Size-Selected Submicron Gold Spheres: Controlled Assembly onto

6 days ago - ... show the synthesis and assembly of size-selected submicron gold spheres using pulsed laser ablation of a gold plate in a supercritica...
0 downloads 0 Views 5MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega XXXX, XXX, XXX−XXX

http://pubs.acs.org/journal/acsodf

Size-Selected Submicron Gold Spheres: Controlled Assembly onto Metal, Carbon, and Plastic Substrates Ken-ichi Saitow,†,‡,* Yoshinori Okamoto,‡ and Hidemi Suemori‡ †

Natural Science Center for Basic Research and Development (N-BARD), and ‡Department of Chemistry, Graduate School of Science, Hiroshima University, 1-3-1 Kagamiyama, Higashi-hiroshima 739-8526, Japan

Downloaded via 91.200.82.58 on August 28, 2019 at 21:51:13 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Size-selected submicron spheres become very useful building blocks if the spheres could be synthesized and integrated at any desired position. In particular, spheres having a similar size to visible-light wavelength have attracted much attention. Here, we show the synthesis and assembly of size-selected submicron gold spheres using pulsed laser ablation of a gold plate in a supercritical fluid. Four findings were obtained in the study. Submicron spheres with a narrow size distribution were generated, and the polydispersity was ≈ 6%. The average diameter was controlled from 600 to 1000 nm. A thermodynamic condition for scalable synthesis was found. The assembly of spheres onto a metal, carbon, or plastic substrate was accomplished.

1. INTRODUCTION Nobody doubts that gold as a material has attracted considerable attention in chemical and material sciences and industrial applications.1 In particular, gold nanomaterials are utilized as substrates for single-molecule detection by surfaceenhanced Raman scattering2 and enhanced fluorescence,3 as optomaterials for ultrasensitive biosensors and medical sensors,4 as catalysts for chemical reactions such as CO oxidation,5 and as photothermal materials for cancer therapy.6 All of these properties are accomplished by nanomaterials with a size < 50 nm, which are synthesized by a conventional chemical synthesis method: reduction of gold (Au) ions in HAuCl4 aqueous solution.7 However, it has been difficult to produce a size-selected metal particle larger than 100 nm. Specifically, submicron particles could become very useful building block materials if submicron-sized particles with a homogeneous size could be synthesized and integrated at any desired position, for example, photonic crystals matching the visible and near-infrared (NIR) light, nanogap electrodes with a special smooth surface, field enhancement driven by high scattering efficiency, and parts of a molecular circuit. Pulsed laser ablation (PLA), a physical synthesis method, is a promising approach used to obtain stable nanoparticles. This method consists of a one-step process conducted at room temperature with a short duration, for example, a few minutes to an hour.8−14 In addition, scalable synthesis of nanoparticles, several grams/hour as products, has been recently realized.11d,e PLA of Au in solution has been extensively investigated in the last decade.11a,b Recently, submicron-sized spherical Au particles were synthesized by PLA.12e,13 We have developed a novel method for nanoparticle synthesis by conducting PLA in a supercritical fluid,12a−f as shown in Figures 1 and S1. Specifically, thermal and dielectric properties of a surrounding medium can be easily tuned by changing the fluid density and pressure. Thus, light-emitting silicon nanocrystals were © XXXX American Chemical Society

Figure 1. Schematic diagram of the system used for nanoparticle generation by laser ablation in supercritical fluids. Photograph represents the high-pressure vessel for preparing supercritical fluids. Photograph courtesy of Saitow. Copyright 2019.

generated with photoluminescence color (RGB12b and whitelight continuum12c) that can be controlled by fluid pressure and/or density during PLA. The morphology of Au nanoparticles was also changed by the cooling rate or permittivity of fluids during PLA.12d,e Furthermore, PLA in a supercritical fluid becomes a popular method to synthesize various nanomaterials, recently.14 Here, we show submicron Au spheres with a narrow size distribution, that is, deviations of diameters ranging 5−8%, using PLA laser synthesis in a supercritical fluid. The average diameter of particles ranged from 600 to 1000 nm, whose size matches the light wavelength in the visible and NIR regions. In Received: July 2, 2019 Accepted: August 2, 2019

