DOI:
10.1039/C6RA21768K
(Paper)
RSC Adv., 2016,
6, 92454-92462
A bright yellow light from a Yb3+,Er3+-co-doped Y2SiO5 upconversion luminescence material†
Received
31st August 2016
, Accepted 21st September 2016
First published on 21st September 2016
Abstract
A bright yellow light from Yb3+ and Er3+ doped Y2SiO5 upconversion materials in particle and fiber form were prepared by a co-precipitation, and electrospinning method. The morphologies of the prepared samples were investigated through FT-IR, FE-SEM, XRD and Raman measurements. The upconversion properties of the samples are carefully studied based on the absorption, fluorescent and decay time measurements. Under 980 nm excitation, the prepared material shows bright yellow emission. By controlling the concentration of Yb3+ or Er3+ and the excitation laser power, the ratio of red to green emission intensity can be varied to adjust the color tuning properties of the phosphor. The upconversion mechanism, and transition probabilities were elucidated owing to Judd–Ofelt theory. The radiative quantum efficiencies for the red (4F9/2) and green (4S3/2) bands were estimated to be 87% and 55%, respectively. The color coordinates of the system were evaluated as a function of the dopant concentrations and plotted on a standard CIE index diagram. The change of band intensity ratio with dopant concentrations gives a promising potential of the current phosphor for lighting application.
Introduction
Upconversion (UC) is an anti-Stokes process in which a longer wavelength radiation is converted to a shorter wavelength via a multiphoton mechanism. In recent years, upconversion materials have attracted considerable interest owing to their unique optical properties.1 The rare-earth (RE)-doped UC materials have been studied extensively, but most reports were focused on fluoride materials, such as NaYF4,2,3 NaGdF4,4,5 NaYbF4,6 BaYF5,7 and LaF3.8 However, the upconversion materials based on other hosts including Y2O3,9 Bi2Ti2O7,10 BiOBr,11 CaTiO3,12 Y2SiO5 (ref. 13) have been rarely reported.
It is well known that Y2SiO5 (YSO) is an ideal host material for both photoluminescent and cathodoluminescent phosphors due to the combined chemical–physical characteristics of the matrix, such as high chemical and thermal stability, and elevated mass absorption coefficient with high stopping power.14–17 For example, Ce3+-doped YSO was considered as a candidate to substitute ZnS.18 Tb3+-doped YSO exhibited a high efficient green phosphor excited by UV light-emitting diode.17 To date, a few studies on the upconversion of RE-doped YSO have been reported, where YSO was doped with Pr3+, Tb3+, Tm3+, Li+.13,19–21 There are studies reporting the UC effect of YSO for converting visible to ultraviolet,19 violet, cyan and yellow to UVC (under polychromatic excitation)13 and IR to blue/red, based on Pr3+-doped YSO.22 However, UC from IR to green in Yb3+,Er3+-co-doped Y2SiO5 (YSO:YbEr) has been rarely investigated.
It is worthwhile to investigate the efficiency of rare-earth ions in host material because the interactions between dopants and host materials play a key role in controlling the excitation dynamics. From a fundamental point of view, the physical understanding of the luminescence properties of rare-earth ions in host material is very important. Furthermore, it is difficult to make direct measurements of radiative rate constants for the emitting states of Er3+ because the emitting states include numerous non-radiative paths (multiphonon relaxation, Er–Er relaxation, and Yb–Er energy transfer, etc.).23 Judd–Ofelt (JO) theory describes the origin of electric-dipole intensity for the nominally parity forbidden f–f transitions. According to that a well-established means to estimate the radiative rates was provided.24,25 It has been extensively used to determine the basic spectroscopy properties of RE-doped materials. To quantify the three JO intensity parameters, Ωλ (λ = 2, 4, 6), an absorption spectrum is often used. However, the Judd–Ofelt theory have been often applied to transparent bulky amorphous or crystals. In order to apply Judd–Ofelt theory to powder samples, the powdered samples were often pressed to a very thin semi-transparent pellet at high pressure and the absorption spectra were obtained.26 Luo et al.27 used the excitation spectra and the proportional relationship between the excitation and absorption spectra for determining JO parameters.
In materials science, controlled synthesis of different morphological forms has been a focus of worldwide research work, since like flower, fiber, hollow porous sphere structure hold higher specific surface area, and better permeation.28–31 Here, we present the synthesis and UC spectroscopic studies of YSO:YbEr material in two different morphologies, namely powder (YSO:YbEr) and fiber form (FYSO:YbEr), prepared by different methods. The UC mechanism of the products has been proposed and discussed in detail. Judd–Ofelt theory has been first time adapted to estimate the transition probabilities, fluorescence branching ratios, and quantum efficiencies of the various emission bands quantitatively for elucidating the energy transfer UC process. The color coordinates of the system were evaluated and plotted on a standard CIE index diagram. To the best of our knowledge, this is the first report on the quantitative analysis of the fluorescent properties and UC process in YSO:YbEr material.
