Next Article in Journal
Effects of Temperature on the Leaching Behavior of Pb from Cement Stabilization/Solidification-Treated Contaminated Soil
Next Article in Special Issue
A Novel Method for Simulating Micro-Scale Water Droplet Movements
Previous Article in Journal
Simvastatin: In Vitro Metabolic Profiling of a Potent Competitive HMG-CoA Reductase Inhibitor
Previous Article in Special Issue
Saline Diffusion Modeling for Sodium Chloride Aqueous Solutions: Freezing for Desalination Purposes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Preparation and Application of Graphene–Based Materials for Heavy Metal Removal in Tobacco Industry: A Review

1
Shanghai Key Lab of Chemical Assessment and Sustainability, School of Chemical Science and Engineering, Tongji University, Shanghai 200092, China
2
Technical Centre of Shanghai Tobacco Group Co., Ltd., Shanghai 200082, China
3
State Key Laboratory of Pollution Control and Resource Reuse, Shanghai Institute of Pollution Control and Ecological Security, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Separations 2022, 9(12), 401; https://doi.org/10.3390/separations9120401
Submission received: 22 October 2022 / Revised: 21 November 2022 / Accepted: 23 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Advances in Separation Engineering)

Abstract

:
Heavy metals are nondegradable in the natural environment and harmful to the ecological system and human beings, causing an increased environmental pollution problem. It is required to remove heavy metals from wastewater urgently. Up until now, various methods have been involved in the heavy metal removals, such as chemical precipitation, chemical reduction, electrochemical, membrane separation, ion exchange, biological, and adsorption methods. Among them, adsorption by graphene–based materials has attracted much more attentions for the removal of heavy metals from wastewater systems in recent years, arising due to their large specific surface area, high adsorption capacity, high removal efficiency, and good recyclability. Therefore, it is quite important to review the heavy metal removal with the graphene–based material. In this review, we have summarized the physicochemical property and preparation methods of graphene and their adsorption property to heavy metals. The influencing parameters for the removal of heavy metals by graphene–based materials have been discussed. In addition, the modification of graphene–based materials to enhance their adsorption capability for heavy metal removal is also reviewed. The heavy metal removal by modified graphene–based materials in the tobacco industry has been especially described in detail. Finally, the future trend for graphene–based materials in the field of heavy metal wastewater treatment is proposed. This knowledge will have great impacts on the field and facilitate the researchers to seek the new functionalization method for graphene–based materials with high adsorption capacity to heavy metals in the tobacco industry in the future.

