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

Quantification of energy transfer processes in LiLa9(SiO4)6O2:Er3+/Yb3+ under selective Er3+ excitation

Open Access Open Access

Abstract

The energy transfer mechanisms between Er3+ and Yb3+ ions have been investigated in LiLa9(SiO4)6O2 under selective Er3+ excitation. IR emission spectra, measured in the CW excitation regime, were used to establish a relationship between the macroscopic transfer and back transfer parameters. These measurements were combined with the results obtained under pulsed excitation to quantify the absolute values of transfer (Yb3+ → Er3+) and back transfer coefficients (Er3+ → Yb3+), C25 = 9.5 × 10−17 cm3s−1 and C52 = 1.4 × 10−17 cm3s−1, respectively. Additionally, it has been observed an energy transfer that reduces the quantum efficiency of the green emitting Er3+ levels. The corresponding macroscopic coefficient has been also determined (CGQ = 6.1 × 10−17 cm3s−1).

© 2014 Optical Society of America

1. Introduction

In today’s technological age, there is a constant demand for new optical materials taking advantage of a variety of their properties. LiLa9(SiO4)6O2 (LLS) belongs to the nesosilicate family that can be easily synthetized in the form of small hexagonal transparent rods with a partially disordered lattice analogous to that of apatite [1,2].

Apart from the good optical, thermal and mechanical characteristics exhibited by apatite-structure materials [3], LLS presents the advantage of containing Lanthanum as one of its cationic lattice. This feature represents an additional advantage in comparison to other materials as it enables the incorporation of active trivalent lanthanide ions in large doping levels without the need of charge compensation, preserving the crystal quality and homogeneity needed for technological applications.

The highly efficient energy transfer processes that occur between Er3+ and Yb3+ ions make them some of the most attractive rare-earths (RE) to activate crystalline and amorphous materials. In fact, this high efficiency in the transfer processes from Yb3+ to Er3+ has played a fundamental role in the development of Er3+/Yb3+ co-doped materials with different functionalities such as optical amplifiers and solid state lasers operating at different wavelengths [47] and efficient up-converting submicron- and nano-sized phosphors [813]. In this wide variety of applications Yb3+ acts as sensitizer, providing its high absorption cross section as well as its broad absorption band (900 – 1100 nm) to excite the Er3+ luminescence via energy transfer. However, these ion-ion interactions can be useful or unsuitable depending on the concentration range and the final desired application, as they can potentiate or quench a luminescent band depending on the concentration range. Therefore, the knowledge and quantification of these cross-relaxation mechanisms between Er3+ and Yb3+ ions is an essential tool to optimize and assure the technological application of the co-doped material.

In this work, the energy transfer between Yb3+ and Er3+ ions in LLS crystals is investigated under selective Er3+ excitation. CW experiments, performed under Ti:Sapphire pumping, reveal an efficient energy transfer from Yb3+ to Er3+ (around ten times larger than the back transfer from Er3+ to Yb3+). The dynamics of the main emission bands have been also explored, in this case by using the second harmonic of a pulsed Nd:YAG laser. These measurements have allowed the full quantification of the macroscopic transfer and back transfer coefficients as well as the detection, and characterization, of a quenching process that affects the relaxation of the Er3+ green emitting levels. The macroscopic parameters have been used to establish a rate-equation model that can be used to describe the dynamics of Er3+/Yb3+ co-doped LLS crystals.

2. Experimental procedure

Co-doped LiLa9(SiO4)6O2:Er3+/Yb3+ single crystals have been grown by the flux technique following the procedure reported elsewhere [14]. The starting melts have a fixed Er3+ concentration (1 mol %) and five different Yb3+ concentrations (0, 1, 2, 5, 10 mol %). Table 1 summarizes the rare-earth concentration in the different samples (1 mol% = 1.54 × 1020 ions/cm3).

Tables Icon

Table 1. Er3+ and Yb3+ concentrations in the studied samples

The grown crystals present a prismatic shape with a size up to 0.5 × 1 × 2 mm3 and good optical quality and homogeneity as it can be observed in the optical microscope photograph shown in Fig. 1.

 figure: Fig. 1

Fig. 1 Optical microscope photograph showing the morphology of the LLS grown crystals.