Results and discussion
Phase and morphology
The XRD patterns of the 1000 °C heat treated YSO and YSO:YbEr (18 mol% Yb3+, 2 mol% Er3+) (YSO:YbEr-1000) are shown in Fig. 1a. There is no difference in the peak positions in XRD patterns of the two samples. This confirms that Yb3+ and Er3+ ions didn't create other phases and entered in the crystal phase of YSO. All diffraction peaks could be indexed to the monoclinic phase of Y2SiO5 (JCPDS no. 52-1810); the crystal belongs to the P21/c space group. Rare earth oxyorthosilicates form two polymorphs, both of which are monoclinic. From La to Gd, they have the P21/c space group called the X1 phases.32 Whereas the elements from Dy to Lu are found in the C2/c space group called the X2 phases, Tb and Y exist in both the X1 and X2 phases.33 The phase transition from low temperature crystalline phase (X1) to high temperature crystalline phase (X2) occurred around 1200–1300 °C.32,33 In this study, the X1 phases always existed when the samples were annealed below 1300 °C (data not shown here). The X1–X2 phase transition appeared above 1300 °C agreed with previous reports. Both phases belong to the same monoclinic crystal system with different space group. For the samples heat treated at 1400 °C (YSO:YbEr-1400) there is an elongation along c axis. The crystal belongs to the I2/a space group, as shown in Fig. 1b (JCPDS no. 74-2011). The c value was changed from 6.64 Å to 12.49 Å resulted the increasing of cell volume from 397.38 Å3 to 852.66 Å3. Liu's32 group proposed that the change in the coordination of two heavy cation sites to oxygen, corresponding to a silicon-bonded oxygen from the isolated tetrahedral and a non-silicon-bound oxygen may be reason for different structures. In X1 phase, Y3+ ions occupied two sites, where they are surrounded by nine and seven oxygen ions, whereas only six oxygen ions are involved in X2 phase to result an elongation c axis.34 The appearance of X2-phase could be originated from the replacement of Y3+ ions with Yb3+ and/or Er3+ ions after the samples were calcined at high temperature. The ionic radii of Y3+, Yb3+, and Er3+ are 1.02, 0.99, and 1.00 Å, respectively. It should be noted that there were no noticeable differences in the XRD data of YSO and YSO:Yb,Er. No second phase is detected at this doping level, indicating that the Yb3+ and Er3+ ions can be efficiently incorporated into the YSO host lattice by substitution for the Y3+ ion or occupying interstitial sites of the crystal lattice. Additional phases or impurity peaks did not appear in both cases. The characteristic peaks were consistent with JCPDS no. 74-2011. The XRD pattern of the fiber is similar.
 |
| Fig. 1 XRD patterns of heat treated samples (a) at 1000 °C with the reference JCPDS no. 52-1810, and (b) at 1400 °C with the reference JCPDS no. 74-2011. | |
Fig. 2a and b show the SEM images of YSO:YbEr-1400 powder samples with low and high magnification, respectively. As can be seen in Fig. 2b, the obtained particles were uniform and like “boomerang form” with average size of 400 nm. It is well known that the morphology and diameter of the fibers can be influenced by many eletrospinning parameters. These parameters include the polymer concentration, the distance between the spinneret and collector, the value of high voltage, and the spinning rate. The results were shown that the diameter of FYSO:YbEr was clearly increased when the voltage was increased from 15 kV to 20 kV. At higher voltage (25 kV), the diameter of fibers change was not significant, as shown in Fig. S1a–c.† The length of the fibers also depends on the distance between the spinneret and collector. It was shortened with increasing the distance at voltage of 20 kV, as shown in Fig. S1d–f.† The shortening might be a result of the complete fiber drying due to evaporation of solvent before reaching the collector. Fig. 2c presents the SEM image of post calcined FYSO:YbEr fiber. As can be observed, the fibers inhibited relative uniform structure without any beads-on-string morphology. Fig. 2d also depicted the average diameter of 150 nm for calcined-FYSO:YbEr fiber. The results from EDS data confirmed the existence of Yb and Er in all prepared samples, as shown in inset of Fig. 2a.
 |
| Fig. 2 SEM images of YSO:YbEr with low magnification (a), with high magnification (b); FYSO:YbEr fiber samples before (c), and after calcination at 1400 °C (d). | |
FT-IR spectrum of the 1400 °C heat treated sample is shown in Fig. S2.† The peak at 1010 cm−1 originates from the absorption of Si–O–Si asymmetric stretching vibrations. Inorganic silicates have a characteristic, strong band cantered around 1100 cm−1 which in some cases appears as multiple bands.35 In our case, the band at 1010 cm−1 of the samples transformed into three well differentiated absorption peaks of SiO4 at 1010, 915, and 870 cm−1 in the sample suggesting the formation of well-crystallized silicate. The band at 590 cm−1 is due to bonding vibrations of the Y–O bonds.35 The obvious bands at 1020–1080, and 846, 930–950 cm−1 may arise from the asymmetric and symmetric stretching vibration of Si–O–Si, respectively.30
Raman spectra of YSO:YbEr samples were investigated, as shown in Fig. 3. The low wavenumber region (up to 300 cm−1) is related to the motion of the cations, the bands between 300 and 800 cm−1 relates to the vibration of RE-O, and higher energy vibrations are due to the motion of Si–O atoms.17,36 Ricci group17 reported that the vibrational frequencies in the 880–930 cm−1 range are a fingerprint that can be used to assign the effective crystal phase of the samples. Besides that, the two vibration modes at 891 and 918 cm−1 can be assigned to the symmetric stretching motion of the O atom with respect to the central cation (ν1 mode).36 The temperature-induced phase transformation from P21/c to C2/c to structure has been observed through a splitting of the stretching mode in that band, as shown in Fig. 3.36 There are slight shifts (∼3 cm−1) and change of relative intensities of these peaks that can be related to the polarized nature of the Raman bands in oxyorthosilicate in the YSO:YbEr samples. In addition, the quite high intensity of bands in range of 632–796 cm−1 in the spectra of both samples can relates to the absence of a cation-deficient type.17
 |
