viscoelectric effects in nanochannel electrokinetics · viscoelectric effects in nanochannel...

6
Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu, 1, * Dalton J. E. Harvie, 2 Malcolm R. Davidson, 2 David E. Dunstan, 2 Junho Hwang, 1 and Hirofumi Daiguji 1 1 Department of Mechanical Engineering, University of Tokyo, Tokyo 113-8656, Japan 2 Department of Chemical and Biomolecular Engineering, University of Melbourne, Victoria 3010, Australia Electrokinetic transport behavior in nanochannels is different to that in larger sized channels. Specifically, molecular dynamics (MD) simulations in nanochannels have demonstrated two little understood phenomena which are not observed in microchannels, being: (i) the decrease of average electroosmotic mobility at high surface charge density, and (ii) the decrease of channel conductance at high salt concentrations, as the surface charge is increased. However, current electric double layer models do not capture these results. In this study we provide evidence that this inconsistency primarily arises from the neglect of the viscoelectric effect (being the increase of local viscosity near charged surfaces due to water molecule orientation) in conventional continuum models. It is shown that predictions of electroosmotic mobility in a slit nanochannel, derived from a viscoelectric- modified continuum model, are in quantitative agreement with previous MD simulation results. Furthermore, viscoelectric effects are found to dominate over ion steric and dielectric saturation effects in both electroosmotic and ion transport processes. Finally, we indicate that mechanisms of the previous MD-observed phenomena can be well-explained by the viscoelectric theory. Electrokinetic transport of aqueous electrolytes in nanofluidic channels is of essential importance to a number of cutting-edge technologies such as capaci- tive deionizaion [1], nanofluidic batteries [2] and bio- nanosensing [3]. To gain a fundamental understanding of electrokinetics in these systems, both molecular dynam- ics (MD) and continuum theory have been conducted to investigate transport processes at the nanoscale [4, 5]. As MD simulations directly utilize atomic properties when predicting transport behavior, one would expect these results to be more accurate than continuum based simu- lations, however high computational cost severely limits the dimensions of investigable systems. Within nanoflu- idic devices, although the channel size (i.e., height or diameter) is just a few nanometers, the length (and/or width) of these channels can be up to several microns or millimeters in many cases, which is computationally in- feasible for MD [6]. Besides, for many nanofluidic appli- cations, nanochannels are integrated with microchannels (or reservoirs) that directly influence the transport be- havior within the nanochannels, further expanding the required simulation domain [7]. Hence under most cir- cumstances, continuum simulations must be relied on. However, a lack of consistency between previous MD and continuum simulation results for nano-length nanochannels has been observed [8, 9], implying that the employed continuum models may be missing criti- cal physical transport phenomena that are relevant in nanochannels. In particular, an increase in solvent vis- cosity and decrease in solute diffusivities (compared to their bulk values) in the vicinity of a charged surface are found in previous MD simulations [9, 10] (properties which are usually assumed constant in continuum mod- els), resulting in an overestimated electroosmotic veloc- ity and channel conductance by continuum theory. In FIG. 1: Schematic of the transport behavior in a pos- itively charged slit nanochannel. Within the viscoelec- tric layers (VELs), the viscosity is increased and ionic diffusivities and dielectric permittivity are decreased, at- tributed to the orientation of water molecules. The solu- tion is effectively immobile in the vicinity of the surface (.. within the viscoelectric (VE) immobile layers) due to the high viscosity. denotes the channel height. addition, previous MD simulations [9] have reported two counter-intuitive transport phenomena in nanochannels: Decreases in (i) electroosmotic mobility at high surface charge levels, and (ii) channel conductance at high salt concentrations, when increasing the surface charge. De- tailed mechanisms of these unique phenomena have not been clarified previously, although it was suspected that they may relate to the increased viscosity [11]. To investigate full range nanofluidic systems with suf- ficient accuracy, and to better understand electrokinetic transport phenomena in nanochannels, a simple contin- uum model that captures the dominant behavior in MD simulations is desired but yet to be available. Thus, the objective of this study is to construct a modified contin- arXiv:1611.02449v2 [physics.flu-dyn] 14 Nov 2016

Upload: others

Post on 07-Jul-2020

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Viscoelectric Effects in Nanochannel Electrokinetics · Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu,1, *Dalton J. E. Harvie, 2Malcolm R. Davidson, David E. Dunstan,2

Viscoelectric Effects in Nanochannel Electrokinetics

Wei-Lun Hsu,1, * Dalton J. E. Harvie,2 Malcolm R. Davidson,2

David E. Dunstan,2 Junho Hwang,1 and Hirofumi Daiguji1

1Department of Mechanical Engineering, University of Tokyo, Tokyo 113-8656, Japan2Department of Chemical and Biomolecular Engineering,

University of Melbourne, Victoria 3010, Australia

Electrokinetic transport behavior in nanochannels is different to that in larger sized channels.Specifically, molecular dynamics (MD) simulations in nanochannels have demonstrated two littleunderstood phenomena which are not observed in microchannels, being: (i) the decrease of averageelectroosmotic mobility at high surface charge density, and (ii) the decrease of channel conductanceat high salt concentrations, as the surface charge is increased. However, current electric doublelayer models do not capture these results. In this study we provide evidence that this inconsistencyprimarily arises from the neglect of the viscoelectric effect (being the increase of local viscositynear charged surfaces due to water molecule orientation) in conventional continuum models. It isshown that predictions of electroosmotic mobility in a slit nanochannel, derived from a viscoelectric-modified continuum model, are in quantitative agreement with previous MD simulation results.Furthermore, viscoelectric effects are found to dominate over ion steric and dielectric saturationeffects in both electroosmotic and ion transport processes. Finally, we indicate that mechanisms ofthe previous MD-observed phenomena can be well-explained by the viscoelectric theory.

Electrokinetic transport of aqueous electrolytes innanofluidic channels is of essential importance to anumber of cutting-edge technologies such as capaci-tive deionizaion [1], nanofluidic batteries [2] and bio-nanosensing [3]. To gain a fundamental understanding ofelectrokinetics in these systems, both molecular dynam-ics (MD) and continuum theory have been conducted toinvestigate transport processes at the nanoscale [4, 5]. AsMD simulations directly utilize atomic properties whenpredicting transport behavior, one would expect theseresults to be more accurate than continuum based simu-lations, however high computational cost severely limitsthe dimensions of investigable systems. Within nanoflu-idic devices, although the channel size (i.e., height ordiameter) is just a few nanometers, the length (and/orwidth) of these channels can be up to several microns ormillimeters in many cases, which is computationally in-feasible for MD [6]. Besides, for many nanofluidic appli-cations, nanochannels are integrated with microchannels(or reservoirs) that directly influence the transport be-havior within the nanochannels, further expanding therequired simulation domain [7]. Hence under most cir-cumstances, continuum simulations must be relied on.

However, a lack of consistency between previousMD and continuum simulation results for nano-lengthnanochannels has been observed [8, 9], implying thatthe employed continuum models may be missing criti-cal physical transport phenomena that are relevant innanochannels. In particular, an increase in solvent vis-cosity and decrease in solute diffusivities (compared totheir bulk values) in the vicinity of a charged surfaceare found in previous MD simulations [9, 10] (propertieswhich are usually assumed constant in continuum mod-els), resulting in an overestimated electroosmotic veloc-ity and channel conductance by continuum theory. In

FIG. 1: Schematic of the transport behavior in a pos-itively charged slit nanochannel. Within the viscoelec-tric layers (VELs), the viscosity is increased and ionicdiffusivities and dielectric permittivity are decreased, at-tributed to the orientation of water molecules. The solu-tion is effectively immobile in the vicinity of the surface(𝑖.𝑒. within the viscoelectric (VE) immobile layers) dueto the high viscosity. 𝐻 denotes the channel height.

addition, previous MD simulations [9] have reported twocounter-intuitive transport phenomena in nanochannels :Decreases in (i) electroosmotic mobility at high surfacecharge levels, and (ii) channel conductance at high saltconcentrations, when increasing the surface charge. De-tailed mechanisms of these unique phenomena have notbeen clarified previously, although it was suspected thatthey may relate to the increased viscosity [11].

To investigate full range nanofluidic systems with suf-ficient accuracy, and to better understand electrokinetictransport phenomena in nanochannels, a simple contin-uum model that captures the dominant behavior in MDsimulations is desired but yet to be available. Thus, theobjective of this study is to construct a modified contin-

arX

iv:1

611.

0244

9v2

[ph

ysic

s.fl

u-dy

n] 1

4 N

ov 2

016

Page 2: Viscoelectric Effects in Nanochannel Electrokinetics · Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu,1, *Dalton J. E. Harvie, 2Malcolm R. Davidson, David E. Dunstan,2

2

uum model that offers close predictions to previous MDresults and, using it as a tool, we examine the mecha-nisms of the observed phenomena in water and ion trans-port.

The channel geometry and solution conditions consid-ered are based on the previous MD study by Qiao andAluru [9] in which a potassium chloride (KCl) aqueoussolution is confined in a long slit channel separated bya distance 𝐻, as illustrated in Figure 1. Both surfacespossess the same amount of positive charge uniformly dis-tributed along each wall. An external electric field 𝐸ext

(|𝐸ext |= 0.2 V/nm) is applied parallel to the surfaces,simultaneously yielding an electroosmotic flow in the op-posite direction of 𝐸ext and an electric current in thesame direction of 𝐸ext , whereby the local nanochannelconductance in the 𝑦-direction per unit length along thechannel 𝐺L(𝑦) is obtained as :

𝐺L(𝑦) = 𝑒

(︂𝑣(𝑦)

|𝐸ext|+

𝑒𝒟K+

𝑘B𝑇

)︂𝑛K+(𝑦)

−𝑒

(︂𝑣(𝑦)

|𝐸ext|− 𝑒𝒟Cl−

𝑘B𝑇

)︂𝑛Cl−(𝑦)

(1)

in which 𝑦 is the direction normal to the surfaces with 𝑦 =0 on the channel centerline, 𝑒 the element charge, 𝑣(𝑦)the electroosmotic velocity, 𝑘B the Boltzmann constant,𝑇 the temperature (= 300 K), 𝒟K+ the ionic diffusivityof K+ , 𝒟Cl− the ionic diffusivity of Cl− , 𝑛K+(𝑦) the K+

concentration and 𝑛Cl−(𝑦) the Cl− concentration.We employ the electric Poisson equation and a modi-

fied Navier-Stokes equation (considering an electric bodyforce from 𝐸ext) to calculate 𝑛K+(𝑦), 𝑛Cl−(𝑦) and 𝑣(𝑦) :

𝑑2𝜑(𝑦)

𝑑𝑦2= − 𝜌𝑒

𝜖r𝜖0= − 𝑧𝑒

𝜖r𝜖0(𝑛K+(𝑦)− 𝑛Cl−(𝑦)) (2)

𝑑

𝑑𝑦(𝜂

𝑑𝑣(𝑦)

𝑑𝑦) + 𝜌𝑒|𝐸ext | = 0 (3)

In these expressions, 𝜑(𝑦) is the electric potential, 𝜌𝑒 thespace charge density, 𝜖r the relative permittivity of thesolution, 𝜖0 the permittivity of vacuum, 𝜂 the viscosityand 𝑧 the ionic valence of binary electrolytes (=1 for theKCl solution).

The presence of ions affects the transport behavior inthree ways:

(i) Steric (S) effects : Since the classical Boltzmannequation assumes the ions are point-like (which becomesinvalid when the channel size is just several nanometers),a modified Boltzmann distribution considering steric ef-fects of ions is employed and 𝜌𝑒 is expressed as [12] :

𝜌𝑒 =2𝑧𝑒𝑛0 sinh(

𝑧𝑒𝜑𝑘B𝑇 )

1 + 2𝜆 sinh2( 𝑧𝑒𝜑2𝑘B𝑇 )

(4)

where 𝑛0 is the bulk KCl concentration and the bulkvolume fraction of ions 𝜆 is given by:

𝜆 = 2𝑛0𝑎3 (5)

where 𝑎 is the size of hydrated ions and given by 𝑎 =6.6 �̊� [9].(ii) Dielectric (DE) effects : Due to the high electric

field near the surfaces, the water molecules are electri-cally saturated and thus the dielectric permittivity atthe interface is lower than the bulk value [13]. A per-mittivity modification was proposed by Booth [14] basedon the Onsager [15] and Kirkwood [16] theories of polardielectrics and re-interpreted by Hunter [17] as :

𝜖r = 𝜖r,0(1− 𝑏|𝐸EDL |2) (6)

where 𝜖r,0 and 𝐸EDL are the relative permittivity ofthe solution in the absence of an electric field (= 81)and local electric field within the electric double layer(EDL), respectively. The coefficient 𝑏 is estimated to be4×10−18 m2/V2 for water [17, 18].(iii) Viscoelectric (VE) effects : As a result of the varia-

tion in vibration frequency of water molecules, the inter-actions of orientated water molecules near the chargedsurface increase. Consequently, the motion of watermolecules is largely inhibited, giving rise to a higher ap-parent viscosity. A formula was proposed by Andradeand Dodd [19] based on experimental observations at lowelectric field magnitude and, later theoretically verifiedby Lyklema and Overbeek [20] for water :

𝜂 = 𝜂0(1 + 𝑓 |𝐸EDL |2) (7)

where 𝜂0 and 𝑓 are the viscosity of the solution in theabsence of an electric field (= 0.743 mPa·s) and VE co-efficient, respectively. Here, we extend eq 7 for arbitraryelectric field magnitude based on the theory of polariza-tion [20] :

𝜂 = 𝜂0 exp(Δ𝐸a

𝑘B𝑇) = 𝜂0 exp(

𝛼𝑚2𝐸2i

𝑘B𝑇)

= 𝜂0 exp(𝑓 |𝐸EDL |2) = 𝜂0

∞∑︁𝑛=0

(𝑓 |𝐸EDL |2)𝑛

𝑛!

(8)

where Δ𝐸a is the increased activation energy due to thepresence of 𝐸EDL , constant 𝛼 a structural coefficient, 𝑚the dipole moment, and 𝐸i the internal electric field mag-nitude which is proportional to |𝐸EDL |. If 𝑓 |𝐸EDL |2 ≪ 1,eq 8 converges to eq 7, however there is no theoretical ba-sis for neglecting the higher order terms of eq 8 (generallyfor EDLs), and indeed, in this paper we find that theireffect is significant (a typical value of surface |𝐸EDL | =1.18 ×108 V/m, when the surface charge density equals80 mC/m2, making 𝑓 |𝐸EDL |2 > 1). In this study, wherewe employ eq 8, it is found that 𝑓 = 2.3 ×10−16 m2/V2

achieves the closest fit to MD simulation results, whichis close to previous experimental estimates of 𝑓 [21].

Page 3: Viscoelectric Effects in Nanochannel Electrokinetics · Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu,1, *Dalton J. E. Harvie, 2Malcolm R. Davidson, David E. Dunstan,2

3

FIG. 2: (a) Ion concentrations 𝑛Cl− (solid curves) and𝑛K+ (dashed curves), (b) electric potential 𝜑 [24], (c)electroosmotic velocity 𝑣 and (d) viscosity 𝜂 profiles alongthe 𝑦-direction at 𝜎 = 80 mC/m2 and 𝐻 = 3.49 nm. Thegray regions indicate the boundaries of Qiao and Aluru’scontinuum model [9].

Based on the Stokes-Einstein equation, the ionic dif-fusivity 𝒟𝑖 (in which 𝑖 denotes K+ or Cl−) concurrentlydecreases due to the increase of hydrodynamic drag forceon ions when migrating in a viscous solution [22] :

𝒟𝑖 =𝒟𝑖,0𝜂0

𝜂(9)

where 𝒟𝑖,0 is the ionic diffusivity in the absence ofan electric field (= 1.96 × 10−9 m2/s for K+ and2.03 × 10−9 m2/s for Cl−). A decrease in diffusion co-efficients of spherical nanoparticles within a quartz (neg-atively charged) nanopillar chip was experimentally ob-served by Kaji et al. [23] supporting eq 9.

On the boundaries, it is assumed that the nanochan-nel walls are non-conductive that the surface charge isentirely balanced by the net charge within the solution,and non-slip [9]. We derive the following boundary con-ditions at 𝑦 = ±𝐻/2:

𝑑𝜑

𝑑𝑦= ∓ 𝜎

𝜖r𝜖0(10)

𝑣 = 0 (11)

in which 𝜎 denotes the surface charge density.For comparison, we employ four continuum models : a

classical Gouy-Chapman (GC) model that ignores S, DEand VE effects (𝑎 = 𝑏 = 𝑓 = 0), a S model (𝑎 = 6.6 �̊�and 𝑏 = 𝑓 = 0), a DE model (𝑏 = 4×10−18 m2/V2 and 𝑎= 𝑓 = 0), and a VE model (𝑓 = 2.3 ×10−16 m2/V2 and𝑎 = 𝑏 = 0). In addition, we also employ Qiao and Aluru’scontinuum model [9]. This model is basically the same

as the GC model except the boundaries of eqs 10 and11 are on the center of the first layer of water moleculesand ions adjacent to walls, respectively. Specifically, theboundaries for 𝑣 and 𝜑 are 𝑦 = (± 𝐻/2 ∓ 1.6) �̊� and 𝑦= (± 𝐻/2 ∓ 3.3) �̊�, respectively, in this model.Figure 2 shows profiles of 𝑛K+ , 𝑛Cl− , 𝜑, 𝑣 and 𝜂 along

the 𝑦-direction for five continuum models and the MDresults. Despite the fact that it provides close estimatesto the MD results in 𝑛K+ , 𝑛Cl− and 𝜑, Qiao and Aluru’scontinuum model greatly overestimates 𝑣. The effect ofthe boundary shifts on these results can be clarified bycomparing the results of Qiao and Aluru’s model andthe GC model. Because the gradient of 𝜑 with respectto 𝑦 for the GC model is smaller than that for Qiao andAluru’s model, 𝑣 for the GC model is smaller than thatfor Qiao and Aluru’s model (still much larger than theMD results, nevertheless). Both S and DE effects do notgreatly change these profiles but slightly increase 𝑣 espe-cially around the channel center in comparison with theGC results. In contrast, when the VE effect is considered,𝑣 is significantly reduced and becomes in close agreementwith the MD simulation results. This indicates that thelow 𝑣 is a consequence of the increased 𝜂 near the surfacesdue to water molecule orientation. Notably, the 𝑣 profileobtained from the MD simulation is greatly suppressednear the surfaces. The increased 𝜂 near the surfaces isessential to this 𝑣 profile and a higher average 𝜂 cannotreproduce it.The average electroosmotic mobility 𝜇 (over the

nanochannel cross sectional area) and local electroos-motic velocity at different 𝜎 are shown in Figure 3, where𝜇 is obtained as [9] :

𝜇 =1

𝐻

∫︁ 𝐻2

−𝐻2

𝑣

|𝐸ext|𝑑𝑦 (12)

As seen in Figure 3a, the results of the VE model quanti-tatively agree with the MD results over the whole rangeof 𝜎. Importantly, 𝜇 becomes much less sensitive to 𝜎 athigh levels of 𝜎. A slight decrease of 𝜇 while increasing𝜎 even occurs when 𝜎 > 90 mC/m2. Note that, despitethe hydrated ion size (𝑎 = 6.6 �̊�) being roughly 19 %of 𝐻 and 𝜖r near the walls dropping over 10% from itsbulk value (as shown in the sub-figure of Figure 3a), itis shown that once VE effects are included, S and DEeffects are significantly suppressed as evidenced by thesmall difference between the results from the VE modeland a continuum model (S/DE/VE) that simultaneouslyconsiders S, DE and VE effects.Figure 3b shows 𝑣 profiles across the nanochannel at

different 𝜎 based on the VE model. In the low 𝜎 regime(𝑖.𝑒. 30 and 60 mC/m2, under which VE effects are rel-atively weak), the centerline electroosmotic velocity 𝑣cincreases with the increase of 𝜎 as does the averaged 𝜇.At higher 𝜎 levels (≥ 90 mC/m2), VE immobile layers(in which the solution becomes immobile) are gradually

Page 4: Viscoelectric Effects in Nanochannel Electrokinetics · Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu,1, *Dalton J. E. Harvie, 2Malcolm R. Davidson, David E. Dunstan,2

4

FIG. 3: (a) Variation of average electroosmotic mobility𝜇 (along with a sub-figure showing variation of relativepermittivity 𝜖r at the wall interfaces as a function of 𝜎under the DE model) and (b) electroosmotic velocity 𝑣distributions along the 𝑦-direction at different levels ofsurface charge 𝜎 at 𝐻 = 3.49 nm and KCl concentration𝑛0 = 0.6 M.

formed in the vicinity of the walls due to the high localviscosity (highlighted in gray).

It is found that this insensitiveness of 𝑣 upon 𝜎 arisesfrom the presence of the VE immobile layers. By dou-ble integrating eq 3 from the centerline (𝑑𝑣/𝑑𝑦 =0 and𝑣 = 𝑣c) to the location at which the solution begins be-coming immobile (𝑖.𝑒. the boundary of the VE immobilelayers, where 𝑣 ≈ 0), we derive a modified Smoluchowskiequation:

𝑣c|𝐸ext|

=𝜖0(𝜖r,0𝜑c − 𝜖r,IL𝜑IL)

𝜂0(13)

where 𝜑c, 𝜖r,IL and 𝜑IL are the electric potential at thecenterline, relative permittivity at the boundary of theVE immobile layers and electric potential at the bound-ary of the VE immobile layers, respectively. This equa-

FIG. 4: Variation of average electroosmotic mobility 𝜇as a function of surface charge density 𝜎 (a) at differentchannel size 𝐻 and KCl concentration 𝑛0 = 0.6 M, and(b) at different 𝑛0 and 𝐻 = 3.49 nm, respectively.

tion indicates that when the VE immobile layers exist, 𝑣cis no longer determined by 𝜑 on the channel walls (whichis a function of 𝜎 by eq 10). Instead, it depends upon a𝜎-independent parameter 𝜖r,IL𝜑IL, rendering almost con-stant 𝑣c when 𝜎 ≥ 90 mC/m2. In contrast, the local𝑣 (in between the centerline and the VE immobile layerboundary) decreases in response to the thickening of theVE immobile layers. As a consequence, the derived 𝜇decreases at larger 𝜎.Given that the average mobility decrease is a result of

the slower electroosmotic velocities within the EDLs, thedegree of this decrease greatly depends on the channelsize and the EDL thickness. At constant 𝑛0 (= 0.6 Min Figure 4a), the decrease appears in the small channels(𝐻 = 2 and 3.49 nm) due to the overlapped EDLs, but italmost vanishes at large 𝐻 (= 10 nm). Similarly, at fixed𝐻 (=3.49 nm in Figure 4b), the decrease is apparent atthe lower concentrations (𝑒.𝑔. at 𝑛0 = 0.1 M) when EDLoverlap occurs. Hence, in summary, the average mobilitydecrease occurs under conditions where the VE immobilelayer is significant at small 𝐻 and low 𝑛0.

We investigate ion transport behavior using thenanochannel conductance per unit length along the chan-nel 𝐺 given by:

𝐺 =

∫︁ 𝐻2

−𝐻2

𝐺L𝑑𝑦 (14)

As seen in Figure 5a, at 𝑛0 = 1 M, 𝐺 decreases as 𝜎increases for the VE-modified models (𝑖.𝑒. the VE andS/DE/VE models). Conversely, in a lower concentrationsolution with 𝑛0 = 0.1 M (Figure 5b), this relationshipreverses at low 𝜎, namely 𝐺 increases with the increaseof 𝜎, before plateauing at 𝜎 ≥ 60 mC/m2. These con-tinuum based results are consistent with previous MDsimulations [9]. Note that, at high 𝑛0 (= 1 M in Fig-ure 5a), S effects, which amplify VE effects due to ionjamming [25], become non-negligible, although the qual-itative conductive behavior in response to the 𝜎 increaseremains similar.We herein define a viscoelectric layer (VEL), as illus-

trated in Figure 1, in which VE effects are significant.

Page 5: Viscoelectric Effects in Nanochannel Electrokinetics · Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu,1, *Dalton J. E. Harvie, 2Malcolm R. Davidson, David E. Dunstan,2

5

FIG. 5: Variation of nanochannel conductance per unitlength 𝐺 as a function of surface charge density 𝜎 at𝐻 = 3.49 nm and KCl concentration 𝑛0 = (a) 1M and(b) 0.1M, respectively. Sub-figures show local nanochan-nel conductance per unit length 𝐺L at different surfacecharge density 𝜎. (c) Variation of 𝐺 as a function of 𝑛0

at different 𝜎 levels. At low 𝑛0 (EDL > VEL) and low𝜎, 𝐺, which increases with 𝜎, is dominated by chargingeffects. At high 𝑛0 (EDL = VEL), VE effects dominateand 𝐺 decreases at larger 𝜎. In inset figures, the VELsare highlighted by the red shadow.

At high concentrations, the size of the region coveredby the VELs is equivalent to the EDL region. On theother hand, at low concentrations, when the EDLs be-come overlapped, the VELs only occupy the wall adja-cent region. As a results, a ‘VE free zone’ remains aroundthe centerline. At high concentrations (e.g. 𝑛0 = 1 M), asseen in the sub-figure in Figure 5a, 𝐺L near the centerlinearea is constant, implying that the EDLs are not over-lapped. Within the EDLs (equivalent to VELs), in which𝐺L is suppressed by 𝜎, 𝐺L decreases with the increase of𝜎, due to higher 𝜂 and thus lower 𝒟𝑖 (based on eq 9). Inconsequence, 𝐺 decreases gradually with the increase of𝜎. At low concentrations (e.g. 𝑛0 = 0.1 M), as seen in thesub-figure in Figure 5b, 𝐺L is altered with 𝜎 across thewhole range in the nanochannels. When 𝜎 increases, 𝐺L

is altered by two competing effects : (i) VE effects and(ii) charging effects. The former suppresses 𝐺L and thelatter, which refers to the increase of net charge withinthe solution, enhances 𝐺L. At low 𝜎, charging effectsdominate over the VE effects, while at high 𝜎, two fac-

tors compete and offset each other.Figure 5c shows the calculated nanochannel conduc-

tance versus KCl concentration (𝐺 − 𝑛0) curves for dif-ferent 𝜎 based on the VE model. At high 𝑛0, 𝐺 decreaseswith the increase of 𝜎 in the same way as the previousMD simulation results [9] .To conclude, we have evidenced that the discrepancy

between the previous MD simulations and continuumtheory is primarily due to the neglect of VE effects inthe previous continuum model. Two unique phenomenaobserved in previous MD simulations have been describedby a modified continuum model that considers VE effects.The success of the VE-modified continuum theory shownhere suggests that a similar re-examination of differentelectrokinetic systems at the nanoscale may also help toreduce discrepancies between results derived from con-tinuum theory, MD simulation and experiment.We acknowledge Prof. Rui Qiao of Virginia Tech for

providing us with the MD simulation results and someadditional information of ref 9.

* Electronic address: [email protected][1] M. E. Suss, T. F. Baumann, W. L. Bourcier, C. M.

Spadaccini, K. A. Rose, J. G. Santiago, and M. Stader-mann, Energy Environ. Sci. 5, 9511 (2012).

[2] H. Daiguji, P. D. Yang, A. J. Szeri, and A. Majumdar,Nano Lett. 4, 2315 (2004).

[3] C. Dekker, Nature Nanotechnol. 2, 209 (2007).[4] K. Shirono, N. Tatsumi, and H. Daiguji, J. Phys. Chem.

B 113, 1041 (2009).[5] H. Daiguji, Chem. Soc. Rev. 39, 901 (2010).[6] D.-K. Kim, C. Duan, Y.-F. Chen, and A. Majumdar,

Microfluid. Nanofluid. 9, 1215 (2010).[7] W.-L. Hsu, D. W. Inglis, H. Jeong, D. E. Dunstan, M. R.

Davidson, E. M. Goldys, and D. J. E. Harvie, Langmuir30, 5337 (2014).

[8] R. Qiao and N. R. Aluru, J. Chem. Phys. 118, 4692(2003).

[9] R. Qiao and N. R. Aluru, Langmuir 21, 8972 (2005).[10] J. B. Freund, J. Chem. Phys. 116, 2194 (2002).[11] P. Wu and R. Qiao, Phys. Fluids 23, 072005 (2011).[12] I. Borukhov, D. Andelman, and H. Orland, Phys. Rev.

Lett. 79, 435 (1997).[13] P. Debye, Polar molecules (Chemical Catalogue Com-

pany, Reinhold, 1929).[14] F. Booth, J. Chem. Phys. 19, 391 (1951).[15] L. Onsager, J Am. Chem. Soc. 58, 1486 (1936).[16] J. G. Kirkwood, J. Chem. Phys. 7, 911 (1939).[17] R. J. Hunter, J. Colloid Interface Sci. 22, 231 (1966).[18] R. J. Hunter, Zeta potential in colloid science: Principles

and applications (Elsevier, 1981).[19] E. N. d. C. Andrade and C. Dodd, Proc. R. Soc. Lond.

A 204, 449 (1951).[20] J. Lyklema and J. T. Overbeek, J. Colloid Sci. 16, 501

(1961).[21] The VE coefficient 𝑓 was experimentally estimated to

be (0.5 − 1) × 10−15 m2/V2 for water by Hunterand Leyendekkers based on eq 7, [26] which is close

Page 6: Viscoelectric Effects in Nanochannel Electrokinetics · Viscoelectric Effects in Nanochannel Electrokinetics Wei-Lun Hsu,1, *Dalton J. E. Harvie, 2Malcolm R. Davidson, David E. Dunstan,2

6

to Lyklema and Overbeek’s theoretical prediction of𝑓 = 1.02 × 10−15 m2/V2 [20]. 𝑓 estimated in this study2.3 ×10−16 m2/V2 is slightly lower than Hunter andLeyendekkers’ estimate, however the difference may bedue to the fact that 𝜂 was several times higher than𝜂0 near the interface in Hunter and Leyendekkers’s ex-periments, resulting in non-negligible higher order fieldterms (which were neglected in their analysis) and a con-sequential overestimation of 𝑓 . Other studies have simi-larly concluded that Lyklema and Overbeek’s coefficientis too high [17, 27].

[22] W.-L. Hsu, H. Daiguji, D. E. Dunstan, M. R. Davidson,and D. J. E. Harvie, Adv. Colloid Interface Sci. 234, 108

(2016).[23] N. Kaji, O. R, O. A, H. Y, T. M, and B. Y, Anal. Bioanal.

Chem. 386, 759 (2006).[24] The MD results of 𝜑 are derived based on the MD re-

sults of 𝑛Cl− in Figure 2a using the classical Boltzmannequation.

[25] P. Olsson and S. Teitel, Phys. Rev. Lett. 99, 178001(2007).

[26] R. J. Hunter and J. V. Leyendekkers, J. Chem. Soc. Fara-day Trans. I 74, 450 (1978).

[27] D. Stinger, J. Phys. Chem. 68, 3600 (1964).