTm), and Tm3+:TeO2-K2O-Nb2O5 (TeNbKTm), primarily for 2-μm laser applications. Tellurite glasses were prepared at different doping concentrations in order to investigate the effect of Tm3+ ion concentration as well as host composition on the stimulated emission cross sections and the luminescence quantum efficiencies. By performing Judd–Ofelt analysis, we determined the average radiative lifetimes of the 3H4 level to be 2.55 ± 0.07 ms, 2.76 ± 0.03 ms and 2.57 ± 0.20 ms for the TeZnTm, TeNbTm and TeNbKTm samples, respectively. We clearly observed the effect of the cross-relaxation, which becomes significant at higher Tm2O3 concentrations, leading to the quenching of 1460-nm emission and enhancement of 1860-nm emission. Furthermore, with increasing Tm2O3 concentrations, we observed a decrease in the fluorescence lifetimes as a result of the onset of non-radiative decay. For the 3H4 level, the highest obtained quantum efficiency was 32% for the samples with the lowest Tm2O3 ion concentration. For the 1860-nm emission band, the average emission cross section was determined to measure around 6.33 ± 0.34 × 10−21 cm2, revealing the potential of thulium-doped tellurite gain media for 2-μm laser applications in bulk and fiber configurations.

Graphical Abstract

1. Introduction

Thulium-doped systems are drawing a great deal of interest for numerous laser applications in the near and mid-infrared regions of the spectrum, since they provide broad emission bands covering the empty region (1400–2700 nm) between neodymium and erbium systems. In this respect, transitions originating from the 3F4 (3F43H4 and 3F43H5 transitions corresponding to 1.5-μm and 2.5-μm emissions) and 3H4 (3H43H6 transition corresponding to 1.9-μm emissions) energy states are worth studying. For this purpose, we choose to employ a 800-nm excitation scheme, which leads two broad emission bands centered near 1.9 μm and 1.5 μm. Employing this pumping scheme is further beneficial because there are many commercial pump sources available at these wavelengths and this pumping scheme provides an additional advantage, namely a “two-for-one” process which favors 1.9-μm emission at higher active ion concentrations.
The broad emission band around 1.9 μm can be utilized to build new mid-infrared laser sources suitable for spectroscopic, chemical, and atmospheric sensing applications [1]. In addition, strong water absorption occurs at these wavelengths, making thulium-doped gain media very attractive in the development of lasers for biomedical applications such as tissue welding and ablation [2,3]. Therefore, investigation of novel laser materials, especially thulium-doped hosts which fluoresce around 2 μm, deserves a great deal of attention, considering the ever-increasing demand and need for new laser sources operating in the near- and mid-IR regions of the spectrum.
So far, detailed studies have been conducted on germinate, silicate, and fluoride glass hosts together with various crystal gain media doped with rare earths for 2-μm laser applications [4,5,6]. Among other laser gain media, tellurite glasses (with network former TeO2) offer several benefits, such as high chemical stability, easy and low-cost production due to their amorphous nature, and a wide transparency range from 0.355 to 5 µm [3,7,8]. Tellurite glasses are known to be moisture-resistant, thermally and mechanically stable materials, making them attractive for fiber laser applications as well [9,10]. Furthermore, tellurite glass hosts have relatively low phonon energies (700–750 cm−1). Hence, non-radiative decay is relatively low which leads to relatively high luminescence quantum efficiencies.
To date, most of the effort has been spent on Nd3+- and Er3+-doped tellurite glass systems [11,12,13,14,15,16] and there are relatively fewer studies conducted on the Tm3+-doped tellurite glasses [17,18,19,20,21,22,23,24,25]. Recently, lasing and mode-locked operation from bulk thulium-doped tellurite glasses were reported [20,26,27,28]. Since Tm3+-doped tellurite glasses offer a great potential for the development of 2-μm laser sources in bulk as well as fiber configurations, there is interest in further exploring the spectroscopic properties of the tellurite glass hosts doped with thulium.
In this study, we performed a comparative spectroscopic analysis of three different tellurite glasses with various Tm3+ concentrations in order to investigate the dependence of the emission cross section and quantum efficiency, which are among the most critical parameters for laser applications. In particular, we studied tellurite glasses with the compositions Tm2O3:TeO2-ZnO, Tm2O3:TeO2-Nb2O5, and Tm3+:TeO2-K2O-Nb2O5. We investigated the influence of doping concentration as well as the host composition, both of which affect the cross relaxation rate and the luminescence quantum efficiency. The analysis was conducted for the two emission bands with peaks at 1460 nm (3F4 level) and 1860 nm (3H4 level), and Judd–Ofelt theory was used in the analysis of the experimental data.

