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

Comparative study of 2.7 μm emission of Ho3+ desensitized Er3+ in tellurite and bismuth glass

Open Access Open Access

Abstract

Energy transfer mechanisms have been quantitatively studied in Er3+/Ho3+ codoped tellurite and bismuth glass. Upon an excitation of 980 nm, an enhanced Er3+:2.7 μm emission was observed when Ho3+ was added to tellurite glass. Using Dexter’s theory, it is clearly shown that the efficiency of Er3+:4I13/2 → Ho3+:5I7 energy transfer is higher in tellurite glass than in bismuth glass. Based on upconversion and near-infrared spectra, the cross relaxation Er3+:4I13/2 + Ho3+:5I6 → Er3+:4I15/2 + Ho3+:5F5 is comparatively effective in tellurite glass, where it depletes the population of Er3+:4I13/2 level. The excellent spectroscopic characteristics indicate Er3+/Ho3+ co-doped tellurite glass has potential application for mid-infrared laser.

© 2017 Optical Society of America

1. Introduction

The 2.7 μm mid-infrared luminescence of the Er3+ ion is very useful for infrared countermeasures as well as in medical surgery owing to the electronic transitions 4I11/24I13/2 [1–3]. A highly doped Er3+ concentration is preferred for the energy recycling of the 4I11/2 level and the fast depletion of the 4I13/2 level [4]. Meanwhile, a suitable codoping of rare earth ions is required to provide a high population inversion. It has been reported that Ho3+ ions can deplete upper-level 4I11/2 and terminal-level 4I13/2 as a deactivator. An enhanced 2.7 μm emission was obtained from Er3+/Ho3+ codoped fluorophosphate glass (FP) using ZBYA glass in the literature [5,6]. However, the opposite results were also observed for Er3+/Ho3+ codoped fluorophosphate [4]. Therefore, the compositional dependence of the energy transfer mechanism between Er3+ and Ho3+ must be further examined.

Unlike the desensitizer Pr3+ and Nd3+ ions, Ho3+ depletes the population on the energy level, assisted by a phonon-involved energy transfer. Thus, the phonon energy distribution of the host glass has a strong influence on the fluorescence behavior of Er3+and Ho3+ ions. With regard to 2–3 μm lasers, research interest has been focused mainly on low-phonon energy glass, such as fluoride and chalcogenide glass, owing to the small energy gaps between the Er3+:4I11/2 and 4I13/2 levels [7–9]. In addition, fluorophosphate with a small amount of phosphate could yield a quantum efficiency that promotes the 2.7 μm transition positively [5]. Chalcogenide glass is characterized by its wide infrared transparency and high refractive index, which makes it a natural candidate for infrared lasers [10, 11]. However, its complicated preparation and processing, and its inferior thermal stability, restrict it from wide application. Fluoride glass possesses the high solubility of rare earth ions, has wide infrared transparency, and can be easily prepared using the melt-quenching method [12]. Unfortunately, fluoride glass suffers from poor chemical stability.

Heavy metal oxide (HMO) glass is a promising host glass owing to its special properties such as a high refractive index, high density, and excellent infrared transmission [13, 14]. Tellurite and bismuth glass are the most important types of HMO glass. The difference in phonon distribution of tellurite and bismuth glass has a great influence on the phonon-assisted energy transfer processes between Er3+ and Ho3+ ions. Based on these qualities, two series of Er3+-doped and Er3+/Ho3+ codoped tellurite and bismuth glass were prepared and characterized to investigate the effect of Ho3+ on Er3+:2.7 μm fluorescence in HMO glass. The Judd–Ofelt theory and the Dexter model are employed to quantitatively examine the kinetic behaviors between the excited levels.

