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

Laser power density dependent energy transfer between Tm3+ and Tb3+: tunable upconversion emissions in NaYF4:Tm3+,Tb3+,Yb3+ microcrystals

Open Access Open Access

Abstract

Energy transfer between Tm3+ and Tb3+ dependent on the power density of pump laser was investigated in NaYF4: Tb3+,Tm3+,Yb3+ microcrystals. Under the excitation of a 976-nm near-infrared laser at various power densities, Tb3+-Tm3+-Yb3+ doped samples exhibited intense visible emissions with tunable color between green and blue. The ratio of blue and green emission were determined by energy transfer between Tm3+ and Tb3+. When the power density of pump laser was low, the energy transfer process from Tm3+ (3F4) to Tb3+ (7F0) occurred efficiently. Upconversion processes in Tm3+ were inhibited, only visible emissions from Tb3+ with green color were observed. When the power density increased, energy transfer from the 3F4 (Tm3+) to 7F0 level (Tb3+) was restrained and population on high energy levels of Tm3+ was increased. Contribution of upconversion emissions from Tm3+ gradually became dominant. The emission color was tuned from green to blue with increasing the power density. Energy transfer processes between low-lying levels of activators, such as Tm3+ will greatly reduce the population on certain levels for further high-order upconversion processes. The Tb3+-Tm3+-Yb3+ doped phosphors are promising materials for detecting the condition of power density of the invisible near-infrared laser.

© 2016 Optical Society of America

1. Introduction

Lanthanide-doped upconversion nanocrystals have attracted much research interest in the recent decades. Under excitation of near-infrared laser, i.e. at 976 nm, upconversion nanocrystals exhibit intense visible emissions that have been widely applied in biolabeling and bioimaging [1, 2]. The energy transfer upconversion (ETU) has been considered as the most efficient approach to achieve upconversion [3]. Ytterbium ions (Yb3+) are used as sensitizers to absorb energy from pump laser and transfer to activators such as Er3+, Tm3+, and Ho3+ by ETU [4–7]. Some high energy levels of other lanthanide ions, i.e. Tb3+ and Eu3+ cannot be excited by ETU from Yb3+. Energy transfer and migration from Tm3+ and Er3+ were extensively used to sensitize other type of lanthanide activators, e.g. Tb3+ and Eu3+ [8–10]. Energy transfer also occurs between low energy levels (i.e. less than 3000 cm−1) in lanthanide ions except for Gd3+ and Y3+ which have no excited states in this region. The energy transfer processes sufficiently change the population of low-lying levels and further affect the population of high energy levels. In other words, the energy transfers among low-lying levels determine the upconversion processes. Unfortunately, there is no clear investigation of sensitization processes from other activators such as Tm3+, especially the energy transfers between low energy levels.

Temperature sensing is one of the potential applications of upconversion phosphors [11–15]. Ishiwada et al. reported that the visible emissions from Tm3+ and Tb3+ can be affected by temperature via energy transfer processes from the 1G4 level of Tm3+ to the 5D4 level of Tb3+ [16]. Although energy level positions are similar, the strongest emission peaks from the 1G4 and 5D4 level are different that they are about 474 nm (blue light) and 544 nm (green light), respectively [17–19]. Owing to the energy transfer process, the color of emission lights can be tuned between green and blue with various thermal conditions under the excitation by 355 nm UV light [16]. If visible emissions from upconversion phosphors can be widely tuned by energy transfer processes between activators, it is meaningful not only for sensing but also for the investigation of energy transfer processes.

Upconversion processes are strongly dependent on the power density of pump laser [20]. Zhao et al. reported that emission ratios in NaYF4:Tm3+,Yb3+ nanocrystals changed when the power density of pump laser increased [21]. If energy transfer processes between two types of lanthanide activators exist, the populations of certain energy levels will have dramatic changes under excitation at various power density. As a result, visible upconversion emissions from the phosphors can be consecutively tuned over a wide range based on the population changing. This type of upconversion phosphors is promising for sensors to indicate power density variation.

Although nanocrystals have many applications, particle size, surface defects and adsorbed ligands determine quantum yield of nanocrystals [22–24]. The decrement of quantum yield is due to the quenching effect caused on the surface. Therefore, microcrystals that have low surface-to-volume ratio are preferred for the investigation on energy transfer to exclude the influence of surface [24].

Here, we report the energy transfer between Tm3+ and Tb3+ dependent on the power density of pump laser in NaYF4: Tb3+,Tm3+,Yb3+ microcrystals. Hexagonal phase NaYF4 is known as one of the most efficient host materials for upconversion [4,25]. The Tb3+,Tm3+,Yb3+-doped NaYF4 microcrystals were synthesized via a hydrothermal method with similar shape and surface-to-volume ratio. The samples showed intense upconversion emissions. The color of upconversion emissions were tuned from green to blue via increasing the power density of 976-nm pump laser. When the power density was low, the energy transfer from 3F4 (Tm3+) to 7F0 (Tb3+) was sufficient. The population on levels of Tm3+ were inhibited. Only emissions from Tb3+ were observed. When the power density of pump laser was increased, the number of Tb3+ ions on the ground state decreased, consequently, the energy transfer from 3F4 (Tm3+) to 7F0 (Tb3+) was relatively reduced. The high energy levels of Tm3+ were populated and emitted through radiative transitions which are dominant in emission spectrum. Energy transfer efficiency from Tm3+ to Tb3+ increased when power density of pump laser became higher. The cross-relaxation between Tm3+ and Tb3+, which determines the populations of energy levels for upconversion emissions, depends on the laser power density. This type of upconversion phosphors can be used for power density sensors in the future.

2. Experiments

The samples were synthesized via a hydrothermal method as previously reported [19]. Typically, 0.5 mmol of RE(NO3)3 aqueous solution (RE = 59 mol% Y, 20 mol% Tb, 1 mol% Tm, and 20 mol% Yb), and 0.5 mmol of EDTA were well mixed under vigorous stirring for 1 h. The rare earth complex (RE-complex) was formed. Then 6 mmol of NaF aqueous solution was added into the RE-complex and kept stirring for 1 h. The mixture was transferred into a 25 mL autoclave and kept at 180 °C for 12 h. The resulting products were collected by centrifugation, washed several times by distilled water, and finally obtained by drying at 60 °C. The procedure of synthesizing other samples were similar, except concentrations of lanthanide ions. The samples doped with 1%Tm3+/20%Yb3+, 10%Tb3+/1%Tm3+/20%Yb3+, 20%Tb3+/1%Tm3+/20%Yb3+, and 20%Tb3+/20%Yb3+ were denoted as TmYb, 10TbTmYb, 20TbTmYb, 20TbYb, respectively.

