Next Article in Journal
Photonics Large-Survey Telescope Internal Motion Metrology System
Next Article in Special Issue
Calcination-Enhanced Laser-Induced Damage Threshold of 3D Micro-Optics Made with Laser Multi-Photon Lithography
Previous Article in Journal
Numerical Simulation of a Light Field Structure in an Integrating Sphere via the Monte Carlo Method
Previous Article in Special Issue
A Deep Learning Framework to Remove the Off-Focused Voxels from the 3D Photons Starved Depth Images
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Crystalline Flat Surface Recovered by High-Temperature Annealing after Laser Ablation

1
Optical Sciences Centre (OSC) and ARC Training Centre in Surface Engineering for Advanced Materials (SEAM), School of Science, Computing and Engineering Technologies, Swinburne University of Technology, Hawthorn, VIC 3122, Australia
2
Research Institute of Electronics, Shizuoka University, Johoku 3-5-1, Hamamatsu 432-8011, Japan
3
World Research Hub Initiative (WRHI), School of Materials and Chemical Technology, Tokyo Institute of Technology, 2-12-1, Ookayama, Meguro-ku, Tokyo 152-8550, Japan
*
Authors to whom correspondence should be addressed.
Photonics 2023, 10(5), 594; https://doi.org/10.3390/photonics10050594
Submission received: 24 March 2023 / Revised: 14 May 2023 / Accepted: 15 May 2023 / Published: 19 May 2023
(This article belongs to the Special Issue Research in Computational Optics)

Abstract

:
Ultra-short laser pulses (1030 nm/230 fs) were used to laser ablate the surface of crystalline sapphire (Al2O3) at high intensity per pulse 20–200 TW/cm 2 /pulse. Laser-ablated patterns were annealed at a high temperature of 1500  ° C. Surface reconstruction took place, removing the ablation debris field at the edges of ablated pits in oxygen flow (O2 flow). Partial reconstruction of ripples was also observed when multi-pulse ablated surfaces were annealed at high temperature in O2 flow. Back-side ablation of a 0.5-mm-thick Al2O3 produced high surface roughness ∼ 1 μ m which was reduced to ∼ 0.2 μ m by high-temperature annealing at 1500  ° C for 2 h in O2. Improvement of surface quality was due to restructuring of the crystalline surface and sublimation, while the defined 3D shape of a micro-lens was not altered after HTA (no thermal morphing).

1. Introduction: High Intensity > TW/cm 2  Crystal Processing

Matter at extreme conditions of pressures above 1 megabar ( 10 11  Pa) and high temperatures as in the interior of planets—the warm dense matter (WDM)—is an active field of research [1]. High-pressure and high-temperature phases of materials can be created by ultra-fast thermal quenching of a small modified volume inside transparent material, e.g., sapphire [2,3]. Internally confined micro-explosions create WDM conditions similar to the centre of the Earth. For point-like energy deposition, it is informative to present Sedov’s solution, where the radius, R, of the shock wave is increasing in time t as [4,5]:
R ( t ) = E p K ( γ ) ρ 0 1 / ( 2 + i ) t 2 / ( 2 + i ) ,
where  E p  is the instantaneous (in space and time) release of energy,  K ( γ ) = ( 0.6 0.8 )  is the constant dependent on the adiabatic coefficient  γ = C p / C v  ( γ = 1.4  for air),  ρ 0  is the mass density in an unperturbed medium, and i is the dimensionality coefficient  i = 3  for spherical and  i = 2  for cylindrical explosion; here,  C p , v  is the heat capacities at constant pressure and volume, respectively. Such a shock wave launched from the focal volume creates highly dynamic conditions of high pressures and temperatures, which lead to formation of new materials [2,3]. In addition, metastable amorphous phases can be produced as shown for sapphire [6] due to ultra-fast thermal quenching of small volumes modified by fs-laser pulse-triggered explosions. Spatially confined, energy deposition of ∼40 fs onto a water surface triggers shock waves inside the water and the surrounding air which causes emissions of X-rays and THz [7]. This can be a single optical cycle (∼1–3 ps) due to the point-like energy deposition in time and space.
The micro-explosion hydrodynamics was shown to follow the established and tested macroscopic versions of strong explosions [8]. Evolution of the micro-explosion is characterized by the independent parameters: energy E (or pressure P), density  ρ , and radius r, while the two other parameters required for the full description of the hydrodynamic processes, the velocity and time, can be expressed as  v = P / ρ E / ( r 3 ρ )  and  t = r 5 / 2 ρ / E , respectively. It is shown that micro-explosions triggered by tightly focused fs-laser pulses are self-similar with macro-explosions made by a  10 21  large energy deposition when the time t and space r are  10 7  times larger [8]. All hydrodynamic phenomena with the same initial pressure and density but differing in space and time scales are described by the same hydrodynamic equations. In astrophysics, these same equations are used to describe the propagation of shock waves in self-gravitating perfect gases for isothermal flows [9], studying the dynamics of the envelope in supernova remnants [10,11], and the expansion of supernova shells [12].
Strong structural modifications induced by high-intensity ultra-short laser pulses could be modified by a post-processing steps of high temperature annealing (HTA) as recently demonstrated for micro-optical elements laser ablated out of crystalline sapphire Al2O3 [13]. HTA is a commonly used process for changing the surface morphology of different materials [14,15,16,17]. It has been used to make the particle size of high-temperature alloy thin films smaller [14], to create ceramic devices from polymers [17], and to smooth out germanium layers on silicon substrates [15] and terraces of chemically etched SrTiO 3 [16]. The roughness evolution of rapidly annealed surfaces is well understood; as the annealing time increases, the roughness decreases linearly [18]. Germanium, silicon, and some alloys have lower melting temperatures, making them easier to control. Sapphire, on the other hand, has a high melting temperature at 2053  ° C requiring more specialised equipment to reach these temperatures. In this study, it is shown that using temperatures between 1400  ° C to 1600  ° C and different gaseous environments, it is possible to alter the surface morphology of crystalline sapphire. In glasses, the softening point, at which a material does not become liquid as it would at the melting point but becomes malleable [19] is used; however, in crystals, only thermally activated surface diffusion on an atomic scale as well as sublimation is occurring. Sublimation is the process by which a solid material turns to gas without passing through the liquid phase. In crystals, sublimation causes them to create step-like structures, [20] as has been shown with GaN(0001) [21], GaAs(001) [22,23], and Si(111) [24,25]. For each of these materials, both Monte Carlo simulation and experimental results showed agreement on the appearance of step-like structures and in the case of Si(111), it was possible to nano-texture the surface with the sublimation method. The theory here is to balance the removal of material through the sublimation process with the growth of the crystalline structure to create the step surface. Sapphire sublimation has been used before on sapphire spheres 1 mm in diameter with much longer annealing times of 7.5 h and 1.5 h and at temperatures of 1500  ° C and 1850  ° C, respectively, in H2 gas [26]. The vaporisation temperature for Al2O3 at a pressure of 1 atmosphere ranges from 2114.35  ° C to 2249.35  ° C [27]. The advantage of HTA of crystals is the preservation of large micro-scale structures and patterns, while reconstruction and smoothing of laser-treated regions takes place at nanoscale.
Here, we show the results of laser ablation of sapphire (Al2O3) and its high-temperature annealing for surface roughness reduction and reconstruction of surfaces with ablated ripples. The conditions to recover a crystalline smooth surface up to the rim of the ablated pits were found, which can be useful for applications to 2D materials. It is shown that back-side ablation of Al2O3 in air which produced roughness of ∼ 1 μ m was reduced to <200 nm after HTA at  1500 ° C for 2 h.

