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

MHz-ultrasound generation by chirped femtosecond laser pulses from gold nano-colloidal suspensions

Open Access Open Access

Abstract

Strong absorption of femtosecond laser pulses in Au nano-colloidal suspensions was used to generate coherent ultrasound signals at 1–20 MHz frequency range. The most efficient ultrasound generation was observed at negative chirp values and was proportional to the pulse duration. Maximization of a dimensionless factor A ≡ αc0tp defined as the ratio of pulse duration tp and the time required for sound at speed c0 to cross the optical energy deposition length (an inverse of the absorption coefficient α) given by 1/(αc0). Chirp controlled pulse duration allows effective enhancement of ultrasound generation at higher frequencies (shorter wavelengths) and is promising for a high spatial resolution acoustic imaging.

© 2016 Optical Society of America

1. Introduction

Laser-induced breakdown through optical absorption is a result of the interaction between high intensity ultra-short laser pulses and matter. The initial energy deposition consists of the generation of localized highly ionized plasma with a typical density of one electron per hundreds-of-molecules in the focal region of a laser beam [1–3]. In liquids, laser breakdown is accompanied by filamentation; an important aspect of femtosecond laser propagation that is characterized by a number of nonlinear effects including self focusing, intensity clamping, propagation at high intensity beyond Rayleigh range and supercontinuum white light generation [4]. It is one considerable interest over the past decades due to its potential applications ranging from micro- and nano-processing of dielectric materials for biomedical and therapeutic applications such as optical-to-acoustic conversion in photoacoustics and bubble dynamics in ultrasound generation [5–7].

Acoustic signals with longer propagation length give rise to localization of thermal expansion in 3D space and time that opens feasibility to probe materials which are optically turbid and not transparent. Moreover, it opens new possibilities to probe IR properties of materials with sub-wavelength nanoscale resolution which cannot be accessed by optical IR spectroscopy techniques due to larger focal spots [8]. For example, a spectral scan with focal spot placed at the atomic force microscope (AFM) tip with simultaneous height detection by AFM delivers absorption spectra via local absorption-heating-expansion [9]. With the current state of the art in 3D triangulation and imaging, there is a strong possibility of photoacoustic techniques to enter wider fields of applications in biomedical and material science engineering.

Femtosecond laser-induced acoustic wave generation contributes to a wider range of frequency emission spectra as compared to the case of nanosecond laser. In the case of nanosecond lasers, longer pulse duration leads to slow heating which is accompanied by a complex mechanism such as thermal expansion, cavitation, and bubble formation. An amplified photoacoustic performance of reduced graphene oxide-coated gold nanorods with photoacoustic amplitudes in the 4–11 MHz range was observed with 5–7 ns laser pulses [10]. The quantitave imaging of microvasculature in deep tissues with spectrum-based microscopy up to 10 MHz was reported using 8–10 ns laser pulses [11]. On the other hand, femtosecond laser pulses with shorter pulse duration induce rapid heating which yields to the generation of acoustic wave with wider frequency range. It can easily generate optical breakdown in water through multiphoton and impact ionization [12]. Two-photon absorption-induced photoacoustic imaging using 50 fs laser pulses showed a wide frequency emission spectra up to 30 MHz [13]. A broadband emission up to 20 MHz from the propagation of intense 45 fs laser pulses in water was reported [14]. Enhanced photoacoustics from gold nano-colloidal suspensions under 40 fs laser excitation was observed with a wide frequency range up to 25 MHz [15]. These aspects of photoacoustic generation over a wide range of frequencies by femtosecond laser pulses are highly beneficial for imaging technologies with high spatial resolution.

Pulse pre-chirping is known to have a strong impact in nonlinear dynamics and has been widely used in recent years. An optimal negative input chirp makes an equal spatial focusing and temporal compression lengths, yielding enhancements in non-linear effects resulting in high intensities, long plasma channels, and broader spectra and possibility to remote spectroscopy [16,17]. Hatanaka, et al. studied the chirp effect on X-ray emission intensity from CsCl aqueous solution jet irradiated by femtosecond laser pulses. The negatively-chirped pulses of approximately 240 fs duration produced up to ten times higher X-ray intensity as compared with transform-limited 160 fs pulses [18]. Similar studies were conducted by Yamaoka, et al. photoacoustic microscopy using ultra-short pulses with two different pulse durations in the range of femtoseconds to picoseconds. It was demonstrated that the required number of photons in photoacoustic microscopy using ultrashort optical pulses (tp = 250 fs) was about 1/1000 lower than sub-nanosecond optical pulses (tp = 600 ps) [19].

