You are on page 1of 10

Environmental Pollution 251 (2019) 344e353

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Nitrogen and sulfur codoped graphene aerogels as absorbents and


visible light-active photocatalysts for environmental remediation
applications+
Yiqun Jiang a, 1, Shamik Chowdhury b, 1, Rajasekhar Balasubramanian a, *
a
Department of Civil & Environmental Engineering, National University of Singapore, 1 Engineering Drive 2, Singapore 117576, Singapore
b
School of Environmental Science and Engineering, Indian Institute of Technology Kharagpur, West Bengal 721302, India

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

Article history: Graphene aerogels (GAs) are increasingly being recognized as high performance multifunctional mate-
Received 1 March 2019 rials to tackle our current and emerging environmental concerns. In order to extend the application
Received in revised form potential of GAs, herein we have successfully synthesized nitrogen (N) and sulfur (S) codoped GAs
28 April 2019
(NSGAs) via a simple, scalable, and inexpensive approach. Owing to their large specific surface area (up to
Accepted 29 April 2019
Available online 30 April 2019
132 m2 g1), profound porosity, superior mechanical properties, and coexistence of N and S atoms with
tunable atomic content and bonding configurations, the as-prepared NSGAs demonstrated exceptional
absorption capacity toward a broad spectrum of oils and organic solvents, with an average absorption
Keywords:
Graphene aerogels
rate many folds higher than conventional absorbents. Further, the NSGAs exhibited excellent photo-
Nitrogen and sulfur codoping catalytic activity for the decomposition of recalcitrant organic compounds under visible light illumina-
Oil absorption tion due to pronounced synergistic coupling effect between the heteroatoms. Specifically, after 5 h of
Photocatalysis exposure to visible light, a degradation efficiency of over 99% was observed and more than 84% of the
Organic pollutants total organic carbon was eliminated. Radical trapping experiments revealed that superoxide anion
Visible light radicals are the predominant oxygen reactive species driving the photocatalytic reactions. More
importantly, the mineralization byproducts did not pose any significant antibacterial activity, illustrating
the environmentally benign nature of these macroscale photocatalysts.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction GAs combined with their versatile surface chemistry, fast electron
conduction, and rapid mass transport enables efficient removal of a
In recent years, three-dimensional (3D) graphene-based mac- broad spectrum of waterborne pollutants such as synthetic dyes,
rostructures (GBMs), such as sponges, foams, hydrogels, aerogels potentially toxic elements, pesticides, herbicides, organic solvents,
and xerogels, have been the subject of intense research and salts, and pathogens as well as greenhouse gases such as CO2
development (R&D) efforts to address our pressing environmental (Chowdhury et al., 2019; Wu et al., 2019; Yousefi et al., 2019).
issues in a sustainable manner (Shen et al., 2014). The structural In order to enhance the environmental performance of GAs,
integrity, mechanical robustness, and interconnected porosity of 3D doping heteroatoms such as nitrogen (N) into their nanoscale
GBMs make them extremely suitable for a wide range of environ- building blocks has attracted enormous research attention in recent
mental remediation applications (Yousefi et al., 2019). Freestanding years (Fang et al., 2015). The incorporation of N leads to polarization
graphene aerogels (GAs), in particular, hold great promise for a in the sp2 conjugated network and creates a band gap around the
range of water treatment and air purification applications (Maleki, Dirac point by repressing the neighboring density of states (Wang
2016; Gorgolis and Galiotis, 2017). The large specific surface area of et al., 2012a, 2014). This local electronic rearrangement in turn
improves the surface reactivity, electrical conductivity, and optical
absorption properties of GAs for a wide array of promising envi-
ronmental cleanup applications (Zhao et al., 2012; Sui et al., 2015;
+
This paper has been recommended for acceptance by Eddy Y. Zeng. Du et al., 2016; Jiang et al., 2017, 2019). For example, N-doped GAs
* Corresponding author.
(NGAs) have shown remarkable performance as absorbents for a
E-mail address: ceerbala@nus.edu.sg (R. Balasubramanian).
1
These authors contributed equally to this work. multitude of organic compounds (Yang et al., 2016), with a

https://doi.org/10.1016/j.envpol.2019.04.132
0269-7491/© 2019 Elsevier Ltd. All rights reserved.
Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353 345

