ArticlePDF Available

Using Al3+ to Tailor Graphene Oxide Nanochannels: Impact on Membrane Stability and Permeability

MDPI
Membranes
Authors:

Abstract and Figures

Graphene oxide (GO) membranes, which form from the lamination of GO sheets, attract much attention due to their unique nanochannels. There is much interest in controlling the nanochannel structures and improving the aqueous stability of GO membranes so they can be effectively used in separation and filtration applications. This study employed a simple yet effective method of introducing trivalent aluminum cations to a GO sheet solution through the oxidation of aluminum foil, which modifies the nanochannels in the self-assembled GO membrane by increasing the inter-sheet distance while decreasing intra-sheet spacing. The Al3+ modification resulted in an increase in membrane stability in water, methanol, ethanol, and propanol, yet decreased membrane permeability to water and propanol. These changes were attributed to strong interactions between Al3+ and the membrane oxygenated functional groups, which resulted in an increase in membrane hydrophobicity and a decrease in the intra-sheet spacing as supported by surface tension, contact angle, atomic force microscopy, and X-ray photoelectron spectroscopy measurements. Our approach for forming Al3+ modified GO membranes provides a method for improving the aqueous stability and tailoring the permeation selectivity of GO membranes, which have the potential to be implemented in vapor separation and fuel purification applications.
Content may be subject to copyright.
Citation: Stehle, Y.Y.; Robertson, E.J.;
Cortez, R.; Vlassiouk, I.V.; Bucinell,
R.B.; Olsson, K.; Kilby, L. Using Al3+
to Tailor Graphene Oxide
Nanochannels: Impact on Membrane
Stability and Permeability. Membranes
2022,12, 871. https://doi.org/
10.3390/membranes12090871
Academic Editor: Mohammad
Rezaul Karim
Received: 24 August 2022
Accepted: 6 September 2022
Published: 9 September 2022
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
published maps and institutional affil-
iations.
Copyright: © 2022 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
membranes
Article
Using Al3+ to Tailor Graphene Oxide Nanochannels: Impact on
Membrane Stability and Permeability
Yijing Y. Stehle 1 ,* , Ellen J. Robertson 2, * , Rebecca Cortez 1, Ivan V. Vlassiouk 3, Ronald B. Bucinell 1,
Katelyn Olsson 1and Luke Kilby 1
1Department of Mechanical Engineering, Union College, Schenectady, NY 12308, USA
2Chemistry Department, Union College, Schenectady, NY 12308, USA
3Center for Nanophase Materials Sciences, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
*Correspondence: stehley@union.edu (Y.Y.S.); roberte2@union.edu (E.J.R.)
Abstract:
Graphene oxide (GO) membranes, which form from the lamination of GO sheets, attract
much attention due to their unique nanochannels. There is much interest in controlling the nanochan-
nel structures and improving the aqueous stability of GO membranes so they can be effectively
used in separation and filtration applications. This study employed a simple yet effective method of
introducing trivalent aluminum cations to a GO sheet solution through the oxidation of aluminum
foil, which modifies the nanochannels in the self-assembled GO membrane by increasing the inter-
sheet distance while decreasing intra-sheet spacing. The Al
3+
modification resulted in an increase in
membrane stability in water, methanol, ethanol, and propanol, yet decreased membrane permeability
to water and propanol. These changes were attributed to strong interactions between Al
3+
and the
membrane oxygenated functional groups, which resulted in an increase in membrane hydrophobicity
and a decrease in the intra-sheet spacing as supported by surface tension, contact angle, atomic force
microscopy, and X-ray photoelectron spectroscopy measurements. Our approach for forming Al
3+
modified GO membranes provides a method for improving the aqueous stability and tailoring the
permeation selectivity of GO membranes, which have the potential to be implemented in vapor
separation and fuel purification applications.
Keywords: graphene oxide; Al3+ modification; nanochannel; permeability; stability
1. Introduction
Graphene oxide (GO) sheets are monolayers of bonded sp
2
-hybridized carbon atoms
decorated with abundant oxygenated functional groups (OFGs). These functional groups
are located on the basal planes (hydroxyl, carbonyl, and carboxyl groups) and edges (car-
boxyl groups), as described by the widely accepted Lerf–Klinowski structural model [
1
3
].
GO membranes form from the lamination of individual GO sheets, which are prone to
align parallel to a casting mold and assemble in a “brick-and-mortar” stacking configura-
tion during drying on a solid surface [
4
,
5
]. During assembly, electrostatic and hydrogen
bonding interactions between OFGs on individual adjacent GO sheets form nanochannels
in GO membranes through which polar solvents and ions can permeate. These unique
nanochannels in the laminated structure of GO membranes make them excellent candidates
for separation and filtration applications [
6
13
]. To be useful materials in these applications,
GO membranes must be tailored to be permeable to specific species of interest and exhibit
good stability and structural integrity.
The presence of OFGs, however, can greatly affect both the permeability and aqueous
stability of GO membranes. On the one hand, nanochannels within the GO membrane have
the potential to be tunable and selective for specific chemical species based on the chemical
modification to the OFGs. On the other hand, the hydrophilic OFGs can be hydrated by
water; thus, the GO membranes readily dissolve in water. While strong
π
π
interactions
Membranes 2022,12, 871. https://doi.org/10.3390/membranes12090871 https://www.mdpi.com/journal/membranes
Membranes 2022,12, 871 2 of 16
exist between individual GO sheets in GO membranes as they do in graphene membranes,
the introduction of OFGs weakens these interactions. Specifically, the OFGs cause the C-O
sp
3
carbons to become distorted, leading to amorphous regions of the membrane [
1
]. These
structural defects weaken the strength and toughness of the laminated membranes, thus
reducing their chemical, thermal, and aqueous stability. The instability of GO membranes
in water prevents them from being used in aqueous environments, which is required in
several filtration and separation applications [
14
21
]. Thus, strategies that could increase
the stability of GO membranes while tailoring the selective permeability at the same time
are of great interest. Previous studies demonstrated that either removing the OFGs or
chemically cross-linking them to each other can increase the aqueous stability and control
the permeability of GO membranes [
3
,
6
,
19
29
]. One of the most feasible and effective
strategies to improve stability, however, is the crosslinking of OFGs on the GO sheets
to multivalent metal cations [
13
16
]. The electrostatic and bonding interactions between
multivalent cations and OFGs on the edges and basal planes of adjacent GO sheets can
increase GO membrane stability by reducing negative surface charges and increasing
adhesion between GO layers. Moreover, the ionic radii and valences of different metal
cations can affect the interlayer spacing in GO membranes, which can in turn be used to
tune the permeability of GO membranes for specific molecules and ions.
Previous studies revealed that trivalent cations such as Al
3+
can interact with GO
sheets through cation-
π
, coordination, and electrostatic interactions, which increases the
interlayer spacing and improves both the integrity and water stability of the GO mem-
brane [
14
,
22
]. Al
3+
has benefits over other trivalent cations for stabilizing GO membranes,
such as the ease and low cost of using aluminum foil as a source of the cation, as we
previously reported [
24
]. Both theoretical and experimental studies on cation modified
and unmodified GO membranes focused on the relationship between interlayer spacing,
water flow, and stability of the GO membranes [
20
23
]. In these previous studies, however,
the impact of intra-layer spacing on the nanochannel permeability to different molecular
vapors was not explored, which could be critical for determining the types of environments
in which these membranes can be used for filtering specific chemical species. Our previ-
ous study revealed that Al
3+
cations can be added to the acidic GO solution with Al foil
through chemical oxidation, which caused the solution pH to increase, the O/C ratio to
decrease, and more wrinkles to form on the dried membrane surface [
24
]. In addition, the
resultant Al
3+
-modified graphene oxide (AGO) membranes showed an increased inter-
sheet distance, and a decreased intra-sheet spacing when compared with the unmodified
GO membranes [24].
In this work, we explored how the structural changes to the GO membrane upon the
addition of Al
3+
affect the as-formed GO membrane stability in and its permeability to
different polar solvents. Based on our previous results, this modification strategy induces
crosslinking between the metal cations and the OFGs and decreases the density of OFGs,
which we hypothesize will act to increase the overall stability of the membranes in polar
solvents. Moreover, the decrease in the O/C ratio and the increase in the surface roughness
as previously observed can increase the hydrophobicity of the GO membranes, which we
hypothesize will specifically increase the aqueous stability of the membrane. Finally, we
hypothesize that the increased inter-sheet distance and decreased intra-sheet spacing can
allow for tailoring both the nanochannel structure and thus the membrane permeability to
different solvents.
To test these hypotheses, we performed a series of experiments to characterize the
properties of GO sheet solutions and membranes both with and without Al
3+
modification.
Specifically, we fabricated Al
3+
-modified graphene oxide (AGO) membranes with 0.3 wt.%
Al, which is the highest addition that does not cause the membrane to crack, for scalable
fabrication [
24
]. The AGO membranes were easily formed by adding Al foil to an acidic
GO solution. In the acidic solution, the Al foil was oxidized, which gradually released Al
3+
cations to the solution. Subsequently, individual GO sheets initially well-dispersed in an
aqueous solution were brought together by crosslinking them to the positively charged
Membranes 2022,12, 871 3 of 16
Al
3+
cations, followed by a slow self-assembly into an ordered laminated structure via
drying the AGO solution on a Teflon evaporation dish. To determine the effect of Al
3+
on the GO sheet solution, we compared the viscosity and surface tension of the AGO
solution to the GO solution. The hydrophilicities and structures of the dried GO and AGO
membranes were also compared using the contact angle, atomic force microscopy (AFM),
and X-ray photoelectron spectroscopy (XPS) measurements. Finally, the stability of the
GO and AGO membranes in different polar solvents as well as the vapor permeability of
each membrane to these solvents were systematically assessed and explained through the
inter- and intra-sheets structure. This study provides valuable guidance for developing GO
membranes with tunable permeability and stability in different polar environments.
2. Materials and Methods
2.1. Materials and Membrane Fabrication
Graphene oxide water dispersions (4 mg/mL) were purchased from MSE Supplies,
Tucson, AZ, USA. The aluminum (Al) foil (0.01 in. thick, Puratronic, 99.997% purity) with
an area density of 0.65 mg/mm
2
was purchased from Alfa Aesar, Haverhill, MA, USA. All
reagents were used as received without further purification. Solutions of GO and Al (Al
concentration = 0.3 wt.%) were prepared by dissolving the Al foil in a GO solution followed
by continuous stirring for the complete dissolution of the Al foil and coordination between
the GO sheets and the Al
3+
cations. Free-standing GO and AGO membranes were obtained
through a slow layer-by-layer self-assembly process in which the GO and AGO solutions
were dried on glass slides. Ambient drying conditions were used to form all GO and AGO
membranes to prevent membrane cracking and air bubble formation.
