Paper The following article is Open access

Galilean relativity for Brownian dynamics and energetics

, , and

Published 6 August 2021 © 2021 The Author(s). Published by IOP Publishing Ltd on behalf of the Institute of Physics and Deutsche Physikalische Gesellschaft
, , Citation Minghao Li et al 2021 New J. Phys. 23 083012 DOI 10.1088/1367-2630/ac15b5

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/23/8/083012

Abstract

We study experimentally the impact of inertial reference frame changes on overdamped Brownian motion. The reference frame changes are implemented by inducing, with a laser, laminar convection flows in a column of fluid where Brownian microspheres are dispersed. The convection flow plays the role of the relative velocity between the laboratory and the fluid comoving frames, and enables us to analyse the consequences of Galilean transformations on Brownian diffusion. We verify in particular how the Brownian dynamics remains 'weakly' Galilean invariant, in agreement with recent discussions Cairoli et al (2018 Proc. Natl Acad. Sci. USA 115 5714). We also carefully look at the consequences of Galilean relativity on the Brownian energetics. This leads us to derive a Galilean invariant expression of the stochastic thermodynamic first law, consistent with existing theoretical results Speck et al (2008 Phys. Rev. Lett. 100 178302). We finally discuss a potential ambiguity of the Galilean relativistic features of diffusive systems that has obvious practical implications in the context of force measurements in external flows.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

Introduction

The laws of classical mechanics are written in inertial reference frames, interconnected by Galilean transformations (GT) that ensure the invariance of the acceleration. If one further assumes that masses and forces are invariant through GT, Newton's laws then become invariant within the whole class of inertial reference frames, according to the principle of Galilean relativity [3].

It is however well-known that stochastic diffusive models break Galilean invariance (GI) because friction and random forces entering the stochastic equations of motion, such as the Langevin equation, emerge from coarse-graining procedures performed in the preferred reference frame where the fluid is at rest [4, 5]. This selection of a reference frame is in contradiction with GI and this difficulty demands proper transformation rules for describing the coarse-grained dynamics in different inertial reference frames [5]. Such rules have been recently given theoretically, from both dynamic [1] and energetic [2] view points. The Langevin description of the Brownian motion between the preferred and a moving inertial frames was shown to differ only by a drift term that corresponds to the difference in the thermal noise statistics between the two frames. With this difference, the Langevin equations written in each frame are related by GT performed on positions and velocities only. This led in particular to maintain a 'weak' GI for the description of the stochastic system with motional probability density functions (PDF) that are only shifted according to the GT that interconnects the two frames [1]. A fundamental consequence in this relativistic framework is the necessity to modify the stochastic energetics in order to account for this drift term with frame invariant definitions of work, heat and entropy (see appendix C) [2]. As a consequence, it is crucial to carefully recognize and evaluate flows in Brownian experiments, such as Brownian dynamic force measurements using the drift model [6], Brownian heat engines [7, 8], Brownian systems with growing domains [9], etc.

In this article, we experimentally explore these issues, by measuring the GT rules on an overdamped diffusive system under a well controlled laminar flow and by assessing the dynamic and energetic consequences of Galilean relativity on Brownian motion. We extend the discussion to the definition of Galilean invariant energetics quantities (work, heat, potential energy, entropy) and to the evaluation of the different PDF associated with the first law. These PDF are built experimentally and their transformations through GT are verified. The analysis reveals the crucial importance in distinguishing the Galilean relativistic drift from an 'induced' force field in order to correctly describe the thermodynamics features of an overdamped Brownian system in a comoving frame. In addition, the control obtained on GT through the stable laminar flow allows a fine-tuning of colloidal levitating regimes that can be exploited in optofluidic systems for colloidal transport [10, 11]. This tuning is also potentially interesting to implement when studying the mechanical response of colloidal ensembles under weak external force fields [1214] or hydrodynamic interactions in the context of ordering effects and phase transitions in colloidal assemblies [15, 16].

1. Experimental inertial frame changes

Our experiment consists in recording under laser illumination the Brownian motion of melamine micron-sized, spherical, beads dispersed in water inside a fluidic cell [14]. The experimental configuration is described in figure 1 and further in appendix A. When the laser is off, the beads simply diffuse and sediment in the laboratory frame along the vertical axis. But as soon as the laser is switched on, water inside the cell heats up, with local modifications of its density ρ. Under conditions detailed below, this laser heating effect brings the fluid into uniform motion that can be precisely controlled along the vertical axis. The beads, dragged by the constant hydrodynamic flow hence generated, diffuse in a reference frame comoving with the fluid and interconnected to the laboratory frame through a GT. We monitor in the laboratory frame the Brownian motion of the beads and describe the dynamic and the energetic features associated with the GT.

Figure 1.

Figure 1. Schematics of the experiment. (a) The cell containing the colloidal suspension of melamine microspheres in water consists of a quartz cuvette of dimensions 10(x) × 2(y) × 35(z) mm3. The counterpropagating laser beams are sent through the centre of the cuvette along the y-axis. The region of interest (ROI) is chosen at the centre of the cell far from any wall, allowing us to neglect any perturbation of the walls on the diffusion dynamics. (b) The profile of the laser beam with a waist of 65 μm is viewed in the (x, z)-plane where the arrows indicate qualitatively the convection flows due to the laser heating water. The DOF determining the dimension of the ROI in the x-direction is of ca. 10 μm, i.e. very far from the cell's walls. Within the ROI, the convection can be therefore considered as laminar and uniform. (c) This laminar convection flow within the ROI act as a Galilean transformation for the Brownian diffusion along the z-axis, interconnecting, with a velocity ${v}_{0}\hat{z}$, the gravitational acceleration $\mathbf{g}=-g\hat{z}$, the two inertial reference frames, the laboratory frame S and the comoving, fluid rest frame S'.

