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

Positive influence of Ce3+ on effective transfer Yb3+:2F5/2 → Ho3+:5I6 in silica-germanate glass for mid-infrared applications

Open Access Open Access

Abstract

A1.5 time enhancement of a 2.0 μm emission was achieved successfully in Yb3+/Ho3+ doped silica-germanate glass with a 0.1 mol% Ce3+ addition, which possesses a larger emission cross section (4.43 × 10−21 cm2). According to the measured absorption spectra, the Judd-Ofelt parameters and radiative properties were calculated and discussed. The energy transfer mechanisms existed in Ho3+, Yb3+ and Ce3+ ions were investigated based on absorption, upconversion and fluorescence spectra. Meanwhile, the decay profiles of several levels were measured to further examine the enhanced mid-infrared emissions. Moreover, a high energy transfer microscopic parameter (7.54 × 10−40cm6/s) of Yb3+→Ho3+ process was obtained when 0.1 mol% Ce3+ ions were introduced into the Ho3+/Yb3+ system. All results indicate that the Yb3+/Ho3+/Ce3+ tri-doped silica-germanate glass is a promising candidate material for improving the Ho3+ 2.0 μm fiber laser performance.

© 2017 Optical Society of America

Corrections

28 February 2017: A correction was made to the title.

1. Introduction

In last decades, extensive work has been done on the 2.0 μm laser materials driven by its wide applications in biomedicine, hazardous chemical detection, pollution monitoring and military system [1–4]. In particular, the trace gases with even ppm level can be detected because of various molecules (CO2, CO, NO, SO2) have strong rovibrational absorption lines in this ‘molecule fingerprint’ region of the electromagnetic spectrum. Tm3+ and Ho3+ are commonly chosen as fiber dopants to generate 2 μm laser due to Tm3+:3F43H6 and Ho3+:5I75I8 transitions, respectively. Ho3+ possesses longer radiative lifetime, longer-operating laser wavelength, higher emission cross section and broad emission band compared with Tm3+ [5–7]. However, there is no energy level in Ho3+ matching with the low cost and readily commercially available 800 or 980 nm laser diodes (LDs) [8,9], which limits its efficient laser operation. Therefore, some of the sensitizers such as Er3+, Yb3+, Tm3+ and Nd3+ with strong absorption bands in commercial LD at wavelength of 808 or 980 nm are taken into consideration to absorb and transfer pumping energy effectively. Up to now, investigations on the 2.0 μm emission from Er3+/Ho3+, Yb3+/Ho3+, Tm3+/Ho3+ co-doped and Er3+/Tm3+/Ho3+, Er3+/Yb3+/Ho3+ tri-doped have been performed in the glasses [10–14].

Among all the sensitizers, Yb3+ plays an important role due to its efficient absorption at 980 nm, and the simple energy level structure for Yb3+ (the ground state 2F7/2 and excited state 2F5/2) can avoid any undesirable excited state absorption [15,16]. However, strong upconversion (UC) emissions occurred in Yb3+/Ho3+ co-doping scheme, especially, when they are incorporated in a host matrix with relatively low phonon energies, leading to a dramatic reduction of the pumping efficiency for the 2.0 μm emission [8,17,18]. Thus, Ho3+ 2.0 μm emission intensity can be enhanced by developing an effective approach to minimize the UC process. Ce3+ has been proved as an ideal candidate in enhancing Er3+ 1.53 μm emission by reducing the UC emission when pumped by a 980 nm LD [19]. A small energy mismatch is found between the Er3+:4I11/24I13/2 and Ce3+:2F7/22F5/2 energy gap, so the cross relaxation (CR) process (Er3+:4I11/2 + Ce3+:2F5/2→Er3+:4I13/2 + Ce3+:2F7/2) between Er3+ and Ce3+ lead to a inhibition of the excited state absorption (ESA) and cooperative upconversion (CUC) processes which occur in Er3+:4I11/2, both of them would further cause a population of Er3+:4F7/2 and result in green UC fluorescence [20]. The energy gap between Ho3+:5I6 and 5I7 levels is similar to that between Er3+:4I11/2 and 4I13/2 levels, and the visible upconverted emissions are also occurred in Yb3+/Ho3+ co-doping under 980 nm LD due to the ESA and CUC processes, in which some of Ho3+ ions in 5I6 were excited to higher manifolds (5S2 + 5F4) [8,17,18]. Hence, the incorporation of Ce3+ would reinforce the Ho3+ 2.0 μm emission intensity through CR process (Ho3+:5I6 + Ce3+:2F5/2→Ho3+:5I7 + Ce3+:2F7/2). In 2011, Tian et al. reported the 2.0 μm emission in Yb3+/Ho3+/Ce3+ tri-doped fluorophosphate glass under 980 nm LD [21]. In the next year, Tao et al. using Ce3+ as sensitize ion obtained an enhanced 2.0 μm emission in Yb3+/Ho3+/Ce3+ tri-doped tellurite glass [8], but it has not been reported in heavy metal oxide glass. In addition, the poor chemical durability, lower damage threshold, low thermal shock resistance in fluoride and tellurite glasses, which limit the stability and output power of fluoride and tellurite fiber lasers [22,23], and the effect of Ce3+ ion on Ho3+ 2.0 μm emission in fluorophosphate glass only resulting in a very weak enhancement.

