Vous êtes sur la page 1sur 12

This article appeared in a journal published by Elsevier.

The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elseviers archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Author's personal copy


Chemical Engineering Science 66 (2011) 19621972

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Investigations of the unsteady diffusion process in microchannels


Diana Broboana a, Catalin Mihai Balan a, Thorsten Wohland b, Corneliu Balan a,n
a b

Department of Hydraulics and Fluid Machineries, REOROM Laboratory, Politehnica University of Bucharest, Splaiul Independentei 313, 060042 Bucharest, Romania National University Singapore, Department of Chemistry, 3 Science Drive 3, Singapore 117543, Singapore

a r t i c l e i n f o
Article history: Received 12 August 2010 Received in revised form 20 January 2011 Accepted 24 January 2011 Available online 2 February 2011 Keywords: Micro-channel hydrodynamics Diffusion process Buttery effect Flow visualizations Confocal microscopy Numerical computation

a b s t r a c t
This paper is concerned with the investigations and modeling of the unsteady diffusion process along a straight micro-channel with a cross section of 380 mm 360 mm. The studied process is characterized clet number (Pe 4 1000). The 3D computations of the by small Reynolds numbers (Re o 10) and high Pe coupled momentum and diffusion equations for isochoric motions are performed with the FLUENT code using the unsteady solvers for both equations. In the limit of stationary solutions, the numerical results are validated by direct ow visualizations and experimental data using confocal microscopy. The computed distributions of concentration provide qualitative and quantitative information on the transitory diffusion process and the rate of solute spreading within the investigated geometry. In particular, the pattern of the buttery effect is represented and analyzed during the non-stationary dynamical process. The work is relevant for the design of novel microuidics applications where the control of diffusion processes at the walls are important (absorption, extraction, capture of molecules or nano-particles). & 2011 Elsevier Ltd. All rights reserved.

1. Introduction One method to determine the diffusion coefcient of a solute dispersion in a liquid solvent is considering the evolution of solute concentration along a micro-channel with well established hydrodynamics. The experimental pattern of concentration is then compared, for a range of the diffusion coefcient Di, either with the corresponding analytical solution of homogeneous diffusion (if it is available), or with the numerical solution of the NavierStokes equation coupled with the diffusivity equation. Finally, the best t of the experimental data with the computational results, in some established xed space domains of the micro-geometry, decides the appropriate value of D (for details see Adeosun and Lawal, 2009; Kamholz et al., 1999; Pan et al., 2007a, 2007b; Tian and Wohland, 2008). One of the most common investigated coupled diffusionmomentum equations is the Taylor dispersion phenomena, where the dispersion of a Dirac-pulse of a passive scalar (tracer) is studied in a channel with laminar ow (Gan et al. 2007; Hartmann and Sasso, 2007; Kamholz et al., 1999). The results show the contribution of the convective transport term to diffu clet number on the process, which sion and the inuence of the Pe determines through its power of two the increase of the effective diffusion coefcient (Ajdari et al., 2006; Stone et al., 2004).

Corresponding author. Tel.: +40 214029705; fax: + 40 214029865. E-mail address: corneliu.balan@upb.ro (C. Balan).

At present, the diffusion phenomena are mostly investigated in bifurcated planar channels of T- or Y-shapes. The cross section of the channels is rectangular, with aspect ratios between the width (W, x-direction) and the height (H, y-direction) starting from one (square section, Sullivan et al., 2007), to very large values (shallow channels, W/H 4 100, Kamholz and Yager, 2001; Kamholz et al., 2001). Two liquid specimens (one enriched with the dispersed particles in small concentration) are introduced through the branches in the main channel of length L (L b W), where the diffusivity of the dispersed particles is quantied by optical measurements along the ow (which denes the z-direction). Using classical visualizations techniques (micro-PIV, Wu and Nguyen, 2005) or molecular tagging velocimetry (Garbe et al., 2008), coupled with confocal microscopy (Ismagilov et al., 2000; Kamholz et al., 1999; Tian and Wohland, 2008), NMR imaging (Akpa et al., 2007), Raman imaging (Dambrine et al., 2009), time-resolved uorescence imaging (Benninger et al., 2005), or micro-interferometry (Garvey et al., 2008), the concentration distribution is measured in the main channel at xed distances downstream the junction (at different heights within the channel, if proper devices are available). The fundamental studies of steady diffusion in the T-sensor with rectangular shallow geometry were presented by Ismagilov et al. (2000), Kamholz et al. (2001) and Kamholz et al. (1999). The main phenomena investigated in these studies is the buttery effect, which characterizes the distribution of the scalar concentration in normal planes to the main ow direction. This asymmetry in concentration distribution is induced by the convective

0009-2509/$ - see front matter & 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2011.01.048

