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

Superbroadband near-IR photoluminescence from Pr3+-doped fluorotellurite glasses

Open Access Open Access

Abstract

Praseodymium(Pr3+)-doped fluorotellurite glasses were synthesized and broadband photoluminescence (PL) covering a wavelength range from 1.30 to 1.67 μm was observed under both 488 and 590 nm wavelength excitations. The broadband PL emission is mainly due to the radiative transition from the manifolds Pr3+: 1D2 to 1G4. The PL line-shape, band width, and lifetime were modified by the Pr3+ dopant concentration, and a quantum efficiency as high as 73.7% was achieved with Pr3+ dopant in a low concentration of 0.05 mol%. The good spectroscopic properties were also predicted by the Judd-Ofelt analysis, which indicates a stronger asymmetry and covalent bonding between the Pr3+ sites and the matrix lifgand field. The large stimulated emission cross-section, long measured lifetime, and broad emission bandwidth confirm the potential of the Pr3+-singly doped fluorotellurite glass as broadband luminescence sources for the broadband near-infrared optical amplifications and tunable lasers.

©2012 Optical Society of America

1. Introduction

Recently, the expansion of the transmission window of silica fibers to a broader range of 1.2~1.7 μm makes it attractive to explore superbroadband luminescence sources for the broadband near-infrared(IR) optical amplifiers and tunable lasers operating in this expanded low loss window [1,2]. Early investigations have demonstrated that rare earth ions of erbium (Er3+), thulium (Tm3+), holmium (Ho3+), and praseodymium (Pr3+) are good candidates for the optical amplifications at separate C-, L-, S-, E-, and O-bands in this expanded window, and considerable progress have been achieved [3]. To further improve the quantum efficiency, novel low phonon energy host (e.g., heavy metal oxide glasses and transparent glass-ceramics), rare earths codoping schemes (e.g., Er3+-Ce3+ for the C-band optical amplification), dual-pump configuration, and nano-structures (e.g., silver nanoparticles) have been investigated [410]. The codoping scheme has been also demonstrated to be an effective approach to achieve broader emission, and broadband near-IR emission covering S + L-band has already been obtained in Er3+-Tm3+ codoped systems [1116]. On the other hand, broadband near-IR luminescence peaking around 1.3 μm from the transition metal (TM) and heavy metal (HM) ions such as active bismuth (Bi), nickel (Ni), chromium (Cr), and lead (Pb) were investigated [1723], and their fiber/waveguide applications have been explored [22,23], although the luminescence origins for some of them (e.g., Bi+ and/or Bi5+) need further studies [24]. We have recently observed the broadband emissions around 1.20 and 1.47 μm wavelengths in Tm3+-singly doped tellurite glass under blue wavelength excitation [25,26], and further achieved the superbroadband near-IR luminescence covering 1.0-1.7 μm wavelength region from Tm-Bi codoped gallogermanate glasses taking advantage of Bi emissions located around 1.3 μm.

Among the rare earth ions, Pr3+ shows potential to yield luminescence at specific near-IR wavelength bands within the expanded low loss window considering its rich multiple energy levels. Apart from the 1.3 μm emission (Pr3+: 1G43H5) [10], an intense emission located at 1.6 μm from the Pr3+: 3F4,33H4 transition was also obtained in selenide glasses [27], which presents a promising candidate for the U-band optical amplification. We recently observed the superbroadband near-IR luminescence from Pr3+-singly doped bismuth gallate glasses [28]. However, the relatively stronger intrinsic absorption of the bismuth gallate glass at the short wavelength side (<500 nm) would result in a serious depression of the blue light pump efficiency. To overcome it, development of host matrix with higher transmission at blue wavelength region together with unique ligand field is necessary.

In this work, we report the observation of broadband near-IR luminescence extended from 1.30 to 1.67 μm in Pr3+-singly doped novel fluorotellurite glasses for the first time to our best knowledge. Tellurite-based glasses show high refractive index, low phonon energy, good mechanical properties and chemical durability [5,25]. More significantly, near-IR tellurite glass fiber/waveguide amplifiers and lasers have been reported [9,12,29]. Some of fluorides were added to modify the local ligand field between the Pr3+ and the host matrix. The high transparency at the short wavelength side (<500 nm) enables to improve the pump efficiency under blue excitation sources. The energy transfer processes involved are proposed and discussed, and the optical amplification is also evaluated.