Meanwhile, the luminescence of Ho3+ is easily quenched by the rapid multiphonon effect when it is incorporated in traditional oxide glasses due to their high phonon energies (e.g., 1500 cm−1 for silica) [17,24]. So the suitable host is as important as the luminescent ions, when considering the gain media to be used in fiber lasers, both good mechanical properties and a high gain are required [25]. Silica-germanate glass combines the merits of relatively low maximum phonon energy and higher refraction index of germanate glass and low cost together with higher thermal stability of silicate glass [26,27]. In 2015, Wei et al. reported broadband 2.0 μm fluorescence in Er3+/ Ho3+ co-doped germanosilicate glass [28]. In this work, we report the sensitization effect of Ce3+ ions on Yb3+/Ho3+ co-doped silica-germanate glass. The Judd-Ofelt intensity parameters, transition and radiative probabilities, absorption and stimulated emission cross sections were calculated and discussed. In addition, the microparameters of the energy transfer processes were calculated.

2. Experimental

The compositions of silica-germanate glasses were 35SiO2-35GeO2-20 Ga2O3-5BaO-5Li2O-1Yb2O3-1Ho2O3-XCeF3 (X = 0, 0.1, 0.25, 0.5 denoted as YHC0, YHC0.1, YHC0.25 and YHC0.5, respectively). Meanwhile, the 1 mol% Ho3+ singly doped silica-germanate glass was prepared and denoted as H. High-purity of SiO2, GeO2, Ga2O3, BaO, Li2CO3, Yb2O3, Ho2O3 and CeF3 powders were used as raw materials. Batches of raw materials (15 g) were well-mixed and melted in a platinum crucible in a Si-Mo resistance electric furnace at 1480 °C for 40 min. Then, they were quenched on preheated stainless steel plate and annealed near the temperature of glass transition for 2 h before they were cooled to room temperature. The annealed samples were finally cut and optically polished to the size of 10 mm × 10 mm × 1 mm for the optical property measurements.

The densities of samples were tested via the Archimedes principle using distilled water as the immersion medium with error of ± 0.001 g/cm3. Refractive indexes were measured by the prism minimum deviation method at the wavelength of 1053 nm with error limit of ± 0.05%. The visible UC (500–800 nm) fluorescence spectra were obtained with a TRIAX550 spectrofluorimeter upon the excitation of 980 nm LD. Absorption spectra were obtained using a PerkinElmer Lambda 900UV-VIS-NIR spectrophotometer in the range of 300~2200 nm with a resolution of 1 nm. Fluorescence spectra from 1000~2200 nm were tested with a computer-controlled Triax 320 type spectrometer excited by a 980 nm LD. All the measurements were carried out at room temperature.

3. Results and discussion

3.1 Absorption spectra and Judd-Ofelt analysis

Figure 1 shows the absorption spectra of Ho3+ singly doped, Yb3+/Ho3+ co-doped and Yb3+/Ho3+/Ce3+ tri-doped silica-germanate in the wavelength region of 300~2200 nm. The absorption bands corresponding to the transitions starting from Yb3+:2F7/2 ground state to 2F5/2 and the Ho3+:5I8 ground state to the higher levels 5I7, 5I6, 5F5, (5F4, 5S2), (5G6, 5F1) and 5G5 are labeled. It can be seen from Fig. 1 that the addition of Ce3+ ions into the glass matrix results in a red-shift of the UV-side absorption edge, and this is assigned to the 4f→5d transition in Ce3+ ions which causes a broad absorption around 400–500 nm [29,30].

 figure: Fig. 1

Fig. 1 Absorption spectra of H, YHC0 and YHC0.1 glasses.

Download Full Size | PDF

Based on the absorption spectrum of Ho3+ and J–O theory [31,32], J–O intensity parameters have been calculated in order to investigate the local environment around Ho3+ ions. Details of the theory and method have been well described earlier [33–35]. Table 1 shows the values measured (fmea) and calculated (fcal) oscillator strengths of sample (H) and other reported glass systems. Good agreement between fmea and fcal can be observed, and the small root mean square deviation δrms of 0.82 × 10−6, which further indicates the reliability and validity of the calculated J-O intensity parameters. In addition, it can be noticed that the oscillator strengths of silica-germanate glass are higher than those of silicate [36] and germanate [34] glasses in Table 1. It results from the different ligand field environment around Ho3+ ions [14]. What’s more, the transition of 5I85F1 + 5G6 possesses the highest oscillator strength called hypersensitive transition (HST), which is sensitive to minor changes of local environment around Ho3+ ions and affects the J-O intensity parameters that characterize the optical properties of the system [35,37]. J–O parameters were obtained for further estimation of the difference of ligand field in glasses below.

Tables Icon

Table 1. Measured (fmea) and calculated (fcal) oscillator strengths of H sample and compared with other systems.

