Influence of solvent polarity on light-induced thermal ... - OSA Publishing

2 downloads 76 Views 2MB Size Report
Apr 30, 2015 - synchronizing, light-induced thermal cycles in plasmonically absorbing .... stage (MTS) varies the entrance depth of light into the nanofluid.
Research Article

Vol. 2, No. 5 / May 2015 / Optica

447

Influence of solvent polarity on light-induced thermal cycles in plasmonic nanofluids J. L. DOMINGUEZ-JUAREZ,1,4,† S. VALLONE,1,† A. LEMPEL,1 M. MOOCARME,1,2 J. OH,1 H. D. GAFNEY,3 AND L. T. VUONG1,2,* 1

Department of Physics, City University of New York, Queens College, Flushing, New York 11367, USA The Graduate Center, City University of New York, New York, New York 10016, USA 3 Department of Chemistry, City University of New York, Queens College, Flushing, New York 11367, USA 4 Centro de Fisica Aplicada y Tecnologia Avanzada, UNAM, Quéretaro 76230, Mexico *Corresponding author: [email protected] 2

Received 15 December 2014; revised 7 April 2015; accepted 9 April 2015 (Doc. ID 228584); published 30 April 2015

Pattern formation often reveals constituent nonlinear mechanisms of a complex system. Here, we study selfsynchronizing, light-induced thermal cycles in plasmonically absorbing nanofluids, whose anomalous thermal, optomechanical, electrochemical, and hydrodynamic responses are not yet well understood. We show that the oscillatory behavior—caused by light grazing the nanofluid meniscus—exhibits a strong dependence on hydrogen bonding in the solvent environment and that there are low-intensity optical thresholds in alcohol–water binary-solvent nanofluids. Moreover, these thermal cycles occur with a periodic, vertically discharging heat-dissipation mechanism, which could be facilitated by nanobubbles or thermophoresis. We show that an incoherent white-light source, such as sunlight, will also induce self-synchronizing thermal cycles; in this demonstration, we illustrate new methods of energy storage, transfer, and harvesting that will not alter the natural carbon cycle of life. © 2015 Optical Society of America OCIS codes: (160.4236) Nanomaterials; (190.3100) Instabilities and chaos; (190.4710) Optical nonlinearities in organic materials; (190.4870) Photothermal effects; (190.5940) Self-action effects; (260.2160) Energy transfer. http://dx.doi.org/10.1364/OPTICA.2.000447

1. INTRODUCTION Metal nanofluids, i.e., liquid mixtures containing plasmonically absorbing nanoparticles, are appealing for their potential uses in numerous light harvesting applications [1,2]; however, control of the associated light-driven dynamics remain a challenge. Nanofluids exhibit complex behavior, with thermal conductivities far greater than those of the solvent alone [3–7]. Local heating effects increase the mean free velocity of both surface and solvent charges that shift the chemical, electrical, and hydrodynamic equilibrium, particularly in the presence of heat or light [5,8–12]. Correspondingly, nanoparticles associate with the solvent environment in a manner similar to molecules during solvation processes [13]. Control of the light-driven nanofluid dynamics is further complicated by the temperature-dependent refractive indices, photoinduced pressure gradients, thermal or chemical diffusion processes, and changing reaction affinities with the nanoparticles [3,5,9,14]. The recently observed memory of nanofluids that occurs as a result of the applied DC magnetic fields [15,16] remains unexplained by conventional Mie scattering or light diffusion dynamics and may be a result of longer-lived changes in the solvent environment of the nanoparticle [17]. This investigation is centered on the oscillations or thermal cycles that spontaneously commence when light grazes the meniscus of the nanofluid [18–20]. The spatiotemporal oscillations 2334-2536/15/050447-07$15/0$15.00 © 2015 Optical Society of America

are previously established and point to a class of instabilities that occur via the asymmetric dissipation of heat near a “free interface,” i.e., the meniscus (see Ref. [19] for a review). A generic description of the periodic heat cycles follows: light illuminates a fluid, is strongly absorbed and converted to heat; the heat is transferred to the surrounding liquid; subsequently, the heated liquid expands and its refractive index is lowered; due to gravity, the hotter fluid rises, resulting in a higher refractive index below the light beam; light refracts toward the cooler fluid, providing an instability or a Marangoni flow that dissipates heat via the meniscus and the neighboring liquid; the liquid immediately above the beam cools; and the light Poynting vector is returned to its starting direction. This periodic behavior is broadly modeled by a “dripping faucet” or continuously charging/threshold-discharging capacitor [21], where limit cycles exhibit bifurcations and hysteresis. Figure 1 illustrates the cycle in four separate processes of the oscillations (light absorption, heat dissipation, light refraction, and convection) even though the processes occur with varying degrees, simultaneously. Further complexity arises in our study of various polar organicaqueous nanofluid mixtures (water with methanol, ethanol, isopropanol, acetone, or acetonitrile). In fact, alcohol–water binary mixtures exhibit numerous intriguing physical properties that may manifest in novel nonlinear effects in nanofluids. Partial

