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

Enhanced 2.0 μm emission and gain coefficient of transparent glass ceramic containing BaF2: Ho3+,Tm3+ nanocrystals

Open Access Open Access

Abstract

Transparent glass ceramic containing BaF2:Ho3+,Tm3+ nanocrystals has been prepared by melt quenching and subsequent thermal treatment. The precipitation of BaF2 nanocrystals was confirmed by X-ray diffraction and high-resolution transmission electron microscopy. Intense 2.0 µm fluorescence originating from Ho3+: 5I75I8 transition was achieved upon excitation with 808 nm laser diode. A large ratio of forward Tm3+ → Ho3+ energy transfer constant to that of backward process indicated high efficient energy transfer from Tm3+(3F4) to Ho3+(5I7), benefited from the reduced ionic distances of Tm3+-Tm3+ and Tm3+-Ho3+ pairs and low phonon energy environment with the incorporation of rare-earth ions into the precipitated BaF2 nanocrystals. The results indicate that glass ceramic is a promising candidate material for 2.0 μm laser.

©2009 Optical Society of America

1. Introduction

Solid state lasers operating at approximately 2.0 μm have been the subject of several spectroscopic investigations for their potential applications in eye-safe laser radar, medical surgery and atmospheric monitoring [13]. In this respect, Ho3+ has been recognized as suitable laser active ion for its high effective stimulated emission cross section and long-lived 5I7 level related to Ho3+: 5I75I8 transition [4]. With the advent and rapid development of AlGaAs laser diodes (LD), diode-pumped Tm3+-Ho3+ system has been proposed as a promising way to achieve efficient 2.0 μm laser. In this context, the drawback of Ho3+, i.e., lack of convenient bands for diode pumping can be circumvented via pumping Tm3+: 3H4 state followed by resonant energy transfer (ET) from Tm3+ to Ho3+. Moreover, a pump quantum efficiency of 200% can be expected from the cross relaxation (CR) between two neighboring Tm3+ ions: 3H4 + 3H63F4 + 3F4 [5]. Indeed, efficient laser operation in the vicinity of 2.0 μm has been demonstrated in Ho3+/Tm3+ co-doped various fluoride crystals and fluoride fibers [68], benefited from low phonon energy, high transparency and good rare-earth (RE) ion solubility [9]. However, the poor chemical and mechanical stability of fluoride fibers has limited their applications [10]. On the other hand, single crystals are often difficult and costly to produce.

Since Wang and Ohwaki reported of greatly enhanced Er3+ upconversion emission in transparent oxyfluoride glass ceramic (GC) in 1993 [11], RE doped transparent oxyfluoride GC as a promising optical material, has attracted much attention for the excellent combined properties of high chemical and mechanical stability related to oxide glass and low phonon energy of the precipitated fluoride nanocrystals [12]. It has been reported that the spectroscopic properties of RE ions were greatly enhanced with the incorporation of RE into fluoride nanocrystals during the creaming process [12,13]. In 2001, Samson et al. reported efficient laser operation in Nd3+-doped GC fiber and further highlighted the potential of such material for a new range of optical devices [14].

Herein, this paper investigates the spectroscopic and gain properties of 2.0 µm emission in transparent GC containing BaF2: Ho3+, Tm3+ nanocrystals to assess its potentiality as 2.0 µm laser host. We have demonstrated intense 2.0 µm fluorescence of Ho3+ upon excitation of 808 nm LD. ET processes between Tm3+ and Ho3+ were analyzed in terms of ET constants CTmHo and CHoTmcalculated by Dexter’s model for dipole-dipole interaction. The results show that transparent GC is a good candidate for 2.0 µm laser.