A

DOI: 10.1021/acsomega.9b01999 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

in a supercritical fluid,12f submicron-sized Au particles are generated within the time shorter than 300 ns after laser irradiation. Thermodynamic calculations indicated that the generation process is attributed to the solidification of hot liquid droplets, by which the lowest surface free energy is obtained with a spherical shape, within a few 100 ns.12d A cavitation bubble, in which nanoparticles are generated, emerges within submicroseconds on the irradiated surface of the target in liquids and high-pressure liquids, according to time-resolved shadowgraph and light scattering measurements.17 On the basis of these experimental evidences in a supercritical fluid and the observations by in situ time-resolved spectroscopy, it was considered that a concerted process of evaporation and cooling of Au liquid droplets are responsible for the particle size. Specifically, the evaporation and cooling rates of the gold liquid droplet in which the cavitation bubble adiabatically expands and shrinks in the high-pressure fluid are key factors for the characterization of the particle size by PLA at high pressures. The amounts of particles produced as a function of fluid pressure/density during PLA were investigated. Figure 3a−d shows the typical SEM images of Au spherical particles fabricated with four representative densities. A significant dependence on the fluid density is observed; a large amount of gold nanospheres was generated by PLA at a high fluid density, whereas that generated at a fluid density lower than ρr = 0.7

addition, a thermodynamic condition for scalable synthesis was found. Furthermore, the integration of gold spheres onto a metal, carbon, or plastic film was accomplished and optimized.

2. RESULTS AND DISCUSSION Figure 2 shows scanning electron microscopy (SEM) images of submicron-sized Au spherical particles generated by PLA at

Figure 2. SEM images of submicron spherical Au particles generated by PLA in supercritical CHF3 at a reduced temperature Tr = T/Tc = 1.02. The Au particles are generated at a reduced density ρr = ρ/ρc of (a) 0.7, (b) 1.7, and (c) 1.9. Diameter D ranges over 600−1200 nm: (a) D(ρr = 0.7) = 890 ± 55 nm, (b) D(ρr = 1.7) = 710 ± 40 nm, and (c) D(ρr = 1.9) = 1050 ± 80 nm. The plus−minus values denote the standard deviation, σ, from the average diameters. The deviations of particle sizes are in the 5−8% range of the average diameters.

three reduced densities ρr = ρ/ρc, where ρc is the critical density, of supercritical trifluoromethane (CHF3) (ρr = 0.7, 1.7, and 1.9). Many submicron-sized spherical Au particles are generated, and the particle size is very uniform. This phenomenon has not been observed by PLA in other fluids, for example, CO2 and SF6. The SEM analysis from each sample provided the average diameter D, of the spherical particles, and the standard deviation, σ. D ranged over 700−1000 nm: D(ρr = 0.7) = 890 ± 55 nm, D(ρr = 1.7) = 710 ± 40 nm, and D(ρr = 1.9) = 1050 ± 80 nm. Their size distributions are shown in Figure S2.15,16 Here, the plus−minus values indicate the standard deviation σ from the average diameters, and the dispersity of particle size is evaluated as the value σ/average. Note that this dispersity of particle sizes is small and estimated as 6%, ranging from 5 to 8%. Here, let us discuss the particle size dependence on the fluid density during PLA, briefly. According to the transient absorption spectra monitoring PLA

Figure 3. SEM images of submicron spherical Au particles generated by PLA in supercritical CHF3 at a reduced temperature, Tr = T/Tc = 1.02. Particles were fabricated at a reduced density ρr = ρ/ρc of (a) 0.3, (b) 0.7, (c) 0.9, and (d) 1.9. Corresponding pressures are denoted in the upper axis. The particles were integrated into the micropores. The number of particles increases with the fluid density during PLA. (e) Numbers of Au particles produced at various ρr. B

