Next Article in Journal
Unexpected Phase Behavior of Pluronic Polymer-Organic Derivative Mixtures Depending on Temperature in Aqueous Solution
Next Article in Special Issue
Organs-on-a-Chip Module: A Review from the Development and Applications Perspective
Previous Article in Journal
Mechanical Fatigue Resistance of Piezoelectric PVDF Polymers
Previous Article in Special Issue
A High-Throughput Electrokinetic Micromixer via AC Field-Effect Nonlinear Electroosmosis Control in 3D Electrode Configurations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Exact Solution for Power-Law Fluids in a Slit Microchannel with Different Zeta Potentials under Electroosmotic Forces

Department of Mechanical Engineering, Korea National University of Transportation, Chungju, 27469, Korea
*
Authors to whom correspondence should be addressed.
Micromachines 2018, 9(10), 504; https://doi.org/10.3390/mi9100504
Submission received: 18 September 2018 / Revised: 3 October 2018 / Accepted: 3 October 2018 / Published: 5 October 2018
(This article belongs to the Special Issue Micro/Nano-Chip Electrokinetics, Volume III)

Abstract

:
Electroosmotic flow (EOF) is one of the most important techniques in a microfluidic system. Many microfluidic devices are made from a combination of different materials, and thus asymmetric electrochemical boundary conditions should be applied for the reasonable analysis of the EOF. In this study, the EOF of power-law fluids in a slit microchannel with different zeta potentials at the top and bottom walls are studied analytically. The flow is assumed to be steady, fully developed, and unidirectional with no applied pressure. The continuity equation, the Cauchy momentum equation, and the linearized Poisson-Boltzmann equation are solved for the velocity field. The exact solutions of the velocity distribution are obtained in terms of the Appell’s first hypergeometric functions. The velocity distributions are investigated and discussed as a function of the fluid behavior index, Debye length, and the difference in the zeta potential between the top and bottom.

1. Introduction

Recently, microfluidic device applications are increasing in the fields of chemical analysis, medical diagnostics, material synthesis, and others [1,2,3]. In the field of microfluidics, flow control in a microchannel is one of the most important issues. The problem with conventional pressure-driven flow is that as the channel size decreases, the hydraulic area becomes extremely small, resulting in a significant increase in the corresponding hydraulic resistance [4]. Electroosmotic flow (EOF) does not suffer from this problem because it is the motion of fluid that depends on the electric field across a microchannel [5,6,7].
Many efforts have been made to study electroosmotic flow (EOF) using Newtonian fluids. However, a few microfluidic devices are used more frequently for processing biological fluids such as blood, saliva, DNA, and polymer solutions, which cannot be treated as Newtonian fluids. To analyze the EOF of such fluids, an approach to non-Newtonian constitutive relations should be considered [8,9,10,11,12,13,14,15].
Among the various constitutive laws for non-Newtonian fluids, the power law model is the most popular because of its simplicity and suitability for analyzing a wide range of fluids. Thus, many researchers have conducted EOF studies using the power law model [16,17,18,19,20,21,22,23,24,25]. Zhao et al. analyzed the EOF of power-law fluids and obtained the approximate solution of the velocity field in a slit microchannel [16]. In addition, they studied a general Smoluchowski slip velocity over a surface [17] and provided an exact solution of the velocity distribution in a slit microchannel [18]. They also analyzed the EOF of power-law fluids in cylindrical [19] and rectangular [20] microchannels. Tang et al. conducted a numerical study of the EOF in microchannels of a power-law fluid using the lattice Boltzmann method [21]. Vasu and De analyzed a mathematical model of the EOF of power-law fluids in a rectangular microchannel at high zeta potential [22]. Babaie et al. and Hadigol et al. numerically analyzed the EOF of power-law fluids in a slit microchannel with pressure gradient [23,24]. Ng and Qi developed a simplified analytical model to describe the electrokinetic flow of a power-law fluid for varying wall potentials and channel heights in a slit channel [25].
Most previous studies have been performed on microchannels with the same zeta potential at the top and bottom walls. To our knowledge, however, many microfluidic devices are made with a combination of different materials, such as silicon dioxide (glass) as the base and polydimethylsiloxane as the top and side-walls. In these cases, asymmetric electrochemical boundary conditions should be applied for reasonable analysis of the EOF. Afonso et al. [26] and Choi et al. [27] analyzed the EOF of viscoelastic fluids in a microchannel with asymmetric zeta potential using the simplified Phan–Thien–Tannar model. Qi and Ng investigated the EOF of a power-law fluid through a slit channel where the walls were asymmetrically patterned with periodic variations in shape and zeta potential [28]. Hadigol et al. numerically investigated the characteristics of electroosmotic micromixing of power-law fluid in a slit microchannel with nonuniform zeta potential distributions along the walls of the channel [29]. Jiménez et al. investigated the start-up from rest of the EOF of Maxwell fluids in a rectangular microchannel with asymmetric high zeta potentials at the walls [30]. Peralta et al. conducted theoretical analysis of the start-up of oscillatory EOF in a parallel-plate microchannel under asymmetric zeta potentials [31]. Recently, Choi et al. presented the EOF in a rectangular microchannel using numerical analysis [32] and suggested an approximate solution for the EOF of power-law fluid with asymmetric zeta potential of a planar channel [33].
Obtaining the exact solution not only provides physical insight into the phenomena but can also serve as a benchmark for experimental, numerical, and asymptotic analyses. Zhao and Yang [18] have reported the exact solution for the EOF of a power-law fluid with a symmetric zeta potential. However, it remains a challenge to obtain the exact solution of the EOF in microchannels with asymmetric electrochemical boundary conditions.
In the present study, exact solutions for EOFs of power-law fluids in a slit microchannel with different zeta potentials at the top and bottom walls are presented. In addition, the key parameters affecting the velocity distribution of EOF, including the fluid behavior index, Debye length, and different zeta potentials at the top and bottom walls are analyzed.

