Вы находитесь на странице: 1из 6

Thermo-Osmotic Flow in Thin Films

Andreas P. Bregulla, Alois Würger, Katrin Günther, Michael Mertig, Frank


Cichos

To cite this version:


Andreas P. Bregulla, Alois Würger, Katrin Günther, Michael Mertig, Frank Cichos. Thermo-
Osmotic Flow in Thin Films. Physical Review Letters, American Physical Society, 2016, 116
(18), pp.188303. <10.1103/PhysRevLett.116.188303>. <hal-01312488>

HAL Id: hal-01312488


https://hal.archives-ouvertes.fr/hal-01312488
Submitted on 6 May 2016

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
Thermo-osmotic flow in thin films

Andreas P. Bregulla1 , Alois Würger2 , Katrin Günther3 , Michael Mertig3,4 , Frank Cichos1∗
1
Molecular Nanophotonics Group, Institute of Experimental Physics I, University of Leipzig, 04103 Leipzig, Germany
2
Laboratoire Ondes et Matière d’Aquitaine, Université de Bordeaux & CNRS, 33405 Talence, France
3
BioNanotechnology and Structure Formation Group, Department of Chemistry and Food Chemistry,
Chair of Physical Chemistry, Measurement and Sensor Technology,
Technische Universität Dresden, 01062 Dresden, Germany
4
Kurt-Schwabe-Institut für Mess- und Sensortechnik e.V. Meinsberg, 04736 Waldheim, Germany

We report on the first micro-scale observation of the velocity field imposed by a non-uniform heat
content along the solid/liquid boundary. We determine both radial and vertical velocity compo-
nents of this thermo-osmotic flow field by tracking single tracer nanoparticles. The measured flow
profiles are compared to an approximate analytical theory and to numerical calculations. From the
measured slip velocity we deduce the thermo-osmotic coefficient for both bare glass and Pluronic
F-127 covered surfaces. The value for Pluronic F-127 agrees well with Soret data for polyethylene
glycol, whereas that for glass differs from literature values and indicates the complex boundary layer
thermodynamics of glass-water interfaces.

charged surface Pluronic coated surface


Osmosis is the passage of a liquid through a semiper- a) b) vs
vs
meable membrane, towards a higher concentration of a
λ
molecular solute or salt. Osmotic processes are funda-
+
λ + + + + ++ + + ++ +
mental for life, for example in selective transport through --- ------ ---- -
cell membranes, and for applications such as desalination
of seawater and power generation from the salinity dif-
ference with river water [1]. In physical terms, osmosis c) 5 300

tracer velocity [µm/s]


250
is driven by the gain in mixing entropy and requires to 4
200

impede solute diffusion. 3 150


z [µm]

100

2 50
Thermo-osmosis relies on the same principle, albeit 0

1
with a liquid of non-uniform heat content instead of a
