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

Efficient ~2.0 μm emission from Ho3+ doped tellurite glass sensitized by Yb3+ ions: Judd-Ofelt analysis and energy transfer mechanism [Invited]

Open Access Open Access

Abstract

The ~2.0μm emission characteristics of Ho3+ both by direct excitation and through Yb3+ sensitization in barium-tellurite glass are reported. The radiative properties of active ions have been evaluated by applying Judd-Ofelt theory on the measured absorption spectrum. A significant enhancement of Ho3+ emission (2.0μm) observed with 12 fold decrease of Yb3+ emission (1008nm) in co-doped sample entrenched the efficient energy transfer from Yb3+:2F5/2→Ho3+:5I6. The host phonon assistance in the energy transfer process has been conferred by using Dexter model. Comparatively better emission properties (Arad, Δλeff, σem) reveal that, the present material could be promising for laser emission at ~2.0μm.

©2011 Optical Society of America

1. Introduction

Solid-state lasers emitting in the ~2.0 μm eye-safe region are promising candidates for various applications in laser medicine surgery, remote chemical sensing, eyesafe laser radar and monitoring of atmospheric pollutions [15]. Ever since Johnson [6] has reported the first laser action at 2.0 μm from Ho3+ doped CaWO4 crystal in 1962, an intense level of research has been focused on the spectroscopic and lasing properties of Ho3+ doped various crystals and glasses [7,8]. To date, ~2 μm solid-state lasers have utilised silica, fluoride and germanate glasses in addition to various crystals as the rare-earth ion host materials. However, yet another material, in which the study of solid-state laser performance is worthwhile, is the tellurite family of glasses. Tellurite glasses exhibit high linear and nonlinear refractive indices, good resistance to corrosion, relatively high mechanical stability, low cut-off phonon energy of around 750 cm−1 among oxide systems which makes their transmission range extend into the mid-IR up to about 5 - 6 μm and it has good capability to accommodate lanthanide dopants. The low phonon energy and broad transmission range of tellurite glasses allow the observation of laser emissions from rare earth ions in a wide optical range. These features, coupled with good resistance to corrosion and high mechanical stability, make tellurite glasses promising hosts for mid-IR lasers at ~2.0 μm wavelength. The most common sensitizer for Ho3+ doped glasses/crystals is Tm3+. The great advantage of sensitization with Tm3+ is the quantum efficiency of 2 that leads to a theoretical maximum achievable efficiency of 75% under 785 nm pumping with lasing at 2.1 μm [9]. Unfortunately Tm3+ has an absorption maximum in at 785 nm which does not correspond well to the emission wavelength of common and powerful laser diodes. On the other hand Yb3+ is also an efficient sensitizer for Ho3+ glass/crystal lasers [10] which has a broad absorption band lying in the emission range of commercially available InGaAs diode lasers that can be used as excitation sources to obtain an efficient energy transfer from Yb3+to Ho3+.

There are innumerable studies on Er3+/ Tm3+ or Yb3+ co-doped tellurite glasses of different compositions as 1.54 μm or 1.4 μm amplifiers for optical communication [11,12] and also as IR to visible frequency upconversion host materials [13,14]. However, there are very few reports on NIR emission at ~2 μm from Ho3+ singly or Yb3+ co-doped tellurite glasses [15]. Earlier, Debnath et. al., have reported upconversion visible emission and spectroscopic properties of Er3+ doped barium tellurite glasses [14,16] but there are no investigations on Ho3+ fluorescence in this glass system. Hence the present paper mainly aims to report on the studies of the 2.0 μm emission characteristics of Ho3+ ions both by direct excitation and by sensitised excitation through energy transfer from Yb3+ ions in the co-doped barium-tellurite glass. Additionally, a systematic characterisation of spectroscopic properties of Ho3+ ions in the present glass by applying the Judd-Ofelt theory has been dealt along with the physical, optical and structural analysis.

2. Sample Preparation and Characterization

The Ho3+/Yb3+ co-doped, Yb3+ singly doped and undoped barium tellurite glasses having chemical composition (mol %) 80 TeO2 – 15 (BaF2 + BaO) – (5-x-y) La2O3 – x Ho2O3 – y Yb2O3, (where x = 0, 1 and y = 0, 1) were prepared by melt quenching method. The size of batches was approximately 10 g, and all the chemicals used were of AR grade (Sigma-Aldrich) with 99.99% purity. The batches were melted using pure platinum crucible as a container in an electrical furnace at 700-750°C for 1 hour with intermittent stirring with thin platinum rod for attaining the homogeneity. The homogeneous melt was poured onto a pre heated graphite mould to form clear glass followed by annealing at 230-250 °C for 12 hours to avoid thermal stress.

