Sie sind auf Seite 1von 5

Colloids and Surfaces A: Physicochem. Eng.

Aspects 522 (2017) 368–372

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Cavitation bubble dynamics and nanoparticle size distributions in


laser ablation in liquids
J. Tomko a , S.M. O’Malley a,b , Cory Trout a , J.J. Naddeo a , Richard Jimenez a ,
Julianne C. Griepenburg a , Wafaa Soliman c , D.M. Bubb a,b,∗
a
Department of Physics, Rutgers University – Camden, 227 Penn Street, Camden, NJ 08102, USA
b
Center for Computational and Integrative Biology, Rutgers University, Camden, NJ 08102, USA
c
Department of Laser Sciences and Interactions, National Institute of Laser Enhanced Science (N.I.L.E.S.), Cairo University, El-Gamaa Street, Giza 12613,
Egypt

h i g h l i g h t s g r a p h i c a l a b s t r a c t

• The height of the liquid column is


found the strongly affect the cav-
itation bubble dynamics in laser
ablation in liquids.
• The cavitation bubble collapse time
determines the size distribution of
nanoparticles produced.
• The results are found to be in good
agreement with simulations of laser
ablation in liquids.

a r t i c l e i n f o a b s t r a c t

Article history: The influence of cavitation bubble dynamics is investigated during nanoparticle synthesis by ps laser
Received 12 January 2017 ablation of Au targets in distilled and deionized water. The height of the liquid column and the ambient
Received in revised form 3 March 2017 pressure is found to strongly influence both the maximum bubble radius and the collapse time, and
Accepted 11 March 2017
thus the nanoparticle size distribution. The bubble radius is fit to the Rayleigh–Plesset equation and the
Available online 16 March 2017
results are compared to earlier studies with ns lasers. These results are interpreted in the light of recent
computational work and the strong body of experimental evidence that shows the fundamental role the
Keywords:
central cavitation bubble plays in determining the size distribution of nanoparticles in laser ablation in
Laser ablation liquids
Nanoparticles
liquids.
Cavitation © 2017 Elsevier B.V. All rights reserved.
Colloids
Size-distribution

∗ Corresponding author at: Department of Physics, Rutgers University – Camden,


227 Penn Street, Camden, NJ 08102, USA.
E-mail address: danny.bubb@rutgers.edu (D.M. Bubb).

http://dx.doi.org/10.1016/j.colsurfa.2017.03.030
0927-7757/© 2017 Elsevier B.V. All rights reserved.
J. Tomko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 522 (2017) 368–372 369