2. Experimental

The set of glass used in this paper has the following nominal molar composition: 54TeO2 - 31WO1.5 - 15LaO1.5 - 2ErO1.5 - xHoO1.5 (x = 0, 2); 55BiO1.5 - 30GeO2 - 15NaO0.5 - 2ErO1.5 - xHoO1.5 (x = 0, 2), henceforth called TWLE, TWLEH, BGNE, and BGNEH, respectively. The undoped tellurite and bismuth glass matrices are denoted as TWL and BGN, respectively. The raw materials were prepared from high-purity TeO2, W2O3, La2O3, Bi2O3, GeO2, Na2CO3, and Er2O3 and Ho2O3 powder. Well-mixed raw materials (30 g) were placed in an alumina crucible and melted at 1050°C for 30 min in an oxygen atmosphere. Bubbling dry oxygen gas in melt was used to minimize the hydroxyl groups. The melts were quickly poured into preheated stainless-steel molds and annealed for 2 h near the glass transition temperature (Tg). The annealed sample was fabricated and polished to a size of 20 × 10 × 1 mm3 for optical property measurements.

Refractive indices were detected at room temperature using a Metricon Model 2010/M Prism Coupler. The Raman spectra of the glasses were measured with a Renishaw InVia Raman spectrophotometer using the 488-nm excitation line from a spectra physics laser. The absorption spectra were recorded with a Perkin-Elmer Lambda 900UV/VIS/NIR spectrophotometer in 1-nm steps. The fluorescence spectra were measured by a TRIAX550 spectrophotometer with a 980-nm LD as an excitation source. All measurements were conducted at room temperature.

3. Results and discussion

3.1 Raman spectra

The Raman spectra of the TWL and BGN samples are depicted in Fig. 1. The glass network structure could be effectively understood by the Raman scattering. Furthermore, the arrangement of the phonon energy of the host glass system can be evaluated from Raman spectroscopy. As shown in Fig. 1, the most prominent band at 780 cm−1 of TWL pure glass comprises the combined vibrations of asymmetric stretching and symmetric stretching of TeO4 trigonal bipyramid (tbp), and stretching vibrations of nonbridging Te-O bands in the TeO3 trigonal pyramid (tp) [15, 16]. The wide band at 345 cm−1 is associated with the vibration of TeO3 tp [17]. With regard to the BGN glass, the Raman spectrum is dominated by an intense and broad band at 384 cm−1, which arises from the vibrations of Bi-O-Bi symmetric stretching of [BiO3] pyramidal units and [BiO6] octahedral units [18,19]. By contrast, the phonon density at a higher frequency range of TWL glass is larger than that of BGN glass, which may increase the probability rate of nonradiative relaxation by multiphonon relaxation.

 figure: Fig. 1

Fig. 1 Raman spectra of TWL and BGN samples.

Download Full Size | PDF

3.2 Absorption spectra and Judd–Ofelt analysis

Absorption spectra were obtained for Er3+ single-doped and Er3+/Ho3+ codoped tellurite and bismuth glass in the 400–2200 nm range (Fig. 2). The absorption bands correspond to the 4f-4f optical excitations assigned to the transitions from the ground state to the higher states of Er3+ and Ho3+ ions. These are indicated in the figure. As observed in Fig. 2, the absorption cut edge shifted to the ultraviolet range for TWL with respect to BGN. This is associated with the higher polarizability of the Te4+ ions (αBi = 1.508 × 10−3 nm3, αTe = 1.595 × 10−3 nm3) [20]. By comparison, there is no obvious difference in the position and shape of the absorption bands between the single and codoped glasses.

 figure: Fig. 2

Fig. 2 Absorption spectra of prepared samples (1 mm thickness).

Download Full Size | PDF

In order to further examine spectroscopic variations of Er3+, the Judd–Ofelt (simplified as J-O) theory is employed to evaluate radiative properties such as the spontaneous transition probability (Arad), branching ratio (β), and radiative lifetime (τR) of excited levels of Er3+ in the single and codoped samples [21, 22]. These values are listed in Table 1. The theory and detailed calculation procedure were described in earlier literature [23]. The parameters Ω2 and Ω4 are widely studied in literature [24]. Ω2 is sensitive to the covalence of chemical bonding around the RE ion and the ligand asymmetry of the local environment of the host glass, while Ω6 is of particular interest since this parameter affects the values of the calculated spontaneous emission probabilities for the electronic transitions of RE3+ ions [25–28]. As shown in Table 1, the calculated Ω2 for TWL glass is lower than that for BGN glass. This may be caused by the weaker covalence of the Bi-O- bond compared with the Te-O- bond. A higher magnitude of Ω6 in bismuth glass entirely favors the larger transition ability. Therefore, owing to the higher Ω2 and Ω6, it is reasonable that the spontaneous transition probability (Arad) of Er3+ transitions for BGN glass is higher than that for TWL glass, as listed in Table 1.