2. Experimental

Glasses of molar composition (59-x)SiO2-20ZnO-20BaF2-1TmF3-xHoF3 (denoted as SZBTx) (x = 0, 0.1, 0.3, 0.6, 1.0, and 1.5) were prepared by melting mixtures of raw materials in covered corundum crucibles at 1460 °C for 1 h in air. The melts were then quenched on a preheated stainless steel plate followed by annealing at 480 °C for 2 h. The obtained glasses were polished and heat-treated at 580 °C for 6 h to fabricate transparent GC. Microstructures of GC were characterized by X-ray powder diffractometer (X’ Pert PROX, Cu Kα) and transmission electron microscope (TEM, JEM-2010). Optical absorption spectra measurements were carried out on a Perkin-Elmer Lambda 900/UV/VIS/NIR spectrophotometer. The fluorescence spectra were recorded on a computer-controlled Triax 320 spectrofluorimeter (Jobin-Yvon Corp.) upon excitation of an 808 nm LD. Lifetimes were determined from the first e-folding time of emission intensities in the decay curves recorded with a digital oscilloscope, a mechanical chopper and an InSb detector.

3. Results and discussion

Figure 1 (a) shows the XRD patterns of Ho3+/Tm3+ co-doped glass and GC. It is clear that the precursor glass is completely amorphous without any diffraction peak. After thermal treatment at 580 °C for 6 h, the XRD pattern exhibits several broad peaks, which can be indexed to the cubic BaF2 phase (JCPDS Card No. 00-004-0452). Moreover, one can observe that the diffraction peaks move towards the higher angle side as compared with the XRD pattern of standard BaF2 crystals, indicative of the occurrence of lattice shrinkage in BaF2 nanocrystals resulted from the replacement of Ba2+ with smaller ions (Tm3+, Ho3+). The lattice parameter of BaF2 nanocrystals is found to be 6.08 Å, which is smaller than that of the standard BaF2 crystal (6.20 Å). TEM image and corresponding selected area electron diffraction pattern shown in Fig. 1(b) demonstrate the composite structure of GC with nanocrystals sized 7-19 nm (see Fig. 1(c)) distributing homogenously among glass matrix. The high-resolution TEM image in Fig. 1(d) shows the detailed lattice structure of an individual BaF2 nanocrystal.

 figure: Fig. 1

Fig. 1 (a). XRD patterns of the precursor glass and GC heat-treated at 580°C for 6 h. (b) TEM bright field image and corresponding selected area electron diffraction pattern (left upper side) of the GC. (c) the size distribution of BaF2 nanocrystals embedded in glass matrix. (d) High-resolution TEM image of a BaF2 nanocrystal.

Download Full Size | PDF

Figure 2 (a) illustrates the absorption spectra of GC samples doped with 1 mol.% Tm3+ and/or 0.6 mol.% Ho3+. Referring to Tm3+-doped sample, the absorption spectrum consists of seven absorption bands centered at 354, 467, 657, 680, 790, 1208 and 1669 nm, ascribed to transitions from the 3H6 ground state to excited states 1D2, 1G4, 3F2, 3F3, 3H4, 3H5 and 3F4, respectively. Moreover, it is noted that the absorption spectrum of Tm3+-doped GC is characteristic of two Stark split peaks around 1669 nm due to the partition of Tm3+ into BaF2 nanocrystals, which is similar to the case in GC containing PbF2 nanocrystals [15]. The absorption spectrum of Ho3+ is characterized by thirteen bands at 333, 345, 360, 385, 417, 452, 472, 485, 541, 641, 890, 1169 and 1977 nm, assigned to transitions from 5I8 to 3F4, 3L9, 3H5 + 3H6, 5G4 + 3K7, 5G5, 5G6 + 5F1, 5F2 + 3K8, 5F3, 5S2 + 5F4, 5F5, 5I5, 5I6 and 5I7 levels, respectively. Meanwhile, it is observed that the absorption spectrum of Ho3+/Tm3+ co-doped sample is simply a superposition of the absorption spectra of Tm3+ and Ho3+. To investigate the absorption variation of the sample after crystallization, the Judd-Ofelt analysis was applied to the precursor glass and the GC. In general, the Judd-Ofelt intensity parameter Ω2 is sensitive to asymmetry around RE ions [16] and will decrease with the increase in symmetry between lanthanide ions and the ligand fields. Besides, it is often related to the nature of bonding between RE ions and the ligand anions. For instances, the Ω2 value decreases with the host changing from oxides to fluorides [17]. From the precursor glass to the GC, the decrease of Ω2 (in units of 10−20 cm2) from 3.33 to 2.79 for Tm3+ and 3.65 to 2.23 for Ho3+ verifies the partition of RE ions into the BaF2 nanocrystals. Absorption cross sections corresponding to the Tm3+:3H63F4 and Ho3+:5I85I7 transitions were calculated from the measured absorption spectra. Fuchtbauer-Ladenburg formula [18] and McCumber theory [19] were used to calculate the corresponding stimulated emission cross sections of Tm3+ and Ho3+, respectively. The calculated cross sections for GC are presented in Fig. 2(b). Significant overlap between Tm3+ (3F43H6) emission cross section and Ho3+ (5I85I7) absorption cross section suggests the great possibility of ET from Tm3+ (3F4) to Ho3+ (5I7).

