This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Thermo-osmotic slip flows around a thermophoretic microparticle
characterized by optical trapping of tracers

Tetsuro Tsuji [email protected]    Satoshi Mei    Satoshi Taguchi Graduate School of Informatics, Kyoto University, Kyoto 606-8501, Japan
Abstract

Thermo-osmotic flow around a microparticle in a liquid is characterized by observing and analyzing the distribution of tiny particles, i.e., tracers, near the microparticle’s surface. First, an optical trapping laser is used to localize the tracer motion along a circular path near the circumference of the microparticle. Then, upon creating an overall temperature gradient in the liquid, the tracers on the circular path, originally uniformly distributed, gather towards the hotter side of the microparticle, indicating a flow along the particle toward the hot. Analyzing the tracer distribution further, it is found that (i) the flow magnitude decreases with the distance from the surface, and (ii) changing the surface property of the microparticle results in a change in the flow magnitude. These show that the observed flow is a thermally induced slip flow along the microparticle’s surface. Then, assuming a simple slip boundary condition for a fluid equation, we evaluate the magnitude of the slip coefficient based on two experimental data: (i) the thermophoretic velocity of the microparticle and (ii) the thermo-osmotic flow around the microparticle. The results of the two approaches are in quantitative agreement. They are also compared with those of theoretical models for a slip flow in existing studies.

Thermo-osmosis, Slip flow, Thermophoresis, Optical trapping, Optical heating, Nanofluidics, Electrokinetics
preprint: , ver. 880

When a fluid is in contact with a solid surface with inhomogeneous temperature distribution, thermally-induced slip flows manifest. These flows are often called thermo-osmosis or thermo-osmotic flows in the literature [1, 2, 3, 4, 5], being considered novel tools for micro/nanoscale material transport [6, 7, 8] and manipulation [9, 10]. Thermo-osmosis has multiple physical origins such as an excess enthalpy variation [11, 12], thermo-electric [13, 14, 15], and Marangoni effects [16, 17, 18].

A growing interest in thermo-osmosis originates from recent development on photothermal effects and thermoplasmonics [19, 20, 21], where fluids and/or solid surfaces can be heated locally in nanoscale. These localized heat sources lead to a steep temperature gradient, inducing thermo-osmotic flows of the detectable order of magnitude, e.g., a few tens of µm s-1 [10]. Nonetheless, due to multi-physics nature and difficulties in experimental evaluation, comparisons between theory and experiment are considered challenging. To tackle this situation, the present paper develops a new tool for the experimental evaluation of thermo-osmotic slip flows, and apply it to the characterization of the flows around microparticles.

Before explaining the developed tools, let us describe why we are interested in the flows around particles. Thermo-osmosis around particles is particularly important among its applications, because it is one of the driving mechanisms of thermophoresis of tiny particles [22, 23, 24, 25] and active matters [26, 27, 28, 29]. Thermophoresis has been widely explored for separating a mixture [30, 31, 32, 33], molecular characterization in bioscience [34, 35, 36], and opto-thermal manipulation of nanomaterial [37, 38] and biomaterial [39]. Nevertheless, to the extent of authors’ knowledge, the profiles of thermo-osmotic flows around thermophoretic particles have not been evaluated, and thus, the contribution of the flow to thermophoresis remains unclear. The motivation of the present paper is to present the direct connection between the thermophoresis of a particle and the thermally-induced flow around it. In Refs. [40] and [26], the thermally-induced slip flows around a micro- and Janus particle were observed, respectively, but the flow profile was not quantified. Thermoosmotic flows on the flat surface were observed in [1, 10] and that around a Janus particle in [41]. However, because the thermally-slip flows occur near the surface, tracers for flow visualization were scarce, preventing a systematic flow characterization.

Refer to caption
Figure 1: (a) Concept of a proposed scheme for the investigation of thermally-induced flow around a microparticle. Tracers are optically trapped around the microparticle, and play the role of flow indicators upon laser heating of a surrounding fluid. (b,c) Schematic of experiments: (b) A side view and (c) a bottom view (i.e., a camera view) of a test section. (b) A target particle with a diameter of 77 µm is immobilized at the upper surface of a microchannel with a height of 1717 µm. A heating laser is irradiated at L=18L=18 µm away from the center of the target. (c) A scanning trapping laser on a circular path with a radius of RR confine fluorescent tracers around the target microparticle.

Now we explain the proposed tool in this paper. To overcome the experimental difficulty above, we use optical trapping (or optical tweezer) [42, 43]. In recent years, optical trapping has been a powerful tool for the control of tiny objects [44, 45]. Here, using the optical tweezer, we trap tracers as flow indicators (Fig. 1(a)). By confining the tracers optically in a test section, which is a circular path around a target microparticle, we facilitate the visualization of thermally-induced flows near the particle surface immersed in a water solution. In particular, we focus on the spatial dependence of the flow, the surface characteristics of the microparticle, and the connection between thermo-osmosis and thermophoresis.

I Experimental method

We briefly explain the present experimental method, leaving details in Appendix A. A target polystyrene particle with a diameter of d=7d=7 µm is immobilized at the upper surface of a microchannel with a height of 1717 µm, as shown in Fig. 1(b). The immobilization is necessary to focus on a flow around the target particle without its translocation.

Then, a focused trapping laser with a wavelength of 1064 nm is irradiated to the upper surface of the microchannel. The trapping laser is scanned circularly around the target particle so that polystyrene fluorescent tracers with a diameter of 500500 nm are optically trapped around the target particle, as shown in Fig. 1(c). A scanning circular path has a radius RR, which is an experimental parameter and chosen as R=4.4R=4.4, 5.45.4, 6.56.5, or 7.57.5 µm. We use a polar coordinate system with a radial position rr and an angle θ\theta measured from a line connecting the center of the target particle and a heating spot (Fig. 1(c)). The optical trapping is loose on the circular path, on which the tracers undergo the Brownian motion in the θ\theta direction (Fig. 1(c)).

After a while, many tracers are held on the circular path so that the spatial distribution of tracers become almost uniform with respect to θ\theta, as shown in Fig. 1(a). Then, at time t=0t=0 s, a focused heating laser with a wavelength of 14801480 nm and with a power PP is irradiated to the upper surface of the microchannel at a position distant from the target particle (Fig. 1(b,c)). The temperature gradient of the fluid is kept by a photothermal conversion: the absorption of a near-infrared laser to water [46, 47, 48, 49]. Such optical heating is necessary to achieve a steep temperature gradient for the generation of the thermo-osmotic flow with a detectable magnitude.

The distance between the target particle and the focus of the heating laser is set to L=18.0±0.1L=18.0\pm 0.1 µm, and the distance from the heating laser is denoted by rr^{\prime} (Fig. 1(c)). The results of the temperature measurement for the heating laser power P=13.5P=13.5, 30.230.2, 46.746.7, and 78.778.7 mW are summarized in the Appendix B; the magnitude of the temperature gradient is 1\sim 1 K µm-1. We observe the change of the tracer distribution on the circular path in the xyx\,y plane (Fig. 1(c)) after the onset of laser heating.

II Model

In this section, we present a simple model that describes the physical situation of the experiment. For simplicity, we consider a steady state, which is realized after a few tens of seconds of the heating laser irradiation. Furthermore, we neglect the presence of microchannel wall and consider the temperature variation directed only to the xx direction.

II.1 Fluids

The temperature of the fluid and that inside the target particle are denoted by TfT_{f} and TsT_{s}, respectively. Since the thermal Péclet number 𝑃𝑒T\mathit{Pe}_{T} is small, i.e., 𝑃𝑒T1\mathit{Pe}_{T}\ll 1, we can neglect the convection term; the temperature fields TfT_{f} and TsT_{s} obey the steady heat conduction equations. For instance, when the reference length is 1010 µm, the reference speed is 11 µm, and the thermal diffusivity is 1.4×1071.4\times 10^{-7} m2 s-1, we have 𝑃𝑒T=7×105\mathit{Pe}_{T}=7\times 10^{-5}.