Research Article

Fig. 1. Illustration of a single oscillation period of the light-induced thermal cycle. Inset: schematic of the light propagation and experimental CCD images of the far-field transmitted patterns. Although absorption, dissipation, refraction, and convection are separately outlined, the processes occur simultaneously in varying degrees.

molar volume, adiabatic compressibility, deviations in the expected heat capacity and viscosity, variable boiling point and light scattering, and ultrasonic speeds [22,23] are some anomalous behaviors of binary mixtures that have been documented for decades. Enhanced Marangoni convective flows with binary mixtures may play a role in the low-threshold illumination intensities necessary to produce the oscillatory modes [23]. Our investigation highlights the solvent-dependent mechanical, chemical, thermal, and optical coupling of energy in alcohol–water nanofluids that are of broad interest and warrant further study. We provide the first study of the thermal cycles in plasmonically absorbing nanofluids, which are produced at low illumination intensities in a stable, collimated-light geometry with alcohol–water nanofluid mixtures. Significantly higher intensities or focused-light geometries are necessary to produce the periodic cycles when instead pure aqueous solvents are chosen or when absorbing nanoparticles are not present. In the first part of our report, we show that the intermolecular hydrogen bonding in alcohol–water mixtures correlates with the reliable low-threshold production of the oscillations, which are subsequently controllable via the choice of polar organic-aqueous solvent. In the second part of this report, we provide numerical calculations of the temperature profiles, from which we extrapolate different heat-discharging mechanisms and associate the driving dynamics of the thermal cycles with alcohol–water binary fluids. In the final section of this report, we demonstrate that it is possible to drive the spontaneous thermal cycles with an incoherent white-light source in an isopropanol–water nanofluid. 2. RESULTS—EXPERIMENTAL The stock aqueous nanofluids [polyvinylpyrrolidone (PVP)coated 80 nm gold, 0.05 mg/ml, Nanocomposix] compose onefifth of each of the samples by volume. The remaining four-fifths are a mixture of either a protic (methanol, ethanol, or isopropanol) or an aprotic (acetone or acetonitrile) organic solvent and deionized water as the base solvent. Figure 2 shows the general experimental setup with a laser light source and further illustrates

Vol. 2, No. 5 / May 2015 / Optica

448

Fig. 2. (a) Side and (b) top views of the experimental setup. The halfwave plate (WP) and polarizer (P) control the beam power and polarization (fixed horizontally), respectively, while the vertical translation stage (MTS) varies the entrance depth of light into the nanofluid. The CCD camera records the imaged beam pattern as a function of time. (c) Side view images of the sample cuvette with 80 nm gold nanoparticles illustrating beam propagation and depth are shown where the two curved lines correspond to menisci with parasitic reflections formed by the residual light scattered. (d) Four images of the far-field pattern that are recorded during a single oscillation period. (e) Plots of the integrated CCD images as a function of time for five different depths.

the periodic behavior that is the focus of this study. Light is polarized horizontally, with 2 ps pulse duration (λ  532 nm), 300 mW average power, 1 MHz repetition rate, and a 0.9 mm beam diameter, and propagates through a 1 cm quartz cuvette containing the nanofluid. In contrast to prior work where light is focused onto or into a liquid sample, here light is collimated when it enters the plasmonically absorbing sample. The collimated-light setup reduces the likelihood of nanofluid degradation [12] and may provide more reliable oscillations. In this system, light simultaneously drives and provides a signature of the nonlinear dynamics: the transmitted far-field beam exhibits power-dependent ring-like interference fringe patterns (Figs. 1 and 2, insets), which provide information about the energy transfer in the nanofluids. The transmitted beam interference patterns are clearly visible on a white screen placed 80 cm from the sample and the projected beam diameter is reduced or enlarged as a function of time. We image and record the far-field beam pattern at 30 frames per second with a CCD camera and quantify oscillations via the intensity-integrated unsaturated CCD images or by the beam power measured by a silicon photodetector in place of the screen (not shown). A motorized translation stage (MTS) translates the sample vertically. The characteristic frequency of the oscillations decreases as a function of depth from the meniscus [Fig. 2(e)]. The generated far-field pattern is stable when the beam depth is greater than 5 mm measured from the nanofluid–air interface; instabilities are observed when the beam grazes the meniscus above this depth. The subsequent oscillations vary with lightentrance depth and with solvent mixture, and in general, the periodic oscillations are not purely sinusoidal. We plot the Fourier decomposition of the integrated CCD images as a function of depth in each nanofluid mixture in Fig. 3. From left to right, the organic solvent contribution for each nanofluid increases from 1 part solvent/3 parts deionized water/1 part aqueous stock nanofluid solution (1∶3∶1) to 3 parts solvent/1 part deionized water/1 part aqueous stock nanofluid solution (3∶1∶1). Figure 3 shows acetone, acetonitrile, methanol, ethanol, and