Tables Icon

Table 1. J-O intensity parameters Ωλ, spontaneous radiative transition probability (Arad), fluorescence branching ratios (β), and radiative lifetime (τR) of excited levels of Er3+ in single and codoped samples.

After the introduction of the codoped Ho3+, Ω2, Ω6, Arad, and τR show considerable variations in the codoped sample. This indicates that the addition of Ho3+ has an important influence on the local surroundings of Er3+. Arad of the Er3+:4I11/24I13/2 transition increases from 56.04 s−1 to 74.89 s−1 in tellurite glass. A similar trend occurs for bismuth glass. This is derived from a partial population transfer from Er3+:4I13/2 to Ho3+:5I7 (shown in Fig. 6). On the other hand, the values of τR corresponding to the Er3+:4I13/24I15/2 transition in codoped samples are lower than those for single-doped samples. This results from the consumption of the population of Er3+:4I13/2 to Ho3+:5I7 based on the desensitized effect of Ho3+. This predicts the lower fluorescence intensity at 1.5 μm in the codoped samples. Interestingly, the branching ratio β of the Er3+:4I11/24I13/2 transition is about 19% for TWL and 15% for BGN, and decreases moderately after the introduction of the desensitizer Ho3+. This trend is different from that of FP and ZBAY [4, 29]. If an additional nonradiative feeding route does not exist, only portions of less than one fifth of the total population of 4I11/2 are radiatively fed to the 4I13/2 level. This has a detrimental effect on the 2.7 μm fluorescence.

3.3 Photoluminescence spectra and energy transfer process

Under the excitation of a 980 nm laser diode, the fluorescence spectra from 2550 nm to 2850 nm of Er3+ single-doped samples and Er3+/ Ho3+ codoped samples are depicted in Fig. 3. The emission band around 2.7 μm assigned to the Er3+:4I11/24I13/2 transition increases significantly with the introduction of Ho3+; however, the opposite phenomenon was observed in the BGNEH samples. These results suggest a great discrepancy between the phonon-assisted energy transfer mechanisms of tellurite and bismuth glass.

 figure: Fig. 3

Fig. 3 Mid-infrared fluorescence spectra of samples pumped by 980 nm laser diode.

Download Full Size | PDF

According to the Fuchtbauer–Ladenburg theory and emission spectra, the 2.7 μm emission cross section (σem) can be calculated using the following equation [30, 31]:

σem=λ4Arad8πcn2×λI(λ)λI(λ)dλ,
where λ represents the wavelength, Arad represents the spontaneous transition probability corresponding to the Er3+:4I11/24I13/2 transition, I(λ) is the emission intensity, and c and n are the light speed in a vacuum and the refractive index of glass, respectively. The maximum of the emission cross section in the TWLEH glass at 2.7 μm reaches 11.4 × 10−21 cm2, which is higher than that of Er3+/Ho3+ codoped ZBYA (10.2 × 10−21 cm2) or ZBLAY (10.8 × 10−21 cm2) [6, 8]. A high σem is beneficial to obtaining a laser action. This suggest that Er3+/Ho3+ codoped TWL glass might be a promising candidate for a mid-infrared laser system.

In order to explore the energy transfer mechanisms of tellurite and bismuth glass, the fluorescence spectra ranging from 1300 nm to 2200 nm and the visual upconversion spectra were measured and shown in Fig. 4 and Fig. 5, respectively.

 figure: Fig. 4

Fig. 4 Measured fluorescence spectrum of samples from 1300 nm to 2200 nm pumped at 980 nm.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Upconversion luminescence spectra of samples under 980 nm LD excitation.

