You are on page 1of 8

Carbon 191 (2022) 563e570

Contents lists available at ScienceDirect

Carbon
journal homepage: www.elsevier.com/locate/carbon

Shear-induced assembly of high-aspect-ratio graphene nanoribbon


nanosheets in a confined microchannel: Membrane fabrication for
ultrafast organic solvent nanofiltration
Ji Hoon Kim, Yunkyu Choi, Junhyeok Kang, Ju Yeon Kim, Jun Hyuk Bae, Ohchan Kwon,
Dae Woo Kim*
Department of Chemical and Biomolecular Engineering, YONSEI University, Yonsei-ro 50, Seodaemun-gu, Seoul, (03722), Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: One-dimensional graphene oxide nanoribbons (GONRs) were self-assembled into two-dimensional (2D)
Received 14 December 2021 nanosheets using the shear and confinement effect during a slot-die coating process. An aqueous GONR
Received in revised form suspension comprising nanostrings made of entangled GONRs at a concentration of 5 mg/mL was used.
5 February 2022
When the GONR suspension was injected into the microchannel of the slot-die coater, the GONR
Accepted 11 February 2022
Available online 16 February 2022
nanostrings self-assembled to form a nanosheet. The thickness of the GONR nanosheet could be
controlled at the nanometer scale by adjusting the injection rate of the GONR suspension into the slot-die
head, and the lateral dimension of the nanosheet was in the range of several tens of micrometers. The
Keywords:
Graphene nanoribbon
GONR nanosheets could be directly and continuously coated on a porous polymer support by the slot-die
Self-assembly coating method. In particular, a 40-nm-thick GONR layer exhibited ultrafast organic solvent nano-
Slot-die coating filtration (OSN) with an isopropyl alcohol permeance of 679 LMH/bar and molecular weight cut-off of
Nanofiltration 961 Da, substantially surpassing the upper limit of the OSN performances of polymeric and 2D-material-
Nanosheet based membranes. Highly efficient diafiltration of mixed organic molecules in an organic solvent is also
feasible using this GONR membrane.
© 2022 Elsevier Ltd. All rights reserved.

1. Introduction multiwalled carbon nanotubes (MWCNTs) in acidic conditions us-


ing a strong oxidizing agent (e.g., KMnO4) [14,15]. As the resulting
Graphene nanoribbons (GNRs) are one-dimensional (1D) ma- unzipped nanoribbons of graphene, referred to as graphene oxide
terials that exhibit the structural properties of both graphene and nanoribbons (GONRs), are decorated with oxygen functional
carbon nanotubes (CNTs) [1e4]. GNRs can be stacked with a narrow groups, they can be easily dispersed in water and various organic
interlayer spacing owing to their graphene-like basal plane [5e7], solvents [16,17].
and they can easily form entangled structures because of their high Most industrial fabrication technologies are based on solution
aspect ratio and narrow widths [8,9]. Because of their unique processing [18]. Therefore, it is critical to understand the dispersion
chemical, electrical, and physical properties such as a controllable behavior of GONRs in various solvents for their practical application
bandgap, abundant edge density, electrical conductivity, and and industrial production [17,19]. Recently, we observed the
chemical stability, GNRs have been used in various applications, concentration-dependent self-assembly of GONRs suspended in
such as in biomedicine, chemical sensors, energy devices, elec- aqueous solvents. According to a previous study, aqueous GONR
tronic devices, electrocatalysis, composites, and membranes suspensions exhibit shear-thinning viscoelastic properties, and the
[10e13]. In addition, GNRs can be prepared scalably by unzipping GONRs entangle to form nanostrings, nanoplates, and macroporous
scaffolds as the GONR concentration increases [20]. However,
despite this finding, there is still a lack of understanding of the
GONR structure in solvents with respect to various factors that
Abbreviations: GONR, Graphene oxide nanoribbon; GO, Graphene oxide; OSN,
Organic solvent nanofiltration; MR, Methyl red; MnB, Methylene blue; BBG, Bril- affect the dispersion state of graphene oxide (GO), which has a
liant blue G; EB, Evans blue; RosB, Rose Bengal; MWCO, Molecular weight cut-off. similar chemical structure [21]. For example, GO exhibits lyotropic
* Corresponding author. liquid crystal (LC) properties and forms nematic, smectic, and
E-mail address: audw1105@yonsei.ac.kr (D.W. Kim).

https://doi.org/10.1016/j.carbon.2022.02.026
0008-6223/© 2022 Elsevier Ltd. All rights reserved.
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

cholesteric phases according to the GO concentration and dopant assembly. The GONR suspension that penetrated through the
type [22,23]. Furthermore, the alignment of GO-based LCs can be 80 mm gap of the slot dies was collected and frozen using liquid
controlled using shear and magnetic/electric forces [24,25], which nitrogen. It was then freeze-dried for 12 h, and the obtained
is useful for improving the performance of GO-based devices such powder was imaged by SEM.
as battery, supercapacitors, membranes, and fibers that require
highly ordered graphene structures [26e29]. 2.4. Nanofiltration using a dead-end setup
Inspired by the assembly of GO, herein, we demonstrate that the
self-assembly of GONRs into a nanosheet is feasible using the shear The nanofiltration test was performed using a customized dead-
force generated during a slot-die coating process. A slot-die head end filtration device. The effective area was 4.52 cm2 and the
with a micro-gap was used to prepare the GONR nanosheet, and the applied pressure was controlled by controlling the feed amount of
as-prepared suspension containing the GONR nanosheet was nitrogen gas. The following dyes were tested: methyl red (MR:
directly loaded onto a porous polymer support using the slot-die electroneutral, 269 Da), methylene blue (MnB; positively charged,
coater. The fabricated GONR membranes were tested for organic 320 Da), brilliant blue G (BBG; negatively charged, 854 Da), Evans
solvent nanofiltration (OSN) and organic molecule diafiltration. In blue (EB; negatively charged, 960 Da), and Rose Bengal (RosB;
particular, we discuss the effects of the slot-die coating conditions negatively charged, 1010 Da). The concentration of the single dye
on the structure of the GONR nanosheet and its coating solution was 10 mg/L. To evaluate the rejection rate and permeance
characteristics. through the membrane, 30 mL of each dye solution was permeated
under a pressure of 1 bar. The permeance (J in L m2 h1 bar1,
2. Experimental LMH/bar) and rejection rate (R, %) were calculated using the
following equations:
2.1. Preparation of a GONR suspension for slot-die coating
Vp
J¼ (1)
GONR was prepared by unzipping the MWCNTs in acidic con- t  A  DP
ditions using KMnO4 as an oxidizer [14,15] and experimental details
can be found in our previous work [20]. Briefly, 2 g of MWCNT was ðCf  Cp Þ
mixed in 130 mL of H2SO4. And then, 10 g of KMnO4 was slowly R¼ *100 (2)
Cf
added in the mixture while stirring the solution in an ice bath. After
cooling the mixture, the solution was stirred at 35  C. Then 150 mL where Vp (L) is the volume of the permeate solution, t (h) is the
DI water and 40 mL H2O2 were added sequentially in an ice bath.
filtration time, A (m2) is the effective area of the membrane, and DP
The resulting solution was centrifuged to remove the acidic solu- is the applied pressure. Cf and Cp are the concentrations of the feed
tion and redispersed in DI water several times. The as-prepared
and permeate solutions, respectively, as calculated by UVevis
GONR slurry was diluted with DI water to a concentration of spectroscopic analysis. To evaluate the mechanical stability of the
5 mg/mL. To verify the GONR concentration, 2 mL of the GONR
GONR membrane with a thickness of 40 nm under high pressure, a
suspension was dried in a vacuum oven at 50  C for 10 h. Then, the 10 mg/L EB solution was used, and the pressure was increased from
weight of the GONR powder was determined using an electronic
6 to 12 bar.
balance. Importantly, ethanol was added to the GONR suspension
to achieve an ethanol concentration of 8 wt%, which is required for
improving the coating quality and decreasing the thickness of the 2.5. Cross-flow filtration using the GONR membrane
GONR layer. Typically, 20 mL of the 5 mg/mL GONR suspension was
used in a single coating process; it was directly injected into the A long-term filtration test was performed in the cross-flow
head of the slot-die coater. filtration mode. The effective membrane area was 7.07 cm2. The
feed solutions with different dyes (MR, EB, MnB, BBG, RosB, and
2.2. GONR membrane fabrication through slot-die coating MR/EB mixture) were prepared in 40 L of isopropyl alcohol (IPA).
The concentration of the single dye solution was 10 mg/L, and the
A lab-scale slot-die coater (DCN: Development Creation Nano- concentration of the mixed dye solution was 20 mg/L (10 mg/L for
technology, YSU-LSCDC-190801) was used in this study. The slot- each dye). The feed flow rate was 2 L/min, and filtration was con-
die coater consists of three parts: an ink injection (feeding) ducted for 56 h at 1 bar using a 40-nm-thick GONR membrane. The
pump, a slot-die, and a moving stage. The GONR suspension was retentate was recycled into the feed tank using a pump. The per-
filled into a feed syringe tank, which was directly connected to the meance values were recorded using a digital balance. The rejection
slot-die head of the coater. Then, a porous polymer support (PES: rate was evaluated by UVevis spectroscopic analysis. To measure
0.03 mm of pore size, the diameter of 9 cm or 4.8 cm; GVS North the pure IPA permeance of the GONR membrane depending on the
America Sandford, ME 04073-USA) was placed on the slot-die thickness of the GONR layer, filtration was conducted for 1 h at
coating stage. The gap between the slot-die head and substrate 1 bar.
was fixed at 0.5 mm. The gap between the slot dies was 80 mm. The
coating was generally performed at a feed injection speed of 1.1 mL 2.6. Characterization
min1 with a coating stage moving rate of 40 mm s1. The slot-die-
coated GONR membrane was further dried in a vacuum oven at The morphology of the GONRs was characterized by trans-
50  C overnight to obtain a well-stacked and stable GONR layer on mission electron microscopy (TEM; JEM-F200, JEOL, Japan). The
the PES support. contact angles of water, ethanol, and IPA were measured using an
optical tensiometer (DYNE TECHNOLOGY, USA). The functional
2.3. Shear-induced assembly of GONR in the slot-die head groups of the GONRs were characterized by Fourier-transform
infrared spectroscopy (FT-IR, Vertex 70, Bruker ALPHA, Germany).
A GONR suspension with a concentration of 5 mg/mL was The surface chemical states of the GONR were characterized by X-
injected into the slot-die head. The injection rate was increased ray photoelectron spectroscopy (XPS, K-alpha, Thermo U.K.). The
from 0.5 to 3.0 mL/min to understand its effect on the GONR self- surface and cross-sectional morphologies of the membranes and
564
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

substrates were imaged by field-emission scanning electron mi- this assembly phenomenon are presented in Figs. 2 and 3.
croscopy (FE-SEM, IT-500 HR JEOL, USA) at a voltage of 10 kV and a Fig. 1B shows a photograph obtained during the slot-die coating
current of 10 mA. The d-spacing of the GONR layer was determined process and Fig. S2 depicts the whole slot-die setup. In this process,
by conducting X-ray diffraction on a high-resolution X-ray diffrac- a 5 mg/mL aqueous GONR suspension was injected into the slot-die
tometer (HR-XRD, SmartLab, Rigaku, Japan). head at the rate of 1.1 mL/min. The moving rate of the coating stage
was typically set to 14 mm/s. The distance between the slot-die lip
3. Results and discussion and substrate was 0.5 mm, and the width of the micro-gap between
the slot-die heads was 80 mm. A uniform ultrathin layer of the
3.1. Overall GONR assembly and membrane fabrication process deposited GONRs is formed in this method as the solvent sponta-
using a slot-die coater neously dries on the target substrate.
Fig. 1C shows a photograph of the as-prepared GONR-coated
Fig. 1A schematizes the fabrication procedure of the GONR polyethersulfone (PES) filter (Fig. S3: pore size, 0.03 mm; diameter,
membrane using a slot-die coater and the assembly of GONRs 9 cm) with a uniform GONR layer. Fig. 1D and E shows the top-view
during the coating process. The GONR coating process can be and cross-sectional scanning electron microscopy (SEM) images of
divided into three parts: (1) the injection of a GONR suspension the GONR membrane. As shown, a uniform 120-nm-thick GONR
from a feed tank to the slot-die head using a syringe pump at a layer was obtained successfully. In general, thick carbon nano-
controlled injection rate, (2) penetration of the GONR suspension material layers are coated at the micrometer scale using conven-
through the micrometer gap of the slot-die head, and (3) deposition tional bar-coating or doctor blade methods, which rely on highly
of the extruded GONR suspension onto a target substrate on a concentrated and viscous carbon nanomaterial suspensions
moving stage. For depositing a uniform and ultrathin GONR layer [25,30]. However, it was possible to obtain a uniform and ultrathin
on a porous polymeric substrate, it is critical to assemble the GONR GONR layer using a GONR suspension of a relatively low concen-
into a nanosheet, which can be achieved by controlling several tration (5 mg/mL) because of the advantage of the slot-die coating
fabrication parameters such as the concentration of the GONR method, which allows direct coating of the nanomaterial suspen-
suspension, injection rate of the suspension into the slot-die head, sion on a substrate at a controlled deposition rate [19,31,32].
and moving rate of the stage. As shown in the right panel of Fig. 1A, To investigate the hydrophilic surface properties of the GONR
GONRs obtained by the unzipping of MWCNTs (Fig. S1) could be layer, the surface contact angles of water, ethanol, and IPA were
dispersed in water at a concentration of 5 mg/mL, and the indi- measured (Fig. 1F). The GONR layer was found to be both hydro-
vidual GONRs assembled to form nanostrings during the ball- philic and oleophilic, as reflected by the low contact angles of 40 ,
milling and stirring process used for preparing the GONR suspen- 5 , and 3 for water, ethanol, and IPA, respectively. These contact
sion (see details in the experimental section). The well-dispersed angles are similar to those on a GO film. While the degree of
GONR nanostrings assembled into a nanosheet structure as they oxidation is different between GO and GONR, both these materials
passed through the micro-gap in the slot-die head. When the GONR contain oxygen functional groups and sp2 carbon domains (Fig. S4),
nanosheets were extruded from the slot-die lip and deposited on a which lead to their similar surface properties. In the case of GONRs,
moving substrate, the thickness of the GONR nanosheet decreased the oxygen functional groups at the edge may contribute to hy-
owing to the shear force induced by the moving stage. The details of drophilicity, while the sp2 carbon domains in the basal plane are

Fig. 1. (A) Schematics of the GONR membrane fabrication using a slot-die coater and GONR assembly during the coating process. (B) Photographs illustrating the slot-die coating
process. (C) Photograph of a GONR membrane. (D) Top-view and (E) cross-sectional SEM images of a GONR membrane. (F) Contact angles of water, ethanol, and IPA on GONR
membranes (right) in comparison with those on GO membranes (left). In the membrane fabrication process, the stage moving rate was set at 14 mm/s and the feeding rate of the
GONR suspension into the slot-die head was set at 1.1 mL/min. (A colour version of this figure can be viewed online.)

565
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

Fig. 2. (A) SEM images of a GONR sample obtained by freeze-drying a 5 mg/mL GONR suspension. (B) SEM images of the GONR sample after it penetrated the micro-gap of the slot-
die head. The feeding rate of the GONR suspension was increased from 0.5 to 3 mL/min, and the gap of the slot-die head was fixed at 80 mm. The obtained suspensions were freeze-
dried to maintain the GONR morphologies in the suspension. (C) XRD patterns of different GONR samples. All the GONR samples were freeze-dried except for the sample labeled
“GONR film 1.1”. The numbers in sample labels indicate the feeding rate (mL/min) of the GONR suspension. (D) Schematic of the GONR assembly in the micro-gap of the slot-die
head. (A colour version of this figure can be viewed online.)

likely responsible for their high affinity for organic solvents [14,33]. a 5 mg/mL GONR suspension before slot-die injection. Thick
The high affinity of the GONR for organic solvents is beneficial for bundle-like GONR strings were formed, as reported previously [20].
achieving ultrafast organic solvent permeance through a GONR- Fig. 2B shows the SEM images of freeze-dried GONR samples after
coated membrane [34], as discussed later. injection into the slot die-head at different feeding rates in the
range of 0.5e3.0 mL/min. As the injection rate increased, the thick
3.2. Assembly of a GONR nanosheet in a confined micro-gap bundle-like GONR strings reorganized to form wider and thinner
nanosheets. In particular, the thickness of the nanosheets
In our previous work, we studied the assembly of GONRs decreased to 150 nm and the lateral size ranged up to 100 mm,
depending on their concentration in aqueous suspensions and resulting in an aspect ratio of ~670. In the magnified SEM images in
observed the development of relatively thick strings and scaffolds Fig. 2B, no defects or cracks are observed in the basal plane of the
with increasing GONR concentration [20]. Therefore, a thick GONR nanosheets, indicating the formation of a dense 2D structure
layer was expected for the coating layer obtained with a 5 mg/mL composed of entangled GONRs.
suspension in this study. However, an ultrathin GONR layer with a The influence of the feeding rate on the interlayer spacing of the
thickness of ~100 nm was obtained, as shown in Fig. 1E. To un- nanosheet was investigated by the X-ray diffraction (XRD) analysis
derstand the origin of this ultrathin layer, the variation in the GONR of the freeze-dried samples (Fig. 2C). All the GONR powders yielded
morphology was observed before and after the injection of the a diffraction peak at a 2q of ~10.2 (8.7 Å); the peak position was
GONR suspension into the slot-die head (Fig. 2A and B). Fig. 2A slightly shifted to a lower angle compared to that of the GONR film
shows the SEM images of a GONR sample obtained by freeze-drying (8.0 Å) prepared by slot-die coating. The results suggest that the

566
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

Fig. 3. (A) Photographs of GONR membranes obtained at different stage moving rates. (B) Corresponding top-view and (C) cross-sectional SEM images of the GONR membranes. A
thinner GONR layer was formed at a higher stage moving rate. The concentration of the GONR suspension used for slot-die coating was 5 mg/mL, and the feeding rate of the
suspension into the slot-die head was fixed at 1.1 mL/min. (A colour version of this figure can be viewed online.)

injection rate of the GONR suspension into the slot-die head does conditions for producing a nanosheet with a thickness of 230 nm,
not change the interlayer structure but significantly modifies the as shown in Fig. 2B. Because the extruded GONR suspension is
assembled GONR morphology and aspect ratio. The slightly wider directly applied to a moving support, another shear effect due to
interlayer spacing of the GONR powder compared to that of the stage movement may exist, and the shear force increases as the
GONR film can be attributed to the freeze-drying process, which stage moving rate increases.
hinders the regular stacking of GONRs during solvent drying. Fig. 3A shows the photographs of the as-prepared GONR layers
Considering the above results, we conclude that the morphology on PES supports, and Fig. 3B shows the corresponding SEM images.
control of the GONR strings to nanosheets is induced by the micro- When the stage moving rate was lower than 9 mm/s, sustained and
gap between the slot-die heads. Fig. 2D shows a schematic of the uncoated areas were observed, and a uniform layer was not formed
GONR assembly in the micro-gap of the slot-die head. Because the because the suspension was highly dilute and flowed on the sub-
size of the GONR strings, ~10 mm, is much smaller than the width of strate. When the stage moving rate was greater than 14 mm/s, a
the micro-gap, 80 mm, the physical extrusion effect on the GONR pale-colored GONR membrane was formed without any defective
assembly is not significant, while the slit may guide their shaping or sustained area. The photographs of samples obtained at stage
into a nanosheet as per the dimensions of the slit gap. This aspect is moving rates of 20e40 mm/s exhibit some stripes owing to the
confirmed by the SEM images depicted in Fig. S5, which shows the local accumulation of the suspension because of the high stage
effects of the slit gap size on the alignment of the GONR nanosheets moving rate; however, the top-view SEM images confirm the for-
prepared with a higher concentration solution (50 mg/ml). The mation of a uniform GONR layer without any cracks or defects
nanosheet thickness is drastically decreased as the slit size is (Fig. 3B).
reduced due to the higher guiding shear force. We note that, The thickness of the GONR layer was further investigated using
although the concentration is increased tenfold, the general the cross-sectional SEM images of the samples, shown in Fig. 3A. As
alignment is well achieved, indicating the feasibility of the slot-die shown in Fig. 3C, surprisingly, the thickness was smaller than that
coater towards high viscosity GONR solutions. Finally, a high in- of the parent GONR nanosheet in suspension with a thickness of
jection rate decreases the thickness of the GONR nanosheet; 230 nm. In addition, the thickness of the GONR layer gradually
therefore, the shear force generated in the micro-gap is expected to decreased from 180 nm to 40 nm with an increase in the stage
be the main driving force for transforming the GONRs into a moving rate, where a higher speed increases the shear force during
nanosheet. In our previous study, the GONR solution was coated by the coating process. The results suggest that further thinning of the
the bar-coating technique, which has a wider gap, and the applied GONR nanosheet occurs owing to the second shear effect due to the
shear force is relatively reduced [20]. As such, nanosheet formation moving stage when the suspension is deposited on the substrate.
was not observed using the bar-coating method. Conversely, the Consequently, even if the same suspension is used as a feed for slot-
slot-die head has a narrower, confined gap that induces a stronger die coating, the combination of the feed injection rate and stage
shear force, which allows the formation of 2D structures. moving rate can dramatically modulate the structure of the GONR
nanosheet and deposited layer. Because interlayer spacing is
3.3. Effect of the stage moving rate on the quality and thickness of responsible for the molecular separation performance of stacked
the GONR membrane on a PES support (pore size: 0.03 mm) graphene membranes [19,35,36], the stability of the interlayer
spacing of the GONR and coating layer was investigated by
After investigating the self-assembly behavior of the GONR in immersing the GONR membrane in water, ethanol, and IPA for 1
the slot-die head, the influence of the stage moving rate was also day (Fig. S6). XRD analyses revealed a stable interlayer structure of
studied by fixing the GONR concentration at 5 mg/mL and the the GONR membranes in the tested solvents, as reflected by the fact
suspension injection rate at 1.1 mL/min (Fig. 3), which are that the d-spacing of the GONR membrane was the same (~8 Å)

567
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

regardless of the solvent type. In addition, the GONR membrane a general trend of increasing permeances with decreasing solvent
remained stable during 7 days of immersion in the tested solvents, viscosity. However, toluene and benzene showed higher per-
while the GO layer was easily delaminated after immersion in meances than acetone and hexane solvents despite having a
ethanol for 1 h (Fig. S6B). slightly higher viscosity, possibly attributed to their planar molec-
ular structures.
3.4. OSN and diafiltration performance of GONR membranes Fig. 4B shows the rejection rates of 40-nm-thick GONR mem-
branes for five dyes dissolved in IPA: methyl red (MR, 269 Da
After confirming the stability of the GONR membrane and its (neutrally charged)), methylene blue (MnB, 320 Da (positively
rigid interlayer structure in the tested organic solvents, the per- charged)), brilliant blue G (BBG, 854 Da (negatively charged)),
formance of the membrane in the nanofiltration of organic solvents Evans blue (EB, 961 Da (negatively charged)), and Rose Bengal
was evaluated (see Fig. 4). First, the permeance of pure IPA was (RosB, 1018 Da (positively charged)). The membrane was tested
studied by changing the thickness of the GONR layer (Figs. 4A and with a 10 mg/L dye solution in IPA at 1 bar. Both dead and cross-
S7). The permeance test was conducted at 1 bar in dead-end and flow filtration experiments were performed. In dead-end filtra-
cross-flow filtration modes to investigate the effect of the filtration tion, the dye rejection rates were 41.3% (MR), 33.3% (MnB), 51.7%
mode on the solvent permeance. In the cross-flow test, the per- (BBG), 93.1% (EB), and 72.8% (RosB). Further, 150-nm-thick mem-
meance was measured under stabilized equilibrium conditions branes also showed a similar dye rejection rate trend (Fig. S10),
(Fig. S7). while their solvent permeance decreased because of the thicker
The IPA permeance increased dramatically with a decrease in selective layer [25,36]. Moreover, a similar rejection rate trend was
the thickness of the GONR layer from 150 nm to 40 nm in both observed in the cross-flow tests, but the dye rejection rates were
modes. Surprisingly, the IPA permeance was enhanced in the cross- slightly different: 10% (MR), 6% (MnB), 25.8% (BBG), 99.9% (EB), and
flow filtration mode even when identical membranes were tested, 14.48% (RosB) (see Fig. S11). In particular, the rejection rate of EB
regardless of their thickness, and the permeance of the 40-nm- increased under cross-flow filtration, while the rejection rates of
thick GONR membrane was ultrafast at 679 LMH/bar. Because de- the other dyes decreased. The lower rejection rate of RosB (1018 Da)
fects can increase the solvent permeance, photographs and SEM than that of EB (961 Da) can be attributed to the shape of the dye
images of the GONR membranes were obtained after cross-flow molecule (Fig. S12): RosB with a planar shape can relatively easily
filtration experiments (Fig. S8), and no defective areas were penetrate the nanochannels of the stacked GONRs, as is frequently
observed in the membranes. This observation indicates that the observed with graphene-based membranes [37]. The 40-nm-thick
enhanced permeance under cross-flow is the nature of the GONR GONR membranes were also tested at higher pressures ranging up
membrane rather than permeation through defective bypaths, to 12 bar; they showed a high and stable rejection rate (~100%) for
although the related mechanism is not clear at this stage. EB (Fig. S13). These results indicate that the ultrathin GONR layer
Additionally, other organic solvent permeances in GONR (thickness, 40 nm) is effective as a molecular sieve in the separation
membrane were tested (Fig. S9). The thickness of the used GONR of subnanometer-sized molecules, and its molecular weight cut-off
membrane was 150 nm. The obtained experimental data exhibited (MWCO) is around 961 Da.

Fig. 4. (A) Permeance of pure IPA depending on the thickness of the GONR layer and filtration modes. (B) Rejection rate of a 40 nm thick GONR membrane for dye molecules
dispersed in IPA. Filtration was conducted at 1 bar using a solution with a dye concentration of 10 mg/L. (C) Comparison of the IPA nanofiltration performances. (D) Schematic of the
diafiltration mechanism of the GONR membrane for a mixed dye solution in IPA. (E) EB/MB mixture separation using a 40 nm thick GONR membrane under cross-flow filtration at
1 bar and (F) the corresponding UVevis absorbance spectra of feed and permeates. Insets show photographs of the feed/permeate solutions after 56 h of cross filtration with the
membrane. The dye concentration was 10 mg/L for the single dye solution. A 20 mg/L binary dye solution was tested for diafiltration. (A colour version of this figure can be viewed
online.)

568
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

The IPA nanofiltration performance of the GONR membrane is CRediT authorship contribution statement
compared with those of other 2D-material-based membranes in
Fig. 4C and Table S1 [38e51]. The performances of the GONR Ji Hoon Kim: Conducted the overall experiment, wrote the
membranes in both dead-end and cross-flow experiments exceed manuscript and all experiments were conducted under the super-
the upper limits of the performances of other membranes. In vision of D.W. Kim. Yunkyu Choi: characterized the structure of
particular, the membrane performance under cross-flow filtration graphene. Junhyeok Kang: characterized the structure of gra-
(679 LMH/bar of IPA permeance, 961 Da of MWCO) far surpasses phene. Ju Yeon Kim: wrote the manuscript and all experiments
those of the previously reported polymeric membranes or gra- were conducted under the supervision of D.W. Kim. Jun Hyuk Bae:
phene membranes, while the MWCO is larger than that of the conducted formal analysis. Ohchan Kwon: conducted formal
nanoporous graphene membrane (nGO) obtained by rapid thermal analysis. Dae Woo Kim: wrote the manuscript and all experiments
activation [45,46]. were conducted under the supervision of D.W. Kim.
According to the aforementioned results for membrane filtra-
tion with a single dye solution, GONR membranes have a sharp Declaration of competing interest
MWCO and high IPA permeance under cross-flow filtration, which
can be beneficial for the diafiltration of mixed organic molecules, as The authors declare that they have no known competing
illustrated in Fig. 4D. In this process, the small molecule in the financial interests or personal relationships that could have
mixed solution permeates the membrane, whereas the large appeared to influence the work reported in this paper.
molecule accumulates on the membrane feed side [46,52,53].
Fig. 4E shows the diafiltration performance of 40-nm-thick GONR Acknowledgments
membranes under cross-flow filtration for an EB/MB mixture. EB
(961 Da) and MR (269 Da) were selected for preparing the mixed This work was supported by a National Research Foundation of
dye solution because the MWCO of the GONR membrane lies be- Korea (NRF) grant funded by the Korean government (MSIT) (NRF-
tween the sizes of EB and MR. These two dyes were mixed to a total 2020R1C1C1003289). This research was also supported by the Basic
concentration of 20 mg/mL in IPA; Fig. S14 shows the UVevis Science Research Program through the National Research Founda-
spectrum of this solution. Filtration was conducted at 1 bar for tion of Korea funded by the Ministry of Education (NRF-
56 h, and the inset in Fig. 4F shows the photographs of the GONR 2019R1A6A1A11055660).
membrane before and after filtration. Fig. 4F shows the UVevis
absorbance spectra of the feed and permeate solutions corre-
Appendix A. Supplementary data
sponding to Fig. 4E. The permeated solution did not show the
absorbance peak of EB molecules, indicating that EB was selectively
Supplementary data to this article can be found online at
removed by the GONR membrane with a 100% rejection rate.
https://doi.org/10.1016/j.carbon.2022.02.026.
Interestingly, the intensity of the absorbance peak of MR in the
permeate solution is identical to that of the mixed solution, indi-
References
cating that MR is highly permeable through the GONR membrane.
A stable separation of the mixture was observed over 56 h, while [1] M. Terrones, Nanotubes unzipped, Nature 458 (2009) 845e846.
the permeance declined slightly from 900 to 280 LMH/bar, possibly [2] M. He, J. Dong, H. Wang, H. Xue, Q. Wu, B. Xin, W. Gao, X. He, J. Yu, H. Sun,
due to compact selective layer, surface fouling, or pore blocking F. Ding, J. Zhang, Advance in close-edged graphene nanoribbon: property
investigation and structure fabrication, Small 15 (2019) 1804473.
caused by the adsorbed dye molecules [31,54,55]. Despite the slight [3] X. Zhou, G. Yu, Modified engineering of graphene nanoribbons prepared via
decline in permeance, the GONR membrane maintained ultrafast on-surface synthesis, Adv. Mater. 32 (2020) 1905957.
IPA/small dye molecule permeation while blocking larger [4] L. Jiao, L. Zhang, X. Wang, G. Diankov, H. Dai, Narrow graphene nanoribbons
from carbon nanotubes, Nature 458 (2009) 16.
molecules.
[5] T. Asano, J. Nakamura, Edge-state-induced stacking of zigzag graphene
nanoribbons, ACS Omega 4 (2019) 22035e22040.
4. Conclusions [6] Y.W. Lan, C. Torres Jr., X. Zhu, C.L. Sun, S. Zhu, C.D. Chen, K. Wang, Self-aligned
graphene oxide nanoribbon stack with gradient bandgap for visible-light
photodetection, Nano Energy 27 (2016) 114e120.
We observed the assembly of GONRs into a GONR nanosheet as [7] L. Tapaszto, G. Dobrik, P. Lambin, L.P. Biro , Tailoring the atomic structure of
a GONR suspension penetrated a narrow microchannel. It is graphene nanoribbons by scanning tunnelling microscope lithography, Nat.
observed that a thin GONR nanosheet can be formed by decreasing Nanotechnol. 3 (2008) 397e401.
[8] D.B. Kosynkin, W. Lu, A. Sinitskii, G. Pera, Z. Sun, J.M. Tour, Highly conductive
the gap width of the microchannel, critically below 50 mm. Further, graphene nanoribbons by longitudinal splitting of carbon nanotubes using
we used the GONR nanosheet to fabricate high-performance OSN potassium vapor, ACS Nano 5 (2011) 968e974.
membranes using a scalable slot-die coating method. In addition [9] E.C. Martínez, J.C. Gonza lez, J. Sovich, M.D. Lima, High temperature structural
transformations of few layer graphene nanoribbons obtained by unzipping
to the shear effect in the microchannel of the slot-die head, the carbon nanotubes, J. Mater. Chem. 2 (2014) 221.
shear effect of the moving stage with the PES support also changed [10] A.P. Johnson, C. Sabu, N.K. Swamy, A. Anto, H.V. Gangadharappa, K. Pramod,
the structure of the coated GONR layer, leading to a thinner Graphene nanoribbon: an emerging and efficient flat molecular platform for
advanced biosensing, Biosens. Bioelectron. 184 (2021) 113245.
deposited layer at a higher stage moving rate. The GONR mem- [11] H. Jin, H. Huang, Y. He, X. Feng, S. Wang, L. Dai, J. Wang, Graphene quantum
brane showed high permeance of 186 LMH/bar for pure IPA in dots supported by graphene nanoribbons with ultrahigh electrocatalytic
dead-end filtration and moderate permeance of 679 LMH/bar in performance for oxygen reduction, J. Am. Chem. Soc. 137 (2015) 7588e7591.
[12] R. Nadiv, M. Shtein, M. Buzaglo, S.P. Damari, A. Kovalchuk, T. Wang, J.M. Tour,
cross-flow filtration with a MWCO of 961 Da. The membrane could O. Regev, Graphene nanoribbon - polymer composites: the critical role of edge
be used in the diafiltration of mixed dyes in organic solutions. Our functionalization, Carbon 99 (2016) 444e450.
observations indicated that the structure of 1D materials such as [13] H.R. Lee, D.W. Kim, V.O. Jones, Y. Choi, V.E. Ferry, M.A. Geller, S.M. Azarin,
Sonosensitizer-functionalized graphene nanoribbons for adhesion blocking
GONRs can be further tuned using a physical treatment, such as by
and sonodynamic ablation of ovarian cancer spheroids, Adv. Healthcare
using the shear effect, without performing an additional chemical Mater. 10 (2021) 2001368.
treatment. This study provides insights into the scalable fabrica- [14] D.V. Kosynkin, A.L. Higginbotham, A. Sinitskii, J.R. Lomeda, A. Dimiev,
tion of 1D-material-based coatings and membranes and enhances B.K. Price, J.M. Tour, Longitudinal unzipping of carbon nanotubes to form
graphene nanoribbons, Nature 458 (2009) 16.
the understanding of the assembly behaviors of 1D materials in [15] A. Hirsch, Unzipping carbon nanotubes: a peeling method for the formation of
their suspensions. graphene nanoribbons, Angew. Chem. Int. Ed. 48 (2009) 6594e6596.

569
J.H. Kim, Y. Choi, J. Kang et al. Carbon 191 (2022) 563e570

[16] J.I. Paredes, S. Villar-Rodil, A. Martinez-Alonso, M.M. Tascon, Graphene oxide separation capability of graphene oxide membranes in organic solvents, ACS
dispersions in organic solvents, Langmuir 24 (19) (2008) 10560. Nano 14 (2020) 6013e6023.
[17] S. Naficy, R. Jalili, S.H. Aboutalebi, R.A. Gorkin III, K. Konstantinov, P.C. Innis, [37] A. Nordenstro €m, N. Boulanger, A. Iakunkov, I. Baburin, A. Klechikov,
G.M. Spinks, P. Poulin, G.G. Wallace, Graphene oxide dispersions: tuning A. Vorobeiv, A. Talyzin, Intercalation of dyes in graphene oxide thin films and
rheology to enable fabrication, Mater. Horiz. 1 (2014) 326. membranes, J. Phys. Chem. C 125 (2021) 6877e6885.
[18] X. Dong, D. Lu, T.L. Harris, I. Escobar, Polymers and solvents used in membrane [38] L. Hao, H. Zhang, X. Wu, J. Zhang, J. Wang, Y. Li, Novel thin-film nano-
fabrication: a review focusing on sustainable membrane development, composite membranes filled with multi-functional Ti3C2Tx nanosheets for
Membranes 11 (2021) 309. task-specific solvent transport, Compos. Part A Appl. Sci. Manuf. 100 (2017)
[19] J.H. Kim, Y. Choi, J. Kang, E. Choi, S.E. Choi, O. Kwon, D.W. Kim, Scalable 139e149.
fabrication of deoxygenated graphene oxide nanofiltration membrane by [39] M.F. Jimenez-Solomon, Q.L. Song, K.E. Jelfs, M.M. Ibanez, A.G. Livingston,
continuous slot-die coating, J. Membr. Sci. 612 (2020) 118454. Polymer nanofilms with enhanced microporosity by interfacial polymeriza-
[20] Y. Choi, S.S. Kim, J.H. Kim, J. Kang, E. Choi, S.E. Choi, J.P. Kim, O. Kwon, tion, Nat. Mater. 15 (2016) 760e767.
D.W. Kim, Graphene oxide nanoribbon hydrogel: viscoelastic behavior and [40] Y. Feng, M. Weber, C. Maletzko, T.S. Chung, Facile fabrication of sulfonated
use as a molecular separation membrane, ACS Nano 14 (2020) 12195e12202. polyphenylenesulfone (sPPSU) membranes with high separation performance
[21] M.Y. Song, Y.S. Yun, N.R. Kim, H.J. Jin, Dispersion stability of chemically for organic solvent nanofiltration, J. Membr. Sci. 549 (2018) 550e558.
reduced graphene oxide nanoribbons in organic solvents, RSC Adv. 6 (2016) [41] Z. Si, Z. Wang, D. Cai, G. Li, S. Li, P. Qin, A high-permeance organic solvent
19389. nanofiltration membrane via covalently bonding mesoporous MCM-41 with
[22] Z. Xu, C. Gao, Graphene chiral liquid crystals and macroscopic assembled fi- polyimide, Separ. Purif. Technol. 241 (2020) 116545.
bres, Nat. Commun. 2 (2011) 571. [42] X. Li, S.D. Feyter, D. Chen, S. Aldea, P. Vandezande, F.D. Prez, F.J. Vankelecom,
[23] R. Narayan, J.E. Kim, J.Y. Kim, K.E. Lee, S.O. Kim, Graphene oxide liquid crystals: Solvent-resistant nanofiltration membranes based on multilayered poly-
discover, evolution and applications, Adv. Mater. 28 (2016) 3045e3068. electrolyte complexes, Chem. Mater. 20 (2008) 3876e3883.
[24] J.E. Kim, T.H. Han, D.S.H. Lee, J.Y. Kim, D.C.W. Ahn, D.J.M. Yun, S.O. Kim, Gra- [43] H. Zhang, Y. Zhang, L. Li, S. Zhao, H. Ni, S. Cao, J. Wang, Cross-linked
phene oxide liquid crystals, Angew. Chem. Int. Ed. 50 (2011) 3043e3047. polyacrylonitrile/polyethyleneimine-polydimethysiloxane composite mem-
[25] A. Akbari, P. Sheath, S.T. Martin, D.B. Shinde, M. Shaibani, P.C. Banerjee, brane for solvent resistant nanofiltration, Chem. Eng. Sci. 106 (2014)
R. Tkacz, D. Bhattacharyya, M. Majumder, Large-area graphene-based nano- 157e166.
filtration membrane by shear alignment of discotic nematic liquid crystals of [44] A.A. Tashvigh, L. Luo, T.S. Chung, M. Weber, C. Maletzko, A novel ionically
graphene oxide, Nat. Commun. 7 (2016) 10891. cross-linked sulfonated polyphenylsulfone (sPPSU) membrane for organic
[26] D. Schmelter, H.H. Bruening, Highly ordered graphene oxide and reduced solvent nanofiltration (OSN), J. Membr. Sci. 545 (2018) 221e228.
graphene oxide based polymer nanocomposites: promise and limits for dy- [45] J. Jang, Y.T. Nam, D. Kim, Y.J. Kim, D.W. Kim, H.T. Jung, Turbostatic nanoporous
namic impacts demonstrated in model organic coatings, ACS Appl. Mater. carbon sheet membrane for ultrafast and selective nanofiltration in viscous
Interfaces 8 (2016) 16328e16338. green solvents, J. Mater. Chem. 8 (2020) 8292.
[27] C. Xiang, N. Behabtu, Y. Liu, H.G. Chae, C.C. Youn, B. Genorio, D.E. Tsentalovich, [46] J. Kang, Y. Choi, J.P. Kim, J.H. Kim, J.K. Kim, O. Kwon, D.I. Kim, D.W. Kim,
C. Zhang, D.V. Kosynkin, J.R. Lomeda, C.C. Hwang, S. Kumar, M. Pasquali, Thermally-induced pore size tuning of multilayer nanoporous graphene for
J.M. Tour, Graphene nanoribbons as an advanced precursor for making carbon organic solvent nanofiltration, J. Membr. Sci. 637 (2021) 119620.
fiber, ACS Nano 7 (2013) 1628e1637. [47] M.H.D.A. Farahani, D. Hua, T.S. Chung, Cross-linked mixed matrix membranes
[28] K.H. Thebo, X. Qian, Q. Zhang, L. Chen, H.M. Cheng, W. Ren, Highly stable (MMMs) consisting of amine-functionalized multi-walled carbon nanotubes
graphene-oxide-based membranes with superior permeability, Nat. Commun. and P84 polyimide for organic solvent nanofiltration (OSN) with enhanced
9 (2018) 1486. flux, J. Membr. Sci. 548 (2018) 319e331.
[29] T. Yun, G.H. Jeong, S.P. Sasikala, S.O. Kim, 2D graphene oxide liquid crystal for [48] Z.F. Gao, G.M. Shi, Y. Cui, T.S. Chung, Organic solvent nanofiltration (OSN)
real-world applications: energy, environment, and antimicrobial, Apl. Mater. membranes made from plasma grafting of polyethylene glycol on cross-linked
8 (2020), 070903. polyimide ultrafiltration substrates, J. Membr. Sci. 565 (2018) 169e178.
[30] S.E. Choi, S.S. Kim, E. Choi, J.H. Kim, Y. Choi, J. Kang, O. Kwon, D.W. Kim, [49] S. Ma, Y. Ji, Y. Dong, S. Chen, Y. Wang, S. Lü, An environmental-friendly
Diamine vapor treatment of viscoelastic graphene oxide liquid crystal for gas pesticide-fertilizer combination fabricated by in-situ synthesis of ZIF-8, Sci.
barrier coating, Sci. Rep. 11 (2021) 9518. Total Environ. 789 (2021) 147845.
[31] J.H. Kim, G.S. Park, Y.J. Kim, E. Choi, J. Kang, O. Kwon, S.J. Kim, J.H. Cho, [50] X. Li, P. Vandezande, I.J. Vankelecom, Polypyrrole modified solvent resistant
D.W. Kim, Large-area Ti3C2Tx-MXene coating: toward industrial-scale fabri- nanofiltration membranes, J. Membr. Sci. 320 (2008) 143e150.
cation and molecular separation, ACS Nano 15 (2021) 8860e8869. [51] A. Akbari, S.E. Meragawi, S.T. Martin, B. Corry, E. Shamsaei, C.D. Easton,
[32] O. Kwon, Y. Choi, E. Choi, M. Kim, Y.C. Woo, D.W. Kim, Fabrication techniques D. Bhattacharyya, M. Majumder, Solvent transport behavior of shear aligned
for graphene oxide-based molecular separation membranes: towards indus- graphene oxide membranes and implications in organic solvent nano-
trial application, Nanomaterials 11 (2021) 757. filtration, ACS Appl. Mater. Interfaces 10 (2018) 2067e2074.
[33] R.R. Aburto, L.B. Alemany, T.K. Weldeghiorghis, S. Ozden, Z. Peng, A. Lherbier, [52] Y. Choi, J. Kang, E. Choi, J.Y. Kim, J.P. Kim, J.H. Kim, O. Kwon, D.W. Kim, Carbon
A.R.B. Me ndez, C.S. Tiwary, J.T. Tijerina, Z. Yan, M. Tabata, J.C. Charlier, nanotube-supported graphene oxide nanoribbon bilayer membrane for high-
J.M. Tour, P.M. Ajayan, Chemical makeup and hydrophilic behavior of gra- performance diafiltration, Chem. Eng. J. 427 (2022) 131805.
phene oxide nanoribbons after low-temperature fluorination, ACS Nano 9 [53] P. Marchetti, M.J. Solomon, G. Szekely, A. Livingston, Molecular separation
(2015) 7009e7018. with organic solvent nanofiltration: a critical review, Chem. Rev. 114 (2014)
[34] D.W. Kim, I. Kim, J. Jang, Y.T. Nam, K. Park, K.O. Kwon, K.M. Cho, J. Choi, D. Kim, 10735e10806.
K.M. Kang, S.J. Kim, Y. Jung, H.T. Jung, One dimensional building blocks for [54] S. Mondal, S. De, A fouling model for steady state crossflow membrane
molecular separation: laminated graphitic nanoribbons, Nanoscale 9 (2017) filtration considering sequential intermediate pore blocking and cake for-
19114. mation, Separ. Purif. Technol. 75 (2010) 222e228.
[35] W. Li, W. Wu, Z. Li, Controlling interlayer spacing of graphene oxide mem- [55] J.Y. Chong, B. Wang, C. Mattevi, K. Li, Dynamic microstructure of graphene
branes by external pressure regulation, ACS Nano 12 (2018) 9309e9317. oxide membranes and the permeation flux, J. Membr. Sci. 549 (2018)
[36] S. Zheng, Q. Tu, M. Wang, J.J. Urban, B. Mi, Correlating interlayer spacing and 385e392.

570

You might also like