Research Article isopropanol mixtures from top to bottom, arranged in order of increasing proticity. Distinct differences between polar protic (alcohols) and aprotic (acetone and acetonitrile) organic–aqueous nanofluids are evident in the Fourier decomposition of the transmitted light patterns (Fig. 3). The light-induced oscillations are limited or negligible with polar aprotic organic–aqueous mixtures. In the alcohol–water mixtures, the oscillations vary with mixture and appear strongest when samples have a net composition ratio of 3∶2 of water to alcohol. This approximate ratio between water and alcohol corresponds to a change in sign of the Soret thermophoretic coefficient [24] and enhanced Marangoni flows [23], which may increase the thermal dissipation. Here, we show that

Fig. 3. Logarithm of the Fourier decomposition of the transmitted light power as a function of depth in polar organic-aqueous mixtures containing 80 nm gold nanosphere stock solution. The three columns represent from left to right organic solvent/water/stock mixtures with ratios of 1∶3∶1, 2∶2∶1, and 3∶1∶1, respectively. Sample data are shown with increasingly protic organic solvent from top to bottom: (a–c) acetone, (d–f) acetonitrile, (g–i) methanol, (j–l) ethanol, (m–o) isopropanol. The DC components of the Fourier transform are omitted.

Vol. 2, No. 5 / May 2015 / Optica

449

the thermal cycles depend strongly on the hydrogen bonding of the alcohols in the nanofluid and correlate with the proticity of the solvent in the mixture. Therefore, the choice of binary mixture yields control of the nanofluid thermal dynamics. The amplitudes of the oscillations are largest when light enters the nanofluid at an intermediate distance from the meniscus [Fig. 2(e)]. At this intermediate light-entrance depth, the Fourier decomposition shows the presence of additional harmonic oscillatory frequencies, or a sharp difference in the characteristic times for heat charging (absorption and storage) or discharging (thermal dissipation). Analysis of the heat equation indicates that the maximal oscillations are related to higher temperatures: if we assume negligible heat dissipation at the meniscus–air interface, then the highest steady-state temperature is produced when a light beam enters at a depth of approximately one beam diameter below the liquid–air interface. The characteristic oscillatory frequencies decrease as the entrance depth increases and could be a result of lower nanofluid temperatures and reduced heat dissipation rates. Figure 3 indicates that the energy storage and transfer in the thermal cycles are tunable with solvent, yet the underlying physical mechanisms remain unclear—particularly in relation to the properties of the binary mixture. In an effort to understand the dynamics of the binary mixture, we scrutinize the concentration-dependent transmission and focus solely on aprotic acetone and protic isopropanol, each in a mixture with water alone, without the plasmonically absorbing 80 nm gold nanoparticles. Figure 4 shows the optical transmissions of aqueous-acetone and aqueous-isopropanol mixtures in 1 cm quartz cuvettes, which are measured at varying volume ratios with 532 nm, 2 ps, 150 nJ laser pulses. Prior to measuring, samples were stirred and visually inspected to ensure homogeneity, and each transmission measurement is averaged over 50 s. In contrast to the measured linear transmissions of acetone–water mixtures associated with a Maxwell–Garnett effective-medium approximation, alcohol–water binary mixtures exhibit a 10% dip in transmission when the volume ratio of alcohol to water is 3∶6. The sharp absorption coincides with strong hydrogen bonding with the reported presence of alcohol–water nanocage

Fig. 4. Transmission of low-intensity laser light (λ  532 nm, average power  100 mW) though isopropyl alcohol–water and acetone–water binary mixtures as a function of volume fraction. The isopropyl alcohol mixture features a sharp dip with 3 parts alcohol to 6 parts water.