Crystal structures were characterized by a X-ray diffraction (XRD) on a LabX XRD-6100 X-ray diffractometer with a Cu Kα radiation source (λ=1.5405 Å) operated at 40 kV and 30 mA. The scan was performed in the arrange from 2θ = 10° to 80° with a scan speed of 0.02°/s in steps of 0.02°. Morphologies were observed on a field emission scanning electron microscope (FE-SEM, JEOL-7000F). Upconversion and near-infrared emission spectra were measured using a monochromator equipped with a photomultiplier tube (Hamamatsu). Absorption spectrum in the near-infrared region was measured on a spectrometer (Perkin Elmer, Lambda 900). A fiber coupled 976 nm laser diode (LD) with tunable power up to 154 mW was used as a pump source. Emission decay curves were recorded by a 200 MHz digital oscilloscope for temporal investigation.

3. Results and discussion

3.1. Crystal structure and morphology

Crystal structures of the as-prepared samples were characterized by a X-ray powder diffraction. Figure 1 shows the XRD patterns of samples TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb. The obtained patterns can be indexed to hexagonal phase NaYF4 (JCPDS 16-0334). Although Tb3+ and Yb3+ concentrations were high, up to 20%, no diffraction peaks of impurities were detected. It indicates that hexagonal phase NaYF4 was the main phase. Sharp diffraction peaks indicate that samples were crystallized well.

 figure: Fig. 1

Fig. 1 XRD patterns of Tb3+-Tm3+-Yb3+ doped NaYF4 microcrystals synthesized by a hydrothermal method: (a) TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb. Black sticks correspond to the diffraction peaks of hexagonal phase NaYF4 (JCPDS 16-0334).

Download Full Size | PDF

Morphologies of samples were all in the shape of hexagonal rod, as shown in Fig. 2. The averaged lengths of the rods were 12.5, 11.1, 9.6, and 7.0 µm and the averaged diameter of the rods were 6.0, 2.8, 4.0, and 2.6 µm for TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb samples, respectively. The surface-to-volume ratios of the as-prepared samples were about 0.0008, 0.0016, 0.0012, and 0.0018 nm−1. In the case of nanocrystals with the size of about 6 nm, the surface-to-volume ratio was about ∼ 0.9 nm−1 that was three orders of magnitude larger than those of microcrystals [23]. We have reported that surface adsorbed ligands had large quenching effect on nanocrystals [23,24]. To reduce the influence of surface quenching, microcrystals are more suitable owing to their low surface-to-volume ratio. The aspect ratio of crystal rods is another crucial factor. Gao et al. reported that the aspect ratio of NaYF4: Er3+,Yb3+ rods affected upconversion emission properties [26]. The aspect ratios were 2.1, 4.0, 2.4, and 2.7 for TmYb, 10TbTmYb, 20TbTmYb, 20TbYb samples, respectively. Since the as-prepared samples have the same shape, similar aspect and surface-to-volume ratio, the influence caused by crystal size and shape can be neglected.

 figure: Fig. 2

Fig. 2 SEM images of as-prepared Tb3+,Tm3+,Yb3+ doped NaYF4 microcrystals: (a) TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb.

Download Full Size | PDF

3.2. Emission properties

The samples showed intense visible and near-infrared emissions excited by a 976-nm laser, as shown in Fig. 3(a). The spectra were normalized according to the intensity of emission at 544 and 1220 nm in the visible and near-infrared region, respectively. In the TmYb sample, the emissions centered at 346, 362, 452, 476, 648, 690, and 803 nm are assigned to the radiative transitions of 1I63F4, 1D23H6, 1D23F4, 1G43H6, 1G43F4, 3F2,33H6, and 3H43H6 of Tm3+ ions, respectively [27,28].

 figure: Fig. 3

Fig. 3 (a) Emission spectra of as-prepared NaYF4 phosphors with various dopant (sample TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb) in the visible and near-infrared region excited by 976 nm laser. (b) Energy level diagram and possible upconversion processes among Tb3+, Tm3+, and Yb3+.

Download Full Size | PDF

In the Tb3+-Yb3+ co-doped sample, the emissions centered at 381, 417, 438, 490, 544, 586, and 621 nm are assigned to the radiative transitions 5D37FJ (J=6, 5, and 4) and 5D47FJ (J=6,5,4, and 3) of Tb3+ ions, respectively [29,30]. In the 20TbTmYb and 10TbTmYb samples, emissions generated from both Tm3+ and Tb3+ ions were observed. However, the intensities of emissions at 346, 360, and 452 nm from the 1I6 1D2, and 1G4 levels of Tm3+ were greatly reduced compared with the TmYb sample. When the Tb3+ concentration increased to 20%, the intensity ratio of emission at 476 nm to that at 544 nm clearly showed a decreasing tendency. The different spectral contribution implies the existence of energy transfer between Tm3+ and Tb3+. The emissions centered at 1170, 1220, 1470, and 1639 nm in the near-infrared region are assigned to the transition of 3F2,33F4, 3H53H6, 3H43F4 and 3F43H6 of Tm3+, respectively. The emission band at 1550 nm is caused by the slight amount of Er3+ impurity in rare materials. The tail near 1100 nm in the Tb3+-doped sample is caused by the strong pump laser and the spontaneous emissions from Yb3+. The TmYb sample shows intense near-infrared emissions while they were greatly quenched in Tb3+-Tm3+-Yb3+ samples. It implies that the Tm3+ might transfer energy to Tb3+.

Figure 3(b) schematically depicts possible upconversion and energy transfer processes among Tb3+, Tm3+, and Yb3+. In this system, Yb3+ ions act as sensitizers to absorb energy from the 976-nm laser irradiation, then transfer to the activators, Tb3+ and Tm3+. Electrons on the ground state of Tm3+ are excited to the 3H5 level by phonon-assisted energy transfer [31], then non-radiatively relaxed to the 3F4 level. The electrons on the 3F4 level are excited to the 3F2,3 level via energy transfer from Yb3+, relaxed to the 3H4 level and then pumped to the 1G4 level. Owing to the large energy mismatch between the 1D2 and 1G4 level of Tm3+ (about 3500 cm−1) [32], the 1D2 level cannot be directly excited from the 1G4 level by ETU process. It is ascribed to cross-relaxation processes between Tm3+ pairs. Four possible cross-relaxation approaches were reported as 1G4+3H41D2+3F4, 1G4+3H43F4+1D2, 3F2+3H43H6+1D2, and 3F3+3F33H6+1D2 [28,33,34]. Population of the 3F2,3 level is strongly quenched by multi-phonon relaxation process because of the small energy gap to the nearest low-lying level (about 1500 cm−1) [32]. Thus, the latter two approaches related to the 3F2,3 level cannot be taken into account [28]. We attribute the 3F2,3 level is populated by the former two cross-relaxation processes. The 1I6 level is excited by energy transfer from Yb3+. Thus, emissions from the 1I6 (346 nm), 1D2 (360 and 452 nm), and 1G4 level (474 and 649 nm) are generated by a five-, four-, and three-photon upconversion process, respectively [27]. Cooperative sensitizing upconversion process by two Yb3+ ions is the main mechanism to populate the 5D4 level of Tb3+. Through an excited state absorption or energy transfer from Yb3+, the 5D3 level was further populated [30,35].