Nanoparticle-facilitated absorption of pulsed laser leads to the production of MHz ultrasound waves used for photoacoustic techniques. In particular, plasmonic Au nanoparticles have been investigated due to their biocompatibility and superior optical properties such as large absorption cross-section and spectra selectivity in the visible and near infrared ranges based on surface plasmon resonance. The mechanism of interaction between pulsed laser and Au nanoparticles is attributed to non-radiative relaxation dynamics of surface plasmon oscillations [19–27]. Tunability of surface plasmon resonance wavelength to the near-IR region where human tissue exhibits high sensitivity and transmission is considered to be highly beneficial. When Au nanoparticles are irradiated with ultra-short laser pulses, a rapid increase in temperature resulting in photothermal and photoacoustic effects was used for highly localized cell damage in photothermal therapy and contrast agents in photoacoustic imaging were used to visualize temperature maps in a desired location in tissues [28–33].

Here, we show the generation of MHz frequency coherent ultrasound by fast energy deposition with ultrashort laser pulses. Chirp effects on amplitude of photoacoustic signal and spectral properties of Au colloidal nanoparticles of spherical and rod-like shapes were systematically studied for generation of augmented photoacoustic signals.

2. Samples and procedures

Colloidal suspensions of Au nanospheres and rods for the ultrasound generation were prepared via syntheses described elsewhere [15]. Au nanospheres with an absorption peak at 520 nm (diameter of 20 nm) and nanorods with transverse and longitudinal absorption peaks at 520 nm and 800 nm (dimensions of 12 nm×35 nm) respectively were used for photoacoustic detection and measurements. The longitudinal band of Au nanorods was controlled by changing the aspect ratio of nanorod depending on the desired wavelength of interest. The 800 nm longitudinal absorption band of Au nanorod with aspect ratio of ~2.92 which coincides with the wavelength of femtosecond laser at 800 nm was chosen for the ultrasound generation. Au nano-colloidal suspensions with atomic concentration of ~1.4×104 mol/L, particle concentration of ∼ 4.21×1018 NPs/mL and volume of ~4×103 nm3 were used in the experiments.

Figure 1(a) shows the experimental set-up. Femtosecond laser pulses (Mantis, Legend, Elite HE USP, Coherent, Inc.) were tightly focused by an objective lens (10×, NA = 0.28, M Plan Apo, Mitutoyo) into a 5-mm glass tube inside the water tank. The sample suspension in the glass tube was circulated continuously by a conventional pump. The photoacoustic measurements were carried out with a hydrophone (HNA-0400, ONDA) in the detection frequency range at 1–20 MHz and an ultrasound preamplifier (5678, Olympus). The distance between the hydrophone and the glass tube was kept constant at 1.5 cm. The signals were recorded and analyzed with a digital oscilloscope (DSO-X 3034A, Agilent Tech.). The geometry of glass tube in water was used to simulate the experiments with biomedical relevance where the photoacoustic signal generation is separated from detection. The white light continuum spectra was measured by a spectrometer (SD2000, Ocean Optics, Inc.).

 figure: Fig. 1

Fig. 1 (a) Generation of ultrasound in aqueous solutions of Au nanoparticles with fs-laser pulses. Laser pulses are focused inside the glass tube using 10× numerical aperture NA = 0.28 objective lens. (b) Side-view of the optical image of white light continuum (WLC) in water at the focal region at different pulse energies Ep = 30,100 μJ/pulse at repetition rate of 1 kHz and pulse duration tp ≃ 40 fs; the corresponding power per pulse Pp = 0.75,2.5 GW/pulse. Distance between focal spot and hydrophone was set 1.5 cm in all experiments. Arrows mark pulse propagation direction.

Download Full Size | PDF

Near-IR femtosecond laser pulses used for ultrasound generation: the pulse duration of t0 = 40 ± 5 fs at the central wavelength of λ = 800 nm, the pulse energy was typically from Ep = 0.1 mJ to 1 mJ, at 1 kHz repetition rate. A pulse can be described by a Gaussian temporal envelope of the form E(t)=E0e2ln2(t/tp)2cos(ωt+βt2), where ω is the cyclic frequency of light, tp is the pulse duration at the full-width at half maximum (FWHM), β [1/fs2] is the linear chirp, and E0=2I0/(cε0n) is the field amplitude, n is the refractive index, c is speed of light, t is time, I0 = 2Iav is the peak intensity which is twice larger than the average, Iav for the Gaussian, and ε0 is permittivity of vacuum. The instantaneous cyclic frequency ωins(t) = ω0 + 2βt, where β > 0 corresponds to the positive chirp with trailing high frequency components.