Author's personal copy


D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972 1963

term of the diffusion equation, due to the parabolic velocity distribution of the solvent in the micro-channel. As consequence of the uid adherence at the walls and non-uniformity velocity distribution in the planes z constant, in the vicinity of the upper and lower walls of the channel (y H/2, respectively, y H/2) the iso-concentration lines are curved, taking the form of a C shape. In this case, the inter-diffusion length d d(y, z) dened by the distance between the original interface at z 0 (assumed to be the line x 0) and the current interface is not only a function of convection length (i.e. z-coordinate), but also a function of the height in the planes with z zi constant, respectively d(H/2, zi) 4 d(0, zi), with d(y, 0) 0. In the last decade, most of the applications of diffusion processes in microchannels are directed to: (i) diffusion-based cell extraction from cell suspension (diffusion with reactive term) (Mata et al., 2008), (ii) transport with stable concentration gradients (Gorman and Wikswo, 2008), (iii) protein deposition in microgeometries (Bransky et al., 2008), or (iv) improving the mixing of the uids in order to increase the diffusivity of the solute in simple or complex microgeometries (Abonnenc et al., 2009; Beutel, 2003; Lee et al., 2005; Truesdell et al., 2003), see also the last review by Aubin et al. (2010). In all these studies (which belong to the Lab-on-a-chip applications, Stone et al., 2004) ow visualizations are normally corroborated with numerical simulations, in order to obtain a better description of the hydrodynamics and diffusion process within the tested microgeometry (Liu et al., 2004; Shih and Chung, 2008; Yamaguchi et al., 2006). Solutions of the coupled equations of motion and diffusion are based on various numerical methods and numerical codes: commercial FLUENT code (Adeosun and Lawal, 2009), Fourier vez-Torres et al., 2007), Lattice Boltzman (Sullivan transform (Este et al., 2007), or FD specialized programs (Morf et al., 2008). Numerical computations are performed normally in 3D microchannel congurations. These investigations are focused on the study of mixing processes (Liu et al., 2004; Shih and Chung, 2008), the detection of the buttery effect or the demonstration of the inuence of the local viscosity dependence on concentration of the solute on the diffusion processes (Sullivan et al., 2007). The main goal of the present paper is to model the unsteady diffusion process in a symmetric Y-geometry, in particular to determine the evolution of the non-stationary buttery effect in a straight channel with the cross section of 380 mm 360 mm. The clet numbers are values of the characteristic Reynolds and Pe Re 1.85 and Pe 3000, respectively. The 3D computations of the coupled momentum and diffusion equations for isochoric motions are performed (in the absence of reaction) with the FLUENT code, using the unsteady solvers for both equations. The applied numerical procedure is similar to the computations presented in the recent paper by Adeosun and Lawal (2009), where transient diffusion of a pulse tracer injection was studied in a 3D T-junction geometry, in order to appreciate the residence time distribution of particles in a micro-channel mixing process under a steady velocity distribution. The transient dispersion regime was also investigated by Goulpeau et al. (2007), but in relation with the control of concentration gradients along microchannels of parabolic crosssections. The unsteady diffusion in the neighborhood of the walls is also a topic of interest in establishing the correct boundary conditions at solid surfaces in the presence of adsorption process, as investigated by Brenner and Ganesan (2000), or the analysis of Brownian diffusion process in the very vicinity of the walls (Sadr et al., 2007). The numerical procedure is tested making a comparison of the stationary computed concentration along the channel with the measured distributions of a uorescent tracer, teramethyl-

rhodamine (TMR), in aqueous buffer solution (phosphate buffered saline (PBS)) (see Tian and Wohland, 2008). The experimental results, represented as distributions of the signal intensity (which is proportional to concentration of the solute) across the width of the channel, are found consistent with the numerical solutions. The unsteady diffusion was investigated exclusively by numerical simulations. We represent the iso-concentration distributions in planes normal to the channel in order to characterize the dynamics of the buttery effect and the rate of solute spreading across the channel width. The computed spectrum of concentration gives original and detailed information about the transitory diffusion which take place at contact with the channel walls channel. These results have potential inuence on the design of novel micro-channel applications based on extraction and absorptions processes, where the control of the solute concentration gradient is needed (Smith et al., 2010) and the quantication of mixing between the contact phases is important (Glasgow et al., 2004; Sun and Sie, 2010).

2. Experimental In this paper, the investigated geometry is a symmetric Y-branching channel with the angle of 751 between the entrances and the length L 100 mm of the main channel, along the z-direction (see Fig. 1). The geometry of the transverse cross section of the channel is W 380 mm (width) and H 360 mm (height), W/H 1.2, so the characteristic space scale is R 370 mm (i.e. four times the hydraulics radius). The Y-microchannels were manufactured from polydimethylsiloxane prepolymer (PDMS) (Dow Corning, Singapore), which was poured into the corresponding mold prole. It was then degassed and cured at 601 C overnight in an oven. Following that, the hardened PDMS was peeled off from the mold. Holes were then made on the inlets and outlet, both PDMS and glass slide were washed with de-ionized water and blow dried with nitrogen. Next, they were placed into the air plasma for 1 min to enable irreversible bonding of the PDMS and glass slide before placing on the hot plate at 105 1C for 30 min. Lastly, needles were introduced into the inlets and outlet and the lines at 5, 15, 20, 25 and 35 mm from the Y-junction were made on hardened PDMS substrate, so that measurements can be taken at these marked positions. In experiments the solvent is a PBS buffer solution and the uorophore solute is Tetramethylrhodamine (TMR) (TMR is a red uorophore with an excitation wavelength of 550 nm and emitting at 580 nm). TMR has a non-planar structure and sterically hindered substituents that prevents the formation of short pp interactions, which allows uorophores to exhibit strong uorescence. The visualizations have been performed using two setups: (i) confocal laser scanning microscope and (ii) inverted microscope and a CCD camera (see for details Pan et al., 2007a, 2007b; Tian and Wohland, 2008). The dye solution, c0 1% volumetric concentration of TMR1 mM in PBS (i.e. uid A + c0), was pumped from one inlet while PBS buffer solution (i.e. uid A) was pumped from the other inlet at the same ow rate (see Fig. 1). There were used two identical syringe pumps, with the ow rate within the range of (0.1, 2) ml/h, which determine an average velocity in the main channel V0A(0.5, 10) mm/s, corresponding to ReA(0.25, 5). Diffusion of the TMR in PBS buffer was detected by monitoring the change in uorescence intensity across the channel width. The 543 nm laser beam of the confocal microscope (FV300, Olympus, Singapore) was scanned in an xz directions from a depth of 0 mm (cover slide) to a depth of 180 mm (middle of the channel). The uorescence observed in the channel is reected as

Author's personal copy


1964 D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972

Fig. 1. Geometry of the channel and location of the local frame {x, y, z}: (a) the Y-geometry and dimensions corresponding to the model used in numerical simulations (the length of main channel in the real geometry is 100 mm instead of 40 mm), (b) steady path lines distributions at the inlet in the main channel, and (c) steady parabolic velocity distribution at z 0.