Some energy levels of Tm3+ and Tb3+ have similar positions, i.e. 1I6-5H4, 1D2-5G5, 1G4-5D4, and 3F4-7F0. The energy transfer between Tm3+ and Tb3+ will occur effectively, which is confirmed from the spectral difference in Fig. 3.

Owing to the effective energy transfer from the 1I6 and 1D2 levels, emission intensity in the 10TbTmYb sample is very weak, and that in the 20TbTmYb sample is hardly to be observed, as shown in Fig. 3(a). The relative intensity of the emission generated from the 3F4 level of Tm3+ decreased in the Tb3+-Tm3+-Yb3+ tri-doped samples, which indicates the energy transfer from Tm3+ to Tb3+. Because of the narrow energy gap to the nearest low-lying levels, nonradiative relaxation is the main depopulating process for the 7F0−5 levels. There are no suitable levels in Tb3+ which match the 3F2,3, and 3H4 level of Tm3+. Thus, the decrements of relative intensity were not caused by resonant energy transfer, but by the cross-relaxation between Tb3+ and Tm3+. This will be discussed in the next subsection.

We attempted to clarify the energy transfer processes by changing concentration of acceptor Tb3+ ions. It is difficult to clearly demonstrate energy transfer processes between Tm3+ and Tb3+, because Tm3+ and Tb3+ both can be pumped by the ETU from Yb3+. As well known, population of electrons on excited states in upconversion processes is proportional to the power density of pump light [20]. If the power density is low, the population of electrons on high levels of Tm3+ and Tb3+ will be relatively low. In this case, the change in electronic population caused by energy transfer on certain levels will be more obvious.

Two approaches were utilized to adjust power density of pump laser: controlling spot area with fixed power and changing power with fixed spot area. First, we fixed power of 976-nm laser about 154 mW, and then changed the distance between sample and the laser. Figure 4 shows upconversion emission spectra measured with various power densities. Spectra of the TmYb sample and all Tb3+-doped samples were normalized according to the intensity of emission at 476 (1G4, Tm3+), and 544 nm (5D4, Tb3+), respectively. Electronic population on the 1D2 and 1I6 level of Tm3+ and the 5D3 level of Tb3+ became large when laser power density increased. The emission intensity is enhanced when the power density of pumping laser increased from 3 to 122 W/cm2 as shown in Fig. 4(a) and 4(d). The emissions from Tm3+ and Tb3+ were observed in the 10TbTmYb and 20TbTmYb samples, as shown in Figs. 4(b) and 4(c). However, the intensity ratios of emission at 476 nm (Tm3+) to 544 nm (Tb3+) were dependent on power density of pump laser. The intensity ratio was used to illustrate the contribution of Tm3+ and Tb3+. When the power density was 3 W/cm2, emissions of Tm3+ can be easily observed in the TmYb sample as shown in Fig. 4(a). However, the intensity of emission at 476 nm in the 20TbTmYb sample was significantly reduced, and only emissions from Tb3+ were observed under the same excitation conditions. It implies that the energy transfer from Tm3+ to Tb3+ is efficient in the case of low power density.

 figure: Fig. 4

Fig. 4 Upconversion emission spectra of samples (a)TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb excited by 976-nm laser with various power density. The chromaticity diagrams of each sample were inset correspondingly. The black vertical arrows indicate the power density increasing from 3 to 122 W/cm2.

Download Full Size | PDF

The chromaticity diagrams of samples are inset in Fig. 4. The intensity of blue emissions from the TmYb and 20TbYb sample increased with increasing the power density. Chromaticity coordinates of the TmYb sample shifts slightly within the blue region and the chromaticity coordinates of 20TbYb sample shifts towards white region. In the 10TbTmYb and 20TbTmYb samples, the contribution of blue emissions from Tm3+ was greatly enhanced with increasing power density. The coordinates of the 20TbTmYb sample will shift from the green towards white region and finally to the blue region with the power density increased from about 3 to 122 W/cm2 as shown in Fig. 4(c). Under similar excitation conditions, however, the coordinates of the 10TbTmYb sample transit from the blue to cyan region. It illustrates that excessive Tb3+ could effectively quench the electronic populations of Tm3+ excited states.

Figure 5 shows the emissions in the near-infrared region depends on power density. These spectra were normalized according to the peak intensity of emission at 1220 nm. The intensities of emission from 3F2,3, 3H4, and 3F4 in the TmYb sample were increased with increasing power density. No emission from Tb3+ was observed in the 20TbYb sample, except weak emission band from Er3+ impurity. In the discussions, we can ignore the influence from Er3+ impurity. Emission peak from Er3+ was observed in samples doped with Tb3+, as shown in Figs. 3(a) and 5. However, its concentration was too low to affect upconversion processes in Tm3+ and Tb3+. As shown in Fig. 4, no upconversion emissions from Er3+ were observed. It seems that only the 4I13/2 level of Er3+ was populated. In the next section we propose energy transfer process from the 3H4 level of Tm3+ to the 7F0 level of Tb3+ which affects their upconversion processes. The 4I13/2 level of Er3+ is higher than the 3H4 level of Tm3+ and the 7F0 level of Tb3+ which means energy transfer from Er3+ to Tm3+ and Tb3+ should be more sufficient than energy back transfer to Er3+. Thus, the tiny amount of Er3+ impurity did not seriously affect energy transfer process between Tm3+ and Tb3+. Although intensities were weak, emissions from the 3F2,3, 3H4, and 3F4 level in the 10TbTmYb sample could be observed when the power density was high. In the 20TbTmYb sample, however, they cannot be detected even under high power density pumping. Only the transition 3H53H6 with weak intensity was recorded. It indicates the energy transfer from Tm3+ to Tb3+ was more sufficient in samples with higher Tb3+ concentration.

 figure: Fig. 5

Fig. 5 Emission spectra of samples (a) TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb in near-infrared region excited by 976-nm laser. The black vertical arrows indicate the power density increasing from 3 to 122 W/cm2.

Download Full Size | PDF

3.3. Energy transfer between Tm3+ and Tb3+

Some cross-relaxation approaches are proposed to explain the power density dependent color tuning feature of the as-prepared sample as follows,

 5D3+ 7F6 5D4+ 7F0(Tb3+Tm3+), 5D4+ 3H6 4F0+ 3F2(Tb3+Tm3+), 3H4+ 7F6 3H5+ 7F3(Tm3+Tb3+), 5F2+ 7F6 3H4+ 7F5(Tm3+Tb3+).