2. Experimental

Pr3+-doped fluorotellurite glass samples were prepared by melting the well-mixed batches of high-purity chemicals with mol% composition of 3LaF3-4BaF2-4BaCO3-9ZnO-80TeO2-xPrF3 (x=0.05, 0.1, 0.3, 0.5, and 1.0) following the procedures described in Ref [25]. For comparison, a sample without Pr3+ doping was also prepared. The as-prepared glasses were cut and optically polished for optical measurements. The refractive indices of the glass samples were measured using a Metricon 2010 prism coupler and they are 2.068 and 2.013 at wavelengths of 632.8 and 1536 nm. The profile of the indices on wavelength was determined to be n(λ)=2.002+26333.26/λ2(the wavelength λ is in unit of nm) using Cauchy formula [30]. The Raman spectrum of undoped glass sample was measured by a HORIBA Jobin Yvon HR800 Raman spectrometer with a 488 nm laser excitation source. The absorption spectra were recorded using a Perkin Elmer UV-VIS-NIR Lambda 19 double beam spectrophotometer. The visible and infrared photoluminescence (PL) spectra were recorded using a photomultiplier tube (PMT) and a near-IR PMT detector, respectively. The excitation sources were tuned from a continuous xenon lamp by a monochromator. The photoluminescence excitation (PLE) spectra were recorded using the same setup with a continuous wavelength xenon lamp as the excitation source. The emission decay curves were recorded using the same setup with a flash xenon lamp as the excitation source. All the measurements were carried out at room temperature.

3. Results and discussion

Figure 1 shows the optical absorption spectrum of the 0.5 mol% Pr3+-doped glass sample. Absorption bands located at 447, 472, 485, 592, 1018, 1448, 1543, 1951, and 2331 nm are assigned to the transitions from the ground state 3H4 to specific excited states 3P2, 1I6 + 3P1, 3P0, 1D2, 1G4, 3F4, 3F3, 3F2, and 3H6, respectively. There are broad and intense absorption bands in 440-490 nm wavelength region, as shown in inset (a) of Fig. 1. This is different from that in bismuth gallate glass where only a very weak absorption band corresponding to the 3P0 was recorded in this wavelength region [28]. The higher transparency of fluorotellurite glass than the bismuth gallate glass in short wavelength side would enable to get more efficient emission under the blue light excitations. The absorption cross-section which determines the absorption efficiency of the active ions on the pump sources can be calculated by σabs(λ)=2.303E(λ)/Nd, where N, d, and E(λ) represent the ion density, the thickness, and the absorbance of the substrate glass, respectively. For instance, σabs was obtained to be 1.43 × 10−20 cm2 for the 3P0 absorption band, this value is much larger than that of Pr3+ in fluoroaluminate and ZBLAN glasses [31].

 figure: Fig. 1

Fig. 1 Optical absorption spectrum of (0.5 mol%)Pr3+-doped fluorotellurite glass. Inset (a) shows the detail of the absorption bands corresponding to the manifolds (3P2,1I6+3P1,3P0); Inset (b) shows the Raman spectrum (pink line) of undoped fluorotellurite glass with Gaussian fitting (green lines).

Download Full Size | PDF

The inset (b) of Fig. 1 shows the Raman spectrum of the undoped host measured using a 488 nm laser excitation. The maximum phonon energy is determined to be ~766 cm−1 by Gaussian fitting, which can be assigned to the stretching vibrations of TeO3+δ polyhedral, and/or tellurium and non-bridging oxygen atoms in TeO3 trigonal pyramid units [32,33]. This value is comparable to many other oxide-based tellurite and fluorotellurite glasses [3236]. The band with the most intense peak at 658 cm−1 is due to the anti-symmetric vibration modes of TeO4, and the bands in lower frequency region at 311 and 451 cm−1 can be attributed to the bending vibration bonds of O-Te-O/Te-O-Te linkages of TeO4 [34]. The Raman spectrum intensity of the peak at 658 cm−1 in comparison to that at ~766 cm−1 in fluorotellurite glass is slightly lower than those of the oxide-based tellurite glasses, this may be ascribed to the breakage effect of F ions on Te-O bonds and/or formation of more TeO3+δ/Te(O,F)3+δ groups [3436]. This indicates that fluorides play a different role as network modifiers in comparison to the oxides.

