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

La3+ and Er3+ co-doped Y2O3 transparent ceramics with a tunable refractive index and long coherence lifetime

Open Access Open Access

Abstract

La3+ is used as an index modifier to tune the refractive index of Er3+ doped Y2O3 transparent ceramics without reducing the coherence lifetime of Er3+ 4I13/24I15/2 transition. La3+ and Er3+ are incorporated into Y2O3 through a solution-phase nanoparticle synthesis and nanoparticles are sintered into transparent ceramics by hot isostatic press. The maximum La3+ doping concentration is about 10%, which increases the refractive index of Y2O3 by 0.009 (Δn/n = 0.48%). La3+ doping doesn’t reduce Er3+ optical coherence lifetime. The homogeneous linewidth of Er3+(20 ppm) 4I13/24I15/2 transition in La3+ doped Y2O3 ceramics is about 10 kHz at 2 K and 0.65 T, which is close to the reported homogeneous linewidth in Er3+ doped Y2O3 ceramics and single crystals. Such La3+ Er3+ co-doped Y2O3 ceramics are proper materials to fabricate optical waveguides for quantum memory applications.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Rare earth ion doped solids have attracted much attention as optical quantum memory materials [16]. The 4f-4f transitions of rare-earth ions have long optical coherence lifetime on the order of tens of microseconds, which enables the transfer of photonic quantum states to atomic states back and forth, constituting a photonic quantum memory. Although research in this field has been focusing on single crystalline materials, rare-earth ion doped polycrystalline ceramics could be alternatives to single crystals [713]. In previous work we reported Er3+ doped Y2O3 transparent ceramics with coherence lifetime (Er3+ 4I13/2-4I15/2 transition) comparable to that of Er3+ doped Y2O3 single crystals [1213]. These Er3+ doped Y2O3 transparent ceramics were made by sintering Er3+ doped Y2O3 nanoparticles which were synthesized via a wet-chemistry approach. Such a bottoms-up process offers convenient and low-cost methods to tune ceramic compositions and refractive index for photonic device fabrications.

Ceramic optical waveguides are highly desirable platform for optical quantum memories. Waveguide structures can offer several advantages such as significantly higher optical density, longer interaction length, more robust and compact device packaging, etc. However, fabricating ceramic waveguide quantum memories is challenging. One barrier is to find the proper core and cladding material combination that provides strong optical confinement, and at the same time does not detrimentally affect the optical coherence properties.

We report here a method to tune optical refractive index of Er3+ doped Y2O3 transparent ceramics without reducing optical coherence lifetime of Er3+ 4I13/24I15/2 transitions. La3+ dopants are used as index modifiers to change the refractive index of transparent Y2O3 ceramics. We choose La3+ because i) La3+ can be dissolved into Y2O3 at a maximum concentration of about 10% without phase separation based on the phase diagram of Y-La-O system [14], and ii) the 4f orbital of La3+ is empty and thus, it has no electronic states or spin states that can strongly interact with Er3+. On the other hand, the dominant La isotope is 139La, which has a natural abundance of 99.9% and a magnetic moment of 2.7832 ${\mu _N}$. Therefore, there would be some super-hyperfine interaction due to the nuclear spin of La. La3+ and Er3+ are doped into Y2O3 nanoparticles via a solution synthesis method and La3+, Er3+ co-doped Y2O3 transparent ceramics are made by sintering nanoparticles using hot isostatic press (HIP). La3+ doping increases optical refractive index of Y2O3 ceramics by about Δn = +0.001 per percent of La3+ (in molar; based on La/(La + Y)). The incorporation of La3+ increases the inhomogeneous linewidth by 2 orders of magnitude, from about 0.5 GHz (0% La3+) to 42 GHz (10% La3+). On the other hand, La3+ doping up to 10% doesn’t increase the homogeneous linewidth. At a temperature of 2 K and in a 0.65 T magnetic field, the homogeneous linewidth of Er3+ in La3+ doped Y2O3 ceramics is about 10 kHz, constituting an inhomogeneous linewidth to homogeneous linewidth ratio of $4\,{\times}\,{10^6}$.