The density of the glasses was measured by Archimedes method using distilled water as buoyancy liquid at room temperature. Refractive indices of glasses were measured at five different wavelengths (473 nm, 532 nm, 632.8 nm, 1060 nm and 1552 nm) on Metricon M2010 Prism Coupler equipped with respective laser sources. The FTIR reflectance spectra of the glasses were recorded on a FTIR spectrometer (Model: Series 1615, Perkin Elmer, Norwalk, CT) in the wavenumber range of 400–2000 cm−1 at a 15° angle of incidence. The measurement of room temperature absorption spectra was performed with optically polished plate shaped samples of co-doped glass (average thickness 1.58 mm) and base glass as a reference on an UV-VIS-NIR absorption spectrophotometer (Model: 3001, Shimadzu, Japan). The emission, excitation spectra and fluorescence decay kinetics of the samples were recorded on spectrofluorimeter (Model: Quantum Master enhanced NIR from PTI, USA) fitted with double monochromators on both excitation and emission channels. The instrument was equipped with LN2 cooled gated NIR photo-multiplier tube (Model: NIR-PMT-R 1.7, Hamamatsu) as well as InGaS detectors for acquiring both steady state spectra and phosphorescence decay. For decay measurements, a 60 W Xenon flash lamp was employed as an excitation source.

3. Results and Discussion

3.1. Physical, Optical and Structural Properties

The measured densities of Yb3+/Ho3+ co-doped and Yb3+ singly doped barium-tellurite glasses along with the average molecular weight of the respective glasses have been used to estimate the rare earth ion concentration (NRE) and the results are presented in Table 1 . The measured refractive indices (n) of both the glasses at five different wavelengths were fitted to sellmeier dispersion equation to determine the linear refractive indices at nD, nF and nC to calculate the abbe number (ν). The non-linear refractive index (n2) is an important parameter for Raman amplifier application has also been computed for both the glasses by using relevant expression [17] and used for the estimation of the third-order nonlinear susceptibility (χ3) yet another important optical parameter for ultra-fast optical switching in optical communication systems. The computed values of all the physical and optical parameters were presented in Table. 1.Figure 1 presents the FTIR reflectance spectra of Yb3+ singly doped and Yb3+-Ho3+ co-doped tellurite glasses. For both the samples, the spectra comprise identical nature suggesting that the incorporation of Ho3+ in glass network does not bring significant change in the glass structure. The spectra revealed a broad band with superimposition of three reflectance bandscentred at 732 cm−1, 680 cm−1 and 595 cm−1 which have been identified due to the Te – O- stretching vibrational band in [TeO3] trigonal pyramid, Te – Oaxial stretching vibrational band of [TeO4] trigonal bi-pyramid and [TeO3+1] polyhedra respectively [1820]. It is observed that, in the present barium tellurite glass these vibrational bands have shifted to lower wavenumber compared to the other tellurite glass systems, this may be due to the fluoride anion effect in the network. Further, it is interesting to note that in this system the magnitude of stretching vibrational mode arising from non-bridging oxygen units belonging to TeO3 trigonal pyramids (732 cm−1) is at lowest and that of connecting bridging bonds of TeO4 bi-pyramides (680 cm−1 and 595 cm−1) is at highest. This is considered to be good indication of stability or resistant to devitrification. From the FTIR reflectance spectra it is clear that, the vibrational mode at 595 cm−1 is found to be dominant over the other modes and hence it can be considered as the phonon energy of the present glass. To the authors’ knowledge, this is the lowest phonon energy reported so far for any tellurite glass system.

Tables Icon

Table 1. Physical and Optical properties of Yb3+/Ho3+ co-doped and Yb3+ singly doped tellurite glass: Density (ρ), Rare earth ion concentration (NRE), linear refractive indices (nD, nF, nC), Abbe number (ν), Non-linear refractive index (n2), Third-order nonlinear susceptibility (χ3).

 figure: Fig. 1

Fig. 1 FTIR reflectance spectra of tellurite glasses.

