how a “pinch of salt” can tune chaotic mixing of colloidal ... · how a “pinch of salt” can...

5
How a “pinch of salt” can tune chaotic mixing of colloidal suspensions Julien Deseigne, a ecile Cottin-Bizonne, a Abraham D. Stroock, b Lyd´ eric Bocquet, a,c and Christophe Ybert a,,* Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX First published on the web Xth XXXXXXXXXX 200X DOI: 10.1039/b000000x Efficient mixing of colloids, particles or molecules is a cen- tral issue in many processes. It results from the complex interplay between flow deformations and molecular diffu- sion, which is generally assumed to control the homoge- nization processes. In this work we demonstrate on the contrary that despite fixed flow and self-diffusion condi- tions, the chaotic mixing of colloidal suspensions can be either boosted or inhibited by the sole addition of trace amount of salt as a co-mixing species. Indeed, this shows that local saline gradients can trigger a chemically-driven transport phenomenon, diffusiophoresis, which controls the rate and direction of molecular transport far more ef- ficiently than usual Brownian diffusion. A simple model combining the elementary ingredients of chaotic mixing with diffusiophoretic transport of the colloids allows to ra- tionalize our observations and highlights how small-scale out-of-equilibrium transport bridges to mixing at much larger scales in a very effective way. Considering chaotic mixing as a prototypal building block for turbulent mix- ing, this suggests that these phenomena, occurring when- ever the chemical environment is inhomogeneous, might bring interesting perspective from micro-systems up to large-scale situations, with examples ranging from ecosys- tems to industrial contexts. Transport and mixing of molecules or particles play a central role in many processes, from large scales phenomena such as industrial chemical reactors, dispersion of pollutants or depo- sition of sediments involved in biochemical cycles 1,2 , down to the rapidly developing Lab-on-a-Chip micro-systems. In this field, strong limitation to mixing involved by small-scales Electronic Supplementary Information (ESI) available: Additional details on experimental and high resolution cross-section images of the mixing pro- cess for different salt configurations. See DOI: 10.1039/b000000x/ a Institut Lumi` ere Mati` ere, Universit´ e Claude Bernard Lyon 1-CNRS, UMR 5306, Universit´ e de Lyon, F-69622 Villeurbanne, France. b School of Chemical and Biomolecular Engineering, Cornell University, Ithaca, New York 14853, USA. c Present address: Department of Civil and Environmental Engineering, Mas- sachusetts Institute of Technology, Cambridge, MA, USA E-mail for correspondence: [email protected] viscous flows have triggered research for specific strategies 3,4 to optimize and facilitate (bio-)chemical analysis 5 or syn- thesis 6 operations in microsystems. The generic route for ef- ficient mixing proceeds through the stretch and fold mecha- nism that generates ever-thinner structures of the substances to mix. This is achieved either from turbulent flow properties 7,8 in large-scales situations, or through laminar chaotic flows in viscous regimes 9,10 and, accordingly, mixing has been mostly approached from the “large” hydrodynamic scales perspec- tive, focusing on flow characteristics. In all such processes however, mixing or homogenization involves a continuous in- terplay of global deformation within the flow, and local molec- ular transport –generically assumed to occur through Brown- ian diffusion 8,9,11,12 . In this communication, we start from this molecular-scale perspective and ask if we can trigger a different molecu- lar transport phenomenon, in place of the canonical diffu- sion, and impact in this way the global mixing properties. To explore this bottom-up approach to mixing, we consider an overlooked phenomenon –colloidal diffusiophoresis– re- cently studied in the context of surface-driven flow and non- equilibrium transport 13,14,15,16,17 . Local gradients of small chemicals –such as a salt– induce a migration of particles with transport properties that can be orders of magnitude higher than bare particles diffusion 18,19 . We demonstrate here how this non-equilibrium phenomenon can be harnessed to take control of the molecular-scale transport involved in mixing and thus modify the overall mixing properties of suspended particles. Indeed we show for the first time that creating an inhomogeneous chemical environment by adding only traces of salt deeply alters the mixing dynamics of colloidal suspen- sion, which can be either boosted or inhibited. Moreover, we present a simple model which allows to bridge between the new non-equilibrium nano-scale transport and the large-scale mixing properties. Experimentally, the mixing of a suspension made of fluo- rescent colloids (200 nm diameter polystyrene, 0.02% w/v) with an aqueous buffer solution (1mM Tris, pH=9) is studied under laminar chaotic flows. This is done thanks to a Y-shaped 1–5 | 1 arXiv:1403.6390v1 [physics.flu-dyn] 25 Mar 2014

Upload: others

Post on 12-Sep-2019

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: How a “pinch of salt” can tune chaotic mixing of colloidal ... · How a “pinch of salt” can tune chaotic mixing of colloidal suspensions† Julien Deseigne,a Cecile Cottin-Bizonne,´

How a “pinch of salt” can tune chaotic mixing of colloidal suspensions†

Julien Deseigne,a Cecile Cottin-Bizonne,a Abraham D. Stroock,b Lyderic Bocquet,a,c and ChristopheYbert a,‡,∗

Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XXFirst published on the web Xth XXXXXXXXXX 200XDOI: 10.1039/b000000x

Efficient mixing of colloids, particles or molecules is a cen-tral issue in many processes. It results from the complexinterplay between flow deformations and molecular diffu-sion, which is generally assumed to control the homoge-nization processes. In this work we demonstrate on thecontrary that despite fixed flow and self-diffusion condi-tions, the chaotic mixing of colloidal suspensions can beeither boosted or inhibited by the sole addition of traceamount of salt as a co-mixing species. Indeed, this showsthat local saline gradients can trigger a chemically-driventransport phenomenon, diffusiophoresis, which controlsthe rate and direction of molecular transport far more ef-ficiently than usual Brownian diffusion. A simple modelcombining the elementary ingredients of chaotic mixingwith diffusiophoretic transport of the colloids allows to ra-tionalize our observations and highlights how small-scaleout-of-equilibrium transport bridges to mixing at muchlarger scales in a very effective way. Considering chaoticmixing as a prototypal building block for turbulent mix-ing, this suggests that these phenomena, occurring when-ever the chemical environment is inhomogeneous, mightbring interesting perspective from micro-systems up tolarge-scale situations, with examples ranging from ecosys-tems to industrial contexts.

Transport and mixing of molecules or particles play a centralrole in many processes, from large scales phenomena such asindustrial chemical reactors, dispersion of pollutants or depo-sition of sediments involved in biochemical cycles1,2, downto the rapidly developing Lab-on-a-Chip micro-systems. Inthis field, strong limitation to mixing involved by small-scales

† Electronic Supplementary Information (ESI) available: Additional detailson experimental and high resolution cross-section images of the mixing pro-cess for different salt configurations. See DOI: 10.1039/b000000x/a Institut Lumiere Matiere, Universite Claude Bernard Lyon 1-CNRS, UMR5306, Universite de Lyon, F-69622 Villeurbanne, France.b School of Chemical and Biomolecular Engineering, Cornell University,Ithaca, New York 14853, USA.c Present address: Department of Civil and Environmental Engineering, Mas-sachusetts Institute of Technology, Cambridge, MA, USA‡ E-mail for correspondence: [email protected]

viscous flows have triggered research for specific strategies3,4 to optimize and facilitate (bio-)chemical analysis5 or syn-thesis6 operations in microsystems. The generic route for ef-ficient mixing proceeds through the stretch and fold mecha-nism that generates ever-thinner structures of the substances tomix. This is achieved either from turbulent flow properties7,8

in large-scales situations, or through laminar chaotic flows inviscous regimes9,10 and, accordingly, mixing has been mostlyapproached from the “large” hydrodynamic scales perspec-tive, focusing on flow characteristics. In all such processeshowever, mixing or homogenization involves a continuous in-terplay of global deformation within the flow, and local molec-ular transport –generically assumed to occur through Brown-ian diffusion8,9,11,12.

In this communication, we start from this molecular-scaleperspective and ask if we can trigger a different molecu-lar transport phenomenon, in place of the canonical diffu-sion, and impact in this way the global mixing properties.To explore this bottom-up approach to mixing, we consideran overlooked phenomenon –colloidal diffusiophoresis– re-cently studied in the context of surface-driven flow and non-equilibrium transport13,14,15,16,17. Local gradients of smallchemicals –such as a salt– induce a migration of particles withtransport properties that can be orders of magnitude higherthan bare particles diffusion18,19. We demonstrate here howthis non-equilibrium phenomenon can be harnessed to takecontrol of the molecular-scale transport involved in mixingand thus modify the overall mixing properties of suspendedparticles. Indeed we show for the first time that creating aninhomogeneous chemical environment by adding only tracesof salt deeply alters the mixing dynamics of colloidal suspen-sion, which can be either boosted or inhibited. Moreover, wepresent a simple model which allows to bridge between thenew non-equilibrium nano-scale transport and the large-scalemixing properties.

Experimentally, the mixing of a suspension made of fluo-rescent colloids (200 nm diameter polystyrene, 0.02% w/v)with an aqueous buffer solution (1mM Tris, pH=9) is studiedunder laminar chaotic flows. This is done thanks to a Y-shaped

1–5 | 1

arX

iv:1

403.

6390

v1 [

phys

ics.

flu-

dyn]

25

Mar

201

4

Page 2: How a “pinch of salt” can tune chaotic mixing of colloidal ... · How a “pinch of salt” can tune chaotic mixing of colloidal suspensions† Julien Deseigne,a Cecile Cottin-Bizonne,´

A.

B.

Fig. 1 Salt effect on chaotic mixing of suspended particles. A.Sketch of chaotic mixing experimental setup: a suspension offluorescent colloids (orange) and a raw buffer solution (black) areinjected at constant flow rate into separate feeding channels (notshown) and merge into the main channel of a StaggeredHerringbone Micromixer9 (w = 200µm, h = 115µm, λ = 2mm,α1 = 0.35) where a chaotic mixing of the two solutions takes place.B. Evolution of the suspension mixing process in the presence ofsalt: cross-section images of the central portion of the mixingchannel at different locations l after solutions’ merging (white scalebar: 50µm; mean flow velocity U = 8.6mm/s). Central row:reference mixing process without salt (saltless configuration); Upperand lower rows: effects of additional salt solute (20mM LiCl) eitherinto colloidal suspension (salt-in configuration) or into theco-flowing buffer solution (salt-out configuration).

microfluidic device (Fig. 1A) where the two solutions to mixmerge into the main channel of a Staggered Herringbone Mi-cromixer (SHM)9,20,21,22. Cross-section images are capturedwith confocal microscope at different channel locations l fromthe inlet (See Supp. Mat. for further details †). This allows tofollow the evolution of the mixing process with elapsed timet = l/U or with the number of stretch and fold cycles l/λ, withU the mean downstream flow velocity and λ the length of oneSHM cycle. This SHM was shown to generate a chaotic mix-ing process9, see Fig. 1B middle.

Now, keeping the global hydrodynamics unchanged, we ob-serve Fig. 1B that adding a small amount of a passive molec-ular solute (20mM LiCl salt) to one of the two solutions hasa very noticeable impact on the mixing process. Indeed whensalt is added to the colloidal suspension (salt-in configuration,Fig 1B top), initial stages of mixing show an increased con-centration of the colloidal suspension associated with thinnerand brighter fluorescent filaments with sharper edges as com-pared to the saltless reference mixing. A “pinch of salt” in theparticles’ phase thus suppresses the homogenization withinthe mixing process, usually provided by the coupling with lo-cal diffusion.

On the contrary, when salt is added to the buffer solution

0 5 10 15 200

0.2

0.4

0.6

0.8

1

Fig. 2 Normalized sttandard deviation σ =√〈c2〉−〈c〉2/〈c〉 of the

colloids concentration c in the channel cross-section as a function ofthe location l/λ for the three mixing configurations (mean flowvelocity U = 8.6mm/s): salt-in (red), saltless (black), salt-out(blue). Three corresponding cross-sections are shown for illustration(saltless case). Mixing length lm (see text) is defined at a standarddeviation threshold set to σc = 0.3.

(salt-out configuration, Fig. 1B bottom), mixing is again mod-ified but with opposite consequences. Initial stages of mix-ing now show thicker and dimmer fluorescent filaments withblurred edges (see e.g. l/λ = 4 in Fig. 1B). There a trace con-centration of salt has a clear enhancing mixing effect. Impor-tantly, we stress that with salt contrasts from∼ 1mM in buffersolvent to 20mM in salty solution, no density or viscosity mis-matches nor any floculation or interparticle interaction can beat stake here.

These observations are confirmed by a more quantitativeanalysis, using the normalized standard deviation of the col-loids concentration field σ =

√〈c2〉−〈c〉2/〈c〉 to quantify

mixing (where the concentration in colloids c was checkedto be proportional to the fluorescent intensity). As shown inFig. 2, this analysis confirms the previous picture: the pres-ence of salt in the outer solution results in a faster decrease ofthe concentration inhomogeneities as compared to the saltlesscase, while, on the opposite, salt added to the colloidal sus-pension delays homogenization. In the following, we define amixing length lm using a threshold value as σ(lm) = σc = 0.3.

With all flow parameters unchanged, our experimentsclearly points to the importance of molecular effects in mixingprocesses. Such effects are known to occur in the homogeniza-tion step, where it is classically assumed that coupling withmolecular diffusion allows for local cross-filament transport.Signature of molecular diffusivity has been reported so far inchaotic9,20 or moderately turbulent8,11 mixing flows. Qual-itatively, mixing is achieved when the diffusional spreadinglength LD compares with the typical width of the stretchedfilaments, say LS. One readily expects LD ∼

√Dlm/U , with

D the molecular diffusivity of the mixed species, while in a

2 | 1–5

Page 3: How a “pinch of salt” can tune chaotic mixing of colloidal ... · How a “pinch of salt” can tune chaotic mixing of colloidal suspensions† Julien Deseigne,a Cecile Cottin-Bizonne,´

A.

B.

100

101

102

103

4

6

8

10

12

Fig. 3 A Reduced mixing length lm/λ as a function of moleculardiffusivity D (mean flow velocity U = 8.6mm/s): (2) Rhodamine Bdye, (3) Rhodamine-PEG5000, (◦) 200 nm diameter colloids(saltless, black) and with 20 mM LiCl salt with colloidal suspension(salt-in, red) or with coflowing solution (salt-out, blue). BDiffusiophoretic transport in chaotic mixing: sketch of theunderlying salt gradients at the edges of mixing filaments ofcolloidal suspension. For salt-in and salt-out configurations, totalcross-filament flux of colloids (red beads) reads ~jc = ~jD +~jDP, withbare diffusive flux ~jD dominated by orders of magnitude16,18 bydiffusiophoretic transport ~jDP. One thus expects boosted-mixing(salt-out, left) or anti-mixing (salt-in, right).

fully chaotic flow LS ∼ (w/2)exp(−lm/δ), with w the chan-nel width and δ the stretching length9. This predicts a weaksignature of molecular effects on the mixing length lm in theform lm ∼ log(Uw/D).

As shown in Fig. 3, this crude argument captures the maineffect of molecular diffusivity on mixing in saltless solutions.In this graph, the mixing length lm is plotted – for a given flowvelocity U – versus the molecular diffusivity of the suspendedparticles, from molecules to colloids (see Supp. Mat. for fur-ther details †), with diffusivities spanning more than 2 ordersof magnitude. The measured evolution is compatible with thelog dependency expected for chaotic flows, as reported previ-ously for this SHM geometry whereby the effect of changingflow velocity U was probed9.

Now, as is evident from Fig. 3, the salt effects as a co-mixing solute do not fit into this Brownian diffusion paradigmfor molecular transport. With colloids diffusivity rigorouslyunchanged, salt-in and salt-out configurations delay or boostmixing as if the molecular transport was changed by about

3 orders of magnitudes (as would be measured in “diffusiv-ity scale” fig. 3A). As we show now, this is a striking mani-festation of the onset and influence of new out-of-equilibriumlocal transport (here diffusiophoresis) overtaking classical dif-fusion.

Diffusiophoretic transport refers to the migration of parti-cles (colloids or macromolecules) induced by gradients of so-lute. This subtle phenomenon of osmotic origins was studiedin pioneering works by Anderson, Prieve, and coworkers13,23

and recently received in-depth characterization of its effects onthe migration, trapping or patterning of particles16,17,18,19,24.For salt as a solute with concentration field Cs, saline gradi-ents induce a diffusiophoretic drift velocity of a particle as

VDP = DDP∇ logCs, (1)

where the diffusiophoretic (DP) mobility DDP has the dimen-sion of a molecular diffusivity23. While the theoretical ex-pression of DDP includes particle’s surface charge and salt-type corrections18,23, it is typically much larger than the barecolloid diffusion coefficient (DDP/D� 1), and close to smallmolecules fast diffusivities (here DDP ' 290µm2/s16).

Coming back to the colloidal suspension mixing problem,it is now possible to propose a rationalization of the observedsalt effects on the basis of this diffusiophoretic mechanism.As sketched in Fig. 3.B, the salt concentration fields Cs ex-hibits strong gradients localized at the edges of the stretchedfilaments, thereby triggering a supplementary DP migration ofthe nearby colloids toward high salt concentrations, see Eq. 1.This will induce an accelerated spreading of the colloids pro-file for salt-out and conversely inhibit mixing for salt-in, infull agreement with the observed behavior in Fig. 3.A. Thiseffect is measured systematically for colloids, for all flow ve-locities or SHM geometries probed as shown in the inset ofFig. 4 where the salt-in and salt-out mixing lengths, normal-ized by the saltless reference case l0

m, are plotted against thePeclet number. Finally we need to emphasize that this effectshows up here although the salt concentration remains quitelow (20mM). This is a key consequence of the log depen-dency in Eq. 1 allowing response even for traces of solutesas recently demonstrated17.

Going beyond this analysis towards a more quantitative de-scription of this effect constitutes an important challenge. Itinvolves coupling two intrinsically complex mechanisms: thedescription of mixing and its complex interplay between flowdeformation and local transport; together with the descriptionof molecular-scale non-equilibrium phenomenon controllinghere local colloids dynamics. A first step in this direction canbe performed by considering the simplified coupling frame-work proposed by Ranz for the chaotic laminar mixing process25. It consists in considering the reference frame of a stretchedfilament, whose axis are aligned with maximal compression

1–5 | 3

Page 4: How a “pinch of salt” can tune chaotic mixing of colloidal ... · How a “pinch of salt” can tune chaotic mixing of colloidal suspensions† Julien Deseigne,a Cecile Cottin-Bizonne,´

and stretching directions, and to reduce the homogenization toa 1D cross-filament dynamics.

For solute species where transport is only ruled by diffusionand flow-advection (i.e. single solute or saltless colloids), theRanz model writes:

∂tc− γx∂xc = D∂2xc, (2)

with x the direction perpendicular to the filament interface andγ ∝ U the principal strain rate. Transverse filament thicknessevolves as s(t)∼ s0 exp(−γt), with s0 = w/2 the initial width.This equation is solved by a change of variables ξ= x/s(t) andτ = D

∫ t0 dt ′/s2(t ′), yielding a simple diffusion equation from

which concentration profiles are obtained25,26. This yields amixing time for purely advection-diffusion dynamics in theform:

t0m =

12γ

log(Pe/2) , (3)

where the Peclet number is defined in the model as Pe =γs2

0/D ∝ Uw/2D. This predicts a logarithmic dependence ofthe mixing length l0

m =U t0m in terms of the Peclet number. De-

spite the simplicity of Ranz model, it is well consistent withour experimental results for single species mixing (Fig. 3.A),in agreement with literature9,20,26.

It is now possible to extend this framework to include thesupplementary colloid DP migration. While the salt concen-tration Cs is still ruled by Eq.(2), the cross-filament dynam-ics of the colloids concentration c in salt-in and salt-out nowwrites:

∂tc− γx∂xc+∂x [VDP c] = Dc∂2xc, (4)

with VDP given by (1), and Dc the colloids diffusivity. In thereduced variables (ξ,τs) associated to salt dynamics, Eq.(4)transforms to

∂τsc+∂ξ

[(sVDP

Ds

)c]' 0, (5)

where a diffusive term of order Dc/Ds � 1 was neglected.Information on the filament interface position ξ = Ξ(τs)can be obtained using the method of characteristics, lead-ing to d Ξ/dτs = s(τs)VDP(Ξ(τs),τs)/Ds. For short times andsmall interface displacements τs,ξ� 1, the salt concentrationCs(ξ,τs) reduces to c0(1+ξ/

√πτs)/2 and one may obtain an

expression for the interface displacement ∆Ξ = Ξ(τs)−Ξ(0)as:

∆Ξ'±2DDP

Ds

√τs

π, (6)

valid for short times; the sign ± corresponds to the salt-outand salt-in cases respectively.

Let us now focus on the salt-out case. One defines the–reduced– mixing time τout

s as the time when the interfacereaches the next colloidal suspension filament (ξ = ±1/2).

This is obtained as τouts = Ds/2Deff, where we introduced an

effective diffusion coefficient defined as Deff = 8D2DP/(πDs),

in line with Abecassis et al.16,18,19. One deduces accordingly,

loutm ≡U× tout

m =U2γ

log(Peeff/2) , (7)

where we defined an effective Peclet number in the formPeeff = 2γs2

0/Deff = Pe× (2Dc/Deff).Comparing with Eq. (3), this suggests that in the salt-out

configuration, the effect of diffusiophoresis can be mappedonto the classical mixing framework, provided the bare molec-ular diffusivity Dc is replaced by an effective diffusivity Deff/2(Deff = 157µm2/s). Accordingly the relevant Peclet numberis defined in terms of the effective diffusion coefficient Deff.Since Deff � Dc, we now expect large particles to mix asfast as small molecules in co-mixing configuration. Whilethe mixing length is only a weak (logarithmic) function of thePeclet number, the huge change in effective diffusion leads aquantitative change in the mixing efficiency. This predictionis tested in Fig. 4 where the experimental mixing lengths lm/λ

obtained for various velocities and salts are plotted againstthe effective Peclet number Peeff defined above. Although wedo not expect this generalized Ranz model to capture the fullcomplexity of the mixing process, it indeed provides an in-sightful framework that well captures the importance and ba-sic mechanisms of this molecular co-mixing phenomenon.

Finally, in contrast to the salt-out configuration, the analy-sis for the salt-in case requires accounting for the –up to nowneglected– diffusive transport, which becomes the dominantmechanism for mixing at long time. While the above frame-work indeed predicts a focusing of the colloid profiles at earlytimes –thus with no possible diffusivity mapping–, a full quan-titative analysis is far more complex and we leave it for futureinvestigations.

In conclusion, we demonstrate that traces of salt can con-siderably impact the chaotic mixing of suspensions: here syn-thetic colloids, but it generalizes to molecules14,16, silica18 orclay particles, etc.. This arises from a shift in molecular trans-port, with canonical diffusion overtaken by diffusiophoresis.Unlike physically-applied fields, the chemical driving forceunder diffusiophoresis evolves under flow as a mirror of thecolloids distribution, thus providing especially relevant andefficient transport unusually bridging across widely separatedscales.

Beyond the here-demonstrated interest at microfluidicscale, we stress that diffusiophoresis should show up when-ever colloids, particles or molecules evolve in an inhomo-geneous chemical background. Combined with the fact thatchaotic mixing can be viewed as a prototypal building blockfor turbulent mixing, this suggests that tuning the local trans-port properties may yield possible implications at upper-scales27, in a broad range of situations from bio-reactors to estuar-

4 | 1–5

Page 5: How a “pinch of salt” can tune chaotic mixing of colloidal ... · How a “pinch of salt” can tune chaotic mixing of colloidal suspensions† Julien Deseigne,a Cecile Cottin-Bizonne,´

105

106

0.5

1

1.5

103

104

105

106

2

4

6

8

10

12

Peeff

Pe

Fig. 4 Reduced mixing length lm/λ as a function of –effective–Peclet number Peeff for different velocities, molecular species andsalt content (SHM geometrical parameters as in fig. 1). No salt:rhodamine B (2), rhodamine PEG5000 (3), saltless 200 nmcolloids (◦) ; 20 mM LiCl salt contrast: salt-out 200 nm colloids(•); Peeff calculated according to eq. (7). The black dashed linerepresents a log evolution of Peeff Inset. Salt effects on mixinglength for colloids lm/l0

m as a function of Pe number, for differentSHM geometrical characteristics. Salt-less mixing length l0

m is usedas a normalization for saltless (red) and salt-out (blue) data. SHMspecifications as in fig. 1 except for α: α1 = 0.35 (◦), α2 = 0.40(4) and α3 = 0.36 (5).

ies environment where fresh water full of sediments meets thesalty seawater, see e.g.28. Indeed, in the context of marine bi-ology, the local non-equilibrium dynamics of bio-organisms,revealed at the micro-scale, has been hinted as a crucial factorfor ocean-scale distribution29,30,31,32,33.

Acknowledgements

We thank F. Raynal, M. Bourgoin and R. Volk for helpfuldiscussions, P. Sundararajan for providing initial SHM mas-ter and design, G. Simon and R. Fulcrand for technical sup-port. Microfabrication was performed using the NanoLyonfacilities (INL). We acknowledge financial support from ANRSYSCOM and from LABEX iMUST (ANR-10-LABX-0064/ANR-11-IDEX-0007) under project MAXIMIX.

References1 M. Dai, J.-M. Martin and G. Cauwet, Marine Chemistry, 1995, 51, 159–

175.2 C. Lee and S. G. Wakeham, Marine Chemistry, 1992, 39, 95–118.3 M. R. Bringer, C. J. Gerdts, H. Song, J. D. Tice and R. F. Ismagilov, Philo-

sophical Transactions of the Royal Society A: Mathematical, Physical andEngineering Sciences, 2004, 362, 1087–1104.

4 F. Schonfeld, V. Hessel and C. Hofmann, Lab on a chip, 2004, 4, 65–69.5 M. A. Burns, C. H. Mastrangelo, T. S. Sammarco, F. P. Man, J. R. Web-

ster, B. N. Johnsons, B. Foerster, D. Jones, Y. Fields and A. R. Kaiser,Proceedings of the National Academy of Sciences, 1996, 93, 5556–5561.

6 P. Watts and S. J. Haswell, Current Opinion in Chemical Biology, 2003,7, 380–387.

7 J. M. Ottino, Annual Review Of Fluid Mechanics, 1990, 22, 207–254.8 E. Villermaux and J. Duplat, Physical Review Letters, 2006, 97, 144506.9 A. D. Stroock, S. Dertinger, A. Ajdari, I. Mezic, H. A. Stone and G. M.

Whitesides, Science (New York, NY), 2002, 295, 647.10 F. Raynal, A. Beuf, F. Plaza, J. Scott, P. Carriere, M. Cabrera, J.-P. Cloarec

and E. Souteyrand, Physics Of Fluids, 2007, 19, 017112.11 E. Villermaux and H. Rehab, Journal Of Fluid Mechanics, 2000.12 P. Sundararajan and A. D. Stroock, Annual Review of Chemical and

Biomolecular Engineering, 2012, 3, 473–496.13 D. C. Prieve and R. Roman, Journal of the Chemical Society, Faraday

Transactions 2, 1987, 83, 1287.14 M. S. Munson, C. R. Cabrera and P. Yager, Electrophoresis, 2002, 23,

2642–2652.15 A. Ajdari and L. Bocquet, Physical Review Letters, 2006, 96, 186102.16 J. Palacci, B. Abecassis, C. Cottin-Bizonne, C. Ybert and L. Bocquet,

Physical Review Letters, 2010, 104, –.17 J. Palacci, C. Cottin-Bizonne, C. Ybert and L. Bocquet, Soft Matter, 2012,

8, 980–994.18 B. Abecassis, C. Cottin-Bizonne, C. Ybert, A. Ajdari and L. Bocquet,

Nature Materials, 2008, 7, 785–789.19 B. Abecassis, C. Cottin-Bizonne, C. Ybert, A. Ajdari and L. Bocquet,

New Journal of Physics, 2009, 11, –.20 A. D. Stroock and G. J. McGraw, Philosophical Transactions of the Royal

Society A: Mathematical, Physical and Engineering Sciences, 2004, 362,971–986.

21 J. Aubin, D. F. Fletcher, J. Bertrand and C. Xuereb, Chemical Engineering& Technology, 2003, 26, 1262–1270.

22 J.-T. Yang, K.-J. Huang and Y.-C. Lin, Lab on a chip, 2005, 5, 1140.23 J. Anderson, Annual Review Of Fluid Mechanics, 1989, 21, 61–99.24 H.-R. Jiang, H. Wada, N. Yoshinaga and M. Sano, Physical Review Let-

ters, 2009, 102, 208301.25 W. Ranz, AIChE Journal, 1979, 25, 41–47.26 E. Villermaux, A. D. Stroock and H. A. Stone, Physical Review E, 2008,

77, 015301.27 R. Volk and F. Raynal, Personnal Communication.28 A. Thill, S. Moustier, J.-M. Garnier, C. Estournel, J.-J. Naudin and J.-Y.

Bottero, Continental Shelf Research, 2001, 21, 2127–2140.29 H. W. Paerl and J. L. Pinckney, Microbial Ecology, 1996, 31, 225–247.30 A. Malits, F. Peters, M. Bayer Giraldi, C. Marrase, A. Zoppini, O. Gua-

dayol and M. Alcaraz, Microbial Ecology, 2004, 48, 287–299.31 R. Stocker, Science (New York, NY), 2012, 338, 628–633.32 J. R. Taylor and R. Stocker, Science (New York, NY), 2012, 338, 675–679.33 W. M. Durham, E. Climent, M. Barry, F. De Lillo, G. Boffetta, M. Cencini

and R. Stocker, Nature Communications, 2013, 4, 1–7.

1–5 | 5