2. Mathematical Formulation

Figure 1 shows a two-dimensional EOF in a slit microchannel of height 2H. The top and bottom walls were charged with zeta potential ψt and ψb, respectively. An external electric field E0 was applied to a power-law fluid with a constant density ρ and electric permittivity ϵ.
The velocity field in the microchannel is governed by the continuity and Cauchy momentum equations given as:
v = 0 ,
ρ D v D t = p + τ + F ,
where v = (u, v) is the velocity vector, p is the pressure, τ is the stress tensor, and F = (Fx, Fy) is the body force. The stress tensor can be given by
τ = μ ( v + v T ) ,
where μ is the effective viscosity.
For a steady, fully developed, unidirectional flow with no applied pressure and negligible gravitational force, the body force acts only along the x-direction and the Cauchy momentum equation in Equation (2) can be simplified as:
d d y ( μ d u d y ) + F x = 0 .
The effective viscosity of the power-law fluid can be expressed as:
μ = m | d u d y | n 1 ,
where m is the flow consistency index, and n is the flow behavior index.
The body force along the x-direction is given by:
F x = ρ e E 0 .
The net charge density ρe can be obtained by the Poisson equation, which takes the form of:
ε d 2 ψ d y 2 = ρ e .
With the assumption of Boltzmann distribution and small zeta potentials, the electrical potential profile in the electrical double layer (EDL) is governed by the linearized Poisson–Boltzmann equation expressed by:
d 2 ψ d y 2 = κ 2 ψ ,
which is subject to the following boundary conditions:
ψ | y = H = ψ t , ψ | y = H = ψ b .
κ−1 is called the Debye length and is defined as κ−1 = (εkBT/2e2z2n)1/2, where n and z are the bulk number concentration and the valence of ions, respectively, e is the fundamental charge, kB is the Boltzmann constant, and T is the absolute temperature.
The solution for the electrical potential distribution is of the form:
ψ ( y ) = ψ t + ψ b 2 cosh ( κ y ) cosh ( κ H ) + ψ t ψ b 2 sinh ( κ y ) sinh ( κ H ) .
Then, the net charge density ρe can be expressed as a function of the EDL potential
ρ e ( y ) = κ 2 ε ψ ( y ) .
With all the aforementioned considerations, the Cauchy momentum equation in Equation (4) is expressed as:
d d y [ m | d u d y | n 1 d u d y ] κ 2 ε E 0 ψ ( y ) = 0 .
This equation is constrained by the following boundary conditions (no-slip conditions)
u | y = H = 0 , u | y = H = 0 .
Substituting Equation (10) into Equation (12) yields:
{ (14a) d d y [ ( d u d y ) n ] = κ 2 ε E 0 m ( ψ t + ψ b 2 cosh ( κ y ) cosh ( κ H ) + ψ t ψ b 2 sinh ( κ y ) sinh ( κ H ) ) , i f   d u d y 0 , (14b) d d y [ ( d u d y ) n ] = κ 2 ε E 0 m ( ψ t + ψ b 2 cosh ( κ y ) cosh ( κ H ) + ψ t ψ b 2 sinh ( κ y ) sinh ( κ H ) ) , i f   d u d y < 0 .
Since most of the materials that make up the microchannels have negative zeta potential [34,35], the wall zeta potentials are assumed to be negative in the present study; thus, the flow occurs in the +x-direction (if E0 > 0). Let yc be the point where d u d y | y = y c = 0 , ( H y c H ) , then the velocity gradient is positive ( d u d y 0 ) in the interval H y y c , and negative ( d u d y < 0 ) in the interval y c y H .
Integrating Equation (14) with y leads to:
d u d y = { (15a) ( κ ε E 0 ψ m m ) 1 n { sinh ( κ y ) cosh ( κ H ) R cosh ( κ y ) sinh ( κ H ) + C + } 1 n ,   if   H y y c , (15b) ( κ ε E 0 ψ m m ) 1 n { ( sinh ( κ y ) cosh ( κ H ) R cosh ( κ y ) sinh ( κ H ) + C ) } 1 n ,   if   y c < y H ,
where ψm and R are the average zeta potential and the dimensionless zeta potential difference between the top and bottom walls, respectively, which are defined by:
ψ m ψ t + ψ b 2 ,
R ψ t ψ b ψ t + ψ b ,
and C+ and C are integral constants.
Both Equations (15a) and (15b) should be zero at y = yc. Therefore,
C + = C = sinh ( κ y c ) cosh ( κ H ) + R cosh ( κ y c ) sinh ( κ H ) C .
Integrating Equation (15) with the corresponding boundary condition in Equation (13) leads to the velocity distribution:
u ( y ) = { (19a) ( κ ε E 0 ψ m m ) 1 n H y { I ( y ) + C } 1 n d y ,   if   H y y c , (19b) ( κ ε E 0 ψ m m ) 1 n H y { I ( y ) C } 1 n d y ,   if   y c < y H ,
where
I ( y ) sinh ( κ y ) cosh ( κ H ) R cosh ( κ y ) sinh ( κ H ) ,
By integrating Equation (19), the velocity distribution can be obtained as:
u ( y ) = { (21a) u s [ V + ( y ) V + ( H ) ] ,   if   H y y c , (21b) u s [ V ( y ) + V ( H ) ] ,   if   y c < y H ,
where us denotes the generalized Smoluchowski velocity for power-law fluids by employing the average zeta potential ψm at the top and bottom walls on the basis of the work of Zhao et al. [16], which is expressed as:
u s = n κ 1 n 1 ( ε E 0 ψ m m ) 1 n ,
and
V + ( y ) 1 ( n + 1 ) C 2 + w 2 [ I ( y ) + C ] n + 1 n F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   I ( y ) + C C + i w , I ( y ) + C C i w ) ,
V ( y ) 1 ( n + 1 ) C 2 + w 2 [ I ( y ) C ] n + 1 n F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   I ( y ) + C C + i w , I ( y ) + C C i w ) .
y c = 1 κ ln ( C + C 2 + w 2 R sinh ( κ H ) + 1 cosh ( κ H ) ) .
w = 1 cos h 2 ( κ H ) R 2 sin h 2 ( κ H ) .
The integral constant C can be obtained from the following equation:
J ( C ) = 0 ,
where J(x) are defined by:
J ( x ) = [ I ( H ) + x ] n + 1 n F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   I ( H ) + x x + i w , I ( H ) + x x i w ) [ I ( H ) x ] n + 1 n F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   I ( H ) + x x + i w , I ( H ) + x x i w ) .
The details of the mathematical derivations are described in Appendix A. It is a challenge to obtain the explicit form for the integral constant C. Thus, in this study, a numerical method was used for evaluating C.
F1(a; b1, b2; c; x, y) in Equation (23) is the Appell’s first hypergeometric function [36], which can be represented as a one-dimensional integral form [37]:
F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ; γ α + i β ,   γ α i β ) = Γ ( 2 + 1 n ) Γ ( 1 + 1 n ) 0 1 t 1 n 1 γ t α i β 1 γ t α + i β d t = n + 1 n 0 1 t 1 n γ 2 t 2 2 α γ t + ( α 2 + β 2 ) α 2 + β 2 d t ,
where α, β, and γ are real values, and Γ(z) is a gamma function. It is evident from Equation (28) that, although the Appell’s first hypergeometric function in Equation (23) has complex arguments, it always has a real value.
Alternatively, Equation (21) can be expressed in the single form using Equation (15) as follows:
u ( y ) = V ( y ) V ( H ) ,
where
V ( y ) 1 κ n n + 1 1 C 2 + w 2 ( d u d y ) [ I ( y ) + C ] F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   I ( y ) + C C + i w , I ( y ) + C C i w ) .
Equation (21) is applicable to the EOF of power-law fluids with arbitrary zeta potentials at the top and bottom walls. If the top and bottom walls have the same zeta potential, then the velocity distribution is expressed as follows:
u ( y ) = u s [ V s y m m ( H ) V s y m m ( y ) ] ,
where
V s y m m ( y ) ( 1 ) n 1 2 n n cosh ( κ y ) cosh 1 n ( κ H ) 2 F 1 ( 1 2 ,   n 1 2 n ;   3 2 ;   cos h 2 ( κ y ) )
which is identical to the result of Zhao and Yang [18] on the EOF of power-law fluid with a symmetrical zeta potential. The detailed derivations are described in Appendix B.