A Judd-Ofelt analysis can be performed to predict the spontaneous transition properties of Pr3+ in host matrix [37,38]. In the present work, absorption bands (3P2, 3P1+1I6, 3P0, 1D2, 1G4, 3F4, 3F3, 3F2, and 3H6,) were used for the Judd-Ofelt calculation, and the overlapped bands were de-convoluted using Gaussian fitting. The reduced matrix elements of the tensor operators used in the calculation of the electric dipole line strengths (Sed) can be considered independent of the host materials and the values of them were quoted from Ref [39]. The phenomenological Judd-Ofelt intensity parameters Ωt(t=2,4,6) were calculated by a least-squares fitting of the electric dipole oscillator strengths from the experiment (Pexp) and calculation (Pcal). The root-mean-square deviation δrms between the Pexp and Pcal was calculated by δrms=[(PexpPcal)2/(NtranNpara)]1/2, where Ntran and Npara are the number of the transition and parameters used in the calculation. The Ωt(t=2,4,6) are obtained to be (3.57, 6.60, 5.18)×10−20 cm2, and δrms] is obtained to be 3.54×10−6. The percentage error of the Judd-Ofelt analysis is 39.4%. The measured and the calculated oscillator strength values are shown in Table 1 . Generally, the asymmetry of the local structure and the covalency of the lanthanide-ligand bonds can be reflected by the Ω2. The Ω2 value of Pr3+ in fluorotellurite glass is much larger in comparison to that in transparent oxyfluoride glass and glass ceramics, tellurofluorophosphate glass, fluorozirconate and mix-halide glasses [4042], indicating a stronger asymmetry and covalent environment between the Pr3+ ions and the ligand in the fluorotellurite glass matrix than in those host.

Tables Icon

Table 1. Measured and calculated oscillator strengths, and electric dipole line strengths of absorption transitions of Pr3+ in fluorotellurite glass

Using Ωt(t=2,4,6) values, the spontaneous transition probability (A), originating from an electrical dipole transition from aJ to bJ' can be calculated by

A(aJ;bJ')=64π4e23hλ3(2J+1)n(n2+2)29×Sed(aJ;bJ'),
where e is the charge of electron, h is the Planck constant, n is the refractive index, λ is the wavelength of the transition, and J is the total angular momentum quantum number. The matrix elements used in calculation of Sed (aJ;bJ') in Eq. (1) are employed from Ref [43]. The spontaneous transition probability from magnetic contribution can be ignored because they are very small compared with the electrical dipole transition. And other important parameters including the branch ratio β (aJ;bJ') and radiative lifetime τrad can be further calculated by β(aJ;bJ')=A(aJ;bJ')/J'A(aJ;bJ') and τrad=1/J'A(aJ;bJ'), where the summation is over all transitions to every terminal state bJ '. The radiative quantum efficiency η is defined as η=τm/τrad, where τm is the measured lifetime. The results for transitions from Pr3+: 3P1, 3P0, 1D2, and 1G4 in fluorotellurite glass are summarized in Table 2 . The transition probability for the Pr3+: 1D21G4 is 880.1−1 with a branch ratio of 11.05%. And for the 3P0, most of the energy in this level would radiate to the ground-state, this is in aggreement with the experimental observations where the blue emission (3P03H4) dominates [41,42].

Tables Icon

Table 2. Spontaneous transition parameters of Pr3+: 3P1, 3P0, 1D2, and 1G4 in fluorotellurite glass

Figure 2(a) shows the near-IR PL spectra of Pr3+-doped samples under 488 nm excitation. Broadband emissions covering 1.30-1.67 μm wavelength region are observed for the samples with lower Pr3+ dopant concentration. With the increasing of Pr3+, the PL intensity decreases after an initial increase, and the emission peak extends to the short wavelength side whilst the emission band at the long wavelength side declines. Figure 2(b) shows the full-width at half-maximum (FWHM) of the emission bands. It can be seen that the FWHM increases with the increasing of Pr3+ and reaches a maximum FWHM bandwidth of 196 nm when Pr3+ concentration is equal to 1.0 mol%. This broadband emission can be assigned to the Pr3+: 1D21G4 transition, the unique ligand of the Pr3+ sites in the fluorotellurite matrix enables the observation of broadband emission from this transition. The emission band at the shorter wavelength side might be also contributed by the Pr3+: 1G43H5 transition [23]. In deed, an intense emission at around 1.33 μm corresponding to this transition is observed under 996 nm wavelength excitation [see Fig. 2(c)]. The PL intensity decline at higher Pr3+ doping level, in particular, at the longer wavelength side can be ascribed to the cross relaxation processes [1D2, 3H4]→[1G4, 3F4,3], considering the spectral overlap of the emission band (1D21G4) and the absorption band (3F4,33H4) as shown in Fig. 2(c). The closer Pr3+-Pr3+ separation at higher Pr3+ doping level makes this process occur easily, leading to a depleting of the emission level Pr3+: 1D2 [28].

 figure: Fig. 2

Fig. 2 (a) Near-IR PL spectra of Pr3+-doped fluorotellurite glasses under 488 nm wavelength excitation. (b) The FWHM bandwidth of the emissions at different Pr3+ dopant concentration. (c) Normalized PL spectra line-shapes and a comparison of them with the Pr3+: 3F4,33H4 absorption band (dotted line) located in this wavelength region. The 1.33 μm emission from the Pr3+: 1G43H5 transition is also displayed in Fig. 2(c).