2.2. Characterization
The viscosities of the GO and AGO solutions were measured using a Brookfield DV-E
Viscometer with an UL adapter spindle. Viscosity measurements were taken with a shear
rate of 0.5~100 rpm. The surface tension measurements were performed using the pendant
drop method on a Biolin Scientific Optical Tensiometer. Here, a hanging drop of the solution
of interest was formed using an air-tight Hamilton syringe equipped with a blunt 22-gauge
needle. Images of the droplet were collected approximately every minute for 30 min. The
droplet shapes were fitted to the Laplace–Young equation using the OneAttension software
provided by Biolin Scientific to determine the shape factor,
β
. The software then computed
the air–fluid surface tension (
γ
, mN/m) using Equation (1), in which
ρ
is the density
difference between the air and solution, R
0
is the radius of the droplet at the apex, and gis
the acceleration due to gravity.
γ=
ρR0g
β, (1)
In all surface tension measurements, the droplets were encased in a quartz cuvette to
prevent drop evaporation and agitation from air currents. The reported surface tension
data are averages of at least 10 measurements. Contact angle measurements of dried
GO and AGO membranes that had been drop-casted on glass microscope slides were
performed using the Biolin Scientific Optical Tensiometer. In these measurements, data
collection began immediately after a ~10
µ
L droplet of deionized water was dispensed
onto a membrane of interest. Images of the droplet were taken over the course of 1 min at a
rate of 0.68 frames per second. Contact angles were determined using the OneAttension
software. The reported contact angle values were the averages of the right and left sides
of the droplets. XPS data were acquired using a Thermo Scientific Model K-Alpha XPS
instrument with a 400
µ
m X-ray spot size. Wide energy range survey spectra (0–1350 eV)
were acquired by performing a subsequent analysis using an analyzer with a pass energy
of 200 eV and energy step size of 1 eV.
Membranes 2022,12, 871 4 of 16
2.3. Stability
To investigate the stability of the as-prepared GO and AGO membranes in differ-
ent polar solvents, the dried membranes were cut into pieces with dimensions of about
0.6 cm ×0.6 cm
and soaked in 15 mL of either ultrapure water, methanol, ethanol, or
propanol at room temperature. Four different stability studies were performed that var-
ied the extent to which the membranes were agitated in the solvents. In the first study,
the membranes were soaked in the solvents with no agitation. In the second study, the
membranes were gently stirred using a 10 mm long
×
3 mm diameter stir bar that rotated
at a rate of 200 rpm. In these studies, which were performed in 20 mL scintillation vials,
the stir bar was small enough that it did not collide with the membranes during stirring.
In the third study, the membranes were vigorously stirred using a 12 mm long
×
5 mm
diameter stir bar that rotated at a rate of 200 rpm. In the 20 mL scintillation vials, the larger
stir bar continuously collided with the membrane during stirring. In the fourth study,
the membranes were sonicated using a 60 Hz sonicator. Images of the membranes were
collected over time for each stability study using a digital camera to assess the extent of
membrane dissolution.
2.4. Permeability
The permeability of GO and AGO membranes to different solvents was assessed
using a simple evaporation device as shown in Figure S1 (in Supplementary Materials).
The free-standing GO and AGO membranes were cut into circular shapes with diameters
of 2 cm and sealed on the top of a glass vial. The setup was placed in an atmospheric
pressure chamber with a constant temperature and humidity. The effective permeation
area (A) for the setup was 1.27 cm
2
. Ultrapure water and high purity methanol, ethanol,
and propanol were used for permeability tests at temperatures of 30, 40, 50, 60, and 70 C.
The weight difference (M, g) of the solvent was measured by periodically monitoring the
mass of the vial for up to seven days and used in the permeation rate calculation. The flux
(J,gm
2
hr
1
) of solvent molecules across membranes was obtained by finding the slope
of the curve between the amount of solute molecule permeated per unit membrane area
(n/A) and the permeation time (t), according to Equation (2). Each recorded permeability
value is the mean of at least 5 experiments at the same temperature. According to Fick’s
first law (Equation (3)), the flux through a thin membrane is proportional to the diffusion
coefficient (D, m2/h), concentration of vapor on the two sides of the membrane (C), and
the reciprocal of the membrane thickness (x).
J=
M
At(2)
J=
DC
x(3)
3. Results and Discussion
It was previously reported that pristine GO membranes are completely impermeable
to all gasses and liquids except water vapor due to their unique nanochannel structure [
3
].
The major inner-membrane structures that dominate the nanochannels are inter-layer
galleries, intra-layer frameworks, and vacancy defects [
9
,
12
]. These nanochannels are more
generally spaces in the membrane through which molecules can permeate, rather than
well-defined tubes that can be experimentally visualized and quantified. The inter-layer
and intra-layers structures play an important role in the stacking of the GO nanosheets
within the membrane. This stacking affects the nanochannel size, which in turn affects the
membrane stability and permeability. The enlarging of the GO nanochannel by increasing
either interlayer or intra-layer spacing can accelerate the water flow or permit larger vapor
molecular to pass through the membrane. It was previously reported that for membranes
modified with cations of the same valency, a larger interlayer spacing results in greater
water permeation if the intra-layer spacing change is negligible [
19
]. Multivalent metal
Membranes 2022,12, 871 5 of 16
cation modification is effective at both improving the aqueous stability of the GO membrane
and increasing the interplanar distance, but this modification does not always result in
increased water permeation [
15
,
19
]. Because it is difficult to experimentally measure the
intra-sheet spacing, previous studies focused on understanding the effects of the interplanar
(inter-layer) distance on the nanochannel structure and the membrane stability [
9
,
20
23
].
In our previous study and as demonstrated in Figure 1, we determined that modification of
the GO membrane with Al
3+
via oxidation of Al foil resulted in an increase in the membrane
inter-layer distance, which was previously reported, and in a decrease in the membrane
intra-layer spacing, which is not often discussed [
24
]. In the studies described below, we
related the changes in the inter- and intra-layer distances upon Al
3+
modification of the GO
membrane to the membrane properties as well as stability in and permeability to different
polar solvents.
Membranes 2022, 12, x FOR PEER REVIEW 5 of 16
cation modification is effective at both improving the aqueous stability of the GO mem-
brane and increasing the interplanar distance, but this modification does not always result
in increased water permeation [15,19]. Because it is difficult to experimentally measure
the intra-sheet spacing, previous studies focused on understanding the effects of the in-
terplanar (inter-layer) distance on the nanochannel structure and the membrane stability
[9,2023]. In our previous study and as demonstrated in Figure 1, we determined that
modification of the GO membrane with Al3+ via oxidation of Al foil resulted in an increase
in the membrane inter-layer distance, which was previously reported, and in a decrease
in the membrane intra-layer spacing, which is not often discussed [24]. In the studies de-
scribed below, we related the changes in the inter- and intra-layer distances upon Al3+
modification of the GO membrane to the membrane properties as well as stability in and
permeability to different polar solvents.
Figure 1. Schematic of the nanochannel in both unmodified (top panel) and Al3+-modified (bottom
panel) GO membranes. The addition of Al3+ to the GO membrane results in an increase in the inter-
layer distance and a decrease in the intra-layer spacing.
3.1. Solution and Membrane Characterization
The GO solution in the absence of Al3+ is a liquid crystal. Its rheological-related prop-
erties such as viscosity determine the applicable fabrication methods, which may include
electrospraying, wet/dry spinning, inkjet printing techniques, etc. [3033]. Unlike the un-
modified GO solution, the AGO solution exhibited a “muddyappearance (Figure S2),
which was initially considered to be due to the formation of a gel. To fully understand the
“muddy” appearance of the AGO solution, the viscosity of the unmodified GO and AGO
solutions (Figure S3) was measured. Both the GO and AGO solutions exhibited a typical
and similar pseudoplastic behavior, or so-called viscoelastic behavior, which means the
viscosity (η) decreases and the shear stress (τ) increases with the shearing rate. Thus, the
solution viscosity cannot explain the “muddyappearance of the AGO solution. Instead,
it is likely that the change in consistency of the sheet solution upon the addition of the Al
foil was due to interactions of the Al3+ cations in solution with the OFGs on the GO sheet
surfaces.
The interactions between Al3+ and the GO sheets is supported by surface tension
measurements of the GO and AGO solutions. These surface tension measurements pro-
vide information about the extent of adsorption of a species in solution to the airwater
interface. If the surface tension of a solution does not decrease much relative to that of the
neat airwater interface (72 mN/m), then the species in the solution is not highly surface
Figure 1.
Schematic of the nanochannel in both unmodified (
top panel
) and Al
3+
-modified (
bottom
panel
) GO membranes. The addition of Al
3+
to the GO membrane results in an increase in the
inter-layer distance and a decrease in the intra-layer spacing.
3.1. Solution and Membrane Characterization
The GO solution in the absence of Al
3+
is a liquid crystal. Its rheological-related prop-
erties such as viscosity determine the applicable fabrication methods, which may include
electrospraying, wet/dry spinning, inkjet printing techniques, etc. [
30
33
]. Unlike the
unmodified GO solution, the AGO solution exhibited a “muddy” appearance (Figure S2),
which was initially considered to be due to the formation of a gel. To fully understand the
“muddy” appearance of the AGO solution, the viscosity of the unmodified GO and AGO
solutions (Figure S3) was measured. Both the GO and AGO solutions exhibited a typical
and similar pseudoplastic behavior, or so-called viscoelastic behavior, which means the
viscosity (
η
) decreases and the shear stress (
τ
) increases with the shearing rate. Thus, the
solution viscosity cannot explain the “muddy” appearance of the AGO solution. Instead,
it is likely that the change in consistency of the sheet solution upon the addition of the
Al foil was due to interactions of the Al
3+
cations in solution with the OFGs on the GO
sheet surfaces.
The interactions between Al
3+
and the GO sheets is supported by surface tension
measurements of the GO and AGO solutions. These surface tension measurements provide
information about the extent of adsorption of a species in solution to the air–water interface.
If the surface tension of a solution does not decrease much relative to that of the neat
air–water interface (72 mN/m), then the species in the solution is not highly surface active.
If the surface tension of a solution significantly decreases relative to that of the neat air–
water interface, then the species is surface active. The average surface tension vs. time
Membranes 2022,12, 871 6 of 16
data were collected for both GO and AGO solutions using a pendant drop tensiometer
and are summarized in Figure 2a. In the absence of Al
3+
, the surface tension of the GO
solution initially increased from approximately 60 mN/m to approximately 70 mN/m over
the course of around one minute. After this rapid increase, the surface tension did not
significantly change over time. The fact that the surface tension is similar to the neat air–
water interface suggests that GO sheets are not highly surface active and likely desorbed
from the interface over the first minute of interface formation. This result is consistent with
the surface structure of GO, in which the negatively charged OFGs cause the material to
be highly water-soluble. In the presence of Al
3+
, the surface tension of the AGO solution
showed a small initial increase followed by a decrease to approximately 30 mN/m over
the course of 1800 s. The decrease in surface tension over time indicates that AGO sheets
continually adsorb species to the air–water interface, potentially forming multiple layers
through OFG-Al
3+
interactions. The surface tension data suggest that the addition of Al
3+
can impact the self-assembling process of the GO sheets at the air–water interface, which
acts to increase the hydrophobicity of the sheets and decreases their solubility in water.
Membranes 2022, 12, x FOR PEER REVIEW 6 of 16
active. If the surface tension of a solution significantly decreases relative to that of the neat
airwater interface, then the species is surface active. The average surface tension vs. time
data were collected for both GO and AGO solutions using a pendant drop tensiometer
and are summarized in Figure 2a. In the absence of Al3+, the surface tension of the GO
solution initially increased from approximately 60 mN/m to approximately 70 mN/m over
the course of around one minute. After this rapid increase, the surface tension did not
significantly change over time. The fact that the surface tension is similar to the neat air
water interface suggests that GO sheets are not highly surface active and likely desorbed
from the interface over the first minute of interface formation. This result is consistent
with the surface structure of GO, in which the negatively charged OFGs cause the material
to be highly water-soluble. In the presence of Al3+, the surface tension of the AGO solution
showed a small initial increase followed by a decrease to approximately 30 mN/m over
the course of 1800 s. The decrease in surface tension over time indicates that AGO sheets
continually adsorb species to the airwater interface, potentially forming multiple layers
through OFG-Al3+ interactions. The surface tension data suggest that the addition of Al3+
can impact the self-assembling process of the GO sheets at the airwater interface, which
acts to increase the hydrophobicity of the sheets and decreases their solubility in water.
Figure 2. Effect of Al3+ modification on GO and AGO sheet solutions and membranes formed via
the layer-by-layer self-assembly method. The influence of Al3+ on the (a) airwater interfacial surface
tension of a GO sheet solution, (b) contact angle of GO membranes assembled on glass slides; AFM
image of (c) pristine GO (height scale bar 600 nm) and (d) AGO (height scale bar 1 µm) membranes
on thermal silicon oxide wafers; XPS spectra of (e) C 1s and (f) O 1s spectra of GO and AGO mem-
branes.
Our previous report showed that releasing Al3+ in the GO solution via the oxidation
of Al foil is a slow process and results in an increase in the solution pH [24]. The protona-
tion of the OFGs is highly sensitive to the solution pH, which in turn affects the extent of
crosslinking between the OFGs and Al3+. Thus, it is likely that the hydrophobicity of the
AGO membrane is impacted by both the Al3+ concentration and the time at which the GO
solution is exposed to the Al foil before membrane formation. The relative hydrophobi-
cities of GO membranes formed from solutions that were exposed to Al foil for varying
times were determined from contact angle measurements, as presented in Figure 2b. Here,
the contact angles of ultrapure water droplets with membranes casted on the glass slide
surfaces were used to estimate the hydrophobicity of the GO and AGO membranes. As
shown in Figure 2b, the contact angle of the GO membrane increased with the Al3+ con-
centration, which indicates an increase in the hydrophobicity of AGO membranes when
compared to the GO membranes. The increase in sheet hydrophobicity, as observed in the
Figure 2.
Effect of Al
3+
modification on GO and AGO sheet solutions and membranes formed via the
layer-by-layer self-assembly method. The influence of Al
3+
on the (
a
) air–water interfacial surface
tension of a GO sheet solution, (
b
) contact angle of GO membranes assembled on glass slides; AFM
image of (
c
) pristine GO (height scale bar 600 nm) and (
d
) AGO (height scale bar 1
µ
m) membranes on
thermal silicon oxide wafers; XPS spectra of (
e
) C 1s and (
f
) O 1s spectra of GO and AGO membranes.
Our previous report showed that releasing Al
3+
in the GO solution via the oxidation of
Al foil is a slow process and results in an increase in the solution pH [
24
]. The protonation
of the OFGs is highly sensitive to the solution pH, which in turn affects the extent of
crosslinking between the OFGs and Al
3+
. Thus, it is likely that the hydrophobicity of the
AGO membrane is impacted by both the Al
3+
concentration and the time at which the GO
solution is exposed to the Al foil before membrane formation. The relative hydrophobicities
of GO membranes formed from solutions that were exposed to Al foil for varying times
were determined from contact angle measurements, as presented in Figure 2b. Here, the
contact angles of ultrapure water droplets with membranes casted on the glass slide surfaces
were used to estimate the hydrophobicity of the GO and AGO membranes. As shown in
Figure 2b, the contact angle of the GO membrane increased with the Al
3+
concentration,
which indicates an increase in the hydrophobicity of AGO membranes when compared
to the GO membranes. The increase in sheet hydrophobicity, as observed in the surface
tension and contact angle measurements, is consistent with the interaction of the positively
charged Al
3+
cations with the negatively charged OFGs on the sheet surfaces, either
through electrostatic screening or chelation. Additionally, the increase in the density
Membranes 2022,12, 871 7 of 16
of wrinkles formed with the addition of Al
3+
can further increase the hydrophobicity of the
GO membranes [24].
The interactions between Al
3+
and the OFGs not only affect membrane hydrophobicity,
but also the presence of membrane features such as wrinkles, shrinkage, and cracks.
Specifically, it was observed that the muddy-like AGO solutions with an Al wt.% of
greater than 0.3 wt.% formed membranes that, upon drying, demonstrated a significant
decrease in diameter and excessive cracking. Significant membrane shrinking and cracking
was not observed for membranes formed from less muddy-like AGO solutions with an Al
wt.% of less than 0.3 wt.% (Figure S4). The decrease in the membrane diameter with an
increase in the Al wt.% supports the previous conclusion that the addition of Al
3+
decreases
the intra-sheet spacing [24].
The “brick-and-mortar” stacking configuration of laminated GO sheets introduces
mixed interactions among the basal planes and edges, which lead to ‘peak and valley’
undulations in the membrane structure. The accumulation of these undulations results in
wrinkles, which increases the surface roughness of the membrane [
1
]. Our previous report
shows that Al
3+
modification via Al foil oxidation causes more wrinkles to form on both GO
sheets and membranes, which can lead to an increase in surface roughness [
24
]. Stronger
interactions between Al
3+
and the OFGs on GO sheet surfaces can increase the ‘peak and
valley’ undulations and thus cause more wrinkles in the AGO membrane compared to the
GO membrane. Three-dimensional AFM images of the GO (Figure 2c) and AGO (Figure 2d)
membranes can be used to estimate the average surface roughness (RA) of as-fabricated GO
and AGO membranes. According to the average RA value collected from five AFM images
of the GO and AGO membranes, the AGO membrane had an RA roughness value of 98nm,
while the value of pure GO membrane was 87nm, which is consistent with the greater
density of wrinkles in the AGO membrane when compared to the GO membrane [24].
To better understand the effect of Al
3+
on the chemical composition of the dried
GO membranes, XPS data were collected and are shown in Figure 2e,f. XPS is a surface-
sensitive technique that can be used to effectively determine the elemental composition
and chemical environment of a GO membrane within the first few nanometers of its
surface. Figure 2e, which provides the overall surface composition of the GO and AGO
membranes, demonstrates that carbon and oxygen comprised more than 98 at. % of
the surface. Moreover, the C/O ratio varied between 2.3 and 2.6 for all GO and AGO
membranes studied. These results are consistent with the presence of hydroxyl, epoxy, and
carboxyl groups on the membrane surfaces, as expected. All membranes showed peaks
due to small amounts of N and S, while only the AGO membranes showed a small peak
due to the presence of Al near 75 eV (Figure 2e). As shown in Figure 2f, the primary C
1s peaks at 284.5 eV and 286.8 eV confirm the presence of C–C (or C
C/C=C) and C-O
(C
O/C
N, C
O
C) functional groups, respectively. Figure 2f also shows two small C 1s
peaks at 288.8 eV and 289.4 eV, which indicate the presence of O-C=O and C=O functional
groups. The primary O 1s peak at 532.3 eV is attributed to the presence of OFGs within
the GO membrane. Importantly, the introduction of Al
3+
to the GO membrane resulted in
a decrease in the C-C peak area and an increase in the C-O peak area in the C 1s spectra.
This result can be attributed to the coordination between Al
3+
and epoxide groups that
underwent ring opening reactions.
3.2. Membrane Stability
The instability of GO membranes in aqueous environments is a long-standing problem
for applications that require GO membranes to be used in polar solvents, which is due to the
chemical structure of the GO sheets that compose the membranes [
10
12
]. Specifically, the
hydrophilic OFGs on GO sheets cause both the sheets and laminated membranes to become
hydrophilic and hold a negative surface potential. The electrostatic repulsions between
negatively charged GO sheets within the membranes can overcome the van der Waals
attractions and hydrogen bonding interactions between sheets in the aqueous environment,
especially in the presence of shear force and sonication energy [
12
]. The low contact angle of
Membranes 2022,12, 871 8 of 16
unmodified GO indicates the high tendency of GO sheets to be solvated by water molecules
after becoming fully soaked, which further decreases the extent of interactions between
adjacent GO sheets within the GO membrane.
In this study, the stabilities of GO and AGO membranes in water and three different
polar solvents were assessed under three different degrees of agitation including soaking,
stirring, and sonication. The results of this study are illustrated in the digital camera photos
shown in Figure 3and Figures S5–S8. Compared to GO membranes typically prepared
through the vacuum filtration process, GO membranes prepared with layer-by-layer self-
assembly demonstrate better stability due to the greater number of hydrogen bonds formed
between individual GO sheets during the slow-drying process [
2
,
3
]. Specifically, previous
studies demonstrated that GO membranes formed via vacuum filtration are only stable
in water for up to an hour after soaking [
2
]. The results here show that the as-fabricated
unmodified GO membranes are stable and maintain their original morphologies after being
soaked in pure water as well as methanol, ethanol, and propanol for more than 6 months
(Figure 3, column 1; Figure S5) [24].
Membranes 2022, 12, x FOR PEER REVIEW 8 of 16
3.2. Membrane Stability
The instability of GO membranes in aqueous environments is a long-standing prob-
lem for applications that require GO membranes to be used in polar solvents, which is
due to the chemical structure of the GO sheets that compose the membranes [1012]. Spe-
cifically, the hydrophilic OFGs on GO sheets cause both the sheets and laminated mem-
branes to become hydrophilic and hold a negative surface potential. The electrostatic re-
pulsions between negatively charged GO sheets within the membranes can overcome the
van der Waals attractions and hydrogen bonding interactions between sheets in the aque-
ous environment, especially in the presence of shear force and sonication energy [12]. The
low contact angle of unmodified GO indicates the high tendency of GO sheets to be solv-
ated by water molecules after becoming fully soaked, which further decreases the extent
of interactions between adjacent GO sheets within the GO membrane.
In this study, the stabilities of GO and AGO membranes in water and three different
polar solvents were assessed under three different degrees of agitation including soaking,
stirring, and sonication. The results of this study are illustrated in the digital camera pho-
tos shown in Figures 3 and S5S8. Compared to GO membranes typically prepared
through the vacuum filtration process, GO membranes prepared with layer-by-layer self-
assembly demonstrate better stability due to the greater number of hydrogen bonds
formed between individual GO sheets during the slow-drying process [2,3]. Specifically,
previous studies demonstrated that GO membranes formed via vacuum filtration are only
stable in water for up to an hour after soaking [2]. The results here show that the as-fabri-
cated unmodified GO membranes are stable and maintain their original morphologies
after being soaked in pure water as well as methanol, ethanol, and propanol for more than
6 months (Figure 3, column 1; Figure S5) [24].
Figure 3. Stability of the GO and AGO membranes in water, methanol, ethanol, and propanol. The
GO and AGO membranes were soaked (left), vigorously stirred (middle), and sonicated (right) in
different polar solvents for different time intervals.
Figure 3.
Stability of the GO and AGO membranes in water, methanol, ethanol, and propanol. The
GO and AGO membranes were soaked (
left
), vigorously stirred (
middle
), and sonicated (
right
) in
different polar solvents for different time intervals.
However, agitation of the membranes decreased their stability over time. Under
a sonication frequency of 60 Hz, the GO membranes in water, methanol, and ethanol
disintegrated within 30 s of sonication(Figure 3, column 5), while the GO membranes in
propanol maintained their integrity during 60 s of sonication (Figure S6). Agitation of the
membranes via stirring (Figure 3, column 3), which is more representative of a filtration
working environment, resulted in GO membranes that were less stable than those soaked
in the solvents with no agitation, but more stable than those subjected to sonication. This
result is consistent with the fact that the shear force generated by stirring is much lower
than the shear force generated through sonication. During stirring, the disintegration of the
GO membrane in water exhibited an edge-to-center (intra-layer) disassembly pattern rather
than a layer-by-layer (inter-layer) disassembly pattern when the time of stirring increased.
Membranes 2022,12, 871 9 of 16
The complete dispersion of the unmodified GO membrane in water required about 96 h
of stirring. The rate of membrane agitation, and thus rate of disintegration, was adjusted
by increasing the size of the stir bar. Because the solvent volume and sample vial used
remained the same, the increase in the stir bar size resulted in collisions between the stir
bar and membrane during each stir bar rotation. For GO membranes in water, membrane
disintegration became visible after 20 min of stirring with a larger stir bar, with complete
dispersion occurring within 1 h (Figure S7). Here, the rate of membrane disintegration
was much faster than for membranes agitated by a smaller stir bar because the membrane
was exposed to both the force of the shear stress and the impact of the stir bar (Figures S7
and S8).
As seen in Figure 3, the stability of the unmodified GO membrane decreased with
decreasing solvent hydrophobicity, such that the membrane was most stable in propanol
and least stable in water. The instability of the GO membrane in aqueous solution results
from the hydration of OFGs that are both exposed on the surface of the membrane and
embedded in the membrane interior [34]. Specifically, hydrogen bonding, ion-dipole, and
dipole–dipole interactions between water molecules and hydrophilic OFGs on the surface
of the GO membrane can overcome the relatively weak van der Waals and hydrogen
bonding interactions that hold the surface-exposed sheets together. Additionally, the size
of a water molecule is likely smaller than the inter-sheet lamellar distance, or intra-sheet
framework space, and so it can permeate into the membrane interior. Here, interactions
between water molecules and OFGs located in the membrane interior can increase the
inter-sheet distances, thereby weakening the
π
-
π
interactions that hold individual sheets
together in the membrane. The increase in membrane stability with increasing solvent
hydrophobicity is likely due to both a decrease in the extent of interactions between the
solvent alcohol groups and surface-exposed membrane OFGs and an inability to permeate
into the membrane interior due to the increasing molecular size beyond that of the critical
lamellar distance.
Previous work demonstrated that the water stability AGO membranes formed via
pre-coordination followed by vacuum filtration is only improved under soaking conditions
for up to a few hours [
14
,
22
]. In our study, AGO membranes that were formed via pre-
coordination followed by a slow layer-by-layer self-assembly process remained stable under
soaking conditions in water, methanol, ethanol, and propanol for more than 6 months
(Figure S5). Thus, to assess the effect of Al
3+
modification on GO membrane stability, the
responses of the as-prepared GO and AGO membranes to stirring and sonication in the
different polar solvents were compared (Figure 3and Figures S6–S8). These results show
that when the membranes are agitated by both stirring and sonication, the AGO membranes
show a better stability and survive for longer than the unmodified GO membranes in all
four different polar solvents.
However, consistent with previous studies of GO membranes formed with vacuum
filtration, improvements in membrane stability were not dramatic when GO membranes
were modified with Al
3+
via the pre-coordination method [
2
]. As seen in Figures S6–S8,
the AGO membranes still disintegrated in water, methanol, and ethanol upon increasing
the time of stirring and sonication. Both GO and AGO membranes showed good stability
in propanol and did not disintegrate after 60 s of sonication. Moreover, the process by
which the GO and AGO membranes dissolved in pure water under stirring and sonication
were observed to be quite different. In water, the GO membrane disintegrated uniformly
into a well-dispersed brown solution with no visible particles or flakes, as observed by the
eye. Under the same conditions as the GO membrane, the AGO membrane disintegrated
non-uniformly into large black flakes consisting of AGO sheet aggregates that could be
seen by eye. In the water solution, the disintegrated AGO membrane existed as large black
colored flakes. After sonication and storage, these flakes persisted in water, methanol,
and ethanol (Figure S6). The persistence of the AGO flakes compared to the uniformly
dispersed GO membranes further demonstrates that Al
3+
modification increases the stabil-
ity of GO membranes in polar solvents. This increased membrane stability was achieved
Membranes 2022,12, 871 10 of 16
by replacing inter-sheet hydrogen bonds with cation-to-OFGs coordinate covalent bonds,
cation-
π
interactions, and electrostatic interactions. In addition to decreasing membrane hy-
drophilicity by screening the negative charges on the membrane surface, these interactions
are overall stronger than inter-sheet hydrogen bonds. Both factors are critical for increasing
the stability of GO membranes in polar solvents.
It is likely that the limited stability improvement of the AGO membrane is in part due
to its reduced water permeability. The AGO membranes modified with Al
3+
from Al foil
exhibited increasing membrane wrinkling, shrinking, and cracking after increasing the Al
wt.% in the GO solution (Figure S4). The shrinking of the AGO membrane supports our
previous report [
24
], which showed a decrease in the intra-sheet spacing after increasing
the Al
3+
wt.%. This membrane diameter decrease is consistent with stronger intra-layer
interactions and increased membrane stability. However, the highly wrinkled and con-
tracted AGO membrane exhibited swelling after being fully soaked in water, suggesting
that the membrane is permeable to water (Figure S9). The fact that water can intercalate
into the membrane interior likely prevents the membrane from being completely stable in
water under all agitation conditions. The extent of AGO membrane swelling compared to
the extent of GO membrane swelling in ultrapure water could not be reliably quantified
(Figure S10); therefore, it was difficult to assess the effect of Al
3+
modification on membrane
permeability purely from soaking in water.
In summary, the unmodified GO membrane prepared using the slow layer-by-layer
self-assembling process can maintain its integrity when soaked in water, methanol, ethanol,
and propanol, but will severely disintegrate and disperse into water, methanol, or ethanol
within 30 s of vigorous sonication or 2 h of stirring. The stability of GO in methanol and
ethanol is better than that in water, while the best stability is observed in propanol. Overall,
AGO membranes demonstrated enhanced stability in water, methanol, and ethanol, but
this improvement was limited.
3.3. Membrane Permeability
In this study, the membrane permeability of both GO and AGO to water, methanol,
ethanol, and propanol vapor was assessed by measuring the loss of solvent mass from a
bottle that was capped with the membrane of interest (Figure S1). From the loss of mass,
the flux, J, was calculated using Equation (2), and subsequently the diffusion coefficient,
D, was derived using Equation (3), which was equated to the membrane permeability.
The flux of water vapor passing through membranes with thicknesses ranging from 6
µ
M
to 12
µ
M and at temperatures ranging from 30 to 70
C was calculated first (Figure 4a).
We were unable to measure the permeability of thinner membranes due to their fragility
and thicker membranes due to their extensive wrinkling and inability to properly seal the
permeability set-up. It is expected that as the temperature increases, the flux of water vapor
through the membrane will increase due to an increase in the kinetic energy, and thus
speed, of the vapor molecules passing through the membrane. Moreover, as membrane
thickness increases, it is expected from Fick’s Law that the flux of water vapor through
the membrane will decrease due to an increase in the amount of time the molecules can
spend in the thicker membrane. As expected, the flux of water vapor through the GO
membrane increased with an increasing temperature; however, the flux also increased with
increasing membrane thickness. The flux data were used to calculate the permeability, D,
for membranes at each thickness and temperature (Figure 4b). These data indicate that the
permeability also increased with increasing temperatures and membrane thickness.
The increase in permeability with increasing membrane thickness was unexpected,
and likely resulted from how the individual GO sheets were structured differently at the
surface of the membrane compared to the interior of the membrane during drying in the
layer-by-layer self-assembly process. As suggested in the surface tension data and observed
during the experiment, GO sheets accumulated at the air–water interface and dried first
from the surface to the interior during the self-assembly process. To explore the effect of
membrane thickness on the inter-sheet structure at the surface of the GO membrane, XRD
Membranes 2022,12, 871 11 of 16
spectra of GO membranes were collected with varying thicknesses (Figure S11). These
results reveal no obvious trend in the average interplanar distance of GO membranes with
different thicknesses. Thus, it is likely that the surface structure of the membrane does not
depend on the membrane thickness. However, it is possible that the membrane surface
structure is significantly different than the structure of the membrane interior, which could
affect the flux of water through membranes with different thicknesses.
Membranes 2022, 12, x FOR PEER REVIEW 11 of 16
membrane increased with an increasing temperature; however, the flux also increased
with increasing membrane thickness. The flux data were used to calculate the permeabil-
ity, D, for membranes at each thickness and temperature (Figure 4b). These data indicate
that the permeability also increased with increasing temperatures and membrane thick-
ness.
Figure 4. Permeability of a GO membrane to water as a function of membrane thickness and tem-
perature. (a) Flux of water vapor through GO membranes with thicknesses varying from 8~12 µm
and temperatures ranging from 30 to70 °C. (b) Corresponding permeability values derived from (a).
(c) XRD spectra of GO membrane before and after exposure to oxygen plasma
The increase in permeability with increasing membrane thickness was unexpected,
and likely resulted from how the individual GO sheets were structured differently at the
surface of the membrane compared to the interior of the membrane during drying in the
layer-by-layer self-assembly process. As suggested in the surface tension data and ob-
served during the experiment, GO sheets accumulated at the airwater interface and dried
first from the surface to the interior during the self-assembly process. To explore the effect
of membrane thickness on the inter-sheet structure at the surface of the GO membrane,
XRD spectra of GO membranes were collected with varying thicknesses (Figure S11).
These results reveal no obvious trend in the average interplanar distance of GO mem-
branes with different thicknesses. Thus, it is likely that the surface structure of the mem-
brane does not depend on the membrane thickness. However, it is possible that the mem-
brane surface structure is significantly different than the structure of the membrane inte-
rior, which could affect the flux of water through membranes with different thicknesses.
To determine the difference of the inter-sheet structure between the GO membrane
surface and GO membrane interior, XRD spectra of membranes were collected before and
after exposure to oxygen plasma for varying times (Figure 4c). Here, a radio-frequency
oxygen plasma at the processing power of 60 W from 10 to 60 min was applied to remove
the surface of the GO membrane before each XRD scan. The resultant spectra show a slight
shift of the peak from 10.01°, corresponding to an average interlayer distance of 8.826 nm,
to 10.11°, corresponding to an average interlayer distance of 8.739 nm. The decrease in the
average inter-sheet distance after removal of the surface suggests that the sheets near the
membrane surface are somewhat more densely packed in the vertical direction than sheets
in the membrane interior. The formation of the densely packed membrane interior occurs
during the membrane drying process. It is likely that the GO sheets that assemble at the
airwater interface form a permeation layer that inhibits the evaporation of water mole-
cules from the bulk solution and slows down the drying process. This can result in a uni-
formly assembled membrane interior structure through which water molecules have to
slowly permeate before reaching the surface.
It is expected that as the membrane thickness increases and the interior structure ra-
ther than the surface structure dominates the properties of the membrane, permeability
decreases due to the decrease in the inter-sheet distance in the membrane interior. How-
ever, the permeation channel is controlled by the inter-sheet distance, the intra-sheet spac-
ing, and vacancy defects on the GO sheet. Specifically, and as we show in our AGO studies
below, an increase in the inter-sheet distance does not increase the membrane permeabil-
ity due to the decreased intra-sheet spacing. The finding that the permeation coefficient
Figure 4.
Permeability of a GO membrane to water as a function of membrane thickness and
temperature. (
a
) Flux of water vapor through GO membranes with thicknesses varying from 8~12
µ
m
and temperatures ranging from 30 to70
C. (
b
) Corresponding permeability values derived from (
a
).
(c) XRD spectra of GO membrane before and after exposure to oxygen plasma.
To determine the difference of the inter-sheet structure between the GO membrane
surface and GO membrane interior, XRD spectra of membranes were collected before and
after exposure to oxygen plasma for varying times (Figure 4c). Here, a radio-frequency
oxygen plasma at the processing power of 60 W from 10 to 60 min was applied to remove
the surface of the GO membrane before each XRD scan. The resultant spectra show a slight
shift of the peak from 10.01
, corresponding to an average interlayer distance of 8.826 nm,
to 10.11
, corresponding to an average interlayer distance of 8.739 nm. The decrease in
the average inter-sheet distance after removal of the surface suggests that the sheets near
the membrane surface are somewhat more densely packed in the vertical direction than
sheets in the membrane interior. The formation of the densely packed membrane interior
occurs during the membrane drying process. It is likely that the GO sheets that assemble
at the air–water interface form a permeation layer that inhibits the evaporation of water
molecules from the bulk solution and slows down the drying process. This can result in a
uniformly assembled membrane interior structure through which water molecules have to
slowly permeate before reaching the surface.
It is expected that as the membrane thickness increases and the interior structure
rather than the surface structure dominates the properties of the membrane, permeability
decreases due to the decrease in the inter-sheet distance in the membrane interior. However,
the permeation channel is controlled by the inter-sheet distance, the intra-sheet spacing, and
vacancy defects on the GO sheet. Specifically, and as we show in our AGO studies below,
an increase in the inter-sheet distance does not increase the membrane permeability due
to the decreased intra-sheet spacing. The finding that the permeation coefficient increases
with membrane thickness and larger inter-sheet distances between sheets near the surface
than in the interior supports our hypothesis that the inter-sheet distance does not dominate
the permeation channel of the GO membrane. However, the intra-sheet structure difference
between the surface and the interior of the membrane is unclear. Moreover, it is also unclear
if the ambient oxidation of the surface during the week-long drying process occurs, which
could also affect the nanochannel structure.
Previous reports revealed that metal cation-decorated GO membranes exhibited selec-
tive permeation to organic solvents [
27
], which is consistent with the results presented here.
To determine the effect of Al
3+
modification on membrane permeability, the permeabilities
of 12
µ
m thick GO and AGO membranes to four different polar solvents at temperatures
ranging from 30
C to 70
C were compared (Figure 5). A summary of the temperature-
dependent data is provided in Figure S12. Here, the temperature dependent permeability
Membranes 2022,12, 871 12 of 16
data for the GO membrane (Figure 5a) and the AGO membrane (Figure 5b) were plot-
ted according to the Arrhenius equation (Equation (4)) to better understand the kinetic
parameters affecting GO and AGO membrane permeability to different polar solvents.
lnD =lnDQd
RT (4)
Membranes 2022, 12, x FOR PEER REVIEW 12 of 16
increases with membrane thickness and larger inter-sheet distances between sheets near
the surface than in the interior supports our hypothesis that the inter-sheet distance does
not dominate the permeation channel of the GO membrane. However, the intra-sheet
structure difference between the surface and the interior of the membrane is unclear.
Moreover, it is also unclear if the ambient oxidation of the surface during the week-long
drying process occurs, which could also affect the nanochannel structure.
Previous reports revealed that metal cation-decorated GO membranes exhibited se-
lective permeation to organic solvents [27], which is consistent with the results presented
here. To determine the effect of Al3+ modification on membrane permeability, the perme-
abilities of 12 μm thick GO and AGO membranes to four different polar solvents at tem-
peratures ranging from 30 °C to 70 °C were compared (Figure 5). A summary of the tem-
perature-dependent data is provided in Figure S12. Here, the temperature dependent per-
meability data for the GO membrane (Figure 5a) and the AGO membrane (Figure 5b) were
plotted according to the Arrhenius equation (Equation (4)) to better understand the kinetic
parameters affecting GO and AGO membrane permeability to different polar solvents.
Figure 5. Arrhenius plots for (a) GO permeability and (b) AGO permeability to water, methanol,
ethanol, and propanol. The lines in (a,b) are fit to Equation (4). (c) Permeability of GO and AGO
membranes to different polar solvents at 30 °C. (d) Activation energies of diffusion for water, meth-
anol, ethanol, and propanol through the GO and AGO membranes.
𝑙𝑛𝐷 =𝑙𝑛𝐷° 𝑄𝑑
𝑅𝑇
(4)
Here, is the pre-exponential factor, Qd is the energy of activation of diffusion, T is
the absolute temperature in Kelvin, and R (8.314 J/mol K) is the gas constant. As expected,
the permeability of both GO and AGO to all solvents increased with increased tempera-
ture. The GO membrane has a water vapor flux of 68 ± 5 g/m2h at 30 , which is consistent
with other reports [11]. It is impressive that the flux of water vapor across the 12 µm thick
GO membranes can go up to 600 g/m2h at 70 , which is almost the same as the vapori-
zation rate of water at the same temperature.
At all temperatures studied, both membranes showed an increase in permeability
with a decreasing molecular size, such that the 12 µ m thick GO and AGO membranes
were most permeable to water and least permeable to propanol. Without the effect of hy-
dration by water vapor present in the atmosphere, the GO membrane permeabilities to
methanol, ethanol, and propanol (Figure 5c) were 1.32 × 10−6, 6.57 × 10−7, and 3.76 × 10−8
m2/h, respectively, at 30 , which were much lower that the water vapor permeability at
the same temperature (6.85 × 10−6 m2/h). This result is consistent with the stability studies,
Figure 5.
Arrhenius plots for (
a
) GO permeability and (
b
) AGO permeability to water, methanol,
ethanol, and propanol. The lines in (
a
,
b
) are fit to Equation (4). (
c
) Permeability of GO and AGO mem-
branes to different polar solvents at 30
C. (
d
) Activation energies of diffusion for water, methanol,
ethanol, and propanol through the GO and AGO membranes.
Here, D
is the pre-exponential factor, Q
d
is the energy of activation of diffusion, Tis
the absolute temperature in Kelvin, and R(8.314 J/mol K) is the gas constant. As expected,
the permeability of both GO and AGO to all solvents increased with increased temperature.
The GO membrane has a water vapor flux of 68
±
5 g/m
2
h at 30
C, which is consistent
with other reports [
11
]. It is impressive that the flux of water vapor across the 12
µ
m
thick GO membranes can go up to 600 g/m
2
h at 70
C, which is almost the same as the
vaporization rate of water at the same temperature.
At all temperatures studied, both membranes showed an increase in permeability
with a decreasing molecular size, such that the 12
µ
m thick GO and AGO membranes were
most permeable to water and least permeable to propanol. Without the effect of hydration
by water vapor present in the atmosphere, the GO membrane permeabilities to methanol,
ethanol, and propanol (Figure 5c) were 1.32
×
10
6
, 6.57
×
10
7
, and 3.76
×
10
8
m
2
/h,
respectively, at 30
C, which were much lower that the water vapor permeability at the
same temperature (6.85
×
10
6
m
2
/h). This result is consistent with the stability studies,
which showed that membranes were most stable in propanol and least stable in water
under vigorous stirring and sonication. A similar explanation to that of the the stability
data can be used for the permeability data. Specifically, the GO membrane was most
permeable to water because it is likely that the nanochannels are larger than the size of a
water molecule. As the molecular size increases relative to the size of the nanochannels,
the vapor becomes less permeable to the membrane. Another possible effect that was
previously reported is the effect of water vapor present in the atmosphere on membrane
permeability [
11
,
27
]. In this case, intercalating water molecules in GO-based membranes
either block or impede the alcohol molecules from passing through the membranes. As
seen in Figure 5c, Al
3+
modification decreased membrane permeability to both water and
propanol vapor, while having little effect on permeability to methanol and ethanol vapor.
Membranes 2022,12, 871 13 of 16
At 30
C, the permeability to water decreased to 1.65
×
10
6
m
2
/h (76% decrease) and the
permeability to propanol decreased to 3.20 ×109m2/h (91% decrease).
To see if the change in membrane permeability to different solvents was related to the
activation energy of diffusion, the lnD vs. 1/T data in Figure 5a,b were fit to Equation (4).
The activation energies, which can be used to quantify the energy barrier, that the vapor
molecule must overcome to diffuse through the GO or AGO membrane are reported
in Figure 5d. From these results, it can be seen that upon Al
3+
modification of the GO
membrane, the activation energy of water vapor diffusion increased from 38
±
2 kJ/mol to
60
±
3 kJ/mol, the activation energy of propanol diffusion increased from 80
±
7 kJ/mol to
100
±
20 kJ/mol, and the activation energies for methanol and ethanol diffusion did not
significantly change. Thus, it was possible to relate a decrease in membrane permeability to
an increase in the activation energy of diffusion for a given solvent through the membrane.
The decrease in GO membrane permeability to water, and the corresponding increase
in the activation energy of diffusion, upon Al
3+
modification, is consistent with our pre-
vious studies of GO and AGO membranes formed via the layer-by-layer self-assembly
method [
24
]. Our studies demonstrated that modification via post-cation intercalation
leads to GO membranes that possess controllable thicknesses and interlayer spacings.
Specifically, it was discovered that crosslinking Al
3+
to the OFGs in the GO membrane
resulted in an increase in the inter-layer distance and a decrease in intra-layer spacing. The
intra-layer framework of the nanochannels impacted the water vapor flow more than the
inter-layer gallery, which resulted in both longer and narrower nanochannels than those in
unmodified membranes as shown in Figure 1. Moreover, the surface tension and contact
angle data presented here indicate that Al
3+
modification increases the hydrophobicity
of the GO membrane due to interactions between the cations and the negatively charged
OFGs. Both a decrease in intra-layer spacing and an increase in membrane hydrophobicity
can decrease membrane permeability to water vapor. In the first case, the narrowing of
the nanochannels due to Al
3+
crosslinking can block water molecules from permeating
through the membrane. In the second case, the edge-to-edge coordination of Al
3+
with the
carboxylate groups on the edges of the GO sheets can inhibit permeability by consuming
the OFGs of GO sheets such that water molecules cannot interact with these hydrophilic
groups. For methanol, ethanol, and propanol, which are all more hydrophobic than water,
it is likely that the increase in membrane hydrophobicity upon Al
3+
modification does not
greatly affect GO membrane permeability to these solvent vapors. For propanol, however,
it is likely that the decrease in intra-layer spacing greatly reduces the GO membrane per-
meability to this relatively large vapor molecule, whereas no significant effect is observed
in the GO membrane permeability to the smaller methanol and ethanol vapor molecules
upon Al3+ modification.
4. Conclusions
In summary, Al
3+
-modified GO membranes were formed by directly dissolving Al foil
into an acidic GO sheet solution. The coordination interactions formed between Al
3+
and
the OFGs on individual GO sheets in solution impacted the assembly process and resulted
in an inter-layer distance increases and an intra-layer spacing decreases. The addition
of Al
3+
affected the properties of both the GO solution and as-formed membrane. When
added to the GO sheet solution, the Al
3+
caused the solution to appear “muddy” and
resulted in a decrease in the solution surface tension. After drying during the layer-by-layer
self-assembly process, AGO membranes showed a higher surface roughness, were more
hydrophobic, and exhibited shrinkage with more cracking than unmodified GO membranes.
Studies of GO membranes under soaking, stirring, and sonication in different polar solvents
revealed an increase in membrane stability with increasing solvent hydrophobicity such
that membranes were most stable in propanol and least stable in pure water. Modification
of the GO membrane with Al
3+
resulted in AGO membranes with improved stability
in aqueous environments. This increased stability is likely the result of both a decrease
in membrane hydrophilicity and an increase in inter and intra-layer interactions due to
Membranes 2022,12, 871 14 of 16
the coordination between Al
3+
and the negatively charged OFGs on the GO sheet edges
and basal planes. The permeability of the GO and AGO membranes to water, methanol,
ethanol, and propanol vapor was explored using a simple apparatus. The results showed
that both GO and AGO membranes were more permeable to smaller solvent molecules.
Moreover, the permeability of the GO membrane to water and propanol decreased with Al
3+
modification, with no effects observed in permeability to methanol or ethanol. The decrease
in permeability was related to an increase in the activation energy of diffusion. The decrease
in membrane permeability to different polar vapors with Al
3+
modification was likely due
to both a decrease in nanochannel size and an increase in membrane hydrophobicity that
resulted from crosslinking between Al
3+
and the OFGs on the edges of the GO sheets within
the membrane. Although an increase in permeability might be expected with an increase
in inter-sheet distance, the decrease in intra-sheet spacing appeared to dominate the GO
membrane nanochannel structure and permeability. A complete understanding of how
Al
3+
modification impacts the GO membrane self-assembling process, its nanochannel
structure, properties, and performance is still limited.Our approach provides a simple and
cost-effective strategy to tailor the stability and permeation selectivity of GO membranes
for potential water treatment and fuel dehydration applications by using Al
3+
to specifically
tune the inner structure of the GO membrane.
Supplementary Materials:
The following supporting information can be downloaded at: https:
//www.mdpi.com/article/10.3390/membranes12090871/s1, Figure S1: Schematic (
a
) and camera
picture (
b
) of the permeability experiment set up; Figure S2: The AGO solution has a “muddy”
look when compared with the GO solution; Figure S3: Viscosity and shear stress at different shear
rates of GO solutions with Al concentrations from 0~0.5 wt.%; Figure S4: Digital camera image of
as-fabricated GO and AGO membrane in the evaporation dish of the same size; Figure S5: Soaking
stability test of GO/AGO membranes in solution with different solvents; Figure S6: Sonication
stability test of GO/AGO membranes in solution with different solvents; Figure S7: Stability test of
GO and AGO membranes in different solvents with a stirring rate of 200 Rpm with a large stir bar
(12 mm Length
×
5 mm Diameter) touches the sample during stirring; Figure S8: Stability test of
GO and AGO membranes in different solvents with a stirring rate of 200 Rpm with a small stir bar
(
10 mm Length ×3 mm
Diameter) bar not touch the sample during stirring; Figure S9: A highly wrin-
kled and contracted AGO membrane (
left
) swelled after soaking in water for 48 h(right); Figure S10:
GO (
left column
) and AGO (
middle
and
right columns
) membrane soaking. The swelling cannot
be visibly observed for the GO membrane and AGO membrane with
Al wt.% < 1wt.%
; Figure S11:
(
a
) XRD spectra of GO membrane with different thicknesses, and (
b
) distribution of interplanar
distance derived from XRD spectra; Figure S12: Permeability of GO and AGO membranes to water,
methanol, ethanol, and propanol water at different temperatures.
Author Contributions:
Conceptualization, Y.Y.S. and E.J.R.; methodology, Y.Y.S. and E.J.R.; validation,
Y.Y.S.; formal analysis, Y.Y.S. and E.J.R.; investigation, Y.Y.S., K.O. and L.K.; resources, Y.Y.S., I.V.V.,
R.C. and E.J.R.; writing—original draft preparation, Y.Y.S. and E.J.R.; writing—review and editing,
R.C., R.B.B. and I.V.V.; visualization, Y.Y.S. and E.J.R.; supervision, Y.Y.S. All authors have read and
agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments:
A portion of this research was conducted at the Center for Nanophase Materials
Sciences, Oak Ridge National Laboratory, which is a DOE Office of Science User Facility.
Conflicts of Interest: The authors declare no conflict of interest.
Membranes 2022,12, 871 15 of 16
References
1.
He, H.; Riedl, T.; Lerf, A.; Klinowski, J. Solid-State NMR Studies of the Structure of Graphite Oxide. J. Phys. Chem.
1996
,100,
19954–19958. [CrossRef]
2.
Yeh, C.-N.; Raidongia, K.; Shao, J.; Yang, Q.-H.; Huang, J. On the origin of the stability of graphene oxide membranes in water.
Nat. Chem. 2015,7, 166–170. [CrossRef]
3. Mi, B. Graphene Oxide Membranes for Ionic and Molecular Sieving. Science 2014,343, 740–742. [CrossRef]
4.
Chowdhury, I.; Mansukhani, N.D.; Guiney, L.M.; Hersam, M.C.; Bouchard, D. Aggregation and Stability of Reduced Graphene
Oxide: Complex Roles of Divalent Cations, pH, and Natural Organic Matter. Environ. Sci. Technol.
2015
,49, 10886–10893.
[CrossRef]
5.
Shen, X.; Lin, X.; Yousefi, N.; Jia, J.; Kim, J.-K. Wrinkling in graphene sheets and graphene oxide papers. Carbon
2014
,66, 84–92.
[CrossRef]
6.
Ye, J.; Zhang, B.; Gu, Y.; Yu, M.; Wang, D.; Wu, J.; Li, J. Tailored Graphene Oxide Membranes for the Separation of Ions and
Molecules. ACS Appl. Nano Mater. 2019,2, 6611–6621. [CrossRef]
7.
Kim, C.-M.; Hong, S.; Li, R.; Kim, I.S.; Wang, P. Janus Graphene Oxide-Doped, Lamellar Composite Membranes with Strong
Aqueous Stability. ACS Sustain. Chem. Eng. 2019,7, 7252–7259. [CrossRef]
8.
Jia, Z.; Shi, W. Tailoring permeation channels of graphene oxide membranes for precise ion separation. Carbon
2016
,101, 290–295.
[CrossRef]
9.
Wei, N.; Peng, X.; Xu, Z. Understanding Water Permeation in Graphene Oxide Membranes. ACS Appl. Mater. Interfaces
2014
,6,
5877–5883. [CrossRef]
10.
Lv, X.-B.; Xie, R.; Ji, J.-Y.; Liu, Z.; Wen, X.-Y.; Liu, L.-Y.; Hu, J.-Q.; Ju, X.-J.; Wang, W.; Chu, L.-Y. A Novel Strategy to Fabricate
Cation-Cross-linked Graphene Oxide Membrane with High Aqueous Stability and High Separation Performance. ACS Appl.
Mater. Interfaces 2020,12, 56269–56280. [CrossRef]
11.
Lyu, J.; Wen, X.; Kumar, U.; You, Y.; Chen, V.; Joshi, R.K. Separation and purification using GO and r-GO membranes. RSC Adv.
2018,8, 23130–23151. [CrossRef]
12.
Jiao, S.; Zhou, K.; Wu, M.; Li, C.; Cao, X.; Zhang, L.; Xu, Z. Confined Structures and Selective Mass Transport of Organic Liquids
in Graphene Nanochannels. ACS Appl. Mater. Interfaces 2018,10, 37014–37022. [CrossRef]
13.
Zhang, Y.; Wang, S.; Tang, P.; Zhao, Z.; Xu, Z.; Yu, Z.-Z.; Zhang, H.-B. Realizing Spontaneously Regular Stacking of Pristine
Graphene Oxide by a Chemical-Structure-Engineering Strategy for Mechanically Strong Macroscopic Films. ACS Nano
2022
,16,
8869–8880. [CrossRef]
14.
Liu, T.; Yang, B.; Graham, N.; Yu, W.; Sun, K. Trivalent metal cation cross-linked graphene oxide membranes for NOM removal in
water treatment. J. Membr. Sci. 2017,542, 31–40. [CrossRef]
15.
Yu, W.; Yu, T.; Graham, N. Development of a stable cation modified graphene oxide membrane for water treatment. 2D Mater.
2017,4, 045006. [CrossRef]
16.
Li, A.; Han, K.; Zhou, Y.; Ye, H.; Liu, G.; Kung, H.H. Incorporating multivalent metal cations into graphene oxide: Towards
highly-aqueous-stable free-standing membrane via vacuum filtration with polymeric filters. Mater. Today Commun.
2017
,11,
139–146. [CrossRef]
17.
Yan, F.; Yu, C.; Zhang, B.; Zou, T.; Zhao, H.; Li, J. Preparation of freestanding graphene-based laminar membrane for clean-water
intake via forward osmosis process. RSC Adv. 2017,7, 1326–1335. [CrossRef]
18.
Liu, Y.; Phillips, B.; Li, W.; Zhang, Z.; Fang, L.; Qiu, J.; Wang, S. Fullerene-Tailored Graphene Oxide Interlayer Spacing for
Energy-Efficient Water Desalination. ACS Appl. Nano Mater. 2018,1, 6168–6175. [CrossRef]
19.
Ching, K.; Lian, B.; Leslie, G.; Chen, X.; Zhao, C. Metal-Cation-Modified Graphene Oxide Membranes for Water Permeation.
Carbon 2020,170, 646–657. [CrossRef]
20.
Willcox, J.A.L.; Kim, H.J. Molecular Dynamics Study of Water Flow across Multiple Layers of Pristine, Oxidized, and Mixed
Regions of Graphene Oxide. ACS Nano 2017,11, 2187–2193. [CrossRef]
21.
Ghaffar, A.; Zhang, L.; Zhu, X.; Chen, B. Scalable graphene oxide membranes with tunable water channels and stability for ion
rejection. Environ. Sci. Nano 2019,6, 904–915. [CrossRef]
22. Xi, Y.-H.; Hu, J.-Q.; Liu, Z.; Xie, R.; Ju, X.-J.; Wang, W.; Chu, L.-Y. Graphene Oxide Membranes with Strong Stability in Aqueous
Solutions and Controllable Lamellar Spacing. ACS Appl. Mater. Interfaces 2016,8, 15557–15566. [CrossRef] [PubMed]
23.
Xi, Y.-H.; Liu, Z.; Liao, Q.-C.; Xie, R.; Ju, X.-J.; Wang, W.; Faraj, Y.; Chu, L.-Y. Effect of Oxidized-Group-Supported Lamellar
Distance on Stability of Graphene-Based Membranes in Aqueous Solutions. Ind. Eng. Chem. Res.
2018
,57, 9439–9447. [CrossRef]
24.
Stehle, Y.Y.; Robertson, E.; Cortez, R.; Bucinell, R.B.; Vlassiouk, I.V. In-plane and through-plane dielectric properties of graphene
oxide membrane: Effect of Al3+ modification. Materialia 2022,23, 101442. [CrossRef]
25.
Chen, L.; Shi, G.; Shen, J.; Peng, B.; Zhang, B.; Wang, Y.; Bian, F.; Wang, J.; Li, D.; Qian, Z.; et al. Ion sieving in graphene oxide
membranes via cationic control of interlayer spacing. Nature 2017,550, 380–383. [CrossRef]
26.
Park, S.; Lee, K.-S.; Bozoklu, G.; Cai, W.; Nguyen, S.T.; Ruoff, R.S. Graphene Oxide Papers Modified by Divalent Ions—Enhancing
Mechanical Properties via Chemical Cross-Linking. ACS Nano 2008,2, 572–578. [CrossRef]
27.
Long, Y.; Wang, K.; Xiang, G.; Song, K.; Zhou, G.; Wang, X. Molecule Channels Directed by Cation-Decorated Graphene Oxide
Nanosheets and Their Application as Membrane Reactors. Adv. Mater. 2017,29, 1606093. [CrossRef]
Membranes 2022,12, 871 16 of 16
28.
Guan, K.; Liu, Q.; Zhou, G.; Liu, G.; Ji, Y.; Liu, G.; Jin, W. Cation-diffusion controlled formation of thin graphene oxide composite
membranes for efficient ethanol dehydration. Sci. China Mater. 2019,62, 925–935. [CrossRef]
29.
Nam, Y.T.; Choi, J.; Kang, K.M.; Kim, D.W.; Jung, H.-T. Enhanced Stability of Laminated Graphene Oxide Membranes for
Nanofiltration via Interstitial Amide Bonding. ACS Appl. Mater. Interfaces 2016,8, 27376–27382. [CrossRef]
30.
Chang, P.; Mei, H.; Zhou, S.; Dassios, K.G.; Cheng, L. 3D printed electrochemical energy storage devices. J. Mater. Chem. A
2019
,7,
4230–4258. [CrossRef]
31.
Jalili, R.; Aboutalebi, S.H.; Esrafilzadeh, D.; Konstantinov, K.; Razal, J.M.; Moulton, S.E.; Wallace, G.G. Formation and processabil-
ity of liquid crystalline dispersions of graphene oxide. Mater. Horiz. 2013,1, 87–91. [CrossRef]
32.
Feng, J.; Su, B.-L.; Xia, H.; Zhao, S.; Gao, C.; Wang, L.; Ogbeide, O.; Feng, J.; Hasan, T. Printed aerogels: Chemistry, processing,
and applications. Chem. Soc. Rev. 2021,50, 3842–3888. [CrossRef] [PubMed]
33. Xu, Z.; Gao, C. Aqueous Liquid Crystals of Graphene Oxide. ACS Nano 2011,5, 2908–2915. [CrossRef]
34.
Zhang, M.; Mao, Y.; Liu, G.; Liu, G.; Fan, Y.; Jin, W. Molecular Bridges Stabilize Graphene Oxide Membranes in Water. Angew.
Chem. Int. Ed. 2020,59, 1689–1695. [CrossRef]
... Metal ions (e.g., Na + , Ca 2+ , Mg 2+ , Fe 3+ , or Al 3+ ) can improve membrane stability aluminum foil to the acidic GO sheet solution before the self-assembly process. These AGO membranes showed improved stability under stirring and sonication in polar solvents as well as reduced permeability to polar solvent vapors [22]. The AGO membrane properties were in part attributed to a decrease in the intra-layer spacing due to cross-linking between Al 3+ and carboxylate groups on the GO sheet edges. ...
... We previously reported that surface tension measurements of GO sheets at the airwater interface are indicative of the extent of sheet hydrophobicity and how the addition of Al 3+ affects the assembly of the GO sheets as the membrane dries [22,23]. Previous studies have also shown that the interfacial activity of GO sheets is highly dependent on the solution pH, and that the surface tension of a GO sheet solution decreases with decreasing pH [28][29][30]. ...
... Thus, there are more available carboxylate groups to interact with Al 3+ cations on GO sheets in the Al foil sample than on sheets in the AlCl 3 sample. Our previous work showed that interactions between Al 3+ and the carboxylate groups on the GO sheet edges act to decrease the intralayer spacing between sheets in the as-formed AGO membranes [22,23]. Here, we believe that a greater extent of interactions between Al 3+ and the GO sheets in the Al foil samples allowed these sheets to pack together to a greater extent at the air-water interface than GO sheets in the AlCl 3 sample. ...
Article
Full-text available
Graphene oxide (GO) membranes are promising materials for water filtration applications due to abundant nanochannels in the membrane structure. Because GO membranes are unstable in water, metal cations such as Al3+ are often introduced to the membrane structure to promote cross-linking between individual GO sheets. Here, we describe a simple yet versatile method to incorporate Al3+ into GO membranes formed via a slow self-assembly process. Specifically, we directly added aluminum to acidic GO sheet solutions from a variety of sources: Al2O3, AlCl3 and Al foil. Each species reacts differently with water, which can affect the GO solution pH and thus the density of carboxylate groups on the sheet edges available for cross-linking to the Al3+ cations. We demonstrate through characterization of the GO sheet solutions as well as the as-formed membranes’ morphologies, hydrophobicities, and structures that the extent to which the Al3+ cross-links to the GO sheet edges vs. the GO sheet basal planes is dependent on the Al source. Our results indicate that greatest enhancements in the membrane stability occur when electrostatic and coordination interactions between Al3+ and the carboxylate groups on the GO sheet edges are more extensive than Al3+–π interactions between basal planes.
... Free-standing GO membranes [12] with varying thicknesses between 10 and 16 µm were adhered on flexible substrates with GO solution and left to dry overnight. The GO membrane was reduced using a conventional CO 2 laser engraver (Epilog Fusion Pro, wavelength: 1065 nm) under ambient conditions (21 • C, 30% RH). ...
Article
Full-text available
Reduced graphene oxide (rGO) has attracted attention as an active electrode material for flexible electrochemical devices due to its high electric conductivity and large surface area. Compared to other reduction processes, laser reduction is a precise, low-cost, and chemical-free process that is directly applied to graphene oxide (GO) membranes. This study aims to develop rGO through laser irradiation for application as electrodes in thin flexible electrochemical sensors. Laser irradiation parameters will be optimized to achieve reduction of a low oxygen to carbon (O/C) ratio and surface impedance. The influence of humidity on the impedance of rGO electrodes will be studied. The observed instability of the rGO electrode is related to incomplete reduction and oxygenated defects involved in reduction. Partially removed oxygenated functional groups not only influence the impedance of the electrode but make it sensitive to the humidity of the working environment. The result provides references for GO’s laser reduction optimization, demonstrates the potential of applying rGO as an electrode in sensing applications, but also reveals the limitation of applying the laser reduced rGO electrode in a non-constant humidity environment.
... Recently, GO have attracted enormous attention as a filter material in membrane technology for desalination, organic solvent nanofiltration, pervaporation, and gas separation applications [6][7][8][9][10][11][12][13]. GO contains various oxygen functional groups on its edges and basal plane. ...
Article
Full-text available
Two-dimensional graphene oxide (GO)-based lamellar membranes have been widely developed for desalination, water purification, gas separation, and pervaporation. However, membranes with a well-organized multilayer structure and controlled pore size remain a challenge. Herein, an easy and efficient method is used to fabricate MoO2@GO and WO3@GO nanocomposite membranes with controlled structure and interlayer spacing. Such membranes show good separation for salt and heavy metal ions due to the intensive stacking interaction and electrostatic attraction. The as-prepared composite membranes showed high rejection rates (˃70%) toward small metal ions such as sodium (Na+) and magnesium (Mg2+) ions. In addition, both membranes also showed high rejection rates ˃99% for nickel (Ni2+) and lead (Pb2+) ions with good water permeability of 275 ± 10 L m−2 h−1 bar−1. We believe that our fabricated membranes will have a bright future in next generation desalination and water purification membranes.
Article
As an emerging and promising two-dimensional (2D) nanoporous membrane material, the K+-cross-linked graphene oxide membrane composed of 2D K+-controlled nanochannels (abbreviated herein as K+-CLGOM) has recently attracted great attention due to its attractive ion-sieving performance. However, the ion-sieving of K+-CLGOM for mono-/mono-valent cations has been still far from satisfactory. In this work, a facile and efficient ion sieving system for mono-/mono-valent cations was built by using the K+-CLGOM, which was firstly applied to supported a simple model of semi-biomembrane, namely oil/water (O/W) interface. As-obtained voltammetric results demonstrate that K+, NH4+ and Rb+ can prefer over Na+ and Li+ to transfer across K+-CLGOM in the trans-membrane system with O/W interface. This work not only provides the membrane-supported O/W interface as a novel platform to explore and understand the ion discrimination of K+-CLGOM, but also help to further exploit the potential applications of K+-CLGOM in the electrochemical ion sieving and extraction. Schematic illustration for the preparation process of a K+-CLGOM and the mechanism of cation-π interaction between K+ and graphene oxide aromatic system.
Article
Full-text available
The osmotic energy conversion properties of biomimetic light‐stimulated nanochannels have aroused great interest. However, the power output performance is limited by the low light‐induced current and energy conversion efficiency. Here, nanochannel arrays with simultaneous modification of ZnO and di‐tetrabutylammonium cis‐bis(isothiocyanato)bis(2,20‐bipyridyl‐4,40‐dicarboxylato) ruthenium (II) (N719) onto anodic aluminum oxide (AAO) to combine the nano‐confined effect and heterojunction is designed, which demonstrate rectified ion transport behavior due to the asymmetric composition, structure and charge. High cation selectivity and ion flux contribute to the high power density of ≈7.33 W m⁻² by mixing artificial seawater and river water. Under light irradiation, heterojunction promoted the production and separation of exciton, enhanced cation selectivity, and improved the utilization efficiency of osmotic energy, providing a remarkable power density of ≈18.49 W m⁻² with an increase of 252% and total energy conversion efficiency of 30.43%. The work opens new insights into the biomimetic nanochannels for high‐performance energy conversion.
Article
Full-text available
As an extraordinarily lightweight and porous functional nanomaterial family, aerogels have attracted considerable interest in academia and industry in recent decades. Despite the application scopes, the modest mechanical durability of aerogels makes their processing and operation challenging, in particular, for situations demanding intricate physical structures. “Bottom-up” additive manufacturing technology has the potential to address this drawback. Indeed, since the first report of 3D printed aerogels in 2015, a new interdisciplinary research area combining aerogel and printing technology has emerged to push the boundaries of structure and performance, further broadening their application scope. This review summarizes the state-of-the-art of printed aerogels and presents a comprehensive view of their developments in the past 5 years, and highlights the key near- and mid-term challenges.
Article
Full-text available
Recent innovations highlight the great potential of two‐dimensional graphene oxide (GO) films in water‐related applications. However, undesirable water‐induced effects, such as the redispersion and peeling of stacked GO laminates, greatly limit their performance and impact their practical application. It remains a great challenge to stabilize GO membranes in water. A molecular bridge strategy is reported in which an interlaminar short‐chain molecular bridge generates a robust GO laminate that resists the tendency to swell. Furthermore, an interfacial long‐chain molecular bridge adheres the GO laminate to a porous substrate to increase the mechanical strength of the membrane. By rationally creating and tuning the molecular bridges, the stabilized GO membranes can exhibit outstanding durability in harsh operating conditions, such as cross‐flow, high‐pressure, and long‐term filtration. This general and scalable stabilizing approach for GO membranes provides new opportunities for reliable two‐dimensional laminar films used in aqueous environments.
Article
Mechanical-electrical properties of macroscopic graphene films derived from graphene oxide (GO) sheets are substantially restricted by their surface wrinkles and structural misalignment. Herein, we propose a chemical-structure-engineering strategy to realize the spontaneously regular stacking of modified GO (GO-m) with trace carboxyl. The highly aligned GO-m film delivers a fracture strength and modulus of nearly 3- and 5-fold higher than a wrinkled film with conventional Hummer's method derived GO (GO-c). The favorable assembly pattern of GO-m sheets is attributed to their decreased interfacial friction on the atomic scale, which weakens their local gelation capability for freer configuration adjustment during the assembly process. The chemical structure of GO-m can be further engineered by an epoxide-to-hydroxyl reaction, achieving a record high tensile strength of up to 631 MPa for the pristine GO film. By exploring the relationship between the surface terminations of GO and its stacking mode, this work proves the feasibility to realize high-performance macroscopic materials with optimized microstructure through the chemical modulation of nanosheet assembly.
Article
A graphene oxide (GO) membrane can be used in structural dielectric capacitors and planar micro-supercapacitors as a dielectric separator material due to its excellent through-plane and in-plane dielectric properties. The ability to tailor the dielectric properties of the GO membrane is crucial for dielectric applications. We believe this to be the first demonstration describing how Al³⁺ ions impact the dielectric properties of the GO membrane through cation to oxygenated functional groups interactions. In this study, the Al³⁺-modification to the GO membrane was realized through a facile pre-coordination method followed by layer-by-layer assembling. The influence of Al³⁺ on the GO membranes’ through-plane and in-plane dielectric properties (dielectric constant and dielectric loss) was investigated in the frequency range from 0.01 Hz to 10⁵ Hz. Plane-to-plane and edge-to-edge Al³⁺ coordination with the functional group on the basal plane or edge of the GO sheets was proposed to understand the metal ion's impact on the dielectric properties. The results demonstrated that the Al³⁺ modification was an effective strategy to tailoring the GO membranes’ dielectric properties in a low frequency regime and further suggested the possibility of designing dielectric GO membranes with controllable dielectric performance. At a low frequency range (0.01-1 Hz), the addition of Al³⁺ increased the through-plane dielectric constant and decreased the in-plane dielectric constant. In terms of dielectric loss, unmodified GO membranes exhibited a higher in-plane dielectric loss than the through-plane dielectric loss. Al³⁺-modified GO membranes exhibited increased through-plane dielectric loss and decreased in-plane dielectric loss compared to the unmodified GO membranes.
Article
Graphene oxide (GO) membranes have shown enormous promise in desalination and molecular/ionic sieving. However, the instability of GO membranes in aqueous solutions seriously hinders their practical applications. Herein, we report a novel and simple strategy to fabricate stable GO membranes in water-based environments through the insertion of various metal cations from metal foils (e.g., copper (Cu), iron (Fe), nickel (Ni), and zinc (Zn) foils) and natural deposition. Based on the cation-π, coordination, and electrostatic interaction between metal cations and GO nanosheets, the aqueous stability and mechanical strength of the membranes are significantly improved. The permeation rates for acetone, toluene, and p-xylene molecules across the GO membrane cross-linked by copper ions with a deposition time of 24 h are 0.966, 0.074, and 0.100 mol m-2 h-1, respectively. Moreover, this membrane displays excellent separation performance, and the separation factor of K+/Mg2+ is up to 68.8 in mono-/multivalent metal cation sieving, which indicate the effective molecular/ionic sieving performance. Meanwhile, the ionic sieving of the GO membrane cross-linked by copper ions has excellent repeatability and long-term stability. The versatility of this natural deposition strategy to fabricate GO membranes cross-linked by metal cations is investigated by using Fe foil, Zn foil, and Ni foil as well as other porous substrates such as polyvinylidene fluoride (PVDF), polyethersulfone (PES), and nylon membranes and filter paper. This fabrication strategy also enables low-cost preparation of large-area GO membranes. Therefore, GO membranes cross-linked by metal cations and prepared by this simple metal cation incorporation strategy have large potential application for molecular/ionic sieving in various solution systems.
Article
The stability of graphene oxide (G-O) membranes in water can be enhanced via the simple addition of metal-cations to the planar laminate structure, however, the relationship between valency and hydration radius and the properties and performance of the membrane structure is unclear. This study examines the impacts of adsorbed monovalent (K⁺ and Na⁺), divalent (Cu²⁺ and Mn²⁺) and trivalent (Al³⁺, Fe³⁺ and La³⁺) cations on permeation for G-O membranes in water vapor permeation and pervaporation. Metal-cations with high electronegativity and a small hydration radius possess greater binding strength that constrains the G-O laminate channels (d-spacing). Both water permeation methods show a positive trend between permeation and interlayer spacing for metal-cations with the same valency. Amongst the different valence states, steric exclusion from cations is dominant for minimal water layers (water vapor permeation), while the ability for cations to attract and hold water molecules inside its outer shell has a greater effect on swollen channels (pervaporation). These findings help pave the road towards G-O based applications in next generation water filtration methods for desalination and wastewater treatment.
Article
Graphene oxide (GO) membranes, with 2D nanochannels and peculiar permeation performance, are promising nanobarriers for ion and molecule separations. However, practical applications have been challenged by dramatic deterioration in mechanical durability, structural stability, and sieving ability under aqueous solution. Thus, enhancement of stability while maintaining species sieving under aqueous conditions are key topics of GO-based membrane research. Herein, the functionalization of GO membranes was investigated by combining imidazolium complexation with electron beam irradiation to enhance membrane stability in aqueous solution, which included improvements in mechanical strength and suppression of interlayer spacing (d-spacing) of GO laminates in aqueous solution. Here, 1-allyl-3-vinylimidazolium chloride ([AVIM]Cl) was first used to complex with GO sheets and were subsequently covalently bonded onto adjacent GO sheets by electron beam irradiation. It was found that prepared GO-hybrid membranes could withstand 15.9 kPa of hydrostatic pressure and that d-spacing enlargement after water immersion was effectively suppressed, compared to original GO membranes. In addition, the properties of GO hybrid membranes were easily controlled via the mass ratio of [AVIM]Cl/GO in the mixed GO dispersions and the absorbed dose and dose rate of electron beam irradiation. Permeation results revealed that electrostatic repulsion between imidazolium cation and metal ions had a crucial influence on metal ion permeation through GO-hybrid membranes. The well-confined d-spacing in the wet state made the resultant membranes completely sieve out species with hydrated diameters >0.99 nm. This study confirmed that imidazolium-complexation coupling with electron beam irradiation was effective for tailoring GO membranes with enhanced mechanical durability and controlled d-spacing, which benefitted its applications for ion and molecule separations in aqueous solution.
Article
Nanostructured graphene oxide (GO) membranes offer outstanding mass transport performances such as ultrahigh water flux and precise molecular sieving and thus have a great potential as a novel filtration platform for energy efficient molecular or ionic separation. However, the structural stability of the GO membranes in aqueous environments remains a challenging problem, which limits its practical application perspective. We, here, report novel GO composite membrane composed of GO and asymmetrically functionalized Janus GO (JGO) sheets. Single-side functionalized JGO was made by amidation of dodecylamine. Incorporation of JGO in GO film exhibited extraordinary stability in water at broad pH values even under agitation. Moreover, JGO-doped GO composite membrane showed robust integrity in acid or base solutions over months. It exhibits high molecular retention above 95 % for uncharged and charged dye molecules, rhodamine B and brilliant blue G while maintaining water permeability comparable with previously reported GO-based membranes under osmotic pressure. This work significantly expand possibilities for new way to enhance structural stability in graphene-based or 2D-material membrane and surface-selective functionalization in membrane fabrication for many applications.
Article
Structural manipulation of graphene oxide (GO) building blocks has been widely researched. Concerning GO membranes for separation applications, the validity and maintenance of their microscopic structures in the chemical environment are pivotal for effective separation at the molecular scale. Cationic interactions with both aromatic rings and oxygenated functional groups of GO make metal ions intriguing for physically and chemically structural reinforcement. By filtrating GO suspension through the substrate loaded with cations, stacking of GO nanosheets and diffusion of cations steadily evolve simultaneously in an aqueous environment without flocculation. Thus, thin and homogeneous GO membrane is obtained. Divalent and monovalent cations were studied regarding their interactions with GO, and the performance of correspondingly functionalized membranes was evaluated. The divalent cation-stabilized membranes have favorable stability in the separation of water/ethanol. This facile fabrication and functionalization method may also be applicable for structure construction of other two-dimensional materials.
Article
3D printing technology, which can be used to design functional structures by combining computer-aided design and advanced manufacturing procedures, is regarded as a revolutionary and greatly attractive process for the fabrication of electrochemical energy storage devices. In comparison to traditional manufacturing methods, 3D printing possesses unique advantages in geometrical shape design as well as rapid prototyping, especially with high surface area complex 3D structure constructions. Recently, a number of 3D-printed electrochemical energy storage devices have been reported, showing an increased interest of the scientific community. To further advance material design and technology development, comprehensive understanding of the strengths and weaknesses of each 3D printing techniques and knowledge of recent progress in 3D-printed electrochemical energy storage, is indispensable. To that end, a literature review of recent advances of 3D printing technology for capacitive energy storage is provided herein. Emphasis is given on the design of printing materials, printing process and electrochemical performance of printed devices. Some 3D printed solid electrolytes for energy storage applications are also summarized. Discussion and outlook on the potential future designs and development of 3D printing for electrochemical energy storage devices are provided in the text.