Download Full Size | PDF

3.2. Optical Absorption Spectra and Judd Ofelt analysis

The room temperature absorption spectra of Yb3+/Ho3+ co-doped tellurite glass sample with base glass corrected and uncorrected along with base glass sample have been presented in Fig. 2 . The spectrum of base glass shows good transparency with no characteristic peaks in the measured wavelength range of 420 nm – 2500 nm having strong band gap absorption starting at 410 nm. While the rare earth co-doped sample exhibits a number of distinct absorption bands in the Vis-IR range at 420, 452, 488, 540, 644, 1154 and 1950 nm for Ho3+ ions and at 978nm for Yb3+ ions. Depending upon their spectral peak energies [21,22], peaks are assigned to the transitions from ground state 5I8 to (5G, 3G) 5, 5G6, 5F3, (5F4 + 5S2), 5I6 and 5I7 excited states of Ho3+ ions and 2F7/22F5/2 for Yb3+ ions respectively. It is observed that the intensity of the 5I85G6 transition is highest among the detected peaks. This is due to its hypersensitive nature in 4f 10 electronic configuration of Ho3+ ions with invariably obeying ΔS = 0, ΔL = 2 and ΔJ = 2 spectral rule possessing strong dependence on the ligand field environment surrounding the rare earth ion. In general, magnitude of the hypersensitivity of a transition mainly linked to the ligand covalency, crystal field symmetry and it is correlated strongly with the Ω2 intensity parameter.The Judd-Ofelt intensity parameters (Ω2, Ω4, Ω6) were calculated using relevant expressions [21], considering integrated absorption coefficient values of well defined absorption bands of Ho3+ ions along with the refractive indices of the glass at the concerned absorption peak wavelengths that derived from dispersion curve (Table 2 ). To calculate the integrated absorption coefficient of the bands, base glass corrected absorption spectrum has been used. A set of equations depending on the number of absorption bands considered for the simulation were generated from the measured electric dipole line strength (Sedmea)and by employing the least square fitting method on those equations, the J-O intensity parameters (Ωt = 2,4,6) are attained. The best-fitted values obtained are Ω2 = 12.73 × 10−20 cm2, Ω4 = 7.06 × 10−20 cm2, Ω6 = 3.94 × 10−20 cm2 respectively. By using these J-O intensity parameters the calculated electric dipole line strengths(Sedcal)for the respective absorption bands were determined. Among all the absorption transitions of Ho3+ ions, 5I85I7 transition has the considerable contribution of magnetic dipole nature. Hence magnetic dipole line strength (Smd) for this transition has also been calculated and the value is given in Table. 2. By using Sedmea and Sedcalvalues, measured (Pmea) and calculated (Pcal) oscillator strengths have been obtained for each transition and are listed in Table. 2. Further to validate the quality of fitting, the root mean square deviation for the measured and the calculated values of electric dipole line strengths (rms-ΔSed) and oscillator strengths (rms-ΔP) have been examined. The minimum the deviation, the more accurate are the obtained values. In the present case, the obtained results demonstrated good reliability of the data.

 figure: Fig. 2

Fig. 2 Absorption spectra of base glass corrected and not corrected Yb3+/Ho3+ co-doped tellurite glass along with base glass.

Download Full Size | PDF

Tables Icon

Table 2. Electric dipole line strengths (measured: Sedmea, calculated: Sedcal), magnetic dipole strength (Smd), Total Oscillator strengths (measured: Pmea, calculated: Pcal), refractive index (n) of different absorption transitions and (Ωt = 2,4,6) J-O intensity parameters of Ho3+ions in the co-doped glass.

Table 3 presents a comparative data of J-O intensity parameters of Ho3+ ions obtained in the present glass system with that of various crystal/glass hosts reported in the literature. In the present glass system the Ωt parameters follow the trend Ω2 > Ω4 > Ω6 which is similar to the most of the tellurite glasses. Generally, Ω2 is closely related to the ligand symmetry and degree of covalency of the host material. The larger Ω2 value represents a higher degree of covalency of the glass network bonding and lower symmetry [22]. The 2 µm emission transition 5I75I8 (Ho3+) is mainly affected by Ω6 intensity parameter due to its larger U(6) matrix element and as the radiative spontaneous emission probability (Ar) is proportionately dependent on Ω6 intensity parameter [23].

Tables Icon

Table 3. Comparative chart of J-O intensity parameters of Ho3+ ion in various hosts.

It should be mentioned that the Ω6 value in the present glass is higher than the other hosts except ZnBS glass. Similarly, it is observed that the spectroscopic quality factor (χ = Ω46), which is an important predictor of stimulated emission is found to be almost similar to any other crystal/glass except YAG crystal and chalcogenide glass where is it higher (Table 3).

Further, by using the Ωt values, various important radiative properties such as spontaneous emission probability (Ar), branching ratios (βR) and lifetime of the radiative transition (τr) were calculated and are listed in Table 4 . As the 5I75I8 transition of Ho3+ ions is magnetic dipole allowed and the contribution of Amd (77.15 s−1 in the present glass) has also been consider for estimating the spontaneous emission probability (Ar = ed + md).

Tables Icon

Table 4. Electric dipole line strength (Sed), Refractive index (n), Spontaneous emission probability (Ar), Radiative rate (ΣAr), Radiative lifetime (τr), Branching ratio (βR), for fluorescent levels of Ho3+ ions in the present tellurite glass.