The condition for the temperature imposed at infinity is TfT0+αxT_{f}\to T_{0}+\alpha x (r|𝒙|)(r\equiv|\bm{x}|\to\infty) with a constant temperature gradient α(>0)\alpha(>0). Note that the xx direction corresponds to θ=0\theta=0 in Fig. 1(c), and T0T_{0} is a reference temperature at the center of the target particle. At the surface of the target particle r=a(=d/2)r=a(=d/2), the temperature and the heat flux in the normal direction to the sphere surface are continuous. Then, we readily obtain

TfT0αa=(1+a3r3ξf)racosθ,ξfκfκs2κf+κs,\displaystyle\frac{T_{f}-T_{0}}{\alpha a}=\left(1+\frac{a^{3}}{r^{3}}\xi_{f}\right)\frac{r}{a}\cos\theta,\quad\xi_{f}\equiv\frac{\kappa_{f}-\kappa_{s}}{2\kappa_{f}+\kappa_{s}}, (1a)
TsT0αa=ξsracosθ,ξs3κf2κf+κs(=ξf+1),\displaystyle\frac{T_{s}-T_{0}}{\alpha a}=\xi_{s}\frac{r}{a}\cos\theta,\quad\xi_{s}\equiv\frac{3\kappa_{f}}{2\kappa_{f}+\kappa_{s}}(=\xi_{f}+1), (1b)

where αa\alpha a is a characteristic temperature difference over a distance aa; κf\kappa_{f} and κs\kappa_{s} are the thermal conductivity of the fluid and the target particle, respectively; ξf\xi_{f} and ξs\xi_{s} are non-dimensional parameters related with the thermal conductivity. In this paper, the temperature dependence of thermal conductivity is not considered, i.e., the constants κf=0.6\kappa_{f}=0.6 W m-1 K-1, κs=0.2\kappa_{s}=0.2 W m-1 K-1, ξf0.286\xi_{f}\approx 0.286, and ξs1.286\xi_{s}\approx 1.286 are used.

The temperature field (1) induces a fluid flow through a slip boundary condition, that is, the flow velocity vector 𝒖\bm{u} satisfies, on the sphere surface,

{ur𝒖𝒆r=0,uθ𝒖𝒆θ=K𝒆θTf,(r=a),\begin{cases}u_{r}\equiv\bm{u}\cdot\bm{e}_{r}=0,\\ u_{\theta}\equiv\bm{u}\cdot\bm{e}_{\theta}=-K\bm{e}_{\theta}\cdot\nabla T_{f},\end{cases}\quad(r=a), (2)

where 𝒆r\bm{e}_{r} and 𝒆θ\bm{e}_{\theta} are the unit vectors in the rr and θ\theta directions, respectively, and KK is a slip coefficient; KK is related to a thermo-osmotic coefficient χ\chi [1, 10] as χ=KTf|r=a\chi=-KT_{f}|_{r=a}. Note that 𝒆θTf=𝒆θTs\bm{e}_{\theta}\cdot\nabla T_{f}=\bm{e}_{\theta}\cdot\nabla T_{s} at r=ar=a. Assuming that the Reynolds number 𝑅𝑒\mathit{Re} is small, i.e., 𝑅𝑒1\mathit{Re}\ll 1 (e.g., a kinematic viscosity of 106\sim 10^{-6} m2 s-1 leads to 𝑅𝑒105\mathit{Re}\sim 10^{-5}), the flow field 𝒖\bm{u} satisfies the Stokes equation, which can be solved explicitly under the boundary condition (2) and the condition imposed at infinity, namely, 𝒖=0\bm{u}=0 at rr\to\infty. The results are

urαK=ξs(1+a2r2)arcosθ,\displaystyle\frac{u_{r}}{\alpha K}=\xi_{s}\left(-1+\frac{a^{2}}{r^{2}}\right)\frac{a}{r}\cos\theta, (3a)
uθαK=ξs(1+a2r2)a2rsinθ,\displaystyle\frac{u_{\theta}}{\alpha K}=\xi_{s}\left(1+\frac{a^{2}}{r^{2}}\right)\frac{a}{2r}\sin\theta, (3b)
pαK(η/a)=ξsa2r2cosθ,\displaystyle\frac{p}{\alpha K(\eta/a)}=-\xi_{s}\frac{a^{2}}{r^{2}}\cos\theta, (3c)

where pp is the pressure, αK\alpha K is the product of the temperature gradient and the slip coefficient, namely, the characteristic speed of thermally-induced flow; η\eta is the viscosity of the fluid. In particular, on the sphere surface, we have

uθ=ussinθ,usξsKα(r=a),\displaystyle u_{\theta}=-u_{s}\sin\theta,\quad u_{s}\equiv-\xi_{s}K\alpha\quad(r=a), (4)

where the sign of usu_{s} is chosen so that usu_{s} is positive when the slip flow is induced in the negative θ\theta direction, i.e., the hotter side of the target particle. When the thermal conductivity satisfies κs=κf\kappa_{s}=\kappa_{f}, it leads to ξs=1\xi_{s}=1 and thus us=Kαu_{s}=-K\alpha. On the other hand, when the thermal conductivity of the target is huge, e.g., the case of a metal particle, we have κsκf\kappa_{s}\gg\kappa_{f} that yields ξs1\xi_{s}\ll 1, resulting in no slip flows on the sphere surface.

II.2 Target particle

The fluid force FiF_{i} (i=x,y,z)(i=x,\,y,\,z) acting on the target particle is obtained from the surface integral Fi=pijnjdSF_{i}=\int p_{ij}n_{j}\mathrm{d}S, where pijp_{ij} denotes the stress tensor for a Newtonian fluid, nin_{i} is a unit vector on the surface pointing to the fluid, dS\mathrm{d}S is an area element, and the integral is carried out over the sphere surface r=ar=a. The xx component FxF_{x} is then readily obtained as Fx=4πaηξsKαF_{x}=4\pi a\eta\xi_{s}K\alpha using Eq. (3). Therefore, in this model, the xx component of the thermophoretic velocity of the target particle, vTtargetv_{T}^{\mathrm{target}}, is computed from the balance between the force FxF_{x} and the Stokes drag 6πaηvTtarget6\pi a\eta v_{T}^{\mathrm{target}}, that is, vTtarget=(2/3)ξsKαv_{T}^{\mathrm{target}}=(2/3)\xi_{s}K\alpha [=(2/3)us][=-(2/3)u_{s}]. This means that the positive usu_{s} (i.e., a slip flow toward the hot) induces the motion of the target particle toward the cold, vTtarget<0v_{T}^{\mathrm{target}}<0.

Recall that, in general, the thermophoretic velocity 𝒗T\bm{v}_{T} of particles is given by 𝒗T=DTTf\bm{v}_{T}=-D_{T}\nabla T_{f} phenomenologically, where DTD_{T} is called a thermophoretic mobility [22, 23]. In this equation, the term Tf\nabla T_{f} should be understood as the temperature gradient at infinity, that is, the effect of temperature deformation due to the presence of the target particle is confined in the mobility DTD_{T}. Therefore, by noting that vTtarget=𝒆x𝒗Tv_{T}^{\mathrm{target}}=\bm{e}_{x}\cdot\bm{v}_{T} and 𝒆xTf=α\bm{e}_{x}\cdot\nabla T_{f}=\alpha with 𝒆x\bm{e}_{x} a unit vector in the xx direction, the thermophoretic mobility of the target, DTtargetD_{T}^{\mathrm{target}}, is obtained as,

DTtargetvTtargetα=23ξsK,\displaystyle D_{T}^{\mathrm{target}}\equiv-\frac{v_{T}^{\mathrm{target}}}{\alpha}=-\frac{2}{3}\xi_{s}K, (5)

which is the relation between DTtargetD_{T}^{\mathrm{target}} and the slip coefficient KK. Experimentally, the slip coefficient KK can be estimated from (5) through the evaluation of DTtargetD_{T}^{\mathrm{target}}, i.e., vTtargetv_{T}^{\mathrm{target}} and α\alpha.

