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

Cumulative defocusing of sub-MHz-rate femtosecond-laser pulses in bulk diamond envisioned by transient A-band photoluminescence

Open Access Open Access

Abstract

High-rate direct femtosecond (fs) laser writing is a well-established technology for fabricating various micro-optical elements in bulk dielectrics. In this technology, the “heat accumulation” effect, occurring during high-repetition rate (∼ 1 MHz) exposure in dielectrics by a fs laser, enables ultralow-energy micro-modification via cumulative heating. Meanwhile, in this work in the transient multi-photon A-band photoluminescence studies, we demonstrate that this effect underlies dynamic thermal lensing even in diamond with its high thermal conductivity, dynamically shifting the laser focus upstream. Our study paves the way for more precise, accurate and robust direct fs-laser writing of advanced three-dimensional structures in diamond and other dielectrics for a variety of photonic applications.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Direct femtosecond laser writing is well-established as a key-enabling, versatile technology for fabricating micro-optical and micro-fluidic elements in bulk dielectrics [13]. Depending on laser fluence and exposure, related material micro-modification regimes and the resulting embedded microstructures vary from 1) nanoscale plasmonic chemical segregation, yielding in transparent bulk birefringent nanogratings [4]; 2) homogeneous melting/recrystallization, producing densified regions of phase-contrast waveguides without material disruption [5]; till 3) local ablation and lateral densification, making hollow voids and channels for optical storage [6] and microfluidics [7], respectively.

Generally, minimal heat-affected zone (HAZ) around fs-laser focus in dielectrics, emerging due to their low thermal conductivity (∼ 0.01 W/cmK [8]), was considered as the main prerequisite for high-precision and resolution in direct fs-laser writing [9], and also as the main reason for marginal energy transport off the focal region. Practically, some time ago one important regime of oscillator-only direct laser writing in bulk dielectrics was found [10], supporting their ultralow-energy modification via multi-shot cumulative heating at multi-MHz repetition rates [10,11]. Due to such multiple cumulative heating, it has been possible to date to modify optically-active point defects in diamond [1214]. Moreover, waveguides were fabricated in a few studies in bulk silica materials in this regime, exploring the repetition-rate effect in the MHz-range [1011, 1517].

Meanwhile, the related issues of inhomogeneous temperature distribution during cumulative fs-laser heating and related thermal lensing effect were not considered so far, comparing to, e.g., local modification of refractive index via generation of self-trapped excitons [18]. Furthermore, potentially these thermal cumulative effects could strongly emerge in crystalline dielectrics as anisotropic wave-front distortions (“thermal lensing”), thermal birefringence etc., which could transiently shift or modify the laser-heating source and thus are highly important for well-controlled and precise direct fs-laser writing in bulk dielectrics.

In this study we observe dynamic displacement of fs-laser focus in bulk natural diamond, the highly-transparent material with very high thermal conductivity, under its sub-MHz exposure on a millisecond timescale, which is visualized by transient laser-excited photoluminescence and appears to be reversible upon switching the laser off for a few milliseconds. This observation rises a question of robust spatial focusing of high-rate fs-laser pulses in bulk dielectrics and direct writing of continuous buried micro-structures while changing writing regime.

2. Experimental details

In this experiment, 1030-nm, 300-fs fiber laser pulses, coming at 100-kHz repetition rate and variable pulse energies E = 4-63 nJ (peak powers of 13-140 kW, sub-filamentary focusing regime for 1030-nm peak laser-pulse powers < 2 MW [19]), were focused at each pulse energy into a new 3-micron wide (1/e-level diameter 2w) and 25-micron long (1/e-level Rayleigh length 2l) region inside a diamond (refractive index n0 (λ = 1030 nm) ≈ 2.4 [20]) by 0.25-NA micro-objective (10×) (Fig. 1(a))