1. Introduction

Generally, heavy metals are metals with a density higher than 4.5 g/cm3, including chromium (Cr), cadmium (Cd), arsenic (As), mercury (Hg), lead (Pb), and so on [1]. Heavy metals are difficult to be degraded in the natural environment [2,3]. Nowadays, a large amount of heavy metal wastewater in mining, metallurgy, machinery manufacturing, chemical industry, electronics, and other industries is directly discharged into the natural environment without treatment, causing severe harm to the ecological environment and human health [4,5]. Heavy metal pollution is becoming a significant global environmental problem so far [6]. Therefore, it is urgent to remove heavy metals from wastewater [7]. In the past few decades, many methods have been developed for the treatment of heavy metals, such as chemical precipitation, chemical reduction, electrochemical, membrane separation, ion exchange, biological, and adsorption methods [8,9,10,11]. These methods are relatively mature and play an important role in the field of water pollution control [12]. Among these methods, adsorption is one of the most effective methods because of simple operation and low cost [13,14], which exhibits a great potential application for heavy metal removal [15].
Normally, the removal efficiency of the adsorption method mainly depends on the type of adsorbent. Recently, various low–cost adsorbents developed from agricultural wastes, industrial by–products, or natural polymer materials have been widely used in the treatment of heavy metal wastewater [16]. However, it is difficult for these common adsorbents to reduce the concentration of heavy metals in wastewater to ppb level. In addition, their subsequent recovery and reuse are also difficult. Therefore, the research and development of new adsorbents are currently a research hotspot in the field of heavy metal wastewater treatment [17].
Carbon nanomaterials possess at least one dimension less than 100 nm. They have the advantages of small size, large pore volume, large specific surface area, high adsorption efficiency, low operation process, and low production cost [18]. These characteristics make them have broad prospects in the field of heavy metal wastewater treatment [19]. At present, common adsorbents, such as lignin, molecular sieve, zeolite, activated carbon, and mesoporous silica, which are widely used in heavy metal wastewater treatment [20], have difficulty in reducing the concentration of heavy metal in wastewater to ppb level. Moreover, the long adsorption time and poor selectivity also limit their applications for heavy metal removal. By contrast, the use of new carbon nanomaterials as adsorbents can solve these problems very well. The new carbon nanomaterial adsorbents mainly include carbon nanotubes, graphene, and other nano–adsorbents with carbon as the main element and sp2 as the main hybrid form [21]. With the unique structure and pore size distribution of these materials, the new carbon nanomaterial adsorbents have very large specific surface areas, high electron transfer rates, outstanding physical properties, and good thermal conductivity [22]. The new carbon nanomaterial adsorbent has outstanding advantages to adsorb heavy metals, such as strong adsorption capacity, high removal efficiency, fast equilibrium speed, low price, no secondary pollution, and good recyclability [23].
Graphene is a two–dimensional carbon material [24,25]. Its crystal lattice has a hexagonal lattice structure with a single layer of crystalline carbon atoms inside. It can combine three adjacent carbon atoms through sp2 hybridization, and the remaining valence electron can form a delocalized large π bond [26]. The discovery of graphene in 2004 by Geim et al. [27] overturned the classical theory that “thermodynamic expansion does not allow two–dimensional crystals to exist freely at a limited temperature” and shocked the entire physics community, which set off a research boom in the field of graphene. As depicted in Figure 1, graphene is the mother for building graphitic materials of all dimensionalities. For example, it can be wrapped into zero–dimensional (0D) fullerene, rolled into the one–dimensional (1D) carbon nanotubes, or stacked into three–dimensional (3D) graphite as well. Currently, graphene is receiving extensive attention in the field of heavy metal wastewater treatment due to its large specific surface area and huge adsorption flux [28,29].
Even though there are some reviews on graphene for heavy metal removal, they mostly focus on the organo–functionalized magnetic graphene oxide nanocomposites [30], graphene oxide and its composites [31], hybrid graphene materials [32], and graphene–based materials for heavy metal removal in the electroplating process [33]. Particularly, Xu et al. [34] reviewed the modification of graphene–based materials and their adsorption efficacy on the heavy metals and radionuclides from wastewater. They also discussed the detailed adsorption mechanism, adsorption dynamics, and isotherms. The work by Kong et al. [35] mostly focused on the improvement of adsorption capacities to heavy metals for graphene–based materials. They indicated three strategies, such as surface functionalization regulation, morphology and structure control, and material composite, to enhance the heavy metal adsorption capacities of graphene–based materials. Unfortunately, the details on the physicochemical properties of graphene and fabrications of graphene–based materials are rarely summarized for heavy metal removal [36]. In addition, the graphene–based materials for heavy metal removal in the tobacco industry have not been summarized yet. Therefore, in this review, we discuss in detail the physicochemical properties and fabrication methods of graphene as well as the modification of graphene to improve their heavy metal removal ability. We also discuss the effect parameters on heavy metal removal, especially in regard to its use in heavy metal removal in the tobacco industry, as well as proposed future perspectives in this field.
Figure 1. Graphene and its derivatives: 0D fullerene, 1D carbon nanotubes, and 3D graphite. Reprinted with permission from [37]. Copyright@ Springer–Nature.
Figure 1. Graphene and its derivatives: 0D fullerene, 1D carbon nanotubes, and 3D graphite. Reprinted with permission from [37]. Copyright@ Springer–Nature.
Separations 09 00401 g001