Besides cross-relaxation between Tm3+ and Tb3+, self-quenching processes between Tm3+ pairs and Tb3+ pairs also affect electronic populations, which are described as follows,

 3H4+ 3H6 3F4+ 3F4(Tm3+Tm3+), 5D3+ 7F6 5D4+ 7F0(Tb3+Tb3+).

Figure 6 schematically shows the cross-relaxation and laser power density dependent energy transfer processes. The existence of energy transfers between Tm3+ and Tb3+ is inferred according to the above subsection. The electrons on a certain level will be depopulated via several approaches: radiative transition and nonradiative relaxation to low-lying levels, being pumped to higher levels, cross-relaxation between ion pairs, and phonon-assisted energy transfer processes. The difference in spectral profiles can be used to investigate and gain insight into the dynamics of electronic populations of energy levels. If the emissions from an excited state cannot be observed, it implies that electrons are depopulated by nonradiative relaxation, cross-relexation, or energy transfer processes.

 figure: Fig. 6

Fig. 6 (a) Energy level diagram, upconversion processes and possible energy transfer between Tb3+, Tm3+, and Yb3+. (b) Schematic illustration of energy transfer between Tm3+ and Tb3+ under low and high power density pumping.

Download Full Size | PDF

In Fig. 4, the TmYb and 20TbYb samples show the upconvesion emissions from Tm3+ and Tb3+ are independent of the power density of pump laser, whereas the 10TbTmYb and 20TbTmYb samples show only Tb3+ emissions when the pump power density is low. The simple speculation that all the elections on Tm3+ were depopulated via energy transfer to Tb3+ by 1I65D3, 1D25D3, and 1G45D4, as reported in [36], cannot perfectly explain this phenomenon. This is because it is impossible to excite only high energy levels without populating low energy levels such as 3F2,3, 3H4, 3H5, and 3F4 of Tm3+, in upconverson processes. The reported energy transfer processes probably have no influence on the low lying levels to change their electronic population. It other words, it is possible to observe the emissions from these levels, for example, at 696, 803, and 1220 nm. This is inconsistent with the experimental results that no emissions from Tm3+ were observed when the pump power density was lower than 5 W/cm2.

The only possibility is that the upconversion processes in Tm3+ were inhibited. For low pump power density, the population of the 3F4 level was too low to excite the 3F2,3 levels and other high energy levels. Energy transfer from 3F4 (Tm3+) to 7F0 (Tb3+) is the main process to reduce the population of 3F4 [37]. It is well known that energy transfer probability depends on the distance between sensitizer and activator and the overlap of emission spectrum of sensitizer and absorption spectrum of activator [38]. Figure 7 shows the overlap of emission spectrum of the TmYb sample and absorption spectrum of TbYb sample in the near-infrared region (1600–2400 nm). The emission band centered at 1639 nm is assigned to the transition 3F43H6 of Tm3+. The absorption bands centered at 1667, 1723, 1932, and 2264 nm are assigned to the transition from the 7F6 to 7F0−3 level of Tb3+, respectively. The spectral overlap (1600–1800 nm) implies the possibility of energy transfer between the 3F4 and 7F0 levels. The cross-relaxation process between Tm3+ pairs (3H4+3H63F4+3F4) concentrates electrons from the low energy levels of Tm3+ on the 3F4 level. According to the multiphonon relaxation, the electrons on the 7F0 level will soon relax to the ground state because of the high multiphonon relaxation rate caused by the narrow energy gaps to the nearest low-lying levels. The energy transfer process from the 3F4 to the 7F0 levels was enhanced. As a result, the 3F4 level was greatly depopulated.

 figure: Fig. 7

Fig. 7 Emission spectrum in the 1600–1700 nm range in TmYb excited by 976-nm laser, and absorption spectrum of 20TbYb in 1600–2400 nm range.

Download Full Size | PDF

The electrons on the 5D3 level of Tb3+ will be quenched to the 5D4 level by cross-relaxation ➀ and self-quenching ⓑ. The electrons on the 5D4 level will be quenched by cross-relaxation ➁ to the 7F0 level, then fast relax to the ground state by multi-phonon relaxation processes. For Tm3+ ions, the 1I6, 1D2, and 1G4 levels are depopulated and transfer the energy to the 5H4, 5G5, and 5D4 levels of Tb3+ by phonon-assisted energy transfer. The electrons on the 3F2,3 and 3H4 level are quenched to the 3F4 level by the cross-relaxation process ➂ and ➃ and self-quenching ⓐ, then depopulated by the energy transfer process, 3F47F0.

Since the measurements were carried out by recording the scattered emission lights, the comparison of peak intensities among samples cannot give reliable information. The integrated intensity ratios were used to analyze the change of emissions from different energy levels. Figure 8(a) shows the ratios of emission at 476 nm relative to that of 544 nm with varying power density of pump laser.

 figure: Fig. 8

Fig. 8 Dependence of integrated intensity ratio of (a) emission at 476 nm versus 544 nm, (b) Tm3+ versus Tb3+ emissions, (c) emissions from the 1I6 level versus those from the 1G4 level, (d) emissions from the 1D2 level versus those from the 1G4 level, (e) emissions from the 1G4 level versus those from the 3H4 level, and (f) emissions from the 5D3 level versus those from the 5D4 level on power density of pump laser.

Download Full Size | PDF

The high ratio shows that the contribution of Tm3+ was larger in the 10TbTmYb sample. Increasing tendency of the ratios in both samples implies the contribution of Tm3+ increases with power density. The ratio of Tm3+ to Tb3+ emission in Fig. 8(b) confirms that the contribution of Tm3+ in 10TbTmYb is higher and it increases with power density in both samples. As power density became higher, more ions were pumped to excited states. If the population on the ground states becomes less, the cross-relaxation processes ➀ – ➃ between Tm3+ and Tb3+ will be insufficient. Thus, contribution of emissions from Tm3+ becomes higher. The excessive amount of Tb3+ in 20TbTmYb guarantees the numbers of ions on the ground state for the cross-relaxation and the sufficient energy transfer from Tm3+ to Tb3+. Most of the electronic populations on the Tm3+ levels are quenched by Tb3+. As a result, the ratio of Tm3+ to Tb3+ emission is low.

