Next Article in Journal
Understanding Graphene Response to Neutral and Charged Lead Species: Theory and Experiment
Next Article in Special Issue
Effects of Pyrazine Derivatives and Substituted Positions on the Photoelectric Properties and Electromemory Performance of D–A–D Series Compounds
Previous Article in Journal
Structural, Elastic, Electronic and Optical Properties of SrTMO3 (TM = Rh, Zr) Compounds: Insights from FP-LAPW Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Temperature-Dependent and Threshold Behavior of Sm3+ Ions on Fluorescence Properties of Lithium Niobate Single Crystals

1
School of Physics, Sun Yat-sen University, Guangzhou 510275, China
2
Sino French Institute of Nuclear Engineering and Technology, Sun Yat-sen University, Zhuhai 519082, China
3
School of Physics Science & Engineering, Institute of Optoelectronic and Functional Composite Materials, State Key Lab Optoelectronic Materials and Technologies, Sun Yat Sen University, Guangzhou 510275, China
*
Authors to whom correspondence should be addressed.
Materials 2018, 11(10), 2058; https://doi.org/10.3390/ma11102058
Submission received: 11 September 2018 / Revised: 5 October 2018 / Accepted: 15 October 2018 / Published: 22 October 2018
(This article belongs to the Special Issue Advanced Semiconductors for Photonics and Electronics)

Abstract

:
Temperature-dependent and threshold behavior of Sm3+ ions on fluorescence properties of lithium niobate (LiNbO3, LN) single crystals were systematically investigated. The test materials, congruent LiNbO3 single crystals (Sm:LN), with various concentrations of doped Sm3+ ions from 0.2 to 2.0 mol.%, were grown using the Czochralski technique. Absorption spectra were obtained at room temperature, and photoluminescence spectra were measured at various temperatures in the range from 73 K to 423 K. Judd–Ofelt theory was applied to calculate the intensity parameters Ωt (t = 2, 4, 6) for 1.0 mol.% Sm3+-doped LiNbO3, as well as the radiative transition rate, Ar, branching ratio, β, and radiative lifetime, τr, of the fluorescent 4G5/2 level. Under 409 nm laser excitation, the photoluminescence spectra of the visible fluorescence of Sm3+ mainly contains 568, 610, and 651 nm emission spectra, corresponding to the energy level transitions of 4G5/26H5/2, 4G5/26H7/2, and 4G5/26H9/2, respectively. The concentration of Sm3+ ions has great impact on the fluorescence intensity. The luminescence intensity of Sm (1.0 mol.%):LN is about ten times as against Sm (0.2 mol.%):LN at 610 nm. The intensity of the fluorescence spectra were found to be highly depend on temperature, as well as the concentration of Sm3+ ions in LiNbO3 single crystals, as predicted; however, the lifetime changed little with the temperature, indicating that the temperature has little effect on it, in Sm:LN single crystals. Sm:LN single crystals, with orange-red emission spectra, can be used as the active material in new light sources, fluorescent display devices, UV-sensors, and visible lasers.

1. Introduction

Lithium niobate (LiNbO3, LN), as a full-feature crystal, possesses many excellent properties, such as electro-optical, nonlinear optical, and acousto-optical properties, that enable it to be widely used in electro-optic (EO) signal modulation, lasers, biological imaging, temperature sensors, and many other examples. In recent years, the optical properties of rare earth (RE)3+-doped LN single crystals, such as Nd3+, Er3+, Ho3+, Yb3+, Pr3+, and Dy3+:LN, have been studied in detail [1,2,3,4,5,6,7,8,9]; some of these crystals have been widely applied in solid-state optical devices. The trivalent samarium ion (Sm3+), which has a complex energy level structure, exhibits emission in the visible region, and is known for its rich orange emission at 610 nm and red emission 651 nm, both of which originate from the 4G5/2 state; such an emission is interesting for many photonic applications. For example, the reddish-orange emission band in the visible region, corresponding to the emission spectra of the 4G5/26H7/2 and 4G5/26H9/2 transitions, is desirable for uses in high-density optical storage, color displays, medical diagnostics, fluorescent display devices, and sensors [2,3,10,11]. More specifically, Sm3+-containing hosts have been offered as infrared counters, as an activated solid, and as a means to increase the photovoltaic solar cell conversion efficiency via down-conversion of the solar spectrum [12,13].
However, most of those reports have mainly focused on fluorosilicate glasses [14], alkyl-barium-bismuth-tellurite (LKBBT) glass [15], and other glasses at room temperature [16]. Dominiak-Dzik [2] reported the spectroscopic properties of a Sm:LN single crystal with a fixed Sm concentration at 0.65 wt.%, while Wang [3] also studied the optical properties of Mg:Sm (0.3 at%):LN and Zn:Sm (0.3 at%):LN at room temperature. Obviously, the doping concentration of rare metal ions has a strong impact on the fluorescence intensity of the doped materials [4]. Moreover, temperature also influences the luminescence property, such as Sm3+-doped oxyfluoride powders [17].
In this paper, all the congruent Sm:LN single crystals were grown using the Czochralski method. Our work is mainly focused on the impact of temperature and Sm3+-doped concentration on the spectral emission properties of Sm:LN single crystals. The effect of temperature and Sm3+-doped concentration variation, on the emission properties of Sm3+-doped LN single crystals, was investigated systematically. To predict the radiative properties, such as the radiative transition probabilities, radiative lifetimes, and branching ratios, a detailed J-O analysis was conducted. The X-ray diffraction (XRD) data and ultraviolet–visible (UV-vis) absorption spectra were used to qualitatively analyze the occupation mechanism of Sm3+ ions in the crystal lattice. For each specimen, we also calculated the stimulated emission cross-section σem, which is an important parameter signifying the energy extraction rate from a lasing material.

