the golden aspect ratio for ion transport - arxiv · pdf filethe golden aspect ratio for ion...

11
1 The golden aspect ratio for ion transport Subin Sahu 1,2 and Michael Zwolak 1,* 1 Center for Nanoscale Science and Technology, National Institute of Standards and Technology, Gaithersburg, Maryland 2 Maryland Nanocenter, University of Maryland, College Park, Maryland ABSTRACT Access resistance indicates how well current carriers from the bulk medium can converge to a pore or open- ing, and is important in many fields, such as cell biology, electronics, electrochemical engineering, thermal transport, among others. In simplified scenarios, it depends only on the resistivity and pore radius when the bulk dimensions are infinite in all directions. These conditions are often not valid in simulations of transport due to the computational cost of large simulation cells, and can even break down in micro- and nano-scale systems due to strong confinement. Here, we examine a scaling theory for the access resistance that predicts there should be a special simulation cell aspect ratio – the golden aspect ratio – where finite size effects are eliminated. We demonstrate that this golden aspect ratio exists and that it takes on a universal value in linear response and moderate concentrations, i.e., regimes of interest to transport through biological ion channels and nanopores. Outside of linear response, it gains an apparent dependence on characteristics of the transport scenario (concentra- tion, voltages, etc.) for small simulation cells, but this dependence vanishes at larger length scales. These results will enable the use of all-atom molecular dynamics to study contextual properties of access resistance – i.e., its dependence on protein and molecular-scale fluctuations, the presence of charges, etc. – and yield the opportunity to quantitatively compare computed and measured resistances. INTRODUCTION Ion transport through narrow constrictions in biological membranes permits the regulation of concentrations – e.g., ion homeostasis – that is vital for physiological functions of cells (1–3). Moreover, ion transport through porous inor- ganic membranes is of interest for technological – e.g., DNA sequencing (4, 5) – and industrial applications (6, 7). The recent progress in the fabrication of atomically thin mem- branes such as graphene (8–10), MoS 2 (11) and hexago- nal boron nitride (12) opens new avenues in the field of ion transport: These membranes are excellent candidates for molecular and ionic sieves (13–15) for, e.g., desalina- tion (16, 17) and gas separation (18, 19). Their atomic thick- ness provides advantages for bio-sensing, such as sequenc- ing (9, 10, 20) and protein folding (21). Synthetic pores – especially ones with atomic thickness – also provide a testing ground for understanding biological ion channels and cre- ating biomimetic pores. In particular, ion currents through pores of controllable size can probe dehydration (22, 23) and directly quantify its effect on selectivity (i.e., experimen- tally determine the free energy barrier for partial dehydra- tion) (24, 25). The access resistance – the resistance for ions to con- verge from the bulk electrolyte to the mouth of the pore (26) – sets the upper limit of current flow through ion chan- nels (27, 28) and can become the dominating resistance at low salt concentration (29). For atomically thin pores with radii sufficiently larger than the solvation shell of ions, the access resistance will be the only significant resistance, even at high salt concentration. It thus becomes an important com- ponent to understand and quantify for, e.g., sensing and sequencing, which require high precision in the measure- ment and analysis of the ion current. Moreover, numerous efforts, especially using water-soluble polymers, have sought to separate pore and access contributions to the resistance in order to characterize different aspects of biological channels (30–32). Ideally, all-atom molecular dynamics (MD) simulations – which give a powerful means to simulate ion channels (33) – would be employed, as only these simulations capture con- textual aspects of the pores, such as molecular-scale fluctu- ations in pore sizes/geometries and local charges (34). For instance, edge fluctuations and the noncircular nature of a pore in graphene, together with van der Waals interactions and dehydration, make the pore radius hard to define (34). However, MD simulations are computationally intensive. Together with the long-range nature of convergence, this makes it difficult to reach the required simulation sizes to quantify the access resistance (35). Here, we examine a scal- ing analysis to extract the access and pore resistances (34), demonstrating that there is indeed a “golden aspect ratio” (different than the golden ratio, 1+ 5 2 ) that removes finite size effects. We use continuum simulations in order to scan the necessary parameter space. However, our findings should be equally applicable to all-atom molecular dynamics, where we have already come close to the golden aspect ratio based on an initial estimated value (34). For an infinitely large, balanced (i.e., infinite in all directions), and homogeneous system, the access resistance depends on the resistivity of the medium, γ , and the pore radius, a, only. Hall’s expression, R Hall = γ 4a , (1) *[email protected] arXiv:1711.00472v1 [cond-mat.mes-hall] 1 Nov 2017

Upload: dinhque

Post on 30-Jan-2018

225 views

Category:

Documents


5 download

TRANSCRIPT

1

The golden aspect ratio for ion transport

Subin Sahu1,2 and Michael Zwolak 1,∗

1Center for Nanoscale Science and Technology, National Institute of Standards and Technology, Gaithersburg, Maryland2Maryland Nanocenter, University of Maryland, College Park, Maryland

ABSTRACT Access resistance indicates how well current carriers from the bulk medium can converge to a pore or open-ing, and is important in many fields, such as cell biology, electronics, electrochemical engineering, thermal transport, amongothers. In simplified scenarios, it depends only on the resistivity and pore radius when the bulk dimensions are infinite in alldirections. These conditions are often not valid in simulations of transport due to the computational cost of large simulationcells, and can even break down in micro- and nano-scale systems due to strong confinement. Here, we examine a scalingtheory for the access resistance that predicts there should be a special simulation cell aspect ratio – the golden aspect ratio– where finite size effects are eliminated. We demonstrate that this golden aspect ratio exists and that it takes on a universalvalue in linear response and moderate concentrations, i.e., regimes of interest to transport through biological ion channels andnanopores. Outside of linear response, it gains an apparent dependence on characteristics of the transport scenario (concentra-tion, voltages, etc.) for small simulation cells, but this dependence vanishes at larger length scales. These results will enablethe use of all-atom molecular dynamics to study contextual properties of access resistance – i.e., its dependence on proteinand molecular-scale fluctuations, the presence of charges, etc. – and yield the opportunity to quantitatively compare computedand measured resistances.

INTRODUCTION

Ion transport through narrow constrictions in biologicalmembranes permits the regulation of concentrations – e.g.,ion homeostasis – that is vital for physiological functionsof cells (1–3). Moreover, ion transport through porous inor-ganic membranes is of interest for technological – e.g., DNAsequencing (4, 5) – and industrial applications (6, 7). Therecent progress in the fabrication of atomically thin mem-branes such as graphene (8–10), MoS2 (11) and hexago-nal boron nitride (12) opens new avenues in the field ofion transport: These membranes are excellent candidatesfor molecular and ionic sieves (13–15) for, e.g., desalina-tion (16, 17) and gas separation (18, 19). Their atomic thick-ness provides advantages for bio-sensing, such as sequenc-ing (9, 10, 20) and protein folding (21). Synthetic pores –especially ones with atomic thickness – also provide a testingground for understanding biological ion channels and cre-ating biomimetic pores. In particular, ion currents throughpores of controllable size can probe dehydration (22, 23)and directly quantify its effect on selectivity (i.e., experimen-tally determine the free energy barrier for partial dehydra-tion) (24, 25).

The access resistance – the resistance for ions to con-verge from the bulk electrolyte to the mouth of the pore (26)– sets the upper limit of current flow through ion chan-nels (27, 28) and can become the dominating resistance atlow salt concentration (29). For atomically thin pores withradii sufficiently larger than the solvation shell of ions, theaccess resistance will be the only significant resistance, evenat high salt concentration. It thus becomes an important com-ponent to understand and quantify for, e.g., sensing and

sequencing, which require high precision in the measure-ment and analysis of the ion current. Moreover, numerousefforts, especially using water-soluble polymers, have soughtto separate pore and access contributions to the resistance inorder to characterize different aspects of biological channels(30–32).