Fig. 2. Diffusion evolution along the ow direction (CCD pictures at V0 2.4 mm/s). The dye solution (uorescent TMR in PBS buffer solution) is represented with white color.

images in the computer. This was done at distances of 5, 15, 20, 25 and 35 mm from the Y-junction along z-direction of the ow. The experiment was also repeated using a normal uorescence microscope, the pictures being recorded with a CCD camera.

However, as the CCD camera cannot capture images at a particular depth, only one image of averaged uorescence intensity across the whole channel depth was taken at each position from the Y-junction.

Author's personal copy


D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972 1965

Data from the linescans are then related to the diffusion coefcient (D) of the uorescent compound (TMR), the graphs being tted with the program Igor Pro (Vers. 6.0, Wavemetrics, Lake Oswego, OR, USA) according to the classical one-dimensional solution of the diffusion equation, i.e.   x cx 0:5erfc p 4Dt 1

Table 1 Variation of diffusion coefcient along z-direction at V0 2.4 mm/s (here tc is the characteristic time) (see Figs. 2 and 3). The values are automatically computed by the program Igor Pro. z (mm) 5 15 20 25 35 tc (s) 2.1 6.2 8.2 10.3 14.4 D (m2/s) (2.95 7 0.03) 10 10 (3.10 7 0.05) 10 10 (3.30 7 0.06) 10 10 (3.42 7 0.06) 10 10 (3.22 7 0.08) 10 10

where c is the solute concentration and t tc (z + a)/V0 is considered the characteristic time of the diffusion along the channel length, V0 being the average velocity in the main channel (see Fig. 1). We have to mention that diffusion time scales are different on the other directions, e.g. the diffusion across the channel width has the characteristic time td W2/4D (Garvey et al., 2008; Hartmann and Sasso, 2007). We also remark that (1) is obtained with imposed innite boundary conditions for the x-direction, but here the formula is used to approximate at t constant (i.e. z constant) the distribution of solute in the vicinity of the symmetry axis of a nite channel (i.e. at x 0). Direct visualization of the TMR diffusion process in PBS is shown in Fig. 2. The pictures are taken at V0 2.4 mm/s in the main channel, the process being considered steady. The measured distributions of the uorescence intensity (proportional to the concentration of the TMR) across the channel width is represented in Fig. 3. From (1), the average computed diffusion coefcient is D (3.2 7 0.2) 10 10 m2/s (see Table 1), the value which ts the best all distributions from Fig. 3 (for details about procedure see also Munson et al. (2005)). The corresponding average value of the diffusion coefcient computed from the confocal measurements data is D (3.71 7 1.1) 10 10 m2/s (Tian and Wohland, 2008). The error is larger in the last case because the average is made not only between the values recorded at different lengths and ow rates, but also between the measurements took at different depths: y 0 (middle of the channel) and y 150 mm. In this case the presence of the buttery effect is evident (see Fig. 4), but using our set-up we could not perform exact measurements of the inter-diffusion length d at the channel walls. At this moment, quantitative results of the inter-diffusion length and the evolution of thebuttery effect along the ow can be obtained only by numerical simulations.

Fig. 4. Experimental uorescence intensity recorded across the depth of the channel with the confocal microscope at z 20 mm (see Fig. 2), between two y-coordinates. The presence of the buttery effect is easily observable in the vicinity of the wall (y 150 mm).

3. Numerical simulations Numerical computations of the diffusion process within the channel were performed using the commercial FLUENT 6.3.26 code, interfaced with the Gambit pre-processor to construct the geometry and the corresponding meshes. At each time step, the NavierStokes equation (2) is solved in parallel with the diffusion equation (3) for the concentration eld c,   @v Re Shi grad vv grad p Dv 2 @t   @c grad cUv Dc Pe Shd @t 3

Fig. 3. Experimental normalized measured uorescence intensity (arbitrary units) across the channel (x-direction), at different z-distances from the junction (see Fig. 2) (steady process); a slight asymmetry of the measurements is observed. From the best tting of the curves with relation (1) the diffusion coefcients from Table 1 are obtained.

where Re rRV =Z is the Reynolds number, Pe RV =D is the clet number, Shi R=Vti and Shi R=Vtd being the correPe sponding Strouhal numbers. The expressions (2) and (3) are nondimensional, scaled with space dimension R , the average velocity (V V0), momentum characteristic time ti rR2 =Z, and the diffusivity characteristic time td R2 =D (here Z is the dynamic viscosity, r is the mass density and D is the diffusion coefcient, all these material parameters are being considered constant within the ow eld). The ratio between the two time scales (td =ti ) denes the Schmidt number, Sc Z=rD, one of the most important parameter which characterizes the diffusion process in a viscous uid. The present simulations have been performed for a single uid phase with constant viscosity and density (details on the numerical simulations and modeling of mixing immiscible uids in microchannels using the VOF code implemented in FLUENT are given in Balan et al. (2008) and Balan et al. (2010)). A solute concentration below 1% (i.e. c r 0:01 [-]) does not change the uid properties and the diffusion coefcient is constant, so the only coupling between the Eqs. (2) and (3) is the convective term from the diffusion equation, i.e. grad cUv, where v is the NavierStokes solution for the velocity eld (see Fig. A1 from the Annex for comparison of results with constant viscosity and variable viscosity). All the simulations are performed with water (density r 1000 kg/m3 and viscosity Z 1 mPa s) in a Y-geometry with length scale R 0.37 mm and diffusion coefcients in the range