2. Materials and Methods

LiNbO3 single crystals doped with Sm2O3 were grown using the Czochralski method, along the c-axis, from a congruent melt composition (Li/Nb = 48.6/51.4). The concentrations of Sm3+ in the melt were 0.2, 0.6, 1.0, and 2.0 mol.%. The raw materials were Li2CO3, Nb2O5, and Sm2O3, and the purity of these raw materials were 99.99%. To remove CO2, the raw samples were heated to 750 °C slowly, and maintained at that temperature for 3 h, and then heated up to 1150 °C and maintained at that temperature for 3 h to synthesize Sm:LN polycrystals in a Pt crucible. The crystals were grown under the optimum conditions of a rotation rate of 15 rpm and pulling rate of 1.0 mm/h. The as-grown Sm:LN single crystals are transparent, crack-free, and faint yellow, as shown in Figure 1a. The crystals were polarized at 1180 °C with a direct current density of 5 mA/cm2. The polarized crystals were cut and double-side polished into c-axis wafers (X × Y × Z = 25 × 25 × 2 mm3), as shown in Figure 1b.
The crystalline phase was identified by an X-ray diffraction (XRD, D-MAX-2200 VPC) apparatus (HITACHI, Hitachi, Japan) equipped with a copper Kα radiation source. The diffraction data were acquired over scattering angles (2θ) from 10° to 90°. The absorption spectra were recorded using a UV–vis–NIR spectrophotometer (UV-3600, SHIMADZU, Kyoto, Japan) at room temperature. We also obtained fluorescence emission spectra at a series of temperatures (73, 173, 298, 373, and 423 K) using a photoluminescence spectrometer (FLS920, Edinburgh, UK) equipped with a temperature controller under 409 nm excitation from a 450 mW Xe lamp. The lifetime of the 4G5/2 level energy was measured using a μF920 flash lamp, and the induced time-resolved curves were recorded at various temperatures.

3. Results

3.1. XRD Studies of Sm:LN Single Crystals

As shown in Figure 2, the triclinic phase diffraction peaks of Sm:LN were predominant in the XRD pattern, and are quite similar to the peaks of pure LiNbO3. Only the LiNbO3 triclinic phase was observed. The results indicate that the doped samarium ion impurity cannot change the lattice structure of LiNbO3. The lattice parameters were calculated by Rietveld refinement, and the unit cell volumes were obtained using the formula V = (abc)·cos30°, as shown in Table 1. With the increasing concentration of doped Sm3+ ions, the unit cell volumes of the crystals initially increase. Since the comparable radii of the samarium ion, lithium ion, and niobium ion are 0.964 Å, 0.60 Å, and 0.70 Å, respectively, the doped samarium ions cannot exist as interstitial ions in the crystal.
According to the Li-site vacancy model [18], there are many Li vacancy (VLi) sites, and some excessive Nb ions occupy vacant Li sites to form anti-site Nb ( Nb Li 4 + ) defects, because the Li/Nb ratio is 0.946 in congruent LiNbO3. In the Sm-0.2 and Sm-0.6 samples, the weakly doped Sm3+ ions replace the anti-site Nb ( Nb Li 4 + ). In Sm-1.0, all the Nb Li 4 + ions are replaced by Sm3+ ions; this doping level has exceeded the “threshold value”. As the concentration of Sm3+ continues to increase, Sm3+ ions may replace both the Li-site and Nb-site, and form Sm Li 2 + Sm Nb 2 defect centers in the samples. Moreover, the c/a of Sm:LN lattice decreases as the Sm3+ ions concentration increases because the Sm Li 2 + Sm Nb 2 has a shorter chemical bond.