Ideally, all-atom molecular dynamics (MD) simulations– which give a powerful means to simulate ion channels (33)– would be employed, as only these simulations capture con-textual aspects of the pores, such as molecular-scale fluctu-ations in pore sizes/geometries and local charges (34). Forinstance, edge fluctuations and the noncircular nature of apore in graphene, together with van der Waals interactionsand dehydration, make the pore radius hard to define (34).However, MD simulations are computationally intensive.Together with the long-range nature of convergence, thismakes it difficult to reach the required simulation sizes toquantify the access resistance (35). Here, we examine a scal-ing analysis to extract the access and pore resistances (34),demonstrating that there is indeed a “golden aspect ratio”(different than the golden ratio, 1+

√5

2 ) that removes finitesize effects. We use continuum simulations in order to scanthe necessary parameter space. However, our findings shouldbe equally applicable to all-atom molecular dynamics, wherewe have already come close to the golden aspect ratio basedon an initial estimated value (34).

For an infinitely large, balanced (i.e., infinite in alldirections), and homogeneous system, the access resistancedepends on the resistivity of the medium, γ, and the poreradius, a, only. Hall’s expression,

RHall =γ

4a, (1)

*[email protected]

arX

iv:1

711.

0047

2v1

[co

nd-m

at.m

es-h

all]

1 N

ov 2

017

2 Sahu and Zwolak

+V0/2

−V0/2

ξ = 0

ξ = l/a

η = 0 η = 0

η=

1

+V0/2

−V0/2

ξ = 0

ξ = l/a

η = 0 η = 0

η=

1

Figure 1: Continuum simulation of ion transport through apore driven by a potential drop V0. The equipotential sur-faces (heat map) are elliptical near the pore – i.e., an access-like region. The finite size of the simulation cell curtails theaccess region, forcing the equipotential surfaces to transi-tion into nearly flat profiles closer to the electrodes – i.e.,a bulk-like region. The top half shows the rotational ellipti-cal coordinates, ξ and η, we employ in modeling the accessresistance.

gives the general form under these simplified condi-tions (26). In fact, the access resistance, Eq. (1), wasalready derived by Maxwell well over a century ago in thecontext of the electronic current through an orifice (36).Access/convergence resistance occurs in heat flow, mass dif-fusion, and other related scenarios, such as electrical (e.g.,disc electrodes) and thermal contacts. Hence, many authorshave derived the same expression for these various physi-cal systems, see, e.g., Gray and Mathews (37), Brown andEscombe (38), Grober (39), Holm (40), and Newmann (41).In the case of electronic transport, the resistance in the bal-listic regime – the Sharvin resistance (42) – crosses over toMaxwell’s expression when the pore is larger than the elec-tron mean free path (43, 44). Although our focus is on iontransport, the general findings should be applicable to othertransport scenarios as well.

Moreover, the fact that the access resistance varies as1/a rather than 1/a2, as it does for the pore resistance inthe diffusive regime, has an interesting consequence: Whenthe pore resistance is negligible, the current through onelarge pore is less than that through several smaller poresof the same total area if the pores do not interfere witheach other. Nature uses this effect, for example, to maximizethe gas exchange rate between the atmosphere and stomatain leaves (38) and one can envision using the same effectfor maximizing permeation through atomically thin mem-branes, such as those through graphene and MoS2, which

naturally have a small pore resistance [above the dehydrationlimit (24, 25)].

MODEL

In addition to an infinitely large, balanced, and homogeneousbulk, the typical derivation of Eq. (1) assumes that a hemi-spherical electrode is at infinity. This means that at largedistances the electric field lines extend radially outward fromthe pore/contact. Such a directionality will emerge, though,from an infinitely large disc electrode at infinity. Within thepore, these field lines have to transition to pointing along thesymmetry axis, which we take as the z-axis. These symme-tries can be seen from the equipotential surfaces close to thepore, as shown in Fig. 1. We will retain the ellipsoidal sym-metry to derive the finite-size corrections. These correctionsare especially important in simulations, but can also be rele-vant to nanofluidic (45) and microelectromechanical systems(MEMS) (46).

We have shown previously (34) how to modify Eq. (1)for a finite-size system [while retaining the symmetry of theproblem (41, 47)]. This setup is easier to solve using rota-tional elliptic coordinates, ξ and η, which relate to cylindricalcoordinates, z and ρ, via

z = a ξ η

ρ = a√

(1 + ξ2)(1− η2). (2)

Laplace’s equation for the electric potential in this coordi-nate system is (48)

∂ξ

[(1 + ξ2

) ∂V∂ξ

]+

∂η

[(1− η2

) ∂V∂η

]= 0. (3)

The boundary conditions are: (i) a constant potential at thepore mouth (V = 0 at ξ = 0), (ii) no radial electric fieldon the membrane surface (∂V/∂η = 0 at η = 0), and (iii)a ellipsoidal electrode of radius l (V = V0 at ξ = l/a).Only condition (iii) is different than that used for an infinitebulk. We stress that these conditions only partially hold inour system. In particular, condition (iii) is really a fictitiouselectrode placed at the end of the access region. We will usethe location as a fitting parameter, as well as fitting parame-ters to capture the effect of the transition region – in betweenthe access and bulk regions – and the location of the start ofthe normal bulk region.

The solution to Eq. (3) is (see Appendix A)

V

V0=

tan−1 ξ

tan−1(l/a)(4)

for the potential. Thus, the access resistance is

Raccess/2 =γ tan−1(l/a)

2πa

4a

(1− 2a

πl+O

[(al

)3])

≈ RHall

(1− 2a

πl

)= RHall −

γ

2πl, (5)

The golden aspect ratio 3

56

58

60

62

64

66

8 16 32 64 128

R(M

Ω)

L (nm)

4γπ

(14.5 nmL2 − 1.1

L

)+ 59.2 MΩ

γ(14.5 nmL2 − 1.2

L

)+ 58.7 MΩ

Extrapolations: - - /

Cylindrical box

Cubical box

R∞

Figure 2: Resistance versus the cross-sectional length L fora cell of fixed electrolyte height H = 14.5 nm, pore radiusa = 1 nm, membrane thickness hp = 1 nm, and resis-tivity γ = 71 MΩ·nm [taken to match our all-atom MDvalue of 1 M KCl from prior results (34)]. The total simu-lation cell height is H + hp. For L & H/f , R is smallerthan the actual resistance for an infinite-size electrolyte, R∞(horizontal dotted line). The scaling form still correctly pre-dicts R∞ when fit for small L (solid lines) but to do so ithas to display non-monotonic behavior (extrapolated, dashedlines). The computed data at large L, however, converges toa value below R∞, as the increasing volume of electrolytehorizontally far away from the pore does not contribute anadditional path for ions to flow (and decrease the resistance)nor can it increase the resistance. We note that obtaining ffrom where R crosses R∞ gives about 1.04 for cylindricaland 1.21 for rectangular cells. This is in close agreementwith the golden aspect ratio we obtain later, with the dif-ference from the exact values likely due to the strong non-monotonic effects here. The standard error of the fit for fand R∞ are less than 0.2 %.

where we ignore higher order corrections since they areO[(a/l)3

]. The notation “access/2” indicates that this is

the access resistance on only a single side of the membrane.When the (simulation) cell size has, e.g., a cross-sectionallength of 10 nm, l has to be less than 5 nm. Hence, theexpression shows that the classical form is off by ≈ 13% for a pore radius of 1 nm. This will get worse for evenmoderately larger pore sizes, not to mention the difficulty inapplying the Maxwell-Hall form when a is ill-defined.