2. Experimental Procedure and Analysis

2.1. Experiment

Tellurite-based glass samples doped with Tm2O3 were prepared by using the melt quenching technique which was described in our previous study [29]. Three kinds of tellurite-based glass samples were prepared with the glass network modifiers niobium oxide, zinc oxide, and a mixture of potassium and niobium oxide. The Tm2O3 concentration was varied between 0.125% and 1%. Two tellurite-based glass samples were prepared by using niobium oxide as a glass modifier: (x)Tm2O3-(95)TeO2-(5-x)Nb2O5, (TeNbTm) where x = 1.0 and 0.25 (1.0 and 0.25 mol %); four samples were prepared at different concentrations using zinc oxide as a glass modifier: (x)Tm2O3-(80)TeO2-(20-x)ZnO (TeZnTm), where x = 1.0, 0.5, 0.25 and 0.125 (1.0, 0.5, 0.25 and 0.125 mol %); and four thulium-doped tellurite glasses with potassium and niobium oxide were prepared at different Tm2O3 concentrations (x)Tm2O3-(70)TeO2-(15)K2O-(15)Nb2O5 (TeNbKTm) where x = 1.0, 0.5, 0.25 and 0.125 (1.0, 0.5, 0.25 and 0.125 mol %).
The absorption spectra of the glass samples were recorded by using a commercial spectrophotometer. In order to measure the emission spectrum, a home-made, 60-ns pulsed, tunable Ti:sapphire laser was employed as the excitation source. The emission spectra and lifetime measurements were conducted as described in detail in [29]. All measurements were carried out at room temperature.

2.2. Judd–Ofelt Analysis

In order to study the effect of the Tm3+ ion concentration on the radiative lifetimes for the 3F4 and 3H6 levels, Judd–Ofelt (J-O) theory was employed. According to the theory, the effect of the electric-dipole transition from the ground state SLJ to an exited state S’L’J on the integrated absorption coefficient ( Σ μ ) c a l c of the rare earth ions can be calculated by using the following equation [30,31],
( Σ μ ) c a l c = 8 π 3 e 2 3 c h ( n 2 + 2 ) 2 9 n λ ¯ ( 2 J + 1 ) N o x t = 2 , 4 , 6 Ω t | S L J U ( t ) S L J | 2
Here, No is the ion concentration in the gain medium, e is the electron charge, c is the speed of light, h is Planck’s constant, n is the refractive index, λ ¯ is the average wavelength of the absorption band, J is the total angular momentum quantum number, the Ωt’s are the J-O intensity parameters, SLJ and S’L’J’ are the ground state and excited state of the dipole transition, respectively, and U(t) is the doubly reduced matrix element of the unit tensor operator. The Ωt’s can be estimated by using the measured integrated absorption coefficient ( Σ μ ) exp and the matrix elements in [32]. ( Σ μ ) exp can be obtained by calculating the integral under the absorption spectrum for each transition Then, Ωt’s can be determined by fitting the experimental integrated absorption coefficient to theoretically calculated value. Then, the spontaneous emission probability (A(J,J’)) of the dipole transition from SLJ to S’L’J’ can be calculated by using the equation,
A ( J , J ) = 64 π 4 e 2 3 h ( 2 J + 1 ) n ( n 2 + 2 ) 2 ν ¯ 3 9 x t = 2 , 4 , 6 Ω t | S L J U ( t ) S L J | 2
In Equation (2), v   ¯ is the average wave number of the corresponding dipole transition. By using the calculated A(J,J’) from the Jth excited state to all possible J’ states, the radiative lifetime τR for the Jth excited state can be calculated from
τ R = 1 j A ( J , J )

3. Results and Discussion

3.1. Absorption Spectroscopy

Figure 1a,b shows the energy level diagram and the absorption spectra of two Tm2O3:TeO2-Nb2O5 glass samples with Tm2O3 concentrations of 1.0 and 0.25 mol % in the range of 450–2050 nm. The other glass samples with different glass modifiers exhibit similar absorption characteristics but different absorption strengths, which will be discussed later. Table 1 summarizes the Ωt’s intensity parameters determined for three different glass hosts with 1 mol % ion concentration by employing J-O theory described above. In the radiative lifetime calculations for the decay from 3F4 and 3H4 levels to the ground state, we only used J-O intensity parameters derived from the absorption spectrum of the 1 mol % doped samples, where the signal to noise ratio is the highest. We also used the fact that the radiative lifetime is independent of active ion concentration. The average radiative lifetimes for the niobium containing glasses were determined as 2.76 ms and 0.37 ms for the sample TeNbTm, and 2.57 ms and 0.35 ms for the sample TeNbKTm. For the glasses containing zinc as a modifier (TeZnTm), we obtained average radiative lifetimes of 2.55 ms and 0.37 ms for the 3H4 and 3F4 levels, respectively.