 figure: Fig. 2

Fig. 2 (a). Absorption spectra of GC samples doped with 1 mol.% Tm3+ and/or 0.6 mol.% Ho3+. (b) Absorption (solid line) and emission cross sections (dash line) corresponding to the 3H6 - 3F4 transition of Tm3+ and the 5I8 - 5I7 transition of Ho3+ in GC.

Download Full Size | PDF

Figure 3 shows the fluorescence spectra of SZBTx glass samples upon excitation of 808 nm LD. For Tm3+ doped sample, the spectrum exhibits two emission bands located at approximately 1.46 and 1.8 μm, attributed to the Tm3+: 3H43F4 and 3F43H6 transitions. Obviously, a trace addition of Ho3+ in the Tm3+ doped sample gives rise to 2.0 µm emission band originated from Ho3+: 5I75I8 transition. Compared with Tm3+-doped sample, the emission intensity at 1.8 μm decreases monotonously with increasing Ho3+ concentration, whereas 2.0 μm emission intensity increases significantly before it reaches the maximum value when 0.6 mol.% Ho3+ is doped. These observations indicate that the 2.0 µm emission of Ho3+ can be achieved through efficient ET from Tm3+ to Ho3+. The principal involved ET mechanisms for Tm3+-Ho3+ ions system have been proposed, as shown in the inset of Fig. 3. Upon excitation at 808 nm, Tm3+: 3H4 state is excited, followed by CR process between two adjacent Tm3+ ions (3H4, 3H63F4, 3F4), resulting in effective “two for one” excitation of Tm3+: 3F4 level [20]. After that, Ho3+: 5I7 level can be populated via ET from Tm3+ to Ho3+, i.e., Tm3+: 3F4 + Ho3+: 5I8 → Tm3+: 3H6 + Ho3+: 5I7. Finally, radiative transition of Ho3+: 5I75I8 occurs, yielding 2.0 µm emission. On the other hand, there might exist a direct route to populate Ho3+: 5I7 state, i.e., CR process of Tm3+: 3H4 + Ho3+: 5I8 → Tm3+: 3F4 + Ho3+: 5I7.

 figure: Fig. 3

Fig. 3 HoF3 concentration dependence of Tm3+ and Ho3+ fluorescence intensities in SZBTx glass samples under 808 nm excitation. The inset shows the simplified energy level diagram for Tm3+-Ho3+ ions system.

Download Full Size | PDF

To examine the suitability of GC as potential 2.0 µm laser material, SZBT0.6 was selected to precede controlled crystallization. The fluorescence spectra of the precursor glass and GC are presented in a comparable way in Fig. 4(a) . Obviously, the emission spectrum of GC exhibits Stark split emission peaks around 2.0 µm due to the partition of Ho3+ ions into BaF2 nanocrystals. Moreover, integrated emission intensity of Ho3+ in GC increases by a factor of about 6 compared with that of the precursor glass. This can be ascribed to the enhanced ET efficiency from Tm3+ (3F4) to Ho3+ (5I7), benefited from the concentration of RE ions in BaF2 nanocrystals. In this case, the distances between RE ions would be shorter than those of the ions distributed in the precursor glass, resulting in the efficient CR and ET processes of Tm3+-Tm3+ and Tm3+-Ho3+ pairs. Meanwhile, the multiphonon relaxation of Tm3+: 3H43H5, which competes with CR process (3H4, 3H63F4, 3F4), can be limited due to the low phonon energy environment of BaF2 nanocrystals. Assuming that electrons in Ho3+ are either in the 5I7 state or in the 5I8 state, the gain coefficient G(λ)of Ho3+ near the 2.0 µm wavelength region can be determined by the following equation [21]:

G(λ)=N[pσe(λ)(1p)σa(λ)]
where, N = 1.482 × 1020 ions/cm3 is the total concentration of Ho3+, p is the population inversion given by the ratio between the population of Ho3+ lasing level (5I7) and the total Ho3+ concentration, σe(λ) is the emission cross section of Ho3+: 5I75I8 transition calculated from the corresponding absorption cross section σa(λ).

 figure: Fig. 4

Fig. 4 (a). Fluorescence spectra of glass and GC co-doped with 0.6 mol.% Ho3+ and 1 mol.% Tm3+ under 808 nm excitation. (b) Calculated gain coefficients corresponding to the 5I8 - 5I7 transition of Ho3+ for the GC sample.

Download Full Size | PDF

Figure 4(b) shows the calculated gain coefficients versus wavelengths of 5I7 - 5I8 transition in Ho3+/Tm3+ co-doped GC by assuming a set of p values ranging from 0 to 1. It is noted that the laser performance wavelengths of the maximum gain coefficient shift to longer wavelengths with decreasing the value of p. This behavior is a typical characteristic of the quasi-three level system. To estimate the extent of ET between Tm3+ and Ho3+, ET constantCDAand critical radiusR0are calculated by Dexter theory [22]. In the case of dipole-dipole interaction, they can be estimated by using the following equations:

CDA=R06τ0, R06=3h4c4Q64π5n4gi(E)gj(E)E4dE
whereτ0is the effective decay time of the donor ion in the absence of an acceptor ion,R0 is defined as the distance (between donor and acceptor ions) at which the ET rate becomes equal to the intrinsic decay rate of the donor. The meaning of other parameters is referred in Ref [23]. The measured fluorescence lifetimes at present research (singly Tm3+ (Ho3+) ion-doped GC) for Tm3+: 3F4 level and Ho3+: 5I7 level are 1.628 ms and 2.829 ms, which are close to those reported in tellurite glass [24,25], but much shorter than the reported values of fluorozirconate glass with lower phonon energy [26]. The values ofR0 for the forward ET of Tm3+: 3F4 → Ho3+: 5I7 and the reverse process were evaluated to be 17.26 Å and 10.8 Å, respectively. Hence, CTmHo for the forward ET in GC was calculated to be 1.624 × 10−38 cm6s−1, while the value of CHoTm corresponding to the backward ET was found to be 0.561 × 10−39 cm6s−1. A high value of CTmHo/CHoTm indicates that ET from Tm3+ (3F4) to Ho3+ (5I7) in GC is quite efficient, which is beneficial to efficient Ho3+ 2.0 µm laser.

4. Conclusion

In summary, a transparent GC containing BaF2:Ho3+,Tm3+ nanocrystals was prepared. Thanks to the incorporation of RE ions into the precipitated BaF2 nanocrystals, efficient 2.0 µm fluorescence ascribed to Ho3+: 5I75I8 transition has been demonstrated upon excitation of 808 nm LD. A large ratio of forward ET constant to that of reverse process further revealed that ET from Tm3+ (3F4) to Ho3+ (5I7) is quite efficient in GC, resulted from the reduced ionic distances of Tm3+-Tm3+ and Tm3+-Ho3+ pairs and limited multiphonon relaxation of Tm3+: 3H43H5 in BaF2 nanocrystals. The results have demonstrated the promising potential of transparent oxyfluoride GC as a host for efficient Ho3+ 2.0 µm laser.

Acknowledgments

This work is financially supported by the NSFC (Grant No. U0934001). We are grateful to Dr. J. H. Zhang of Postech for lifetime measurements.