Scaling AnalysisThe simple boundary conditions and ellipsoidal symmetryallow us to derive the exact expression, Eq. (5), for the accessresistance for an idealized finite size system. However, inpractice neither the boundary conditions nor the symmetryhold. Simulations usually have parallel disc electrodes (orhomogeneous applied fields) and uniform cross-section, andnanopore experiments have even more complicated arrange-ments. In simulations, a rectangular or cylindrical cell arethe natural choices. However, as shown in Fig. 1, the poten-tial still has ellipsoidal symmetry near the pore and it startsto become flat away from the pore. This is also the case inall-atom MD simulations (34).

For bulk of heightH (not including the membrane thick-ness) and cross-sectional length L, there are three regions oneach side: an ellipsoidal access-like region extending fromξ = 0 to ξ = l, a flat bulk-like region extending fromapproximately z = l to z = H/2 (this takes the upper mem-brane surface to be at z = 0), and an intermediate regionbetween the two. The length l will be some fraction of thecross-sectional length L and the resistance of the interme-diate region should decay as 1/L. Thus, the total resistanceis

R =Raccess +Rbulk +Rintermediate +Rpore

=2

(RHall −

γ

πf1L

)+ γ

(H − f2LGL2

)+γf3GL +Rpore,

(6)

where GL2 is the cross-sectional area of the bulk cell (G = 1for rectangular and G = π/4 for cylindrical). We intro-duce factors f1, f2, and f3 due to the uncertainty in theextent/contribution of these regions and the transitory natureof the boundaries between them. For instance, the accessregion ends at l = f1L/2 and we expect f1 is O(1), i.e.,this region encompasses a substantial fraction of the cellwidth (when H > L). As well, the normal bulk-like regionhas a height H − 2f2L/2, where f2L/2 is subtracted fromeach side of the membrane (the membrane thickness is notincluded inH), and we expect f2 isO(1). When the cell sizeis infinite in all directions, the resistance reduces to

R→ R∞ = 2RHall +Rpore. (7)

Thus, we get a simple scaling form

R =γ

G

(H

L2− f

L

)+R∞, (8)

wheref = 2G/πf1 + f2 − f3 (9)

gives a single fitting factor. This relation shows that the finitesize – and confined – correction to the access resistancedepends on both height and the cross-section of the cell. Inthe context of heat flow (49, 50), others have shown that, foran infinitely tall cell, access resistance has functional depen-dence on a/L, which reduces to the classical form whenL a.

We note that we do not require that R∞ = 2RHall +Rpore, rather only that the resistance convergences to someR∞. We expect that conditions, such as the presence ofcharges or an ill-defined pore radius, that give an access con-tribution other than the Maxwell-Hall form will still obeythe scaling law, so long as the cell is large enough to removenon-scaling finite size effects. We will show this explicitly inthe case of non-linear response.

The difficulty with the scaling form, Eq. (8), is that Rdepends on two geometric variables, H and L. On the onehand, if we hold L constant and increase H , then R willincrease linearly with H (e.g., this would apply both in sim-ulations and experiments where a very narrow nanofluidicconstriction leads up to the membrane/pore). On the other

4 Sahu and Zwolak

50

60

70

80

90

50

60

70

80

90

50

60

70

80

90

50

60

70

80

90

50

60

70

80

90

50

60

70

80

90

50

60

70

80

90

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

50

60

70

80

90

8 16 32 64 128

Cylindrical box

Rectangular box

R(M

Ω)

4γ(0.25 − 1.30)/πL + 57.8 MΩ

R(M

Ω) 4γ(0.50 − 1.11)/πL + 59.1 MΩ

R(M

Ω)

4γ(0.80 − 1.07)/πL + 59.2 MΩ

R(M

Ω)

4γ(1.08 − 1.07)/πL + 59.2 MΩ

R(M

Ω)

4γ(1.60 − 1.07)/πL + 59.2 MΩ

R(M

Ω)

4γ(2.00 − 1.07)/πL + 59.2 MΩ

R(M

Ω)

4γ(3.00 − 1.07)/πL + 59.2 MΩ

R(M

Ω)

L (nm)

γ (0.25 − 1.28) /L + 57.4 MΩ

R(M

Ω)

L (nm)

γ (0.50 − 1.08) /L + 58.6 MΩ

R(M

Ω)

L (nm)

γ (0.80 − 1.12) /L + 59.0 MΩ

R(M

Ω)

L (nm)

γ (1.20 − 1.20) /L + 59.0 MΩ

R(M

Ω)

L (nm)

γ (1.60 − 1.29) /L + 59.0 MΩ

R(M

Ω)

L (nm)

γ (2.00 − 1.37) /L + 59.0 MΩ

R(M

Ω)

L (nm)

γ (3.00 − 1.59) /L + 59.0 MΩ

Figure 3: Resistance versus L for different aspect ratios α,with a = 1 nm, hp = 1 nm, and γ = 71 MΩ·nm. For smallα, dR/dL > 0 and for large α, dR/dL < 0. At special valueof α in between – the “golden aspect ratio”, dR/dL = 0 andthe resistance is constant (i.e., no finite size effects). Here,the gold line is very close, but not quite at, the golden aspectratio. The solid lines show the fits R = γ

GL (α− f) + R∞,where G = π/4 for cylindrical cells (top panel) and G = 1for rectangular cells (bottom panel). The dashed lines extrap-olate these fits, which match well with the calculations forlargeL and yield consistent values forR∞. The performanceof the scaling analysis using small L indicates that the sim-ulation cell sizes achievable with all-atom MD should besufficient to find the total resistance (pore plus access). Theerror of the fits for the f ’s and the R∞’s are about 0.5 % and0.1 % respectively (except for α = 0.25 where the respectiveerrors are about about 3 % and 0.5 % ).

hand, if we keep H constant and increase L, the finite-sizecorrection in Eq. (8) becomes negative once L & H/f andR becomes smaller than R∞, as seen in Fig. 2. A furtherincrease in L does not converge R to R∞, as implied byEq. (8), because at aspect ratios less than one (H . L) thepotential surfaces become flat horizontally (rather than ver-tically) away from the pore (46), nullifying the assumptionsthat went into Eq. (8). One can still apply the scaling form,Eq. (8), so long as the fit is for a range of L . H/f , whichFig. 2 shows it still extracts the correct R∞ (found by usingthe approach we describe below).

In order for R to converge to R∞ monotonically (in theabsence of nonlinearities), both L and H need to increasesimultaneously. An intuitive method to do so is to keep the

aspect ratio, α = H/L, constant, which further simplifiesEq. (8) to

R =γ

GL (α− f) +R∞. (10)

Not only does this yield an elegant scaling form, this equa-tion shows that if we choose α = f then the finite sizecorrection is eliminated and R is independent of L. In otherwords, at this “golden aspect ratio”, R = R∞ for any L.