Pulse pre-chirping is implemented to increase pulse duration up to ten times using 1D array of liquid crystal cells (FemtoJock, Biophotonics Solutions, Inc.). For the slow frequency with ω varying spectral phase φ(ω) it can be expanded into the Taylor series around the central frequency ω0 with the few first terms as φ(ω)=φ(ω0)+Φ1ω+Φ22!ω2+, where φ(ω0) is the absolute phase of the pulse in time domain, the first derivative φ′(ω0) = Φ1 is the group delay (GD) which defines a shift of envelope in time domain, φ″(ω0) = Φ2[fs2] is the group delay dispersion (GDD) or the second order dispersion which defines the chirp in time domain. Duration of the time broadened pulse at FWHM is defined as tp=t01+[4ln2Φ2/t02]2t01+β2 where t0 = 40 fs is the shortest spectral bandwidth limited pulse duration. For this study higher orders as well as Φ1 were set to zero.

3. Coherent ultrasound generation: theory

Generation of coherent ultrasound by laser pulses is not an efficient process and depends on the absorbed energy, Eabs, and acoustic energy of coherent ultrasound, Eac [34]:

η=EacEabs4β2ρ0cp2I5×1014(I/[W/cm2]),
where β [1/°C] is the temperature expansion coefficient for the volume, ρ0 is the unperturbed mass density of the medium, cp is the heat capacity at constant presssure. The acoustic energy for parameter A = 1 (see Eq. (2) below):
EacS|pmax|2ρ0cp2×2tp,
where S is the surface area of the focal spot; Eq 1 is derived for the case |pmax| = 2p1 (see Eq. (3) below).

Despite low efficiency, a substantial pressure amplitude of the coherent sound can be generated [34]:

|pmax|=2p1A/(1+A)[bar],
where p1 (1−R)I0c0β/cp is the characteristic pressure in the acoustic wave with R reflectivity coefficient and I0 incident light intensity; A ≡ αc0tp is the parameter defined by the absorption coefficient α, speed of sound in media c0, and the laser pulse duration tp (note, A≪1). The non-dimensional parameter A is the ratio of laser pulse duration and time 1/(αc0) required for sound to pass through the light absorption region. These analytical expressions are derived for the laser pulse described by function f(t/tp) = exp(−|t|/tp) [34].

4. Results and discussion

4.1. Water

Generation of white light continuum (WLC) by ultra-short laser pulses in wide bandgap materials is an efficient process when photon energy is less than the half of the bandgap of the host matrix; for shorter wavelengths two-photon absorption becomes an efficient absorption mechanism. Water absorption at ∼6.2 eV energy was for a long time assumed as the band gap energy. Recently, it was demonstrated that this energy corresponds to production of solvated electron complex out of the valence band in pure water [35] while the bandgap of the water is ∼9.5 eV. The use of fs-laser pulses with central wavelength of 800 nm is appropriate for efficient WLC generation. Side-view WLC images from the focal region reveals localization of emission around the focal region which would have an axial extent of 2zR=πw02/λ24μm with a waist of the beam (radius) w0 = 0.61λ/NA ≃ 1.7 μm for the NA = 0.28 objective lens; the estimation assumes a Gaussian beam profile and aberration free linear focusing. The average power used in experiments for the Ep = 100 μJ pulse was Iav = 2.5 GW. Nonlinear WLC and filamentation are both related to self focusing. In water and glass, a strong WLC generation is observed upon excitation by near-IR fs-laser pulses. For a pulse duration tp ≃ 40 fs, pulse power becomes larger than the critical self-focusing threshold Pcr = α1λ2/(4πn0n2), where n2 is the nonlinear refractive index due to Kerr effect n(I) = n0 + n2I with constant α1 = 1.8962 for a Gaussian pulse [36]. For water at λ = 800 nm, n2 = 4.1×1016 cm2/W [37], refractive index of n = 1.33, the critical power of self focusing is Pcr ≃ 1.77 MW. However, filamentation along the propagation was not very strong L ∝ 100 μm (Fig. 1(b)). Filaments for photoacoustic generation of sound by WLC in glass (for silica n2 = 3.5 × 1016 cm2/W [37]) using longer tp ≃ 150 fs pulses at the same λ = 800 nm wavelength [38] showed considerably longer traces. The better axial localization of energy delivery in water can be due to a strong light scattering and bubble formation along light propagation in the focal region. Such localization and energy deposition is helpful for photoacoustic applications due to a better localized pressure wave time-arrival onto the hydrophone.