1. Introduction column over the target. The volume of liquid was kept constant
and a stir-bar was utilized to minimize screening from highly con-
Laser ablation in liquids (LAL) is a technique used to produce col- centrated nanoparticles right above the target surface.
loidal nanoparticle dispersions by focusing an intense laser beam While the pressure range used in this study is admittedly nar-
onto a target surface immersed in liquid. Despite a wealth of exper- row, the maximum cavitation bubble radius and the collapse time
imental results (c. f . [1]), there is significantly less in the way of are observed to change by a factor of more than two. This change
theoretical/computational framework to understand the separate in bubble dynamics strongly influences the nanoparticle size dis-
influence of liquid properties, ambient conditions, materials prop- tribution obtained from LAL and the results can be understood in
erties, and laser parameters in LAL [2,3]. It is a very challenging terms of the models that follow from the Rayleigh–Plesset equation
problem to computationally model as there is all the attendant [10,11].
complexity of laser ablation with the added challenges of model-
ing a highly confined plasma and plume hydrodynamics in a dense 2. Results and discussion
medium. When one adds the cavitation bubble dynamics that occur
immediately over the ablated laser spot, it is understandable why 2.1. Experimental details
computational studies lag somewhat behind experimental results.
Itina, Povarnitsyn, et al., have provided a comprehensive overview The procedure for obtaining particles has been extensively
of the first nanosecond following the arrival of a 100 fs pulse [4,5], described previously [13]. Briefly, a 1064 nm laser with a 25 ps pulse
but it is clear that molecular dynamical simulations must be linked duration and 250 Hz repetition rate (Ekspla PL-2241) was focused
with continuum models to capture the rich variety of length and onto an Au target immersed in distilled and deionized (DDI) water
time scales present during this process. Such knowledge is key to (R > 18 M). 4.7 mJ laser pulses with a beam waist of 100 ␮m (deter-
the control of size and shape necessary for application-related work mined via knife-edge method) were utilized, leading to a fluence
in this area [6]. of 20 J/cm2 . The platform was rastered to avoid excessive structur-
Understanding the role of the cavitation bubble and the ing of the target. Ablation occurred for 2 min and the solutions were
target–bubble interaction during collapse appears to be key to immediately scanned to obtain UV–Vis spectra (Cary 20). Individual
manipulating the size and shapes of nanostructures formed during drops of solution were placed on transmission electron microscope
the LAL process [2,7]. Ibrahimkutty, Wagener, et al., have pro- (TEM) (Zeiss 902) grids where they evaporated to leave behind
mulgated the idea that one must control the cavitation bubble nanoparticles. To obtain histograms, 400 particles were counted
dynamics if one wishes to specify the size distribution of nanopar- for each sample with ImageJ software. Finally, images of the cav-
ticle produced by LAL [8,9]. Through the use of small X-ray and light itation bubble were obtained in the same fashion as earlier work
scattering, they have shown that nascent nanoparticles form within [13,14].
the central cavitation bubble and that the dispersity index of the
particles is greatly influenced by the rate of collapse of the bubble. 2.2. Experimental results
In particular, they find that particles are re-deposited on the target
surface during ablation with a ns laser and that they can be irre- 2.2.1. Influence of liquid height
versibly fused to produce larger aggregates in the high temperature . If the liquid height above the target is increased as shown in
and pressure of the collapsed bubble. Since their measurements are Fig. 2, we see that the nanoparticle diameter increases up until a
occurring within the bubble, they can obtain an accurate assess- value of 8 mm and remains roughly constant after. As in our pre-
ment of the nanoparticle size distribution that is not influenced by vious studies [13], we plot the geometric mean and the error bars
laser-nanoparticle interactions that complicate the understanding reflect the same range as a normal distribution plus or minus one
of ex-situ analyses. standard deviation would. While Ibrahimkutty et al. [9] argue for
In the present work, we have either utilized a small chamber the use of the mass-weighted average, we believe that the geomet-
as shown in Fig. 1 in which we have varied the ambient pressure ric mean most accurately captures the average particle size and we
within the chamber from 20 kPa to 180 kPa, or we have placed the do not observe enough large nanoparticles to assert the existence
target on an elevated platform to specify the height of the liquid

Fig. 2. Nanoparticle diameter and histograms as a function of the height of the liquid
column above the target surface. All data points are obtained from TEM measure-
Fig. 1. Schematic of chamber used in these experiments. The camera lens is visible ments except for the first point in the back panel which was obtained by UV–Vis
in the left of the frame. fitting [12]. The lines in the back panel serve as visual guides.
370 J. Tomko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 522 (2017) 368–372

Fig. 3. (a) Bubble radius as a function of time and resulting Rayleigh–Plesset fits, (b) temperature within bubble according to the Van der Waals hard core model. Here, 2 mm
and 15 mm refer to the height of the liquid column.

of a bimodal distribution in this case. It may be that these discrep- determination of the maximum bubble height and subsequent par-
ancies are rooted in the fact that we are utilizing a 25 ps laser as ticle distributions.
opposed to the ns laser in the aforementioned study. It is also pos- The so-called Rayleigh time is the interval between the bubble
sible that we utilize too low of a fluence to observe this phenomena. reaching maximum radius and the eventual collapse, and it is given
Finally, it could be that the incoming laser pulse strongly interacts as:
with already-formed nanoparticles and reduces the average par- 