$$w = \frac{\lambda }{{\pi {n_0}}}\frac{{\sqrt {n_0^2 - N{A^2}} }}{{NA}},l = \frac{\lambda }{{\pi {n_0}}}\frac{{n_0^2 - N{A^2}}}{{N{A^2}}}$$
In the focal region, a bluish ellipsoidal emission emerged in each pulse and was visualized at the right angle by a 0.2-NA quartz/fluorite LOMO micro-objective (10×), with the optical signal acquired through a flip-mirror by a broadband (190-1100 nm) spectrometer ASP-150F (Avesta) or color CCD camera (calibration scale ≈ 1 µm/pixel). The natural monocrystalline diamond sample was IR-spectrally characterized nitrogen-doped (>700 ppm) IaAB-type crystal with the 650 ppm of A-centers (substitutional N-dimers, 2N), 50 ppm of B1-centers (4NV) and very minor content of B2-centers (platelets).

 figure: Fig. 1.

Fig. 1. (a) Schematic of the experimental setup. (b) Photoluminescence spectra at variable fs-laser pulse energies.

Download Full Size | PDF

At the maximal 500-kHz exposures and reduced pulse energy of 35 nJ, side-view imaging of the bluish emission region, spatially evolving far beyond the waist range 2l, was performed by the color CCD camera as a function of millisecond-scale exposure times, interrupted from time to time to explore the reversibility of its sub-millimeter upstream (back to the laser) displacements.

3. Results

3.1 PL spectral assignment and pulse energy yield

The observed photoluminescence spectra in Fig. 1(b) demonstrate a main structureless emission A-band [21], starting at ≈ 350 nm and extending till ≈ 700 nm, with its intensity peak at ≈ 430 nm monotonously raising versus the increasing pulse energy of 4-42 nJ (peak intensity of 0.2-1.8 TW/cm2). Indeed, the uv-vis PL origin can be rather unambiguously identified as A-band with its peak at 415-445 nm [21,22], which is especially very intense in low-nitrogen diamonds because of its strong nitrogen A-center damping [23,24]. The corresponding PL excitation band is known to be symmetric to A-band, occupying the range of 3-4 eV [25]. Importantly, A-bands may have different origins [21], where in the first model the relatively narrow A-band peaked at 440 nm in low-nitrogen type II diamonds is associated with radiative recombination at dislocations (donor-acceptor (D-A) pairs decorating dislocations [22], vacancies bound to dislocations [26,27], dislocations decorated with D-A pairs [28], on pure non-decorated dislocations [2932]). In the second model, A-band with a maximum at 480 nm observed in natural type I diamonds is intra-center transitions at the B2-centers (platelets). In the third model A-band is also assigned to electron-hole recombination at deep centers, the energy levels of which lie in the middle of the bandgap [33,34]: EA = (Eg/2) ± 0.75 eV, occurring through the formation of free excitons [35] as a two-stage process [36,37]. In our study, the observed A-band extends in the range of 2-4 eV with its peak at ≈ 3 eV at the presence of high (650 ppm) A-center density and marginal B2-center density in the high-nitrogen natural diamond, representing, according to the diamond specification and our recent spectral studies [38], the radiative relaxation in the intra-gap energy set of molecular-like states of the unknown optical center. Additionally, at shorter wavelengths, there is a preceding periodical vibronic progression with the diamond LO/TO-phonon wavenumber of ≈ 1300 cm-1. Meanwhile, at higher pulse energies of 50 and 63 nJ (peak intensity ≈ 2-2.5 TW/cm2) an additional broad photoluminescence band appears just above the second-harmonic position (515 nm) in the spectra (Fig. 2(a)), slightly reducing the main band intensity.

 figure: Fig. 2.

Fig. 2. (a) Laser energy/intensity dependence of PL peak intensity and its dual-range linear fits in lgΦ-lgE coordinates. (b) Schematic spectral structure, explaining the dual-band photoluminescence spectra and dual-slope fitting in (a). The upright red arrows indicate the 1030-nm laser photons, blue and green arrows indicate the corresponding photoluminescence bands, and other arrows indicate radiation and non-radiative (dashed) relaxation processes.

