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

High performance laser-driven flyers based on a refractory metamaterial perfect absorber

Open Access Open Access

Abstract

Laser-driven flyers (LDFs), which can drive metal particles to ultra-high speeds by feeding high-power laser, have been widely used in many fields, such as ignition, space debris simulation, and dynamic high-pressure physics. However, the low energy-utilization efficiency of the ablating layer hinders the development of LDF devices towards low power consumption and miniaturization. Herein, we design and experimentally demonstrate a high-performance LDF based on the refractory metamaterial perfect absorber (RMPA). The RMPA consists by a layer of TiN nano-triangular array, a dielectric layer and a layer of TiN thin film, and is realized by combing the vacuum electron beam deposition and colloid-sphere self-assembled techniques. RMPA can greatly improve the absorptivity of the ablating layer to about 95%, which is comparable to the metal absorbers, but obviously larger than that of the normal Al foil (∼10%). This high-performance RMPA brings a maximum electron temperature of ∼7500 K at ∼0.5 µs and a maximum electron density of ∼1.04 × 1016 cm-3 at ∼1 µs, which are higher than that the LDFs based on normal Al foil and metal absorbers due to the robust structure of RMPA under high-temperature. The final speed of the RMPA-improved LDFs reaches to about 1920 m/s measured by the photonic Doppler velocimetry system, which is about 1.32 times larger than the Ag and Au absorber-improved LDFs, and about 1.74times larger than the normal Al foil LDFs under the same condition. This highest speed unambiguously brings a deepest hole on the Teflon slab surface during the impact experiments. The electromagnetic properties of RMPA, transient speed and accelerated speed, transient electron temperature and density have been systematically investigated in this work.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Laser-driven flyer (LDF) have attracted a significant attention owing to their security, reliability and synchronism in various applications, such as the detonation [13], space scrap metal science [46], and dynamic high-pressure physics [79]. The LDF is a device, where the high temperature and pressure plasma generated from the pulse laser-ablated metal film is used to drive remaining thin plate forming a flyer to several km/s. A typical LDF is composed of an ablating layer, an insulation layers, a flyer layer and an accelerated chamber, where the ablating layer can absorb the incident laser pulse, and generate high temperature and pressure plasma to drive the flyer layer. Optimizing precisely composition materials or the device structure parameters can improve the flyer performance, where the ablating layer plays a crucial role in improving the energy utilization efficiency [10,11]. Over the past decades, LDF have made great progress in both of civilian and military applications [12,13]. However, the generally-employed ablating layer of Al foil own large reflectance and thus low energy utilization efficiency. This greatly hiders the progress of LDF towards the miniaturization and industrialization.

To improve the energy utilization efficiency of LDF, researchers have made a great effort to improve the performance of the ablating layer. Lots of high-absorption materials, such as C, Ge, Ti, Zn, and even the commercial-grade black lacquer paint, have been investigated as the anti-reflection layers to improve the ablating layer absorptivity, and the final flyer speed has been successfully improved [14,15]. But the speed-enhanced factor is usually limited to smaller than 30%. Employing the energetic materials to fabricate the ablating layer is another scheme to improve the flyer speed, where the reactive multilayer films can release additional chemical energy through aluminothermic reaction [1618]. The absorptivity of this reactive layer still doesn't exceed 50% yet. Benefit from the rapid process in nanophotonics, the perfect metamaterial absorbers (PMAs), which are composed by the artificial structures, shows a nearly 100% absorptivity in the designed waveband through properly engineering the optical resonators, and inspire a new scheme for designing high-performance LDF [19]. Some skillful designs can achieve perfect absorption from visible light to near-infrared, and have been widely used [2022]. Our previous work firstly demonstrated this scheme by employing the PMA with an Ag nano-pillar array, and the final flyer speed was enhanced for about 1.4 times due to the improved high electron temperature and plasma density in the ablating layer [23]. In addition, Aluminum nanoparticles based on direct writing technique also have been investigated as potential PMA for improving the flyer speed [24]. However, the ablating layer will experience melting and vaporization and generating plasma under the irradiation of high-power laser. And all of the previously-reported PMAs used in LDFs are made by noble metals, for instance, Ag, Au and Al, and can't keep the structure stable and high absorptivity under high temperature, which will certainly limit the energy utilization efficiency.

In this work, we design and experimentally realize a kind of refractory PMA (RPMA) to improve the optical absorbance stability of LDF and thus the final flyer speed. Titanium nitride (TiN) has been reported as a novel refractory plasmonic material which can sustain high temperatures and exhibits a plasmonic resonance in the visible-NIR range, potentially opening the door for high-power system applications [2527]. The PMA used in this work owns a classical metal-dielectric-metal structure composed by a TiN nano-triangular layer, a SiO2 layer, and TiN layer, which is fabricated by combing the electron beam evaporation and colloid-sphere assembled techniques. Strong absorption in the NIR region is demonstrated in both simulation and experiment, and the perfect absorption mainly come from the surface plasmonic resonances and the intrinsic loss of the material. In particular, due to its high melting point of TiN, the RPMA can support continuous high absorption for a long time under high temperature condition. And this leads to an enhancement on the electron temperature and plasma density formed in the ablating layer, and results in a maximum velocity enhancement of about 70% when comparing with the Al foil LDF, and 30% when comparing the normal PMA-improved LDFs. The electromagnetic properties of TiN absorber, laser-induced plasma spectrum, the flyer velocity, accelerated velocity and impact results of the flyer have been systematically investigated.

