High-productivity synthesis of ligand-free nanoparticles by laser ablation in liquids

Metal and metal oxide nanoparticles are produced on the gram scale with a novel ablation setup that includes a high-power pulsed-laser system, an ultrafast laser scanner, and a customized ablation chamber.
06 January 2017
René Streubel and Bilal Gökce

In typical wet chemical synthesis methods for nanoparticles (such as precipitation and sol-gel techniques) organic adsorbates are used, or marks of the residual reagents are left. Unfortunately, both these factors cause the deactivation of nanoparticle surfaces. That is, the chemisorbed or physisorbed ligands weaken the catalytic activity, interfere with surface-sensitive spectroscopy, and hinder surface-assisted mass spectrometry. In contrast, the physico-chemical synthesis method of pulsed laser ablation in liquid (PLAL) produces ligand-free, highly pure nanoparticles. In recent years, this technique has thus emerged as a way to produce a variety of nanoparticles with unique properties. Research in this field has progressed rapidly with regard to the design of innovative composites1–3 and their application in laser-assisted material processing techniques (e.g., rapid prototyping4 and laser lithography5). In comparison to ligand-coated nanoparticles, the laser-generated nanoparticles exhibit higher conjugation efficiency,6 higher grafting density,7 high electroaffinity toward charged biomolecules,8 and novel preparation routes for heterogeneous catalysts.9 With the current laser-based synthesis methods, however, it is difficult to achieve the large amounts of nanoparticles required for industrial applications.

Purchase Nanotechnology: A Crash Course In PLAL methods, a laser pulse interacts with a target material and ignites a plasma, which in turn leads to the formation of a cavitation bubble (of vapor gas). Nanoparticles form inside this bubble10 and then disperse into the liquid after its collapse. In addition, the cavitation bubble causes the scattering of consecutive laser pulses if they arrive before the bubble collapses (which usually takes about 200 microseconds). The lifetime, as well as the dimensions (diameter of up to 1mm), of the cavitation bubble therefore limit the energy input and the nanoparticle productivity. Accordingly, temporal bypassing of the cavitation bubble is only possible for low (i.e., kHz) repetition-rate laser systems.

In our work, we have therefore performed PLAL with a novel ablation setup (see Figure 1).11 Indeed, we take advantage of recent advances in high-power pulsed-lased systems and laser scanning technologies in our unique upscaling concept for nanoparticle production.12, 13 In particular, our setup consists of a high-power ultrafast laser system (500W, 0.9–5ps, 1.2-40.5MHz, 1030nm), a polygon scanner, and a customized ablation chamber. With the polygon scanner, we scan in the ‘fast axis’—see Figure 1(b)—with an ultrasonic speed of 500m/s. The slow axis is controlled by a galvanometer scanner, which scans the target material at a rate of about 10m/s. We produce nanoparticles as the laser beam is directed onto the target material, which is placed in the water-filled ablation chamber. Furthermore, to enable continuous nanoparticle synthesis, we pump liquid through the chamber at a flow volume of up to 25 liters per hour. With our PLAL concept we can thus spatially bypass the cavitation bubbles with interpulse distances of up to 400μm (at a repetition rate of 1.2MHz), as shown in Figure 1(b).


Figure 1. (a) Schematic illustration of the novel pulsed laser ablation in liquid (PLAL) setup used to produce nanoparticles. The apparatus consists of a high-power laser system coupled with a polygon scanning system (which can reach ultrasonic speeds) in a specially designed liquid flow reactor. vfocus: Scanning speed of the focus. (b) Schematic diagram of the flow-through reactor, showing how the laser pulses spatially bypass the cavitation bubbles with interpulse distances of up to 400μm.