2. Samples and Methods

2.1. Sapphire

The PHAROS (Light Conversion Ltd., Vilnius, Lithuania) laser system was used to irradiate the surface of the Al2O3 on the front and the back-side. The Al2O3 samples have ∼0.5 mm thicknesses. Al2O3 is a hard and inert material that requires ∼10 TW/cm 2  intensities for ablation at near-IR 800–1000 nm wavelengths where multi-photon absorption is required to seed an efficient avalanche generation of free electrons for a wide bandgap ∼8–9 eV dielectric [28]. Among its virtues is high transparency in UV-to-IR spectral range (200–4000 nm) and high heat conductivity. This makes sapphire optics a promising target for industrial and space applications.

2.2. Laser Ablation Conditions

The PHAROS (Light Conversion Ltd.) laser system with  λ = 1030  nm wavelength,  t p = 230  fs pulse duration, focusing with objectives lenses with numerical apertures of  N A = 0.26  and  N A = 0.14 , resulted in focal spots calculated by Airy disk formula  2 r = 1.22 λ / N A  of ≈ 4.8 μ m and ≈ 8.9 μ m, respectively. The Rayleigh length is  z R = π r 2 / λ = 35.62 μ m and  122.6 μ m, which is half of the depth of focus (in air). For example, for the pulse energy of  E p = 1 μ J (on the sample), fluence was  F p = 5.45  J/cm 2  and average intensity  I p = 24  TW/cm 2 . Laser repetition rate was  f r = 200  kHz and linear density of pulses was varied for larger or smaller overlap (typically  10 2 10 3  pulses/mm), which affected the material removal rate. The laser writing velocity was up to 15 mm/s for a position sensitive output, which was important for the high-precision ablation of micro-lenses in a multi-pass scanning mode. In general, this is a 2D/3D sculpturing method that can reach sub-micron precision when it is performed in the light-induced back-side wet etching (LIBWE) mode [29]. Specific details of laser patterning, if different from those listed above, are shown where applicable.

2.3. High-Temperature Annealing

High-temperature annealing (HTA) was carried out using a tube furnace (Model RHTC, Nabeterhm, Germany). The recipe starts by ramping the temperature up to 1500  ° C at 150  ° C an hour. The next step is holding the temperature at 1500  ° C for 2 hours and afterwards ramping down to 1400  ° C at 200  ° C per hour. The final step is cooling down to 30  ° C at 100  ° C for an hour to finish the procedure. The extra step down to 1400  ° C before cooling to 30  ° C is used as a test for the limits of the tube furnace. This may be important if rapid quenching is required for further experimentation on the lenses. Currently, rates of up to 500  ° C per hour have been tested.