Author's personal copy


1966 D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972

10 10 o D o 10 9 (m2/s). The boundary conditions are specied as follows: constant velocity at the entrances in upper/lower branches and zero relative pressure at the exit of the main branch, and constant trace concentration c c0 0.01 at the entrance of

Table 2 Numbers of nodes and cells size for the used meshes. Mesh No. of nodes x-axis (width) (mm) 40 20 15 y-axis (high) (mm) 40 30 20 z-axis (length) (mm) 40 20 16

M1 M2 M3

164,400 857,571 1,843,532

the upper branch and c 0 at the entrance of the lower one, with zero diffusive uxes at the wall. The analyzed nominal case corresponds to D0 6 10 10 m2/s and V0 4.8 mm/s, therefore the ow is characterized by the following values of the non-dimensional parameters: Re 1.85, Pe 3000, and Sc 1600 (with Sc Pe=Re). The similitude criteria clet number (the same geometry is with experiments is the Pe used, but the values of the diffusion coefcient and the velocity is doubled in order to speed up the computations). The motion within the channel is purely laminar and the velocity prole is stabilized much faster than the concentration distribution (Sc b 1). For that reason in similar studies the unsteady part of the momentum equation is neglected in comparison to the transient diffusion (Adeoson and Lawal, 2009).

Mesh M1 0.010 Concentration [-] 0.008 0.006 0.004 0.002 0.000 -0.0002 -0.0001 0.0000 0.0001 Channel width [m] t = 10s 0.010 Concentration [-] 0.008 0.006 0.004 0.002 0.000 -0.0002 -0.0001 0.0000 0.0001 Channel width [m] t = 30s 0.010 Concentration [-] 0.008 0.006 0.004 0.002 0.000 -0.0002 -0.000 10.0000 0.0001 Channel width [m] 0.0002
0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.005 0.010 0.015 0.020 0.025 0.030 0.035

Mesh M2 0.010 Concentration [-] 0.008 0.006 0.004 0.002 0.000 0.0002 -0.0002 -0.0001 0.0000 0.0001 Channel width [m] t = 10s 0.0002
0.005 0.010 0.015 0.020 0.025 0.030 0.035

t = 5s

t = 5s

0.010 Concentration [-] 0.008 0.006 0.004 0.002 0.000

x = 0.005 x = 0.010 x = 0.015 x = 0.020 x = 0.025 x = 0.030 x = 0.035

0.0002

-0.0002

-0.0001 0.0000 0.0001 Channel width [m] t = 30s

0.0002

0.010 Concentration [-] 0.008 0.006 0.004 0.002 0.000

0.005 0.010 0.015 0.020 0.025 0.030 0.035

-0.0002

-0.0001 0.0000 0.0001 Channel width [m]

0.0002

Fig. 5. Time variation of concentration, c c(x), in the middle of the channel (y 0), at z zi constant (inuence of the mesh size; comparison between M1 and M2 meshes).

Author's personal copy


D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972 1967

Since the ow is perfectly symmetric and the input branches are long enough, the ow reaches a stable velocity parabola at the entrance in the main channel at 10 ms from the onset of the motion (at Re o 10). The tracer starts to be present in the main channel, at z 0 mm, not before 1 s, so the process is basically time dependent only in the diffusion equation (3). Only after 30 s the solute concentration has a steady distribution along the channel. The coupling between (2) and (3) becomes relevant either at small Schmidt numbers or for non-symmetric entrance conditions in the main channel and unsteady velocity distributions. These ows are not explored in this paper, but we mention that the latest case is at the moment under study in our laboratory. The numerical simulations were performed with the following options: laminar solver (Green Gauss cell based gradient option), unsteady formulation (second order implicit scheme), pressure velocity coupling (PISO) and QUICK scheme as the interpolation method for convective terms. The computations used a time step of 10 4 s and a 10 10 convergence criteria, Fluent code being installed on a 64-bit server (Dual 2.66 GHz with 16 GB RAM memory). The inuence of the grid size was investigated using three meshes (see Table 2).

Fig. 8. Comparison between the analytic solutions (continuous line in the detail), experimental data (Fig. 3) and the corresponding 3D solutions (scatter) at z 20 mm. The qualitative conrmation of the computations are given by the experimental picture from Fig. 4.

Fig. 6. Time variation of concentration, c c(x), in the middle of the channel (y 0), at z zi constant; comparison between M2 and M3 meshes, at time 5 s, respectively 10 s.

Fig. 7. Flow geometry at z constant. The buttery effect is represented by the curvature of the iso-concentration line c c(x, y) 0.1c0.

Fig. 9. Comparison between the experiments (Fig. 3) and the corresponding 3D solutions in the center of the channel (a) and at the wall (b) (t 30 s, Pe 3000; experiment: Re 0.925, numerics: Re 1.85).