Download Full Size | PDF

To optimize the excitation wavelengths, the PLE spectra were measured at 1480 nm and are shown in Fig. 3(a) . It can be seen that, apart from the excitation at around 590 nm, broad excitation bands extending from 440 to 490 nm also have contributions to the near-IR emission. The excitation band corresponds to the manifolds (3P0,1,2,1I6), on which the energy can decay rapidly to the next lower emission level 1D2 mainly through the cross relaxation [3P0, 3H4]→[1D2, (3H6,3F2)]; the radiative decay can be ignored because the 3P01D2 transition rate is very low according to the Judd-Ofelt result. It can be also found that this excitation band exhibits similar change compared with the near-IR PL spectra as the Pr3+ concentration increases whilst the excitation band (3P0,1,2,1I6) gives a relatively rapider decrease in comparison to the 590 nm (1D2) excitation band [see Fig. 3(a)]. The excitation band around 590 nm shows advantage, in particular, at higher Pr3+ dopant concentration, is primarily due to the resonant excitation route by which the emission level Pr3+: 1D2 is populated directly.

 figure: Fig. 3

Fig. 3 (a) PLE spectra of Pr3+-doped fluorotellurite glass samples monitored at 1480 nm. (b) A comparison of the PLE spectrum and the absorption spectrum with respect to the 590 nm band for the 0.1 mol% Pr3+-doped sample. Inset of (b) shows the absorption/PLE spectra intensity ratio of the 3P0 band to the 1D2 band.

Download Full Size | PDF

By comparing the PLE spectra with the absorption spectrum [see Fig. 3(b) and the inset], it can be found that the absorptions in 440-490 nm wavelength region with respect to the 590 nm band is much more intense than that in the PLE spectra. This result indicates that there are some other dacays for the energy residing in the (3P0,1,2,1I6) manifolds. To further understand this phenomenon, the visible PL spectra under 445 nm excitation were recorded and are shown in Fig. 4 . It should be mentioned that there is no obvious change in the near-IR emission line-shapes measured under 445 nm compared with those measured under 488 nm wavelength excitation. Intense visible emissions peaking around 490, 528, 611, and 643 nm are observed, and they were assigned to the Pr3+: 3P03H4, 3P03H5, 3P03F3, and 3P03F3 transitions, respectively. This observation is in aggrement with the Judd-Ofelt calculation result that these transtitions posses larger branch ratios than those transitions terminating to other levels. The recorded visible PL spectra show similar altering tendency as the broadband near-IR emission; they decrease after an initial increase with the further increment of Pr3+ in concentration. The mechanism responsible for this can be attributed to the cross relaxation process [3P0, 3H4]→[1D2, (3H6,3F2)], higher Pr3+ dopant concentration enables an easy occurrence of this process, leading to a population of the broadband emission level 1D2. The less population of Pr3+: 3P0 results in a decreasing change of the visible emissions.

 figure: Fig. 4

Fig. 4 Visible PL spectra of Pr3+-doped fluorotellurite glasses under 445 nm excitation. Emission bands located around 490, 528, 611, and 643 nm correspond to the Pr3+ transitions of (3P1,3P0)→3H4, 3P03H5, (3P1,3P0)→3H6, and 3P03F2, respectively. Inset (a) compares the peak wavelengths of Pr3+: (3P1,3P0)→3H4 emission at different Pr3+ dopant concentration (solid lines) as well as the absorption band corresponding to Pr3+: (1I6 + 3P1,3P0)←3H4 (dotted line). Inset (b) shows the normalized PL intensity from the Pr3+: (3P1,3P0)→3H6 emission (solid lines) and the absorption band from the Pr3+: 1D23H4 transition (dotted line).