The J-O intensity parameters Ωλ (λ = 2, 4, 6) of Ho3+ ions in various glass hosts are listed in Table 2. Previous studies have proved that parameters Ω2 is closely related to the covalency parameter through the nephelauxetic effect and is strongly dependent on the local environments of Ho3+ ions [13]. It can be obtained that the Ω2 value of Ho3+ doped sample (H) is larger than those of fluoride and fluorophosphate glasses [38,39] but lower than those of silicate and germanate glasses [40,41], which can be explained by the higher polarizability of O2- ions than F- ions. It is illustrated that the silica-germanate glass possesses lower polarizability and covalency. Meanwhile, the Ω6 parameter is a vibronic dependent parameter, related to the rigidity and the viscosity of the host glass and the radio of Ω46 determines the spectroscopy quality of the materials [42,43]. The prepared sample (H) possesses highest Ω46 value as shown in Table 2. The larger Ω46 value means the higher lasing efficiency for the laser transition of Ho3+ ions in a given matrix. Therefore, the largest Ω46 value of H indicates that it is a great matrix for 2.0 μm emission [44].

Tables Icon

Table 2. J–O intensity parameters Ωλ (λ = 2, 4, 6) ( × 10−20 cm2) of Ho3+ ions in various glass hosts.

The radiative properties of Ho3+ in silica-germanate glass such as radiative transition probability (Arad), fluorescence branching ratios (β) and radiative lifetimes (τrad) calculated based on J-O intensity parameters are listed in Table 3. High spontaneous transition probability is beneficial for high gain and more opportunity to achieve laser action [43]. It is found that the Arad of Ho3+:5I75I8 transition is 85.89 s−1 of the prepared sample, which is larger than those of fluorophosphate (73.54 s−1) [39], silicate (78.71s−1) [40] and germanate glasses (70.15 s−1) [41].

Tables Icon

Table 3. The calculated radiative transition probability (Arad), fluorescence branching ratios (β) and radiative lifetimes (τrad) of Ho3+ ions for different transitions.

3.2 Fluorescence spectra and energy transfer mechanism

The fluorescence spectra of the YHC (0, 0.1, 0.25, 0.5) glasses in the wavelength of 1050 to 2200 nm pumped at 980 nm are shown in Fig. 2. Because of the Ho3+:5I65I8 transition, the 1.2 μm emission is obtained. The 1.2 μm emission intensity decreases with increasing Ce3+ concentration, while the emission intensity of 2.0 μm increases with incorporation of Ce3+, and reaches the maximum when mol percentage of Ce3+ reaches to 0.1% at which the emission intensity almost enhances twice compared with YHC0. The 2.0 μm emission intensity was restrained with further Ce3+ ions adding. Insets in Fig. 2 show the decay curves of the 5I6 and 5I7 levels monitoring at 1200 and 2055 nm respectively, which are measured in both YHC0 and YHC0.1. It is found that the lifetime of 5I7 level in YHC0.1 is 727 μs, which shows a clear increase compared with YHC0 (565 μs).

 figure: Fig. 2

Fig. 2 Fluorescence spectra of YHC (0, 0.1, 0.25, 0.5) glasses. Insets in Fig. 2 show the decay curves of the YHC0 and YHC0.1 at 1200 and 2055 nm.

Download Full Size | PDF

The UC emission spectra of the samples were measured to further understand the enhancement of Ho3+ 2.0 μm emission. As shown in Fig. 3, intensive red emission (662 nm) dominates the UC luminescence, corresponding to the Ho3+:5F55I8 transitions. Two weak emission peaks located at 552 and 755 nm originating from Ho3+:(5F4 + 5S2)→5I8 and (5F4 + 5S2)→5I7 are also observed. Different from the variation tendency of 662 nm emission, the intensities of 552 and 755 nm emissions decrease with increasing Ce3+ concentration.

 figure: Fig. 3

Fig. 3 Upconversion spectra of YHC (0, 0.1, 0.25, 0.5) glasses.

Download Full Size | PDF

According to the absorption, upconversion, fluorescence spectra in YHC (0, 0.1, 0.25, 0.5) glasses, the involved energy transfer mechanisms are indicated in Fig. 4. Ions on the Yb3+:2F7/2 level are excited to the 2F5/2 state by ground state absorption (GSA) when sample is pumped by 980 nm LD. Then the energy is transferred to a nearby Ho3+ in the ground state resulting in a population of the Ho3+:5I6 by energy transfer process (ET) under the assistance of host phonons. Some of the Ho3+:5I6 level are excited to the (5F4 + 5S2) manifolds via an energy transfer upconversion (ETU1) and excited state absorption (ESA1), resulting the green emission when it radiatively decays to the ground state. Meanwhile, the red upconversion emission centered at 662 nm shown in Fig. 4 is originated from the 5F55I8 transition. Population of the 5F5 level is derived from two different processes which are nonradiative relaxation process from Ho3+:5F4 + 5S2 level to 5F5 level and ETU2 or ESA2 processes [8,21]. The rest of ions in Ho3+:5F4 + 5S2 radiatively decay to Ho3+:5I7 state with 755 nm emission. The CR process (Ho3+:5I6 + Ce3+:2F5/2→Ho3+:5I7 + Ce3+:2F7/2) between Ho3+ and Ce3+, the Ho3+ in 5I6 state decays non-radiatively to the next lower Ho3+ 2.0 μm emission level 5I7 with the energy transferred to a nearby Ce3+ in the ground state. Thus, the emissions occurred in the Ho3+:5I6 level were reduced, which is confirmed by the decreased intensity of 552, 662, 755 and 1200 nm emissions. It should be noticed that the red emission at 662 nm shows a slight increment after initial decrease and the peak dropped again with increasing Ce3+ concentration while the green emission (552 nm) intensity reduces continuously. It is caused by the energy transfer upconversion (ETU2) and excited state absorption (ESA2) that takes place in the Ho3+:5I7 level, which can also explain the variation tendency of 2.0 μm emission. When Ce3+ equals to 0.1 mol%, the Ho3+:5I7 level is populated by CR process and few ions are excited to the Ho3+:5F5 state; when adding 0.25% Ce3+, a portion of Ho3+:5I7 level are excited to the Ho3+:5F5 state; when Ce3+ concentration reaches to 0.5%, a large part of Ho3+:5I7 level is excited to the Ho3+:5F5 state. The reason why the emission at 662 nm in YHC0.5 is weakest may because of the Ho3+:5F5 level is almost only derived from Ho3+:5I7 level by ETU2 and ESA2 process.