When the power density increases, the quenching effect from Tb3+ was inhibited, and the contribution of Tm3+ emission increased significantly. The quenching effect was investigated by the intensity ratios of emissions from different energy levels in Tm3+, as shown in Figs. 8(b) and 8(c). Similar to the above discussion, the ratio of 1I6 and 1D2 emissions to 1G4 emissions decreased as increasing the Tb3+ concentration from 0 to 20%. The energy transfer from the 1I6 and 1D2 levels of Tm3+ to the 5H4 and 5G5 levels of Tb3+ effectively quenched their populations. As the power density was increased, the upconversion process in Tm3+ was enhanced. The ratio of emissions from the 1I6 and 1D2 levels increased in the Tb3+-doped samples, but the slope values became lower. However, the ratios of emissions from the 1G4 level versus those from the 3H4 level in Tb3+-doped samples are large. It implies the increment of emissions from the 1G4 level or the decrement of emissions from the 3H4 level. The population on the 1G4 level will be quenched by energy transfer to 5D4 level of Tb3+. Thus, the cross-relaxation process ➃ greatly reduces the population on the 3H4 level.

Figure 8(f) shows the 5D3 to 5D4 emission intensity ratio. As power density increases, contribution of emissions from the 5D3 level gradually increased. In the low power density case, the higher slope value in 20TbTmYb implies that the Tm3+ energy transfer contribute to the population of the 5D3 level. When power density was high, the ratio was saturated in the 20TbYb but shows decreasing tendency in 20TbTmYb. Self-quenching ⓑ from Tb3+ occurs in both samples. Owing to the cross-relaxation processes ➀ and ➁, however, the population on the 5D3 and 5D4 level will decrease. The decreasing tendency shows that 5D3 level was quenched more effectively.

The upconversion process can be confirmed by fitting required photon number based on the dependence of integrated emission intensity on the power of pump laser [20,39]. The number of required photons to populate a certain energy level can be obtained, according to the relation If = Pn, where If is the emission intensity, P is the pump power, and n is the number of photons. The n values for the emissions at 345 (1I63F4 in Tm3+), 452 (1D23F4 in Tm3+), 476 (1G43H6 in Tm3+), and 803 nm (3H43H6 in Tm3+) in TmYb were 4.3 ± 0.3, 3.9 ± 0.3, 2.9 ± 0.2, and 1.9 ± 0.1, respectively. It confirms these levels were populated by five-, four-, three-, and two-photon upconversion processes [40].

The n values of the emission at 381 (5D37F6 in Tb3+), and 544 nm (5D47F6 in Tb3+) in 20TbYb were about 2.7 ± 0.2 and 2.0 ± 0.1. It shows that the 5D3 and 5D4 level were populated by three- and two-photon upconversion processes. The n values for the Tm3+ emissions in the 10TbTmYb and 20TbTmYb samples were similar to those in TmYb sample, as shown in the inset values in Fig. 9. It implies that energy transfer back to the 1I6, 1D2, and 1G4 level of Tm3+ has little influence. The 3H4 level could be populated by electrons relaxed from the 3F2,3 level excited by cross-relation process ➁ involving three photons. Thus, the n value of 803 nm emission in the Tb3+-doped sample slightly increased. Moreover, the n values of Tb3+ emissions were increased with increasing the molar ratio of Tm3+ to Tb3+. It was 3.3 ± 0.3 and 3.4 ± 0.2 for emissions at 381 nm and 2.3 ± 0.2 and 3.2 ± 0.3 for emissions at 544 nm in 20TbTmYb and 10TbTmYb, respectively. It indicates that the 5D3 level requires more than three photons to be excited and the 5D4 level needs more than two photons to be excited. The energy transfers from Tm3+ to Tb3+ are confirmed.

 figure: Fig. 9

Fig. 9 Log-Log plot of upconversion emission intensity versus excitation power density for the (a) 1I63F4 (345 nm), (b) 1D23F4 (452 nm), (c) 1G43H6 (476 nm), and (d) 3H43H6 (803 nm) transition of Tm3+, the (e) 5D37F6 (381 nm) and (f) 5D47F6 (544 nm) transition of Tb3+ in TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb samples.

Download Full Size | PDF

The fluorescence lifetimes were obtained by fitting the measured emission decay curves with double-exponential function. Averaged lifetimes (τav) were estimated using the following expression [41,42],

τav=Aiτi2Aiτi,
where τi and Ai are the multiple lifetime values and the corresponding weighed amplitudes. Figure 10 shows the averaged fluorescence lifetimes of emissions at 348, 452, 476, and 803 nm for Tm3+ emissions, 381 and 544 nm for Tb3+ emissions dependent on power density of pump laser. A total decay rate from an energy level, which is equal to a reciprocal of measured fluorescence lifetime (τm) can be written as [43,44],
1τm=AR+WNR+WCR,
where AR, WNR, WCR represents the radiative, nonradiative transition, and cross-relaxation rate. WCR involves the contribution of cross-relaxation between Tm3+ and Tb3+ and self-quenching in Tm3+ and Tb3+. Because of the strong influence of cross-relaxation between Tm3+ and Tb3+, fluorescence lifetimes in Tm3+-Tb3+-Yb3+ tri-doped significantly reduced, compared to the Tm3+-Yb3+ and Tb3+-Yb3+ dual-doped samples, as shown in Fig. 10.

 figure: Fig. 10

Fig. 10 Fluorescent lifetime of the emissions at (a) 345 nm, (b) 452 nm, (c) 476 nm, and (d) 803 nm, (e) 381 nm and (f) 544 nm versus excitation power density in TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb samples.

Download Full Size | PDF

The analysis of fluorescence lifetime confirms the presence of energy transfer and cross-relaxation processes between Tb3+ and Tm3+, which are inferred according to the analysis in Fig. 8. The Tm3+-Tb3+-Yb3+ tri-doped samples were considered as doping Tb3+ into Tm3+-Yb3+ dual-doped sample and doping Tm3+ into Tb3+-Yb3+ dual-doped sample. Although complicated energy transfer processes occurred among the three types of lanthanide ions, we discuss the quenching effect of Tb3+ to the Tm3+-Yb3+ system and Tm3+ to the Tb3+-Yb3+ system, respectively.

Energy transfer efficiency from Tb3+ to Tm3+ and backward from Tm3+ to Tb3+ could be calculated based on the obtained fluorescent lifetime by,

ηTbTm=1τTmYbTbτTmYb,
ηTmTb=1τTbYbTmτTbYb,
where ηTb−Tm and τTmYb represents the energy transfer efficiency from Tb3+ to Tm3+ and from Tm3+ to Tb3+, respectively. The τTmYbTb and τTbYb represents the lifetime in the presence and absence of Tb3+, and τTmYbTb and τTbYb represents the lifetime in the presence and absence of Tm3+.