3.2. Absorption Spectra and Judd–Ofelt Analysis

The optical absorption spectra of various concentrations of Sm:LN single-crystal samples are shown in Figure 3a for the UV-vis region, and in Figure 3b for the NIR region; these spectra are also related to the transition of Sm3+ from the ground state 6H5/2 to the excited state.
To study the issue of defects and ion occupation in Sm:LN crystals, we calculated the polarizability of the ions by using the equation F = (Z*)2/r, where Z* = Z − Σs, and Z*, Z, Σs, and r represent the effective nuclear charge number, the nuclear charge number, the shield factor, and the radius of the ion, respectively. The calculated values of the polarizability of Li+, Nb5+, and Sm3+ ions were 2.82, 58.51, and 62.31, respectively. The basal optical absorption edge of the lithium niobate crystal is determined by the valence electron transition energy from the 2p orbital of O2− to the 4d orbital of Nb5+ [19]; this absorption edge can be used to analyze the ion occupation issue of the Sm:LN crystal qualitatively.
When the polarizability of the doped ion is lower than that of the replaced ion, the polarizability of O2− decreases, and the transition energy from the 2p orbital of O2− to the 4d orbital of Nb5+ increases. Additionally, the absorption edge shifts to the ultraviolet band and, conversely, the absorption edge shifts to the infrared band [20,21]. Figure 3a shows that the absorption edges of the samples, Sm-0.2 and Sm-0.6, redshift; however, when the concentration reaches 0.6 mol.%, the absorption edge shifts toward the ultraviolet. According to the Li-site vacancy model, when Sm3+ ions enter the LiNbO3 crystal lattice, they will first replace Nb Li 4 + and, then, occupy the normal Li-site. Since the polarization ability is in the order of Sm3+ > Nb5+ > Li+, the transition energy from the 2p orbital of O2− to the 4d orbital of Nb5+ decreases, and results in a shift of the absorption edge to the longer wavelengths when Nb Li 4 + defects were replaced by Sm3+ ion. After Nb Li 4 + are completely replaced by Sm3+, the doped ions may replace Li-site and Nb-site to form Sm Li 2 + Sm Nb 2 center [22,23,24], and give rise to the absorption edge that goes toward the shorter wavelengths; hence, the inversion of the absorption edge illustrates that the threshold concentration of Sm3+ in the LiNbO3 single crystal is between 0.6 mol.% and 1.0 mol.%. This is due to the interionic distances between O2− and Nb5+ becoming expanded with the doped Sm3+ concentration, which is caused by the lattice of Sm:LN crystals slightly increasing, as shown in Table 1. The absorption edge will shift to shorter wavelengths when the transition energy from the 2p orbital of O2− to the 4d orbital of Nb5+ decreases. The absorption edge shifts of Sm:LN crystals is different with anti-photorefractive ion, such as Mg2+ and Zr4+ [25,26], because the polarizability of Mg2+ and Zr4+ ions is less than that of Nb5+, while the polarizability of Sm3+ ion is bigger than that of the Nb5+ ion.
The J-O theory was independently developed by Judd [27] and Ofelt [28]. We describe some of the relevant and important formulas, as follows.
The calculated oscillator strength of an induced electric dipole transition, from the ground state, Ψ1J1, to the excited state, Ψ2J2, can be given by
f c a l = 8 π 2 m c 3 h ( 2 J 1 + 1 ) λ ( n 2 + 2 ) 2 9 n t = 2 , 4 , 6 Ω t | Ψ 1 J 1 U ( t ) Ψ 2 J 2 | 2 ,
where n is the refractive index of the lithium niobate wafer; 2(n2 + 2)/9n is the Lorentz local field correction and accounts for the dipole–dipole transition; h is Planck’s constant; Ωt (t = 2, 4, and 6) denotes the J-O intensity parameters; and ‖U(t)2 is the doubly reduced matrix elements of the unit tensor operator [29] evaluated in the intermediate coupling scheme for the transition Ψ1J1→Ψ2J2. The experimental oscillator strength for the transition from the ground state to the excited state (fexp) is directly proportional to the area under the absorption curve, and is expressed by Equations (2) and (3) [30]:
f e x p ( J 1 J 2 ) = mc 2 N π e 2 λ 2 α ( λ ) d λ ,
f e x p ( J 1 J 2 ) = 1 0.43 d OD ( λ ) d ( λ ) = 4.318 × 10 9 ε ( ν ) d ν ,
where m and e are the mass and charge of the electron, respectively; c is the speed of velocity of light; N is the concentration of ions doped in the crystal; and ε(υ) is the molar absorptivity of a band at a wave number, υ (cm−1). A least-squares fit method is then used for Equation (1) to determine the Ωt parameters that provide the best fit between the experimental and calculated oscillator strengths. The theoretical oscillator strength, fcal, is then obtained using Equation (1) and Ωt.
The root mean square deviation of the fit can be expressed as
δ r m s e = ( f c a l f e x p ) 2 L 3 ,
where L is the number of analyzed groups of the transition bands, and 3 is the number of fitted parameters.
As described above, with the obtained Judd–Ofelt parameters, Ωt, the radiative properties for the rare earth-doped materials could be evaluated through the relevant expressions given below. The spontaneous emission probability (Ar) from the initial manifold (Ψ1J1) to the final manifold (Ψ2J2) is
A r = 64 π 4 e 2 3 h ( 2 J 1 + 1 ) λ 3 n ( n 2 + 2 ) 2 9 t = 2 , 4 , 6 Ω t | Ψ 1 J 1 U ( t ) Ψ 2 J 2 | 2 ,
and the excited state radiant lifetime can be described as
τ r ( Ψ 1 J 1 ) = 1 Ψ 2 J 2 A r ( Ψ 1 J 1 , Ψ 2 J 2 ) .
Once the transition probability is obtained, the laser property-defining parameters can be estimated; the expression for the branching ratio is given as
β r ( Ψ 1 J 1 , Ψ 2 J 2 ) = A r ( Ψ 1 J 1 , Ψ 2 J 2 ) Ψ 2 J 2 A r ( Ψ 1 J 1 , Ψ 2 J 2 ) .
Combining the UV–vis–NIR absorption spectra at room temperature and Equations (1)–(3), fexp is acquired and tabulated in Table 2. For fcal, the refractive indices of Sm:LN were measured to be n = 2.24 [2]. The values of fexp and fcal, shown in Table 2, indicate that Sm-1.0 possesses highest oscillator strengths for all the transitions compared with other samples. The value of Judd–Ofelt parameters, Ω2, Ω4, and Ω6, increased as the concentration of Sm3+ ions increased, and reached the max value at 1.0 mol.%, then decreased sharp at 2.0 mol.%; despite the change of Ω, all the samples have the same relation Ω4 > Ω2 > Ω6. Judd et al. [31] have reported that the hypersensitivenesses are associated with the large value of the reduced matrix elements ‖U2‖, and the hypersensitiveness is mainly described by the parameter Ω2. Compared with ‖U2‖, ‖U4‖ and ‖U6‖ have little effect on the hypersensitive transition. The hypersensitive degree of the rare-earth ion can be measured from the relative variation of the Ω2 for a rare-earth ion in different host environments. Table 2 shows the relation of the oscillator strength of the hypersensitive transition and Ω2, with the concentration of Sm3+ doped in Sm:LN. The values of Ω2 could be used to explain, with consideration of the peculiarity of the hypersensitive transition 6H5/26P3/2 in absorption spectra of Sm3+ ions, and the exceptionally large oscillator strength of the transition 6H5/26P3/2 changed with the Sm3+ concentration in Table 2, which clearly indicated the arising of the hypersensitive transition phenomenon. The value of hypersensitive transition 6H5/26P3/2 and Ω2, varies in a same manner as the Sm3+ concentration changed from 0.2 to 2.0 mol.%.
Next, the radiative transition rates and branching ratio for emission, from 4G5/26HJ, were calculated at room temperature by using Equations (5)–(7); the results are displayed in Table 3. Obviously, the radiative transition rates and branching ratio of 4G5/26H7/2 and 4G5/26H9/2 transition are much higher than others, which is similar to that in Ref. [2]. The radiative lifetimes go down from 2465.54 to 1309.07 μs with the increasing of Sm3+ ion concentration from 0.2 to 1.0 mol.%, and then go up to 4480.89 μs when the concentration of Sm3+ ions reach 2.0 mol.%, which is similar to that of Sm3+-doped oxyfluoroborate glasses [32]. It is clear that the 4G5/26H7/2 transition, the orange-red radiative transition, was strengthened due to the increasing of doped Sm3+ ion concentration in the crystal. The “threshold concentration” of Sm3+-doped LN crystal is more superior in orange-red emission, which is quite useful for orange-red laser devices.
When compared to fluorozincate glass [14], oxyfluoride glass [17], silicate glass [33], PbFPSm10 and SNbKZFSm10 [34], the Sm (1.0 mol.%):LN single crystal possesses a higher value of Ω2, confirming the higher Sm–O covalency, and more asymmetry at the Sm3+ ion site. This result also indicates higher mixing of the opposite parity electronic configurations that are responsible for the spectral intensities. However, Ω4 and Ω6 are structure-dependent parameters that are related to the bulk properties of materials [14]. The spectroscopic quality factor, X, which is defined by the ratio of the intensity parameter Ω4 to Ω646), is an important predictor for stimulated emission in a laser-active material. The spectroscopic quality factor X of Sm (1.0 mol.%):LN was calculated to be 2.66, that is larger than that of the other samples, suggesting that Sm (1.0 mol.%):LN single crystal is a promising laser material.