References and links

1. O. A. Louchev, Y. Urata, M. Yumoto, N. Saito, and S. Wada, “Thermo-optical modeling of high power operation of 2 μm codoped Tm, Ho solid-state lasers,” J. Appl. Phys. 104(3), 033114 ( 2008). [CrossRef]  

2. A. Taniguchi, T. Kuwayama, A. Shirakawa, M. Musha, K. Ueda, and M. Prabhu, “1212 nm pumping of 2 μm Tm-Ho-codoped silica fiber laser,” Appl. Phys. Lett. 81(20), 3723–3725 ( 2002). [CrossRef]  

3. Y. Tsang, B. Richards, D. Binks, J. Lousteau, and A. Jha, “Tm3+/Ho3+ codoped tellurite fiber laser,” Opt. Lett. 33(11), 1282–1284 ( 2008). [CrossRef]   [PubMed]  

4. C. J. Lee, G. Han, and N. P. Barnes, “Ho:Tm Lasers II: Experiments,” IEEE J. Quantum Electron. 32(1), 104–111 ( 1996). [CrossRef]  

5. R. M. Percival, D. Szebesta, S. T. Davey, N. A. Swain, and T. A. King, “Thulium sensitised holmium-doped CW fluoride fiber laser of high efficiency,” Electron. Lett. 28(24), 2231–2232 ( 1992). [CrossRef]  

6. J. Yu, A. Braud, and M. Petros, “600-mJ, double-pulse 2-μm laser,” Opt. Lett. 28(7), 540–542 ( 2003). [CrossRef]   [PubMed]  

7. J. Yu, B. C. Trieu, E. A. Modlin, U. N. Singh, M. J. Kavaya, S. Chen, Y. Bai, P. J. Petzar, and M. Petros, “1 J/pulse Q-switched 2 μm solid-state laser,” Opt. Lett. 31(4), 462–464 ( 2006). [CrossRef]   [PubMed]  

8. S. D. Jackson, “8.8W diode-cladding-pumped Tm3+, Ho3+-doped fluoride fibre laser,” Electron. Lett. 37(13), 821–822 ( 2001). [CrossRef]  

9. D. C. Tran, G. H. Sigel, and B. Bendow, “Heavy metal fluoride glasses and fibers: a review,” J. Lightwave Technol. 2(5), 566–586 ( 1984). [CrossRef]  

10. Q. Y. Zhang, T. Li, Z. H. Jiang, X. H. Ji, and S. Buddhudu, “980 nm laser-diode-excited intense blue upconversion in Tm3+/Yb3+-codoped gallate-bismuth-lead glasses,” Appl. Phys. Lett. 87(17), 171911 ( 2005). [CrossRef]  

11. Y. Wang and J. Ohwaki, “New transparent vitroceramics codoped with Er3+ and Yb3+ for efficient frequency upconversion,” Appl. Phys. Lett. 63(24), 3268–3270 ( 1993). [CrossRef]  

12. D. Chen, Y. Wang, Y. Yu, and P. Huang, “Intense ultraviolet upconversion luminescence from Tm3+/Yb3+: β-YF3 nanocrystals embedded glass ceramic,” Appl. Phys. Lett. 91(5), 051920 ( 2007). [CrossRef]  

13. X. Qiao, X. Fan, and M. Wang, “Spectroscopic properties of Er3+ doped glass ceramics containing Sr2GdF7 nanocrystals,” Appl. Phys. Lett. 89(11), 111919 ( 2006). [CrossRef]  

14. B. N. Samson, P. A. Tick, and N. F. Borrelli, “Efficient neodymium-doped glass-ceramic fiber laser and amplifier,” Opt. Lett. 26(3), 145–147 ( 2001). [CrossRef]   [PubMed]  

15. M. Mattarelli, V. K. Tikhomirov, A. B. Seddon, M. Montagna, E. Moser, A. Chiasera, S. Chaussedent, G. Nunzi Conti, S. Pelli, G. C. Righini, L. Zampedri, and M. Ferrari, “Tm3+-activated transparent oxy-fluoride glass-ceramics: structural and spectroscopic properties,” J. Non-Cryst. Solids345&346, 354–358 ( 2004). [CrossRef]  

16. S. Tanabe, T. Ohyagi, N. Soga, and T. Hanada, “Compositional dependence of Judd-Ofelt parameters of Er3+ ions in alkali-metal borate glasses,” Phys. Rev. B 46(6), 3305–3310 ( 1992). [CrossRef]  