It is, however, likely not possible to find a general expres-sion for the factor f = 2G/πf1+f2−f3, since the transitionfrom access-like to bulk-like is complicated and the factorsf1, f2 and f3 will in general depend on geometric detailsof the pore and the boundary conditions. Nevertheless, withsome reasonable approximations we can give an estimate off . We expect that f1 ≈ f2 ≈ 1 (i.e., a boundary betweenaccess and bulk-like regions at l = L/2) and f3 to dependon f1 and f2. To estimate f3 we assume that the intermedi-ate region is about an equal contribution from access- andbulk-like behavior. This intermediate region can be assumedto have a hemispherical boundary of radius f1L/2 and a flatboundary of cross-section GL2 at height f2L/2 on each sideof the pore. Thus the resistance of the intermediate region isestimated to be

γf3GL =

1

2

(2γ

πf1L− 2γ

πL

)+

1

2

(2γf2L/2

GL2

), (11)

giving

f3 =Gf1π− Gπ

+f22. (12)

The estimate of f3 likely overestimates the bulk-like con-tribution, while either under- or over-estimating the accesscontribution. Using it in Eq. (9), simplifies f to

f =Gπf1

+f22

+Gπ. (13)

Table 1 shows the value of f for various values off2 = f1. The value of f is quite insensitive to f1 and f2in a reasonable range (due to the fact that one appears in adenominator and the other in a numerator). The equipoten-tial surfaces in Fig. 1 suggest f1 ≈ 2/3 and f2 ≈ 3/4, andthus we estimate

f =

1.2 for rectangular box1.0 for cylindrical box

(14)

for the golden aspect ratio.We will also examine the sensitivity of f to different

conditions, including whether one uses the Poisson-Nernst-Planck (PNP) equations or just Laplace’s equation (i.e.,Ohm’s law). The latter – the homogeneous medium approx-imation – is valid when the applied potential is small andthe ion concentration is large (8). Poisson’s equation andthe stationary Nernst-Planck equation express the spatialdependence of the potential V and current density J as

∇2V =∑

ν

qνcνε

(15)

The golden aspect ratio 5

16

32

64

128

4 8 16 32 64 128

(a)

16

32

64

128

4 8 16 32 64 128

(a)

16

32

64

128

4 8 16 32 64 128

(a)

R(M

Ω)

L (nm)

a = 0.5 nm

4γπL

(1.07 − 1.07) + 163.2 MΩ

R(M

Ω)

L (nm)

a = 1.0 nm

4γπL

(1.07 − 1.07) + 59.2 MΩ

R(M

Ω)

L (nm)

a = 2.0 nm

4γπL

(1.07 − 1.07) + 23.9 MΩ

10

30

50

70

1 2 3 4

−4

0

4

8

0.25 1 4

(b)

R(M

Ω)

a (nm)

γ(

12a + 1.05 nm

πa2

)+ ∆R

∆R

(MΩ

)

α

0

5

10

15

20

25

1 2 3 4

−2

0

2

4

0.25 1 4

(c)

R(M

Ω)

a (nm)

γ4a + ∆R

∆R

(MΩ

)

α

Figure 4: (a) The resistance versus L for cylindrical cells with α = 1.07, hp = 1 nm, γ = 71 MΩ·nm, and various poreradii. The golden aspect ratio is fairly constant over this range of a. The solid lines show the fits R = 4γ

πL (α− f) + R∞and the dashed lines show the extrapolations – which make good predictions of R at large L. The errors of the fits for f ’sand R∞’s are less than 0.5 % and 0.1 % respectively. As the pore radius increases, we examine fits for larger L, to roughlykeep proportionality between L and a as there are (non-scaling) finite-size effects that will affect the simulations when theradius starts to become comparable to the cell cross-sectional length. The range of the fitting can influence the extracted f (weexpect that ultra-precise calculations at really large L will reveal the exact golden aspect ratio, which will be close to 1.07).(b) The resistance versus a for cylindrical cells with L = 32 nm, hp = 1 nm, and the “golden aspect ratios” α = 1.07. Theresistance fits into the model,R = γ

(1/2a+ heffp /πa2

)+∆R with heffp and ∆R as fitting parameters. The inset shows that

∆R ≈ 0 for α = 1.07 (dashed lines) but is non-zero at smaller and larger values of α. The effective height, heffp , is (5±1) %larger than the 1 nm membrane thickness, see Fig. 5. (c) R versus a for a half cylinder with the pore mouth set at potentialV = 0. The resistance fits into the model, R = γ/4a+ ∆R. The inset shows ∆R for different aspect ratios with ∆R ≈ 0 forα = 1.07 (dashed lines). The interpolated lines in the insets are for visual clarity only.

andJν = −qν (Dν∇cν + µνcν∇V) , (16)

where qν , cν , Dν , µν are the charge, concentration, diffusiv-ity, and mobility of ion species ν. We use a commercial finiteelement solver for both cases.

f1 1.00 0.80 0.60 0.40f2 1.00 0.80 0.60 0.40

f (rectangular) 1.14 1.12 1.15 1.31f (cylindrical) 1.00 0.96 0.97 1.08

Table 1: Estimates of the golden aspect ratio for variousvalues of f1 and f2.

RESULTS

We first test the scaling form, Eq. (10), for a homogeneousmedium by solving Laplace’s equation for the electric poten-tial. All simulations take a silica membrane with dielectricconstant 2.1. Figure 3 shows that, for both cylindrical andrectangular cells, the resistance increases with L for smallaspect ratios and decreases for large aspect ratios as pre-dicted from Eq. (10). Physically, we can understand this asfollows: When the aspect ratio (H/L) is small, the access-like region covers most of the height of the cell. Thus, whenincreasing L, the access contribution increases, as the finite

size correction – a negative correction – in Eq. (5) is beingeliminated. When the aspect ratio is large, the access-likeregion is localized near the pore and the bulk-like regioncovers the remainder of the height of the cell. This bulk-likeregion gives a large positive contribution to the resistance,but is decreasing with L (due to the fact that it decreaseswith area). This explains the decrease in resistance with L.This transition from increasing to decreasing resistance withL suggests that there should be a constant R at some specialvalue of the aspect ratio. At this special value, the two com-peting effects cancel. Figure 3 shows that there is an aspectratio that shows little variation as L increases. Thus, indeed,the golden aspect ratio exists and is approximately f ≈ 1.07for cylindrical cell and f ≈ 1.2 for rectangular cell, in linewith the estimate in Eq. (14).

In Fig. 4(a), we examine four pore sizes and use cylindri-cal cell with aspect ratio α = 1.07 where we get a visuallyflat profile ofR versus L. Inappropriately chosen cell dimen-sions, however, will have R dependent on L and disagreewith the classical form of the access resistance. Figure 4(b)shows that, around α = 1.07 and at finite L (32 nm), theform R = γ

(1/2a+ hp/πa

2)

+ ∆R, fits the data with∆R ≈ 0, thus coinciding with the classical form. For α = 2,a reasonable value from a computational standpoint, the fitgives ∆R ≈ 3 MΩ, which is ≈ 30 % of R∞ at a = 4 nm.

Notice that we obtain hp = 1.05 nm from the fit, which isdifferent than the 1 nm membrane thickness. The expressionγhp/πa

2 assumes the potential surfaces are flat within the

6 Sahu and Zwolak

8

16

32

64

128

1 2 3 48

16

32

64

128

1 2 3 48

16

32

64

128

1 2 3 48

16

32

64

128

1 2 3 4

R(M

Ω)

a (nm)

γ(

12a + 1.05 nm

πa2

)+ 0.2 MΩ

R(M

Ω)

a (nm)

γ(

12a + 2.05 nm

πa2

)+ 0.2 MΩ

R(M

Ω)

a (nm)

γ(

12a + 3.05 nm

πa2

)+ 0.3 MΩ

R(M

Ω)

a (nm)

γ(

12a + 4.05 nm

πa2

)+ 0.3 MΩ

Figure 5: The resistance versus a for cylindrical cells withL = 32 nm, the golden aspect ratio α = 1.07, and hp =1.0 nm, 2.0 nm, 3.0 nm and 4.0 nm. The resistance fits intothe model, R = γ

(1/2a+ heffp /πa2

)+ ∆R with heffp and

∆R as fitting parameters. The fitted heffp is slightly largerthan the actual membrane thickness but its relative differencediminishes as the membrane thickness is increased. This isdue to the curvature of the potential at the pore mouth andnot due to numerical grid-size errors (a ten times smaller gridincrease this value to 7 % and thus we expect this height cor-rection to converge to something of this order.). The standarderrors of heffp and ∆R are about 0.01 nm and 0.04 MΩ ineach case.

full length of the pore. However, in actual pores, these sur-faces are not perfectly flat, especially near the pore mouth,as is visible in Fig. 1. This difference in curvature will give acorrection to γhp/πa2. This correction becomes negligiblefor thick membranes, as shown in Fig. 5, where the correc-tion comes out to be 1.0 % for a 4 nm membrane (comparedto 5 % for the 1 nm membrane).