2. Results and discussion

La, Er co-doped Y2O3 nanoparticles. La3+, Er3+ co-doped Y2O3 nanoparticles are synthesized by the reaction of urea, LaCl3, ErCl3, and YCl3 in boiled water, followed by thermal annealing. In a typical synthesis, a urea solution (6.7 mol urea in 0.35 L water at 100 oC) is poured into a metal salts solution (a mixture of LaCl3, ErCl3, and YCl3, 0.22 mol in 4L water at 100 °C). The reaction is kept at 100 °C for 1 hour. The thermal decomposition of urea provides OH- and CO32- which bond to metal ions and produce Y1-x-yLaxEry(OH)CO3·H2O amorphous nanoparticles. The amorphous nanoparticles are collected by filtration and are converted into crystalline (Y1-x-yLaxEry)2O3 nanoparticles by thermal annealing in air at 750 °C. The chemical yield is about 95%. The Er3+ and La3+ doping concentration can be tuned by varying Er/Y and La/Y ratios in the starting materials. All samples discussed in this work have the same Er3+ doping concentration of 20 ppm.

Figure 1(a) shows an SEM image of Y2O3 nanoparticles doped by 20 ppm Er3+ and 2% La3+. The nanoparticles are individually about 40 nm and are agglomerated. The particle size doesn’t show a clear change when the La3+ doping concentration is varied in a range of 0-10%. Powder X-ray diffraction of 10% La3+ doped Y2O3 nanoparticles reveals a pure cubic phase. This pattern matches well with a cubic Y2O3 phase, but all peaks are slightly shifted (Fig. 1(b)). This indicates that La3+ dopants are incorporated into Y2O3 lattice and expand the unit cell (La3+ is larger than Y3+). The unit cell parameter (a-axis) was calculated by Rietveld refinement of XRD patterns. The incorporation of La3+ into Y2O3 lattice expands cell parameters from 10.6143 ± 0.001 Å (Y2O3 nanoparticles), to 10.6275 ± 0.001 Å (2% La3+ doping) and 10.6728 ± 0.001 Å (10% La3+ doping) (Fig. 1(c)). This is due to the 14% larger ionic radius of La3+ (1.160 Å) compared with Y3+ (1.019 Å) [15]

 figure: Fig. 1.

Fig. 1. a) Scanning electron microscopy images of La3+(2%), Er3+(20 ppm) co-doped Y2O3 NPs. b) Powder X-ray diffraction pattern of La3+(10%), Er3+(20 ppm) co-doped Y2O3 nanoparticles. The red stick pattern corresponds to PDF 00-041-1105 (cubic Y2O3). c) Unit cell parameters of Y2O3 nanoparticles doped with different concentrations of La3+. The unit cell parameters are calculated by Rietveld refinement of XRD patterns.

Download Full Size | PDF

La3+, Er3+ co-doped Y2O3 transparent ceramics. Hot isostatic pressing (HIP) is used to sinter La3+, Er3+ co-doped Y2O3 nanoparticles into transparent ceramics. In a typical process, La3+, Er3+ co-doped Y2O3 nanoparticles (2 grams) are pressed into a ¾ inch diameter pellet in a steel die at approximately 8 klb force. This pellet is isostatically pressed at 25 kpsi in a latex isopressing sheath at room temperature, followed by sintering at 1600 °C in air for 2 h. The pellet is then hot isostatically pressed at 1590 °C for 16 h at 29 kpsi under an argon atmosphere. Figure 2(a) shows the optical transmission of a sample with 2% La3+ and 20 ppm Er3+ after HIP process. The axial transmission through 1.6 mm thickness (50 nm surface roughness) is about 80% in the wavelength range of 1250-2000 nm. This is close to the theoretical limit of Y2O3 at 1535 nm (82.7%) when only considering the surface reflections.

 figure: Fig. 2.

Fig. 2. a) Axial optical transmission through the thickness (1.6 mm) of a La3+(2%), Er3+(20 ppm) co-doped Y2O3 ceramic sample. A dip at about 850 nm is due to detector switch. b) Optical refractive index of Y2O3 transparent ceramics doped with different concentrations of La3+.

Download Full Size | PDF

La3+ doping increases the optical refractive index of transparent Y2O3 ceramics by about Δn = +0.001 per molar percent of La3+, as shown in Fig. 2(b). The highest La3+ doping we have achieved is 10%, which increases the index by 0.009 (Δn/n = 0.48%). Such an index difference can be used to fabricate optical waveguides where the core is La3+, Er3+ co-doped Y2O3 and the cladding is Y2O3.