Download Full Size | PDF

Spectroscopic measurements were performed by using two different excitation sources, a Millenia pumped Ti:Sapphire laser for exploring the CW-regime and a Nd:YAG laser (Spectra Physics model DCR 2/2A 3378, linked to a harmonic generator) to investigate the temporal dynamics. Independently from the excitation regime, the signal was dispersed through an ARC monochromator model SpectraPro 500-I and then detected by a Thorn Emi photomultiplier (QB9558) or an InGaAs Judson photodiode depending on the spectral region (visible or IR radiation, respectively). Additionally, for lifetime measurements, the signal was synchronously detected and recorded by a TEKTRONIX TDS 420 oscilloscope.

3. Results and discussion

3.1 CW pumping: ratio between the energy transfer coefficients

The 4f11 electronic configuration of Er3+ ions splits into a wide energetic distribution of 2S + 1LJ multiplets, when they are embedded in a solid. Such variety of energy levels makes possible the use of several excitation schemes to explore the spectroscopic properties of Er3+/Yb3+ co-doped samples. In the present work, we have chosen CW excitation at λ = 800 nm for several reasons. First, it minimizes sizeable up-conversion processes at sufficiently low pump powers; second, it provides a convenient way of excitation to Er3+ ions that does not interfere with the relevant emissions of interest in near infrared and, finally but not less important, it is adequate from the experimental point of view, considering the availability of suitable excitation sources (Ti:Sapphire laser or semiconductor laser diodes). For all these reasons, 800 nm becomes an excellent pumping wavelength for the present study.

Figure 2(a) shows the emission bands in the 900 – 1700 nm spectral range measured in samples #1 and #2 (singly Er3+-doped and Er3+/Yb3+ co-doped, respectively), under selective excitation at 800 nm (4I15/24I9/2 Er3+ absorption band). As it is sketched in Fig. 2(b), the excitation to the 4I9/2 level efficiently populates the lower lying level, 4I11/2, via non radiative relaxation [15]. In Er3+ singly doped samples, as it is the case of sample #1, the decay from this level is basically governed by the non-radiative connection to the 4I13/2 manifold. As a consequence, the most intense band in its IR emission spectrum corresponds to the 4I13/24I15/2 transition (λemi ~1.5 μm) while only a weak luminescence from the 4I11/24I15/2 transition (λemi ~1.0 μm) can be detected (note that in the figure the 1.0 μm band has been enlarged by a factor × 10).

 figure: Fig. 2

Fig. 2 (a) Infrared emission spectra measured in samples #1 and #2 under selective Er3+ excitation at 800 nm (blue and red lines respectively). (b) Partial energy level diagram showing the dominant emission bands and energy transfer processes observed after excitation at 800 nm.

Download Full Size | PDF

In Er3+/Yb3+ co-doped samples, the presence of Yb3+ ions opens an additional relaxation route for the 4I11/2 Er3+ manifold: the energy transfer process 4I11/24I15/2 (Er3+): 2F7/22F5/2 (Yb3+) [1618]. This transfer mechanism excites the Yb3+ ions to the 2F5/2 level, which relaxes to the ground state giving luminescence that partially overlaps with the 4I11/24I15/2 Er3+ transition. Then, in the case of sample #2, the intensity of the luminescence band around 1.0 μm becomes substantially higher than in the Er3+ singly doped sample. The energetic coincidence of the 4I11/2 (Er3+) and 2F5/2 (Yb3+) excited states makes it possible that the transfer operates not only from Er3+ to Yb3+ but also in the opposite direction, that is from Yb3+ to Er3+. Under CW pumping, the populations of the resonant 4I11/2 (Er3+) and 2F5/2(Yb3+) levels reach a dynamical equilibrium, and the relative intensities of the emissions at 1.0 μm in Er3+- doped and Er3+/Yb3+ co-doped samples allows the determination of the ratio between transfer (C25) and back transfer (C52) coefficients, as will be shown next.