II.3 Tracers

Now, suppose that the tracers are confined in the circular path r=Rr=R because of the trapping laser (see, Fig. 1(c)). The number density of the tracers is denoted by cc. The velocity of the tracer, 𝒗\bm{v}, is the sum of the flow velocity 𝒖\bm{u} and the thermophoretic velocity 𝒗TtracerDTtracerTf\bm{v}_{T}^{\mathrm{tracer}}\equiv-D_{T}^{\mathrm{tracer}}\nabla T_{f} with DTtracerD_{T}^{\mathrm{tracer}} the thermophoretic mobility of the tracer. In this equation, Tf\nabla T_{f} should be interpreted as a local quantity due to the smallness of the tracer, i.e., Tf\nabla T_{f} is computed from (1a). Then, the tracer distribution cc satisfies a steady drift-diffusion equation with a drift velocity 𝒗𝒖+𝒗Ttracer\bm{v}\equiv\bm{u}+\bm{v}_{T}^{\mathrm{tracer}}. The θ\theta component of 𝒗\bm{v} can be computed explicitly as 𝒗𝒆θ=vsinθ\bm{v}\cdot\bm{e}_{\theta}=-v^{\ast}\sin\theta, where vv^{\ast} is the flow magnitude in the direction toward hotter region, i.e.,

v\displaystyle v^{\ast} =ξs(1+a2r2)a2rαK(1+a3r3ξf)αDTtracer.\displaystyle=-\xi_{s}\left(1+\frac{a^{2}}{r^{2}}\right)\frac{a}{2r}\alpha K-\left(1+\frac{a^{3}}{r^{3}}\xi_{f}\right)\alpha D_{T}^{\mathrm{tracer}}. (6)

Because of the scanning trapping laser at r=Rr=R, we assume that the tracer’s radial motion is suppressed and only the motion in the θ\theta direction is significant. Further assuming that there is no steady flux of the tracers, the drift-diffusion equation is reduced to the balance between the drift and diffusion in the θ\theta direction at r=Rr=R:

c(𝒗𝒆θ)DRcθ=0,\displaystyle c(\bm{v}\cdot\bm{e}_{\theta})-\frac{D}{R}\frac{\partial c}{\partial\theta}=0, (7)

where DD is the diffusion coefficient of the trapped tracers. Equation (7) can be solved explicitly as

cc0=exp(γcosθ),\displaystyle\frac{c}{c_{0}}=\exp(\gamma\cos\theta), (8a)
γv|r=RD/R,\displaystyle\gamma\equiv\frac{v^{\ast}|_{r=R}}{D/R}, (8b)

where c0c_{0} is a normalization constant and the pre-factor γ\gamma (8b) is the ratio between the magnitude of the drift velocity vv^{\ast} and the diffusive speed D/RD/R. Note that γ\gamma depends on the radius of the circular path RR, the thermophoretic mobility DTtargetD_{T}^{\mathrm{target}} (or KK) and DTtracerD_{T}^{\mathrm{tracer}}, and the temperature gradient α\alpha. When γ>0\gamma>0 (or γ<0\gamma<0), the number density of the tracers cc takes maximum at θ=0\theta=0 (or π\pi), i.e., the hotter (or colder) side of the target particle. When |γ|0|\gamma|\approx 0, the diffusion dominates and cc tends to be distributed uniformly on the circular path.

Model (8) is used to fit and analyze the experimental results. Technical details of the fitting procedure are given in the Appendix C. Once γ\gamma, DTtracerD_{T}^{\mathrm{tracer}}, and α\alpha are obtained experimentally, one can evaluate the slip coefficient KK using Eqs. (6) and (8b). Later, we will compare KK thus obtained with that estimated in the method described in Sec. II.2 previously.

III Results and Discussion

III.1 Control experiments

Before going into the main part of experimental results, we show the results of some control experiments.

Refer to caption
Figure 2: Thermophoresis of tracers. Snapshots at (a) t=0t=0, (b) 55, (c) 1010, (d) 2020, and (e) 3030 s, where a heating laser with P=13.5P=13.5 mW is irradiated at the center of images. (f) Relative intensity distribution at t=5t=5, 1010, 2020, and 3030 s averaged over 55 s as the function of the distance rr^{\prime} from the laser.

III.1.1 Thermophoresis of the tracers

Firstly, we observe the thermophoretic motion of tracers in the absence of the target particle. In the previous studies, the tracers are known to move against the temperature gradient with the present experimental condition [32, 50]. That is, DTtracerD_{T}^{\mathrm{tracer}} is positive.

Figure 2(a–e) shows snapshots of the thermophoresis of tracers (whites dots) near the upper surface of the microchannel at time t=0t=0, 55, …, 3030 s, where the heating laser with a power P=13.5P=13.5 mW is irradiated at the center of the images. We observe two characteristic features in the figure: (i) the fluorescent intensity increases near the radial distance r0r^{\prime}\approx 0 µm, where rr^{\prime} is the distance from the heating laser, and (ii) the dark region with low intensity spreads over r10r^{\prime}\gtrsim 10 µm.

The former feature (i) is attributed to the tracers accumulated at the bottom surface. Such accumulation is due to thermophoresis in the negative zz direction [51], which occurs due to a cooling at the bottom glass substrate with a rather high thermal conductivity compared with water and/or the upper surface made of polydimethylsiloxane (PDMS). These are out of the scope of the present paper.

On the other hand, the latter feature (ii) indicates the onset of the thermophoresis of tracers in the radial direction. Figure 2(f) shows the time-development of the normalized intensity distribution in panels (a–e), and the depletion of the tracer, i.e., positive DTtracerD_{T}^{\mathrm{tracer}}, is clearly presented.

Refer to caption
Figure 3: Bright field imaging of the thermophoresis of a target particle with a carboxylate-modified surface. Snapshot at (a) t=0t=0, (b) 11, (c) 33, (d) 77, and (e) 2020 s, where a heating laser is irradiated at the center of images. The target particle is indicated by an arrow and a yellow circle with a diameter of 7 µm in panel (a) for eye guide. Grid spacing is 20 µm. (f) Moving distance as a function of time tt.

III.1.2 Thermophoresis of the target particle

Figure 3 shows the thermophoresis of the target particle before immobilization. Here and in what follows, the target has carboxylate-surface modification unless otherwise stated. Note that a target, without immobilization, settles out onto the bottom surface of the microchannel because of its large sedimentation velocity (>1>1 µm s-1).

As shown in Fig. 3(a) at t=0t=0 s, the target particle is placed at a distance r=19.4r^{\prime}=19.4 µm away from the heating laser. As time goes on (see panels (b–e)), the target particle is repelled from the heated spot, moving over almost 2020 µm in the duration of 2020 s. Figure 3(f) shows the moving distance of the target particle as a function of time, obtained using the particle tracking analysis of an open-source image-processing software, ImageJ.

The average speed of the target particle up to t=2t=2 s is 0.890.89 µm s-1 and the average temperature and temperature gradient over the trajectory is 328328 K and 1.77±0.08-1.77\pm 0.08 K µm-1, respectively, leading to the thermophoretic mobility of the target, DTtarget=0.50±0.17D_{T}^{\mathrm{target}}=0.50\pm 0.17 µm2 s-1 K-1.

III.1.3 Thermophoresis of the trapped tracers

In this section, we show the tracer motion which are trapped on the circular path in the absence of the target particle, as shown in Fig. 4(a). This is a control experiment to assure that the irradiation of the scanning trapping laser does not affect the direction of thermophoresis of tracers. The distance between the center of the scanning trapping laser and the heating laser is set to L=15.5L=15.5 µm and the laser power of the trapping laser is set to 7171 mW.

Figure 4(b–g) are time-averaged images of the tracer with the radius of the circular path R=4.4R=4.4 µm. For instance, panel (b) is the time-averaged images over 5-5 s t0\leq t\leq 0 s. Since these are time-averaged images, if a tracer looks not blurry, it is not a Brownian particle, that is, the tracer adheres to a channel wall (see, e.g., the tracers indicated by arrows in panel (e)).

It is seen that the tracers near θ=0\theta=0 are depleted as time goes on, i.e., the intensity reduces near θ=0\theta=0. Moreover, the overall intensity decreases, indicating that some tracers escape from the optical trapping and move away from the hot region. Figure 4(h) shows the relative intensity distribution near r=Rr=R and t=15t=15 s (i.e., panel (e)) as a function of θ\theta. The fitting is made by Eq. (8) with γ=0.52\gamma=-0.52 (see Appendix C), indicating the motion of tracers toward the colder side. In the absence of the target particle, the temperature gradient is uniform (see, Eq. (1a)) and the slip flow is absent. Therefore, |γ||\gamma| corresponds to the phoretic Péclet number [52], that is, the magnitude |γ|O(1)|\gamma|\approx O(1) indicates that the thermophoresis of the tracers is competing with the diffusion of themselves. Noting that ξf=0\xi_{f}=0 (or κf=κs\kappa_{f}=\kappa_{s}) and K=0K=0 without the target particle, the value of γ=0.52\gamma=-0.52 results in DTtracer=0.10D_{T}^{\mathrm{tracer}}=0.10 µm2 s-1 K-1 using Eqs. (8b) and (6).