The productivities we achieve with our PLAL system—for several different nanomaterials—are as high as 4g/h (see Figure 2). In addition, we reach 265 times higher ablation rates (with an increase in average laser power of only a factor of 10) compared with previous results recorded in the literature.14 Although the ablation of noble metals (e.g., platinum, gold, and silver) leads to the formation of metal nanoparticles, other metals (such as aluminum, copper, and titanium) react more strongly with the solvent in our approach and form oxide and suboxide nanoparticles.


Figure 2. Productivity of the PLAL synthesis approach for a number of different metal (left) and metal oxide (right) nanoparticles, after one hour of laser ablation. Pt: Platinum. Au: Gold. Ag: Silver. Ni: Nickel. NiO: Nickel oxide. Al2O3: Aluminum oxide. CuO: Copper oxide.

We have also analyzed the nanoparticle quality that we produce during the one-hour ablation period. Our results show that the ablation process is stable, in terms of crystallite size (see Figure 3) and surface composition (see Figure 4).13 Interestingly, we also find that by increasing the repetition rate of the laser system (to more than 1MHz), we can increase the ablation efficiency and simultaneously decrease the ablation threshold by a significant level. Such observations can be explained by incubation effects.15


Figure 3. Crystallite size of (a) Ni and NiO, and (b) Pt nanoparticles (as determined from x-ray diffraction measurements) at different points during the one-hour ablation period.

Figure 4. Oxidation state of the (a) Ni and (b) Pt nanoparticles (determined from x-ray photoelectron spectroscopy measurements) at different points during the one hour of ablation. Ni(HO)2: Nickel hydroxide.

Furthermore, our experimental results show that—at high ablation speeds—the scaling factors of ablation productivity in liquid start to converge with those of laser ablation in air. Indeed, we have used a model16—which is based on the ablation law, as developed for ablation in air—to show that the ablation rate can be optimized with a maximum specific ablation rate (defined as the ablate rate per average power) at a specific optimum fluence. As laser manufacturing—after automation and efficiency optimization—has led to industrial series production, we believe that our method for decoupling ablation from cavitation effects in the liquid may lead to the same maturity for PLAL as for laser ablation in air.

In summary, we have developed a novel PLAL setup for nanoparticle production in which we use a high-power pulsed-laser system, an ultra-high-speed laser scanner, and a customized ablation chamber. In this way, we can spatially bypass the cavitation bubbles (with interpulse distances of up to 400μm) and have thus addressed the current major drawback with PLAL methods, i.e., the productivity. With our work, we can therefore scale up the synthesis process to the multigram/hour regime. Our experimental results show that we can achieve productivities (for a variety of metal and metal oxide nanoparticles) of up to 4g/h. The difference in productivities for the different materials, however, cannot be explained by material or laser parameters (e.g., thermal conductivity, heat capacity, or optical penetration depth). The ablation phenomenon in liquid is therefore a combination of different—probably competing—parameters in the laser–target–liquid system. In our future studies we will thus focus on the identification of the relevant parameters and the theoretical prediction of productivities for different material systems.


René Streubel, Bilal Gökce
Technical Chemistry I and Center for Nanointegration
University of Duisburg-Essen
Essen, Germany

René Streubel is a PhD student. He received his Dipl.-Ing. (FH) in technical physics from the University of Applied Sciences Münster (Germany) in 2011. His current research interests include strategies and applications for high-power ultrashort pulsed lasers.

Bilal Gökce obtained his physics degree from RWTH Aachen University (Germany) in 2008 and his PhD in physics from North Carolina State University in 2012. He joined the chemistry faculty at University of Duisburg-Essen in 2014. His group focuses on the functionalization of laser-generated nanoparticles and applications for high-power ultrafast lasers.