Author's personal copy


1968 D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972

The comparisons between the results computed with meshes M1, M2, and M3 are shown in Figs. 5 and 6. One can observe that the grid size inuence is remarkable only in the transitory regime (see Fig. 5). In the limit of the steady state (t Z 30 s) the results are almost identical. The transient numerical results computed with M1 and M2 meshes differ signicantly (see Fig. 5) (t o 30 s). The precision of the numerical calculus is directly inuenced by the clet number; it is doubled for the M1 magnitude of the local Pe grid, so the errors generated during the computations of the unsteady and convective terms from (3) are larger by a factor of two. Since the meshes M2 and M3 (Fig. 6) produce minor differences also during the transitory regime, the mesh M2 was considered the optimum for the next computations (the computation time is signicant larger for M3 and 1 s of t-time ow simulation for mesh M2 consumes up to 12 h of computing time). However, it is well known that calculus of diffusion uxes (i.e. grad cUv) through the cell faces involves (especially at high global clet numbers) a truncation error, referred to as numerical or false Pe diffusion (see Ferziger and Peric, 1999; Chung, 2003). The present simulations are characterized by a CFL factor less than one and grid clet number computed lines aligned with the ow, but the local Pe with the grid dimension is much larger then one (100 o Pelocal o 200), so the accuracy of the numerical prediction is limited (for details see Barz et al., 2008; Soleymani et al., 2008). Since it is not feasible to rene the grid in order to decrease signicantly the clet number and we do not have access to other numerical local Pe codes, the only possibility to appreciate the degree of computational predictions is the direct comparison with measured data (comparison between the performances of different commercial CFD codes

5s

10 s

15 s

30 s

Fig. 10. Inter-diffusion lengths, dw (wall) and d0 (central), at different diffusion coefcients, along the ow direction (z-coordinate).

Fig. 12. Time dependence of the concentration distributions in the main channel along the ow direction, in sections z 5/10/15/20/25/30/35 mm; c 4 0.5c0 (red), c o 0.1c0 (blue); D 6 10 10 m2/s, V0 4.8 mm/s (simulations performed at Pe 3000, respectively, Re 1.85). (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

c > 0.1c0 c = 0.1 c0 c < 0.1c0 c = 0.1c0

Fig. 11. Iso-concentration line c 0.1c0 at t 35 s, for different diffusion coefcients, in section z 5 mm (average velocity in the channel: V0 4.8 mm/s).

Author's personal copy


D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972 1969

and experiments performed in similar microudic applications are discussed by Glatzel et al., 2008).

4. Comparison between experiments and simulations Since the numerical simulations are performed at the same Pe-number as the experiments from Section 2, the steady state numerical patterns of the concentration spectrum have to reproduce the experiments, with the remark that the corresponding iso-concentration lines are reached faster in simulations (since the Reynolds number is twice larger in the simulation than in experiments, see Fig. A2 from the Annex). One phenomena with a major contribution to the complexity of computation is the convection of concentration, and as a consequence the presence of the buttery effect mentioned in Section 1 (see also Kamholz and Yager, 2001). Beyond the length of stabilization of the parabolic velocity distribution vz vz x, y, along the main channel at z constant, the solute concentration c is not only a function of the x-direction (channel width), but also dependents on the channel height (y-direction), i.e. c c(x, y). In steady state, diffusion in x-direction is higher in the vicinity of the walls then in the center (y 0), because convection is lower where velocity is lower, hence dw 4 d0 (see Fig. 7) (see the experimental qualitative conrmation from Fig. 4). Here dw is the inter-diffusion length at the wall (y 7 H/2) and d0 (y 0) is the inter-diffusion length at the middle of the channel, both lengths correspond to an iso-concentration line (in this case, c(x, y) 0.1c0) within the plane z constant. In Figs. 8 and 9 the numerical 3D results of the steady concentration eld (t Z 30 s) are compared with the analytic

solutions, relation (1), and with the experimental data from Fig. 3, respectively. The phenomena is well represented by the numerical solutions, and both the analytical solution and the experimental data from Fig. 3 being between the values computed in the middle and at the wall of the channel (see Fig. 8). One observes in Fig. 9 that 3D numerical solutions at y 0 t better the experimental data at small and medium z-coordinates, since at large distances from the junction (z 4 20 mm) the measured concentration is fair tted by the numerical solution at the wall (y H/2). This fact might be explained by the presence of the buttery effect, which increases the width of the separation band observed by direct visualizations between the marked and un-marked regions with uorophore solute. However, some discrepancies between experiments and computation might be generated also by two other factors: (i) propagation of the numerical errors along the ow direction (numerical diffusion), or (ii) lack of isotropy of the real diffusion coefcient within the channel and possible dependence of diffusion on the residence time. In Fig. 10 are represented the computed inter-diffusion lengths for different values of diffusion coefcients, as function of the z-coordinate, at t 35 s (considered the steady state for all cases). For the geometry under investigation, the dependence of interdiffusion lengths discloses an increase with constant slope (n) at small values of D (D o 5 10 10 m2/s), nw o n0 (where dw pznw , respectively, d0 pzn0 : We have to remark that the exponents nw and n0 are in this case smaller than the values reported by Ismagilov et al. (2000) and Kamholz and Yager (2001): nw 0.24 (instead of 0.33) and n0 0.27 (instead of 0.5), but in the mentioned papers the exponents were obtained for different parameters, i.e. D 10-9 m2/s, W/H 2.5 and Pe b z/H b 1 (see also the review of Stone et al., 2004).

x y

t = 35 s c < 0.1c0 c < 0.1 c0

t = 35 s

10 s

t6s

t 6.3 s

10 s t = 35 s c < 0.1c0 t = 35 s c < 0.1c0


Fig. 13. Time evolution (t r 35 s) of the iso-concentration line c 0.1c0, at different z-coordinates: (a) z 5 mm, (b) z 10 mm, (c) z 20 mm, and (d) z 35 mm. The domains with concentration c o 0.1c0 at t 35 s (steady state) are marked (simulations performed at Pe 3000, respectively, Re 1.85).

15 s

Author's personal copy


1970 D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972

From the present numerical results, one observes that at D 4 5 10 10 m2/s the exponents are increasing, but their values are not maintained constants along the ow direction of the tested channel. As was expected, the increase of the diffusion clet number, if velocity is kept constant) coefcient (i.e. smaller Pe determines the increase of the inter-diffusion lengths and, consequently, the spreading of solute is faster in the channel (see also Fig. 11) where the buttery effect is represented. One concludes that the 3D simulations produced fairly consistent results with the measured data and offer a real description of the investigated diffusion process, at least from the qualitative point of view. The numerical distributions of concentration can also provide valuable information about the transitory diffusion process within the channel, in particular to characterize the dynamics of the buttery effect and the rate of solute spreading at the wall.

5. Transient buttery effect One goal of the present study is to investigate the transitory regime (t o 35 s), in particular to determine the dynamics of the iso-concentration patterns. In Figs. 12 and 13 are shown the computed concentration distributions for the transitory diffusion process (nominal case; at time t 0 the solute enters the upper branch of the Y-geometry). The results shown in Fig. 13 disclose a different pattern of the inter-diffusion length distribution (corresponding to c 0.1c0), in comparison to the steady results from Fig. 10. At each constant z-coordinate, we remark the existence of time thresholds which separate qualitatively the unsteady distributions of d0 and dw , e.g. at t o tcr1 the concentration in the very vicinity of the walls y 7 H/2 is c o 0.1c0, so the solute actually does not reach the wall. The most important aspect to be mentioned is the existence of the second critical time value, tcr2, which denes the time corresponds to dw d0 (e.g. tcr2 15 s in Fig. 13d). For tcr1 o t o tcr2 there coexists two values for dw , both smaller than d0 , at least one being negative (see the topology of iso-concentration line at t 10 s in Fig. 13d). Therefore, always a close domain with c o 0.1c0 is present in vicinity of the upper corners of the ow section, beside the region with c o 0.1c0 located near the wall x W/2. For t 4 tcr2 the pattern of iso-concentration lines became qualitatively similar to the steady case and the region of c o 0.1c0 is present only in the vicinity of x W/2. The unsteady spectrum of concentration from Fig. 13, corroborated with the corresponding velocity and wall shear stress distributions, offer the possibility to obtain detailed information about transitory diffusion processes which take place at the contact with the channel walls, relevant for applications as: (i) surface adsorption (Das and Chakraborty, 2010), nano-particles (molecules) capture (Munir et al., 2009), liquidliquid microextraction (Kikutani et al., 2009).

Fig. 14. Concentration rate variation along the width channel at y 0, with z-coordinate parameter at t 5 s and with time parameter at z 0.01 m. The original data are presented in Fig. 5.

6. Final remarks and conclusions The modeling of diffusion processes in micro-channel congurations is important not only to determine the diffusion coefcient of a specic constituent in a solution or to obtain the desired mixing between different solutes, but also to control the diffusion in the case of active interfaces laid on the channel wall. In the latest case, the time variation of the concentration distribution during the transitory ow regime is a phenomenon of major interest. The present paper brings an original contribution in this direction, especially in the modeling of the transitory buttery

Fig. 15. Numerical computed rate along the width channel at y 0, with z-coordinate parameter (steady state).

Author's personal copy


D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972 1971

effect and the analyses of time evolution of the inter-diffusion length at the micro-channel walls. After the validation of the numerical procedure by the steady experimental results, the study investigated numerically the evolution of concentration in the analyzed Y-branching geometry, from the onset of the ow until the stabilization of the stationary spreading of the solute. The rates of concentration on different direction within the channel and the corresponding uxes of the dispersed solute can be also computed. In Fig. 14 are shown the variations in space and time of the rate dc=dx: Of course, as computations are approaching the steady state, the graphs are approaching the normal symmetric distribution, more and more at as the z-coordinate is increasing (see Fig. 15). A special aim of the paper was to obtain the description of the transitory buttery effect along the main channel. The time evolution of the iso-concentration patterns in the ow domain is a qualitative characteristic pattern of any diffusion process. In the view of the authors, the simulation and calculus of transitory diffusion in micro-channels has a direct application in developing new techniques and applications, as the modeling, design and control of cell extraction/absorption processes through porous walls of micro-channels (see also Zhu, 2009).

The further investigations are focused on two directions: (i) to establish the inuence on transitory diffusion of the non-symmetric and unsteady velocities distributions at the entrance in the channel and (ii) to describe the transitory hydrodynamics and diffusion in the vicinity of porous surface, at the micro-channel walls.

Acknowledgements This work was nancially supported through CEEX-NANOINT project (Romanian National Authority for Scientic Research ANCS) and CNCSIS grant BD 73. Corneliu Balan acknowledges also the nancial support of the National University of Singapore (EERSS program). Thorsten Wohland acknowledges funding by the Singaporean Ministry of Education (R-143-000-358-112). The authors are very thankful to Dr. Tiberiu Barbat for his advices and support in obtaining the numerical results.

Annex See Figs. A1 and A2 for more details.

(1 + 2.5 c)

7.5 s

7.5 s

6.3 s

6.5 s

10 s 15 s 20 s 30 s

10 s 15 s 20 s 30 s

Fig. A1. Time evolution of the iso-concentration line c 0.1c0 at z 35 mm (D 6 10 10 m2/s, V0 4.8 mm/s), for constant viscosity coefcient (Z0) and variable viscosity (dependent on the local concentration c, the Einstein formula).

D = 610-10 m2/s, V = 2.4 mm/s

D = 1210-10 m2/s, V = 4.8 mm/s

10 s

5s

15 s 10 s 30 s 15 s 30 s

Fig. A2. Time evolution of the iso-concentration line c 0.1c0 at z 20 mm. For the same Pe-number, the iso-concentration lines have the same topology at time t (V0/V)t0 (compare the corresponding iso-concentration lines from the two gures at the bold marked times).

Author's personal copy


1972 D. Broboana et al. / Chemical Engineering Science 66 (2011) 19621972

References
Adeosun, J.T., Lawal, A., 2009. Numerical and experimental studies of mixing characteristics in a T-junction microchannel using residence-time distribution. Chem. Eng. Sci. 64, 24222432. Abonnenc, M., Josserand, Jacques, Girault, Hubert H., 2009. Sandwich mixer reactor: inuence of the diffusion coefcient and ow rate ratios. Lab Chip 9, 440448. Ajdari, A., Bontoux, N., Stone, H.A., 2006. Hydrodynamic dispersion in shallow microchannels: the effect of cross-sectional shape. Anal. Chem. 78, 387392. Akpa, B.S., Matthews, S.M., Sederman, A.J., Yunus, K., Fisher, A.C., Johns, M.L., Gladden, L.F., 2007. Study of miscible and immiscible ows in a microchannel using magnetic resonance imaging. Anal. Chem. 79, 61286234. Aubin, J., Ferrando, M., Jiricny, V., 2010. Current methods for characterising mixing and ow in microchannels. Chem. Eng. Sci. 65, 20652093. Balan, C., Marculescu, C., Calin, A., 2008. Modelling of interfaces between inmiscible uids with the VOF method. In: Susan-Resiga, R., Bernad, S., Muntean, S. (Eds.), Vortex Hydrodynamics and Applications. Eurostampa Publ., Timisoara, pp. 195203. Balan, C.M., Broboana, D., Balan, C., 2010. Mixing process of immiscible uids in micro-channels. Int. J. Heat Fluid Flow. doi:10.1016/j.ijheatuidow.2010.06.008. Barz, D.P.J., Zadeh, H.F., Ehrhard, P., 2008. Laminar ow and mass transport in a twicefolded microchannel. AIChE J. 54, 381393. Benninger, R.K.P., Hofmann, O., McGinty, J., Requejo-Isidro, J., Munro, I., Neil, M.A.A., deMello, A.J., French, P.M.W., 2005. Time-resolved uorescence imaging of solvent interactions in microuidic devices. Opt. Express 13 (16), 62756285. Beutel, D., 2003. Mixing in microchannels. Ph.D. Thesis, Harvey Mudd College, USA. Bransky, A., Korin, N., Levenberg, S., 2008. Experimental and theoretical study of selective protein deposition using focused micro laminar ows. Biomed. Microdevices 10, 421428. Brenner, H., Ganesan, V., 2000. Molecular wall effects: are conditions at a boundary boundary conditions? Phys. Rev. E 61, 68796897. Chung, T.J., 2003. Computational Fluid Dynamics. Cambridge University Press. raud, B., Salmon, J.-B., 2009. Interdiffusion of liquids of different Dambrine, J., Ge viscosities in a microchannel. New J. Phys. 11, 075015. Das, S., Chakraborty, S., 2010. Augmented surface adsorption characteristics by employing patterned microuidic substrates in conjunction with transverse electric elds. Microuid Nanouid 8, 313327. vez-Torres, A., Gosse, C., Le Saux, T., Allemand, J.-F., Croquette, V., BerthouEste mieux, H., Lemarchand, A., Jullien, L., 2007. Fourier analysis to measure diffusion coefcients and resolve mixtures on a continuous electrophoresis chip. Anal. Chem. 79, 82228231. Ferziger, J.H., Peric, M., 1999. Commputational Methods for Fluid Dynamics. Springer, Berlin. Gan, H.Y., Lam, Y.C., Nguyen, N.T., Tam, K.C., Yang, C., 2007. Efcient mixing of viscoelastic uids in a microchannel at low Reynolds number. Microuid Nanouid 3, 101108. Garbe, C.S., Roetmann, K., Beushausen, V., Jahne, B., 2008. An optical ow MTV based technique for measuring microuidic ow in presence of diffusion and Taylor dispersion. Exp. Fluids 44, 439450. Garvey, J., Newport, D., Lakestani, F., Whelan, M., Joseph, S., 2008. Full eld measurement at the micro-scale using licro-interferometry. Microuid Nanouid 5, 7787. Glasgow, I., Lieber, S., Aubry, N., 2004. Parameters inuencing pulsed ow mixing in microchannels. Anal. Chem. 76 (16), 48254832. Gorman, B.R., Wikswo, J.P., 2008. Characterization of transport in microuidic gradient generators. Microuid Nanouid 4, 273285. Goulpeau, J., Lonetti, Barbara, Trouchet, Daniel, Ajdari, Armand, Tabeling, Patrick, 2007. Building up longitudinal concentration gradients in shallow microchannels. Lab Chip 7, 11541161. Glatzel, T., Litterst, C., Cupelli, C., Lindemann, T., Moosmann, C., Niekrawietz, R., Streule, W., Zengerle, R., Koltay, P., 2008. Comput. Fluids 37, 218235. Hartmann, C., Sasso, L., 2007. Convectiondiffusion in microchannels, Internal report, DTU, Lyngby. Ismagilov, R.F., Stroock, A.D., Kenis, P.J.A., Whitesides, G., Stone, H.A., 2000. Experimental and theoretical scaling laws for transverse diffusive broadening in two-phase laminar ows in microchannels. Appl. Phys. Lett. 76, 23762378.