La3+ doping significantly reduces the grain size of Y2O3 ceramics. Figure 3 shows the grain images of ceramic samples prepared by the same thermal process with different La3+ doping concentrations. The average grain area was reduced from 1072 µm2 (Y2O3) to 398 µm2 (1% La3+) and 1 µm2 (4% La3+).

 figure: Fig. 3.

Fig. 3. Electron back scattering diffraction (EBSD) images of grains in Er3+(20 ppm)-Y2O3 transparent ceramics doped with different concentrations of La3+. The ceramics are prepared by the same thermal process. The black dots at grain boundaries are where EBSD couldn’t identify crystal orientations. a) Y2O3; the average grain area is 1072 µm2. b) 1% La3+ doping; the average grain area is 398 µm2. c) 4% La3+ doping; the average grain area is 1 µm2. The average grain area is calculated by > 1500 grains. The scale bar is 100 µm in Fig. 3(a) and Fig. 3(b), 10 µm in Fig. 3(c).

Download Full Size | PDF

Optical dephasing properties of Er3+ in La3+ co-doped Y2O3 transparent ceramics. Incorporation of La3+ significantly increases the static structural disorder of the crystalline structure, due to the different sizes of La3+ and Y3+ ions and the random substitution. This can be directly observed from the inhomogeneous linewidth increase due to the La3+ co-doping (Fig. 4(a)). Here one can see that the inhomogeneous linewidth increases from 0.5 GHz without La3+ to about $42\, \pm \,3$ GHz at 10% doping, a 80 times increase. The initial increase is about 8 GHz broadening per percent La3+ doping, similar to those observed in Eu3+ doped Er3+:Y2SiO5 [16].

 figure: Fig. 4.

Fig. 4. a) Inhomogeneous linewidth ${\Gamma _{inh}}$ of Er3+(20 ppm)-Y2O3 transparent ceramics co-doped with different concentrations of La3+, measured at 1.7 K and zero magnetic field. b) Temperature dependence of the homogeneous linewidth ${\Gamma _h}$ (B = 0.65 T) of Er3+(20 ppm)-Y2O3 transparent ceramics co-doped with different concentrations of La3+.

Download Full Size | PDF

To see if La3+ co-doping affects the optical dephasing properties of Er3+, we measured the homogeneous linewidth $({\Gamma _h})$ of Er3+ 4I13/24I15/2 transition around 1535 nm in Er3+, La3+ co-doped Y2O3 transparent ceramics by two-pulse photon echo spectroscopy (details see Method section) [17], and compared it to the homogeneous linewidth of the Er3+ doped Y2O3 ceramics without La3+ co-dopants. The concentration of Er3+ is 20 ppm in all these samples, and all the thermal treatment procedures are the same. As shown in Fig. 4(b), doping La3+ into Y2O3 does not increase the homogeneous linewidth of Er3+, which is about 10 kHz at a temperature of 2 K and in a 0.65 T magnetic field. There is a small variation of linewidth for different La3+ doping concentration. However, it does not show a definite trend, and we suspect it is mainly caused by sample to sample variation. Nevertheless, previous reports have observed decreased linewidths in Eu3+ co-doped Er3+:Y2SiO5 and Sr3+ co-doped Eu3+/Y2O3 [16,1819]. They were attributed to increased broadening of spin transition, which resulted in less resonant spin-spin interaction between Er3+ or Eu3+ ions.

3. Conclusions

La3+ is used as an index modifier to tune refractive index of Er3+ doped Y2O3 transparent ceramics. La3+ and Er3+ are incorporated into Y2O3 through a solution-phase nanoparticle synthesis and transparent ceramics are made by sintering La3+, Er3+ co-doped Y2O3 nanoparticles. A maximum of 10% La3+ can be dissolved in Y2O3 ceramics without phase separation. La3+ doping increases optical refractive index of Y2O3 by about 0.001 per molar percent of La3+, but doesn’t influence the homogeneous linewidth of Er3+ 4I13/2-4I15/2 transition. The homogeneous linewidth of La3+, Er3+ co-doped Y2O3 transparent ceramics is about 10 kHz at 2 K and in a 0.65 T magnetic field, which is close to the reported homogeneous linewidth of Er3+ doped Y2O3 ceramics and single crystals. This work opens broad opportunities for making ceramic optical waveguides for quantum memory applications.