Refer to caption
Figure 4: (a) Schematic of an experiment in the absence of a target particle. (b–g) Time-averaged images over 5 s. Arrows in panel (e) indicates tracers stuck to the channel wall. These stuck tracers can be noise in the intensity analysis. Scale bar: 44 µm. (h) Relative intensity profile at r=Rr=R with respect to θ\theta.
Refer to caption
Figure 5: (a) Schematic of experiments in the presence of a target particle with a carboxylate-modified surface (ζ=46.0\zeta=-46.0 mV). RR is the radius of a scanning trapping laser and PP is the power of a heating laser. (b–f) Time-averaged images over 5 s for various PP and RR: (b,c,f) P=78.7P=78.7 mW, (d) P=46.7P=46.7 mW, and (e) P=30.2P=30.2 mW; (b) R=4.4R=4.4 µm, (c–e) R=5.4R=5.4 µm, and (f) R=6.5R=6.5 µm. Since these are time-averaged images, if a tracer looks not blurry, it adheres to the upper channel wall. (g,h) Relative intensity profile as a function of θ\theta; (g) P=78.7P=78.7 mW (t=25t=25 s) for various RR and (h) R=5.4R=5.4 µm for various PP (t=30t=30 s). See also a movie“Movie-Fig5(b)-carboxylate-mod.avi” for panel (b).

III.2 Observation of thermally-induced flows

Let us describe the main result of the present paper, the thermally-induced flow around the target particle. Firstly, the results for a target particle with a carboxylate-modified surface are presented. The zeta potential ζ\zeta of target particles is measured from the electrophoretic mobility (ELSZ-2000Z, Otsuka Electronics) as ζ=46\zeta=-46 mV. Figure 5(a) shows the schematic of the situation presented in panels (b–f), which are the recorded images for various parameters P=78.7P=78.7. 46.746.7, and 30.230.2 mW (the power of the heating laser) and R=4.4R=4.4, 5.45.4, 6.56.5, and 7.57.5 µm (the radius of the scanning trapping laser). One can compare the effect of RR by inspecting panels (b,c,f) and that of PP by panels (c,d,e). Lowering the laser power to P=13.5P=13.5 mW did not show a significant change in the tracer distribution from the initial state.

Firstly, it is seen from panel (b) that the tracers, which are initially rather uniform in the θ\theta variable, tend to gather toward the hotter side (i.e., the right side, see panel (a)) as time goes on. This tendency is the same for other panels (c–f), and should be contrasted with Fig. 4 for a case without a target particle; the tracers gather toward the colder side.

As can be seen from the comparison between panels (b) R=4.4R=4.4 µm, (c) R=5.4R=5.4 µm, and (f) R=6.5R=6.5 µm, the tracers’ movement is more intense and faster for smaller RR. To be more precise, at time t=5t=5 s, panel (b) starts to exhibit the depletion of the tracers near θ=π\theta=\pi (the colder side; see panel (a)); however panels (c) and (f) show the apparent depletion near θ=π\theta=\pi only for t10t\geq 10 s. At t=30t=30 s, where the distributions can be considered steady, the intensity near θ=0\theta=0 seems to reach maximum, indicating the accumulation near the hot side. To compare more quantitatively, we show in panel (g) the relative intensity distribution (symbols) and the fitting using Eq. (8) (lines) (see also Appendix C). Although the results are qualitative, the plots in panel (g) capture the observation made in panels (b,c,f).

It should be noted that some tracers seem trapped in the region between the upper wall of the microchannel and the target particle (see the region “𝖠\mathsf{A}” in Fig. 1(b)), as clearly seen in Fig. 5(f) at t10t\geq 10 s. These tracers can stay there even when the scanning trapping laser is not irradiated. We suppose that the thermo-osmotic flow and the thermophoresis in the zz direction may cause this trapping, but further discussion is difficult using only the present two-dimensional information on the xyx\,y plane, and thus future work.

One can compare the effect of the heating laser power PP by seeing (c) P=78.7P=78.7 mW, (d) P=46.7P=46.7 mW, and (e) P=30.2P=30.2 mW. The corresponding temperature-gradient measurement given in Appendix B shows that |Tf||\nabla T_{f}| is approximately (b) 2.12.1 K µm-1, (c) 1.61.6 K µm-1, and (d) 1.01.0 K µm-1 near rL=18r^{\prime}\approx L=18 µm, at which the target particle is fixed. Therefore, the tracer movement is more intense for the case with larger PP. Figure 5(h) shows the corresponding relative intensity distribution at t=30t=30 s. Sharper shape near θ=0\theta=0 for larger PP indicates the more intense accumulation of tracers near the hotter side.

Next, let us describe the thermally-induced flow around a target particle with an amine-modified surface, the zeta potential ζ\zeta of which is measured as ζ=23.7\zeta=-23.7 mV. The thermophoretic mobility of a target particle with the amine-modified surface is estimated in the same manner as Sec. III.1.2, and we obtain DTtarget=0.29±0.08D_{T}^{\mathrm{target}}=0.29\pm 0.08 µm2 s-1 K-1 where the average temperature and temperature gradient are 320320 K and 1.18-1.18 K µm-1, respectively. This value of DTtargetD_{T}^{\mathrm{target}} is smaller than that of a target particle with the carboxylate-modified surface (see Sec. III.1.2).

Figure 6(a) shows the schematic of the situation presented in panels (b–g), which are recorded images for P=78.7P=78.7 mW and R=4.4R=4.4 µm with a target particle with the amine-modified surface. See Fig. 5(b) for comparison with the case of a carboxylate-modified surface. Remarkably, the case of the amine-modified target results in a faint accumulation of tracers. As shown in panels (b-g), the tracers slightly gather toward the hotter side, but the accumulation is very weak compared with Fig. 5(b) for a carboxylate-modified target. This observation is further verified by the profile of the relative intensity distribution shown in panel (h) for both amine (circle) and carboxylate (square) modified surfaces. Since the surface modification is the only difference between these two cases, panel (h) is a clear evidence that the observed tracer movement originates from surface phenomena.

Refer to caption
Figure 6: (a) Schematic of experiments in the presence of a target particle with an amine-modified surface (ζ=23.7\zeta=-23.7 mV). (b–g) Time-averaged images over 5 s. Scale bar: 44 µm. (h) Relative intensity profile at r=Rr=R with respect to θ\theta for amine-modified (circle) and carboxylate-modified (square; see also Fig. 5(b) for microscope images). See also a movie“Movie-Fig6(b-g)-amine-mod.avi” for panel (b–g).

III.3 Discussion

To compare the experimental results with existing models of thermally-induced slip flows, we use the model proposed in Ref. [13], in which a thin electric-double-layer (EDL) approximation for a charged surface of a spherical colloid is applied under the Debye-Hückel approximation. The resulting slip velocity is written as

vB=v0sinθ(r=a),\displaystyle v_{B}=-v_{0}\sin\theta\quad(r=a), (9a)
v0εψ028η(1dlnεdlnTf)ξs|Tf|Tf,\displaystyle v_{0}\equiv\frac{\varepsilon\psi_{0}^{2}}{8\eta}\left(1-\frac{\mathrm{d}\ln\varepsilon}{\mathrm{d}\ln T_{f}}\right)\xi_{s}\frac{|\nabla T_{f}|_{\infty}}{T_{f}}, (9b)

where |Tf||(=α)|\nabla T_{f}|_{\infty}|(=\alpha) is the temperature gradient at infinity, ε\varepsilon is the temperature-dependent electric permittivity, and ψ0\psi_{0} is the electrostatic potential at the surface. We assume that ψ0ζ\psi_{0}\approx\zeta. The temperature dependence of ε\varepsilon for water is modeled as ε/ε0=305.7×exp(Tf/217)\varepsilon/\varepsilon_{0}=305.7\times\exp(-T_{f}/217) [53] where ε0=8.85×1012\varepsilon_{0}=8.85\times 10^{-12} F m-1 is the electric permittivity in vacuum. Furthermore, the empirical model of viscosity η=(2.761×106)×exp(1713/Tf)\eta=(2.761\times 10^{-6})\times\exp(1713/T_{f}) Pa s [54] is used. For the above model, the slip coefficient KK in Eq. (4) is expressed as K=v0/(ξsα)K=-v_{0}/(\xi_{s}\alpha) because us=v0u_{s}=v_{0}. To be more specific, we have

K=εζ28ηTf(1dlnεdlnTf).\displaystyle K=-\frac{\varepsilon\zeta^{2}}{8\eta T_{f}}\left(1-\frac{\mathrm{d}\ln\varepsilon}{\mathrm{d}\ln T_{f}}\right). (10)

In Ref. [1], a similar form is proposed based on the Derjaguin’s model [55] with an enthalpy excess given by EDL enthalpy with the Debye-Hückel approximation. This model was considered to analyze the thermo-osmosis on a bare glass substrate. The result was

K=χTf=εζ28ηTf.\displaystyle K=-\frac{\chi}{T_{f}}=-\frac{\varepsilon\zeta^{2}}{8\eta T_{f}}. (11)

The theoretical prediction KK obtained by Eqs. (10) and (11) can be compared with KK obtained from the experiments via two approaches. (Method 1) First, KK can be obtained from the thermophoretic motion of the target, Eq. (5), because we have obtained DTtargetD_{T}^{\mathrm{target}} experimentally by the particle tracking. See Sec. III.1.2. (Method 2) Second, KK can be obtained from the observation of slip flows in Sec. III.2, namely, the tracer accumulation behavior (8), through the fitting of the value γ\gamma (see Figs. 5 and 6 for the result of fitting). Then, equation (6) and γ\gamma yield KK since we have already evaluated DTtracerD_{T}^{\mathrm{tracer}} in Sec. III.1.1.