DOI: 10.1021/acsomega.9b01999 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

was significantly decreased. To quantify the dependence of the amount of particles on the fluid density, the number of particles was estimated both for the number density of particles and the particle-deposited area in the SEM image. Figure 3e shows the number of particles as a function of reduced densities during PLA. Many particles were generated by PLA at ρr ≥ 0.7, whereas few particles were generated at ρr < 0.7. Thus, Figure 3e reveals that the thresholds of fluid density and pressure for the generation of particles are ρr = 0.7 and P = 5.3 MPa, respectively. From the Supporting Information, it is noted that many gold nanonetworks, composed of smaller nanospheres with 30 nm diameter, are generated by PLA at a lower density of supercritical CHF3 but disappear at a higher density of ρr = 0.7 (Figures S3 and S4). The density for the emergence of submicron particles is in good agreement with that for the disappearance of the nanonetworks, that is, liquid gold droplets as precursors are fragmented to form smaller nanospheres at lower densities giving the gold nanonetwork, whereas at higher densities, the solidification of liquid droplets produces submicron-sized particles. This precursor model of gold liquid droplets has been confirmed by measuring the amounts/laser pulse of both spherical particles and nanonetworks generated by PLA in supercritical CO2.12d Briefly, the amount of gold nanonetworks increased, whereas that of spherical particles with diameter ≅ 800 nm simultaneously decreased, as the number of laser pulses for PLA was increased. The density dependence of the amount of spherical particles is attributed to the branching ratio, which determines whether the liquid gold droplet solidifies or fragments to the nanonetwork.12e Thus, the threshold density for solidification via cooling was ρr = 0.7. Another distinct feature evident in Figures 3 and S3 is that almost all the synthesized particles are not present on a flat region of the substrate but are selectively collected in a pore of size ≈ 100 μm on the substrate. Briefly, let us describe the mechanism of collection of particles into the pore, and the details are given elsewhere (vide infra). As a first step, particles are synthesized in supercritical CHF3. In this step, the particles disperse in the supercritical fluid. As a second step, the pressure of supercritical CHF3 is decreased to retrieve the synthesized particles from the high-pressure vessel to the air. In this process, large particles with the size of 600−1000 nm do not disperse in the gas but sediment into the liquid region of the substrate, which is located at the bottom of the vessel. As a third step, all liquid region evaporates by decreasing further pressure, and the concentration of particle in the liquid becomes higher. The final liquid involving many spherical particles drops into a pore, which is located at the lower position of the substrate. To further investigate the collection in the pore, we analyzed the amounts of particles by preparing a brass substrate with pores of different depths. Figure 4a,b show the top and side views of the substrate with the micropores. The pore diameter was set to 300 μm and the depth ranged from 50 to 1000 μm. The typical cross sections of pores measured by a laser microscope are also displayed in Figure 4c,d. PLA of a gold plate substrate in supercritical CHF3 was conducted at a pressure of 5.8 MPa, which corresponds to ρr = 1.25, and the generated particles in the pores were observed using a scanning electron microscope. The amount of particles was counted and the results are shown in Figure 4d as a function of pore depth. These values correspond to the summation of the numbers of particles synthesized by the experiments performed five

Figure 4. Laser microscope images of the (a) top and (b) side of the brass substrate used to collect submicron spherical Au particles. There are eight different pores with the size of 500 μm and different depths. Laser microscope images of the cross sections of pores with the depths of (c) 50 μm and (d)1000 μm. (e) Number of submicron spherical Au particles in pores as a function of pore depth. Particles were fabricated at ρr = ρ/ρc = 1.25.

times.18 The maximum integration of particles into a pore is observed when the pore depth is 400 μm. Here, we describe the integration processes of submicron particles into the pore in detail (vide supra). Supercritical CHF3 with a pressure of 5.8 MPa was released to retrieve the substrate from the highpressure vessel after the synthesis, during which the fluid temperature decreased from 305 to 285 K with the adiabatic expansion of the high-pressure fluid. Thus, the temperature T of CHF3 became lower than the critical temperature19 (Tc = 299.3 K), that is, tentative temperature ≈ 10 °C, so that the supercritical CHF3 became a liquid. Liquid CHF3 was successively evaporated until its pressure was equal to the pressure of the atmosphere. These successive phase transition processes caused liquid CHF3 to settle at the bottom of the vessel. The final liquid, including many Au particles, was collected in the pores of the substrate, and many particles were integrated in the bottom of the pores. Multiple concerted factors, such as viscosity and surface tension depending on the pressure and temperature of CHF3, wettability of the substrate, mass transfer driven by capillary flow,20,21 and pinning and depinning at the interface between the liquid and the substrate22 can also govern the collection efficiency. Note that the integration into the pore was also established using the substrate made of other materials, such as stainless steel (SUS), copper, carbon, and polyethylene terephthalate (PET). The data are displayed in Figure 5. Thus, the pores in these substrates can also collect many submicron particles.