4. Methods

Materials. YCl3·6H2O (99.9%), ErCl3·6H2O (99.995%), LaCl3·7H2O (99.999%) and urea (99.0-100.5%) are purchased from Aldrich and used as received.

Optical Inhomogeneous Broadening Measurement. The inhomogeneous linewidth $({\Gamma _{inh}})$ is measured directly using an optical spectrum analyzer (OSA), except the one without La3+ co-doping, which was measured using a scanning laser. For all inhomogeneous linewidth measurement, the sample is measured at zero field and 1.7 K to minimize the broadening due to inhomogeneous Zeeman splitting or phonon broadening.

Optical Homogeneous Broadening Measurement. The homogeneous linewidth (Γh) of Er3+ 4I13/2-4I15/2 transition at 1534.94 nm in Er3+, La3+ co-doped Y2O3 transparent ceramics is measured by the two-pulse photon echo technique. The sample is placed in the Oxford Instruments Spectromag SM4000, with a pair of superconducting magnets. The light source is a tunable diode laser (Toptica DLpro) with <50 kHz linewidth. The light is temporally shaped by an acoustic-optic modulator (AOM, pulse width 50 ns), before being amplified by a fiber amplifier. The energy of the input pulse is carefully controlled so that there is no gain depletion seen from pulse to pulse. The output is filtered through a 1 nm bandpass filter and an electro-optic modulator (pulse width 50 ns) to cutoff amplified spontaneous emission from the amplifier. The beam is focused down to a spot of ∼50 µm. The input peak intensity on the sample is about a few kW/cm2. The resulting echo is filtered through a second electro-optic modulator, and detected by an InGaAs avalanche photo-diode and averaged on the oscilloscope.

Disclosures

The authors declare no conflicts of interest.

References

1. H. de Riedmatten and M. Afzelius, “Quantum light storage in solid state atomic ensembles,” https://arxiv.org/abs/1502.00307 [quant-ph] (2015).

2. A. I. Lvovsky, B. C. Sanders, and W. Tittel, “Optical quantum memory,” Nat. Photonics 3(12), 706–714 (2009). [CrossRef]  

3. J. J. Longdell, E. Fraval, M. J. Sellars, and N. B. Manson, “Stopped light with storage times greater than one second using electromagnetically induced transparency in a solid,” Phys. Rev. Lett. 95(6), 063601 (2005). [CrossRef]  

4. T. Zhong, J. M. Kindem, E. Miyazono, and A. Faraon, “Nanophotonic coherent light-matter interfaces based on rare-earth-doped crystals,” Nat. Commun. 6(1), 8206 (2015). [CrossRef]  

5. W. Tittel, M. Afzelius, T. Chaneliére, R. L. Cone, S. Kröll, S. A. Moiseev, and M. Sellars, “Photon-echo quantum memory in solid state systems,” Laser Photonics Rev. 4(2), 244–267 (2010). [CrossRef]  

6. C. Thiel, T. Bottger, and R. Cone, “Rare-earth-doped materials for applications in quantum information storage and signal processing,” J. Lumin. 131(3), 353–361 (2011). [CrossRef]  

7. A. Ferrier, C. W. Thiel, B. Tumino, M. O. Ramirez, L. E. Bausá, R. L. Cone, A. Ikesue, and P. Goldner, “Narrow Inhomogeneous and homogeneous optical linewidths in a rare earth doped transparent ceramic,” Phys. Rev. B 87(4), 041102 (2013). [CrossRef]  

8. N. Kunkel, A. Ferrier, C. W. Thiel, M. O. Ramírez, L. E. Bausá, R. L. Cone, A. Ikesue, and P. Goldner, “Rare-earth doped transparent ceramics for spectral filtering and quantum information processing,” APL Mater. 3(9), 096103 (2015). [CrossRef]  

9. N. Kunkel, J. Bartholomew, S. Welinski, A. Ferrier, A. Ikesue, and P. Goldner, “Dephasing mechanisms of optical transitions in rare-earth-doped transparent ceramics,” Phys. Rev. B 94(18), 184301 (2016). [CrossRef]  

10. N. Kunkel, J. Bartholomew, L. Binet, A. Ikesue, and P. Goldner, “High-resolution optical line width measurements as a material characterization tool,” J. Phys. Chem. C 120(25), 13725–13731 (2016). [CrossRef]  