Download Full Size | PDF

In Fig. 4, two obvious emission bands at 1.53 μm and 2.02 μm are attributed to the Er3+:4I13/24I15/2 transition and Ho3+: 5I75I8 transition, respectively. Since Ho3+ cannot be pumped efficiently by 980 nm LD, the emergence of a 2.02 μm emission accompanied by a descending 1.53 μm emission gives evidence to the energy transfer from Er3+ to Ho3+. Note that the intensity ratio of the 1.53 μm emission from codoped glass to single-doped glass (C/S) in tellurite glass (9.2%) is larger than that in bismuth glass (2.6%). The remaining 2.6% 1.5 μm emission intensity in bismuth glass indicates a more efficient energy transfer from Er3+:4I13/2 to Ho3+. This ratio is helpful in estimating the discrepancies between the energy transfer mechanisms of Er3+ and Ho3+ in different glass hosts. It is noted that the 1.5μm emission intensity was found almost two times larger in tellurite glass than in bismuth glass, indicating the more energy resided in the Er3+:4I11/2 level are transferred to Ho3+ in tellurite glass.

Under excitation of a 980 nm LD, the visual upconversion emission spectra of the samples are shown in Fig. 5. The spectra contain two intense bands in the green and red regions. Each assignment corresponding to the 4f-4f electronic transition is labeled for the present case. Owing to the incorporation of the Ho3+:5F4, 5S25I8 and Ho3+:5F55I8 transitions, the emission band around 550 nm generally shifts to the longer-wavelength side, while the 670 nm emission band shifts toward the shorter-wavelength side. With regard to the emission intensity, a weaker green emission and a stronger red emission are observed after the introduction of Ho3+ in the codoped samples, which is different from that of FP glass but similar to that of ZBYA glass [5, 29]. It is worthwhile to trace the differences in the spectral features of the two glasses. The C/S intensity ratios of the 550 nm emission and 670 nm emission in tellurite glass are larger than those in bismuth glass. This is depicted in Fig. 5.

On the basis of the above analysis, the details of the energy transfer channels between Er3+ and Ho3+ are shown in Fig. 6. For the Er3+ single-doped sample, the green and red emission bands correspond to the 4F7/2, 2H11/2, 4S3/24I15/2, and 4F9/24I15/2 transitions, while the 2.7 μm emission contributes to the 4I11/24I13/2 transition [29, 32]. After the introduction of Ho3+, owing to the small energy gap, some particles in Er3+:2H11/2 and Er3+:4S3/2 are transferred to Ho3+:5F4 and Ho3+:5S2, respectively, then nonradiatively de-excite to Ho3+:5F5, yielding the red emission. The energy transfer process can explain the wavelength shift of the upconversion luminescence peaks in the codoped samples. This is also responsible for the weaker green emission and stronger red emission in the codoped samples. According to the enhanced 2.0 μm emission and reduced 1.5 μm emission, there exists a shortcut from Er3+:2H13/2 to Ho3+:5I7. Similarly, a fraction of Er3+:2H11/2 excitation can be transferred to Ho3+:5I6, which has a significant influence on the 2.7 μm emission. It is noted that some phonons are needed to assist the two energy transfer processes in hosts.

 figure: Fig. 6

Fig. 6 Energy level diagram of Er3+ and Ho3+ ions with involved energy transfer mechanism.

Download Full Size | PDF

Except for direct energy transfers, the interactions between ions at higher concentrations contain cross-relaxation channels. Two main cross-relaxation processes are considered to be responsible for the luminescence quenching and shorter lifetimes [33]. On the one hand, energy on the Er3+:4I11/2 level can be transferred to the Ho3+:5I6 level, resulting an increase in populations on the Ho3+:5F4 level. On the other hand, the Ho3+:5F5 level could be populated through the cross relaxation if the Ho3+:5I6 level absorbs the energy stored on the Er3+:4I11/2 level.

The main energy transfer processes under excitation at 980 nm are as follows:

Er3+:2H11/2+Ho3+:5I8Er3+:4I15/2+Ho3+:5F4 (ET1)
Er3+:4S3/2+Ho3+:5I8Er3+:4I15/2+Ho3+:5S2 (ET2)
Er3+:4I11/2+Ho3+:5I8Er3+:4I15/2+Ho3+:5I6 (ET3)
Er3+:4I13/2+Ho3+:5I8Er3+:4I15/2+Ho3+:5I7 (ET4)
Er3+:4I11/2+Ho3+:5I6Er3+:4I15/2+Ho3+:5F4 (CR1)
Er3+:4I13/2+Ho3+:5I6Er3+:4I15/2+Ho3+:5F5 (CR2)