3.3. Fluorescence Analysis

Due to the strong absorption peak of the 6H5/26P3/2 transition, compared with that of other level transitions in the UV–vis region, a 409 nm laser is used to measure the emission spectra. Simultaneously, we use the area of the peaks in the NIR regions to calculate the spectral parameters. The fluorescence spectra of various Sm:LN single crystals were measured at room temperature, as shown in Figure 4a.
Figure 4a clearly shows that the luminescence intensity increases with the increasing concentration of Sm3+, and reaches a maximum at a concentration of 1.0 mol.% Sm3+. To study the emission spectra in more detail, we present a simple energy level transition diagram of the Sm3+ ion in Figure 5 that includes two energy transition patterns: a single ion of Sm3+ and cross-relaxation among multiple Sm3+ ions [14].
There are main three transitions under 409 nm laser excitation corresponding to 4G5/26HJ (J = 5/2, 7/2, 9/2) in Figure 5. Based on the level energies of the samarium ion, we found the transitions 4G5/26F11/2 = 6H5/26F5/2, 4G5/26F5/2 = 6H5/26F11/2 occur in the crystals, and indicate a cross-relaxation process, such as with 4G5/26F9/26H5/26F7/2 and 4G5/26F7/26F9/2, as shown in Figure 5. This transition mechanism may explain the relationship between the concentration and fluorescence intensity.
Sm3+ ions have two energy transition patterns: (1) when the concentration of Sm3+ in the crystal is less than 1 mol.%, the radiation processes of Sm3+ are independent; i.e., the luminescence intensity enhances as the concentration increases; (2) when the concentration of Sm3+ exceeds the threshold value, the emission spectra mainly rely on energy transfer between Sm3+ ions; i.e., the emission spectra of Sm3+ is primarily caused by the increased energy transfer between Sm3+ ions via nonradioactive processes in heavily Sm3+-doped LiNbO3 single crystals. The prominent emission spectral peaks at 610 and 651 nm may be attributed to the multiphonon-assisted nonradiative relaxation in the crystal [34]. Hence, we have determined the relationship between Sm3+ and fluorescence intensity, and the mechanism of the energy level transitions of Sm3+. This mechanism can explain the fluorescence intensity of 2.0 mol.% Sm:LN being weaker than that of 1.0 mol.% Sm:LN well.
To investigate the effect of temperature on the fluorescence property of the Sm3+ ions in Sm:LN, we selected the 1.0 mol.% Sm:LN sample, and measured the fluorescence spectra under 409 nm laser excitation as the temperature changed from 73 to 423 K, as shown in Figure 4b. We found that the peak intensities of all samples decreased with the temperature increase, and the split peak shoulders of 4G5/26H7/2 and 4G5/26H9/2 are more apparent at a lower temperature than a higher temperature in the emission spectra at 610 nm [35]. Figure 4c shows the fluorescence intensity peak of 4G5/26H7/2, corresponding to the center of the peak at 610 nm. Figure 4d shows the fluorescence intensity peak of 4G5/26H9/2, corresponding to the center of the peak at 651 nm. From Figure 4c,d), it is shown that concentration of Sm3+ ions has great effect on the fluorescence intensity. The fluorescence intensity of Sm (1.0 mol.%):LN is 10 times and 13 times that in Sm (0.2 mol.%):LN at 610 nm and 651 nm, respectively. We can see that all the magnitudes of the fluorescent peaks are maximized when the concentration of Sm3+ is 1.0 mol.%, and the peaks for each concentration of Sm:LN crystals follow the same rule; i.e., the fluorescent peaks gradually decrease as the temperature increases. The reason for this behavior is the cross-relaxation process of Sm3+ ions, as shown in Figure 5, indicating that multiphonon relaxation of the 4G5/2 level plays an important role in the measured fluorescent spectra.
The experimental lifetimes of 1.0 mol.% Sm:LN were measured by monitoring the emission at 651 nm, corresponding to the 4G5/26H9/2 transition upon 409 nm pulsed laser excitation at various temperatures, as presented in Figure 6. The lifetime was found to gradually increase with the increasing temperature. However, the maximum change is only approximately 40 µs, which indicated that the lifetime of Sm3+ in Sm:LN single crystals is not sensitive to temperature.
The stimulated emission cross-section (σem) is an important parameter for predicting the energy extraction efficiency of a laser material. The stimulated emission cross-sections, σem, can be calculated using the formula
σem = λ5βI(λ)/(8πcn2τr∫λI(λ)),
where I(λ) is the experimental emission intensity as a function of the peak position, c is the light velocity, n is the mean refractive index, β is the branching ration of the transition, and τr is the radiative lifetime of the fluorescence level. The values of β and τr were calculated earlier in this paper using the Judd–Ofelt method. For Sm (1.0 mol.%):LN, the transitions of 4G5/26H7/2, 4G5/26H7/2, and 4G5/26H9/2 correspond to the center of the peaks at 568, 610, and 651 nm, respectively. The corresponding stimulated emission cross-sections, σem, of the peaks at 568, 610, and 651 nm, are 0.32 × 10−20 cm2, 1.257 × 10−20 cm2, and 1.081 × 10−20 cm2, respectively. Those values are much higher than the stimulated emission cross-section in the glass materials, such as Sm3+ in fluoroborate glasses, LKBBT, PbFPSm10, PKBASm10, PKBFASm10, and oxyfluoroborate [2,14,15,16,17,36].
The higher branching ratios and stimulated emission cross-sections for the 4G5/26H7/2 and 4G5/26H9/2 transitions suggest that the Sm:LN single crystals can be useful for laser applications. The higher stimulated emission cross-section is also favorable for low threshold and high gain laser applications, and can be utilized to obtain continuous-wave laser action [36]. Since the Sm:LN single crystals exhibit larger stimulated emission cross-sections and branching ratios, they are suitable for use in the development of visible lasers and optical fiber amplifiers.