3.2. Emission Spectroscopy and Analysis

Figure 2a shows the fluorescence decay signal of the 3H4 level and 3F4 level, corresponding to the 1860-nm and 1460-nm transitions of the two TeO2-K2O-Nb2O5 samples with Tm2O3 concentrations of 1.0 and 0.125 mol %. The fluorescence signal has a sharp peak following an exponential decay. As can be seen from Figure 2b, the curves corresponding to the 1860-nm transition have a smooth peak in comparison with the other transition. That is mainly due to increase in the population in 3H4 level resulting from the decay from the 3F4 to 3H4 level as illustrated in the energy level diagram in Figure 1a. Similar behavior was observed for the other glass samples with different glass modifiers. In the case of 0.125 mol % doping concentration for the 1860-nm transition (Figure 2b), the fluorescence signal exhibits an initial rise followed by an exponential decay. This can be attributed to the slow decay rate from the 3F4 to 3H4 level, filling the population in the 3H4 level. In the case of the sample with 1.0 mol % doping concentration, this effect is negligible due to the faster decay rate from 3F4 to 3H4. The fluorescence lifetimes (τF) were obtained by fitting a single exponential decay curve to the tail of the measured signal, where the cross-relaxation effect is negligible. From the calculated radiative lifetimes and experimentally measured fluorescence lifetimes, the luminescence quantum efficiency (η) can then be determined by using η = τ F / τ R .
Table 2 summarizes the obtained fluorescence lifetimes and quantum efficiencies of the glass samples for the two transitions peaking at 1860 nm and 1460 nm. With increasing Tm2O3 concentration from 0.125 to 1.0 mol %, we observed a decrease in the fluorescence lifetime due to the increasing role of nonradiative relaxation mechanisms. The Te2O3-Nb2O5 glass with 0.25 mol % Tm2O3 doping had the longest fluorescence lifetime (893 μs) for the 1860-nm emission band among the glass samples investigated in this paper. The highest quantum efficiency for the sample TeZnTm was obtained as 32% for the 1860-nm emission band and 89% for the 1460-nm emission band when the ion concentration was 0.125 mol %. We obtained the highest quantum yields of 32% and 76% for the 1860 and 1460 nm emission bands, respectively, in the sample TeNbTm, when the ion concentration was 0.25 mol %.
Figure 3a shows the emission spectra of the two transitions of Tm3+ ion originating from 3F4 and 3H4 levels to the ground level for the TeZnTm glass host with Tm2O3 concentrations of 1.0 and 0.25 mol %. The emission spectra clearly show the two emission bands centered around 1860 nm and 1460 nm and further verify the role of cross relaxation, which is quite common for the Tm3+-doped systems [33,34]. Cross relaxation is a non-radiative energy transfer process, which becomes significant at high ion concentrations, leading to the quenching of the 1460-nm emission and enhancement of the 1860-nm emission band (Figure 3a). On the other hand, the quantum efficiency drops from 32 to 6% for TeO2-ZnO glass host. As the Tm3+ concentration increased, similar enhancement behavior was observed for the other tellurite glass hosts. It is important to find an optimum concentration level here where quantum yield and pump absorption are not dramatically low and still cross-relaxation is significant. Regarding this fact, 0.25 or 0.5 mol % Tm2O3-doped samples are potential candidates as hosts in laser applications for bulk as well as fiber applications. Figure 3b shows the emission spectra of the glass hosts with different glass modifiers at 1.0 mol % dopant concentration. As can be seen, the 1460-nm emission band remains similar for different glass modifiers. On the other hand, glass modifier changes produce a notable change in the 1860-nm emission band. In particular, the TeNbKTm glass has the highest emission peak at 1860 nm, whereas the lowest peak of 1795 nm was observed in the TeNbTm host.
In order to obtain the emission cross section, we used both the Fuchtbauer–Ladenburg equation [35,36],
σ e m ( λ ) = λ 5 I ( λ ) A ( J , J ' ) 8 π n 2 c I ( λ ) λ d λ ,
and the McCumber formula [37],
σ e m ( λ ) = σ a ( λ ) [ Z l Z u ] exp [ h c k T ( 1 λ Z P 1 λ ) ] ,
where σ a ( λ ) is the absorption cross section, [ Z l Z u ] represents the ratio of the parity functions of lower and upper states (13/9 for the Tm3+ ion), h is the Planck constant, k is the Boltzmann constant, T is the absolute temperature, and λ Z P is the wavelength of the zero phonon line. From the overlap of the absorption and emission bands, λ Z P was estimated to be 1780 nm. McCumber formalism provides an accurate estimate of the emission cross section as well as its spectral distribution for narrow band emission [38]. On the other hand, the Fuchtbauer–Ladenburg formalism is proven to be accurate for low ion-doping concentrations, where the reabsorption of the emitted photons is negligible and the analysis is applicable to relatively broadband emissions in comparison with the case in McCumber approach, since there is no assumption of the band shape. In Table 3, the emission cross sections obtained by using both formalisms are shown for the two emission bands of the tellurite glasses at different Tm2O3 concentrations. By using the Fuchtbauer–Ladenburg equation for the TeNbTm sample, we obtained average stimulated emission cross section values of 2.62 ± 0.02 × 10−21 and 5.63 ± 0.02 × 10−21 cm2 for the 1460-nm and 1860-nm bands, respectively. As can be further seen from the table for the sample containing zinc as a glass modifier (TeZnTm), the average emission cross sections were determined to be 2.78 ± 0.39 × 10−21 and 6.33 ± 0.34 × 10−21 cm2 for the 1460- and 1860-nm bands, respectively. These results compare well with values reported for other tellurite hosts [9,39] and it can also be concluded that for the TeZnTm sample, we obtained higher stimulated emission cross sections (as high as 6.33 ± 0.34 × 10−21 cm2), especially for the 1860-nm emission band. As can be seen from Table 3, the TeNbKTm glass host has the highest average absorption cross section (8.18 × 10−21 cm2) among the three hosts, at 794 nm. The TeZnTm glass host also has a reasonably high absorption cross section (7.85 × 10−21 cm2). Furthermore, at low doping concentration levels, the quantum efficiencies for both transitions are quite high in comparison with the other hosts, suggesting that the TeZnTm glasses are the most promising laser gain media among the materials explored in this study.
As can be seen from Table 3, both formalisms agree reasonably well for all glasses investigated in this study. Figure 4 further shows the absorption and emission cross sections as a function of wavelength for the TeNbKTm samples with 1.0 mol % doping concentration, based on the McCumber analysis. As can be seen, the emission cross section has a broad, smooth peak, which makes the glass system a promising candidate for ultrafast laser applications.