Before going into the comparison between Eq. (10), Eq. (11), Method 1, and Method 2, we show the values of γ\gamma used in Method 2. Figure 7(a) is the values of γ\gamma averaged from t=10t=10 s to 3030 s for the heating laser power P=78.7P=78.7 mW (circle), 46.746.7 mW (square), and 30.230.2 mW (triangle) as a function of the scanning radius RR. Filled symbols indicate the results of the carboxylate-modified target particles and empty symbols those of the amine-modified target particles. Although qualitative, we see that γ\gamma tends to decrease as RR, that is, the magnitude of thermally-induced flow decreases. Moreover, comparison over different laser power PP shows that γ\gamma decreases with PP. The cases of amine-modified target yield significantly smaller γ\gamma than those of the carboxylate-modified target. These values of γ\gamma can be converted into KK through Eqs. (8b) and (6); the results are summarized in Fig. 7(b). For the case of carboxylate-modified surface, KK seems to take the value around 0.4-0.4 µm2 s-1 K-1, while that of amine-modified surface results in K0.2K\approx-0.2 µm2 s-1 K-1.

Finally, we summarize in Table 1 the results of KK obtained by Eq. (10), Eq. (11), Method 1, and Method 2, where the average over different RR is taken for Method 2 (see Fig. 7(b)). Furthermore, Table 2 shows the values of KK computed from the existing experimental results of thermo-osmotic flows [1, 10] with different surface details. All the values of KK in both tables are negative, i.e., the flow is induced from cold to hot side. The magnitude of KK in Table 1 has the same order O(1)O(1) µm2 s-1 K-1 but quantitative agreement is not obtained between theory and experiments. However, importantly, Methods 1 and 2 show similar values of KK and capture the difference of surface modification predicted by Eqs. (10) and (11): KK for the carboxylate-modified case is larger than that for the amine-modified case. Comparing Tables 1 and 2, it is seen that the slip coefficients obtained in the present experiments are close to that for bare glass.

The thermo-osmotic coefficients have been also computed using molecular dynamics simulation in Refs. [2] and [56] for Lennard-Jones fluid/solid system and water/quartz system, respectively. The results were χO(108)\chi\sim O(10^{-8}) m2 s-1 and one or two orders magnitude larger than theory and experiments.

The reason for the quantitative discrepancy between theory and experiment cannot be fully clarified at the moment, but let us point out some possibilities. First, Eq. (10) was derived under the thin-EDL and Debye–Hückel approximation. The former is valid because we use an electrolyte solution with 10 mM ionic concentration. But the latter may be violated due to the non-negligible magnitude of the zeta potential compared with kBTfk_{B}T_{f}. Experiments, i.e., Method 1 or Method 2, may underestimate vTtargetv_{T}^{\mathrm{target}} or KK because the target particle stays close to the microchannel walls due to a technical limitation in the experiment. Further improvement of the proposed method and its application to other systems is future work.

Refer to caption
Figure 7: (a) The values of γ\gamma as a function of a scanning radius RR for carboxylate-modified (COOH, filled symbols) and amine-modified (NH2, empty symbols) surfaces with various heating power P=78.7P=78.7 (circle), 46.746.7 (square), and 30.230.2 mW (triangle). (b) Slip coefficient KK computed from γ\gamma in panel (a) through Eqs. (8b) and (6)\eqref{eq:vast}.
Table 1: Slip coefficient KK [µm2 s-1 K-1] obtained from the model (10) and (11) based on Refs. [13, 1] and the present experimental results with Method 1 and Method 2.
Eq. (10) Eq. (11) Method 1 Method 2
Ref. [13] Ref. [1] P=78.7P=78.7 mW P=46.7P=46.7 mW P=30.2P=30.2 mW
carboxylate-mod. (ζ=46.0\zeta=-46.0 mV) 2.36-2.36 0.94-0.94 0.58±0.20-0.58\pm 0.20 0.50±0.02-0.50\pm 0.02 0.44±0.02-0.44\pm 0.02 0.45±0.03-0.45\pm 0.03
amine-mod. (ζ=23.7\zeta=-23.7 mV) 0.58-0.58 0.23-0.23 0.34±0.10-0.34\pm 0.10 0.25±0.01-0.25\pm 0.01
Table 2: Slip coefficient K=χ/TfK=-\chi/T_{f} [µm2 s-1 K-1] reported in Ref. [1] for a glass substrate with or without non-ionic copolymer surfactant (Pluronic F-127) coating and Ref. [10] for a thin gold film on a glass. In the references, the thermo-osmotic coefficient χ\chi is reported and the values shown in this table are obtained for Tf=300T_{f}=300 K and 330330 K.
Wall condition (experiments) χ\chi [m2 s-1] KK at Tf=300T_{f}=300 K KK at Tf=330T_{f}=330 K
Glass coated with Pluronic F-127 [1] 13×101013\times 10^{-10} 4.3-4.3 3.9-3.9
Bare glass [1] 1.8×10101.8\times 10^{-10} 0.6-0.6 0.5-0.5
Gold-thin film on a glass [10] 10×101010\times 10^{-10} 3.3-3.3 3.0-3.0

IV Conclusion

In this paper, we proposed an experimental method to characterize thermo-osmotic slip flows around a microparticle. The method utilized the optical trapping of tracers around a microparticle, which indicated the generation of thermally-induced flows near the particle surface through the change of tracer’s spatial distribution. We applied the proposed method to the flow characterization around a microparticle under a strong temperature gradient created by the optical heating of a water solution.

The slip coefficient was estimated experimentally by using two methods: a thermophoretic velocity measurement of a target particle (Method 1) and a thermo-osmotic flow measurement around a target particle (Method 2), where the two methods were related through a simple physical model assuming a thermally-induced slip boundary condition. The values of slip coefficients obtained by the two methods were in reasonable agreement, that is, we demonstrated that a flow was indeed induced along the surface of the particle under the temperature gradient and the agreement provides an evidence that the flow was the major driving mechanism of thermophoresis. Furthermore, the experiments captured the difference of the surface modification predicted by the theory of charged surface [13] with temperature variation.

The use of the optical trapping of tracers, as demonstrated in the present paper, can also be applied to other flows near surfaces, such as those along the surface of a heated Janus particle. Investigating not only the motion of active matter but also the induced flows around them will explore a further understanding of nanofluidics and their novel applications.

Acknowledgements.
This work was partly supported by JSPS KAKENHI grant No. JP18H05242, JP20H02067, and JP22K18770, JP22K03924, and also by JST PRESTO grant No. JPMJPR22O7.
Refer to caption
Figure 8: An overview of the experimental setup. AF: absorption filters, AOD: acousto-optic deflector, DM: dichroic mirrors, EF: excitation filter, HWP: half-wave plate. An inset shows a degassing protocol.

Appendix A Experimental details

A.1 Overview

A schematic of the experiment is shown in Fig. 8. A microfluidics device made of PDMS block and a cover glass is mounted on an inverted microscope (IX–73, Olympus). The fabrication process of the microchannel was given in [32]. The straight microchannel has a rectangular cross section with a height of H=17H=17 µm in the zz direction, a width of 400400 µm in the xx direction and the length >3>3 mm in the yy direction (see also Fig. 1). The height H=17H=17 µm is small enough to suppress the effect of thermal convection.