For the case of an almost resonant transfer, ET1 and ET2, few phonons are needed to assist the processes in hosts. The energy of the Er3+:2H11/2 and Er3+:4S3/2 levels are transferred to the Ho3+:5F4 and Ho3+:5S2 levels, respectively, resulting in a weaker green emission and a wavelength shift to the shorter sideband. The CR1 process can feed the population on the Ho3+:5F4 level, consuming the energy of the upper level corresponding to the 2.7 μm emission. There is energy competition between the transition luminescence of the Er3+:4I11/2 level and energy transfer processes ET1, ET2, and CR1 in glass. The observed weaker green emission in the codoped BGN glass (C/S = 0.22) in Fig. 5 proves that the energy transfer from the Er3+:4I11/2 level to the Ho3+ level is more efficient than that in TWL glass (C/S = 0.42). It is noted that the green emission intensity was found almost 40 times larger in tellurite glass than in bismuth glass.

On the other hand, benefiting from the ET3 and ET4 process, a 2.0 μm emission assigned to the Ho3+:5I75I8 transition is enhanced at the expense of the decreased 1.5 μm emission of Er3+. For the energy stored on the Er3+:2H13/2 level, there are two channels for population feeding from the Er3+:2H13/2 level to Ho3+: ET4 and CR2. The reduction in the 1.5 μm emission intensity is accompanied by an enhanced 2.0 μm emission and red upconversion emission. Furthermore, comparing the C/S intensity ratio of the 1.5 μm emission and the red upconversion emission in tellurite and bismuth glass, it is found that the population density of the Er3+:2H13/2 level is drastically reduced in tellurite glass.

According to Dexter and Förster, the microscopic constants of phonon-assisted energy transfer processes ET3 and ET4 can be determined [34, 35]. The related calculation methods were described in the literature [36]. The energy transfer microparameters of the above two processes are listed in Table 2. The N-phonon contribution terms (%) to the total probability rates are also listed. The results prove that the ET4 process is more effective than the ET3 process in the two systems, which are favorable to a population inversion between the upper Er3+:4I11/2 level and the lower Er3+:4I13/2 level. In addition, the forward energy transfer microparameters CDA of the ET4 process in tellurite glass is 11.89 × 10−40cm6/s. This is higher than that in bismuth glass, indicating that the ~2.7 μm Er3+:4I13/2 lower level is depopulated drastically. At least three phonons in bismuth glass are needed to assist the ET4 process, implying that this process barely takes place in the BGN samples [37, 38].

Tables Icon

Table 2. Calculated microparameters CDA, CAD for the two Er3+ → Ho3+ energy transfer processes (ET3 and ET4). Number (#) of phonons needed to assist energy transfer, as well as the percentage of each phonon participation (%) in the process.

In summary, two facts may be responsible for the discrepancies of the 2.7 μm spectroscopic properties in tellurite and bismuth glass. First, through the three processes ET1, ET2, and CR1, the energy stored on the Er3+:4I11/2 level is transferred efficiently to the excited states of Ho3+ in bismuth glass. Second, the ET4 process has a very harmful impact on the population of the Er3+:4I13/2 level, resulting in a strong population inversion in tellurite glass.

4. Conclusions

Based on a description of visual, near-infrared, and mid-infrared spectroscopic properties in Er3+/Ho3+ codoped tellurite and bismuth glass, some conclusions have been derived to understand the energy interaction between Er3+ and Ho3+ ions. Meanwhile, the radiative properties of RE3+ in heavy metal oxide glass were quantitatively analyzed using the Judd–Ofelt theory and Dexter–Förster model. The results demonstrate that the Ho3+ desensitization effect on the Er3+:2.7 μm fluorescence exhibits a great discrepancy between the tellurite and bismuth glass matrices. In particular, the Er3+:2.7 μm mid-infrared emission is characterized by a higher spontaneous transition probability (74.89 s−1) and larger calculated emission cross section (11.4 × 10−21cm2) in the TWLEH sample. The codoping introduction of Ho3+ depresses the Er3+:2.7 μm emission in bismuth glass. According to the Judd–Ofelt theory, the spontaneous transition probability of the Er3+:4I11/24I13/2 transition corresponding to the 2.7 μm emission increases after Ho3+ codoping in the two systems. By comparing the ratios of integrated emission intensity in the codoped doped sample to those of the single-doped sample, it is found that green emissions are decreased drastically in bismuth glass, while the increase of red emissions are relatively weak. For the nonresonant energy transfer mechanism between the Er3+ and Ho3+ ions, the phonon energy distribution of the glass host plays an important role. Benefiting from its relatively high phonon energy, tellurite glass has a larger energy transfer microparameter for the Er3+:4I13/2 → Ho3+:5I7 transition, which promotes the 2.7 μm emission.

Funding

National Natural Science Foundation of China (NSFC) (No. 61605115, 51472162); Shanghai Sailing Project (No.15YF1411800); Program for the Professor of Special Appointment (Eastern Scholar) at Shanghai Institutions of Higher Learning (No. TP2014061).

References and links

1. G. Wu, S. Fan, Y. Zhang, G. Chai, Z. Ma, M. Peng, J. Qiu, and G. Dong, “2.7 μm emission in Er3+:CaF2 nanocrystals embedded oxyfluoride glass ceramics,” Opt. Lett. 38(16), 3071–3074 (2013). [CrossRef]   [PubMed]  

2. G. Chai, G. Dong, J. Qiu, Q. Zhang, and Z. Yang, “2.7 μm emission from transparent Er3+,Tm3+ codoped yttrium aluminum garnet (Y3Al5O12) nanocrystals–tellurate glass composites by novel comelting technology,” J. Phys. Chem. C 116(37), 19941–19950 (2012). [CrossRef]  

3. M. Peng and L. Wondraczek, “Photoluminescence of Sr2P2O7:Bi2+ as a red phosphor for additive light generation,” Opt. Lett. 35(15), 2544–2546 (2010). [CrossRef]   [PubMed]  

4. L. Zhang and T. Xue, “Discrepancies between Pr3+ and Ho3+ de-sensitized Er3+:2.7 μm emission,” J. Lumin. 178, 22–26 (2016). [CrossRef]  

5. L. Zhang, Z. Yang, Y. Tian, J. Zhang, and L. Hu, “Comparative investigation on the 2.7 μm emission in Er3+/Ho3+ codoped fluorophosphate glass,” J. Appl. Phys. 110(9), 093106 (2011). [CrossRef]  

6. F. Huang, X. Liu, Y. Zhang, L. Hu, and D. Chen, “Enhanced 2.7- and 2.84-μm emissions from diode-pumped Ho3+/Er3+-doped fluoride glass,” Opt. Lett. 39(20), 5917–5920 (2014). [CrossRef]   [PubMed]  

7. H. Guo, L. Liu, Y. Wang, C. Hou, W. Li, M. Lu, K. Zou, and B. Peng, “Host dependence of spectroscopic properties of Dy 3+ - doped and Dy 3+, Tm 3+ -codped Ge-Ga-S-CdI 2 chalcohalide glasses,” Opt. Express 17(17), 15350–15358 (2009). [CrossRef]   [PubMed]  

8. Y. Tian, T. Wei, X. Jing, J. Zhang, and S. 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]  

9. P. Kuan, X. Fan, W. Li, X. Liu, C. Yu, L. Zhang, and L. Hu, “High-efficiency ∼2 μm laser in a single-mode Tm-doped lead germanate composite fiber,” Chin. Opt. Lett. 14(8), 081601 (2016). [CrossRef]  

10. Y. Yang, Z. Yang, P. Lucas, Y. Wang, Z. Yang, A. Yang, B. Zhang, and H. Tao, “Composition dependence of physical and optical properties in Ge-As-S chalcogenide glasses,” J. Non-Cryst. Solids 440, 38–42 (2016). [CrossRef]  

11. Z. Li, T. Xu, X. Shen, S. Dai, X. Wang, Q. Nie, and X. Zhang, “Investigation of 2.9 μm luminescence properties and energy transfer in Tm3+/Dy3+ co-doped chalcohalide glasses,” J. Rare Earths 29(2), 105–108 (2011). [CrossRef]  

12. T. Hu, D. D. Hudson, and S. D. Jackson, “Actively Q-switched 2.9 μm Ho3+Pr3+-doped fluoride fiber laser,” Opt. Lett. 37(11), 2145–2147 (2012). [CrossRef]   [PubMed]  

13. G. Zhao, Y. Tian, H. Fan, J. Zhang, and L. Hu, “Efficient 2.7μm emission in Er3+-doped bismuth germanate glass pumped by 980-nm laser diode,” Chin. Opt. Lett. 10(9), 091601 (2012). [CrossRef]  

14. S. C. Ardelean, R. C. Lucacel, and O. Hulpus, “EPR and FT-IR spectroscopic studies of B2O3Bi2O3MnO glasses,” Solid State Sci. 7(11), 1438–1442 (2005). [CrossRef]  

15. A. Chagraoui, A. Tairi, K. Ajebli, H. Bensaid, and A. Moussaoui, “New tellurite glasses and crystalline phases in the Bi2O3–CaO–TeO2 system: Synthesis and characterization,” J. Alloys Compd. 495(1), 67–71 (2010). [CrossRef]  

16. G. Zhao, Y. Tian, H. Fan, J. Zhang, and L. Hu, “Properties and Structures of Bi2O3–B2O3–TeO2 Glass,” J. Mater. Sci. Technol. 29(3), 209–214 (2013). [CrossRef]  

17. A. K. Yadav and P. Singh, “A review of the structures of oxide glasses by Raman spectroscopy,” RSC Advances 5(83), 67583–67609 (2015). [CrossRef]  

18. H. Fan, G. Gao, G. Wang, and L. Hu, “Infrared, Raman and XPS spectroscopic studies of Bi2O3–B2O3–GeO2 glasses,” Solid State Sci. 12(4), 541–545 (2010). [CrossRef]  

19. B. L. Shivachev, T. Petrov, H. Yoneda, R. Titorenkova, and B. Mihailova, “Synthesis and nonlinear optical properties of TeO2–Bi2O3–GeO2 glasses,” Scr. Mater. 61(5), 493–496 (2009). [CrossRef]  

20. V. Dimitrov and S. Sakka, “Electronic oxide polarizability and optical basicity of simple oxides. I,” J. Appl. Phys. 79(3), 1736–1740 (1996). [CrossRef]  

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

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

23. G. Zhao, P.-W. Kuan, H. Fan, and L. Hu, “Enhanced green and red upconversion and 2.7μm emission from Er3+/Tm3+ co-doped bismuth germanate glass,” Opt. Mater. 35(5), 910–914 (2013). [CrossRef]  

24. H. Ebendorff-Heidepriem, D. Ehrt, M. Bettinelli, and A. Speghini, “Effect of glass composition on Judd–Ofelt parameters and radiative decay rates of Er3+ in fluoride phosphate and phosphate glasses,” J. Non-Cryst. Solids 240(1–3), 66–78 (1998). [CrossRef]  

25. T. Izumitani, H. Toratani, and H. Kuroda, “Radiative and nonradiative properties of neodimium doped silicate and phosphate glasses,” J. Non-Cryst. Solids 47(1), 87–99 (1982). [CrossRef]  

26. S. Tanabe, T. Ohyagi, N. Soga, and T. Hanada, “Compositional dependence of Judd-Ofelt parameters of Er3+ ions in alkali-metal borate glasses,” Phys. Rev. B Condens. Matter 46(6), 3305–3310 (1992). [CrossRef]   [PubMed]  

27. S. Tanabe, T. Hanada, T. Ohyagi, and N. Soga, “Correlation between 151Eu Mössbauer isomer shift and Judd-Ofelt Ω 6 parameters of Nd3+ ions in phosphate and silicate laser glasses,” Phys. Rev. B Condens. Matter 48(14), 10591–10594 (1993). [CrossRef]   [PubMed]  

28. H. Takebe, Y. Nageno, and K. Morinaga, “Effect of network modifier on spontaneous emission probabilities of Er3+ in oxide glasses,” J. Am. Ceram. Soc. 77(8), 2132–2136 (1994). [CrossRef]  

29. F. Huang, X. Li, X. Liu, J. Zhang, L. Hu, and D. Chen, “Sensitizing effect of Ho3+ on the Er3+: 2.7 μm-emission in fluoride glass,” Opt. Mater. 36(5), 921–925 (2014). [CrossRef]  