Figure 2 shows WLC spectra for different pulse durations controlled by GDD, Φ2. Negative Φ2 values correspond to a spectrally blue-shifted wavelengths preceding the red-shifted ones, while for the positive chirp (Φ2 > 0) the blue-part is trailing the red. As expected, the most spectrally broad spectra were generated when the pulse duration was the shortest, or Φ2 = 0 (shaded region in Fig. 2). The highest peak intensities are reached at such setting; i.e., for the shortest pulse. There was a small difference between spectra of WLC generated with positive and negative chirp with the same absolute values indicating that the phenomenon is mostly dependent on intensity rather ionization which has a strong wavelength dependence.

 figure: Fig. 2

Fig. 2 Generation of white light continuum (WLC) with positively Φ2 > 0 (a) and negatively Φ2 < 0 (b) chirped ultra-short laser pulses in water. Spectra are plotted as lg(Intensity). Arrows mark trend with increase of pulse duration. Pulse energy was Ep = 100 μJ/pulse.

Download Full Size | PDF

The representative photoacoustic signals in time domain generated by Au nanosphere colloidal suspensions were shown in Fig. 3; negatively-chirped pulse (800 fs) (a), transform-limited pulse (40 fs) (b), and positively-chirped pulse (800 fs) (c). The first peak represents the fundamental ultrasound signal (a presssure wave) and the second peak corresponds to the internal reflection of ultrasound waves from the inner surface of the glass tube. The time interval between the two peaks is ascribed to the difference in travel distances of the two signals in water solutions where the speed of sound is 1500 m/s [15]. The first peak signal amplitude was used as a measure of photoacoustic response from femtosecond laser-irradiated sample solutions. It was observed that higher photoacoustic signal amplitude was exhibited by negatively-chirped pulses rather than positively-chirped pulses. On the other hand, irradiation of transform-limited pulses result in lowest photoacoustic signal amplitude.

 figure: Fig. 3

Fig. 3 Representative photoacoustic signals in time domain generated by Au nanosphere colloidal suspensions: negatively-chirped pulse (800 fs) (a), transform-limited pulse (40 fs) (b), and positively-chirped pulse (800 fs) (c) at pulse energy of Ep = 100 μJ/pulse.

Download Full Size | PDF

Figure 4 shows amplitude of photoacoustic signal of hydrophone at different frequencies over 1–20 MHz range. It is clearly observed that the shortest chirp-free pulses induce the least efficient ultrasound generation in pure water as well as in Au nano-colloidal solutions. This is consistent with the acoustic pressure scaling with parameter A = αc0tp (Theory sec. 3). Longer pulses generated stronger acoustic (pressure) signals. The central wavelength was at 800 nm which was close to the longitudinal (L-mode) resonance of nanorods [Fig. 4(c)]. The insets in Figs. 4(b) and 4(c) shows the calculation results of the light intensity distribution around single spherical and rod-shaped nanoparticles. Up to ten times more intense light field localization is expected for the resonance case Fig. 4(c). Au nanoparticles are homogeneously distributed and randomly oriented in the solution phase. In the case of Au nanorods, the relative orientation of nanorod particles and the incident laser polarization is not always the same as expected in the calculation, while in the case of Au nanospheres the calculation reflects the real condition of the solution. Therefore, as shown in Figs. 4(b) and 4(c), it is reasonable that the ultrasound intensity of nanorod is lower than expected from the calculation.

 figure: Fig. 4

Fig. 4 Generation of photoacoustic (PA) signal at different values of positive and negative chirp values in water (a), Au nanosphere (b), and nanorod (c) solutions. Pulse energy was Ep = 100 μJ/pulse. Insets in (b,c) shows a light intensity |E|2 distribution around nanosphere and nanorod for linearly polarized incident field E = 1 at a close to resonant condition (L-mode) for the nanorod (finite difference time domain code FDTD-Lumerical).

Download Full Size | PDF

In our previous work, it was demonstrated that higher photoacoustic intensity was exhibited by Au nanorods which is attributed to non-radiative relaxation dynamics of surface plasmon oscillations [15]. Au nanorods maximize the absorption of pulsed light which leads to an enhanced photoacoustic intensity. Similarly, Hatanaka, et al. [39] performed a study on the influence of temporal chirp on X-ray emission. The highest X-ray intensity was observed at the longer pulse width tp = 372 fs as compared to the shortest pulse width tp = 45 fs. The negative chirp was, also, slightly more efficient in X-ray generation. Spectral shape of the photoacoustic signal was different for different colloidal solutions as discussed next.

4.2. Au nano-colloidal suspensions

Spectrally broad WLC could be considered advantageous in delivering energy by absorption to Au nanoparticles in water, since Au has interband transitions at shorter wavelengths and spectrally broad extinction peak which is size and shape dependent. Apparently, in pure water and colloidal solutions sound generation followed similar scaling at different acoustic frequencies and showed stronger pressure waves formed for longer pulses (Fig. 4).

Interestingly, the photoacoustic signal was the smallest then WLC was the most spectrally broad at Φ2 = 0 (Fig. 5). Figure 5 shows efficiency of photoacoustic response for different chirp settings (correspondingly different pulse durations). Negative chirp was slightly more efficient in sound generation as can be evidenced by steeper slope of the dependence (Fig. 5). However, the key control parameter to obtain strong photoacoustic signal was the pulse duration. Longer pulses were more efficient in sound production as follows from analytical prediction Eq. (2).

 figure: Fig. 5

Fig. 5 (a) White light continuum (WLC) spectral width at different pulse durations controlled by chirp. (b) The photoacoustic signal integrated over the range of hydrophone 1 – 20 MHz as a function of pulse duration defined by the GDD, Φ2. Duration of the time broadened pulse at FWHM is defined as tp=t01+[4ln2Φ2/t02]2, where t0 = 40 fs is the shortest spectral bandwidth limited pulse duration. Note, the vertical axis of WLC spectra in Figure 2 is logarithmic. Diameter of Au spheres was 20 nm and rods were 35-nm-long and 12-nm-wide. The slope difference between the dashed lines is 1.79 times. Pulse energy was Ep = 100 μJ/pulse.

Download Full Size | PDF

Speed of sound in gold c0 = 3240 m/s is 2.19 times higher than in water, 1482 m/s. The slope difference of photo-acoustic signal in the case of water and Au nanorods is approximately 1.89 for the long pulse side [see lines at the negative chirps; Fig. 5(b)]. This might indicate a changing contribution to thermal expansion detected by transducer from water to increasingly from gold nanorods as pulse duration gets longer. Speed of sounds enters parameter Aαc0tp which governs photo-acoustic amplitude (see, Eq. 3).

In pure water, there was strong nonlinearity in the photoacoustic signal generation around Φ2 = 0. It was almost absent for Au nanoparticle solution. This can be understood by highly nonlinear absorption mechanism in water and linear absorption and avalanche evolution via free electron absorption in the case of Au nanoparticles. By changing chirp, an instantaneous laser frequency is changing accordingly, ωins(t) = ω0 + 2βt. As governed by Eqn. 3, the strongest photo-acoustic signal is expected for the longer pulse and stronger absorbance, α. The steepest changes in PA amplitude where, indeed, observed for the Au nanorods at negative chirp. Since the nanorod had extinction (absorption and scattering losses) around 800 nm wavelengths, pre-chirping was also affecting spectral overlap of plasmonic resonance and the WLC spectrum centered around 800 nm.

5. Conclusions and outlook

It is demonstrated that negative chirp of ultra-short laser pulses at the wavelength overlapping with an extinction resonance of Au nanoparticles achieves the most efficient ultrasound generation at MHz-frequency range. Accordingly, it has been shown that size homogenization of Au colloidal dispersions can be achieved by WLC reabsorption after an Au foil ablation in pure water and a surfactant-free stable colloidal solution can be obtained [40]. Pulse pre-chirping effect on photoacoustic intensity and WLC emission spectra exhibited an opposite tendency; longer pulse duration generates higher photoacoustic intensity and narrower WLC spectra, while shorter pulse duration generates lower photoacoustic intensity and broader WLC spectra. This tendency is beneficial for the determination of the optimum pulse duration that can be used in photoacoustic imaging techniques.

Funding

SJ is grateful for partial support via the Australian Research Council DP130101205 Discovery project and by the nanotechnology ambassador fellowship program at the Melbourne Centre for Nanofabrication (MCN) in the Victorian Node of the Australian National Fabrication Facility (ANFF).

Acknowledgments

TY acknowledges the partial support of Murata Foundation.

References and links

1. K. Kennedy, D. X. Hammer, and B. A. Rockwell, “Laser-induced breakdown in aqueous media,” Prog. Quant. Elect. 21, 155–248 (1997). [CrossRef]  

2. E. Abraham, K. Minoshima, and H. Matsumoto, “Femtosecond laser-induced breakdown in water: time-resolved shadow imaging and two-color interferometric imaging,” Opt. Comm. 176, 441–452 (2000). [CrossRef]  

3. G. Toker, V. Bulatov, T. Kovalchuk, and I. Schechter, “Micro-dynamics of optical breakdown in water induced by nanosecond laser pulses of 1064 nm wavelength,” Chem. Phys. Lett. 471, 244–248 (2009). [CrossRef]  

4. H. Zhang, Z. Zhou, A. Lin, J. Cheng, L. Yan, J. Si, F. Chen, and X. Hou, “Chirp structure measurement of a supercontinuum pulse based on transient lens effect in tellurite glass,” J. Appl. Phys. Lett. 113, 113106 (2013).

5. C. Milian, A. Jarnac, Y. Brelet, V. Jukna, A. Houard, A. Mysyrowicz, and A. Couairon, “Effect of input pulse chirp on nonlinear energy deposition and plasma excitation,” J. Opt. Soc. Am. B. 31, 2829–2837 (2014). [CrossRef]  

6. D. Tan, K. N. Sharafudeen, Y. Yue, and J. Qiu, “Femtosecond laser induced phenomena in transparent solid materials: fundamentals and applications,” Prog. Mat. Sci. 76, 154–228 (2016). [CrossRef]  

7. T. Winkler, C. Sarpe, N. Nelzow, L. H. Lillevang, N. Gotte, B. Zielinski, P. Balling, A. Seftleben, and T. Baumert, “Probing spatial properties of electronic excitation in water after interaction with temporally shaped femtosecond laser pulses: experiments and simulations,” Appl. Surf. Sci. 374, 235–242 (2015). [CrossRef]  

8. A. Dazzi, C. B. Prater, Q. Hu, D. B. Chase, J. F. Rabolt, and C. Marcott, “Afm-air: combining atomic force microscopy and infrared spectroscopy for nanoscale chemical charaterization,” Appl. Spectrosc. 66, 1365–1384 (2012). [CrossRef]   [PubMed]  

9. A. Dazzi, F. Glotin, and R. Carminati, “Theory of infrared nanospectroscopy by photothermal induced resonance,” J. Appl. Phys. 107, 124519 (2010). [CrossRef]  

10. H. Moon, D. Kumar, H. Kim, C. Sim, J.-H. Chang, H.-M. Kim, H. Kim, and D. K. Lim, “Amplified photoacoustic performance and enhanced photothermal stability of reduced graphene oxide coated gold nanorods for sensitive photoacoustic imaging”, ACS Nano. 9(3), 2711–2719 (2015). [CrossRef]   [PubMed]  

11. X. Gao, C. Tao, X. Wang, and X. Liu, “Quantitative imaging of microvasculature in deep tissue with a spectrum-based photo-acoustic microscopy”, Opt. Lett. 4(6), 970–973 (2015). [CrossRef]  

12. Y. Brelet, A. Jarnac, J. Carbonnel, Y. Andre, A. Mysyrowicz, and A. Houard, “Underwater acoustic signals induced by intense ultrashort laser pulse”, J. Acoust. Soc. Am. 137(4), 288–292 (2015). [CrossRef]  

13. G. Langer, K. D. Bouchal, H. Grun, P. Burgholzer, and T Berer, “Two-photon absorption-induced photoacoustic imaging of Rhodamine B dyed polyethylene spheres using femtosecond laser"”, Opt. Express 21(19), 22410–22422 (2013). [CrossRef]   [PubMed]  

14. A. Houard, Y. Brelet, A. Jarnac, J. Carbonnel, A. Mysyrowicz, C. Milian, A. Couarion, R. Guillermin, and J. P. Sessaego, “Propagation of intense femtosecond laser pulse in water and acoustic waves generation”, in Conference on Lasers and Electro-Optics (Optical Society of America, 2014) paper STh1E.8.

15. F. C. P. Masim, H.-L. Liu, M. Porta, T. Yonezawa, A. Balčytis, S. Juodkazis, W.-H. Hsu, and K. Hatanaka, “Enhanced photoacoustics from gold nano-colloidal suspensions under femtosecond laser excitation,” Opt. Express 24(13), 14781–14792 (2016). [CrossRef]  

16. P. J. S. V. Capel, E. Peronne, and J. I. Dijkhuis, “Nonlinear ultrafast acoustics at the nanoscale,” Ultrasonics 56, 136–151 (2015).