Download Full Size | PDF

It can be observed in Fig. 4 that the blue wavelength emission gives a deeper decrease than the green and red emissions although they originate from the common level 3P0. This means that some other quenching mechanism occurred on the Pr3+: 3P03H4 emission. Considering the spectral overlap between the high-energy emission side and the low-energy absorption side of Pr3+: 3P03H4, some energy in the higher sub-levels of Pr3+: 3P0 can be absorbed by a nearby Pr3+ in the ground state when they are in a close separation, resulting in a depleting of the higher energy sub-levels of Pr3+: 3P0. This is confirmed by the changes involved in the emission peak wavelength that presents a red-shift from 485 to 495 nm in this process, as shown in Fig. 4 inset (a). In addition, the transition from the 3P1 to 3H4 may contribute to this blue emission considering the observation of a shoulder around 470 nm which decreases with the increasing of Pr3+ concentration and nearly to be completely quenched when it reaches 1.0 mol%. This can be attributed to the process [3P13H4]:[3H43P0], taking into account that Pr3+: 3P0 shows intense absorption around 470 nm [see Fig. 4 inset (a)]. It should be also noted that the shoulder at the short wavelength side of the 613 nm emission, which has been demonstrated due to the hypersensitive Pr3+: 3P13F3 transition [31], shows a similar decrease tendency as the shoulder around 470 nm. This can be ascribed to the cross relaxation [3P1, 3H4]→[3H6, 1D2], by which some of Pr3+ in 3P1 non-radiatively decay to 3H6 with the energy being absorbed by Pr3+ ions in the ground state. The matched energy difference, as shown in inset (b) of Fig. 4, enables an easy occurrence of this process especially at higher Pr3+ doping level. The above explanations are also confirmed by the lifetime changes of the 3P0 and 1D2 states; the both decrease as the increment of Pr3+ in concentration, as shown in Figs. 5(a) and 5(b) and the insets of them. The lifetime values are determined using τm=tφ(t)/φ(t), where φ(t) is the dependence of decay curve on the time t [30], since the decay curves show a slight deviation from the single exponential function. All the aforementioned energy transfer processes together with the excitation routes are schematically illustrated in Fig. 6 .

 figure: Fig. 5

Fig. 5 Decay curves of Pr3+-doped glass samples monitored at (a) 1480 nm and (b) 495 nm. Insets (a) and (b) show the lifetimes of the both emissions as a function of Pr3+ dopant concentration.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Schematic energy-level diagram of Pr3+ in fluorotellurite glass and energy transfer processes involved. Notations (i), (ii), and (iii) stand for the cross relaxation processes [3P0, 3H4]→[1D2, (3H6,3F2)], [3P1, 3H4]→[3H6, 1D2], and [1D2, 3H4]→[1G4, 3F4,3] among Pr3+ ions, respectively.

Download Full Size | PDF

To further evaluate the gain performance of this glass system, it is of significance to determine the stimulated emission cross-section σem. Considering that the Pr3+ near-IR emission occurs between two excited states, the profiles of σem can be described by the equation [28]:

σem(λ)=Ajβjiλji5I(λji)8πcn2λjiI(λji)dλ,
where c is the light speed in vacuum, λji is the wavelength of a given transition from j to i which leads to the fluorescence I(λji). Here, Ajβji = 880.1 s−1 for the Pr3+: 1D21G4 transition according to the Judd-Ofelt calculation. The peak value of σem is obtained to be 0.90 × 10−20 cm2 at 1487 nm. This value is larger than that in bismuth gallate glass and of Ni2+ (0.63 × 10−20 cm2) in glass-ceramics [44], although smaller than the bismuth doped aluminum germanate glasses (4.435×10−20 cm2) [26] and bismuth and tantalum codoped germanium oxide glasses (1.59×10−20 cm2) [45]. The production σemτm is obtained to be 8.31 × 10−26 cm2 s, which is proportional to the amplification gain and inversed laser oscillation threshold, suggesting that this host matrix is promissing for the gain media as broadband amplifiers and tunable lasers.

4. Conclusions

In conclusion, broadband NIR emissions in the wavelength region of 1.30-1.67 μm were achieved from Pr3+-singly doped fluorotellurite glasses. The bandwidth (FWHM) was improved by the increase of Pr3+ concentration although it resulted in a quenching of the emission at higher dopant concentration. The spontaneous transition properties analyzed using the Judd-Ofelt theory indicate a stronger covalent bonding between the Pr3+ sites and the matrix lifgand field. The results suggest that the Pr3+-doped fluorotellurite glass system is promising for broadband near-IR luminescence sources and optical amplifiers/lasers operating at the expanded low-loss telecommunication transmission window. Further investigations and experiments are underway.

Acknowledgments

Y. H. Tsang acknowledges support from The Hong Kong Polytechnic University under Grants G-YH91, G-YJ20, and A-PK72. This work was supported by the Research Grants Council of the Hong Kong SAR, China, under project CityU 119708.

References and links

1. G. A. Thomas, B. I. Shraiman, P. F. Glodis, and M. J. Stephen, “Towards the clarity limit in optical fibre,” Nature 404(6775), 262–264 (2000). [CrossRef]   [PubMed]  

2. S. Kasap, “Optoelectronics,” in The Optics Encyclopedia, T. Brown, K. Creath, H. Kogelnik, M. A. Kriss, J. Schmit, and M. J. Weber, eds. (Wiley-VCH, Weihein, Germany, 2004), Vol. 4., pp. 2237–2284.

3. See, for example, M. J. F. Digonnet, ed., Rare-Earth-Doped Fiber Lasers and Amplifiers (Second Edition, Revised and Expanded), (Marcel Dekker, New York, 2009), and references therein.