Figure 11 shows that the estimated energy transfer efficiencies in 10TbTmYb and 20TbTmYb samples increase with increasing power density of pump laser. The energy transfer efficiency from the 3H4 to 7F0 level was around 70%. When the energy level position of Tm3+ enhanced from low to high (3H4, 1G4, 1D2, and 1I6), the transfer efficiencies were reduced. The electrons on the 3H4 level was sufficiently depopulated via cross-relaxation processes ➂ and ➃. Therefore, the upconversion processes were inhibited. The efficiencies from the 5D3 and 5D4 levels, which are approximately 20–30%, demonstrate the presence of cross-relaxation process ➀ and ➁. The average lifetimes of the 5D4 and 3H4 level are invariant around 3700 and 650 µs, respectively. The ratio of Tm3+ ions on the ground state should be relatively larger. Consequently, cross-relaxation processes ➀ and ➁ are effective, the electrons on the 5D3 and 5D4 level will be relaxed to the 7F0 level, and then finally nonradiatively relaxed to the ground state. This process helps to increase the number of Tb3+ on the ground state, which will enhance the energy transfer and cross-relaxation processes from Tm3+ to Tb3+. Therefore, the energy transfer efficiency from Tm3+ to Tb3+ and Tb3+ to Tm3+ of all levels increase with increasing power density of pump laser.

 figure: Fig. 11

Fig. 11 Energy transfer efficiency between Tm3+ and Tb3+ calculated from fluorescent lifetime of emissions generated from various energy levels in (a) 10TbTmYb and (b) 20TbTmYb sample.

Download Full Size | PDF

Quantum yield Φ is given by the radiative and nonradiative transition rate as,

Φ=τmτR,
where τR is radiative lifetime and the reciprocal of radiative rate AR. The quantum yield can be rewritten according to Eq. (2) as,
Φ=ARAm=ARAR+WNR+WCR.

Owing to the strong quenching effect caused by the cross-relaxation between Tm3+ and Tb3+, the radiative rate is determined by crystal field on dopant ions; the nonradiative transition rate is decided by phonon energy of host, temperature, and energy gap to the nearest low-lying level. Although they will be affected after inducing the third type of lanthanide ions, the strong influence of quenching effect caused by the cross-relaxation between Tm3+ and Tb3+ can dominate. The WCR term is proportional to the energy transfer efficiency, which increases with increasing power density as shown in Fig. 11. Therefore, it is predicted that quantum yield in tri-doped sample will be much lower than that in dual-doped samples and will be reduced when the power density of pump laser becomes higher.

4. Conclusions

In conclusion, Tb3+,Tm3+,Yb3+ tri-doped NaYF4 microcrystal rods were synthesized via a hydrothermal method. The as-prepared samples showed intense visible upconversion emissions. The energy transfer process between Tm3+ and Tb3+ depends on the power density of pump laser. Upconversion emissions of samples can be tuned by the laser power density dependent energy transfer process. When the power density was low, energy transfer from the 3F4 (Tm3+) to 7F0 (Tb3+) level was sufficient. As a result, upconversion processes in Tm3+ were inhibited. Because of the low population on the 3F4 level, green color from Tb3+ was observed. When the power density became high, the energy transfer processes between low energy levels of Tm3+ and Tb3+ were insufficient, the blue upconversion emissions of Tm3+ were significantly enhanced. These phosphors with tunable visible upconversion emissions dependent on power density of pump laser would be utilized for novel laser power density sensors in the future.

Funding

This work was supported by JSPS and CAS under the JSPS – CAS Joint Research Program, JSPS KAKENHI Grant Number 26889058, 15H02250, and National Natural Science Foundation of China (NSFC) (Grant Nos. 11374084 and 61307056).

References and links

1. A. Gnach and A. Bednarkiewicz, “Lanthanide-doped up-converting nanoparticles: Merits and challenges,” Nano Today 7, 532–563 (2012). [CrossRef]  

2. H. H. Gorris and O. S. Wolfbeis, “Photon-upconverting nanoparticles for optical encoding and multiplexing of cells, biomolecules, and microspheres,” Angew. Chem., Int. Ed. 52, 3584–3600 (2013). [CrossRef]  

3. F. Auzel, “Upconversion and anti-stokes processes with f and d ions in solids,” Chem. Rev. 104, 139–174 (2004). [CrossRef]   [PubMed]  

4. K. W. Krämer, D. Biner, G. Frei, H. U. Güdel, M. P. Hehlen, and S. R. Lüthi, “Hexagonal sodium yttrium fluoride based green and blue emitting upconversion phosphors,” Chem. Mater. 16, 1244–1251 (2004). [CrossRef]  

5. J. Suyver, A. Aebischer, D. Biner, P. Gerner, J. Grimm, S. Heer, K. Krämer, C. Reinhard, and H. Güdel, “Novel materials doped with trivalent lanthanides and transition metal ions showing near-infrared to visible photon upconversion,” Opt. Mater. 27, 1111–1130 (2005). [CrossRef]  

6. W. Gao, R. Wang, Q. Han, J. Dong, L. Yan, and H. Zheng, “Tuning red upconversion emission in single LiYF4:Yb3+/Ho3+ microparticle,” J. Phys. Chem. C 119, 2349–2355 (2015).

7. D. Li, W. Qin, D. Zhao, T. Aidilibike, H. Chen, S. Liu, P. Zhang, and L. Wang, “Tunable green to red upconversion fluorescence of water-soluble hexagonal-phase core-shell CaF2@NaYF4 nanocrystals,” Opt. Mater. Express 6, 270–278 (2016). [CrossRef]  

8. F. Wang, R. Deng, J. Wang, Q. Wang, Y. Han, H. Zhu, X. Chen, and X. Liu, “Tuning upconversion through energy migration in core–shell nanoparticles,” Nat. Mater. 10, 968–973 (2011). [CrossRef]   [PubMed]  

9. L. Wang, X. Xue, H. Chen, D. Zhao, and W. Qin, “Unusual radiative transitions of Eu3+ ions in Yb/Er/Eu tri-doped NaYF4 nanocrystals under infrared excitation,” Chem. Phys. Lett. 485, 183–186 (2010). [CrossRef]  

10. L. Wang, X. Xue, F. Shi, D. Zhao, D. Zhang, K. Zheng, G. Wang, C. He, R. Kim, and W. Qin, “Ultraviolet and violet upconversion fluorescence of europium (III) doped in YF3 nanocrystals,” Opt. Lett. 34, 2781–2783 (2009). [CrossRef]   [PubMed]  

11. S. F. León-Luis, U. R. Rodríguez-Mendoza, E. Lalla, and V. Lavín, “Temperature sensor based on the Er3+ green upconverted emission in a fluorotellurite glass,” Sens. Actuators B 158, 208–213 (2011). [CrossRef]  

12. B. P. Singh, A. K. Parchur, R. S. Ningthoujam, P. V. Ramakrishna, S. Singh, P. Singh, S. B. Rai, and R. Maalej, “Enhanced up-conversion and temperature-sensing behaviour of Er3+ and Yb3+ co-doped Y2Ti2O7 by incorporation of Li+ ions,” Phys. Chem. Chem. Phys. 16, 22665–22676 (2014). [CrossRef]   [PubMed]  