Two lasers with the wavelengths of 1064 nm (Powerwave–1064, NPI lasers) and 1480 nm (FOL1439R50–667–1480, Furukawa Electric) are irradiated to the microchannel through an objective (×\times50, numerical aperture (NA) = 0.65; LCPLN50xIR, Olympus); the former/latter is called a trapping/heating laser in this paper. The trapping laser is scanned in the focal plane by using a two-axis acousto-optic deflector (AOD; DTD-274HD6M & DE-272M, Intra Action) and captures tracers for flow visualization. The heating laser creates an inhomogeneous temperature field by using a photothermal effect of a water solution. Note that the absorption coefficient β\beta of the water at the wavelength of 1480 nm is β=2.4×103\beta=2.4\times 10^{3} m-1 [47, 48]. This value of β\beta is large enough for the necessary temperature elevation with a moderate laser power <100<100 mW.

A background fluid flow in the microchannel in the yy direction is controlled by adjusting a pressure difference ΔP\Delta P between inlet and outlet reservoirs (Fig. 8). To be more specific, ΔP\Delta P is controlled with a resolution 0.010.01 Pa by the water-level difference Δh\Delta h between the two reservoirs using a zz-stage. Before every experiment, we confirm that the background fluid flow is absent. After an experiment, there arises a non-uniform tracer distribution. Therefore, we induce a background flow to supply a fresh solution with a uniform tracer distribution to the test section.

Any residue bubbles in the microchannel, connecting tubes, and/or reservoirs can lead to the loss of reproducibility of the experiments. To be more precise, if there are any residue bubbles, (a) the background flow never stops and there is always a very slow flow (<1<1 µm) that can hinder the observation of thermally-induced flows; (b) undesired and uncontrollable tracer motion occurs just after an onset of the heating laser irradiation. Therefore, we pay special attention to the removal of the bubbles. More specifically, before mounting the microfluidic device on the microscope, the whole device is degassed by putting only the PDMS-glass part in a vacuum chamber, as shown in the inset of Fig. 8. Since the PDMS can absorb a gas [57], the degassing can avoid the creation of bubbles and also can remove the residue bubble in the microchannel. By further degassing the empty outlet (see the inset of Fig. 8) using a hand pump, we can successfully fill the whole channel (including tubes) with a sample solution without residue bubbles.

A.2 Sample solution

Target particles are non-fluorescent polystyrene spheres with a diameter of 77 µm and with carboxylate (01-02-703, Micromod) or amine (01-01-703, Micromod) surface modification. Tracer particles are fluorescent polystyrene spheres with yellow-green fluorescence (F8813, Molecular Probes) and have a diameter of 500500 nm and a carboxylate surface modification. These as-purchased particles are dispersed in a tris(hydroxymethyl)amino-methane hydrochloride (tris-HCl) buffer solution (pH=8.0\mathrm{pH}=8.0, Nippon Gene) with a concentration of 10 mM. The concentration of the target particles and tracers are 2.4×1042.4\times 10^{-4} % and 4×1024\times 10^{-2} %, respectively. The sample solution is sonicated for 5 minutes before use to avoid the agglomeration of tracers.

A.3 Optical setup

Two collimated lasers (for trapping and heating) are expanded using a set of plano-convex lenses, as shown in Fig. 8. A half-wave plate (HWP) is used for the optical path of the trapping laser to adjust the direction of linear polarization to that of the specification of AOD. The lenses L1, L2, and AOD are placed at the conjugate planes of the back aperture of the objective lens, that is, the motion of L1 and L2 perpendicular to the lenses and the beam deflection at AOD can rotate the beam at the back aperture. In this way, we can move the laser focus in the focal plane in the xx and yy directions with minimal beam clipping [58].

Mercury lamp (U-HGLGPS, Olympus) with a band path filter (460 nm–495 nm) is used for an excitation light source in fluorescent imaging. A fluorescence from the tracers is collected by the same objective and is recorded by a scientific complementary-metal-oxide-semiconductor (sCMOS) camera (Zyla 5.5, Andor Technology), after removing undesired stray lights using absorption filters. The field depth in the zz direction for this observation setup is 66 µm, as measured in our previous experiments with the same objective [50]. To observe the non-fluorescent target particle, we use a transmission light from a halogen lamp.

A.4 Experimental procedure

To immobilize the target particle to the upper surface in the microchannel, we use the trapping laser, following procedures (1)–(4) shown in Fig. 9: (1) Target particles have a non-negligible sedimentation speed. Thus, they tend to stay at a bulk part and rarely come to the microchannel part. (2) We lift a target particle in the bulk by irradiating the trapping laser. (3) We induce a background flow to bring the target particle to the center of the microchannel part. (4) We again lift the target particle by irradiating the trapping laser and push it in the positive zz direction to the upper surface of the microchannel, waiting for about 10 minutes until the target particle sticks to the upper surface.

Note that the yy position of the immobilized particle should be about the center of the microchannel. Otherwise, we observe very slight but non-negligible tracer motion in the yy direction just after the initiation of laser heating. We consider that this unwanted tracer motion does not originate from the residue bubbles’ existence, because the magnitude is much weaker than that of the case with the residue bubbles. In Fig. 1(b), we take the yy axis, i.e., the direction of the background flow, perpendicular to the line of θ=0\theta=0 to assure that the accumulation (or depletion) of tracers (see Sec. III) is not caused by such bias if any.

After the immobilization of the target particle, we scan the trapping laser in a circular path with the radius RR with a power of 178178 mW (after the objective). The scanning frequency is set to 200 Hz, which is high enough to avoid forcing tracers to orbit around the target particle. After the realization of the homogeneous distribution of tracers with respect to θ\theta, the heating laser with a power of PP (after the objective) is irradiated, where P=13.5P=13.5, 30.230.2, 46.746.7, and 78.778.7 mW are used in this paper. The motion of tracers are recorded with a frame rate of 2020 fps for >25>25 s.

Refer to caption
Figure 9: Schematic of an immobilization process. (1) Many target particles are sedimented at the bottom of the bulk part, but (2) they are optically pushed by the laser with a wavelength of 1064 nm so that they disperse again. (3) The dispersed target particles are then driven into the microchannel. (4) The target particles are optically pushed to the ceil of the microchannel until they stick there.
Table 3: Parameters used in the fitting of Fig. 10.
PP [mW] ΔT\Delta T [K] σ\sigma [µm] T0T_{0} [K]
78.778.7 75.7 16.9 24.6
46.746.7 59.5 14.0 24.6
30.230.2 41.6 13.1 24.1
13.513.5 21.4 11.7 21.6

Appendix B Temperature profile

We visualize the temperature profile of the sample solution in the xyx\,y plane using a laser-induced fluorescence method [49, 59, 47, 48, 50]. The measurement process is exactly same as that in Refs. [49, 50] and the details are omitted here. Note that the measured temperature profile below is the temperature of the fluid integrated over the depth (i.e., the zz) direction.

Figure 10(a) shows the result of temperature measurement for the laser power PP used in this paper. The experimental data are shown by symbols and we fit these plots using half Lorentzian curves, T0+ΔT/(1+(r/σ)2)T_{0}+\Delta T/(1+(r^{\prime}/\sigma)^{2}), where ΔT\Delta T is the temperature increase at the beam center, T0T_{0} is the temperature at infinity, σ\sigma is the half-width at the midheight value of the temperature profile, and rr^{\prime} is the distance from the heating spot. The parameters used in the fitting are summarized in Table 3. The corresponding radial gradient is also shown in Fig. 10(b).

Refer to caption
Figure 10: (a) Temperature of a fluid and (b) a radial temperature gradient as a function of a radial distance rr^{\prime} from a heating spot (see, Fig. 1(c)). The power PP of the laser with a wavelength 1480 nm are set to P=78.7P=78.7, 46.746.7, 30.230.2, and 13.513.5 mW.