30. T. Schweizer, D. W. Hewak, B. N. Samson, and D. N. Payne, “Spectroscopic data of the 1.8-, 2.9-, and 4.3-microm transitions in dysprosium-doped gallium lanthanum sulfide glass,” Opt. Lett. 21(19), 1594–1596 (1996). [CrossRef]   [PubMed]  

31. G. Tang, X. Wen, Q. Qian, T. Zhu, W. Liu, M. Sun, X. Chen, and Z. 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]  

32. G. Chai, G. Dong, J. Qiu, Q. Zhang, and Z. Yang, “Phase transformation and intense 2.7 μm emission from Er3+ doped YF3/YOF μm -crystals,” Sci. Rep. 3, 1598 (2013). [CrossRef]   [PubMed]  

33. R. Chen, Y. Tian, B. Li, F. Wang, X. Jing, J. Zhang, and S. Xu, “2 μm fluorescence of Ho3+:5I75I8 transition sensitized by Er3+ in tellurite germanate glasses,” Opt. Mater. 49, 116–122 (2015). [CrossRef]  

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

35. T. Förster, “Zwischenmolekulare energiewanderung und fluoreszenz,” Ann. Phys. 437(1–2), 55–75 (1948). [CrossRef]  

36. Y. Fang, G. Zhao, J. Xu, N. Zhang, Z. Ma, and L. Hu, “Energy transfer and 1.8μm emission in Yb3+/Tm3+ co-doped bismuth germanate glass,” Ceram. Int. 40(4), 6037–6043 (2014). [CrossRef]  

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

38. D. F. de Sousa and L. A. O. Nunes, “Microscopic and macroscopic parameters of energy transfer betweenTm3+ ions in fluoroindogallate glasses,” Phys. Rev. B 66(2), 024207 (2002). [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 (6)

Fig. 1
Fig. 1 Raman spectra of TWL and BGN samples.
Fig. 2
Fig. 2 Absorption spectra of prepared samples (1 mm thickness).
Fig. 3
Fig. 3 Mid-infrared fluorescence spectra of samples pumped by 980 nm laser diode.
Fig. 4
Fig. 4 Measured fluorescence spectrum of samples from 1300 nm to 2200 nm pumped at 980 nm.
Fig. 5
Fig. 5 Upconversion luminescence spectra of samples under 980 nm LD excitation.
Fig. 6
Fig. 6 Energy level diagram of Er3+ and Ho3+ ions with involved energy transfer mechanism.

Tables (2)

Tables Icon

Table 1 J-O intensity parameters Ωλ, spontaneous radiative transition probability (Arad), fluorescence branching ratios (β), and radiative lifetime (τR) of excited levels of Er3+ in single and codoped samples.

Tables Icon

Table 2 Calculated microparameters CDA, CAD for the two Er3+ → Ho3+ energy transfer processes (ET3 and ET4). Number (#) of phonons needed to assist energy transfer, as well as the percentage of each phonon participation (%) in the process.

Equations (7)

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

σ e m = λ 4 A r a d 8 π c n 2 × λ I ( λ ) λ I ( λ ) d λ ,
Er 3 + : 2 H 11 / 2 + Ho 3 + : 5 I 8 Er 3 + : 4 I 15 / 2 + Ho 3 + : 5 F 4
Er 3 + : 4 S 3 / 2 + Ho 3 + : 5 I 8 Er 3 + : 4 I 15 / 2 + Ho 3 + : 5 S 2
Er 3 + : 4 I 11 / 2 + Ho 3 + : 5 I 8 Er 3 + : 4 I 15 / 2 + Ho 3 + : 5 I 6
Er 3 + : 4 I 13 / 2 + Ho 3 + : 5 I 8 Er 3 + : 4 I 15 / 2 + Ho 3 + : 5 I 7
Er 3 + : 4 I 11 / 2 + Ho 3 + : 5 I 6 Er 3 + : 4 I 15 / 2 + Ho 3 + : 5 F 4
Er 3 + : 4 I 13 / 2 + Ho 3 + : 5 I 6 Er 3 + : 4 I 15 / 2 + Ho 3 + : 5 F 5
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.