ticle diameter. These possible differences will be explored more tc = 0.915Rmax × (1)
Pamb − Pv
below.
Both the mean diameter of the particles and the width of the where  is the density of the fluid, Rmax is the radius of the bub-
distribution increase with liquid height. To estimate the conditions ble at its maximum, Pamb is the pressure in the fluid, and Pv is the
within the bubble we performed fitting to the Rayleigh–Plesset vapor pressure. In the experiments with varying liquid height, this
equation [10] for the bubble radius as a function of time for liq- relationship is not obeyed. Something delays the collapsing bub-
uid column heights of 2 mm and 15 mm. These results are shown ble when the liquid column’s height is low. In order to find an
in Fig. 3 and several salient points immediately commend them- explanation, we return to the plume shadowgraph images.
selves. Clearly, the bubble radius is strongly dependent on the Examining images of the cavitation bubble and shockwaves
liquid column height for a fixed fluence. Second, for comparable flu- present during LAL experiments [18], several types of shockwaves
ences, ps laser – produced bubbles are predicted to be much cooler are observed. First, one is created by the laser ablation event and
than bubbles produced with a ns laser [10] (c. f . 5000–8000 K with appears after 100 ns. It is a weak shock front and moves at the
20,000 K). This is at least partly attributable to the strong inverse speed of sound in the liquid (∼1.5 km/s in water). This shock front
bremsstrahlung effect that is present for ns laser ablation in liquid. eventually reflects off of the water–air interface, and because of the
While the calculation of the temperatures according to the hard large mechanical impedance (Z = cs where ␳ is the density and cs is
core Van der Waals model seems reasonable [15], the predicted the sound velocity) mismatch between these two layers, very little
pressures are too high by at least one order of magnitude (∼1 TPa). sound wave intensity is transmitted into the air (T ∼ 0.001 where
Conversely, if one assumes the bubble expansion and contraction T is the transmission coefficient). Virtually of the original shock-
to be adiabatic, the pressure is more realistic, but the temperature wave is reflected back towards the target. There are also stress
is predicted to be ∼1 K at maximum radius. waves created in the Au target that reflect off of the bottom inter-
De Giacomo [15] calculates the temperature and pressure from face and propagate into the liquid after passing through the top
the actual data, but rightly notes that since the bubble dynamics (metal–liquid) interface. These stress waves first make an appear-
occur on the low ns scale, the temperatures are pressures are much ance approximately 700–800 ns after the arrival of the laser pulse
higher during events like collapse. Most shadowgraphy experimen- and they continue to be emitted from the target surface in regular
tal apparatuses have temporal resolutions in the range of 10–50 ns. intervals thereafter. These waves are less intense than those in the
So it seems reasonable to use the data generated from the fit in order interior of the target as T ≈ 0.1 in this case.
to calculate the temperature and pressure during collapse. The fail- The downwardly propagating shockwave compresses the bub-
ure is not with Rayleigh–Plesset fitting, but more in our inability to ble, thus causing a temperature rise and a concomitant increase in
find a simple analytical model for the collapsing bubble that gives maximum radius. This delays the eventual collapse and the growth
physically reasonable results. and nucleation time of the particles. A heuristic model based upon
Explaining why the cavitation bubble radius depends on the liq- La Mer’s diagram [19] can be invoked to explain the dependence
uid column is more easily accomplished. First, given that each mm of the nanoparticle size in this case. When the bubble is at its
of liquid only introduces another 10 Pa of pressure to the column, maximum radius, the concentration is not high enough to induce
inertial confinement is not a likely explanation, particularly when rapid homogeneous nucleation. However, upon collapse, the con-
one considers earlier results related to pressure [16,10]. Addition- centration rises to a value greater than the lower critical density
ally, the bubble pressure during its expansion phase tends to be and particles rapidly grow. If the bubble rebounds weakly with-
on the order of 103 Pa. Given the bulk modulus of water, this leads out ejecting a large fraction of its mass, then the concentration
to a volume change of the surrounding liquid well below 0.01%, will remain above the lower critical value and allowing particles
regardless of the liquid height. Correspondingly, translation of the to enlarge. In this fashion, we propose a strategy to control particle
liquid layer should be constant for all layer heights. Finally, while size during LAL through external pressure modification (next sec-
prior work has shown the effect of liquid height on fluence on tion) or and/or changes in the liquid height column above the target.
the target surface [17], other mechanisms are clearly involved in This could also help explain the differences observed between
J. Tomko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 522 (2017) 368–372 371

Fig. 4. (a) Early time dynamics of cavitation bubble growth for 2 mm liquid height column. Upward facing arrows mark the appearance of new shock fronts. These fronts are
marked by a dashed line in the images in the inset. Downward facing arrows mark the arrival of more intense shockwaves reflected off of the air-water boundary. The delay
time of the images (ns) is indicated in blue and a 0.18 mm scale bar is indicated in red. (b) Heuristic version of LaMer’s diagram showing concentration and bubble radius and
their influence on growth. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of the article.)

particles produced with a ps laser and those formed with a ns laser. nucleation and growth. Bubbles formed at lower pressure are able
In the ps case, the bubble rebounds one time and appears to stop to expand more and the concentration remains more in line with
oscillating whereas in the ns case, the oscillations occur for sev- the conditions for moderate growth and nucleation.
eral more cycles. These last cycles of the bubble are much smaller
than the initial oscillation and the total lifetime can stretch over 2.2.3. Comparison with earlier results
hundreds of  s (Fig. 4). . In our earlier work, we reported the effect of laser fluence on
the average nanoparticle size and the cavitation bubble radius [13].
Here, there was a clear and strong dependence of the bubble radius
2.2.2. Influence of ambient pressure on fluence, but there was large scatter in the nanoparticle diame-
. The ambient pressure strongly influences the maximum cavi- ter results. While it is difficult to directly compare the work done
tation bubble and time of collapse. Perhaps most surprising is that here with different liquid heights and ambient pressure to earlier
the range over which this occurs is fairly narrow. In Fig. 5b, one can work with varied fluences, it may be instructive to compare the
see that as the pressure is reduced, the bubble’s maximum diam- fluence-dependent data with the computational results of Povar-
eter increases as does the time until first collapse. In the range of nitsyn et al. [4,(especially Figs. 2, 4, 6, and 8)]. If we assume t1/2 –
our observations, the maximum bubble radius depends linearly on scaling [20] between the computational results that utilize a 100 fs
the ambient pressure while the collapse time falls off more rapidly. laser and the experimental results obtained with a 25 ps laser in
In the lower panel of Fig. 5b, we fit the data to Eq. (1) to confirm order to compare fluences, we are able to identify low, medium,
1
tc
that Rmax ∼P − 2 . We can neglect the vapor pressure at the extremely and high fluence ranges that correspond to distinctly different ejec-
low temperatures inside of the fully expanded bubble [10]. Just as tion mechanisms. In this study, we operate solely in the “high”
in the data from a liquid column, the mean particle diameter and fluence regime (>10 J/cm2 ). At these fluences, the computational
distribution size are found to be strongly correlated with the maxi- model predicts that a secondary molten layer forms after energy
mum bubble radius and collapse time. A ready interpretation of our is reflected back into the target. This is thought to lead to larger
pressure-dependent results can also be found in the above adap- primary particles with a bimodal distribution. We do not observe
tation of LaMer’s diagram. Here we associate the smaller bubble this in our experimental results in the present work or the past.
volume found at high pressure with the conditions that favor rapid One possible mechanism for breaking apart this secondary molten

tC
Fig. 5. (a) TEM histograms of particles made at different pressures. The back panel shows the nanoparticle diameter as a function of the cavitation bubble radius. (b) Rmax
as
a function of ambient pressure. The line is a fit to Fig. (1).
372 J. Tomko et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 522 (2017) 368–372

layer could be the collapse of the cavitation bubble and the violent [6] T. Shoji, J. Saitoh, N. Kitamura, F. Nagasawa, K. Murakoshi, H. Yamauchi, S. Ito,
secondary ejection that often follows. It could also be that we still H. Miyasaka, H. Ishihara, Y. Tsuboi, permanent fixing or reversible trapping
and release of DNA micropatterns on a gold nanostructure using
at too low of a fluence to clearly observe the bimodal nature of the continuous-wave or femtosecond-pulsed near-infrared laser light, J. Am.
particles. In any event, we regard the overall agreement between Chem. Soc. 135 (17) (2013) 6643–6648.
our experimental work and the earlier computational work to be [7] Z. Yan, R. Bao, R.N. Wright, D.B. Chrisey, Hollow nanoparticle generation on
laser-induced cavitation bubbles via bubble interface pinning, Appl. Phys.
good. Lett. 97 (12) (2010) 124106.
[8] P. Wagener, S. Ibrahimkutty, A. Menzel, A. Plech, S. Barcikowski, Dynamics of
3. Conclusion silver nanoparticle formation and agglomeration inside the cavitation bubble
after pulsed laser ablation in liquid, Phys. Chem. Chem. Phys. 15 (9) (2013)
3068–3074.
The size distribution of nanoparticles formed in LAL experi- [9] S. Ibrahimkutty, P. Wagener, T. dos Santos Rolo, D. Karpov, A. Menzel, T.
ments can be greatly affected by the height of the liquid column Baumbach, S. Barcikowski, A. Plech, A hierarchical view on material formation
during pulsed-laser synthesis of nanoparticles in liquid, Sci. Rep. 5 (2015)
above the target and the ambient pressure. It is also found that there
16313.
is a strong correlation between the size of the cavitation bubble at [10] W. Soliman, T. Nakano, N. Takada, K. Sasaki, Modification of Rayleigh–Plesset
its maximum radius and the collapse time. The time required for theory for reproducing dynamics of cavitation bubbles in liquid-phase laser
ablation, Jpn. J. Appl. Phys. 49 (11) (2010) 116202.
bubble collapse also appears to strongly influence the particle size
[11] W. Soliman, Fundamental Studies on the Synthesis Dynamics of Nanoparticles
and we hypothesize that a heuristic model based upon La Mer’s dia- by Laser Ablation in Pressurized Water (PhD thesis), Nagoya, 2011.
gram might be used to help target desired nanoparticle properties [12] V. Amendola, M. Meneghetti, Size evaluation of gold nanoparticles by UV–vis
during LAL. spectroscopy, J. Phys. Chem. C 113 (2009) 4277–4285.
[13] J. Tomko, J.J. Naddeo, R. Jimenez, Y. Tan, M. Steiner, J.M. Fitz-Gerald, D.M.
Bubb, S.M. O’Malley, Size and polydispersity trends found in gold
Acknowledgments nanoparticles synthesized by laser ablation in liquids, Phys. Chem. Chem.
Phys. 17 (2015) 16327–16333.
[14] M. Amin, J. Tomko, J.J. Naddeo, R. Jimenez, D.M. Bubb, M. Steiner, J.
D. M. Bubb and S. M. O’Malley acknowledge Awards CMMI- Fitz-Gerald, S.M. O’Malley, Laser-assisted synthesis of ultra-small anatase
0727713, CMMI-0922946, and CMMI-1531783 from the National TiO2 nanoparticles, Appl. Surf. Sci. 348 (2015) 30–37.
Science Foundation and support from the Charles & Johanna Busch [15] A. De Giacomo, M. Dell’Aglio, A. Santagata, R. Gaudiuso, O. De Pascale, P.
Wagener, G.C. Messina, G. Compagnini, S. Barcikowski, Cavitation dynamics of
Biomedical Grant Program. laser ablation of bulk and wire-shaped metals in water during nanoparticles
production, Phys. Chem. Chem. Phys.: PCCP 15 (9) (2013)
References 3083–3092.
[16] K. Sasaki, T. Nakano, W. Soliman, N. Takada, Effect of pressurization on the
dynamics of a cavitation bubble induced by liquid-phase laser ablation, Appl.
[1] V. Amendola, M. Meneghetti, What controls the composition and the Phys. Express 2 (2009) 046501.
structure of nanomaterials generated by laser ablation in liquid solution? [17] A. Men, P. Wagener, S. Barcikowski, Transfer-matrix method for efficient
Phys. Chem. Chem. Phys. 15 (9) (2013) 3027–3046. ablation by pulsed laser ablation and nanoparticle generation in liquids, 2011,
[2] A.V. Simakin, V.V. Voronov, G.V. Shafeev, Nanoparticle formation during laser pp. 5108–5114.
ablation of solids in liquids, Phys. Wave Phenom. 15 (4) (2007) 218–240. [18] T. Tsuji, Y. Okazaki, Y. Tsuboi, M. Tsuji, Nanosecond time-resolved
[3] Z. Yan, D.B. Chrisey, Pulsed laser ablation in liquid for micro-/nanostructure observations of laser ablation of silver in water, Jpn. J. Appl. Phys. 46 (4A)
generation, J. Photochem. Photobiol. C: Photochem. Rev. 13 (3) (2012) (2007) 1533–1535.
204–223. [19] V.K. LaMer, R.H. Dinegar, Theory, production and mechanism of formation of
[4] M.E. Povarnitsyn, T.E. Itina, P.R. Levashov, K.V. Khishchenko, Mechanisms of monodispersed hydrosols, J. Am. Chem. Soc. 72 (8) (1950)
nanoparticle formation by ultra-short laser ablation of metals in liquid 4847–4854.
environment, Phys. Chem. Chem. Phys.: PCCP 15 (9) (2013) 3108–3114. [20] E.G. Gamaly, A.V. Rode, B. Luther-Davies, V.T. Tikhonchuk, Ablation of solids
[5] M.E. Povarnitsyn, T.E. Itina, Hydrodynamic modeling of femtosecond laser by femtosecond lasers: ablation mechanism and ablation thresholds for
ablation of metals in vacuum and in liquid, Appl. Phys. A 117 (1) (2014) metals and dielectrics, Phys. Plasmas (1994-present) 9 (3) (2002)
175–178. 949–957.

Das könnte Ihnen auch gefallen