Download Full Size | PDF

Moreover, the corresponding PL yield versus laser energy/intensity was observed to change its slope (power) in the lgΦ-lgE coordinates from 2.9 ± 0.5 for E ≤ 10 nJ (I ≤ 0.4 TW/cm2) till 1.2 ± 0.1 for higher energies (intensities) (Fig. 2(b)). The low-energy three-photon process can be directly related to the intra-center excitation of the top excited state, radiatively relaxing via the “blue” photoluminescence (Fig. 1(b)). Then, in together with the high-energy appearance of the second, “green” luminescence band in Fig. 1(b), the second, high-energy linear “blue” PL yield could be associated with radiative/non-radiative relaxation of the top excited state into the intermediate one (Fig. 2(a)), yielding further either radiative “green” relaxation, or one-photon laser re-excitation back to the top excited state. In this way, both the linear high-energy PL yield and the high-energy appearance of the “green” PL band can be tied together. This suggestion of the intermediate excited state of this poorly explored A-band-related optical center is consistent with our previous temperature-dependent studies of 515-nm fs-laser excited A-band photoluminescence, showed enhanced damping even at minor heating (<55 °C) [38], as typical for A-band [39,40]. Specifically, 1) thermal activation of dissociation in the excited state, thermal overcoming of an activation barrier from the excited state 2) to the ground state, or 3) to the intermediate state are considered as the main mechanisms of thermal damping of photoluminescence, and the latter one probably takes the place in the optical center of the diamond, similarly to, e.g., NV-center in this material [41].

3.2 Reversible spatial displacements of the PL region during ms-scale cumulative exposures

Using the synchronized color CCD-camera, side-view imaging of the 1030-nm fs-laser excited emission region was performed as a function of ms-scale exposure time (Fig. 3(a)). Specifically, its upstream (back to the laser) monotonic multi-micrometer displacement was observed. Moreover, at the later time instants the main peak drops in its intensity and splits off into two low-intensity peaks (Fig. 3(b)), which could be related to 1) intra-pulse dynamics, including the spatiotemporal transition in the abovementioned three/one-photon A-band photoexcitation mechanism, 2) longitudinal instabilities in PL yield during the fs-laser exposure, etc.

 figure: Fig. 3.

Fig. 3. (a) Upstream evolution of the focal region versus exposure time. (b) Longitudinal profiles of emission intensity at different time instants, showing the double-peak structure separated by the vertical dashed line. (с) Enlarged image of PL in 2.5x with deferred parameters half-lengths and half-widths.

Download Full Size | PDF

More exactly, the first peak shifts downstream and completely disappears in some point, giving rise to the second peak (Fig. 4(a)). Its simultaneous drop in intensity is more pronounced over the first 10-15 ms (Fig. 4(b)). The PL intensity center-of-mass in the emission regime decreases very rapidly in time from one emission region to another one (Fig. 4(c)). Finally, despite the visible elongation of the emission regions in time, their lengths at half-maximum (half-lengths, Fig. 4(d)) drop twice rather quickly versus time, similarly to the corresponding half-widths.

 figure: Fig. 4.

Fig. 4. Spatial (a) and temporal (b) evolution of the first and second peak intensity in longitudinal profiles (Fig. 3(b)). (c) Velocity of the first and second peak center-of-mass displacements versus exposure time. (d) Half-lengths and half-widths of emission regions (Fig. 3(a)) versus exposure time.

Download Full Size | PDF

The electromechanical shutter was used to explore the reversibility of the downstream displacement versus the interruption time in laser exposure (Fig. 5). First, Fig.5a indicates the almost complete recovery (upstream displacement) of the emission regime after 12 ms, with the emission region emerging instantaneously again in the same initial position of the focus (Z = 110 µm) upon the start of a new exposure. The next Fig.5b shows that the intermediate velocities of the downstream and upstream displacements are almost identical, as well as the corresponding transient positions (Fig.5c).

 figure: Fig. 5.