4. Conclusions

Congruent Sm:LN single crystals, with concentrations of trivalent samarium ions changing from 0.2 to 2.0 mol.%, were grown using the Czochralski method. Based on XRD data and UV absorption spectra, the locations of the Sm3+ ions and the shifts in the absorption edges of Sm:LN were analyzed; after examining the emission spectra, we found that the optimal concentration of Sm3+ ions is 1.0 mol.%. The Sm Li 2 + Sm Nb 2 centers existed in highly doped Sm:LN single crystals, and play an important role in the process of luminescence. The fluorescence spectra obtained over a large temperature range demonstrated that the fluorescence intensity decreased with increasing temperature. However, the lifetime of the transition 4G5/26H9/2 only changed slightly with the increasing temperature. This observation indicates that the lifetime is not sensitive to the working temperature. Cross-relaxation among Sm3+ ions strongly depends on the material composition, because the lifetime of Sm3+ ions is very different in different materials. Under 409 nm laser excitation, the visible fluorescence spectrum has green light at 568 nm, orange bands at 610 nm, and reddish-orange light at 651 nm. The fluorescence branching ratio, β, for the4G5/26H7/2 and 4G5/26H9/2 transitions, is 0.4347 and 0.3846, respectively, indicating that Sm:LN single crystals can be an attractive laser material to exhibit efficient visible lasing emissions in the orange spectral region. We can adjust and control the ratio of the orange-red light through the concentration of Sm3+ ions and the environmental temperature. Such control also provides a method to regulate and control spectral components in visible-light region.

Author Contributions

Conceptualization, M.Y.; Methodology, S.L. (Shaopeng Lin); Formal Analysis, Y.Z., X.Y. and S.L. (Siwei Long); Investigation, M.Y.; Resources, B.W.; Writing—Original Draft Preparation, M.Y; Writing—Review & Editing D.M. and Y.Z.; Visualization, D.M.; Supervision, D.M. and B.W.; Project Administration, D.M. and B.W.; Funding Acquisition, B.W.