Research Article structures, and with an anomalous and increased thermal conductivity [13,25–30], as discussed further in Section 4. We experimentally observe that ethanol and methanol aqueous mixtures exhibit similar but less-pronounced transmission dips. The transmission dips are also less pronounced at higher illumination powers. An analysis of the potential heat dissipation mechanisms, as extrapolated from the transmitted optical behavior, follows in the next section. 3. RESULTS—NUMERICAL We analyze the interference fringes, which are produced by shifts in the nanofluid refractive index as a result of variations in temperature or high light intensity; the intensity of the far-field fringe patterns are approximated by Fraunhofer diffraction: 2 ZZ   −2αd  iΔϕx;y −ixξyν e dxdy ; (1) I inc x; ye I FF ξ; ν  e  where x and y (ξ and ν) are non-dimensionalized spatial coordinates of the incident I inc (far-field I FF ) electric fields and are scaled to the incident Gaussian width such that the incident in2 2 tensity I inc x; y  e −2x y  ; α is the characteristic absorption, d is the nanofluid thickness, and Δϕx; y denotes the phase shift accumulated in the slab. Equation (1) assumes light propagation in short optical-path-length samples and averages dynamics in the direction of light propagation. Temperature-dependent changes in refractive index dominate the outlines of the far-field interference patterns; the effects of self-focusing and diffraction are present but are reduced in the presence of plasmonically absorbing nanoparticles. In fact, temperature-dependent phase shifts are more than an order of magnitude larger than Kerr effects. The temperature-dependent phase shift is ∂n T x; y; (2) Δϕx; y  kd ∂T ∂n where k is the wavenumber in the nanofluid, ∂T is the temperature-dependent change in refractive index, and T x; y is the temperature profile of the nanofluid. The heat and light that are absorbed depend on the distribution of nanoparticles, whose dynamics are challenging to model. Our quasi-steady-state thermal “snapshots” of the nanofluid assume that the absorption parameter α varies over the nanofluid thermal cycle, but not spatially as a function of x or y. Here, we approximate that the nanoparticle density is uniform in the nanofluid, although we expect that several effects lead to the contrary. Gradient and scattering forces are not expected to significantly influence the nanoparticle density in our experimental setup, which employs a low-duty-cycle laser in a collimated geometry; gradient forces are several orders of magnitude smaller than scattering forces hF scat i [31,32], which are of the same order as Brownian forces hF Brownian i [32], or hF scat i∕hF Brownian i ∼ 1 for 1 W∕cm2 average illumination intensities. In contrast, it remains an open question as to whether thermophoretic effects are significant. We expect that variations in the nanoparticle density would influence the optical absorption and the thermal profiles calculated here (see Section 4). In fact, experimentally, we observe that the gold nanoparticles move toward the laser beam profile; however, the thermophoretic coefficients for nanofluids are nonlinear and even change sign with mixed-solvent mixtures [22].

Vol. 2, No. 5 / May 2015 / Optica

450

We calculate the average temperature profile from the far-field interferometric patterns and allow asymmetric thermal diffusion with the assumption of slow thermal changes governed by transverse diffusion and light absorption:   dT ∂2 ∂2 απP 0  C T 2  C nv  C T  2 T  I ≈ 0; (3) ∂x ∂y ρC ρ dt where C T is the thermal diffusion coefficient of the stationary nanofluid, C nv is the vertical thermal discharging coefficient that is nonzero when y > 0, C ρ is the heat capacity of the solvent, ρ is the mass density of the nanofluid, and P 0 is the average beam power. The temperature profile is numerically calculated in the Fourier domain with continuity at y  0. We set the unilluminated nanofluid temperature as 293 K, kd  1.5e5, ∂n 3 ∂T  1e − 4∕K, P 0  300 mW, and ρC ρ  4.2 J∕cm K. The larger far-field beam size and flat-topped asymmetry are produced by higher absorption α and an appreciable vertical heat diffusion (a large C nv with respect to C T ), as shown in Fig. 5. As C nv increases and heat is increasingly dissipated in the vertical direction, the thermal gradient for y > 0 is reduced. Subsequently, the far-field diffraction is also truncated as y increases, leading to a flattened far-field intensity profile, I FF , as illustrated in Figs. 5(b)–5(c). Moreover, the wider far-field patterns correspond with sharper heat gradients and higher light absorption. Marangoni and Rayleigh–Bénard convective flows partially explain how heat is dissipated in the vertical direction due to the rise of heated solvent toward the meniscus. A question remains as to the physical mechanism that leads to the extreme changes in absorption and heat dissipation in the vertical direction. There is evidence that as gold nanoparticles are illuminated, the surrounding solution can undergo a phase transition to a superheated metastable liquid state as an alternative to direct vaporization, which, upon cooling, collapses into mixed states of higher and lower density; this could account for the vertical motion of the nanoparticles [31]. The dynamics of the water– solute interactions at this stage must both enhance and play a critical role in the low-light threshold heat oscillations. We also observe a cyclic nanoparticle flow in the vertical plane of the light via side-camera images that would be facilitated by nanobubbles, thermal blooming, and thermophoresis [10,11,32–35]. From the numerical calculations and the far-field patterns, we extrapolate that C nv varies by three orders of magnitude over the duration of one thermal cycle. The large and sudden changes in C nv suggest a threshold heat-discharging mechanism that sharply accelerates the Marangoni flows, which could be explained by the

Fig. 5. Numerically calculated far-field interference patterns I FF (above) associated with the temperature profiles of the nanofluid (below) for (a) α  1.7e − 2∕cm, C nv  10C T , (b) α  1.3e − 2∕cm, C nv  1000C T , (c) α  8e − 3∕cm, C nv  3000C T , and (d) α  0.5e − 3∕cm, C nv  C T .