17. A. Couairon and A. Mysyrowicz, “Femtosecond filamentation in transparent media,” Physics Reports 441, 47–189 (2007). [CrossRef]  

18. K. Hatanaka, T. Ida, H. Ono, S. Matsushima, H. Fukumura, S. Juodkazis, and H. Misawa, “Chirp effect in hard x-ray generation from liquid target when irradiated by femtosecond laser pulses,” Opt. Express 16(17), 12650–12657 (2008). [CrossRef]   [PubMed]  

19. Y. Yamaoka, Y. Harada, M. Sakakura, T. Minamikawa, S. Nishino, S. Maehara, S. Hamano, H. Tanaka, and T. Takamatsu, “Photoacoustic microscopy using ultra-short pulses with two different pulse durations,” Opt. Express 22(14), 17063–17072 (2014). [CrossRef]   [PubMed]  

20. B. Song, A. Fiorino, E. Meyhofer, and P. Redyy, “Near-field radiative thermal transport: From theory to experiment,” AIP Advances 5, 053503 (2015). [CrossRef]  

21. V. P. Zharov, “Ultrasharp nonlinear photothermal and photoacoustic resonances and holes beyond the spectral limit,” Nature Photonics 5, 110–116 (2011). [CrossRef]   [PubMed]  

22. Y. S. Chen, W. Frey, S. Kim, K. Homan, P. Kruizinga, K. Sokolov, and S. Emalianov, “Enhanced thermal stability of silica-coated gold nanorods for photoacoustic imaging and image-guided therapy,” Opt. Express 9(18), 8867–8878 (2013).

23. T. A. El-Brossy, T. Abdallah, M. B. Mohamed, S. Abdallah, K. Easawi, S. Negm, and H. Talaat, “Shape and size dependence of gold nanoparticles studied by photoacoustic techniques,” Eur. Phys. J. Special Topics 153, 361–364 (2008). [CrossRef]  

24. S. Link and M. A. El-Sayed, “Spectral properties and relaxation dynamics of surface plasmon resonance oscillations in gold and silver nanodots and nanorods,” J. Phys. Chem. B. 103, 8410–8426 (1999). [CrossRef]  

25. Y. Yamaoka and T. Takamatsu, “Enhancement of multiphoton excitation-induced photoacoustic signals by using gold nanoparticles surrounded by fluorescent dyes,” Proc. of SPIE 7177, 71772 (2012). [CrossRef]  

26. J.-W. Liaw, S.-W. Tsai, H. H. Lin, T.-C. Yen, and B.-R. Chen, “Wavelength-dependent Faraday-Tyndall effect on laser-induced microbubble in gold colloid”, J. Quant. Spectro. Rad. Trans. 113, 2234–2242 (2012). [CrossRef]  

27. S.-Y. Yu, S.-W. Tsai, Y.-J. Chen, and J.-W. Liaw, “Pulsed laser induced microbubble in gold nanorod colloid”, Microelec. Eng. 138, 102–106 (2015). [CrossRef]  

28. C. Tarapacki, C. Kumaradas, and R. Karshafian, “Enhancing laser-thermal therapy using ultrasound-microbubbles and gold nanorods of in vitro cells,” Ultrasonics 53, 793–798 (2013). [CrossRef]   [PubMed]  

29. J. Baumgart, L. Humbert, E. Boulais, R. Lachaine, J. J. Lebrun, and M. Meunier, “Off-resonance plasmonic enhanced femtosecond laser optoporation and transfection of cancer cells,” Biomaterials33 (2012). [CrossRef]  

30. R. J. Zemp, R. Bitton, M. L. Li, k. K. Shung, G. Stoica, and L. V. Wang, “Photoacoustic imaging of the microvasculature with a high-frequency ultrasound array transducer,” J. Biomed. Optics 12, 010501 (2007). [CrossRef]  

31. H. F. Zhang, K. Maslov, M. L. Li, G. Stoica, and L. V. Wang, “In vivo volumetric imaging of subcutaneous microvasculature by photoacoustic microscopy,” Opt. Express 14, 9317 (2006). [CrossRef]   [PubMed]  

32. W. Song, W. Zheng, R. Liu, H. Huang, X. Gong, S. Yang, and L. Song, “Reflection mode in vivo photoacoustic microscopy with subwavelength lateral resolution,” Biomed. Opt. Express 5, 4235 (2014). [CrossRef]  

33. H.-W. Yang, H.-L. Liu, M.-L. Li, I.-W. Hsi, C. T. Fan, C.-Y. Huang, Y.-J. Lu, M.-Y. Hua, H.-Y. Chou, J.-W Liaw, C.-C. Ma, and K.-C. Wei, “Magnetic gold nanorod/PNIPAAmMA nanoparticles for dual magnetic resonance and photoacoustic imaging and targeted photothermal therapy”, Biomater. 34, 5651–5660 (2013). [CrossRef]  

34. N. I. Koroteev and I. L. Shumay, Physics of High Power Laser Radiation (NaukaMoscow, 1991).

35. N. Linz, S. Freidank, X.-X. Liang, H. Vogelmann, T. Trickl, and A. Vogel, “Wavelength dependence of nanosecond infrared laser-induced breakdown in water: Evidence for multiphoton initiation via an intermediate state,” Phys. Rev. B 91, 134114 (2015). [CrossRef]  

36. G. Fibich and A. L. Gaeta, “Critical power of self-focusing in buld media and in hollow waveguides,” Opt. Lett. 25, 335–337 (2000). [CrossRef]  

37. R. W. Boyd, S. G. Lukishova, and Y. R. Shen, Topics in Applied Physics (Springer, 2009). [CrossRef]  

38. S. I. Kudryashov, V. D. Zvorykin, A. A. Ionin, V. Mizeikis, S. Juodkazis, and H. Misawa, “Acoustic monitoring of microplasma formation and filamentation of tightly focused femtosecond laser pulses in silica glass,” Appl. Phys. Lett. 92, 101916 (2008). [CrossRef]  

39. K. Hatanaka, M. Porta, F. C. P. Masim, W. H. Hsu, M. T. Nguyen, T. Yonezawa, A. Balčytis, and S. Juodkazis, “Efficient X-ray generation from gold colloidal solutions”, https://arxiv.org/abs/1604.07541.

40. R. Kubiliūtė, K. Maximova, A. Lajevardipour, J. Yong, J. S. Hartley, A. S. M. Mohsin, P. Blandin, J. W. M. Chon, A. H. A. Clayton, M. Sentis, P. R. Stoddart, A. Kabashin, R. Rotomskis, and S. Juodkazis, “Ultra-pure, water-dispersed au nanoparticles produced by femtosecond laser ablation and fragmentation,” Int. J. Nanomed. 8, 2601–2611 (2013).

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 (5)

Fig. 1
Fig. 1 (a) Generation of ultrasound in aqueous solutions of Au nanoparticles with fs-laser pulses. Laser pulses are focused inside the glass tube using 10× numerical aperture NA = 0.28 objective lens. (b) Side-view of the optical image of white light continuum (WLC) in water at the focal region at different pulse energies Ep = 30,100 μJ/pulse at repetition rate of 1 kHz and pulse duration tp ≃ 40 fs; the corresponding power per pulse Pp = 0.75,2.5 GW/pulse. Distance between focal spot and hydrophone was set 1.5 cm in all experiments. Arrows mark pulse propagation direction.
Fig. 2
Fig. 2 Generation of white light continuum (WLC) with positively Φ2 > 0 (a) and negatively Φ2 < 0 (b) chirped ultra-short laser pulses in water. Spectra are plotted as lg(Intensity). Arrows mark trend with increase of pulse duration. Pulse energy was Ep = 100 μJ/pulse.
Fig. 3
Fig. 3 Representative photoacoustic signals in time domain generated by Au nanosphere colloidal suspensions: negatively-chirped pulse (800 fs) (a), transform-limited pulse (40 fs) (b), and positively-chirped pulse (800 fs) (c) at pulse energy of Ep = 100 μJ/pulse.
Fig. 4
Fig. 4 Generation of photoacoustic (PA) signal at different values of positive and negative chirp values in water (a), Au nanosphere (b), and nanorod (c) solutions. Pulse energy was Ep = 100 μJ/pulse. Insets in (b,c) shows a light intensity |E|2 distribution around nanosphere and nanorod for linearly polarized incident field E = 1 at a close to resonant condition (L-mode) for the nanorod (finite difference time domain code FDTD-Lumerical).
Fig. 5
Fig. 5 (a) White light continuum (WLC) spectral width at different pulse durations controlled by chirp. (b) The photoacoustic signal integrated over the range of hydrophone 1 – 20 MHz as a function of pulse duration defined by the GDD, Φ2. Duration of the time broadened pulse at FWHM is defined as t p = t 0 1 + [ 4 ln 2 Φ 2 / t 0 2 ] 2, where t0 = 40 fs is the shortest spectral bandwidth limited pulse duration. Note, the vertical axis of WLC spectra in Figure 2 is logarithmic. Diameter of Au spheres was 20 nm and rods were 35-nm-long and 12-nm-wide. The slope difference between the dashed lines is 1.79 times. Pulse energy was Ep = 100 μJ/pulse.

Equations (3)

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

η = E a c E a b s 4 β 2 ρ 0 c p 2 I 5 × 10 14 ( I / [ W / cm 2 ] ) ,
E a c S | p m a x | 2 ρ 0 c p 2 × 2 t p ,
| p m a x | = 2 p 1 A / ( 1 + A ) [ bar ] ,
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.