Kamholz, A.E., Weigl, B.H., Finlayson, B.A., Yager, P., 1999. Quantitative analysis of molecular interaction in a microuidic channel: T-sensor. Anal. Chem. 71, 53405347. Kamholz, A.E., Yager, P., 2001. Theoretical analysis of molecular diffusion in pressure-driven laminar ow in microuidic channels. Biophys. J. 80, 155160. Kamholz, A.E., Schilling, E.A., Yager, P., 2001. Optical measurement of transverse molecular diffusion in microchannel. Biophys. J. 80, 19671972. Kikutani, Y., Mawatari, K., Hibara, A., Kitamori, T., 2009. Circulation microchannel for liquidliquid microextraction. Microchim. Acta 164, 241247. Lee, N.Y., Yamada, M., Seki, M., 2005. Development of a passive micromixer based on repeated uid twisting and attening, and its application to DNA purication. Anal. Bioanal. Chem. 383, 776782. Liu, Y.Z., Kim, B.J., Sung, H.J., 2004. Two-uid mixing in a microchannel. Int. J. Heat Fluid Flow 25, 986995. Mata, C., Longmire, E.K., McKenna, D.H., Glass, K.K., Hubel, A., 2008. Experimental study of diffusion based extraction from a cell suspension. Microuid Nanouid 5, 529540. Morf, W.E., van der Wal, P., de Rooij, N.F., 2008. Computer simulation and theory of the diffusion- and ow-induced concentration dispersion in microuidic devices and HPLC systems based on rectangular microchannels. Anal. Chim. Acta 622, 175181. Munir, A., Wang, J., Li, Z., Zhou, H.S., 2009. Numerical analysis of a magnetic nanoparticle-enhanced microuidic surface-based bioassay. Microuid Nanouid. doi:10.1007/s10404-009-0497-3. Munson, M.S., Hawkins, Kenneth R., Hasenbank, Melissa S., Yager, Paul, 2005. Diffusion based analysis in a sheath ow microchannel: the sheath ow T-sensor. Lab Chip 5, 856862. Pan, X.T., Yu, H., Shi, X.K., Korzh, V., Wohland, T., 2007a. Characterization of ow direction in microchannels and zebrash blood vessels by scanning uorescence correlation spectroscopy. J. Biomed. Opt. 12 (1), 014034. Pan, X.T., Foo, W., Lim, W., Fok, M.H.Y., Liu, P., Yu, H., Maruyama, I., Wohland, T., 2007b. Multifunctional uorescence correlation microscope for intracellular and microuidic measurements. Rev. Sci. Instr. 78, 053711. Sadr, R., Hohenegger, Ch., Li, H., Mucha, P.J., Yoda, M., 2007. Diffusion-induced bias in near-wall velocimetry. J. Fluid Mech. 577 (443456). Shih, T.R., Chung, C.K., 2008. A high-efciency planar micromixer with convection and diffusion mixing over a wide Reynolds number range. Microuid Nanouid 5, 175183. Soleymani, A., Kolehmainen, E., Turunen, I., I., 2008. Numerical and experimental investigations of liquid mixing in T-type micromixers. Chem. Eng. J. 135S, S219S228. Smith, R.L., Demers, R.L., Collins, S.D., C.J., 2010. Microuidic device for the combinatorial application and maintenance of dynamically imposed diffusional gradients. Microuid Nanouid 9, 613622. Stone, H.A., Stroock, A.D., Ajdari, A., 2004. Engineering ows in small devices: microuidics toward a Lab-on-a-chip. Annu. Rev. Fluid Mech. 36, 381411. Sullivan, S.P., Akpa, B.S., Matthews, S.M., Fisher, A.C., Gladden, L.F., Johns, M.L., 2007. Simulation of miscible diffusive mixing in microchannels. Sensors Actuators B 123, 11421152. Sun, C.-L., Sie, J.-Y., 2010. Active mixing in diverging microchannels. Microuid Nanouid 8, 485495. Tian, L.Q., Wohland, T., 2008. Measuring reactions in microchannels. NUS Internal Research Report, P07265. Truesdell, R.A., Vorobieff, P.V., Sklar, L.A., Mammoli, A.A., 2003. Mixing a continuous ow of two uids due to unsteady ow. Phys. Rev. E 67, 066304. Wu, Z., Nguyen, N.-T., 2005. Hydrodynamics focusing in microchannels under consideration of diffusive dispersion: theories and experiments. Sensors Actuators B 107, 965974. Yamaguchi, Y., Ogura, D., Yamashit, K., Miyazaki, M., Nakamura, H., Maeda, H., 2006. A method for DNA detection in a microchannel: uid dynamics phenomena and optimization of microchannel structure. Talanta 68, 700707. Zhu, X., 2009. Micro/nanoporous membrane based gaswater separation in microchannel. Microsyst. Technol. 15, 14591465.

Vous aimerez peut-être aussi