Research Article finite heat of vaporization required to produce the nanobubbles. The formation of nanobubbles or vapor enclosures around hot nanoparticles would produce buoyant lifting forces and accelerate the removal of heat from the nanofluid center. Moreover, the simultaneous removal of plasmonically absorbing particles from the light path would also rapidly decrease the light that is absorbed. The presence of nanobubbles agrees experimentally with the time-dependent changes in the far-field patterns, where the diffraction pattern simultaneously shrinks as it gains the flattopped asymmetry (insets in Figs. 1 and 2). 4. DISCUSSION Knowledge of the thermal profiles and the presence of the heat dissipation alone does not adequately explain why the thermal cycles commence with organic protic-aqueous nanofluids (i.e., isopropanol) and not organic aprotic-aqueous nanofluids (i.e., acetone): acetone and isopropanol have similar viscosities and volatilities; in fact, acetone has a lower heat of vaporization than isopropanol, from which one would extrapolate a lower threshold for heat dissipation if facilitated by the presence of nanobubbles. We turn our attention to the nature of solute–solvent interactions, which directly influence the physical behavior of liquids [17,31,36] and account for the existence of water–alcohol clusters [13,27]. By contrasting the behavior of nanoparticles in aqueous solutions of acetone versus isopropyl alcohol, we note that acetone, unable to function as a proton donor either to other acetone molecules or to its solvent environment, is fully miscible in water with a correspondingly low entropy of mixing [37]. We claim that the concentration-dependent drop in the transmission of alcohol–water mixtures (Fig. 4) reflects an essential component of the low-intensity thermal cycles of the alcohol– water nanofluids studied here. With the addition of nanoparticles under laser illumination, these variations in concentration would manifest as noticeable changes in the optical properties. We suggest that the dynamic changes in sample transmission and conductivity occur via concentration variations in the local mixture, which toggle the heat storage and dissipation mechanisms: during the discharge or heat-dissipation process, alcohol is preferentially vaporized and ejected with hot nanoparticles from the beam center. The discharge mechanism would decrease the ratio of alcohol to water as the temperature drops; subsequently during the reset process, close to the critical threshold of 3∶6 alcohol:water, the generation of nanobubbles is clamped, the transmission of the matrix decreases significantly; more heat is absorbed and dissipated by the solvent and less heat is absorbed by the nanoparticles; and the charging or heat storage cycle recommences. Isopropyl alcohol and water show inhomogeneities in mixing and show evidence for supermolecular structure formation at the critical threshold of 3∶6 alcohol:water. Although there is not yet consensus on the exact nature of these structures, it has been shown that these structures are long-lived, i.e., stable beyond the time scales of experiments [13,27,38]. Alcohol–water mixtures are often considered to be miscible, but are in fact anisotropic; cluster radii are theoretically calculated and experimentally verified to be as large as 2.6 Å [39,40]. Through Kirkwood–Buff integration theory, the critical volume fraction of 0.32 isopropanol:water maximizes the intermolecular self-interaction energies [39] that provide a cluster volume surrounding the isopropanol molecules.