4. H. Takebe, K. Yoshino, T. Murata, K. Morinaga, J. Hector, W. S. Brocklesby, D. W. Hewak, J. Wang, and D. N. Payne, “Spectroscopic properties of Nd3+ and Pr3+ in gallate glasses with low phonon energies,” Appl. Opt. 36(24), 5839–5843 (1997). [CrossRef]   [PubMed]  

5. M. Naftaly, S. Shen, and A. Jha, “Tm3+-doped tellurite glass for a broadband amplifier at 1.47 μm,” Appl. Opt. 39(27), 4979–4984 (2000). [CrossRef]   [PubMed]  

6. B. Zhou, H. Lin, D. Yang, and E. Y. B. Pun, “Emission of 1.38 μm and gain properties from Ho3+-doped low-phonon-energy gallate bismuth lead oxide glasses for fiber-optic amplifiers,” Opt. Lett. 35(2), 211–213 (2010). [CrossRef]   [PubMed]  

7. C. Strohhöfer and A. Polman, “Silver as a sensitizer for erbium,” Appl. Phys. Lett. 81(8), 1414–1416 (2002). [CrossRef]  

8. K. Driesen, V. K. Tikhomirov, C. Görller-Walrand, V. D. Rodriguez, and A. B. Seddon, “Transparent Ho3+-doped nano-glass-ceramics for efficient infrared emission,” Appl. Phys. Lett. 88(7), 073111 (2006). [CrossRef]  

9. J. Dong, Y. Q. Wei, A. Wonfor, R. V. Penty, I. H. White, J. Lousteau, G. Jose, and A. Jha, “Dual-pumped tellurite fiber amplifier and tunable laser using Er/Ce codoping scheme,” IEEE Photon. Technol. Lett. 23(11), 736–738 (2011). [CrossRef]  

10. S. Tanabe, “Rare-earth-doped glasses for fiber amplifiers in broadband telecommunication,” C. R. Chim. 5(12), 815–824 (2002). [CrossRef]  

11. S. Y. Seo, J. H. Shin, B. S. Bae, N. Park, J. J. Penninkhof, and A. Polman, “Erbium-thulium interaction in broadband infrared luminescent silicon-rich silicon oxide,” Appl. Phys. Lett. 82(20), 3445–3447 (2003). [CrossRef]  

12. L. Huang, A. Jha, S. Shen, and X. Liu, “Broadband emission in Er3+-Tm3+ codoped tellurite fibre,” Opt. Express 12(11), 2429–2434 (2004). [CrossRef]   [PubMed]  

13. Z. Xiao, R. Serna, C. N. Afonso, and I. Vickridge, “Broadband infrared emission from Er-Tm:Al2O3 thin films,” Appl. Phys. Lett. 87(11), 111103 (2005). [CrossRef]  

14. D. Chen, Y. Wang, F. Bao, and Y. Yu, “Broadband near-infrared emission from Tm3+/Er3+ co-doped nanostructured glass ceramics,” J. Appl. Phys. 101(11), 113511 (2007). [CrossRef]  

15. Y. Xu, Q. Zhang, C. Shen, D. Chen, H. Zeng, and G. Chen, “Broadband near-IR emission in Tm/Er-codoped GeS2-In2S3-based chalcohalide glasses,” J. Am. Ceram. Soc. 92(12), 3088–3091 (2009). [CrossRef]  

16. B. Zhou and E. Y. B. Pun, “Broadband near-infrared photoluminescence and energy transfer in Tm3+/Er3+ codoped low phonon energy gallate bismuth lead glasses,” J. Phys. D Appl. Phys. 44(28), 285404 (2011). [CrossRef]  

17. K. Murata, Y. Fujimoto, T. Kanabe, H. Fujita, and M. Nakatsuka, “Bi-doped SiO2 as a new laser material for an intense laser,” Fusion Eng. Des. 44(1-4), 437–439 (1999). [CrossRef]  

18. M. Y. Sharonov, A. B. Bykov, V. Petricevic, and R. R. Alfano, “Spectroscopic study of optical centers formed in Bi-, Pb-, Sb-, Sn-, Te-, and In-doped germanate glasses,” Opt. Lett. 33(18), 2131–2133 (2008). [CrossRef]   [PubMed]  

19. V. G. Truong, L. Bigot, A. Lerouge, M. Douay, and I. Razdobreev, “Study of thermal stability and luminescence quenching properties of bismuth-doped silicate glasses for fiber laser applications,” Appl. Phys. Lett. 92(4), 041908 (2008). [CrossRef]  

20. M. A. Hughes, T. Akada, T. Suzuki, Y. Ohishi, and D. W. Hewak, “Ultrabroad emission from a bismuth doped chalcogenide glass,” Opt. Express 17(22), 19345–19355 (2009). [CrossRef]   [PubMed]  

21. B. Wu, S. Zhou, J. Ruan, Y. Qiao, D. Chen, C. Zhu, and J. Qiu, “Enhanced near-infrared emission from Ni2+ in Cr3+/Ni2+ codoped transparent glass ceramics,” Appl. Phys. Lett. 92(15), 151102 (2008). [CrossRef]  

22. I. A. Bufetov and E. M. Dianov, “Bi-doped fiber lasers,” Laser Phys. Lett. 6(7), 487–504 (2009). [CrossRef]  

23. N. D. Psaila, R. R. Thomson, H. T. Bookey, A. K. Kar, N. Chiodo, R. Osellame, G. Cerullo, G. Brown, A. Jha, and S. Shen, “Femtosecond laser inscription of optical waveguides in Bismuth ion doped glass,” Opt. Express 14(22), 10452–10459 (2006). [CrossRef]   [PubMed]  

24. M. Peng, G. Dong, L. Wondraczek, L. Zhang, N. Zhang, and J. Qiu, “Discussion on the origin of NIR emission from Bi-doped materials,” J. Non-Cryst. Solids 357(11–13), 2241–2245 (2011). [CrossRef]  

25. B. Zhou, H. Lin, and E. Y. B. Pun, “Tm3+-doped tellurite glasses for fiber amplifiers in broadband optical communication at 1.20 µm wavelength region,” Opt. Express 18(18), 18805–18811 (2010). [CrossRef]  

26. B. Zhou, H. Lin, B. Chen, and E. Y. B. Pun, “Superbroadband near-infrared emission in Tm-Bi codoped sodium-germanium-gallate glasses,” Opt. Express 19(7), 6514–6523 (2011). [CrossRef]   [PubMed]  

27. Y. G. Choi, K. H. Kim, B. J. Park, and J. Heo, “1.6 μm emission from Pr3+: (3F3, 3F4)→3H4 transition in Pr3+- and Pr3+/Er3+-doped selenide glasses,” Appl. Phys. Lett. 78(9), 1249–1251 (2001). [CrossRef]  

28. B. Zhou and E. Y. B. Pun, “Superbroadband near-IR emission from praseodymium-doped bismuth gallate glasses,” Opt. Lett. 36(15), 2958–2960 (2011). [CrossRef]   [PubMed]  

29. P. Nandi, G. Jose, C. Jayakrishnan, S. Debbarma, K. Chalapathi, K. Alti, A. K. Dharmadhikari, J. A. Dharmadhikari, and D. Mathur, “Femtosecond laser written channel waveguides in tellurite glass,” Opt. Express 14(25), 12145–12150 (2006). [CrossRef]   [PubMed]  

30. B. Zhou, E. Y. B. Pun, H. Lin, D. Yang, and L. 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]  

31. M. Naftaly, C. Batchelor, and A. Jha, “Pr3+-doped fluoride glass for a 589 nm fiber laser,” J. Lumin. 91(3-4), 133–138 (2000). [CrossRef]  

32. 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]  

33. R. Jose, Y. Arai, and Y. Ohishi, “Raman scattering characteristics of the TBSN-based tellurite glass system as a new Raman gain medium,” J. Opt. Soc. Am. B 24(7), 1517–1526 (2007). [CrossRef]  

34. V. Nazabal, S. Todoroki, A. Nukui, T. Matsumoto, S. Suehara, T. Hondo, T. Araki, S. Inoue, C. Rivero, and T. Cardinal, “Oxyfluoride tellurite glasses doped by erbium: thermal analysis, structural organization and spectral properties,” J. Non-Cryst. Solids 325(1-3), 85–102 (2003). [CrossRef]  

35. G. Liao, Q. Chen, J. Xing, H. Gebavi, D. Milanese, M. Fokine, and M. Ferraris, “Preparation and characterization of new fluorotellurite glasses for photonic application,” J. Non-Cryst. Solids 355(7), 447–452 (2009). [CrossRef]  

36. A. Lin, A. Ryasnyanskiy, and J. Toulouse, “Fabrication and characterization of a water-free mid-infrared fluorotellurite glass,” Opt. Lett. 36(5), 740–742 (2011). [CrossRef]   [PubMed]  

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

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

39. 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]  

40. R. T. Génova, I. R. Martin, U. R. Rodriguez-Mendoza, F. Lahoz, A. D. Lozano-Gorrin, P. Nunez, J. Gonzalez-Platas, and V. Lavin, “Optical intensities of Pr3+ ions in transparent oxyfluoride glass and glass-ceramic. Applications of the standard and modified Judd-Ofelt theories,” J. Alloy. Comp. 380(1-2), 167–172 (2004). [CrossRef]  