4. Conclusions

We have conducted a detailed spectroscopic analysis on three new types of tellurite glass materials (Tm2O3-TeO2-ZnO (TeZnTm), Tm2O3-TeO2-Nb2O5 (TeNbTm) and Tm2O3-TeO2-K2O-Nb2O5 (TeNbKTm)) in order to investigate the effect of the host composition as well as Tm3+ ion concentration on the cross sections and the luminescence quantum efficiencies. From the data, it can be concluded that J-O intensity parameters vary slightly among tellurite glasses with different compositions, as expected. Yet, our calculations showed that for the sample containing zinc as a glass modifier (TeZnTm), we obtained shorter radiative lifetimes for the transition corresponding to the 1860-nm band. We obtained average radiative lifetimes of 2.55 ± 0.07 ms and 2.76 ± 0.03 ms for the TeZnTm and TeNbTm samples for the 3H4 level.
From the lifetime measurements and the radiative lifetime calculations based on the J-O theory, we determined the luminescence quantum efficiencies of the samples. The highest quantum efficiency for the sample TeZnTm was 32% for the 1860-nm emission band and 89% for the 1460-nm emission band. For the sample TeNbTm, we obtained the highest quantum yields as 32% and 76% for the 1860-nm and 1460-nm emission bands, respectively. For higher doping concentrations, we observed a cross-relaxation effect, which enhanced the 1860-nm emission band. Samples with low concentration provide higher quantum efficiencies that are favorable for lasing applications but on the other hand, exhibit weak absorption. When the ion concentration is increased, quantum efficiencies decrease but the absorption near 800 nm together with the cross relaxation effect enhancing the 1860 nm band increase, which suggests that 0.25 or 0.5 mol % samples are potential candidates as hosts in laser applications for bulk as well as fiber applications.
For the TeNbTm sample, we obtained average emission cross sections of 2.62 ± 0.02 × 10−21 and 5.63 ± 0.02 × 10−21 cm2 for the 1460-nm and 1860-nm bands, respectively. For the sample containing zinc as a glass modifier (TeZnTm), the average emission cross sections were determined to be 2.78 ± 0.39 × 10−21 and 6.33 ± 0.39 × 10−21 cm2 for the 1460-nm and 1860-nm bands, respectively. These results compare well with values reported for other tellurite hosts [9,39] and it can also be pointed out that for the TeZnTm sample, we obtained higher stimulated emission cross sections, especially for the 1860-nm emission band. Among the hosts we analyzed, TeZnTm glass stands as a good candidate for the development of 2-μm lasers with high emission cross sections near 1860 nm.