0
non-uniform solute concentration; accordingly, the un- 0 1 2 3 4 5
r [µm]
derlying thermodynamic force is the temperature gradi-
ent rather than a concentration gradient. Since there Figure 1. Schematic view of the boundary layer at an interface
are no heat-selective membranes, thermo-osmosis occurs with non-uniform temperature. a) Thermo-osmotic velocity
in open geometries only, where heat and liquid flow in profile close to a charged solid boundary with an excess en-
opposite directions along a solid boundary, similarly to thalpy within an interaction length λ; for z > λ the velocity
electro-osmosis in capillaries or nanofluidic diodes [2]. saturates at vs . b) The same for a Pluronic-coated surface.
Water flow due to a temperature gradient was first ob- c) Streamlines in the liquid film resulting from slip velocities
at the upper and lower boundaries, reconstructed by numeri-
served by Derjaguin and Sidorenkov through porous glass
cal calculations and the experimental results for the Pluronic
[3]. F-127 coated surface.
In the last decade thermal gradients have become a ver-
satile means of manipulating colloidal dispersions, e.g.,
self-propulsion of metal-capped Janus particles [4, 5],
cluster formation through hydrodynamic interactions
Here we report on the first micro-scale observation of
[7, 8], force-free steering through dynamical feedback
the velocity field imposed by thermo-osmosis along the
[6, 9], sieving by size and stretching macromolecules [10–
solid boundary. Both radial and vertical velocity compo-
12], dynamical trapping of nano-particles [13, 14], and
nents are determined by tracking single tracer nanopar-
detection of DNA through functionalized gold nanopar-
ticles. The measured flow profiles are compared to an
ticles (AuNP) [15]. In all these examples, the mo-
analytical theory and to numerical calculations.
tion arises from a superposition of thermo-osmosis and
molecular osmosis [16]: The temperature gradient in- Thermo-osmotic slip velocity. A bulk liquid in a tem-
duces heat flow along the colloidal surface, whereas its perature gradient reaches a non-equilibrium stationary
companion fields, e.g. composition [5, 17] and ion con- state, with a steady non-uniform composition but zero
centration [18–21], drive molecular currents. Finally, matter flow. A solid boundary, however, exerts addi-
thermo-osmotic flow could be relevant for particle motion tional forces on the liquid; the corresponding excess spe-
through hot nanostructures, in addition to thermophore- cific enthalpy h in the boundary layer results in a creep
sis and thermoconvection[22, 23]. flow parallel to the surface, with the effective slip velocity
2

[24] a)
glass
=250 nm
Pluronic F127

∇T ∇T
Z
1
vs = − dzzh(z) ≡χ , (1) water Au tracer
=150 nm
5 µm
η 0 T T
Pluronic F127
where η is the viscosity. An enthalpy excess, h > 0, glass

leads to a negative χ and liquid flow towards the cold b) 5 c)

16
6

2
side, as observed for glass capillaries, clays, and silica 4

temperature increment [K]


4 2
gels [3, 24], whereas a negative enthalpy (h < 0) drives 10
1/r

3 6
the liquid towards the hot, e.g., through various synthetic

z [µm]
4
2
membranes [25, 26]. 2
1
In terms of Onsager’s reciprocal relations, χ is the 1
1 6
4 surface temperature profile at the
top glass slide
mechano-caloric cross-coefficient, which describes equally 2
0.1
botton glass slide
0
0 1 2 3 4 5 2 3 4 5 67 2 3 4 5
well the excess heat carried by liquid flow at constant r [µm]
0.1 1
radial distance to heat source [µm]
temperature [24]. Its explicit form can also be derived
from the principles of non-equilibrium thermodynamics
Figure 2. a) Principal design of the experiment. The fluid is
[27]: A solid boundary modifies the specific chemical contained between two glass cover slides in a sandwich struc-
potential µ within an interaction length λ. Plugging ture with a gap of about 5 µm. A R = 125 nm AuNP is
the thermodynamic force −T ∇(µ/T ) in Stokes’ equation immobilized at the top surface and used as the heat source.
and using the Gibbs-Helmholtz relation d(µ/T )/dT = A focused laser beam (λ = 532 nm) is heating the particle.
−h/T 2 , one obtains eq. 1. R = 75 nm AuNPs were used to measure the velocity field.
Experimental To measure the thermo-osmotic flow b) Temperature map around the heated AuNP in cylindrical
coordinates r and z. c) Temperature profile along the upper
field we have constructed a sample cell that consists (black curve) and the lower glass (red curve) surface.
of two glass cover slides (Roth) enclosing a water film
of about 5 µm thickness (Fig. 2a). Both slides were
either used untreated or covered with Pluronic F-127 per plate decreases with the inverse square of the radius,
(see supplementary material[28]). An AuNP of radius
a = 125 nm is used as a heat source to generate a well de- ∆TAu a
vs = −χ . (3)
fined temperature gradient along the glass/water (poly- T r2
mer/water) interface (eq. 2). The AuNP was immobi- For χ > 0 the surface flow is oriented towards the ori-
lized at the upper glass surface to avoid convection. gin. A corresponding but considerably weaker surface
flow (vs0 ) is also present at the other glass cover slide.
Pinc a These surface flows induce a radially symmetric volume
T (r) = T0 + = T0 + ∆TAu . (2) flow w(r, z), which is traced by the AuNP.
4πκr r
In Fig. 3 the experimentally obtained velocity of the
The AuNP is heated with a λ = 532 nm laser at a tracers close to the upper boundary of the liquid film as a
power of Pinc = 5 mW. The temperature increment of the function of the radial distance from the heated AuNP is
AuNP ∆TAu ≈ 80 K was estimated in a separate exper- plotted. The velocity profile has been measured for a bare
iment [28]. Fig. 2 c) displays the expected temperature glass plate (black curve) and one coated with Pluronic
profiles at the liquid/solid interface corresponding to this F-127 (red curve). At distances larger than 2 microns,
nanoparticle temperature at the upper slide (black curve) both agree well with the power law ∝ r−2 (green dashed
and the lower slide (red curve). The velocity field is mea- lines). For both systems, the flow is towards higher tem-
sured by tracking single AuNP of R = 75 nm radius. peratures. The measured velocities differ by a factor
Because of its high thermal conductivity, the particle is of 7.5 where the larger one is found for the Pluronic
almost isothermal and thus does not migrate in the tem- F-127 coated surface. This suggests that the mobility
perature gradient. As the sample cell is thicker than the parameter χ is considerably stronger for the non-ionic
focal depth (≈ 1 µm) of the microscopy setup, the scat- block-copolymer as compared to the charged glass sur-
tering intensity of the tracers varies with the z−position. face. A value for vs can be extracted when considering
Selecting only the brightest particles for the analysis for that the tracer velocity represents an average hw(r, z)i∆r
example allows access to the velocity field close to the over twice the diffusion length (∆r = 620 nm) during
solid boundary. the exposure time (see SI [28]). Accordingly we find a
Velocity field at the boundary Fig. 1a,b) give a maximum velocity of vsglass = 40 µm/s and vsPluronic =
schematic view of the velocity profile which saturates at 300 µm/s
distances beyond λ in the slip velocity vs (eq. 1). As Thermo-osmotic coefficient χ. The surface velocities
the temperature gradient is not constant along the solid and the temperature increments ∆TAu yield the thermo-
boundary one expects that the slip velocity along the up- osmotic coefficients. Table 1 compares the coefficient χ
3

2
cess enthalpy of PEG in water results from the bal-
10
ance of the hydrogen bridging of the oxygen (h < 0)
8
6
and the opposite hydrophobic effect of the poly-ethylene.
radial velocity [µm/s] 4 The measured enthalpy of mixing at low PEG content
is ∆H = −0.66 × 10−20 J per monomer [31]. Taking
2
the polymer as a rod of radius b and unit length d,
1 and assuming that the enthalpy density h is constant
8
6 within the interaction length λ and zero beyond, we ob-
4 tain h = ∆H/2πbdλ and a thermo-osmotic coefficient
creep flow on
2 pluronic coated glass slide
∆H λ
χ=− ∼ 14 × 10−10 m2 /s, (4)
bare glass slide η 4πbd
0.1
2.0 3.0 4.0 5.0 with b = λ and the numerical value d = 3.5Å. This esti-
radial distance to heat source r [µm] mate agrees well with the present measurement and pre-
vious thermophoresis data for PEG.
Figure 3. Temperature profile and slip velocity along the up- The situation is less clear concerning thermo-osmosis
per boundary. The temperature agrees very well the power on a bare glass surface. Our experiment shows a slip ve-
law ∝ r−2 . The slip velocity on both glass and Pluronic- locity towards the heated spot, implying a negative en-
coated glass follows this law beyond 2 microns from the heated thalpy h, and the analysis of the experimental data gives
spot. χ = 1.8×10−10 m2 /s. This sign agrees with a very recent
experiment on silica Janus particles [37], whereas previ-
ous works reported the opposite effect: In their thermo-
osmosis experiments on porous glass, Derjaguin et al.
Table I. Thermo-osmotic parameter χ. Results are com-
pared with data of previous work for χ and the reduced ther- found χ < 0 [35], and the Soret data of Rusconi et al.
mophoretic mobility T DT , and with theoretical values as dis- on Ludox particles provided a negative thermophoretic
cussed in the main text. DT is obtained from Soret data for mobility DT [34].
poly-ethylene glycol [32, 33] and Ludox particles [34]. The The excess enthalpy h in the vicinity of a glass surface
scatter of previous data is due to differences between porous is a complex quantity. Besides electrostatic and van der
glass specimens [24] and the dependence of DT on pH and Waals forces, structuration of water e.g. due to surface
salinity [34]. We also mention that DT of PEG depends on
OH-groups seems to play a major role [35]. The latter
both molecular weight [32] and temperature [33].
results in an increase of enthalpy (h > 0), whereas the
χ (10−10 m2 /s) this work theory prev. work T DT electric-double layer contribution is negative. This would
Pluronic F-127 13 ∼ 14 15 [32, 33] suggest that the positive coefficient χ of the present ex-
Glass 1.8 ∼ 1 −(0.2...1.5) [24] −(1...9) [34] periment results from surface charges, and that previous
findings are dominated by a strong structuration con-
tribution (χ < 0). This is supported by the tempera-
obtained for the Pluronic coated and bare glass slides, ture dependence of Derjaguin’s data [24]: The negative
with previously reported thermo-osmotic data and the thermo-osmotic coefficient χ has a large positive deriva-
reduced thermophoretic mobility T DT of colloidal sus- tive dχ/dT > 0; this is expected for the structuration of
pensions. For large particles the coefficients are related water which is strong at low temperatures yet weakens at
through T DT = 32 χ [29], whereas for small particles and higher T . On the other hand, the dependence of DT on
polymers one rather has T DT ' χ. salinity and pH reported in [34], indicates a significant
Pluronic F-127 is a non-ionic triblock copolymer con- electrostatic contribution, in addition to the dominant
sisting of a central hydrophobic block of polypropylene structuration effect.
glycol (PPG, 56 repeat units, bright yellow in Fig. 1 b) The above discussion suggests that the present data
and two hydrophilic blocks of poly-ethylene glycol (PEG, (χ > 0) are related to the electrostatic contribution.
101 units, dark red). Both molecules show complex ther- Glass in contact with water in general carries surface
modynamic behavior in water; PEG is soluble under the charges, which give rise to an electric double layer with
present conditions, whereas PPG is not [30]. Thus the Debye screening length λ, as sketched in Fig. 1. The
PPG block is assumed to stick to the glass slides, whereas electric-double layer enthalpy is negative and thus results
the more hydrophilic PEG parts form a molecular brush, in a flow towards the heat source (vs < 0). In the Debye-
as illustrated in Fig. 1 b). As a consequence, there is no Hückel approximation one has h = − 21 ε(ζ/λ)2 e−2z/λ ,
well defined plane-of-shear. where ε is the permittivity and ζ the surface potential,
Since the water-polymer interface consists essentially resulting in
of hydrophilic blocks, one expects the parameter χ to εζ 2
be mainly determined by the PEG properties. The ex- χ= . (5)

4

With ζ ∼ 30 mV, one finds χ ∼ 10−10 m2 /s, which agrees a) b)


with both the sign and the order of magnitude of the
measured value.
Bulk velocity field In order to fully characterize
the thermo-osmotic flow field, we measured radial and
vertical components wr and wz of the bulk velocity
field w(r, z) and compare it to numerical simulations
3 µm
(COMSOL)[28]. The vertical separation is achieved by
collecting signals of different intensity, which correspond c) 15 pluronic coating
80 d) 2 20
back flow
10

temperature increment ∆T [K]


creep flow 0 0

radial velocity wr [µm/s]


to tracer particles at different z. Figure 4 a) depicts the

vertical velocity wz [µm/s]


60

intensity change ∆I
5 -2 -20
flow field detected for the brightest tracer particles at the 0 40 -4
-40

interface, which is directed towards the heat source. At -5 -6 experiment


-60
20 numerical simulation
intermediate intensities the flow is directed away from -10 -8
-80

-100
the heat source (Fig. 4 b). This reversal of the flow di- -15
0 1 2 3 4 5
0 -10
0 1 2 3 4 5

rection is obvious from the mass conservation; the inward radial distance [µm] radial distance [µm]

slip velocity at the boundaries requires an outward flow


in the center of the liquid film. The corresponding radial Figure 4. a) Two dimensional velocity map of the tracer
motion at the upper surface. The color indicates the verti-
dependencies are plotted in Fig. 4 c) and agree well with
cal velocity with blue being a flow in the negative z-direction
the 1/r2 dependence of eq. 3 at distances beyond 2 µm. away from the heat source and red in the positive z-direction.
At shorter distances, the velocity profile reveals a ver- b) Two-dimensional velocity map of the backflow within the
tical component wz . This vertical component averaged sample. c) Displays the average radial velocity at the upper
over the complete film thickness is detected by count- surface (red curve) and the backflow (blue curve) as a function
ing the number of particles which decrease or increase of the radial distance from the heat source. Negative veloc-
their scattering intensity within neighboring frames. A ities represent a flow towards the heat source. The dashed
line displays the vertically averaged radial velocity obtained
decrease in intensity corresponds to a downward motion
from numerical calculations. In addition the temperature in-
of the particles, away from the heat source, while an in- crement in the surrounding of the heat source is displayed
creased intensity reveals an upward motion. Figure 4 a) (purple curve). d) The average change of the tracer intensity
depicts in the color underlay that close to heat source and the corresponding vertical velocity in comparison to the
particles on average move vertically away from the heat numerical calculation.
source (blue). At about r = 2 µm distance from the heat
source a motion in the opposite direction is observed,
while even farther away no net transport in the vertical
direction is found. The corresponding intensity changes The discrepancy probably results from additional contri-
are plotted in Figure 4 d). This detected vertical flow butions to the excess enthalpy, such as structuration of
can only be due to the thermo-osmosis as convection is water or rearrangement of the hydrogen bond network,
in the used sample geometry absent. Further, the radia- as discussed in Ref. [29]. Our results further suggest
tion pressure acting in the region of the heating laser is that thermo-osmotic flows along solid/liquid interfaces
pointing against the detected vertical flow. The complete may contribute considerably to thermophoretic measure-
calculated flow field is depicted in Fig. 1 c). ments in thin film geometries and may be harnessed for
Summary We measured the velocity field caused by microfluidic applications.
thermo-osmosis due to a heated AuNP fixed at the sur- The authors acknowledge financial support by Sonder-
face of a glass slide. The inferred interfacial flow veloci- forschungsbereich TRR 102, the DFG priority program
ties reach values of up to 300 µm/s and set up a parabolic 1726 ”Microswimmers”, the Sächsische Forschergruppe
flow field in the water film at large distances from the FOR 877, the DFG the joint DFG/ANR project ”Ther-
heat source. The slip velocities are found much stronger moelectric effects at the nanoscale” and Leibniz Program
at interfaces covered with non-ionic block-copolymers as at the Universität Leipzig.
compared to bare charged glass interfaces. The good
agreement of the numbers gathered in Table 1 leads us
to the conclusion that the slip velocity on a Pluronic-
coated surface is driven by thermo-osmosis on the PEG-

water interface. More generally, they confirm the role cichos@physik.uni-leipzig.de; http://www.uni-
leipzig.de/˜physik/mona
of the interaction enthalpy for the thermal separation of
[1] B. E. Logan, M. Elimelech, Nature 488, 313 (2012).
molecular mixtures [36] and suggest that simple mod- [2] C. B. Picallo, S. Gravelle, L. Joly, E. Charlaix, L. Boc-
els such as eq. 4 provide a good description for Soret quet, Phys. Rev. Lett. 111, 244501 (2013).
data of polymers. Regarding the thermo-osmosis on a [3] B. V. Derjaguin,. G.P. Sidorenkov, Doklady Akad. Nauk.
glass surface, our findings differ from previous results. SSSR 32, 622 (1941).
5