Fig. 5. (a) Reverse upstream displacements (recovery of the initial position) of emission region versus interruption time in laser exposure. The horizontal dashed lines indicate the starting point (Z = 110 µm) and the maximal displacement (Z = 210 µm) during the laser heating. (b) Velocity of down- and upstream displacements versus exposure time (“pumping”) and interruption (“relaxation”) time in laser exposure. (c) Downstream and upstream displacements of emission region (center-of mass) during the first 12 ms of exposure time (“pumping”) and interruption (“relaxation”) time in laser exposure.

Download Full Size | PDF

4. Discussion

Though previously only local heat accumulation was observed in dielectrics during high-rate fs-laser exposures [1011,1517], in this work we observed reversible displacement of the PL-tracked fs-laser focal region, which could be related to local modification of the refractive index, Δn, along the laser path in the diamond, resulting in its earlier focusing, i.e., self-focusing. Usually, prompt electronic or delayed picosecond-scale orientational self-focusing mechanisms with the nonlinear index n2 are taken into account [18]

$$n \approx {n_0} + {n_2}I + \frac{{{N_{STE}}{e^2}}}{{2{n_0}m{\varepsilon _0}}}\frac{1}{{\omega _{tr}^2 - {\omega ^2}}}$$
which have ultrafast, (sub)picosecond relaxation times, comparing to self-trapped excitons (STE) and other atomistic damage features in dielectrics, contributing to refractive index as follows in Eq.(2) [18] and metastable over much longer times – up to millisecond scales [42]. Hence, similarly to the previous studies [1011,1517] and despite the high room-temperature thermal conductivity ∼10-20 W/cmK in diamond comparing to ∼10−2 W/cmK in glasses [8]), one can alternatively consider cumulative and reversible slow process of heating in the focal region in the diamond at the ∼1-MHz repetition rate [1011,1517], resulting in diamond in the local pre-focus raise of the refractive index according to the thermo-optical coefficient value (dn/dT)/n ≈ 4 × 10 –6 K-1 > 0 at room and higher temperatures [20,21]) and self-focusing of the succeeding laser pulses in the transient lens [43]. Moreover, the gradual damping of the PL intensity is also consistent with the known thermal damping of PL intensity [38].

5. Conclusions

Though cumulative heating effect, occurring during MHz-rate direct femtosecond laser modification in dielectrics, is rather well known, dynamic lensing effect during such multi-shot exposures was revealed in our study for the first time. Such reversible upstream displacements of the focal region by almost ten Rayleigh lengths reveals the important details of non-linear “laser-matter” optical interaction and could strongly affect direct fs-laser local nano- and micro-fabrication of various micro-optical elements in bulk dielectrics.

Funding

Russian Science Foundation (21-79-30063).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are available from the authors upon reasonable request.

References

1. K. Sugioka and Y. Cheng, “Ultrafast lasers—reliable tools for advanced materials processing,” Light Sci. Appl 3(4), e149 (2014). [CrossRef]  

2. M. Malinauskas, A. Žukauskas, S. Hasegawa, Y. Hayasaki, V. Mizeikis, R. Buividas, and S. Juodkazis, “Ultrafast laser processing of materials: from science to industry,” Light Sci. Appl 5(8), e16133 (2016). [CrossRef]  

3. M. Krečmarová, M. Gulka, T. Vandenryt, J. Hrubý, L. Fekete, P. Hubík, A. Taylor, V. Mortet, R. Thoelen, E. Bourgeois, and M. Nesládek, “A label-free diamond microfluidic DNA sensor based on active nitrogen-vacancy center charge state control,” ACS Appl. Mater. Interfaces (2021)