 figure: Fig. 4

Fig. 4 Energy level diagrams and energy transfer mechanism among Yb3+, Ho3+ and Ce3+ ions.

Download Full Size | PDF

3.3 Cross sections and emission parameters

Emission cross section is an important parameter to estimate the possibility for achieving laser action. Higher emission cross section means that better laser gain can be achieved in prepared glass [45]. The emission cross section (σem) can be calculated according to Fuchtbauer–Ladenburg (FL) equation [46]

σem(λ)=λ4Arad8πcn2×λI(λ)λI(λ)dλ
where λ is the emission wavelength, Arad is the spontaneous radiative transition probability of Ho3+:5I75I8 transition, I(λ) is the 2.0 μm fluorescence intensity, c is the velocity of light in vacuum, n is the refractive index of glass host.

The absorption cross section (σabs) can be derived from the calculated σem using the McCumber (MC) formula [47]

σem(λ)=σabs(λ)×ZlZu×exp[hckt×(1λZL1λ)]
where Zl and Zu are the partition functions for the lower and the upper levels, respectively. Parameter h, k, T are the Planck constant (6.63 × 10−34 J·S), Boltzmann constant (1.38 × 10−23 J/K) and the temperature (here it is the room temperature), respectively, and λZL is the wavelength for the transition between the lower Stark sublevels of the emitting multiplets and the lower Stark sublevels of the receiving multiplets.

The absorption and emission cross sections of Ho3+ in YHC0.1 are shown in Fig. 5(a). The maximal value of the calculated emission cross section spectrum centers at 2055 nm and reaches as high as 4.43 × 10−21 cm2. This value is higher than the reported ones of germanate (4.0 × 10−21 cm2) [48], silicate (3.07 × 10−21 cm2) [12] and fluoride (4.3 × 10−21 cm2) glasses [5]. The Yb3+/Ho3+/Ce3+ tri-doped silica-germanate glasses possessing large emission cross section can be an excellent candidate in achieving intense 2.0 μm laser.

 figure: Fig. 5

Fig. 5 (a) The calculated absorption and emission cross sections (b) Gain coefficient for prepared glass (YHC0.1).

Download Full Size | PDF

Gain coefficient is another characteristic parameter to evaluate 2.0 μm emission properties quantitatively. The wavelength dependent gain coefficient is calculated in detail as a function of population inversion for the upper laser level on the basis of the σabs and σem. The gain cross-section spectrum G(λ), can be calculated by

G(λ)=N[Pσem(λ)(1P)σabs(λ)]
where P is the population inversion given by the ratio between the population of Ho3+:5I7 level and the total Ho3+ concentration and N is the total concentration of Ho3+. Figure 5(b) presents the calculated gain coefficient versus wavelength for different population inversion parameter P. Obviously, the positive gain is obtained when P > 0.4, which is similar to the case in other glasses with Ho3+ doped [10,14]. Meanwhile, the maximum gain coefficient is up to 1.9 cm−1 around 2055 nm, which is higher than those in fluoride (0.58 cm−1) [5], silicate (1.75 cm−1) [49] and germanate-tellurite glasses (0.53 cm−1) [14]. These results indicate that Yb3+/Ho3+/Ce3+ tri-doped silica-germanate glass can be used as a suitable optical material to obtain 2.0 μm infrared emission.

3.4 Energy transfer microscopic parameters between Yb3+, Ho3+ and Ce3+

In order to further investigate energy transfer processes between Yb3+, Ho3+ and Ce3+ ions quantitatively, energy transfer microscopic parameters were calculated by the determination of their absorption and emission cross sections and the extended integral method is used to analyze energy transfer microscopic parameters.

The probability rate of energy transfer between Yb3+ and Ho3+ can be estimated as [50,51]