maximum absorption uptake of up to 210 g g1 (Rahmani et al., repeatedly with copious amounts of ultrapure water to remove all
2018), which is over four times greater than that of pristine GA leftover chemicals, and ultimately freeze dried at 50  C for 24 h to
(Wang et al., 2012b). Moreover, it has recently been reported that obtain the aerogel. By varying the mass of NH4SCN in 80 mg GO, a
the dye removal efficiency of NGAs prepared using a hydrothermal series of NSGAs were prepared. The aerogels with different GO to
method is enhanced by about 68% over that of undoped GA (Jiang NH4SCN mass ratio of 1:0.5, 1:0.33, 1:0.25 and 1:0.20 were labeled
et al., 2017). Very recently, the impressive photocatalytic activity as NSGA-I, NSGA-II, NSGA-III and NSGA-IV, respectively. For com-
of NGAs has been demonstrated for degradation of recalcitrant parison, pristine graphene aerogel (GA) (without adding any
waterborne pollutants (Jiang et al., 2019; Wang et al., 2019). Usu- doping agent) and N-doped graphene aerogel (NGA) (using 8 mL of
ally, NGAs have been particularly designed and fabricated for spe- ammonium hydroxide (28e30% ammonia basis; Sigma-Aldrich) as
cific environmental remediation applications with different N the N elemental precursor) was also prepared employing the same
atomic concentrations and molecular configurations. In order to procedure.
advance their practical applications, the fabrication of multifunc-
tional NGAs is highly desirable (Kotal et al., 2016). 2.2. Instrumental characterization
In this regard, the incorporation of sulfur (S) together with N
into the hexagonal lattice of GAs could yield advanced and versatile A broad range of state-of-the-art microscopy (field emission
material systems with enhanced sorption capacities, desired pho- scanning electron microscopy, FESEM; transmission electron mi-
tocatalytic properties and novel functionalities because of the croscopy, TEM), spectroscopy (Raman spectroscopy; energy-
synergistic coupling effect between the heteroatoms (Su et al., dispersive X-ray spectroscopy, EDX; X-ray photoelectron spectros-
2013; Huang et al., 2016; Saana Amiinu et al., 2016; Tian et al., copy, XPS; diffuse reflectance spectroscopy) and diffraction (X-ray
2016; Ma et al., 2017). Specifically, S doping would likely result in diffraction, XRD) techniques as well as classical surface-specific
a localized structural distortion due to the large atomic size of S methods, such as basic measurements of zeta potential, surface
compared with carbon (C) and the long SeC bond (Serp and area, average pore volume and pore size, were used to examine the
Machado, 2015), activating the adjacent sp2-C into coordinatively bulk and surface properties of the developed materials. For details
unsaturated sp3-C. Further, the difference between the outermost on these characterization procedures, please see the Supplemen-
orbital energies of S and C atoms would probably induce a tary information.
nonuniform spin density distribution (Wang et al., 2014), convert-
ing the S atom into a catalytic center (Chen et al., 2016). This feature 2.3. Environmental applications
of S-doped graphene together with the enhanced electron donation
capability of nearby C atoms due to N doping is thought to syner- In order to illustrate the multifunctional properties of the as-
gistically boost the surface reactivity of GAs for a myriad of envi- prepared monolithic materials, the NSGAs were evaluated for two
ronmental applications (Kicin  ski et al., 2014). However, to the best selected environmental remediation applications: (1) absorption of
of our knowledge, the effect of simultaneously introducing N and S oils (diesel oil, engine oil, pump oil, and soybean oil), petroleum
dopants on the environmental application of GAs remains relatively products (kerosene), and organic solvents (acetone, n-hexane,
unexplored. phenol, and toluene), and (2) visible light-induced photodecom-
To fill this knowledge gap, herein we have prepared N and S position of a recalcitrant dye pollutant (acridine orange). The
codoped GAs (NSGAs) via a simple, scalable, and ecofriendly detailed description of the corresponding experimental procedures
method. The resulting lightweight N/S dual-doped carbonaceous is provided in the Supplementary information.
aerogels exhibit hierarchically interconnected meso/macroporosity
with large specific surface area and superior mechanical properties. 3. Results and discussion
Such unique textural features, along with the coexistence of N and S
dopants with different types of bonding configurations, not only 3.1. Materials characterization
offer a large number of accessible sites for molecular binding, but
also provide abundant reactive regions for catalytic reactions. In this work, a facile one-pot strategy, involving hydrothermal
Attributing to these beneficial structural features, our versatile processing of a homogeneous aqueous mixture of GO and NH4SCN,
NSGAs demonstrate tremendous potential for two selected envi- was adopted to fabricate mechanically strong and lightweight GAs
ronmental remediation applications: (1) absorption of oils and doped with both N and S atoms (Fig. S1). Specifically, upon hy-
organic solvents, and (2) visible light-induced photodecomposition drothermal treatment at 180  C, NH4SCN decomposed into highly
of recalcitrant organic pollutants from aqueous phase. reactive nitrogen and sulfur compounds (such as ammonia,
hydrogen sulfide, carbon disulfide, etc.), which subsequently
2. Experimental reacted with the defective sites and oxygenic functionalities (such
as carboxyl, carbonyl, epoxy, etc.) of GO (Su et al., 2013). As such,
2.1. Synthesis of nitrogen/sulfur codoped graphene aerogels some of the C atoms in the graphitic lattice were replaced with
heteroatoms (N and S). Additionally, the GO nanosheets underwent
NSGAs were fabricated by a simple one-step hydrothermal reduction and eventually coalesced via supramolecular interactions
approach. In a typical procedure, a weighed amount of ammonium (such as pep stacking and hydrogen bonding), thus forming N/S
thiocyanate (NH4SCN, 99 wt%; Sigma-Aldrich) was added into codoped 3D GBMs (Fig. S2). The microstructure of the as-obtained
40 mL aqueous dispersion of graphene oxide (GO, 2 mg mL1 NSGAs was characterized via FESEM and TEM imaging (Fig. 1,
dispersion in water; Sigma-Aldrich) and continuously stirred for Figs. S3 and S4). Similar to the morphology observed for pristine
20 min at room temperature, followed by ultrasonic treatment at monolithic GA (Fig. S3a), the NSGAs possessed an extensive
30  C for 30 min in an ultrasonic bath operating at 37 kHz (Elma- network of circular and/or elliptical shaped pores (Fig. 1a and
sonic S 60 H, Elma Schmidbauer GmbH, Germany). The as-obtained Fig. S3b‒d), interconnected through wrinkled graphene nanosheets
homogenous mixture was then sealed in a stainless steel Teflon- (Fig. 1b and Fig. S4). Moreover, the vast majority of the pores were
coated autoclave (100 mL) and treated hydrothermally at 180  C in the meso/macropore range, as inferred from the corresponding
for 12 h. The autoclave was then allowed to cool naturally to BJH pore size distribution plots (Fig. S5), with the average pore
ambient temperature. The resultant hydrogel was rinsed diameter spanning from 3.78 to 11.56 nm. The surface area and the
346 Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353

Fig. 1. (a) FESEM and (b) TEM image of NSGA-IV. Scale bars in (a) and (b) represent 20 mm and 200 nm, respectively. (c) Wide-angle XRD profile of the as-synthesized GA, NGA and
NSGAs. The interlayer distance (d) between the 2D graphene building blocks in the aerogels, calculated using the Bragg's equation (Xing et al., 2016; Liu et al., 2017), is listed in the
inset. (d) Raman spectroscopic signatures of the as-prepared GA, NGA and NSGAs.

total pore volume of the NSGAs are listed in Table S1. Interestingly, boundaries/edges (May et al., 2013). In comparison with undoped
there was a dramatic decline in the surface area and pore volume of and single doped samples prepared under similar conditions, the
the bulk materials at higher NH4SCN concentrations which could be peak positions of G band were downshifted by 6e13 cm1, probably
attributed to greater doping-induced defects. The defects doubled because of the tensile (stretching) strain on the CeC bonds
up as interlocking sites for the flexible graphene sheets during the resulting from the incorporation of heteroatoms into the sp2 con-
self-assembling process to form a porous network with higher jugated backbone of our as-developed N/S dual-doped GAs. This is
interconnectivity. conceivable since the bond lengths of CeC (1.42 Å) and CeN (1.41 Å)
Next, the effect of incorporating N and S dopants into the hon- are quite similar, whereas the CeS (1.74 Å) bond distance is usually
eycomb lattice of the fabricated 3D graphene-based macro- about 25% longer than that of the CeC bond (Chen et al., 2016).
architectures was examined through a quantitative analysis of their However, there was no direct correlation between the ID/IG values
wide-angle diffraction data. Irrespective of their doping level, the (i.e., the intensity ratio of the D to G band) and the bulk doping level
NSGAs exhibited exemplary scattering patterns of semicrystalline of the aerogels. Moreover, the ID/IG ratio computed for NSGA-III was
materials (Fig. 1c). The broad peak situated in the 2q range of the lowest amongst all the samples, implying that both the dopants
21 e28 , with maxima centred at around 2q ¼ 24 , can be indexed preferred to occupy the edge planes of this material.
to (0 0 2) facets of graphitic carbon, while the minor hump-like In order to explore the elemental composition of our chemically
feature appearing at about 2q ¼ 43 can be assigned to the modified graphene monoliths, EDX chemical mapping was per-
graphitic (1 0 0) crystal plane (Chowdhury et al., 2018). Moreover, formed with different colors representing the distribution of
the relatively weak intensities of these peak indicate that the as- various elements in the sample (Fig. S6). As can be seen, the as-
prepared NSGAs were made up of highly disordered and prepared NSGAs were primarily composed of C and O. In addi-
randomly oriented graphene sheets, with the interlayer distance tion, trace amounts of N and S were detected in all the samples,
increased by almost 10% compared to their singly N-doped coun- confirming that N and S atoms were successfully codoped into the
terpart (Fig. 1c). This can be rationalized by the fact that the cova- constituent graphene sheets of the aerogels. However, the
lent radius of S (102 pm) is much larger than those of C (77 pm) and elemental distribution of N was higher than S in all the samples
N (75 pm) (Qie et al., 2015). As such, the introduction of S atoms can (Table S2). Specifically, the N doping level increased according to
considerably enlarge the interlayer spacing between graphene the sequence: NSGA-IV (1.44 wt%) < NSGA-II (1.73 wt%) < NSGA-III
sheets by steric occupancy (Li et al., 2017). Such an expanded (2.38 wt%) < NSGA-I (2.42 wt%), while the total S content decreased
interlayer spacing has been reported to endow 2D-based materials in the following order: NSGA-I (1.69 wt%) > NSGA-III (1.49 wt
with beneficial structural and electronic modulations for high %) > NSGA-IV (1.37 wt%) > NSGA-II (1.24 wt%). These trends suggest
performance catalytic applications (Gao et al., 2015). that the atomic fraction of both N and S in the as-prepared samples
Additionally, Raman spectroscopic characterization was carried was not directly proportional to the amounts of NH4SCN added to
out to probe the lattice distortion and electronic charge localization GO during the hydrothermal reaction.
induced by N and S codoping. The comparative Raman scattering Further, the precise bonding state of N and S atoms in our dual
spectra of the various synthesized samples is shown in Fig. 1d. The heteroatom-doped graphene monoliths was established through
G band at around 1570 cm1 is a characteristic feature of in-plane interpretation of their XPS spectra. The survey scan generated the
stretching mode of all sp2 bonded pairs (including CeC, CeN and characteristic C1s, O1s, N1s and S2p photoelectron peaks at binding
CeS bonds), while the D band at about 1330 cm1 is mainly asso- energies (B.E.) of about 284, 532, 400 and 166 eV, respectively
ciated with defective/disordered graphitic structures, including in- (Fig. S7). Employing the iterative Shirley method, the high resolu-
plane heteroatom substitutions, vacancies, and/or grain tion N1s peak can be deconvoluted into two prominent signals
Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353 347

centred at ~398 and ~400 eV (Fig. 2a and Fig. S8), which can be the form of thiophene-like structures. This was corroborated by the
assigned to the well documented pyridinic (two CeN bonds in a peak-fitting resulting from the spineorbit coupling position of
hexagon), and pyrrolic (two CeN bonds in a pentagon) type N S2p1/2 of the CeSeC covalent bond for thiophene S at 164.8 eV
structural defects, respectively (Wang et al., 2012a). Surprisingly, (Yang et al., 2015) (Fig. S9). Last but not the least, the deconvolution
the emergence of a new peak at ~396 eV possibly indicates the of an intense but relatively smaller photoelectric peak at 168.2 eV
presence of substitutional N (Kusano et al., 2017); however, its exact can be attributed to the bonding of oxidized sulfur moieties
chemical identity is unknown. Fig. 2c compiles the fitting results for (CeSOXeC) such as sulfonate groups (‒SO‒3) to the 3D graphene
the N1s spectra of the as-prepared N/S dual-doped monolithic framework of the NSGA-IV monolith (Calvo et al., 2009; Yang et al.,
aerogels. The pyridinic N species was invariably present in all the 2012). In order to confirm the above findings, the O1s peak of each
samples and was also the dominant N-bonding configuration in the material was also deconvoluted into its various components, as
graphene core of all the materials, except NSGA-II wherein the shown in Fig. S10. These curve fitting results for the O1s spectrum
spectrum consisted primarily of pyrrolic N. Interestingly, no provide further evidence that S atoms were embedded in the form
decomposition peak representative of pyrrolic N was found in of oxidized sulfur species (such as sulfonic acid and sulfonate
NSGA-III; instead, substitutional N was the dominant defect functional groups) on the surface of the 2D nanoscale building
structure. Meanwhile, the high resolution S2p spectra can be fitted blocks of the aerogels. A schematic demonstrating the various to-
into several distinct subpeaks corresponding to the different pological defects in our bulk NSGA samples, as comprehended on
chemical environments of the embedded S atoms in the hexagonal the basis of XPS core-level analysis, is given in Fig. 3.
carbon skeleton of the aerogels (Fig. 2b and Fig. S9). The major peak Additionally, the light absorption properties of the NSGAs were
situated at ~166 eV for all doping levels originates from oxidized investigated through UVeVis diffuse reflectance spectroscopy. As
sulfur forms, principally sulfoxide (Upare et al., 2013). The shown in Fig. S11, the absorption spectra of the NSGAs did not vary
appearance of the rather weak but broad peak at the lower B.E. of significantly and presented high reflectance across the full visible
160.5 eV implies that some of the sulfoxide functionalities in NSGA- light portion of the electromagnetic spectrum. Interestingly, pris-
III were reduced to sulfides in the high temperature-pressure hy- tine GA, even without the presence of a sizeable energy gap, dis-
drothermal environment (Fantauzzi et al., 2015). In addition, the played sufficient absorption in the visible region. However, after
peak located at around 162 eV is indicative of the presence of incorporation of N and S heteroatoms, the intensity of the ab-
organosulfur (thiol) ligands (Wang et al., 2015). However, the thiol sorption increased by several folds. Applying the KubelkaeMunk
(eSH) signal was not detected in the S2p XPS spectrum of NSGA-III (KeM) function for direct allowed transition (Lόpez and Gόmez,
(Fig. 2d), suggesting that the S atoms mainly existed at the edge 2012; Ntholeng et al., 2016), i.e., [F(R)hv]2 ¼ A (hv e Eg) where
plane and defect sites of the graphene building blocks of NSGA-III in F(R) is proportional to the absorption coefficient (a), hv is the

Fig. 2. High resolution (a) N1s core level XPS spectrum and (b) S2p core level XPS spectrum of NSGA-IV. Relative distribution of the different (c) N- and (d) S-structural motifs in the
NSGAs.
348 Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353

Fig. 3. The various types of N and S functionalities observed in our as-fabricated NSGA materials: (1a) pyridinic N, (1b) pyridinic N with monovacancy, (1c) triple vacancy pyridinic
N, (2) pyrrolic N, (3) substitutional N, (4) sulfide, (5) thiol, (6) sulfoxide, (7) thiophenic and (8) sulfonic.

photon energy, Eg is the energy of the optical transition corre-


sponding to the optical energy band gap, and A is the proportion-
ality constant, the band gap energy could be calculated from the
intercept of the linear portion of the plot of [F(R)hv]2 versus hv
(Fig. S12). The Eg values estimated by this method were 1.51, 1.51,
1.52 and 1.42 eV for NSGA-I, NSGA-II, NSGA-III and NSGA-IV,
respectively.

