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

Spectroscopy of thulium and holmium co-doped silicate glasses

Open Access Open Access

Abstract

In this study, the spectroscopic properties of Tm3+/Ho3+ co-doped silicate glasses under an 808 nm diode laser excitation are reported to discover their potential laser performance. To confirm the best candidates for glass fiber drawing, the optimal ratio of Tm3+ and Ho3+ is 1: 0.3. We calculate and discuss the J-O parameters (Ωt), the transition probability (A) of the transition from 5I7 to 5I8 is 129.89s−1 and the calculated lifetime (τrad) is 7.70 ms, respectively. The maximum emission cross section of the transition from Ho: 5I75I8 is 7.59 × 10−21 cm2 at 2065 nm as well as the gain coefficient of the STH glasses is discussed. The energy transfer between Tm3+ and Ho3+ plays an important role in the luminescence process. The sample doped with 1 mol% Tm2O3 and 0.3 mol% Ho2O3 presents a broad band spectrum with a full width at half-maximum of 189 nm, the transfer efficiency from Tm3+ to Ho3+ is 0.6702. Energy transfer constant is 63.2 × 10−40 cm6 /s and the measured fluorescence lifetimes of the sample is 0.637 ms, and the ΔT is 167 °C.These values indicate that the Tm3+/Ho3+ co-doped silicate is a promising way to achieve 2 μm laser emissions.

© 2016 Optical Society of America

1. Introduction

In order to achieve fiber lasers emitting in the near infrared region (NIR), numerous of researchers focus their interests on the emission and interaction features of rare earth (RE) ions such as Er3+, Nd3+, Pr3+, Ho3+, Tm3+, and Yb3+. Among these ions, Ho3+ has been widely used in medicine, range monitoring, and sensing [1–3] based on the energy level transition 5I75I8 lying in the 1.95-2.15 μm wavelength range. Nevertheless, exciting this transition directly is infeasible owing to the mismatching of absorption peak with the wavelength of available commercial diode lasers. Thus sensitized by other ions (Tm3+, Yb3+ et al.) is an alternative way to achieve 2 μm emissions [4, 5]. In Tm3+/Ho3+ co-doped system, pumping the photon of Tm3+ at ~790 nm to 3H4 level offers highly efficient energy transfer (ET) to activator ions and several transition processes [6]. The cross-relaxation process is conducive to the ions gather of level 3F4, a portion of energy of 3F4 can be transfer to neighboring Ho3+, then according to the 5I75I8 transition, emit the ~2050 nm laser [7]. Not only traditional rare earth ions but also other activators are excited to emit in mid infrared. For activators such as Bi+ or Cr+, although they are able to achieve ultra-broadband luminescence in 1000-1700 nm [8], they are sensitive to coordination environment. When coordination number is changed, the valence state of luminescence center is affected [9]. In Bi+ doped glass, the Bi+ cannot be dispersed to the glass structure uniformly, which will induced cluster or quenching in glass, which is harmful to luminescence [10]. Besides, the Bi+ fluorescence peak is located in 1300, 1330, 1470 nm, which is different with the peak of Ho3+ and the energy transfer process is not clearly, the most doping concentration is lower than Ho3+, it is difficult to achieve high power laser emission, which limited the further development of Bi+. For glass fiber, the laser efficiency is decreased obviously with the rising of temperature [11]. Thus, the Ho3+ doped glass sensitized by other ions is still a promising candidate to realize 2 μm laser output. The difficulties to emit earth doped fiber laser are selecting the optimal population, high efficient pumping source and reduction of depletion processes (energy transfer up-conversion from the upper energy level and back energy transfer processes).

Several studies related to Tm3+/Ho3+ co-doped fiber lasers in silica or heavy metal glass hosts have been reported in recent years. 2 μm Tm3+/Ho3+ co-doped silica fiber laser have achieved [12]. Enhanced 2 μm emission of bismuth germanate glasses doped with Ho3+/Tm3+ ions are published in 2015 [13]. Besides, Tm3+/Ho3+ co-doped are applied to α-NaYF4 single crystals in order to produce 2 μm emissions [14]. Compare with these matrixes, silicate glasses are able to dissolve much higher concentration of rare earth ions, which is helpful to efficient cross relaxation energy transfer, improving the quantum efficiency and reducing device length [15], own better ability of anti-crystallization, higher refractive index, and relatively low thermal expansion coefficient [16]. Besides, the maximum phonon energy of silicate (~1050 cm−1) is much lower compared to those of borate glasses (~1350 cm−1) and phosphate glasses (~1300 cm−1) [15], so quantum efficiency can be less influenced by the multiphonon relaxation process. Thus silicate glass possesses incomparable advantages in glass fiber field. Recently, the luminescence properties of Er3+/Yb3+ silicate glasses [17] is demonstrated and some researchers have interest in lithium iron silicate [18]. Emission of Co2+ ions with Bi3+ ions in lead silicate glasses has been discussed [19].

From all we know, there is few studies focus on the silicate glass co-doped with Tm3+/Ho3+. In our previous work [20], we reported highly Ho3+/Yb3+ co-doped silicate glasses and discussed the optimal dopant concentration for ~2 μm lasers, the results shown that it is a potential way to achieve ~2 μm laser. In this study, we extend the previous work by introducing Tm3+ in the same silicate glass host to investigate the spectroscopy properties of Ho3+ and Tm3+.

2. Experimental techniques

2.1 Material synthesis

In this study, the silicate glass samples are composed (in mol %) of (99-x) (SiO2-Al2O3-CaO-BaO- BaF2- La2O3)-1Tm2O3-xHo2O3 (where x = 0.1, 0.2, 0.3, 0.4), with denoted as STH-1, STH-2, STH-3, STH-4, respectively. The mixed row powders are melted at 1400°C for 1h, and then poured onto a preheated (750°C) steel plate, and annealed for 4h at 600°C. Annealed glasses are fabricated and polished to 10 × 10 × 1 mm3 sizes for optical property measurements.

2.2. Measurement

According to the Archimedes principle, the densities of STH-3 (3.78 g/cm3) had been measured using distilled water as immersion liquid. The absorption spectrum had been obtained by JASCOV 570UV/VIS spectrophotometer. The emission spectrum had been tested by pumping the sample with 808 nm laser diode. The lifetime had been determined by a combined fluorescence lifetime and steady state spectrometer (FLSP-920) (Edingburg Co.). All measurements were at room temperature.

3. Experimental results and discussion

3.1 Absorption spectrum and Judd-Ofelt analysis

The absorption spectrum of Tm3+, Ho3+ singly doped and co-doped glass samples in the wavelength region of 370-2000 nm are shown in Fig. 1. The absorption bands of Tm3+/Ho3+ co-doped sample corresponding to the transitions starting from the 5I8 ground state to the higher levels 5I7, 5I6, 5F5, (5F4, 5S2), 5F3, (5G6, 5F1), the absorption peak of these energy levels are centered at the wavelengths of 1932, 1140, 642, 540, 452, and 420 nm, respectively. Contrasting the absorption curve in Fig. 1, it is worth noting that for Tm3+ singly and Tm3+/Ho3+ co-doped doped silicate glass, the shape and peak positions of each transition are extremely similar to glasses with other matrix doped Tm3+ and Ho3+ [21, 22]. Besides, although Ho3+ ions do not have any absorption bands near 808 nm or 980 nm, Tm3+ has a strong absorption band around 808 nm, which can be excited by commercial laser diode. At the same time, the absorption band of 3F4 (Tm3+) is close to that of 5I7 (Ho3+), means that it is likely to exist efficient energy transfer between them. It is obvious that for different concentrations rate of Tm3+/Ho3+ co-doped glass samples, there is no obvious distinguish in the position of the absorption peaks, only the intensity grows stronger with Ho3+ concentration increasing.

 figure: Fig. 1

Fig. 1 Absorption spectrum of silicate glass doped Tm3+, Ho3+ singly and Tm3+/Ho3+ co-doped.

Download Full Size | PDF

The Judd-Ofelt theory [23, 24] is comprehensively accepted to evaluate the spectroscopic properties of rare-earth doped glasses. Based on the absorption spectra, the Judd-Ofelt parameters represent the local environment and bond property of rare-earth ions. The parameters Ω2, Ω4, and Ω6 of Ho3+ in STH-3 are 4.87 × 10−20, 2.05 × 10−20, and 1.39 × 10−20 cm2, respectively, and the comparison with other glasses is listed in Table 1. In this table, we can find that the ZBLAN glass host possesses comparatively small Ω2, where as germanate, tellurite, and silicate glasses show relatively the larger ones, because Ω2 is relative to covalent bonding in micromechanics, heavily metal and oxide glasses have a stronger covalent bonding than that of fluoride glass. At the same time, Ω2 of Ho3+ in STH-3 is larger than in a previously reported silicate glass [15], which is due to the variation in the local environment of Ho3+ in different glasses. Diverse kinds of modifying elements tend to be distributed in modifier regions, thus the modifier ions affect Judd-Ofelt parameters directly. There are no alkali ions in this silicate glasses component, the distortion around Ho3+ ion is larger than that with alkali ions, that is the reason STH-3 owns a larger Ω2 [18].

Tables Icon

Table 1. Judd-Ofelt parameters Ωt of Ho3+ in various glasses

The transition probability A and radiative lifetime τrad can be calculated from Judd-Ofelt parameters. These two parameters can be determined by the following expressions:

A=1τrad=64π4e23h(2J+1)λ3[n(n2+1)29SED+n2SMD]
SED=λ=2,4,6Ωλ|<S,L,JU(λ)S',L',J'>|2
SMD=(ħ2mc)|<S,L,JL+2SS',L',J'>|2
In above formula, n is the refractive index (Here is 1.649 in STH-3 glass sample), m and e are the mass and charge of the electron, J is the total angular momentum at the ground state, |U(λ)| is the reduced matrix element in sensitive to the host environment, and SED and SMD are the line strengths for electric and magnetic dipole transitions, respectively. The transition probability (A) of the transition from 5I7 to 5I8 is 129.89s−1 and the calculated lifetime (τrad) of STH-3 is 7.70 ms, respectively. These values are larger than those in two other previously reported silicate glasses (71.64s−1 [15] and 61.65s−1 [22]). Higher transition probability provides a better opportunity for lasers, thus we infer STH glasses maybe show better emission performances.

3.2 Fluorescence spectra and hydroxyl analysis

Fluorescence spectra of the STH samples are measured from 1600 nm to 2200 nm and shown in Fig. 2. In this figure, the emission peaks of Tm3+ and Ho3+ are located around 1838 and 2018 nm, respectively. With the increasing of Ho3+, the emission of Ho3+ from 5I7 is enhanced, which indicates existence effective ET from Tm: 3F4 to Ho: 5I7 in glass samples. However, when the concentration of Ho3+ as high as 0.4 mol%, it appears the decline of the fluorescence intensity, which led by concentration quenching. STH-3 has the maximum emission intensity of Ho3+: 5I75I8. That is the reason why we choose the STH-3 as the object of study. A broad band spectrum with a full width at half-maximum (FWHM) of 189 nm are presented, which is little smaller than that in tellurite glass [26].

 figure: Fig. 2

Fig. 2 the fluorescence spectra of STH glass.

Download Full Size | PDF

The factors that affect the emission of silicate glass in MIR are mainly concentration quench and hydroxyl quench. The concentration quench have occurred measured by fluorescence spectrum. Microscopically, when the critical distances of rare earth ions are broken, it will happen cascade effect, energy are transferred from one center to another, emission is weaken. That is the reason why the fluorescence intensity of STH-4 is lower than that of STH-3. Besides, OH- is the mainly quench center in laser glass, the vibration of OH- groups (2450-3700 cm−1) matches with the maximum phonon energy of silicate glass (~1080 cm−1), only two or three phonon can weak the luminance of rare earth ions. However, with conventional melting processes, the glasses containing little OH- groups cannot be prepared without an elaborate setup. A feasible way of removing the OH- groups from the glass during melting process is to add a fluoride to liberate hydrogen species from the melt. The added fluorine ions will react with OH- groups according to [24]:

OH+F=O2+HF

Figure 3 is the transmittance spectra when alkali earth metal oxide is replaced by fluoride. It is obvious that with the increasing of F-, the content of OH- is decreased sharply. The content of OH- groups in the glass can be evaluated by the absorption coefficient αOH- of the OH- vibration band at 3 μm, which can be given by αOH- = ln(T0/T)/l where l is the thickness of the sample equal to 0.15 cm, and the T0 and T are the incident and transmitted transmittance intensities, respectively. The value of αOH- is 1.401, 1.025, 0.545, it can be demonstrated by the effect of introduced fluorine ion on the structure of glass: a small amount introduction of fluorine ions will destroy the stable structure of silicate glass and the equilibrium between fluorine and oxygen ions were achieve with the increase of fluorine ions. Then, the glasses’ structure tends to be much more stable and homogeneous, which is registered as better thermal stability and transmission. The introduced fluorine ions can eliminate the OH- content and the BaF2/CaF2 sample possesses the lowest αOH- in all samples. The excellent properties hints that the STH glass matrix might possesses better emission properties especially 2 μm emission.

 figure: Fig. 3

Fig. 3 The transmittance spectrum of silicate glass samples with different fluoride content.

Download Full Size | PDF

3.3 Cross-sections and gain coefficient of Ho3+

To evaluate the possibility of achieving laser effects, the simulated emission cross-section (σem) defined as the intensity gain of a laser beam per unit of population inversion is calculated. Based on the obtained fluorescence spectra, the emission cross-section can be calculated using the Füchtbauer-Ladenburg formula [28, 29]:

σem(λ)=λ4Arad8πcn2λI(λ)λI(λ)
Where λ is the wavelength of the measured fluorescence spectra, Arad is the transition probability calculated by the Judd-Ofelt theory (here is 129.89 s−1), I (λ) is the fluorescence intensity, n and c are the light speed and the index of refraction, respectively. The absorption cross-section (σabs) can be calculated according to the Beer-Lambert theory and displayed as [28, 30]:
σa=2.303log(I0I)Nl
Where I0 (λ) and I (λ) are the incident and outgoing light intensities of the samples respectively, l is the sample thickness and N is the rare-earth ion concentration. Absorption and emission cross-sections of transitions from Tm3+: 3F4 to 3H6 and from Ho3+: 5I7 to 5I8 are calculated and shown in the same figure (Fig. 4). The maximum emission cross section of the transition from Ho: 5I7 to 5I8 is 7.59 × 10−21 cm2 at 2069 nm. This value is much larger than the reported value (3.07 × 10−21 cm2) in silicate glass [15], and slightly larger than that (7.00 × 10−21 cm2) in another glass [22]. Energy transfer process also can be evaluated by calculating the absorption and emission cross sections. From Fig. 4, we notice there is a large overlap between the cross sections of Tm3+ emission and Ho3+ absorption, the bigger overlap area, the more possibility to occur the effective ET process from Tm3+: 3F4 to Ho3+: 5I7 [31].

 figure: Fig. 4

Fig. 4 Calculated absorption cross sections and emission cross sections corresponding to the 5I75I8 transition of the Ho3+ doped and to the 3F43H6 transition of the Tm3+ doped glasses.

Download Full Size | PDF

The gain spectrum can be obtained using the absorption and emission cross sections of Ho3+ according to the following expression [31]:

G(λ)=N[Pσe(λ)(1P)σa(λ)]
Where P is the population at the upper laser level, and N is the total rare earth ion concentration. We choose the P is 0, 0.2, 0.4, 0.6, 0.8 and 1, the results is shown in Fig. 5. When P>0.4, the gain coefficient is positive at wavelengths >1983 nm. When P>0.6, the gain coefficient is positive from 1850 nm to 2100 nm.

 figure: Fig. 5

Fig. 5 Gain coefficient with various population inversion values P ranging from 0 to1 for Ho3+ 5I75I8 transition.

Download Full Size | PDF

3.4 Energy transfer and lifetimes

To further understand the mechanism of energy transfer between Tm3+ and Ho3+ when pumped by 808 nm LD, the energy level diagram for silicate glass samples is shown in Fig. 6. We have known that at 808 nm LD excitation, Tm3+ ions can be excited to 3H4 state from the 3H6 ground state. When the qualities of Tm3+ ions are accumulated to a certain degree, a part of the Tm3+ ions in 3H4 states is transferred to 3H6, while most of the ions decay from the 3H4 state to the 3F4 state producing 1.47 μm emissions. Meanwhile, other ions are excited from the 3H6 ground state to the 3F4 state, which is the cross relaxation between Tm3+: state 3H4 and Tm3+: 3H6 state (3H4 + 3H63F4 + 3F4). Once Tm3+ ions are saturated in 3F4 level, on one aspect, they decay to 3H6 ground state with strong 1.8 μm emission via 3F43H6 transition; furthermore, Tm3+ ions in 3F4 state transfer their energy to Ho3+: 5I7 state via energy transferring (ET) (3F4 + 5I83H6 + 5I7). When the ions are populated in 5I7 state, the Ho3+ ions decay to 5I8 ground state producing strong 2 μm emissions. Besides, ESA in Tm3+ and Ho3+ also produce fluorescence in weak visible at 480 nm and 550 nm from the Tm3+: 1G43H6 and Ho3+: 5S2, 5F45I8 transitions, respectively [32].

 figure: Fig. 6

Fig. 6 Energy level diagram of Tm3+ and Ho3+ in silicate glass samples.

Download Full Size | PDF

The energy transfer efficiency η from Tm3+ to Ho3+ ions can be measured by the donor lifetime values according to the formula [33]

η =1τTmτTm0
Where τTm and τTm0 are the lifetimes of glasses that co-doped with Tm3+/Ho3+ and doped Tm3+ singly. The lifetimes of Tm3+ for 1 mol% Tm2O3 singly silicate glass sample and 1mol% Tm2O3/0.3 mol% Ho2O3 co-doped silicate glass samples are 470 ms and 155 ms, respectively. Then the energy transfer efficiency from Tm3+ to Ho3+ ions is calculated to be 67.02%, which is higher than the values in previous reported. Though both donor excited levels are same (3F4 of Tm3+ ions), the intrinsic lifetimes differ, which also proves that there are interactions in Ho3+ and Tm3+ indirectly [25, 30]. In the Tm3+/Ho3+ co-doped silicate glass, the higher energy transfer efficient is, the more opportunities that energy of Tm3+ transfer to Ho3+, which is beneficial to obtain 2 μm emission.

It is common to use extended integral method analyze ET processes. The energy transfer probability rate between donor and acceptor (here is Tm3+ ions and Ho3+ ions respectively) can be valued as [34]:

Wda=Cda/R6
Where R is the distance between the donor and acceptor, and Cd-a is the transfer constant calculated by [34]:
 CDA=6cglowD(2π)4n2gupDm=0e(2n¯+1)S0S0mm!(n+1)mσemsDλm+σabsA(λ)
where n¯ is the average occupancy of the phonon mode at temperature T, c is the light speed, n is the refractive index, gdup and gdlow are the degeneracies of the upper and lower levels of the donor, respectively, and λm+=1/(1/λmħω0) is the wavelength with m phonon creation. The critical radius of each interaction can be obtained by [34]:
Rc6=CDAτD
τD is the intrinsic lifetime of the donor without the acceptor. The calculated transfer constants and critical radii of energy migration (EM) are listed in Table 2. These transfer constants are larger than those in silicate glass with alkali ion and fluoride glass but smaller than those in germanate glass [15, 35, 25]. The critical radii of these ET processes are smaller than those in fluoride and germanate glasses [35, 25]. The intrinsic lifetimes are shorter in silicate glass than in fluoride and germanate glasses because of the larger phonon energy of silicate glass. The energy gap between Tm3+ and Ho3+ is nearly ~700 cm−1. As is known the phonon energy of the silicate glasses is ~1050 cm−1, so only about one or less phonon is required to bridge the energy gap.

Tables Icon

Table 2. energy transfer constant and critical radius of each energy transfer process

Figure 7 is the decay curves obtained under excitation at 808 nm LD. And the radiative lifetimes of the 5I7 level (τm) in STH glasses samples are valued by a single exponential function to a good approximation. Since a long lifetime is beneficial high population inversion under steady state, τm of 5I7 is a critical factor to achieve the 2 μm laser [36]. As is shown in Fig. 6, the τm of present glass is larger than those of hosts [15, 22], which makes this Ho3+/Tm3+ co-doped silicate glass possess more probably to develop 2 μm lasers. In decay curves, the lifetimes of Ho3+ are increased with the sensitizer concentration increment. Due to the influence of concentration quenching, the lifetime of STH-4 is much smaller than that of STH-3, which is in keeping with the result of fluorescence spectra. In this process, Tm3+ ions are transferred from 3H4 to 3H5, and then relax to 3F4 through multiphonon relaxation. The energy of there are transferred to 5I7 (Ho3+) state.

 figure: Fig. 7

Fig. 7 The decay curves of 5I7 level in the STH glass samples.

Download Full Size | PDF

In the view of the above results, the prepared Tm3+/Ho3+ co-doped silicate glass is a promising candidate to achieve mid-infrared laser emission around the wavelength of 2 μm with high energy transfer.

3.5 Raman spectra and thermal stability analysis

The spectroscopic technique is used to analyze the glass structure, as well as the vibrational modes. In addition, the corresponding Raman spectra of silicate glass samples are shown in Fig. 8. As is shown in this spectrum, three obvious Raman peaks are labeled. Peak A occurring around 566 cm−1 in a low-frequency region may be attributed to breathing modes of the three-ring structures of SiO4 tetrahedron. Peak B around 660 cm−1 observed in a high-frequency region are related to the convolution of bands related to flexural vibrations of Si-O units [37]. With the appearance of F-, the bridging oxide is disappeared and appears non-bridging oxide. Peak C is ascribed to Si-O-Si symmetric stretching [38]. In addition, the Si-O-Si asymmetric stretching intensity in this glass matrix is stronger than that in other glass, indicating the further destruction of the asymmetric stretching. Due to the appearance of fluoride, the phonon energy of present silicate glass is reduced to 958 cm−1, which is lower the average value (1080 cm−1) [39] of silicate glass. As we known, the smaller phonon energy is the smaller non-radiative relation rate, which means this glass possesses excellent luminescence productiveness.

 figure: Fig. 8

Fig. 8 The Raman spectra of silicate glass.

Download Full Size | PDF

Combing the spectrum prosperities, STH-3 glass sample is considered to be the promising candidate to achieve laser emission. For drawing fiber, the thermal stability is a significant factor that will affect the quality of glass fibers. Just as Fig. 9 shows, the values of Tg, Tx, are found to be 619, 786 °C, respectively. The glass criterion, ΔT = Tx-Tg [40], introduced by Dietzel, is often regarded as an important parameter for evaluating the glass-forming ability. In this glass, the ΔT is 167°C. The glass formation factor of the materials is given by the formula kgl = (Tx-Tg)/(Tm-Tg), which is more suitable for estimating the glass thermal stability than ΔT. Larger kgl represents better forming ability of the glass. The existing ability criterion parameters kgl is 0.214. The values reveal that the STH-3 glass possesses good forming ability. Obviously, the STH-3 glass with a high glass transition temperature Tg is desired and has a potential for high power laser application.

 figure: Fig. 9

Fig. 9 DSC curve of silicate glass (STH-3).

Download Full Size | PDF

4. Conclusion

In summary, we have prepared samples of Tm3+/Ho3+ co-doped silicate glass by the melt-quenching method. The infrared emission of Tm3+/Ho3+ co-doped silicate glass is obtained under 808 nm excitation. The Judd-Ofelt theory has been used to calculate the intensity parameters and emission cross sections. In consideration of the emission property, fluorescence lifetimes, the content of hydroxyl and other factors, we select the STH-3 as the best candidate. The predicted spontaneous transition probability and emission cross section of Ho3+: 5I75I8 transition in STH-3 reach 128.89 s−1 and 7.59 × 10−21 cm2 at the wavelength of 2065 nm, respectively. It is found that Tm3+ ions can transfer their energy to Ho3+ with high efficiency (67.02%), which can enhance the pump absorption. The energy transfer constant from Tm3+ to Ho3+ can reach as high as 63.2 × 10−40 cm6/s and fluorescence lifetime of STH-3 is 0.637 ms. Besides, the thermal stable is nice. The above results indicate that the present glasses should have potential technological applications in 2 μm laser materials.

Acknowledgments

This research was financially supported by the Chinese National Natural Science Foundation (No. 51372235, 51272243, 51472225 and 61308090), Zhejiang Provincial Natural Science Foundation of China (No.LR14E020003), the International Science & Technology Cooperation Program of China (Grant no. 2013DFE63070), and Public Technical International Cooperation project of Science Technology Department of Zhejiang Province(2015c340009)

References and links

1. C. Guo, D. Shen, J. Long, and F. Wang, “High-power and widely tunable Tm doped fiber laser at 2 μm,” Chin. Opt. Lett. 10(9), 091406 (2012). [CrossRef]  

2. B. Yao, W. Wang, K. Yu, G. Li, and Y. Wang, “Passively mode-locked Tm, Ho: YVO4 laser based on a semiconductor saturable absorber mirror, Chin,” Opt. Lett. 10(7), 071402 (2012). [CrossRef]  

3. B. M. Walsh, “Review of Tm and Ho materials spectroscopy and lasers,” Laser Phys. 19(4), 855–866 (2009). [CrossRef]  

4. F. Fusari, A. A. Lagatsky, G. Jose, S. Calvez, A. Jha, M. D. Dawson, J. A. Gupta, W. Sibbett, and C. T. A. Brown, “Femtosecond mode-locked Tm3+ and Tm3+-Ho3+ doped 2 μm glass lasers,” Opt. Express 18(21), 22090–22098 (2010). [CrossRef]   [PubMed]  

5. G. Ozen and B. DiBartolo, “The microscopic interaction parameter for Tm-to-Ho resonant energy transfer in LiYF4,” J. Phys. Condens. Matter 13(1), 195–202 (2001). [CrossRef]  

6. L. L. Yang, J. F. Tang, J. H. Huang, X. H. Gong, Y. J. Chen, Y. F. Lin, Z. D. Luo, and Y. D. Huang, “Spectral properties of Ho3+/Tm3+ co-doped β′-Gd2(MoO4)3 crystal as laser gain medium around 2.0 μm,” Opt. Mater. 35(12), 2188–2193 (2013). [CrossRef]  

7. X. Wang, X. K. Fan, S. Gao, K. F. Li, and L. L. Hu, “Spectroscopic properties of Tm3+/Ho3+ co-doped SiO2-Al2O3-CaO-SrO glasses,” Ceram. Int. 40(7), 9751–9756 (2014). [CrossRef]  

8. L. Wang, Y. Zhao, S. Xu, and M. Peng, “Thermal degradation of ultrabroad bismuth NIR luminescence in bismuth-doped tantalum germanate laser glasses,” Opt. Lett. 41(7), 1340–1343 (2016). [CrossRef]   [PubMed]  

9. I. Razdobreev, L. Bigot, V. Pureur, A. Favre, G. Bouwmans, and M. Douay, “Efficient all-fiber bismuth-doped laser,” Appl. Phys. Lett. 90(3), 031103 (2007). [CrossRef]  

10. V. V. Dvoyrin, V. M. Mashinsky, L. I. Bulatov, I. A. Bufetov, A. V. Shubin, M. A. Melkumov, E. F. Kustov, E. M. Dianov, A. A. Umnikov, V. F. Khopin, M. V. Yashkov, and A. N. Guryanov, “Bismuth-doped-glass optical fibers--a new active medium for lasers and amplifiers,” Opt. Lett. 31(20), 2966–2968 (2006). [CrossRef]   [PubMed]  

11. E. M. Dianov, “Bismuth-doped optical fibers: a challenging active medium for near-IR lasers and optical amplifiers,” Light Sci. Appl. 1(5), e12 (2012). [CrossRef]  

12. T. Feng, F. P. Yan, W. J. Peng, Q. Li, S. Y. Tan, J. Wang, and X. D. Wen, “Theoretical analysis of characteristics for 2 μm Tm3+: Ho3+ co-doped silica fiber laser pumped by a 1550 nm fiber laser,” Opt. Fiber Technol. 18(4), 204–208 (2012). [CrossRef]  

13. J. Tang, K. Lu, S. Zhang, P. Zhang, F. Chen, S. Dai, and Y. Xu, “Surface Plasmon resonance-enhanced 2 μm emission of bismuth germanate glasses doped with Ho3+/Tm3+ ions,” Opt. Mater. 54, 160–164 (2016). [CrossRef]  

14. Z. Feng, S. Yang, H. Xia, C. Wang, D. Jiang, J. Zhang, X. Gu, Y. Zhang, B. Chen, and H. Jiang, “Energy transfer and 2.0 μm emission in Tm3+/Ho3+ co-doped α-NaYF4 single crystals,” Mater. Res. Bull. 76, 279–283 (2016). [CrossRef]  

15. M. Li, Y. Guo, G. Bai, Y. Tian, L. Hu, and J. Zhang, “2 μm luminescence and energy transfer characteristics in Tm3+/Ho3+ co-doped silicate glass,” J. Quant. Spectrosc. Radiat 127, 70–77 (2013). [CrossRef]  

16. D. Dorosz, J. Zmojda, and M. Kochanowicz, “Investigation on broadband near-infrared emission in Yb3+/Ho3+ co-doped antimony-silicate glass and optical fiber,” Opt. Mater. 35(12), 2577–2580 (2013). [CrossRef]  

17. S. Stanek, P. Nekvindova, B. Svecova, S. Vytykacova, M. Mika, J. Oswald, A. Mackova, P. Malinsky, and J. Spirkova, “The influence of silver-ion doping using ion implantation on the luminescence properties of Er-Yb silicate glasses,” Nucl. Instrum. Methods Phys. Res. B 371, 350–354 (2016). [CrossRef]  

18. T. Togashi, T. Honma, K. Shinozaki, and T. Komatsu, “Electrochemical performance as cathode of lithium iron silicate, borate and phosphate glasses with different Fe2+ fractions,” J. Non-Cryst. Solids 436, 51–57 (2016). [CrossRef]  

19. B. Suresh, M. Srinivasa Reddy, J. Ashok, A. S. S. Reddy, P. V. Rao, V. R. Kumar, and N. Veeraiah, “Enhancement of orange emission of Co2+ ions with Bi3+ ions in lead silicate glasses,” J. Lumin. 172, 47–52 (2016). [CrossRef]  

20. R. Cao, M. Cai, Y. Lu, Y. Tian, F. Huang, S. Xu, and J. Zhang,“Ho3+/Yb3+ co-doped silicate glasses for 2 μm emission performances,” Appl. Opt. 55, 2065–2070 (2016). [CrossRef]   [PubMed]  

21. Y. Zhou, L. Wu, B. Huang, F. Yang, S. Peng, Y. Qi, and D. Yin, “Enhanced 2 μm fluorescence and thermal stability in Ho3+/Tm3+ co-doped WO3 modified tellurite glasses,” Mater. Lett. 142, 277–279 (2015). [CrossRef]  

22. B. Peng and T. Izumitani, “Optical properties, fluorescence mechanisms and energy transfer in Tm3+, Ho3+ and Tm3+/Ho3+ doped near-infrared laser glasses, sensitized by Yb3+,” Opt. Mater. 4(6), 797–810 (1995). [CrossRef]  

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

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

25. Q. Zhang, J. Ding, Y. Shen, G. Zhang, G. Lin, J. Qiu, and D. Chen, “Infrared emission properties and energy transfer between Tm3+ and Ho3+ in lanthanum aluminum germanate glasses,” J. Opt. Soc. Am. B 27(5), 975–980 (2010). [CrossRef]  

26. K. Li, G. Wang, J. Zhang, and L. Hu, “Broadband 2 μm emission in Tm3+/Ho3+ co-doped TeO2-WO3-La2O3 glass,” Solid State Commun. 150, 1915–1918 (2010). [CrossRef]  

27. F. Auzel, D. Meichenin, and H. Poignant, “Laser cross section and quantum yield of Er3+ at 2.7 μm in a ZrF4 based fluoride glass,” Electron. Lett. 24(15), 909–910 (1988). [CrossRef]  

28. B. Jacquier, L. Bigot, S. Guy, and A. M. Jurdyc, “Rare earth doped confined structures for lasers and amplifiers,” The Series Springer Series in Materials Science 83, 430–461 (2005). [CrossRef]  

29. A. Jouini, A. Brenier, Y. Guyot, G. Boulon, H. Sato, A. Yoshikawa, K. Fukuda, and T. Fukudat, “Spectroscopic and laser properties of the near-infrared tunable laser material Yb3+ doped CaF2 crystal,” Cryst. Growth Des. 8(3), 808–811 (2008). [CrossRef]  

30. G. X. Chen, Q. Y. Zhang, G. F. Yang, and Z. H. Jiang, “Mid-infrared emission characteristic and energy transfer of Ho3+-doped tellurite glass sensitized by Tm 3+.,” J. Fluoresc. 17(3), 301–307 (2007). [CrossRef]   [PubMed]  

31. H. Gebavi, D. Milanese, R. Balda, S. Taccheo, J. Fernandez, J. Lousteau, and M. Ferraris, “Spectroscopy of thulium and holmium heavily doped tellurite glasses,” J. Lumin. 132(2), 270–276 (2012). [CrossRef]  

32. 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]  

33. L. Zhang, H. Hu, C. Qi, and F. Lin, “Spectroscopic properties and energy transfer in Yb3+/Er3+ doped phosphate glasses,” Opt. Mater. 17(3), 371–377 (2001). [CrossRef]  

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

35. Y. Tian, X. F. Jing, and S. Q. Xu, “Spectroscopic analysis and efficient diode-pumped 2.0 μm emission in Ho3+/Tm3+ co-doped fluoride glass,” Spectro chemic Acta Part A: Molecular and Bimolecular Spectroscopy 115, 33–38 (2013). [CrossRef]  

36. R. Rolli, A. Chiasera, M. Montagna, E. Moser, S. Ronchin, S. Pelli, G. C. Righini, A. Jha, V. K. Tikhomirov, S. A. Tikhomirova, C. Duverger, P. Galinetto, and M. Ferrari, “Rare-earth-activated fluoride and tellurite glasses: optical and spectroscopic properties,” Proc. SPIE 4282, 10 (2001). [CrossRef]  

37. F. Huang, Y. Guo, Y. Ma, L. Hu, and J. Zhang, “2.7 μm Emission Properties of Er3+ Doped Fluorozirconate Glass,” Glass Phys. Chem. 40(3), 277–282 (2014). [CrossRef]  

38. T. Wei, F. Chen, X. Jing, F. Wang, Y. Tian, and S. Xu, “Structure and spectroscopic properties of Er3+ doped germanate glass for mid-infrared application,” Solid State Sci. 31, 54–61 (2014). [CrossRef]  

39. Y. Guo, M. Li, L. Hu, and J. Zhang, “Intense 2.7 μm emission and structural origin in Er3+ doped bismuthate glass,” Opt. Lett. 37(2), 268–270 (2012). [CrossRef]   [PubMed]  

40. Y. P. Peng, X. Yuan, J. Zhang, and L. Zhang, “The effect of La₂O₃ in Tm³⁺-doped germanate-tellurite glasses for ~2 μm emission,” Sci. Rep. 4, 5256 (2014). [CrossRef]   [PubMed]  

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 (9)

Fig. 1
Fig. 1 Absorption spectrum of silicate glass doped Tm3+, Ho3+ singly and Tm3+/Ho3+ co-doped.
Fig. 2
Fig. 2 the fluorescence spectra of STH glass.
Fig. 3
Fig. 3 The transmittance spectrum of silicate glass samples with different fluoride content.
Fig. 4
Fig. 4 Calculated absorption cross sections and emission cross sections corresponding to the 5I75I8 transition of the Ho3+ doped and to the 3F43H6 transition of the Tm3+ doped glasses.
Fig. 5
Fig. 5 Gain coefficient with various population inversion values P ranging from 0 to1 for Ho3+ 5I75I8 transition.
Fig. 6
Fig. 6 Energy level diagram of Tm3+ and Ho3+ in silicate glass samples.
Fig. 7
Fig. 7 The decay curves of 5I7 level in the STH glass samples.
Fig. 8
Fig. 8 The Raman spectra of silicate glass.
Fig. 9
Fig. 9 DSC curve of silicate glass (STH-3).

Tables (2)

Tables Icon

Table 1 Judd-Ofelt parameters Ωt of Ho3+ in various glasses

Tables Icon

Table 2 energy transfer constant and critical radius of each energy transfer process

Equations (11)

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

A= 1 τ rad = 64 π 4 e 2 3h( 2J+1 ) λ 3 [ n ( n 2 +1 ) 2 9 S ED + n 2 S MD ]
S ED = λ=2,4,6 Ω λ | <S, L, J U (λ) S ' , L ' , J ' > | 2
S MD =( ħ 2mc )| <S, L, J L+2S S ' , L ' , J ' > | 2
OH + F = O 2 +HF
σ em ( λ )= λ 4 A rad 8 πcn 2 λI(λ) λI(λ)
σ a = 2.303log( I 0 I ) Nl
G( λ )=N[ P σ e ( λ )( 1P ) σ a ( λ ) ]
η =1 τ Tm τ Tm 0
W da = C da / R 6
 C DA = 6 cg low D ( 2π ) 4 n 2 g up D m=0 e ( 2 n ¯ +1 ) S 0 S 0 m m! ( n+1 ) m σ ems D λ m + σ abs A ( λ )
R c 6 = C DA τ D
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.