Funding

This research was funded by the National Natural Science Foundation of China (NSFC) (Nos. 11372361, 11302268, 11232015, 11472321, 11572355), the National Natural Science Foundation of Guangdong province (2017A030310426), the Guangdong Science & Technology Project (2015B090927005) and the Fundamental Research Funds for the Central Universities (17lgpy37).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Qian, Y.N.; Wang, R.; Xu, C.; Xu, W.; Wu, X.H.; Yang, C.H. Spectroscopic characteristics of 1.54 μm emission in Er/Yb:LiNbO3 crystals tridoped with In3+ ions. J. Alloy Compd. 2012, 527, 152–156. [Google Scholar] [CrossRef]
  2. Dominiak-Dzik, G. SM3+-doped LiNbO3 crystal, optical properties and emission cross-sections. J. Alloy Compd. 2005, 391, 26–32. [Google Scholar] [CrossRef]
  3. Wang, Y.H.; Wang, B.A.; Wu, D.; Lin, S.P.; Zhou, Z.F.; Zhu, Y.Z. Optical properties of Sm3+ doped Mg:LiNbO3 and Zn:LiNbO3 single crystals. Opt. Mater. 2012, 34, 845–849. [Google Scholar] [CrossRef]
  4. Long, S.W.; Ma, D.C.; Zhu, Y.Z.; Lin, S.P.; Wang, B. Effects of Zr4+ co-doping on the spectroscopic properties and yellow light emissions of Dy3+ in LiNbO3 single crystals. Opt. Mater. Express 2016, 6, 3354–3365. [Google Scholar] [CrossRef]
  5. Bloemen, M.; Vandendriessche, S.; Goovaerts, V.; Brullot, W.; Vanbel, M.; Carron, S.; Geukens, N.; Parac-Vogt, T.; Verbiest, T. Synthesis and Characterization of Holmium-Doped Iron Oxide Nanoparticles. Materials 2014, 7, 1155–1164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Xie, R.-J.; Hirosaki, N.; Li, Y.; Takeda, T. Rare-Earth Activated Nitride Phosphors: Synthesis, Luminescence and Applications. Materials 2010, 3, 3777–3793. [Google Scholar] [CrossRef] [Green Version]
  7. Dai, L.; Liu, C.R.; Yan, Z.H.; Han, X.B.; Wang, L.P.; Tan, C.; Xu, Y.H. Effect of dopant concentration on the spectra characteristic in Zr4+ doped Yb:Nd:LiNbO3 crystals. J. Rare Earth 2017, 35, 761–766. [Google Scholar] [CrossRef]
  8. Xing, L.; Wu, X.; Wang, R.; Xu, W.; Qian, Y. Upconversion white-light emission in Ho3+/Yb3+/Tm3+ tridoped LiNbO3 single crystal. Opt. Lett. 2012, 37, 3537–3539. [Google Scholar] [CrossRef] [PubMed]
  9. Zhang, P.X.; Chen, Z.Q.; Hang, Y.; Li, Z.; Yin, H.; Zhu, S.Q.; Fu, S.H.; Li, A.M. Enhanced 2.7 μm mid-infrared emissions of Er3+ via Pr3+ deactivation and Yb3+ sensitization in LiNbO3 crystal. Opt. Express 2016, 24, 25202–25210. [Google Scholar] [CrossRef] [PubMed]
  10. Kaur, P.; Kaur, S.; Singh, G.P.; Singh, D.P. Sm3+ doped lithium aluminoborate glasses for orange coloured visible laser host material. Solid State Commun. 2013, 171, 22–25. [Google Scholar] [CrossRef]
  11. Nakano, H.; Suehiro, S.; Furuya, S.; Hayashi, H.; Fujihara, S. Synthesis of New RE3+ Doped Li1+xTa1-xTixO3 (RE: Eu, Sm, Er, Tm, and Dy) Phosphors with Various Emission Colors. Materials 2013, 6, 2768–2776. [Google Scholar] [CrossRef] [PubMed]
  12. Krasutsky, N.J. 10-μm samarium based quantum counter. J. Appl. Phys. 1983, 54, 1261–1267. [Google Scholar] [CrossRef]
  13. Li, Q.B.; Lin, J.M.; Wu, J.H.; Lan, Z.; Wang, Y.; Peng, F.G.; Huang, M.L. Improving photovoltaic performance of dye-sensitized solar cell by downshift luminescence and p-doping effect of Gd2O3:Sm3+. J. Lumin. 2013, 134, 59–62. [Google Scholar] [CrossRef]
  14. Linganna, K.; Basavapoornima, C.; Jayasankar, C.K. Luminescence properties of Sm3+-doped fluorosilicate glasses. Opt. Commun. 2015, 344, 100–105. [Google Scholar] [CrossRef]
  15. Lin, H.; Wang, X.Y.; Lin, L.; Yang, D.L.; Xu, T.K.; Yu, J.Y.; Pun, E.Y.B. Spectral parameters and visible fluorescence of Sm3+ in alkali-barium-bismuth-tellurite glass with high refractive index. J. Lumin. 2006, 116, 139–144. [Google Scholar] [CrossRef]
  16. Kindrat, I.I.; Padlyak, B.V.; Lisiecki, R. Judd-Ofelt analysis and radiative properties of the Sm3+ centres in Li2B4O7, CaB4O7, and LiCaBO3 glasses. Opt. Mater. 2015, 49, 241–248. [Google Scholar] [CrossRef]
  17. Liu, F.C.; Han, Q.; Liu, T.G.; Yao, Y.Z.; Chen, Y.F.; Du, Y. Spectroscopic properties of Sm3+-doped oxy-fluoride powders with various Sm3+ concentration and sintering temperature. Opt. Mater. 2015, 45, 239–245. [Google Scholar] [CrossRef]
  18. Lerner, P.; Legras, C.; Dumas, J.P. Stoechiométrie des monocristaux de métaniobate de lithium. J. Cryst. Growth 1968, 3, 231–235. [Google Scholar] [CrossRef]
  19. Tang, Z.H.; Lin, S.P.; Ma, D.C.; Wang, B. Blue and green upconversion emissions of Zr:Nd:LiNbO3 single crystals. Int. J. Mod. Phys. B 2015, 29, 1550080. [Google Scholar] [CrossRef]