17. X. Zou and T. Izumitani, “Spectroscopic properties and mechanisms of excited state absorption and energy transfer upconversion for Er3+-doped glasses,” J. Non-Cryst. Solids 162(1-2), 68–80 ( 1993). [CrossRef]  

18. B. F. Aull and H. P. Jenssen, “Vibronic interactions in Nd:YAG resulting in nonreciprocity of absorption and stimulated emission crosss sections,” IEEE J. Quantum Electron. 18(5), 925–930 ( 1982). [CrossRef]  

19. D. E. McCumber, “Theory of phonon-terminated optical masers,” Phys. Rev. 134(2A), A299–A306 ( 1964). [CrossRef]  

20. S. D. Jackson, A. Sabella, and D. G. Lancaster, “Application and development of high-power and highly efficient silica-based fiber lasers operating at 2 μm,” IEEE J. Sel. Top. Quantum Electron. 13(3), 567–572 ( 2007). [CrossRef]  

21. X. Zou and H. Toratani, “Spectroscopic properties and energy transfers in Tm3+ singly- and Tm3+/Ho3+doubly-doped glasses,” J. Non-Cryst. Solids 195(1-2), 113–124 ( 1996). [CrossRef]  

22. D. L. Dexter, “A theory of sensitized luminescence in solids,” J. Chem. Phys. 21(5), 836–850 ( 1953). [CrossRef]  

23. J. H. Song, J. Heo, and S. H. Park, “1.48-μm emission properties and energy transfer between Tm3+ and Ho3+/Tb3+ in Ge-Ga-As-S-CsBr glasses,” J. Appl. Phys. 97(8), 083542 ( 2005). [CrossRef]  

24. L. Huang, S. Shen, and A. Jha, “Near infrared spectroscopic investigation of Tm3+-Yb3+ co-doped tellurite glasses,” J. Non-Cryst. Solids345&346, 349–353 ( 2004). [CrossRef]  

25. B. Richards, S. Shen, A. Jha, Y. Tsang, and D. Binks, “Infrared emission and energy transfer in Tm(3+), Tm3+-Ho3+ and Tm3+-Yb3+-doped tellurite fibre,” Opt. Express 15(11), 6546–6551 ( 2007). [CrossRef]   [PubMed]  

26. L. D. da Vila, L. Gomes, L. V. G. Tarelho, S. J. L. Ribeiro, and Y. Messaddeq, “Dynamics of Tm-Ho energy transfer and deactivation of the 3F4 low level of thulium in fluorozirconate glasses,” J. Appl. Phys. 95(10), 5451–5463 ( 2004). [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 (4)

Fig. 1
Fig. 1 (a). XRD patterns of the precursor glass and GC heat-treated at 580°C for 6 h. (b) TEM bright field image and corresponding selected area electron diffraction pattern (left upper side) of the GC. (c) the size distribution of BaF2 nanocrystals embedded in glass matrix. (d) High-resolution TEM image of a BaF2 nanocrystal.
Fig. 2
Fig. 2 (a). Absorption spectra of GC samples doped with 1 mol.% Tm3+ and/or 0.6 mol.% Ho3+. (b) Absorption (solid line) and emission cross sections (dash line) corresponding to the 3H6 - 3F4 transition of Tm3+ and the 5I8 - 5I7 transition of Ho3+ in GC.
Fig. 3
Fig. 3 HoF3 concentration dependence of Tm3+ and Ho3+ fluorescence intensities in SZBTx glass samples under 808 nm excitation. The inset shows the simplified energy level diagram for Tm3+-Ho3+ ions system.
Fig. 4
Fig. 4 (a). Fluorescence spectra of glass and GC co-doped with 0.6 mol.% Ho3+ and 1 mol.% Tm3+ under 808 nm excitation. (b) Calculated gain coefficients corresponding to the 5I8 - 5I7 transition of Ho3+ for the GC sample.

Equations (3)

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

G ( λ ) = N [ p σ e ( λ ) ( 1 p ) σ a ( λ ) ]
C D A = R 0 6 τ 0
R 0 6 = 3 h 4 c 4 Q 64 π 5 n 4 g i ( E ) g j ( E ) E 4 d E
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.