3. Results and Discussion

3.1. Roughness Reduction of Ablated Sapphire Surfaces

Front-side ablation was used with laser pulses traveling from air-into-Al2O3; back-side ablation occurs when the focus is placed onto the exit plane (Al2O3-to-air). This is usually possible with samples thinner than 0.5 mm and at less tight focusing with  N A < 0.4 ; see Section 3.3 for the back-side ablation and HTA. It took ∼30 min to ablate the profile of a 1-mm-diameter lens (Figure 1a) with a height of  h = 15 μ m. The surface roughness of the aspherical ablated surface was ∼ 200 ± 50  nm and it was reduced to  70 ± 30  nm after 2 h annealing (Figure 1c). However, light scattering was still strong and the surface remained visibly white. Initial roughness was also dependent on the focusing conditions and pulse overlap during ablation. Other smoothing strategies need to be tested, e.g., ablation with more focused pulses. When ablation took place at the very top of the lens profile (surface of sapphire), there was a strong change in ablation conditions, which resulted in a non-ablated plateau, which is ≈ 150 μ m in diameter (Figure 1). This is a known issue and can be circumvented by placing a micro-layer of a matching refractive index material, e.g., a silica film, that allows to ablate sapphire [30]; the added layer is later washed out in a solvent/acid.
The effect of HTA was examined through the ablation of crystalline flat c-plane Al2O3, which led to better understanding of surface changes. Since the maximum temperature used was  1500 ° C, which is much lower than the melting point of Al2O3 2050 ° C, surface restructuring by surface tension-driven flow of the molten phase was absent and other contributions from surface diffusion at atomic level, coarsening of debris nanoparticles, and sublimation took place. Sapphire has a melting point of 2053  ° C and boiling temperature of 2980  ° C. It has recently been shown that lower temperatures with larger times can also yield surface re-flow, which can be understood from length of diffusion scaling  D s × t , where  D s [cm 2 /s] is the surface diffusion coefficient and t is time. Temperatures of 1100  ° C and 1200  ° C at 3, 6, 9, 18, and 30 h (in air ambience) have been tested and showed smoothing of c-plane sapphire terraces in step configurations [31]. Laser-modified regions under high-intensity irradiation in air might also have surface formation of ceramic ALON phase (AlN) x · (Al2O3) 1 x  with x = 0.30–0.37 or AlN [32]. When sapphire was placed in N2 flow, the crystalline flat c-plane surface of Al2O3 changed to a nanotextured surface as shown by the SEM images in Figure A1a, and the entire piece of Al2O3 looks frosted and translucent as shown by the optical image in Figure A1b.
Generation of defects, vacancies, and interstitials under high intensity fs-pulses is taking place, e.g., F +  color center of the oxygen vacancy with a trapped electron created by fs-laser machining of sapphire is fully annealed at 1100 ° [33]. Frenkel pair defects interstitial—vacancy observed earlier by MeV-electron irradiated high-pressure cubic-BN was induced by fs-laser pulses [34] at similar conditions as used in this study. Chemical bond-breaking by high-intensity laser pulses can create metal-enriched regions on the surface of oxides and nitrides due to formation of volatile O2 and N2. Such laser-ablated surfaces of oxides rich with defects and debris and depleted of oxygen are expected to be restructured by HTA. We chose O2 flow in order to facilitate surface migration of Al atoms to form terraces of Al2O3 under oxygen supply. This is one probable contribution to the formation of smooth crystalline surface.

3.2. Recovery of Crystalline Smooth Surface on Ablated Sapphire

When focusing using a less tight objective lens  N A < 0.14  and carrying out single-pulse ablation with HTA in N2 flow, there were more debris left after annealing as compared with samples annealed in O2 (Figure 2). Nitrogen-rich atmosphere was not favorable to recovery of crystalline surface as compared with oxygen flow (see discussion above). Also, a pattern of single pulse ablation showed non-periodic structure with close-to-wavelength feature length. In addition, there was no typical ablation crater, rather a nano-textured surface resembling thermally quenched molten phase. The pulse fluence used was  F p 10  J/cm 2  or average intensity of  I p 43  TW/cm 2 , which is smaller by more than an order of magnitude than the self-clamping intensity of fs-pulse ∼ 10 14  W/cm 2  when multi-filaments are formed in air [35], and could cause a random pattern of dot-like modifications similar to those in the SEM image (Figure 2). Formation and quenching of a molten phase is more probable. Under such high intensity 43 TW/cm 2 /pulse exceeding the ablation threshold of sapphire by ∼5 times, formation of a highly excited plasma skin layer of tens of nanometers was expected during the laser pulse [36]. Light-matter interaction takes place as if there is a nano-film of different optical permittivity  ε n ˜  placed on dielectric sapphire;  n ˜ = n + i κ  is the complex refractive index. Such transient material state described by changing permittivity  ε ( t )  is called Die-Met (from dielectric and metal-like plasma) [36]. It has strong absorption due to reduced n and increased  κ  due to electron generation and is also referred to as epsilon-near-zero (ENZ) state  ε = ( n κ ) 2 + i 2 n κ [37].
The ablation of thin films, which proceeds after the end of the laser pulse, follows an imprint of an fs-laser pulse intensity. Such modulation has  ( λ / n ) / 2  period and, under multi-pulse exposure, reveals itself as ripples [38]. After HTA, a more orderly ripple pattern emerged with period smaller than  Λ < λ . The period of ripple-like pattern which evolved after HTA (Figure 2) resembles ripples ablated under multi-pulse exposure. We put forward here a conjecture (for future validation) that light intensity modulation ∼ λ / 2  occurred due to the presence of highly excited plasma nano-layer. Such modification of sub-surface region at the sapphire-plasma interface is hidden from SEM observation in Figure 2, which shows surface of a quenched molten phase. HTA was responsible for restructuring the surface and the imprinted modification was revealed. Figure 3 shows closeup SEM views of the ripples with evident re-crystallisation following c-plane of Al2O3 (hexagonal pits). Even though HTA temperature protocol was the same, the surface quality of HTA-treated ablated surface had more debris as compared with that made under O2 flow (Figure 4).
Figure 4 shows scanning electron microscopy (SEM) images of single and five-pulse ablated sites after HTA in O2 flow. A single pulse of  E p = 11.25 μ J which corresponds to a fluence of  F p = 61.32  J/cm 2  and an average intensity of  I p = 266  TW/cm 2  ablated a hole of 300 nm depth and a re-healed area of 4.4  μ m diameter, which almost perfectly matches the focal diameter of 4.8  μ m. After annealing, a perfectly smooth surface was observed near the rim of the ablated pit. When the number of pulses was two or more, formation of ripples [38] was observed (see structure of period  Λ 1 μ m at the bottom (Figure 4b)). Recovering a smooth surface during HTA in O2 flow is a promising outcome for opto-electronic applications of 2D materials.