2. Physicochemical Properties and Fabrications of Graphene

2.1. Physicochemical Property

The special chemical structure of graphene determines its unique chemical properties. The basic building block of graphene is the extremely stable six–membered benzene ring, which makes the structure of graphene particularly stable, and the general chemical methods cannot destroy its structure [38]. Furthermore, graphene also contains boundary groups and plane defects, which determine that graphene can provide adsorption sites for chemical reactions. Additionally, owing to its own large π conjugated system, graphene has a relatively negatively charged system and can react with many oxidants. These unique physicochemical properties enable graphene to undergo chemical reactions through boundary functional groups while well–maintaining the stability of its main framework [37]. Therefore, nowadays, researchers mostly apply graphene oxide (GO) and functionalize the oxygen–containing functional groups, such as carboxyl groups and hydroxyl groups, on the surface of GO to prepare novel GO–based materials [39] with higher adsorption capacities to heavy metals. The chemical structure of GO is disclosed in Figure 2.
The structure characterizations of graphene and GO, including Fourier–transform infrared spectroscopy (FTIR), Raman spectroscopy, X–ray diffraction (XRD), X–ray photoelectron spectroscopy (XPS), Brunauer–Emmett–Teller (BET), scanning electron microscope (SEM), and transmission electron microscopy (TEM), are depicted in Figure 3 [41].
As depicted in Figure 3A, there is no obvious absorption peak in the FTIR spectrum of graphene. By contrast, the absorption peaks at 1735, 1617, 1220, and 1047 cm−1 are visibly observed in the FTIR spectrum of GO, originating from the stretching vibration of C=O in the carboxyl group, C=C, C–O, and C–O–C, respectively [42]. In the Raman spectrum of graphene, Figure 3B, the intensity of the G–band (sp2 C=C bond stretching vibration) is significantly higher than that of the D–band (structural defects), whereas in the Raman spectrum of GO, the D–band is clearly increased and G–band is declined, illustrating the oxidation of graphene [43]. In the XRD pattern of graphene, Figure 3C, there is a characteristic diffraction peak at 2θ = 26.38°, while that of GO is at 2θ = 10.13°. This arises from the increase of interlayer distance by the introduction of oxygen–containing functional groups into graphene sheets because of oxidation [44]. Figure 3D,E shows the high–resolution C1s XPS spectra of graphene and GO, respectively. It is found that only two peaks at 284.8 and 285.3 eV, attributed to C=C and C–C, respectively, are deconvoluted in the high–resolution C 1s XPS spectrum of graphene, while four peaks at 284.6, 286.4, 287.5, and 289.1 eV, corresponding to C=C, C–O, C=O, and O–C=O, respectively, are noticed in that of GO [45]. Figure 3F provides the nitrogen adsorption–desorption curve of GO obtained from the BET test. This curve exhibits a typical IV adsorption–desorption profile, implying mesoporous characteristics of GO. Figure 3G,H is the SEM images of GO at different magnifications. Many diverse micropores are seen in the SEM image of GO because of the freeze–drying process, which may be beneficial for the adsorption of heavy metals. The TEM image of GO is drawn in Figure 3I. It is noted that there are many folds on the graphene sheets of GO, which are from the covalent bond or physical interaction between the oxygen–containing functional groups formed during the oxidation process [41].

2.2. Fabrications of Graphene

Generally, preparation methods for graphene include the mechanical exfoliation method, chemical redox method, chemical vapor deposition method, epitaxial growth method, electrochemical method, and organic synthesis method, etc. [46].

