Next Article in Journal
Diffusion of Tracer Atoms in Al4Ba Phases Studied Using Perturbed Angular Correlation Spectroscopy
Previous Article in Journal
Corrosion Mechanism and the Effect of Corrosion Time on Mechanical Behavior of 5083/6005A Welded Joints in a NaCl and NaHSO3 Mixed Solution
Previous Article in Special Issue
Optical Properties of Yttria-Stabilized Zirconia Single-Crystals Doped with Terbium Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Saturation Spectroscopic Studies on Yb3+ and Er3+ Ions in Li6Y(BO3)3 Single Crystals

Wigner Research Centre for Physics, Konkoly-Thege Miklós út 29-33, H-1121 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(8), 1151; https://doi.org/10.3390/cryst12081151
Submission received: 27 July 2022 / Revised: 11 August 2022 / Accepted: 13 August 2022 / Published: 16 August 2022
(This article belongs to the Special Issue Optical and Spectroscopic Properties of Rare-Earth-Doped Crystals)

Abstract

:
The results of a series of pump–probe spectral hole-burning experiments are presented on Yb3+- or Er3+-doped Li6Y(BO3)3 (LYB) single crystals in the temperature range of 2–14 K and 9–28 K, respectively. The spectral hole has a complex structure for Yb3+ with superposed narrow and broad bands, while a single absorption hole has been observed for Er3+. Population relaxation times (T1) at about 850 ± 60 μs and 1010 ± 50 μs and dipole relaxation times (T2) with values of 1100 ± 120 ns and 14.2 ± 0.3 ns have been obtained for the two components measured for the Yb3+:2F7/22F5/2 transition. T1 = 402 ± 8 μs and T2 = 11.9 ± 0.2 ns values have been found for the Er3+:4I15/24I11/2 excitation. The spectral diffusion rate at about 1 and 5 MHz/ms has been determined for the narrow and broad spectral line in Yb3+-doped crystal, respectively. The temperature dependence of the spectral hole halfwidth has also been investigated.

1. Introduction

Coherent quantum optical processes in rare-earth (RE) ion-doped dielectric crystals are of great interest in the quest for physical systems in order to realize quantum information processing devices. The successful realization of coherent quantum control procedures raises serious demands of atomic systems, such as a sufficiently long population, T1 and dipole relaxation, T2 times of the involved quantum states. Among several choices, e.g., magnetic- or laser-trapped ultracold single atoms or atomic clouds or coherent spin dynamics in quantum dots, a potential way to localize atomic systems is doping a single crystal with RE ions. The primary motivation of choosing Er3+ dopant ion is that its 4I13/24I15/2 transition is resonant with the frequency of the telecommunication laser field with a wavelength of about ~1.5 μm. A number of oxide crystals have been tested as a host of erbium to obtain a long coherence time in this wavelength range [1]. Moreover, the coherence time depends not only on the host material but on the concentration of the dopant, the temperature of the sample and the applied magnetic field as well [2,3,4,5]. Erbium-doped crystals have been used in several coherent quantum optical applications, such as the control of dispersion and group velocity [6], diode laser frequency stabilization [7], interference detection of spontaneous emission of light from two solid state atomic ensembles [8], electromagnetically induced transparency [9], ultraslow light propagation via coherent population oscillation [10], and quantum memories [11,12,13,14,15].
For practical applications, there are other relevant wavelength intervals beside the telecommunication domain. One of them is the near infrared range between 980–1100 nm. The ytterbium-doped crystal and fiber lasers radiate in the wavelength range between 1030–1080 nm, which can be pumped with ~980 nm light. There are several optical devices optimized to this wavelength domain [16,17].
Yb3+ and Er3+ dopant ions are perspective candidates for coherent quantum optical applications because the transitions 2F5/22F7/2 and 4I11/24I15/2, respectively, lay in the wavelength range 960–980 nm. Our goal is to investigate the use of the LYB crystal as a host for these ions according to the previous properties. Additionally, the solution of the problem about the interaction between the RE ions and the borate lattice could show us new research directions to develop borate crystals with targeted compositions.
Our aim was to study the ytterbium and erbium ions as dopants in a novel host crystal: lithium yttrium borate Li6Y(BO3)3 (LYB). This member of the borate crystal family can be grown by both the Bridgman and the Czochralski methods [18,19,20], which have a monoclinic structure with the P21/c space group. The advantage of LYB crystal is that rare-earth dopants can incorporate into Y site with C1 symmetry in the lattice without any distortion because of the charge neutral substitution and the similar ion radius. LYB:Yb3+ and LYB:Er3+ have already been investigated by optical spectroscopic methods [21,22,23,24,25] to determine the energy terms of the dopant ions. In addition, LYB:Yb3+ has been used for short pulse laser applications [26] and as a diode-pumped laser with Er co-doping [27].
A detailed description of the spectral hole burning procedure carried out by using sequential pump and probe pulses derived from a single laser beam passing through lithium niobate crystals can be found, e.g., in our previous papers [28,29]. The RE ions can be considered as effective two-level systems, where the laser-induced transition between the ground and excited states can be considered as a dipole transition. Thus, we suppose a set of effective two-state atoms interacting with a long and intensive pump and then a short and sufficiently weaker read-out probe pulse with a much longer time delay than the dipole relaxation time. In addition, the probe pulse needs to be much shorter than the population relaxation time but much longer than the dipole relaxation time. If these requirements are met, then by scanning the frequency of the probe field around the pumping frequency the attenuation of the probe field determines a Lorentzian curve, which is called spectral hole. The lifetime of the spectral hole was found to be single exponential with a characteristic time constant T1. The halfwidth of the spectral hole depends on the intensity of the pumping pulse, thus the value of T2 comes out only in the low intensity limit, which can be achieved by the Z-scan method described in the next section.