Author Contributions

Huseyin Cankaya and Adil Tolga Gorgulu performed the spectroscopic studies, consisting of absorption measurements, emission measurements, and data analysis by using the Judd–Ofelt and McCumber theory. Huseyin Cankaya prepared the manuscript, which was edited by Alphan Sennaroglu, and Marco Bettinelli, Alphan Sennaroglu and Marco Bettinelli designed and supervised the spectroscopic studies. Alphan Sennaroglu built the laser used during the experiments and prepared the codes for data analysis. Adnan Kurt prepared the computer-controlled emission measurement system and the related data fitting programs. Marco Bettinelli and Adolfo Speghini fabricated the glasses under investigation and contributed to the analysis of the experimental data. Adolfo Speghini also measured the preliminary absorption spectra.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Carrig, T.J.; Hankla, A.K.; Wagner, G.J.; Rawle, C.B.; Kinnie, I.T.M. Tunable Infrared Laser Sources for Dial. In Proceedings of the SPIE Laser Radar Technology and Applications VII, AEROSENSE 2002, Orlando, FL, USA, 1–5 April 2002; pp. 147–155. [Google Scholar]
  2. Bilici, T.; Tabakoglu, H.O.; Topaloglu, N.; Kalaycioglu, H.; Kurt, A.; Sennaroglu, A.; Gulsoy, M. Modulated and continuous-wave operations of low-power thulium (Tm:YAP) laser in tissue welding. J. Biomed. Opt. 2010, 15, 038001. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, J.S.; Snitzer, E.; Vogel, E.M.; Sigel, J.G.H. 1.47, 1.88, and 2.8 μm emission of Tm3+ and Tm3+-Ho3+-codoped tellurite glasses. J. Lumin. 1994, 60, 145–149. [Google Scholar] [CrossRef]
  4. McAleavey, F.J.; O’Gorman, J.; Donegan, J.F.; MacCraith, B.D.; Hegarty, J.; Maze, G. Narrow linewidth, tunable Tm—Doped fluoride fiber laser for optical-based hydrocarbon gas sensing. IEEE J. Sel. Top. Quantum Electron. 1997, 3, 1103–1111. [Google Scholar] [CrossRef]
  5. Tsanga, Y.H.; Colemana, D.J.; King, T.A. High power 1.9 μm Tm3+-silica fibre laser pumped at 1.09 μm by a Yb3+-silica fibre laser. Opt. Commun. 2004, 231, 357–364. [Google Scholar] [CrossRef]
  6. Wu, J.F.; Yao, Z.D.; Zong, J.; Jiang, S.B. Highly efficient high-power thulium-doped germanate glass fiber laser. Opt. Lett. 2007, 32, 638–640. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, J.S.; Vogel, E.M.; Snitzer, E. Tellurite glass: A new candidate for fiber devices. Opt.Mater. 1994, 3, 187–203. [Google Scholar] [CrossRef]
  8. Yamane, M.; Asahara, Y. Glasses for Photonics; Cambridge University Press: Cambridge, UK, 2000. [Google Scholar]
  9. Kalaycioglu, H.; Cankaya, H.; Cizmeciyan, M.N.; Sennaroglu, A.; Ozen, G. Spectroscopic investigation of Tm3+:TeO2-WO3 glass. J. Lumin. 2008, 128, 1501–1506. [Google Scholar] [CrossRef]
  10. Zou, X.L.; Toratani, H. Spectroscopic properties and energy transfers in Tm3+ singly- and Tm3+/Ho3+ doubly-doped glasses. J. Non Cryst. Solids 1996, 195, 113–124. [Google Scholar] [CrossRef]
  11. Azkargorta, J.; Iparraguirre, I.; Balda, R.; Fernández, J. On the origin of bichromatic laser emission in Nd3+-doped fluoride glasses. J. Non Cryst. Solids 2008, 16, 11894–11906. [Google Scholar]
  12. Lei, N.; Xu, B.; Jiang, Z.H. Ti:Sapphire laser pumped nd:Tellurite glass laser. Opt. Commun. 1996, 127, 263–265. [Google Scholar] [CrossRef]
  13. Mori, A.; Sakamoto, T.; Kobayashi, K.; Shikano, K.; Oikawa, K.; Hoshino, K.; Kanamori, T.; Ohishi, Y.; Shimizu, M. 1.58 um broad-band erbium-doped tellurite fiber amplifier. J. Lightwave Technol. 2002, 20, 794. [Google Scholar] [CrossRef]