3.3. Fluorescence Spectra

Figure 3 shows the NIR emission spectra of Yb3+/Ho3+ co-doped sample when excited by 980 nm through 2F5/2 excited state of sensitizer, Yb3+ ions and by direct Ho3+ ion excitation through 5F5 level with 658nm and 5I6 level with 1191nm wavelength. These respective excitation wavelengths have been chosen from the recorded excitation spectrum of co-doped sample by monitoring 2050 nm emission of Ho3+ ions as seen in inset of Fig. 3. The emission peaks detected at 1008, 1209, 1487 and 2050 nm are assigned to the transitions of Yb3+: 2F5/22F7/2, Ho3+: 5I65I8, Ho3+: 5I55I6 and Ho3+: 5I75I8 respectively.

 figure: Fig. 3

Fig. 3 Fluorescence spectra of Yb3+/Ho3+ co-doped tellurite glass (Inset: Excitation spectra for λemi = 2050nm.)

Download Full Size | PDF

Further to mention that the emission peak at 1487 nm corresponding to the transition Ho3+: 5I55I6 was observed only under direct excitation (by 658nm) and not detected under sensitizer excitation with 980 nm due to the fact that the energy is not sufficient to excite to higher level (5F5). From the emission spectra it can be clearly seen that the intensity of the peak at 2050 nm (Ho3+: 5I75I8) is highest among all the measured emissions transitions. Hence, its stimulated emission cross-section (σe) has been calculated by using the following expression [24]

σe(λ)=λp48πcn2ΔλeffAr
where λp is fluorescence peak wavelength, Δλeff is the effective band width, n is refractive index, c is velocity of light and Ar is its spontaneous emission probability.

The stimulated emission cross-section, effective band width and spontaneous emission probability of Ho3+: 5I75I8 transition in the present barium tellurite glass is presented in Table 5 in comparison with the other host systems. It is evident from this data that the effective band width of 5I75I8 (2050 nm) fluorescence band is higher (160 nm) when compared to other glass hosts which can be attributed to the existence of varied structural units TeO4, TeO3+1 and TeO3 of tellurite glass. The different Te-O bond lengths associated with these three different structural units result in multiplicity of ion-host field strengths and therefore, yield a range of electro-static fields around a rare earth ion in a tellurite glass leading to an inhomogeneous broadening of the fluorescence band [25]. This could result in an enhanced tuning range for solid state lasers utilising a tellurite glass material. Also, it is observed that, the value of stimulated emission cross section (σe) of this transition in present host has the higher (1.45 × 10−20 cm2) when compared to other glasses systems (Table 5) but, slightly lesser than that of chalcogenide glass (1.54 × 10−20 cm2) [26].

Tables Icon

Table 5. Comparative chart of peak wavelength (λP), concentration of Ho2O3 (in wt%), radiative transition probability (Ar), effective bandwidth (Δλp) and stimulated emission cross section (σe) for ~2.0µm emission of Ho3+ ions in various glass hosts.

The larger emission cross-section is mainly due to the high spontaneous emission probability resulting from high refractive index. From Table 5 it is worthwhile to notice that the value of effective band width is higher than any other host material, which is highly useful for laser tuning. Another interesting observation was that, the emission intensity of Ho3+: 5I75I8 had increased 8 fold under the excitation with 980 nm through the energy level 2F5/2 of Yb3+ ions when compared to the direct excitations of Ho3+. This is because of high absorption cross section of Yb3+ ions (σa = 1.889 × 10−20cm2 in the present glass) accompanied with efficient sensitised energy transfer from Yb3+ to Ho3+ ions.

3.4. Energy transfer mechanism and Decay analysis

In order to understand the energy transfer process from Yb3+ to Ho3+, the emission characteristics of Yb3+ ions with and without Ho3+ ions, have been examined in the glass system undertaken. Figure 4 compares the room temperature emission spectra of Yb3+ singly doped and co-doped samples under 924 nm excitation. The second intense excitation peak at924 nm has been selected instead of 980 nm to excite upon Yb3+ ions, only to obtain a full gaussian shape of spectral transition 2F5/22F7/2 at 1008 nm and also because of relatively small energy separation between excited state multiplets at 924 nm and 980 nm. From this Fig. 4, it is clear that the relative peak intensity of Yb3+ at 1008 nm has significantly decreased up to 12 folds in the Yb3+/Ho3+ co-doped tellurite glass compared to Yb3+ singly doped glass, showing an evidence of energy transfer from Yb3+: 2F5/2 → Ho3+: 5I6. The partial energy level diagram shown in Fig. 5 explains the mechanism involved in the energy transfer based emission process.

 figure: Fig. 4

Fig. 4 Emission spectra of Sensitizer, Yb3+ ions in absence (black) and in the presence (red) of activator, Ho3+ ions, (Inset: Decay profiles for the same.)