Appendix C Data analysis procedure

Let fijnf_{ij}^{n} be the intensity of an obtained image, where the indices ii and jj are the pixel numbers in the xx and yy directions; nn is the frame number. Firstly, we average the images over n0n_{0} frames with n0=100n_{0}=100 (over 5 s) to reduce noise and define them as f¯ijn=m=nn0nf¯ijm\bar{f}^{n}_{ij}=\sum_{m=n-n_{0}}^{n}\bar{f}^{m}_{ij}. Then, we subtract the background value, which is the minimum value over whole pixel of f¯ijn\bar{f}^{n}_{ij}. The θ\theta direction is divided into intervals and we introduce θp=pΔθ\theta_{p}=p\Delta\theta (p=p0,,p0)(p=-p_{0},...,p_{0}) with p0=10p_{0}=10 and Δθ=π/p0\Delta\theta=\pi/p_{0}. To analyze the result with r=Rr=R, we average the intensity f¯ijn\bar{f}^{n}_{ij} over a domain (r,θ)[RΔr,R+Δr]×[θp,θp+Δθ](r,\theta)\in[R-\Delta r,R+\Delta r]\times[\theta_{p},\theta_{p}+\Delta\theta] with Δr258\Delta r\approx 258 nm; This averaged value is named as FpnF_{p}^{n} and we then compute the spatial average over whole θ\theta, i.e., F¯n=12πall pFpnΔθ\bar{F}^{n}=\frac{1}{2\pi}\sum_{\text{all }p}F_{p}^{n}\Delta\theta. The results shown in the main text (e.g., Sec. III.2) are the relative intensity, namely, Fpn/F¯nF_{p}^{n}/\bar{F}^{n} [== (intensity near θ=θp\theta=\theta_{p}) / (intensity averaged over whole θ\theta)].

Next, we describe details of the fitting procedure. In practice, the value of the normalization constant c0c_{0} in Eq. (8) is not free to choose but should be determined by the following manner. We fit the experimental data (i.e., the relative intensity explained above) using the function cfit=c0exp(γcos(θθ))+cc_{\mathrm{fit}}=c_{0}\exp(\gamma\cos(\theta-\theta^{\prime}))+c^{\prime} to obtain γ\gamma, where θ\theta^{\prime} and cc^{\prime} are auxiliary fitting parameters to compensate the unavoidable shifts in the experimental results. Here, c0c_{0} is determined so that ππcfitdθ=1\int_{-\pi}^{\pi}c_{\mathrm{fit}}\mathrm{d}\theta=1, because the experimental data are normalized by the average over whole θ\theta as described above. Thus, c0c_{0} is a function of γ\gamma and cc^{\prime}. By noting the identity ππexp(γcosθ)dθ=2π0(γ)\int_{-\pi}^{\pi}\exp(\gamma\cos\theta)\mathrm{d}\theta=2\pi\mathcal{I}_{0}(\gamma), where 0\mathcal{I}_{0} is the modified Bessel function of the first kind, we get c0=(12πc)/(2π0(γ))c_{0}=(1-2\pi c^{\prime})/(2\pi\mathcal{I}_{0}(\gamma)). The function 0\mathcal{I}_{0} can be approximated by an elementary function [60], which is used in the actual fitting. The above-described procedure is carried out using the software ImageJ.

The adequateness of the fitting is evaluated by the coefficient of determination 𝖱2\mathsf{R}^{2}, where 𝖱2(<1)\mathsf{R}^{2}(<1) close to unity indicates the successful fitting. The values of 𝖱2\mathsf{R}^{2} for the fitting procedure are summarized in Table 4. For the cases with the carboxylate-modified target particles, the signal-to-noise ratio is good and the model (8) fits the experimental data well, i.e., 𝖱21\mathsf{R}^{2}\approx 1. For the cases with the amine-modified target and the case without the target particle, the value of 𝖱2\mathsf{R}^{2} is decreased to >0.6>0.6 for P=78.7P=78.7 mW, but still the model (8) can be considered to fit the experimental data. In the analysis of the present paper, the data with bad values of 𝖱2<0.6\mathsf{R}^{2}<0.6 due to stochastic noise are not used in the case without the target particle.

Table 4: The coefficient of determination 𝖱2\mathsf{R}^{2} for the fitting procedure used in the present paper. For typical examples of the actual fitting curves, see Figs. 4, 5, and 6. For the actual fitting values, see Fig. 7(a).
carboxylate-modified amine-modified no target particle
RR [µm] P=78.7P=78.7 mW P=46.7P=46.7 mW P=30.2P=30.2 mW P=78.7P=78.7 mW P=78.7P=78.7 mW
4.44.4 0.970.97 0.980.98 0.970.97 0.610.61 0.730.73
5.45.4 0.970.97 0.980.98 0.850.85 0.250.25
6.56.5 0.940.94 0.920.92 0.840.84 0.010.01
7.57.5 0.860.86 0.690.69 0.660.66 0.010.01