Figure 2(b) illustrates the most relevant optical processes after selective excitation of Er3+ ions (λexc = 800 nm) via the 4I15/24I9/2 absorption band. Excitation power is kept sufficiently low to assure that: i) there is negligible ground state depletion, so that the population of the ground state of Er3+ and Yb3+ ions remains basically constant and coincident with the corresponding RE concentrations (NEr and NYb), and ii) it is possible to neglect up-conversion processes.

Under these conditions, the population dynamics of the infrared levels can be described by the following rate equations:

dN5dt=RPN3+C25N2N3C52N5N1A5mN5
dN4dt=(A54+W54NR)N5A4mN4
dN2dt=C52N5N1C25N2N3A2mN2
N3=NErN4N5NEr
N1=NYbN2NYb
where Ni (i = 1, …, 5) is the population density of i-th level (in units of ions/cm3), NEr and NYb are the total Er3+ and Yb3+ doping levels. Aij and WijNR symbolize the radiative and non-raditive transition probabilities associated to the ij transition and Aim = τ−1exp,i is the total de-excitation probability, which can be calculated from the reciprocal of the measured lifetime of the i-th level. Finally, C25 and C52 (in units of cm3s−1) represent the macroscopic parameters associated to the transfer (Yb3+ → Er3+) and back transfer (Er3+ → Yb3+) processes (see Fig. 2(b)).

Under CW pumping, after reaching steady state conditions (dNi/dt = 0), from Eqs. (3)(5):

0C52N5NYbC25N2NErA2mN2
and then:

N2N5C52NYbC25NEr+A2m

That is, after reaching the stationary regime, the ratio of populations of the resonant 4I11/2 (Er3+) and 2F5/2(Yb3+) manifolds is governed, apart from RE concentrations by the ratio between the transfer and back transfer coefficients that can be experimentally determined. In order to do that, it is usual to define (Pr) as the ratio between the Er3+ emission at around 1.0 μm (P53) over the total Er3+/Yb3+ emission also at 1.0 μm (P53 + P21) [16,18]:

Pr=P53P53+P21
where PijIij(λ), being Iij(λ) the measured emission intensity associated to the i → j transition. Considering that the non-radiative de-excitation probability of Yb3+ ions is negligible [15], then A21 ≈A2m, and the last equation can be rearranged and expressed in terms of the transition probabilities as:

1PrPr=A2mA53C52C25NEr+A2mNYbβNYb

Thus, measuring Pr in samples with fixed Er3+ concentration and variable Yb3+ concentrations, the quantity (1- Pr)/Pr should give a linear dependence on ytterbium concentration and the slope of such fit (β) links the transfer and back transfer coefficients by the relationship:

C52=A53A2mβ(C25NEr+A2m)

Nevertheless, from the experimental point of view, in co-doped samples it is impossible to measure the 4I11/24I15/2 (Er3+) and 2F5/22F7/2 (Yb3+) emissions around 1.0 μm independently, see Fig. 2(a). Therefore, the standard procedure [1618] is to compare the overall Er3+/Yb3+ emission at 1.0 μm (P21 + P53) with the Er3+ emission at 1.5 μm (P43, 4I13/24I15/2), which ratio with the 4I11/24I15/2 can be evaluated from Er3+ singly doped samples [16].

Pr=P43P53+P21(P53'P43')
where (P’53/ P’43) is the emission intensities ratio between the 4I11/24I15/2 and 4I13/24I15/2 transitions measured in the singly doped sample.

Then, according to Eq. (11), the experimental Pr values were determined from the emission bands at 1.0 μm and 1.5 μm measured for different doping levels (samples #1 - #5), and the calculated (1- Pr)/Pr values are represented in Fig. 3.

 figure: Fig. 3

Fig. 3 Least squares fit of (1 - Pr)/Pr as function of Yb3+ concentration.

Download Full Size | PDF

As it can be observed, the experimental data can be conveniently fit by a linear dependence with a slope value β = (1.63 ± 0.05) × 10−20 cm3 which together with the known spectroscopic parameters (Table 2), gives the numerical values that link the transfer coefficients C25 and C52 through Eq. (10).

Tables Icon

Table 2. Radiative, Aij, and total, Aim = τi,exp−1 = Σj Aij + WijNR , transition probabilities.

In order to obtain the absolute values of these transfer parameters we shall examine the temporal dynamics under pulsed excitation.