| Fig. 3 Raman spectra at 300 K of YSO:YbEr-1000 and YSO:YbEr-1400 samples. | |
The absorption spectrum of the YSO:YbEr sample was shown in Fig. 4 along with standard notations of the observed spectral transitions of both Er3+ and Yb3+. It can be seen that a strong band below 280 corresponds to the intrinsic absorption band edge of the YSO crystal.37
 |
| Fig. 4 Absorption spectrum of YSO:YbEr-1400 sample. | |
Upconversion study
Fig. 5 displays the upconversion emission spectra of YSO:YbEr sample under the excitation with a 980 nm laser. Spectra shows two strong bands: one band located at the green spectral range is due to 4f–4f radiative relaxation from the thermally coupled states 2H11/2, 4S3/2 to the ground state 4I15/2 of Er3+, and another intense band located at the red spectral range due to 4f–4f radiative relaxation from the excited state 4F9/2 to the ground state 4I15/2 of Er3+ ion.38 Results show that upconversion fluorescent spectra depend on calcination temperature of the samples, as shown in Fig. 5a. It can be seen that the upconversion intensity linearly increases with the annealing temperature. The increase in intensity is due to the increase in crystallinity due to the change in the coordination of two heavy cations sites to oxygen, corresponding to a silicon-bonded oxygen from the isolated tetrahedral and a non-silicon-bounded oxygen.32 These changes result in different alterations of the crystal-field splitting in the energy levels of the erbium due to the change of coordination number around the Yb3+, Er3+. Besides, the transition 4F9/2 → 4I15/2 is more sensitive to the annealing temperature than the transitions 2H11/2 → 4I15/2, 4S3/2 → 4I15/2. The yellow emission (4F9/2 → 4I15/2) arises from two main possible mechanisms in YSO. First one, the non-radiative relaxations originating from the 4S3/2 state of Er3+ result in the increasing population of the 4F9/2. Other possible mechanism is the energy transfer from Yb3+ (2F5/2) to Er3+ (4I11/2). The 4I13/2 level is populated owing to the non-radiative relaxation from the upper 4I11/2, excited-state absorption 4I13/2 + a photon → 4F9/2 and cross relaxation between Er3+ ions 4I13/2 + 4I11/2 → 4I15/2 + 4F9/2.39
 |
| Fig. 5 Upconversion spectra as a function of (a) heat treatment temperature, (b) Yb3+ concentration, and (c) Er3+ concentration. | |
As shown in Fig. 5b and c the concentration of Yb3+ and Er3+ has a strong influence on the relative intensities of the green (G) and red (R) UC bands. Upconversion processes in all Yb/Er doped samples are due the multiphoton processes happening from the upper excited sates which involve mainly the 4I13/2 and 4F11/2 states of Er. Under different concentrations these levels are populated in different ways and the emission intensity is proportional to the population of these excited levels. Since green and red are the major upconversion emission bands in all Yb/Er based upconversion phosphors the relative emission intensities of these emission band determine the G/R ratio. It can be seen that the R/G ratio increases with increasing Yb3+ concentration. The (R/G) ratio obtained for 5, 10, 18, 23 and 27 mol% of Yb3+ are 2.4, 4.5, 8.0, 5.8, and 5.8 respectively. In case, of fixed-Yb3+ concentration of 18 mol%, the R/G values are 5.1, 5.9, 8.0, and 3.7 corresponding to 0.5, 1, 2, and 3 mol% of Er3+ (Fig. 5c). Both these results indicate that a strong energy transfer is occurring between Yb3+ and Er3+ ions. The UC intensity increased with increasing Yb3+ concentration and got the maximum value at 18 mol% and decreased with further increase in Yb3+ concentration. The decrease of UC intensity can be related to the excess Yb3+ ions in the lattice (concentration quenching effect). Too much Yb3+ ion will end up in forming Yb–Yb pairs or more complex clusters competing with the Yb–Er interactions.
The UC spectra of FYSO:YbEr fiber and YSO:YbEr powder are the same except that the intensity of fiber was lower than that of the powder form. The reason relates to the surface quenching effect40 because the fiber consists of the small particles (some ten nanometers) whereas the size of YSO:YbEr powder was big (some hundreds nanometer).
In order to verify the multi-photon excitation process, the dependence of the UC intensity of YSO:YbEr samples on pump power were examined. The relationship between UC intensity IUC and pump power Ip can be described as:
where,
m- is the number of photons participating in the UC process. The plot of log
IUC vs. log
Ip yields a straight line with slope
n. The relationship between pump power of laser and UC intensity is shown in
Fig. 6. The results shows that the UC intensity of 1400 °C heat treated YSO:YbEr samples are more sensitive to laser pump power than that of 1000 °C heat treated samples. The slope
n = 2 indicate that the two photon process is primary mechanism of up-conversion luminescence in YSO:YbEr material.
 |
| Fig. 6 log–log plots of UC fluorescence emission as a function of the 980 nm pump power. | |
Judd–Ofelt analyses
Many researchers have applied the Judd–Ofelt theoretical analyses to determine important spectroscopic parameters. The absorption spectrum of YSO:YbEr is shown in Fig. 4. Most of the bands can be attributed to the 4f–4f transition of Er3+ from the 4I15/2 ground state to different excited states and only one band corresponding to the transition from the 2F7/2 level to 2F5/2 of Yb3+. Six bands can be assigned to the ground state of Er3+ (4I15/2) to the excited state 4I9/2, 4F9/2, 4S3/2, 2H11/2, 4F7/2 and 4F5/2 transitions.
The oscillator strength, fexp, of these absorption bands were determined experimentally using the following formula:41,42
|
 | (1) |
where
ε is the molar extinction coefficient [L mol
−1 cm
−1] at energy
ν [cm
−1]. Judd–Ofelt theory provides the calculated oscillator strength
fcal of a transition from the ground state to an excited state. By Judd–Ofelt model, the calculated oscillator strength for an electric dipole transition from the ground state to an excited state is given by
24,25 |
 | (2) |
where
J is total angular momentum quantum number of the ground state,
ν is the energy of the transition,
Ωλ is Judd–Ofelt intensity parameter and ‖
Uλ‖
2 is the squared double reduced matrix element of the unit tensor operator of rank
λ = 2, 4, 6 calculated from the intermediate coupling approximation.