2. Experiment section

2.1 TiN-absorber enhanced flyer design and achievement

The designed RPMA-improved LDF is composed of a confining layer, a TiN-absorber, an Al foil and an accelerated chamber, as shown in Fig. 1(a). In general, the confining substrate is an optically transparent and lossless dielectric substrate, which is Al2O3 with refractive index of about 1.7 in this work. This confining substrate can let light transmit with low-loss, and prevent the plasma to expand towards the substrate side. The Al foil, together with the RPMA, work as both the ablating layer and the flyer layer, where the high temperature and pressure plasma can be generated immediately when the pulse laser illuminates on the ablating layer. The Al foil is used here due to its advantages of high tenacity and low density for shearing an integrated flyer. The accelerated chamber provides a shearing fore for the formation of Al flyer, and also bounds the generated detonation wave and the plasma to accelerate the flyer. And the RPMA consists of a TiN nano-triangular array layer, a SiO2 layer and a TiN layer, which is placed between the confining substrate and the Al foil layer. The optimized absorber not only improve the absorbance, but also enhance the ablation efficiency effectively, as a consequence of the higher melting point causing higher laser energy deposition of the absorption layer under laser ablation before the structure melting.

 figure: Fig. 1.

Fig. 1. (a) The schematic diagram of the RPMA-improved LDF, include a confining window (sapphire substrate), the ablating layer (TiN absorber), the flyer layer (Al foil) and an accelerated chamber (barrel). (b) The fabrication and packaging processing of the integrated flyer, the fabrication of PMA by the colloid lithography and thin-film deposition technique firstly, and then the Al foil and the accelerated chamber are packed closely behind the PMA orderly.

Download Full Size | PDF

The detail fabrication process of the RPMA-improved LDF is shown in Fig. 1(b). The cylindrical sapphire with thickness of 2 mm and diameter of 5 mm was used as the confining layer. The sapphire substrate was washed with acetone and alcohol under the ultrasound condition for 20 minutes. The PS array monolayer with different size of 430 nm, 600 nm and 880 nm were prepared on the sapphire substrate though the method in Ref. [28]. And then, 150 nm TiN film was deposited vertically on the PS monolayer-covered substrate by the electron beam evaporation, and the triangular array was successfully formed on substrates after removing the PS monolayer by immersing the sample in acetone under ultrasound for 20 seconds. It's worth noting that the deposition rate could be properly controlled to prevent PS deformation caused by high chamber temperature. After that, 250 nm SiO2 and 150 nm TiN were successively deposited sequentially on the triangular array surface through the same method. After these steps, the RPMA is fabricated successfully. The normal novel metal PMAs were fabricated by the similar methods except for the vacuum thermal evaporation system. The LDF fabrication started by adhering a 25 µm Al foil to the PMAs by using a silica gel carefully, and then attaching the accelerated chamber with thickness of 2 mm and the internal diameter of 0.8 mm tightly to the Al foil. Finally, the PMA-improved LDF was packaged by a metal shell.

2.2 Spectrum and morphology characterization

The scanning electron microscopy (SEM, Apreo S, USA) was used to characterize the morphology of the fabricated PS sphere monolayer and the triangular array. To improve the conductivity of the samples, a layer of 2 nm gold film was deposited on the top surface of the samples, and a conductive packaging tape was used to link the substrate and the specimen mount to improve the conductivity. The measured SEM images are shown in the Results section. The UV-VIS-NIR spectrophotometer (Cary 5000, Agilent, China) was used to measure the reflectivity spectra of the fabricated samples in the near infrared regions (900-1500 nm). The absorbance A was calculated by A = 1-R, where R is the reflectance.

2.3 Transient plasma emissing spectra measurement

The Q-switched Nd:YAG laser working at 1064 nm with a pulse duration of 14 ns was employed to illuminate the RPMA-improved LDF, and the laser energy was set to 45 mJ. The laser beam was vertically focused onto the sample surface through a quartz convex lens of 12 cm focal length, and the diameter of laser spot was 0.8 mm on the ablation layer. The high temperature and pressure plasma was generated immediately after the laser illuminated on the surface of the ablating layer. The optical emission spectrum from the plasma was collected by an optical fiber of 100-µm core diameter, which is connected to the slit entrance of a spectrometer (Andor Tech., Shamrock 303i, UK). A grating of 600 grooves/mm was chosen to provide a spectral resolution of 0.07 nm. The analysis of emission spectra was performed using a computer.