3. CONCLUSIONS In summary, nanosecond PLA of an Au plate was conducted in supercritical CHF3. Many spherical Au particles with diameters from 600 to 1000 nm were synthesized over the threshold density and/or pressure, that is, ρr = 0.7 and P = 5.3 MPa. The particle diameter was controlled according to the density and/ or pressure used during PLA, and the polydispersity was as low C

DOI: 10.1021/acsomega.9b01999 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

where vt is the sedimentation velocity depending on the fluid density, z is the depth, t is the time, ρ is the density of Au, ρ0 is the fluid density, D is the particle diameter, g is the gravitational constant, and η is the fluid viscosity depending on the fluid density. The sedimentation time was adjusted to the amount of time required for a 100 nm diameter sphere to sink to a depth of 1 cm. The same sedimentation time was applied to every experiment to prepare SEM samples for the evaluation of the amount of Au nanoparticles according to the pressure and/or density during PLA. The critical constant of CHF3 is reported to be Tc = 299.3 K, Pc = 4.83 MPa, and ρc = 0.527 g cm−3. We represent the density of CHF3 by the reduced density ρr = ρ/ρc, as shown in Figure S3.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01999. Instrument of supercritical vessel, size analysis of Au particles, SEM images of particles and their amount analysis, and phase diagram of supercritical CHF3 (PDF)

Figure 5. Submicron spherical Au particles generated by PLA in supercritical CHF3 at a reduced temperature Tr = T/Tc = 1.02. The Au particles are generated at a reduced density ρr = ρ/ρc of 1.25 and gathered at the micron-sized pore of each substrate. SEM images of the particles on the substrates of (a) carbon tape, (b) copper, and (c) SUS. Optical microscope image of the particles on the substrate of (d) PET. The scale bar in (d) is 50 μm.



AUTHOR INFORMATION

ORCID

as 6%. The selective integration of particles into various substrates was accomplished by both PLA in a supercritical fluid and successive phase changes from a high-pressure fluid.

Ken-ichi Saitow: 0000-0003-2405-222X

4. EXPERIMENTAL SECTION An instrument to fabricate particles was constructed, each equipment of which is illustrated in Figures 1 and S1. A highpressure vessel, a high-performance liquid chromatography (HPLC) pump, and a gas cylinder were used to prepare the supercritical state, and optics with a Q-switched frequencydoubled Nd:YAG laser were used to conduct PLA. The temperature of the fluid in the vessel was controlled using a set of heaters, a proportional−integral−derivative controller, and a thermocouple. The pressure of the fluid was increased with the HPLC pump. The Nd:YAG laser serves as the PLA light source and was operated at an excitation wavelength of 532 nm, an energy of 19 mJ/pulse, a repetition rate of 20 Hz, a fluence of 0.8 J cm−2, and a pulse width of 8 ns. A gold plate (99.95%, Tanaka Co.) immersed in supercritical fluid (CHF3, 99.995%) is irradiated with the laser for 10 min at the isotherm corresponding to a reduced temperature Tr = T/Tc = 1.02, where Tc is the critical temperature. The pressure for PLA ranged from 3.81 to 14.9 MPa. The fluid density was calculated from the empirical equations of state, using the measured values of P and T. The density ranged from 0.158 to 1.00 g cm−3 and is expressed as 0.3 ≤ ρr = ρ/ρc ≤ 1.9, where ρr and ρc are the reduced and critical densities, respectively, as shown in Figure S5. The Au particles generated were deposited on a substrate immersed in supercritical CHF3, SUS, copper, carbon, and PET. After the sedimentation of the particles in the fluid, the substrate was retrieved from the vessel and examined using a field emission scanning electron microscope (Hitachi N-3400) and a laser microscope (Shimadzu OLS4000). The sedimentation time was calculated using the pressure-dependent viscosity of the fluid12d

ACKNOWLEDGMENTS K.S. acknowledges financial support from the Funding Program for the Next Generation World-Leading Researchers (GR073) of the Japan Society for the Promotion of Science (JSPS), a Grant-in-Aid for Scientific Research (A) (15H02001) and (B) (19H02556) from JSPS, and the PRESTO Structure Control and Function program of the Japan Science and Technology Agency (JST).

vt = dz /dt = ((ρ − ρ0 )D2g )/18η

Notes

The authors declare no competing financial interest.