References:
1. S. Barcikowski, T. Baranowski, Y. Durmus, U. Wiedwald, B. Gökce, Solid solution magnetic FeNi nanostrand--polymer composites by connecting-coarsening assembly, J. Mater. Chem. C 3, p. 10699-10704, 2015.
2. D. Zhang, B. Gökce, C. Notthoff, S. Barcikowski, Layered seed-growth of AgGe football-like microspheres via precursor-free picosecond laser synthesis in water, Sci. Rep. 5, p. 13661, 2015.
3. C. Schmitz, B. Gökce, J. Jakobi, S. Barcikowski, B. Strehmel, Integration of gold nanoparticles into NIR-radiation curable powder resin, Chem. Sel. 1, p. 5574-5578, 2016.
4. J. T. Sehrt, S. Kleszczynski, C. Notthoff, M. Lau, B. Gökce, S. Barcikowski, Laser powder bed fusion of nano-WC-modified and nano-TiO2-modified metal powders, Proc. Int'l Conf. Addit. Technol. , 2016.
5. L. Jonušauskas, M. Lau, P. Gruber, B. Gökce, S. Barcikowski, M. Malinauskas, A. Ovsianikov, Plasmon assisted 3D microstructuring of gold nanoparticle-doped polymers, Nanotechnology 27, p. 154001, 2016.
6. S. Petersen, S. Barcikowski, Conjugation efficiency of laser-based bioconjugation of gold nanoparticles with nucleic acids, J. Phys. Chem. C 113, p. 19830-19835, 2009.
7. D. Zhang, B. Gökce, Perspective of laser-prototyping nanoparticle-polymer composites, Appl. Surf. Sci. 392, p. 991-1003, 2017.
8. L. Gamrad, C. Rehbock, J. Krawinkel, B. Tumursukh, A. Heisterkamp, S. Barcikowski, Charge balancing of model gold-nanoparticle-peptide conjugates controlled by the peptide's net charge and the ligand to nanoparticle ratio, J. Phys. Chem. C 118, p. 10302-10313, 2014.
9. J. Zhang, M. Chaker, D. Ma, Pulsed laser ablation based synthesis of colloidal metal nanoparticles for catalytic applications, J. Colloid Interface Sci., 2016. doi:10.1016/j.jcis.2016.07.050
10. S. Reich, P. Schönfeld, P. Wagener, A. Letzel, S. Ibrahimkutty, B. Gökce, S. Barcikowski, A. Menzel, T. dos Santos Rolo, A. Plech, Pulsed laser ablation in liquids: impact of the bubble dynamics on particle formation, J. Colloid Interface Sci., 2016. doi:10.1016/j.jcis.2016.08.030
11. B. Gökce, S. Barckowski, Process-stable, highly pure nanomaterials by pilot-scale, continuous high-power laser production. Presented at SPIE Photonics West 2017.
12. R. Streubel, S. Barcikowski, B. Gökce, Continuous multigram nanoparticle synthesis by high-power, high-repetition-rate ultrafast laser ablation in liquids, Opt. Lett. 41, p. 1486-1489, 2016.
13. R. Streubel, G. Bendt, B. Gökce, Pilot-scale synthesis of metal nanoparticles by high-speed pulsed laser ablation in liquids, Nanotechnology 27, p. 205602, 2016.
14. A. Schwenke, P. Wagener, S. Nolte, S. Barcikowski, Influence of processing time on nanoparticle generation during picosecond-pulsed fundamental and second harmonic laser ablation of metals in tetrahydrofuran, Appl. Phys. A 104, p. 77-82, 2011.
15. F. Di Niso, C. Gaudiuso, T. Sibillano, F. P. Mezzapesa, A. Ancona, P. M. Lugarà, Role of heat accumulation on the incubation effect in multi-shot laser ablation of stainless steel at high repetition rates, Opt. Express 22, p. 12200-12210, 2014.
16. B. Neuenschwander, B. Jaeggi, M. Schmid, V. Rouffiange, P.-E. Martin, Optimization of the volume ablation rate for metals at different laser pulse-durations from ps to fs, Proc. SPIE 8243, p. 824307, 2012. doi:10.1117/12.908583
Recent News
PREMIUM CONTENT
Sign in to read the full article
Create a free SPIE account to get access to
premium articles and original research