2. Materials and Methods

Lithium yttrium borate single crystals doped with 0.1 mol% Yb3+, 0.05 mol% or 1.0 mol% Er3+ were grown by the Czochralski method [20]. LYB has a monoclinic structure of the P21/c space group. The dielectric x-axis coincides with the crystallographic b-axis <010>, while the dielectric z-axis is perpendicular to the optical plane (-102). Although the crystal is biaxial, it behaves as uniaxial positive at λ = 1064 nm [21]. Yb3+-doped sample was prepared with incident plane (010) and a thickness of 0.6 mm, while the Er3+-doped sample was prepared with incident plane (-102) and a thickness of 2 mm. The high resolution absorption spectra of Er3+- or Yb3+-doped LYB crystals have been measured at 9 K by a Bruker IFS 66v/S or a Bruker IFS 120 FTIR spectrometer, respectively. In case of the Er3+ dopant, the 1 mol% concentration sample was more suitable for spectroscopy than the one containing fewer Er3+ ions, while for saturation spectroscopy, lower dopant concentration is required. The absorption spectra between 10,250 and 10,450 cm−1 wavenumbers are shown in Figure 1. In the case of the Yb3+-doped crystal, a single absorption peak was found at 10,283.5 cm−1 with a halfwidth of 0.18 cm−1, corresponding to the transition from the lowest Stark level of 2F7/2 multiplet to the lowest crystal field level of 2F5/2 multiplet of the Yb3+ ions in LYB at T = 9 K. The other two crystal field components of the 2F5/2 excited state lie out of this range and are not suitable for our laser spectroscopic measurements. However, all possible six absorption bands of the 4I15/24I11/2 transition of Er3+ can be identified in this wavenumber range. The positions and halfwidths of the investigated absorption lines are listed in Table 1.
The measurement setup used for the spectral hole-burning experiments can be seen in Figure 2. The light source was a stabilized external cavity diode laser (Sacher Lasertechnik, Manually Tunable Littrow Laser System–Lynx S3, tunable between 930 and 985 nm) with current, temperature, mechanical and piezo wavelength-tuning capability. For the narrow spectral hole measurements, a Toptica Photonics DL Pro grating stabilized tunable single-mode diode laser was used. This laser has a typical linewidth of 25 kHz with 5 μs integration time and less than 100 kHz in a millisecond time scale.
The wavelength of the laser beam was measured real time with a High Finesse WS-6 laser wavelength meter (WLM). The further details of the setup and the measurement method are described in our earlier paper [28]. The measurements presented in the following section were carried out by using an acousto-optic modulator (AOM) for laser frequency modulation, except for the temperature dependence of the broad spectral hole and all experiments on Er3+-doped sample, where laser frequency was modulated by the piezo modulation capability of the laser. The samples were mounted in a closed-cycle helium cryostat and cooled down between 9–30 K temperatures for the laser spectroscopic measurements. For the measurement of the narrow spectral hole of the LYB:Yb3+ sample, a liquid helium cryostat was used with an additional helium pump in order to cool down the sample to near 2 K.
To determine the low-intensity limit of the spectral hole halfwidth, a Z-scan measurement setup was established. The pumping laser intensity was varied from 1.5 to 190 W/cm2 for Yb3+- and from 9 to 700 W/cm2 for the Er3+-doped LYB, while the confocal lens pair in front and behind the cryostat was moved together from the starting position to shift the focal plane passing through the sample. All intensity data given here are the peak value calculated for a two-dimensional Gaussian beam profile. The probing pulse with an intensity of about one tenth of the pumping pulse was delayed at about 10 and 30 μs for Yb3+ and Er3+ ions, respectively. Assuming a Gaussian laser beam profile, a theoretical function can be fitted to the halfwidth (FWHM) of the spectral hole as the function of the sample position to determine the homogeneous linewidth in the low-intensity limit:
σ 2 ( z ) = f 2 2 2 + k f 2 2 [ 1 + ( z z 0 z R ) 2 ] + f 2 4 4 + k f 2 3 2 [ 1 + ( z z 0 z R ) 2 ] ,
where σ is the spectral hole halfwidth, f2 = 4/T2, zR is the Rayleigh length, and k depends on the material parameters, T1, optical power, and beam waist:
k = 4 T 1 | d e g | 2 μ 0 c P 2 n π w 0 2 .
Here, deg is the dipole momentum associated with the transition, μ0 is the permeability of the vacuum, c is the velocity of the light, P is the power of the focused beam, ℏ is the reduced Planck constant, n is the effective refractive index of the sample, and w0 is the beam waist.