  14. Wang, J.S.; Machewirth, D.P.; Wu, F.; Snitzer, E.; Vogel, E.M. Neodymium-doped tellurite single-mode fiber laser. Opt. Lett. 1994, 19, 1448–1449. [Google Scholar] [CrossRef] [PubMed]
  15. Kalaycioglu, H.; Cankaya, H.; Ozen, G.; Ovecoglu, L.; Sennaroglu, A. Lasing at 1065 nm in bulk Nd3+-doped telluride-tungstate glass. Opt. Commun. 2008, 281, 6056–6060. [Google Scholar] [CrossRef]
  16. Cankaya, H.; Sennaroglu, A. Bulk Nd3+-doped tellurite glass laser at 1.37 μm. Appl. Phys. B Lasers Opt. 2010, 99, 121–125. [Google Scholar] [CrossRef]
  17. Li, K.; Zhang, G.; Hu, L. Watt-level 2 μm laser output in Tm3+-doped tungsten tellurite glass double-cladding fiber. Opt. Lett. 2010, 35, 4136–4138. [Google Scholar] [CrossRef] [PubMed]
  18. Richards, B.; Tsang, Y.; Binks, D.; Lousteau, J.; Jha, A. 2 μm Tm3+/Yb3+-Doped Tellurite Fibre Laser. In Proceedings of the 5th International Conference on Optical, Optoelectronic and Photonic Materials and Applications, London, UK, 29 July–3 August 2007; pp. 317–320. [Google Scholar]
  19. Tsang, Y.; Richards, B.; Binks, D.; Lousteau, J.; Jha, A. Tm3+/Ho3+ codoped tellurite fiber laser. Opt. Lett. 2008, 32, 1282–1284. [Google Scholar] [CrossRef]
  20. Richards, B.; Shen, S.; Jha, A.; Tsang, Y.; Binks, D. Infrared emission and energy transfer in Tm3+, Tm3+-Ho3+ and Tm3+-Yb3+-doped tellurite fibre. Opt. Express 2007, 15, 6546–6551. [Google Scholar] [CrossRef] [PubMed]
  21. Debnath, R.; Bose, S. A comprehensive phononic sof phonon assisted energy transfer in the Yb3+ aided upconversion luminescence of Tm3+ and Ho3+ in solids. J. Lumin. 2015, 161, 103–109. [Google Scholar] [CrossRef]
  22. Ma, Y.; Wang, X.; Zhang, L.; Huang, F.; Hu, L. Increased radiative lifetime of Tm3+:3F4 3H6 transition in oxyfluoride tellurite glass. Mater. Res. Bull. 2015, 64, 262–266. [Google Scholar] [CrossRef]
  23. Wang, W.C.; Zhang, W.J.; Li, L.X.; Liu, Y.; Chen, D.D.; Qian, Q.; Zhang, Q.Y. Spectroscopic and structural characterization of barium tellurite glass fibers for mid-infrared ultra-broad tunable fiber lasers. Opt. Mater. Express 2016, 6, 2095–2107. [Google Scholar] [CrossRef]
  24. Shen, L.; Cai, M.; Lu, Y.; Wang, N.; Huang, F.; Xu, S.; Zhang, J. Preparation and investigation of Tm3+/Ho3+ co-doped germanate-tellurite glass as promising materials for ultrashort pulse laser. Opt. Mater. 2017, 67, 125–131. [Google Scholar] [CrossRef]
  25. Zhou, D.; Bai, X.; Zhou, H. Preparation of Ho3+/Tm3+ co-doped lanthanum tungsten germanium tellurite glass fiber and its laser performance for 2.0 μm. Sci. Rep. 2017, 7, 44747. [Google Scholar] [CrossRef] [PubMed]
  26. Fusari, F.; Vetter, S.; Lagatsky, A.A.; Richards, B.; Calvez, S.; Jha, A.; Dawson, M.D.; Sibbett, W.; Brown, C.T.A. Tunable laser operation of a Tm3+-doped tellurite glass laser near 2 μm pumped by a 1211 nm semiconductor disk laser. Opt. Mater. 2010, 32, 1007–1010. [Google Scholar] [CrossRef]
  27. Fusari, F.; Lagatsky, A.A.; Richards, B.; Jha, A.; Sibbett, W.; Brown, C.T.A. Spectroscopic and lasing performance of Tm3+- doped bulk tzn and tzng tellurite glasses operating around 1.9 µm. Opt. Express 2008, 16, 19146–19151. [Google Scholar] [CrossRef] [PubMed]
  28. Fusari, F.; Lagatsky, A.A.; Jose, G.; Calvez, S.; Jha, A.; Dawson, M.D.; Gupta, J.A.; Sibbett, W.; Brown, C.T.A. Femtosecond mode-locked Tm3+ and Tm3+-Ho3+ doped 2 μm glass lasers. Opt. Express 2010, 18, 22090–22098. [Google Scholar] [CrossRef] [PubMed]
  29. Tolga Gorgulu, A.; Cankaya, H.; Kurt, A.; Speghini, A.; Bettinelli, M.; Sennaroglu, A. Spectroscopic characterization of Tm3+:TEO2-K2O-Nb2O5 glasses for 2-μm lasing applications. J. Lumin. 2012, 132, 110–113. [Google Scholar] [CrossRef]