We can remove this source of ambiguity completely,though, by examining the resistance on a half cylinder –between the pore mouth set at potential V = 0 and anelectrode at one end of the cylinder. Figure 4(c) shows thisresistance versus pore radius. Since the half cylinder doesnot have the pore resistance, we fit it to R = γ/4a + ∆R.Once again we get ∆R ≈ 0 for the aspect ratio α = 1.07and substantially different than zero for other values of α.These calculations using Laplace’s equation pin the goldenaspect ratio to be around 1.1 for the cylindrical box (andaround 1.2 for a rectangular box). These values of thegolden aspect ratio are remarkably close to the estimate inEq. (14). However, we also want to examine more realisticsimulations, and thus we now examine the PNP solution.

PNP SolutionWe test Eq. (10) for PNP simulations with two different con-centrations, 0.1 mol/L and 1.0 mol/L of KCl solution, andapplied voltages of 10 mV, 100 mV and 300 mV. Figure 6shows that for 10 mV and 100 mV we obtain a similar golden

aspect ratio as for the homogeneous medium solution. How-ever, this special ratio seems to decrease for the larger volt-age of 300 mV. Clearly, there are nonlinear effects cominginto play when the voltage increases and/or the concentrationis insufficient to screen the field without substantially per-turbing the medium. We will show that this decrease is justan apparent decrease due to non-scaling [in the context ofEq. (10)] finite-size effects. Larger L will restore the scalingform.

We first want to understand its origin, for which weneed to take into consideration the formation of concentra-tion gradients (51). As seen in Fig. 7, the resistivity (γ =1/∑cνµν) is fairly constant along the axis of the simula-

tion cell for small voltages and thus we get the same resultas in the homogeneous case. For higher voltages, however,γ is smaller near the pore as a result of classical Wien effect(52, 53). However, the scaling form, Eq. (10), can still beemployed as long as the resistivity of different regions donot change drastically. Relaxing the homogeneity assump-tion that went into the derivation of Eq. (10), we set γb,γa, and γp to be the resistivity in the bulk-like region, theaccess-like region, and in the pore, respectively. Then, thetotal resistance is

R =Raccess +Rbulk +Rintermediate +Rpore

=2

(γa4a− γaπf1L

)+ γb

(αL− f2LGL2

)+γbf3GL +

γphpπa2

=γbGL

[α− f

(γaγb

)]+R′∞, (17)

where R′∞ = γa/2a+ γphp/πa2 and

f

(γaγb

)=

2Gγaπf1γb

+ f2 − f3. (18)

Within this simplified – “compartmentalized” – inhomoge-neous system, the resistance of the infinite system and thegolden aspect ratio are linear function of γa, decreasing as γadecreases. This qualitatively explains why we see the appar-ent decrease in the golden aspect ratio, it is the depression ofthe access contribution that decreases the required height ofthe bulk-like region that “balances” it. Of course, the resis-tivity is not constant throughout the access-like region and,at very high voltage and low concentration, not even in thebulk-like region. Thus, this simplified inhomogeneous modelwill not accurately determine the apparent golden aspectratio. However, it does not need to: The extensive decreasein resistivity is due to the fact that the simulation cells thatare really small can not sustain large potential drops with-out a global perturbation to the medium. Examining muchlarger cells shows that the scaling form is restored, see Fig. 8,which is due to the fact that the perturbation to the mediumbecomes localized (in dimensionless terms) around the pore,see the curves for different L in Fig. 7.

In other words, there are finite-size effects and there arefinite-size effects: The perturbation to the medium is a non-scaling finite size effect and the simulation cells need to besufficiently large to accommodate the perturbed region. Wedo not rule out, however, that an alternative scaling ansatz

The golden aspect ratio 7

500

550

600

6500.1

mol/L

1.0

mol/L

500

550

600

6500.1

mol/L

1.0

mol/L

500

550

600

6500.1

mol/L

1.0

mol/L

0.1

mol/L

1.0

mol/L

0.1

mol/L

1.0

mol/L

0.1

mol/L

1.0

mol/L

0.1

mol/L

1.0

mol/L

0.1

mol/L

1.0

mol/L

0.1

mol/L

1.0

mol/L

50

55

60

65

8 16 32 64 128

0.1

mol/L

1.0

mol/L

50

55

60

65

8 16 32 64 128

0.1

mol/L

1.0

mol/L

50

55

60

65

8 16 32 64 128

0.1

mol/L

1.0

mol/L

8 16 32 64 128

0.1

mol/L

1.0

mol/L

8 16 32 64 128

0.1

mol/L

1.0

mol/L

8 16 32 64 128

0.1

mol/L

1.0

mol/L

8 16 32 64 128

0.1

mol/L

1.0

mol/L

8 16 32 64 128

0.1

mol/L

1.0

mol/L

8 16 32 64 128

0.1

mol/L

1.0

mol/L

10 mV

4γ(0.8 − 1.1)/πL + 594.7 MΩ

10 mV

4γ(1.1 − 1.1)/πL + 594.8 MΩ

10 mV

4γ(1.4 − 1.1)/πL + 594.9 MΩ

100 mV

4γ(0.8 − 1.1)/πL + 583.3 MΩ

100 mV

4γ(1.1 − 1.1)/πL + 583.8 MΩ

100 mV

4γ(1.4 − 1.1)/πL + 584.0 MΩ

300 mV

4γ(0.8 − 0.9)/πL + 528.3 MΩ

300 mV

4γ(1.1 − 0.9)/πL + 530.8 MΩ

300 mV

4γ(1.4 − 0.9)/πL + 532.3 MΩ

R(M

Ω)

4γ(0.8 − 1.1)/πL + 59.5 MΩ

R(M

Ω)

4γ(1.1 − 1.1)/πL + 59.6 MΩ

R(M

Ω)

4γ(1.4 − 1.1)/πL + 59.6 MΩ

L (nm)

4γ(0.8 − 1.1)/πL + 59.4 MΩ

L (nm)

4γ(1.1 − 1.1)/πL + 59.4 MΩ

L (nm)

4γ(1.4 − 1.1)/πL + 59.4 MΩ

4γ(0.8 − 1.1)/πL + 58.2 MΩ

4γ(1.1 − 1.0)/πL + 58.3 MΩ

4γ(1.4 − 1.0)/πL + 58.3 MΩ

Figure 6: PNP solution of the resistance versus L for cylindrical cells with hp = 1 nm, a = 1 nm, and various aspect ratios(α = 1.4, 1.1 and 0.8 from top to bottom). The label at the top shows the applied potential and the labels at the right showthe concentration of KCl. The golden aspect ratio remains the same as in homogeneous medium calculation for the smallervoltages (10 mV and 100 mV) and for both 1 mol/L and 0.1 mol/L concentrations. For 300 mV applied voltage, the goldenaspect ratio seems to take on a lower value due to the increased concentration of ions near the membrane, which is morepronounced for lower concentrations due to the decreased ability to screen and generate local fields. The error of the fits forf ’s and R∞’s are about 0.5 % and 0.1 % respectively.

could be developed to handle such cases, e.g., including theperturbing effect of local fields. Its applicability would bemore limited (unlike the general forms, Eq. (10) or Eq. (8),which, however, may require larger L to reach the scalingregime) but still helpful. Finite-size effects related to trun-cating the access region or asymmetric dimensions can beremoved via Eq. (10) or Eq. (8), respectively.

DISCUSSION

Ion transport properties are difficult to compute with all-atom molecular dynamics. In the dehydration limit, forinstance, large free energy barriers exist, which entails verylong (microseconds or more) simulations to accumulateenough ion crossing events to determine the current withreasonable accuracy (24, 25). This applies even at elevatedvoltages [simulations are often done at 3 V or higher (54)],which risks introducing nonlinearities and prohibits a directcomparison with experiment. Contextual properties of iontransport – fluctuations of the pore or membrane, localizeddipoles or charges, van der Waals interactions and dehy-dration, geometrically “imperfect” pore shapes, etc. – makethese simulations even more difficult to understand and con-trol (e.g., slow fluctuations of pore structure require longtimes to acquire the requisite statistics).

We have further developed a scaling theory to extractboth the pore and access contributions to resistance with-out going to excessively large simulation cells (i.e., L < 16

nm for linear response). This scaling theory was originallyemployed to determine the access contribution – includinghow to properly define the pore size in the presence of con-textual properties – to graphene nanopores with all-atomMD (34). Here, we demonstrated that the “golden aspectratio” – a special aspect ratio where finite size effects areeliminated – exists, where previously we used an aspect ratioclose to an initial estimate of the golden aspect ratio (and,indeed, we had little variation of R with L) but large enoughto ensure we decrease from above to R∞ (34). Within con-tinuum simulations, this aspect ratio gives a completely flat(i.e., without non-monotonic/oscillatory behavior) resistanceversus simulation cell size. The value for the golden aspectratio is about 1.1 for cylindrical cells and 1.2 for rectangular.It is relatively insensitive to voltages and ion concentrationswithin experimentally relevant regimes (i.e., around 0.1 Vapplied voltages and 0.1 mol/L to 1.0 mol/L concentrations).For large voltages and small concentrations, the simulationcells have to be sufficiently large to remove non-scalingfinite-size effects, after which the scaling and golden aspectratio are restored.

This scaling approach will be most useful when exam-ining pores that do not have really high pore resistances, asin that case, the access resistance becomes a dominant orequal contributor rather than a correction to the resistance.Graphene and other atomically thin pores have a dominant orsubstantial access contribution all the way down to the dehy-dration limit (34), and thus the scaling approach is essential.

8 Sahu and Zwolak

0.6

0.8

1

0.6

0.8

1

0.6

0.8

1

−0.6 −0.4 −0.2 0.0 0.2 0.4 0.6

0.1 V

L = 8 nm 32 nm 128 nm

0.3 V

γ/γ

b

z/L

1.0 V

Figure 7: The resistivity along the z-axis for various appliedpotentials and cell dimensions L. The pore radius is a = 1nm and the KCl concentration is 0.1 mol/L (γb = 710MΩ·nm). Large applied voltages enhance the density ofcharge carriers near the pore, decreasing the resistivity. Sincethe access resistance contribution is largest near the pore,see Eq. (5), this enhancement lowers the overall access resis-tance and gives an apparent depression to the golden aspectratio when examining small simulation cells.

Biological channels can fall into either category, sometimesrequiring a correct determination of the access contributionand sometimes not. For instance, α-hemolysin pores havea 1 GΩ resistance (at 1 mol/L KCl) (55), compared to anestimated [from Eq. (1)] 40 MΩ access contribution. On theother hand, a sodium channel of radius 0.3 nm and effec-tive length 0.5 nm has both an access and pore resistancethat are nearly equal (27). In low salt concentrations, accessresistance may become the dominating resistance in a widevariety of cases (29).

We expect that all-atom MD will have a similar valuefor the golden aspect ratio under the same conditions (lowvoltages and moderate concentrations, and in the absenceof fixed charges/dipoles which introduce inhomogeneitiesin resistivity). However, the primary power of MD is tostudy contextual properties of pores, such as the effect ofatomic/molecular fluctuations of the pore/membrane and thepresence of (partial) charges. Also, having periodic cell,which is the most common technique in MD, may change thegolden aspect ratio. Hence, in practice, one can not simplyset the aspect ratio to 1.2 (for a rectangular cell) and performthe simulation. Rather, one has to apply the scaling form,Eq. (10), by performing simulations for different L. How-ever, one should use a reasonable guess of the golden aspectratio, i.e., 1.2 under normal conditions (entailing simulationcells of height ≈ 1.2 · L+ hp for the rectangular cells oftenused in MD). Moreover, we note that even if one knew aheadof time the golden aspect ratio, one has to be sure not to work

360

370

380

390

360

370

380

390

360

370

380

390

51

52

53

54

32 64 128 256 512 1024 2048

0.1 mol/L

1.0 mol/L51

52

53

54

32 64 128 256 512 1024 2048

0.1 mol/L

1.0 mol/L51

52

53

54

32 64 128 256 512 1024 2048

0.1 mol/L

1.0 mol/L

4γ(0.8 − 1.1)/πL+ 361.3 MΩ

4γ(1.1 − 0.9)/πL+ 361.4 MΩ

4γ(1.4 − 0.7)/πL+ 361.4 MΩ

R(M

Ω)

L (nm)

4γ(0.8 − 1.2)/πL+ 52.5 MΩ

R(M

Ω)

L (nm)

4γ(1.1 − 1.1)/πL+ 52.5 MΩR(M

Ω)

L (nm)

4γ(1.4 − 1.0)/πL+ 52.5 MΩ

Figure 8: PNP solution of the resistance versus L at appliedpotential of 1 V and KCl concentrations 0.1 mol/L (toppanel) and 1.0 mol/L (bottom panel) for cylindrical cellswith hp = 1 nm, a = 1 nm, and aspect ratios (α =1.4, 1.1and 0.8). The solid lines show the fit with f and R∞ as fit-ting parameters and the dashed lines show the extrapolation.Since the applied field is large, the resistance at smaller L (<100 nm) is highly nonlinear and cannot be scaled to extractR∞ with Eq. (10) as is done for the smaller applied voltagesin Fig. 6. Hence, a largerL is necessary to employ the scalingform when the electric field is large and ion concentrationsare small. Note that the resistance for the 0.1 mol/L solutionis substantially smaller than at lower voltages. For 0.1 mol/Lsolution, the error of the fit for f and R∞ are about 3 %and 0.03 % respectively, whereas for 1.0 mol/L solution therespective errors are less than 1 % and 0.01 %.

with too small L in all-atom MD. For instance, a simulationcell that is too small may not have enough ions to properlyscreen localized dipoles or charges. As well, the partition-ing of the voltage drop across the cell can yield a strongfield across the pore for small simulation cells (as there isless voltage drop in the bulk medium), which can introducenon-linearities in the transport properties. This further sup-ports the use of the scaling analysis, as it will also allowone to identify these “non-scaling” finite size effects andeliminate them when going to large enough simulation cells.Moreover, it goes without saying that computations shouldbe done as close to experimental and biologically-relevantconditions as possible. Our work gives new reasons to dosimulations in the linear response regime, as it is only thenthat some non-scaling finite size effects are negligible forsimulation cell sizes in the 16 nm regime.

In addition, the classical form of access resistanceassumes an infinite bulk, which fails for micro- and nano-scale systems. In particular, if one has a narrow (micro- ornano-fluidic) constriction leading up to a membrane/pore orsome other “active” region, then access resistance no longerfollows the Maxwell-Hall expression, but rather has to becorrected for the geometric setup (56). Often simulationsof such scenarios will be performed by using finite-element

The golden aspect ratio 9

solvers with the exact setup and sizes mapped from exper-iment. This will be fine when the pore/membrane are notcontextual. Our results show that one may be able to modelsuch a setup by using Eq. (6), but, instead of using the goldenaspect ratio, to properly enlarge H and L (or both cross-sectional lengths) to match the device. This is modifying thescaling theory to capture the convergence of the current tothe actual device, so one can use results from smaller sim-ulation sizes (e.g., on the scale of 10 nm for a single lineardimension) to model a device that is on a larger scale (e.g.,100’s of nm). Essentially, this is important when there areactually finite-size (confinement) effects, as well as contex-tual properties that continuum simulations cannot capture, inthe device one wants to simulate.

The golden aspect ratio will likely depend on contextualproperties of the pores. For instance, charges present on/inthe pore will decrease the resistivity within the Debye length,reducing the access contribution to the resistance. Hence,this potentially can decrease the golden aspect ratio (and alsogive a non-scaling finite size effect). Further scaling analysesof different classes of pores (i.e., charged, strongly fluctuat-ing, etc.) will give a set of heuristic values of this specialratio. Using such values will minimize the variation of theresistance with L, reducing errors in the scaling analysis.This approach will allow the quantitative extraction of bothpore and access contributions to the resistance. Its employ-ment has already shed light on how to define pore size forgeometrically imperfect pores, including edge fluctuationsand dehydration/van der Waals interactions (34). Its furtheremployment will give the first calculated values of accessresistance in the presence of, e.g., molecular-scale/proteinfluctuations in biological ion channels and other contex-tual properties (charges/dipoles) and open a new era incomparing computed and measured values of resistance.

Appendix A: Access resistance in a finite region

Here, we calculate access resistance for a finite size regionwhile keeping the ellipsoidal symmetry. This problem canbe solved using the rotational elliptic coordinates. We beginwith oblate spheroidal coordinates, (µ, ν, φ), which aredefined in terms of the Cartesian coordiates as

x = a coshµ cos ν cosφ

y = a coshµ cos ν sinφ

z = a sinhµ sin ν (19)

Rotational elliptic coordinates, ξ = sinhµ and η = sin ν,are related to the cylindrical coordinates as

z = aξη

ρ = a√

(1 + ξ2)(1− η2), (20)

where a is the radius of the pore. Laplace’s equation in thiscoordinate system is

∂ξ

[(1 + ξ2)

∂V

∂ξ

]+

∂η

[(1− η2)

∂V

∂η

]= 0 (21)

This equation can be solved via separation of variables,V = H(η)Ξ(ξ), which gives

∂ξ

[(1 + ξ2)

∂Ξ

∂ξ

]+ λΞ = 0 (22)

and∂

∂η

[(1− η2)

∂H

∂η

]− λH = 0, (23)

where λ is a constant. The boundary condition (ii), ∂V/∂η =0 at η = 0, yields λ = 0. Integrating Eq. (22) thus gives

Ξ =

∫b1

1 + ξ2dξ = b1 tan−1 ξ + b2, (24)

where b1 and b2 are the constants of integration. Using theboundary condition (i), V (ξ = 0) = 0, we get b2 = 0.This simplifies the potential to V = Hb1 tan−1 ξ. Theboundary condition (iii), V (ξ = l/a) = V0, gives V0 =Hb1 tan−1(l/a). Taking the ratio of these two expressionsgives Eq. (4) of the main text,

V

V0=

tan−1 ξ

tan−1(l/a). (25)

The current through the pore is

I =2π

γ

∫ a

0

∂V

∂z

∣∣∣z=0

ρdρ =2π

γ

∫ a

0

1

∂V

∂ξ

∣∣∣ξ=0

ρdρ

=2πaVo

γ tan−1(l/a). (26)

Equations (25) and (26) give Eq. (5).

AUTHOR CONTRIBUTION

M.Z. proposed the project. S.S. performed all calculations.Both authors wrote the manuscript and clarified the ideas.

ACKNOWLEDGMENTS

We thank J. Elenewski, D. Gruss, and C. Rohmann forhelpful comments. S. Sahu acknowledges support underthe Cooperative Research Agreement between the Univer-sity of Maryland and the National Institute of Standardsand Technology Center for Nanoscale Science and Tech-nology, Award 70NANB14H209, through the University ofMaryland.

References

1. Hille, B., 2001. Ion channels of excitable membranes, volume507. Sinauer Sunderland, MA.

2. Kandel, E. R., J. H. Schwartz, T. M. Jessell, S. A. Siegelbaum,A. J. Hudspeth, et al., 2000. Principles of neural science,volume 4. McGraw-hill New York.

3. Doyle, D. A., J. a. M. Cabral, R. A. Pfuetzner, A. Kuo, J. M.Gulbis, S. L. Cohen, B. T. Chait, and R. MacKinnon, 1998.The Structure of the Potassium Channel: Molecular Basis ofK+ Conduction and Selectivity. Science 280:69–77.

10 Sahu and Zwolak

4. Zwolak, M., and M. Di Ventra, 2008. Colloquium: Physicalapproaches to DNA sequencing and detection. Rev. Mod. Phys.80:141–165.

5. Heerema, S. J., and C. Dekker, 2016. Graphene nanodevicesfor DNA sequencing. Nat. Nanotechnol. 11:127–136.

6. Baker, R. W., 2000. Membrane Technology. Wiley OnlineLibrary.

7. Mulder, J., 2012. Basic principles of membrane technology.Springer Science & Business Media.

8. Garaj, S., W. Hubbard, A. Reina, J. Kong, D. Branton, andJ. Golovchenko, 2010. Graphene as a subnanometre trans-electrode membrane. Nature 467:190–193.

9. Merchant, C. A., K. Healy, M. Wanunu, V. Ray, N. Peterman,J. Bartel, M. D. Fischbein, K. Venta, Z. Luo, A. T. C. Johnson,and M. Drndic, 2010. DNA translocation through graphenenanopores. Nano Lett. 10:2915–2921.

10. Schneider, G. F., S. W. Kowalczyk, V. E. Calado, G. Pandraud,H. W. Zandbergen, L. M. Vandersypen, and C. Dekker, 2010.DNA Translocation through Graphene Nanopores. Nano Lett.10:3163–3167.

11. Heiranian, M., A. B. Farimani, and N. R. Aluru, 2015. Waterdesalination with a single-layer MoS2 nanopore. Nat. Com-mun. 6.

12. Walker, M. I., K. Ubych, V. Saraswat, E. A. Chalklen,P. Braeuninger-Weimer, S. Caneva, R. S. Weatherup, S. Hof-mann, and U. F. Keyser, 2017. Extrinsic Cation Selectivity of2D Membranes. ACS Nano 11:1340–1346.

13. Joshi, R., P. Carbone, F. Wang, V. Kravets, Y. Su, I. Grig-orieva, H. Wu, A. Geim, and R. Nair, 2014. Precise and ultra-fast molecular sieving through graphene oxide membranes.Science 343:752–754.

14. Abraham, J., K. S. Vasu, C. D. Williams, K. Gopinadhan,Y. Su, C. T. Cherian, J. Dix, E. Prestat, S. J. Haigh, I. V. Grig-orieva, A. K. Geim, and R. R. Nair, 2017. Tunable sievingof ions using graphene oxide membranes. Nat. Nanotechnol.12:546–550.

