An official website of the United States government
Here's how you know
Official websites use .gov
A
.gov website belongs to an official
government organization in the United States.
Secure .gov websites use HTTPS
A lock (
) or https:// means you've safely
connected to the .gov website. Share sensitive
information only on official, secure websites.
As a library, NLM provides access to scientific literature. Inclusion in an NLM database does not imply endorsement of, or agreement with,
the contents by NLM or the National Institutes of Health.
Learn more:
PMC Disclaimer
|
PMC Copyright Notice
Permits the broadest form of re-use including for commercial purposes, provided that author attribution and integrity are maintained (https://creativecommons.org/licenses/by/4.0/).
Hybrid nanoplasmonic structures composed of subwavelength
apertures
in metallic films and nanoparticles have recently been demonstrated
as ultrasensitive plasmonic sensors. This work investigates the electrokinetically
driven propagation of the assembly mechanism of the metallic nanoparticles
through nanoapertures. The Debye–Hückel approximation
for a symmetric electrolyte solution with overlapping electrical double
layers (EDLs) is used to obtain an analytical solution to the problem.
The long-term silver nanoparticle concentration response is derived
using the homogenization method and a multiscale analysis. The results
indicate that uncharged nanoparticles will flow through the nanohole
array if the nanochannel height is larger than the Debye length (h0 > λD), while a trapping
mechanism
occurs, due to the overlapping of the EDL, when h0 ∼ 3.8λD. For charged nanoparticles,
the response to the electric field occurs locally with the walls of
the nanochannel, regardless of its height. For a critical value of
the nanochannel length, the leading order of the concentration field
becomes purely diffusive.
Introduction
Optofluidics, the emergent field arising
from the synergistic combination
of photonics and microfluidics, has enabled the development of ultracompact
devices with applications in different fields, such as label-free
sensing.1 Ordered arrays subwavelength
apertures, known as nanohole arrays (NHAs), have been demonstrated
as nanoplasmonic sensors that can be fabricated with accurate periodicities
and aperture shapes due to recent advances in nanofabrication techniques.2 NHAs feature relatively long channels (in the
order of 101 to 102 nm)3,4 with
customizable pore sizes.5,6 These nanoapertures
have been used for fluidic transport of considerably small volumes,
and the enrichment of electrocharged analytes results in improved
sensor response7−12 and controllable nanoinjection across the nanostructured substrate.
Unlike microfluidics, nanofluidics involve nanostructures where the
solid boundaries are extremely close, and overlapping of electrical
double layers (EDLs) is possible,13 which
invalidates the use of Boltzmann distribution for ionic charge density.
Recent studies have presented analytical models for overlapped
EDLs. Qu and Li14 derived a model to determine
the electrical potential and ionic concentration distributions between
two infinitely large flat plates, establishing corrected boundary
conditions for these distributions. Golovnev and Trimper15 obtained an analytical solution for the Poisson–Nernst–Planck
equation when Faradaic processes are discarded, revealing different
ion concentration behaviors in both the short- and the long-time regimes.
Zachariah et al.16 analyzed the repulsive
forces that appear during the collapse of the EDLs using the DLVO
theory, concluding that the hydration force is due to multiple layering
of hydrate ions, which subsequently undergo transitions between different
confined adsorbed ion states. Such a case occurs when a microchannel
is connected to a nanochannel, and a micro–nano interface is
formed, leading to charge transport and ion concentration polarization
(ICP).17 ICP enables the enrichment of
charged particles, such as biomarkers and ions, as well as rectification
effects on ionic current.18,19 In this regard, Mani
et al.20 and Yaroshchuk and Bondarenko21 established transport analytical models considering
the area average for ICP, with emphasis on the dominance of axial
diffusion in determining the extent of diffuse layers at micro-to-nanochannel
interfaces when the ion reservoir is large. In the context of sensing,
the transport and placement of metallic nanoparticles within and around
NHAs to construct hybrid nanoplasmonic structures with improved sensitivity
have been demonstrated experimentally.11,22,23 However, the reported analytical work has focused
on ionic concentration and does not address the transport scenarios
with charged colloidal systems.
In this work, we investigate
the role of the EDL overlapping on
both charged and uncharged metallic nanoparticles in nanoconfinement
using an analytical approach. The analysis is focused on the creation
of hybrid nanostructures as surface-enhanced Raman scattering (SERS)
substrates, as previously reported experimentally.11 The investigation involves flow-through metallic nanoapertures
that support surface plasmon resonance and metallic nanoparticles
that naturally exhibit localized surface plasmon resonance, within
a microfluidic environment under an applied electric field. We analyze
the significance of the field distribution and the influence of the
overlapping EDL on it, elucidating the critical impact of some parameters
leading to novel perspectives on the physics of nanoparticle transport.
Hydrodynamic Formulation
The system used in this study
encompasses a microfluidic chip assembly
with an embedded metallic NHA under the influence of a direct current
(DC) electric field applied externally, as shown in Figure 1a. A Newtonian fluid is assumed
to contain small ions (average size ∼1 Å) and large silver
nanoparticles (Ag NPs), flowing through an NHA. Both the ionic and
Ag NP concentrations are initially uniform throughout the system.
An electroosmotic flow (EF) emerges from an externally imposed electric
field and the induced field in the EDL. The imposed
external field yields the transference of electrons to particles,
resulting in the acquisition of a net negative surface electric charge.
Some particles may undergo polarization while retaining their electrical
neutrality; in this manner, consideration of both charged and uncharged
nanoparticles in the present analysis is warranted. Figure 1b depicts a cross-sectional
view of a single nanohole within the NHA. Notably, silver nanoparticles
exhibit a distinct Gaussian distribution concentration close to the
nanohole’s entrance, a phenomenon attributed to the localized
enhancement of the electric field.8 The
effect of the electric field gradient at the rim of the nanoholes
on both the ions and the Ag NPs is of particular interest for producing
hybrid nanostructures that enable SERS. This system, along with the
assumption of NHA axis symmetry regarding electrokinetic phenomena,
is the main domain for the analytical solution presented in this work.
The nanohole is modeled as an isothermal flat nanochannel of height h0 and length L. A 2D Cartesian
system of coordinates is adopted at the nanochannel left inlet,
where · indicates that the variable has dimensional units. The
extension of the model to cylindrical coordinates is trivial, as reported
by Pennathur and Santiago.3
Governing Equations
The governing equations that serve
as a starting point to investigate the EF in this study are the continuity
equation
1
and the Navier–Stokes equations
2
where ρm is the fluid density, considered constant, is the velocity vector, is the time, is the pressure, ∇ is the Nabla
operator defined as , and is the stress tensor for a Newtonian fluid,
given by
3
where μ is the dynamic
viscosity of the fluid. The hydrodynamic boundary conditions related
to the impermeability and no-slip condition at the walls are given
by
4
and
5
respectively. n represents the unit vector normal to the microelectrode surface
pointing toward the fluid. The pressure is at = 0,L. The electrical
body force in eq 2 consists
of the electric charge density ρe and the total electrical potential , which is governed by Poisson’s
equation
6
where ϵm denotes the dielectric permittivity of the medium. Here, is split into the induced nonuniform equilibrium
potential in the EDL, , and the potential describing the external
electric field , where Ex = ϕ0/L and ϕ0 is the voltage provided by the generator. The boundary condition
for at the walls is = ζ, where ζ is the zeta potential.
The electrical charge density ρe is proportional to the local concentration difference between the
cations and anions. For a symmetric (z/z) electrolyte solution, the charge density can be defined as
7
where z is the valence of
the electrolyte, e represents the fundamental charge
of an electron, and represents the concentration of cations
and anions, respectively. Transport of ions in a dilute solution is
described by the Nernst–Planck equation24
8
where Di is the diffusion coefficient of the ions, kB is the Boltzmann constant, and T is
the absolute temperature. Equation 8 is subject to the following boundary conditions14
9
The concentration field of the diffusing charged NPs is governed
by the convective diffusion equation25
10
where the molecular diffusion coefficient
of nanoparticles, denoted as D, is determined through
the Stokes–Einstein equation26
11
where Rp is the hydrodynamic radius of the nanoparticles. In eq 10, the second right-hand
term corresponds to an electromigration phenomenon, which consists
of nanoparticle motion under the influence of Coulomb force. The Coulomb
force appears between the gradient of the total electrical potential
and the electrically charged NPs. Therefore, this term should only
be considered for charged NPs. The boundary and initial conditions
associated with eq 10 are
12
13
14
where is the rate of disappearance of NPs due
to an irreversible first-order reaction between the solution and the
walls,27C* is the initial
concentration of NPs, , and a is a constant.
In boundary conditions 12 we consider two cases:
(i) the nonpenetration boundary for uncharged nanoparticles and (ii)
the so-called perfect-sink model for charged nanoparticles. The nonpenetration
boundary is obtained when = 0, which specifies that there is no penetration
of the particles at the boundary, ensuring that any change in concentration
at the wall is solely due to diffusion and not advection.28 The boundary acting as a perfect sink, which
is used in theories of diffusion of charged particles,29 is obtained when . This model assumes that all particles
arriving at the wall will be irreversibly adsorbed immediately and
subsequently disappear from the system.30
Nondimensional Mathematical Model
The governing equations
together with their corresponding boundary conditions can be written
in a nondimensional form by introducing the dimensionless variables x = /L, , , , , , , ψ = /ζ, , , and . Here, Uc = ϵmζ2/μL
is the characteristic velocity,31tc = LλD/D is the harmonic time,32 is the Debye length, and n∞ is the ionic number in the concentration in the
bulk solution. When defining the nondimensional pressure p′, we have introduced the useful definition P = p′ – (1/2)(dψ/dy)2 introduced by Ajdari,33 which serves to eliminate the electric terms in the momentum
equation in the y direction. Therefore, the expanded
nondimensional forms of the hydrodynamic, electric, and concentration
governing eqs 1–3, 6–8, and 10 are as follows
15
16
17
18
19
20
In eqs 15–20, ϵ = λD/L, η = h0/L, k = h0/λD, α = −ζ/ϕ0, δ = −ζze/kBT, DT = D/Di, Pe = UcλD/D is the Péclet
number, and Re = ρmUch0/μ is the Reynolds numbers. The dimensionless boundary
conditions of eqs 15–20 are
21
22
23
24
25
26
and
27
where
28
and . In conditions 22, the induced potential gradient is zero at x =
0,1 due to symmetry. In nanofluidic systems, typical values of the
parameters ϵ, DT, and Re are very small (λD = 10
nm, L = 200 nm, Di = 10–9 m2/s, D = 8.5844 × 10–12 m2/s, Uc = 0.002 m/s, ϵ = 0.05, DT = 10–3,
and Re = 2.5 × 10–4), while
η < 1 (η = 0.5). Therefore, a simplified version of
the nondimensional governing equations, as well as a regular perturbation
technique,34 can be used to solve the set
of mentioned equations for small values of the parameter ϵ.
Thus, a regular expansion is proposed for each dependent variable
(say, X) in the following form
29
where X = u, v, P, ψ. Substituting the expansion eq 29 into the nondimensional
governing eqs 15–19, and collecting terms of , we obtain the following problems
30
31
32
33
34
From eq 32, P0 = P0(x) should be determined as a part of
the hydrodynamics problem together
with u0 and v0. The solution of Nernst–Planck eq 34 is given by14
35
and
36
where ψc is the unknown electrical potential at the center of the nanochannel.
In eqs 35 and 36 it is assumed that the concentrations of cations
and anions at the center are the same (n0,+ = n0,– = 1 at y = 0). Further applying the Debye–Hückel approximation
yields
The solution of Poisson eq 39, considering the boundary conditions [eq 23], is given by
40
To recover the Boltzmann equation (ψ
= 0 at y = 0 and k ≫ 1),
we have assumed that ψc = ζ/k, obtaining
the following simplification
41
The general hydrodynamic solution of eq 31 is obtained by substituting eq 41 in the momentum equation
and integrating
the result twice with respect to y; considering eq 21 as the boundary condition,
we obtain that the velocity profile in the lowest order is given as
42
In the above equation, the pressure
gradient is unknown and can
be obtained using the continuity equation, eq 30. The suggested procedure is to substitute eqs 42 into 30, obtaining a solution for v0 and P0. However, one should find that the leading
order for the variables v0 and P0 is zero. Therefore, the above physically means
that there are no induced pressure terms in this order and that the
velocity field is hydrodynamically developed. Therefore, substituting eqs 41 in 42 yields
43
The next step is to assess the solutions for ψ1, u1, v1, and P1. In this order, the ion concentration does
not get affected by convection, thus making the Poisson equation to yield ψ1 = 0, leading to u1 = v1 = P1 = 0 due to its role
as the primary force in the momentum equations.
Homogenization Method
To obtain an analytical solution
for the concentration field of Ag NPs, the homogenization method35 is proposed to derive an expression that allows
us to solve the convective diffusion equation. Thus, three distinct
time scales are involved in the analysis of Ag NPs, which are as follows:
the harmonic time,32t0 ∼ LλD/D, the transverse diffusion time, t1 ∼ 4h02/D, and the longitudinal diffusion time, t2 ∼ L2/D.
From typical values of the previous times (L = 200 nm, h0 = 100 nm, t0 = 2 × 10–4 s, and t1 = t2 = 4 × 10–3 s), the following two time scales can be introduced
44
and using eq 29 to expand for the dimensionless concentration
45
where cj = cj(x,y,t0,t2) and j = 0,1,2. The original
time derivate becomes, according to the chain rule
To obtain a solution for eq 48, we consider two cases: uncharged
nanoparticles, which neglect
the electromigration term (second left-hand term) in eq 48, and charged nanoparticles. In
addition, both cases must satisfy the following boundary condition
due to the symmetry of the nanochannel
49
In both cases, the solution at is c0 = Cx(x,t0,t2). The procedure
that determines the function Cx is given in the lines below. Taking the from eq 47 yields the governing equation for c1
50
The next step is to take the cross-sectional
average of eq 50, defined
as for any function f, where
· will indicate the averaged function. In this context, the first
right-hand term in eq 50 becomes zero as a consequence of its symmetry with respect to the y-axis, which is known from the inflection point at y = 0 [eq 49]. Similarly, the second right-hand term becomes zero due to the
product of two odd functions. Thus, the cross-sectional average of eq 50 is
Considering the linearity of eq 53, the solution c1 can be
expressed as
54
and its substitution in eq 53 leads to a second-order ordinary differential
equation for B(y) as follows
55
where
56
First, eq 55 is solved
for uncharged nanoparticles using boundary condition 25 considering β = 0 as follows
57
Solving eq 55 for
uncharged NPs, neglecting the electromigration term (second left-hand
term) in eq 55, yields
58
For charged nanoparticles (β
≫ 1), eq 55 was
solved using the fourth-order
Runge–Kutta method with the aid of the shooting approach together
with the following boundary condition [eq 25]
Taking the cross-sectional average
of eq 61, the following
governing equation is obtained
62
where
63
For uncharged nanoparticles, was calculated by substituting eqs 56 and 58 in 63, obtaining the following equation
64
For charged nanoparticles, was calculated by using numerical methods.
Finally, eq 51 is added
to 62, where the artifice of two times is no
longer needed and can be removed35
65
where
66
The second right-hand term in eq 66 is part of electromigration
and should only be considered
for charged nanoparticles. First, we propose a solution for Cx that eliminates the convective
term in eq 65, i.e.,
the first right-hand term, as follows
The initial and boundary conditions
of eq 68 are taken from eqs 26–28 as follows
69
and
70
The general solution for the leading
order, using the Fourier method
for eqs 68 and substituting 67, is
71
Results and Discussion
In the Nondimensional
Mathematical Model and Homogenization Method sections, the
nondimensional potential in the EDL, velocity vector, and concentration
field of silver nanoparticles, subjected to the electromigration effect,
were calculated. To estimate the values of dimensionless parameters
involved in the analysis, we consider values of physical and geometrical
parameters that have been reported in previous work:11h0 = 100 nm, L = 200 nm, Rp = 25 nm, T = 293 K, ϵm = 7.8 ×
10–10 C/V m, ρm = 997 kg/m3, μ = 1 × 10–3 kg/ms, ϕ0 = 3 V, ζ = −25.4 ×
10–3 V, z = 1, n∞ = 6.022 × 1023 m–3, λD = 10 nm, Uc = 2.5 × 10–3 m/s, D = 8.58 × 10–12 m2/s, and Di = 1.65 × 10–9 m2/s. With the previous physical domain, the dimensionless
parameters for the calculations assume the following values: ϵ
= 0.05, k = 10, η = 0.5, α = 8.4 ×
10–3, δ = 1, Re = 2.5 ×
10–4, Pe = 2.93, and . For the analytical process, we consider
uncharged and charged Ag NPs, obtaining and at k = 10. Besides, the
nondimensional concentration field is governed by the following equation
72
In Figure 2c, the
nondimensional concentration field [eq 72] is shown for both charged and uncharged Ag NPs at k = 10. The selected times are determined using the time-dependent
diffusive component in eq 71 to counteract the condition . As t increases beyond
these selected values, the nondimensional concentration field converges
to a constant value, i.e., . The first noticeable effect in Figure 2a is the propagation
of NPs from their initial concentration at t = 0,
which occurs rapidly throughout the entire system. This phenomenon
is primarily attributed to diffusion, and notably, it conserves the
original distribution of NPs but elongates along the system. The concentration
distribution at x = 0.2 at this initial time is exclusively
influenced by the initial boundary condition [eq 70]. Figure 2b shows the concentration field for uncharged nanoparticles
at t = 10–5, where a distinctive
negative concentration is observed at the walls of the nanochannel.
This negative concentration indicates a deficit of nanoparticles close
to the walls. On the other hand, positive concentration values at
the entrance, middle, and exit of the nanochannel indicate that uncharged
nanoparticles, initially located near the entrance, are driven toward
the center of the channel, flow through it, and are eventually expelled
at the opposite end. Figure 2d shows that the concentration field for charged nanoparticles
is presented at t = 10–5. In this
case, a concentration value of c = 0 indicates the
occurrence of the reaction of NPs with the walls, as can be appreciated
from eqs 13 and 59. This outcome suggests that most charged NPs react
primarily at the entrance and exit regions of the nanochannel, while
the excess of NPs that cannot react at the walls is concentrated at
the central region. Furthermore, the coefficient , as defined in eq 66, may become zero for charged nanoparticles
implying that, under certain nanochannel dimensions, no convective
transport can take place for the leading order of the concentration
field. An analytical expression for the critical nanochannel length,
denoted as Lcrit, is derived by eq 66, yielding Lcrit = 12πRpλD3n∞f(k), where [eq 52]. For instance, at k = 10, this results
in Lcrit = 474 nm. However, it is noteworthy
to mention that for the current ratio , no significant changes in the concentration
fields are discernible even at the critical length Lcrit. To improve the concentration field with convection,
it is necessary to increase the ratio. Our analysis, using eqs 64–66, reveals that this can only be achieved by increasing the parameter
α and/or decreasing k = h0/λD. The parameter α = −ζ/ϕ0 can be increased by subjecting the system to an external
heat flux,36 or by reducing the applied
voltage from the generator. Caution must be exercised when decreasing
ϕ0 since this would cause a quadratic reduction in
the dielectrophoretic force and thus is not recommended. Considering
the reduction of h0, a lower limit of k = 2.5 is deduced. This requirement ensures that the height
of the nanochannel allows the passage of at least one nanoparticle
through it, i.e., h0 = Rp = 25 nm.
In Figure 3, the
nondimensional concentration field at k = 2.5 is
shown. In Figure 3a,
a pronounced trapping mechanism for uncharged nanoparticles is evident,
whereby a significant quantity of Ag NPs becomes trapped near the
center of the nanochannel. This change in behavior is governed by
the variable B(y), which, in return,
is a consequence of the overlap within the EDL. This phenomenon can
be elucidated by considering the representation of eqs 58 with 39 and 42, as follows
73
The trapping mechanism is observed to manifest
when tanh(k) <
0.999 which is obtained when k = 3.8. In Figure 3b, the concentration
field for charged Ag NPs is depicted, where it is observed that these
nanoparticles undergo electrical reactions predominantly at the exit
of the nanochannel while filling the nanohole in a counter-flow manner.
For the specified value of k = 2.5, the critical
length is calculated to be Lcrit = 212
nm.
Conclusions
Propagation of uncharged and charged nanoparticles
due to an EF
and electromigration in a nanochannel with overlapping EDLs has been
studied by deriving an analytical expression for the ionic distribution,
hydrodynamic forces, and Ag NP concentration. From the current analysis,
the following major points are obtained: (i) for charged nanoparticles,
colloidal transport convection is countered by electromigration, where
a critical length of the nanochannel will produce a pure diffusion
process for the leading concentration field solution; (ii) for uncharged
nanoparticles, a trapping mechanism can be achieved due to overlapping
of the EDL at k = 3.8; and (iii) in addition to modifying
the nanochannel dimension, the propagation of colloids can be achieved
by increasing the surface potential ζ through an external heat
source. This last finding requires that the energy equation be coupled
with the governing equations. Further studies on the propagation of
colloids in nanoconfinement would be required to investigate, experimentally,
the nanoaperture dimension and the variation in zeta potential. The
latter could be achieved by using an external heat source36 or by modifying the ionic concentration of the
solvent,14 as both approaches invalidate
the Debye–Hückel approximation.
Acknowledgments
C.V. acknowledges the support from the DGAPA program
for a postdoctoral fellowship at UNAM. Carlos Escobedo gratefully
acknowledges the financial support from the Natural Sciences and Engineering
Research Council of Canada (NSERC), from the Canada Foundation for
Innovation (CFI), and from Queen’s University for a FEAS Excellence
in Research Award.
Author Contributions
The manuscript
was written through all the contributions of all authors. All authors
have approved the final version of the manuscript.
Natural Sciences
and Engineering Research Council of Canada (NSERC), no. RGPIN-201-05138.
Canada Foundation for Innovation (CFI), no. 31967. Queen’s
University FEAS Excellence in Research Award.
The authors
declare no competing financial interest.
References
Wang J.; Maier S. A.; Tittl A.
Trends in Nanophotonics-Enabled Optofluidic
Biosensors. Adv. Opt. Mater.
2022, 10, 2102366. 10.1002/adom.202102366. [DOI] [Google Scholar]
Escobedo C.
On-chip nanohole
array based sensing: a review. Lab Chip
2013, 13, 2445–2463. 10.1039/c3lc50107h.
[DOI] [PubMed] [Google Scholar]
Pennathur S.; Santiago J. G.
Electrokinetic Transport in Nanochannels. 1. Theory. Anal. Chem.
2005, 77, 6772–6781. 10.1021/ac050835y.
[DOI] [PubMed] [Google Scholar]
Pennathur S.; Santiago J. G.
Electrokinetic transport
in nanochannels. 2. Experiments. Anal. Chem.
2005, 77, 6782–6789. 10.1021/ac0508346.
[DOI] [PubMed] [Google Scholar]
Striemer C. C.; Gaborski T. R.; McGrath J. L.; Fauchet P. M.
Charge- and size-based
separation of macromolecules using ultrathin silicon membranes. Nature
2007, 445, 749–753. 10.1038/nature05532.
[DOI] [PubMed] [Google Scholar]
Mukaibo H.; Wang T.; Perez-Gonzalez V.; Getpreecharsawas J.; Wurzer J.; Lapizco-Encinas B.; McGrath J. L.
Ultrathin nanoporous
membranes for insulator-based dielectrophoresis. Nanotechnology
2018, 29, 235704. 10.1088/1361-6528/aab5f7.
[DOI] [PubMed] [Google Scholar]
Eftekhari F.; Escobedo C.; Ferreira J.; Duan X.; Girotto E. M.; Brolo A. G.; Gordon R.; Sinton D.
Nanoholes as nanochannels:
Flow-through plasmonic sensing. Anal. Chem.
2009, 81, 4308–4311. 10.1021/ac900221y.
[DOI] [PubMed] [Google Scholar]
Escobedo C.; Brolo A. G.; Gordon R.; Sinton D.
Optofluidic Concentration:
Plasmonic Nanostructure as Concentrator and Sensor. Nano Lett.
2012, 12, 1592–1596. 10.1021/nl204504s.
[DOI] [PubMed] [Google Scholar]
Ertsgaard C. T.; McKoskey R. M.; Rich I. S.; Lindquist N. C.
Dynamic
placement of plasmonic hotspots for super-resolution surface-enhanced
Raman scattering. ACS Nano
2014, 8, 10941–10946. 10.1021/nn504776b.
[DOI] [PubMed] [Google Scholar]
Larson S.; Luong H.; Song C.; Zhao Y.
Dipole Radiation-Induced
Extraordinary Optical Transmission for Silver Nanorod-Covered Silver
Nanohole Arrays. J. Phys. Chem. C
2019, 123, 5634–5641. 10.1021/acs.jpcc.9b00477. [DOI] [Google Scholar]
Bdour Y.; Beaton G.; Gomez-Cruz J.; Cabezuelo O.; Stamplecoskie K.; Escobedo C.
Hybrid plasmonic metasurface
as enhanced
Raman hot-spots for pesticide detection at ultralow concentrations. Chem. Commun.
2023, 59, 8536–8539. 10.1039/D3CC01015E. [DOI] [PubMed] [Google Scholar]
Khosravi B.; Gordon R.
Accessible Double Nanohole
Raman Tweezer Analysis of
Single Nanoparticles. J. Phys. Chem. C
2024, 128, 15048–15053. 10.1021/acs.jpcc.4c03536. [DOI] [PMC free article] [PubMed] [Google Scholar]
De
Leebeeck A.; Sinton D.
Ionic dispersion in nanofluidics. Electrophoresis
2006, 27, 4999–5008. 10.1002/elps.200600264.
[DOI] [PubMed] [Google Scholar]
Qu W.; Li D.
A model for
overlapped EDL Fields. J. Colloid
Interface Sci.
2000, 224, 397–407. 10.1006/jcis.1999.6708.
[DOI] [PubMed] [Google Scholar]
Golovnev A.; Trimper S.
Analytical solution
of the Poisson-Nernst-Planck equations
in the linear regime at an applied dc-voltage. J. Chem. Phys.
2011, 134, 154902. 10.1063/1.3580288.
[DOI] [PubMed] [Google Scholar]
Zachariah Z.; Espinosa-Marzal R. M.; Spencer N. D.; Heuberger M. P.
Stepwise
collapse of highly overlapping electrical double layers. Phys. Chem. Chem. Phys.
2016, 18, 24417–24427. 10.1039/c6cp04222h.
[DOI] [PubMed] [Google Scholar]
Yeh L.-H.; Zhang M.; Qian S.; Hsu J.-P.; Tseng S.
Ion concentration
polarization in Polyelectrolyte-modified nanopores. J. Phys. Chem. C
2012, 116, 8672–8677. 10.1021/jp301957j. [DOI] [Google Scholar]
Lapizco-Encinas B. H.
Microscale
electrokinetic assessments of proteins employing insulating structures. Curr. Opin. Chem. Eng.
2020, 29, 9–16. 10.1016/j.coche.2020.02.007. [DOI] [Google Scholar]
Wu Z.-Q.; Li Z.-Q.; Ding X.-L.; Hu Y.-L.; Xia X.-H.
Influence
of Asymmetric Geometry on the Ion Transport of tandem nanochannels. J. Phys. Chem. C
2021, 125, 24622–24629. 10.1021/acs.jpcc.1c06884. [DOI] [Google Scholar]
Mani A.; Zangle T. A.; Santiago J. G.
On the propagation of concentration
polarization from microchannel-nanochannel interfaces Part I. Analytical
model and characteristic analysis. Langmuir
2009, 25, 3898–3908. 10.1021/la803317p.
[DOI] [PMC free article] [PubMed] [Google Scholar]
Yaroshchuk A.; Bondarenko M. P.
Current-Induced concentration polarization of nanoporous
media: Role of electroosmosis. Small
2018, 14, 1703723. 10.1002/smll.201703723. [DOI] [PubMed] [Google Scholar]
Chen L.; Hu C.; Dong Y.; Li Y.; Shi Q.; Liu G.; Long R.; Xiong Y.
Tunable Layered Gold Nanochips for
High Sensitivity and Uniformity in SERS. J.
Phys. Chem. C
2023, 127, 8167–8174. 10.1021/acs.jpcc.3c01359. [DOI] [Google Scholar]
Sakamoto M.; Saitow K.
Large-Area Plasmon
Mapping via an Optical Technique:
Silver Nanohole Array and Nano-Sawtooth Structures. J. Phys. Chem. C
2023, 127, 13105–13111. 10.1021/acs.jpcc.3c01919. [DOI] [Google Scholar]
Probstein R. F.Physicochemical Hydrodynamics; Wiley-Interscience, 2005. [Google Scholar]
Philipse A. P.Brownian Motion. Elements
of Colloid Dynamics; Springer, 2018. [Google Scholar]
Vacík J.Electrophoresis
a Survey of Techniques and Applications; Elsevier, 1979. [Google Scholar]
Elimelech M.; Gregory J.; Jia X.; Williams R. A.. Particle Deposition and Aggregation.
Measurement, Modelling and Simulation; Butterworth-Heinemann, 1995. [Google Scholar]
Leonard G.; Mitchner M.; Self S. A.
Particle
transport in electrostatic
precipitators. Atmos. Environ.
1980, 14, 1289–1299. 10.1016/0004-6981(80)90230-9. [DOI] [Google Scholar]
Adamcyzk Z.
Particle deposition
from flowing suspensions. Colloids Surf.
1989, 39, 1–37. 10.1016/0166-6622(89)80176-3. [DOI] [Google Scholar]
Miloh T.; Boymelgreen A.
Travelling
wave dipolophoresis of ideally polarizable
nanoparticles with overlapping electric double layers in cylindrical
pores. Phys. Fluids
2014, 26, 072101. 10.1063/1.4884956. [DOI] [Google Scholar]
Bazant M.
Z.; Thornton K.; Ajdari A.
Diffuse-charge dynamics in electrochemical
systems. Phys. Rev. E:Stat., Nonlinear, Soft
Matter Phys.
2004, 70, 021506. 10.1103/PhysRevE.70.021506. [DOI] [PubMed] [Google Scholar]
Ajdari A.
Electro-osmosis
on Inhomogeneously charged surfaces. Phys. Rev.
Lett.
1995, 75, 755–758. 10.1103/PhysRevLett.75.755.
[DOI] [PubMed] [Google Scholar]
Bender C. M.; Orzag S. A.. Advanced Mathematical
Methods for Scientists and Engineers I: Asymptotic Methods and Perturbation
Theory; Springer Science and Business, 2013. [Google Scholar]
Mei C. C.; Vernescu B.. Homogenization Methods
for Multiscale Mechanics; World Scientific
Publishing Co., 2010. [Google Scholar]
Vargas C.; Bautista O.; Méndez F.
Effect of temperature-dependent properties
on electroosmotic mobility at arbitrary zeta potentials. Appl. Math. Model.
2019, 68, 616–628. 10.1016/j.apm.2018.11.050. [DOI] [Google Scholar]