3.3. HTA Treatment of Back-Side Ablated Al2O3

Since front-surface ablation caused extensive deposition of debris, a back-side ablation was used to mitigate this issue. No liquid immersion was used as in a previous study [13]. Ablation by circular scan started from the largest diameter and was closing towards the center in concentric circles following trajectory ablating an aspherical micro-lens of  125 μ m radius. After all-layer ablation, the depth of focus was changed and the next layer (into the sample from the back-side) was ablated. This protocol was robust against formation of debris and free from defects such as random non-ablated regions. The results are shown in Figure 5. Bruker ContourGT Optical Profiler was used to map surface geometry and surface roughness. The ablation steps according to the protocol of focal position are clearly discernable together with the roughness. The height step of  Δ z = 0.8 μ m was used for most reliable ablation of asphetric dome; the step inside Al2O3 becomes  n × Δ z  due to refractive index  n 1.7  at the wavelength of laser printing. The fluence per pulse  F p = 4.7  J/cm 2  was approximately twice the threshold of sapphire ablation for the focal spot of  r = 0.61 λ / N A = 1.4 μ m radius. Up to  1 μ m roughness (min-max) was observed on as ablated surfaces after ultrasonic washing. Samples were Cr-coated for SEM observation which corroborated reduction of roughness after HTA. Approximately five-times reduction of surface roughness down to ∼200 nm was observed for the laser ablation at  N A = 0.45  focusing. Fabrication of the micro-lens shown in Figure 5 took ∼15 min at laser repetition rate of 200 kHz and linear scan speed of ∼15 mm/s (pulse-to-pulse overlap was  Δ x = 1.3 μ m as well as the line-to-line separation and was comparable with the radius r of the focal spot). The designed aspherical lens has a height profile:
Z ( s ) = C s 2 1 + 1 1 ( 1 + k ) C 2 s 2 + A 4 s 4 + A 6 s 6 + A 8 s 8 + A 10 s 10 ,
where the coefficients  C = 1 / 0.15 k = 0 A 4 = 0.0093 A 6 = 0.005 A 8 = 0.0056 A 10 = 0.0005 , and s is the radial distance from the optical axis.
HTA made the surface smoother and the steps in z-direction less steep. HTA is responsible for the evolution of the overall aspheric shape becoming closer to the designed shape. There was almost no change in the base diameter and height of the lens (Figure 5a,b). Ablation at a fluence only twice that of the ablation threshold of the crystalline host material contributed to a good definition of the shape of the micro-lens. In glasses, melting and surface tension-induced reshaping is taking place even without HTA [39].
The main cause of the reduction in roughness was surface restructuring during the HTA process, as illustrated in Figure 6. Sublimation could be contributing to the discernible difference in the height change of the optical profiles (Figure 5). The steps along the ablation scans become less steep. This study shows that smoother surfaces of crystalline lenses can be achieved using HTA in O2. Smaller height steps will be tested to achieve tens-of-nm surface roughness required for optical applications.

4. Conclusions and Outlook

Sapphire ablation with different numerical apertures ( N A = 0.26  and 0.14) and similar pulse energies ( 10 μ J and  6.25 μ J) are compared. SEM images showed the formation of ripples with period close to the wavelength of the laser irradiation for the  N A = 0.26  and interestingly, it was half of the wavelength of the laser PHAROS (Light Conversion Ltd.) for  N A = 0.14 . After the samples had undergone high-temperature annealing at 1500  ° C for 2 to 4 h in either an O2 or N2 environment, these ripples underwent restructuring and crystalline step-like structures become apparent. It was also shown that annealing in an O2 environment reduced the amount of debris and the surface became smooth. In the case of HTA in N2, however, pieces of debris were still present and their coarsening took place. Detailed X-ray Photoelectron Spectroscopy (XPS) experiments are planned for detailed understanding of material modification on the surface during HTA in O2 and N2 flow. Apparently, the O2 flow was an important factor for restoration of a smooth c-plane surface by HTA.
We show that the restructuring of laser-ablated crystalline surfaces can be applied to the fabrication of micro-optical elements in reducing the surface roughness and allowing for tighter control over the surface profile while maintaining the precision of a shape definition. This is different from glasses [40], where HTA causes a surface tension-driven reflow and formation of spherical surfaces which might be different from the required shape for a defined optical function. Back-side ablation in air was used to define an aspherical micro-lens profile, which was not altered by HTA while surface roughness was markedly reduced. The true extent of the control over surface roughness is yet to be fully explored using this technique. The optimal time of annealing and the optimal temperature to achieve the most restructuring without sacrificing the surface profile will be determined in future studies. Sublimation via solid-to-gas occurring along with solid-fusion-vaporisation stages during HTA needs better understanding and can be different in other crystals.