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Partial energy level diagram of Yb3+/Ho3+ co-doped tellurite glass showing the energy transfer mechanism under 980nm (red) and direct 1191nm, 658nm excitations (black)

Download Full Size | PDF

It is observed that, the measured fluorescence decay of Yb3+ emission is independent of excitation wavelength (980 nm/954 nm/924 nm). By studying the decay kinetics of Yb3+ ions in the presence and absence of Ho3+ ions, the energy transfer rate (WET) and the energy transfer efficiency (ηET) were estimated by using the following expressions [27]

WET=1τ1τ0,
ηET=1ττ0,
where, τ, τ0 are lifetime of donor, Yb3+ ions in the presence and absence of acceptor, Ho3+ ions. The derived energy transfer rate is found to be 9557.3 s−1 and with the energy transfer efficiency of 85.8%. This is highly beneficial for the design of 2.0 µm emitting laser under readily available high power, compact diode laser pumping. Although a two photon assisted upconversion emissions have been observed at 550 nm and 665 nm, the relative intensities were poor compared to 2050 nm emission in the present tellurite glass host.

3.5. Phonon assisted energy transfer process

According to the Forster’s resonant energy transfer (FRET) theory, the extent of energy transfer depends on the spectral overlap of donor’s emission with acceptor’s absorption. Hence, in the case of resonant energy transfer among donor-acceptor, the energy transfer probability (PET) exhibits a linear relationship with the spectral overlap integral as [28],

PETfD(E)fA(E)E2dE,
where, fD(E), fA(E) are the line-shape functions of donor’s emission and acceptor’s absorption respectively. In the present Yb3+ - Ho3+ doped tellurite glass, the spectral overlap between Yb3+ emission and Ho3+ absorption is very poor with an energy gap of about 1500 cm−1 due to the non-resonant transitions. However the observed efficient energy transfer with a transfer efficiency of 85% despite meagre spectral overlap suggests that the energy transfer in present glass system may be assisted with host phonons. For such non-resonant energy transfer, the Dexter model in Eq. (4) is generalized for phonon assisted energy transfer process by considering the energy of phonons involved in the energy transfer. Thus, transfer probability (PET) can be estimated from the phonon modified spectral overlap integral, I(Eph) as given by [28],
PETI(Eph)=eEphkBTeEphkBT1fD(EEph)fA(E)E2dE,
where, Eph is the host phonon energy, kB is Boltzmann constant and T is absolute temperature. Based on the Yb3+ emission and Ho3+ absorption cross-section spectra, the energy transfer probability has been calculated as a function of phonon energy in the range of 0 – 2700 cm−1 and presented in Fig. 6 . It can be seen that the normalized energy transfer probability increases with an increase in phonon energy and it reaches a maximum for the phonons with energy of around 1480 cm−1. For further increase in the phonon energy, the energy transfer probability decreases and diminishes. Thus, it is evident that, the energy difference between the energy levels of 2F5/2 (Yb3+) and 5I6 (Ho3+) has been bridged by the host phonons for an efficient energy transfer. As the phonon energy of present telluride glasses is 595 cm−1, about two to three phonons are required to bridge the energy gap. Though the high phonon energy hosts like silicate, phosphate or borate are suitable to promote the energy transfer from Yb3+ to Ho3+ with less number of phonons but they also possess higher non-radiative relaxation of 5I7 level of Ho3+ leading to less probability of 2 μm emission.

 figure: Fig. 6

Fig. 6 Energy transfer probability (Vs) Phonon energy in the co-doped glass.

Download Full Size | PDF

3.6. Energy transfer microparameters

The energy transfer microparameters for a donor - acceptor energy transfer can be estimated using Forster’s spectral overlap model given by [29],

CDA=3c8π4n2σemD(λ)σabsA(λ)dλ.
However, in present Yb3+ - Ho3+ telluride glasses, the energy transfer mechanism is phonon assisted and it has to be taken into consideration while calculating the energy transfer microparameters. It can be done by constructing the Stokes phonon sidebands to the absorption and emission cross section spectra from an exponential law of Auzel as given below [30],
σStokes=σelectexp(αSΔE),
where, ΔE is the energy mismatch between electronic and vibronic transitions and αS is the host dependent parameter for Stokes transitions represented as,
αS=(hν)1(ln{(N¯/S0)[1exp(hνmax/kT)}1),
where, N¯is the number of phonons required for bridging the energy gap, S0 is the electron-phonon coupling constant (~0.04), max is the maximum phonon energy of host and k is Boltzmann constant. An average two and half phonons assistance has been assumed in energy transfer for constructing the phonon sidebands. Figure 7 presents the emission and absorption cross-section spectra of Yb3+ and Ho3+ ions respectively by using Eq. (7) with corresponding phonon sidebands. By applying Eq. (6) to the spectral overlap function in Fig. 7, the energy transfer microparameter has been calculated and found to be 1.39 × 10−40 cm6.sec−1. Such a reasonably good value of microparameter suggests the host phonon contribution in the energy transfer process.