13. S. Zhou, K. Deng, X. Wei, G. Jiang, C. Duan, Y. Chen, and M. Yin, “Upconversion luminescence of NaYF4: Yb3+, Er3+ for temperature sensing,” Opt. Commun. 291, 138–142 (2013). [CrossRef]  

14. H. Zheng, B. Chen, H. Yu, J. Zhang, J. Sun, X. Li, M. Sun, B. Tian, S. Fu, and H. Zhong, “Microwave-assisted hydrothermal synthesis and temperature sensing application of Er3+/Yb3+ doped NaY(WO4)2 microstructures,” J. Colloid Interface Sci. 420, 27–34 (2014). [CrossRef]   [PubMed]  

15. H. Zheng, B. Chen, H. Yu, X. Li, J. Zhang, J. Sun, L. Tong, Z. Wu, H. Zhong, and R. Hua, “Rod-shaped NaY(MoO4)2:Sm3+/Yb3+ nanoheaters for photothermal conversion: Influence of doping concentration and excitation power density,” Sens. Actuators B 234, 286–293 (2016). [CrossRef]  

16. N. Ishiwada, S. Fujioka, T. Ueda, and T. Yokomori, “Co-doped Y2O3:Tb3+/Tm3+ multicolor emitting phosphors for thermometry,” Opt. Lett. 36, 760–762 (2011). [CrossRef]   [PubMed]  

17. X. Zhai, S. Liu, Y. Zhang, G. Qin, and W. Qin, “Controlled synthesis of ultrasmall hexagonal NaTm0.02Lu0.98–xYbxF4 nanocrystals with enhanced upconversion luminescence,” J. Mater. Chem. C 2, 2037–2044 (2014). [CrossRef]  

18. X. Xue, L. Wang, L. Huang, D. Zhao, and W. Qin, “Effect of alkali ions on the formation of rare earth fluoride by hydrothermal synthesis: structure tuning and size controlling,” Cryst. Eng. Commun. 15, 2897–2903 (2013). [CrossRef]  

19. X. Xue, T. Cheng, T. Suzuki, and Y. Ohishi, “Upconversion emissions from high energy levels of Tb3+ under near-infrared laser excitation at 976 nm,” Opt. Mater. Express 5, 2768–2776 (2015). [CrossRef]  

20. M. Pollnau, D. R. Gamelin, S. R. Lüthi, H. U. Güdel, and M. P. Hehlen, “Power dependence of upconversion luminescence in lanthanide and transition-metal-ion systems,” Phys. Rev. B 61, 3337–3346 (2000). [CrossRef]  

21. J. Zhao, D. Jin, E. P. Schartner, Y. Lu, Y. Liu, A. V. Zvyagin, L. Zhang, J. M. Dawes, P. Xi, J. A. Piper, E. M. Goldys, and T. M. Monro, “Single-nanocrystal sensitivity achieved by enhanced upconversion luminescence,” Nat. Nanotechnol. 8, 729–734 (2013). [CrossRef]   [PubMed]  

22. J. Zhao, Z. Lu, Y. Yin, C. McRae, J. A. Piper, J. M. Dawes, D. Jin, and E. M. Goldys, “Upconversion luminescence with tunable lifetime in NaYF4:Yb,Er nanocrystals: role of nanocrystal size,” Nanoscale 5, 944–952 (2013). [CrossRef]  

23. X. Xue, T. Suzuki, R. N. Tiwari, M. Yoshimura, and Y. Ohishi, “Quenching effect of surface adsorbed ligands on luminescence of α-NaYF4:Nd3+ nanocrystals,” Jpn. J. Appl. Phys. 53, 075001 (2014). [CrossRef]  

24. X. Xue, S. Uechi, R. N. Tiwari, Z. Duan, M. Liao, M. Yoshimura, T. Suzuki, and Y. Ohishi, “Size-dependent upconversion luminescence and quenching mechanism of LiYF4: Er3+/Yb3+ nanocrystals with oleate ligand adsorbed,” Opt. Mater. Express 3, 989–999 (2013). [CrossRef]  

25. J. Suyver, J. Grimm, M. van Veen, D. Biner, K. Krämer, and H. Güdel, “Upconversion spectroscopy and properties of NaYF4 doped with Er3+, Tm3+ and/or Yb3+,” J. Lumin. 117, 1–12 (2006). [CrossRef]  

26. D. Gao, X. Zhang, and W. Gao, “Tuning upconversion emission by controlling particle shape in NaYF4:Yb3+/Er3+ nanocrystals,” J. Appl. Phys. 111, 033505 (2012). [CrossRef]  

27. G. Wang, W. Qin, L. Wang, G. Wei, P. Zhu, and R. Kim, “Intense ultraviolet upconversion luminescence from hexagonal NaYF4:Yb3+/Tm3+ microcrystals,” Opt. Express 16, 11907–11914 (2008). [CrossRef]   [PubMed]  

28. B. Zhou, L. Tao, W. Jin, and Y. H. Tsang, “Intense near-uv upconversion luminescence in Tm3+/Yb3+ co-doped low-phonon-energy lithium gallogermanate oxide glass,” IEEE Photonics Technol. Lett. 24, 1726–1729 (2012). [CrossRef]  

29. H. Liang, G. Chen, L. Li, Y. Liu, F. Qin, and Z. Zhang, “Upconversion luminescence in Yb3+/Tb3+-codoped monodisperse NaYF4 nanocrystals,” Opt. Commun. 282, 3028–3031 (2009). [CrossRef]  

30. X. Xue, M. Liao, R. N. Tiwari, M. Yoshimura, T. Suzuki, and Y. Ohishi, “Intense ultraviolet and blue upconverison emissions in Tb3+/Yb3+ codoped KY3F10 nanocrystals,” Appl. Phys. Express 5, 092601 (2012). [CrossRef]  

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

32. W. Carnall, H. Crosswhite, and H. M. Crosswhite, “Energy level structure and transition probabilities in the spectra of the trivalent lanthanides in LaF3,” Tech. Rep. (1978).