3.2 Pulsed excitation: full quantification of the energy transfer coefficients

The second harmonic of a pulsed Nd:YAG laser (λexc = 532 nm) has been used as excitation source to explore the temporal dynamics of the Er3+ and Yb3+ excited states. The emission spectra in the visible range, measured in samples #1 and #2, are depicted in Fig. 4(a).

 figure: Fig. 4

Fig. 4 (a) Visible emission spectra measured in samples #1 and #2 under selective Er3+ excitation at 532 nm (blue and red solid lines respectively). (b) Partial energy level diagram showing the dominant emission bands and energy transfer processes observed after excitation at 532 nm.

Download Full Size | PDF

Now the pumping wavelength populates the 2H11/2:4S3/2 Er3+ thermally coupled manifolds through the 4I15/22H11/2 absorption band giving rise to the occurrence not only of the previously mentioned IR emission bands (λemi ~1.0 μm and λemi ~1.5 μm), but also an additional luminescent band in the visible spectral range (λemi ~550 nm) associated to the radiative decay from the 2H11/2:4S3/2 levels to the ground state. Then, as it is sketched in Fig. 4(b), this excitation scheme results adequate to explore the temporal dynamics of 2H11/2:4S3/2 and 4I13/2 Er3+ manifolds as well as that corresponding to the resonant levels, 4I11/2 (Er3+):2F5/2 (Yb3+).

The temporal evolution of 2H11/2:4S3/24I15/2 transition as function of Yb3+ concentration measured under resonant pumping (samples #1 to #5), is presented in Fig. 5(a).

 figure: Fig. 5

Fig. 5 (a) Temporal evolution of the 2H11/2:4S3/24I15/2 emission band as function of Yb3+ concentration. (b) Estimation of the CGQ energy transfer coefficient associated to the cross relaxation process 4S3/24I13/2 (Er3+):2F7/22F5/2 (Yb3+).

Download Full Size | PDF

As it can be observed, after pulsed laser excitation (pulse width ~10 ns), the luminescent signals decay following a single exponential dependence with a lifetime value that decreases as Yb3+ concentration increases. Within the Yb3+ co-doping levels explored in this work, the 2H11/2:4S3/2 lifetime is slightly reduced in comparison to the intrinsic lifetime value, τ0 = 5.0 μs, measured in the Er3+ singly doped LLS sample (sample #1). This reduction in the lifetime of the green emitting manifolds has been previously reported in other Er3+/Yb3+ co-doped materials and it is due to the occurrence of a quenching process involving Yb3+ ions [1921].

The presence of Yb3+ ions opens a new de-excitation channel that partially depopulates the 2H11/2:4S3/2 manifolds transferring the excitation to the 4I13/2 (Er3+) and 2F5/2 (Yb3+) excited states, via the non-resonant cross relaxation process 4S3/24I13/2 (Er3+):2F7/22F5/2 (Yb3+). This additional mechanism has been also included in Fig. 4(b), and its probability will be quantified by a new transfer coefficient, CGQ.

In order to model the dynamics of the LLS:Er3+/Yb3+ system under this pulsed excitation scheme (λ = 532 nm, see Fig. 4(b)), it is necessary to consider a new set of rate-equations including the additional de-excitation paths operative under this pumping, the quenching mechanism and the 4F9/2 Er3+ level.

dN7dt=RPN3CGQN7N1A7mN7
dN6dt=(A76+W76NR)N7A6mN6
dN5dt=A75N7+(A65+W65NR)N6+C25N2N3C52N5N1A5mN5
dN4dt=C74N7N1+A74N7+A64N6+(A54+W54NR)N5A4mN4
dN2dt=C52N5N1+CGQN7N1C25N2N3A2mN2
N3=NErN4N5N6N7
N1=NYbN2

Equation (12) contains now the pumping term that populates the 2H11/2:4S3/2 manifolds during the laser pulse width and it predicts a single exponential decay when the pulse ends. This equation implies that, for moderate pumping levels (N1 ≈NYb), the experimental lifetime of these coupled manifolds depends on Yb3+ concentration as:

1τexpA7m+CGQNYb
or, alternatively:
1τexpA7m=CGQNYb
where A7m = τ0−1 is the reciprocal of the intrinsic lifetime value (sample #1 in Fig. 5(a)).