Author Contributions

Conceptualisation, S.J. and S.H.N.; investigation, D.S., A.T., D.M. and T.K.; writing—original draft preparation, S.J. and D.S.; review and editing, all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Australian Research Council, grant number LP190100505. S.J. is grateful for startup funding from the Nanotechnology Facility at Swinburne and to the Workshop-on-Photonics for the technology transfer project, which installed the first industrial-grade microfabrication setup in Australia. Part of this study was made during a Grand Challenge undergraduate project of A.T. Mechanisms of light-matter interactions were discussed with Prof. Eugene G. Gamaly. We are grateful to Laser Systems Ltd., Sapporo, Japan and Tecdia Ltd., Tokyo, Japan companies for providing ultra-thin 30 μm-thick Al2O3 samples.

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.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. HTA of Al2O3 in N2

The Al2O3 HTA in O2 flow caused restructuring of surface at the ablated regions and sublimation of debris (Figure 5) contrasting the surface annealed in N2 flow (Figure A1). The phase diagrams single out ALN and ALON [32] for surface modification shown in Figure A1.
Figure A1. (a) SEM images of ALON nanotextured surfaces around the ablated regions. (b) The ablated areas becomes translucent (Al2O3 was placed on a green-colored pad for photo). Ablation was carried out at similar conditions to those shown in Figure 4. Thickness of Al2O3 was ∼ 30 μ m; initially, it was two-side polished to a high optical transparency.
Figure A1. (a) SEM images of ALON nanotextured surfaces around the ablated regions. (b) The ablated areas becomes translucent (Al2O3 was placed on a green-colored pad for photo). Ablation was carried out at similar conditions to those shown in Figure 4. Thickness of Al2O3 was ∼ 30 μ m; initially, it was two-side polished to a high optical transparency.
Photonics 10 00594 g0a1