Standard image High-resolution image

Technically, specific requirements have to be met that determine the configuration schematized in figure 1. First, the fluidic cell is chosen sufficiently large so that it is possible to define in its centre an imaging ROI far from the cell walls. The hydrodynamics within the ROI is thus described without the influence of boundary-wall conditions. The cell therefore is traversed by collinear, counterpropagating Gaussian laser beams of common waist w0 ≃ 65 μm with a Rayleigh length zR = 18 mm much larger than the cell width. Within the ROI, the Gaussian profiles of the laser beams can be considered as uniform along the optical y-axis. The waist is also much larger than the diameter of a single bead (d = 0.94 μm) so that large statistical ensembles of displacements can be measured within the ROI [17]. The dimension of the ROI along the observation x-axis is set by the depth-of-field (DOF) of the imaging objective (in our case, ca. 10 μm) which is smaller than the laser waist. Finally, the small volume fraction ϕ ∼ 10−6 of the micron-sized colloidal dispersion used in the experiment is such that the Brownian motion monitored within the ROI can be described in the absence of any hydrodynamic interaction between the beads. The practically plane-wave illumination conditions minimize any gradient contribution in the optical force field thus only determined by scattering contributions, i.e. radiation pressure. The counterpropagating beam configuration allows to cancel any effect of radiation pressure [14] by simply tuning the intensities in each beams to even values.

With such dimensions and conditions of illumination, the laser sets the fluid into laminar motion inside the ROI by a heat convection effect. Such a convective dynamics is described by coupling the equation of heat under laser illumination to the Navier–Stokes (NS) equation for the transport of fluid momentum per unit volume ρ v in the presence of a diluted, homogeneous, colloidal dispersion. Within the Boussinesq's approximation [18], the change in density associated with the laser heating of the fluid δρ = −αρδT is assumed to be such that δρρ, where α is the thermal expansion coefficient of water and δT = T(r) − T0 the difference between the local temperature and the background temperature of water inside the cell.

Under such an approximation, the heat and NS coupled equations write as:

Equation (1)

Equation (2)

with the convective derivative Dt = ∂t + v ⋅ ∇, cp the specific heat capacity of water, k its thermal conductivity, μ its shear viscosity—we assume standard (room temperature) values cp = 4.18 × 103 J K−1 kg−1, k = 0.62 W K−1 m−1 and μ = 0.87 × 10−3 Pa s. On the heat transport equation, ${\dot {q}}_{L}=2A{P}_{0}/\pi {w}_{0}^{2}\enspace \mathrm{exp}(-2({x}^{2}+{z}^{2})/{w}_{0}^{2})$ is the volumetric heat rate generated, at its waist, by the Gaussian laser of power P0, taking for water an absorption coefficient A = 0.3 m−1 at a wavelength of 633 nm. On the NS equation, ϕΔρ g corresponds to the external body force exerted on a unit volume of water by the sedimenting ensemble of colloidal spheres with $\mathbf{g}=-g\hat{z}$ the gravitational acceleration, Δρ the density difference (513 kg m−3) between a single melamine sphere and water and ϕ the volume fraction inside the ROI (ϕ ∼ 10−6).

Because of the very thin fluid layer defined within the imaging DOF thickness DOF ∼ 10 μm—see figure 1(b)—the velocity v(r) of the convection flow is such that vx ∼ 0. We also assume vy ∼ 0 within the ROI positioned far from all walls. This is fully consistent with our observations that reveal a convection flow laminar within the DOF layer in the vertical z-direction, yielding therefore $\mathbf{v}(\mathbf{r})={v}_{0}(\mathbf{r})\hat{z}$ with v0(r) corresponding to the velocity of frame change. With the Boussinesq's approximation that preserves the incompressibility condition ∇ ⋅ v = 0, we further write $\mathbf{v}(\mathbf{r})={v}_{0}(x,y)\hat{z}$. This cancels the convective contribution in the Lagrangian derivative where Dt v ∼ ∂t v in the NS equation. Since the measured convection flow are of the order of 10−6 m s−1 and temperature changes δT under ca. 100 mW laser irradiation that we evaluate to be at the mK level, the convective contribution to Dt δT can also be neglected in the heat equation. As a consequence, the two equations are decoupled and can be solved in the steady-state, as detailed in appendix. The corresponding thermal and velocity profiles are plotted in figure 2. Temperature and velocity boundary layer are evaluated as δT ∼ 0.9 mm and δv ∼ 2 mm respectively. With δv DOF, the convection flow is strictly laminar within the DOF.

Figure 2.

Figure 2. (a) Laser-induced temperature difference δT and frame change velocity v0 profiles due to the laser-induced convection, calculated within the ROI and along the x-axis by the approximated solution of equations (1) and (2) detailed in appendix B for a laser power of 36 mW and a mean volume fraction ϕ ∼ 6.8 × 10−7 which corresponds to the order of magnitude of our experimental conditions. Temperature and velocity boundary layer thicknesses are marked as δT and δv , respectively. (b) Profile maxima δTmax and frame change velocity within our ROI v0(x = 0) as a function of the illumination laser power.

Standard image High-resolution image

Within our approximation, the fluid along the vertical z-axis can be considered as fully uniform across the ROI since the height of the cell is much larger than that of the ROI. Of course, this upward flow is eventually compensated by a downward flow, as expected from mass conservation and incompressibility of the fluid. But this downward flow is close to the cell walls and far away from our ROI, with therefore no impact on our tracking experiments all performed within the ROI, as seen in figures 1(a) and (b).

The central point of our scheme is that in the steady-state regime, inside the ROI, this flow of velocity v0 uniformly carries the colloidal beads and hence defines an inertial reference frame (z', t') where the fluid is at rest. This comoving frame S' is related to the laboratory frame S by a GT where the velocity of S' with respect to S is set and controlled by the laser illumination power P0.

2. Brownian dynamics in different inertial frames

In the comoving inertial frame S', the Brownian motion of each bead is performed under the constant sedimentation force field $\mathcal{M}\mathbf{g}=(\pi {\mathrm{d}}^{3}{\Delta}\rho /6)\mathbf{g}$ resulting from the competition between gravity and buoyancy with $\mathcal{M}$ the effective buoyant mass. The motion is described by the Langevin equation written along the z'-axis as:

Equation (3)

where ξ' is the random thermal force with $\left\langle {\xi }^{\prime }({t}^{\prime })\right\rangle =0$ and $\left\langle {\xi }^{\prime }({t}_{1}^{\prime }){\xi }^{\prime }({t}_{2}^{\prime })\right\rangle =\delta ({t}_{1}^{\prime }-{t}_{2}^{\prime })$, γ is the friction coefficient, kB is the Boltzmann constant and T is the temperature.