Vol. 2, No. 5 / May 2015 / Optica

451

Our report marks the beginning of understanding how alcohol–water binary mixtures may be exploited for optical applications. The interplay of alcohol–alcohol, alcohol–water, and water–water hydrogen bonding leads to unusual optical characteristics that are underexplored. Further investigations with threshold-generated nanobubbles and azeotropic mixtures may yield new control over complex nanofluid dynamics associated with the excess entropy that arises from the incomplete mixing at the molecular level [13]. The robust thermal cycles studied here would be useful for characterizing the solvent mixture and thermal diffusivity of binary liquids. The heat cycles may characterize optically induced modes for heat and energy transfer and may provide a mechanism by which the anomalous properties of liquids and mixtures may be further explored. The optical transmission spectra of alcohol–water mixtures indicate strong intermolecular interactions that might be exploited or utilized to more effectively absorb and convert the energy in broadband low-intensity light sources such as sunlight, which we further demonstrate experimentally. Oscillations are produced when a low-average-power, 1.3 W, continuous incoherent xenon-lamp source (collimated to a 1.4 mm beam) grazes the meniscus of a silver nanofluid. The silver nanofluid contains 20% aqueous stock (equal parts of PVPcoated 60 nm nanospheres, 650 nm resonant nanoplatelets, and 750 nm resonant nanoplatelets, 0.05 mg/ml, Nanocomposix) with 40% isopropanol and 40% deionized water. The transmission spectrum of the silver nanofluid is shown in Fig. 6(a). With a setup similar to that in Fig. 1 (replacing the laser with a collimated lamp source), we measure the transmitted power as a function of time and show the depth-dependent decrease of the characteristic frequency. The fluctuations of the transmitted power are plotted at different depths in Fig. 6(b), where the peak-to-peak amplitudes vary between 1% and 3%. The Fourier transform of the measured transmission is plotted as a function of depth in Fig. 6(c); from this plot, we extrapolate that the broadband light excitation provides a comparable effect to that achieved using a pulsed coherent-source laser even though the time-varying diffraction patterns are not observable. The data presented in Fig. 6 establish detectable oscillations that arise when the fluid is pumped with low-power, incoherent white light. The whitelight-induced thermal cycles would be stronger by increasing the broadband nanoparticle absorption, tightening the collimation of the input beam, as well as by tuning the solvent mixture ratios.

Fig. 6. (a) Transmission spectra for isopropanol:water:stock nanoparticle sample in a ratio of 2∶2∶1. (b) White-light-induced oscillations in the transmitted power through a silver nanofluid of the same ratio. (c) Logarithm of the Fourier decomposition of the oscillations (DC components removed), which shows a depth-dependent characteristic frequency.

Research Article In conclusion, we have shown that low-intensity light-pumped oscillations in nanofluids are enabled with the presence of alcohol as the polar organic solvent in deionized water mixtures. We propose that the preferential vaporization of alcohol in nanobubbles, in combination with the anomalous optical transmission and thermal behavior of alcohol–water binary mixtures, enables the thermal cycles to be produced at low light intensity thresholds. The corresponding thermal cycling, which characterizes a complex nonlinear system, provides a mechanism that can identify intermolecular hydrogen bonding associated with organicaqueous binary mixtures, where small amounts of alcohol added to a nanofluid sample lead to the onset of the low-powerthreshold heat oscillations. There exists further demand for investigations of the transmitted optical behavior with the thermal response of nanofluids and for determining the ideal environment that induces the instabilities in the nanofluids, i.e., investigating the contribution of gravity-induced concentration gradients [41], thermodiffusion, and Soret coefficients in water–solvent mixtures. Our understanding of the hydrogen bonding in alcoholaqueous solutions may be relevant to the future designs of ion switches, optical pulse shaping techniques [42], engineered microfluidic flows [43,44], biomimetics [45], or ionic gradients [46–50] for osmotic energy harvesting. †These authors contributed equally to this work. Division of Materials Research (DMR) (1151783).

REFERENCES 1. R. Taylor, S. Coulombe, T. Otanicar, P. Phelan, A. Gunawan, W. Lv, G. Rosengarten, R. Prasher, and H. Tyagi, “Small particles, big impacts: a review of the diverse applications of nanofluids,” J. Appl. Phys. 113, 011301 (2013). 2. I. Gur, K. Sawyer, and R. Prasher, “Searching for a better thermal battery,” Science 335, 1454–1455 (2012). 3. N. Shalkevich, W. Escher, T. Bürgi, B. Michel, L. Si-Ahmed, and D. Poulikakos, “On the thermal conductivity of gold nanoparticle colloids,” Langmuir 26, 663–670 (2010). 4. J. Fan and L. Wang, “Heat conduction in nanofluids: structure–property correlation,” Int. J. Heat Mass Transfer 54, 4349–4359 (2011). 5. E. Greenfield, C. Rotschild, A. Szameit, J. Nemirovsky, R. El-Ganainy, D. N. Christodoulides, M. Saraf, E. Lifshitz, and M. Segev, “Light-induced self-synchronizing flow patterns,” New J. Phys. 13, 053021 (2011). 6. L. Wang and J. Fan, “Toward nanofluids of ultra-high thermal conductivity,” Nanoscale Res. Lett. 6, 153 (2011). 7. R. Prasher, P. Bhattacharya, and P. Phelan, “Thermal conductivity of nanoscale colloidal solutions (nanofluids),” Phys. Rev. Lett. 94, 025901 (2005). 8. F. B. Hicks and C. Franck, “Low-power laser driving of a binary liquid,” Phys. Rev. E 58, 4582–4586 (1998). 9. Y. E. Geints, N. S. Panamarev, and A. Zemlyanov, “Transient behavior of far-field diffraction patterns of a Gaussian laser beam due to the thermooptical effect in metal nanocolloids,” J. Opt. 13, 055707 (2011). 10. S. Mubeen, J. Lee, N. Singh, S. Krämer, G. D. Stucky, and M. Moskovits, “An autonomous photosynthetic device in which all charge carriers derive from surface plasmons,” Nat. Nanotechnol. 8, 247–251 (2013). 11. Z. Fang, Y.-R. Zhen, O. Neumann, A. Polman, F. J. García de Abajo, P. Nordlander, and N. J. Halas, “Evolution of light-induced vapor generation at a liquid-immersed metallic nanoparticle,” Nano Lett. 13, 1736–1742 (2013). 12. Y. Takeuchi, T. Ida, and K. Kimura, “Colloidal stability of gold nanoparticles in 2-propanol under laser irradiation,” J. Phys. Chem. B 101, 1322– 1327 (1997). 13. S. Dixit, W. C. K. Poon, and J. Crain, “Hydration of methanol in aqueous solutions: a Raman spectroscopic study,” J. Phys. Condens. Matter 12, L323–L328 (2000).