References

  1. Batani, K.; Batani, D.; He, X.; Shigemori, K. Recent progress in matter in extreme states created by laser. Matter Radiat. Extrem. 2022, 7, 013001. [Google Scholar] [CrossRef]
  2. Juodkazis, S.; Nishimura, K.; Tanaka, S.; Misawa, H.; Gamaly, E.E.; Luther-Davies, B.; Hallo, L.; Nicolai, P.; Tikhonchuk, V. Laser-Induced Microexplosion Confined in the Bulk of a Sapphire Crystal: Evidence of Multimegabar Pressures. Phys. Rev. Lett. 2006, 96, 166101. [Google Scholar] [CrossRef] [PubMed]
  3. Vailionis, A.; Gamaly, E.G.; Mizeikis, V.; Yang, W.; Rode, A.; Juodkazis, S. Evidence of super-dense Aluminum synthesized by ultra-fast micro-explosion. Nat. Commun. 2011, 2, 445. [Google Scholar] [CrossRef] [PubMed]
  4. Zelovich, Y.B.; Raizer, Y.P. Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena; Dover, Dover Books on Physics: New York, NY, USA, 2002. [Google Scholar]
  5. Campanella, B.; Legnaioli, S.; Pagnotta, S.; Poggialini, F.; Palleschi, V. Shock waves in laser-induced plasmas. Atoms 2019, 7, 57. [Google Scholar] [CrossRef]
  6. Juodkazis, S.; Nishimura, K.; Misawa, H.; Ebisui, T.; Waki, R.; Matsuo, S.; Okada, T. Control over the State of Crystallinity: Sapphire. Adv. Mat. 2006, 18, 1361–1364. [Google Scholar] [CrossRef]
  7. Huang, H.H.; Juodkazis, S.; Gamaly, E.; Nagashima, T.; Yonezawa, T.; Hatanaka, K. Spatio-temporal control of THz emission. Commun. Phys. 2022, 5, 134. [Google Scholar] [CrossRef]
  8. Juodkazis, S.; Misawa, H.; Gamaly, E.G.; Luther-Davis, B.; Hallo, L.; Nicolai, P.; Tikhonchuk, V. Is the nano-explosion really microscopic? J. Non-Crystall. Solids 2009, 355, 1160–1162. [Google Scholar] [CrossRef]
  9. Nath, G. Approximate analytical solution for the propagation of shock waves in self-gravitating perfect gas via power series method: Isothermal flow. J. Astrophys. Astron. 2020, 41, 21. [Google Scholar] [CrossRef]
  10. Sharova, Y.S.; Glazyrin, S.I.; Gasilov, V.A. Study of the Influence of the Background Neutral Component on the Dynamics of the Envelope in Supernova Remnants. Astron. Lett. 2021, 47, 746–753. [Google Scholar] [CrossRef]
  11. Petruk, O.; Kuzyo, T.; Orlando, S.; Pohl, M.; Brose, R. Magneto-hydrodynamic simulations of young supernova remnants and their energy-conversion phase. Mon. Not. R. Astron. Soc. 2021, 505, 755–770. Available online: https://academic.oup.com/mnras/article-pdf/505/1/755/39549986/stab1319.pdf (accessed on 1 May 2023). [CrossRef]
  12. Hirashima, K.; Moriwaki, K.; Fujii, M.; Hirai, Y.; Saitoh, T.; Makino, J. Predicting the Expansion of Supernova Shells Using Deep Learning toward Highly Resolved Galaxy Simulations. Proc. Int. Astron. Union 2020, 16, 209–214. [Google Scholar] [CrossRef]
  13. Hua, J.G.; Liang, S.Y.; Chen, Q.D.; Juodkazis, S.; Sun, H.B. Free-Form Micro-Optics Out of Crystals: Femtosecond Laser 3D Sculpturing. Adv. Func. Mat. 2022, 32, 2200255. [Google Scholar] [CrossRef]
  14. Marupalli, B.C.; Adhikary, T.; Sahu, B.; Mitra, R.; Aich, S. Effect of annealing temperature on microstructure and mechanical response of sputter deposited Ti-Zr-Mo high temperature shape memory alloy thin films. Appl. Surf. Sci. Adv. 2021, 6, 100137. [Google Scholar] [CrossRef]
  15. Shklyaev, A.; Latyshev, A. Surface Morphology Transformation Under High-Temperature Annealing of Ge Layers Deposited on Si(100). Nanoscale Res. Lett. 2016, 11, 366. [Google Scholar] [CrossRef] [PubMed]
  16. Woo, S.; Jeong, H.; Lee, S.; Seo, H.; Lacotte, M.; David, A.; Kim, H.; Prellier, W.; Kim, Y.; Choi, W.S. Surface properties of atomically flat poly-crystalline SrTiO3. Sci. Rep. 2015, 5, 8822. [Google Scholar] [CrossRef]
  17. Gailevicius, D.; Padolskyte, V.; Mikoliunaite, L.; Sakirzanovas, S.; Juodkazis, S.; Malinauskas, M. Additive-manufacturing of 3D glass-ceramics down to nanoscale resolution. Nanoscale Horiz. 2019, 4, 647–651. [Google Scholar] [CrossRef]
  18. Acosta-Alba Pablo, E.; Kononchuk, O.; Gourdel, C.; Claverie, A. Surface self-diffusion of silicon during high temperature annealing. J. Appl. Phys. 2014, 115, 134903. [Google Scholar] [CrossRef]
  19. Liu, M.; Zhou, H.; Zhu, H.; Yue, Z.; Zhao, J. Microstructure and dielectric properties of glass/Al2O3 composites with various low softening point borosilicate glasses. J. Mater. Sci. Mater. Electron. 2012, 23. [Google Scholar] [CrossRef]
  20. Uwaha, M.; Saito, Y. Kinetic smoothing and roughening of a step with surface diffusion. Phys. Rev. Lett. 1992, 68, 224–227. [Google Scholar] [CrossRef]
  21. Zaluska-Kotur, M.A.; Krzyzewski, F.; Krukowski, S.; Czernecki, R.; Leszczynski, M. Structures Built by Steps Motion during Sublimation from Annealed GaN(0001) Surface. Cryst. Growth Des. 2013, 13, 1006–1013. [Google Scholar] [CrossRef]
  22. Kazantsev, D.; Akhundov, I.; Rudaya, N.; Kozhukhov, A.; Alperovich, V.; Latyshev, A. Thermal roughening of GaAs surface by unwinding dislocation-induced spiral atomic steps during sublimation. Appl. Surf. Sci. 2020, 529, 147090. [Google Scholar] [CrossRef]
  23. Kazantsev, D.; Akhundov, I.; Karpov, A.; Shwartz, N.; Alperovich, V.; Terekhov, A.; Latyshev, A. Monte Carlo simulation of GaAs(001) surface smoothing in equilibrium conditions. Appl. Surf. Sci. 2015, 333, 141–146. [Google Scholar] [CrossRef]
  24. Mundschau, M.; Bauer, E.; Telieps, W.; Swieh, W. Atomic steps on Si100 and step dynamics during sublimation studied by low-energy electron microscopy. Surf. Sci. 1989, 223, 413–423. [Google Scholar] [CrossRef]