  30. Judd, B.R. Optical absorption intensities of rare-earth ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef]
  31. Ofelt, G.S. Intensities of crystal spectra of rare-earth ions. J. Chem. Phys. 1962, 37, 511. [Google Scholar] [CrossRef]
  32. Kaminskii, A.A. Laser Crystals; Springer: Berlin, Germany, 1990. [Google Scholar]
  33. Gebavi, H.; Milanese, D.; Balda, R.; Chaussedent, S.; Ferrari, M.; Fernandez, J.; Ferraris, M. Spectroscopy and optical characterization of thulium doped TZN glasses. J. Phys. D Appl. Phys. 2010, 43, 13. [Google Scholar] [CrossRef] [Green Version]
  34. Sennaroglu, A.; Kabalci, I.; Kurt, A.; Demirbas, U.; Ozen, G. Spectroscopic properties of Tm3+:TeO2-PbF2 glasses. J. Lumin. 2006, 116, 79–86. [Google Scholar] [CrossRef]
  35. Moulton, P.F. Spectroscopic and laser characteristics of ti-Al2O3. J. Opt. Soc. Am. B Opt. Phys. 1986, 3, 125–133. [Google Scholar] [CrossRef]
  36. Sudesh, V.; Goldys, E.M. Spectroscopic properties of thulium-doped crystalline materials including a novel host, La2Be2O5: A comparative study. J. Opt. Soc. Am. B Opt. Phys. 2000, 17, 1068–1076. [Google Scholar] [CrossRef]
  37. McCumber, D.E. Einstein relations connecting broadband emission and absorption spectra. Phys. Rev. 1964, 136, A954. [Google Scholar] [CrossRef]
  38. Digonnet, M.J.F.; Murphy-Chutorian, E.; Falquier, D.G. Fundamental limitations of the mccumber relation applied to er-doped silica and other amorphous-host lasers. IEEE J. Quantum Electron 2002, 38, 1629–1637. [Google Scholar] [CrossRef]
  39. Sennaroglu, A.; Kurt, A.; Özen, G. Effect of cross relaxation on the 1470 and 1800 nm emissions in Tm3+:TeO2-CdCl2 glass. J. Phys. Condens. Matter 2004, 16, 2471–2478. [Google Scholar] [CrossRef]
Figure 1. (a) Energy level diagram of the Tm3+ ion. CR: Cross-relaxation (b) Absorption spectra of two Tm2O3: TeO2-Nb2O5 glass samples with Tm2O3 concentrations of 1.0 and 0.25 mol % in the range of 450–2050 nm. The corresponding absorption bands from the ground state (3H6) are indicated.
Figure 1. (a) Energy level diagram of the Tm3+ ion. CR: Cross-relaxation (b) Absorption spectra of two Tm2O3: TeO2-Nb2O5 glass samples with Tm2O3 concentrations of 1.0 and 0.25 mol % in the range of 450–2050 nm. The corresponding absorption bands from the ground state (3H6) are indicated.
Applsci 08 00333 g001
Figure 2. Measured fluorescence decay curves and the exponential fits of the (a) 3F4 level and (b) 3H4 level for the two Tm2O3: TeO2-K2O-Nb2O5 (TeNbKTm) samples with Tm2O3 concentrations of 1.0 and 0.125 mol % [29].
Figure 2. Measured fluorescence decay curves and the exponential fits of the (a) 3F4 level and (b) 3H4 level for the two Tm2O3: TeO2-K2O-Nb2O5 (TeNbKTm) samples with Tm2O3 concentrations of 1.0 and 0.125 mol % [29].
Applsci 08 00333 g002
Figure 3. (a) Emission spectra of the two transitions of the Tm3+ ion originating from 3F4 and 3H4 levels to the ground level for the glass host TeZnTm with Tm2O3 concentrations of 1.0 and 0.25 mol %; (b) Emission spectra of the three glass hosts TeNbKTm, TeZnTm and TeNbTm, each with 1.0 mol % Tm2O3 doping.
Figure 3. (a) Emission spectra of the two transitions of the Tm3+ ion originating from 3F4 and 3H4 levels to the ground level for the glass host TeZnTm with Tm2O3 concentrations of 1.0 and 0.25 mol %; (b) Emission spectra of the three glass hosts TeNbKTm, TeZnTm and TeNbTm, each with 1.0 mol % Tm2O3 doping.
Applsci 08 00333 g003