2.4 Flyer velocity and impact features characterization

The flyer speed in the accelerated chamber was characterized by using a photonics Doppler velocimetry (PDV) system. In the PDV system, the detection light signal is collected by a fiber bundle located on the end of the accelerated chamber and then transferred to the oscillograph for further analysis. To measure the impact features of the high-speed flyers, the cylindrical Teflon with height of 4 mm and diameter of 4 mm was placed tightly behind the accelerated chamber. After illuminating the high-power laser, as the high velocity impacted, the craters were formatted due to the impact of the high-speed flyers. The appearance and depth of crater were recorded by a Microscope (Olympus BX51) and a step profile (AmbiosXP + 200) respectively after washing off the impact residue with alcohol.

2.5 Optical simulation of the absorber

The electromagnetic simulations were carried out by using the finite-difference time-domain (FDTD) simulation method. A plane wave with the wavelength from 800 nm to 1600 nm was used to illuminate the primitive unit cell of RPMA along Z + direction, and the reflectivity was collected by a monitor placed in the reflection space. The periodic boundary conditions were used in the X and Y directions, and the perfectly matched layer boundary conditions were used in the Z direction. The permittivities of TiN, Au and Ag used in simulations were from the previously measured values [29,30]. The refractive index of Al2O3 and SiO2 were set as 1.7, 1.45, respectively. The absorption (A) was calculated by A = 1 − R. The calculated absorption spectra and the electric-field and electromagnetic energy dissipation distributions are shown in Fig. 2 and 3.

 figure: Fig. 2.

Fig. 2. (a) The perspective view, top view and side view of a unit cell of the absorber. The simulated absorption spectra of the RPMA samples with varying different structure parameters are shown in (b) diameter of PS sphere, D; (c) thickness of the silver triangular array, t; (d) thickness of the SiO2, m; (e) the refractive index of the dielectric medium, n. The inset dotted line marks the interface of the diffraction order. (f) Absorption spectrum of TiN, Ag and Au absorber with different D, where the optimum parameters are t = 150 nm, m = 250 nm, b = 150 nm.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. Calculated electric field distribution on the cross section of RPMA at the wavelength of 1064 nm: (a) the X-Y plane (triangular array -SiO2 spacer interface), (c) the Y-Z plane (X = 0). The electromagnetic energy dissipation distribution: (b) and (f) the X-Y plane, (d) the Y-Z plane. (e)-(i) Corresponding calculation results of Ag PMA.

Download Full Size | PDF

3. Results and discussions

3.1 Optical properties of the absorber

Figure 2(a) shows the detail structure of the designed RPMA. To optimize the optical properties, we make a systematical investigation on the influence of the key parameters, i.e., the diameter of PS sphere D, the thickness of triangle particle t, the thickness of the cavity and triangle m, and the refractive index of cavity n, as shown in Fig. 2(b-e). As expect, increasing D from 400 to about 900 nm leads to the absorption peaks red-shifts, where t = 150 nm and m = 250 nm. Due to the working wavelength of laser in experiment is 1064 nm, the diameter of 600 nm is chosen which the maximum absorbance can reach to about 99%. While increasing t from 20 nm to 150 nm, the absorption efficiency increases significantly in the waveband of from 0.93 µm to 1.6 µm, and the thickness of 150 nm is chosen as an optimized parameter. The parameter m shows a little weak influence on the absorbance, and the maximum absorbance happens at about m = 250 nm for the laser wavelength of 1064 nm. The refractive index n shows a similar influence to that of D. And considering the practical conditions, we choose SiO2 (n = 1.45) as the optimized dielectric material. Finally, we conclude that the optimal parameter is D = 600 nm, t = 150 nm and m = 250 nm, respectively. The optimized spectra are shown in Fig. 2(f), where the absorption spectra of the Au and Ag PMAs are also included to make a comparison. It is obvious that the absorbance of TiN PMA is higher and broader than that of Au and Ag PMAs under the same conditions. In addition, it should be noted that the diffraction mode greatly limits the absorption bandwidth, as indicated by the dashed lines in figures. The diffraction modes are extracted from the formula [31]

$${\lambda _{mn}} = {n_{med}}{[{{{({{m / {{P_x}}}} )}^2} + {{({{n / {{P_y}}}} )}^2}} ]^{{{ - 1} / 2}}}, $$
where nmed is the refractive index in diffraction space, m and n are the diffraction orders in X and Y directions, and Px and Py are the period in X and Y directions, respectively. The calculation results are marked on Fig. 2(b-e) with dotted lines.

