scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Steric-effect-induced enhancement of electrical-double-layer overlapping phenomena

20 Jul 2011-Physical Review E (American Physical Society)-Vol. 84, Iss: 1, pp 012501-012501
TL;DR: It is demonstrated that nontrivial interactions between steric effect and electrical-double-layer (EDL) overlap phenomena may augment the effective extent of EDL overlap in narrow fluidic confinements to a significant extent by virtue of rendering the channel centerline potential tending to the ζ potential in a limiting sense as the stericEffect progressively intensifies.
Abstract: In this paper, we demonstrate that nontrivial interactions between steric effect and electrical-double-layer (EDL) overlap phenomena may augment the effective extent of EDL overlap in narrow fluidic confinements to a significant extent by virtue of rendering the channel centerline potential tending to the ζ potential in a limiting sense as the steric effect progressively intensifies. Such a behavior may result in a virtually uniform (undiminished) magnitude of the EDL potential across the entire channel height and may cause lowering of the total charge within the EDL.

Content maybe subject to copyright    Report

PHYSICAL REVIEW E 84, 012501 (2011)
Steric-effect-induced enhancement of electrical-double-layer overlapping phenomena
Siddhartha Das
1
and Suman Chakraborty
2
1
Physics of Fluids Group, J. M. Burgers Centre for Fluid Dynamics, University of Twente, P.O. Box 217,
7500 AE Enschede, The Netherlands
2
Department of Mechanical Engineering, Indian Institute of Technology, Kharagpur, Kharagpur-721302, India
(Received 3 March 2011; revised manuscript received 18 May 2011; published 20 July 2011)
In this paper, we demonstrate that nontrivial interactions between steric effect and electrical-double-layer
(EDL) overlap phenomena may augment the effective extent of EDL overlap in narrow fluidic confinements
to a significant extent by virtue of rendering the channel centerline potential tending to the ζ potential in a
limiting sense as the steric effect progressively intensifies. Such a behavior may result in a virtually uniform
(undiminished) magnitude of the EDL potential across the entire channel height and may cause lowering of the
total charge within the EDL.
DOI: 10.1103/PhysRevE.84.012501 PACS number(s): 82.45.Gj
Recent advances toward understanding nanoscale elec-
trochemical hydrodynamics have led to the development of
several widely applicable miniaturized devices and systems
tuned for practical applications [18]. The intricate dynamics
within the surface-bound double layer of charge, popularly
known as the electric-double layer (EDL) [9], often dictates
the functionalities of these systems to an extremely significant
extent. Theoretical electromechanical analysis of the EDL, on
the other hand, has attracted great interest for close to half
a century [1012], and well known solutions exist for EDL
charge distribution in a fluidic confinement, obtained under
classical approximations, such as mean field assumption,
assumption of nonoverlapping EDLs, assumption of EDL ions
as point charges, etc.
A growing need to understand the electromechanics of
systems with overlapping EDLs, and/or with very high
ionic concentration, and/or with very large ζ potential, has
necessitated the underlying analysis to be extended beyond
the above-mentioned simplified assumptions, even within the
framework of mean field approximations. For example, several
recent investigations demonstrate remarkable influences of
EDL overlap on nanoscale electrokinetic transport [1321],
and there has been considerable effort in the theoretical
modeling of the concerned phenomena [2233], obeying
the governing principle of net charge electroneutrality of an
isolated channel.
Interestingly, independent of the considerations of EDL
overlap, there has been a continuous endeavor toward relaxing
the treatment of ionic species as point charges by aptly taking
the effects of their finite sizes into account. Such finite size
effects of ions have been shown to have tremendous influence
on different applications of nanoelectrokinetic phenomena
[3442]. Stern, in his celebrated work [43], introduced cor-
rections to the Poisson-Boltzmann (PB) equation to account
for the unphysical divergences of the Gouy-Chapman model
of the double layer and indicated the volume constraints
of ions in the electrolyte phase [43,44]. However, the first
complete ion-size-effect-induced modified Poisson Boltzmann
(mPB) model was developed by Bikerman [45]. Over the next
several years, Bikerman’s mPB equation was reformulated
by Grimley and Mott [46], Grimley [47], Dutta and Bagchi
[48], Bagchi [49,50], Dutta and Sengupta [51], Wicke and
Eigen [5254], Wiegel and Strating [55], Strating and Wiegel
[56,57], and Bohinc and co-workers [58,59]. For a flat double
layer, Kilic et al.[60] used the lattice approach to obtain
Bikerman’s mPB equation with a simple composite diffuse
layer model. Issues delineating the concerned consequences
on the resultant diffuse layer capacitances were also addressed
by some other researchers [44,60,61], in contrast with the
predictions obtained using the Gouy-Chapman theory [62].
A significant step forward in the development of the effect
of ion size in PB description was achieved by the use of the
hard-sphere (HS) model for the ions, which Biesheuvel and
van Soestbergen, in their pioneering paper [63], showed to be
capable of addressing several inconsistencies and limitations
in the lattice model. Their approach is distinct from the
earlier approaches in the sense that, while Bikerman [45]
assumed the local available volume to be the total volume
minus the volume of all ions plus their hydration shell, and
Sparnaay [64] replaced the volume of the hydrated ions
by the excluded volume, Biesheuvel and van Soestbergen
empirically included an extended Carnahan-Starling equation
of state to describe HS interactions in EDLs containing ions
of different sizes and charges. In this way, they attempted
to overcome certain important limitations of the traditional
lattice based models, which invariably underestimated the
excluded volume effect as well as erroneously considered
the size of a lattice as equivalent to the size of a solvent
molecule and not the hydrated site. Their results were in
excellent agreement with experimental findings [65], and
their model was successfully extended to study the case of
mixture of ions of different sizes [6669]. However, none of
the above studies have reported any analysis based on the
simultaneous considerations of steric effects and EDL overlap
phenomenon.
Here, we demonstrate that, by simultaneous considerations
of finite ion-size effect and EDL overlap, one can remarkably
have a completely geometric phenomenon (i.e., the EDL
overlap induced by the physical presence of a confinement
having a length scale below a threshold limit) influenced by a
completely thermoelectrochemical transport phenomenon as
governed by the ionic-size effects. In this perspective, it is
important to pinpoint the distinctive fundamental contributions
from our Brief Report as compared to some statistical mechan-
ics based investigations (postulated on the density functional
formalism) of Urbanija et al. [70] and Perutkova et al. [71] in
012501-1
1539-3755/2011/84(1)/012501(4) ©2011 American Physical Society

BRIEF REPORTS PHYSICAL REVIEW E 84, 012501 (2011)
which they have considered finite ion sizes and confinement
simultaneously. In their studies, the ions considered are giant
macroions that are randomly placed inside a confined channel.
Minimization of the free energy of the system results in govern-
ing integrodifferential equations for the potential distribution
within the confinement. Such an approach, however, may be
intrinsically complex in addressing complicated scenarios, and
in many cases, reduces analytical tractability so that nontrivial
numerical treatment may be necessary to address even a simple
problem [60]. Moreover, their approach never considers the
condition of EDL overlap (realized when the confinement
size goes below a threshold limit in relation to the Debye
length so that a nonzero centerline potential is realized that
cannot be prescribed apriori). In this context, it is important
to mention that mere consideration of a confined geometry
does not implicate the consideration of an EDL overlap
phenomenon, since, if the characteristic length scale is beyond
a critical limit (typically on the order of the Debye length),
then far-stream boundary conditions on the electrostatic
potential can effectively be employed at the channel centerline,
rendering the situation qualitatively identical to the one with
a single charged plate, despite the physical existence of a
confinement.
The mathematical model employed in the present Brief
Report follows from the modified space charge model de-
veloped by Cervera and co-researchers [72,73] in which the
ionic-size effect is explicitly incorporated in the ionic flux
density expression in the governing species transport equation
as follows:
J
i
=un
i
D
i
n
i
(ln a
i
+
z
i
RT
), where u is the
advective velocity, F is the Faraday constant, z
i
is the valency
of the ith ionic species, T is the absolute temperature, D
i
is
the diffusivity of the ith ionic species, n
i
is the number density
of the ith ionic species, φ is the resultant electrical potential,
and a
i
is the activity of the ith ionic species and is expressed
in terms of the ionic concentrations n
i
as a
i
=
n
i
/n
r
1v
k
n
k
/n
r
.
Here, n
r
is the reference ionic number density, and v is the
partial molal volume of the ionic species (may alternatively be
termed the steric factor). This model is applied to a geometrical
confinement in which the characteristic EDL thickness is
specifically chosen to be larger than the channel half height.
The resulting EDL overlap phenomenon is assessed in terms
of the difference between the centerline potential and the
ζ potential, as a combined consequence of electrochemical,
thermodynamic, and a pure geometrical artifact of the system
configuration. In fact, the channel centerline potential is not
aprioriknown in the present context, which needs to be
calculated in consistency with the ζ potential and the potential
gradient developed within the geometrical confinement as a
function of the partial molal volume of the ionic species. Thus,
the channel centerline potential derived in a self-consistent
fashion, effectively represents the extent EDL overlap. The
central result of our calculation is that the finite ion-size
effects may remarkably enhance the effective extent of EDL
overlap, quantitatively delineated by the proximity of the
channel centerline potential with the ζ potential, as compared
to the cases without involving finite ion-size effects. The
resulting effective extent of the EDL overlap, under critical
circumstances, may be so significantly enhanced that one may
recover the ζ potential virtually undiminished across the entire
channel height. This implies that one can completely ignore
the EDL and can end up, quite dramatically, with an ideally
homogeneously charged nanofluidic system.
For the present Brief Report, we consider a nanoflu-
idic system (essentially a slit type nanochannel connected
with its end reservoirs) in which the steady state transport
phenomenon of each ionic species may be represented by
the following [72,73]:
∇·
J
i
= 0. In the absence of the
bulk advective transport of ions, the same consideration
implicates
k
B
T ln
(
a
i
)
+ ez
i
ψ = constant, (1)
where ψ is the EDL potential, e is the protonic charge, and k
B
is
the Boltzmann constant. Considering equilibrium between the
nanochannel and its connecting reservoirs (where n
+
= n
=
n
and ψ = 0) and assuming a z : z symmetric electrolyte for
simplicity, one gets
n
±
=
n
1 +
2vn
n
r
cosh
ezψ
k
B
T
1
exp
ezψ
k
B
T
. (2)
Substituting the above in the Poisson equation
d
2
ψ
dy
2
=
ρ
e
ε
=
ez(n
+
n
)
ε
(where ε is the permittivity of the ionic
medium and y is the channel wall-normal coordinate) and
integrating the same from the Stern layer-diffuse layer inter-
face to the channel centerline (where
dy
= 0 and ψ = ψ
c
=
0 under overlapped EDL conditions), one gets
dy
ψ
r
λ
v
ln
1 +
cosh
ψ
ψ
r
1
1 + 2v
cosh
ψ
c
ψ
r
1
, (3)
where λ is the Debye length (
1
λ
2
=
2n
e
2
z
2
εk
B
T
) and ψ
r
=
k
B
T
ez
is a
reference potential.
In an effort to obtain the variations in ψ,weemployan
iterative solution technique, where we start with a guessed
value of ψ
c
and iteratively obtain a corrected value of the
same by using the EDL potential gradient as given by Eq. (3)
and a specified ζ potential, until convergence is achieved. It is
important to mention here that instead of a given ζ potential
boundary condition, one could as well employ a given surface
charge density boundary condition or, more appropriately, a
chemical equilibrium based interfacial treatment [74]. While
the later ones effectively include more parameters into the
iterative framework mentioned as above, the same does not
introduce any essential new physics in the context of the
physical issue addressed in this Brief Report. Accordingly, the
specified ζ potential boundary condition is used in the present
simulations for simplicity, without sacrificing the essential
physics.
The central result of this Brief Report is the variation in
the ratio ψ
c
/ζ with the partial molal volume or the effective
steric factor ν, for different λ/H ratios (see Fig. 1). The
effective extent of EDL overlap is essentially governed by
the parameter ψ
c
/ζ (the closer its value to unity, the more
intense the EDL overlap). From Fig. 1, it is evident that an
increase in ν enhances ψ
c
/ζ , implicating a greater extent
012501-2

BRIEF REPORTS PHYSICAL REVIEW E 84, 012501 (2011)
10
−2
10
−1
10
0
0
0.2
0.4
0.6
0.8
1
ν
ψ
c
/ζ
−1 −0.8 −0.6 −0.4 −0.2 0
0
1
2
3
x 10
y/H
dψ/dy (V/m)
ν = 1
λ /H = 2
λ /H = 1
λ /H = 3/4
λ /H = 1/4
ν = 1/100
ν = 1/10
FIG. 1. (Color online) Variation of the ratio ψ
c
/ζ with steric factor
ν for different λ/H ratios. In the inset of the figure, we show the
transverse variation of dψ/dy for the channel bottom half for λ/H =
1 for different values of steric factor ν. Results are provided for H =
5nmandζ =−100 mV.
of EDL overlap (theoretically, the maximum extent of EDL
overlap occurs when ψ
c
/ζ 1). This enhancement is most
prominent for channels where the EDL overlap phenomenon
can be predicted even in the absence of steric effects. For such
systems, realizable values of the steric factor ensure that a state
is reached where ψ
c
/ζ 1. Such a prominent influence of the
steric effect in magnifying the extent of EDL overlap, however,
is not witnessed for channels that have nonoverlapping EDLs
in the absence of steric effects. To understand such a role of the
steric effect, we study the variation of the transverse gradient
of the EDL (see the inset of Fig. 1)asafunctionofν. Larger
ν leads to a weaker EDL potential gradient [this is clearly
noticeable from the analytical expression in Eq. (3)], causing
a much weaker decay (with height) of the wall conditions
and making it penetrate much deeper into the nanochannel
bulk. Consequently, the channel centerline potential has a
value significantly closer to the ζ potential, which, in effect,
signifies a more enhanced EDL overlap. The manifestation
of pronounced influence of the steric effect, based on the
steric-effect-independent extent of EDL interactions, can be
argued from steric-effect-driven weaker crowding of ions in
the EDL. For an overlapped EDL even in the absence of steric
effects, the EDL counterions are in a spars condition so that
the average attractive pull of the charged wall is substantially
weak. For such cases, the contribution of the entropic mixing
of the ions (caused by their finite sizes, s ee Ref. [60])
can easily overcome the electrostatic attractive pull of the
wall, ensuring a homogeneous mixing and, hence, a uniform
distribution (rather than a wall-potential-dictated distribution)
of the ions across the channel height. This effectively leads
to a significantly enhanced value of the channel centerline
potential, implying a high effective degree of EDL overlap.
On the contrary, for an initially nonoverlapping system (even
for finitely large EDL thicknesses), the ions are always
−1 −0.8 −0.6 −0.4 −0.2 0
0.4
0.6
0.8
1
y/H
ψ/ζ
−1 −0.8 −0.6 −0.4 −0.2 0
10
10
10
10
y/H
n /n , n /n
n /n ,
ν=1/100
n
/n ,
ν=1/10
ν = 1/10
ν = 1
ν = 1/100
n /n ,
ν=1/100
n
/n ,
ν=1/10
FIG. 2. (Color online) Transverse variation of the EDL potential
(ψ) for the channel bottom half, corresponding to λ/H = 1for
different values of the steric factor ν. In the inset of the figure, we
show the corresponding transverse variation of the number density of
the counterions (n
+
/n
) and coions (n
/n
). Results are provided for
H = 5nmandζ =−100 mV.
tightly bound within the attractive range of the channel wall
so that the entropic considerations can never lead to uniform
distribution of the ions across the channel. Hence, for such
cases, only minor enhancement of the extent of EDL overlap
can be noted.
The consequences of such a steric-effect-induced en-
hancement of EDL overlap are r emarkable. For example,
the EDL potential, under such conditions, remains virtually
undiminished across the entire channel (see Fig. 2). This
implies that it is possible to achieve a uniformly charged
nanochannel where the behavior at any location in the conduit
is exactly identical to that interfacing the wall. Implicitly, this
signifies that the role of t he added ions toward dictating how
the bulk sees and how the EDL screens the wall ceases to be
important. This can straightaway be related to the fact that there
is a significant lowering of the number of ions in the EDL, as a
consequence of a lowering of the interfacial potential gradient
(see t he inset of Fig. 2).
To summarize, we have shown that the steric effect can
substantially enhance the extent of effective EDL overlap in
narrow confinements. These enhancements are most promi-
nent for channels that have some degree of steric-effect-
independent EDL overlap (i.e., EDL overlap under hypotheti-
cal conditions of the ionic species taken as point charges). Such
a phenomenon may turn out to be immensely significant, since
it leads to the paradigm of a homogeneously charged nanoflu-
idic system, where the concept of strong near-wall gradients
due to EDL phenomena virtually appears to be nonexistent.
This, in turn, may affect several electrokinetic phenomena that
depend on the extent of EDL interactions in nanochannels,
such as streaming potential and energy transfer effects in
narrow fluidic confinements [75 80], confinement-influenced
macromolecular separation [8183], and many others.
[1] A. Watson, Science 289, 850 (2000).
[2] S. K. Griffiths and R. H. Nilson, Anal. Chem. 78, 8134 (2006).
[3] F. H. J. van der Heyden, D. Stein, K. Besteman, S. G. Lemay,
and C. Dekker, Phys.Rev.Lett.96, 224502 (2006).
[4] L. Bocquet and E. Charlaix, Chem.Soc.Rev.39, 1073
(2010).
[5] M. Napoli, J. C. T. Eijkel, and S. Pennathur, Lab Chip 10, 957
(2010).
012501-3

BRIEF REPORTS PHYSICAL REVIEW E 84, 012501 (2011)
[6] W. Sparreboom, A. van den Berg, and J. C. T. Eijkel, Nat.
Nanotechnol. 4, 713 (2009).
[7] W. Sparreboom, A. van den Berg, and J. C. T. Eijkel, New J.
Phys. 12, 015004 (2010).
[8] P. Abgrall and N. T. Nguyen, Anal. Chem. 80, 2326 (2008).
[9] R. J. Hunter, Zeta Potential in Colloid Science: Principles and
Applications (Academic, New York, 1981).
[10] D. Burgreen, and F. R. Nakache, J. Phys. Chem. 68, 1084 (1964).
[11] C. L. Rice and R. Whitehead, J. Phys. Chem. 69, 4017 (1965).
[12] S. Levine, J. R. Marriott, G. Neale, and N. Epstein, J. Colloid
Interface Sci. 52, 136 (1975).
[13] H. Ban, B. Lin, and Z. Song, Biomicrofluidics 4, 014104 (2010).
[14] R. Chein, C. Liao, and H. Chen, Nanoscale Microscale
Thermophys. Eng. 14, 75 (2010).
[15] J. Jamaati, H. Niazmand, and M. Renksizbulut, Int. J. Therm.
Sci. 49, 1165 (2010).
[16] X. Wang and W. Hong, Phys. Rev. E 81, 041803 (2010).
[17] E. C. Yusko, R. An, and M. Mayer, ACS Nano 4, 477 (2010).
[18] M. L. Kovarik, K. Zhou, and S. C. Jacobson, J. Phys. Chem.
113, 15960 (2009).
[19] R. Chein, H. Chen, and C. Liao, J. Electroanal. Chem. 1, 630
(2009).
[20] S. Chakraborty and A. K. Srivastava, Langmuir 23, 12421
(2007).
[21] S. Talapatra and S. Chakraborty, Eur. J. Mech. B 27, 297 (2008).
[22] W. Qu and D. Li, J. Colloid Interface Sci. 224, 397 (2000).
[23] C. L. Ren, Y. D. Hu, D. Q. Li, and C. Werner, J. Adhes. 80, 831
(2004).
[24] A. T. Conlisk, J. McFerran, Z. Zheng, and D. Hansford, Anal.
Chem. 74, 2139 (2002).
[25] Z. Zheng, D. J. Hansford, and A. T. Conlisk, Electrophoresis 24,
3006 (2003).
[26] S. Bhattacharyya, Z. Zheng, and A. T. Conlisk, J. Fluid Mech.
540, 247 (2005).
[27] D. Stein, M. Kruithof, and C. Dekker, Phys.Rev.Lett.93,
035901 (2004).
[28] E. H. Kwak and E. F. Hasselbrink,
J. Colloid Interface Sci. 284,
753 (2005).
[29] S. Kang and Y. K. Suh, J. Mech. Sci. Technol. 20, 2250
(2006).
[30] R. B. Schoch and P. Renaud, Appl. Phys. Lett. 86, 253111
(2005).
[31] R. B. Schoch, H. van Lintel, and P. Renaud, Phys. Fluids 17,
100604 (2005).
[32] F. Tessier and G. W. Slater, Electrophoresis 27, 686 (2006).
[33] F. Baldessari and J. G. Santiago, J. Colloid Interface Sci. 325,
526 (2008).
[34] A. Garai and S. Chakraborty, Electrophoresis 31, 843 (2010).
[35] C. K. Hua, I. S. Kang, K. H. Kang, and H. A. Stone, Phys. Rev.
E 81, 036314 (2010).
[36] H. J. Zhao, J. Phys. Chem. C 114, 8389 (2010).
[37] A. S. Khair and T. M. Squires, J. Fluid Mech. 343, 640 (2009).
[38] B. D. Storey, L. R. Edwards, M. S. Kilic, and M. Z. Bazant,
Phys.Rev.E77, 036317 (2008).
[39] L. H. Olesen, M. Z. Bazant, and H. Bruus, Phys. Rev. E 82,
011501 (2010).
[40] M. Z. Bazant, M. S. Kilic, B. D. Storey, and A. Ajdari, Adv.
Colloid Interface Sci. 152, 48 (2009).
[41] M. Z Bazant, M. S. Kilic, B. D. Storey, and A. Ajdari, New J.
Phys. 11, 075016 (2009).
[42] M. Z. Bazant, B. D. Storey, and A. A. Kornyshev, Phys. Rev.
Lett. 106, 046102 (2011).
[43] O. Stern, Z. Elektrochem. 30, 508 (1924).
[44] V. Freise, Z. Elektrochem. 56, 822 (1952).
[45] J. J. Bikerman, Philos. Mag. 33, 384 (1942).
[46] T. B. Grimley and N. F. Mott, Discuss. Faraday Soc. 1, 3 (1947).
[47] T. B. Grimley, Proc. R. Soc. London, Sec. A 201, 40 (1950).
[48] M. Dutta and S. N. Bagchi, Indian J. Phys. 24, 61 (1950).
[49] S. N. Bagchi, Indian J. Chem. Soc. 27, 199 (1950).
[50] S. N. Bagchi, Indian J. Chem. Soc. 27, 204 (1950).
[51] M. Dutta and M. Sengupta, Proc. Natl. Inst. Sci. India 20,1
(1954).
[52] E. Wicke and M. Eigen, Naturwiss. 38
, 453 (1951).
[53] E. Wicke and M. Eigen, Z. Elektrochem. 56, 551 (1952).
[54] E. Wicke and M. Eigen, J. Phys. Chem. 58, 702 (1954).
[55] F. W. Wiegel and P. Strating, Mod. Phys. Lett. B 7, 483 (1993).
[56] P. Strating and F. W. Wiegel, J. Phys. A 26, 3383 (1993).
[57] P. Strating and F. W. Wiegel, Physica A 193, 413 (1993).
[58] K. Bohinc, V. Kralj-Iglic, and A. Iglic, Electrochim. Acta 46,
3033 (2001).
[59] K. Bohinc, A. Iglic, T. Slivnik, and V. Kralj-Iglic,
Bioelectrochemistry 57, 73 (2002).
[60] M. S. Kilic, M. Z. Bazant, and A. Ajdari, Phys.Rev.E75,
021502 (2007).
[61] A. A. Kornyshev, J. Phys. Chem. B 111, 5545 (2007).
[62] D. L. Chapman, Philos. Mag. 25, 475 (1913).
[63] P. M. Biesheuvel and M. van Soestbergen, J. Colloid Interface
Sci. 316, 490 (2007).
[64] M. J. Sparnaay, Recl. Trav. Chim. Pays-Bas 77, 872 (1954).
[65] V. L. Shapovalov and G. Brezesinski, J. Phys. Chem. B 110,
10032 (2006).
[66] S. Lamperski and C. W. Outhwaite, J. Colloid Interface Sci. 328,
458 (2008).
[67] T. Goel, C. N. Patra, S. K. Ghosh, and T. M ukherjee, J. Chem.
Phys. 132, 194706 (2010).
[68] V. B. Chu, Y. Bai, J. Lipfert, D. Herschlag, and S. Doniach,
Biophys. J. 93, 3202 (2007).
[69] J. G. Ibarra-Armenta, A. Mart
´
ın-Molina, and M. Quesada-P
´
erez,
Phys. Chem. Chem. Phys. 11, 309 (2009).
[70] J. Urbanija, K. Bohinc, A. Bellen, S. Maset, A. Iglic, V. Kralj-
Iglic, and P. B. S. Kumar, J. Chem. Phys. 129, 105101 (2008).
[71] S. Perutkova, M. Frank, K. Bohinc, G. Bobojevic, J. Zelko, B.
Rozman, V. Kralj-Iglic, and A. Iglic, J. Membr. Biol. 236,43
(2010).
[72] J. Cervera, V. Garc
´
ıa-Morales, and J Pellicer, J. Phys. Chem. B
107, 8300 (2003).
[73] J. Cervera, P. Ramirez, J. A. Manzanare, and S. Mafe, Microfluid.
Nanofluid. 9, 41 (2010).
[74] M. Wang and A. R evil, J. Colloid Interface Sci.
343, 381 (2010).
[75] S. Chakraborty and S. Das, Phys.Rev.E77, 037303 (2008).
[76] S. Das and S. Chakraborty, Langmuir 26, 11589 (2010).
[77] T. Das, S. Das, and S. Chakraborty, J. Chem. Phys. 130, 244904
(2009).
[78] M. Wang and Q. Kang, Microfluid. Nanofluid. 9, 181 (2010).
[79] C.-C. Chang and R.-J. Yang, Microfluid. Nanofluid. 9, 225
(2010).
[80] P. Goswami and S. Chakraborty, Langmuir 26, 581 (2010).
[81] S. Das and S. Chakraborty, Electrophoresis 29, 1115 (2008).
[82] S. Das and S. Chakraborty, Langmuir 24, 7704 (2008).
[83] S. Das and S. Chakraborty, Langmuir 25, 9863 (2009).
012501-4
Citations
More filters
Posted Content
TL;DR: This two-part series considers steric effects on diffuse charge dynamics (in the absence of electro-osmotic flow) and two simple models for the charging of a thin double layer, which must form a condensed layer of close-packed ions near the surface at high voltage.
Abstract: The classical Poisson-Boltzmann (PB) theory of electrolytes assumes a dilute solution of point charges with mean-field electrostatic forces. Even for very dilute solutions, however, it predicts absurdly large ion concentrations (exceeding close packing) for surface potentials of only a few tenths of a volt, which are often exceeded, e.g. in microfluidic pumps and electrochemical sensors. Since the 1950s, several modifications of the PB equation have been proposed to account for the finite size of ions in equilibrium, but in this two-part series, we consider steric effects on diffuse charge dynamics (in the absence of electro-osmotic flow). In this first part, we review the literature and analyze two simple models for the charging of a thin double layer, which must form a condensed layer of close-packed ions near the surface at high voltage. A surprising prediction is that the differential capacitance typically varies non-monotonically with the applied voltage, and thus so does the response time of an electrolytic system. In PB theory, the capacitance blows up exponentially with voltage, but steric effects actually cause it to decrease above a threshold voltage where ions become crowded near the surface. Other nonlinear effects in PB theory are also strongly suppressed by steric effects: The net salt adsorption by the double layers in response to the applied voltage is greatly reduced, and so is the tangential "surface conduction" in the diffuse layer, to the point that it can often be neglected compared to bulk conduction (small Dukhin number).

603 citations

01 Jul 2010
TL;DR: In this article, the authors studied the response of a microelectrochemical cell to a large ac voltage of frequency comparable to the inverse cell relaxation time, and showed that at very large voltage, the quasiequilibrium structure of the double layers decays.
Abstract: We study the response of a model microelectrochemical cell to a large ac voltage of frequency comparable to the inverse cell relaxation time. To bring out the basic physics, we consider the simplest possible model of a symmetric binary electrolyte confined between parallel-plate blocking electrodes, ignoring any transverse instability or fluid flow. We analyze the resulting one-dimensional problem by matched asymptotic expansions in the limit of thin double layers and extend previous work into the strongly nonlinear regime, which is characterized by two features---significant salt depletion in the electrolyte near the electrodes and, at very large voltage, the breakdown of the quasiequilibrium structure of the double layers. The former leads to the prediction of ``ac capacitive desalination'' since there is a time-averaged transfer of salt from the bulk to the double layers, via oscillating diffusion layers. The latter is associated with transient diffusion limitation, which drives the formation and collapse of space-charge layers, even in the absence of any net Faradaic current through the cell. We also predict that steric effects of finite ion sizes (going beyond dilute-solution theory) act to suppress the strongly nonlinear regime in the limit of concentrated electrolytes, ionic liquids, and molten salts. Beyond the model problem, our reduced equations for thin double layers, based on uniformly valid matched asymptotic expansions, provide a useful mathematical framework to describe additional nonlinear responses to large ac voltages, such as Faradaic reactions, electro-osmotic instabilities, and induced-charge electrokinetic phenomena.

128 citations

Journal ArticleDOI
TL;DR: In this paper, the authors developed the theory of the electrical double layer in ionic liquids as proposed earlier by Kornyshev (2007), and derived a modified expression for differential capacitance, which makes the results closer to typical experimental observations, and makes it appealing to use this formalism for treatment of experimental data.

123 citations

Journal ArticleDOI
01 May 2020-Fuel
TL;DR: In this paper, the authors summarized relevant mechanisms for predicting responses of low salinity waterflooding (LSW) and an overview of the optimal wettability for oil recovery is given with relevant pore-scale fluid mechanisms.

86 citations

Journal ArticleDOI
TL;DR: A model of the electro-diffusion of ions in porous electrodes is applied to analyze the dynamics of capacitive-mixing extraction of energy from salinity gradients with carbon porous electrodes and the complex time-evolution of the cell voltage observed in experiments is satisfactorily described.
Abstract: A model of the electro-diffusion of ions in porous electrodes is applied to analyze the dynamics of capacitive-mixing extraction of energy from salinity gradients with carbon porous electrodes. The complex time-evolution of the cell voltage observed in experiments is satisfactorily described. The asymmetry on the duration of the solution-change steps performed in open circuit is found to be due to the nonlinear voltage–concentration relationship of the electric double layers and to a current that redistributes the counterions along the depth of the electrode leading to nonuniform charge and salt adsorption. The validated model is an essential tool for the design and optimization of renewable energy extraction by this technique.

85 citations

References
More filters
Book
01 Jan 1981

2,324 citations


"Steric-effect-induced enhancement o..." refers background in this paper

  • ...The intricate dynamics within the surface-bound double layer of charge, popularly known as the electric-double layer (EDL) [9], often dictates the functionalities of these systems to an extremely significant extent....

    [...]

Journal ArticleDOI
TL;DR: A contribution to the theory of electrocapillarity is given in this paper, where it is shown that the theory can be applied to the field of electrochemysics as well as physics.
Abstract: (1913). LI. A contribution to the theory of electrocapillarity. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science: Vol. 25, No. 148, pp. 475-481.

2,252 citations

Journal ArticleDOI
TL;DR: This critical review will explore the vast manifold of length scales emerging for fluid behavior at the nanoscale, as well as the associated mechanisms and corresponding applications, and in particular explore the interplay between bulk and interface phenomena.
Abstract: Nanofluidics has emerged recently in the footsteps of microfluidics, following the quest for scale reduction inherent to nanotechnologies. By definition, nanofluidics explores transport phenomena of fluids at nanometer scales. Why is the nanometer scale specific? What fluid properties are probed at nanometric scales? In other words, why does ‘nanofluidics’ deserve its own brand name? In this critical review, we will explore the vast manifold of length scales emerging for fluid behavior at the nanoscale, as well as the associated mechanisms and corresponding applications. We will in particular explore the interplay between bulk and interface phenomena. The limit of validity of the continuum approaches will be discussed, as well as the numerous surface induced effects occurring at these scales, from hydrodynamic slippage to the various electro-kinetic phenomena originating from the couplings between hydrodynamics and electrostatics. An enlightening analogy between ion transport in nanochannels and transport in doped semi-conductors will be discussed (156 references).

1,111 citations

Journal ArticleDOI
TL;DR: A detailed critique of the present work on the potential dependence of double-layer capacitance at a planar metal-ionic liquid interface is provided, promoting the point of view that future considerations of ionic liquids should be based on the modern statistical mechanics of dense Coulomb systems, or density-functional theory, rather than classical electrochemical theories which hinge on a dilute-solution approximation.
Abstract: Applications of ionic liquids at electrified interfaces to energy-storage systems, electrowetting devices, or nanojunction gating media cannot proceed without a deep understanding of the structure and properties of the interfacial double layer. This article provides a detailed critique of the present work on this problem. It promotes the point of view that future considerations of ionic liquids should be based on the modern statistical mechanics of dense Coulomb systems, or density-functional theory, rather than classical electrochemical theories which hinge on a dilute-solution approximation. The article will, however, contain more questions than answers. To trigger the discussion, it starts with a simplified original result. A new analytical formula is derived to rationalize the potential dependence of double-layer capacitance at a planar metal−ionic liquid interface. The theory behind it has a mean-field character, based on the Poisson−Boltzmann lattice-gas model, with a modification to account for the...

1,092 citations