33. D. Chen, Y. Wang, Y. Yu, and P. Huang, “Intense ultraviolet upconversion luminescence from Tm3+/Yb3+:β-YF3 nanocrystals embedded glass ceramic,” Appl. Phys. Lett. 91, 051920 (2007). [CrossRef]  

34. F. Pandozzi, F. Vetrone, J.-C. Boyer, R. Naccache, J. A. Capobianco, A. Speghini, and M. Bettinelli, “A spectroscopic analysis of blue and ultraviolet upconverted emissions from Gd3Ga5O12:Tm3+, Yb3+ nanocrystals,” J. Phys. Chem. B 109, 17400–17405 (2005). [CrossRef]  

35. J. Zhang, Z. Hao, X. Zhang, Y. Luo, X. Ren, X.-j. Wang, and J. Zhang, “Color tunable phosphorescence in KY3F10:Tb3+ for x-ray or cathode-ray tubes,” J. Appl. Phys. 106, 034915 (2009). [CrossRef]  

36. L. Wang, D. Zhang, D. Zhao, G. Qin, and W. Qin, “Sensitized upconversion emissions of Tb3+ by Tm3+ in YF3 and NaYF4 nanocrystals,” J. Nanosci. Nanotechnol. 11, 9580–9583 (2011). [CrossRef]  

37. T. Sakamoto, M. Shimizu, M. Yamada, T. Kanamori, Y. Ohishi, Y. Terunuma, and S. Sudo, “35-db gain Tm-doped ZBLYAN fiber amplifier operating at 1.65 µ m,” IEEE Photonics Technol. Lett. 8, 349–351 (1996). [CrossRef]  

38. E. Nakazawa and S. Shionoya, “Energy transfer between trivalent rare-earth ions in inorganic solids,” J. Chem. Phys. 47, 3211–3219 (1967). [CrossRef]  

39. J. Suyver, J. Grimm, K. Krämer, and H. Güdel, “Highly efficient near-infrared to visible up-conversion process in NaYF4:Er3+,Yb3+,” J. Lumin. 114, 53–59 (2005). [CrossRef]  

40. X. Chen and Z. Song, “Study on six-photon and five-photon ultraviolet upconversion luminescence,” J. Opt. Soc. Am. B 24, 965–971 (2007). [CrossRef]  

41. A. Sillen and Y. Engelborghs, “The correct use of “average” fluorescence parameters,” Photochem. Photobiol. 67, 475–486 (1998).

42. N. M. Sangeetha and F. C. J. M. van Veggel, “Lanthanum silicate and lanthanum zirconate nanoparticles co-doped with Ho3+ and Yb3+: Matrix-dependent red and green upconversion emissions,” J. Phys. Chem. C 113, 14702–14707 (2009). [CrossRef]  

43. G. Liu and B. Jacquier, Spectroscopic Properties of Rare Earths in Optical Materials (Springer, 2006).

44. H. Toratani, T. Izumitani, and H. Kuroda, “Compositional dependence of nonradiative decay rate in nd laser glasses,” J. Non-Cryst. Solids 52, 303–313 (1982). [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 (11)

Fig. 1
Fig. 1 XRD patterns of Tb3+-Tm3+-Yb3+ doped NaYF4 microcrystals synthesized by a hydrothermal method: (a) TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb. Black sticks correspond to the diffraction peaks of hexagonal phase NaYF4 (JCPDS 16-0334).
Fig. 2
Fig. 2 SEM images of as-prepared Tb3+,Tm3+,Yb3+ doped NaYF4 microcrystals: (a) TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb.
Fig. 3
Fig. 3 (a) Emission spectra of as-prepared NaYF4 phosphors with various dopant (sample TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb) in the visible and near-infrared region excited by 976 nm laser. (b) Energy level diagram and possible upconversion processes among Tb3+, Tm3+, and Yb3+.
Fig. 4
Fig. 4 Upconversion emission spectra of samples (a)TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb excited by 976-nm laser with various power density. The chromaticity diagrams of each sample were inset correspondingly. The black vertical arrows indicate the power density increasing from 3 to 122 W/cm2.
Fig. 5
Fig. 5 Emission spectra of samples (a) TmYb, (b) 10TbTmYb, (c) 20TbTmYb, and (d) 20TbYb in near-infrared region excited by 976-nm laser. The black vertical arrows indicate the power density increasing from 3 to 122 W/cm2.
Fig. 6
Fig. 6 (a) Energy level diagram, upconversion processes and possible energy transfer between Tb3+, Tm3+, and Yb3+. (b) Schematic illustration of energy transfer between Tm3+ and Tb3+ under low and high power density pumping.
Fig. 7
Fig. 7 Emission spectrum in the 1600–1700 nm range in TmYb excited by 976-nm laser, and absorption spectrum of 20TbYb in 1600–2400 nm range.
Fig. 8
Fig. 8 Dependence of integrated intensity ratio of (a) emission at 476 nm versus 544 nm, (b) Tm3+ versus Tb3+ emissions, (c) emissions from the 1I6 level versus those from the 1G4 level, (d) emissions from the 1D2 level versus those from the 1G4 level, (e) emissions from the 1G4 level versus those from the 3H4 level, and (f) emissions from the 5D3 level versus those from the 5D4 level on power density of pump laser.
Fig. 9
Fig. 9 Log-Log plot of upconversion emission intensity versus excitation power density for the (a) 1I63F4 (345 nm), (b) 1D23F4 (452 nm), (c) 1G43H6 (476 nm), and (d) 3H43H6 (803 nm) transition of Tm3+, the (e) 5D37F6 (381 nm) and (f) 5D47F6 (544 nm) transition of Tb3+ in TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb samples.
Fig. 10
Fig. 10 Fluorescent lifetime of the emissions at (a) 345 nm, (b) 452 nm, (c) 476 nm, and (d) 803 nm, (e) 381 nm and (f) 544 nm versus excitation power density in TmYb, 10TbTmYb, 20TbTmYb, and 20TbYb samples.
Fig. 11
Fig. 11 Energy transfer efficiency between Tm3+ and Tb3+ calculated from fluorescent lifetime of emissions generated from various energy levels in (a) 10TbTmYb and (b) 20TbTmYb sample.

Equations (8)

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

  5 D 3 +   7 F 6   5 D 4 +   7 F 0 ( Tb 3 + Tm 3 + ) ,   5 D 4 +   3 H 6   4 F 0 +   3 F 2 ( Tb 3 + Tm 3 + ) ,   3 H 4 +   7 F 6   3 H 5 +   7 F 3 ( Tm 3 + Tb 3 + ) ,   5 F 2 +   7 F 6   3 H 4 +   7 F 5 ( Tm 3 + Tb 3 + ) .
  3 H 4 +   3 H 6   3 F 4 +   3 F 4 ( Tm 3 + Tm 3 + ) ,   5 D 3 +   7 F 6   5 D 4 +   7 F 0 ( Tb 3 + Tb 3 + ) .
τ a v = A i τ i 2 A i τ i ,
1 τ m = A R + W N R + W C R ,
η T b T m = 1 τ T m Y b T b τ T m Y b ,
η T m T b = 1 τ T b Y b T m τ T b Y b ,
Φ = τ m τ R ,
Φ = A R A m = A R A R + W N R + W C R .
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.