WDA=(2π)|HDA|2SDAN
where \HDA\ is the matrix element of the perturbation Hamiltonian between initial and final states in energy transfer process, SNDA is the integral overlap between the m-phonon emission sideband of donor ions (D, here D stands for Yb3+) and k-phonon absorption line shapes of acceptor ions (A, here A stands for Ho3+) and N is the total phonons in the transfer process (m + k = N). In the case of weak electron-phonon coupling which is suitable for rare earth ions, SNDA can be approximated by
SDANe(S0D+S0A)×[(S0D+S0A)NN!]SDA(0,0,E)δ(N,ΔE/w0)
where SD0 and SA0 are the Huang-Rhys factors of the Yb3+ and Ho3+ in silica-germanate glass, SDA(0,0,E) is the overlap between the zero phonon lines shape of Yb3+ emission and Ho3+ absorption. Then the integral overlap in the case of m-phonon emission by the donor and no phonon involvement by the acceptor can be expressed as
SDA(m,0,E)=[S0mm!eS0gemisD(EΔE)]gabsA(E)dE
where ΔE = mħw0. Since the measurements are carried out at some finite temperature T, the multi-phonon probability must be included, the emission cross section (σemis) of Yb3+ with m phonon emission and absorption cross section (σabs) of Ho3+ with k phonon absorption can be proposed as
σemis(mphonon)D=σemisD(λm+)S0meS0m!(n¯+1)mσemisD(EE1)
σabs(kphonon)A=σabsA(λk)S0keS0k!(n¯)kσabsA(E+E2)
where E1 = mħw0, E2 = kħw0. λ+m denotes the translation of Yb3+ emission cross section wavelength by m-phonon emission, and λk represents the translation of Ho3+ absorption cross section spectra wavelength due to k-phonon absorption
λm+=1/(1/λmw0)
λk=1/(1/λ+kw0)
If we ignore the k phonon annihilation process and just focus on the m phonon creation process, the probability rate of energy transfer can be obtained using the following direct transfer equation
WDA(R)=6cglowD(2π)4n2R6gupDm=0e(2n¯+1)S0S0mm!×(n¯+1)mσemisD(λm+)σabsA(λ)dλ=CDAR6
where R is the distance of separation between donor and acceptor, CD-A is the energy transfer coefficient (cm6/s), = (1/eħw0/Kt-1) is the average occupancy of phonon mode at temperature T. The energy transfer coefficient is then expressed by
CDA=6cglowD(2π)4n2gupDm=0e(2n¯+1)S0S0mm!×(n¯+1)mσemisD(λm+)σabsA(λ)dλ
The energy transfer properties of Yb3+→Ho3+ in YHC0 and YHC0.1 are calculated using the Eqs. (4)–(12) and listed in Table 4. The result shows the direct transfer microscopic parameter of Yb3+→Ho3+ in YHC0 is found to be 5.50 × 10−40 cm6/s in a quasi resonant process with non-phonon (8.82%), having a participation of 1 (89.06%) and 2 (2.12%) phonons. The transfer microscopic parameter of Yb3+→Ho3+ in YHC0.1 is as high as 7.54 × 10−40 cm6/s, which is larger than that in YHC0 without Ce3+, the result indicates the high energy transfer efficiency from Yb3+ to Ho3+ after the introduction of Ce3+.

Tables Icon

Table 4. Calculated microscopic parameters CD-A ( × 10−40 cm6/s) for Yb3+→Ho3+ energy transfer in YHC0 and YHC0.1 glasses. The number # of phonons necessary to assist the energy transfer process is also revealed with their percent contributions.

4. Conclusions

By incorporating Ce3+, the upconversion occurred in Ho3+:5I6 state is remarkably reduced through the cross relaxation process (Ho3+:5I6 + Ce3+:2F5/2→Ho3+:5I7 + Ce3+:2F7/2), the luminescence intensity of Ho3+ 2.0 μm emission is nearly two times enhanced in YHC0.1 compared with YHC0. The emission cross section and the maximum gain coefficient in YHC0.1 is calculated to be 4.43 × 10−21 cm2 and 1.9 cm−1 around 2055 nm. Meanwhile, the measurements of decay lifetimes in several levels were carried out to further examine energy transfer mechanism and enhanced mid-infrared emissions. In addition, the energy transfer coefficient from Yb3+→Ho3+ in YHC0.1 is 7.54 × 10−40 cm6/s, which is larger than that in YHC0 (5.50 × 10−40 cm6/s) and almost has 1.5 fold increased. All results suggest that the Yb3+/Ho3+/Ce3+ tri-doped silica-germanate glass is promising for the improvement of Ho3+ 2.0 μm fiber laser performance.

Funding

National Natural Science Foundation of China (NSFC) (No. 61605192, 51472225, 51401197, and 51502022); Open Fund of the Guangdong Engineering Technology Research and Development Center of Special Optical Fiber Materials and Devices (South China University of Technology).

Reference and links

1. S. D. Jackson and A. Lauto, “Diode-pumped fiber lasers: a new clinical tool?” Lasers Surg. Med. 30(3), 184–190 (2002). [CrossRef]   [PubMed]  

2. P. Zhou, X. Wang, Y. Ma, H. Lü, and Z. J. Liu, “Review on Recent Progress on Mid-Infrared Fiber Lasers,” Laser Phys. 22(11), 1744–1751 (2012). [CrossRef]  

3. M. Bernier, V. Michaud-Belleau, S. Levasseur, V. Fortin, J. Genest, and R. Vallée, “All-fiber DFB laser operating at 2.8 μm,” Opt. Lett. 40(1), 81–84 (2015). [CrossRef]   [PubMed]  

4. S. D. Jackson, “Towards high-power mid-infrared emission from a fibre laser,” Nat. Photonics 6(7), 423–431 (2012). [CrossRef]  

5. F. Huang, J. Cheng, X. Liu, L. Hu, and D. Chen, “Ho3+/Er3+ doped fluoride glass sensitized by Ce3+ pumped by 1550 nm LD for efficient 2.0 μm laser applications,” Opt. Express 22(17), 20924–20935 (2014). [CrossRef]   [PubMed]  