Moving to the laboratory frame (z, t) can be simply done by the GT $z=\mathcal{G}[{z}^{\prime }]={z}^{\prime }+{v}_{0}t$, $t=\mathcal{G}[{t}^{\prime }]={t}^{\prime }$ performed on the velocity of the Langevin equation (3) while leaving the noise ξ' unchanged to give:

Equation (4)

As explained in [1], the possibility to do so is physically rooted in the fact that the drift term γv0 induced by the GT fundamentally corresponds to the modification of the thermal noise statistics between the two inertial frames, with $\sqrt{2{k}_{\mathrm{B}}T\gamma }{\xi }^{\prime }({t}^{\prime })+\gamma {v}_{0}=\sqrt{2{k}_{\mathrm{B}}T\gamma }\xi (t)$. This connection leads to a 'weak' GI of the Langevin equation. This additional drift has important thermodynamic consequences that we discuss below.

We first analyse the Brownian motion in the laboratory frame under different laser illumination powers. All the corresponding trajectories of the microspheres within the ROI are recorded from successive images that give access to the succession of vertical displacements Δzi (tk ) = zi (tk+1) − zi (tk ) measured, for one trajectory i, at a fixed frame rate f = 1/(tk+1tk ). When the laser power is weak, the microspheres collectively sediment in the cell with an ensemble of trajectories displayed in figure 3(a). By increasing the laser power, the convection flow is induced and drags the spheres upward. It is easy to find a power value (i.e. a convection velocity) that can practically compensate sedimentation, leaving the spheres suspended in the laboratory frame. The trajectories corresponding to this case are shown in figure 3(b). Convection can even take over sedimentation if the laser power is further increased, as seen clearly in figure 3(c). These trajectories can be analysed, as a function of the time lag Δ, by the mean square displacement (MSD) averaged on the ensemble of N trajectories recorded within the ROI

Equation (5)

a stationary quantity independent of the initial time t. The three MSD plotted on figure 3(d) clearly reveal the dynamics in the three different cases, with parabolic Δ2 MSD in both sedimentation and convection regimes. For the suspended case, the MSD is practically linear in Δ, a feature that corresponds to the free-like Brownian dynamics observed in figure 3(b). We also show in figure 3(e) that the MSD evaluated from the ensemble of displacements recorded along the y-axis remains perfectly linear, confirming that the convection flow is only induced by the laser along the vertical z-axis, implementing the GT detailed above.

Figure 3.

Figure 3. (a)–(c). Individual experimental trajectories recorded in the (y, z) plane of the ROI for three different regimes controlled by the laser power: collective sedimentation regime (low laser power at 10 mW), suspension regime (laser power tuned at 24 mW) and lift regime (high laser power at 36 mW), with the starting point of each trajectory marked by a red dot. The global flow direction and its relative strength are indicated by the blue arrows on the left-hand side of the graphs. (d)–(e) MSDs calculated in both the z-panel (d) and y-panel (e) directions from the trajectories observed in the three regimes presented in (a) blue circles, (b) red triangles and (c) orange crosses. The MSD along the y direction in the three regimes remains linear, revealing a normal Brownian diffusion with no external flow or force acting on this direction. In contrast, the MSD along the z direction in regimes (a) and (c) are parabolic. But for a well-tuned laser power, the MSD along the z direction can remain linear, corresponding to the remarkable suspension regime observed in the laboratory frame on the trajectories displayed in (b).

Standard image High-resolution image

From these ensembles, it is also possible to construct the displacement PDF associated with each regime of convection. In the comoving fluid frame, the Brownian spheres diffuse in the gravity force field characterized by a single particle sedimentation velocity

Equation (6)

along the z'-axis according to equation (3). In this frame therefore, the PDF for vertical displacements is given by

Equation (7)

Here, Δz' is a displacement measured along the z'-axis within a time lag Δt', and D = kB T/γ the diffusion coefficient in the vertical direction. In the laboratory frame, equation (4) yields

Equation (8)

where vz = v0 + vsed is the mean velocity along the z-axis in the laboratory frame, resulting from sedimentation and convection flows. The comparison between P'(Δz', Δt') and Pz, Δt) verifies the relation between the PDF acquired in different inertial frames—see below [1]:

Equation (9)

The experimental PDF constructed in the laboratory frame are displayed in figures 4(a)–(c) for the three different values of laser illumination power. We also extract from these PDF the mean velocity ${v}_{z}=\left\langle {\Delta}z/{\Delta}t\right\rangle $ whose evolution as a function of the laser power is plotted in figure 4(d). The dispersions in the vz values measured for fixed power levels show that vz is relatively stable in time when averaged throughout the ROI, confirming that the combination of sedimentation and convection flows describes a GT. Since the PDF plotted in the laboratory frame mix displacements measured on different trajectories and at different times, it is interesting to extract the diffusion coefficient D from a fit of the PDF variances at each time lag Δt. All D thus extracted are plotted in figure 4(e) for all different laser powers. The agreement between these values extracted from experimental data and the theoretically expected diffusion coefficient (including uncertainties associated with the size dispersion of the colloidal suspension) demonstrates that the laser induced convection flow corresponds to a genuine drift term and does not affect the Brownian noise spectrum. This constitutes an experimental proof that a change of inertial reference frames modifies the noise spectrum as $\xi (t)={\xi }^{\prime }({t}^{\prime })+{v}_{0}\gamma /\sqrt{2{k}_{\mathrm{B}}T\gamma }$. As discussed above, this drift term exactly corresponds to the GT performed directly on the velocity $\dot {{z}^{\prime }}$, preserving the fluctuation–dissipation theorem [19].

Figure 4.