3.2. Absorption of oils and organic solvents

With frequent occurrence of oil spill incidents and accidental


leakage of hydrophobic organic solvents (such as benzene, toluene,
cyclohexane, dichloromethane, etc.), the development of highly
efficient, lightweight, and mechanically strong absorbents has
become a top priority (Ge et al., 2016; Chowdhury et al., 2019). As
noted in the preceding section, our fabricated NSGAs have a highly
interconnected meso/macroporous structure with enlarged pores
Fig. 4. Absorption capacity of NSGA-IV for various oils and organic solvents. The digital
and large pore volumes, and these properties endow aerogels with images in the inset show the absorption of soybean oil by NSGA-IV over time.
good absorption properties. Furthermore, the aerogels developed
in this study have densities as low as 44.71 mg cm3. Assuming that
the carbon density is 2 g cm3, up to 97.76% of the aerogels can be displayed excellent absorption ability for various classes of organic
filled with air, which is extremely beneficial for absorption appli- compounds. In particular, depending on the density and surface
cations. Indeed, when our NSGA-IV aerogel was brought into con- tension of the hydrocarbons, the NSGA-IV material could absorb
tact with soybean oil, the oil layer immediately began to shrink and oils and organic solvents up to 22e63 times its own weight. This
disappeared completely within a minute (inset of Fig. 4). More performance is superior to that of other contemporary absorbents
importantly, the adsorbed oil could be removed from the such as activated carbon (<1 times) (Lillo-Ro  denas et al., 2005),
thermally-stable NSGA-IV absorbent through direct combustion in marshmallow-like macroporous gels (6e14 times) (Hayase et al.,
air for repeated use (inset of Fig. 4). To further evaluate the ab- 2013), polyacrylonitrile foam (<18 times) (Zhang et al., 2018), gra-
sorption potential of the NSGA-IV aerogel, we investigated its ab- phene aerogel (19e26 times) (Ren et al., 2017), spongy graphene
sorption capacities for several commercial petroleum products aerogel (29e54 times) (Luo et al., 2017), conjugated microporous
(e.g., diesel oil, engine oil, pump oil and kerosene) and common polymers (<33 times) (Li et al., 2011), polyurethane sponge (15e20
organic solvents (e.g., acetone, n-hexane, phenol and toluene). The times) (Zhu et al., 2011), graphene foam (10e37 times) (Niu et al.,
absorption efficiency was assessed by percentage weight gain, i.e., 2012), graphene/a-FeOOH aerogel (13e27 times) (Cong et al.,
the ratio of the mass of the organic liquid absorbed to that of the 2012), and nanocellulose aerogels (20e40 times) (Korhonen et al.,
dry aerogel. 2011). These results clearly demonstrate the application potential
As illustrated in Fig. 4, the freestanding NSGA-IV monolith of our synthesized multiple heteroatom doped graphene-based
Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353 349

aerogels for oil spill response and cleanup operations. We believe photocatalyst surface rather than in the bulk, the dye/photocatalyst
that the organic compounds permeate via capillary action into the suspension was first stirred for 1 h without any light exposure in
interconnected void spaces of NSGA-IV, and are subsequently order to establish adsorption‒desorption equilibrium between the
entrapped within its bimodal meso/macroporous matrix, contrib- dye and the photocatalyst surface (‘Light off’ period of Fig. 5a).
uting to the high absorption capacity. Furthermore, it is interesting Compared to their undoped and single doped counterparts, the
to note that the absorption ability of NSGA-IV for phenol and adsorption capacity of the NSGAs toward AO appeared to be sub-
toluene was higher than that for acetone and n-hexane. This is stantially higher, which was expected to provide excellent photo-
likely because aromatic hydrocarbons have greater affinity for catalytic performance. The ‘Light on’ period in Fig. 5a presents the
graphene-based materials and wet them better than aliphatic visible light driven photocatalytic degradation of AO over the as-
compounds via pep bonding, as has been reported previously by prepared graphene-based macroscale structures as a function of
Qian et al. (2014). irradiation time. Even in the absence of any photocatalyst, the dye
molecules underwent a photodegradation of about 34% upon
3.3. Photocatalytic degradation of textile dyes exposure to visible light, which is a typical case of direct photolytic
transformation. Meanwhile, the bulk GA and NGA samples, both
The direct discharge of dye-containing effluents from textile exhibited fairly weak photocatalytic activity, affording only ~49%
mills and leather processing industries is detrimental to aquatic and ~70% degradation of AO under visible light illumination for 5 h,
environments. Heterogeneous photocatalysis hold enormous po- respectively. In contrast, almost complete decolorization
tential to transform a broad spectrum of water borne pollutants (95.29e99.23%) was achieved within 4 h over the freestanding
into relatively innocuous end products (e.g., CO2 and H2O). We, NSGAs. These results clearly ascertain that codoping with two
therefore, also studied systematically the dye degradation ability of different elements is an appealing approach to boost the photo-
our NSGA samples under visible light exposure. Acridine orange catalytic efficiency and thus the environmental performance of
(AO) was selected as a model dye contaminant due to its frequent GAs, due to the redistribution of charge (spin) density and newly
occurrence in industrial wastewaters and mutagenic activities in created active sites induced by the synergistic interactions between
humans (Chowdhury et al., 2018). Since photocatalysis is essentially the heteroatoms. Nevertheless, in order to confirm that the dye
a surface phenomenon, occurring predominantly on the removal using our as-developed NSGAs is based predominantly on

Fig. 5. (a) Visible light-induced photocatalytic degradation of AO by GA, NGA and NSGAs (experimental conditions: C0 ¼ 20 mg L1; catalyst dose ¼ 0.33 g L1; temperature ¼ 25  C).
(b) Percentage removal of TOC during photocatalytic degradation of AO over the NSGAs. (c) Effect of initial dye concentration on the photocatalytic degradation of AO by NSGA-IV
under visible light irradiation (experimental conditions: catalyst dose ¼ 0.33 g L1; temperature ¼ 25  C). (d) Effect of pH on the photocatalytic degradation of AO by NSGA-IV under
visible light irradiation (experimental conditions: C0 ¼ 20 mg L1; catalyst dose ¼ 0.33 g L1; temperature ¼ 25  C).
350 Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353

the photocatalytic degradation mechanism, we also studied the diffusion of AO within the 3D porous structure of the NSGA-IV
adsorption and photocatalytic decolorization of AO over NSGA-IV material, and ultimately resulting in faster photocatalytic degra-
under identical experimental conditions (Fig. S13). The use of dation of AO. Upon further increasing the initial mass of the dye in
NSGA-IV in the dark resulted in a dye removal level of approxi- the reaction medium up to 25 mg L1, the degradation curves
mately 38%. On the other hand, the use of NSGA-IV under visible became shallower and it took longer to completely degrade AO for
light illumination led to a dye elimination efficiency of close to the same amount of light absorbance. It could be explained that the
100%. These findings confirm that photocatalysis was the dominant net flux of light reaching the photocatalyst surface decreases at a
mechanism governing the purification process. higher initial concentration of AO. This in turn partially offsets the
It should be noted, however, that the apparent disappearance of number of electronehole pairs participating in the photocatalytic
the dye from the reaction medium does not necessarily imply a reaction, thus slowing down the visible light-induced degradation
complete degradation of the pollutant but rather the formation of reaction to some extent. Moreover, at higher initial dye concen-
transformation products. These transformation products can trations, the concentrations of the intermediates generated during
sometimes be even more persistent and/or toxic than the parent the photocatalytic process are also expected to be higher. These
compound. Therefore, the complete mineralization of the dye prior intermediates would compete with dye molecules for interaction
to wastewater discharge to the sewer is of significant interest (Yu with primary oxidants, which eventually tend to lower the AO
et al., 2012). To this end, the total organic carbon (TOC) is an degradation rate.
important water quality parameter and its reduction is highly From an application point of view, it is equally important to
desirable. Fig. 5b compares the rate of TOC disappearance during study the effect of operating pH on the photocatalytic process, since
the photocatalytic degradation of AO with the as-prepared NSGAs the pH of the reaction medium dictates the surface charge char-
under visible light irradiation. It can be observed that the highest acteristics of the photocatalyst; and hence the overall rate of
experimentally obtained TOC elimination rate was ~84%, reflecting oxidative degradation. Fig. 5d shows the photodecomposition
the superior dye photomineralization potential of our N/S codoped profile of AO as a function of irradiation time under acidic and
hierarchically porous aerogels. Further, the TOC removal percent- alkaline conditions. Evidently, increasing the initial pH of the dye
ages decreased according to the following sequence: NSGA-IV solution from pH 4 increased the removal efficiency of AO up to pH
(84.30%) > NSGA-III (76.40%) > NSGA-II (68.82%) > NSGA-I 7 and did not show any significant changes thereafter. The observed
(34.60%). In spite of its relatively low doping level amongst all the trend can be attributed to the protonation of the pyridinic N moi-
prepared samples, the excellent photocatalytic effect of NSGA-IV eties and the oxidized sulfur species (sulfoxides and sulfonate) on
can be ascribed to the following factors. First, the highest specific NSGA-IV upon exposure to acidic environments (Su et al., 2015;
surface area and the largest pore size of NSGA-IV (Fig. S5 and Rauf et al., 2016). As a consequence, the adsorption of AO on the
Table S1) provide more active sites for dye adsorption/activation. As photocatalyst surface is hindered by electrostatic repulsion, thus
a prerequisite for photocatalyzed reactions, the enhanced adsorp- potentially compromising the photocatalytic activity of our mate-
tivity augment the photocatalytic activity. Second, the smallest rial. Conversely, with rising pH, deprotonation of the concerned N
band gap energy of the NSGA-IV sample facilitates a more efficient and S functionalities allows the cationic pollutants to be readily
utilization of visible light, increasing the number of both electrons adsorbed onto the positively charged surface of NSGA-IV through
and holes taking part in the photocatalytic reaction. This in turn electrostatic attraction, resulting in higher photocatalytic perfor-
builds up the concentration of primary oxidants (viz., hydroxyl mance. The plausibility of this explanation is corroborated by
radicals and superoxide anions) on the photocatalyst surface, thus measuring the surface zeta potentials as a function of pH (Fig. S14).
enhancing the photocatalytic performance. Last but not the least, The zeta potential of NSGA-IV became more negative with
unlike the other bulk materials, up to 30% of the S atoms exists increasing pH, confirming that higher pH is more favorable for
exclusively as sulfonic acid groups within the host lattice of NSGA- adsorption of the cationic AO dye and its subsequent
IV (Fig. 2d). Like pyridinic N, these oxidized sulfur clusters, situated photodegradation.
primarily at the edges or defect sites, serve as a preferential site for It is well documented that various reactive oxygen species
absorbing oxygen (key electron acceptor in aqueous photocatalysis) (ROS), such as singlet molecular oxygen (1O2), hydroxyl radicals
due to their large charge (or spin) density (Zhang et al., 2014). (,OH) and superoxide anions (O, 2 ), are formed when the photo-
Moreover, when N and S atoms are simultaneously incorporated induced electrons (e‒) and holes (hþ) interact with surface adsor-
into the sp2 network of graphene, the corresponding charge (or bed O2, and these active species in turn play a crucial role in the
spin) densities are further improved (Su et al., 2013), and conse- actual oxidation mechanism of photocatalytic reactions (Nosaka
quently promote the generation of more primary oxidants. As a and Nosaka, 2017). Therefore, in order to assess the individual
result, NSGA-IV even with a fairly low amount of N and S exhibited contributions of hþ, 1O2, O, ,
2 , and OH to the AO photodegradation
the best photocatalytic performance. Subsequently, the NSGA-IV process, their respective quenchers or scavengers were separately
sample was more thoroughly evaluated for its practical applica- spiked into the reaction medium. The results thus obtained are
tions in photocatalysis, as described in the following paragraphs. summarized in Fig. S15. When the hole sacrificial agent potassium
Since the initial concentration of the target recalcitrant organic iodide (KI) was added, the AO photodecolorization efficiency was
compound (AO) can have a profound influence on the reaction virtually unaffected, confirming that holes were not the main active
dynamics, the photocatalytic ability of NSGA-IV toward degrada- agent in our system. Upon introducing the selective 1O2 radical
tion of AO was explored over the concentration range of scavenger sodium azide (NaN3), the degradation percentage only
5e25 mg L1. Fig. 5c presents the time-dependent photo- declined slightly. On the other hand, the removal rate of AO drop-
decolorization curves of AO as a function of initial dye concentra- ped to 93.36% in the presence of isopropyl alcohol (IPA), a well-
tions. Clearly, when the initial AO concentration was amplified from known specific quencher for ,OH radicals. These findings suggest
5 mg L1 to 15 mg L1, the AO photodecomposition rate increased that the aqueous phase 1O2 and ,OH radicals were not the key
sharply and it took less time to achieve the same photocatalytic factors affecting the visible light-induced photocatalytic decom-
performance. This is likely because enhancement of the concen- position of AO. In the meantime, the photodegradation of AO was
tration gradient augments the net driving force to effectively sup- greatly suppressed by the addition of the O, 2 sacrificial reagent
press the overall mass transfer resistance of the dye molecules benzoquinone (BQ) into the reaction system, implying that O, 2
between the aqueous and solid phases, thus facilitating rapid anions were the primary oxidant governing the photocatalytic
Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353 351

oxidation of AO over NSGA-IV under visible light irradiation. On the material. As can be seen in Fig. S17, the AO degradation efficiency
basis of these observations, the plausible pathways for the photo- reached over 99% after the first run. During the second run, the
degradation of AO can be proposed as follows: same photocatalyst could give as much as 92% decolorization,
whereas the third run resulted in almost 88% AO removal. The
AO þ hv / AO* (1) observed drop in photocatalytic efficiency can be attributed to the
accumulation of oxidation byproducts on the surface of the catalyst
with time. Nevertheless, the NSGA-IV photocatalyst could still
AO* þ NSGA  IV / AOþ þ NSGA  IV ðe
CB Þ (2)
retain at least 87% of its initial activity even after three catalytic
cycles, demonstrating its long-term durability and good recycla-
NSGA  IV þ hv / NSGA  IV ðhþ
VB þ e
CB Þ (3) bility for continuous operation.


O2 þ e
CB / O2 (4)
4. Conclusion

O
2 þ hþ
VB / 1
O2 (5) In summary, extremely lightweight, yet mechanically resilient,
N/S codoped GAs have been developed through a facile and scalable
H2 O þ hþ
VB / OH

þ Hþ (6) one-pot hydrothermal method. The as-prepared monolithic NSGAs
possessed continuously interconnected meso/macroporous struc-
tures with large specific surface area (107e132 m2 g1) and abun-
hþ þ OH / 
OH (7)
VB dant pore space (0.19e0.25 cm3 g1). Due to these beneficial
structural features, our freestanding NSGAs exhibited outstanding
O  1
2 ; OH; O2 þ AOþ / CO2 þ H2 O absorption properties and could absorb hydrocarbons (e.g., engine
þ Mineralization Products (8) oil, diesel oil, pump oil, soybean oil and kerosene) and different
types of organic solvents (such as acetone, n-hexane, phenol and
Initially, the surface adsorbed dye molecules are promoted to an toluene) up to several times their own weight, and may therefore
electronically excited state upon exposure to visible light (eq. (1)). be considered for industrial chemical and oil spill cleanup.
Immediately after photoexcitation, the electrons (e‒) in excited Furthermore, the NSGAs displayed remarkable visible light-driven
states are quickly injected into the conduction band (CB) of NSGA- photocatalytic activities. Particularly, the aerogel with relatively
IV (eq. (2)). Concurrently, by absorbing photons having energies low concentrations of N and S atoms in its graphene core (i.e.,
equal to or greater than the band gap of NSGA-IV, some electrons NSGA-IV) exhibited the highest photocatalytic efficiency for the
(e‒) from the valence band (VB) are photoexcited to the CB (eq. (3)). degradation of recalcitrant organic pollutants such as AO. Nearly
This direct interband transition leaves an equal number of holes complete mineralization of AO with over 84% TOC removal was
(hþ) in the VB (eq. (3)). Thereafter, the photoinduced electrons at observed following exposure to visible light for 5 h. Moreover,
the CB reduce the surface adsorbed oxygen molecules to O, 2 rad- compared to the parent compound, the oxidation byproducts were
icals (eq. (4)). Likewise, the photogenerated holes oxidize the sur- fairly innocuous, as inferred from toxicity assessment using E. coli
face bound water molecules, or OH groups to ,OH radicals (eqs. (6) as a model organism. The superior photocatalytic performance
and (7)). Alongside, a fraction of the produced O, 2 is converted to could be ascribed to the synergistic effect produced by N/S
1
O2 through direct oxidization by VB holes (as in eq. (5)). Eventu- codoping and 3D porous networks of interlinked 2D graphene
ally, due to their strong oxidizing ability, these ROS (O, ,
2 , OH, and nanosheets. Additionally, the photocatalytic mechanism was clari-
1
O2) cleave the dye molecules, disintegrate them into a range of fied by employing scavengers or quenchers of reactive oxygen and/
transient aromatic intermediates, and finally mineralize them into or radical species, which revealed that superoxide anions played a
H2O, CO2 and other inorganic anions (eq. (8)). However, the precise key role in the surface-mediated photocatalytic reactions. Ulti-
estimation of the lifetime of the photogenerated radicals as well as mately, our preliminary findings explicitly demonstrate that the
identification of the transient reaction intermediates and end coupling of N and S dopants with different geometrical configura-
products is imperative in order to validate the proposed tions in the graphene lattice synergistically enhance the application
mechanism. potential of GAs for addressing anthropogenic environmental
Furthermore, the toxicity of the photodegradation products of degradation.
AO was evaluated by monitoring the viability of the Gram negative
bacterium Escherichia coli (E. coli) using the conventional plate
counting technique. The untreated dye solution (25 mg L1) Acknowledgements
induced obvious toxicity to E. coli, as evident from the very low cell
viability (~18%) depicted in Fig. S16. On the contrary, the survival This work was financially supported by the National Research
rate of E. coli remained almost unaltered (more than 95%) in the Foundation (NRF), Prime Minister's Office, Republic of Singapore
presence of the photodegradation end products (Fig. S16). There- through its Campus for Research Excellence and Technological
fore, it is reasonable to infer that our as-developed NSGA-IV could Enterprise (CREATE) programme (Grant No. R-706-002-101-281) as
effectively transform recalcitrant organic pollutants into harmless well as the National University of Singapore (Grant No. R-302-000-
compounds, and hence warrants further consideration for the 209-133). Ms. Yiqun Jiang further acknowledges the financial sup-
treatment of dye bearing wastewaters. port provided by the China Scholarship Council, Ministry of Edu-
Apart from exceptional photocatalytic activity, the easy recovery cation of the People's Republic of China for her Doctoral study.
and repeated use of the spent photocatalyst without substantial
loss in catalytic ability are essential for its prolonged cyclic opera-
tion, which in turn has a direct influence on the total cost of the Appendix A. Supplementary data
treatment process. We, therefore, performed three sequential
photocatalytic runs toward the degradation of AO under identical Supplementary data to this article can be found online at
experimental conditions to explore the recyclability of our NSGA-IV https://doi.org/10.1016/j.envpol.2019.04.132.
352 Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353

References graphene oxide foams. Adv. Mater. 24, 4144e4150.


Nosaka, Y., Nosaka, A.Y., 2017. Generation and detection of reactive oxygen species
in photocatalysis. Chem. Rev. 117, 11302e11336.
Calvo, A., Yameen, B., Williams, F.J., Azzaroni, O., Soler-Illia, G.J.A.A., 2009. Facile
Ntholeng, N., Mojela, B., Gqoba, S., Airo, M., Govindraju, S., Moloto, M.J., Van Wyk, J.,
molecular design of hybrid functional assemblies with controllable transport
Moloto, N., 2016. Colloidal synthesis of pure CuInTe2 crystallites based on the
properties: mesoporous films meet polyelectrolyte brushes. Chem. Commun.
HSAB theory. New J. Chem. 40, 10259e10266.
2553e2555.
Qian, Y., Ismail, I.M., Stein, A., 2014. Ultralight, high-surface-area, multifunctional
Chen, J.-F., Mao, Y., Wang, H.-F., Hu, P., 2016. Theoretical study of heteroatom doping
graphene-based aerogels from self-assembly of graphene oxide and resol.
in tuning the catalytic activity of graphene for triiodide reduction. ACS Catal. 6,
Carbon 68, 221e231.
6804e6813.
Qie, L., Chen, W., Xiong, X., Hu, C., Zou, F., Hu, P., Huang, Y., 2015. Sulfur-doped
Chowdhury, S., Jiang, Y., Muthukaruppan, S., Balasubramanian, R., 2018. Effect of
carbon with enlarged interlayer distance as a high-performance anode material
boron doping level on the photocatalytic activity of graphene aerogels. Carbon
for sodium-ion batteries. Adv. Sci. 2, 1500195.
128, 237e248.
Rahmani, Z., Rashidi, A.M., Kazemi, A., Samadi, M.T., Rahmani, A.R., 2018. N-doped
Chowdhury, S., Pan, S., Balasubramanian, R., Das, P., 2019. Three-dimensional
reduced graphene oxide aerogel for the selective adsorption of oil pollutants
graphene-based macroscopic assemblies as super-absorbents for oils and
from water: isotherm and kinetic study. J. Ind. Eng. Chem. 61, 416e426.
organic solvents. In: Naushad, M. (Ed.), A New Generation Material Graphene:
Rauf, M., Zhao, Y.-D., Wang, Y.-C., Zheng, Y.-P., Chen, C., Yang, X.-D., Zhou, Z.-Y.,
Applications in Water Technology. Springer, Cham, pp. 43e68.
Sun, S.-G., 2016. Insight into the different ORR catalytic activity of Fe/N/C be-
Cong, H.-P., Ren, X.-C., Wang, P., Yu, S.-H., 2012. Macroscopic multifunctional
tween acidic and alkaline media: protonation of pyridinic nitrogen. Electro-
graphene-based hydrogels and aerogels by a metal ion induced self-assembly
chem. Commun. 72, 71e74.
process. ACS Nano 6, 2693e2703.
Ren, R.-P., Li, W., Lv, Y.-K., 2017. A robust, superhydrophobic graphene aerogel as a
Du, X., Liu, H.-Y., Mai, Y.-W., 2016. Ultrafast synthesis of multifunctional N-doped
recyclable sorbent for oils and organic solvents at various temperatures.
graphene foam in an ethanol flame. ACS Nano 10, 453e462.
J. Colloid Interface Sci. 500, 63e68.
Fang, Q., Shen, Y., Chen, B., 2015. Synthesis, decoration and properties of three-
Saana Amiinu, I., Zhang, J., Kou, Z., Liu, X., Asare, O.K., Zhou, H., Cheng, K., Zhang, H.,
dimensional graphene-based macrostructures: a review. Chem. Eng. J. 264,
Mai, L., Pan, M., Mu, S., 2016. Self-organized 3D porous graphene dual-doped
753e771.
with biomass-sponsored nitrogen and sulfur for oxygen reduction and evolu-
Fantauzzi, M., Elsener, B., Atzei, D., Rigoldi, A., Rossi, A., 2015. Exploiting XPS for the
tion. ACS Appl. Mater. Interfaces 8, 29408e29418.
identification of sulfides and polysulfides. RSC Adv. 5, 75953e75963.
Serp, P., Machado, B., 2015. Nanostructured Carbon Materials for Catalysis. The
Gao, M.-R., Chan, M.K.Y., Sun, Y., 2015. Edge-terminated molybdenum disulfide with
Royal Society of Chemistry, Cambridge.
a 9.4-Å interlayer spacing for electrochemical hydrogen production. Nat.
Shen, Y., Fang, Q., Chen, B., 2014. Environmental applications of three-dimensional
Commun. 6, 7493.
graphene-based macrostructures: adsorption, transformation, and detection.
Ge, J., Zhao, H.-Y., Zhu, H.-W., Huang, J., Shi, L.-A., Yu, S.-H., 2016. Advanced sorbents
Environ. Sci. Technol. 49, 67e84.
for oil-spill cleanup: recent advances and future perspectives. Adv. Mater. 28,
Su, Y., Zhang, Y., Zhuang, X., Li, S., Wu, D., Zhang, F., Feng, X., 2013. Low-temperature
10459e10490.
synthesis of nitrogen/sulfur co-doped three-dimensional graphene frameworks
Gorgolis, G., Galiotis, C., 2017. Graphene aerogels: a review. 2D Mater. 4, 032001.
as efficient metal-free electrocatalyst for oxygen reduction reaction. Carbon 62,
Hayase, G., Kanamori, K., Fukuchi, M., Kaji, H., Nakanishi, K., 2013. Facile synthesis of
296e301.
marshmallow-like macroporous gels usable under harsh conditions for the
Su, S., Wei, J., Zhang, K., Qiu, J., Wang, S., 2015. Thermo- and pH-responsive fluo-
separation of oil and water. Angew. Chem. Int. Ed. 52, 1986e1989.
rescence behaviors of sulfur-functionalized detonation nanodiamond-poly(N-
Huang, H., Zhu, J., Zhang, W., Tiwary, C.S., Zhang, J., Zhamg, X., Jiang, Q., He, H.,
isopropylacrylamide). Colloid Polym. Sci. 293, 1299e1305.
Wu, Y., Huang, W., Ajayan, P.M., Yan, Q., 2016. Controllable codoping of nitrogen
Sui, Z.-Y., Meng, Y.-N., Xiao, P.-W., Zhao, Z.-Q., Wei, Z.-X., Han, B.-H., 2015. Nitrogen-
and sulphur in graphene for highly efficient Li-oxygen batteries and direct
doped graphene aerogels as efficient supercapacitor electrodes and gas adsor-
methanol fuel cells. Chem. Mater. 28, 1737e1745.
bents. ACS Appl. Mater. Interfaces 7, 1431e1438.
Jiang, Y., Chowdhury, S., Balasubramanian, R., 2017. Nitrogen-doped graphene
Tian, W., Zhang, H., Duan, X., Sun, H., Tade, M.O., Ang, H.M., Wang, S., 2016. Ni-
hydrogels as potential adsorbents and photocatalysts for environmental
trogen- and sulfur-codoped hierarchically porous carbon for adsorptive and
remediation. Chem. Eng. J. 327, 751e763.
oxidative removal of pharmaceutical contaminants. ACS Appl. Mater. Interfaces
Jiang, Y., Chowdhury, S., Balasubramanian, R., 2019. New insights into the role of
8, 7184e7193.
nitrogen-bonding configurations in enhancing the photocatalytic activity of
Upare, P.P., Yoon, J.-W., Kim, M.Y., Kang, H.-Y., Hwang, D.W., Hwang, Y.K., Kung, H.H.,
nitrogen-doped graphene aerogels. J. Colloid Interface Sci. 534, 574e585.