  25. Sheglov, D.V.; Sitnikov, S.V.; Fedina, L.I.; Rogilo, D.I.; Kozhukhov, A.S.; Latyshev, A.V. From Self-Organization of Monoatomic Steps on the Silicon Surface to Subnanometer Metrology. Optoelectron. Instrum. Data Process. 2020, 56, 533–544. [Google Scholar] [CrossRef]
  26. Readey, D.; Kuczynski, G. Sublimation of Aluminum Oxide in Hydrogen. J. Am. Ceram. Soc. 2006, 49, 26–29. [Google Scholar] [CrossRef]
  27. Andre, P.; Abbaoui, M.; Augeard, A.; Desprez, P.; Singo, T. Study of Condensed Phases, of Vaporization Temperatures of Aluminum Oxide and Aluminum, of Sublimation Temperature of Aluminum Nitride and Composition in an Air Aluminum Plasma. Plasma Chem. Plasma Process. 2016, 36, 1161–1175. [Google Scholar] [CrossRef]
  28. Gamaly, E.E.; Juodkazis, S.; Nishimura, K.; Misawa, H.; Luther-Davies, B.; Hallo, L.; Nicolai, P.; Tikhonchuk, V. Laser-matter interaction in a bulk of a transparent solid: Confined micro-explosion and void formation. Phys. Rev. B 2006, 73, 214101. [Google Scholar] [CrossRef]
  29. Hua, J.G.; Ren, H.; Huang, J.; Luan, M.L.; Chen, Q.D.; Juodkazis, S.; Sun, H.B. Laser-Induced Cavitation-Assisted True 3D Nano-Sculpturing of Hard Materials. Small 2023, 2207968. [Google Scholar] [CrossRef]
  30. Liu, X.Q.; Zhang, Y.L.; Li, Q.K.; Zheng, J.X.; Lu, Y.M.; Juodkazis, S.; Chen, Q.D.; Sun, H.B. Biomimetic sapphire windows enabled by inside-out femtosecond laser deep-scribing. PhotoniX 2022, 3, 1. [Google Scholar] [CrossRef]
  31. Mahesh, J.Y.; Aravindan, S.; Rao, P.V.; Yogita, M.; Singh, J.P. Surface morphology evolution of C-plane sapphire during multi-step thermal annealing. Surf. Topogr. Metrol. Prop. 2022, 10, 045015. [Google Scholar] [CrossRef]
  32. Prosvirnin, D.V.; Kolmakov, A.G.; Larionov, M.D.; Prutskov, M.E.; Alikhanyan, A.S.; Samokhin, A.V.; Lysenkov, A.S.; Titov, D.D. Effect of sintering methods and temperatures on porosity of the ceramics from aluminum oxinitride. IOP Conf. Ser. Mater. Sci. Eng. 2018, 347, 012030. [Google Scholar] [CrossRef]
  33. Kudrius, T.; Šlekys, G.; Juodkazis, S. Surface-texturing of sapphire by femtosecond laser pulses for photonic applications. J. Phys. D: Appl. Phys. 2010, 43, 145501. [Google Scholar] [CrossRef]
  34. Buividas, R.; Aharonovich, I.; Seniutinas, G.; Wang, X.W.; Rapp, L.; Rode, A.V.; Juodkazis, S. Photoluminescence from voids created by femtosecond-laser pulses inside cubic-BN. Opt. Lett. 2015, 40, 5711–5713. [Google Scholar] [CrossRef] [PubMed]
  35. Hsu, W.H.; Masim, F.; Balcytis, A.; Juodkazis, S.; Hatanaka, K. Dynamic position shifts of X-ray emission from a water film induced by a pair of time-delayed femtosecond laser pulses. Opt. Express 2017, 25, 24109–24118. [Google Scholar] [CrossRef]
  36. Gamaly, E.G.; Rode, A.V. Ultrafast re-structuring of the electronic landscape of transparent dielectrics: New material states (Die-Met). Appl. Phys. A 2018, 124, 278. [Google Scholar] [CrossRef]
  37. Ng, S.; Juodkazis, S. Nanoscale Plasmonic Printing in Ultrafast Laser Nanostructuring: The Pursuit of Extreme Scales; Stoian, R., Bonse, J., Eds.; Springer Series in Optical Science; Springer: Berlin/Heidelberg, Germany, 2023; Volume 239, Chapter 25. [Google Scholar]
  38. Buividas, R.; Mikutis, M.; Juodkazis, S. Surface and bulk structuring of materials by ripples with long and short laser pulses: Recent advances. Prog. Quantum Electron. 2019, 38, 119–156. [Google Scholar] [CrossRef]
  39. Vanagas, E.; Kudryashov, I.; Tuzhilin, D.; Juodkazis, S.; Matsuo, S.; Misawa, H. Surface nanostructuring of borosilicate glass by femtosecond nJ energy pulses. Appl. Phys. Lett. 2003, 82, 2901–2903. [Google Scholar] [CrossRef]
  40. Bach, H.; Krause, D. Technical Glasses—Physical and Technical Properties. In Technical Report; SCHOTT AG: Mainz, Germany, 2020. [Google Scholar]
Figure 1. (a) A 1-mm-diameter lens ablated on sapphire Al2O3 c-plane with fs-laser pulses 1030 nm/230 fs. (b) 3D Bruker ContourGT Optical Profiler scan. (c) Roughness reduction by a high-temperature annealing (HTA) at 1500  ° C for 2 h in N2 flow of 200 mL/hour; same micro-lens. The center of the lens was not ablated for this front-side ablation mode due to a  π -phase change of the reflected pulse. Ablation was made at a pulse density of 750 pulses/mm;  N A = 0.14  (Mitutoyo). Linear polarisation of the E-field of laser pulses was horizontal. Pulse energy was  E p = 4.375 μ J (on sample). Fluence was  F p = 6.91  J/cm 2 .
Figure 1. (a) A 1-mm-diameter lens ablated on sapphire Al2O3 c-plane with fs-laser pulses 1030 nm/230 fs. (b) 3D Bruker ContourGT Optical Profiler scan. (c) Roughness reduction by a high-temperature annealing (HTA) at 1500  ° C for 2 h in N2 flow of 200 mL/hour; same micro-lens. The center of the lens was not ablated for this front-side ablation mode due to a  π -phase change of the reflected pulse. Ablation was made at a pulse density of 750 pulses/mm;  N A = 0.14  (Mitutoyo). Linear polarisation of the E-field of laser pulses was horizontal. Pulse energy was  E p = 4.375 μ J (on sample). Fluence was  F p = 6.91  J/cm 2 .
Photonics 10 00594 g001
Figure 2. Single-pulse ablation and HTA of Sapphire (Al2O3) c-plane. HTA at 1500  ° C for 2 h in N2 flow. Ablation was made by one pulse per ablation site. Focusing of 1030 nm/230 fs pulses was carried out with an objective lens of numerical aperture  N A = 0.14  (Mitutoyo). Linear polarisation of the E-field of laser pulses is shown by arrow marker. Pulse energy was  E p = 6.25 μ J (on sample). Fluence was  F p = 9.87  J/cm 2 .