Figure 4. Absorption and emission cross section of the 1.0% Tm2O3-doped TeNbKTm samples deduced by using the McCumber analysis.
Figure 4. Absorption and emission cross section of the 1.0% Tm2O3-doped TeNbKTm samples deduced by using the McCumber analysis.
Applsci 08 00333 g004
Table 1. Judd–Ofelt (J-O) intensity parameters for three different glass hosts with 1 mol % Tm ion concentration.
Table 1. Judd–Ofelt (J-O) intensity parameters for three different glass hosts with 1 mol % Tm ion concentration.
J-O ParametersΩ2Ω4Ω6
(10−20 cm2)(10−20 cm2)(10−20 cm2)
TeNbKTm4.51 ± 0.790.76 ± 0.461.13 ± 0.67
TeZnTm4.02 ± 0.560.93 ± 0.331.12 ± 0.47
TeNbTm4.09 ± 0.030.69 ± 0.011.11 ± 0.01
Table 2. Measured fluorescence lifetimes (τF), calculated radiative lifetimes (τR), and the corresponding luminescence quantum yields (η) for the two transitions at 1860 nm and 1460 nm.
Table 2. Measured fluorescence lifetimes (τF), calculated radiative lifetimes (τR), and the corresponding luminescence quantum yields (η) for the two transitions at 1860 nm and 1460 nm.
Fluorescence and Radiative Lifetimes and Quantum YieldsτF (μs)η (%)τRaveraged (ms)
3H43F43H43F4τR (1860 nm)τR (1460 nm)
TeNbKTm [29]1860 nm1460 nm1860 nm1460 nm(ms)(ms)
0.125 mol %81425832742.570.35
0.25 mol %5721992257--
0.5 mol %5451482143--
1.0 mol %439471714--
TeZnTm τR (1860 nm)τR (1460 nm)
0.125 mol %80632732892.550.37
0.25 mol %7222502868--
0.5 mol %5111712047--
1.0 mol.%14535610--
TeNbTm τR (1860 nm)τR (1460 nm)
0.25 mol %89328432762.760.37
1.0 mol %382761420--
Table 3. Emission cross sections and bandwidths at full-wave half maximum (FWHM) of the 1460-nm and 1860-nm bands, and absorption cross sections at 794 nm for the Tm3+:TeO2-based glass hosts at different Tm2O3 concentrations.
Table 3. Emission cross sections and bandwidths at full-wave half maximum (FWHM) of the 1460-nm and 1860-nm bands, and absorption cross sections at 794 nm for the Tm3+:TeO2-based glass hosts at different Tm2O3 concentrations.
Δλ (FWHM, nm)σem (10−21 cm2)σa (10−21 cm2)
1460 nm1860 nm1460 nm1860 nm1860 nm794 nm
TeNbKTm [29] Fucht.-Laden.McCumber
1.0 mol %1221702.67 ± 0.535.61 ± 0.426.219.07
0.5 mol %1162312.74 ± 0.546.12 ± 0.456.048.33
0.25 mol %1333192.40 ± 0.485.46 ± 0.405.798.26
0.125 mol %1282403.00 ± 0.596.85 ± 0.516.107.07
Average 2.70 ± 0.546.01 ± 0.456.04 ± 0.188.18
TeZnTm
1.0 mol %1061922.74 ± 0.405.48 ± 0.306.008.08
0.5 mol %1192002.52 ± 0.375.65 ± 0.315.007.79
0.25 mol %1111722.91 ± 0.376.85 ± 0.334.987.65
0.125 mol %1062002.95 ± 0.437.36 ± 0.405,607.89
Average 2.78 ± 0.396.33 ± 0.345.40 ± 0.57.85
TeNbTm
1.0 mol %1052372.65 ± 0.024.99 ± 0.025.007.68
0.25 mol %1001032.60 ± 0.026.28 ± 0.026.206.90
Average 2.62 ± 0.025.63 ± 0.025.60 ± 0.857.29

Share and Cite

MDPI and ACS Style

Cankaya, H.; Gorgulu, A.T.; Kurt, A.; Speghini, A.; Bettinelli, M.; Sennaroglu, A. Comparative Spectroscopic Investigation of Tm3+:Tellurite Glasses for 2-μm Lasing Applications. Appl. Sci. 2018, 8, 333. https://doi.org/10.3390/app8030333

AMA Style

Cankaya H, Gorgulu AT, Kurt A, Speghini A, Bettinelli M, Sennaroglu A. Comparative Spectroscopic Investigation of Tm3+:Tellurite Glasses for 2-μm Lasing Applications. Applied Sciences. 2018; 8(3):333. https://doi.org/10.3390/app8030333

Chicago/Turabian Style

Cankaya, Huseyin, Adil Tolga Gorgulu, Adnan Kurt, Adolfo Speghini, Marco Bettinelli, and Alphan Sennaroglu. 2018. "Comparative Spectroscopic Investigation of Tm3+:Tellurite Glasses for 2-μm Lasing Applications" Applied Sciences 8, no. 3: 333. https://doi.org/10.3390/app8030333

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop