Subscriber access provided by Kaohsiung Medical University
Critical Review
Porous PVdF/GO Nanofibrous Membranes for Selective Separation and Recycling of Charged Organic Dyes from Water Abdul Ghaffar, Lina Zhang, Xiaoying Zhu, and Baoliang Chen Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b06081 • Publication Date (Web): 28 Feb 2018 Downloaded from http://pubs.acs.org on February 28, 2018
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 34
Environmental Science & Technology
1
Porous PVdF/GO Nanofibrous Membranes for Selective Separation
2
and Recycling of Charged Organic Dyes from Water
3 4
Abdul Ghaffar 1, 2, Lina Zhang 1,2, Xiaoying Zhu 1, 2 and Baoliang Chen* 1, 2
5
1. Department of Environmental Science, Zhejiang University, Hangzhou, Zhejiang 310058,
6
China.
7
2. Zhejiang Provincial Key Laboratory of Organic Pollution Process and Control, Hangzhou
8
310058, China.
9 10 11
*Corresponding authors email:
12
Dr. Chen Baoliang
[email protected] 14
Tel:
+86 – 571 – 8898 – 2587
15
Fax:
+86 – 571 – 8898 – 2587
13
16 17 18
Coauthor emails
19
Abdul Ghaffar:
[email protected] 20
Lina Zhang:
[email protected] 21
Dr. Xiaoying Zhu:
[email protected] 22
1 ACS Paragon Plus Environment
Environmental Science & Technology
23
Abstract
24
Graphene oxide (GO) membranes are robust and continue to attract great attention due to their
25
fascinating properties, despite their potential issues regarding stability and selectivity in aqueous-
26
phase processing. That being said, however, the functional moieties of GO could be used for
27
membrane surface modification, while ensuring simultaneous removal and recycling of industrial
28
organic dyes. Herein, we present a versatile porous structured polyvinylidene fluoride-graphene
29
oxide (PVdF-GO) nanofibrous membranes (NFMs), prepared by using simple and
30
straightforward electrospinning approach for selective separation and filtration. The GO
31
nanosheets were distributed homogenously throughout the PVdF nanofiber, regulating the
32
surface morphology and performance of PVdF-GO NFM. The PVdF-GO NFMs possesses high
33
mechanical strength and surface free energy (SFE), consequently resulting high permeation and
34
filtration efficiency as compared to PVdF NFM. The selectivity towards positively charged dyes
35
(99%) based on electrostatic attraction, while maintaining rejection (100%) for negatively
36
charged dye from mixed solutions highlight the role of GO in PVdF-GO NFM, owing to uniform
37
pores and negatively charged surface. In addition, the actual efficiency of NFMs could be
38
recovered easily up to three consecutive filtration cycles by regeneration, thereby assuring high
39
stability. The high permeation, purification and filtration efficiency, good stability and recycling
40
of PVdF-GO NFMs are promising for use in practical water purification and applications,
41
particularly for selective filtration and recycling of dyes.
42
Keywords: PVdF-GO nanofibrous membrane, selectivity, molecular filtration, dye pollutant,
43
regeneration.
44 45
2 ACS Paragon Plus Environment
Page 2 of 34
Page 3 of 34
46
Environmental Science & Technology
1. Introduction
47
GO as verified to be two-dimensional and one atom thick sp2 carbon lattice, has attracted
48
great attention in recent years because of its superior properties such as extremely large surface
49
area with chemical inertness and strong mechanical properties.1–3 Membranes based on GO has
50
great potential for separation and purification with promising flux, antifouling and high
51
selectivity for gases and liquids.4–9 Unfortunately, the negatively charged functional moieties on
52
GO causes repulsion in aqueous solution, resulting disintegration and destabilization but also
53
decrease the selectivity of GO membranes.2,
54
achieve asymmetric structure of GO membranes, the transfer process, large volume of liquid, GO
55
alignment, layer-by-layer (LbL) assembly and dip coating, following CVD or restacking process
56
arguably limits the scalability of GO membranes with rapid productivity issues for industrial
57
applications.5, 13–15 Nonetheless, in practice, the abundant oxygen containing functional groups
58
on GO can be used for surface modification of membranes and charge based separation. In
59
addition, organic pollutants were mainly adsorbed on basal plane of GO nanosheets, while water
60
enters in GO modified membrane surface primarily around the oxidized edges of GO
61
nanosheets.7 Consequently, it is more important to recycle organic pollutants rather than
62
degradation in order to minimize their environmental impact. For example, environmental
63
problems associated with dye baths are related to wide variety of different components, often in
64
relatively high concentrations. In future, many of dye processing units will face the requirement
65
of reuse and recycling of industrial wastewater, because traditionally used methods are
66
insufficient for obtaining the required water quality. So far, simultaneous removal and recycling
67
has been paid very little attention.
10–12
Despite all this considerable progress to
3 ACS Paragon Plus Environment
Environmental Science & Technology
68
The separation and filtration technology play crucial roles in industrial water treatment,
69
biomedical engineering, food processing, organic pollutants recovery, textile processing and
70
pharmaceutical purification.16–20 The conventional separation technology like distillation and
71
evaporation require immense space and are energy intensive, which makes the industrial
72
processing costly in terms of operation, energy and resources.18 To date, the membranes made up
73
of flexible polymers such as polysulfone, polyamide and cellulose are being utilized on large
74
scale because of simple and easy processing despite their limitations like swelling,
75
decomposition beyond critical temperature and pH, poor tolerance to corrosive solvents, oxidants,
76
or strong acidic/alkaline reagents under mild operative conditions.21–23 In addition, the routine
77
polymeric membranes can barely filter the aqueous mixtures and later are hard to recycle with
78
prolonged durability.13–24 Nonetheless, charge based selectivity and separation for similar
79
organic molecules with identical properties demand less energy and space, while being
80
productive, durable and are considered to be environmental friendly and cheaper.25 Therefore,
81
facile fabrication of membranes for charge based filtration with adequate thermal or chemical
82
stability and effective processing under different environments are challenging. Besides, the
83
composite membranes may overcome the drawbacks of polymeric membranes and could be good
84
candidates for charge based filtration and selectivity under critical environmental conditions.
85
Electrospinning technology allows a promising and straightforward approach to fabricate
86
consecutive and perpetual nonwoven NFMs containing organic and inorganic fibers or both with
87
proportions ranging from micro to nanoscale.26–29 Although, the NFMs exhibit porous structure
88
and high surface area under different compositions, which have been reported earlier for
89
separation and purification. Currently, some NFMs such as silica-titanium dioxide, carbon-silica,
90
cyclodextrin, polyethersulfone, biochar-polyvinylidene fluoride and hollow zinc oxide were used
4 ACS Paragon Plus Environment
Page 4 of 34
Page 5 of 34
Environmental Science & Technology
91
in separation, filtration and photocatalysis.24, 30–35 However, despite all this considerable research
92
and progress regarding NFMs, the existing NFMs are mostly tenuous resulting easy breakage,
93
which may put possible difficulties during practical applications.13, 24, 30, 36, 37 Therefore, the facile
94
fabrication of GO into PVdF NFM would be of both technological and scientific importance. To
95
date, the NFMs containing PVdF-GO has not been utilized for charge based separation and
96
filtration of complex mixtures, which would be of great interest for industrial applications.
97
Herein, we report a porous structured PVdF-GO NFM fabricated by electrospinning
98
technology. The premise of our strategy is that GO is sustainably incorporated into NFMs for
99
desirable recycling of organic dyes rather than degradation from mixed organic dye solutions,
100
based on electrostatic attraction/repulsion besides molecular size. More importantly, the knotty
101
porous structure in between twisted fibers, GO embedded NFMs significantly gives rise to high
102
permeation while simultaneously maintaining filtration efficiency even after three filtration
103
cycles. The selective separation and recycling based on NFMs may provide a new approach to
104
manufacture molecular filters for selectivity, filtration efficiency and high solvent permeate flux
105
toward mixed organic molecules within fields of practical molecular purification and separation
106
science.
107 108
2. Materials and Methods
109
2.1. Materials.
110
Natural graphite flake (325 mesh, 99.8%) was purchased from Alfa Aesar. PVdF (Mw ~
111
180000 by GPC) was obtained from Sigma-Aldrich. Analytical grade dimethylacetamide
112
(DMAc >99%), formamide, diiodomethane, ethylene glycol, glycerol, methylene blue (MB),
113
neutral red (NR), basic fuchsin (BF), rhodamine B (RB) and methyl orange (MO) were
5 ACS Paragon Plus Environment
Environmental Science & Technology
114
purchased from Aladdin chemicals reagent company (Shanghai, China). HPLC grade acetone
115
(>99.5%) was obtained from Tedia, U.S.A. Ultrapure water was produced by CascadaTM IX-
116
water purification system, PALL Corporation. All chemicals were used as obtained without
117
further refinement.
118
2.2. Synthesis of GO and Preparation of PVdF-GO NFMs.
119
GO was synthesized from natural Graphite flake (325 mesh, 99.8%, Alfa Aesar) by modified
120
Hummers method.38, 39 The detailed process regarding GO synthesis can be found in supporting
121
information (SI).1,
122
straightforward electrospinning technique with subsequent PVdF and GO mixtures. The PVdF-
123
GO precursor solution was obtained by adding PVdF and GO into DMAc and acetone under
124
sonication for 1h and stirring at 70°C for 12h. The composition of electrospinning solution was
125
20% PVdF, 64% DMAc and 16% acetone for PVdF NFM, 17.5% PVdF, 2.5% GO, 64% DMAc
126
and 16% acetone for PVdF-GO 2.5% NFM, and 15% PVdF, 5% GO, 64% DMAc and 16%
127
acetone for PVdF-GO 5.0% NFMs, respectively. The resultant homogeneous solutions were then
128
filled into 10mL syringe with 18 gauge needle. The syringe was positioned horizontally for 30
129
min and air was removed completely. The solutions were fed at controllable propulsion velocity
130
of 0.5mL h−1 with an automatic pump. The electrospinning was performed by using QZNT–E01
131
spinning equipment with a voltage supply of 15kV at 8cm distance. The entire spinning process
132
was performed under ambient operating temperature and relative humidity of 25°C and 40%,
133
respectively. The NFMs were then thermally treated in between glass plates at 120°C for 8h in
134
dry-oven to assure pore diameter and preserve physical property. Finally, the NFMs were rinsed
135
with methanol and ultrapure water to remove any remaining impurities completely.
136
2.3. Characterization of NFMs
39
The PVdF and PVdF-GO NFMs were prepared by combining the
6 ACS Paragon Plus Environment
Page 6 of 34
Page 7 of 34
Environmental Science & Technology
137
The surface morphologies of NFMs were examined by field emission scanning electron
138
microscopy (FE-SEM, S-4800, Hitachi, Tokyo). Transmission electron microscopy (JEM-1230,
139
JEOL Ltd., Japan) and energy dispersive X-ray spectroscopy (EDX) were used to get insight into
140
the microstructure and composition of NFMs. Raman spectra were collected at room temperature
141
with LabRamHRUV Raman spectrometer (JDbin-yvon, FR); equipped with thermoelectrically
142
cooled charge-coupled device (CCD), with an excitation source by Ar+ laser at 514 nm
143
wavelength. The crystal phase of PVdF and PVdF-GO NFMs were analyzed by X-ray diffraction
144
(XRD, Rigaku D/MAX/PC 2550, Japan) equipped with Cu Kα radiation source. The mechanical
145
properties in terms of tensile stress and strain were obtained by Universal material experiment
146
machine (Zwick/Reoll Z020). The surface area (SA) was measured via Brunauer, Emmett and
147
Teller (BET) method by N2 (0.162 nm2) gas sorption at 77 K, on NOVA-2000E surface area
148
analyzer (Quantachrome Instruments). The nonlocal density functional theory (NLDFT) model
149
was applied to analyze pore size distribution (PSD). An electro-kinetic analyzer Anton Paar
150
SurPASS (GmbH, Austria) was used to evaluate streaming potentials of NFMs. The
151
measurements were carried out using 200mL of 0.1 mM KCl solution as electrolyte and pH was
152
adjusted by 0.1M HCl or 0.1M NaOH. The water contact angles (CA) of NFMs were obtained
153
by OSA200–T Optical Contact Angle analyzer following air drop (~10 µL) method with a
154
micro-syringe to the targeted NFM surface. Attenuated total reflectance fourier transformed
155
infrared (ATR-FTIR) spectra were obtained using a Thermo-Fisher Scientific; with 64 scans in
156
range of 4000-400 cm-1 at resolution of 4 cm-1 for each NFM.
157
2.4. Adsorption Kinetics and Filtration
158
Five different organic dyes namely MB, NR, BF, RB and MO (purity > 98%) were selected
159
as model representative pollutants to get insight into adsorption kinetics and dynamic molecular
7 ACS Paragon Plus Environment
Environmental Science & Technology
160
filtration. These five compounds were selected based on their variation in electrostatic charge
161
and molecular size and detailed physicochemical properties are listed in Table S1. Adsorption
162
kinetics of dyes were obtained using a batch equilibration experiment at an initial concentration
163
of 5mg L-1. Briefly, the sorbates were dissolved in water and 5mg NFM samples were immersed
164
in five kind of solutions to probe higher affinity and adsorption capacity toward organic dyes.
165
Furthermore, the molecular filtration experiments were carried out by using Millipore filtration
166
device equipped with N2 cylinder for pressure, a digital balance and computer. Prior to filtration
167
process, the NFM samples were initially compacted for 15min at 1 bar to get steady flux. Four
168
mixtures of organic dyes with opposite electrostatic charge (positive/negative), i.e., MB/MO,
169
NR/MO, BF/MO and RB/MO were prepared, each dye with 40mg L−1 concentration (1:1
170
volume), respectively. Three mixtures with same electrostatic charge (positive/positive), i.e.,
171
MB/NR, MB/BF and MB/RB were also prepared to further evaluate the molecular filtration
172
efficiency among positively charged molecules. Following, the mixed dye solutions were forced
173
to flow through NFMs, at an average pressure of 1bar by N2 gas. The permeate concentration
174
was measured by UV–Vis spectrophotometer (UV-2550, Shimadzu), at selected wavelengths
175
(Table S1). The separation efficiency (η) of filtration process was calculated by the following
176
equation:40
177
𝜂% =
[𝑂𝑀]𝐹 × 100 [𝑀𝑂]𝐹 + [𝑂𝑀]𝐹
178
where [OM] represents the organic molecules including MB, NR, BF, RB, MO and [MO] F and
179
[OM]F are the concentrations of MO and the other four organic dyes in filtrate, respectively.
180
After molecular filtration process of mixed dyes, the NFMs were washed thoroughly by mixed
181
solution of hydrochloric acid and ethanol (1:1 by volume) as eluent. Subsequently, the washed
182
NFMs were dried at 80°C to remove residual solvent, thus NFMs can be easily regenerated. The 8 ACS Paragon Plus Environment
Page 8 of 34
Page 9 of 34
Environmental Science & Technology
183
regenerated NFMs were recycled three times and used again for filtration up to three consecutive
184
runs to examine the regeneration and reusability of as-prepared PVdF-GO 5.0% NFM.
185
2.5. Data Analysis
186
The fractal dimension, SFE, work of adhesion between solid-liquid surfaces was examined
187
for characterization of NFMs. Pseudo-first-order and pseudo-second-order kinetic models were
188
used to analyze the adsorption kinetics and capacity of NFMs. More details about data analysis
189
equations for fractal dimension, SFE, work of adhesion, kinetic models, are presented in the SI. Table 1. The structural characteristics and properties of NFMs. Parameter
PVdF
PVdF-GO 2.5%
PVdF-GO 5.0%
60.71/39.14/0.15
67.44/30.81/1.75
69.02/28.76/2.22
127.56 ±7.23
149.12 ±12.45
174.68 ±9.87
ID/IG c
N/A
0.98
0.96
Tensile strength (MPa)
4.39 ±1.3
7.43 ±0.84
11.7 ±2.25
Tensile strain (%)
30.90 ±13.31
33.10 ±3.85
40.10 ±10.96
12.43
13.09
51.47
0.030
0.046
0.077
2.434
2.435
2.377
3.212
3.493
3.902
116.60 ±4.1
72.80 ±4.52
52.10 ±2.21
Surface free energy (mJ m )
51.50
57.36
88.50
Pure water flux (LMH)
193.02 ±6.95
439.00 ±10.23
663.54 ±54
Single fiber composition % (C/F/O) Mean fiber dia
a
b
2
-1 d
Surface area (m g )
Pore volume (cc g-1) e Fractal dimension (D)
f
Isoelectric point (pI) Water contact angle (°) -2 g
a
EDX, b Based on 10 measurements from SEM images, c Raman spectra, d BET method, e NLDFT, f from
N2 isotherms, g Neumann’s method.
190 191
3. Results and Discussions
192
3.1. Morphologies and Structure of NFMs
193
The facile fabrication of NFMs for effective adsorption of different organic molecules, while
194
being selective requires intellectual functional design of filtration membrane, based on controlled 9 ACS Paragon Plus Environment
Environmental Science & Technology
195
relation between structural geometry and pore selectivity. We constructed NFMs for multilevel
196
functionality based on the following criteria: (a) Effective pore geometry thereby assuring
197
maximum adsorption for selected organic dyes while minimum sacrifice for pure water
198
permeation, (b) porous structure with controlled thickness to facilitate the practical application,
199
and (c) gradient selectivity for complex aqueous mixtures. The required criteria for NFMs were
200
fulfilled by scalable and versatile strategy of high speed continuous stirring followed by
201
advanced electrospinning approach. The homogenous solution and controlled ratio of GO can
202
ensure an effective performance of NFMs. Thus, uniform synthesis and tunable geometry of
203
NFMs can be alleviated.
204
The surface morphology of NFMs was examined thoroughly by SEM and representative
205
images are shown in Figures 1a, 1b and 1c. The morphologies of NFMs were changed gradually
206
with GO and became brownish in color with close distribution of GO throughout NFMs. Notably,
207
the NFMs exhibited typically nonwoven structure with random and uniform arrangement of
208
consistent fibers. The NFMs exhibited obvious porous geometry with three dimensional pores
209
interconnected with each other through triangular junctions. The corresponding skeletonized
210
images reveal the triangular junctions with numerous nano, meso and micropores (Fig. 1d, 1e
211
and 1f). The significant difference between PVdF NFM and PVdF-GO NFMs is the unique
212
surface of fibers that consist of GO nanosheets (~1µm). The NFMs with uniform interconnected
213
pores endows the stable structure benefiting for high permeation.23 The morphology of pristine
214
fibers was observed to be smoother without any apparent difference, while maintaining the
215
porous geometry. The mean diameter of PVdF fiber was 127.56nm and gradually increased with
216
GO (149.12 and 174.68nm, Table 1). This phenomenon could be ascribed to viscosity and
10 ACS Paragon Plus Environment
Page 10 of 34
Page 11 of 34
Environmental Science & Technology
217
electrical conductivity of spinning solution, resulting in irregular whipping and larger stretching
218
of jet during electrospinning process.41
Figure 1. SEM images of electrospun NFMs (a) PVdF, (b) PVdF-GO 2.5%, and (c) PVdF-GO 5.0%. The corresponding skeletonized transformation of NFMs (d) PVdF, (e) PVdF-GO 2.5%, and (f) PVdF-GO 5.0%, respectively. 219 220
The representative TEM images provide spontaneous insight into microstructure of single
221
PVdF and PVdF-GO fibers (Fig. 2a, 2b and 2c). Notably, the NFMs were composed of smooth
222
fibers without any significant disordered geometry, which is also consistent with SEM images
223
(Fig. 1a, 1b and 1c). As observed in Figures 2b, and 2c, GO nanosheets were dispersed
224
uniformly, both internally and externally within fibers, endowing the PVdF-GO matrix with
225
classical rough structure. The GO nanosheets were attached on fiber surface homogeneously at
226
nanometer level, potentially as active sites. In order to confirm the existence of GO nanosheets
227
on nanofiber surface, the EDX was performed on single fiber (Fig. 2d, 2e and 2f). As can be seen
228
roughly from single fiber composition (Table 1), the molar ratio of oxygen increased with GO
229
ratio. The EDX images verify the successful incorporation of GO nanosheets on fiber surface. 11 ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 34
230
(d)
(e)
(f)
Figure 2. TEM images of NFMs (a) PVdF, (b) PVdF-GO 2.5%, and (c) PVdF-GO 5.0%. Elemental mapping images of C, F and O of selected area (arrow) from single fiber at 25nm scale (d) PVdF, (e) PVdF-GO 2.5% and (f) PVdF-GO 5.0%. The symbols ,
and
represent
C, F and O elements, respectively. 231 232
The surface chemistry of NFMs was further examined by Raman spectroscopy and XRD
233
patterns. In Raman spectra, two prominent peaks D-band (1349cm−1) and G-band (1582cm−1),
234
were observed for GO (Fig. 3a). The D-band intensity originated from sp3 hybridization and G-
235
band was induced by crystalline graphitic/sp2 carbon atoms, respectively. The Raman bands
236
observed at 795cm-1 and 839cm-1 in PVdF NFM corresponds to α and β phases, which are often
237
used for phase identification in PVdF. The PVdF-GO 2.5% and PVdF-GO 5.0% NFMs clearly
238
showed the presence of D-band and G-band (Fig. 3a), confirming the presence of adjacent GO
239
nanosheets, thereby preserving the π-stacking on fiber surface. The intensity of D peak was
240
comparable to G peak and ratio of D-band to G-band (ID/IG) provides a sensitive measure of
241
disorder and crystallite size of graphitic layers. The D-band intensity was lower than G-band, 12 ACS Paragon Plus Environment
Page 13 of 34
Environmental Science & Technology
242
indicating the structural defects and disorder, because of oxidation and reduction process. The
243
intensity ratio of D-band to G-band (ID/IG) of GO (1.01) was larger than PVdF-GO NFMs (0.98
244
and 0.96, Table 1), suggesting the existence of deeper sp2 domains. A clear transition of PVdF
245
with GO could also be observed by SEM images of NFMs (Fig. 1b and 1c). In addition, XRD
246
patterns of as-prepared PVdF and PVdF-GO NFMs are shown in Figure 3b. The XRD pattern of
247
GO reveals an intense and sharp peak at 9.74°, related to the diffraction peak of GO. The XRD
248
pattern of PVdF NFM consists predominantly α-phase (nonpolar) with characteristic (020), (110)
249
and (021) peaks at 2θ = 18.4°, 19.9° and 26.6°, respectively. The PVdF-GO NFMs show a broad
250
peak around 2θ = 9.76°, indicative of polar β-phase (200) and (110) peaks and progressively
251
increased with GO content. However, the XRD peaks in PVdF-GO NFMs were broad with
252
presence of semi crystalline and GO nanosheets implying α-, β- and γ-phases. The presence of
253
GO in PVdF matrix does not change the structural crystal phase compared with PVdF, despite
254
different α-, β- and γ-phase compositions. As the maximum concentration of GO is 5wt% in
255
NFMs, their corresponding peaks in XRD patterns were weaken, but progressively lead to peak
256
formation. Thus, the Raman and XRD analysis direct confirmation of graphitic structure,
257
anchored on PVdF fiber surfaces. The substantial arrangement of graphitic structure is promising
258
for efficient water permeation and better filtration performance.23
13 ACS Paragon Plus Environment
Environmental Science & Technology
Figure 3. (a) Raman spectra of GO and PVdF-GO NFMs. (b) XRD patterns of GO, PVdF and PVdF-GO NFMs. (c) Tensile strength of PVdF and PVdF-GO NFMs. 259 260
3.2. Mechanical Properties of NFMs
261
The mechanical properties of PVdF-GO NFMs in terms of tensile stress and strain were
262
measured to probe the ability to withstand the transmembrane pressure during filtration. The
263
corresponding tensile stress and strain were listed in Table 1 and Figure 3c, respectively. The
264
tensile stress of PVdF NFM was 4.39 MPa, with an elongation of 30.9%, both being the lowest
265
among prepared NFMs. The NFMs with GO i.e., PVdF-GO 2.5% and PVdF-GO 5.0% had
266
higher tensile strength (7.43 and 11.7MPa), which could be attributed to contribution of GO
267
within polymer matrix, thereby decreasing the slippage of fibers during testing process, in
268
comparison to PVdF-NFM. Besides the slippage, GO may cause higher load transfer efficiencies
269
at interface, thus possibly result in higher tensile stress. In addition, tensile strain of PVdF-GO
270
2.5% and PVdF-GO 5.0% NFMs at break (33.10 and 40.10%) was also higher than PVdF NFM
271
(30.90%) (Fig. S1). The tensile stress and strain of NFMs increased with GO ratio, owing to the
272
transformation of fibrous structure. Nonetheless, the NFMs exhibited enhanced rigidity along
273
with stronger ductile properties following GO addition.
14 ACS Paragon Plus Environment
Page 14 of 34
Page 15 of 34
274
Environmental Science & Technology
3.3. Quantitative Analysis of Porous Structure
275
The effective BET-SA and PSD of NFMs were systematically analyzed by N2 gas
276
adsorption-desorption, to probe molecular affinity and selectivity. The corresponding N2 gas
277
adsorption-desorption isotherms were typically type I (Fig. 4a), showing rapid condensation and
278
monolayer adsorption of N2 molecules inside porous NFMs. Generally, N2 adsorption was
279
increased following relative pressure (P/P0) approaching < 0.9, and central region of P/P0
280
indicated the existence of micropores in NFMs. The measured SA and PV of PVdF NFM was
281
12.43m2 g-1 and 0.03cc g-1, with maximum adsorption as high as 30.16cc g-1 STP while pressure
282
approaching saturation (P/P0 = 0.99), respectively (Table 1). The PVdF-GO 2.5% NFM had
283
almost similar SA as that of PVdF NFM, i.e., 13.09m2 g-1 with 0.046cc g-1 PV (Table 1). In
284
addition, PVdF-GO 5.0% NFM exhibited higher SA (51.47m2 g-1) and PV (0.077cc g-1) with
285
gradual increase in GO content with maximum adsorption for N2 as 62.94cc g-1 STP (Fig. 4a and
286
Table 1).
287
The geometric complexity and hierarchical textures of NFMs could be further analyzed by
288
fractal dimension (D) analysis following Frenkel-Halsey-Hill (FHH) theory based on N2
289
adsorption data. The plots of ln(V/Vmono) against ln(ln(P/P0)) were reconstructed from N2
290
isotherms for each NFM (Fig. 4b). The corresponding FHH plots had the same coverage region
291
for PVdF and PVdF-GO 2.5% NFMs, while a higher coverage region for PVdF-GO 5.0% was
292
observed comparatively. The calculated D was in order of PVdF (2.434), PVdF-GO 2.5% (2.435)
293
and PVdF-GO 5.0% (2.377), respectively, which further confirms the structural difference of
294
NFMs mainly composed of meso and micropores. The distinct PSD for PVdF-GO 5.0% NFM is
295
in good agreement with its higher SA and PV as compared to PVdF and PVdF-GO 2.5% NFM
296
(Fig. 4c and Table 1). Notably, most of the mesopores were observed within the range of
15 ACS Paragon Plus Environment
Environmental Science & Technology
297
1−10nm, with narrow size distribution and PVdF-GO 5.0% had much different mesopores.
298
Hence, BET SA and PSD reveals the significance of GO, thereby increasing the N2 gas
299
penetration, with 3D hierarchical architectures providing active sites, and may serve as transport
300
path to accelerate mass diffusion.
Page 16 of 34
Figure 4. (a) N2 adsorption–desorption isotherms of NFMs. (b) Plots of ln(V/Vmono) against ln(ln(P/P0)) reconstructed from N2 isotherms. (c) NLDFT PSD curves of relevant NFMs. 301 302
3.4. Surface Composition and Properties of NFMs
303
The surface charge properties of NFMs were examined from streaming potential and current
304
measurements over a pH range of 3 to 11 (Fig. 5a). The PVdF NFM was found to be less
305
negative in comparison to PVdF-GO NFMs. The GO functionalized PVdF NFMs possessed
306
significantly different surface charge characteristics, and became more negative owing to the
307
increase in density of oxygen containing functional moieties. The overall zeta (ζ) potential values
308
progressively became less negative due to the deprotonation of surface functional groups (Fig.
309
5a). The ζ potential behavior is in accordance with surface functional groups present in both GO
310
and PVdF. The pI of PVdF NFM was approximately 3.01 as resulted from the adsorption of
311
hydroxide ions originating from the self-ionization of water (Table 1). The pI for PVdF-GO 2.5% 16 ACS Paragon Plus Environment
Page 17 of 34
Environmental Science & Technology
312
and PVdF-GO 5.0% NFMs were found to be 3.63 and 3.68, following GO ratio (Table 1). The
313
surface charge of NFMs in presence of GO significantly changed, thereby indicating the
314
dominance of GO on NFM surface. Furthermore, the ζ potential values indirectly confirms the
315
presence of GO on NFM surface, thereby regulating the surface properties.
316
Figure 5. (a) Plots of ζ potential versus pH values of NFMs. (b) Plots of contact angle for different solvents. (c) ATR-FTIR spectra of GO and NFMs. 317 318
The surface wettability of NFMs were evaluated by SFE based on five probe liquids namely
319
water, formamide, diiodomethane, ethylene glycol and glycerol (Fig. 5b). The surface wettability
320
based on CA hysteresis for PVdF-GO NFMs were expected to vary as compared to PVdF NFM
321
due to the presence of hydrophilic functional groups in GO. The PVdF NFM exhibited
322
hydrophobic nature (CA > 90°) and had the highest average water CA of 116.5°(Table 1). PVdF
323
does not have significant polar groups; therefore it tends to be hydrophobic. The average water
324
CA of NFMs was decreased in presence of GO (72.5°and 52°, Table 1) and thus considered to
325
be hydrophilic. This trend is consistent with the presence of surface functional moieties such as –
326
OH and –CH on GO, as observed in ATR-FTIR spectra (Fig. 5c). Hence, NFMs were shifted to
17 ACS Paragon Plus Environment
Environmental Science & Technology
327
hydrophilic surface following GO addition. In order to further observe the dynamic wettability of
328
NFMs, organic solvents with different polar group were used. It is noteworthy; that PVdF-NFM
329
was also non-wettable to glycerol, because CA was found to be higher than 90°(Table S2). In
330
addition, the NFMs with GO were wettable to all tested polar solvents (Table S2), suggesting
331
some polar groups were present on NFMs surface. The NFMs were less wettable to glycerol in
332
comparison to formamide, diiodomethane and ethylene glycol (Table S2). The surface wetting
333
behavior was further examined by solid-liquid interaction force depending on both polar and
334
dispersive parts. The PVdF NFM had the lowest SFE 51.50 mJ m-2, and SFE was increased with
335
GO following ratio up to 57.36 and 88.50mJ m-2, for PVdF-GO 2.5% and PVdF-GO 5.0% NFMs
336
(Table 1). The SFE of NFMs is consistent with work of adhesion (Table S3) for solid-liquid
337
interface. The higher SFE is attributed to increased surface polarity, thereby offering more polar
338
interactions for NFMs. The pure water flux of PVdF NFM was 193±6.95 LMH, which is
339
according to its hydrophobic nature (Table 1). The pure water flux of NFMs containing GO was
340
increased, i.e., 439±10.23 and 663±54 LMH for PVdF-GO 2.5% and PVdF-GO 5.0%. The
341
higher flux could be explained by the hydrophilic nature of PVdF-GO NFMs.
342
The changes in surface composition of NFMs were further analyzed by ATR-FTIR spectra
343
(Fig. 5c). For GO, several bands at 3371cm−1 for aliphatic –OH stretching, at 2850cm-1 for –CH
344
stretching, at 1719cm-1 and 1044cm-1 for C=C, C=O stretching were observed. The
345
representative peak in NFMs at 1400 cm-1 corresponds to –CH2 scissoring, wagging and
346
vibrating of vinyldine. The observed peak intensities at 1275, 1181, 1070, 883 and 840cm-1
347
represents –CF2 stretching, symmetrical stretching of –CF, bending of –CH bond, –CH2 wagging
348
of vinyldine and –CH2 rocking. The peak at 1070cm-1 became progressive due to crosslinking of
349
carboxylic acid functional group (–COOH) and asymmetric stretch of carboxylate (–COO–)
18 ACS Paragon Plus Environment
Page 18 of 34
Page 19 of 34
Environmental Science & Technology
350
originated from GO. In addition, the band at 3300cm-1 was assigned to aliphatic –OH stretching,
351
corroboratively from GO.
352 353
3.5. Adsorption Kinetics of NFMs
354
The as-prepared NFMs had porous structure, large SA, enhanced mechanical strength with
355
prosperous flexibility as discussed above, thus could ensure promising results during molecular
356
filtration and selectivity process. Pseudo-first order and Pseudo-second order kinetic models
357
were used to investigate molecular adsorption and mass transfer on NFMs for selected organic
358
molecules. The NFMs were negatively charged and became more negative with GO (Fig. 5a),
359
because of oxygen containing functional moieties on surface of fiber along with pore orifice.
360
Therefore, we choose positively charged (MB, NR, BF and RB) and negatively charged (MO) as
361
representative organic dyes for selective adsorption and filtration experiments. The properties
362
and 3D molecular structure of organic dyes were presented in Table S1 and Figure S2,
363
respectively. The corresponding parameters and related fitting of kinetic models are listed in
364
Table S4 and Figure S3. The pseudo-second order model had a better fit to kinetics with higher
365
adsorption capacity as suggested by their correlation coefficients (r2) than pseudo-first order
366
model. The adsorption rate for positively charged molecules (MB, NR, BF and RB) increased
367
rapidly during first 24h and reached adsorption equilibrium in approximately 24h (Fig. S3a, S3b
368
and S3c), indicating physisorption process on heterogeneous fiber surface. In addition, GO
369
nanosheets provided additional active sites and contributed to adsorption uptake due to
370
compositional and structural difference in PVdF and PVdF-GO NFMs. These results are clearly
371
attributed to electrostatic attraction in between positively charged molecules and fiber matrix due
372
to improved ζ potential. The experimental adsorption capacity (Qe) followed the order of MB
19 ACS Paragon Plus Environment
Environmental Science & Technology
373
(96.48) > NR (43.74) > BF (31.04) > RB (30.39), with maximum capacity for PVdF-GO 5.0%
374
NFM (Table S4), respectively. In contrast, the NFMs had negligible adsorption capacity for
375
negatively charged MO molecule i.e., 1.18, 1.33 and 1.56 mg g-1, respectively, for PVdF, PVdF-
376
GO 2.5% and PVdF-GO 5.0% (Table S4 and Fig. S3d), because of electrostatic repulsion.
377
Therefore, it is believed that the electrostatic interaction was the dominant force in between
378
positive charged organic molecules and negatively charged NFMs, since NFMs exhibited
379
negative as observed from ζ potentials.
380
To further evaluate the difference in adsorption capacities (Fig. 6a), three-dimensional (3D)
381
molecular volume was considered, which was in order of MB < NR < BF < RB.24, 42 The smaller
382
molecular volume may account in pore diffusion and will quickly adjust in nanopores to occupy
383
the adsorption sites, whereas, larger molecular volume would be sterically hindered for quick
384
adsorption, resulting in blockage of nanopores orifice, thus decreasing the adsorption uptake.1
385
Additionally, MB and NR have similar 3D molecular size with a difference in charged group,
386
which is smaller in NR, thus could result in decreased electrostatic attraction in between NFMs
387
and organic molecules.43 The negatively charged dye (MO), being the smallest one, could ingress
388
into the NFMs nanopores, thus resulted in smallest uptake.
389 390
3.6.Molecular Filtration Performance
391
The separation of organic molecules during dynamic purification process needs to be reliable
392
and predictable for practical applications featuring applied chemical purification based on
393
selectivity and molecular filtration. To check the practical performance and molecular selectivity
394
of PVdF-GO NFMs, we selected five different organic dyes with different electrostatic charge
395
and size to evaluate separation efficiency of NFMs as prototypical application.
20 ACS Paragon Plus Environment
Page 20 of 34
Page 21 of 34
Environmental Science & Technology
396
Filtrate Filtrate
Filtrate
Filtrate
Filtrate
Filtrate
Filtrate
Filtrate
Filtrate
Figure 6. (a) Adsorption capacity of PVdF and PVdF-GO NFMs for five representative watersoluble dyes. The UV−Vis spectra of (b) MB (positive) and MO (negative) filtrate, oppositely charged dye mixtures (c) MB/MO, (d) NR/MO, (e) BF/MO (f) RB/MO, same charge dye mixtures (g) MB/NR, (h) MB/BF and (i) MB/RB; before and after filtration process. The inset photographs show the colors of aqueous solutions before and after (arrow) filtration process, respectively. 397 398
First, molecular filtration was executed with simple aqueous dye solutions at an average
399
pressure of 1bar. The colored feed solutions of positively charged organic dyes (MB, NR, BF
400
and RB) were turned into pure water color effectively, when passed through PVdF-GO 2.5% and
401
PVdF-GO 5.0% NFMs (Fig. 6b and S4). The PVdF-GO successfully adsorbed the positively
402
charged organic dyes efficiently and resulted permeate had 100% purity. In contrast, there was 21 ACS Paragon Plus Environment
Environmental Science & Technology
403
negligible adsorption for negatively charged MO molecules with 100% permeation (Fig. 6b).
404
The UV-Vs spectra for all organic dyes are also consistent with efficient performance of NFMs;
405
as there was no peaks observed for positively charged dyes (Fig. 6b and S4); while same peak for
406
negatively charged feed solution (Fig. 6b) after permeation. In comparison to PVdF-GO NFMs,
407
PVdF NFM had obvious peaks for positively charged organic molecules, but concentration of
408
each dye came down. It means that PVdF NFM was not as efficient as PVdF-GO NFMs, and
409
thus permeated some of the positively charged organic molecules once saturated, but passed
410
negatively charged MO molecules thoroughly.
411
Second, the molecular filtration performance of as-prepared NFMs was further examined
412
with mixtures besides pure aqueous solution of various organic dyes. For mixed molecular
413
filtration, four mixtures based on opposite charge i.e., MB/MO, NR/MO, BF/MO and RB/MO
414
were filtered through each NFMs. The UV-Vis spectra of organic mixtures, before and after
415
filtration process are shown in Figures 6c, 6d, 6e and 6f. The colored feed solutions turned into
416
the original color of MO molecules after permeation and the color of NFMs changed to the color
417
of dye molecules being penetrated. This indicates that NFMs adsorbed positively charged
418
molecules quickly and allowed only MO molecules to pass through, highlighting the selective
419
affinity and separation from a mixture only for positively charged molecules. The average
420
concentration of MO after filtration through NFMs was approximately 98.37%, comparatively to
421
initial MO feed solution with simultaneous adsorption of positively charged molecules. The
422
concentration of positively charged molecules decreased absolutely from 40 mg L-1 to 0.16 (MB)
423
< 0.31 (NR) < 0.37 (BF) < 0.48 (RB) mg L-1, after permeation through PVdF-GO 5.0% NFMs,
424
because of prominent adsorption capacities during filtration process. The average filtration
425
efficiency for four mixed solutions (MB/MO, NR/MO, BF/MO and RB/MO) through PVdF-GO
22 ACS Paragon Plus Environment
Page 22 of 34
Page 23 of 34
Environmental Science & Technology
426
5.0% NFM was 99.20%, 98.47%, 98.17% and 97.63%, respectively. The PVdF-GO 2.5% and
427
PVdF NFMs also captured positively charged molecules and residual concentration in permeate
428
was 3.44 (MB) < 3.47 (NR) < 3.61 (BF) < 3.72 (RB) mg L-1 and 5.52 (MB) < 5.72 (NR) < 5.94
429
(BF) < 6.21 (RB) mg L-1, respectively. The filtrates of PVdF-GO 5.0% and PVdF-GO 2.5%
430
NFMs had no sign of absorption peaks for MB, NR, BF and RB molecules but only the
431
representative peak of MO molecule in UV-Vis spectra for MB/MO, NR/MO, BF/MO and
432
RB/MO mixtures, whereas obvious peaks when passed through PVdF NFMs for both organic
433
molecules (Fig. 6c, 6d, 6e and 6f). The filtration efficiency of PVdF-GO 2.5% NFMs for mixed
434
solutions was 85.30%, 85.19%, 84.66% and 84.29% respectively and PVdF NFMs had the lower
435
filtration efficiency as 78.28%, 77.67% 76.96% and 76.15%, respectively, which suggests that
436
NFMs reached saturation capacity and no further adsorption was possible. Therefore, it is
437
reasonable to speculate that GO was actively involved to sorb positively charged molecules
438
being at PVdF fiber surface and thus exhibited better performance during filtration process, in
439
comparison to PVdF NFM. Filtration curves of dye solutions and their mixtures through NFMs
440
at 1 bar pressure are displayed in Fig. S5, respectively. The PVdF-GO NFM had lower
441
permeability for dye solutions and their mixtures (Fig. S5a) than that of PVdF-GO 2.5% and
442
PVdF-GO 5.0% NFMs (Fig. S5b and Fig. S5c), reasonably due to hydrophobic nature.
443
Finally, the molecular filtration was also performed by mixed organic molecules with same
444
electrostatic charge i.e., NR/MB, BF/MB and RB/MB. The NFMs exhibited almost 100%
445
efficiency during filtration process of NR/MB, BF/MB and RB/MB mixtures. The concentration
446
of MB was found to be slightly lower than NR, BF and RB for each NFMs filtrate following
447
PVdF-GO 5.0% < PVdF-GO 2.5% < PVdF NFMs, confirming the higher affinity of NFMs
448
towards MB molecules. The residual concentration in filtrate was in order of MB < NR < BF
NR > BF > RB, following GO addition, respectively.
452
The corresponding UV-Vis absorption peaks of these three mixtures are shown in Figures 6g, 6h
453
and 6i, and the concentration of both positively charged organic dyes decreased. The filtration
454
processes of selected organic dyes by PVdF-GO NFMs indicate the acceptable filtration
455
efficiency and thus are of great potential for selective separation and purification for waste-water
456
treatment.
457
The proposed mechanism of filtration process for selected organic dyes is illustrated in
458
Figure 7. During molecular filtration, porous structure of NFMs, size exclusion and electrostatic
459
attraction-repulsion for charged organic dyes, simultaneously resulted for high efficient
460
performance.2,
461
linkers forming a continuous mask within PVdF network, thus acted as potential active sites for
462
interaction. The GO nanosheets also enabled pores affinity for charge selectivity within NFMs.
463
Presumably; the positively charged organic molecules were adsorbed quickly by electrostatic
464
forces during penetration within pores, whereas negatively charged organic molecules are
465
expectable to permeate through NFMs. In addition, the specific pore size of NFMs could
466
essentially stop or decelerate the permeation process of large size organic molecules gradually,
467
highlighting the size exclusion evident by hindered theory. Moreover, the stable pores of NFMs
468
consequently result efficient mass transport within inter-fiber network, empowering a powerful
469
permeation. Therefore, the open fibrous architecture and controlled PSD of NFMs regulated the
470
filtration performance and promoted high permeation.
13, 20
GO nanosheets were preserved well on surface of nanofibers like cross-
24 ACS Paragon Plus Environment
Page 24 of 34
Page 25 of 34
Environmental Science & Technology
Figure 7. (a) Preparation of PVdF-GO NFMs. (b) Schematic illustration for distribution of GO nanosheets on PVdF nanofiber and proposed filtration process (before and after) towards the mixture of organic dyes by NFMs. 471 472
3.7. Regeneration of NFMs
473
The potential stability and reusability for molecular filtration process is essential for practical
474
application with long-term durability, whereas continuous recyclability and proper regeneration
475
remains challenging, thus are inevitable issues. The entire process of reusability and regeneration
476
can provide an insight into structural stability and long term reusability of as-prepared PVdF-GO
477
NFMs. The regeneration and recyclability performance of PVdF-GO 5.0% NFM was
478
systematically observed by ATR-FTIR and BET-N2 adsorption-desorption, before and after
479
filtration (Fig. S7a and S7b). ATR-FTIR spectra (Fig. S7a) of NFMs after permeation of MB and 25 ACS Paragon Plus Environment
Environmental Science & Technology
480
NR at ~1400 and 1334 cm-1, were attributed to electrostatic attraction, and then completely
481
retrieved original spectra after regeneration by the eluent, confirming the total removal of MB
482
and NR molecules from the NFMs. Furthermore, the NFMs retains almost original filtration
483
efficiency after three consecutive regeneration cycles for further application, suggesting that the
484
NFM is effective and stable for further adsorption of MB or NR molecules from aqueous media.
485
In addition, the BET-N2 adsorption-desorption after regeneration cycles were almost the same as
486
original (Fig. S7b), suggesting the stable fibrous morphology and excellent reversibility of NFMs.
487
The inset of Fig. S7b show the permeate concentration of MB and NR from mixed solutions up
488
to three regeneration cycles, and the observed separation efficiency was almost the same as
489
original, highlighting the effective regeneration, stability and filtration performance. Table S5
490
lists the comparison of graphene-incorporated membranes for flux and dyes removal
491
performance from water. The separation efficiency with high permeation, simultaneous recycling,
492
while avoiding common problems associated with GO and common polymeric membranes are
493
important for industrial processing. The results achieved are desirable for large scale practical
494
applications, in industrial waste-water treatment plants. The stable performance throughout the
495
system indicates the effective role of GO nanosheets for environmental pollutant management,
496
thus have potential to be used for field specific applications.
497
GO can be used effectively for modification of membrane surface and structural geometry as
498
portrayed by electrospun PVdF-GO NFMs, thereby utilization the properties of GO, but
499
overwhelming the problems caused by functional moieties in aqueous-phase. The GO provided
500
prolonged durability and reusability for continuous regeneration cycles, which were hindered by
501
routine polymeric membranes for aqueous mixtures and are hard to recycle. Considering the
502
excellent chemical stability of GO, the reported membrane can be used for filtration of complex
26 ACS Paragon Plus Environment
Page 26 of 34
Page 27 of 34
Environmental Science & Technology
503
organic mixtures with desirable separation and recycling, accompanied by high permeation, with
504
textile and pharmaceutical industries being the potential beneficiaries. The simple fabrication,
505
mimetic structure, unique properties and existence of GO, will enable more fundamental research
506
and real applications for selective filtration and recycling of environmental pollutants.
507 508
Appendix A. Supplementary data
509
The synthesis of GO and data analysis equations are detailed in SI. The physico-chemical
510
properties of selected organic dyes, average CA, work of adhesion, adsorption kinetics
511
parameters, comparison of graphene-incorporated membranes are listed in Tables S1 to S5,
512
respectively. The stress-strain curves of NFMs, 3D molecular structures of representative dyes,
513
adsorption kinetics graphs, UV-vis spectra (NR, BF and RB), filtration curves, removal
514
efficiency, ATR-FTIR and N2 isotherms of PVdF-GO 5.0% NFM before and after regeneration,
515
are presented in Figures S1 to S7, respectively. The supporting information is available free of
516
charge on the ACS Publications website.
517 518
Author Information
519
Corresponding Author
520
*Phone: 0086-571-88982587; fax: 0086-571-88982587; e-mail:
[email protected].
521
Author disclosure statement
522
The authors declare that no competing financial conflicts exist.
523 524
Acknowledgments
27 ACS Paragon Plus Environment
Environmental Science & Technology
525
This project was supported by the National Natural Science Foundation of China (21425730,
526
21537005, 21621005, and 21607124), the National Key Technology Support Program of China
527
(2015BAC02B01), and National Basic Research Program of China (2014CB441106). Abdul
528
Ghaffar also acknowledges the financial support for doctoral study from Zhejiang University.
529 530
References
531
(1) Ghaffar, A.; Zhu, X.; Chen, B. Structural characteristics of biochar-graphene nanosheet
532
composites and their adsorption performance for phthalic acid esters. Chem. Eng. J. 2017,
533
319, 9–20.
534
(2) Joshi, R.K.; Carbone, P.; Wang, F.C.; Kravets, V.G.; Su, Y.; Grigorieva, I.V.; Wu, H.A.;
535
Geim, A.K.; Nair, R.R. Precise and ultrafast molecular sieving through graphene oxide
536
membranes. Science 2014, 343 (6172), 752–754.
537 538
(3) Kim, J.E.; Han, T.H.; Lee, S.H.; Kim, J.Y.; Ahn, C.W.; Yun, J.M.; Kim, S.O. Graphene oxide liquid crystals. Angew. Chem. Int. Ed. 2011, 50 (13), 3043–3047.
539
(4) Bunch, J.S.; Verbridge, S.S.; Alden, J.S.; van der Zande, A.M.; Parpia, J.M.; Craighead, H.G.;
540
McEuen, P.L. Impermeable atomic membranes from graphene sheets. Nano Lett. 2008, 8
541
(8), 2458–2462.
542
(5) Celebi, K.; Buchheim, J.; Wyss, R.M.; Droudian, A.; Gasser, P.; Shorubalko, I.; Kye, J.I.;
543
Lee, C.; Park, H.G. Ultimate permeation across atomically thin porous graphene. Science
544
2014, 344 (6181), 289–292.
545 546
(6) Cohen-Tanugi, D.; Grossman, J.C. Water desalination across nanoporous graphene. Nano Lett. 2012, 12 (7), 3602–3608.
28 ACS Paragon Plus Environment
Page 28 of 34
Page 29 of 34
Environmental Science & Technology
547
(7) Han, J.L.; Xia, X.; Tao, Y.; Yun, H.; Hou, Y.N.; Zhao, C.W.; Luo, Q.; Cheng, H.Y.; Wang,
548
A.J. Shielding membrane surface carboxyl groups by covalent-binding graphene oxide to
549
improve anti-fouling property and the simultaneous promotion of flux. Water Res. 2016,
550
102, 619–628.
551
(8) Huang, K.; Liu, G.; Lou, Y.; Dong, Z.; Shen, J.; Jin, W. A graphene oxide membrane with
552
highly selective molecular separation of aqueous organic solution. Angew. Chem. Int. Ed.
553
2014, 53 (27), 6929–6932.
554
(9) Pastrana-Martinez, L.M.; Morales-Torres, S.; Figueiredo, J.L.; Faria, J.L.; Silva, A.M.
555
Graphene oxide based ultrafiltration membranes for photocatalytic degradation of organic
556
pollutants in salty water. Water Res. 2015, 77, 179–190.
557
(10) Crock, C.A.; Rogensues, A.R.; Shan, W.; Tarabara, V.V. Polymer nanocomposites with
558
graphene-based hierarchical fillers as materials for multifunctional water treatment
559
membranes. Water Res. 2013, 47 (12), 3984–3996.
560 561
(11) Mi, B. Graphene oxide membranes for ionic and molecular sieving. Science 2014, 343 (6172), 740–742.
562
(12) Qiu, L.; Zhang, X.; Yang, W.; Wang, Y.; Simon, G.P.; Li, D. Controllable corrugation of
563
chemically converted graphene sheets in water and potential application for nanofiltration.
564
Chem. Commun. 2011, 47 (20), 5810–5812.
565 566
(13) Han, Y.; Xu, Z.; Gao, C. Ultrathin graphene nanofiltration membrane for water purification. Adv. Funct. Mater. 2013, 23 (29), 3693–3700.
567
(14) Senyuk, B.; Behabtu, N.; Martinez, A.; Lee, T.; Tsentalovich, D.E.; Ceriotti, G.; Tour, J.M.;
568
Pasquali, M.; Smalyukh, II. Three-dimensional patterning of solid microstructures through
29 ACS Paragon Plus Environment
Environmental Science & Technology
569
laser reduction of colloidal graphene oxide in liquid-crystalline dispersions. Nat. Commun.
570
2015, 6, 7157.
571 572
(15) 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.
573
(16) Huang, H.; Song, Z.; Wei, N.; Shi, L.; Mao, Y.; Ying, Y.; Sun, L.; Xu, Z.; Peng, X.
574
Ultrafast viscous water flow through nanostrand-channelled graphene oxide membranes.
575
Nat. Commun. 2013, 4, 2979.
576
(17) Lin, X.; Yang, Q.; Ding, L.; Su, B. Ultrathin silica membranes with highly ordered and
577
perpendicular nanochannels for precise and fast molecular separation. ACS Nano 2015, 9
578
(11), 11266–11277.
579
(18) Marchetti, P.; Jimenez Solomon, M.F.; Szekely, G.; Livingston, A.G. Molecular separation
580
with organic solvent nanofiltration: a critical review. Chem. Rev. 2014, 114 (21), 10735–
581
10806.
582
(19) Prince, J.A.; Bhuvana, S.; Anbharasi, V.; Ayyanar, N.; Boodhoo, K.V.K.; Singh, G. Ultra-
583
wetting graphene-based PES ultrafiltration membrane - A novel approach for successful oil-
584
water separation. Water Res. 2016, 103, 311–318.
585
(20) Striemer, C.C.; Gaborski, T.R.; McGrath, J.L.; Fauchet, P.M. Charge- and size-based
586
separation of macromolecules using ultrathin silicon membranes. Nature 2007, 445, 749–
587
753.
588
(21) Chen, P.; Liang, H.W.; Lv, X.H.; Zhu, H.Z.; Yao, H.B.; Yu, S.H. Carbonaceous nanofiber
589
membrane functionalized by beta-cyclodextrins for molecular filtration. ACS Nano 2011, 5
590
(7), 5928–5935.
30 ACS Paragon Plus Environment
Page 30 of 34
Page 31 of 34
591 592
Environmental Science & Technology
(22) Pendergast, M.M.; Hoek, E.M.V. A review of water treatment membrane nanotechnologies. Energy Environ. Sci. 2011, 4 (6), 1946–1971.
593
(23) Soyekwo, F.; Zhang, Q.; Gao, R.; Qu, Y.; Lin, C.; Huang, X.; Zhu, A.; Liu, Q. Cellulose
594
nanofiber intermediary to fabricate highly-permeable ultrathin nanofiltration membranes for
595
fast water purification. J. Membr. Sci. 2017, 524, 174–185.
596
(24) Wen, Q.; Di, J.; Zhao, Y.; Wang, Y.; Jiang, L.; Yu, J. Flexible inorganic nanofibrous
597
membranes with hierarchical porosity for efficient water purification. Chem. Sci. 2013, 4
598
(12), 4378–4382.
599 600
(25) Vandezande, P.; Gevers, L.E.; Vankelecom, I.F. Solvent resistant nanofiltration: separating on a molecular level. Chem. Soc. Rev. 2008, 37 (2), 365–405.
601
(26) Deng, H.; Li, X.; Ding, B.; Du, Y.; Li, G.; Yang, J.; Hu, X. Fabrication of polymer/layered
602
silicate intercalated nanofibrous mats and their bacterial inhibition activity. Carbohydr.
603
Polym. 2011, 83 (2), 973–978.
604 605
(27) Li, D.; Xia, Y. Electrospinning of nanofibers: Reinventing the wheel? Adv. Mater. 2004, 16 (14), 1151–1170.
606
(28) Obaid, M.; Mohamed, H.O.; Yasin, A.S.; Yassin, M.A.; Fadali, O.A.; Kim, H.; Barakat,
607
N.A.M. 2017. Under-oil superhydrophilic wetted PVdF electrospun modified membrane for
608
continuous gravitational oil/water separation with outstanding flux. Water Res. 2017, 123,
609
524–535.
610
(29) Wang, X.; Ding, B.; Yu, J.; Si, Y.; Yang, S.; Sun, G. Electro-netting: fabrication of two-
611
dimensional nano-nets for highly sensitive trimethylamine sensing. Nanoscale 2011, 3 (3),
612
911–915.
31 ACS Paragon Plus Environment
Environmental Science & Technology
613
(30) Bae, J.; Baek, I.; Choi, H. Mechanically enhanced PES electrospun nanofiber membranes
614
(ENMs) for microfiltration: The effects of ENM properties on membrane performance.
615
Water Res. 2016, 105, 406–412.
616
(31) Choi, S.H.; Ankonina, G.; Youn, D.Y.; Oh, S.G.; Hong, J.M.; Rothschild, A.; Kim, I.D.
617
Hollow ZnO nanofibers fabricated using electrospun polymer templates and their electronic
618
transport properties. ACS Nano 2009, 3 (9), 2623–2631.
619
(32) Ghaffar, A.; Zhu, X.; Chen, B. Biochar composite membrane for high performance pollutant
620
management: Fabrication, structural characteristics and synergistic mechanisms. Environ.
621
Pollut. 2018, 233, 1013–1023.
622
(33) Tai, M.H.; Gao, P.; Tan, B.Y.; Sun, D.D.; Leckie, J.O. Highly efficient and flexible
623
electrospun carbon-silica nanofibrous membrane for ultrafast gravity-driven oil-water
624
separation. ACS Appl. Mater. Interfaces 2014, 6 (12), 9393–9401.
625
(34) Uyar, T.; Havelund, R.; Nur, Y.; Hacaloglu, J.; Besenbacher, F.; Kingshott, P. Molecular
626
filters based on cyclodextrin functionalized electrospun fibers. J. Membr. Sci. 2009, 332 (1-
627
2), 129–137.
628
(35) Zhu, J.; Sun, G. Facile fabrication of hydrophilic nanofibrous membranes with an
629
immobilized metal-chelate affinity complex for selective protein separation. ACS Appl.
630
Mater. Interfaces 2014, 6 (2), 925–932.
631
(36) Huang, L.W.; Arena, J.T.; Manickam, S.S.; Jiang, X.Q.; Willis, B.G.; McCutcheon, J.R.
632
Improved mechanical properties and hydrophilicity of electrospun nanofiber membranes for
633
filtration applications by dopamine modification. J. Membr. Sci. 2014, 460, 241–249.
634 635
(37) Kanehata, M.; Ding, B.; Shiratori, S. Nanoporous ultra-high specific surface inorganic fibres. Nanotechnology 2007, 18 (31), 315602.
32 ACS Paragon Plus Environment
Page 32 of 34
Page 33 of 34
636 637 638 639
Environmental Science & Technology
(38) Hummers, Jr., W.S.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80 (6), 1339–1339. (39) Wang, J.; Chen, Z.; Chen, B. Adsorption of polycyclic aromatic hydrocarbons by graphene and graphene oxide nanosheets. Environ. Sci. Technol. 2014, 48 (9), 4817–4825.
640
(40) Fu, G.; Su, Z.; Jiang, X.; Yin, J. Photo-crosslinked nanofibers of poly(ether amine) (PEA)
641
for the ultrafast separation of dyes through molecular filtration. Polym. Chem. 2014, 5 (6),
642
2027–2034.
643
(41) Li, Y.; Zhu, Z.; Yu, J.; Ding, B. Carbon nanotubes enhanced fluorinated polyurethane
644
macroporous membranes for waterproof and breathable application. ACS Appl. Mater.
645
Interfaces 2015, 7 (24), 13538–13546.
646
(42) Zhuang, X.; Wan, Y.; Feng, C.M.; Shen, Y.; Zhao, D. Highly efficient adsorption of bulky
647
dye molecules in wastewater on ordered mesoporous carbons. Chem. Mater. 2009, 21 (4),
648
706–716.
649
(43) Yan, A.X.; Yao, S.; Li, Y.G.; Zhang, Z.M.; Lu, Y.; Chen, W.L.; Wang, E.B. Incorporating
650
polyoxometalates into a porous MOF greatly improves its selective adsorption of cationic
651
dyes. Chem. Eur. J. 2014, 20 (23), 6927–6933.
652 653 654 655 656 657 658
33 ACS Paragon Plus Environment
Environmental Science & Technology
659
TOC
660 661
662
34 ACS Paragon Plus Environment
Page 34 of 34