3. Results and Discussions

The key parameters that affect velocity distribution are the fluid behavior index n, electrokinetic parameter κH, and dimensionless zeta potential difference R = (ψtψb)/(ψt + ψb) between ψt and ψb. In this section, the effects of these parameters on velocity distribution are investigated.
Figure 2 shows the dimensionless velocity (u/us) distributions from Equation (21) for different values of fluid behavior index n at a fixed κH of 15. Figure 2a represents the velocity distributions with same zeta potentials (R = 0) at the bottom and top, while Figure 2b indicates those of asymmetric zeta potentials with R of 0.2 (ψt/ψb = 1.5). In both cases of symmetric and asymmetric zeta potentials, as the fluid behavior index n decreases, the velocity gradient near the wall increases, and the plug-like characteristics of velocity distribution are enhanced. This is because the fluid with smaller fluid behavior index is less viscous, and the velocity can easily change from zero at the wall to the Smoluchowski velocity at the core region.
Figure 3 shows the dimensionless velocity (u/us) distributions for different values of κH. Figure 3a shows the velocity distributions of the shear thinning fluid (n = 0.8) and Figure 3b shows those of the shear thickening fluid (n = 1.2). In both cases, as κH increases, the velocity distribution changes from parabolic type to plug-like type. The increase of κH means a decrease in Debye length. In other words, the EDL thickness, on which the electrostatic body force is applied, decreases and the velocity distribution changes to a plug-like type.
Figure 4 shows the dimensionless velocity (u/us) distributions according to the dimensionless zeta potential difference R at the bottom and top for the shear thinning fluid (Figure 4a) and shear thickening fluid (Figure 4b). For comparison, the symmetric case (R = 0) is also included in the figure. The velocity distributions near the top and bottom walls develop from zero (on the wall) to close to the generalized Smoluchowski velocity determined by the corresponding zeta potentials; in the core region, these two velocity distributions near the walls are almost linearly connected. Therefore, as the difference in zeta potential between the top and bottom walls increases, the velocity gradient in the core region increases. The velocity gradient in the core region decreases and increases the viscosity of the shear thinning fluid and shear thickening fluid, respectively. As a result, as the dimensionless zeta potential difference R increases, the velocity at the center (y/H = 0) increases for shear thinning fluids and decreases for shear thickening fluids.

4. Conclusions

In this study, the exact solutions are proposed for fully developed two-dimensional steady unidirectional EOFs of power-law fluids with different zeta potentials at the top and bottom walls. The exact solutions are expressed in terms of Appell’s first hypergeometric functions. The effects of parameters such as the fluid behavior index n, electrokinetic parameter κH, and zeta potential ψt and ψb on the velocity distribution are investigated.

Author Contributions

All authors formulated the mathematical solution and discussed the results. D.-S.C. wrote the original manuscript. S.Y. and W.S.C. reviewed and edited the manuscript. All authors read and approved the final manuscript.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP; Ministry of Science, ICT & Future Planning) (NRF-2017R1C1B5017472 and NRF-2017R1C1B5017734).

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

The general integral formula for Equation (19) is
( a sinh ( x ) + b cosh ( x ) + c ) n d x = 1 a ( n + 1 ) 1 b 2 a 2 sech ( tanh 1 ( b a ) + x )   a 1 b 2 a 2 i   a sinh ( x ) i   b cosh ( x ) a 1 b 2 a 2 + i   c a 1 b 2 a 2 + i   a sinh ( x ) i   b cosh ( x ) a 1 b 2 a 2 i   c ( a sinh ( x ) + b cosh ( x ) + c ) n + 1 F 1 ( 1 + n ;   1 2 ,   1 2 ;   2 + n ;   i ( a sinh ( x ) + b cosh ( x ) + c ) a 1 b 2 a 2 i   c , i ( a sinh ( x ) + b cosh ( x ) + c ) a 1 b 2 a 2 + i   c ) + c o n s t .
Using Equation (A1), the primitive function of { I ( y ) + C } 1 n in Equation (19) can be evaluated as
{ I ( y ) + C } 1 n d y = 1 κ n n + 1 1 C 2 + w 2 | ψ ( y ) | ψ ( y ) [ I ( y ) + C ] n + 1 n F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   I ( y ) + C C + i w , I ( y ) + C C i w ) + c o n s t .
where ψ(y), I(y) and w are defined in Equations (10), (20), and (25), respectively. The term | ψ ( y ) | ψ ( y ) means the sign of the electrical potential. Since the wall zeta potentials are assumed to be negative (ψt, ψb < 0), ψ(y) always has a negative value; therefore, the term | ψ ( y ) | ψ ( y ) is −1. With all the aforementioned considerations, the velocity distribution is obtained as Equation (21).
At point y = yc, the two equations according to the interval in Equation (21) should have the same value.
u s [ V + ( y c ) V + ( H ) ] = u s [ V ( y c ) + V ( H ) ] .
Since I ( y c ) + C = 0 by Equation (18), V + ( y c ) = V ( y c ) = 0 .
Therefore, the integral constant C can be obtained from the following equation
V + ( H ) + V ( H ) = 0
which is simplified as Equation (27).

Appendix B

For a symmetrical zeta potential, the R in Equation (17) is zero and the integral constant C is also zero because d u d y | y = 0 = 0 . Owing to the symmetry, the velocity distribution can be considered only in the interval 0 y H . Then, Equation (23b) becomes
V ( y ) = 1 n + 1 c o s h ( κ H ) [ sinh ( κ y ) cosh ( κ H ) ] n + 1 n F 1 ( 1 + 1 n ;   1 2 ,   1 2 ;   2 + 1 n ;   i sinh ( κ y ) ,   i sinh ( κ y ) ) = 1 n + 1 s i n h ( κ y ) [ sinh ( κ y ) cosh ( κ H ) ] 1 n 2 F 1 ( 1 2 , 1 2 n + 1 2 ;   ,   1 2 n + 3 2 ;   sin h 2 ( κ y ) ) ,
where Appell’s first hypergeometric function is reduced to the hypergeometric function 2F1(a1, a2; b; y).
Applying Euler’s hypergeometric transformations [38,39] to the hypergeometric function in Equation (B1) gives
V ( y ) = ( 1 ) n 1 2 n n cosh ( κ y ) cosh 1 n ( κ H ) 2 F 1 ( 1 2 , n 1 2 n ; 3 2 ; cosh 2 ( κ y ) ) + π 2 n   Γ ( n + 1 2 n )   Γ ( 2 n + 1 2 n ) ( 1 ) n 1 2 n cos h 1 n ( κ H ) .
Then the velocity distribution can be obtained as
u ( y ) = u s [ V ( y ) + V ( H ) ] = u s [ V s y m m ( H ) V s y m m ( y ) ] .
where Vsymm(y) is expressed in Equation (32).