By equating the measured and calculated values of the oscillator strength (fexp = fcal) and solving the system of equations by the method of least squares, the Judd–Ofelt intensity parameters Ωλ have been calculated numerically. The resulting Ωλ (×10−20 cm2) parameters for this YSO sample are found to be Ω2 = 1.29, Ω4 = 0.29, and Ω6 = 2.78 × 10−20 cm2. These results are represented in Table 1 together with the Judd–Oflet parameters of other samples for comparison. Jørgensen and Reisfeld43 noted that the Ω2 parameter is indicative of the amount of asymmetry and covalent bonding while the Ω6 parameter is related to the rigidity of the host. Generally, it is found that the magnitude of Ω2 is much greater than Ω4 and Ω6 in most of the Er3+ doped glasses due to the high asymmetry of the amorphous state. A comparative study of the intensity parameters for complexes containing Er3+ allowed Jorgensen and Reisfeld to conclude that the Ω6 parameter is related to the rigidity of the medium in which the Er3+ ions are embedded and the Ω6 parameter increases with increasing vibrational amplitude of the RE-X distance (X = O, F…), which in turn is related to the vibrational phonon energy. Phonon energy of the host increases in the order fluorite < aluminate < silicate.
Table 1 Judd–Ofelt parameters (Ωλ) in comparison with other hosts. All parameters are in units of 10−20 cm2
Host |
Ω2 |
Ω4 |
Ω6 |
This work. |
Y2SiO5a |
1.29 |
0.29 |
2.78 |
Germo-tellurite44 |
4.99 |
1.47 |
0.76 |
Zinc tellurite50 |
0.85 |
0.01 |
0.02 |
Ag-doped tellurite41 |
6.63 |
2.47 |
2.92 |
NaYF4 (ref. 51) |
4.65 |
0.82 |
1.11 |
Fluorindate52 |
2.91 |
1.27 |
1.11 |
Lu2SiO5 (ref. 53) |
4.57 |
3.11 |
1.42 |
It was observed that Ω2 is related with the symmetry of host, while Ω6 decreases with the increasing of the covalence of the Er–O bond.44 The value of Ω2 is smaller in magnitude as compared to some other hosts, as shown in Table 1. The smaller values of Ω2 in this system indicate that the Er3+ ions occupy ionic sites with small distortion. In comparison with the Lu2SiO5 glass sample, the value of Ω2 is smaller than that of Lu2SiO5 because the distortion of the glass system is higher. The value of Ω6 is larger than those of germo-tellurite, fluoride. This suggests that the covalency of Er–O bond in YSO:YbEr is lower than that of germo-tellurite, fluoride.
The phenomenological parameters Ωλ were used in the calculation of radiative decay rate A(J → J′) according to the formula
|
 | (3) |
The reciprocal of the sum of the radiative decay rate yields the radiative decay times of the excited states. The calculated values of the emission spectroscopic parameters of Er3+ doped YSO are presented in Table S1.† The 4S3/2 → 4I15/2 and 4F9/2 → 4I15/2 transitions are especially interesting in this study. The radiative transition rate AR (s−1), branching ratio β (%) and the radiative lifetime of the excited level τ (ms) of these transitions are compared with those of the same material reported by Li et al.45
Fluorescence decay curves of the red and green emission bands in YSO:YbEr-1400 sample are shown in Fig. 7. In order to estimate the accuracy of the JO analysis in this case, the calculated life time is compared with the measured life time obtained from the decay curves. The calculated radiative decay times of the green and red bands are respectively 0.187 ms and 0.405 ms. The experimental values of the life times of these transitions are 0.164 ms and 0.226 ms, respectively. From the calculated and measured life times of the 4F9/2 and 4S3/2 states, the radiative quantum efficiency for the red (4F9/2) and green (4S3/2) band are 87% and 55%, respectively.46,47
 |
| Fig. 7 Decay curves of green (a) and red (b) emissions in YSO:YbEr-1400 sample. | |
It should be noted that the emission bands around 525 and 545 nm correspond to the 2H11/2 → 4I15/2 and 4S3/2 → 4I15/2 transitions, respectively, which are originated from levels 4H11/2 and 4S3/2. At room temperature, thermalization of these two levels occurs and therefore, a simple three-levels system comprised of 4I15/2 (level 0), 4S3/2 (level 1) and 2H11/2 (level 2) could be used to describe the thermalization of the 2H11/2 state by the following relation:48
|
 | (4) |
where
A(
4S
3/2) = 5360 s
−1,
A(
2H
11/2) = 3940 s
−1 are the calculated total spontaneous emission rate,
hν1 = 18
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
351 cm
−1 and
hν2 = 18
![[thin space (1/6-em)]](https://www.rsc.org/images/entities/char_2009.gif)
692 cm
−1 are the highest experimental Stark energy level of the luminescence band
4S
3/2 →
4I
11/2 and the lowest experimental Stark energy level of the luminescence band
2H
11/2 →
4I
15/2 respectively.
g1 and
g2 are degeneracies (2
J + 1) = 4 and 12 of
4S
3/2 and
2H
11/2 levels, respectively. The
I(
4S
3/2) and
I(
2H
11/2) represent the integrated measured emission intensities of the observed transitions at room temperature,
k is Boltzmann constant, and Δ
E is the energy gap between the
2H
11/2 and
4S
3/2 levels.
Using the experimental values of emission spectra and the predicted results of the Judd–Ofelt analysis and using eqn (4), the energy gap ΔE is predicted to be 465 cm−1. This predicted value is in good agreement with the energy separation of 341 cm−1 between two nearest Stark levels of the 4S3/2 → 4I15/2 and 2H11/2 → 4I15/2 bands. The difference of 120 cm−1 could be explained by the inherent error of about 15–20% in the Judd–Ofelt calculations.49 From these remarks, we conclude that the Judd–Ofelt analysis performed in this study could be used to explain satisfactorily the obtained experimental results.
Since Yb3+ has a much larger absorption cross-section than Er3+, the GSA of Er3+ from 4I15/2 → 4I11/2 is not the main reason for the population of the 4I11/2 level of Er3+. The energy transfer process contributes to the population accumulation on the 4I11/2. This level serves as the intermediate state for the upconversion process. The life time of the 4I11/2 level was high enough to confirm its role. The 4S3/2 state lie close to 2H11/2 (ΔE = 465 cm−1) and resulted in the existence of thermal population process. In this process, the electrons can migrate between 4S3/2 and 2H11/2, so that both level states have a common transition probability.54,55 This transition probability was expressed by
|
 | (5) |
with Δ
E = 465 cm
−1,
AR(
2H
11/2) = 3940 s
−1, and
AR(
4S
3/2) = 5360 s
−1 Boltzmann constant
kB = 0.695 cm
−1 K
−1,
T = 300 K. Using above results, the calculated common transition probability
At is 5025 s
−1 and the life time of
4S
3/2 state was corrected to be 0.199 ms.