  20. Zhang, T.; Wang, B.; Ling, F.-R.; Fang, S.-Q.; Xu, Y.-H. Growth and optical property of Mg, Fe co-doped near-stoichiometric LiNbO3 crystal. Mater. Chem. Phys. 2004, 83, 350–353. [Google Scholar] [CrossRef]
  21. Zhu, Y.Z.; Lin, S.P.; Zheng, Y.; Ma, D.C.; Wang, B. Improvement of pyroelectric figures of merit in zirconia-doped congruent lithium niobate single crystals. J. Mater. Sci. 2016, 51, 3155–3161. [Google Scholar] [CrossRef]
  22. Kostritskii, S.M.; Maring, D.B.; Tavlykaev, R.F.; Ramaswamy, R.V. Energy transfer upconversion in Er-doped LiTaO3. Appl. Phys. Lett. 2000, 76, 2161–2163. [Google Scholar] [CrossRef]
  23. Qian, Y.N.; Wang, R.; Wang, B.A.; Xu, C.; Xu, W.; Wu, X.H.; Xu, Y.L.; Xing, L.L.; Yang, C.H. Growth of Zr codoped Er:LiNbO3 and Er/Yb:LiNbO3 single crystal. J. Cryst. Growth 2012, 361, 85–88. [Google Scholar] [CrossRef]
  24. Zhang, P.X.; Yin, J.G.; Zhang, L.H.; Liu, Y.C.; Hong, J.Q.; Ning, K.J.; Chen, Z.; Wang, X.Y.; Shi, C.J.; Hang, Y. Efficient enhanced 1.54 μm emission in Er/Yb: LiNbO3 crystal codoped with Mg2+ ions. Opt. Mater. 2014, 36, 1986–1990. [Google Scholar] [CrossRef]
  25. Polgár, K.; Kovács, L.; Földvári, I.; Cravero, I. Spectroscopic and electrical conductivity investigation of Mg doped LiNbO3 single crystals. Solid State Commun. 1986, 59, 375–379. [Google Scholar] [CrossRef]
  26. Abarkan, M.; Aillerie, M.; Kokanyan, N.; Teyssandier, C.; Kokanyan, E. Electro-optic and dielectric properties of Zirconium-doped congruent lithium–niobate crystals. Opt. Mater. Express 2014, 4, 179–189. [Google Scholar] [CrossRef]
  27. Judd, B.R. Optical Absorption Intensities of Rare-Earth Ions. Phys. Rev. 1962, 127, 750–761. [Google Scholar] [CrossRef] [Green Version]
  28. Ofelt, G.S. Intensities of Crystal Spectra of Rare-Earth Ions. J. Chem. Phys. 1962, 37, 511–520. [Google Scholar] [CrossRef]
  29. Carnall, W.T.; Fields, P.R.; Rajnak, K. Electronic Energy Levels in the Trivalent Lanthanide Aquo Ions. I. Pr3+, Nd3+, Pm3+, Sm3+, Dy3+, Ho3+, Er3+, and Tm3+. J. Chem. Phys. 1968, 49, 4424–4442. [Google Scholar] [CrossRef]
  30. Suhasini, T.; Kumar, J.S.; Sasikala, T.; Jang, K.W.; Lee, H.S.; Jayasimhadri, M.; Jeong, J.H.; Yi, S.S.; Moorthy, L.R. Absorption and fluorescence properties of Sm3+ ions in fluoride containing phosphate glasses. Opt. Mater. 2009, 31, 1167–1172. [Google Scholar] [CrossRef]
  31. Jørgensen, C.K.; Judd, B.R. Hypersensitive pseudoquadrupole transitions in lanthanides. Mol. Phys. 1964, 8, 281–290. [Google Scholar] [CrossRef]
  32. Mahamuda, S.; Swapna, K.; Venkateswarlu, M.; Rao, A.S.; Shakya, S.; Prakash, G.V. Spectral characterisation of Sm3+ ions doped Oxy-fluoroborate glasses for visible orange luminescent applications. J. Lumin. 2014, 154, 410–424. [Google Scholar] [CrossRef]
  33. Annapurna, K.; Dwivedi, R.N.; Kumar, A.; Chaudhuri, A.K.; Buddhudu, S. Temperature dependent luminescence characteristics of Sm3+-doped silicate glass. Spectrochim. Acta A 2000, 56, 103–109. [Google Scholar] [CrossRef]
  34. Kesavulu, C.R.; Jayasankar, C.K. Spectroscopic properties of Sm3+ ions in lead fluorophosphate glasses. J. Lumin. 2012, 132, 2802–2809. [Google Scholar] [CrossRef]
  35. Jørgensen, C.K.; Reisfeld, R. Judd-Ofelt parameters and chemical bonding. J. Less Common Metals 1983, 93, 107–112. [Google Scholar] [CrossRef]
  36. Saisudha, M.B.; Ramakrishna, J. Optical absorption of Nd3+, SM3+ and Dy3+ in bismuth borate glasses with large radiative transition probabilities. Opt. Mater. 2002, 18, 403–417. [Google Scholar] [CrossRef]
Figure 1. (a) As-grown and homogeneous single crystals orange 1.0 mol.% Sm3+-doped lithium niobate (LN) with the diameter of 25 mm and the length of 35 mm. (b) The double-side polished sample wafers, with Sm3+-doped concentration from 0.2 mol.% to 2.0 mol.%, were marked as Sm-0.2, Sm-0.6, Sm-1.0, and Sm-2.0, respectively.
Figure 1. (a) As-grown and homogeneous single crystals orange 1.0 mol.% Sm3+-doped lithium niobate (LN) with the diameter of 25 mm and the length of 35 mm. (b) The double-side polished sample wafers, with Sm3+-doped concentration from 0.2 mol.% to 2.0 mol.%, were marked as Sm-0.2, Sm-0.6, Sm-1.0, and Sm-2.0, respectively.
Materials 11 02058 g001
Figure 2. X-ray powder diffraction patterns of congruent LiNbO3 and series Sm3+-doped LiNbO3 single crystals.
Figure 2. X-ray powder diffraction patterns of congruent LiNbO3 and series Sm3+-doped LiNbO3 single crystals.
Materials 11 02058 g002