Figure 4. (a)–(c) PDF for the displacements measured along the z-axis in the laboratory frame Pz, Δt) evolving with time lag Δt from 1 s to 4 s—colored data points—and PDF for the vertical displacements of Galilean transformed reconstructed trajectories (see main text) shown as grey data points evolving with the time lag Δt from 1 s to 4 s. Such a PDF corresponds to the dynamics expected in the fluid rest frame (S') P'(Δz', Δt'). One clearly sees that all the PDF hence reconstructed are all similar, while they correspond each to different dynamical regimes when observed from the laboratory frame. These regimes are measured for the different particle velocities vz shown in (d) for the corresponding laser powers. These velocities vz = ⟨Δzt⟩ are measured by time ensemble averages of vertical displacements Δz in the laboratory frame—within the smallest time lag Δt = 1/120 corresponding to the inverse of frame rate—divided by Δt with the standard error indicated in the data as an error bar. Positive values for vz correspond to convection flows moving upwards. When vz < 0, the convection flow is not strong enough (low laser power regimes) to compensate the collective sedimentation flow. (e) Brownian diffusion coefficients along the z-axis (red points) obtained by fitting linearly the variances of the PDF for each Δt with error bars coming from the 99% linear fitting confidence interval. Each variance is obtained by a Gaussian fit made on each PDF. The expected value of the diffusion coefficient derived from the temperature (T = 299 K), viscosity (η = 0.87 × 10−3 N s m−2) and particle diameter is also displayed (black solid line) with the inaccuracy coming from the remaining uncertainty in the temperature determination (δT = ±1 K), viscosity (δη = ±0.02 × 10−3 N s m−2) and particle diameter (δd = ±50 nm). This inaccuracy in the determination of the diffusion coefficient is represented by the grey shaded zone.

Standard image High-resolution image

With such a flow stable in time and homogeneous in space within the ROI, the comoving frame velocity v0 can be estimated from the determination of vz using the PDF Pz, Δt) and the knowledge of vsed assuming that the sole external force field exerted along the z-axis results from buoyancy. With this, the trajectories in the comoving fluid frame can be reconstructed by applying a GT to each trajectory recorded in the laboratory frame. The PDF in this frame can then be built and they are plotted in panels (a)–(c). The comparison clearly shows that the laboratory PDF are related to the comoving PDF by a GT, ${\Delta}z=\mathcal{G}[{\Delta}{z}^{\prime }]={\Delta}{z}^{\prime }+{v}_{0}{\Delta}{t}^{\prime }$ with $P({\Delta}z,{\Delta}t)=\mathcal{G}[{P}^{\prime }({\Delta}{z}^{\prime },{\Delta}{t}^{\prime })]={P}^{\prime }(\mathcal{G}[{\Delta}{z}^{\prime }],{\Delta}t)$ with P'(Δz', Δt') defined above. This is in perfect agreement with the 'weak' GI principle proposed by [1].

3. GI for Brownian stochastic energetics

We now look carefully at the energetic consequences of such a change of reference frame. To do so, the influence of the GT induced by an external flow must be accounted for in the definition of the production rates associated with the stochastic thermodynamics of Brownian motion, while keeping in mind that the drift term related to GT fixes the choice of the reference frame.

Following Sekimoto's [20] and Seifert's approaches [2], the stochastic work production rate in the presence of an external flow v0 within a potential U and a nonconservative force f is defined as

Equation (10)

Accordingly, the first law leads to a heat production rate written as:

Equation (11)

Here, the sign of the heat is taken to be positive if the energy is dissipated into the bath, following Seifert's convention.

As discussed above, an external flow induced in a chosen frame is equivalent to moving the observer with respect to the fluid according to the principle of Galilean relativity. Therefore, the external flow v0 present in equations (10) and (11) can be associated with the flow connecting the laboratory frame to the fluid reference frame through a GT. With this connection, equations (10) and (11) can be used to describe inertial reference frame changes and can be specified to the case of our experiments as soon as the external flow velocity v0 is measured, as it is done in section 2.

The force $\mathcal{M}\mathbf{g}$ in equation (3), resulting from gravity and buoyancy, derives from a potential energy defined in the laboratory frame $U(z,t)=\mathcal{M}gz$. The principle of Galilean relativity enforces that U(z, t) = U'(zv0 t, t). This implies that the potential in the fluid comoving frame reads as ${U}^{\prime }({z}^{\prime },{t}^{\prime })=\mathcal{M}g({z}^{\prime }+{v}_{0}{t}^{\prime })$.

Using equation (10), the expression of work production rate in the fluid frame where v0 = 0 is simply ${\dot {w}}^{\prime }={\partial }_{t}{U}^{\prime }({z}^{\prime },{t}^{\prime })=\mathcal{M}g{v}_{0}$. In the laboratory frame, $\dot {w}=({\partial }_{t}+{v}_{0}\cdot \nabla )U(z,t)=\mathcal{M}g{v}_{0}$. In this way, GI is verified with

Equation (12)

Similarly, from equation (11), the heat production rate in the fluid reference frame is $\dot {{q}^{\prime }}=-\nabla {U}^{\prime }({z}^{\prime },{t}^{\prime })\cdot \dot {{z}^{\prime }}=-\mathcal{M}g\dot {{z}^{\prime }}$ and $\dot {q}=-\mathcal{M}g(\dot {z}-{v}_{0})$ in the laboratory frame. As expected from GI:

Equation (13)

Finally, we recover the expression for the rate of potential energy change using the first law:

Equation (14)

Equations (12)–(14) manifest that a change of inertial reference frame performed by changing v0 leaves invariant heat and work production rates, and eventually thus the first law.

The above analysis makes also clear that the invariance of heat corresponds to the fact that in the fluid reference frame, the stochastic heat production rate is independent of v0 and only depends on the displacement z' measured in the fluid frame. In contrast, the work production rate depends on v0 since it accounts for the deterministic energy contribution injected by the illuminating laser in the system in order to bring the whole column of fluid into motion in a constant potential defined in the laboratory frame. The work does not come from an external force exerted on the particle itself.

As discussed further below, this distinction is important to appreciate in the energetic context. Our setup therefore yields a stochastic energetics different from the one at play on colloidal suspensions under shear flows ([21], e.g.) or when a Brownian particle is optically trapped and the trap is dragged through the fluid at a constant velocity. In this case, the fluid actually works on the confined particle, bringing it out of equilibrium. Such a scheme has been described in details in [22].