[4] H.-R. Jiang, N. Yoshinaga, M. Sano, Phys. Rev. Lett. [23] A. Cuche, A. Canaguier-Durand, E. Devaux, J.A.
105, 268302 (2010). Hutchison, C. Genet, T.W. Ebbesen, Nano Lett. 13, 4230
[5] I. Buttinoni, G. Volpe, F. Kümmel, G. Volpe, C. (2013).
Bechinger, J. Phys.: Condens. Matter 24, 284129 (2012). [24] B.V. Derjaguin, N.V. Churaev, V.M. Muller, Surface
[6] B. Qian, D. Montiel, A. Bregulla, F. Cichos, H. Yang, forces, Plenum, New York (1987).
Chem. Sci. 4, 1420 (2013). [25] J. P. G. Villaluenga, B. Seoane, V. M. Barragán, C. Ruiz-
[7] F. M. Weinert, D. Braun, Phys. Rev. Lett. 101, 168301 Bauzá, J. Membrane Sci. 274, 116 (2006).
(2008). [26] S. Kim, M.M. Mench, J. Membrane Sci. 328, 113 (2009).
[8] R. Di Leonardo, F. Ianni, G. Ruocco, Langmuir 25, 4247 [27] S. R. de Groot, P. Mazur, Non-equlibrium Thermody-
(2009). namics, North Holland Publishing, Amsterdam (1962).
[9] A. Bregulla, H. Yang, F. Cichos, ACS Nano 8, 6542 [28] See Supplemental Material at [URL will be inserted by
(2014). publisher], which includes Refs. [38-41]
[10] S. Duhr and D. Braun, Phys. Rev. Lett. 97, 038103 [29] J. L. Anderson, Ann. Rev. Fluid Mech. 21, 61 (1989).
(2006). [30] R. Kjellander, E. Florin, J. Chem. Soc., Faraday Trans.
[11] Y. T. Maeda, A. Buguin, A. Libchaber, Phys. Rev. Lett. 1 24, 2053 (1981).
107, 038301 (2011). [31] G. N. Malcolm, J.S. Rowlinson, Trans. Faraday Soc. 53,
[12] J. N. Pedersen, C.J. Luscher, R. Marie, L.H. Tham- 921 (1957).
drup, A. Kristensen, H. Flyvbjerg, Phys. Rev. Lett. 113, [32] J. Chan, J. Popov, S. Kolisnek-Kehl, D. Leaist, J. Solut.
268301 (2014). Chem. 32, 197 (2003).
[13] M. Braun, F. Cichos, ACS Nano 7, 11200 (2013). [33] R. Kita, S. Wiegand, J. Luettmer-Strathmann, J. Chem.
[14] M. Braun, A. Bregulla, K. Günther, M. Mertig, F. Ci- Phys. 121, 3874 (2004).
chos, Nano Lett. 15, 5499 (2015). [34] R. Rusconi, L. Isa, R. Piazza, J. Opt. Soc. Am. B 21,
[15] L.-H. Yu, Y.-F. Chen, Anal. Chem. 87, 2845 (2015). 605 (2004).
[16] A. Würger, Rep. Prog. Phys. 73, 126601(2010). [35] B. V. Derjaguin, Pure Appl. Chem. 52, 1163 (1980).
[17] A. Würger, Phys. Rev. Lett. 115, 188304 (2015). [36] A. Würger, J. Phys. Condens. Matt. 26, 035105 (2014).
[18] K. A. Eslahian, A. Majee, M. Maskos, A. Würger, Soft [37] S. Nedev, S. Carretero-Palacios, P. Kühler, T. Lohmüller,
Matter 10, 1931 (2014). A. S. Alexander, L.J.E. Anderson, J. Feldmann, ACS
[19] D. Vigolo, S. Buzzaccaro and R. Piazza, Langmuir 26, Photonics 2, 491 (2015).
7792 (2010). [38] R. G. Horn, J. Phys. France 39, 105 (1978).
[20] A. Brown, W. Poon, Soft Matter 10, 4016 (2014). [39] M. Marinelli, F. Mercuri, U. Zammit and F. Scudieri,
[21] M. Reichl, M. Herzog, A. Götz, D. Braun, Phys. Rev. Phys. Rev. E 58, 5860 (1998).
Lett. 112, 198101 (2014). [40] T. Sakai and P. Alexandridis, J. Phys. Chem. B 109 7766
[22] J. Chen, Z. Kang, S.K. Kong, H.-P. Ho, Opt. Lett. 40, (2005).
3926 (2015). [41] M.R. Nejadnik, A.L.J. Olsson, P.K. Sharma, H.C. van
der Mei, W. Norde and H.J. Busscher, Langmuir 25, 6245
(2009).

Вам также может понравиться