To reveal the strong optical absorption behavior in RPMA, we calculate electric-field distribution and the energy dissipation distribution. Figure 3(a) shows the electric-field intensity profiles in the X-Y plane at the triangular-SiO2 spacer interface for the RPMA at 1064 nm, where the strong electric-field is localized near the edge and corner of the triangular structure due to the surface plasmon resonance. The weak field coupling in the sandwich structures is confirmed when we see that the electric field is confined in the SiO2 layer as shown in Fig. 2(c). These strong electric field enhancements can induce instant field electron emission and producing several seed electron sources in the laser spot, which is beneficial for ablation to generate high-pressure plasma for driving flyer. Meanwhile, the energy dissipation distribution is displayed in the X-Y and Y-Z plane as shown in Fig. 3(b) and (d) respectively. Most of the energy is thus consumed in the triangular arrays and a part of energy is dissipated at the bottom TiN film due to the intrinsic Ohmic loss, where the high energy dissipation indicates high temperature when illuminated by laser, and thus easily-destroyed parts. This is why RPMA is preferred in LDF. In addition, we also compared and analyzed Ag absorbers with the same parameters as show in Fig. 3(e-i). The Ag PMA shows similar electric-field and energy dissipation distributions with that of RPMA, indicating that the PMA is easily destroyed.

3.2 Preparation and characterization of TiN absorber

The optical properties and geometrical morphology of the experimentally realized RPMA are shown in Fig. 4, where the detailed fabrication process can be found in the Experimental Section. As show in Fig. 4(a), a hexagonally close-packed PS monolayer is successfully formed on the sapphire substrate. After depositing TiN and removing the PS sphere monolayer, the nano-triangular arrays is fabricated on substrate as shown in Fig. 4(b-d), where the thickness is about 120 nm, and the diameters are about 430 nm, 600 nm, and 880 nm respectively. The results show that triangle nanoparticle array have a highly ordered arrangement although there were some disorders and point defect, which is unavoidable and should be attributed to the self-assembled process in micro-scale. The complete absorber processing is achieved after depositing 150 nm SiO2 and 150 nm TiN on the triangular nanoparticle array orderly. Figure 4(e) presents the measured absorbance of absorber from 900 nm to 1500 nm at normal incidence. All of the TiN-absorber samples own a large absorbance (> 85%) in a broad waveband of from 900 nm to 1500 nm, which is far larger than that from Ag, Au-absorbers samples. The TiN-absorber (D = 880 nm) have a maximum absorbance reaching to 96% at the wavelength of 1064 nm. This extremely-high absorptivity is anticipated for LDF to improve the utilization efficiency of laser energy.

 figure: Fig. 4.

Fig. 4. (a) The SEM image of a section of mono-layers closely packed PS array with a diameter of 430 nm. (b)-(d) The SEM images of TiN triangle nanoparticle array samples, where the diameters of the removing PS are 430, 600, and 880 nm respectively. (e) The measured absorption spectra of TiN absorbers with using different diameters PS and Au, Ag absorber using PS with D = 600 nm.

Download Full Size | PDF

3.3 Laser-induced plasma temperature in the ablating layer

The time-resolved emission spectra of the absorber improved-ablated layer are measured to understand the temporal revolutions of the plasmas, and the measured schematic diagram for the transient optical emission spectra is depicted in Fig. 5(a), where the injected laser energy was 45 mJ, and the spectra emitted from the generated plasmas are collected by a fiber and analyzed by a high-speed spectrometer. The measured emission spectra at different delay times of 0.0, 0.5, 1 and 1.5 µs for different absorbers are illustrated in Fig. 5(b-e). The emission intensities of atomic spectral lines in the presence of the absorbers are initially higher than those in the pure Al foil, especially for the TiN absorber with the highest absorptivity. The emission intensity decreases regularly with the delay time in the pure Al foil sample. However, all of the absorber samples show an increase first at the beginning 0.5 µs for RPMA and 1 µs for Ag PMA and then decrease in the emission intensity, indicating a clear energy transferring process from PMA to the Al foil.

 figure: Fig. 5.

Fig. 5. (a) Schematic diagram of the experimental setup for the optical emission spectroscopy. (b)-(d) Time-resolved emission spectra of Al foil, Ag absorber and TiN absorber samples in the wavelength ranges of 390-400 nm with different delay time. (e) The calculated electron temperature in plasma in terms of the peak intensity profile via the Boltzmann plot method. (f) The calculated electron density.

Download Full Size | PDF

The electron temperature and density are critical parameters to characterize the plasma generated process in the ablating layer. The electron temperature in the hot plasma can be calculated using the Boltzmann plot method by the following equation [32]

$$\ln \left( {\frac{{{I_1}}}{{{I_2}}}} \right) = \ln \left( {\frac{{{A_1}{g_1}{\lambda_2}}}{{{A_2}{g_2}{\lambda_1}}}} \right) + \frac{1}{{KT}}({{E_2} - {E_1}} ), $$
where I1 and I2 are the emission intensity of the specified emission lines, A1 and A2 are the transition probability of electrons located on the specified energy levels, g1 and g2 are the statistical weight of electrons located on the specified upper levels, λ1 and λ2 are the emission wavelengths, and E1 and E2 are the energy of excited state. The parameters of Ai, gi, and Ei used in this study for the wavelengths of 308.2 nm, 309.3 nm, 394.4 nm, and 396.2 nm are shown in Table 1.

Tables Icon

Table 1. The physical parameters of Al(I) with emission spectrum of laser-induced plasma.

