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

Sensitization of Er3+ ions in silicon rich oxynitride films: effect of thermal treatments

Open Access Open Access

Abstract

The optical properties of reactive co-sputtered erbium doped silicon rich oxynitride (Er:SRON) films are studied as a function of annealing temperatures (Ta). The sensitization mechanism of Er3+ is found to evolve with Ta: excess Si related localized states play the essential role in samples when Ta is below 700 °C, while silicon nanoclusters (Si-NCs) become the dominate sensitizers when Ta exceeds 800 °C. Our results show that higher density of sensitized Er3+ could be acquired via energy transfer from localized states, and thus provide an alternative way for the engineering of light sources based on Er:SRON.

© 2014 Optical Society of America

1. Introduction

Erbium doped silicon-based materials provide a valuable approach for optoelectronic devices that can be monolithically integrated atop the widespread silicon electronics platform. Erbium, in its trivalent Er3+ state, could emit sharp luminescence at the standard telecommunication wavelength of 1.5 μm, and has hence attracted much attention [1]. Particularly, the observation of efficient energy transfer from Si nanoclusters (Si-NCs) embedded in SiO2 matrix to Er3+ ions has made a major breakthrough in this subject, increasing the Er3+ excitation cross section by 3~4 orders of magnitude as well as broadening the excitation band [24]. Recently, intense 1.54 μm luminescence of Er3+ ions has been observed in Er doped silicon rich oxynitride (Er:SRON), and non-resonant Er excitation via Si-NCs has been demonstrated [5,6]. Compared with silicon rich oxide (SRO), the widely studied host matrix for Er3+ ions, the narrower band gap of silicon rich oxynitride (SRON) favors the carrier injection, and the higher refractive index is beneficial for optical mode confinement. Additionally, the band structure of SiOxNy can be easily modulated by changing the oxygen/nitrogen ratio, and equivalent carrier injections could be achieved in Si/SiOxNy system [7], shedding light on the potential applications in bipolar devices for SRON. Generally, the Er excitation in the electroluminescence process in unipolar devices is quite different from that in bipolar devices. In unipolar devices [810], e.g. most of the reported SiO2 based light emitting devices under direct current conditions where electron conduction is dominant, the excitation probability of sensitizers via capture of electrons and holes is low and the efficient route to excite Er3+ ions is via impact excitation by hot carriers when the applied electric field is large enough. While in bipolar devices [11,12], injected electrons and holes can be efficiently captured by sensitizers and create excited states, which upon recombination transfer their energy to Er3+ ions and significantly increase the Er3+ excitation cross section, in a low electric field which is not sufficient to generate hot carriers. The onset voltage of Er3+ EL for the unipolar devices is usually high and thus incompatible with current CMOS technology. It would be reduced by bipolar injection, increasing consequently the reliability of devices. SRON containing Si-NCs is thus an ideal host matrix for Er3+ doping, taking into account both the optical and electrical pumping.

On the other hand, recent studies have shown that the presence of Si-NCs is not necessary to obtain intense Er-related luminescence. In parallel to the sensitization mechanism via Si-NCs, competing models based on luminescent centers (LCs) in SRO [13,14] and localized states in silicon rich nitride (SRN) [15] have been invoked. Due to the atomic size scale of LCs and localized states, sensitizers with much higher densities can be obtained, as opposed to Si-NCs. Indeed, there is now consensus in the literature, that the principal limitation for achievement of optical gain in Er-doped Si-NCs is the small fraction of Er3+ ions sensitized by Si-NCs [16,17], mainly due to the short interaction distance between them and the finite density of Si-NCs [18,19]. Hence, it is believed that the use of sensitizers with high densities, e.g. LCs or localized states, would increase the fraction of sensitized Er3+ ions and improve the photoluminescence (PL) yield, as demonstrated in [13]. However, up to now, there is little information regarding the sensitization mechanism of Er3+ ions in SRON. Obviously, this issue deserves intensive investigation to engineer efficient light emitting devices based on Er:SRON.

In this letter, we study the optical properties of Er:SRON films prepared by reactive co-sputtering. Post-deposition annealing is performed in a wide range of temperatures (Ta), from 600 °C where the films are amorphous, to 1100 °C where phase separation occurs. The evolution of sensitization mechanism with Ta has been particularly examined.