11. K. Nathalie and G. Philippe, “Recent advances in rare earth doped inorganic crystalline materials for quantum information processing,” Z. Anorg. Allg. Chem. 644(2), 66–76 (2018). [CrossRef]  

12. H. Zhang, J. Yang, S. Gray, J. A. Brown, T. D. Ketcham, D. E. Baker, A. Carapella, R. W. Davis, J. G. Arroyo, and D. A. Nolan, “Transparent Er3+-doped Y2O3 ceramics with long optical coherence lifetime,” ACS Omega 2(7), 3739–3744 (2017). [CrossRef]  

13. J. Yang, H. Zhang, S. Gray, T. D. Ketcham, and D. A. Nolan, “Er3+ doped Y2O3 transparent ceramic for quantum memory applications,” Proc. SPIE 10771, 1077109 (2018). [CrossRef]  

14. O. Fabrichnaya, M. Zinkevich, and F. Aldinger, “Thermodynamic assessment of the systems La2O3-Al2O3 and La2O3-Y2O3,” Int. J. Mater. Res. 97(11), 1495–1501 (2006). [CrossRef]  

15. R. D. Shannon, “Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides,” Acta Cryst. 32(5), 751–767 (1976). [CrossRef]  

16. T. Böttger, C. W. Thiel, R. L. Cone, and Y. Sun, “Controlled compositional disorder in Er3+:Y2SiO5 provides a wide-bandwidth spectral hole burning material at 1.5 µm,” Phys. Rev. B 77(15), 155125 (2008). [CrossRef]  

17. N. A. Kurnit, I. D. Abella, and S. R. Hartmann, “Observation of a photon echo,” Phys. Rev. Lett. 13(19), 567–568 (1964). [CrossRef]  

18. C. W. Thiel, W. R. Babbitt, and R. L. Cone, “Optical decoherence studies of yttrium oxyorthosilicate Y2SiO5 codoped with Er3+ and Eu3+ for optical signal processing and quantum information applications at 1.5 microns,” Phys. Rev. B 85(17), 174302 (2012). [CrossRef]  

19. S. Welinski, C. W. Thiel, J. Dajczgewand, A. Ferrier, R. L. Cone, R. M. Macfarlane, T. Chanelière, A. Louchet-Chauvet, and P. Goldner, “Effects of disorder on optical and electron spin linewidths in Er3+, Sc3+:Y2SiO5,” Opt. Mater. 63, 69–75 (2017). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. a) Scanning electron microscopy images of La3+(2%), Er3+(20 ppm) co-doped Y2O3 NPs. b) Powder X-ray diffraction pattern of La3+(10%), Er3+(20 ppm) co-doped Y2O3 nanoparticles. The red stick pattern corresponds to PDF 00-041-1105 (cubic Y2O3). c) Unit cell parameters of Y2O3 nanoparticles doped with different concentrations of La3+. The unit cell parameters are calculated by Rietveld refinement of XRD patterns.
Fig. 2.
Fig. 2. a) Axial optical transmission through the thickness (1.6 mm) of a La3+(2%), Er3+(20 ppm) co-doped Y2O3 ceramic sample. A dip at about 850 nm is due to detector switch. b) Optical refractive index of Y2O3 transparent ceramics doped with different concentrations of La3+.
Fig. 3.
Fig. 3. Electron back scattering diffraction (EBSD) images of grains in Er3+(20 ppm)-Y2O3 transparent ceramics doped with different concentrations of La3+. The ceramics are prepared by the same thermal process. The black dots at grain boundaries are where EBSD couldn’t identify crystal orientations. a) Y2O3; the average grain area is 1072 µm2. b) 1% La3+ doping; the average grain area is 398 µm2. c) 4% La3+ doping; the average grain area is 1 µm2. The average grain area is calculated by > 1500 grains. The scale bar is 100 µm in Fig. 3(a) and Fig. 3(b), 10 µm in Fig. 3(c).
Fig. 4.
Fig. 4. a) Inhomogeneous linewidth ${\Gamma _{inh}}$ of Er3+(20 ppm)-Y2O3 transparent ceramics co-doped with different concentrations of La3+, measured at 1.7 K and zero magnetic field. b) Temperature dependence of the homogeneous linewidth ${\Gamma _h}$ (B = 0.65 T) of Er3+(20 ppm)-Y2O3 transparent ceramics co-doped with different concentrations of La3+.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.