Evaluation of color coordinate
The luminous color was estimated by studying color co-ordinates and color ratios of the 1400 °C heat treated samples under 980 nm excitation. From the emission spectra of the samples, the value of chromaticity co-ordinate of the powder was estimated from CIE-1931 system (as shown in Table S2†). The values were found to locate in the yellow or near white region, as shown in Fig. 8. The values shifted from the bright yellow region to the yellow region with increasing Er3+ ions (fixed Yb3+ ions), as shown in Fig. 8a. The values slightly changed with a change in Yb3+ ions (at 2 mol% Er3+ ions) and located on near white region, as shown in Fig. 8b. Overall the effect of Er3+ concentration is found to be more prominent that Yb3+ concentrations. This is clear from the emission data of Fig. 3b and c where emission intensity shows sudden decrease after 2 mol% of Er3+. The pump power influenced the value of the chromaticity co-ordinate, as show in Fig. 8c. The change of R/G ratios by pump power is a reason for the change of yellow color intensity. Therefore, by adjusting to a suitable Er3+ concentration, the chromaticity coordinates of the sample can be adjusted to near white light zone. This clearly shows that the prepared YSO:YbEr-1400 samples will be more suitable candidates for light emitting applications.
 |
| Fig. 8 CIE chromaticity coordinates diagrams for YSO:Yb,Er as a function of (a) Er3+ concentration, (b) Yb3+ concentration, and (c) pump power. | |
Experiment
Yttrium oxide (Y2O3), ytterbium oxide (Yb2O3), erbium oxide (Er2O3), tetraethylorthosilicate (TEOS), and anhydrous ethanol were obtained from Aldrich Chemical Co. Ammonia solution, nitric acid (HNO3), were obtained from Alfa Co. All the reagents were Analar grade and used without further purification.
Synthesis
Synthesis of powder sample. Yb3+,Er3+-co-doped Y2SiO5 powder samples were synthesized according to the following procedure. First, stoichiometric amount of Y2O3, Yb2O3 and Er2O3 (1 mmol) were dissolved in 15 mL of HNO3 0.1 N at 70 °C and the solution was allowed to cool to room temperature. Second, 1 mmol TEOS in 12 mL ethanol were added to the cooled rare-earth solution with mechanical stirring for 1 h resulting in the formation of a transparent solution. Droplet of ammonia solution was subsequently added to form a white precipitate immediately. This process was continued until no further precipitation occurred. The mixture was centrifuged at 4000 rpm for 15 min and washed 5 times thoroughly with water, followed by subsequent wash with 2-isopropanol in order to remove excess reagents. The product was dried in a vacuum oven at 60 °C for 12 h. The powder samples were annealed at different temperatures in range 1000–1400 °C with heating rate of 5 °C min−1 for 3 h.
Synthesis of fiber sample. 1.1 g of Y2O3, Yb2O3 and Er2O3 (Y
:
Yb
:
Er = 80
:
18
:
2 mol%) were dissolved in 10 mL of solution H2O
:
HNO3 (1
:
1 v/v), then evaporated to powder form. The RE (NO3)3 powder was dissolved in 3.5 g of DMF. TEOS (360 mg) was added to the mixture under stirring for 6 h, and then 2.1 g of PVP (MW: 360
000) was added under vigorous stirring for 6 h to form electrospun sol. In the electrospinning process, the mixture was injected through a stainless steel needle (26G in orifice diameter of 450 nm) that was connected to a high voltage DC supply. The parameters for preparing FYSO:Yb,Er, namely the distance between the spinneret and collector, the value of high voltage, and the spinning rate, were set at 15 cm, 20 kV, 0.7 mL h−1, respectively. The fibers were heated up to 1400 °C with heating rate 2 °C min−1 and kept at this temperature for 3 h in order to remove the polymer and maintain the fiber form.
Characterization
X-ray powder diffraction data was recorded by using an X'pert PRO MPD X-ray diffractometer (Panalytical, Netherlands). The morphology of the samples was investigated by using field emission scanning electron microscopy (FE-SEM), MIRA3 (Tescan, Czech). FT-IR spectra were observed by using a Nicolet iS10 FT-IR spectrometer (Thermo Fisher Scientific, Madison, Wisconsin, USA). The absorption spectra were recorded using a UV-VIS/NIR spectrometer (V-670, Jasco, Japan). The prepared samples were compressed into near transparent pellets of 0.4 mm thickness using a hydraulic press for the absorption measurements.
The UC luminescence emission spectra were measured with a fluorescence spectrophotometer (Acton SpectraPro 750-Triplet Grating Monochromator) from a CCD detector (Princeton EEV 1024 × 1024 and PI-Max 133 Controller), using a 980 nm semiconductor CW laser diode as an excitation source, which was placed at an angle of 45° in front of the sample holder. All of the measurements were performed at room temperature. Photoluminescence decay curves were measured on a Quanta Master 40 system (Photon Technology International Inc., US) using a pulsed 980 nm laser as the excitation source. The collected decay curve was analyzed using built in software provided by PTI.
Raman scattering measurements were performed using a Raman spectrophotometer (NRS-3300 Jasco, Japan) at room temperature in a single monochromator. For excitation, the 532 nm line of a solid-state laser system was focused on the sample by mean of a 100× objective, while the laser power was kept below 2 mW in order to avoid laser-heating effects. The accumulation time was typically two collections of 2 s.