In addition, the electron densities can be calculated according to the Stark broadening mechanisms and the Saha–Eggert equation [33]

$${\lambda _{FWHM}} = 2\omega \left( {\frac{{{n_e}}}{{{{10}^{16}}}}} \right), $$
where λFWHM is the full width at half maximum (FWHM) of the Stark broadening lines, ω is the electron Stark broadening parameters, and ne is the electron density. The parameters of ω are shown in Table 2.

Tables Icon

Table 2. Stark broadening parameters tables for Al I lines at 396.2 nm.

Figure 5(e) shows the calculated electron temperature of plasma. The electron temperature from the Al foil sample is the highest at the beginning, but decreases from 5707 K to 2837 K in the beginning 1.5 µs. In contrast, the electron temperature of TiN-absorber sample increases from 4796 K to 7511 K at 0.5 µs, and then decrease to 5905 K at 1.5 µs. The Ag-absorber sample shows a similar trend of the election temperature to the RPMA, expect for the the maximum value of 4480 K happening at 1 µs. This changing trend in electron temperature is also found in the transient electron density. As shown in Fig. 5(f), the electron density from the TiN and Ag absorbers increases at the beginning 1 µs and then decrease, while the electron density from the Al foil sample decreases regularly. The maximum electron density of about 1.04 × 1016 cm-3 is achieved in the TiN absorber samples at 1 µs, which is higher than that from the Ag-absorber samples (9 × 1015 cm-3), and that from the Al foil samples (4.1 × 1015 cm-3). Thus, it can be expected that the higher maximum electron temperature and density and longer plasma holding time will lead to high flyer speed for the RPMA-improved LDF.

3.4 Transient velocity of flyer in the accelerated chamber

The transient flyer speed is measured by the photonic Doppler velocimetry system shown in Fig. 6(a). The measured flyer speeds from the pure Al foil LDFs and the absorber improved LDFs were shown in Fig. 6(b). The flyer velocity increases sharply at the beginning 5 ns. In the following 45 ns, the flyer speed increases slowly and reaches the maximum speed. The terminal velocity of all the PMA-improved LDFs are higher than the pure Al foil sample, indicating that improving the absorbance can effectively enhance the final speed of the flyers. This enhanced mechanism can be further verified by the flyers from RPMA-improved LDFs with different PS sphere diameters D, where both the absorbance and the final flyer speed increase from 1730 m/s to 1830 m/s and 1923 m/s as D increases from 430 nm to 600 nm and 880 nm, respectively. It should be noted that the flyer speed from the RPMA-improved LDFs is higher than that of both Ag- and Au-absorber(1474 m/s) simproved LDFs, even for the RPMA-improved LDF with D = 430 nm. Thus, we achieve a maximum speed of 1923 m/s from the RPMA-improved LDFs with D = 880 nm, which is about 1.32 times larger than that from both of the Au- and Ag-PMA improved LDF, and 1.76 times than that from the Al foil LDFs.

 figure: Fig. 6.

Fig. 6. (a) The velocity measurement schematic diagram of the PMA enhanced flyers, where a photonic Doppler velocimetry method is used. (b) and (c) The measured transient velocity and calculated acceleration of the flyer in the chamber from the flyers of different structural parameters and materials.

Download Full Size | PDF

To reveal the the detail accelerated process, the acceleration speed from the samples are extracted from the speed curves shown in Fig. 6(b), and summarized in Fig. 6(c). It is obvious that all of the flyers show a very high acceleration speed at the beginning 0.1 ns, and then the acceleration speed drops slowly in the following 10 ns, and keeps almost unchanged at the beginning 50 ns, where the acceleration process for all of the flyers is completed in 50 ns. When comparing with the Al foil LDFs, the acceleration speeds from all of the PMA-improved LDFs own a larger value at the beginning 10 ns and also a longer maintaining time on high values. Especially for the RPMA-improved LDFs with D = 880 nm, the acceleration speed at the beginning 10 ns reach to about 2.43 × 107 m/s2, which is much larger than that of Al foil sample (1.5 × 107 m/s2). When combing with the previous electron temperature and density evolution curves, we can get that the RPMA-improved LDF can absorb more laser energy to generate high temperature and pressure plasma for generating a greater driving force for flyers.

3.5 Impact characteristics of flyers

In practical applications, especially in the field of initiation, the kinetic energy of flyers is the most important factor to evaluate the LDF performance. But the flyer mass is very difficult to be measured in experiment due to the uncertain ablation mass and the incomplete flyers. To estimate the performance of flyers, the impact test was implemented, where a cylindrical Teflon was placed tightly behind the accelerator chamber as the impacted object. The crater profiles impacted by different LDFs are shown in Fig. 7(b-e). The crater is near circular with the outer diameter of about 0.4 mm, which is smaller than the diameter of accelerated chamber (0.8 mm), indicating that the flyer is deformed in the flying process. The relationship between impact depth and flyer energy follows the Empirical formula [34]

$$d = a{E^b}({a,b > 0} ), $$
where d is maximum depth, E is the energy of flyer and a, b is constant which depends on the shape of the flyer. It is obvious that the depth of craters impacted by TiN, Ag, Au-absorber improved flyer is 57, 38, 50 µm, which is larger than that of Al flyers (35 µm), meaning that the flyer launched by RPMA-improved LDFs obtains the maximum kinetic energy when comparing with all of other LDFs. This result further verifies the superiority of the RPMA-improved LDFs.

 figure: Fig. 7.