 figure: Fig. 7

Fig. 7 Emission and absorption cross-section spectra of Yb3+ and Ho3+ ions with corresponding phonon sidebands

Download Full Size | PDF

4. Conclusion

A detailed spectroscopic analysis has been performed by applying the standard J-O theory and the energy transfer mechanism involved in 2.0µm emission of Ho3+ under Yb3+ sensitization upon 980 nm excitation in Yb3+/Ho3+ co-doped tellurite glass has satisfactorily been explained. The evaluated J-O intensity parameters have been used to estimate the oscillator strengths, radiative properties of emission transitions and also the spectroscopic quality factor which is an important predictor in understanding the suitability of this material as a laser material. Further, the important laser parameters such as the stimulated emission cross-section, effective band width and spontaneous emission probability of the Ho3+: 5I75I8 transition in the present glass have been found to be comparatively high. The energy transfer rate of 9557.3 s−1 with 85.8% efficiency and 8 fold increase in the 2.0µm emission under 980 nm excitation clearly indicates the sensitizer excitation is more effective than the direct excitation. A comparative assessment of the data suggests that the present glass having got low energy (595 cm−1) phonons could be identified as a promising optical material for an efficient NIR emission at 2.0 µm.

Acknowledgment

The authors would like to thank the Director, CGCRI for his continued support and encouragement. This work was carried out under an In-house project OLP 0288.

References and links

1. A. Godard, “Infrared (2–12 μm) solid-state laser sources: a review,” C. R. Phys. 8(10), 1100–1128 (2007). [CrossRef]  

2. P. Crump, S Patterson, W Dong, M Grimshaw, J Wang, S Zhang, S Elim, M Bougher, J. Patterson, S Das, D Wise, T Matson, D Balsley, J Bell, M DeVito, and R Martinsen, “Room temperature high power mid-IR diode laser bars for atmospheric sensing applications,” Proc. SPIE 6552(655214), 1–6 (2007).

3. F. Cornacchia, A. Toncelli, and M. Tonelli, “2-μm2-μm lasers with fluoride crystals: Research and development,” Prog. Quantum Electron. 33(2-4), 61–109 (2009). [CrossRef]  

4. B. Richards, A. Jha, Y. Tsang, D. Binks, J. Lousteau, F. Fusari, A. Lagatsky, C. Brown, and W. Sibbett, “Tellurite glass lasers operating close to 2 µm,” Laser Phys. Lett. 7(3), 177–193 (2010). [CrossRef]  

5. S. W. Henderson, C. P. Hale, J. R. Magee, M. J. Kavaya, and A. V. Huffaker, “Eye-safe coherent laser radar system at 21 µm using Tm,Ho:YAG lasers,” Opt. Lett. 16(10), 773–775 (1991). [CrossRef]   [PubMed]  

6. L. J. Johson, G. D. Boyd, and K. Nassau, “Optical maser characteristics of Ho3 + in CaWO4,” Proc IAE 50, 87–90 (1962).

7. B. M. Walsh, N. P. Barnes, M. Petros, J. Yu, and U. N. Singh, “Spectroscopy and modeling of solid state lanthanide lasers: Application to trivalent Tm[sup 3+] and Ho[sup 3+] in YLiF[sub 4] and LuLiF[sub 4],” J. Appl. Phys. 95(7), 3255–3271 (2004). [CrossRef]  

8. L. V. Klimov, I. A. Shcherbakov and V. B. Tsvetkov, “Room Temperature 2-µm Laser Action of Ho3+-Doped YSGG, GSAG, YSAG and YAG Crystals,” OSA proc. Adv. Solid State Lasers 15, 419–423 (1993).