Following Eq. (12), the measured lifetime values (τexp) have been used to evaluate the quantity (τexp−1- A7m) as function of Yb3+ concentration. In fact for a given Yb3+ concentration, this value gives the energy transfer probability, WTR, at this doping level, see for instance [22]. Therefore, Eq. (19b) can be used to quantify the CGQ transfer coefficient. The least squares fit of the transfer probability values are presented in Fig. 5(b). As it can be seen, in good accordance with Eq. (19b), WTR increases linearly with ytterbium concentration and the slope of the linear fit gives directly the transfer coefficient that quantifies the energy transfer between 4S3/2(Er3+) and 2F7/2 (Yb3+):

CGQ=(6.1±0.1)×1017cm3s-1

Then, it is possible now to use the rate-equation model, Eqs. (12)(18), to describe the temporal evolution of all Er3+ and Yb3+ emission bands, after 532 nm pulsed excitation. The spectroscopic parameters indicated in Eqs. (12)(18) are summarized in Table 2, including the lifetime of LLS:Yb3+ measured also in the present work (sample #0).

The set of coupled differential equations were integrated as function of Yb3+ concentration by using a fourth order Runge-Kutta algorithm and the transition probabilities and transfer coefficients shown in the Table. The numerical integration allows the calculation of the temporal evolution corresponding to the different luminescent emissions, including their decay times as well as the associated rise-times. The only fitting parameter used is the value of one of the Er3+ ↔ Yb3+ transfer coefficients, given that the other one remains linked by the relationship Eq. (10) determined from CW experiments.

Figure 6 shows the experimental lifetimes of the resonant 4I11/2 (Er3+):2F5/2 (Yb3+) manifolds (dots) as function of Yb3+ concentration. As it can be seen, the presence of Yb3+ ions induces a lengthening of the lifetime value, from τ = (20 ± 1) μs (intrinsic value of 4I11/2 level) up to τ = (86 ± 5) μs, when Yb3+ doping level is 10 times higher than the Er3+ content (sample #5).

 figure: Fig. 6

Fig. 6 Experimental (dots) and predicted lifetimes (lines) of the resonant levels, 4I11/2(Er3+):2F5/2(Yb3+), as function of Yb3+ concentration.

Download Full Size | PDF

The experimental data can be fit by using a constant transfer coefficient using a least square criterion; the best fit is obtained for C25 = 9.5 × 10−17 cm3s−1. The back transfer coefficient C52 = 1.4 × 10−17 cm3s−1 is determined by inserting the C25 value in the experimental relationship given by Eq. (10). The calculated lifetime values are quite sensitive to small variations in the transfer coefficients. Estimated dependencies with ± 5% variations are included in the Fig. 6 (dotted lines).

To illustrate this fact some calculated dependencies using slightly different values for the transfer parameter (C25) have been included in the figure.

The obtained transfer parameters indicate that the efficiency of the transfer process (Yb3+ → Er3+) is significantly higher than that associated to the back transfer (Er3+ → Yb3+). Then, it can be expected that Yb3+ ions were ideal candidates to act as sensitizers of Er3+ luminescence in LLS crystals.

In Fig. 7, the temporal evolution of the luminescence bands measured in sample #4 (symbols) are compared with those predicted by the rate-equation model (solid lines) using the calculated transfer coefficients; that is, CGQ = 6.1 × 10−17 cm3s−1, C25 = 9.5 × 10−17cm3s−1 and C52 = 1.4 × 10−17cm3s−1.

 figure: Fig. 7

Fig. 7 Temporal evolution of the dominant emission bands in sample #4: measured after pulsed excitation at 532 nm (symbols) and calculated by numerical integration (solid lines).

Download Full Size | PDF

The numerical integration, lines in Fig. 7, represents the predictions of the rate equations without any further recalculation of the parameters. It can be observed that the predictions are in excellent agreement with the observed experimental dynamics: the decays as well as the rise-times of the IR emission bands are very well reproduced. These rise-times are originated by the time required to activate the decay paths (energy transfer and the radiative and non-radiative connections) that populate the 2F5/2, 4I11/2 and 4I13/2 manifolds from the upper levels.