Fig. 7. (a) Schematic diagram of the flyer impacting Teflon. (b)-(e) Depth profile across the crater diameter of flyer impacting by the Al foil and Ag, Au, and TiN absorber improved LDF, respectively. The insert images are the corresponding damage morphology.

Download Full Size | PDF

4. Conclusion

In this paper, we have designed and experimentally demonstrated high performance LDFs based on RPMA through the low-cost and large area colloid lithography and the electron beam vacuum deposition technologies. The fabricated RPMAs show a large absorbance of ∼95% in the NIR wavebands. The in-depth analysis reveals the absorbance mechanism that the surface plasmon resonance is excited in RPMA under normal incidence, and the strong absorption happens on the corner and edge of the TiN triangle structure and the TiN layer. The Laser-induced breakdown spectra results indicate that the RPMA can effectively improve the ablating efficiency, and result in high electron temperature and density for the generating plasma. This high temperature and high pressure plasma brings a large acceleration speed with long maintaining time, and thus the final high speed of 1920 m/s, which is larger than the Al foil LDFs and the Au- and Ag-PMA improved LDFs. These results are further confirmed by the impacting experiments, where the flyers launched by the RPMA-improved LDFs own the largest kinetic energy. The optical properties of absorber, the transient electron temperature and density, the velocity and acceleration, the impact feature are systematically investigated in this work. We believe that the high-performance make the RPMA-improved LDF as a competitive candidate for future applications.

Funding

National Natural Science Foundation of China (11604227); International Visiting Program for Excellent Young Scholars of SCU (20181504); International Science and Technology Innovation Cooperation of Sichuan Province (21GJHZ0230).

Disclosures

The authors declare that they have no competing interests.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. M. D. Bowden and M. P. Maisey, “Determination of critical energy criteria for hexanitrostilbene using laser-driven flyer plates,” Proc. SPIE 7070, 707004 (2008). [CrossRef]  

2. M. W. Greenaway and J. E. Field, “The development of a laser-driven flyer system,” AIP Conf. Proc. 706, 1389–1392 (2004). [CrossRef]  

3. W. Guo, L. Z. Wu, N. B. He, S. J. Chen, W. Zhang, R. Q. Shen, and Y. H. Ye, “Efficiency relationship between initiation of HNS-IV and nanosecond pulsed laser-driven flyer plates of layered structure,” Laser Part. Beams 66, 29–40 (2018). [CrossRef]  

4. R. Roybal, C. Stein, C. Miglionico, and J. Shively, “Laboratory simulation of hypervelocity debris,” Int. J. Impact Eng. 17(4-6), 707–718 (1995). [CrossRef]  

5. T. Liu, Z. Zeng, X. Zhang, X. Qiu, Z. A. Cheng, L. Wang, S. Jia, and J. Cai, “Performance of polyimide film under hypervelocity impact of micro flyer: Experiments and simulations,” Acta Astronaut. 159, 452–470 (2019). [CrossRef]  

6. D. D. Mallick, M. Zhao, J. Parker, V. Kannan, B. T. Bosworth, D. Sagapuram, M. A. Foster, and K. T. Ramesh, “Laser-driven flyers and nanosecond-resolved velocimetry for spall studies in thin metal foils,” Exp. Mech. 59(5), 611–628 (2019). [CrossRef]  

7. M. D. Knudson, D. L. Hanson, J. E. Bailey, C. A. Hall, J. R. Asay, and C. Deeney, “Principal Hugoniot, reverberating wave, and mechanical reshock measurements of liquid deuterium to 400 GPa using plate impact techniques,” Phys. Rev. B 69(14), 144209 (2004). [CrossRef]  

8. S. Duwal, C. A. McCoy, P. F. Weck, P. Kalita, H. L. Hanshaw, K. Cochrane, T. Ao, and S. Root, “High-precision equation of state data for TiO2: A structural analog of SiO2,” Phys. Rev. B 102(2), 024105 (2020). [CrossRef]  

9. M. D. Knudson, J. R. Asay, and C. Deeney, “Adiabatic release measurements in aluminum from 240- to 500-GPa states on the principal Hugoniot,” J. Appl. Phys. 97(7), 073514 (2005). [CrossRef]  

10. L. Wang, M. T. Ai, W. L. Zhang, H. C. Jiang, Y. Wang, W. Z. Qin, L. Zhang, X.B. Ji, Y. Li, and D. Tang, “MgAl hydride films for enhanced energy conversion efficiency of a laser-driven flyer,” Adv. Eng. Mater. 23(2), 2000745 (2021). [CrossRef]  