6. S. D. Jackson, F. Bugge, and G. Erbert, “Directly diode-pumped holmium fiber lasers,” Opt. Lett. 32(17), 2496–2498 (2007). [CrossRef]   [PubMed]  

7. S. D. Jackson, “The spectroscopic and energy transfer characteristics of the rare earth ions used for silicate glass fibre lasers operating in the shortwave infrared,” Laser Photonics Rev. 3(5), 466–482 (2009). [CrossRef]  

8. L. L. Tao, Y. H. Tsang, B. Zhou, B. Richards, and A. Jha, “Enhanced 2.0 μm emission and energy transfer in Yb3+/Ho3+/Ce3+ triply doped tellurite glass,” J. Non-Cryst. Solids 358(14), 1644–1648 (2012). [CrossRef]  

9. Y. Tian, T. Wei, X. F. Jing, J. J. Zhang, and S. Q. Xu, “Enhanced 2.7- and 2.9-μm emissions in Er3+/Ho3+ doped fluoride glasses sensitized by Pr3+ ions,” Mater. Res. Bull. 76, 67–71 (2016). [CrossRef]  

10. G. W. Tang, X. Wen, Q. Qian, T. T. Zhu, W. W. Liu, M. Sun, X. D. Chen, and Z. M. Yang, “Efficient 2.0 μm emission in Er3+/Ho3+ co-doped barium gallo-germanate glasses under different excitations for mid-infrared laser,” J. Alloys Compd. 664, 19–24 (2016). [CrossRef]  

11. K. Biswas, A. D. Sontakke, R. Sen, and K. Annapurna, “Enhanced 2 μm broad-band emission and NIR to visible frequency up-conversion from Ho3+/Yb3+ co-doped Bi2O3-GeO2-ZnO glasses,” Spectrochim. Acta A Mol. Biomol. Spectrosc. 112, 301–308 (2013). [CrossRef]   [PubMed]  

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

13. M. Wang, L. X. Yi, G. N. Wang, L. L. Hu, and J. J. Zhang, “2 μm emission performance in Ho3+ doped fluorophosphate glasses sensitized with Er3+ and Tm3+ under 800 nm excitation,” Solid State Commun. 149(29–30), 1216–1220 (2009). [CrossRef]  

14. Y. Lu, M. Z. Cai, R. J. Cao, Y. Tian, F. F. Huang, S. Q. Xu, and J. J. Zhang, “Ho3+ doped germanate-tellurite glass sensitized by Er3+ and Yb3+ for efficient 2.0 μm laser material,” Mater. Res. Bull. 84, 124–131 (2016). [CrossRef]  

15. Y. Z. Fang, G. Y. Zhao, J. Y. Xu, N. Zhang, Z. F. Ma, and L. L. Hu, “Energy transferand 1.8 μm emissioninYb3+/Tm3+ co-doped bismuth germanate glass,” Ceram. Int. 40(4), 6037–6043 (2014). [CrossRef]  

16. B. Huang, Y. X. Zhou, P. Cheng, Z. Z. Zhou, J. Li, and W. Jin, “Tm3+/Yb3+ co-doped tellurite glass with silver nanoparticles for 1.85 μm band laser material,” Opt. Mater. 60, 341–349 (2016). [CrossRef]  

17. B. Zhou, E. Y. B. Pun, H. Lin, D. L. Yang, and L. H. Huang, “Judd–Ofelt analysis, frequency upconversion, and infrared photoluminescence of Ho3+-doped and Ho3+/Yb3+-codoped lead bismuth gallate oxide glasses,” J. Appl. Phys. 106(10), 103105 (2009). [CrossRef]  

18. X. Y. Wang, H. Lin, D. L. Yang, L. Lin, and E. Y. B. Pun, “Optical transitions and upconversion fluorescence in Ho3+/Yb3+ doped bismuth tellurite glasses,” J. Appl. Phys. 101(11), 113535 (2007). [CrossRef]  

19. D. M. Shi, Y. G. Zhao, and X. F. Wang, “Enhanced 1.53-μm and lowered upconversion luminescence in Er3+-doped Ga2O3–GeO2–Bi2O3–Na2O glass by codoping rare earths,” Physica B 406(13), 2588–2593 (2011). [CrossRef]  

20. Y. G. Choi, K. H. Kim, S. H. Park, and J. Heo, “Comparative study of energy transfers from Er3+ to Ce3+ in tellurite and sulfide glasses under 980 nm excitation,” J. Appl. Phys. 88(7), 3832–3839 (2000). [CrossRef]  

21. T. Ying, R. R. Xu, L. Y. Zhang, L. L. Hu, and J. J. Zhang, “Enhanced effect of Ce3+ ions on 2 μm emission and energy transfer properties in Yb3+/Ho3+ doped fluorophosphate glasses,” J. Appl. Phys. 109(8), 083535 (2011). [CrossRef]  

22. J. L. He, Z. G. Zhou, H. Zhan, A. D. Zhang, and A. X. Lin, “2.85 µm fluorescence of Ho-doped water-free fluorotellurite glasses,” J. Lumin. 145, 507–511 (2014). [CrossRef]  

23. B. Richards, Y. Tsang, D. Binks, J. Lousteau, and A. Jha, “Efficient ~2 μm Tm3+-doped tellurite fiber laser,” Opt. Lett. 33(4), 402–404 (2008). [CrossRef]   [PubMed]  