2. Experiment

Er:SRON films with thicknesses of approximately 150 nm were deposited onto p-type silicon substrates by reactive co-sputtering from Er, Si and Si3N4 targets. The deposition was conducted under the plasma of Ar-diluted 1% O2 atmosphere and the Si substrates were maintained at 100 °C. The concentrations of Si, O, N and Er in the films were obtained by Rutherford backscattering analysis, and the composition of the samples was found to be Si0.479O0.366N0.151Er0.004, with the Er concentration ~1.8 × 1020 at./cm3. The refractive index of the as-deposited sample is about 1.74 at 1540 nm measured by spectroscopic ellispometry. We have also measured the absorption coefficient of the as-deposited sample and derived the optical bandgap (2.3 eV). After deposition, the films were annealed under N2 flux for 1 h at temperatures in the range of 600-1100 °C. The microstructures of the annealed films were analyzed by transmission electron microscopy (TEM, Tecnai F30G2, FEI Company, USA). Matrix-related luminescence measurements in the visible range were carried out by pumping with a He-Cd laser at 325 nm and detecting with a charge coupled device (PIXIS: 100BR, Princeton Instruments, USA). For Er-related luminescence around 1540 nm, the samples were excited resonantly by a laser diode at 980 nm or non-resonantly by a Xe lamp at 325 nm. The PL signal in the near infrared range was measured by a liquid nitrogen cooled photomultiplier tube (PMT, Hamamatsu R5509, Japan). Time-resolved PL was excited by a microsecond lamp at 325 nm and recorded with a multichannel photon counting system (Edinburg Photonics, UK). A Xe lamp was used as light source in the PL excitation (PLE) spectroscopy measurements. All the PL measurements were taken at room temperature unless otherwise specified.

3. Results and discussion

Figure 1 shows the matrix-related PL spectra from Er:SRON films annealed at different temperatures. When Ta is below 700 °C, the PL intensity of samples is nearly indistinguishable. Though intense PL from LCs has been observed in low temperature annealed silicon rich oxide [13,14], cases seem to be different in our films, which indicate the absence of LCs in SRON. When Ta exceeds 800 °C, a broad PL band peaking between 700 and 800 nm can be clearly observed. The PL peaks redshift with the increase of Ta. Figure 1(b) shows the typical time-resolved PL spectra recorded at 700 nm from the specimen annealed at 800 °C. By fitting the spectra with a stretched exponential function, we can obtain that the decay time is about 42 μs. Such slow PL at these wavelengths is commonly attributed to radiative recombination of excitons confined in Si-NCs. These observations reveal the formation of Si-NCs after annealing at 800 °C. Since no lattice fringes can be detected in the cross sectional TEM image of 800 °C annealed specimen (figures not shown), Si-NCs formed at this temperature should be amorphous rather than crystalline. Si-NCs will continue to grow and gradually crystallize with the increase of Ta. In the TEM image of 1100 °C annealed specimen, lattice fringes of Si-NCs can be observed, as shown in the inset of Fig. 1(a). Redshift of the PL peaks is due to an increase of Si-NC diameter with Ta, and a resulting reduction of the band gap.

 figure: Fig. 1

Fig. 1 (a) Matrix-related PL spectra from Er:SRON films annealed at temperatures in the range of 600-1100 °C. The inset shows the cross sectional TEM image from the specimen annealed at 1100 °C. (b) Time resolved PL spectra recorded at 700 nm from the specimen annealed at 800 °C. The excitation wavelength is 325 nm.

Download Full Size | PDF

Figure 2(a) shows the PL excitation (PLE) spectra of Er3+ measured at 1540 nm. Interestingly, all the PLE spectra show a broad band in the range of 250-500 nm, with no direct Er3+ absorption peaks superimposed on. These results clearly show non-resonant excitation of Er3+ in all the samples. In addition, the PLE spectra shift to shorter wavelengths by increasing Ta from 600 °C to 900 °C, while they show an identical shape, except for an increase in the intensity, when further increasing Ta to 1100 °C. For annealing temperatures lower than 700 °C, there seems to be “dark sensitizers” of Er3+ ions in our samples. Indeed, sensitization of Er3+ via “dark” triplet Si-related oxygen deficient centers (ODCs) has been demonstrated in Er doped SRO, with a reversed temperature dependence in which the integrated Er3+ PL intensity increases upon heating [20]. However, we have measured the temperature dependence of integrated Er3+ PL intensity in the 600 °C annealed specimen and find it quenches upon heating (shown in the inset of Fig. 2(c)), indicating that ODCs may not be the main sensitizers in our samples. Nevertheless, taking the PLE peak shift into consideration, we suggest that the excess Si related localized states in the band gap of silicon oxynitride may play the dominant role in sensitizing Er3+ in low temperature annealed samples [15]. Generally, the annealing process would lead to the microstructure reconfiguration of SRON:Er films. The excess Si would concentrate to form Si-rich regions upon annealing, and phase separation would start when the annealing temperature exceeds 800 °C. Since the solubility of Er in SiO2 is much higher than that in Si, it is reasonable to expect that Er3+ ions tend to stay in Si-poor regions. The blue shift of PLE peaks is then attributed to the degenerate absorption of low-energy photons in Si-poor regions. Note that no noticeable PLE peak shifts can be observed when Ta exceeds 900 °C, indicating that the phase separation has completed or another sensitization mechanism takes place.

 figure: Fig. 2

Fig. 2 (a) PL excitation spectra of Er3+ measured at 1540 nm. The PLE spectra measured at 750 nm for the 1100 °C annealed specimen is denoted by empty squares.(b) Er-related PL spectra of samples annealed at different temperatures for excitation at 325 nm. (c) Er3+ PL decay time τdec as a function of the annealing temperatures. The inset shows the evolution of integrated Er3+ PL intensity with temperatures in the 600 °C annealed specimen. (d) IPL/τdec as a function of the annealing temperatures for excitation at 325 nm and 980 nm (right hand scale).

Download Full Size | PDF

The dependence of Er-related PL with Ta is shown in Fig. 2(b) for excitation at 325 nm. The spectra show the typical features of Er3+ emission, originated from the intra-4f 4I13/24I15/2 transition, with a main peak at 1540 nm. The thermal treatments improve gradually the Er3+ emission by a factor of ~2 as Ta is increased from 600 to 1100 °C. Simultaneously, the full width at half maximum (FWHM) of Er3+ related PL peaks around 1540 nm decreases monotonically by ~12%, suggesting the modulation of Er3+ local environment, resulting from the continuous microstructure evolvement such as phase separation and lattice relaxation during thermal treatments. The decay traces of Er3+ PL at 1540 nm are fitted with a stretched exponential function and the extracted PL decay times τdec are plotted in Fig. 2(c) as a function of the annealing temperatures. τdec shows a monotonic increase with Ta, from 0.11 ms to 0.47 ms, as Ta is increased from 600 to 1100 °C. Meanwhile, the extracted exponent beta value increases initially from 0.69 in the 600 °C annealed specimen to 0.75 in the 800 °C annealed specimen, and then decreases eventually to 0.63 in the 1100 °C annealed specimen. The relatively short decay time is probably due to the abundant nonradiative channels, i.e. Si dangling bonds in our samples [21]. During the thermal treatments, the gradual removal of non-radiative channels leads to the increase of Er3+ decay time.

In the linear excitation regime, we have [22]

IPLφσErNEr,senτdecτrad
where φ is the photon flux, σEr is the effective excitation cross section of Er3+ ions, NEr,sen is the density of sensitized Er3+ ions, τdec is the decay time and τrad is the radiative lifetime. After a simple transformation of Eq. (1), we have Eq. (2) as follows:
NEr,senτradφσErIPLτdec
Thus the ratio IPL/τdec gives information on NEr,sen in relative unites, assuming that the Er3+ effective excitation cross section σEr and radiative lifetime τrad don’t vary significantly. Figure 2(d) shows the evolution of IPL/τdec as a function of Ta for excitation at 325 nm and it can be divided into three parts. In the range of 600-700 °C, this quantity remains almost constant; in the range of 700-900 °C a sudden drop is observed; while in the range of 900-1100 °C it varies slightly. The result is different from that in SRO [13], which is a continuously decreasing function of the annealing temperature, as well as that in SRN [23], which gradually decreases for annealing temperatures up to 1000 °C and strongly increases for higher annealing temperatures. Since Er ions clustering, which significantly reduces the density of optically active Er3+ ions, was observed in high temperature (> 800 °C) annealed Er:SRO [24], the degradation of NEr,sen may be ascribed to this phenomenon. To clarify this issue, we have measured the PL intensity of Er3+ resonantly excited by a laser diode at 980 nm so that the sensitization of Er3+ could be excluded. Consequently, by assuming that the Er3+ resonant excitation cross section doesn’t vary significantly in our samples, IPL/τdec for excitation at 980 nm gives information on the density of optically active Er3+ ions, no matter whether they couple to sensitizers or not. It is interesting to find that the ratio of IPL/τdec excited at 980 nm shows a monotonic increase with the annealing temperatures, as shown in Fig. 2(d) (right hand scale). This indicates that the high temperature thermal treatments may not lead to the detrimental Er ions clustering in our cases, but would rather promote the diffusion of Er3+ ions to optically active sites. Thus, the degradation of NEr,sen is not ascribed to Er ions clustering, but a result of change of sensitizers. It is worth noting here, that the mediated excitation cross section of Er3+ is several orders of magnitude higher than the direct one, so the data in Fig. 2(d) are only considered in their trends for different excitations, but not the absolute values. On the basis of the experimental results, we suggest that two different sensitization processes of Er3+ ions may occur in Er:SRON, depending on the annealing temperatures. For Ta lower than 700 °C, excess Si related localized states in the band gap of SRON play the dominant role in sensitizing Er3+ ions. However, such states are not stable during thermal treatments, and their density degrades significantly when Ta exceeds 800 °C, simultaneously with the appearance of clearly observed matrix-related PL, i.e., the formation of Si-NCs. Indeed, the growth of Si-NCs leaves behind a significant number of Er3+ ions in regions with a reduced local concentration of excess Si, as convinced by the continuously decreasing FWHM of Er3+ PL with Ta. Hence, we believe that Er3+ ions are mainly excited via Si-NCs for annealing temperatures higher than 900 °C. This is further supported by the PLE spectra of Si-NCs, which show similar wavelength dependence to that of Er3+ ions, as shown in Fig. 2(a). For the purpose of clarity, only the PLE spectrum measured at 750 nm for the 1100 °C annealed specimen is provided (denoted by empty squares), with the intensity multiplied by a prompt factor for the convenience of reading.

Note that the ratio of IPL/τdec excited at 325 nm in low temperature (700 °C ≤) annealed samples is more than twice of that in high temperature (≥ 800 °C) annealed samples, indicating that higher NEr,sen can be acquired via sensitization by excess Si related localized states, as opposed to Si-NCs. This is due to an increase in the density of sensitizers, taking into account the degradation of optically active Er3+ ions by decreasing Ta. The main limitation of Er3+ PL in low temperature (700 °C ≤) annealed samples seems to lie in the relatively short lifetime, which is due to the unpassivated nonradiative channels. Different from SRO [25], high density of Si dangling bond, an efficient nonradiative channel for Er3+, would be obtained in SRN as well as SRON [21]. By further optimization such as hydrogen passivation, it is believed that intense Er3+ PL via energy transfer from localized states, comparable to that sensitized by Si-NCs, could be achieved when reducing the annealing temperature to 600-700 °C, a temperature range which is more compatible with the CMOS technology.

4. Conclusions

In conclusion, sensitization of Er3+ luminescence was studied in Er doped silicon rich oxynitride films fabricated by reactive co-sputtering with different annealing temperatures. It is demonstrated that the sensitization mechanisms evolve with Ta as well as the microstructures of the films. In the low temperature (700 °C ≤) annealed samples, the films are amorphous and the “dark” excess Si related localized states would dominate in sensitizing Er3+. However, such states are unstable during thermal treatments and their density degrades significantly when annealing temperature exceeds 800 °C, where phase separation starts and Si-NCs play an essential role in sensitizing Er3+ ions. Though encountered by a reduction in the density of optically active Er3+ ions in low temperature annealed samples, higher density of sensitized Er3+ ions is acquired via sensitization by excess Si related localized states, due to an increase in the density of sensitizers, as opposed to Si-NCs.

Acknowledgments

This work was supported by the 973 Program (No. 2013CB632102), and the Innovation Team Project of Zhejiang Province (No. 2009R5005).

References and links

1. A. Polman, “Erbium implanted thin film photonic materials,” J. Appl. Phys. 82(1), 1–39 (1997). [CrossRef]   [PubMed]  

2. A. J. Kenyon, P. F. Trwoga, M. Federighi, and C. W. Pitt, “Optical properties of PECVD erbium-doped silicon-rich silica: evidence for energy transfer between silicon microclusters and erbium ions,” J. Phys. Condens. Matter 6(21), L319–L324 (1994). [CrossRef]  

3. M. Fujii, M. Yoshida, Y. Kanzawa, S. Hayashi, and K. Yamamoto, “1.54 μm photoluminescence of Er3+ doped into SiO2 films containing Si nanocrystals: Evidence for energy transfer from Si nanocrystals to Er3+,” Appl. Phys. Lett. 71(9), 1198–1200 (1997). [CrossRef]  

4. F. Priolo, G. Franzò, D. Pacifici, V. Vinciguerra, F. Iacona, and A. Irrera, “Role of the energy transfer in the optical properties of undoped and Er-doped interacting Si nanocrystals,” J. Appl. Phys. 89(1), 264–272 (2001). [CrossRef]  

5. S. Cueff, C. Labbé, L. Khomenkova, O. Jambois, P. Pellegrino, B. Garrido, C. Frilay, and R. Rizk, “Silicon-rich oxynitride hosts for 1.5 μm Er3+ emission fabricated by reactive and standard RF magnetron sputtering,” Mater. Sci. Eng. B: Adv. Funct. Solid-State Mater. 177(10), 725–728 (2012). [CrossRef]  

6. L. Xu, L. Jin, D. Li, and D. Yang, “Effects of excess silicon on the 1540 nm Er3+ luminescence in silicon rich oxynitride films,” Appl. Phys. Lett. 103(7), 071101 (2013). [CrossRef]  

7. X. Wang, R. Huang, C. Song, Y. Guo, and J. Song, “Effect of barrier layers on electroluminescence from Si/SiOxNy multilayer structures,” Appl. Phys. Lett. 102(8), 081114 (2013). [CrossRef]  

8. A. Nazarov, J. M. Sun, W. Skorupa, R. A. Yankov, I. N. Osiyuk, I. P. Tjagulskii, V. S. Lysenko, and T. Gebel, “Light emission and charge trapping in Er-doped silicon dioxide films containing silicon nanocrystals,” Appl. Phys. Lett. 86(15), 151914 (2005). [CrossRef]  

9. O. Jambois, J. M. Ramírez, Y. Berencén, D. Navarro-Urrios, A. Anopchenko, A. Marconi, N. Prtljaga, A. Tengattini, P. Pellegrino, N. Daldosso, L. Pavesi, J.-P. Colonna, J.-M. Fedeli, and B. Garrido, “Effect of the annealing treatments on the electroluminescence efficiency of SiO2 layers doped with Si and Er,” J. Phys. D Appl. Phys. 45(4), 045103 (2012). [CrossRef]  

10. S. Cueff, J. Manel Ramírez, J. A. Kurvits, Y. Berencén, R. Zia, B. Garrido, R. Rizk, and C. Labbé, “Electroluminescence efficiencies of erbium in silicon-based hosts,” Appl. Phys. Lett. 103(19), 191109 (2013). [CrossRef]  

11. A. Anopchenko, A. Tengattini, A. Marconi, N. Prtljaga, J. M. Ramírez, O. Jambois, Y. Berencén, D. Navarro-Urrios, B. Garrido, F. Milesi, J.-P. Colonna, J.-M. Fedeli, and L. Pavesi, “Bipolar pulsed excitation of erbium-doped nanosilicon light emitting diodes,” J. Appl. Phys. 111(6), 063102 (2012). [CrossRef]  

12. J. M. Ramíırez, F. Ferrarese Lupi, O. Jambois, Y. Berencén, D. Navarro-Urrios, A. Anopchenko, A. Marconi, N. Prtljaga, A. Tengattini, L. Pavesi, J. P. Colonna, J. M. Fedeli, and B. Garrido, “Erbium emission in MOS light emitting devices: from energy transfer to direct impact excitation,” Nanotechnology 23(12), 125203 (2012). [CrossRef]   [PubMed]  

13. O. Savchyn, F. R. Ruhge, P. G. Kik, R. M. Todi, K. R. Coffey, H. Nukala, and H. Heinrich, “Luminescence-center-mediated excitation as the dominant Er sensitization mechanism in Er-doped silicon-rich SiO2 films,” Phys. Rev. B 76(19), 195419 (2007). [CrossRef]  

14. L. Jin, D. Li, L. Xiang, F. Wang, D. Yang, and D. Que, “Energy transfer from luminescent centers to Er3+ in erbium-doped silicon-rich oxide films,” Nanoscale Res. Lett. 8(1), 366 (2013). [CrossRef]   [PubMed]  

15. S. Yerci, R. Li, S. O. Kucheyev, T. van Buuren, S. N. Basu, and L. Dal Negro, “Energy transfer and 1.54 μm emission in amorphous silicon nitride films,” Appl. Phys. Lett. 95(3), 031107 (2009). [CrossRef]  

16. M. Wojdak, M. Klik, M. Forcales, O. B. Gusev, T. Gregorkiewicz, D. Pacifici, G. Franzò, F. Priolo, and F. Iacona, “Sensitization of Er luminescence by Si nanoclusters,” Phys. Rev. B 69(23), 233315 (2004). [CrossRef]  

17. C. J. Oton, W. H. Loh, and A. J. Kenyon, “Er3+ excited state absorption and the low fraction of nanocluster-excitable Er3+ in SiOx,” Appl. Phys. Lett. 89(3), 031116 (2006). [CrossRef]  

18. B. Garrido, C. García, P. Pellegrino, D. Navarro-Urrios, N. Daldosso, L. Pavesi, F. Gourbilleau, and R. Rizk, “Distance dependent interaction as the limiting factor for Si nanocluster to Er energy transfer in silica,” Appl. Phys. Lett. 89(16), 163103 (2006). [CrossRef]  

19. B. Garrido, C. García, S. Y. Seo, P. Pellegrino, D. Navarro-Urrios, N. Daldosso, L. Pavesi, F. Gourbilleau, and R. Rizk, “Excitable Er fraction and quenching phenomena in Er-doped SiO2 layers containing Si nanoclusters,” Phys. Rev. B 76(24), 245308 (2007). [CrossRef]  

20. F. Lenz, A. Hryciw, R. DeCorby, and A. Meldrum, “Reversing the temperature dependence of the sensitized Er3+ luminescence intensity,” Appl. Phys. Lett. 95(9), 091909 (2009). [CrossRef]  

21. F. Rebib, E. Tomasella, M. Dubois, J. Cellier, T. Sauvage, and M. Jacquet, “Structural and optical investigations of SiOxNy thin films deposited by R.F. sputtering,” Surf. Coat. Tech. 200(1–4), 330–333 (2005). [CrossRef]  

22. K. Hijazi, R. Rizk, J. Cardin, L. Khomenkova, and F. Gourbilleau, “Towards an optimum coupling between Er ions and Si-based sensitizers for integrated active photonics,” J. Appl. Phys. 106(2), 024311 (2009). [CrossRef]  

23. E. Steveler, H. Rinnert, X. Devaux, M. Dossot, and M. Vergnat, “Indirect excitation of Er3+ ions in silicon nitride films prepared by reactive evaporation,” Appl. Phys. Lett. 97(22), 221902 (2010). [CrossRef]  

24. N. Prtljaga, D. Navarro-Urrios, A. Tengattini, A. Anopchenko, J. M. Ramírez, J. M. Rebled, S. Estradé, J. P. Colonna, J. M. Fedeli, B. Garrido, and L. Pavesi, “Limit to the erbium ions emission in silicon-rich oxide films by erbium ion clustering,” Opt. Mater. Express 2(9), 1278–1285 (2012). [CrossRef]  

25. O. Savchyn, P. G. Kik, R. M. Todi, and K. R. Coffey, “Effect of hydrogen passivation on luminescence-center-mediated Er excitation in Si-rich SiO2 with and without Si nanocrystals,” Phys. Rev. B 77(20), 205438 (2008). [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 (2)

Fig. 1
Fig. 1 (a) Matrix-related PL spectra from Er:SRON films annealed at temperatures in the range of 600-1100 °C. The inset shows the cross sectional TEM image from the specimen annealed at 1100 °C. (b) Time resolved PL spectra recorded at 700 nm from the specimen annealed at 800 °C. The excitation wavelength is 325 nm.
Fig. 2
Fig. 2 (a) PL excitation spectra of Er3+ measured at 1540 nm. The PLE spectra measured at 750 nm for the 1100 °C annealed specimen is denoted by empty squares.(b) Er-related PL spectra of samples annealed at different temperatures for excitation at 325 nm. (c) Er3+ PL decay time τdec as a function of the annealing temperatures. The inset shows the evolution of integrated Er3+ PL intensity with temperatures in the 600 °C annealed specimen. (d) IPL/τdec as a function of the annealing temperatures for excitation at 325 nm and 980 nm (right hand scale).

Equations (2)

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

IPLφσErNEr,sen τdec τrad
NEr,sen τrad φσEr IPL τdec
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.