Radiative properties from Judd–Ofelt theory
To quantitatively estimate the radiative properties of Er3+, Judd–Ofelt theory was adapted based on the oscillator strengths of the absorption band (fexp) from absorption spectra. The experimental oscillator strengths were fitted with theoretical oscillator strengths (fcal) following the Judd–Ofelt procedure.41,56 By fitting the measured oscillator strengths to the theoretical oscillator strengths using a least-squares method, the three Judd–Ofelt parameters (Ω2, Ω4, and Ω6) were obtained. The radiative transition probability (Arad), radiative lifetime (τrad), and branch ratios of a transition were estimated from these Ω2, Ω4, and Ω6 parameters.
Conclusions
In summary, a bright yellow emitting Yb3+,Er3+-co-doped Y2SiO5 up-conversion material was prepared through a co-precipitation, and electrospinning procedure. The X1–X2 phases were determined through XRD and Raman measurements. The upconversion emission originates from two photon process. Results from the Judd–Ofelt theory are in good agreement with the experimental results. The color coordinates of the system were evaluated and plotted on a standard CIE index diagram. By varying the concentrations of Yb3+ or Er3+ and pump power, the red to green band ratio of YSO:YbEr can be adjusted so as to obtain a color tunable phosphor for various lighting applications. Using co-precipitation method, large amount of YSO:YbEr powder sample can be obtained. The electrospinning technique has been used to make YSO:YbEr hollow fiber that will be studied for some applications, such as drug delivery, cell imaging in future.
Acknowledgements
This work was supported by the NRF-2015K2A1A2070664 project through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology. Authors GAK would like to acknowledge the financial support from the National Science Foundation Partnerships for Research and Education in Materials (NSF-PREM) grant No. DMR-0934218.
Notes and references
- G. Chen, H. Qiu, P. N. Prasad and X. Chen, Upconversion Nanoparticles: Design, Nanochemistry, and Applications in Theranostics, Chem. Rev., 2014, 114(10), 5161–5214 CrossRef CAS PubMed.
- X. Huang, Giant enhancement of upconversion emission in (NaYF4:Nd3+/Yb3+/Ho3+)/(NaYF4:Nd3+/Yb3+) core/shell nanoparticles excited at 808 nm, Opt. Lett., 2015, 40(15), 3599–3602 CrossRef PubMed.
- X. Huang and J. Lin, Active-core/active-shell nanostructured design: an effective strategy to enhance Nd3+/Yb3+ cascade sensitized upconversion luminescence in lanthanide-doped nanoparticles, J. Mater. Chem. C, 2015, 3(29), 7652–7657 RSC.
- X. Huang, Dual-model upconversion luminescence from NaGdF4:Nd/Yb/Tm@NaGdF4:Eu/Tb core–shell nanoparticles, J. Alloys Compd., 2015, 628, 240–244 CrossRef CAS.
- S. B. Wang, Q. Li, L. Z. Pei and Q.-F. Zhang, Branch-shaped NaGdF4:Eu3+ nanocrystals: Selective synthesis, and photoluminescence properties, Mater. Charact., 2010, 61(8), 824–830 CrossRef CAS.
- J. A. Damasco, G. Chen, W. Shao, H. Ågren, H. Huang, W. Song, J. F. Lovell and P. N. Prasad, Size-Tunable and Monodisperse Tm3+/Gd3+-Doped Hexagonal NaYbF4 Nanoparticles with Engineered Efficient Near Infrared-to-Near Infrared Upconversion for In Vivo Imaging, ACS Appl. Mater. Interfaces, 2014, 6(16), 13884–13893 CAS.
- T. Grzyb, S. Balabhadra, D. Przybylska and M. Węcławiak, Upconversion luminescence in BaYF5, BaGdF5 and BaLuF5 nanocrystals doped with Yb3+/Ho3+, Yb3+/Er3+ or Yb3+/Tm3+ ions, J. Alloys Compd., 2015, 649, 606–616 CrossRef CAS.
- X. Huang, Enhancement of near-infrared to near-infrared upconversion luminescence in sub-10-nm ultra-small LaF3:Yb3+/Tm3+; nanoparticles through lanthanide doping, Opt. Lett., 2015, 40(22), 5231–5234 CrossRef PubMed.
- A. N. Meza-Rocha, E. F. Huerta, U. Caldiño, S. Carmona-Téllez, M. Bettinelli, A. Speghini, S. Pelli, G. C. Righini and C. Falcony, Dependence of the up-conversion emission of Li+ co-doped Y2O3:Er3+ films with dopant concentration, J. Lumin., 2015, 167, 352–359 CrossRef CAS.
- Y. Cun, Z. Yang, J. Li, B. Shao, J. Yang, Y. Wang, J. Qiu and Z. Song, Enhanced upconversion emission of three dimensionally ordered macroporous films Bi2Ti2O7:Er3+, Yb3+ with silica shell, Ceram. Int., 2015, 41(9), 11770–11775 CrossRef CAS.
- Y. Li, Z. Song, Z. Yin, Q. Kuang, R. Wan, Y. Zhou, Q. Liu, J. Qiu and Z. Yang, Investigation on the upconversion emission in 2D BiOBr:Yb3+/Ho3+ nanosheets, Spectrochim. Acta, Part A, 2015, 150, 135–141 CrossRef CAS PubMed.
- S. P. Tiwari, M. K. Mahata, K. Kumar and V. K. Rai, Enhanced temperature sensing response of upconversion luminescence in ZnO–CaTiO3:Er3+/Yb3+ nano-composite phosphor, Spectrochim. Acta, Part A, 2015, 150, 623–630 CrossRef CAS PubMed.
- E. L. Cates and J.-H. Kim, Upconversion under polychromatic excitation: Y2SiO5:Pr3+,Li+ converts violet, cyan, green, and yellow light into UVC, Opt. Mater., 2013, 35(12), 2347–2351 CrossRef CAS.
- T. Hirai and Y. Kondo, Preparation of Y2SiO5:Ln3+ (Ln = Eu, Tb, Sm) and Gd9.33(SiO4)6O2:Ln3+ (Ln = Eu, Tb) Phosphor Fine Particles Using an Emulsion Liquid Membrane System, J. Phys. Chem. C, 2006, 111(1), 168–174 CrossRef.
- F. Zhang, Photon Upconversion Nanomaterials, Springer-Verlag, Berlin Heidelberg, 2015 Search PubMed.
- Z. Sun, M. Li and Y. Zhou, Thermal properties of single-phase Y2SiO5, J. Eur. Ceram. Soc., 2009, 29(4), 551–557 CrossRef CAS.
- P. C. Ricci, C. M. Carbonaro, R. Corpino, C. Cannas and M. Salis, Optical and Structural Characterization of Terbium-Doped Y2SiO5 Phosphor Particles, J. Phys. Chem. C, 2011, 115(33), 16630–16636 CAS.
- P. J. Marsh, J. Silver, A. Vecht and A. Newport, Cathodoluminescence studies of yttrium silicate:cerium phosphors synthesised by a sol–gel process, J. Lumin., 2002, 97(3–4), 229–236 CrossRef CAS.
- C. Hu, C. Sun, J. Li, Z. Li, H. Zhang and Z. Jiang, Visible-to-ultraviolet upconversion in Pr3+:Y2SiO5 crystals, Chem. Phys., 2006, 325(2–3), 563–566 CrossRef CAS.
- N. Rakov, S. A. Vieira, R. B. Guimarães and G. S. Maciel, Cooperative upconversion luminescence in Tb3+:Yb3+ co-doped Y2SiO5 powders prepared by combustion synthesis, J. Solid State Chem., 2014, 211, 32–36 CrossRef CAS.
- N. Rakov and G. S. Maciel, Three-photon upconversion and optical thermometry characterization of Er3+:Yb3+ co-doped yttrium silicate powders, Sens. Actuators, B, 2012, 164(1), 96–100 CrossRef CAS.
- J. Li, C. Sun, X. Wang, Y. Yu, H. Yang, C. Hu and Z. Jiang, IR-to-blue and red wavelength upconversion in Pr3+:Y2SiO5 crystals, Phys. B, 2007, 398(1), 144–147 CrossRef CAS.
- G. Yao, C. Lin, Q. Meng, P. Stanley May and M. T. Berry, Calculation of Judd–Ofelt parameters for Er3+ in β-NaYF4:Yb3+,Er3+ from emission intensity ratios and diffuse reflectance spectra, J. Lumin., 2015, 160, 276–281 CrossRef CAS.
- B. R. Judd, Optical Absorption Intensities of Rare-Earth Ions, Phys. Rev., 1962, 127(3), 750–761 CrossRef CAS.
- G. S. Ofelt, Intensities of Crystal Spectra of Rare-Earth Ions, J. Chem. Phys., 1962, 37(3), 511–520 CrossRef CAS.
- M. C. Tan, G. A. Kumar, R. E. Riman, M. G. Brik, E. Brown and U. Hommerich, Synthesis and optical properties of infrared-emitting YF3:Nd nanoparticles, J. Appl. Phys., 2009, 106(6), 063118 CrossRef.
- W. Luo, J. Liao, R. Li and X. Chen, Determination of Judd–Ofelt intensity parameters from the excitation spectra for rare-earth doped luminescent materials, Phys. Chem. Chem. Phys., 2010, 12(13), 3276–3282 RSC.
- Y. Han, S. Gai, P. A. Ma, L. Wang, M. Zhang, S. Huang and P. Yang, Highly Uniform α-NaYF4:Yb/Er Hollow Microspheres and Their Application as Drug Carrier, Inorg. Chem., 2013, 52(16), 9184–9191 CrossRef CAS PubMed.
- M. Wang, Y. Shi, Y. Tang and G. Jiang, Hierarchical nanowires-assembled YVO4 microspheres: Synthesis, characterization and photocatalytic properties, Mater. Lett., 2014, 132, 236–239 CrossRef CAS.
- L. Wang, Z. Hou, Z. Quan, C. Li, J. Yang, H. Lian, P. Yang and J. Lin, One-Dimensional Ce3+- and/or Tb3+-Doped X1-Y2SiO5 Nanofibers and Microbelts: Electrospinning Preparation and Luminescent Properties, Inorg. Chem., 2009, 48(14), 6731–6739 CrossRef CAS PubMed.
- H. Yu, Y. Jia, C. Yao and Y. Lu, PCL/PEG core/sheath fibers with controlled drug release rate fabricated on the basis of a novel combined technique, Int. J. Pharm., 2014, 469(1), 17–22 CrossRef CAS PubMed.
- Y. Liu, C. N. Xu, H. Chen and H. Tateyama, Influence of calcining temperature on photoluminescence and thermal quenching in europium-doped Y2SiO5 using the MOD process, J. Lumin., 2002, 97(2), 135–140 CrossRef CAS.
- M. D. Dramićanin, B. Viana, Ž. Andrić, V. Djoković and A. S. Luyt, Synthesis of Y2SiO5:Eu3+ nanoparticles from a hydrothermally prepared silica sol, J. Alloys Compd., 2008, 464(1–2), 357–360 CrossRef.
- M. Yin, C. Duan, W. Zhang, L. Lou, S. Xia and J.-C. Krupa, Site selectively excited luminescence and energy transfer of X1-Y2SiO5:Eu at nanometric scale, J. Appl. Phys., 1999, 86(7), 3751–3757 CrossRef CAS.
- R. Krsmanović, Ž. Andrić, M. Marinović-Cincović, I. Zeković and M. D. Dramićanin, Optical and Thermal Investigation of Sol–Gel Derived Eu3+:Y2SiO5 Nanoparticles, Acta Phys. Pol., A, 2007, 112(5), 975–980 CrossRef.
- P. C. Ricci, D. Chiriu, C. M. Carbonaro, S. Desgreniers, E. Fortin and A. Anedda, Pressure effects in lutetium yttrium oxyorthosilicate single crystals, J. Raman Spectrosc., 2008, 39(9), 1268–1275 CrossRef CAS.
- X. Wang, J. Qiu, J. Song, J. Xu, Y. Liao, H. Sun, Y. Cheng and Z. Xu, Simultaneous three-photon absorption induced ultraviolet upconversion in Pr3+:Y2SiO5 crystal by femtosecond laser irradiation, Opt. Commun., 2008, 281(2), 299–302 CrossRef CAS.
- J. D. Suter, N. J. Pekas, M. T. Berry and P. S. May, Real-Time-Monitoring of the Synthesis of β-NaYF4:17% Yb,3% Er Nanocrystals Using NIR-to-Visible Upconversion Luminescence, J. Phys. Chem. C, 2014, 118(24), 13238–13247 CAS.
- Y. Wu, X. Shen, S. Dai, Y. Xu, F. Chen, C. Lin, T. Xu and Q. Nie, Silver Nanoparticles Enhanced Upconversion Luminescence in Er3+/Yb3+ Codoped Bismuth–Germanate Glasses, J. Phys. Chem. C, 2011, 115(50), 25040–25045 CAS.
- F. Wang, J. Wang and X. Liu, Direct Evidence of a Surface Quenching Effect on Size-Dependent Luminescence of Upconversion Nanoparticles, Angew. Chem., Int. Ed., 2010, 49(41), 7456–7460 CrossRef CAS PubMed.
- R. J. Amjad, M. R. Dousti and M. R. Sahar, Spectroscopic investigation and Judd–Ofelt analysis of silver nanoparticles embedded Er3+-doped tellurite glass, Curr. Appl. Phys., 2015, 15(1), 1–7 CrossRef.
- B. T. Huy, M.-H. Seo, J.-M. Lim, Y.-I. Lee, N. T. Thanh, V. X. Quang, T. T. Hoai and N. A. Hong, Application of the Judd–Ofelt Theory to Dy3+-Doped Fluoroborate/Sulphate Glasses, J. Korean Phys. Soc., 2011, 59(5), 3300–3307 CAS.
- C. K. Jørgensen and R. Reisfeld, Judd–Ofelt parameters and chemical bonding, J. Less-Common Met., 1983, 93(1), 107–112 CrossRef.
- Y. Gao, Q.-H. Nie, T.-F. Xu and X. Shen, Thermal stability, Judd–Ofelt theory analysis and spectroscopic properties of a new Er3+/Yb3+-codoped germano-tellurite glass, Spectrochim. Acta, Part A, 2005, 61(13–14), 2822–2826 CrossRef PubMed.
- C. Li, C. Wyon and R. Moncorge, Spectroscopic properties and fluorescence dynamics of Er3+ and Yb3+ in Y2SiO5, IEEE J. Quantum Electron., 1992, 28(4), 1209–1221 CrossRef CAS.
- H. Desirena, E. De la Rosa, V. H. Romero, J. F. Castillo, L. A. Díaz-Torres and J. R. Oliva, Comparative study of the spectroscopic properties of Yb3+/Er3+ codoped tellurite glasses modified with R2O (R = Li, Na and K), J. Lumin., 2012, 132(2), 391–397 CrossRef CAS.
- M. Pokhrel, G. A. Kumar, S. Balaji, R. Debnath and D. K. Sardar, Optical characterization of Er3+and Yb3+ co-doped barium fluorotellurite glass, J. Lumin., 2012, 132(8), 1910–1916 CrossRef CAS.
- D. C. Yeh, W. A. Sibley, M. Suscavage and M. G. Drexhage, Multiphonon relaxation and infrared to visible conversion of Er3+ and Yb3+ ions in barium thorium fluoride glass, J. Appl. Phys., 1987, 62(1), 266–275 CrossRef CAS.
- C. G. Walrand and K. Binnemans, Handbook on Physics and Chemistry of Rare Earths, ed. K. A. Gschneidner and J. L. Eyring, Elsevier, 1998 Search PubMed.
- A. Awang, S. K. Ghoshal, M. R. Sahar, M. Reza Dousti, R. J. Amjad and F. Nawaz, Enhanced spectroscopic properties and Judd–Ofelt parameters of Er-doped tellurite glass: Effect of gold nanoparticles, Curr. Appl. Phys., 2013, 13(8), 1813–1818 CrossRef.
- H. Ping, D. Chen, Y. Yu and Y. Wang, Judd–Ofelt analyses and luminescence of Er3+/Yb3+ co-doped transparent glass ceramics containing NaYF4 nanocrystals, J. Alloys Compd., 2010, 490(1–2), 74–77 CrossRef CAS.
- A. Florez, Y. Messaddeq, O. L. Malta and M. A. Aegerter, Optical transition probabilities and compositional dependence of Judd–Ofelt parameters of Er3+ ions in fluoroindate glass, J. Alloys Compd., 1995, 227(2), 135–140 CrossRef CAS.
- Q. Dong, G. Zhao, J. Chen, Y. Ding, B. Yao and Z. Yu, Spectroscopic properties of Ho3+ ions in biaxial Lu2SiO5 crystal, Opt. Mater., 2010, 32(9), 873–877 CrossRef CAS.
- B. Klimesz, G. Dominiak-Dzik, M. Żelechower and W. Ryba-Romanowski, Optical study of GeO2–PbO–PbF2 oxyfluoride glass single doped with lanthanide ions, Opt. Mater., 2008, 30(10), 1587–1594 CrossRef CAS.
- L. Riseberg and H. Moos, Multiphonon Orbit-Lattice Relaxation of Excited States of Rare-Earth Ions in Crystals, Phys. Rev., 1968, 174(2), 429–438 CrossRef CAS.
- N. T. Thanh, V. X. Quang, V. P. Tuyen, N. V. Tam, T. Hayakawa and B. T. Huy, Role of charge transfer state and host matrix in Eu3+-doped alkali and earth alkali fluoro-aluminoborate glasses, Opt. Mater., 2012, 34(8), 1477–1481 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available: SEM images of different parameters prepared fiber, FT-IR spectra, table: calculated radiative properties; CIE calculate. See DOI: 10.1039/c6ra21768k |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.