3. Results

For the LYB:Yb3+ sample a special double spectral hole structure consisting of a narrow and a broad spectral hole (see Figure 3) was observed at about 10,283.58 cm−1, similarly to our measurement on LiNbO3:Yb3+ [28]. A potential reason for such a double spectral hole system can be the instantaneous spectral diffusion (ISD) [30]. In the LYB:Er3+ crystal, a single absorption hole was generated at the 4I15/24I11/2 electronic transition at about 10,275.36 cm−1 (Figure 4). For the higher wavenumber spectral lines in LYB:Er3+ (see absorption spectra in Figure 1), we found only weak spectral hole burning effect with a very broad halfwidth (>1400 MHz).
Using the Z-scan technique, homogeneous linewidths of about 0.29 ± 0.03 and 22.5 ± 0.5 MHz, corresponding to the T2 dipole relaxation times of about 1100 ± 120 and 14.2 ± 0.3 ns, were determined in LYB:Yb3+ crystal for the narrow and broad spectral hole components, respectively (for the narrow one see Figure 5). In Figure 5, the Rabi frequency Ωw of the pump pulse in the middle of the beam, calculated from the fitted parameters, is indicated on the right scale. For Er3+-doped LYB crystal a dipole relaxation time of about 11.9 ± 0.2 ns (corresponding to a homogeneous linewidth of about 26.7 ± 0.4 MHz) was determined by weighted averaging of the results of two measurements in different polarization directions. Although there is a strong polarization dependence in the absorption and the spectral hole depth, any significant polarization dependence was obtained neither in spectral hole halfwidth nor in relaxation times.
The determination of population relaxation time (T1) in LYB:Yb3+ and LYB:Er3+ crystals is shown in Figure 6, where the points represent the depths of the spectral holes as a function of delay time. In case of Yb3+ dopant, the linear behavior of the measured data in logarithmic scale shows a single exponential decay process both for the narrow (T1 = 850 ± 60 μs) and broad (T1 = 1010 ± 50 μs) spectral holes. In addition, the line broadening as a function of delay time, i.e., the spectral diffusion rates at about 1 and 5 MHz/ms, have also been determined for the narrow and broad spectral lines during these measurements, respectively, at a temperature of 9 K and an intensity of 40 mW/cm2 in the center of the beam at position z = −17 mm (see Figure 7). For this reason, the initial width is larger than its minimal value. The measurement of the population relaxation in the LYB:Er3+ crystal exhibits a single exponential decrease with a time constant of 402 ± 8 μs, as a weighted average of results for polarizations parallel to the dielectric axes x and y, since there is no reason to expect polarization-dependence for T1.
In Figure 8, the temperature dependence of the narrow spectral hole width measured in the LYB:Yb3+ crystal is shown, where the solid (red) line is a fitted curve using the direct one-phonon coupling model [31,32,33]:
Γ ( T ) = Γ 0 + a e x p ( ω p h k B T ) [ e x p ( ω p h k B T ) 1 ] 2
where Γ(T) is the halfwidth (i.e., full width at half maximum FWHM) of the spectral hole at a temperature T; Γ0 = Γ(T)T→0; a = 2(δω)2/γ; δω is the coupling constant; γ is the bandwidth; and ωph is the frequency of the phonon band.
Because of the very different halfwidths, the broad and narrow spectral holes were measured by using the piezo and AOM scanning methods, respectively. As a consequence, the measurements of the broad and narrow spectral holes were carried out with a pumping intensity of 2.6 W/cm2 and 9 W/cm2 and with a delay time of 200 and 20 μs, respectively. Similar temperature dependence was obtained for the LYB:Er3+ crystal. One can see that lowering the temperature of the sample below 9 K does not cause further significant decrease of the spectral hole halfwidth. The one-phonon coupling model describes the temperature dependence well for both crystals. The indicated ω0 parameter, characterizing the phonons playing a role during the relaxation, indicates similar relaxation processes through the lattice vibrations in LYB:Er3+ and for the narrow spectral hole in LYB:Yb3+ crystals.
For comparison, the T1 and T2 parameters measured on Yb3+- and Er3+-doped LiNbO3 [28,29] and Yb3+-doped Y2SiO5 crystals [34] are collected in Table 2.