 ski, W., Szala, M., Bystrzejewski, M., 2014. Sulfur-doped porous carbons: syn- Chang, J.-S., 2013. Chemical conversion of biomass-derived hexose sugars to
Kicin
levulinic acid over sulfonic acid-functionalized graphene oxide catalysts. Green
thesis and applications. Carbon 68, 1e32.
Chem. 15, 2935e2943.
Korhonen, J.T., Kettunen, M., Ras, R.H.A., Ikkala, O., 2011. Hydrophobic nanocellulose
Wang, H., Maiyalagan, T., Wang, X., 2012a. Review on recent progress in nitrogen-
aerogels as floating, sustainable, reusable, and recyclable oil absorbents. ACS
doped graphene: synthesis, characterization, and its potential applications.
Appl. Mater. Interfaces 3, 1813e1816.
ACS Catal. 2, 781e794.
Kotal, M., Kim, J., Oh, J., Oh, I.-K., 2016. Recent progress in multifunctional graphene
Wang, J., Shi, Z., Fan, J., Ge, Y., Yin, J., Hu, G., 2012b. Self-assembly of graphene into
aerogels. Front. Mater. 3, 29.
three-dimensional structures promoted by natural phenolic acids. J. Mater.
Kusano, D., Emori, M., Sakama, H., 2017. Influence of electronic structure on visible
Chem. 22, 22459e22466.
light photocatalytic activity of nitrogen-doped TiO2. RSC Adv. 7, 1887e1898.
Wang, X., Sun, G., Routh, P., Kim, D.-H., Huang, W., Chen, P., 2014. Heteroatom-
Li, A., Sun, H.-X., Tan, D.-Z., Fan, W.-J., Wen, S.-H., Qing, X.-J., Li, G.-X., Li, S.-Y.,
doped graphene materials: syntheses, properties and applications. Chem. Soc.
Deng, W.-Q., 2011. Superhydrophobic conjugated microporous polymers for
Rev. 43, 7067e7098.
separation and adsorption. Energy Environ. Sci. 4, 2062e2065.
Wang, Y., Zhang, B., Xu, M., He, X., 2015. Tunable ternary (P, S, N)-doped graphene as
Li, Z., Bommier, C., Chong, Z.S., Jian, Z., Surta, T.W., Wang, X., Xing, Z., Neuefeind, J.C.,
an efficient electrocatalyst for oxygen reduction reaction in an alkaline medium.
Stickle, W.F., Dolgos, M., Greaney, P.A., Ji, X., 2017. Mechanism of Na-ion storage
RSC Adv. 5, 86746e86753.
in hard carbon anodes revealed by heteroatom doping. Adv. Energy Mater. 7,
Wang, J., Duan, X., Dong, Q., Meng, F., Tan, X., Liu, S., Wang, S., 2019. Facile synthesis
1602894.
denas, M.A., Cazorla-Amoro  s, D., Linares-Solano, A., 2005. Behaviour of of N-doped 3D graphene aerogel and its excellent performance in catalytic
Lillo-Ro
degradation of antibiotic contaminants in water. Carbon 144, 781e790.
activated carbons with different pore size distributions and surface oxygen
Wu, Y., Zhu, J., Huang, L., 2019. A review of three-dimensional graphene-based
groups for benzene and toluene adsorption at low concentrations. Carbon 43,
materials: synthesis and applications to energy conversion/storage and envi-
1758e1767.
ronment. Carbon 143, 610e640.
Liu, Z., Li, D., Dai, H., Huang, H., 2017. Enhanced properties of tea residue cellulose
Xing, L.-B., Xi, K., Li, Q., Su, Z., Lai, C., Zhao, X., Vasant Kumar, R., 2016. Nitrogen,
hydrogels by addition of graphene oxide. J. Mol. Liq. 244, 110e116.
sulfur-codoped graphene sponge as electroactive carbon interlayer for high-
Lόpez, R., Gόmez, R., 2012. Band-gap energy estimation from diffuse reflectance
energy and -power lithiumesulfur batteries. J. Power Sources 303, 22e28.
measurements on solegel and commercial TiO2: a comparative study. J. Sol. Gel
Yang, Z., Yao, Z., Guifa, L., Fang, G., Nie, H., Liu, Z., Zhou, X., Chen, X., Huang, S., 2012.
Sci. Technol. 61, 1e7.
Sulfur-doped graphene as an efficient metal-free cathode catalyst for oxygen
Luo, Y., Jiang, S., Xiao, Q., Chen, C., Li, B., 2017. Highly reusable and superhydrophobic
reduction. ACS Nano 6, 205e211.
spongy graphene aerogels for efficient oil/water separation. Sci. Rep. 7, 7162.
Yang, Z., Dai, Y., Wang, S., Cheng, H., Yu, J., 2015. In situ incorporation of a S, N doped
Ma, X.-X., Dai, X.-H., He, X.-Q., 2017. Co9S8-modified N, S, and P ternary-doped 3D
carbon/sulfur composite for lithium sulfur batteries. RSC Adv. 5, 78107e78025.
graphene aerogels as high-performance electrocatalyst for both the oxygen
Yang, W., Gao, H., Zhao, Y., Bi, K., Li, X., 2016. Facile preparation of nitrogen-doped
reduction reaction and oxygen evolution reaction. ACS Sustain. Chem. Eng. 5,
graphene sponge as a highly efficient oil absorption material. Mater. Lett. 178,
9848e9857.
95e99.
Maleki, H., 2016. Recent advances in aerogels for environmental remediation ap-
Yousefi, N., Lu, X., Elimelech, M., Tufenkji, 2019. Environmental performance of
plications: a review. Chem. Eng. J. 300, 98e118.
graphene-based 3D macrostructures. Nat. Nanotechnol. https://doi.org/10.1038/
May, P., Lazzeri, M., Venezuela, P., Herziger, F., Callsen, G., Reparaz, J.S., Hoffmann, A.,
s41565-018-0325-6.
Mauri, F., Maultzsch, J., 2013. Signature of the two-dimensional phonon
Yu, K., Yang, S., Liu, C., Chen, H., Li, H., Sun, C., Boyd, S.A., 2012. Degradation of
dispersion in graphene probed by double-resonant Raman scattering. Phys. Rev.
organic dyes via bismuth silver oxide initiated direct oxidation coupled with
B 87, 075402.
sodium bismuthate based visible light photocatalysis. Environ. Sci. Technol. 46,
Niu, Z., Chen, J., Hng, H.H., Ma, J., Chen, X., 2012. A leavening strategy to prepare
7318e7326.
Y. Jiang et al. / Environmental Pollution 251 (2019) 344e353 353

Zhang, L., Niu, J., Li, M., Xia, Z., 2014. Catalytic mechanisms of sulfur-doped graphene Mater. Interfaces 10, 41871e41877.
as efficient oxygen reduction reaction catalysts for fuel cells. J. Phys. Chem. C Zhao, Y., Hu, C., Hu, Y., Cheng, H., Shi, G., Qu, L., 2012. A versatile, ultralight, nitrogen-
118, 3545e3553. doped graphene framework. Angew. Chem. Int. Ed. 51, 11371e11375.
Zhang, Q., Wang, H., Wang, L., Zhuang, Y., Li, W., Zhou, Y., Gu, S., Wang, X., Yang, H., Zhu, Q., Pan, Q., Liu, F., 2011. Facile removal and collection of oils from water sur-
Xu, W., 2018. Wood-inspired fabrication of polyacrylonitrile solid foam with faces through superhydrophobic and superoleophilic sponges. J. Phys. Chem. C
superfast and high absorption capacity for liquid without selectivity. ACS Appl. 115, 17464e17470.

You might also like