■ ■

REFERENCES

(1) (a) Daniel, M.-C.; Astruc, D. Gold Nanoparticles: Assembly, Supramolecular Chemistry, Quantum-Size-Related Properties, and Applications toward Biology, Catalysis, and Nanotechnology. Chem. Rev. 2004, 104, 293−346. (b) Saha, K.; Agasti, S. S.; Kim, C.; Li, X.; Rotello, V. M. Gold Nanoparticles in Chemical and Biological Sensing. Chem. Rev. 2012, 112, 2739−2779. (2) Pieczonka, N. P. W.; Aroca, R. F. Single molecule analysis by surfaced-enhanced Raman scattering. Chem. Soc. Rev. 2008, 37, 946− 954. (3) Kinkhabwala, A.; Yu, Z.; Fan, S.; Avlasevich, Y.; Müllen, K.; Moerner, W. E. Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna. Nat. Photonics 2009, 3, 654−657. (4) Qian, X.-M.; Nie, S. M. Single-molecule and single-nanoparticle SERS: from fundamental mechanisms to biomedical applications. Chem. Soc. Rev. 2008, 37, 912−920. (5) (a) Haruta, M.; Daté, M. Advances in the catalysis of Au nanoparticles. Appl. Catal., A 2001, 222, 427−437. (b) Stratakis, M.; Garcia, H. Catalysis by Supported Gold Nanoparticles: Beyond Aerobic Oxidative Processes. Chem. Rev. 2012, 112, 4469−4506. (6) Dreaden, E. C.; Alkilany, A. M.; Huang, X.; Murphy, C. J.; ElSayed, M. A. The golden age: gold nanoparticles for biomedicine. Chem. Soc. Rev. 2012, 41, 2740−2779. (7) Grzelczak, M.; Pérez-Juste, J.; Mulvaney, P.; Liz-Marzán, L. M. Shape control in gold nanoparticle synthesis. Chem. Soc. Rev. 2008, 37, 1783−1791.