15. Esfandiar, A., B. Radha, F. C. Wang, Q. Yang, S. Hu, S. Garaj,R. R. Nair, A. K. Geim, and K. Gopinadhan, 2017. Sizeeffect in ion transport through angstrom-scale slits. Science358:511–513.

16. Cohen-Tanugi, D., and J. C. Grossman, 2012. Water Desalina-tion across Nanoporous Graphene. Nano Lett. 12:3602–3608.

17. Surwade, S. P., S. N. Smirnov, I. V. Vlassiouk, R. R. Uno-cic, G. M. Veith, S. Dai, and S. M. Mahurin, 2015. Waterdesalination using nanoporous single-layer graphene. Nat.Nanotechnol. 10:459–464.

18. Jiang, D., V. R. Cooper, and S. Dai, 2009. Porous grapheneas the ultimate membrane for gas separation. Nano Lett.9:4019–4024.

19. Kim, H. W., H. W. Yoon, S.-M. Yoon, B. M. Yoo, B. K. Ahn,Y. H. Cho, H. J. Shin, H. Yang, U. Paik, S. Kwon, et al.,2013. Selective gas transport through few-layered grapheneand graphene oxide membranes. Science 342:91–95.

20. Clarke, J., H.-C. Wu, L. Jayasinghe, A. Patel, S. Reid, andH. Bayley, 2009. Continuous base identification for single-molecule nanopore DNA sequencing. Nat. Nanotechnol.4:265–270.

21. Si, W., and A. Aksimentiev, 2017. Nanopore Sensing ofProtein Folding. ACS Nano 11:7091–7100.

22. Zwolak, M., J. Lagerqvist, and M. Di Ventra, 2009. Quan-tized Ionic Conductance in Nanopores. Phys. Rev. Lett.103:128102.

23. Zwolak, M., J. Wilson, and M. Di Ventra, 2010. Dehydrationand ionic conductance quantization in nanopores. J. Phys.:

Condens. Matter 22:454126.24. Sahu, S., M. Di Ventra, and M. Zwolak, 2017. Dehydration

as a Universal Mechanism for Ion Selectivity in Graphene andOther Atomically Thin Pores. Nano Lett. 17:4719–4724.

25. Sahu, S., and M. Zwolak, 2017. Ionic selectivity and fil-tration from fragmented dehydration in multilayer graphenenanopores. Nanoscale 9:11424–11428.

26. Hall, J. E., 1975. Access Reistance of a Small Circular Pore.J. Gen. Physiol. 66:531–532.

27. Hille, B., 1968. Pharmacological modifications of the sodiumchannels of frog nerve. J. Gen. Phsiol. 51:199–219.

28. Lauger, P., 1976. Diffusion-limited ion flow through pores.Biochim. Biophys. Acta 455:493–509.

29. Alcaraz, A., M. L. Lopez, M. Queralt-Martın, and V. M.Aguilella, 2017. Ion Transport in Confined Geometries belowthe Nanoscale: Access Resistance Dominates Protein ChannelConductance in Diluted Solutions. ACS Nano .

30. Vodyanoy, I., and S. Bezrukov, 1992. Sizing of an ion pore byaccess resistance measurements. Biophys. J. 62:10–11.

31. Bezrukov, S. M., and I. Vodyanoy, 1993. Probing alamethicinchannels with water-soluble polymers. Effect on conductanceof channel states. Biophys. J. 64:16–25.

32. Bezrukov, S. M., I. Vodyanoy, R. A. Brutyan, and J. J.Kasianowicz, 1996. Dynamics and free energy of poly-mers partitioning into a nanoscale pore. Macromolecules29:8517–8522.

33. Maffeo, C., S. Bhattacharya, J. Yoo, D. Wells, and A. Aksi-mentiev, 2012. Modeling and simulation of ion channels.Chem. Rev. 112:6250–6284.

34. Sahu, S., and M. Zwolak, 2017. Maxwell-Hall access resis-tance in graphene nanopores. arXiv:1708.03327 .

35. Yoo, J., and A. Aksimentiev, 2015. Molecular dynamics ofmembrane-spanning DNA channels: conductance mechanism,electro-osmotic transport, and mechanical gating. J PhysChem Lett 6:4680–4687.

36. Maxwell, J. C., 1881. A treatise on electricity and magnetism,volume 1. Clarendon press.

37. Gray, A., and G. B. Mathews, 1895. A treatise on Besselfunctions and their applications to physics. Macmillan andCompany.

38. Brown, H. T., and F. Escombe, 1900. Static diffusion ofgases and liquids in relation to the assimilation of carbon andtranslocation in plants. Proc. Roy. Soc. London 67:124–128.

39. Grober, H., 1921. Die Grundgesetze der Warmeleitung unddes Warmeuberganges: ein Lehrbuch fur Praxis und technis-che Forschung. Springer-Verlag, Berlin.

40. Holm, R., 1958. The contact resistance. General theory.Springer.

41. Newman, J., 1966. Resistance for Flow of Current to a Disk.J. Electrochem. Soc. 113:501–502.

42. Sharvin, Y. V., 1965. A Possible Method for Studying FermiSurfaces. Sov. Phys. JETP 21:655.

43. Wexler, G., 1966. The size effect and the non-local Boltzmanntransport equation in orifice and disk geometry. Proc. Phys.Soc. 89:927.

44. Nikolic, B., and P. B. Allen, 1999. Electron transport througha circular constriction. Phys. Rev. B 60:3963–3969.

45. Panday, N., G. Qian, X. Wang, S. Chang, P. Pandey, and J. He,2016. Simultaneous Ionic Current and Potential Detection ofNanoparticles by a Multifunctional Nanopipette. ACS Nano10:11237–11248.

46. Read, M. B., J. H. Lang, A. H. Slocum, and R. Martens, 2009.Contact Resistance in Flat Thin Films. In Proceedings ofthe 55th IEEE Holm Conference on Electrical Contacts. IEEE,

The golden aspect ratio 11

303–309.47. Braunovic, M., N. K. Myshkin, and V. V. Konchits, 2006.

Electrical contacts: Fundamentals, applications and technol-ogy. CRC press.

48. Abramowitz, M., and I. A. Stegun, editors, 1964. Hand-book of mathematical functions: with formulas, graphs, andmathematical tables. Courier Corporation.

49. Mikic, B. B., 1967. Thermal contact resistance. Ph.D. thesis,Massachusetts Institute of Technology.

50. Cooper, M., B. Mikic, and M. Yovanovich, 1969. Thermalcontact conductance. Int J Heat Mass Transfer 12:279–300.

51. Levadny, V., V. M. Aguilella, and M. Belaya, 1998. Accessresistance of a single conducting membrane channel. Biochim-ica et Biophysica Acta (BBA) - Biomembranes 1368:338–342.

52. Robinson, R. A., and R. H. Stokes, 2002. Electrolyte solutions.Courier Corporation.

53. Onsager, L., and S. K. Kim, 1957. Wien effect in simple strongelectrolytes. J. Phys. Chem. 61:198–215.

54. Sathe, C., X. Zou, J.-P. Leburton, and K. Schulten, 2011.Computational investigation of DNA detection using graphenenanopores. ACS Nano 5:8842–8851.

55. Aksimentiev, A., and K. Schulten, 2005. Imaging α-hemolysinwith molecular dynamics: ionic conductance, osmotic per-meability, and the electrostatic potential map. Biophys. J.88:3745–3761.

56. Green, Y., R. Eshel, S. Park, and G. Yossifon, 2016. Interplaybetween nanochannel and microchannel resistances. NanoLett. 16:2744–2748.