11. S. Watson, M. J. Gifford, and J. E. Field, “The initiation of fine grain pentaerythritol tetranitrate by laser-driven flyer plates,” J. Appl. Phys. 88(1), 65–69 (2000). [CrossRef]  

12. H. M. Wang and Y. L. Wang, “Laser-driven flyer application in thin film dissimilar materials welding and spalling,” Opt. Laser Eng. 97, 1–8 (2017). [CrossRef]  

13. N. Ozaki, M. Koenig, A. Benuzzi-Mounaix, T. Vincil, A. Ravasio, M. Esposito, S. Lepape, E. Henry, G. Hüser, K.A. Tanaka, W. Nazarov, K. Nagai, and M. Yoshida, “Laser-driven flyer impact experiments at the LULI 2000 laser facility,” J. Phys. IV France 133, 1101–1105 (2006). [CrossRef]  

14. S. J. Chen, R. Q. Shen, L. Z. Wu, Y. H. Ye, and Y. Hu, “Laser driven performance of a multilayer flyer with carbon absorption layer,” Proc. SPIE 9543, 95430W (2015). [CrossRef]  

15. H. R. Brierley, D.M. Williamson, and T. A. Vine, “Improving laser-driven flyer velocity with high absorptance layers,” AIP Conf. Proc. 1426(1), 315–318 (2012). [CrossRef]  

16. X. Zhou, R. Q. Shen, Y. H. Ye, P. Zhu, Y. Hu, and L. Z. Wu, “Influence of Al/CuO reactive multilayer films additives on exploding foil initiator,” J. Appl. Phys. 110(9), 094505 (2011). [CrossRef]  

17. W. Guo, L. Z. Wu, N. X. Meng, Y. R. Chen, Z. P. Ma, X. Zhou, W. Zhang, R. Q. Shen, and Y. H. Ye, “Optimisation of modulation period of TiO2/Al reactive multilayer films for laser-driven flyer plates,” Chem. Eng. J. 360, 1071–1081 (2018). [CrossRef]  

18. L. Wang, Y. C. Yan, M. T. Ai, W. L. Zhang, H. C. Jiang, W. Z. Qin, Y. Wang, L. Zhang, and D. Tang, “Promoted plasma generation for laser-driven flyer by MgAl hydride films,” Vacuum 184, 109949 (2021). [CrossRef]  

19. N. I. Landy, S. Sajuyigbe, J. J. Mock, D. R. Smith, and W. J. Padillal, “Perfect metamaterial absorber,” Phys. Rev. Lett. 100(20), 207402 (2008). [CrossRef]  

20. R. M. H. Bilal, S. Zakir, M. A. Naveed, M. Z., M. Q. Mehmood, and Y. Massoud, “Nanoengineered nickel-based ultrathin metamaterial absorber for the visible and short-infrared spectrum,” Opt. Mater. Express 13(1), 28–40 (2023). [CrossRef]  

21. S. Mehrabi, R. M. H. Bilal, M.A. Naveed, and M. M. Ali, “Ultra-broadband nanostructured metamaterial absorber based on stacked square-layers of TiN/TiO2,” Opt. Mater. Express 12(6), 2199–2211 (2022). [CrossRef]  

22. M. A. Naveed, R.M. H. Bilal, M. A. Baqir, M. M. Bashir, M. M. Ali, and A. A. Rahim, “Ultrawideband fractal metamaterial absorber made of nickel operating in the UV to IR spectrum,” Opt. Express 29(26), 42911–42923 (2021). [CrossRef]  

23. X. B. Ji, W. Z. Qin, X. N. Wu, Y. Wang, F. F. Gao, L. Wang, and Y. D. Hou, “Perfect metamaterial absorber improved laser-driven flyer,” Nanophotonics 10(10), 2683–2693 (2021). [CrossRef]  

24. L. Wang, Y. C. Yan, X. B Ji, W. L. Zhang, H. C. Jiang, W. Z. Qin, Y. Wang, and D. Tang, “Improving the energy conversion efficiency of a laser-driven flyer by an in-situ fabricated nano-absorption Layer,” Nanoscale Res. Lett. 15(1), 125 (2020). [CrossRef]  

25. U. Guler, J. C. Ndukaife, G. V. Naik, A. G. A. Nnanna, A. V. Kildishev, V. M. Shalaev, and A. Boltasseva, “Local heating with lithographically fabricated plasmonic titanium nitride nanoparticles,” Nano Lett. 13(12), 6078–6083 (2013). [CrossRef]  

26. W. Li, U. Guler, N. Kinsey, G.V. Naik, A. Boltasseva, J. G. Guan, V. M. Shalaev, and A. V. Kildishev, “Refractory plasmonics with titanium nitride: broadband metamaterial absorber,” Adv. Mater. 26(47), 7959–7965 (2014). [CrossRef]  