Figure 2. Single-pulse ablation and HTA of Sapphire (Al2O3) c-plane. HTA at 1500  ° C for 2 h in N2 flow. Ablation was made by one pulse per ablation site. Focusing of 1030 nm/230 fs pulses was carried out with an objective lens of numerical aperture  N A = 0.14  (Mitutoyo). Linear polarisation of the E-field of laser pulses is shown by arrow marker. Pulse energy was  E p = 6.25 μ J (on sample). Fluence was  F p = 9.87  J/cm 2 .
Photonics 10 00594 g002
Figure 3. Single-pulse ablated Sapphire (Al2O3) c-plane (Figure 2) after HTA at 1500  ° C for 4 h in N2 flow at different magnifications. Focusing of 1030 nm/230 fs pulses was carried out with an objective lens of numerical aperture  N A = 0.14  (Mitutoyo). Linear polarisation of the E-field of laser pulses is shown by arrow marker. Pulse energy was  E p = 6.25 μ J (on sample). Fluence was  F p = 9.87  J/cm 2 .
Figure 3. Single-pulse ablated Sapphire (Al2O3) c-plane (Figure 2) after HTA at 1500  ° C for 4 h in N2 flow at different magnifications. Focusing of 1030 nm/230 fs pulses was carried out with an objective lens of numerical aperture  N A = 0.14  (Mitutoyo). Linear polarisation of the E-field of laser pulses is shown by arrow marker. Pulse energy was  E p = 6.25 μ J (on sample). Fluence was  F p = 9.87  J/cm 2 .
Photonics 10 00594 g003
Figure 4. Sapphire (Al2O3) c-plane after fs-laser ablation and high-temperature annealing (HTA) at 1500  ° C for 2 h in O2 flow. Ablation was made by: (a) one pulse and (b) five pulses per ablation site. Focusing of 1030 nm/230 fs pulses was carried out with an objective lens of numerical aperture  N A = 0.26  (Mitutoyo). Linear polarisation of the E-field of laser pulses was horizontal; shown by marker. Pulse energy was  E p = 10 μ J (on sample). Fluence was  F p = 54.5  J/cm 2 .
Figure 4. Sapphire (Al2O3) c-plane after fs-laser ablation and high-temperature annealing (HTA) at 1500  ° C for 2 h in O2 flow. Ablation was made by: (a) one pulse and (b) five pulses per ablation site. Focusing of 1030 nm/230 fs pulses was carried out with an objective lens of numerical aperture  N A = 0.26  (Mitutoyo). Linear polarisation of the E-field of laser pulses was horizontal; shown by marker. Pulse energy was  E p = 10 μ J (on sample). Fluence was  F p = 54.5  J/cm 2 .
Photonics 10 00594 g004
Figure 5. Back-side ablation of Al2O3 and HTA. (a) Optical Profiler surface map (Bruker ContourGT Optical Profiler). (b) Cross-sectional trace of surface of micro-lens before and after HTA. SEM image inset shows top-view of the lens. Top-of-lens cross-section profile is shown in the inset. (c) SEM images of the top-central region of the lens before and after HTA. Laser fabrication conditions: pulse energy  E p = 0.29 μ J ( F p = 4.7  J/cm 2 I p = 20.5  TW/cm 2 ), pulse-to-pulse separation along scan 1.3  μ m (3 pulses per irradiation site), line-to-line separation (in plane) was also 1.3 nm and along the height direction  0.8 μ m, wavelength  λ = 1030  nm, focusing  N A = 0.45  (Olympus, 20×), repetition rate 0.2 MHz. Circularly polarised beam was set onto sample by fine tuning of  λ / 2 -plate in front of  λ / 4 -plate used as a circular polariser.
Figure 5. Back-side ablation of Al2O3 and HTA. (a) Optical Profiler surface map (Bruker ContourGT Optical Profiler). (b) Cross-sectional trace of surface of micro-lens before and after HTA. SEM image inset shows top-view of the lens. Top-of-lens cross-section profile is shown in the inset. (c) SEM images of the top-central region of the lens before and after HTA. Laser fabrication conditions: pulse energy  E p = 0.29 μ J ( F p = 4.7  J/cm 2 I p = 20.5  TW/cm 2 ), pulse-to-pulse separation along scan 1.3  μ m (3 pulses per irradiation site), line-to-line separation (in plane) was also 1.3 nm and along the height direction  0.8 μ m, wavelength  λ = 1030  nm, focusing  N A = 0.45  (Olympus, 20×), repetition rate 0.2 MHz. Circularly polarised beam was set onto sample by fine tuning of  λ / 2 -plate in front of  λ / 4 -plate used as a circular polariser.
Photonics 10 00594 g005
Figure 6. SEM imaging. Nanoscale reconstruction of the c-plane crystalline Al2O3 surface after HTA; central part of the lens.
Figure 6. SEM imaging. Nanoscale reconstruction of the c-plane crystalline Al2O3 surface after HTA; central part of the lens.
Photonics 10 00594 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Smith, D.; Ng, S.H.; Tang, A.; Katkus, T.; Moraru, D.; Juodkazis, S. Crystalline Flat Surface Recovered by High-Temperature Annealing after Laser Ablation. Photonics 2023, 10, 594. https://doi.org/10.3390/photonics10050594

AMA Style

Smith D, Ng SH, Tang A, Katkus T, Moraru D, Juodkazis S. Crystalline Flat Surface Recovered by High-Temperature Annealing after Laser Ablation. Photonics. 2023; 10(5):594. https://doi.org/10.3390/photonics10050594

Chicago/Turabian Style

Smith, Daniel, Soon Hock Ng, Amanda Tang, Tomas Katkus, Daniel Moraru, and Saulius Juodkazis. 2023. "Crystalline Flat Surface Recovered by High-Temperature Annealing after Laser Ablation" Photonics 10, no. 5: 594. https://doi.org/10.3390/photonics10050594

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