41. L. R. Moorthy, M. Jayasimhadri, A. Radhapathy, and R. V. S. S. N. Ravikumar, “Lasing properties of Pr3+-doped tellurofluorophosphate glasses,” Mater. Chem. Phys. 93(2-3), 455–460 (2005). [CrossRef]  

42. M. A. Newhouse, R. F. Bartholomew, B. G. Aitken, L. J. Button, and N. F. Borrelli, “Pr-doped mixed-halide glasses for 1300 nm amplification,” IEEE Photon. Technol. Lett. 6(2), 189–191 (1994). [CrossRef]  

43. M. J. Weber, “Spontaneous emission probabilities and quantum efficiencies for excited states of Pr3+ in LaF3,” J. Chem. Phys. 48(10), 4774–4780 (1968). [CrossRef]  

44. T. Suzuki, G. S. Murugan, and Y. Ohishi, “Optical properties of transparent Li2O-Ga2O3-SiO2 glass-ceramics embedding Ni-doped nanocrystals,” Appl. Phys. Lett. 86(13), 131903 (2005). [CrossRef]  

45. M. Peng, J. Qiu, D. Chen, X. Meng, and C. Zhu, “Superbroadband 1310 nm emission from bismuth and tantalum codoped germanium oxide glasses,” Opt. Lett. 30(18), 2433–2435 (2005). [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 (6)

Fig. 1
Fig. 1 Optical absorption spectrum of (0.5 mol%)Pr3+-doped fluorotellurite glass. Inset (a) shows the detail of the absorption bands corresponding to the manifolds (3P2,1I6+3P1,3P0); Inset (b) shows the Raman spectrum (pink line) of undoped fluorotellurite glass with Gaussian fitting (green lines).
Fig. 2
Fig. 2 (a) Near-IR PL spectra of Pr3+-doped fluorotellurite glasses under 488 nm wavelength excitation. (b) The FWHM bandwidth of the emissions at different Pr3+ dopant concentration. (c) Normalized PL spectra line-shapes and a comparison of them with the Pr3+: 3F4,33H4 absorption band (dotted line) located in this wavelength region. The 1.33 μm emission from the Pr3+: 1G43H5 transition is also displayed in Fig. 2(c).
Fig. 3
Fig. 3 (a) PLE spectra of Pr3+-doped fluorotellurite glass samples monitored at 1480 nm. (b) A comparison of the PLE spectrum and the absorption spectrum with respect to the 590 nm band for the 0.1 mol% Pr3+-doped sample. Inset of (b) shows the absorption/PLE spectra intensity ratio of the 3P0 band to the 1D2 band.
Fig. 4
Fig. 4 Visible PL spectra of Pr3+-doped fluorotellurite glasses under 445 nm excitation. Emission bands located around 490, 528, 611, and 643 nm correspond to the Pr3+ transitions of (3P1,3P0)→3H4, 3P03H5, (3P1,3P0)→3H6, and 3P03F2, respectively. Inset (a) compares the peak wavelengths of Pr3+: (3P1,3P0)→3H4 emission at different Pr3+ dopant concentration (solid lines) as well as the absorption band corresponding to Pr3+: (1I6 + 3P1,3P0)←3H4 (dotted line). Inset (b) shows the normalized PL intensity from the Pr3+: (3P1,3P0)→3H6 emission (solid lines) and the absorption band from the Pr3+: 1D23H4 transition (dotted line).
Fig. 5
Fig. 5 Decay curves of Pr3+-doped glass samples monitored at (a) 1480 nm and (b) 495 nm. Insets (a) and (b) show the lifetimes of the both emissions as a function of Pr3+ dopant concentration.
Fig. 6
Fig. 6 Schematic energy-level diagram of Pr3+ in fluorotellurite glass and energy transfer processes involved. Notations (i), (ii), and (iii) stand for the cross relaxation processes [3P0, 3H4]→[1D2, (3H6,3F2)], [3P1, 3H4]→[3H6, 1D2], and [1D2, 3H4]→[1G4, 3F4,3] among Pr3+ ions, respectively.

Tables (2)

Tables Icon

Table 1 Measured and calculated oscillator strengths, and electric dipole line strengths of absorption transitions of Pr3+ in fluorotellurite glass

Tables Icon

Table 2 Spontaneous transition parameters of Pr3+: 3P1, 3P0, 1D2, and 1G4 in fluorotellurite glass

Equations (2)

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

A( aJ;bJ' )= 64 π 4 e 2 3h λ 3 ( 2J+1 ) n ( n 2 +2 ) 2 9 × S ed ( aJ;bJ' ),
σ em ( λ )= A j β ji λ ji 5 I( λ ji ) 8πc n 2 λ ji I( λ ji )dλ ,
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.