27. J. T. Hu, X. C. Ren, A. N. Reed, T. Reese, D. Rhee, B. Howe, L. J. Lauhon, A. M. Urbas, and T. W. Odom, “Evolutionary design and prototyping of single crystalline titanium nitride lattice optics,” ACS Photonics 4(3), 606–612 (2017). [CrossRef]  

28. C. D. Tang, F. L. Chen, J. L. Du, and Y. D. Hou, “Large-area cavity-enhanced 3D chiral metamaterials based on angle-dependent deposition technique,” Nanoscale 12(16), 9162–9170 (2020). [CrossRef]  

29. E. D. Palik, Handbook of Optical Constants of Solids (Academic Press, 1985).

30. P. B. Johnson and R. W. Christy, “Optical constants of the noble metals,” Phys. Rev. B 6(12), 4370–4379 (1972). [CrossRef]  

31. X. Guo, C. Liu, and H. Ong, “Generalization of the circular dichroism from metallic arrays that support Bloch-like surface plasmon polaritons,” Phys. Rev. Appl. 15(2), 024048 (2021). [CrossRef]  

32. G. Bekefi, “Principles of laser plasmas,” J. Electrochem. Soc. 124(12), 435C (1977). [CrossRef]  

33. M. Hanif, M. Salik, and M. A. Baig, “Quantitative studies of copper plasma using laser induced breakdown spectroscopy,” Opt. Laser Eng. 49(12), 1456–1461 (2011). [CrossRef]  

34. M. J. Burchell and E. Johnson, “Impact craters on small icy bodies such as icy satellites and comet nuclei,” Mon. Not. R. Astron. Soc. 360(2), 769–781 (2005). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained 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 (7)

Fig. 1.
Fig. 1. (a) The schematic diagram of the RPMA-improved LDF, include a confining window (sapphire substrate), the ablating layer (TiN absorber), the flyer layer (Al foil) and an accelerated chamber (barrel). (b) The fabrication and packaging processing of the integrated flyer, the fabrication of PMA by the colloid lithography and thin-film deposition technique firstly, and then the Al foil and the accelerated chamber are packed closely behind the PMA orderly.
Fig. 2.
Fig. 2. (a) The perspective view, top view and side view of a unit cell of the absorber. The simulated absorption spectra of the RPMA samples with varying different structure parameters are shown in (b) diameter of PS sphere, D; (c) thickness of the silver triangular array, t; (d) thickness of the SiO2, m; (e) the refractive index of the dielectric medium, n. The inset dotted line marks the interface of the diffraction order. (f) Absorption spectrum of TiN, Ag and Au absorber with different D, where the optimum parameters are t = 150 nm, m = 250 nm, b = 150 nm.
Fig. 3.
Fig. 3. Calculated electric field distribution on the cross section of RPMA at the wavelength of 1064 nm: (a) the X-Y plane (triangular array -SiO2 spacer interface), (c) the Y-Z plane (X = 0). The electromagnetic energy dissipation distribution: (b) and (f) the X-Y plane, (d) the Y-Z plane. (e)-(i) Corresponding calculation results of Ag PMA.
Fig. 4.
Fig. 4. (a) The SEM image of a section of mono-layers closely packed PS array with a diameter of 430 nm. (b)-(d) The SEM images of TiN triangle nanoparticle array samples, where the diameters of the removing PS are 430, 600, and 880 nm respectively. (e) The measured absorption spectra of TiN absorbers with using different diameters PS and Au, Ag absorber using PS with D = 600 nm.
Fig. 5.
Fig. 5. (a) Schematic diagram of the experimental setup for the optical emission spectroscopy. (b)-(d) Time-resolved emission spectra of Al foil, Ag absorber and TiN absorber samples in the wavelength ranges of 390-400 nm with different delay time. (e) The calculated electron temperature in plasma in terms of the peak intensity profile via the Boltzmann plot method. (f) The calculated electron density.
Fig. 6.
Fig. 6. (a) The velocity measurement schematic diagram of the PMA enhanced flyers, where a photonic Doppler velocimetry method is used. (b) and (c) The measured transient velocity and calculated acceleration of the flyer in the chamber from the flyers of different structural parameters and materials.
Fig. 7.
Fig. 7. (a) Schematic diagram of the flyer impacting Teflon. (b)-(e) Depth profile across the crater diameter of flyer impacting by the Al foil and Ag, Au, and TiN absorber improved LDF, respectively. The insert images are the corresponding damage morphology.

Tables (2)

Tables Icon

Table 1. The physical parameters of Al(I) with emission spectrum of laser-induced plasma.

Tables Icon

Table 2. Stark broadening parameters tables for Al I lines at 396.2 nm.

Equations (4)

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

λ m n = n m e d [ ( m / P x ) 2 + ( n / P y ) 2 ] 1 / 2 ,
ln ( I 1 I 2 ) = ln ( A 1 g 1 λ 2 A 2 g 2 λ 1 ) + 1 K T ( E 2 E 1 ) ,
λ F W H M = 2 ω ( n e 10 16 ) ,
d = a E b ( a , b > 0 ) ,
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.