9. Th. Rothacher, W. Lüthy, and H. P. Weber, “Spectral properties of a Tm:Ho:YAG laser in active mirror configuration,” Appl. Phys. B 66(5), 543–546 (1998). [CrossRef]  

10. Th. Rothacher, W. Lüthy, and H. P. Weber, “Diode pumping and laser properties of Yb:Ho:YAG,” Opt. Commun. 155(1-3), 68–72 (1998). [CrossRef]  

11. Y. Tsang, B. Richards, D. Binks, J. Lousteau, and A. Jha, “A Yb3+/Tm3+/Ho3+ triply-doped tellurite fibre laser,” Opt. Express 16(14), 10690–10695 (2008). [CrossRef]   [PubMed]  

12. Lihui Huang, Shaoxiong Shen, and Animesh Jha, “Near infrared spectroscopic investigation of Tm3+-Yb3+ co-doped tellurite glass”, J. Non-Cryst. Solids 345&346, 349–353 (2004).

13. Y. Dwivedi, A. Roy, and S. B. Rai, “Intense white upconversion emission in Pr/Er/Yb codoped tellurite glass,” J. Appl. Phys. 104(043509), 1–4 (2008).

14. R. Debnath, A. Ghosh, and S. Balaji, “Synthesis and luminescence properties of an (Er2Te4O11) nanocrystals dispersed highly efficient upconverting lead free tellurite glass,” Chem. Phys. Lett. 474(4-6), 331–335 (2009). [CrossRef]  

15. K. Li, Q. Zhang, S. Fan, L. Zhang, J. Zhang, and L. Hu, “Midinfrared luminescence and energy transfer characteristics of Ho3+/Yb3+ codoped lanthanum-tungsten-tellurite glasses,” Opt. Mater. 33(1), 31–35 (2010). [CrossRef]  

16. A. Ghosh and R. Debnath, “Judd-Ofelt analysis of Er3+ activated lead free fluoro-tellurite glass,” Opt. Mater. 31(4), 604–608 (2009). [CrossRef]  

17. N. L. Boling, A. J. Glass, and A. Owyoung, “Empirical relationship for predicting nonlinear refractive index changes in optical solids,” IEEE J. Quantum Electron. 14(8), 601–608 (1978). [CrossRef]  

18. K. Annapurna, R. Chakrabarti, and S. Buddhudu,“Absorption and emission spectral analysis of Pr3+: tellurite glasses,” J. Mater. Sci. 42(16), 6755–6761 (2007). [CrossRef]  

19. D. Tatar, G. Özen, F. B. Erim, and M. L. Övecoğlu, “Raman characterization and structural properties of the binary TeO2-WO3, TeO2-CdF2 and ternary TeO2-CdF2-WO3 glasses,” J. Raman. Spectrosc. 41, 797–807 (2010).

20. J. Lin, W. Huang, Z. Sun, C. S. Ray, and D. E. Day, “Structure and non-linear optical performance of TeO2-Nb2O5-ZnO glasses,” J. Non-Cryst. Solids 336(3), 189–194 (2004). [CrossRef]  

21. A. A. Kaminskii, Laser Crystals: Their Physics and Properties, Second Edition, (Springer-Verlag, 1990), pp. 156.

22. W. T. Carnall, P. R. Fields, and K. Rajnak, “Electronic energy levels in the trivalent lanthanide aquo ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+ and Tm3+,” J. Chem. Phys. 49(10), 4424–4442 (1968). [CrossRef]  

23. T. Schweizer, B. N. Samson, J. R. Hector, W. S. Brocklesby, D. W. Hewak, and D. N. Payne, “Infrared emission and ion–ion interactions in thulium- and terbium-doped gallium lanthanum sulfide glass,” J. Opt. Soc. Am. B 16(2), 308–316 (1999). [CrossRef]  

24. D. Yeh, R. Petrin, W. Sibley, V. Madigou, J. Adam, and M. Suscavage, “Energy transfer between Er3+ and Tm3+ ions in barium fluoride – thorium fluoride glass,” Phys. Rev. B 39(1), 80–90 (1989). [CrossRef]  

25. A. Jha, S. Shen, and M. Naftaly, “Structural origin of spectral broadening of 1.5-µm emission in Er3+-doped tellurite glasses,” Phys. Rev. B 62(10), 6215–6227 (2000). [CrossRef]  

26. Y. S. Kim, W. Y. Cho, Y. B. Shin, and J. Heo, “Emission characteristics of Ge-Ga-S glasses doped with Tm3+/Ho3+,” J. Non-Cryst. Solids 203, 176–181 (1996). [CrossRef]  

27. G. A. Hebbink, L. Grave, L. A. Woldering, D. N. Reinhoudt, and F. C. M. J. Van Veggel, “Unexpected Sensitization Efficiency of the Near-Infrared Nd3+, Er3+ and Yb3+ Emission by Fluorescein Compared to Eosin and Erythrosin,” J. Phys. Chem. A 107(14), 2483–2491 (2003). [CrossRef]  

28. D. Jaque, M. O. Ramirez, L. E. Bausà, J. García Solé, E. Cavalli, A. Speghini, and M. Bettinelli, “Nd3+ → Yb3+ energy transfer in YAl3(BO3)4 nonlinear laser crystal,” Phys. Rev. B 68,035118 (2003).