Figure 3. (a)UV–vis absorption spectra of Sm:LN single crystals. (b) Near-infrared absorption spectra of the Sm:LN single crystals.
Figure 3. (a)UV–vis absorption spectra of Sm:LN single crystals. (b) Near-infrared absorption spectra of the Sm:LN single crystals.
Materials 11 02058 g003
Figure 4. (a) The emission spectra of various concentration Sm:LN single crystals. (b) The emission spectra of 1.0 mol.% Sm:LN single crystals under series temperature. (c) The fluorescence intensity peaks of 4G5/26H7/2 corresponding to wavelength center 610 nm. (d) The fluorescence intensity peaks of 4G5/26H9/2 corresponding to wavelength center 651 nm.
Figure 4. (a) The emission spectra of various concentration Sm:LN single crystals. (b) The emission spectra of 1.0 mol.% Sm:LN single crystals under series temperature. (c) The fluorescence intensity peaks of 4G5/26H7/2 corresponding to wavelength center 610 nm. (d) The fluorescence intensity peaks of 4G5/26H9/2 corresponding to wavelength center 651 nm.
Materials 11 02058 g004
Figure 5. Simply energy levels diagram of Sm3+ ions in LN single crystals showing excitation and emission processes.
Figure 5. Simply energy levels diagram of Sm3+ ions in LN single crystals showing excitation and emission processes.
Materials 11 02058 g005
Figure 6. The fluorescence decay curves of the 4G5/26H9/2 transition in 1.0 mol.% Sm:LN single crystal at different temperatures.
Figure 6. The fluorescence decay curves of the 4G5/26H9/2 transition in 1.0 mol.% Sm:LN single crystal at different temperatures.
Materials 11 02058 g006
Table 1. The lattice constants of the crystal samples.
Table 1. The lattice constants of the crystal samples.
Samplea (Å)b (Å)c (Å)V3)c/a
CLN5.121575.1215713.80294313.552.69348
Sm-0.25.147985.1479813.85091317.892.69055
Sm-0.65.150595.1505913.84371318.052.68779
Sm-1.05.153225.1532213.84783318.472.68722
Sm-2.05.159725.1597213.85047319.342.68435
Table 2. Experimental and calculated oscillator strengths (10−6) for Sm:LN and the Judd–Ofelt intensity parameters Ωλ (10−20), the root mean square deviation △δrms, and the spectroscopic quality factor X46).
Table 2. Experimental and calculated oscillator strengths (10−6) for Sm:LN and the Judd–Ofelt intensity parameters Ωλ (10−20), the root mean square deviation △δrms, and the spectroscopic quality factor X46).
Transition 6H5/2Energy
(cm−1)
Oscillator Strengths
Sm-0.2Sm-0.6Sm-1.0Sm-2.0
fexpfcalfexpfcalfexpfcalfexpfcal
4D3/227,322----1.201.45--
6P3/2, 6P5/224,1504.864.895.045.319.319.992.902.93
4I13/2, 4I11/220,8331.150.441.560.462.602.010.890.27
6F11/210,4710.340.290.360.300.570.880.140.18
6F9/291521.991.871.981.924.502.911.141.14
6F7/279642.452.642.473.165.005.161.351.83
6F5/2, 6F3/2, 6F1/2, 6H15/266803.382.893.143.156.035.381.611.74
Ω2 = 1.39Ω2 = 1.52Ω2 = 2.37Ω2 = 0.84
Ω4 = 2.36Ω4 = 2.56Ω4 = 4.82Ω4 = 1.41
Ω6 = 1.21Ω6 = 1.24Ω6 = 1.81Ω6 = 0.74
X = 1.95X = 2.06X = 2.66X = 1.91
△δrms = 0.52△δrms = 0.79△δrms = 0.99△δrms = 0.46
Table 3. The peak wavelength (λem), radiative transition probability (Ar), branching ratio (β), total radiative transition probability (∑Ar), and radiative lifetime (τr).
Table 3. The peak wavelength (λem), radiative transition probability (Ar), branching ratio (β), total radiative transition probability (∑Ar), and radiative lifetime (τr).
Transition
4G5/2
λemSm-0.2Sm-0.6Sm-1.0Sm-2.0
(nm)A(S−1)β(%)A(S−1)β(%)A(S−1)β(%)A(S−1)β(%)
6H5/256815.510.038216.850.038730.930.04059.290.0416
6H7/2610178.340.4397189.920.4360332.050.434786.720.3886
6H9/2651160.380.3954173.860.3991296.810.384696.480.4323
6H11/271751.360.126654.960.1262104.110.136330.680.1375
∑Ar(S−1) 405.59435.59763.90223.17
τr (μs) 2465.542295.741309.074480.89

Share and Cite

MDPI and ACS Style

Yang, M.; Long, S.; Yang, X.; Lin, S.; Zhu, Y.; Ma, D.; Wang, B. Temperature-Dependent and Threshold Behavior of Sm3+ Ions on Fluorescence Properties of Lithium Niobate Single Crystals. Materials 2018, 11, 2058. https://doi.org/10.3390/ma11102058

AMA Style

Yang M, Long S, Yang X, Lin S, Zhu Y, Ma D, Wang B. Temperature-Dependent and Threshold Behavior of Sm3+ Ions on Fluorescence Properties of Lithium Niobate Single Crystals. Materials. 2018; 11(10):2058. https://doi.org/10.3390/ma11102058

Chicago/Turabian Style

Yang, Mingming, Siwei Long, Xin Yang, Shaopeng Lin, Yunzhong Zhu, Decai Ma, and Biao Wang. 2018. "Temperature-Dependent and Threshold Behavior of Sm3+ Ions on Fluorescence Properties of Lithium Niobate Single Crystals" Materials 11, no. 10: 2058. https://doi.org/10.3390/ma11102058

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop