diffusive transport of two charge equivalent and structurally similar

19
Diffusive transport of two charge equivalent and structurally similar ruthenium complex ions through graphene oxide membranes Mike Coleman 1 , and Xiaowu (Shirley) Tang 1,2 () Nano Res., Just Accepted Manuscript • DOI: 10.1007/s12274-014-0593-x http://www.thenanoresearch.com on September 24, 2014 © Tsinghua University Press 2014 Just Accepted This is a “Just Accepted” manuscript, which has been examined by the peer-review process and has been accepted for publication. A “Just Accepted” manuscript is published online shortly after its acceptance, which is prior to technical editing and formatting and author proofing. Tsinghua University Press (TUP) provides “Just Accepted” as an optional and free service which allows authors to make their results available to the research community as soon as possible after acceptance. After a manuscript has been technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Please note that technical editing may introduce minor changes to the manuscript text and/or graphics which may affect the content, and all legal disclaimers that apply to the journal pertain. In no event shall TUP be held responsible for errors or consequences arising from the use of any information contained in these “Just Accepted” manuscripts. To cite this manuscript please use its Digital Object Identifier (DOI®), which is identical for all formats of publication. Nano Research DOI 10.1007/s12274-014-0593-x

Upload: doandat

Post on 31-Dec-2016

218 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Diffusive transport of two charge equivalent and structurally similar

Nano Res

1

Diffusive transport of two charge equivalent and

structurally similar ruthenium complex ions through

graphene oxide membranes

Mike Coleman1, and Xiaowu (Shirley) Tang1,2 ()

Nano Res., Just Accepted Manuscript • DOI: 10.1007/s12274-014-0593-x

http://www.thenanoresearch.com on September 24, 2014

© Tsinghua University Press 2014

Just Accepted

This is a “Just Accepted” manuscript, which has been examined by the peer-review process and has been

accepted for publication. A “Just Accepted” manuscript is published online shortly after its acceptance,

which is prior to technical editing and formatting and author proofing. Tsinghua University Press (TUP)

provides “Just Accepted” as an optional and free service which allows authors to make their results available

to the research community as soon as possible after acceptance. After a manuscript has been technically

edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP

article. Please note that technical editing may introduce minor changes to the manuscript text and/or

graphics which may affect the content, and all legal disclaimers that apply to the journal pertain. In no event

shall TUP be held responsible for errors or consequences arising from the use of any information contained

in these “Just Accepted” manuscripts. To cite this manuscript please use its Digital Object Identifier (DOI®),

which is identical for all formats of publication.

Nano Research

DOI 10.1007/s12274-014-0593-x

Page 2: Diffusive transport of two charge equivalent and structurally similar

Nano Res.

TABLE OF CONTENTS (TOC)

Diffusive transport of two charge equivalent and

structurally similar ruthenium complex ions through

graphene oxide membranes

Mike Coleman, Xiaowu (Shirley) Tang*

Department of Chemistry & Waterloo Institute for

Nanotechnology, University of Waterloo, Canada

.

This paper presents a detailed study of the diffusive transport of

Ru(bpy)32+ and Ru(phen)3

2+ across graphene oxide membranes, which

help identify the effects of steric interactions on ion transport, apart

from a matrix of other mechanisms.

Page 3: Diffusive transport of two charge equivalent and structurally similar
Page 4: Diffusive transport of two charge equivalent and structurally similar

Nano Res.

Diffusive transport of two charge equivalent and

structurally similar ruthenium complex ions through

graphene oxide membranes

Mike Coleman1, and Xiaowu (Shirley) Tang1,2 ()

Received: day month year

Revised: day month year

Accepted: day month year

(automatically inserted by

the publisher)

© Tsinghua University Press

and Springer-Verlag Berlin

Heidelberg 2014

KEYWORDS

Graphene oxide; ion

transport; hindered

diffusion; nanofluidics

ABSTRACT

Here, we report a study of ion transport across GO membranes of various

thicknesses, made by vacuum filtration of GO aqueous solutions. The diffusive

transport rates of two charge-equivalent Ruthenium complex ions Ru(bpy)32+

and Ru(phen)32+, with a sub-angstrom size difference, are distinguishable

through GO membranes and their ratio can be a unique tool for probing the

transport-relevant pore structures. Pore and slit-dominant hindered diffusion

models are presented and correlated to experimental results. Our analysis

suggests that ion transport is mostly facilitated by large pores (larger than 1.75

nm in diameter) in the relatively thin GO membranes, while slits formed by GO

stacking (less than 1.42 nm in width) become dominant only in thick

membranes. By grafting PEG molecules to the lateral plane of GO sheets,

membranes with enlarged interlayer spacing were engineered, which showed

drastically increased ion transport rates and lower distinction among the two

Ruthenium complex ions, consistent with the prediction by the slit-dominant

steric hindered diffusion model.

Nano Research

DOI (automatically inserted by the publisher)

Research Article

Page 5: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

2 Nano Res.

1. Introduction

Recently there has been an increased use of porous

carbon nanomaterials to replace current water and

air purification materials as well as the increased

interest in their use for nanofluidic applications such

as separations, DNA manipulation and sensing [1-4].

The synthesis of conventional mesoporous carbon

materials requires silica or other inorganic material

templates as well as multi-step synthesis procedures

using select carbon precursors [5-7]. The resulting

pore geometry is determined by a variety of

synthesis parameters, such as template material and

concentration, carbon precursor, synthesis

temperature, and catalyst concentration [5, 7, 8]. On

the other hand, carbon nanomaterials such as carbon

nanotubes and graphene offer intrinsic nanoporous

geometries that do not require precise control of

synthesis parameters and are easier to fabricate with

user controlled pores, lateral dimensions and

thicknesses. Vast functionalization options are also

available to alter their chemical properties and the

use of graphene materials offers an intrinsic,

well-defined interlayer spacing as well as the

possibility for additional user-defined pores when

subjected to electron beams or oxidative etching [3,

9-13]. Graphene oxide (GO), a chemical derivative of

graphene, consists of crystalline graphitic domains,

highly disordered oxygenated regions, and

nanosized pores [14-16]. These structural

characteristics make GO a 2D amphiphilic and

solution-processable material which can be easily

incorporated into 3D macrostructures. GO sheets can

be stacked to construct desalination and molecular

separation membranes with massive arrays of

nanofluidic channels showing enhanced

hydrodynamic flow, surface-governed transport

phenomena and ion selectivity. This implicates GO as

an ideal material for membrane applications due to

its natural, well-defined geometry and chemical

versatility that can alter the permeability and

transport characteristics for a variety of functions.

Currently, the many mechanisms governing ion

permeation and selectivity through GO membranes

remain debatable [1, 2, 9, 17-24]. GO membranes

have shown ion selectivity and enhanced cation

transport due to electrostatic attraction from the

negatively charged functional groups [17, 18, 20].

Contradicting these claims, though, it has been

identified that cations and anions diffuse through

the membranes stoichiometrically in order to

maintain charge neutrality [19]. In this situation, the

highly disordered sp3 oxygen-containing functional

groups act as steric blockers and do not contribute

to the transport characteristics, only maintaining

the interlayer spacing between GO layers and

allowing the intercalation of water where the

pristine sp2 graphitic regions provide enhanced

flow conditions [19]. Additional studies have

challenged this contradiction and have shown that

the sp3 oxygen-containing functional groups not

only interact with the permeating ions and

contribute to the selectivity, but also reduce the

expected flow enhancement of water through these

nanochannels [1, 22, 23, 25]. In addition, the

permeation and selectivity has been associated to

the interaction strength between the metal ions and

GO membranes, where the permeation of heavy

metal salts are significantly lower than ionic salts

due to the tight coordination between the

heavy-metal ion and the GO functional groups [1].

Furthermore, the balance between cation-π

interactions with the pristine sp2 graphitic regions

and ion desolvation has also been claimed as the

reason for ion selectivity through GO membranes.

For ions with high solvation energies, the hydration

shell will effectively screen and neutralize the ionic

charge and thus lower cation-π interactions,

allowing for faster ion permeation [23].

Furthermore, two mechanisms are commonly

credited for enhanced ion transport, which are the

capillary force in nanochannels formed by the

stacking of individual GO sheets [1,2,9,17-25], and

the slip flow of water in pristine graphitic regions

similar to that inside carbon nanotubes

Address correspondence to Xiaowu (Shirley) Tang, [email protected]

Page 6: Diffusive transport of two charge equivalent and structurally similar

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

3 Nano Res.

[2,9,19,23-25].

2. Experimental

GO was synthesized using a modified Hummers

method from natural graphite, as reported

previously [30-32]. To be brief, graphite flakes (3.0 g,

100 mesh, Sigma-Aldrich) and 1.5 g NaNO3 were

mixed with 70 mL concentrated H2SO4, stirred for 30

minutes and transferred to an ice bath. Following this,

15.0 g KMnO4 was added slowly to keep the

temperature below 20 oC. After all the KMnO4 was

added, the mixture was heated to 35 oC and mixed

for 2 hours. The mixture was carefully diluted with

140 mL deionized (DI) water and stirred for another

hour. The flask was then removed from the water

bath and the mixture was cooled to room

temperature, filtered, and washed several times with

5% HCl and DI water. The final solution had a GO

concentration of 3.4 mg/mL. GO membranes were

prepared by vacuum filtration of certain volumes of

diluted aqueous GO solution (0.1 mg/mL) onto

polycarbonate (PC) membranes (25 mm diameter, 0.1

μm pore size; Whatman) (Fig. 1a).

A set of GO membranes was made with 0.1, 0.3,

0.5, 1.0, 1.5 mg of GO respectively, and air dried at

room temperature. Fig. 1b is a scanning electron

microscopy (SEM) image of a 1.0 mg membrane,

Figure 1 GO membrane characteristics. (a) Optical images of a GO aqueous solution and a GO membrane (2.5 cm in diameter) prepared

by vacuum filtration. A schematic illustration and an optical image (inset) of the ion transport measurement setup. (b) Cross-sectional

SEM image of a 1.0 mg GO membrane. (c) GO membrane thickness vs. GO mass. The line is a linear fit to the data points. Error bars

are ±1 standard deviation. (d) X-ray diffraction spectrum of a 1.0 mg GO membrane.

Page 7: Diffusive transport of two charge equivalent and structurally similar

Nano Res.

showing it as a highly compact layered structure

made of individually stacked GO sheets. The GO

membrane thickness, determined based on

cross-sectional SEM images, increases linearly with

GO mass, to a maximum of 3.5 μm at 1.5 mg (Fig. 1c).

The X-ray diffraction (XRD) spectrum of the same 1.0

mg membrane showed a single 2θ peak at 11°,

corresponding to a interlayer spacing of 0.7 nm

between stacked GO sheets, which is in agreement

with values in previous literature reports [17,19].

The experimental setup for measuring ion

transport through a GO membrane is also shown in

Fig. 1a. For the ease of handling, each GO membrane

was sandwiched between two PC filter membranes

and two polydimethylsiloxane (PDMS, Sylgard 184,

1:10 catalyst: resin ratio) O-rings with a 4 mm inner

diameter, which defined the active membrane area to

be 12.6 mm2. For ion transport measurements, the

GO-PC-PDMS assembly was clamped between two

polystyrene (PS) cuvettes, which served as the feed

and permeate reservoirs. Both reservoirs were filled

with 2 mL DI water for 24 h to effectively hydrate the

GO membranes prior to transport studies. At the

start of each transport experiment, the DI water in

the feed reservoir was replaced by 2 mL 20 mM

Ru(bpy)32+ or Ru(phen) 32+ solution. Ru(bpy) 32+ and

Ru(phen)32+ solutions were made by dissolving

tris(2,2-bipyridyl) dichlororuthenium(II) hexahydrate

(Sigma-Aldrich) and dichlorotris(1,10-phenanthroline)

ruthenium(II) hydrate (Sigma-Aldrich) into DI water

respectively. Small aliquots of permeate solution

were sampled at various time points over 24 h.

Optical absorbance at 450 nm and 448 nm, measured

using a UV-vis spectrometer (PerkinElmer Lambda

25), were used to determine the concentrations of

Ru(bpy)32+ and Ru(phen)32+ in the permeate solutions.

A reference device with a PC-PDMS assembly

clamped between the PS reservoirs was also

constructed and ion transport through it was

measured to identify the contributions from the

supporting PC membranes alone (without GO).

3. Results and Discussion

The diffusive transport of both Ru(bpy)32+ and

Ru(phen)32+ through various GO membranes were

evident, since the ion concentrations in the

permeate reservoirs increased steadily over time.

Using the permeate ion concentrations measured,

the moles of ions permeated through the GO

membranes over 24 h were calculated and plotted

in Fig. 2a & b. The net flow of ions from the feed to

permeate reservoir was driven solely by the

concentration difference (ΔC) across the membrane,

which was imposed to be 20 mM initially and

assumed to be constant (less than 10% change)

during the transport measurements. Ion permeation

was recorded over a much shorter period (2 h) on

the reference device (without GO), due to the

substantially higher net flow of ions observed (Fig.

2c). By applying linear regression to the data points

in Fig. 2a & b, and with a known active membrane

area, molar flow rate per unit area (mmol/h·m2), i.e.

molar flux (J), was obtained for each membrane and

ion for quantitative comparison (Table 1). For the

1.5 mg membrane, only the data point at 24 h was

used. The molar fluxes of both Ru(bpy)32+ (Jbpy) and

Ru(phen)32+ (Jphen) are inversely proportional to the

GO membrane thickness (Fig. 2d) as predicted by

diffusive ion transport. The permeation rate of

Ru(bpy)32+ through GO membranes is on the same

order of magnitude as the rate through aligned

carbon nanotube membranes [33].

Table 1 Molar fluxes of Ru(bpy)32+ and Ru(phen)3

2+ through

GO membranes

GO mass

(mg)

Jbpy

(mmol/h·m2)

Jphen

(mmol/h·m2)

0.1 14.16 8.83

0.3 8.91 6.50

0.5 7.43 4.91

1.0 3.39 2.15

1.5 0.73 0.29

Page 8: Diffusive transport of two charge equivalent and structurally similar

Nano Res.

Figure 2 Permeation of Ru(bpy)32+ and Ru(phen)3

2+ through GO membranes. (a & b) Moles of Ru(bpy)32+ and Ru(phen)3

2+ transported

from the feed to the permeate reservoir through GO membranes of various masses over 24 h. (c) Moles of Ru(bpy)32+ and Ru(phen)3

2+

permeated through the reference PC membrane (no GO) over 2 h. Error bars are ± 1 standard deviation and the lines are linear fits. (d)

Molar fluxes are inversely proportional (fitted lines) to GO membrane thickness.

However, the results shown in Table 1 contradict

the recently reported findings by Joshi et al., which

showed no Ru(bpy)32+ permeation through GO

membranes [19]. In this work, the thickest

membrane examined was 3.5 μm, which is

considerably thinner than the 5 μm GO films used

by Joshi et al.. In addition, the lateral size of our GO

sheets is around 220 nm as determined by dynamic

light scattering (DLS, Fig. S1), which is smaller than

the 1 μm sheet size employed in the theoretical

calculation of Ru(bpy)32+ permeation rate by Joshi

et al. Both factors could lead to higher molar flux

through our membranes and therefore allow

experimental detection of permeated ions within

hours in this work.

Using the Stokes-Einstein equation and the

Wilke-Chang correlation, the theoretical Stokes

radii and bulk diffusivities of the two ions can be

calculated, which are 0.633 nm and 3.872x10-10 m2/s

for Ru(bpy)32+, 0.664 nm and 3.696x10-10 m2/s for

Ru(phen)32+ [34]. The experimentally determined

bulk diffusivity of Ru(bpy)32+ using cyclic

Page 9: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

6 Nano Res.

voltammetry is similar to the theoretically

calculated value.35 Following the approach used by

Joshi et al., theoretical molar fluxes were calculated

at ΔC = 20 mM for a GO sheet size L=220 nm [19].

The experimental molar fluxes are slightly higher

than (up to 10 times) the theoretical values. This

result indicates that both capillary force [18] and

low-friction hydrodynamic flow [23-25], which lead

to much enhanced (3 orders and higher) ion

transport through carbon nanotube and graphene

membranes reported in literature, might not be

significant in the permeation of the relatively large

sized ions through GO membranes, which is

studied in this work. A recent study by Wei et al.

also revealed that the enhancement to water

permeation through GO membranes by boundary

slip is not significant [25].

Besides the absolute molar fluxes of the two

Ruthenium complex ions, we looked into the ratio of

their flow rates, i.e. Jphen/Jbpy. For 0.1 – 1.0 mg GO

membranes, the ratios vary from 0.64 to 0.66, while

the thicker 1.5 mg GO membrane shows a lower ratio

of 0.41±0.07 respectively. Noteworthy is that the two

ions, Ru(bpy)32+ and Ru(phen)32+, are charge

equivalent and structurally similar. They are

different by only 5% in bulk diffusivity and 10% in

molecular weight. Consequently, the effects of

electrostatic interaction [1, 17, 18, 20, 23] ion

desolvation [23], and capillary force [19, 24] on ion

permeation should be about the same for the two

ions, and cannot account for the significant deviation

of Jphen/Jbpy from unity. Since the pores in a GO

membrane are of the same order as those of the

Ruthenium ions, steric hindrance to ion diffusion

should be important and very sensitive to small

variations in ion size. Our analysis, as presented in

the following paragraphs, shows that the large

difference in the flow rates of Ru(bpy)32+ and

Ru(phen)32+ can be explained by sterically hindered

diffusion, and the ratios (Jphen/Jbpy) could give us more

insight into the pore structures of the GO

membranes.

Three types of pores inside a GO membrane

can facilitate ion transport [2, 17, 19, 22, 23, 25, 36,

37]. First, ions can translocate through nanopores

created by vacancies, edges and cracks within

individual GO sheets. Mechanical handling and/or

the harsh chemical exfoliation and oxidation

processes involved in GO synthesis can create holes

in the graphitic sheets between 1 to 15 nm in lateral

size [15, 16, 36, 37]. Second, ions can transport

through the voids between individual GO sheets

that are formed during the filtration process due to

unmatched or misaligned edge geometries, which

has been speculated to be only a few GO layers

thick [19]. Third, ions can translocate through the

2D nanochannels formed between stacked GO

sheets due to the ordered assembly of GO during

the filtration process. The first and second types can

be modeled as cylindrical pores (Fig. 3a) and the

third type can be modeled as slits with a width

dictated by the GO interlayer spacing (Fig. 3b),

which is reported to be 0.7 to 1.42 nm [17,19,38].

Based on Fick’s First law, the diffusive molar

flux (J) of a solute driven by a concentration (C)

gradient along the pore axis (x) is as following,

( )dC

J H Ddx

(1)

where D is the bulk diffusion coefficient, H(λ) is the

overall hindrance coefficient, and λ is the

dimensionless number calculated by dividing the

radius of the solute a by the radius or half-width of

the pore h, as schematically shown in Fig. 3. H(λ)

takes into account the diffusive hindrance Kd(λ) and

the equilibrium partition coefficient Φ(λ) [39,40],

( ) ( ) ( )dH K (2)

The hard-shell approximation of Φ(λ), which is the

ratio of the average concentration in the pore to the

bulk concentration, is defined as,

2(1 )

( )(1 )

cylindrical pores

slit pores

(3)

when only steric interaction is considered.

Page 10: Diffusive transport of two charge equivalent and structurally similar

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

7 Nano Res.

Figure 3 Pore structure and steric hindrance to ion diffusion. (a & b) Schematic drawings of the pores in a GO membrane modelled as

either cylindrical pores or slits. The ion radius, pore half width, and relative ion size are denoted as a, h, and λ = a/h, respectively. (c)

Comparison of theoretical hindrance coefficient ratios for slit and cylindrical geometries as a function of pore size (left panel) to the

experimental ratios of Ru(phen)32+ to Ru(bpy)3

2+ molar flow rates (right panel). The dashed lines correlate the experimentally

determined ratios to pore sizes.

For H(λ) in cylindrical pores, Equation (3a) was

used to calculate Φ(λ) and the Bungay and Brenner

centreline approximation was used to caluculate

Kd(λ), which has been shown to provide the best

estimation for nanofiltration devices and an

accurate correlation for a wide range of pore sizes

[39-42]. For nanoslit analysis, a cross-sectional

average least-squares fit approximation was used to

calculate H(λ), since it has been shown to work best

for the slit geometry [40]. Using the Stokes radii of

Page 11: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

8 Nano Res.

the two ions, H(λ) values for Ru(bpy)32+ (Hbpy) and

Ru(phen)32+ (Hphen) were calculated for pores up to 5

nm (i.e. h up to 2.5 nm). It is expected that steric

hindrance is negligible for pores greater than 5 nm.

For cylindrical pores of 1.35-5 nm in size, one

obtains Hbpy ranging from 2.4x10-6 to 0.27 and Hphen

ranging from 8.1x10-9 to 0.25. For slit pores, one

obtains Hbpy ranging from 0.014 to 0.507 and Hphen

ranging from 0.003 to 0.491.

There are some caveats to note when applying

those calculations for ions, instead of an uncharged

solute. Both Φ(λ) and Kd(λ) are known to be

modulated by electrostatic interactions. It has been

shown that Φ(λ) decreases away from the hard-shell

approximation (Eq. 3) with decreasing ionic strength

(defined by the dimensionless parameter κa, where κ

is the Debye-Hückel parameter and a is the solute

radius) for repulsive electrostatic interactions

whereas it increases towards the hard-shell

approximation, or becomes even greater than one, for

attractive interactions [43-45]. Graphene oxide

consists of epoxide and hydroxyl functional groups

that populate the basal planes and carboxylic, phenol,

and hydroxyl groups along the edges [14]. These

negatively charged functional groups have shown

surface-charge-governed ion transport characteristics

[17, 18, 22]. As a result, the solute-wall interaction will

be attractive and it is expected that Φ(λ) will be equal

to or greater than the hard-shell approximation for

Ru(bpy)32+ and Ru(phen)32+ diffusion. Furthermore,

capillary force enacting on the ions in the

nanochannel could also make Φ(λ) greater than 1 [19].

In addition, for attractive ion-wall interactions, Kd(λ)

should be slightly lower than the value calculated by

centreline approximation, since ions would be drawn

off the centerline and experience more drag from the

pore wall. Therefore, it is anticipated that the

aforementioned calculated H(λ) values would be

lower than their true values. However, the calculated

ratio of Hphen to Hbpy should be accurate and only

dependent on steric hindrance, since contributions

from electrostatic interactions are identical for the

two charge-equivalent ions.

Fig. 3c (left panel) shows the theoretical ratio

(Hphen/Hbpy) as a function of pore size for both

cylindrical and slit pores. This H ratio is equivalent to

the theoretical ratio of Jphen to Jbpy, as long as the ion

concentration gradients across the pores are the same.

Regardless of pore shape, the ratio increases rapidly

towards unity as the pore size increases, which is

expected. For pores greater than 3 nm, the ratio

approaches unity and any distinction between the

permeation of Ru(bpy)32+ and Ru(phen)32+ as a result

of steric interactions is lost. The theoretical values

were further compared to the experimentally

measured ratio of flow rates through GO membranes,

as plotted in Fig. 3d (right panel). For GO

membranes 0.1 mg up to 1 mg, Jphen/Jbpy ranged from

0.63 to 0.66, which are consistent with the theoretical

ratios for slit pores with a width around 1.45 nm

(0.68) and/or cylindrical pores with diameters around

1.75 nm (0.66). Since the highest reported value for

the interlayer spacing of fully-hydrated GO in

literature is 1.42 nm [17], we anticipate that the ion

transport pathways in these relatively thin

membranes are mainly through holes within GO

sheets, voids formed between GO sheet edges, and

large slits formed by folds/wrinkles. For the thickest

membranes (1.5 mg) tested in this study, the ratio

Jphen/Jbpy dropped significantly to 0.41, indicating that

high-resistance pathways consisting of narrower slits

(width around 1.375 nm) and/or smaller holes/voids

(diameter around 1.61 nm) come into play. It has

been shown that the independent stacking of 5 or

more GO layers exponentially reduces the effects of

micrometer-sized tears and nanometer-sized intrinsic

defects on gas permeation through GO [37]. Defects

on one GO layer are covered up by subsequent layers

unless the defects randomly align. In the 1.5 mg

membrane, the effective area and length of the pores

formed by aligned porous defects and non-uniform

geometries should be minimal compared to those of

the massive array of slits formed by GO sheet

stacking. As a result, ion permeation is primarily

Page 12: Diffusive transport of two charge equivalent and structurally similar

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

9 Nano Res.

limited by the intrinsic interlayer spacing (< 1.42 nm)

and would follow the theoretical predictions of

slit-dominant behaviour.

Further, we explored the possibility to modulate

ion permeability of GO membranes through

engineering of the interlayer spacing, i.e. slit width.

GO was functionalized with linear polyethylene

glycol (PEG-L) using carbodiimide chemistry (see

ESM), as shown schematically in Fig. 4a.

Figure 4 Engineering of interlayer spacing by PEGylation of GO. (a) Schematic of GO grafted with linear PEG via carbodiimide

chemistry. (b) AFM image of individual GO-PEG-L sheets on a mica substrate. (c) XRD spectra of GO-PEG-L and GO membranes,

with the interlayer diffraction peak of GO-PEG-L identified. (d) Cross-sectional SEM image of a 1.0 mg GO-PEG-L membrane. (e)

Moles of Ru(bpy)32+ and Ru(phen)3

2+ transported through 1.5 mg GO-PEG-L films over 24 h.

The grafting of PEG-L onto the lateral surface of

the GO sheets, which is shown evidently in the AFM

image of GO-PEG-L (Fig. 4b), is expected to increase

the interlayer spacing using PEG as the steric

blockers to the stacking of GO. GO-PEG-L

membranes were made following the same vacuum

filtration procedure. XRD spectra of GO and

GO-PEG-L membranes with the same GO mass (1.0

mg) showed a shift in the 2θ peak from 11° to 7.5°

(Fig. 4c), corresponding to an increase in the dry

interlayer spacing from 0.7 to 1.1 nm. The SEM image

of a GO-PEG-L membrane (Fig. 4d) showed that it

maintained a layered morphology similar to that of

pristine GO films. Comparing the thicknesses of

membranes with the same GO mass (1.0 mg), the

PEGylation of GO significantly increased the overall

membrane thickness from approximately 3 μm (GO)

to 4.5 μm (GO-PEG-L). Both the XRD and the SEM

data support the effective enlargement of interlayer

spacing by using PEGylated GO.

Following the same ion transport

measurement procedure as done on GO membranes,

the moles of Ru(bpy)32+ and Ru(phen)32+ permeated

through thick 1.5 mg GO-PEG-L membranes over

24 h were recorded and shown in Fig. 4e. The molar

fluxes Ru(bpy)32+ and Ru(phen)32+ were derived to

be 18.71 and 12.74 mmol/h·m2 respectively.

Compared to the pristine 1.5 mg GO films, the ion

permeation rates increased by 2500-4300% despite

the increased membrane thickness. This can be

Page 13: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

10 Nano Res.

attributed to the increased interlayer spacing

facilitating Ru(bpy)32+ and Ru(phen)32+ permeation.

Furthermore, taking the ratio Jphen/Jbpy for the 1.5 mg

GO-PEL-L membrane resulted in a ratio of 0.68, an

increase of 0.26 compared to pristine 1.5 mg GO

films which had a ratio of 0.41. This result strongly

supports our conclusion that thick GO membranes

possess a slit-dominant pore structure and steric

hindrance is a dominant mechanism governing ion

permeation through GO membranes.

4. Conclusions

In summary, we report a study of ion transport

across GO membranes of various thicknesses, made

by vacuum filtration of GO aqueous solutions. The

diffusive transport rates of two charge-equivalent

and structurally similar Ruthenium complex ions,

Ru(bpy)32+ and Ru(phen)32+, with a sub-angstrom size

difference, through GO membranes are

distinguishable. Taking the ratio of the molar fluxes

of the two ions (Jphen/ Jbpy) allowed us to identify the

effect of steric interactions on ion transport, apart

from the many other mechanisms proposed in

literature, such as electrostatic interactions, capillary

force, boundary-slip enhanced water flow, ion

desolvation, etc. Furthermore, by correlating the

experimentally determined ratio Jphen/ Jbpy with the

theoretically calculated ratio of hindrance coefficients

(Hphen/Hbpy) in cylindrical and slit pore models, we

were able to gain some insight into the pore

structures of GO membranes. Our analysis suggests

that ion transport is mostly facilitated by large pores

( ≥ 1.75 nm in diameter) in the relatively thin GO

membranes, while slits formed by GO stacking (less

than 1.42 nm in width) become dominant only in

thick membranes, such as the 3.5 μm membranes (1.5

mg) shown in this work. By grafting PEG molecules

to the lateral plane of GO sheets, membranes with

enlarged interlayer spacing were engineered, which

showed drastically increased ion transport rates and

lower distinction among the two Ru ions, consistent

with the prediction by the slit-dominant steric

hindered diffusion model. This work expands the

versatility of GO membranes for molecular sieving

and separation applications by adding another class

of molecules in addition to alkali and alkaline earth

ions and the larger organic dyes already reported.

Furthermore, our findings could give guidance to the

rational design of GO membranes for high-precision

ion selectivity and molecular sieving.

Acknowledgements

This research is financially supported by a Discovery

grant from the Natural Sciences and Engineering

Research Council of Canada (NSERC).

Electronic Supplementary Material: Supplementary

material (GO lateral dimensions obtained by

dynamic light scattering, theoretical molar flow rate

and enhancement factor calculations, hindered

diffusion models, and synthesis and characterization

of PEGylated GO films) is available in the online

version of this article at

http://dx.doi.org/10.1007/s12274-***-****-*

(automatically inserted by the publisher). References

[1] Sun, P.; Zhu, M.; Wang, K.; Zhong, M.; Wei, J.; Wu, D.; Xu,

Z.; Zhu, H. Selective Ion Penetration of Graphene Oxide

Membranes. ACS Nano 2013, 7, 428-437.

[2] Hu, M.; Mi, B. Enabling graphene oxide nanosheets as

water separation membranes. Environ. Sci. Technol. 2013, 47,

3715-3723.

[3] Schneider, G.; Kowalczyk, S.; Calado, V.; Pandraud, G.;

Zandbergen, H.; Vandersypen, L.; Dekker, C. DNA translocation

through graphene nanopores. Nano. Lett. 2010, 10, 3163-3167.

[4] Wells, D.; Belkin, M.; Comer, J.; Aksimentiev, A. Assessing

Graphene Nanopores for Sequencing DNA. Nano. Lett. 2012,

12, 4117-4123.

[5] Ignat, M.; Van Oers, C.; Vernimmen, J.; Mertens, M.;

Potgieter-Vermaak, S.; Meynen, V.; Popovici, E.; Cool, P.

Textural property tuning of ordered mesoporous carbon

obtained by glycerol conversion using SBA-15 silica as

template. Carbon 2010, 48, 1609-1618.

Page 14: Diffusive transport of two charge equivalent and structurally similar

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

11 Nano Res.

[6] Han, S.; Hyeon, T. Novel silica-sol mediated synthesis of high

surface area porous carbons. Carbon 1999, 37, 1645-1647.

[7] Ryoo, R.; Joo, S.; Kruk, M.; Jaroniec, M. Ordered

mesoporous carbon. Adv. Mater. 2001, 13, 677-681.

[8] Bagshaw, S.; Prouzet, E.; Pinnavaia, T. Templating of

mesoporous molecular sieves by non-ionic polyethylene oxide

surfactants. Science 1995, 269, 1242-1244.

[9] Mi, B. Graphene oxide membranes for ionic and molecular

sieving. Science 2014, 343, 740-742.

[10] Georgakilas, V.; Otyepka, M.; Bourlinos, A.; Chandra, V.;

Kim, N.; Kemp, K.; Hobza, P.; Zboril, R.; Kim, K.

Functionalization of Graphene: Covalent and Non-Covalent

Approaches, Derivatives and Applications. Chem. Rev. 2012, 112,

6156-6214.

[11] Girit, C.; Meyer, J.; Erni, R.; Rossell, M.; Kisielowski, C.;

Yang, L.; Park, C.-H.; Crommie, M.; Cohen, M.; Louie, S.; Zettl,

A. Graphene at the edge: stability and dynamics. Science 2009,

323, 1705-1708.

[12] Koenig, S.; Wang, L.; Pellegrino, J.; Bunch, J. Selective

molecular sieving through porous graphene. Nat. Nanotechnol.

2012, 7, 728-732.

[13] O'Hern, S.; Boutilier, M.; Idrobo, J.-C.; Song, Y.; Kong, J.;

Laoui, T.; Atieh, M.; Karnik, R. Selective ionic transport

through tunable subnanometer pores in single-layer graphene

membranes. Nano Lett. 2014, 14, 1234-1241.

[14] Lerf, A.; He, H.; Forster, M.; Klinowski, J. Structure of

graphite oxide revisited. J. Phys. Chem. B 1998, 102, 4477-4482.

[15] Gómez-Navarro, C.; Meyer, J.; Sundaram, R.; Chuvilin, A.;

Kurasch, S.; Burghard, M.; Kern, K.; Kaiser, U. Atomic structure

of reduced graphene oxide. Nano Lett. 2010, 10, 1144-1148.

[16] Erickson, K.; Erni, R.; Lee, Z.; Alem, N.; Gannett, W.;

Zettl, A. Determination of the Local Chemical Structure of

Graphene Oxide and Reduced Graphene Oxide. Adv. Mater.

2010, 22, 4467-4472.

[17] Raidongia, K.; Huang, J. Nanofluidic ion transport through

reconstructed layered materials. J. Am. Chem. Soc. 2012, 134,

16528- 16531.

[18] Guo, W.; Cheng, C.; Wu, Y.; Jiang, Y.; Gao, J.; Li, D.;

Jiang, L. Bio-Inspired Two-Dimensional Nanofluidic

Generators Based on a Layered Graphene Hydrogel Membrane.

Adv. Mater. 2013, 25, 6064-6068.

[19] Joshi, R.K.; Carbone, P; Wang, F.C.; Kravets, V.G.; Su, Y.;

Grigorieva, I.V.; Wu, H.A.; Geim, A.K.; Nair, R.R. Precise and

ultrafast molecular sieving through graphene oxide membranes.

Science 2014, 343, 752-754.

[20] Han, Y.; Xu, Z.; Chao, G. Ultrathin Graphene

Nanofiltration Membrane for Water Purification. Adv. Funct.

Mater. 2013, 23, 3693–3700.

[21] Huang, H.; Song, Z.; Wei, N.; Shi, L.; Mao, Y.; Wing, Y.;

Sun, L.; Xu, Z.; Peng, X. Ultrafast viscous water flow through

nanostrand-channelled graphene oxide membranes. Nat.

Commun. 2013, 4:2979 doi: 10.1038/ncomms3979.

[22] Huang, H.; Mao, Y.; Liu, Y.; Sun, L.; Peng, X. Salt

concentration, pH and pressure controlled separation of small

molecules through lamellar graphene oxide membranes. Chem.

Commun. 2013, 49, 5963-5965.

[23] Sun, P.; Zheng, F.; Zhu, M.; Song, Z.; Wang, K.; Zhong,

M.; Wu, D.; Little, R.; Zhu, H. Selective Trans-Membrane

Transport of Alkali and Alkaline Earth Cations through

Graphene Oxide Membranes Based on Cation−π Interactions.

ACS Nano 2014, 8, 850-859.

[24] Nair, R. R.; Wu, H. A.; Jayaram, P. N.; Grigorieva, I. V.;

Geim, A.K. Unimpeded Permeation of Water Through

Helium-Leak-Tight Graphene-Based Membranes. Science 2012,

335, 442−444.

[25] Wei, N.; Peng, X.; Xu, Z. Understanding water permeation

in graphene oxide membranes. ACS Appl. Mater. Interfaces

2014, 6, 5877-5883.

[26] Maloney, D.J.; MacDonnell, F.M.

Λ-Tris(1,10-phenanthroline-N,N’)-ruthenium(II)

Bis(hexafluorophosphate)-Acetonitrile-Diethyl Ether (1/1/0.5).

Acta Cryst Sect. C. 1997, 53, 705-707.

[27] Rillema, D.P. and Jones, D.S. Structure of

Tris(2,2’-bipyridyl) ruthenium(II) hexafluorophosphate,

[Ru(bipy)3] [PF6]2; X-Ray Crystallographic Determination. J.

Chem. Soc. Chem. Comm 1979, 19, 849-851.

[28] Moret, M.-E.; Tavernelli, I.; Rothlisberger, U. A Combined

QM/MM and Classical Molecular Dynamics Study of

[Ru(bpy)3]2+ in Water. J. Phys Chem. B. 2009, 113, 7737-7744.

[29] Szymczak, J.J.; Hofmann, F.D.; Meuwly, M. Structure and

dynamics of solvent shells around photoexcited metal complexes.

Phys. Chem. Chem. Phys. 2013, 15, 6268-6277.

[30] Hummers, W.; Offeman, R. Preparation of graphitic oxide.

J. Am. Chem. Soc. 1958, 80, 1339-1339.

Page 15: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

12 Nano Res.

[31] Kovtyukhova, N.; Ollivier, P.; Martin, B.; Mallouk, T.;

Chizhik, S.; Buzaneva, E.; Gorchinskiy, A. Layer-by-Layer

Assembly of Ultrathin Composite Films from Micron-Sized

Graphite Oxide Sheets and Polycations. Chem. Mater. 1999, 11,

771-778.

[32] Gao, X.; Tang, X.-W. Effective reduction of graphene

oxide thin films by a fluorinating agent: Diethylaminosulfur

trifluoride. Carbon 2014, 76, 133-140.

[33] Majumder, M.; Chopra, N.; Hinds, B. Effect of Tip

Functionalization on Transport through Vertically Oriented

Carbon Nanotube Membranes. J. Am. Chem. Soc. 2005, 127,

9062-9070.

[34] Wilke, C.; Chang, P. Correlation of diffusion coefficients

in dilute solutions. A.I. Ch.E. 1955, 1, 264-270.

[35] Martin, C.; Rubinstein, I.; Bard, A. The Heterogeneous Rate

Constant for the Ru(bpy)32+/3+ Couple at a Glassy Carbon

Electrode in Aqueous Solution. J. Electroanal. Chem. 1983,

151, 267-271.

[36] O'Hern, S.; Stewart, C.; Boutilier, M.; Idrobo, J.-C.;

Bhaviripudi, S.; Das, S.; Kong, J.; Laoui, T.; Atieh, M.; Karnik,

R. Selective molecular transport through intrinsic defects in a

single layer of CVD graphene. ACS Nano 2012, 6,

10130-10138.

[37] Boutilier, M.S.H.; Sun, C.; O’Hern, S.C.; Au, H.;

Hadjiconstantinou, N.G.; Karnik, R. Implications of permeation

through intrinsic defects in graphene on the design of

defect-tolerant membranes for gas separation. ACS Nano 2014, 8,

841-849.

[38] Wei, N.; Peng, X.; Xu, Z. Breakdown of fast water

transport in graphene oxides. Phys.Rev.E. 2014, 89, 012133.

[39] Deen, W. Hindered Transport of Large Molecules in

Liquid-Filled Pores. A.I. Ch.E 1987, 33, 1409-1425.

[40] Dechadilok, P.; Deen, W. Hindrance Factors for Diffusion

and Convection in Pores. Ind. Eng. Chem. Res. 2006, 45,

6953-6959.

[41] Silva, V.; Pradános, P.; Palacio, L.; Hernández, A.

Alternative pore hindrance factors: What one should be used for

nanofiltration modelization? Desalination 2009, 245, 606-613.

[42] Bungay, P.M.; Brenner, H. The motion of a closely-fitting

sphere in a fluid-filled tube. Int. J. Multiph. Flow 1973, 1,

25-56.

[43] Chun, M.-S.; Phillips, R.J. Electrostatic Partitioning in Slit

Pores by Gibbs Ensemble Monte Carlo Simulation. A.I. Ch.E

1997, 43, 1194-1203.

[44] Dechadilok, P.; Deen, W. Electrostatic and electrokinetic

effects on hindered diffusion in pores. J. Membr. Sci. 2009, 336,

7-16.

[45] Chen, S.B. Electrostatic Interaction and Hindered Diffusion

of Ion-Penetrable Spheres in a Slit Pore J. Colloid Interf.

Sci.1998, 205, 354-364.

Page 16: Diffusive transport of two charge equivalent and structurally similar

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

Nano Res.

Electronic Supplementary Material

Diffusive transport of two charge equivalent and

structurally similar ruthenium complex ions through

graphene oxide membranes

Mike Coleman1, and Xiaowu (Shirley) Tang1,2 ()

Supporting information to DOI 10.1007/s12274-****-****-* (automatically inserted by the publisher)

1. Supporting Figures

Figure S1 Size distribution of GO sheets comprising the membranes determined by dynamic light scattering (DLS).

Page 17: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

Nano Res.

Figure S2 (a) Adsorption spectra of (a) Ru(bpy)32+ and (b) Ru(phen)3

2+ in permeate solutions compared with those of references in DI

water.

The permeate solutions were collected from devices with a 1 mg GO membrane at 24 h and diluted 100 times

prior to UV-vis measurement. The absorption peak at λ = 450±2 nm for Ru(bpy)32+ and 447±2 nm for

Ru(phen)32+ are associated with a metal to ligand charge transfer (MLCT) between ruthenium and the

2,2’-bipyridine (or 1,10-phenanthroline) ligand [1]. The overlap of peak locations between permeate and

reference spectra indicates that the Ru complex ions remained intact through the GO membranes and there

was no ligand dissociation caused by the interactions with GO. The magnitude difference between the two

complexes for the same reference concentration is due to the difference in their molar extinction coefficients,

which are 14 600 L M-1 cm-1 for Ru(bpy)32+ and 19 000 L M-1 cm-1 for Ru(phen)32+ [1].

2. Determination of Theoretical Molar Flow Rates and Enhancement Factors

To calculate the theoretical molar flow rates (MFR) through the network of nanochannels in GO membranes,

analysis was performed similar to Joshi et al. [2]. Using the following equation,

eff

eff

AMFR D C

L

(S1)

where D is the diffusion coefficient of Ru(bpy)32+ or Ru(phen)32+ in water, ΔC is the concentration gradient

across the membrane, Aeff and Leff are the effective area and length of the membrane, respectfully. All films

have an exposed membrane area of 0.1256 cm2, thus Aeff is 0.1256 cm2 x d/L = 5.7x10-4 cm2 where L ~ 220.2 nm

(the lateral sheet dimensions determined by Figure S1), and d ~ 1 nm (approximate width of the

nanochannels formed by the interlayer spacing of GO). Furthermore, Leff is 220.2 nm x h/d = 6.2x10-2 cm for a

1.0 mg film where h is the membrane thickness obtained by SEM as shown in Figure 1c in the main text. For

Ru(bpy)32+ permeation through a 1.0 mg GO membrane using a 20 mM feed, we obtain a molar flow rate of

6.63×10-13 mol/s or a molar flux of 0.19 mmol/hr m2. This results in an enhancement factor of ~17, which is

obtained by dividing the experimentally obtained molar flow rates by the theoretical flow rate, shown in

Table S1 below. Including the effect of confinement in the theoretical calculation (which increases the overall

diffusion by a factor of 3/2), we still obtain an enhancement factor of ~10.

Page 18: Diffusive transport of two charge equivalent and structurally similar

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

Nano Res.

3. Determination of Theoretical Molar Flow Rates and Enhancement Factors

To calculate the theoretical molar flow rates (MFR) through the network of nanochannels in GO membranes,

analysis was performed similar to Joshi et al. [2]. Using the following equation,

eff

eff

AMFR D C

L

(S1)

where D is the diffusion coefficient of Ru(bpy)32+ or Ru(phen)32+ in water, ΔC is the concentration gradient

across the membrane, Aeff and Leff are the effective area and length of the membrane, respectfully. All films

have an exposed membrane area of 0.1256 cm2, thus Aeff is 0.1256 cm2 x d/L = 5.7x10-4 cm2 where L ~ 220.2 nm

(the lateral sheet dimensions determined by Figure S1), and d ~ 1 nm (approximate width of the

nanochannels formed by the interlayer spacing of GO). Furthermore, Leff is 220.2 nm x h/d = 6.2x10-2 cm for a

1.0 mg film where h is the membrane thickness obtained by SEM as shown in Figure 1c in the main text. For

Ru(bpy)32+ permeation through a 1.0 mg GO membrane using a 20 mM feed, we obtain a molar flow rate of

6.63×10-13 mol/s or a molar flux of 0.19 mmol/hr m2. This results in an enhancement factor of ~17, which is

obtained by dividing the experimentally obtained molar flow rates by the theoretical flow rate, shown in

Table S1 below. Including the effect of confinement in the theoretical calculation (which increases the overall

diffusion by a factor of 3/2), we still obtain an enhancement factor of ~10.

Table S2: Enhancement Factors for Ru(bpy)32+ and Ru(phen)3

2+ through GO membranes

GO mass (mg) Enhancement (Ru(bpy)32+) Enhancement (Ru(phen)3

2+)

0.1 0.67 0.44

0.3 5 3.82

0.5 6.99 4.89

1.0 17.53 11.64

1.5 4.5 1.94

4. Hindered Diffusion Models

For the calculation of H(λ) in cylindrical pores, the Bungay and Brenner centreline approximation (Equations

S2 & S3) was used along with Φ(λ) = (1-λ)2 as it has been shown to provide the best estimation for

nanofiltration devices and it provides a complete correlation for the dimensionless parameter [3-6],

6d

t

KK

(S2)

with

2 4

5/223

1 0

92 1 1 (1 )

4

n nn n

n n

Kt a a

(S3)

where λ is the dimensionless number calculated by dividing the radius of the solute a by the radius or

half-width of the pore h. The constants used for the calculations are found in Table S3 below.

Table S3: Constants for the Bungay and Brenner centreline approximation of Kd in cylindrical pores

a1 a2 a3 a4 a5 a6 a7

-1.2167 1.5336 -22.5083 -5.6117 -0.3363 -1.216 1.647

For slit nanopore analysis, a cross-sectional average least-squares fit approximation (Equation S4) is used [4].

Page 19: Diffusive transport of two charge equivalent and structurally similar

| www.editorialmanager.com/nare/default.asp

Nano Res.

3 4 59( ) 1 ln 1.19358 0.4285 0.3192 0.08428

16H (S4)

5. Synthesis and Characterization of linear Poly(ethylene) glycol functionalized Graphene Oxide

(GO-PEG-L)

GO was grafted using conventional carbodiimide chemistry which has been successfully used to graft

amine-terminated polypeptides and other small molecules onto oxidized carbon nanotubes [7-9]. In a typical

experiment, 25g mg of GO (~7 mL at [GO] = 3.4 mg/mL) was added to 13 mL of

2-(N-morpholino)ethanesulfonic acid (MES) buffer containing 50 mg amine-terminated linear poly(ethylene)

glycol (Poly(ethylene glycol) bis(amine), Mw 20 000, Sigma-Aldrich) as well as equimolar

1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) (GBiosciences) and N-Hydroxysuccinimide (NHS)

(Alfa Aesar) in excess. The mixture was kept stirring at 25°C for 24 hours. The GO-PEG-L solution was

dialyzed against Milli-Q water for 3 days with daily water changes and stored for future use.

For Atomic Force Microscopy (AFM) analysis, GO-PEG-L solution was drop casted onto freshly cleaved mica

substrates. AFM measurements were performed with a Nanoscope MultiMode™ AFM instrument (Veeco) in

tapping mode at a scan rate of 1.5 Hz using a silicon probe with a resonant frequency of 300 kHz. For X-ray

diffraction (XRD) measurements of GO and GO-PEG-L films, a Bruker AXS D8 Advance X-ray diffractometer

was used.

References

[1] Kalyanasundarum, K. Photophysics, photochemistry and solar energy conversion with

tris(bipyridyl)ruthenium(II) and its analogues. Coord. Chem. Rev. 1982, 46, 159-244

[2] Joshi, R.K.; Carbone, P; Wang, F.C.; Kravets, V.G.; Su, Y.; Grigorieva, I.V.; Wu, H.A.; Geim, A.K.; Nair, R.R.

Precise and ultrafast molecular sieving through graphene oxide membranes. Science 2014, 343, 752-754

[3] Deen, W. Hindered Transport of Large Molecules in Liquid-Filled Pores. A.I. Ch.E 1987, 33, 1409-1425

[4] Dechadilok, P.; Deen, W. Hindrance Factors for Diffusion and Convection in Pores. Ind. Eng. Chem. Res.

2006, 45, 6953-6959

[5] Silva, V.; Pradános, P.; Palacio, L.; Hernández, A. Alternative pore hindrance factors: What one should be

used for nanofiltration modelization? Desalination 2009, 245, 606-613

[6] Bungay, P.M.; Brenner, H. The motion of a closely-fitting sphere in a fluid-filled tube. Int. J. Multiph. Flow

1973, 1, 25-56

[7] Majumder, M.; Chopra, N.; Hinds, B. Effect of tip functionalization on transport through vertically oriented

carbon nanotube membranes. J. Am. Chem. Soc. 2005, 127, 9062-9070.

[8] Hinds, B.; Chopra, N.; Rantell, T.; Andrews, R.; Gavalas, V.; Bachas, L. Aligned multiwalled carbon nanotube

membranes. Science 2004, 303, 62-65.

[9] Nednoor, P.; Gavalas, V.; Chopra, N.; Hinds, B.; Bachas, L. Carbon nanotube based biomimetic

membranes: mimicking protein channels regulated by phosphorylation. J. Mater. Chem. 2007, 17, 1755-1757.