References

  1. Squires, T.M.; Quake, S.R. Microfluidics: Fluid physics at the nanoliter scale. Rev. Mod. Phys. 2005, 77, 977–1026. [Google Scholar] [CrossRef] [Green Version]
  2. Sackmann, E.K.; Fulton, A.L.; Beebe, D.J. The present and future role of microfluidics in biomedical research. Nature 2014, 507, 181–189. [Google Scholar] [CrossRef] [PubMed]
  3. Gravesen, P.; Branebjerg, J.; Jensen, O.S. Microfluidics-a review. J. Micromech. Microeng. 1993, 3, 168. [Google Scholar] [CrossRef]
  4. Akers, A.; Gassman, M.; Smith, R. Hydraulic Power System Analysis; CRC Press: Boca Raton, FL, USA, 2006; ISBN 978-0-8247-9956-4. [Google Scholar]
  5. Lyklema, J. Fundamentals of Interface and Colloid Science: Solid-Liquid Interfaces, 1st ed.; Academic Press: London, UK, 1995; ISBN 978-0-12-460524-4. [Google Scholar]
  6. Hunter, R.J. Introduction to Modern Colloid Science, 1st ed.; Oxford University Press: Oxford, NY, USA, 1994; ISBN 978-0-19-855386-1. [Google Scholar]
  7. Li, D. Electrokinetics in Microfluidics; Academic Press: Cambridge, MA, USA, 2004; ISBN 978-0-08-053074-1. [Google Scholar]
  8. Das, S.; Chakraborty, S. Analytical solutions for velocity, temperature and concentration distribution in electroosmotic microchannel flows of a non-Newtonian bio-fluid. Anal. Chim. Acta 2006, 559, 15–24. [Google Scholar] [CrossRef]
  9. Devarakonda, S.B.; Han, J.; Ahn, C.H.; Banerjee, R.K. Bioparticle separation in non-Newtonian fluid using pulsed flow in micro-channels. Microfluid. Nanofluid. 2007, 3, 391–401. [Google Scholar] [CrossRef]
  10. Kolodner, P. Oscillatory convection in viscoelastic DNA suspensions. J. Non-Newton. Fluid Mech. 1998, 75, 167–192. [Google Scholar] [CrossRef]
  11. Owens, R.G. A new microstructure-based constitutive model for human blood. J. Non-Newton. Fluid Mech. 2006, 140, 57–70. [Google Scholar] [CrossRef]
  12. Yilmaz, F.; Gundogdu, M.Y. A critical review on blood flow in large arteries; relevance to blood rheology, viscosity models, and physiologic conditions. Korea-Aust. Rheol. J. 2008, 20, 197–211. [Google Scholar]
  13. Park, H.M.; Lee, W.M. Helmholtz–Smoluchowski velocity for viscoelastic electroosmotic flows. J. Colloid Int. Sci. 2008, 317, 631–636. [Google Scholar] [CrossRef] [PubMed]
  14. Olivares, M.L.; Vera-Candioti, L.; Berli, C.L.A. The EOF of polymer solutions. Electrophoresis 2009, 30, 921–928. [Google Scholar] [CrossRef] [PubMed]
  15. Zhao, C.; Yang, C. Electrokinetics of non-Newtonian fluids: A review. Adv. Colloid Interface Sci. 2013, 201–202, 94–108. [Google Scholar] [CrossRef] [PubMed]
  16. Zhao, C.; Zholkovskij, E.; Masliyah, J.H.; Yang, C. Analysis of electroosmotic flow of power-law fluids in a slit microchannel. J. Colloid Interface Sci. 2008, 326, 503–510. [Google Scholar] [CrossRef] [PubMed]
  17. Zhao, C.; Yang, C. Nonlinear Smoluchowski velocity for electroosmosis of power-law fluids over a surface with arbitrary zeta potentials. Electrophoresis 2010, 31, 973–979. [Google Scholar] [CrossRef] [PubMed]
  18. Zhao, C.; Yang, C. An exact solution for electroosmosis of non-Newtonian fluids in microchannels. J. Non-Newton. Fluid Mech. 2011, 166, 1076–1079. [Google Scholar] [CrossRef]
  19. Zhao, C.; Yang, C. Electroosmotic flows of non-Newtonian power-law fluids in a cylindrical microchannel. Electrophoresis 2013, 34, 662–667. [Google Scholar] [CrossRef] [PubMed]
  20. Zhao, C.; Zhang, W.; Yang, C. Dynamic Electroosmotic Flows of Power-Law Fluids in Rectangular Microchannels. Micromachines 2017, 8, 34. [Google Scholar] [CrossRef]
  21. Tang, G.H.; Li, X.F.; He, Y.L.; Tao, W.Q. Electroosmotic flow of non-Newtonian fluid in microchannels. J. Non-Newton. Fluid Mech. 2009, 157, 133–137. [Google Scholar] [CrossRef]
  22. Vasu, N.; De, S. Electroosmotic flow of power-law fluids at high zeta potentials. Colloids Surf. A Physicochem. Eng. Asp. 2010, 368, 44–52. [Google Scholar] [CrossRef]
  23. Babaie, A.; Sadeghi, A.; Saidi, M.H. Combined electroosmotically and pressure driven flow of power-law fluids in a slit microchannel. J. Non-Newton. Fluid Mech. 2011, 166, 792–798. [Google Scholar] [CrossRef]
  24. Hadigol, M.; Nosrati, R.; Raisee, M. Numerical analysis of mixed electroosmotic/pressure driven flow of power-law fluids in microchannels and micropumps. Colloids Surf. A Physicochem. Eng. Asp. 2011, 374, 142–153. [Google Scholar] [CrossRef]
  25. Ng, C.-O.; Qi, C. Electroosmotic flow of a power-law fluid in a non-uniform microchannel. J. Non-Newton. Fluid Mech. 2014, 208–209, 118–125. [Google Scholar] [CrossRef] [Green Version]
  26. Afonso, A.M.; Alves, M.A.; Pinho, F.T. Electroosmotic flow of viscoelastic fluids in microchannels under asymmetric zeta potentials. J. Eng. Math. 2011, 71, 15–30. [Google Scholar] [CrossRef]
  27. Choi, W.; Joo, S.W.; Lim, G. Electroosmotic flows of viscoelastic fluids with asymmetric electrochemical boundary conditions. J. Non-Newton. Fluid Mech. 2012, 187–188, 1–7. [Google Scholar] [CrossRef]
  28. Qi, C.; Ng, C.-O. Electroosmotic flow of a power-law fluid through an asymmetrical slit microchannel with gradually varying wall shape and wall potential. Colloids Surf. A Physicochem. Eng. Asp. 2015, 472, 26–37. [Google Scholar] [CrossRef] [Green Version]
  29. Hadigol, M.; Nosrati, R.; Nourbakhsh, A.; Raisee, M. Numerical study of electroosmotic micromixing of non-Newtonian fluids. J. Non-Newton. Fluid Mech. 2011, 166, 965–971. [Google Scholar] [CrossRef]
  30. Jiménez, E.; Escandón, J.; Bautista, O.; Méndez, F. Start-up electroosmotic flow of Maxwell fluids in a rectangular microchannel with high zeta potentials. J. Non-Newton. Fluid Mech. 2016, 227, 17–29. [Google Scholar] [CrossRef]
  31. Peralta, M.; Arcos, J.; Méndez, F.; Bautista, O. Oscillatory electroosmotic flow in a parallel-plate microchannel under asymmetric zeta potentials. Fluid Dyn. Res. 2017, 49, 035514. [Google Scholar] [CrossRef]
  32. Choi, W.; Yun, S.; Choi, D.-S. Electroosmotic flows of power-law fluids with asymmetric electrochemical boundary conditions in a rectangular microchannel. Micromachines 2017, 8, 165. [Google Scholar] [CrossRef]
  33. Choi, W.; Yun, S.; Choi, D.-S. Approximate solution for electroosmotic flow of power-law fluids in a planar microchannel with asymmetric electrochemical boundary conditions. Micromachines 2018, 9, 265. [Google Scholar] [CrossRef]
  34. Kirby, B.J.; Hasselbrink, E.F. Zeta potential of microfluidic substrates: 1. Theory, experimental techniques, and effects on separations. Electrophoresis 2004, 25, 187–202. [Google Scholar] [CrossRef] [PubMed]
  35. Gallardo-Moreno, A.M.; Vadillo-Rodríguez, V.; Perera-Núñez, J.; Bruque, J.M.; González-Martín, M.L. The zeta potential of extended dielectrics and conductors in terms of streaming potential and streaming current measurements. Phys. Chem. Chem. Phys. 2012, 14, 9758–9767. [Google Scholar] [CrossRef] [PubMed]
  36. Appell, P. Sur les séries hypergéométriques de deux variables et sur des équations différentielles linéaires aux dérivées partielles. C. R. Hebd. Seances Acad. Sci. 1880, 90, 296–298, 731–735. [Google Scholar]
  37. Picard, É. Sur une extension aux fonctions de deux variables du problème de Riemann relativ aux fonctions hypergéométriques. Annales Scientifiques de l’École Normale Supérieure Série 2 1881, 10, 305–322. (In French) [Google Scholar] [CrossRef]
  38. Barnes, E.W. A New Development in the Theory of the Hypergeometric Functions. Proc. Lond. Math. Soc. 1908, 6, 141–177. [Google Scholar] [CrossRef]
  39. Bailey, W.N. Generalised Hypergeometric Series; Cambridge University Press: Cambridge, UK, 1935; pp. 3–4. [Google Scholar]
Figure 1. Schematic diagram of electroosmotic flow in a slit microchannel.
Figure 1. Schematic diagram of electroosmotic flow in a slit microchannel.
Micromachines 09 00504 g001
Figure 2. Dimensionless velocity distributions u/us for different values of the fluid behavior index n under κH = 15. (a) Symmetric zeta potentials (ψt/ψb = 1). (b) Asymmetric zeta potentials (ψt/ψb = 1.5).
Figure 2. Dimensionless velocity distributions u/us for different values of the fluid behavior index n under κH = 15. (a) Symmetric zeta potentials (ψt/ψb = 1). (b) Asymmetric zeta potentials (ψt/ψb = 1.5).
Micromachines 09 00504 g002
Figure 3. Dimensionless velocity distributions u/us for different values of κH under ψt/ψb = 1.5. (a) Shear thinning fluid (n = 0.8). (b) Shear thickening fluid (n = 1.2).
Figure 3. Dimensionless velocity distributions u/us for different values of κH under ψt/ψb = 1.5. (a) Shear thinning fluid (n = 0.8). (b) Shear thickening fluid (n = 1.2).
Micromachines 09 00504 g003
Figure 4. Dimensionless velocity distributions u/us as a function of the dimensionless zeta potential difference R. The values −2/3, −1/3, 0, 1/3, and 2/3 of R, correspond to 0.2, 0.5, 1.0, 2.0, and 5.0 of the zeta potential ratio (ψt/ψb), respectively. (a) Shear thinning fluid (n = 0.8). (b) Shear thickening fluid (n = 1.2).
Figure 4. Dimensionless velocity distributions u/us as a function of the dimensionless zeta potential difference R. The values −2/3, −1/3, 0, 1/3, and 2/3 of R, correspond to 0.2, 0.5, 1.0, 2.0, and 5.0 of the zeta potential ratio (ψt/ψb), respectively. (a) Shear thinning fluid (n = 0.8). (b) Shear thickening fluid (n = 1.2).
Micromachines 09 00504 g004

Share and Cite

MDPI and ACS Style

Choi, D.-S.; Yun, S.; Choi, W. An Exact Solution for Power-Law Fluids in a Slit Microchannel with Different Zeta Potentials under Electroosmotic Forces. Micromachines 2018, 9, 504. https://doi.org/10.3390/mi9100504

AMA Style

Choi D-S, Yun S, Choi W. An Exact Solution for Power-Law Fluids in a Slit Microchannel with Different Zeta Potentials under Electroosmotic Forces. Micromachines. 2018; 9(10):504. https://doi.org/10.3390/mi9100504

Chicago/Turabian Style

Choi, Du-Soon, Sungchan Yun, and WooSeok Choi. 2018. "An Exact Solution for Power-Law Fluids in a Slit Microchannel with Different Zeta Potentials under Electroosmotic Forces" Micromachines 9, no. 10: 504. https://doi.org/10.3390/mi9100504

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop