You are on page 1of 5

week ending

PRL 110, 184503 (2013) PHYSICAL REVIEW LETTERS 3 MAY 2013

Nonlinear Amplification in Electrokinetic Pumping in Nanochannels in the Presence of


Hydrophobic Interactions
Suman Chakraborty,1,* Dipankar Chatterjee,2 and Chirodeep Bakli1
1
Department of Mechanical Engineering, Indian Institute of Technology, Kharagpur 721302, India
2
Simulation and Modeling Laboratory, CSIR-Central Mechanical Engineering Research Institute, Durgapur 713209, India
(Received 7 February 2012; revised manuscript received 27 February 2013; published 3 May 2013)
We discover a nonlinear coupling between the hydrophobicity of a charged substrate and electrokinetic
pumping in narrow fluidic confinements. Our analyses demonstrate that the effective electrokinetic
transport in nanochannels may get massively amplified over a regime of bare surface potentials and
may subsequently get attenuated beyond a threshold surface charging condition because of a complex
interplay between reduced hydrodynamic resistance on account of the spontaneous inception of a less
dense interfacial phase and ionic transport within the electrical double layer. We also show that the
essential physics delineated by our mesoscopic model, when expressed in terms of a simple mathematical
formula, agrees remarkably with that portrayed by molecular dynamics simulations. The nontrivial
characteristics of the initial increment followed by a decrement of the effective zeta potential with a
bare surface potential may open up the realm of hitherto-unexplored operating regimes of electro-
hydrodynamically actuated nanofluidic devices.

DOI: 10.1103/PhysRevLett.110.184503 PACS numbers: 47.61.k, 47.65.d

The physics of electrically modulated fluidic transport in and nonmonotonic amplifications in electrokinetic pump-
narrow confinements is central to the understanding of ing in nanofluidic channels, beyond the above regimes
several processes of practical relevance, with applications hitherto addressed? Unlike the concerned linear regime
ranging from the transport and separation of charged col- addressed in the previous studies, here we demonstrate
loids to the realization of miniaturized energy conversion that the same is indeed a possibility, although not a trivi-
devices and systems. Many of these challenging applica- ality. Our studies do reveal that the concerned relative
tions involve the formation of a charged interfacial layer in enhancement in the effective zeta potential, in general,
the vicinity of the solid substrates, also known as the may be characterized through a complex nonlinear depen-
electrical double layer (EDL) [1]. Under the action of an dence on the bare surface potential, in sharp contrast to the
external driving force, excess ionic charges present within classical linear regime. We arrive at this intriguing con-
this layer may be preferentially transported across the jecture by interpreting the concerned complex physical
channel. In certain cases, this pumping effect may get interactions over reduced length scales in a combined
further amplified by an effective reduction of the resistive phase field-lattice Boltzmann (PFLB) framework (which
forces in the interfacial layer, as attributed to the inception is different from standard LB models for electrokinetics
of a less dense phase [2–4]. The reduced interfacial friction, [6]), instead of following computationally more intensive
in turn, may amplify the rate of ionic pumping, leading to multiscale simulation routes [7]. Irrespective of the under-
the possibilities of realizing an apparently hyperfluidic lying complexities in the physical issues prevalent within
electrohydrodynamic transport over small scales. While the newly discovered nonlinear regime, we finally establish
related consequences for microscale devices [5] sound an apparently universal and generalized functional form
more trivial, several aspects of the corresponding implica- for the amplification in the effective zeta potential by using
tions over nanometer length scales remain unclear. phase-field thermodynamics [8]. Our prediction agrees
Considering the hydrodynamics within the EDL over remarkably with the effective zeta potential predicted
slipping surfaces, researchers have demonstrated that an directly from MD simulations, albeit in a computationally
effective interfacial slipping phenomenon may amplify efficient paradigm [9,10].
the extent of electrokinetic pumping [2] in nanochannels To begin with our PFLB model description, we first
to an extent, as manifested through augmentations in the consider generic distribution functions (f , fu , f c ) for
effective interfacial (zeta) potential, eff (which is linearly the phase fraction () evolution, velocity (u) field, and
related to the bare surface potential). While this kind of the electrical potential ( c ) distribution, respectively. Here,
linear response has been verified by extensive molecular we capture the phase fraction evolution on account of
dynamics (MD) simulations, its validity over a broader hydrophobic interactions through an order parameter,
regime has by far remained questionable. defined as  ¼ ðn1  n2 Þ=ðn1 þ n2 Þ, where n1 and n2 rep-
Can complex electrochemical-hydrodynamic interac- resent the number densities of the depleted and liquid
tions over the relevant physical scales give rise to nonlinear phases, respectively. In the absence of any external forcing

0031-9007=13=110(18)=184503(5) 184503-1 Ó 2013 American Physical Society


week ending
PRL 110, 184503 (2013) PHYSICAL REVIEW LETTERS 3 MAY 2013

parameters, these distribution functions may evolve accord- We set up a model problem which we first investigate
ing to the discretized LB equation as [11] fi ðxþci t;tþ through MD simulations. The problem essentially addresses
tÞfi ðx;tÞ¼ð1= Þ½fi ðx;tÞfi;eq ðx;tÞ, where  2 electrokinetic transport of a sodium chloride-water solution
of a given concentration through a nanofluidic channel of a
ð; u; c Þ, fi;eq ðx; tÞ are the local equilibrium distribution
given dimension as well as substrate wettability [14].
functions, and ci are the discrete lattice velocity vectors.
To implement the above problem from a PFLB perspec-
The distribution functions are related to the pertinent trans-
P P P tive, we choose the following parameter values [11]: B ¼
port parameters as  ¼ i fiu , u ¼ i fiu ci ,  ¼ i fi . A ¼ 0:003, k ¼ 0:01, ug ¼ 1, ul ¼ 0:8. This yields the
The corresponding equilibrium distribution functions may following
be derived by expanding those in terms of power series in qffiffiffiffiffiffifor the liquid-vapor surface tension coefficient:
the local velocity [12]. We take the parameters ci following lv ¼ 8kB 9 . Further, as measured in lattice units, lv 
lu

a standard D2Q9 model [11] with weighting functions for 0:105  0:002 [3]. One may thus obtain the lattice spacing
the nine lattice directions w0 ¼ 4=9, wi ¼ 1=9 for i 2 in physical units by relating lv with lu lv in the following
ð1; 4Þ and wi ¼ 1=36 for i 2 ð5; 8Þ, whereas we relate the manner [3]: lv ¼ kB T=x2 lu lv , so that x  0:3 nm for
pffiffifficonstant (x) and time step (t) as c ¼ x=t;
lattice T ¼ 300 K. This dimension is indeed comparable with the
c ¼ 3cs ; cs being the sonic speed. For implementing atomic ranges of potentials used in typical MD simula-
position-dependent body forces dictating the momentum tions. Further, we consider  c ¼ 1 [6].
transport [13], we add a forcing term to the right-hand side We first depict a representative variation of slip length
of the evolution equation of the velocity distribution (ls ) with a nondimensional bare surface potential ( ¼
u ¼ wi ð1  ð1=2 u 
equation as fsource P ÞÞ½3ðci  u Þ þ 9ðci  zeo =kB T). The complex variations in slip length, as por-
u Þci   FðxÞ, where u ¼ ð1=Þ½ i fiu ci þ ðF=2Þ. trayed in Fig. 1, essentially stem from the fact that an
Physically consistent prescription of FðxÞ, mentioned uncharged hydrophobic surface would not have favored
as above, turns out to be a critically important considera-
tion for implementing the present model, in consistency
with the electrochemical-thermodynamic interactions in
narrow confinements subjected to hydrophobic interac-
tions. Fundamentally, momentum transport due to electri-
cal effects is explicitly linked with an electrokinetic body
force, as
 
1 1 @"el
FEK ¼ e E  E  Er"el þ r E  E  ;
2 2 @
where E is the local electric field (¼ r c ), "el is the local
permittivity of the medium, and e is the electrical charge
density. This forcing term, thus, is directly linked with the
Poisson equation for electrical potential distribution. In
order to ensure that an unsteady term does not feature in
the governing equation for electrical potential distribution
[7], we append an additional source term in the right-hand
side of the discretized LB form of the evolution equation
for fic , in the form: tw i RD, where R is the right-hand
side of the Poisson equation (which cooperates the wall- FIG. 1 (color online). Variation of slip length, ls , with non-
adjacent density variations [2] in the description of the dimensional bare surface potential shown for a representative
concentration profiles, over and above the traditional case relating salt concentration to the surface charge by the ratio
Poisson-Boltzmann picture), D ¼ c2 ð c  ð1=2ÞÞðt=3Þ, Nþ =Nw ¼ 3:2 (Case 1) and Nþ =Nw ¼ 2:4 (Case 2), where Nþ is
the number of Naþ ions in an unit cell and Nw e is the amount of
and w i ¼ wi for i 2 ð1; 8Þ; w i ¼ ðw0  1Þ for i ¼ 0. The charge on a channel wall confining the unit cell. The channel
corresponding equilibrium distribution function is given by wettability is defined by a static contact angle of 110. The slip
f0c ;eq ¼ ðw=0  1Þf0c ; fic ;eq ¼ w=i fic (i  0), where w=0 ¼ lengths tend to gradually decrease with an increase of bare surface
1=9, w=i ¼ 1=8 (i  0). The resultant electrical potential potential, beyond a critical limit. Extremely large bare surface
P
distribution: c ¼ i fic =ð1  w=0 Þ, satisfies the Poisson
potential causes an abrupt drop in slip length, making the slip
independent of salt concentration, wall charge, or further rise in
equation, which may be recovered by executing a zeta potential. Markers represent MD simulation data, and solid
Chapman-Enskog expansion. We augment the interaction lines represent corresponding PFLB model predictions. Several
forces in the PFLB model further to address some addi- other values of Nþ =Nw are studied, resulting no qualitative
tional facets of electromechanical interactions over changes in the trends observed (not reported in the figure, for
reduced scales [14,15]. the sake of clarity).

184503-2
week ending
PRL 110, 184503 (2013) PHYSICAL REVIEW LETTERS 3 MAY 2013

accumulation of water in the near-surface region and the surface potential. Interestingly, beyond a threshold limiting
same would get depleted having a lower than bulk density value of the surface potential, this additional driving force
[16]. However, introduction of a surface potential favors tends to saturate. Since an increase (or decrease) in this
increments in the local number density of water at the interfacial driving force may be interpreted to result in an
solvent-substrate interface, beyond a threshold limit of increase (or decrease) in the extent of the effective slip,
the surface potential. It is important to mention in this this clearly explains the observed trends in slip length,
context that despite an inherently mesoscopic character purely from mesoscopic simulation considerations. These
of the PFLB model employed here, it implicitly takes trends may further be justified from the interfacial density
into account some subgrid scale physical considerations variations [14].
depicting certain facets of substrate-solvent-solute interac- How do the above variations in slip length translate in
tions, through the description of an augmented free energy the form of an effective zeta potential? Towards answering
functional. Importantly, these considerations include the this question, we first obtain the effective zeta potential
solvation (hydration) interactions between the ionic con- directly from the MD simulations by calculating the ionic
stituents [17], resulting in a modification in the description current as [2] Ie ¼ ðð"d eff Þ=ÞAf0 , where  is the shear
of the free energy. In our PFLB model, the corresponding viscosity, A is the fluid slab cross area, and f0 is the force
effective chemical potential (obtained from the variation per unit volume applied in streaming current experiment.
in the additional component of the free energy with respect The consequent variation is essentially plotted in Fig. 2.
to the order parameter), eff , eventually gives rise to an From Fig. 2, it is apparent that there occurs an initial
amplification in eff with increments in ,  followed by a
additional effective forcing function: Feff ¼ eff r in
the momentum equation (where  is the order parameter dip in the same. This kind of trend appears to be a con-
field), so as to address the corresponding physical impli- sequence of complicated nonlinear interactions between
cations in a mesoscopic sense. the potential profile and the slip length, both being sensi-
From Fig. 1 it is evident that extremely low surface tively dependent on the ionic concentration, as well as the
potentials do not seem to induce any variation in the slip bare surface potential, for a given substrate wettability.
We next formulate a simple yet insightful continuum-
length. As the potential is increased, slip lengths gradually
 slip lengths based semianalytical platform. Towards that, we first reit-
decrease. Beyond a threshold increase in ,
erate that intricate interfacial interactions over small scales
start decreasing abruptly. These avalanches in the slip
may alter the near-wall arrangement of water molecules.
length values are observed for all values of salt concen-
This, in turn, modifies the Boltzmann distribution of
trations and occur almost at the same bare surface potential
ions by appending an extra potential term as n ¼
value for different values of ionic concentration [18].
no expððze c  c ext Þ=kB TÞ, where c is electrostatic
Finally, at very high surface charge densities, the slip
lengths are observed not only to become insensitive to
salt concentration but also to a further rise of wall charge
density. A further increase of surface potential, in effect,
leads to overcrowding of the surface with counterions and
slip length variation tends to plateau. In effect, with a large
counterion population guided by the high bare surface
potential, the near-wall fluid-ion layering attains saturation
and a further increase in surface potential does not induce
any further appreciable change in the interfacial density
distribution. The high electrostatic force for a charged
surface, under such circumstances, overrules the solvation
forces (which could otherwise have introduced a concen-
tration dependency), leading to slip lengths being indepen-
dent of the solution concentration. As is evident from Fig. 1,
very good agreements may indeed be observed between the
PFLB simulation predictions and MD predictions.
In order to delve deeper into the underlying mesoscopic
physics addressed through our PFLB simulations, we have
clearly revealed that below a critical magnitude of the
FIG. 2 (color online). Comparison between the effective zeta
surface potential, the value of Feff due to additional subgrid potential predictions as obtained from MD simulations and the
scale interactions implicitly accounted for by the PFLB same obtained from semianalytical (MPB) theory in conjunction
model does not vary appreciably, leading to nearly invari- with PFLB simulation database for slip length as well as water
ant slip lengths. However, beyond this critical limit, the density fluctuations. Simulation data are the same as those
value of Feff decreases with an increase in the value of the considered for plotting Fig. 1.

184503-3
week ending
PRL 110, 184503 (2013) PHYSICAL REVIEW LETTERS 3 MAY 2013

potential, c ext ¼ kB T logðf ðyÞ=fb Þ [2], fb being the hydrodynamic influences, as manifested through the inter-
bulk density of water and f is the water density variation facial physics delineated above.
along the y (transverse) direction. no , nþ , and n are the How can we explain the nontrivial physics associated
bulk ion density, cation density, and anion density, respec- with the nonmonotonic trends in the effective interfacial
tively. Thus, we obtain a modified Poisson-Boltzmann potential? Towards addressing this question, it is first
(MPB) equation, for a z:z symmetric electrolyte, as imperative to note that the same is essentially decided by
    an intricate interplay between molecular slip and the den-
d dc f ðyÞ ze c sity fluctuations over interfacial scales. As shown in Fig. 1,
"d ¼ 2zeno sinh :
dz dy fb kB T the slip lengths remain virtually unaltered for low bare
surface potentials, and subsequently tend to decrease
Traditionally, the ratio ðf ðyÞÞ=fb is assumed to be unity with increments in the same till a ‘‘sticking’’ condition is
[2], which very well reproduces the simulation results for attained. On the other hand, the interfacial density fluctua-
low values of zeta potential (typically, within Debye- tions, as appearing in the MPB equation, first increase with
Hückel limits). However, here we assess the specific increasing surface charge, until a threshold charging con-
contribution from this term by appealing to the PFLB dition is reached [14]. Below this threshold condition incre-
simulation predictions that offer predictions to the near- ments in surface charge induce more fluctuations in the
wall density profiles. Based on this variation, we obtain the water density profile, which in turn mobilizes the layering
gradient of c , numerically, by integrating the modified PB of the ions adjacent to the interface and generates greater
equation described as above. Next, we integrate the current in the flow. However, beyond the threshold limit, the
Navier-Stokes equation with neglected acceleration terms fluctuations are reduced owing to the overcrowding of the
and in presence of an applied electric field, Ex , along the channel with salt ions. The resultant counteracting effects of
channel axis: l ðd2 u=dy2 Þ"l ðd2 c =dy2 ÞEx ¼0. Applying slip and density profiles essentially give rise to the effective
the partial slip boundary condition as obtained from com- pumping characteristics as depicted in Fig. 2.
prehensive numerical simulations, a solution to the above To summarize, we have discovered a hitherto-
equation (with centerline symmetry) yields: unexplored nonlinear regime of massive amplification in
  
c l dc   electrokinetic pumping in nanochannels. Our studies have
" E 

u¼ l x 1  s  : demonstrated that despite the underlying complexities, the
l   dy  y¼0 corresponding amplification factor may be expressed in the
form of a universal nonlinear functional dependence, in
Noting that along the center line, uc ¼ ðð"l eff Ex Þ=l Þ, it
which the slip lengths as well as the EDL potential gra-
follows that
dients may be obtained successfully through a mesoscopic
  
eff c c ls d c  

 formalism. This, in turn, provides an effective tool for the
¼ 1   ;
   dy y¼0 design and synthesis of nanofluidic devices with enhanced
capabilities, and possibly leading towards the practical
where y ¼ 0 represents a transverse location where the realizability of hypertransportive electrohydrodynamic
partial slip boundary condition is applied. This simple systems over nanoscopic scales.
estimate of effective amplification in electrokinetic pump-
ing, with ls data obtained through the PFLB route,
compares excellently with the MD based simulation
predictions, as evident from Fig. 2. The implications of *suman@mech.iitkgp.ernet.in
this elusively obvious agreement, however, are not far from [1] R. J. Hunter and L. R. White, Foundations of Colloid
being nontrivial. This may be attributed to the fact that ls Science (Clarendon, Oxford, 1989).
itself is a parameter that is sensitively dependent on , [2] L. Joly, C. Ybert, E. Trizac, and L. Bocquet, Phys. Rev.
ionic concentration, and substrate wettability. We have Lett. 93, 257805 (2004); J. Chakraborty, S. Pati,
already established that this dependence may be success- S. K. Som, and S. Chakraborty, Phys. Rev. E 85, 046305
fully captured through PFLB formalism (see Fig. 1). Once (2012); C. Bakli and S. Chakraborty, J. Chem. Phys. 138,
a substitution of this ls dependence in Eq. (1) has been 054504 (2013); A. Bandopadhyay and S. Chakraborty,
made, the data corresponding to a wide range of combina- Langmuir 28, 17 552 (2012).
tions of substrate wettability, ionic concentration, and bare [3] M. Sbragaglia, R. Benzi, L. Biferale, S. Succi, and
F. Toschi, Phys. Rev. Lett. 97, 204503 (2006).
surface potential are found to remarkably converge to
[4] A. Ajdari and L. Bocquet, Phys. Rev. Lett. 96, 186102
identical characteristics of eff =, in line with the corre- (2006); E. Lauga, M. P. Brenner, and H. A. Stone,
sponding MD simulation predictions. Such universal Handbook of Experimental Fluid Dynamics (Springer,
trends are also physically reminiscent of the fact that New York, 2007).
electrokinetic pumping in the presence of hydrophobic [5] V. Tandon and B. J. Kirby, Electrophoresis 29, 1102
interactions is a generic nonlinear function of the intrinsic (2008).
interactions between the underlying electromechanical and [6] Z. Chai and B. Shi, Phys. Lett. A 364, 183 (2007).

184503-4
week ending
PRL 110, 184503 (2013) PHYSICAL REVIEW LETTERS 3 MAY 2013

[7] R. Qiao and N. R. Aluru, Int. J. Multiscale Comput. Eng. [11] J. Horbach and S. Succi, Phys. Rev. Lett. 96, 224503 (2006).
2, 173 (2004). [12] A. J. Briant and J. M. Yeomans, Phys. Rev. E 69, 031603
[8] A. Fakhari and M. H. Rahimian, Phys. Rev. E 81, 036707 (2004).
(2010). [13] Z. Guo, C. Zheng, and B. Shi, Phys. Rev. E 65, 046308
[9] Despite incurring fewer approximations than other meth- (2002).
ods, MD simulations on electrokinetic transport are re- [14] See Supplemental Material at http://link.aps.org/
strained by the following challenges [10]: (i) if the ionic supplemental/10.1103/PhysRevLett.110.184503 for addi-
concentration is in order of magnitude lower than the tional discussion on electro-mechanical interactions, and
solvent concentration then statistical accuracy is achieved details of MD simulations.
only at the expense of long simulation runs, and [15] D. M. Huang, C. Cottin-Bizonne, C. Ybert, and L.
(ii) typical electrokinetic speeds are much slower than Bocquet, Phys. Rev. Lett. 98, 177801 (2007); D. M.
the characteristic thermal speed, necessitating additional Huang, P. L. Geissler, and D. Chandler, J. Phys. Chem.
care to extract the signal amidst the background noise. B 105, 6704 (2001); L. Onsager and N. N. T. Samaras,
These considerations, in addition to favorable computa- J. Chem. Phys. 2, 528 (1934).
tional costs, impart the motivation for developing a PFLB [16] S. Chakraborty, Phys. Rev. Lett. 99, 094504 (2007); 100,
model that can be closely sensitive to MD simulation 097801 (2008); 101, 184501 (2008); D. R. Berard, P.
data. As a quantified measure, a system containing 7176 Attard, and G. N. Patey, J. Chem. Phys. 98, 7236 (1993).
water molecules simulated for 5 ns takes 100 physical [17] Y. Burak and D. Andelman, Phys. Rev. E 62, 5296 (2000);
hours on a 3.4 GHz Core i7-2600 processor for simulating J. Chem. Phys. 114, 3271 (2001).
the MD results presented here, whereas the corresponding [18] Here we express the salt concentration in terms of the
computational time is close to 10 physical hours for the channel height to Debye length ratio. The Debye length is
PFLB simulation. an equivalent representation
P of the ionic concentration,
[10] U. M. B. Marconi and S. Melchionna, Langmuir 28, 13727 given by D ¼ ð"l kB T= N n q2 Þ0:5 ; n being the con-
j¼1 j j j
(2012). centration of species j with charge qj .

184503-5

You might also like