As a summary, the dependencies of the lifetimes corresponding to 4I13/2, 4I11/2 (Er3+):2F5/2 (Yb3+) and 2H11/2:4S3/2 manifolds on Yb3+ concentration are shown in Fig. 8, where the experimental data (dots) are compared with the predictions of the model (solid lines). As it can be seen, model and experiments indicate that while the three cross relaxation mechanisms modify the temporal dynamics of the green emitting levels and the resonant manifolds, the dynamics of the Er3+ metastable level remains unaltered.

 figure: Fig. 8

Fig. 8 Experimental (symbols) and calculated (solid lines) lifetimes, under excitation at 532 nm, for the three main luminescence emissions of LLS:Er3+/Yb3+ as function of Yb3+ concentration.

Download Full Size | PDF

Finally it should be noted that, although the transfer coefficients have been quantified and checked under specific excitation schemes, the knowledge of the CGQ, C25 and C52 values will represent an essential tool to understand other pumping wavelength dependent effects such as the infrared-to-visible energy conversion under Yb3+-excitation.

4. Conclusions

The analysis of the infrared luminescence of Er3+/Yb3+ co-coped LLS samples under CW Er3+ excitation indicates that transfer and back transfer coefficients are not independent but are linked by a linear relationship including decay rates of the emitting levels and Er concentration, given by Eq. (10), that can be determined by comparing the IR emissions at 1.0 μm and 1.5 μm.

The absolute values of these coefficients can be calculated examining the temporal dynamics under pulsed excitation. By comparing the experimental data with the predictions of a model based on the rate-equation formalism, the transfer and back transfer coefficients have been determined: C25 = 9.5 × 10−17cm3s−1 and C52 = 1.4 × 10−17 cm3s−1.

Additionally, the experimental results obtained for the green emission reveal that the presence of Yb3+ ions opens a quenching mechanism that reduces the lifetime of the 2H11/2:4S3/2 Er3+ manifolds. The macroscopic coefficient associated to this non-resonant cross-relaxation has been also quantified: CGQ = 6.1 × 10−17cm3s−1.

The three transfer coefficients above mentioned can be used to predict the temporal dynamics of the main emission bands of Er3+/Yb3+ co-doped LLS.

Acknowledgments

Work partially supported by Ministerio de Economía y Competitividad and Comunidad de Madrid under projects SONAMFIBIOS (MAT2012-34919) and MICROSERES-CM (P2009/TIC-1476). The authors gratefully thank Erica Viviani (Univ. Verona) for expert technical assistance.

References and links

1. M. Bettinelli, A. Speghini, D. Falcomer, E. Cavalli, G. Calestani, M. Quintanilla, E. Cantelar, and F. Cussó, “Crystal structure and optical spectra of LiLa9(SiO4)2 crystals activated with Er3+,” J. Lumin. 128(5–6), 738–740 (2008). [CrossRef]  

2. E. Cavalli, G. Calestani, A. Belletti, M. Bettinelli, and A. Speghini, “Optical spectroscopy of Nd3+ in LiLa9(SiO4)2 crystals,” Opt. Mater. 31(9), 1340–1342 (2009). [CrossRef]  

3. K. B. Steinbruegge, T. Henningsen, R. H. Hopkins, R. Mazelsky, N. T. Melamed, E. P. Riedel, and G. W. Roland, “Laser properties of Nd+3 and Ho+3 doped crystals with the apatite structure,” Appl. Opt. 11(5), 999–1012 (1972). [CrossRef]   [PubMed]  

4. P. E. A. Möbert, E. Heumann, G. Huber, and B. H. Chai, “Green Er3+:YLiF4 upconversion laser at 551nm with Yb3+ codoping: a novel pumping scheme,” Opt. Lett. 22(18), 1412–1414 (1997). [CrossRef]   [PubMed]  

5. Q. Wang, R. Ahrens, and N. K. Dutta, “Optical gain of single mode short Er/Yb doped fiber,” Opt. Express 12(25), 6192–6197 (2004). [CrossRef]   [PubMed]  

6. S. C. Zeller, T. Südmeyer, K. J. Weingarten, and U. Keller, “Passively modelocked 77 GHz Er:Yb:glass laser,” Electron. Lett. 43(1), 32–33 (2007). [CrossRef]  

7. I. Pavlov, E. Ilbey, E. Dülgergil, A. Bayri, and F. Ö. Ilday, “High-power high-repetition-rate single-mode Er-Yb-doped fiber laser system,” Opt. Express 20(9), 9471–9475 (2012). [CrossRef]   [PubMed]  

8. Z. Li, L. Zheng, L. Zhang, and L. Xiong, “Synthesis, characterization and upconversion emission properties of the nanocrystals of Yb3+/Er3+-codoped YF3-YOF-Y2O3 system,” J. Lumin. 126(2), 481–486 (2007). [CrossRef]  

9. N. O. Nuñez, M. Quintanilla, E. Cantelar, F. Cusso, and M. Ocaña, “Uniform YF3:Yb,Er up-conversion nanophosphors of various morphologies synthetised in polyol media through an ionic liquid,” J. Nanopart. Res. 12(7), 2553–2565 (2010). [CrossRef]  

10. C. Wang, H. Tao, L. Cheng, and Z. Liu, “Near-infrared light induced in vivo photodynamic therapy of cancer based on upconversion nanoparticles,” Biomaterials 32(26), 6145–6154 (2011). [PubMed]  

11. M. Haase and H. Schäfer, “Upconverting nanoparticles,” Angew. Chem. Int. Ed. Engl. 50(26), 5808–5829 (2011). [CrossRef]   [PubMed]  

12. J. Zhou, Z. Liu, and F. Li, “Upconversion nanophosphors for small-animal imaging,” Chem. Soc. Rev. 41(3), 1323–1349 (2012). [CrossRef]   [PubMed]  

13. M. Marin-Dobrincic, E. Cantelar, and F. Cusso, “Temporal dynamics of IR-to-visible up-conversion in LiNbO3:Er3+/Yb3+: a path to phosphors with tunable chromaticity,” Opt. Mater. Express 2(11), 1529–1537 (2012). [CrossRef]  

14. M. Setoguchi, “Crystal growth of silicate apatites by flux method,” J. Cryst. Growth 99(1–4), 879–884 (1990). [CrossRef]  

15. E. Cantelar, M. Quintanilla, F. Cussó, E. Cavalli, and M. Bettinelli, “Optical transition probabilities in Er3+- and Tm3+-doped LiLa9(SiO4)6O2 crystals,” J. Phys. Condens. Matter 22(21), 215901 (2010). [CrossRef]   [PubMed]  

16. L. F. Johnson, H. J. Guggenheim, T. C. Rich, and F. W. Ostermayer, “Infrared-to-visible conversion by rare-earth ions in crystals,” J. Appl. Phys. 43(3), 1125–1137 (1972). [CrossRef]  

17. B. Simondi-Teisseire, B. Viana, D. Vivien, and A. M. Lejus, “Yb3+ to Er3+ energy transfer and rate-equations formalism in the eye safe laser material Yb:Er:Ca2Al2SiO7,” Opt. Mater. 6(4), 267–274 (1996). [CrossRef]  

18. E. Cantelar, J. A. Muñoz, J. A. Sanz-García, and F. Cusso, “Yb3+ to Er3+ energy transfer in LiNbO3,” J. Phys. Condens. Matter 10(39), 8893–8903 (1998). [CrossRef]  

19. E. Okamoto, H. Masui, K. Muto, and K. Awazu, “Nonresonant energy transfer from Er3+ to Yb3+ in LaF3,” J. Appl. Phys. 43(5), 2122–2125 (1972). [CrossRef]  

20. Y. Mita, H. Yamamoto, K. Katayanagi, and S. Shionoya, “Energy transfer processes in Er3+- and Yb3+-doped infrared upconversion materials,” J. Appl. Phys. 78(2), 1219–1223 (1995). [CrossRef]  

21. R. E. Di Paolo, E. Cantelar, X. M. Wang, T. Tsuboi, and F. Cussó, “Determination of the Er3+ to Yb3+ energy transfer efficiency in Er3+/Yb3+ -codoped YVO4 crystals,” J. Phys. Condens. Matter 13(35), 7999–8006 (2001). [CrossRef]  

22. F. W. Ostermayer, J. P. van der Ziel, H. M. Marcos, L. G. Van Uiter, and J. E. Geusic, “Frequency upconversion in YF3:Yb3+,Tm3+,” Phys. Rev. B 3(8), 2698–2705 (1971). [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 (8)

Fig. 1
Fig. 1 Optical microscope photograph showing the morphology of the LLS grown crystals.
Fig. 2
Fig. 2 (a) Infrared emission spectra measured in samples #1 and #2 under selective Er3+ excitation at 800 nm (blue and red lines respectively). (b) Partial energy level diagram showing the dominant emission bands and energy transfer processes observed after excitation at 800 nm.
Fig. 3
Fig. 3 Least squares fit of (1 - Pr)/Pr as function of Yb3+ concentration.
Fig. 4
Fig. 4 (a) Visible emission spectra measured in samples #1 and #2 under selective Er3+ excitation at 532 nm (blue and red solid lines respectively). (b) Partial energy level diagram showing the dominant emission bands and energy transfer processes observed after excitation at 532 nm.
Fig. 5
Fig. 5 (a) Temporal evolution of the 2H11/2:4S3/24I15/2 emission band as function of Yb3+ concentration. (b) Estimation of the CGQ energy transfer coefficient associated to the cross relaxation process 4S3/24I13/2 (Er3+):2F7/22F5/2 (Yb3+).
Fig. 6
Fig. 6 Experimental (dots) and predicted lifetimes (lines) of the resonant levels, 4I11/2(Er3+):2F5/2(Yb3+), as function of Yb3+ concentration.
Fig. 7
Fig. 7 Temporal evolution of the dominant emission bands in sample #4: measured after pulsed excitation at 532 nm (symbols) and calculated by numerical integration (solid lines).
Fig. 8
Fig. 8 Experimental (symbols) and calculated (solid lines) lifetimes, under excitation at 532 nm, for the three main luminescence emissions of LLS:Er3+/Yb3+ as function of Yb3+ concentration.

Tables (2)

Tables Icon

Table 1 Er3+ and Yb3+ concentrations in the studied samples

Tables Icon

Table 2 Radiative, Aij, and total, Aim = τi,exp−1 = Σj Aij + WijNR , transition probabilities.

Equations (21)

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

d N 5 dt = R P N 3 + C 25 N 2 N 3 C 52 N 5 N 1 A 5m N 5
d N 4 dt =( A 54 + W 54 NR ) N 5 A 4m N 4
d N 2 dt = C 52 N 5 N 1 C 25 N 2 N 3 A 2m N 2
N 3 = N Er N 4 N 5 N Er
N 1 = N Yb N 2 N Yb
0 C 52 N 5 N Yb C 25 N 2 N Er A 2m N 2
N 2 N 5 C 52 N Yb C 25 N Er + A 2m
P r = P 53 P 53 + P 21
1 P r P r = A 2m A 53 C 52 C 25 N Er + A 2m N Yb β N Yb
C 52 = A 53 A 2m β( C 25 N Er + A 2m )
P r = P 43 P 53 + P 21 ( P 53 ' P 43 ' )
d N 7 dt = R P N 3 C GQ N 7 N 1 A 7m N 7
d N 6 dt =( A 76 + W 76 NR ) N 7 A 6m N 6
d N 5 dt = A 75 N 7 +( A 65 + W 65 NR ) N 6 + C 25 N 2 N 3 C 52 N 5 N 1 A 5m N 5
d N 4 dt = C 74 N 7 N 1 + A 74 N 7 + A 64 N 6 +( A 54 + W 54 NR ) N 5 A 4m N 4
d N 2 dt = C 52 N 5 N 1 + C GQ N 7 N 1 C 25 N 2 N 3 A 2m N 2
N 3 = N Er N 4 N 5 N 6 N 7
N 1 = N Yb N 2
1 τ exp A 7m + C GQ N Yb
1 τ exp A 7m = C GQ N Yb
C GQ =( 6.1±0.1 )× 10 17 c m 3 s -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.