sample preconcentration in nanochannels with …sample preconcentration in nanochannels with...

8
Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden *1 , C. McCallum 1 , B. Storey 2 , S. Pennathur 1 , C. D. Meinhart 1 1 University of California, Santa Barbara, 2 Olin College *Corresponding author: Department of Mechanical Engineering, University of California, Santa Barbara, Santa Barbara, CA 93106-5070. Email: [email protected] Abstract: We present a novel method for concentrating and focusing small analytes by taking advantage of the nonuniform ion distributions produced by thick electric double layers (EDLs) in nanochannels with heterogeneous surface charge. Specifically, we apply a voltage bias to gate electrodes embedded within the channel walls, tuning the surface charge in a region of the channel and subsequently altering the ionic strength and charge density in that region relative to the rest of the channel. The resulting nonuniform electromigration fluxes in the different regions serve to stack charged sample ions near an interface where a step change in zeta potential occurs, providing enhancement ratios superior to those exhibited in traditional microchannel field amplified sample stacking (FASS). Numerical simulations are performed to demonstrate the phenomenon, and resulting velocity and salt concentration profiles show good agreement with analytical results. Keywords: nanofluidics, electrokinetic preconcentration, electroosmosis, electrophoresis 1. Introduction Recent advances in micro- and nanoscale fabrication technologies have spurred the development of myriad novel devices for bioassays, DNA separation/amplification, and other lab-on-chip processes [1-6]. However, the small size scale of these devices introduces several obstacles that must be overcome through engineering prowess. A primary concern remains the necessity for sample analyte preconcentration in bioanalytical micro and nanofluidic devices [7,8]. Many innovative focusing techniques utilizing electrokinetic phenomena such as FASS, ion concentration polarization, isotachophoresis, isoelectric focusing, and concentration gradient focusing have been introduced in previous works to address and attempt to resolve this prevalent issue in on-chip applications [1-11]. Devices employing these mechanisms often exploit the competition between electroosmotic flow (EOF) and electrophoresis in micro- and nanofluidic systems in order to create regions of localized ion enrichment. These enriched ions can then be used for further downstream processing once the level of sample molecules reaches the threshold limit detectable by modern sensing capabilities. Traditional microfluidic FASS involves the injection of a low concentration sample plug solution into a channel in order to create a conductivity gradient between the plug and the bulk fluid. These conductivity gradients produce electric field gradients which drive sample ions to “stack” at an equilibrium position where the various forces balance. Bharadwaj and Santiago comprehensively summarized the driving forces behind the stacking mechanism in microchannels [8], while Sustarich et al presented new findings of increased sample enhancement in the nanoscale regime due to pressure gradient induced flow focusing and electrostatic repulsion from finite-sized electric double layers (EDLs) relative to the channel height [9]. We extend these works by investigating the effects of thick, overlapped EDLs on the behavior of background electrolyte ions and sample ions in a nanochannel with nonuniform EOF. Other authors have previously investigated the effects of nonuniform EOF realized through means ranging from field effect control to EOF suppressing surface treatments [12-14]. Such channels have been shown to exhibit tremendous promise when it comes to controlling the behavior of bulk fluid and individual ions in applications such as nanofluidic diodes and field effect transistors [15-17]. Further, we have found that it is theoretically possible to create regions of nonuniform conductivity and electric fields within a single buffer solution by simply tailoring the surface charge in select regions of a nanochannel. We have therefore designed a tunable nanofluidic preconcentration system with embedded gate electrodes, allowing for field effect control of Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Upload: others

Post on 09-Jun-2020

18 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

Sample Preconcentration in Nanochannels with Nonuniform Surface

Charge and Thick Electric Double Layers

A. Eden*1

, C. McCallum1, B. Storey

2, S. Pennathur

1, C. D. Meinhart

1

1University of California, Santa Barbara,

2Olin College

*Corresponding author: Department of Mechanical Engineering, University of California, Santa Barbara, Santa

Barbara, CA 93106-5070. Email: [email protected]

Abstract: We present a novel method for

concentrating and focusing small analytes by

taking advantage of the nonuniform ion

distributions produced by thick electric double

layers (EDLs) in nanochannels with

heterogeneous surface charge. Specifically, we

apply a voltage bias to gate electrodes embedded

within the channel walls, tuning the surface

charge in a region of the channel and

subsequently altering the ionic strength and

charge density in that region relative to the rest

of the channel. The resulting nonuniform

electromigration fluxes in the different regions

serve to stack charged sample ions near an

interface where a step change in zeta potential

occurs, providing enhancement ratios superior to

those exhibited in traditional microchannel field

amplified sample stacking (FASS). Numerical

simulations are performed to demonstrate the

phenomenon, and resulting velocity and salt

concentration profiles show good agreement with

analytical results.

Keywords: nanofluidics, electrokinetic

preconcentration, electroosmosis, electrophoresis

1. Introduction

Recent advances in micro- and nanoscale

fabrication technologies have spurred the

development of myriad novel devices for

bioassays, DNA separation/amplification, and

other lab-on-chip processes [1-6]. However, the

small size scale of these devices introduces

several obstacles that must be overcome through

engineering prowess. A primary concern remains

the necessity for sample analyte preconcentration

in bioanalytical micro and nanofluidic devices

[7,8]. Many innovative focusing techniques

utilizing electrokinetic phenomena such as

FASS, ion concentration polarization,

isotachophoresis, isoelectric focusing, and

concentration gradient focusing have been

introduced in previous works to address and

attempt to resolve this prevalent issue in on-chip

applications [1-11]. Devices employing these

mechanisms often exploit the competition

between electroosmotic flow (EOF) and

electrophoresis in micro- and nanofluidic

systems in order to create regions of localized

ion enrichment. These enriched ions can then be

used for further downstream processing once the

level of sample molecules reaches the threshold

limit detectable by modern sensing capabilities.

Traditional microfluidic FASS involves the

injection of a low concentration sample plug

solution into a channel in order to create a

conductivity gradient between the plug and the

bulk fluid. These conductivity gradients produce

electric field gradients which drive sample ions

to “stack” at an equilibrium position where the

various forces balance. Bharadwaj and Santiago

comprehensively summarized the driving forces

behind the stacking mechanism in microchannels

[8], while Sustarich et al presented new findings

of increased sample enhancement in the

nanoscale regime due to pressure gradient

induced flow focusing and electrostatic repulsion

from finite-sized electric double layers (EDLs)

relative to the channel height [9].

We extend these works by investigating the

effects of thick, overlapped EDLs on the

behavior of background electrolyte ions and

sample ions in a nanochannel with nonuniform

EOF. Other authors have previously investigated

the effects of nonuniform EOF realized through

means ranging from field effect control to EOF

suppressing surface treatments [12-14]. Such

channels have been shown to exhibit tremendous

promise when it comes to controlling the

behavior of bulk fluid and individual ions in

applications such as nanofluidic diodes and field

effect transistors [15-17].

Further, we have found that it is theoretically

possible to create regions of nonuniform

conductivity and electric fields within a single

buffer solution by simply tailoring the surface

charge in select regions of a nanochannel. We

have therefore designed a tunable nanofluidic

preconcentration system with embedded gate

electrodes, allowing for field effect control of

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 2: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

Figure 1:

in the channel with

embedded electrode (bottom right).

surface charge heterogeneity on the top and

bottom channel walls with the application of a

gate voltage

generate nonuniform EOF and regions of

controllable electrophoretic sample motion

within the channel. Our technique is

demonstrated with 2D numerical simulations

using COMSOL Multiphysics finite element

software. The numeri

ion distributions in fully developed regions of

the flow compare favorably with results from 1D

analytical models for both thick and thin EDL

cases. The parti

sample enhancement are shown to only occ

channels with sufficiently large

layers relative to the channel height, and in

channels with nonuniform surface charge.

Sample

simulations

The unique characteristics o

interactions in

thick EDLs suggest

investigation and optimization

2. Theory

2.1 Electric Double Layer

When aqueous electrolyte solutions come

into contact with a

chemical interactions often leave the surface

with a net charge. Free ions in solution are

subsequently attracted to the charged surface,

forming an electric double layer of ions. In the

first layer adjacent to the wall, often referr

as the Stern Layer, a thin layer of oppositely

charged

attracted to the charged surface

they are essentially adsorbed on

Figure 1: Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

in the channel with a slightly

embedded electrode (bottom right).

surface charge heterogeneity on the top and

bottom channel walls with the application of a

gate voltage (see Figure

generate nonuniform EOF and regions of

controllable electrophoretic sample motion

within the channel. Our technique is

demonstrated with 2D numerical simulations

using COMSOL Multiphysics finite element

software. The numeri

ion distributions in fully developed regions of

the flow compare favorably with results from 1D

analytical models for both thick and thin EDL

cases. The particular mechanisms which cause

sample enhancement are shown to only occ

channels with sufficiently large

layers relative to the channel height, and in

channels with nonuniform surface charge.

Sample enhancement ratios

simulations exceed those from traditional FASS

unique characteristics o

interactions in nano

thick EDLs suggest

investigation and optimization

Theory

2.1 Electric Double Layer

When aqueous electrolyte solutions come

into contact with a

chemical interactions often leave the surface

with a net charge. Free ions in solution are

subsequently attracted to the charged surface,

forming an electric double layer of ions. In the

first layer adjacent to the wall, often referr

as the Stern Layer, a thin layer of oppositely

charged counter-ion

attracted to the charged surface

ey are essentially adsorbed on

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

slightly modified surface charge (top center) and the EDL potential and velocity profiles near an

embedded electrode (bottom right).

surface charge heterogeneity on the top and

bottom channel walls with the application of a

(see Figure 1). This is used to

generate nonuniform EOF and regions of

controllable electrophoretic sample motion

within the channel. Our technique is

demonstrated with 2D numerical simulations

using COMSOL Multiphysics finite element

software. The numerical velocity, potential, and

ion distributions in fully developed regions of

the flow compare favorably with results from 1D

analytical models for both thick and thin EDL

cular mechanisms which cause

sample enhancement are shown to only occ

channels with sufficiently large electric double

layers relative to the channel height, and in

channels with nonuniform surface charge.

enhancement ratios from our

exceed those from traditional FASS

unique characteristics of the

nanochannels with nonuniform

thick EDLs suggest that there is room for further

investigation and optimization of this process

2.1 Electric Double Layer

When aqueous electrolyte solutions come

into contact with a solid surface

chemical interactions often leave the surface

with a net charge. Free ions in solution are

subsequently attracted to the charged surface,

forming an electric double layer of ions. In the

first layer adjacent to the wall, often referr

as the Stern Layer, a thin layer of oppositely

ions are electrostatically

attracted to the charged surface so strongly

ey are essentially adsorbed on

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

modified surface charge (top center) and the EDL potential and velocity profiles near an

surface charge heterogeneity on the top and

bottom channel walls with the application of a

). This is used to

generate nonuniform EOF and regions of

controllable electrophoretic sample motion

within the channel. Our technique is

demonstrated with 2D numerical simulations

using COMSOL Multiphysics finite element

cal velocity, potential, and

ion distributions in fully developed regions of

the flow compare favorably with results from 1D

analytical models for both thick and thin EDL

cular mechanisms which cause

sample enhancement are shown to only occur in

electric double

layers relative to the channel height, and in

channels with nonuniform surface charge.

from our 2D

exceed those from traditional FASS.

f the ion-EDL

channels with nonuniform,

there is room for further

of this process.

When aqueous electrolyte solutions come

surface, various

chemical interactions often leave the surface

with a net charge. Free ions in solution are

subsequently attracted to the charged surface,

forming an electric double layer of ions. In the

first layer adjacent to the wall, often referred to

as the Stern Layer, a thin layer of oppositely

s are electrostatically

so strongly that

ey are essentially adsorbed on the surface.

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

modified surface charge (top center) and the EDL potential and velocity profiles near an

surface charge heterogeneity on the top and

bottom channel walls with the application of a

). This is used to

generate nonuniform EOF and regions of

controllable electrophoretic sample motion

within the channel. Our technique is

demonstrated with 2D numerical simulations

using COMSOL Multiphysics finite element

cal velocity, potential, and

ion distributions in fully developed regions of

the flow compare favorably with results from 1D

analytical models for both thick and thin EDL

cular mechanisms which cause

ur in

electric double

layers relative to the channel height, and in

channels with nonuniform surface charge.

2D

.

EDL

,

there is room for further

When aqueous electrolyte solutions come

, various

chemical interactions often leave the surface

with a net charge. Free ions in solution are

subsequently attracted to the charged surface,

forming an electric double layer of ions. In the

ed to

as the Stern Layer, a thin layer of oppositely

s are electrostatically

that

the surface.

Slightly farther away from the wall is a second

layer of

surface but are free to move due to diffusive

effects. Outside this diffuse layer, the ions in the

bulk solution are shielded from the charged

surface by the presence of the ions in the EDL.

When an external

ions

lines.

of the fluid within the EDL

along with it, generating an electroosmotic flow

profile.

The resulting distribution of ion

near the charged surface

potential

the wall

bulk fluid

potential distribution

related to the spatial free charge densi

in solution using

where

the

solution, and

density. The charge density effectively

represents the net charge present due to a local

imbalance of cations and anions in solution,

which for

as KCl

where

concentration of cations, and

concentration of anions

statistics to account for the relative energy

content of the ions in so

ions near a charged surface is approximated as

an exponentially decaying function

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

modified surface charge (top center) and the EDL potential and velocity profiles near an

Slightly farther away from the wall is a second

layer of counter-ions that are still attracted to the

surface but are free to move due to diffusive

effects. Outside this diffuse layer, the ions in the

bulk solution are shielded from the charged

surface by the presence of the ions in the EDL.

When an external

ions within the EDL

. The resulting shear stress

of the fluid within the EDL

along with it, generating an electroosmotic flow

profile.

The resulting distribution of ion

near the charged surface

potential profile which is largest in m

the wall zeta potential

bulk fluid for large channels

potential distribution

related to the spatial free charge densi

in solution using Poisson

��

where�� is the permittivity of free space,

the relative permittivity of the electrolyte

solution, and �� is the local volumetric charge

density. The charge density effectively

represents the net charge present due to a local

imbalance of cations and anions in solution,

which for a binary monovalent electrolyte

as KCl can be expressed as

where � is Faraday’s constant,

concentration of cations, and

concentration of anions

statistics to account for the relative energy

content of the ions in so

ions near a charged surface is approximated as

exponentially decaying function

� �

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

modified surface charge (top center) and the EDL potential and velocity profiles near an

Slightly farther away from the wall is a second

ions that are still attracted to the

surface but are free to move due to diffusive

effects. Outside this diffuse layer, the ions in the

bulk solution are shielded from the charged

surface by the presence of the ions in the EDL.

When an external electric field is applied, the

within the EDL will migrate

he resulting shear stress from the motion

of the fluid within the EDL pulls the bulk flu

along with it, generating an electroosmotic flow

The resulting distribution of ion

near the charged surface gives rise to an electric

which is largest in m

zeta potential ζ and decays to zero in the

for large channels. This transverse

potential distribution Ψ in the EDL can be

related to the spatial free charge densi

Poisson’s equation,

��� ��Ψ ��, is the permittivity of free space,

permittivity of the electrolyte

is the local volumetric charge

density. The charge density effectively

represents the net charge present due to a local

imbalance of cations and anions in solution,

binary monovalent electrolyte

can be expressed as ��is Faraday’s constant,

concentration of cations, and �concentration of anions. Using Boltzmann

statistics to account for the relative energy

content of the ions in solution, the distribution of

ions near a charged surface is approximated as

exponentially decaying function

,� exp �������� �

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

modified surface charge (top center) and the EDL potential and velocity profiles near an

Slightly farther away from the wall is a second

ions that are still attracted to the

surface but are free to move due to diffusive

effects. Outside this diffuse layer, the ions in the

bulk solution are shielded from the charged

surface by the presence of the ions in the EDL.

ield is applied, the

migrate along field

from the motion

pulls the bulk fluid

along with it, generating an electroosmotic flow

The resulting distribution of ions in the EDL

gives rise to an electric

which is largest in magnitude at

and decays to zero in the

. This transverse

in the EDL can be

related to the spatial free charge density of ions

equation,

(1

is the permittivity of free space, � permittivity of the electrolyte

is the local volumetric charge

density. The charge density effectively

represents the net charge present due to a local

imbalance of cations and anions in solution,

binary monovalent electrolyte such

� ���� � ��is Faraday’s constant, �� is the molar �� is the molar

. Using Boltzmann

statistics to account for the relative energy

lution, the distribution of

ions near a charged surface is approximated as

exponentially decaying function

�, (2

Diagram of the device with embedded electrodes (bottom left), with insets showing example ion distributions

modified surface charge (top center) and the EDL potential and velocity profiles near an

Slightly farther away from the wall is a second

ions that are still attracted to the

surface but are free to move due to diffusive

effects. Outside this diffuse layer, the ions in the

bulk solution are shielded from the charged

surface by the presence of the ions in the EDL.

ield is applied, the

along field

from the motion

id

along with it, generating an electroosmotic flow

s in the EDL

gives rise to an electric

agnitude at

and decays to zero in the

. This transverse

in the EDL can be

ty of ions

1)

is

permittivity of the electrolyte

is the local volumetric charge

density. The charge density effectively

represents the net charge present due to a local

imbalance of cations and anions in solution,

such

��, is the molar

e molar

. Using Boltzmann

statistics to account for the relative energy

lution, the distribution of

ions near a charged surface is approximated as

2)

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 3: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

where � is the ion valence, �� is the elementary

charge, ��� represents the thermal energy, and �,� is the bulk concentration of cations and

anions sufficiently far from the surface. The

characteristic length scale over which the

potential and ion distributions decay within the

EDL is referred to as the Debye length, which

for a symmetric, monovalent electrolyte can be

expressed as

λ! "��� ���2����� . (3)

For low concentration electrolytes (�� < 1

mM) in nanochannels, this characteristic length

can approach and even exceed the height of the

channel. In such situations, it is no longer

appropriate to assume that the ion concentrations

far away from the channel walls (i.e. along the

channel center) are the same as those in the bulk

fluid within the reservoirs supplying the channel.

The centerline concentration in the channel is

instead determined by the potential Ψ' at the

centerline, and the Boltzmann distribution

modifies to � �� exp �� ���Ψ'��� � exp (� ����Ψ � Ψ'���� ). (4)

Using the definition �� ���� � ��� and

assuming a 1D potential profile, we use the

method presented by Baldessari to integrate

Poisson’s equation and find a resulting potential

profile in the different regions of the channel

[18]. It should be noted here that for channels

with modified zeta potentials in the middle

region (see Figure 2), the “bulk” concentration �� of ions in this region will not necessarily be

equal to the supplying inlet or outlet reservoir

concentrations, and must be solved as an

additional unknown using channel-to-well

equilibrium via a species conservation equation.

If the surface charge is known instead of the zeta

potential in a given region of the channel, the

two can be related through the potential gradient

at the wall through the relation

*+ ���� ,-Ψ-./01�. (5)

2.2 Velocity and Electric Fields

The inherently low Reynolds number

associated with flows in nanofluidic devices

allows us to use the incompressible forms of the

Stokes equation and the continuity equation to

describe the conservation of momentum within a

nanochannel for a fluid experiencing an electric

body force,

0 3∇�5 � �6 + ��8; � ∙ 5 0, (6)

where 6 is the pressure within the fluid, 3 is the

dynamic viscosity of the fluid, 5 is the velocity

field, and 8 is the externally applied electric

field. Substituting the charge density from

Poisson’s equation into the electric body force

term in (6) and assuming a 1D fully developed

flow, we obtain the following equation

0 3 ;�<;.� � ;6;= � ��� >? ;�Ψ;.� . (7)

This equation can be directly integrated with a

midplane symmetry condition and a no slip

condition at the wall, resulting in the traditional

1D EOF profile

<�.� �123 -6-= �.� � @.� � �>?ζ3 (1 � Ψ�.�ζ ). (8)

We use the 1D model of Sustarich et al to find

the internal pressure gradients, electric fields,

and resulting flow profiles in the various regions.

This model applies continuity and a known

pressure drop along the channel to solve for the

internal pressure gradients in terms of the electric

fields. The electric fields are then solved with the

constraint that the area averaged electrolyte ion

fluxes due to convection and electromigration

must remain constant throughout the channel

<�AAA + B�̅> �DEFG., (9)

where B HIJKLIMNO [7]. These equations are solved for

the unknown electric fields and the unknown

“bulk” concentration �� of ions in the modified

region. The resulting potential and BGE ion

profiles are then found from the method

described in [18], and the velocity profiles are

calculated from (8).

3. 2D Numerical Model

Commercial finite element simulation

software such as COMSOL Multiphysics has

been used extensively in previous works to

numerically model electrokinetic flows in micro-

and nanofluidics [1,4,7,10,19]. In this paper, we

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 4: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

use COMSOL

electroosmotically driven flow and

electrophoretic stacking of

nanochannel with a modified zeta potential in the

middle 50% of the channel. The governin

equations are similar

in sections 2.1 and 2.2

more highly coupled Poisson

system of eq

the local ion distributions on the EDL potential.

As such, we don’t explicitly assume a Boltzmann

distribution of ions in our simulation. Instead, we

simply represent the spatial free charge density �� as a local imbal

solution, the concentrations of which are

governed by the Nernst

species

electromigration. Poisson’s equation for the EDL

transverse potential

and the Nern

;�;G

We model (1

PDE” module, and (

species

module within COMSOL. The convection term

in (11

Figure

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

and electroosmotic flow within the channel and reservoirs.

use COMSOL v5.1 to simulate a 2D model of

electroosmotically driven flow and

electrophoretic stacking of

nanochannel with a modified zeta potential in the

middle 50% of the channel. The governin

equations are similar

in sections 2.1 and 2.2

more highly coupled Poisson

system of equations to account for the effect of

the local ion distributions on the EDL potential.

As such, we don’t explicitly assume a Boltzmann

distribution of ions in our simulation. Instead, we

simply represent the spatial free charge density

as a local imbalance of cations and anions in

solution, the concentrations of which are

governed by the Nernst

species experiencing

electromigration. Poisson’s equation for the EDL

transverse potential distribution then bec

���� ��Ψand the Nernst-Planck equation is given by

�;G + 5 ∙ �� P�We model (10) using a “Coefficient Form

PDE” module, and (

species using the “Transport of Diluted Species”

module within COMSOL. The convection term

1) is due to the advective transport of ions

Figure 2: Diagram of the

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

and electroosmotic flow within the channel and reservoirs.

v5.1 to simulate a 2D model of

electroosmotically driven flow and

electrophoretic stacking of sample ions in a

nanochannel with a modified zeta potential in the

middle 50% of the channel. The governin

equations are similar to the equations presented

in sections 2.1 and 2.2; however, we use the

more highly coupled Poisson-

uations to account for the effect of

the local ion distributions on the EDL potential.

As such, we don’t explicitly assume a Boltzmann

distribution of ions in our simulation. Instead, we

simply represent the spatial free charge density

ance of cations and anions in

solution, the concentrations of which are

governed by the Nernst-Planck equation for a

experiencing convection, diffusion, and

electromigration. Poisson’s equation for the EDL

distribution then bec

Ψ ���� � ���, Planck equation is given by

� ∙ ��� + ��������) using a “Coefficient Form

PDE” module, and (11) for the various ionic

“Transport of Diluted Species”

module within COMSOL. The convection term

due to the advective transport of ions

Diagram of the numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

and electroosmotic flow within the channel and reservoirs.

v5.1 to simulate a 2D model of

electroosmotically driven flow and the

sample ions in a

nanochannel with a modified zeta potential in the

middle 50% of the channel. The governing

e equations presented

; however, we use the

-Nernst-Planck

uations to account for the effect of

the local ion distributions on the EDL potential.

As such, we don’t explicitly assume a Boltzmann

distribution of ions in our simulation. Instead, we

simply represent the spatial free charge density

ance of cations and anions in

solution, the concentrations of which are

Planck equation for a

convection, diffusion, and

electromigration. Poisson’s equation for the EDL

distribution then becomes

� (10)

Planck equation is given by

�V + ��� �. (11)

) using a “Coefficient Form

for the various ionic

“Transport of Diluted Species”

module within COMSOL. The convection term

due to the advective transport of ions

numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

and electroosmotic flow within the channel and reservoirs.

v5.1 to simulate a 2D model of

the

sample ions in a

nanochannel with a modified zeta potential in the

g

e equations presented

; however, we use the

Planck

uations to account for the effect of

the local ion distributions on the EDL potential.

As such, we don’t explicitly assume a Boltzmann

distribution of ions in our simulation. Instead, we

simply represent the spatial free charge density

ance of cations and anions in

solution, the concentrations of which are

Planck equation for a

convection, diffusion, and

electromigration. Poisson’s equation for the EDL

) using a “Coefficient Form

for the various ionic

“Transport of Diluted Species”

module within COMSOL. The convection term

due to the advective transport of ions

from the electroosmotic flow,

the Stokes equation and mass continuity

This

COMSOL using the “Creeping Flow” module.

Finally, Ohm’s Law

modeled with the “Electric Currents” module

describe the electric field within the fluid

where

electrical conductivity, and

applied

depends on the local number of charge carriers,

and contains components related to the

convective and conductive

channel [18

terms of the local ba

as

The first term

to a nonzero charge being advected downstream

by the EOF, while the second term represents the

current generated by the electromigration of

individual

the presence of the external field.

numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

and electroosmotic flow within the channel and reservoirs.

from the electroosmotic flow,

the Stokes equation and mass continuity

R0 3∇�5This electroosmotic

COMSOL using the “Creeping Flow” module.

Finally, Ohm’s Law

modeled with the “Electric Currents” module

describe the electric field within the fluid

RS *�8where S is the local current density,

electrical conductivity, and

applied potential. The electrical conductivity

depends on the local number of charge carriers,

and contains components related to the

convective and conductive

channel [18]. The conductivity can be written in

terms of the local ba

*T <?���� �>?

The first term in (14)

to a nonzero charge being advected downstream

by the EOF, while the second term represents the

current generated by the electromigration of

individual background electrolyte (

the presence of the external field.

numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

from the electroosmotic flow, and

the Stokes equation and mass continuity

5 � �6 + ���� �� ∙ 5 0.electroosmotic flow field is modeled in

COMSOL using the “Creeping Flow” module.

Finally, Ohm’s Law and current continuity are

modeled with the “Electric Currents” module

describe the electric field within the fluid

8; 8 ��� ∙ S 0,is the local current density,

electrical conductivity, and U is the externally

. The electrical conductivity

depends on the local number of charge carriers,

and contains components related to the

convective and conductive currents in the

]. The conductivity can be written in

terms of the local background ion concentrations

� � ��� + ����� ����in (14) represents the current due

to a nonzero charge being advected downstream

by the EOF, while the second term represents the

current generated by the electromigration of

background electrolyte (

the presence of the external field.

numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

and is coupled to

the Stokes equation and mass continuity

���8, (12

flow field is modeled in

COMSOL using the “Creeping Flow” module.

and current continuity are

modeled with the “Electric Currents” module to

describe the electric field within the fluid

�U, (13

is the local current density, *� is the

is the externally

. The electrical conductivity

depends on the local number of charge carriers,

and contains components related to the

currents in the

]. The conductivity can be written in

ckground ion concentrations

� + ���PV. (14

represents the current due

to a nonzero charge being advected downstream

by the EOF, while the second term represents the

current generated by the electromigration of

background electrolyte (BGE) ions in

For thin EDLs,

numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

340,000 mesh elements were used in the various regions to properly resolve the electric double layers, ion transport,

to

2)

flow field is modeled in

COMSOL using the “Creeping Flow” module.

and current continuity are

to

3)

is the

is the externally

. The electrical conductivity

depends on the local number of charge carriers,

and contains components related to the

currents in the

]. The conductivity can be written in

ckground ion concentrations

14)

represents the current due

to a nonzero charge being advected downstream

by the EOF, while the second term represents the

current generated by the electromigration of

ions in

For thin EDLs,

numerical mesh and boundary conditions used in the COMSOL simulations. Approximately

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 5: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

most of

where

compared to the second. However, we are

investigating nanochannels with electrolyte

concentrations sufficiently low to cause

significant overlapping of the EDLs, so the

contribution of the local charge density to the

electrical conductivity

A

340,000 distributed mapped

triangular elements was employed throughout

the channel and reservoirs

was used to

gradients withi

regions while still allowing for computations that

were both time and memory efficient.

depicts

and boundary conditions for the nanochannel and

supplying reservoirs.

simulations was KCl, with

and Pdiffusivity was

unless otherwise noted, the sample charge was

fixed at

4. Results

In order to verify that our numerical model

produces reasonable results, we first compare the

2D and 1D models for a relatively simple case of

non-overlapping EDLs. We simulated a 0.5 mm

long, 100 nm tall channel filled with a solution

of 2.5mM KCl, corres

of about

potential

EOF with a nominal electric field of 10,000 V/m.

The unmodified and modified zeta potentials

were -

Figure

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

region 2 (red). Discrepancies in (d) the centerline BGE

most of the fluid

where �� ��, and the first ter

compared to the second. However, we are

investigating nanochannels with electrolyte

concentrations sufficiently low to cause

significant overlapping of the EDLs, so the

contribution of the local charge density to the

electrical conductivity

finite element mesh of approximately

340,000 distributed mapped

triangular elements was employed throughout

the channel and reservoirs

was used to ensure proper resolution of fine

gradients within the EDLs and near transition

regions while still allowing for computations that

re both time and memory efficient.

depicts a diagram of the 2D model with the mesh

and boundary conditions for the nanochannel and

supplying reservoirs.

simulations was KCl, with PWX 2.03 x 10

diffusivity was fixed at

unless otherwise noted, the sample charge was

fixed at z = -2.

Results

In order to verify that our numerical model

produces reasonable results, we first compare the

2D and 1D models for a relatively simple case of

overlapping EDLs. We simulated a 0.5 mm

long, 100 nm tall channel filled with a solution

of 2.5mM KCl, corres

of about λ! 6.1 nm. An externally applied

potential of 5 V was used to drive the simulated

EOF with a nominal electric field of 10,000 V/m.

The unmodified and modified zeta potentials

-25 mV and -5 mV, respectively.

Figure 3: Comparison of thin EDL ana

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

region 2 (red). Discrepancies in (d) the centerline BGE

the fluid remains electrically neutral

and the first term is negligible

compared to the second. However, we are

investigating nanochannels with electrolyte

concentrations sufficiently low to cause

significant overlapping of the EDLs, so the

contribution of the local charge density to the

electrical conductivity cannot be neglected [

finite element mesh of approximately

340,000 distributed mapped elements and

triangular elements was employed throughout

the channel and reservoirs. This custom mesh

ensure proper resolution of fine

n the EDLs and near transition

regions while still allowing for computations that

re both time and memory efficient.

a diagram of the 2D model with the mesh

and boundary conditions for the nanochannel and

supplying reservoirs. The electrolyte us

simulations was KCl, with PY 1.96 x 10

2.03 x 10-9 m2/s.

fixed at PZ 1 x 10

unless otherwise noted, the sample charge was

In order to verify that our numerical model

produces reasonable results, we first compare the

2D and 1D models for a relatively simple case of

overlapping EDLs. We simulated a 0.5 mm

long, 100 nm tall channel filled with a solution

of 2.5mM KCl, corresponding to a Debye length

6.1 nm. An externally applied

of 5 V was used to drive the simulated

EOF with a nominal electric field of 10,000 V/m.

The unmodified and modified zeta potentials

5 mV, respectively.

Comparison of thin EDL ana

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

region 2 (red). Discrepancies in (d) the centerline BGE

remains electrically neutral

m is negligible

compared to the second. However, we are

investigating nanochannels with electrolyte

concentrations sufficiently low to cause

significant overlapping of the EDLs, so the

contribution of the local charge density to the

cannot be neglected [19].

finite element mesh of approximately

elements and

triangular elements was employed throughout

This custom mesh

ensure proper resolution of fine

n the EDLs and near transition

regions while still allowing for computations that

re both time and memory efficient. Figure 2

a diagram of the 2D model with the mesh

and boundary conditions for the nanochannel and

olyte used in all

1.96 x 10-9 m2/s

The sample

1 x 10-9 m2/s and,

unless otherwise noted, the sample charge was

In order to verify that our numerical model

produces reasonable results, we first compare the

2D and 1D models for a relatively simple case of

overlapping EDLs. We simulated a 0.5 mm

long, 100 nm tall channel filled with a solution

ponding to a Debye length

6.1 nm. An externally applied

of 5 V was used to drive the simulated

EOF with a nominal electric field of 10,000 V/m.

The unmodified and modified zeta potentials

5 mV, respectively.

Comparison of thin EDL analytical (solid lines) vs. simulation results (open circles). The profiles show

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

region 2 (red). Discrepancies in (d) the centerline BGE

remains electrically neutral

m is negligible

compared to the second. However, we are

investigating nanochannels with electrolyte

concentrations sufficiently low to cause

significant overlapping of the EDLs, so the

contribution of the local charge density to the

finite element mesh of approximately

elements and

triangular elements was employed throughout

This custom mesh

ensure proper resolution of fine

n the EDLs and near transition

regions while still allowing for computations that

Figure 2

a diagram of the 2D model with the mesh

and boundary conditions for the nanochannel and

ed in all

/s

The sample

/s and,

unless otherwise noted, the sample charge was

In order to verify that our numerical model

produces reasonable results, we first compare the

2D and 1D models for a relatively simple case of

overlapping EDLs. We simulated a 0.5 mm

long, 100 nm tall channel filled with a solution

ponding to a Debye length

6.1 nm. An externally applied

of 5 V was used to drive the simulated

EOF with a nominal electric field of 10,000 V/m.

The unmodified and modified zeta potentials

Figure

vertical profiles of the EDL potential, BGE salt

concentration

centerline BGE salt concentration

axial transport equation

The profiles demonst

between the simulated and analytical results,

with perhaps the slight exception of the

centerline profile near the zeta potential

transitions. This is to be expected, however, as

the 1D analytical model cannot account for

localized

dispersion near the transition

We validated our model further for a case

with thick electric double layers, comparing

various profiles with the 1D analytical model as

the zeta potential in the middle of the channel

was varied

show exce

charge density profiles, although the velocity

profiles seem to increasingly deviate for the

more uniform cases. This is likely due to the fact

that the analytical r

difference between channel inlet and outlet,

whereas in the simulation and in

situation

decelerating in the respective

cause an

channel

would become more prominent for thicker EDL

cases due to the nonzero el

everywhere, as well as for faster

hence the higher deviation for the more uniform

cases.

Ou

that the sample stacking only occurs when the

nonuniformities in velocity, potential, and ion

distributions throughout the channel are

lytical (solid lines) vs. simulation results (open circles). The profiles show

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

region 2 (red). Discrepancies in (d) the centerline BGE ion profile are likely due to 2D effects such as dispersion.

Figure 3 shows a comparison of the resulting

vertical profiles of the EDL potential, BGE salt

concentration, and

centerline BGE salt concentration

axial transport equation

The profiles demonst

between the simulated and analytical results,

with perhaps the slight exception of the

centerline profile near the zeta potential

transitions. This is to be expected, however, as

the 1D analytical model cannot account for

localized 2D effects such as flow focusing and

dispersion near the transition

We validated our model further for a case

with thick electric double layers, comparing

various profiles with the 1D analytical model as

the zeta potential in the middle of the channel

as varied in a 0.01mM solution

show excellent agreement for the

charge density profiles, although the velocity

profiles seem to increasingly deviate for the

more uniform cases. This is likely due to the fact

that the analytical r

difference between channel inlet and outlet,

reas in the simulation and in

situations the effect of the fluid accelerating and

decelerating in the respective

cause an adverse pressure

channel and slow the fluid slightly. This effect

would become more prominent for thicker EDL

cases due to the nonzero el

everywhere, as well as for faster

hence the higher deviation for the more uniform

cases.

Our simulation results in Figure 4 also show

that the sample stacking only occurs when the

nonuniformities in velocity, potential, and ion

distributions throughout the channel are

lytical (solid lines) vs. simulation results (open circles). The profiles show

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

ion profile are likely due to 2D effects such as dispersion.

shows a comparison of the resulting

vertical profiles of the EDL potential, BGE salt

and axial velocity. A plot of the

centerline BGE salt concentration

axial transport equation from [7] is also included.

The profiles demonstrate excellent agreement

between the simulated and analytical results,

with perhaps the slight exception of the

centerline profile near the zeta potential

transitions. This is to be expected, however, as

the 1D analytical model cannot account for

2D effects such as flow focusing and

dispersion near the transitions.

We validated our model further for a case

with thick electric double layers, comparing

various profiles with the 1D analytical model as

the zeta potential in the middle of the channel

in a 0.01mM solution

llent agreement for the

charge density profiles, although the velocity

profiles seem to increasingly deviate for the

more uniform cases. This is likely due to the fact

that the analytical results assumed zero pressure

difference between channel inlet and outlet,

reas in the simulation and in

the effect of the fluid accelerating and

decelerating in the respective reservoirs would

adverse pressure gradient

and slow the fluid slightly. This effect

would become more prominent for thicker EDL

cases due to the nonzero electric body force

everywhere, as well as for faster

hence the higher deviation for the more uniform

r simulation results in Figure 4 also show

that the sample stacking only occurs when the

nonuniformities in velocity, potential, and ion

distributions throughout the channel are

lytical (solid lines) vs. simulation results (open circles). The profiles show

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

ion profile are likely due to 2D effects such as dispersion.

shows a comparison of the resulting

vertical profiles of the EDL potential, BGE salt

velocity. A plot of the

centerline BGE salt concentration based on a 1D

is also included.

rate excellent agreement

between the simulated and analytical results,

with perhaps the slight exception of the

centerline profile near the zeta potential

transitions. This is to be expected, however, as

the 1D analytical model cannot account for

2D effects such as flow focusing and

We validated our model further for a case

with thick electric double layers, comparing

various profiles with the 1D analytical model as

the zeta potential in the middle of the channel

in a 0.01mM solution. The results

llent agreement for the potential and

charge density profiles, although the velocity

profiles seem to increasingly deviate for the

more uniform cases. This is likely due to the fact

esults assumed zero pressure

difference between channel inlet and outlet,

reas in the simulation and in experimental

the effect of the fluid accelerating and

reservoirs would

gradient across the

and slow the fluid slightly. This effect

would become more prominent for thicker EDL

ectric body force

everywhere, as well as for faster-moving flows;

hence the higher deviation for the more uniform

r simulation results in Figure 4 also show

that the sample stacking only occurs when the

nonuniformities in velocity, potential, and ion

distributions throughout the channel are

lytical (solid lines) vs. simulation results (open circles). The profiles show

excellent agreement for (a) the transverse potential, (b) BGE ion, and (c) velocity profiles in both region 1 (blue) and

ion profile are likely due to 2D effects such as dispersion.

shows a comparison of the resulting

vertical profiles of the EDL potential, BGE salt

velocity. A plot of the

1D

is also included.

rate excellent agreement

between the simulated and analytical results,

with perhaps the slight exception of the

centerline profile near the zeta potential

transitions. This is to be expected, however, as

the 1D analytical model cannot account for

2D effects such as flow focusing and

We validated our model further for a case

with thick electric double layers, comparing

various profiles with the 1D analytical model as

the zeta potential in the middle of the channel

. The results

potential and

charge density profiles, although the velocity

profiles seem to increasingly deviate for the

more uniform cases. This is likely due to the fact

esults assumed zero pressure

difference between channel inlet and outlet,

experimental

the effect of the fluid accelerating and

reservoirs would

across the

and slow the fluid slightly. This effect

would become more prominent for thicker EDL

ectric body force

moving flows;

hence the higher deviation for the more uniform

r simulation results in Figure 4 also show

that the sample stacking only occurs when the

nonuniformities in velocity, potential, and ion

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 6: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

sufficiently large. This requires thick electric

double layers, such that the background ion

distributions and resulting electromigrative

fluxes are modified sufficiently near the

centerline.

electrostatic interactions of the EDL are confined

near the wall and a majority of the fluid is at the

bulk concentration and bulk conductivity, which

leads to uniform electric fields and sample

fluxes. Additionally, a cha

potential and conductivity distributions will have

a uniform electrochemical potential and not

allow for any local ion enrichment or depletion

along the length of the channel. A nonuniform

surface potential distribution and thick EDL

thus essential to induce the heterogeneity

required within the single buffer solution to

produce the gradients that lead to stacking.

Highly charged sample ions are more strongly

affected by these field gradients and can be

effectively concentrated t

Figure 5 depicts the

2D sample concentration profile within the

channel, showing

sample

step change in zeta potential occurs.

thick electr

repelled towards the center of the channel,

increasing the local enhancement along the

centerline.

the centerline is nonzero due to the thick EDLs,

so there is an additional

along the channel

one region to the next.

Figure

(a) the potential, (b) the charge density, and (c) the

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

potential and sample charge. The zeta potential in region 1 was fixed a

sufficiently large. This requires thick electric

double layers, such that the background ion

distributions and resulting electromigrative

fluxes are modified sufficiently near the

centerline. For thin electric double layers, the

electrostatic interactions of the EDL are confined

near the wall and a majority of the fluid is at the

bulk concentration and bulk conductivity, which

leads to uniform electric fields and sample

fluxes. Additionally, a cha

potential and conductivity distributions will have

a uniform electrochemical potential and not

allow for any local ion enrichment or depletion

along the length of the channel. A nonuniform

surface potential distribution and thick EDL

thus essential to induce the heterogeneity

required within the single buffer solution to

produce the gradients that lead to stacking.

Highly charged sample ions are more strongly

affected by these field gradients and can be

effectively concentrated t

Figure 5 depicts the

2D sample concentration profile within the

channel, showing

sample stacks near an interface where a smooth

step change in zeta potential occurs.

thick electric double layers,

repelled towards the center of the channel,

increasing the local enhancement along the

centerline. The intrinsic EDL potential

the centerline is nonzero due to the thick EDLs,

so there is an additional

along the channel as the flow transitions from

one region to the next.

Figure 4: Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

(a) the potential, (b) the charge density, and (c) the

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

potential and sample charge. The zeta potential in region 1 was fixed a

sufficiently large. This requires thick electric

double layers, such that the background ion

distributions and resulting electromigrative

fluxes are modified sufficiently near the

thin electric double layers, the

electrostatic interactions of the EDL are confined

near the wall and a majority of the fluid is at the

bulk concentration and bulk conductivity, which

leads to uniform electric fields and sample

fluxes. Additionally, a channel with uniform wall

potential and conductivity distributions will have

a uniform electrochemical potential and not

allow for any local ion enrichment or depletion

along the length of the channel. A nonuniform

surface potential distribution and thick EDL

thus essential to induce the heterogeneity

required within the single buffer solution to

produce the gradients that lead to stacking.

Highly charged sample ions are more strongly

affected by these field gradients and can be

effectively concentrated to higher levels.

Figure 5 depicts the temporal evolution of

2D sample concentration profile within the

channel, showing how a negatively charged

near an interface where a smooth

step change in zeta potential occurs.

ic double layers, the sample is

repelled towards the center of the channel,

increasing the local enhancement along the

The intrinsic EDL potential

the centerline is nonzero due to the thick EDLs,

so there is an additional axial potentia

as the flow transitions from

one region to the next. This additional potential

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

(a) the potential, (b) the charge density, and (c) the

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

potential and sample charge. The zeta potential in region 1 was fixed a

sufficiently large. This requires thick electric

double layers, such that the background ion

distributions and resulting electromigrative

fluxes are modified sufficiently near the

thin electric double layers, the

electrostatic interactions of the EDL are confined

near the wall and a majority of the fluid is at the

bulk concentration and bulk conductivity, which

leads to uniform electric fields and sample

nnel with uniform wall

potential and conductivity distributions will have

a uniform electrochemical potential and not

allow for any local ion enrichment or depletion

along the length of the channel. A nonuniform

surface potential distribution and thick EDLs are

thus essential to induce the heterogeneity

required within the single buffer solution to

produce the gradients that lead to stacking.

Highly charged sample ions are more strongly

affected by these field gradients and can be

o higher levels.

temporal evolution of a

2D sample concentration profile within the

how a negatively charged

near an interface where a smooth

step change in zeta potential occurs. Due to the

the sample is

repelled towards the center of the channel,

increasing the local enhancement along the

The intrinsic EDL potential Ψ along

the centerline is nonzero due to the thick EDLs,

axial potential gradient

as the flow transitions from

This additional potential

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

(a) the potential, (b) the charge density, and (c) the velocity profiles for varying zeta potentials in region 2. Sample

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

potential and sample charge. The zeta potential in region 1 was fixed a

sufficiently large. This requires thick electric

double layers, such that the background ion

distributions and resulting electromigrative

fluxes are modified sufficiently near the

thin electric double layers, the

electrostatic interactions of the EDL are confined

near the wall and a majority of the fluid is at the

bulk concentration and bulk conductivity, which

leads to uniform electric fields and sample

nnel with uniform wall

potential and conductivity distributions will have

a uniform electrochemical potential and not

allow for any local ion enrichment or depletion

along the length of the channel. A nonuniform

s are

thus essential to induce the heterogeneity

required within the single buffer solution to

produce the gradients that lead to stacking.

Highly charged sample ions are more strongly

affected by these field gradients and can be

a

2D sample concentration profile within the

how a negatively charged

near an interface where a smooth

Due to the

the sample is

repelled towards the center of the channel,

increasing the local enhancement along the

along

the centerline is nonzero due to the thick EDLs,

gradient

as the flow transitions from

This additional potential

gradient

axial gradient in the applied potential, and the net

electric field can change direction locally

result

electrophoretic trapping region, as shown in

Figure 5.

diffusion

the transport of the sample through the channel.

The diffusi

electromigrative flux and the

the enhancement region

steady state.

transition, the opposite

potential gradient

external field

accelerated, creating a local

5. Conclusions

In this work, we use

Multiphysics to model the 2D electroosmotic

flow of a

electrophoretic stacking of sample ions in a

nanochannel with selectively modified surface

charge. We show

and focusing effects only occur in channels with

sufficiently large step changes in zeta potential,

and for cases i

large relative to the channel height. Our

approach can potentially achieve

hundred

higher than those limited by conductivity ratios

in conventional FASS[

field, potential,

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

velocity profiles for varying zeta potentials in region 2. Sample

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

potential and sample charge. The zeta potential in region 1 was fixed a

gradient can be larger in magnitude than the

axial gradient in the applied potential, and the net

electric field can change direction locally

result. This enhances the s

electrophoretic trapping region, as shown in

Figure 5. Once the

diffusion becomes a significant driving factor in

the transport of the sample through the channel.

The diffusive flux

electromigrative flux and the

the enhancement region

steady state. Near the other zeta potential

transition, the opposite

potential gradient is in the same

external field. Sample ions here are

accelerated, creating a local

Conclusions

In this work, we use

Multiphysics to model the 2D electroosmotic

flow of a dilute background electrolyte and

electrophoretic stacking of sample ions in a

nanochannel with selectively modified surface

charge. We showed

and focusing effects only occur in channels with

sufficiently large step changes in zeta potential,

and for cases in which the EDLs are sufficiently

large relative to the channel height. Our

approach can potentially achieve

hundred-fold sample enhancement ratios, notably

higher than those limited by conductivity ratios

in conventional FASS[

, potential, and ion distributions were

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

velocity profiles for varying zeta potentials in region 2. Sample

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

potential and sample charge. The zeta potential in region 1 was fixed at -100 mV for these simulations.

can be larger in magnitude than the

axial gradient in the applied potential, and the net

electric field can change direction locally

This enhances the stacking

electrophoretic trapping region, as shown in

nce the sample starts to

becomes a significant driving factor in

the transport of the sample through the channel.

ve flux eventually balances the

electromigrative flux and the convective flux in

the enhancement region as the system reaches a

Near the other zeta potential

transition, the opposite effect occurs as the

is in the same

Sample ions here are

accelerated, creating a localized depletion region.

In this work, we use

Multiphysics to model the 2D electroosmotic

background electrolyte and

electrophoretic stacking of sample ions in a

nanochannel with selectively modified surface

ed that these particular stacking

and focusing effects only occur in channels with

sufficiently large step changes in zeta potential,

n which the EDLs are sufficiently

large relative to the channel height. Our

approach can potentially achieve

fold sample enhancement ratios, notably

higher than those limited by conductivity ratios

in conventional FASS[7,8]. Resulting veloc

and ion distributions were

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

velocity profiles for varying zeta potentials in region 2. Sample

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

100 mV for these simulations.

can be larger in magnitude than the

axial gradient in the applied potential, and the net

electric field can change direction locally as

acking and creates an

electrophoretic trapping region, as shown in

sample starts to accumulate,

becomes a significant driving factor in

the transport of the sample through the channel.

eventually balances the

convective flux in

as the system reaches a

Near the other zeta potential

effect occurs as the EDL

is in the same direction as the

Sample ions here are further

depletion region.

In this work, we used COMSOL

Multiphysics to model the 2D electroosmotic

background electrolyte and the

electrophoretic stacking of sample ions in a

nanochannel with selectively modified surface

that these particular stacking

and focusing effects only occur in channels with

sufficiently large step changes in zeta potential,

n which the EDLs are sufficiently

large relative to the channel height. Our

approach can potentially achieve several

fold sample enhancement ratios, notably

higher than those limited by conductivity ratios

]. Resulting velocity

and ion distributions were

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

velocity profiles for varying zeta potentials in region 2. Sample

enhancement ratios are shown in (d) for varying zeta potential ratios and EDL thicknesses, and in (e) for varying zeta

100 mV for these simulations.

can be larger in magnitude than the

axial gradient in the applied potential, and the net

a

and creates an

electrophoretic trapping region, as shown in

accumulate,

becomes a significant driving factor in

the transport of the sample through the channel.

eventually balances the

convective flux in

as the system reaches a

Near the other zeta potential

EDL

as the

further

depletion region.

COMSOL

Multiphysics to model the 2D electroosmotic

the

electrophoretic stacking of sample ions in a

nanochannel with selectively modified surface

that these particular stacking

and focusing effects only occur in channels with

sufficiently large step changes in zeta potential,

n which the EDLs are sufficiently

large relative to the channel height. Our

several

fold sample enhancement ratios, notably

higher than those limited by conductivity ratios

ity

Simulation results for thick EDLs. Analytical (solid lines) vs. simulation (solid circles) results are shown for

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 7: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

Figure

electrophoretic trapping of particles in a simulation, (c) steady state fluxes along the c

averaged time

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

locations in (c),

validated through comparison with existing

analytical techniques to solve the 1D Poisson

Boltzmann equation, Stokes equation, and

Nernst

These results provide encouraging

indications that it is possible to perform

stationary field

preconcentration in nanochannels without using

multiple electrolyte solutions, but by simply

inducing electric field gradients through

tailoring of

embedded electrodes. The resulting enhancement

in our simulations

microchannel

rivals

11], revealing

sample preconcentration in nanofluidic devices.

6. References 1. Yu

nanofluidic concentrator

(2015)

2. Hsu W, Harvie D J, Davidson M R, Jeong H,

Goldys E M, and Ing

focusing and separation in a silica nanofluidic channel

with a non

14, 3539

3. Stein D, Deurvorst Z, van der Heyden F H J,

Koopmans W

Electrokinetic Concentration of DNA Polymers in

Nanofluidic Channels

Figure 5: (a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

electrophoretic trapping of particles in a simulation, (c) steady state fluxes along the c

averaged time-dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

ons in (c), and flux lines near the electrophoretic trap in

validated through comparison with existing

analytical techniques to solve the 1D Poisson

Boltzmann equation, Stokes equation, and

Nernst-Planck equation [

These results provide encouraging

indications that it is possible to perform

stationary field

preconcentration in nanochannels without using

multiple electrolyte solutions, but by simply

inducing electric field gradients through

ing of wall surface charge uniformity via

embedded electrodes. The resulting enhancement

in our simulations can

microchannel-based

rivals that of similar

revealing another pro

sample preconcentration in nanofluidic devices.

. References

1. Yu M, Hou Y, Zhou H, and Yao S,

nanofluidic concentrator

2. Hsu W, Harvie D J, Davidson M R, Jeong H,

Goldys E M, and Inglis

focusing and separation in a silica nanofluidic channel

with a non-uniform electroosmotic flow

3539-49 (2014)

3. Stein D, Deurvorst Z, van der Heyden F H J,

Koopmans W J A, Gabel A, and Dekker C,

Electrokinetic Concentration of DNA Polymers in

Nanofluidic Channels Nano Letters

(a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

electrophoretic trapping of particles in a simulation, (c) steady state fluxes along the c

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

and flux lines near the electrophoretic trap in

validated through comparison with existing

analytical techniques to solve the 1D Poisson

Boltzmann equation, Stokes equation, and

Planck equation [7,18].

These results provide encouraging

indications that it is possible to perform

stationary field-amplified sample

preconcentration in nanochannels without using

multiple electrolyte solutions, but by simply

inducing electric field gradients through

wall surface charge uniformity via

embedded electrodes. The resulting enhancement

can exceed that of

stacking mechanisms

that of similar nanoscale methods [

another promising technique for

sample preconcentration in nanofluidic devices.

M, Hou Y, Zhou H, and Yao S,

nanofluidic concentrator Lab on Chip

2. Hsu W, Harvie D J, Davidson M R, Jeong H,

lis D W, Concentration gradient

focusing and separation in a silica nanofluidic channel

uniform electroosmotic flow

3. Stein D, Deurvorst Z, van der Heyden F H J,

J A, Gabel A, and Dekker C,

Electrokinetic Concentration of DNA Polymers in

Nano Letters 10

(a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

electrophoretic trapping of particles in a simulation, (c) steady state fluxes along the c

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

and flux lines near the electrophoretic trap in

validated through comparison with existing

analytical techniques to solve the 1D Poisson-

Boltzmann equation, Stokes equation, and

These results provide encouraging

indications that it is possible to perform

amplified sample

preconcentration in nanochannels without using

multiple electrolyte solutions, but by simply

inducing electric field gradients through the

wall surface charge uniformity via

embedded electrodes. The resulting enhancement

that of traditional

mechanisms and

methods [1,2,7-

mising technique for

sample preconcentration in nanofluidic devices.

M, Hou Y, Zhou H, and Yao S, An on-demand

Lab on Chip 15, 1524-32

2. Hsu W, Harvie D J, Davidson M R, Jeong H,

Concentration gradient

focusing and separation in a silica nanofluidic channel

uniform electroosmotic flow Lab on Chip

3. Stein D, Deurvorst Z, van der Heyden F H J,

J A, Gabel A, and Dekker C,

Electrokinetic Concentration of DNA Polymers in

10, 765-72 (2010)

(a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

electrophoretic trapping of particles in a simulation, (c) steady state fluxes along the c

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

and flux lines near the electrophoretic trap in (b).

validated through comparison with existing

Boltzmann equation, Stokes equation, and

These results provide encouraging

indications that it is possible to perform

amplified sample

preconcentration in nanochannels without using

multiple electrolyte solutions, but by simply

the

wall surface charge uniformity via

embedded electrodes. The resulting enhancement

traditional

and

-

mising technique for

demand

32

2. Hsu W, Harvie D J, Davidson M R, Jeong H,

Concentration gradient

focusing and separation in a silica nanofluidic channel

Lab on Chip

3. Stein D, Deurvorst Z, van der Heyden F H J,

J A, Gabel A, and Dekker C,

Electrokinetic Concentration of DNA Polymers in

4. Napoli M, Ei

Nanofluidic technology for biomolecule applications,

a critical review

5. Kim

concentration devices for biomolecules utilizing ion

concentration polarization: theory, fabrication, and

applications

6. Sparreboom

Principles and

Nature Nanotechnology

7.

amplified sample stacking and focusing in nanofluidic

channels

8.

field-

57-92

9. Hsu

Davidson

Silica Nanofluidic Channel: Effects of

Electromigratio

86, 8711

10.

Electropreconcentration with Charge

Nanochannels,

11.

van der Linden

Isotachophoresis Using a Nanochannel

Depletion Zone,

12. Herr

Kenny

Nonuniform Zeta Potential,

(2000)

13.

flow in micro/nanochannels with surface potential

heterogeneity: An analysis through the Nernst

(a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

electrophoretic trapping of particles in a simulation, (c) steady state fluxes along the channel centerline, and (d) area

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

4. Napoli M, Ei

Nanofluidic technology for biomolecule applications,

a critical review Lab on Chip

Kim S J, Song Y, and Han J,

concentration devices for biomolecules utilizing ion

concentration polarization: theory, fabrication, and

applications Chem. Soc. Rev.

Sparreboom W, van den Berg

Principles and applications of nanofluidic transport,

Nature Nanotechnology

Sustarich J, Storey B, and Pennathur S

amplified sample stacking and focusing in nanofluidic

channels Phys Fluids

Bharadwaj R and Santiago J G

-amplified sample stacking

92 (2005)

Hsu W, Inglis D W,

Davidson M R, Harvie

Silica Nanofluidic Channel: Effects of

Electromigration and Electroosmosis,

8711-18 (2014)

. Plecis A, Nanteuil

Electropreconcentration with Charge

Nanochannels, Anal. Chem.,

. Quist J, Janssen

van der Linden

Isotachophoresis Using a Nanochannel

Depletion Zone, Anal. Chem.

Herr A E, Molho

Kenny T W, Electroosmotic Capillary Flow with

Nonuniform Zeta Potential,

(2000)

. Bhattacharyya

flow in micro/nanochannels with surface potential

heterogeneity: An analysis through the Nernst

(a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

hannel centerline, and (d) area

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the t

4. Napoli M, Eijkel J C T, and Pennathur S,

Nanofluidic technology for biomolecule applications,

Lab on Chip 10, 957

S J, Song Y, and Han J,

concentration devices for biomolecules utilizing ion

concentration polarization: theory, fabrication, and

Chem. Soc. Rev. 39, 912

W, van den Berg

applications of nanofluidic transport,

Nature Nanotechnology, 4, 713-20 (2009)

Sustarich J, Storey B, and Pennathur S

amplified sample stacking and focusing in nanofluidic

Phys Fluids 22, 112003 (2010)

R and Santiago J G

amplified sample stacking J. Fluid Mech.

D W, Startsev M A,

Harvie D J, Isoelectric Focusing in a

Silica Nanofluidic Channel: Effects of

n and Electroosmosis,

Nanteuil C, Haghiri-Gosnet

Electropreconcentration with Charge

Anal. Chem., 80, 9542

Janssen K G H, Vulto

van der Linden H J, Single

Isotachophoresis Using a Nanochannel

Anal. Chem., 83, 7910

Molho J I, Santiago J G,

, Electroosmotic Capillary Flow with

Nonuniform Zeta Potential, Anal. Chem.,

S, Nayak A K

flow in micro/nanochannels with surface potential

heterogeneity: An analysis through the Nernst

(a) Transient evolution of 2D sample concentration profile within the channel, (b) a diagram depicting

hannel centerline, and (d) area

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

gradient is visible in the positive electrophoretic flux of the negative sample in (d), the peaks near the transition

jkel J C T, and Pennathur S,

Nanofluidic technology for biomolecule applications,

957-85 (2010)

S J, Song Y, and Han J, Nanofluidic

concentration devices for biomolecules utilizing ion

concentration polarization: theory, fabrication, and

912-22 (2010)

A, Eijkel J C T

applications of nanofluidic transport,

20 (2009)

Sustarich J, Storey B, and Pennathur S, Field

amplified sample stacking and focusing in nanofluidic

(2010)

R and Santiago J G, Dynamics of

J. Fluid Mech. 543

M A, Goldys E M, M

, Isoelectric Focusing in a

Silica Nanofluidic Channel: Effects of

n and Electroosmosis, Anal. Chem.,

Gosnet A, Chen Y

Electropreconcentration with Charge-Selective

, 9542-50 (2008)

P, Hankemeier T,

, Single-Electrolyte

Isotachophoresis Using a Nanochannel-Induced

7910-15 (2011)

J G, Mungal M G,

, Electroosmotic Capillary Flow with

Anal. Chem., 72, 1053-57

A K, Electroosmotic

flow in micro/nanochannels with surface potential

heterogeneity: An analysis through the Nernst-Planck

dependent sample fluxes at the location of maximum enhancement. The effect of the axial EDL potential

jkel J C T, and Pennathur S,

Nanofluidic technology for biomolecule applications,

Nanofluidic

concentration devices for biomolecules utilizing ion

concentration polarization: theory, fabrication, and

J C T,

applications of nanofluidic transport,

Field-

amplified sample stacking and focusing in nanofluidic

Dynamics of

543,

E M, M

, Isoelectric Focusing in a

Silica Nanofluidic Channel: Effects of

Anal. Chem.,

Y,

Selective

T,

Electrolyte

Induced

M G,

, Electroosmotic Capillary Flow with

57

, Electroosmotic

flow in micro/nanochannels with surface potential

Planck

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston

Page 8: Sample Preconcentration in Nanochannels with …Sample Preconcentration in Nanochannels with Nonuniform Surface Charge and Thick Electric Double Layers A. Eden*1, C. McCallum1, B

model with convection effect, Colloids and Surfaces

A: Physicochem. Eng. Aspects, 339, 167-77 (2009)

14. Hughes C, Yeh L, Qian S, Field Effect

Modulation of Surface Charge Property And

Electroosmotic Flow in a Nanochannel: Stern Layer

Effect, J. Phys. Chem., 117, 9322-31 (2013)

15. Daiguji H, Ion transport in nanofluidic channels,

Chem. Soc. Rev., 39, 901-11 (2010)

16. Cheng L, Guo L J, Nanofluidic diodes, Chem.

Soc. Rev., 39, 923-38 (2010)

17. Guan W, Li S X, Reed M A, Voltage gated ion

and molecule transport in engineered nanochannels:

theory, fabrication and applications, Nanotechnology,

25, 122001 (2014)

18. Baldessari F, Electrokinetics in nanochannels

Part I. Electric double layer overlap and channel-to-

well-equilibrium, J. Colloid and Interface Science,

325(2) , 526-38 (2008)

19. McCallum C, Pennathur S, Accounting for

electric double layer and pressure gradient-induced

dispersion effects in microfluidic current monitoring,

Microfluid Nanofluid, 20:13 (2016)

7. Acknowledgements

We greatly acknowledge our funding

sources, the Institute for Collaborative

Biotechnologies through grant W911NF-09-001

from the U.S. Army Research Office, and grant

DAAD19-03-D-0004 from the U.S. Army

Research Office.

Excerpt from the Proceedings of the 2016 COMSOL Conference in Boston