4. Discussion

In the case of the Yb3+ ion, the population relaxation time (T1) of both absorption holes is at least twice larger in LYB than in LiNbO3. Additionally, though the dipole relaxation time of the broad hole is essentially similar in all cases, that of the narrow hole is much larger in the case of LYB. Similarly, the dipole and population relaxation times of the Er3+ electronic transition in LYB crystal are almost double those measured in LiNbO3. The increase of the population relaxation time can be explained through the higher level of defects in the lattice of LiNbO3 due to the incorporation of RE3+ ions as compared to the LYB crystal, where the incorporation into Y3+ sites can be established without charge compensation and lattice distortion. The difference between the dipole relaxation times may originate from the different magnetic environment of the incorporated RE ions. In the LiNbO3 crystal, the fluctuating magnetic field of the Nb5+ ions may influence the electronic state of the RE ions in a much stronger manner than any of the ions in the LYB crystal, resulting in significantly shorter relaxation time.
In addition, the temperature dependence of the spectral hole parameters can also give information about the RE ion-lattice coupling, e.g., phonon coupling parameters. Although the phonon band structure of the LiNbO3 and LYB crystals differs significantly, the similar ω0 coupling frequencies may show that some of the low frequency vibrational modes, probably localized ones, play an important role in the relaxation process in both crystals. In case of Er3+, the transition between the 2nd and 3rd Stark level of the 4I11/2 excited state corresponds to 60.7 cm−1 [26] (see Table 1.), which is very near to the fitted phonon frequency at 58 cm−1 and may confirm that the temperature dependence of the spectral hole width can be explained by the direct one-phonon coupling relaxation process. On the other hand, in case of Yb3+, a coupling phonon frequency of 58 ± 6 and 145 ± 8 cm−1 for the narrow and the broad spectral hole components was obtained, respectively, although both are quite far from even the smallest crystal field splitting value of 194 cm−1 of the 2F5/2 excited state [22]. It has to be mentioned, however, that low energy phonons around 130–160 cm−1 have been detected in the Raman spectra of Ce-doped LYB crystals [35], which is in good agreement with the value of ω0 obtained for the broad spectral hole component.
Among the other phonon-coupling relaxation processes, a possible one is the Raman relaxation. Its contribution to the spectral hole broadening is the following:
Δ Γ R = α ( T T D ) 7 0 x 0 x 6 e x p ( x ) ( e x p ( x ) 1 ) 2 d x ,
where α is the electron–phonon coupling parameter for the Raman process, T D = ω D /kB is the effective Debye temperature, ω D is the Debye cut-off frequency, kB is the Boltzmann-constant, and x 0 = T D / T   [32]. The integral has been calculated numerically, and we found that the single phonon relaxation formula in Equation (3), with appropriate change of the parameter ω p h   ω D , is almost identical with Equation (4) in the range of T/TD = 0–0.2. The parameter fitting of this equation describes the temperature dependence of the spectral hole broadening both for the narrow and the broad components quite well, and results in a value for TD equal to 111 ± 6 K and 131 ± 3 K, respectively. Since the numerical value of the expression in Equation (4) is practically the same as that given by the formula for the direct one-phonon model in Equation (3), the contribution of the direct one-phonon and the Raman relaxation cannot be separated only by fitting the temperature dependence of the spectral hole broadening. The temperature dependence of the spectral hole width may consist of contributions of more than one effect with similar temperature dependence. Orbach and multiphonon processes do not likely give significant contribution in relaxation for trivalent lanthanide ions [33].

5. Conclusions

A pump–probe type saturation spectroscopic experiment has been successfully employed to measure the relaxation parameters of the Yb3+: 2F7/22F5/2 and Er3+: 4I15/24I11/2 transitions in LYB single crystals in the temperature range of 2–14 and 9–28 K, respectively. The dipole and population relaxation times were found to be at about twice those measured in stoichiometric and congruent lithium niobate single crystals, except for the dipole relaxation time of the narrow component in Yb3+, where the increase is much larger. The small number of defects due to the neutral charge substitution of Y for the RE ions in LYB crystals explains the longer population relaxation time as compared to lithium niobate. The difference in the dipole relaxation times can be understood by the fluctuating magnetic field, which seems to be stronger in LiNbO3 due to the high nuclear magnetic moment of the Nb5+ ion. Although LYB as a host material seems to be better than LiNbO3, the coherence time of excitation of the Yb3+ dopant is still much shorter than in Y2SiO5, and it is likely that there is a similar situation with the Er3+ dopant as well as for the 4I11/24I15/2 transition.

Author Contributions

Conceptualization, Z.K.; methodology, Z.K.; software, G.M.; validation, G.M. and Z.K.; formal analysis, G.M.; investigation, G.M., K.L. and É.T.-R.; resources, Z.K. and É.T.-R.; data curation, G.M.; writing—original draft preparation, G.M.; writing—review and editing, Z.K., K.L., L.K. and É.T.-R.; visualization, G.M.; supervision, Z.K. and L.K.; project administration, Z.K.; funding acquisition, Z.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Innovációs és Technológiai Minisztérium” (the Ministry of Innovation and Technology) and the “Nemzeti Kutatási Fejlesztési és Innovációs Hivatal” (National Research, Development and Innovation Office) within the Quantum Information National Laboratory of Hungary.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors are grateful to Gábor Corradi for useful discussions and to Ivett Hajdara for participating in the early experiments.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sun, Y.; Thiel, C.W.; Cone, R.L.; Equall, R.W.; Hutcheson, R.L. Recent progress in developing new rare earth materials for hole burning and coherent transient applications. J. Lumin. 2002, 98, 281–287. [Google Scholar] [CrossRef]
  2. Bottger, T.; Thiel, C.W.; Sun, Y.; Cone, R.L. Optical decoherence and spectral diffusion at 1.5 μm in Er3+:Y2SiO5 versus magnetic field, temperature, and Er3+ concentration. Phys. Rev. B 2006, 73, 075101. [Google Scholar] [CrossRef]
  3. Bottger, T.; Sun, Y.; Thiel, C.W.; Cone, R.L. Spectroscopy and dynamics of Er3+:Y2SiO5 at 1.5 μm. Phys. Rev. B 2006, 74, 075107. [Google Scholar] [CrossRef]
  4. Guillot-Noel, O.; Vezin, H.; Goldner, P.; Beaudoux, F.; Vincent, J.; Lejay, J.; Lorgere, I. Direct observation of rare-earth-host interactions in Er:Y2SiO5. Phys. Rev. B 2007, 76, 180408. [Google Scholar] [CrossRef]
  5. Thiel, C.W.; Macfarlane, R.M.; Bottger, T.; Sun, Y.; Cone, R.L.; Babbitt, W.R. Optical decoherence and persistent spectral hole burning in Er3+:LiNbO3. J. Lumin. 2010, 130, 1603–1609. [Google Scholar] [CrossRef]
  6. Wang, G.; Xue, Y.; Wu, J.-H.; Gao, J.-Y. Phase dependences of optical dispersion and group velocity in an Er3+-doped yttrium aluminium garnet crystal. J. Phys. B At. Mol. Opt. Phys. 2006, 39, 4409–4417. [Google Scholar] [CrossRef]
  7. Bottger, T.; Sun, Y.; Pryde, G.J.; Reinemer, G.; Cone, R.L. Diode laser frequency stabilization to transient spectral holes and spectral diffusion in Er3+:Y2SiO5 at 1536 nm. J. Lumin. 2001, 9495, 565–568. [Google Scholar] [CrossRef]
  8. Afzelius, M.; Staudt, M.U.; de Riedmatten, H.; Simon, C.; Hastings-Simon, S.R.; Ricken, R.; Suche, H.; Sohler, W.; Gisin, N. Interference of spontaneous emission of light from two solid-state atomic ensembles. New J. Phys. 2007, 9, 413. [Google Scholar] [CrossRef]
  9. Xu, H.; Dai, Z.; Jiang, Z. Effect of concentration of the Er3+ ion on electromagnetically induced transparency in Er3+:YAG crystal. Phys. Lett. A 2002, 294, 19–25. [Google Scholar] [CrossRef]
  10. Baldit, E.; Bencheikh, K.; Monnier, P.; Levenson, J.A.; Rouget, V. Ultraslow Light Propagation in an Inhomogeneously Broadened Rare-Earth Ion-Doped Crystal. Phys. Rev. Lett. 2005, 95, 143601. [Google Scholar] [CrossRef]
  11. Staudt, M.U.; Hastings-Simon, S.R.; Nilsson, M.; Afzelius, M.; Scarani, V.; Ricken, R.; Suche, H.; Sohler, W.; Tittel, W.; Gisin, N. Fidelity of an Optical Memory Based on Stimulated Photon Echoes. Phys. Rev. Lett. 2007, 98, 113601. [Google Scholar] [CrossRef] [PubMed]
  12. Lauritzen, B.; Minar, J.; de Riedmatten, H.; Afzelius, M.; Gisin, N. Approaches for a quantum memory at telecommunication wavelengths. Phys. Rev. A 2011, 83, 012318. [Google Scholar] [CrossRef]
  13. Thiel, C.W.; Bottger, T.; Cone, R.L. Rare-earth-doped materials for applications in quantum information storage and signal processing. J. Lumin. 2011, 131, 353–361. [Google Scholar] [CrossRef]
  14. Thiel, C.W.; Sun, Y.; Macfarlane, R.M.; Bottger, T.; Cone, R.L. Rare-earth-doped LiNbO3 and KTiOPO4 (KTP) for waveguide quantum memories. J. Phys. B At. Mol. Opt. Phys. 2012, 45, 124013. [Google Scholar] [CrossRef]
  15. Dajczgewand, J.; Ahlefeldt, R.; Bottger, T.; Louchet- Chauvet, A.; Le Gouet, J.-L.; Chaneliere, T. Optical memory bandwidth and multiplexing capacity in the erbium telecommunication window. New J. Phys. 2015, 17, 023031. [Google Scholar] [CrossRef]
  16. Paschotta, R.; Nilsson, J.; Tropper, A.C.; Hanna, D.C. Ytterbium-doped fiber amplifiers. IEEE J. Quantum Electron. 1997, 33, 1049–1056. [Google Scholar] [CrossRef]
  17. Kränkel, C.; Fagundes-Peters, D.; Fredrich, S.T.; Johannsen, J.; Mond, M.; Huber, G.; Bernhagen, M.; Uecker, R. Continuous wave laser operation of Yb3+:YVO4. Appl. Phys. B 2004, 79, 543–546. [Google Scholar] [CrossRef]
  18. Shekhovtsov, A.N.; Tolmachev, A.V.; Dubovik, M.F.; Dolzhenkova, E.F.; Korshikova, T.I.; Grinyov, B.V.; Baumer, V.N.; Zelenskaya, O.V. Structure and growth of pure and Ce3+-doped Li6Gd(BO3)3 single crystals. J. Cryst. Growth 2002, 242, 167–171. [Google Scholar] [CrossRef]
  19. Yang, F.; Pan, S.K.; Ding, D.Z.; Chen, X.F.; Feng, H.; Ren, G.H. Crystal growth and luminescent properties of the Ce-doped Li6Lu(BO3)3. J. Cryst. Growth 2010, 312, 2411–2414. [Google Scholar] [CrossRef]
  20. Peter, A.; Polgar, K.; Toth, M. Synthesis and crystallization of lithium-yttrium orthoborate Li6Y(BO3)3 phase. J. Cryst. Growth 2012, 346, 69–74. [Google Scholar] [CrossRef]
  21. Zhao, Y.W.; Gong, X.H.; Chen, Y.J.; Huang, L.X.; Lin, Y.F.; Zhang, G.; Tan, Q.G.; Luo, Z.D.; Huang, Y.D. Spectroscopic properties of Er3+ ions in Li6Y(BO3)3 crystal. Appl. Phys. B 2007, 88, 51–55. [Google Scholar] [CrossRef]
  22. Sablayrolles, J.; Jubera, V.; Guillen, F.; Decourt, R.; Couzi, M.; Chaminade, J.P.; Garcia, A. Infrared and visible spectroscopic studies of the ytterbium doped borate Li6Y(BO3)3. Opt. Commun. 2007, 280, 103–109. [Google Scholar] [CrossRef]
  23. Zhao, Y.; Xie, Y.; Li, Z.; Gong, X.; Chen, Y.; Lin, Y.; Huang, Y. Polarized Spectral Properties of Er-Yb-Codoped Li6Y(BO3)3 Single Crystal. In Proceedings of the Symposium on Photonics and Optoelectronics (SOPO), Shanghai, China, 21–23 May 2012; pp. 1–3. [Google Scholar] [CrossRef]
  24. Skvortsov, A.P.; Poletaev, N.K.; Polgar, K.; Peter, A. Absorption spectra of Er3+ ions in Li6Y(BO3)3 crystals. Tech. Phys. Lett. 2013, 39, 424–426. [Google Scholar] [CrossRef]
  25. Kovács, L.; Arceiz Casas, S.; Corradi, G.; Tichy-Rács, É.; Kocsor, L.; Lengyel, K.; Ryba-Romanowski, W.; Strzep, A.; Scholle, A.; Greulich-Weber, S. Optical and EPR spectroscopy of Er3+ in lithium yttrium borate, Li6Y(BO3)3:Er single crystals. Opt. Mater. 2017, 72, 270–275. [Google Scholar] [CrossRef]
  26. Delaigue, M.; Jubera, V.; Sablayrolles, J.; Chaminade, J.-P.; Garcia, A.; Manek-Honninger, I. Mode-locked and Q-switched laser operation of the Yb-doped Li6Y(BO3)3 crystal. Appl. Phys. B 2007, 87, 693–696. [Google Scholar] [CrossRef]
  27. Zhao, Y.; Lin, Y.F.; Chen, Y.J.; Gong, X.H.; Luo, Z.D.; Huang, Y.D. Spectroscopic properties and diode-pumped 1594 nm laser performance of Er:Yb:Li6Y(BO3)3 crystal. Appl. Phys. B 2008, 90, 461–464. [Google Scholar] [CrossRef]
  28. Kis, Z.; Mandula, G.; Lengyel, K.; Hajdara, I.; Kovacs, L.; Imlau, M. Homogeneous linewidth measurements of Yb3+ ions in congruent and stoichiometric lithium niobate crystals. Opt. Mat. 2014, 37, 845–853. [Google Scholar] [CrossRef]
  29. Mandula, G.; Kis, Z.; Kovacs, L.; Szaller, Z.; Krampf, A. Site-selective measurement of relaxation properties at 980 nm in Er3+-doped congruent and stoichiometric lithium niobate crystals. Appl. Phys. B 2016, 122, 72. [Google Scholar] [CrossRef]
  30. Huang, J.; Zhang, J.M.; Lezama, A.; Mossberg, T.W. Excess dephasing in photon-echo experiments arising from excitation-induced electronic level shifts. Phys. Rev. Lett. 1989, 63, 78–81. [Google Scholar] [CrossRef]
  31. Yen, W.M.; Scott, W.C.; Schawlow, A.L. Phonon-Induced Relaxation in Excited Optical States of Trivalent Praseodymium in LaF3. Phys. Rev. 1964, 136, A271–A283. [Google Scholar] [CrossRef]
  32. Henderson, B.; Imbusch, G.F. Chapter 5: Electronic Centres in a Vibrating Crystalline Environment. In Optical Spectroscopy of Inorganic Solids; Monographs on the Physics and Chemistry of Materials 44, 1st ed.; Henderson, B., Imbusch, G.F., Eds.; Oxford University Press Inc.: New York, NY, USA, 1989; pp. 183–257. ISBN 9780199298624. [Google Scholar]
  33. Ellens, A.; Andres, H.; Meijerink, A.; Blasse, G. Spectral-line-broadening study of the trivalent lanthanide-ion series. I. Line broadening as a probe of the electron-phonon coupling strength. Phys. Rev. B 1997, 55, 173–179. [Google Scholar] [CrossRef]
  34. Lim, H.J.; Welinski, S.; Ferrier, A.; Goldner, P.; Morton, J.J.L. Coherent spin dynamics of ytterbium ions in yttrium orthosilicate. Phys. Rev. B 2018, 97, 064409. [Google Scholar] [CrossRef]
  35. Yuan, D.; Víllora, E.G.; Shimamura, K. Flux-growth of Ce:LiY6O5(BO3)3 single crystals during the Czochralski pulling from nearly congruent Ce:Li6Y(BO3)3 and their optical properties. J. Solid State Chem. 2021, 300, 122286. [Google Scholar] [CrossRef]
Figure 1. Absorption spectra of LYB:Er3+, cEr = 1 mol% (black, with numbered peaks) and LYB: Yb3+, cYb = 0.1 mol% (red, dotted) between 10,250 and 10,450 cm−1, measured at 9 K.
Figure 1. Absorption spectra of LYB:Er3+, cEr = 1 mol% (black, with numbered peaks) and LYB: Yb3+, cYb = 0.1 mol% (red, dotted) between 10,250 and 10,450 cm−1, measured at 9 K.
Crystals 12 01151 g001
Figure 2. Experimental setup for spectral hole-burning measurements at 9 K. The red box labelled as LASER involves an external cavity diode laser, a beam profiler, a Faraday isolator, a space filtering beam expander, and a lambda half-wave plate to prepare vertical polarization. M = mirror, GW = glass wedge, WLM = wavelength meter, PBS = polarizing beam splitter cube, AOM = acousto-optical modulator, λ/4 = lambda quarter-wave plate, L = lens, CRY = cryostat with crystal sample, and PD = avalanche photodiode with focusing lens system. For the lower temperature measurements, polarization maintaining a single mode optical fiber was used instead of the beam profiler and space filtering beam expander.
Figure 2. Experimental setup for spectral hole-burning measurements at 9 K. The red box labelled as LASER involves an external cavity diode laser, a beam profiler, a Faraday isolator, a space filtering beam expander, and a lambda half-wave plate to prepare vertical polarization. M = mirror, GW = glass wedge, WLM = wavelength meter, PBS = polarizing beam splitter cube, AOM = acousto-optical modulator, λ/4 = lambda quarter-wave plate, L = lens, CRY = cryostat with crystal sample, and PD = avalanche photodiode with focusing lens system. For the lower temperature measurements, polarization maintaining a single mode optical fiber was used instead of the beam profiler and space filtering beam expander.
Crystals 12 01151 g002
Figure 3. Double peak system of the spectral hole in LYB:Yb3+. A peak of 30–50 MHz width is superposed with a narrow peak of 2–3 MHz width. Black squares indicate the measured points, olive and blue lines indicate the fitted broad and narrow Lorentzian curves, and the red line shows the sum of two Lorentzians.
Figure 3. Double peak system of the spectral hole in LYB:Yb3+. A peak of 30–50 MHz width is superposed with a narrow peak of 2–3 MHz width. Black squares indicate the measured points, olive and blue lines indicate the fitted broad and narrow Lorentzian curves, and the red line shows the sum of two Lorentzians.
Crystals 12 01151 g003
Figure 4. Spectral hole burned in LYB:Er3+. The transmittance curve can be fitted by the superposition of the inhomogeneously broadened spectral line and a pure Lorentzian spectral hole.
Figure 4. Spectral hole burned in LYB:Er3+. The transmittance curve can be fitted by the superposition of the inhomogeneously broadened spectral line and a pure Lorentzian spectral hole.
Crystals 12 01151 g004
Figure 5. The width of the narrow spectral hole as a function of distance between the focal plane of the lens and the sample (Z-scan measurement) measured in LYB:Yb3+ at T = 2.2 K. The fit of Equation (1) is shown as a solid curve (left scale). The right scale approximately shows the Rabi frequency calculated from the fitted parameters.
Figure 5. The width of the narrow spectral hole as a function of distance between the focal plane of the lens and the sample (Z-scan measurement) measured in LYB:Yb3+ at T = 2.2 K. The fit of Equation (1) is shown as a solid curve (left scale). The right scale approximately shows the Rabi frequency calculated from the fitted parameters.
Crystals 12 01151 g005
Figure 6. The depth of the spectral holes as a function of delay time between the pump and probe pulses for the Yb3+- and Er3+-doped LYB crystals.
Figure 6. The depth of the spectral holes as a function of delay time between the pump and probe pulses for the Yb3+- and Er3+-doped LYB crystals.
Crystals 12 01151 g006
Figure 7. The impact of the spectral diffusion on the width of the spectral hole components in LYB:Yb3+ at T = 9 K.
Figure 7. The impact of the spectral diffusion on the width of the spectral hole components in LYB:Yb3+ at T = 9 K.
Crystals 12 01151 g007
Figure 8. The width of the narrow spectral hole as a function of temperature measured in LYB:Yb3+. The dots with error bars are the measured values, whereas the solid curve is the fitted exponential function, which originates from the direct one-phonon relaxation model. The measurement was taken at z = −25 mm position of the Z-scan setup, with 500 μs pumping time. For this reason, the halfwidth values and their zero-temperature limit are larger than its zero-intensity limit.
Figure 8. The width of the narrow spectral hole as a function of temperature measured in LYB:Yb3+. The dots with error bars are the measured values, whereas the solid curve is the fitted exponential function, which originates from the direct one-phonon relaxation model. The measurement was taken at z = −25 mm position of the Z-scan setup, with 500 μs pumping time. For this reason, the halfwidth values and their zero-temperature limit are larger than its zero-intensity limit.
Crystals 12 01151 g008
Table 1. Positions and halfwidths of the absorption lines in the spectrum of LYB:Er3+ and LYB:Yb3+ crystals at T = 9 K shown in Figure 1, and Stark energies of the crystal field splittings are determined by L. Kovács et al. [25] and J. Sablayrolles et al. [22].
Table 1. Positions and halfwidths of the absorption lines in the spectrum of LYB:Er3+ and LYB:Yb3+ crystals at T = 9 K shown in Figure 1, and Stark energies of the crystal field splittings are determined by L. Kovács et al. [25] and J. Sablayrolles et al. [22].
Wavenumber, Stark Energy (cm−1)Halfwidth (cm−1)Reference
Er3+4I15/210 [25]
247 [25]
376 [25]
4113 [25]
5368 [25]
6466 [25]
7497 [25]
8524 [25]
4I11/2110,275.340.14present work
210,295.970.15present work
310,356.690.51present work
410,394.971.17present work
510,402.081.13present work
610,410.850.77present work
Yb3+2F7/210 [22]
2367 [22]
3507 [22]
4676 [22]
2F5/2110,283.580.18present work
210,475≈50present work
310,810≈80present work
Table 2. Results of spectral hole-burning experiments on LYB:Yb3+ and LYB:Er3+, compared with those of Yb3+- and Er3+-doped stoichiometric (sLN) and congruent lithium niobate (cLN), and with Y2SiO5:Yb3+. L is the length of transmitted light path through the sample, i.e., the thickness of the sample; T1 and T2 are the population relaxation and dipole relaxation time constants of the transition, respectively; and ω0 is the characteristic phonon energy for the direct one-phonon coupling process in spatial frequency units.
Table 2. Results of spectral hole-burning experiments on LYB:Yb3+ and LYB:Er3+, compared with those of Yb3+- and Er3+-doped stoichiometric (sLN) and congruent lithium niobate (cLN), and with Y2SiO5:Yb3+. L is the length of transmitted light path through the sample, i.e., the thickness of the sample; T1 and T2 are the population relaxation and dipole relaxation time constants of the transition, respectively; and ω0 is the characteristic phonon energy for the direct one-phonon coupling process in spatial frequency units.
MaterialYb/Er Conc (mol%)α
(cm−1)
L
(mm)
T1
(μs)
T2
(ns)
ω0
(cm−1)
LYB:Yb3+
narrow
0.1 *9.60.60850 ± 601100 ± 12058 ± 6
broad 1010 ± 5014.2 ± 0.3145 ± 8
sLN:Yb3+ [28]
narrow
0.09260.57266 ± 17134 ± 7143 ± 20
broad 438 ± 2918.2 ± 0.554 ± 9
cLN:Yb3+ [28]
narrow
0.1811.70.48386 ± 34240 ± 2065 ± 8
broad 420 ± 2016 ± 189 ± 7
Y2SiO5:Yb3+ [34]0.005 4920 ± 10 103,000 ± 10,000
LYB:Er3+0.05 *3.12.00402 ± 811.9 ± 0.258 ± 3
sLN:Er3+ [29]0.13.41.85270 ± 1406.84 ± 0.1344 ± 3
cLN:Er3+ [29]0.10.441.89160 ± 805.83 ± 0.1137 ± 4
* in melt.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mandula, G.; Kis, Z.; Lengyel, K.; Kovács, L.; Tichy-Rács, É. Saturation Spectroscopic Studies on Yb3+ and Er3+ Ions in Li6Y(BO3)3 Single Crystals. Crystals 2022, 12, 1151. https://doi.org/10.3390/cryst12081151

AMA Style

Mandula G, Kis Z, Lengyel K, Kovács L, Tichy-Rács É. Saturation Spectroscopic Studies on Yb3+ and Er3+ Ions in Li6Y(BO3)3 Single Crystals. Crystals. 2022; 12(8):1151. https://doi.org/10.3390/cryst12081151

Chicago/Turabian Style

Mandula, Gábor, Zsolt Kis, Krisztián Lengyel, László Kovács, and Éva Tichy-Rács. 2022. "Saturation Spectroscopic Studies on Yb3+ and Er3+ Ions in Li6Y(BO3)3 Single Crystals" Crystals 12, no. 8: 1151. https://doi.org/10.3390/cryst12081151

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