Vol. 2, No. 5 / May 2015 / Optica

452

14. Y. Lamhot, A. Barak, O. Peleg, and M. Segev, “Self-trapping of optical beams through thermophoresis,” Phys. Rev. Lett. 105, 163906 (2010). 15. N. D. Singh, M. Moocarme, B. Edelstein, N. Punnoose, and L. T. Vuong, “Anomalously-large photo-induced magnetic response of metallic nanocolloids in aqueous solution using a solar simulator,” Opt. Express 20, 3054–3065 (2012). 16. M. Moocarme and L. T. Vuong, “Ultralow-intensity magneto-optical and mechanical effects in metal nanocolloids,” Nano Lett. 14, 1178–1183 (2014). 17. G. Gouesbet and J. A. Lock, “List of problems for future research in generalized Lorenz–Mie theories and related topics, review and prospectus,” Appl. Opt. 52, 897–916 (2013). 18. V. Y. Bazhenov, M. V. Vasnetsov, M. S. Soskin, and V. B. Tarenenko, “Self-oscillations of a liquid near a free interface during continuous local heating,” JETP Lett. 49, 376–379 (1989). 19. G. Gouesbet and C. Roze, “Instabilities by local heating below an interface,” J. Non-Equilib. Thermodyn. 25, 337–379 (2001). 20. E. Greenfield, J. Nemirovsky, R. El-Ganainy, N. Demetri, and M. Segev, “Shockwave based nonlinear optical manipulation in densely scattering opaque suspensions,” Opt. Express 21, 2382–2384 (2013). 21. B. Ambravaneswaran, S. D. Phillips, and O. Basaran, “Theoretical analysis of a dripping faucet,” Phys. Rev. Lett. 85, 5332–5335 (2000). 22. A. Mialdun, V. Yasnou, V. Shevtsova, A. Königer, W. Köhler, D. Alonso de Mezquia, and M. M. Bou-Ali, “A comprehensive study of diffusion, thermodiffusion, and Soret coefficients of water–isopropanol mixtures,” J. Chem. Phys. 136, 244512 (2012). 23. A. Bergeon, D. Henry, H. Benhadid, and L. S. Tuckerman, “Marangoni convection in binary mixtures with Soret effect,” J. Fluid Mech. 375, 143–177 (1998). 24. F. Franks and D. J. G. Ives, “The structural properties of alcohol–water mixtures,” Q. Rev. Chem. Soc. 20, 1–44 (1966). 25. J. Glinski ´ and A. Burakowski, “Compressibility of aqueous solutions of nonelectrolytes: an equilibrium model,” J. Chem. Phys. 132, 124507 (2010). 26. G. Roux, D. Roberts, G. Perron, and J. Desnoyers, “Microheterogeneity in aqueous-organic solutions: heat capacities, volumes and expansibilities of some alcohols, aminoalcohol and tertiary amines in water,” J. Solution Chem. 9, 629–647 (1980). 27. S. Dixit, J. Crain, W. C. K. Poon, J. L. Finney, and A. K. Soper, “Molecular segregation observed in a concentrated alcohol–water solution,” Nature 416, 829–832 (2002). 28. A. Würger, “Thermal non-equilibrium transport in colloids,” Rep. Prog. Phys. 73, 126601 (2010). 29. U. Buhvestov, F. Rived, C. Ra, E. Bosch, and A. Rose, “Solute–solvent and solvent–solvent interactions in binary solvent mixtures. Part 7. Comparison of the enhancement of the water structure in alcohol–water mixtures measured by solvatochromic indicators,” J. Phys. Org. Chem. 11, 185–192 (1998). 30. M. Sedlák and D. Rak, “Large-scale inhomogeneities in solutions of low molar mass compounds and mixtures of liquids: supramolecular structures or nanobubbles?” J. Phys. Chem. B 117, 2495–2504 (2013). 31. K. Svoboda and S. M. Block, “Optical trapping of metallic Rayleigh particles,” Opt. Lett. 19, 930–932 (1994). 32. M. Dienerowitz, M. Mazilu, and K. Dholakia, “Optical manipulation of nanoparticles: a review,” J. Nanophoton. 2, 021875 (2008). 33. S. Baral, A. Green, M. Y. Livshits, A. O. Govarov, and H. H. Richardson, “Comparison of vapor formation of water at the solid/water interface to colloidal solutions using optically excited gold nanostructures,” ACS Nano 8, 1439–1448 (2014). 34. A. Häbich, W. Ducker, D. E. Dunstan, and X. Zhang, “Do stable nanobubbles exist in mixtures of organic solvents and water?” J. Phys. Chem. B 114, 6962–6967 (2010). 35. F. Jin, J. Ye, L. Hong, H. Lam, and C. Wu, “Slow relaxation mode in mixtures of water and organic molecules: supramolecular structures or nanobubbles?” J. Phys. Chem. B 111, 2255–2261 (2007). 36. M. Chaplin, “Theory vs experiment: what is the surface charge of water?” Water 1, 1–28 (2009). 37. A. Pinke and P. Jedlovsky, “Modeling of mixing acetone and water: how can their full miscibility be reproduced in computer simulations?” J. Phys. Chem. B 116, 5977–5984 (2012).