In our configuration, the quantities experimentally determined are the PDF associated with the cumulated energetics calculated from the production rates in equations (12)–(14). These PDF can be directly built from the displacements measured in any chosen reference frame over a time lag Δt. For example in the laboratory frame, the cumulated energetics can be calculated by choosing a displacement Δz measured over Δt:

Equation (15)

Equation (16)

Equation (17)

From equation (8), we easily calculate:

Equation (18)

Equation (19)

Equation (20)

Considering that W has one value for a given convection velocity v0. We experimentally set three different convection flows dragging the Brownian particles whose corresponding trajectories are displayed in figures 5(a)–(c). The associated PDF are shown in panels (d)–(f). They are measured over a time lag Δt = 1 s chosen for providing enough statistics with variances of the order of kB T.

Figure 5.

Figure 5. (a)–(c). Experimental trajectories presented in three different regimes of external flow induced using laser powers of 16, 24 and 36 mW respectively with the starting point of each trajectory marked with a red spot. The associated energetic PDF of W, Q and ΔU are calculated from these trajectories with a time lag of 1 s, and plotted in units of kB T for the two following cases. Panels (d)–(f) correspond to the real, physical, case where the drift is performing a GT between the laboratory frame and the fluid rest frame. In this case, the PDF of W has the form of a Dirac distribution showing its deterministic nature. Panels (g)–(i) display the energetic PDF for the case where one considers the drift stemming from an external force field. The differences clearly reveal the ambiguity discussed in the main text.

Standard image High-resolution image

With GI of the energetic quantities, the PDF are identical in all inertial frames. These PDF perfectly obey the first law with ⟨ΔU⟩ = −⟨Q⟩ + ⟨W⟩, knowing ${v}_{\mathrm{s}\mathrm{e}\mathrm{d}}=\langle {\dot {z}}^{\prime }\rangle $, ${v}_{z}=\langle \dot {z}\rangle $ and vz = vsed + v0. As seen on the data of figure 5(d) with a position of P(W, Δt) in the negative energy scale, we start at the lowest laser power with a laser-induced convection not strong enough to compensate for the collective sedimentation effect of the body force. Increasing the laser power reverses the landscape, by passing through the remarkable regime already discussed above in figure 3(b), where the Brownian motion, seen in the laboratory frame, appears as practically free.

We finally want to address an ambiguity of equation (8) that has far-reaching consequences. The ambiguity stems from the fact that the coarse-grained description does not give the possibility for discriminating the drift term that acts as a GT from an external force field added to the sedimentation one. Indeed, equation (8) takes exactly the form of a motional PDF in the presence of a 'resulting' force field shifting the PDF of the free Brownian motion by (v0 + vsedt. In that case, i.e. if one assigns γv0 to a force (a non-conservative force for instance) instead of a flow, the energetic production rates calculated from equations (10) and (11) are modified as: $\dot {\tilde{q}}=(\gamma {v}_{0}-\mathcal{M}g)\dot {z}$, $\dot {\tilde{w}}=\gamma {v}_{0}\dot {z}$, and $\dot {\tilde{u}}=\mathcal{M}g\dot {z}$ expressed in the laboratory frame. As clearly seen in this case when γv0 is considered as an external force, the stochastic energetic production rates, evaluated in different inertial frame always with a different v0, are no longer GI.

The energetic PDF corresponding to this 'wrong-case' scenario write as

Equation (21)

Equation (22)

Equation (23)

and they are plotted in figure 5 panels (g)–(i). Although, these PDF do not violate the first law, the contrast with respect to the frame invariant PDF of equations (13)–(15) gives a striking illustration of the energetic consequences of assigning to the drift term the role of an external force rather than treating it as induced by a change of reference frame.

These discrepancies correspond to the fundamental ambiguity in the interpretation of the trajectories measured in the laboratory frame and displayed in figures 5(a)–(c). Indeed, without an exhausting knowledge of the forces acting on the system or a full characterization of the flow dynamics within the cell, for instance using a fluid tracker, it is impossible to discriminate a force signal from a Galilean transform looking merely at the displacement probability density distributions, in other words to anticipate that a change of reference frame is operating on the system. One can thus be led to analyse the modified dynamics of the recorded trajectories as due to a drag force exerted by the illuminating laser on the colloidal ensemble of spheres and acting against sedimentation. The problem resides precisely in the fact that this viewpoint leads, as we just showed, to a totally wrong energetic balance. It remains problematic as long as one models in the laboratory reference frame diffusion and transport by an external force field, without having recognized before hand the comoving frame in which the analysis must be set, with a drift term properly treated in relation with GT. This however is not always possible and can impact dramatically the energetic analysis of force measurements performed within currents, as found for instance in micro- and nanofluidics when studying colloidal transport phenomena from mechanosensitive and mechanoresponsive points of view.

4. Conclusion

Our setup has given us the possibility to induce a controllable, stationary and laminar convection flow inside a fluidic cell by locally heating the fluid (water) under laser illumination. This flow, dragging against the sedimentation a colloidal ensemble dispersed inside the cell, corresponds to a Galilean transform interconnecting the comoving fluid reference frame to the laboratory frame where the Brownian trajectories are recorded.

We have verified experimentally the principle of 'weak' GI for coarse-grained diffusive systems, and we have derived the expressions for the stochastic energetics production rates and the associated PDFs that yield a frame invariant formulation of the first law. We emphasized the crucial importance of recognizing in the drift term the signature of a Galilean transform in order not to interpret it as an external force field acting on the colloidal ensemble. We explicitly evaluated the energetic balance in this wrong-case scenario in order to illustrate its strong difference with respect to the frame invariant energetic balance.

This led us to conclude that misinterpreting the actual role of currents in Brownian experiments eventually leads to violate Galilean relativity that remains central for diffusive systems despite their coarse-grained structure. Our experimental scheme has therefore given us the opportunity to show how mistaking reference frame changes for external force fields leads to nonphysical conclusions. This obviously has practical implications in the context of force measurements in external flows, situations found for instance in soft matter physics and biophysics. Finally, we would like to highlight that our experiments can be performed in more complex fluids and thus appear appropriate for testing the principle of 'weak' GI for more sophisticated models of diffusion [1].