24. S. D. Jackson and S. Mossman, “Diode-cladding-pumped Yb3+, Ho3+-doped silica fiber laser operating at 2.1-microm,” Appl. Opt. 42(18), 3546–3549 (2003). [CrossRef]   [PubMed]  

25. Y.-W. Lee, H.-W. Chien, C.-H. Cho, J.-Z. Chen, J.-S. Chang, and S. B. Jiang, “Heavily Tm3+-doped silicate fiber for high-gain fiber amplifiers,” Fibers. 1(3), 82–92 (2013). [CrossRef]  

26. S. S. Bayya, G. D. Chin, J. S. Sanghera, and I. D. Aggarwal, “VIS-IR Transmitting windows,” Proc. SPIE 5786, 262–271 (2005). [CrossRef]  

27. T. Wei, F. Z. Chen, Y. Tian, and S. Q. Xu, “Broadband 1.53 μm emission property in Er3+ doped germa-silicate glass for potential optical amplifier,” Opt. Commun. 315, 199–203 (2014). [CrossRef]  

28. T. Wei, C. Tian, M. Z. Cai, Y. Tian, X. F. Jing, J. J. Zhang, and S. Q. Xu, “Broadband 2 μm fluorescence and energy transfer evaluation in Ho3+/Er3+ codoped germanosilicate glass,” J. Quant. Spectrosc. RA. 161, 95–104 (2015). [CrossRef]  

29. X. Liu, Y. Teng, Y. Zhuang, J. Xie, Y. Qiao, G. Dong, D. Chen, and J. Qiu, “Broadband conversion of visible light to near-infrared emission by Ce3+, Yb3+-codoped yttrium aluminum garnet,” Opt. Lett. 34(22), 3565–3567 (2009). [CrossRef]   [PubMed]  

30. T. Sasikala, L. R. Moorthy, K. Pavani, and T. Chengaiah, “Spectroscopic properties of Er3+ and Ce3+ co-doped tellurite glasses,” J. Alloys Compd. 542(25), 271–275 (2012). [CrossRef]  

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

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

33. R. R. Xu, Y. Tian, L. L. Hu, and J. J. Zhang, “Origin of 2.7 μm luminescence and energy transfer process of Er3+: 4I11/24I13/2 transition in Er3+/Yb3+ doped germanate glasses,” Appl. Phys. (Berl.) 111(3), 033524 (2012). [CrossRef]  

34. Q. Zhang, J. Ding, Y. L. Shen, G. Zhang, G. Lin, J. R. Qiu, and D. P. 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]  

35. P. Babu, H. J. Seo, C. R. Kesavulu, K. H. Jang, and C. K. Jayasankar, “Thermal and optical properties of Er3+-doped oxyfluorotellurite glasses,” J. Lumin. 129(5), 444–448 (2009). [CrossRef]  

36. M. Li, X. Q. Liu, Y. Y. Guo, L. L. Hu, and J. J. Zhang, “Energy transfer characteristics of silicate glass doped with Er3+, Tm3+, and Ho3+ for 2 μm emission,” J. Appl. Phys. 114(24), 243501 (2013). [CrossRef]  

37. P. Nachimuthu and R. Jagannathan, “Judd–Ofelt parameters, hypersensitivity, and emission characteristics of Ln3+ (Nd3+, Ho3+, and Er3+) ions doped in PbO–PbF2 glasses,” J. Am. Ceram. Soc. 82(2), 387–392 (1999). [CrossRef]  

38. L. X. Yi, M. Wang, S. Y. Feng, Y. K. Chen, G. N. Wang, L. L. Hu, and J. J. Zhang, “Emissions properties of Ho3+: 5I75I8 transition sensitized by Er3+ and Yb3+ in fluorophosphate glasses,” Opt. Mater. 31(11), 1586–1590 (2009). [CrossRef]  

39. Y. Tian, L. Y. Zhang, S. Y. Feng, R. R. Xu, L. L. Hu, and J. J. Zhang, “2 μm Emission of Ho3+-doped fluorophosphate glass sensitized by Yb3+,” Opt. Mater. 32(11), 1508–1513 (2010). [CrossRef]  

40. R. Cao, M. Cai, Y. Lu, Y. Tian, F. Huang, S. Xu, and J. Zhang, “Ho3+/Yb3+ codoped silicate glasses for 2 μm emission performances,” Appl. Opt. 55(8), 2065–2070 (2016). [CrossRef]   [PubMed]  

41. J. T. Fan, Y. Y. Fan, Y. Yang, D. P. Chen, L. Calveza, X. H. Zhang, and L. Zhang, “Spectroscopic properties and energy transfer in Yb3+–Ho3+ co-doped germanate glass emitting at 2.0 μm,” J. Non-Cryst. Solids 357(11–12), 2431–2434 (2011). [CrossRef]  

42. P. R. Watekar, S. Ju, and W.-T. Han, “Optical properties of Ho-doped alumino–germano-silica glass optical fiber,” J. Non-Cryst. Solids 354(14), 1453–1459 (2008). [CrossRef]  

43. J. Heo, Y. B. Shin, and J. N. Jang, “Spectroscopic analysis of Tm3+ in PbO-Bi2O3-Ga2O3 glass,” Appl. Opt. 34(21), 4284–4289 (1995). [CrossRef]   [PubMed]  