Research Article 38. M. Sedlak and D. Rak, “On the origin of mesoscale structures in aqueous solutions of tertiary butyl alcohol: the mystery resolved,” J. Phys. Chem. B 118, 2726–2737 (2014). 39. E. Ruckenstein and I. Shulgin, “Hydrophobic self-assembling in dilute aqueous solutions of alcohols and hydrocarbons,” Chem. Eng. Sci. 56, 5675–5680 (2001). 40. H. Hayashi, K. Nishikawa, and T. Iijima, “Small-angle x-ray scattering study of fluctuations in 1-propanol–water and 2-propanol–water systems,” J. Phys. Chem. 94, 8334–8338 (1990). 41. B. L. Smorodin, I. N. Cherepanov, B. I. Myznikova, and M. I. Shliomis, “Traveling-wave convection in colloids stratified by gravity,” Phys. Rev. E 84, 026305 (2011). 42. I. M. Vellekoop and A. P. Mosk, “Focusing coherent light through opaque strongly scattering media,” Opt. Lett. 32, 2309–2311 (2007). 43. A. Bricard, J.-B. Caussin, N. Desreumaux, O. Dauchot, and D. Bartolo, “Emergence of macroscopic directed motion in populations of motile colloids,” Nature 503, 95–98 (2013). 44. H. Amini, E. Sollier, M. Masaeli, Y. Xie, B. Ganapathysubramanian, H. A. Stone, and D. Di Carlo, “Engineering fluid flow using sequenced microstructures,” Nat. Commun. 4, 1826 (2013).

Vol. 2, No. 5 / May 2015 / Optica

453

45. E. Palleau, D. Morales, M. D. Dickey, and O. D. Velev, “Reversible patterning and actuation of hydrogels by electrically assisted ionoprinting,” Nat. Commun. 4, 2257 (2013). 46. R. Rica, R. Ziano, D. Salerno, F. Mantegazza, R. van Roij, and D. Brogioli, “Capacitive mixing for harvesting the free energy of solutions at different concentrations,” Entropy 15, 1388–1407 (2013). 47. N. Y. Yip and M. Elimelech, “Thermodynamic and energy efficiency analysis of power generation from natural salinity gradients by pressure retarded osmosis,” Environ. Sci. Technol. 46, 5230–5239 (2012). 48. J. W. Post, J. Veerman, H. V. M. Hamelers, G. J. W. Euverink, S. J. Metz, K. Nymeijer, and C. J. N. Buisman, “Salinity-gradient power: evaluation of pressure-retarded osmosis and reverse electrodialysis,” J. Membr. Sci. 288, 218–230 (2007). 49. D. Brogioli, “Extracting renewable energy from a salinity difference using a capacitor,” Phys. Rev. Lett. 103, 058501 (2009). 50. L. F. Range, J. C. Santamarina, and D. O. Fratta, “Dynamic electrical– mechanical energy coupling in electrolyte–mineral systems,” Transp. Porous Media 50, 153–178 (2003).