4. Y. Shimotsuma, P. Kazansky, J. Qiu, and K. Hirao, “Self-organized nanogratings in glass irradiated by ultrashort light pulses,” Phys. Rev. Lett. 91(24), 247405 (2003). [CrossRef]  

5. S. Eaton, H. Zhang, P. Herman, F. Yoshino, L. Shah, J. Bovatsek, and A. Arai, “Heat accumulation effects in femtosecond laser-written waveguides with variable repetition rate,” Opt. Express 13(12), 4708–4716 (2005). [CrossRef]  

6. M. Hong, B. Luk’Yanchuk, S. Huang, T. Ong, L. Van, and T. Chong, “Femtosecond laser application for high capacity optical data storage,” Appl. Phys. A 79(4-6), 791–794 (2004). [CrossRef]  

7. D. Hwang, T. Choi, and C. Grigoropoulos, “Liquid-assisted femtosecond laser drilling of straight and three-dimensional microchannels in glass,” Appl. Phys. A 79(3), 605–612 (2004). [CrossRef]  

8. I. S. Grigor’ev and E. Z. Meylikhov, Physical Quantities (Energoatomizdat, 1991).

9. R. Gattass and E. Mazur, “Femtosecond laser micromachining in transparent materials,” Nat. Photonics 2(4), 219–225 (2008). [CrossRef]  

10. R. Gattass, L. Cerami, and E. Mazur, “Micromachining of bulk glass with bursts of femtosecond laser pulses at variable repetition rates,” Opt. Express 14(12), 5279–5284 (2006). [CrossRef]  

11. S. Eaton, H. Zhang, M. Ng, J. Li, W. Chen, S. Ho, and P. Herman, “Transition from thermal diffusion to heat accumulation in high repetition rate femtosecond laser writing of buried optical waveguides,” Opt. Express 16(13), 9443–9458 (2008). [CrossRef]  

12. Y. Liu, G. Chen, M. Song, X. Ci, B. Wu, E Wu, and H. Zeng, “Fabrication of nitrogen vacancy color centers by femtosecond pulse laser illumination,” Opt. Express 21(10), 12843–12848 (2013). [CrossRef]  

13. Y. C. Chen, P. S. Salter, S. Knauer, L. Weng, A. C. Frangeskou, C. J. Stephen, S. N. Ishmael, P. R. Dolan, S. Johnson, B. L. Green, G. W. Morley, M. E. Newton, J. G. Rarity, M. J. Booth, and J. M. Smith, “Laser writing of coherent colour centres in diamond,” Nat. Photonics 11(2), 77–80 (2017). [CrossRef]  

14. Y. Chen, B. Griffiths, L. Weng, S. Nicley, S. N. Ishmael, Y. Lekhai, S. Johnson, C. J. Stephen, B. L. Green, G. W. Morley, M. E. Newton, M. J. Booth, P. S. Salter, and J. M. Smith, “Laser writing of individual nitrogen-vacancy defects in diamond with near-unity yield,” Optica 6(5), 662–667 (2019). [CrossRef]  

15. T. Tamaki, W. Watanabe, and K. Itoh, “Laser micro-welding of transparent materials by a localized heat accumulation effect using a femtosecond fiber laser at 1558 nm,” Opt. Express 14(22), 10460–10468 (2006). [CrossRef]  

16. T. Baldacchini, S. Snider, and R. Zadoyan, “Two-photon polymerization with variable repetition rate bursts of femtosecond laser pulses,” Opt. Express 20(28), 29890–29899 (2012). [CrossRef]  

17. R. Weber, T. Graf, P. Berger, V. Onuseit, M. Wiedenmann, C. Freitag, and A. Feuer, “Heat accumulation during pulsed laser materials processing,” Opt. Express 22(9), 11312–11324 (2014). [CrossRef]  

18. S. Mao, F. Quéré, S. Guizard, X. Mao, R. Russo, G. Petite, and P. Martin, “Dynamics of femtosecond laser interactions with dielectrics,” Appl Phys A 79(7), 1695–1709 (2004). [CrossRef]  