44. G. J. Gao, G. N. Wang, C. L. Yu, J. J. Zhang, and L. L. Hu, “Investigation of 2.0 μm emission in Tm3+ and Ho3+ co-doped oxyfluoride tellurite glass,” J. Lumin. 129(9), 1042–1047 (2009). [CrossRef]  

45. Y. Tian, R. R. Xu, L. L. Hu, and J. J. Zhang, “2.7 μm fluorescence radiative dynamics and energy transfer between Er3+ and Tm3+ ions in fluoride glass under 800 nm and 980 nm excitation,” J. Quant. Spectrosc. RA. 113(1), 87–95 (2012). [CrossRef]  

46. S. A. Payne, L. Chase, L. K. Smith, W. L. Kway, and W. F. Krupke, “Infrared cross-section measurements for crystals doped with Er3+, Tm3+, and Ho3+,” IEEE J. Quantum Electron. 28(11), 2619–2630 (1992). [CrossRef]  

47. D. E. McCumber, “Einstein relations connecting broadband emission and absorption spectra,” Phys. Rev. 136(4A), A954–A957 (1964). [CrossRef]  

48. 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]  

49. X. Q. Liu, F. F. Huang, J. M. Cheng, X. K. Fan, S. Gao, J. J. Zhang, L. L. Hu, and D. P. Chen, “Investigation on Er3+/Ho3+ co-doped silicate glass for ~2 μm fiber lasers,” J. Lumin. 162, 197–202 (2015). [CrossRef]  

50. T. Miyakawa and D. L. Dexter, “Phonon sidebands, multiphonon relaxation of excited states, and phonon-assisted energy transfer between ions in solids,” Phys. Rev. B 1(7), 2961–2969 (1970). [CrossRef]  

51. L. V. G. Tarelho, L. Gomes, and I. M. Ranieri, “Determination of microscopic parameters for nonresonant energy-transfer processes in rare-earth- doped crystals,” Phys. Rev. B 56(22), 14344–14351 (1997). [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 (5)

Fig. 1
Fig. 1 Absorption spectra of H, YHC0 and YHC0.1 glasses.
Fig. 2
Fig. 2 Fluorescence spectra of YHC (0, 0.1, 0.25, 0.5) glasses. Insets in Fig. 2 show the decay curves of the YHC0 and YHC0.1 at 1200 and 2055 nm.
Fig. 3
Fig. 3 Upconversion spectra of YHC (0, 0.1, 0.25, 0.5) glasses.
Fig. 4
Fig. 4 Energy level diagrams and energy transfer mechanism among Yb3+, Ho3+ and Ce3+ ions.
Fig. 5
Fig. 5 (a) The calculated absorption and emission cross sections (b) Gain coefficient for prepared glass (YHC0.1).

Tables (4)

Tables Icon

Table 1 Measured (fmea) and calculated (fcal) oscillator strengths of H sample and compared with other systems.

Tables Icon

Table 2 J–O intensity parameters Ωλ (λ = 2, 4, 6) ( × 10−20 cm2) of Ho3+ ions in various glass hosts.

Tables Icon

Table 3 The calculated radiative transition probability (Arad), fluorescence branching ratios (β) and radiative lifetimes (τrad) of Ho3+ ions for different transitions.

Tables Icon

Table 4 Calculated microscopic parameters CD-A ( × 10−40 cm6/s) for Yb3+→Ho3+ energy transfer in YHC0 and YHC0.1 glasses. The number # of phonons necessary to assist the energy transfer process is also revealed with their percent contributions.

Equations (12)

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

σ e m ( λ ) = λ 4 A r a d 8 π c n 2 × λ I ( λ ) λ I ( λ ) d λ
σ e m ( λ ) = σ a b s ( λ ) × Z l Z u × exp [ h c k t × ( 1 λ Z L 1 λ ) ]
G ( λ ) = N [ P σ e m ( λ ) ( 1 P ) σ a b s ( λ ) ]
W D A = ( 2 π ) | H D A | 2 S D A N
S D A N e ( S 0 D + S 0 A ) × [ ( S 0 D + S 0 A ) N N ! ] S D A ( 0 , 0 , E ) δ ( N , Δ E / w 0 )
S D A ( m , 0 , E ) = [ S 0 m m ! e S 0 g e m i s D ( E Δ E ) ] g a b s A ( E ) d E
σ e m i s ( m p h o n o n ) D = σ e m i s D ( λ m + ) S 0 m e S 0 m ! ( n ¯ + 1 ) m σ e m i s D ( E E 1 )
σ a b s ( k p h o n o n ) A = σ a b s A ( λ k ) S 0 k e S 0 k ! ( n ¯ ) k σ a b s A ( E + E 2 )
λ m + = 1 / ( 1 / λ m w 0 )
λ k = 1 / ( 1 / λ + k w 0 )
W D A ( R ) = 6 c g l o w D ( 2 π ) 4 n 2 R 6 g u p D m = 0 e ( 2 n ¯ + 1 ) S 0 S 0 m m ! × ( n ¯ + 1 ) m σ e m i s D ( λ m + ) σ a b s A ( λ ) d λ = C D A R 6
C D A = 6 c g l o w D ( 2 π ) 4 n 2 g u p D m = 0 e ( 2 n ¯ + 1 ) S 0 S 0 m m ! × ( n ¯ + 1 ) m σ e m i s D ( λ m + ) σ a b s A ( λ ) d 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.