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

Spectroscopic properties and energy transfers in Cr, Tm, Ho triple-doped Y3Al5O12 transparent ceramics

Open Access Open Access

Abstract

Highly transparent Cr, Tm, Ho triple-doped Y3Al5O12 (YAG) ceramics were prepared using advanced ceramic technology and their spectroscopic properties were studied for infrared laser applications. 2.09 μm emission was observed by exciting at 430 nm, which indicated the realization of energy transition from Cr3+ to Ho3+ with Tm3+ acted as an intermediate media. The efficiency between the energy transfer (Tm3+) 3F4→ (Ho3+) 5I7 and its back-transfer process was 7.87, which was comparable to that of YAG single crystal and YLF. Studies on the optical gain and stimulated emission characteristics suggested that this triple-doped YAG ceramic could be an appropriate material for 2.09 μm laser application.

© 2013 Optical Society of America

1. Introduction

Ho3+ ions lasers emitting around 2.09 μm are attractive because of extensive applications in remote sensing, military technology [1,2], and especially in medical field [3]. It is believed that Ho3+ ions laser is a perfect operation source, since it not only penetrates the tissue shallowly but also has the advantages of high precision and hemostasis effect [4]. However, laser operation in singly doped Ho3+ materials is not easy to realize due to the lack of suitable absorption feature in the 780 – 980 nm regions, which leads to the failure of being pumped directly by traditional diode lasers [5,6]. Methods to circumvent the above problem have been carried out by additionally doping Tm3+, Cr3+, Cr3+ and Tm3+, Er3+ and Tm3+, etc, into the Ho3+ doped laser materials. Among these materials, Ho3+:YAG sensitized by Cr3+ and Tm3+ simultaneously is considered as an ideal material for gain medium, due to the high energy transfer efficiency among rare earth ions, optional pumping source, and excellent physicochemical property of YAG [7]. To date, Cr, Tm, Ho triple-doped YAG single crystal has been grown [8], even so, it is still a hard work because of the serious lattice defects [9]. Besides, high cost of the equipment for the single crystal growth also restricts its mass production and commercial application. On the other hand, as the development of advanced ceramic technique [10], transparent Cr, Tm, Ho triple-doped YAG ceramics can be fabricated and have the potential to replace the single crystals. The ceramic laser materials could be much cheaper than the single crystals while their optical qualities are almost the same as (and even better than) those of single crystals [11]. Furthermore, high doping concentration and large-size samples can be easily fabricated without the adoption of complicated facilities and critical techniques. However, so far, there have been no efforts reported on the Cr:Tm:Ho:YAG laser ceramic, though other ceramic laser materials, such as Nd:YAG [12,13], Yb:YAG [14,15], Tm:YAG [16,17] and co-doped Tm:Ho:YAG [10], Yb:Er:YAG [18] etc, have been intensively investigated

In this paper, we report on the fabrication of Cr:Tm:Ho:YAG transparent ceramic by advanced ceramic technology, and study the spectroscopic properties of the host with the interest in infrared laser application at 2.09 μm.

2. Experimental details

High-purity Al2O3 (99.99%), Y2O3 (99.99%), Cr2O3 (99.999%), Tm2O3 (99.99%) and Ho2O3 (99.99%) commercial powders were used as starting materials (Fig. 1) and weighed according to the designed nominal dopant concentrations of 0.56 at% for Cr3+, 5.80 at% for Tm3+, and 0.36 at% for Ho3+. Meanwhile, 0.45 wt% tetraethoxysilane (purity > 99.99%) was used as a sintering aid and 0.25 wt% Oletic Acid (purity > 99%) was chosen as a dispersant. Then the chemicals were mixed thoroughly in ethanol for 24 h using a planetary-milling machine to prepare ceramic slurries. After drying at 65 °C for 48 h, the mixtures were sieved through 300-mesh screen and uniaxially pressed into φ10 mm disks. The green bodies were pre-sintered in the air at 800 °C for 10 h to remove the organic ingredients before cold isostatic pressed (CIP) under 200 Mpa. The as-obtained plates were sintered at 1703 °C under vacuum condition of 10−5 Pa for 15 h, accompanied by annealing at 1400 °C in air and polishing to optical grade..

 figure: Fig. 1

Fig. 1 SEM images of the starting oxide powders. (a) Cr2O3; (b) Tm2O3; (c) Ho2O3; (d) Y2O3; (e) Al2O3; (f) mixed powders after ball milling for 24 h.

Download Full Size | PDF

The structural properties and phase purity were tested by X-ray diffraction (XRD, Dmax2500, Rigaku), while the microstructure was investigated by scanning electron microscopy (SEM, JSM-6700F, JEOL). The optical transmittance and absorption spectra were measured on a Perkin-Elmer UV-visible-near infrared (NIR) spectrometer (Lambda-900). The fluorescence spectra and decay curves were measured by an Edinburgh FSP920 spectrometer equipped with an OPO laser as the excitation source.

3. Results and discussion

Figures 1(a)-1(e) display the SEM images of the starting powders (i.e., Cr2O3, Tm2O3, Ho2O3, Y2O3, α-Al2O3, respectively). All powders have a high purity level, but different morphologies. The Al2O3 powders employed in our experiment have a regular morphology and uniform size, whereas Cr2O3, Tm2O3, Ho2O3 and Y2O3 powders are quite coarse, which are generally considered as unsuitable for the fabrication of transparent ceramics. However, a SEM image shown in Fig. 1(f) suggests that the modified morphology and a fine particle size distribution have been obtained by a sufficient ball-milling process. This is beneficial to obtain a pore-free microstructure during sintering, especially for the ceramics doped with several rare earth ions.

Figure 2 shows the XRD pattern of the Cr:Tm:Ho:YAG ceramic, which is well indexable under Ia-3d lattice symmetry and is consistent with the standard JCPDF file [No. 79-1891 for Y3Al5O12 (YAG)]. Figure 3(a) and its inset display the transmittance spectrum and the photograph of the Cr:Tm:Ho:YAG ceramic with diameter of 8 mm and thickness of 1.2 mm, respectively. This sample shows good transparency, with the in-line transmittances in both the visible and infrared region are almost 81%. The transmittance reaches 81.4% at 2.09 nm, which is helpful to the 2.09 μm laser emission. The SEM surface morphology of the ceramic is shown in Fig. 3(b). Obviously, the density is quite high, and neither pores nor second phases can be observed. The average grain size is 15 μm, and no abnormal grain growth arises.

 figure: Fig. 2

Fig. 2 The XRD pattern of the Cr, Tm, Ho triple-doped YAG ceramic.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 (a) Transmittance spectrum and (b) SEM image of the transparent ceramic. The inset in (a) is the photograph of the sample.

Download Full Size | PDF

The room temperature ground state absorption spectrum of the sample was shown in Fig. 4.The spectrum is well resolved with almost all the Stark components corresponding to different manifold of Cr3+, Tm3+ and Ho3+ are observed. The absorption band positions of Tm3+ and Ho3+ are well matched with the spectral data of Tm3+, Ho3+ co-doped YAG ceramic [10]. The inset of Fig. 4 shows the detail of absorption positions ranging from 372 nm to 850 nm, from which four typical transitions for Cr3+ and Tm3+ are indicated. In the range of 372-650 nm, two absorption bands around 430 nm (4A24T1) and 600 nm(4A24T2)are of the absorption levels of Cr3+ ions, and the calculated absorption coefficients are 3.46 cm−1 and 1.82 cm−1, respectively. The full widths at half maximum (FWHM) of these two absorption bands are very broad (69 nm and 80 nm, respectively), which are consistent with the Cr:Tm:Ho:YAG single crystal and beneficial to the flash lamp pumping [19]. Other two intense absorption bands around 680 nm (3H63F2, 3) and 782 nm (3H63H4) are attributed to the transitions of Tm3+. Due to the lack of appropriate pumping source at 680 nm, the absorption band around 782 nm is the main location for discussion. This absorption has a peak value of 3.68 cm−1 with FWHM of 15 nm, which matches well with the emitting wavelength of high-power AlGaAs laser diodes. Obviously, this Cr, Tm, Ho triple-doped YAG ceramic has the potential advantage of optional pumping source, flash lamp or diode.

 figure: Fig. 4

Fig. 4 Optical absorption spectrum of the mirror polished Cr:Tm:Ho:YAG ceramic. The detail of absorption positions range from 372 nm to 850 nm is shown in the inset.

Download Full Size | PDF

The room temperature fluorescence spectrum of the sample in the range of 1700-2200 nm, obtained with excitation under an OPO laser at 430 nm, is shown in Fig. 5. It consists of 6 main lines range from 2000 to 2200 nm, likely arising from the transition 5I75I8 and the maximum emission is at 2.088 μm with FWHM of 40 nm. This emission decays exponentially with the decay time of 8.93 ms (standard error of 0.09 ms, inset of Fig. 5), which is in good agreement with the literature data [20]. The acquisition of fluorescence at 2.088 μm indicates that the energy transfer from Cr3+ to Ho3+ is realized with Tm3+ acted as an intermediate media. The energy transfer mechanism is schematized in Fig. 6.Energy can be absorbed by Cr3+ in either the 4T1 or 4T2 bands and relaxes to the 2E level with a high quantum efficiency, since Cr is a transition metal atom with overlapping bands [20]. Then, energy transfers from 2E level of Cr3+ to 3F3 and 3H4 levels of Tm3+ (where 3F3 decays nonradiatively to 3H4). At the same time, the cross-relaxation interaction between excited and ground states of Tm3+ gives rise to two Tm3+ ions in the 3F4 excited state according to the schematic process. Subsequently, the Tm atoms in the 3F4 level transfer the energy to the upper laser level (5I7) of Ho3+. This energy transfer process is consistent with Cr:Tm:Ho:YAG single crystal [7], indicating the feasibility of fabricating Cr:Tm:Ho:YAG materials for infrared laser application by ceramic technology.

 figure: Fig. 5

Fig. 5 Emission spectrum of Cr:Tm:Ho:YAG ceramic obtained by exciting the Cr3+ at 430 nm. The decay profile of this emission is shown in the inset.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Scheme of the energy transfer processes.

Download Full Size | PDF

The efficiencies of the energy transfer (Tm3+) 3F4→ (Ho3+) 5I7 and its back-transfer process are important parameters in this co-doped system [21], which deeply affect the laser performance at 2.09 μm. The values of these transfer coefficients can be calculated from the emission and absorption cross sections by means of the Forster- Dexter theory [22]:

CTmHo=9χ2c16π4n2σabs,Ho(λ)σem,Tm(λ)dλ,
CHoTm=CTmHoexp(Ezl:HoEzl:TmkT),
where χ2 includes the orientational average and is usually taken to be 2/3, c is the speed of light, n is the refractive index, σem,Χand σabs,Χare the emission and absorption cross sections of X ion, respectively, and Ezl:X is the zero phonon line of ion X. We define Ezl:X as the energy of the first absorption peak in the absorption spectra and obtain Ezl:Tm = 5875 cm−1 and Ezl:Ho = 5449 cm−1. To calculate the energy transfer parameters, the measured emission spectra of Tm3+ 3F4 manifold and the absorption spectra of Ho3+ 5I7 manifold are introduced, as shown in Fig. 7.The values obtained for the Tm - Ho and Ho - Tm transfer coefficients are CTm→Ho = 6.61 × 10−40 cm6 s−1 and CHo→Tm = 0.84 × 10−40 cm6 s−1. The main uncertainty of the results is from σem,Tm(λ) using the Forster- Dexter theory, which is calculated by the reciprocity method [23]:
σEM(λ)=A(JJ)λ5I(λ)8πn2cI(λ)λdλ,
where A(JJ) is the averaged spontaneous emission probability, I(λ) indicates the fluorescence intensity, n and c are the refractive index and velocity of light. Among these values, A(JJ) is calculated based on the Judd-Ofelt (J-O) theory, which introduces the room-mean-square (RMS) deviation between the calculated and experimental results. Therefore, the uncertainty of the results can be simply interpreted as the RMS deviation. Due to the value of RMS deviation is always very small, the uncertainty of the results is in the acceptable range.

 figure: Fig. 7

Fig. 7 YAG-based Emission cross section of Tm3+ 3F4 manifold and Absorption cross section of Ho3+ 5I7 manifold relevant to energy transfer parameters.

Download Full Size | PDF

The values of the transfer coefficients are reported in Table 1 together with the values calculated with the same method for other hosts. The table shows that our values are similar to those obtained from YAG single crystal [24], a garnet that has already shown laser effect in the 2 μm region when co-doped with Tm3+ and Ho3+ [25]. Furthermore, the ratio between the direct Tm - Ho energy transfer and its back-process for Cr:Tm:Ho:YAG ceramic is 7.87, slightly lower than that of YAG single crystal, but still higher than that of YLF [26]. It is worth noting that a high value of the ratio between the Tm - Ho transfer and its back-process is an important condition for an efficient Ho 2 μm laser [21]. Therefore, this Cr:Tm:Ho:YAG ceramic is a promising material for the realization of efficient diode pumped laser devices.

Tables Icon

Table 1. Comparison of the transfer coefficient between Tm3+ and Ho3+ in our ceramic and in other laser hosts.

In order to obtain more reliable information about the perspective of application of our material, we have calculated the emission cross section by means of the McCumber theory [27]:

σe(λ)=σa(λ)ZlZuexp[εEkT]
where σais the absorption cross-section which can be calculated byα/N(αis the absorption coefficient shown in Fig. 4 and N is the ionic concentration); Zl and Zu are the partition functions for the lower and the upper levels, respectively, involved in the considered optical transition; T is the temperature (here is the room temperature); k is the Boltzmann constant, and ε is the energy for the transition between the lower Stark sublevels of the emitting multiplets and the lower Stark sublevels of the receiving multiplets (zero-phonon line); E is the energy corresponding to the peak wavelength of the absorption. The result is shown in Fig. 8(a) with the relevant absorption cross section at this region. The peak cross section is 7.4 × 10−21 cm2 at 2088 nm, rather similar to that reported by Fan et al [28]. In addition, the spectral overlap between the absorption and emission cross section indicates that the infrared laser could only be realized within 2000-2200 nm.

 figure: Fig. 8

Fig. 8 (a) Absorption and emission cross section spectra of Ho3+ in YAG ceramic corresponding to the transition from 5I7 to 5I8 level. (b) Simulated optical gain spectra for this transition under various population inversion (P) conditions.

Download Full Size | PDF

A meaningful assessment about the potentiality as laser medium can be obtained from the curves of the potential laser gain versus wavelength, which are calculated according to the formula [29]:

G(λ)=N[Pσe(λ)(1P)σa(λ)]
where P is the population of the emitting level. The gain spectra for P ranging from 0 to 1.0 in the 1800 – 2200 nm spectra regions are shown in Fig. 8(b). The gain increases with the population inversion and positive gain in the system can be achieved for a population inversion of 40% and above. At 100% population inversion, a maximum gain coefficient of 6.66 cm−1 is obtained. According to the laser principle, laser can be realized whenG(λ)0. Therefore, this kind of Cr:Tm:Ho:YAG ceramic is suitable for laser gain medium.

Preliminary laser test was carried out to verify the possibility of Cr:Tm:Ho:YAG ceramic used as laser gain medium. Laser operation at 2.09 μm had been acquired with sparks observed when the laser hit a seared wood. It is believed that better laser performance would be realized by the improvement of ceramic fabrication details and optimization laser experiments in the further study.

4. Conclusions

Cr, Tm, Ho triple-doped Y3Al5O12 laser ceramic, which was fabricated by the one step vacuum sintering method, displays excellent quality and optical characteristics comparable to the single crystal. The 5I75I8 fluorescence at 2.09 μm indicates an excellent infrared emission under flash lamp or diode pumping with maximum emission cross section of 7.4 × 10−21 cm2 and optical gain coefficient of 6.66 cm−1. The efficiencies of the energy transfer (Tm3+) 3F4 → (Ho3+) 5I7 and its back-transfer process were also calculated. The obtained ratio between the Tm – Ho transfer and its back-process was 7.87, which was comparable to that of YAG and YLF. It is therefore favorable to consider this material as a potential host to make flash lamp or diode pumped solid state laser operating at 2.09 μm.

Acknowledgments

This work was financially supported by the Key Program in Major Research Plan of National Natural Science Foundation of China (91022035).

References and links

1. G. Rustad and K. Stenersen, “Low threshold laser-diode side-pumped Tm:YAG and Tm:Ho:YAG lasers,” IEEE J. Sel. Top. Quant. 3(1), 82–89 (1997). [CrossRef]  

2. A. S. Kurkov, V. V. Dvoyrin, and A. V. Marakulin, “All-fiber 10 W holmium lasers pumped at lambda=1.15 microm,” Opt. Lett. 35(4), 490–492 (2010). [CrossRef]   [PubMed]  

3. T. R. W. Herrmann, E. N. Liatsikos, U. Nagele, O. Traxer, A. S. Merseburger, and EAU Guidelines Panel on Lasers, Technologies, “EAU guidelines on laser technologies,” Eur. Urol. 61(4), 783–795 (2012). [CrossRef]   [PubMed]  

4. R. M. Kuntz, “Current role of lasers in the treatment of benign prostatic hyperplasia (BPH),” Eur. Urol. 49(6), 961–969 (2006). [CrossRef]   [PubMed]  

5. W. X. Zhang, J. Zhou, W. B. Liu, J. Li, L. Wang, B. X. Jiang, Y. B. Pan, X. J. Cheng, and J. Q. Xu, “Fabrication, properties and laser performance of Ho:YAG transparent ceramic,” J. Alloy. Comp. 506(2), 745–748 (2010). [CrossRef]  

6. H. Chen, D. Y. Shen, J. Zhang, H. Yang, D. Y. Tang, T. Zhao, and X. F. Yang, “In-band pumped highly efficient Ho:YAG ceramic laser with 21 W output power at 2097 nm,” Opt. Lett. 36(9), 1575–1577 (2011). [CrossRef]   [PubMed]  

7. Y. Kalisky, J. Kagan, D. Sagie, A. Brenier, C. Pedrini, and G. Boulon, “Spectroscopic properties, energy-transfer, and laser operation of pulsed holmium lasers,” J. Appl. Phys. 70(8), 4095–4100 (1991). [CrossRef]  

8. K. S. Lim, C. W. Lee, S. T. Kim, H. J. Seo, and C. D. Kim, “Infrared to visible up-conversion in Cr:Tm:Ho:YAG,” J. Lumin. 87-89(9), 1008–1010 (2000). [CrossRef]  

9. C. Zaldo, M. J. Martin, R. Sole, M. Aguilo, F. Diaz, P. Roura, and M. Lopez de Miguel, “Optical spectroscopy of Ho3+ and Tm3+ ions in KTiOPO4 single crystals,” Opt. Mater. 10(1), 29–37 (1998). [CrossRef]  

10. G. A. Kumar, M. Pokhrel, D. K. Sardar, P. Samuel, K. I. Ueda, T. Yanagitani, and H. Yagi, “2.1 mum emission spectral properties of Tm and Ho doped transparent YAG ceramic,” Sci. Adv. Mater. 4(5), 617–622 (2012). [CrossRef]  

11. J. Kong, D. Y. Tang, B. Zhao, J. Lu, K. Ueda, H. Yagi, and T. Yanagitani, “9.2-W diode-end-pumped Yb:Y2O3 ceramic laser,” Appl. Phys. Lett. 86(16), 161116 (2005). [CrossRef]  

12. A. Ikesue and Y. L. Aung, “Ceramic laser materials,” Nat. Photonics 2(12), 721–727 (2008). [CrossRef]  

13. S. H. Lee, S. Kochawattana, G. L. Messing, J. Q. Dumm, G. Quarles, and V. Castillo, “Solid-state reactive sintering of transparent polycrystalline Nd:YAG ceramics,” J. Am. Ceram. Soc. 89(6), 1945–1950 (2006). [CrossRef]  

14. J. Dong, A. Shirakawa, K. I. Ueda, H. Yagi, T. Yanagitani, and A. A. Kaminskii, “Efficient Yb3+: Y3Al5O12 ceramic microchip lasers,” Appl. Phys. Lett. 89(9), 091114 (2006). [CrossRef]  

15. F. Tang, J. Q. Huang, W. Guo, W. C. Wang, B. J. Fei, and Y. G. Cao, “Photoluminescence and laser behavior of Yb:YAG ceramic,” Opt. Mater. 34(5), 757–760 (2012). [CrossRef]  

16. W. L. Gao, J. Ma, G. Q. Xie, J. Zhang, D. W. Luo, H. Yang, D. Y. Tang, J. Ma, P. Yuan, and L. J. Qian, “Highly efficient 2 μm Tm:YAG ceramic laser,” Opt. Lett. 37(6), 1076–1078 (2012). [CrossRef]   [PubMed]  

17. B. J. Fei, J. Q. Huang, W. Guo, Q. F. Huang, J. Chen, F. Tang, W. C. Wang, and Y. G. Cao, “Spectroscopic properties and laser performance of Tm:YAG ceramics,” J. Lumin. 142, 189–195 (2013). [CrossRef]  

18. J. Zhou, W. X. Zhang, T. D. Huang, L. A. Wang, J. Li, W. B. Liu, B. X. Jiang, Y. B. Pan, and J. K. Guo, “Optical properties of Er, Yb co-doped YAG transparent ceramics,” Ceram. Int. 37(2), 513–519 (2011). [CrossRef]  

19. N. Karadimitriou, B. Klinkenberg, D. N. Papadopoulos, and A. A. Serafetinides, “Development and performance characteristics of flash lamp pumped Yb:YAG, Cr:Tm:Ho:YAG, Er:Tm:Ho:YLF laser sources and investigation of their potential biological applications,” SPIE-OSA 6633, 66331H, 66331H-7 (2007). [CrossRef]  

20. N. P. Barnes, K. E. Murray, and M. G. Jani, “Flash-lamp-pumped Ho:Tm:Cr:YAG and Ho:Tm:Er:YLF lasers: modeling of a single, long pulse length comparison,” Appl. Opt. 36(15), 3363–3374 (1997). [CrossRef]   [PubMed]  

21. S. Bigotta, A. Toncelli, M. Tonelli, E. Cavalli, and E. Bovero, “Spectroscopy and energy transfer parameters of Tm3+- and Ho3+-doped Ba2NaNb5O15 single crystals,” Opt. Mater. 30(1), 129–131 (2007). [CrossRef]  

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

23. Z. D. Luo, Y. D. Huang, and X. Y. Chen, Spectroscopy of Solid-State Laser and Luminescent Materials (Nova Science Publishers, 2007).

24. B. M. Walsh, N. P. Barnes, and B. P. Bartolo, “On the distribution of energy between the Tm 3F4 and Ho 5I7 manifolds in Tm-sensitized Ho luminescence,” J. Lumin. 75(2), 89–98 (1997). [CrossRef]  

25. K. J. Yang, H. Bromberger, H. Ruf, H. Schäfer, J. Neuhaus, T. Dekorsy, C. V. B. Grimm, M. Helm, K. Biermann, and H. Künzel, “Passively mode-locked Tm,Ho:YAG laser at 2 microm based on saturable absorption of intersubband transitions in quantum wells,” Opt. Express 18(7), 6537–6544 (2010). [CrossRef]   [PubMed]  

26. S. A. Payne, L. K. Smith, W. L. Kway, J. B. Tassano, and W. F. Krupke, “The mechanism of Tm to Ho energy transfer in LiYF4,” J. Phys. Condens. Matter 4(44), 8525–8542 (1992). [CrossRef]  

27. D. E. McCumber, “Theory of Phonon-Terminated Optical Masers,” Phys. Rev. 134(2A), A299–A306 (1964). [CrossRef]  

28. T. Y. Fan, G. Huber, R. L. Byer, and P. Mitzscherlich, “Spectroscopy and diode laser-pumped operation of Tm, Ho-Yag,” IEEE J. Quantum Electron. 24(6), 924–933 (1988). [CrossRef]  

29. X. L. 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]  

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 (8)

Fig. 1
Fig. 1 SEM images of the starting oxide powders. (a) Cr2O3; (b) Tm2O3; (c) Ho2O3; (d) Y2O3; (e) Al2O3; (f) mixed powders after ball milling for 24 h.
Fig. 2
Fig. 2 The XRD pattern of the Cr, Tm, Ho triple-doped YAG ceramic.
Fig. 3
Fig. 3 (a) Transmittance spectrum and (b) SEM image of the transparent ceramic. The inset in (a) is the photograph of the sample.
Fig. 4
Fig. 4 Optical absorption spectrum of the mirror polished Cr:Tm:Ho:YAG ceramic. The detail of absorption positions range from 372 nm to 850 nm is shown in the inset.
Fig. 5
Fig. 5 Emission spectrum of Cr:Tm:Ho:YAG ceramic obtained by exciting the Cr3+ at 430 nm. The decay profile of this emission is shown in the inset.
Fig. 6
Fig. 6 Scheme of the energy transfer processes.
Fig. 7
Fig. 7 YAG-based Emission cross section of Tm3+ 3F4 manifold and Absorption cross section of Ho3+ 5I7 manifold relevant to energy transfer parameters.
Fig. 8
Fig. 8 (a) Absorption and emission cross section spectra of Ho3+ in YAG ceramic corresponding to the transition from 5I7 to 5I8 level. (b) Simulated optical gain spectra for this transition under various population inversion (P) conditions.

Tables (1)

Tables Icon

Table 1 Comparison of the transfer coefficient between Tm3+ and Ho3+ in our ceramic and in other laser hosts.

Equations (5)

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

C T m H o = 9 χ 2 c 16 π 4 n 2 σ a b s , H o ( λ ) σ e m , T m ( λ ) d λ ,
C H o T m = C T m H o exp ( E z l : H o E z l : T m k T ) ,
σ E M ( λ ) = A ( J J ) λ 5 I ( λ ) 8 π n 2 c I ( λ ) λ d λ ,
σ e ( λ ) = σ a ( λ ) Z l Z u exp [ ε E k T ]
G(λ)=N[ P σ e ( λ )(1P) σ a (λ) ]
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.