(1) D

DOI: 10.1021/acsomega.9b01999 ACS Omega XXXX, XXX, XXX−XXX

ACS Omega

Article

S.; Kato, T.; Sasaki, T.; Terashima, K. Synthesis of higher diamondoids by pulsed laser ablation plasmas in supercritical CO2. J. Appl. Phys. 2011, 109, 123304. (d) Nakahara, S.; Stauss, S.; Miyazoe, H.; Shizuno, T.; Suzuki, M.; Kataoka, H.; Sasaki, T.; Terashima, K. Pulsed Laser Ablation Synthesis of Diamond Molecules in Supercritical Fluids. Appl. Phys. Express 2010, 3, 096201. (e) Koizumi, M.; Kulinich, S. A.; Shimizu, Y.; Ito, T. Slow dynamics of ablated zone observed around the density fluctuation ridge of fluid medium. J. Appl. Phys. 2013, 114, 214301. (f) Urabe, K.; Kato, T.; Stauss, S.; Himeno, S.; Kato, S.; Muneoka, H.; Baba, M.; Suemoto, T.; Terashima, K. Dynamics of pulsed laser ablation in high-density carbon dioxide including supercritical fluid state. J. Appl. Phys. 2013, 114, 143303. (g) Minaev, N. V.; Arakcheev, V. G.; Rybaltovskii, A. O.; Firsov, V. V.; Bagratashvili, V. N. Dynamics of formation and decay of supercritical fluid silver colloid under pulse laser ablation conditions. Russ. J. Phys. Chem. B 2015, 9, 1074. (h) Stauss, S.; Muneoka, H.; Terashima, K. Review on plasmas in extraordinary media: Plasmas in cryogenic conditions and plasmas in supercritical fluids. Plasma Sources Sci. Technol. 2018, 27, 023003. (i) Singh, A.; Salminen, T.; Honkanen, M.; Vihinen, J.; Hyvärinen, L.; Levänen, E. Multiphase TixOy nanoparticles by pulsed laser ablation of titanium in supercritical CO2. Appl. Surf. Sci. 2019, 476, 822−827. (15) To estimate accurate size distributions of particles, we used the particles on the front side on SEM image and did not use the particles on the back sides, because the particles in the back side are seen as smaller size than real size, owing to 3-dimensional configuration. (16) Irani, R. R.; Callis, C. F. Particle Size: Measurement, Interpretation, and Application; Wiley: New York, 1963. (17) (a) Soliman, W.; Takada, N.; Sasaki, K. Growth Processes of Nanoparticles in Liquid-Phase Laser Ablation Studied by Laser-Light Scattering. Appl. Phys. Express 2010, 3, 035201. (b) Lam, J.; Lombard, J.; Dujardin, C.; Ledoux, G.; Merabia, S.; Amans, D. Dynamical study of bubble expansion following laser ablation in liquids. Appl. Phys. Lett. 2016, 108, 074104. (18) After the synthesis of Au particles, the ambient around the particles drastically changes by reducing the fluid pressure. Under this situation, the depth of pore becomes deeper, the repeatability of collection of particles becomes higher, as shown in Figure S6. This is because the spherical particles at the bottom of the deep pore are insensitive to the ambient such as strong convection caused by decreasing fluid pressure, whereas those at shallow pore can detrap from the pore by that convection, as shown Figure 4c,d. (19) Penoncello, S. G.; Lemmon, E. W.; Jacobsen, R. T.; Shan, Z. A Fundamental Equation for Trifluoromethane (R-23). J. Phys. Chem. Ref. Data 2003, 32, 1473−1499. (20) Deegan, R. D.; Bakajin, O.; Dupont, T. F.; Huber, G.; Nagel, S. R.; Witten, T. A. Capillary flow as the cause of ring stains from dried liquid drops. Nature 1997, 389, 827−829. (21) Li, H.; Fowler, N.; Struck, C.; Sivasankar, S. Flow triggered by instabilities at the contact line of a drop containing nanoparticles. Soft Matter 2011, 7, 5116−5119. (22) Hu, H.; Larson, R. G. Analysis of the Microfluid Flow in an Evaporating Sessile Droplet. Langmuir 2005, 21, 3963−3971.