References

  • Bregulla et al. [2016] A. P. Bregulla, A. Würger, K. Günther, M. Mertig, and F. Cichos, Thermo-osmotic flow in thin films, Physical Review Letters 116, 188303 (2016).
  • Fu et al. [2017] L. Fu, S. Merabia, and L. Joly, What controls thermo-osmosis? Molecular simulations show the critical role of interfacial hydrodynamics, Physical Review Letters 119, 214501 (2017).
  • Proesmans and Frenkel [2019] K. Proesmans and D. Frenkel, Comparing theory and simulation for thermo-osmosis, The Journal of Chemical Physics 151, 124109 (2019).
  • Anzini et al. [2019] P. Anzini, G. M. Colombo, Z. Filiberti, and A. Parola, Thermal forces from a microscopic perspective, Physical Review Letters 123, 028002 (2019).
  • Anzini et al. [2022] P. Anzini, Z. Filiberti, and A. Parola, Fluid flow at interfaces driven by thermal gradients, Physical Review E 106, 024116 (2022).
  • Mousavi et al. [2019] S. M. Mousavi, I. Kasianiuk, D. Kasyanyuk, S. K. P. Velu, A. Callegari, L. Biancofiore, and G. Volpe, Clustering of janus particles in optical potential driven by hydrodynamic fluxes, Soft Matter 15, 5748 (2019).
  • Trivedi et al. [2022] M. Trivedi, D. Saxena, W. K. Ng, R. Sapienza, and G. Volpe, Self-organized lasers from reconfigurable colloidal assemblies, Nature Physics 18, 939–944 (2022).
  • Xu et al. [2023] H. Xu, X. Zheng, and X. Shi, Surface hydrophilicity-mediated migration of nano/microparticles under temperature gradient in a confined space, Journal of Colloid and Interface Science 637, 489 (2023).
  • Lou et al. [2018] X. Lou, N. Yu, R. Liu, K. Chen, and M. Yang, Dynamics of a colloidal particle near a thermoosmotic wall under illumination, Soft Matter 14, 1319 (2018).
  • Fränzl and Cichos [2022] M. Fränzl and F. Cichos, Hydrodynamic manipulation of nano-objects by optically induced thermo-osmotic flows, Nature Communications 13, 656 (2022).
  • Derjaguin et al. [1987] B. V. Derjaguin, N. V. Churaev, and M. M. Muller, Surface Forces (Springer, 1987).
  • Anderson [1989] J. L. Anderson, Colloid transport by interfacial forces, Annual Review of Fluid Mechanics 21, 61 (1989).
  • Fayolle et al. [2008] S. Fayolle, T. Bickel, and A. Würger, Thermophoresis of charged colloidal particles, Physical Review E 77, 041404 (2008).
  • Rasuli and Golestanian [2008] S. N. Rasuli and R. Golestanian, Soret motion of a charged spherical colloid, Physical Review Letters 101, 108301 (2008).
  • Würger [2008] A. Würger, Transport in charged colloids driven by thermoelectricity, Physical Review Letters 101, 108302 (2008).
  • Ruckenstein [1981] E. Ruckenstein, Can phoretic motions be treated as interfacial tension gradient driven phenomena?, Journal of Colloid and Interface Science 83, 77 (1981).
  • Würger [2007] A. Würger, Thermophoresis in colloidal suspensions driven by Marangoni forces, Physical Review Letters 98, 138301 (2007).
  • Kim et al. [2022] Y. Kim, H. Ding, and Y. Zheng, Investigating water/oil interfaces with opto-thermophoresis, Nature Communications 13, 3742 (2022).
  • Baffou [2017] G. Baffou, Thermoplasmonics (Cambridge University Press, 2017).
  • Baffou et al. [2020] G. Baffou, F. Cichos, and R. Quidant, Applications and challenges of thermoplasmonics, Nature Materials 19, 946 (2020).
  • Kotsifaki and Chormaic [2022] D. G. Kotsifaki and S. N. Chormaic, The role of temperature-induced effects generated by plasmonic nanostructures on particle delivery and manipulation: a review, Nanophotonics 11, 2199 (2022).
  • Piazza and Parola [2008] R. Piazza and A. Parola, Thermophoresis in colloidal suspensions, Journal of Physics: Condensed Matter 20, 153102 (2008).
  • Piazza [2008] R. Piazza, Thermophoresis: moving particles with thermal gradients, Soft Matter 4, 1740 (2008).
  • Würger [2010] A. Würger, Thermal non-equilibrium transport in colloids, Reports on Progress in Physics 73, 126601 (2010).
  • Tsuji et al. [2017] T. Tsuji, K. Kozai, H. Ishino, and S. Kawano, Direct observations of thermophoresis in microfluidic systems, Micro & Nano Letters 12, 520 (2017).
  • Jiang et al. [2010] H.-R. Jiang, N. Yoshinaga, and M. Sano, Active motion of a Janus particle by self-thermophoresis in a defocused laser beam, Physical Review Letters 105, 268302 (2010).
  • Bickel et al. [2013] T. Bickel, A. Majee, and A. Würger, Flow pattern in the vicinity of self-propelling hot Janus particles, Physical Review E 88, 012301 (2013).
  • Auschra et al. [2021] S. Auschra, A. Bregulla, K. Kroy, and F. Cichos, Thermotaxis of Janus particles, The European Physical Journal E 44, 90 (2021).
  • Paul et al. [2022] D. Paul, R. Chand, and G. V. P. Kumar, Optothermal evolution of active colloidal matter in a defocused laser trap, ACS Photonics 9, 3440 (2022).
  • Vigolo et al. [2010] D. Vigolo, R. Rusconi, H. A. Stone, and R. Piazza, Thermophoresis: Microfluidics characterization and separation, Soft Matter 6, 3489 (2010).
  • Maeda et al. [2011] Y. T. Maeda, A. Buguin, and A. Libchaber, Thermal separation: Interplay between the soret effect and entropic force gradient, Physical Review Letters 107, 038301 (2011).
  • Tsuji et al. [2018a] T. Tsuji, S. Saita, and S. Kawano, Dynamic pattern formation of microparticles in a uniform flow by an on-chip thermophoretic separation device, Physical Review Applied 9, 024035 (2018a).
  • Sanjuan et al. [2022] A. Sanjuan, A. Errarte, and M. M. Bou-Ali, Analysis of thermophoresis for separation of polystyrene microparticles in microfluidic devices, International Journal of Heat and Mass Transfer 189, 122690 (2022).
  • Wienken et al. [2010] C. J. Wienken, P. Baaske, U. Rothbauer, D. Braun, and S. Duhr, Protein-binding assays in biological liquids using microscale thermophoresis, Nature Communications 1, 100 (2010).
  • Seidel et al. [2012] S. A. I. Seidel, C. J. Wienken, S. Geissler, M. Jerabek-Willemsen, S. Duhr, A. Reiter, D. Trauner, D. Braun, and P. Baaske, Label-free microscale thermophoresis discriminates sites and affinity of protein–ligand binding, Angewandte Chemie International Edition 51, 10656 (2012).
  • Liu et al. [2019] C. Liu, J. Zhao, F. Tian, L. Cai, W. Zhang, Q. Feng, J. Chang, F. Wan, Y. Yang, B. Dai, et al., Low-cost thermophoretic profiling of extracellular-vesicle surface proteins for the early detection and classification of cancers, Nature Biomedical Engineering 3, 183 (2019).
  • Lin et al. [2018] L. Lin, M. Wang, X. Peng, E. N. Lissek, Z. Mao, L. Scarabelli, E. Adkins, S. Coskun, H. E. Unalan, B. A. Korgel, L. M. Liz-Marzán, E.-L. Florin, and Y. Zheng, Opto-thermoelectric nanotweezers, Nature Photonics 12, 195 (2018).
  • Setoura et al. [2019] K. Setoura, T. Tsuji, S. Ito, S. Kawano, and H. Miyasaka, Opto-thermophoretic separation and trapping of plasmonic nanoparticles, Nanoscale 11, 21093 (2019).
  • Lin et al. [2017] L. Lin, X. Peng, X. Wei, Z. Mao, C. Xie, and Y. Zheng, Thermophoretic tweezers for low-power and versatile manipulation of biological cells, ACS Nano 11, 3147 (2017).
  • Weinert and Braun [2008] F. M. Weinert and D. Braun, Observation of slip flow in thermophoresis, Physical Review Letters 101, 168301 (2008).
  • Bregulla and Cichos [2019] A. P. Bregulla and F. Cichos, Flow fields around pinned self-thermophoretic microswimmers under confinement, The Journal of Chemical Physics 151, 044706 (2019).
  • Ashkin and Dziedzic [1971] A. Ashkin and J. M. Dziedzic, Optical levitation by radiation pressure, Applied Physics Letters 19, 283 (1971).
  • Ashkin et al. [1986] A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu, Observation of a single-beam gradient force optical trap for dielectric particles, Optics Letters 11, 288 (1986).
  • Tsuji et al. [2022] T. Tsuji, K. Doi, and S. Kawano, Optical trapping in micro- and nanoconfinement systems: Role of thermo-fluid dynamics and applications, Journal of Photochemistry and Photobiology C: Photochemistry Reviews , 100533 (2022).
  • Volpe et al. [2023] G. Volpe et al., Roadmap for optical tweezers 2023, Journal of Physics: Photonics 5, 022501 (2023).
  • Duhr and Braun [2006] S. Duhr and D. Braun, Why molecules move along a temperature gradient, Proceedings of the National Academy of Sciences 103, 19678 (2006).
  • Cordero et al. [2009] M. L. Cordero, E. Verneuil, F. Gallaire, and C. N. Baroud, Time-resolved temperature rise in a thin liquid film due to laser absorption, Physical Review E 79, 011201 (2009).
  • Rivière et al. [2016] D. Rivière, B. Selva, H. Chraibi, U. Delabre, and J.-P. Delville, Convection flows driven by laser heating of a liquid layer, Physical Review E 93, 023112 (2016).
  • Tsuji et al. [2018b] T. Tsuji, Y. Sasai, and S. Kawano, Thermophoretic manipulation of micro- and nanoparticle flow through a sudden contraction in a microchannel with near-infrared laser irradiation, Physical Review Applied 10, 044005 (2018b).
  • Tsuji et al. [2021] T. Tsuji, S. Taguchi, and H. Takamatsu, Switching between laser-induced thermophoresis and thermal convection of liquid suspension in a microgap with variable dimension, Electrophoresis 42, 2401 (2021).
  • Duhr and Braun [2005] S. Duhr and D. Braun, Two-dimensional colloidal crystals formed by thermophoresis and convection, Applied Physics Letters 86, 131921 (2005).
  • Mayer et al. [2023] D. B. Mayer, T. Franosch, C. Mast, and D. Braun, Thermophoresis beyond local thermodynamic equilibrium, Physical Review Letters 130, 168202 (2023).
  • Xuan et al. [2004] X. Xuan, B. Xu, D. Sinton, and D. Li, Electroosmotic flow with joule heating effects, Lab on a Chip 4, 230 (2004).
  • Knox and McCormack [1994] J. H. Knox and K. A. McCormack, Temperature effects in capillary electrophoresis. 1: Internal capillary temperature and effect upon performance, Chromatographia 38, 207 (1994).
  • Derjaguin and Yalamov [1965] B. V. Derjaguin and Y. Yalamov, Theory of thermophoresis of large aerosol particles, Journal of Colloid Science 20, 555 (1965).
  • Chen et al. [2021] W. Q. Chen, M. Sedighi, and A. P. Jivkov, Thermo-osmosis in hydrophilic nanochannels: mechanism and size effect, Nanoscale 13, 1696 (2021).
  • Hosokawa et al. [2004] K. Hosokawa, K. Sato, N. Ichikawa, and M. Maeda, Power-free poly (dimethylsiloxane) microfluidic devices for gold nanoparticle-based dna analysis, Lab on a Chip 4, 181 (2004).
  • Neuman and Block [2004] K. C. Neuman and S. M. Block, Optical trapping, Review of Scientific Instruments 75, 2787 (2004).
  • Braun and Libchaber [2002] D. Braun and A. Libchaber, Trapping of DNA by thermophoretic depletion and convection, Physical Review Letters 89, 188103 (2002).
  • Olivares et al. [2018] J. Olivares, P. Martin, and E. Valero, A simple approximation for the modified bessel function of zero order I0(x){I}_{0}(x)Journal of Physics: Conference Series 1043, 012003 (2018).