29. A. D. Sontakke, K. Biswas, A. K. Mandal, and K. Annapurna, “Concentration quenched luminescence and energy transfer analysis of Nd3+ doped Ba-Al-metaphosphate laser glasses,” Appl. Phys. B 101(1-2), 235–244 (2010). [CrossRef]  

30. M. C. Nostrand, R. H. Page, S. A. Payne, L. I. Isaenko, and A. P. Yelisseyev, “Optical properties of Dy3+- and Nd3+ -doped KPb2Cl5,” J. Opt. Soc. Am. B 18(3), 264–276 (2001). [CrossRef]  

31. K. L. Nash, R. C. Dennis, N. J. Ray, J. B. Gruber, and D. K. Sardar, “Absorption intensities, emission cross sections and crystal field analysis of selected intermanifold transitions of Ho3+ in Ho3+:Y2O3 nanocrystals,” J. Appl. Phys. 106(063117), 1–8 (2009).

32. 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]  

33. 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]  

34. 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]  

35. M. Wang, G. Wang, L. Yi, S. Li, L. Hu, and J. Zhang, “2-µm emission performance of Tm3+-Ho3+ co-doped tellurite glass,” Chin. Opt. Lett. 8(1), 78–81 (2010). [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 (7)

Fig. 1
Fig. 1 FTIR reflectance spectra of tellurite glasses.
Fig. 2
Fig. 2 Absorption spectra of base glass corrected and not corrected Yb3+/Ho3+ co-doped tellurite glass along with base glass.
Fig. 3
Fig. 3 Fluorescence spectra of Yb3+/Ho3+ co-doped tellurite glass (Inset: Excitation spectra for λemi = 2050nm.)
Fig. 4
Fig. 4 Emission spectra of Sensitizer, Yb3+ ions in absence (black) and in the presence (red) of activator, Ho3+ ions, (Inset: Decay profiles for the same.)
Fig. 5
Fig. 5 Partial energy level diagram of Yb3+/Ho3+ co-doped tellurite glass showing the energy transfer mechanism under 980nm (red) and direct 1191nm, 658nm excitations (black)
Fig. 6
Fig. 6 Energy transfer probability (Vs) Phonon energy in the co-doped glass.
Fig. 7
Fig. 7 Emission and absorption cross-section spectra of Yb3+ and Ho3+ ions with corresponding phonon sidebands

Tables (5)

Tables Icon

Table 1 Physical and Optical properties of Yb3+/Ho3+ co-doped and Yb3+ singly doped tellurite glass: Density (ρ), Rare earth ion concentration (NRE ), linear refractive indices (nD, nF, nC ), Abbe number (ν), Non-linear refractive index (n2 ), Third-order nonlinear susceptibility (χ3 ).

Tables Icon

Table 2 Electric dipole line strengths (measured: Sed mea, calculated: Sed cal), magnetic dipole strength (Smd), Total Oscillator strengths (measured: Pmea, calculated: Pcal), refractive index (n) of different absorption transitions and (Ωt = 2,4,6) J-O intensity parameters of Ho3+ions in the co-doped glass.

Tables Icon

Table 3 Comparative chart of J-O intensity parameters of Ho3+ ion in various hosts.

Tables Icon

Table 4 Electric dipole line strength (Sed), Refractive index (n), Spontaneous emission probability (Ar), Radiative rate (ΣAr), Radiative lifetime (τr), Branching ratio (βR), for fluorescent levels of Ho3+ ions in the present tellurite glass.

Tables Icon

Table 5 Comparative chart of peak wavelength (λP), concentration of Ho2O3 (in wt%), radiative transition probability (Ar), effective bandwidth (Δλp) and stimulated emission cross section (σe) for ~2.0µm emission of Ho3+ ions in various glass hosts.

Equations (8)

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

σ e ( λ ) = λ p 4 8 π c n 2 Δ λ e f f A r
W E T = 1 τ 1 τ 0 ,
η E T = 1 τ τ 0 ,
P E T f D ( E ) f A ( E ) E 2 d E ,
P E T I ( E p h ) = e E p h k B T e E p h k B T 1 f D ( E E p h ) f A ( E ) E 2 d E ,
C D A = 3 c 8 π 4 n 2 σ e m D ( λ ) σ a b s A ( λ ) d λ .
σ S t o k e s = σ e l e c t exp ( α S Δ E ) ,
α S = ( h ν ) 1 ( ln { ( N ¯ / S 0 ) [ 1 exp ( h ν max / k T ) } 1 ) ,
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.