(8) Barcikowski, S.; Devesa, F.; Moldenhauer, K. Impact and structure of literature on nanoparticle generation by laser ablation in liquids. J. Nanopart. Res. 2009, 11, 1883−1893. (9) Barcikowski, S.; Compagnini, G. Advanced nanoparticle generation and excitation by lasers in liquids. Phys. Chem. Chem. Phys. 2013, 15, 3022−3026. (10) (a) Yang, G. Laser Ablation in Liquids: Principles and Applications in the Preparation of Nanomaterials; Pan Stanford Publishing: Singapore, 2012. (b) Saitow, K. Nanoparticle Generation by Laser Ablation in Liquid and Supercritical Fluid; Taylor & Francis Group, 2012; Chapter 12. (11) (a) Amendola, V.; Meneghetti, M. Laser ablation synthesis in solution and size manipulation of noble metal nanoparticles. Phys. Chem. Chem. Phys. 2009, 11, 3805−3821. (b) Zhang, D.; Gökce, B.; Barcikowski, S. Laser Synthesis and Processing of Colloids: Fundamentals and Applications. Chem. Rev. 2017, 117, 3990−4103. (c) Special issue on colloids with lasers, Gökce, B.; Amendola, V.; Barcikowski, S. ChemPhysChem 2017, 18, 983. (d) Sajti, C. L.; Sattari, R.; Chichkov, B. N.; Barcikowski, S. J. Phys. Chem. C 2010, 114, 2421−2427. (e) Ishikawa, Y.; Koshizaki, N. Sci. Rep. 2018, 8, 14208. (12) (a) Saitow, K. Silicon Nanoclusters Selectively Generated by Laser Ablation in Supercritical Fluid. J. Phys. Chem. B 2005, 109, 3731−3733. (b) Saitow, K.; Yamamura, T. Effective Cooling Generates Efficient Emission: Blue, Green, and Red Light-Emitting Si Nanocrystals. J. Phys. Chem. C 2009, 113, 8465−8470. (c) Wei, S.; Yamamura, T.; Kajiya, D.; Saitow, K. White-Light-Emitting Silicon Nanocrystal Generated by Pulsed Laser Ablation in Supercritical Fluid: Investigation of Spectral Components As a Function of Excitation Wavelengths and Aging Time. J. Phys. Chem. C 2012, 116, 3928−3934. (d) Saitow, K.; Yamamura, T.; Minami, T. Gold Nanospheres and Nanonecklaces Generated by Laser Ablation in Supercritical Fluid. J. Phys. Chem. C 2008, 112, 18340−18349. (e) Saitow, K.; Okamoto, Y.; Yano, Y. F. Fractal of Gold Nanoparticles Controlled by Ambient Dielectricity: Synthesis by Laser Ablation as a Function of Permittivity. J. Phys. Chem. C 2012, 116, 17252−17258. (f) Wei, S.; Saitow, K. In situ multipurpose timeresolved spectrometer for monitoring nanoparticle generation in a high-pressure fluid. Rev. Sci. Instrum. 2012, 83, 073110. (g) Kitasako, T.; Saitow, K. Si quantum dots with a high absorption coefficient: analysis based on both intensive and extensive variables. Appl. Phys. Lett. 2013, 103, 151912. (h) Xin, Y.; Nishio, K.; Saitow, K. Appl. Phys. Lett. 2015, 106, 201102. (i) Kajiya, D.; Saitow, K. Si-Nanocrystal/ P3HT Hybrid Film with 50- and 12Fold enhancement of hole mobility and density: films prepared by successive drop casting. Nanoscale 2015, 7, 15780−15788. (j) Xin, Y.; Kitasako, T.; Maeda, M.; Saitow, K. Chem. Phys. Lett. 2017, 674, 90−97. (k) Sakaki, S.; Saitow, K.; Sakamoto, M.; Wada, H.; Swiatkowska-Warkocka, Z.; Ishikawa, Y.; Koshizaki, N. Comparison of picosecond and nanosecond lasers for the synthesis of TiN sub-micrometer spherical particles by pulsed laser melting in liquid. Appl. Phys. Express 2018, 11, 035001. (l) Kajiya, D.; Saitow, K. Si Nanocrystal Solution Stable for One Year. RSC Adv. 2018, 8, 41299−41307. (13) (a) Tsuji, T.; Yahata, T.; Yasutomo, M.; Igawa, K.; Tsuji, M.; Ishikawa, Y.; Koshizaki, N. Preparation and investigation of the formation mechanism of submicron-sized spherical particles of gold using laser ablation and laser irradiation in liquids. Phys. Chem. Chem. Phys. 2013, 15, 3099−3107. (b) Tsuji, T.; Sakaki, S.; Fujiwara, H.; Kikuchi, H.; Tsuji, M.; Ishikawa, Y.; Koshizaki, N. StabilizerConcentration Effects on the Size of Gold Submicrometer-Sized Spherical Particles Prepared Using Laser-Induced Agglomeration and Melting of Colloidal Nanoparticles. J. Phys. Chem. C 2018, 122, 21659−21666. (14) (a) Kuwahara, Y.; Saito, T.; Haba, M.; Iwanaga, T.; Sasaki, M.; Goto, M. Nanosecond Pulsed Laser Ablation of Copper in Supercritical Carbon Dioxide. Jpn. J. Appl. Phys. 2009, 48, 040207. (b) Takada, N.; Machmudah, S.; Goto, H.; Wahyudiono; Goto, M.; Sasaki, K. Characteristics of optical emission intensities and bubblelike phenomena induced by laser ablation in supercritical fluids. Jpn. J. Appl. Phys. 2014, 53, 010213. (c) Nakahara, S.; Stauss, E

DOI: 10.1021/acsomega.9b01999 ACS Omega XXXX, XXX, XXX−XXX