Acknowledgments

This work of the Interdisciplinary Thematic Institute QMat, as part of the ITI 2021 2028 program of the University of Strasbourg, CNRS and Inserm, was supported by IdEx Unistra (ANR 10 IDEX 0002), by SFRI STRAT'US project (ANR 20 SFRI 0012), by the Labex NIE (ANR-11-LABX-0058 NIE) and CSC (ANR-10-LABX-0026 CSC) projects, and by the University of Strasbourg Institute for Advanced Study (USIAS) (ANR-10-IDEX-0002-02) under the framework of the French Investments for the Future Program.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Appendix A.: Experimental setup and sample preparation

Our experiment consists in illuminating with horizontally two counterpropagating laser beams a colloidal dispersion of micron-sized melamine spheres, diffusing and sedimenting inside a cuvette filled with water. The same setup has already been described and exploited in a weak force measurement context in [14]. With balanced power in the counterpropagating beams, the laser induces convection within the fluid along the vertical z-direction with no radiation pressure effect at play. This perfect cancelation of radiation pressure makes the diffusion dynamics along the y-axis look like a free normal Brownian motion. Along the z-axis, the Brownian motion is performed within a laminar flow. This diffusion dynamics can be analysed by looking at real-time colloidal trajectories using the optical setup shown in figure 6 and recorded by tracking the successive positions of the particles using an algorithm adapted from [23].

Figure 6.

Figure 6. Schematic of the optical experimental setup already detailed in [14]. Using a halfwave plate (HWP), a polarizing beam-splitter and two mirrors (M1, M2), a linearly polarized single-mode laser beam (diode laser, wavelength 633 nm, 100 mW, TEM00) can be split into two noninterfering (cross-polarized) counterpropagating beams of identical intensity (using the HWP for the fine intensity balance). A microscope objective (NA = 0.25, 20×) collects the fluorescence of dye-doped melamine spheres (diameter d = 0.94 ± 0.05 μm, excitation/emission: 636 nm/686 nm) diffusing in water inside the cell with the help of a long-pass filter F with cutoff wavelength at 650 nm. The particles are imaged on a CCD camera at a frame rate f = 120 Hz. The cell consists of a quartz cuvette of dimensions 10(x) × 2(y) × 35(z) mm3 chosen such that the imaging ROI is located far away from any wall, allowing us to neglect safely any perturbation of the walls on the diffusion dynamics. Adapted figure with permission from [14], Copyright (2019r) by the American Physical Society.

Standard image High-resolution image

Our setup has important features. First, the cuvette has large dimensions compared to the size of the imaging ROI that only extends over a small central region far away from all walls. This, together with the small-volume fraction of the colloidal dispersion allows us to neglect the influence of possible boundary-wall and inter-particle-interaction effects on the diffusion dynamics. Also, this far-from-walls feature is a key point for the laminar flow generated by collective sedimentation and the laser induced convection detailed in appendix B. Second, the illumination conditions are set so that within the imaging ROI, the Gaussian profile of the laser beam is uniform along the horizontal optical y-axis considering that the Rayleigh length is set much larger than the width of the cuvette. Since the waist of laser is much larger than the diameter of a colloidal sphere, the close to plane-wave illumination conditions minimize any gradient contribution to the optical force field, yielding no external force along the z-axis except gravity.

The samples are prepared from an initial dispersion (2.5% mass–volume ratio) of melamine microparticles of diameter d = 0.94 ± 0.05 μm purchased from MicroParticle GmbH (FluoRed), weakly doped with a dye fluorescing at 686 nm for a most efficient detection in water. We dilute the dispersion ∼104 times with ultra-pure water and fill the cuvette with the colloidal dispersion to ca. 10−6 low-volume fraction. The filled cuvette is covered and sealed with a vacuum grease to prevent water evaporation and to isolate the fluid from other environmental influence. In addition, the cuvette and its cover are exposed at least 1 h to UV light to ensure the absence of any bacterial contaminant. The sample is also grounded to remove any electrostatic charge on the surface of the cuvette which generate a resulting external force on particles. Before performing our experiments, we leave for about 1 h the sample relaxing in its holder until well thermalised with the environment. This also ensures that potential colloidal aggregates have sedimented at the bottom of the cuvette. The temperature of the laboratory is controlled with a thermal precision better than 1 K at room temperature.

These illumination conditions, together with the great care of the sample preparation, allows us to exclude, as much as possible, all potential perturbing influences on the colloidal diffusion dynamics. This ensures us the capacity to quantitatively analyse the diffusion dynamics along the vertical z direction.

Appendix B.: Laser induced convection

The temperature and velocity profiles of the laser-induced convection flow within our experimental ROI—see figures 1 and 6—can be calculated analytically by exploiting the symmetries of our system and its illumination conditions that enable to decouple the two heat transfer and NS equations.

With the coordinate setting shown in figure 7, we first look at the heat transport equation (1) in the steady state (∂t T = 0):

Equation (24)

As discussed in the main text, the convection flow inside the ROI is essentially induced along the z-axis and measured to reach velocities of the order of 10−6 m s−1. Considering that a laser of ca. 100 mW will lead to an ca. mK increase in the local temperature on a typical scale of the order of the waist (in our case, ca. 100 μm), a simple scaling argument leads to a convective contribution of the Lagrangian derivative ( v ⋅ ∇)δT ∼ 10−5 much smaller than the contribution associated with the volumetric heat rate $\dot {q}/(\rho {c}_{p})\sim 1{0}^{-1}$ K s−1 at such laser powers. This scaling simplifies the heat equation to:

Equation (25)

with ${\dot {q}}_{L}=A{P}_{0}/\pi {a}^{2}\enspace \mathrm{exp}(-{r}^{2}/{a}^{2})$ and r2 = x2 + z2. In such polar coordinates, equation (25) can be solved through the following steps, starting with:

Equation (26)

first integrated into

Equation (27)

with the limit condition rr δT|r=0 = 0. By reminding that ln'(r) − E1'(r) = 1/r − exp(−r)/r with E1 the exponential integral function, we can further integrate from r to the temperature boundary layer δT/2 defined as the radial distance at which δT(δT/2) = 0, giving

Equation (28)

Figure 7.