19. S. Kudryashov, A. Levchenko, P. Danilov, N. Smirnov, and A. Ionin, “IR femtosecond laser micro-filaments in diamond visualized by inter-band UV photoluminescence,” Opt. Lett. 45(7), 2026–2029 (2020). [CrossRef]  

20. E. D. Palik, Handbook of Optical Constants of Solids (Academic press, 1998).

21. A. M. Zaitsev, Optical Properties of Diamond: A Data Handbook (Springer Science & Business Media, 2013).

22. N. Yamamoto, J. C. H. Spence, and D. Fathy, “Cathodoluminescence and polarization studies from individual dislocations in diamond,” Philos. Mag. B 49(6), 609–629 (1984). [CrossRef]  

23. Y. Yokota, H. Kotsuka, T. Sogi, J. S. Ma, A. Hiraki, H. Kawarada, K. Matsuda, and M. Hatada, “Formation of optical centers in CVD diamond by electron and neutron irradiation,” Diamond and Related Materials 1(5-6), 470–477 (1992). [CrossRef]  

24. S. C. Lawson, H. Kanda, K. Era, and Y. Sato, “A study of broad band cathodoluminescence from boron-doped high-pressure synthetic and chemical vapor deposited diamond,” (personal communication, 1994).

25. K. Iakoubovskii and G. J. Adriaenssens, “Luminescence excitation spectra in diamond,” Phys. Rev. B 61(15), 10174–10182 (2000). [CrossRef]  

26. J. F. Prins, “Ion-implanted n-type diamond: optical evidence,” in Int. Conf. Diamond Films (1995).

27. J. F. Prins, “Increased band A cathodoluminescence after carbon ion implantation and annealing of diamond,” Diamond and Related Materials 5(9), 907–913 (1996). [CrossRef]  

28. J. Bruley and P. E. Batson, “Electronic structure near individual dislocations in diamond by energy-loss spectroscopy,” Mat. Res. Soc. Symp. Proc. 162, 255 (1989). [CrossRef]  

29. J. Ruan, K. Kobashi, and W. J. Choyke, “On the “band-A” emission and boron related luminescence in diamond,” Appl. Phys. Lett. 60(25), 3138–3140 (1992). [CrossRef]  

30. P. J Dean., and J. C. Male, “Luminescence and birefringence in a semiconducting diamond,” Br. J. Appl. Phys. 15(1), 101–102 (1964). [CrossRef]  

31. E. V. Sobolev and Yu. I. Dubov, “On the nature of X-ray luminescence,” Fiz. Tverd. Tela 17, 1142–1148 (1975).

32. E. V. Sobolev and Yu. I. Dubov, “On some features of the X-ray luminescence of natural and synthetic diamonds,” Synthetic Diamonds 2, 3–11 (1979).

33. R. Heiderhoff, “Analyse der elektronischen Eigenschaften von Diamant mittels rastermikroskopischer Methoden,” PhD thesis, University of Wuppertal (1997) [in German].

34. C. Manfredotti, F. Wang, P. Polesello, E. Vittone, and F. Fizzotti, “Blue-violet electroluminescence and photocurrent spectra from polycrystalline chemical vapor deposited diamond film,” Appl. Phys. Lett. 67(23), 3376–3378 (1995). [CrossRef]  

35. H. Kawarada and A. Yamaguchi, “Excitonic recombination radiation as characterization of diamonds using cathodoluminescence,” Diamond and Related Materials 2(2-4), 100–105 (1993). [CrossRef]  

36. E. F. Martynovich, L. V. Morozhnikova, Yu. A. Klyuev, and S. P. Plotnikova, “X-ray luminescence in various kinds of natural diamonds,” in: Problems of Theory and Practice of Diamond Treatment (NIIMASH, Moscow, 1977) [in Russian].

37. A.V. Kurdumov, V.G. Malogolovets, N.V. Novikov, A.H. Piljankevich, and L.A. Shulman, Polymorphous Modification of Carbon and Boron Nitride (Metallurgija, 1994), p.318.

38. S. Kudryashov, P. Danilov, N. Smirnov, A. Levchenko, M. Kovalev, Y. Gulina, O. Kovalchuk, and A. Ionin, “Femtosecond-laser excited luminescence of A-band in diamond and its thermal control,” Opt. Mater. Express (to be published).

39. T. Kurita, Y. Shimotsuma, M. Fujiwara, M. Fujie, N. Mizuochi, M. Shimizu, and K. Miura, “Direct writing of high-density nitrogenvacancy centers inside diamond by femtosecond laser irradiation,” Appl. Phys. Lett. 118(21), 214001 (2021). [CrossRef]  

40. B. Griths, A. Kirkpatrick, S. Nicley, R. Patel, J. Zajac, G. Morley, M. J. Booth, P. Salter, and J. Smith, “Microscopic processes during ultra-fast laser generation of Frenkel defects in diamond,” (to be published).

41. I. Fedotov and A. Zheltikov, “Background-free two-photon fluorescence readout via a three-photon charge-state modulation of nitrogen-vacancy centers in diamond,” Opt. Lett. 44(15), 3737–3740 (2019). [CrossRef]  

42. K.S. Song and R.T. Williams, Self-trapped Excitions, 105, Springer series in solid-state sciences (Springer, 1993).

43. M. Sakakura, M. Terazima, Y. Shimotsuma, K. Miura, and K. Hirao, “Heating and rapid cooling of bulk glass after photoexcitation by a focused femtosecond laser pulse,” Opt. Express 15(25), 16800–16807 (2007). [CrossRef]  

Data availability

Data underlying the results presented in this paper are available from the authors upon reasonable request.

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 (5)

Fig. 1.
Fig. 1. (a) Schematic of the experimental setup. (b) Photoluminescence spectra at variable fs-laser pulse energies.
Fig. 2.
Fig. 2. (a) Laser energy/intensity dependence of PL peak intensity and its dual-range linear fits in lgΦ-lgE coordinates. (b) Schematic spectral structure, explaining the dual-band photoluminescence spectra and dual-slope fitting in (a). The upright red arrows indicate the 1030-nm laser photons, blue and green arrows indicate the corresponding photoluminescence bands, and other arrows indicate radiation and non-radiative (dashed) relaxation processes.
Fig. 3.
Fig. 3. (a) Upstream evolution of the focal region versus exposure time. (b) Longitudinal profiles of emission intensity at different time instants, showing the double-peak structure separated by the vertical dashed line. (с) Enlarged image of PL in 2.5x with deferred parameters half-lengths and half-widths.
Fig. 4.
Fig. 4. Spatial (a) and temporal (b) evolution of the first and second peak intensity in longitudinal profiles (Fig. 3(b)). (c) Velocity of the first and second peak center-of-mass displacements versus exposure time. (d) Half-lengths and half-widths of emission regions (Fig. 3(a)) versus exposure time.
Fig. 5.
Fig. 5. (a) Reverse upstream displacements (recovery of the initial position) of emission region versus interruption time in laser exposure. The horizontal dashed lines indicate the starting point (Z = 110 µm) and the maximal displacement (Z = 210 µm) during the laser heating. (b) Velocity of down- and upstream displacements versus exposure time (“pumping”) and interruption (“relaxation”) time in laser exposure. (c) Downstream and upstream displacements of emission region (center-of mass) during the first 12 ms of exposure time (“pumping”) and interruption (“relaxation”) time in laser exposure.

Equations (2)

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

w = λ π n 0 n 0 2 N A 2 N A , l = λ π n 0 n 0 2 N A 2 N A 2
n n 0 + n 2 I + N S T E e 2 2 n 0 m ε 0 1 ω t r 2 ω 2
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.