Figure 7. Scheme of the coordinate system used for estimating the laser induced convection velocity profile. As the system under consideration has a symmetry along the optical axis, the problem is solved within the (x, z) plane. Therefore, the elementary volume is set as dx ⋅ 1 ⋅ dz. The waist of the laser projected in the (x, z) plane is ${w}_{0}=\sqrt{2}a$. The displayed temperature boundary layer thickness δT and the velocity boundary layer thickness δv correspond to the region beyond which the quantity (temperature or velocity) decays to that of the unperturbed bath.

Standard image High-resolution image

The temperature profile is then injected into the NS equation equation (2) within the Boussinesq's approximation and the condition of incompressibility in the vicinity of our ROI shown in figure 7 that yield the convection velocity field $\mathbf{v}(\mathbf{r})={v}_{0}(x,y)\hat{z}$ and cancel the convective contribution (v ⋅ ∇)v = 0 to the Lagrangian derivative. With $\mathbf{g}=-g\hat{z}$, the NS equation becomes in the steady-state:

Equation (29)

where ν = μ/ρ the kinematic viscosity. We will further exploit the consequence of incompressibility with ∂z v0 = 0 in order to solve equation (29) at z ∼ 0 where we can simplify the problem to a one-dimensional one with $\delta T(r=\sqrt{{x}^{2}+{z}^{2}})\sim \delta T(x)$. The translational invariance (within the ROI) of δT(x) along the y-axis allows us to apply a separation of variable

Equation (30)

that consists in decomposing convection into two drives: one thermal with the source term −gαδT(x) that only depends on x and a second associated with the collective body force ϕΔρg/ρ determining for the fluid the y-dependence of the convection velocity according to

Equation (31)

Equation (32)

Applying on wz a boundary condition of the first type

Equation (33)

where thus wz (y = 0) = vbath is extremal, and a boundary condition of the second type

Equation (34)

where 2ly corresponds to the length of the cuvette along y (ly = 1 mm). With these, the solution reads directly as:

Equation (35)

Considering that water has a Prandtl number Pr larger than one, the velocity boundary layer thickness is larger than the temperature boundary layer thickness according to ${\delta }_{\text{v}}=\sqrt{\mathrm{Pr}}{\delta }_{\text{T}}{ >}{\delta }_{\text{T}}$. Keeping in mind that δT(x) is non-zero only within δT, the uz (x) solution is piecewise, defined on the matching intervals:

Equation (36)

with a piecewise differential equation:

Equation (37)

with first type boundary conditions

Equation (38)

Equation (39)

Equation (40)

and second type boundary conditions

Equation (41)

Equation (42)

Equation (43)

Defining

Equation (44)

simplifies the equation to be solved into:

Equation (45)

The first integration gives

Equation (46)

with the error function erf(x) defined as

Equation (47)

Using the second type boundary condition of equation (41) determines C1 = 0. For u2 and u3, we have,

Equation (48)

Equation (49)

Using the additional second type boundary conditions of equations (42) and (43), fixes

Equation (50)

Equation (51)

Further integrating equation (46) gives:

Equation (52)

where

Equation (53)

and

Equation (54)

This yields the final solutions that read for u1 as:

Equation (55)

and for u2 and u3 as:

Equation (56)

Equation (57)

The constants are now determined by using the first type of boundary conditions:

Equation (58)

Equation (59)

These solutions determine the complete convection velocity v0(x, y) = uz (x) + wz (y) within the velocity boundary layer thickness δv. Accordingly, the velocity measured inside the ROI corresponds to the velocity evaluated at (x = 0, y = 0):

Equation (60)

Evaluating this velocity and its evolution with the laser power P0 demands to determine the boundary layer thicknesses δT and δv together with vbath. vbath is evaluated by solving the NS equation in the absence of any laser heating, under the sole influence of collective sedimentation. The boundary layer thicknesses can be determined through a simple scale analysis, where according to [24]:

Equation (61)

taking for the characteristic length L ∼ 2 × w0 and with Ra the Rayleigh number defined by

Equation (62)

For a medium with large Prandtl number Pr > 1 just like water is:

Equation (63)

According to equation (28), δTmax = δT(0). Using the known identity for the exponential integral function

Equation (64)

where γ is the Euler–Mascheroni constant, we expand E1(x) in the x → 0 limit as:

Equation (65)

Therefore, δTmax reads

Equation (66)

In the limit where $\frac{{\delta }_{\text{T}}^{2}}{4{a}^{2}}\gg 1$, ${E}_{1}\left(\frac{{\delta }_{\text{T}}^{2}}{4{a}^{2}}\right)\sim 0$, so that by substituting the expression of δT—cf equation (61)—into equation (66) yields an implicit equation for determining δTmax as:

Equation (67)

that is solved and which solution δTmax is used for determining δT according to equation (61) and δv through equation (63). Finally, the velocity at the centre of the ROI, i.e. at the centre of the cuvette is

Equation (68)

This is the velocity v0 engaged in the GT that connects the two lab S and comoving S' reference frames. The particle mean velocity noted vz in the main text, can be estimated by adding to the convection flow velocity v0 the single particle sedimentation velocity vsed according to vz = v0 + vsed. The profile of this convection velocity through the ROI is displayed in figure 2 in the main text, using values that correspond to our experimental conditions: α = 0.0003 K−1, A = 0.3 m−1, k = 0.61 W K−1 m−1, ν = 8.72 × 10−7 m2 s−1, cp = 4.18 × 103J kg−1 K−1, Δρ = 513 kg m−3, μ = 0.87 × 10−3 N s m−2. The boundary layer thicknesses δT and δv take the smallest value between the ones obtained by equations (61) and (63) and the minimum length found in the cuvette ly (=2 mm). The mean volume fraction ϕ can be estimated by comparing the experimental data in figure 4(d). The value of ϕ is ca. 6.8 × 10−7 is in good agreement with the expected value (ϕ ∼ 10−6) and the evolution of the observed velocity combining convection (v0) and single sedimentation velocity (vsed) with the illuminating laser power P0 is displayed in figure 8. Negative values at low laser powers correspond to the situation where the laser-induced convection flow is not strong enough to compensate for the collective sedimenting effect of the body force acting on the colloidal dispersion within the ROI.

Figure 8.

Figure 8. Comparison between the evolution of the particle velocity vz evaluated by equation (68) with the illuminating laser power P0—using ϕ = 6.8 × 10−7—(red line) and values of the mean particle velocity measured experimentally (blue dots).

Standard image High-resolution image

Appendix C.: Stochastic thermodynamics under Galilean transformation

The modification of stochastic thermodynamic quantities under GT can be also derived from Galilean relativity, giving the same results obtained in [2]. Following Sekimoto's approach [20], the heat transferred to the fluid by the Brownian system is described, in the fluid reference frame, by a force ${F}_{b}^{\prime }=-\gamma \dot {{z}^{\prime }}+\sqrt{2{k}_{\mathrm{B}}T\gamma }{\xi }^{\prime }({t}^{\prime })$ that the thermal bath (the fluid) exerts on the beads along the vertical axis. This force combines both the friction and the random force, and it exactly compensates the external force field acting on the system according to the Langevin equation. Most generally, this external force field writes as F'(z', t') = −∇U'(z', t') + f'(z', t') with f' a non-conservative force. In the fluid reference frame therefore, the heat production rate can be expressed as a function of the external force as [25] (same sign convention):

Equation (69)

In the laboratory frame, the stochastic force Fb written as ${F}_{b}=-\gamma \dot {z}+\sqrt{2{k}_{\mathrm{B}}T\gamma }\xi (t)$ will account for the difference in the noise statistics $\sqrt{2{k}_{\mathrm{B}}T\gamma }{\xi }^{\prime }(t)+\gamma {v}_{0}=\sqrt{2{k}_{\mathrm{B}}T\gamma }\xi (t)$ between both frames discussed above. Therefore, with the heat bath at equilibrium in the fluid rest frame, the force writes as

Equation (70)

showing that the relevant displacement to be accounted for when describing heat exchanges between the sphere and the fluid thermal bath is the actual displacement zv0 t of the Brownian sphere with respect to the fluid. Just like for Newton's laws, Galilean relativity enforces f(z, t) = f'(zv0 t, t) for the non-conservative force, U(z, t) = U'(zv0 t, t) for the potential energy, and therefore ∇U(z, t) = ∇U'(zv0 t, t). This implies Fb = −F(z, t) = −F'(zv0 t, t) and the final expression of the heat production rate evaluated in the laboratory frame

Equation (71)

The heat production rate is clearly GI with $\dot {q}=\dot {{q}^{\prime }}$ where $\dot {q}=\mathcal{G}[\dot {{q}^{\prime }}]$. Replacing the expressions of the potential energy and the non-conservative force in the comoving fluid frame by those in the laboratory frame yields $\dot {q}=[-\nabla U(z,t)+f(z,t)]\cdot (\dot {z}-{v}_{0})$. With the same approach, we can write the rate of potential energy change as $\dot {{u}^{\prime }}=\mathrm{d}{U}^{\prime }({z}^{\prime },{t}^{\prime })/\mathrm{d}{t}^{\prime }=[\dot {{z}^{\prime }}\cdot \nabla +{\partial }_{{t}^{\prime }}]{U}^{\prime }({z}^{\prime },{t}^{\prime })$ in the comoving frame. In the laboratory frame, $\dot {u}=[\dot {z}\cdot \nabla +{\partial }_{t}]U(z,t)$ gives:

Equation (72)

reminding that U(z, t) = U'(zv0 t, t). This shows that the rates are GI and connected from one reference frame to another by a GT performed on the position and velocity. That is $\dot {u}=\mathcal{G}[\dot {{u}^{\prime }}]=\dot {{u}^{\prime }}$. This leads us to a GI description of the stochastic work production rate and therefore of the first law of thermodynamics. Indeed, in the fluid reference frame, according to the first law, $\dot {{u}^{\prime }}=\dot {{w}^{\prime }}-\dot {{q}^{\prime }}$ so that $\dot {{w}^{\prime }}={\partial }_{{t}^{\prime }}{U}^{\prime }({z}^{\prime },{t}^{\prime })+{f}^{\prime }({z}^{\prime },{t}^{\prime })\cdot {z}^{\prime }$. In the laboratory inertial frame, $\dot {w}=\dot {u}+\dot {q}={\partial }_{t}{U}^{\prime }(z-{v}_{0}t,t)+{f}^{\prime }(z-{v}_{0}t,t)\cdot (\dot {z}-{v}_{0})$ shows that $\dot {w}=\mathcal{G}[\dot {{w}^{\prime }}]$. The work production rate eventually can be written as

Equation (73)

emphasizing the GI of the rate. Following [26], the stochastic entropy for single particle stot can be split into an entropy associated with the single particle trajectory (particle configuration) sp and an entropy sm associated with heat dissipated into the thermal bath. Both are initially defined as:

Equation (74)

Equation (75)

where P( x , t) is the PDF of position for the single particle motion, and q is the dissipated heat from the particle to the thermal bath, with T the temperature of this surrounding bath. This way, the total entropy production rate can be written as:

Equation (76)

The entropy production rates in the two different inertial reference frames S and S' are related by a GT x ' = x v 0 t and can be calculated for each entropy part. For the trajectory-dependent entropy production rate, the results demonstrated in the main text related to the 'weak' GI for the PDF in the different inertial reference frames—see equation (9)—lead to an entropy in the co-moving reference frame sp ' that can be written as:

Equation (77)

For the production rate of that trajectory-dependent entropy, we easily demonstrate the GI:

Equation (78)

As for the entropy production rate due to heat dissipation, considering that we have already demonstrated that the heat production is the same in the different inertial reference frame, we can easily obtain ${\dot {s}}_{m}={\dot {{s}^{\prime }}}_{m}$. Therefore, combining both rates, we directly arrive to the conclusion that the total stochastic entropy production is Galilean invariant.

Please wait… references are loading.