2.2.1. Mechanical Exfoliation Method

The mechanical exfoliation method serves as an effective method manufacture graphene [47]. The basic principle of mechanical exfoliation is to exfoliate graphene from graphite sheets through physical force and other external forces [48]. It contains many exfoliation methods, such as micromechanical exfoliation, ball milling, fluid dynamics, and ultrasonic exfoliation [49]. The most classic mechanical exfoliation method is the directional isocationic pyrolysis of graphite material by Novoselov et al. [27], then the products were transferred with photoresist to the substrate also with photoresist, and the substrate was repeatedly peeled off with tape. After that, the substrate was immersed in an acetone solution, and then the graphene was taken out by using a single crystal silicon wafer, and the acetone was ultrasonicated to obtain the first graphene. Ball milling is also a more commonly used mechanical exfoliation method. Deng et al. [50] used wet ball milling to disperse graphite into N–methylpyrrolidone (NMP) solution system with NMP as the dispersion medium and successfully separated graphene through the interaction with the surface of graphene. The experimental results showed that the graphene sheets were decreased rapidly with the increase of ball milling time. Figure 4 depicts the illustration for graphene exfoliation during the ball milling process.
In this process, two forces, which were named shear force and compression force, were generated. The graphite could be separated by shear force from their outer surfaces and made smaller by the compression force. Finally, the graphene exfoliation was achieved under the action of these two forces. The mechanical exfoliation method is simple and effective, and the prepared graphene is of high quality, which could be widely used in the study of graphene. However, this method consumes a lot of time and energy, is difficult to be controlled accurately, and is not suitable for large–scale production of graphene.

2.2.2. Chemical Redox Method

Currently, the chemical redox method is the most widely used preparation method of graphene [51]. The mechanism of the chemical redox method is to insert small molecules of strong acid into the middle of the graphite layer by treating the graphite powder with strong acid and then adding a strong oxidant into the reaction system for oxidation [52]. This thereby destroys the complete crystal structure of the graphite and introduces oxygen–containing functional groups on the surface of graphene to gain the GO with good dispersibility, and finally reduces GO by different reduction methods according to requirements for obtaining the graphene with desired size and thickness [53].
Oxidation is a key step in the preparation of graphene by the chemical redox method. The preparation methods of GO mainly include the Brodie, Staudenmaier, and Hummer’s methods [54,55]. Researchers are constantly researching and improving the preparation method of GO, especially Hummer’s method. For example, Marcano et al. [56] reported the improved Hummer’s method for the preparation of graphene. They increased the dosage of concentrated sulfuric acid and potassium permanganate without adding sodium nitrate and oxidized graphite in a mixture of sulfuric and phosphoric acid. The detailed procedure is listed in Figure 5. The advantage of this improved Hummer’s method was that it would not release toxic gases, such as NO2, N2O4, and had relatively higher yields, but the addition of phosphoric acid would increase the production cost and cause a series of environmental problems.

2.2.3. Chemical Vapor Deposition Method

Chemical vapor deposition (CVD) is a controllable method for preparing graphene [57]. The principle is that the reaction occurs under high temperature and gaseous conditions by using thin metal films and single metal crystals or polycrystals as a plane substrate, then carbon atoms are deposited on the surface of substrate through high–temperature annealing, and finally, the plane substrate is removed by corrosion and other methods to obtain graphene [58]. Nam et al. [59] utilized the different platinum (Pt) substrates, including sputter–deposited films, e–beam deposited films, and polycrystalline Pt foil, to serve as the substrate for the low–pressure CVD growth of graphene. The results illustrated that the Pt thin films deposited by high–temperature sputtering were the best Pt substrate for the growth of homogeneous, single–layer graphene with a thickness down to 25 nm. Dato et al. [60] proposed a novel method to synthesize graphene sheets in the gas phase in a substrate–free pressure microwave plasma reactor. Graphene sheets were synthesized by passing liquid ethanol droplets into an argon plasma. The results demonstrated that the graphene could be fabricated without three–dimensional materials or substrates, suggesting a possible route to large–scale synthesis of graphene. CVD is suitable for the large–scale preparation of graphene and is the most widely used method in the industry [61]. This method has the characteristics of easy operation and high controllability and is currently the main preparation method in the field of graphene growth.

2.2.4. Epitaxial Growth Method

Epitaxial growth is a method for growing new crystal layers based on crystal layers as the substrate [62]. The volatility and coagulation of the substrate play a decisive role in the growth of graphene on the substrate. When the matrix leaks and becomes more volatile than condensable, the graphene could be generated from carbon atoms on the matrix [63]. Berger et al. [64] found that the silicon atoms on the surface of silicon carbide were vaporized when heated under high vacuum or normal pressure over 1000 °C, and the remaining carbon atoms would re–aggregate during the cooling process and form the graphene on the surface of the inner layer of silicon carbide. This is the main mechanism for the epitaxial growth method. According to the types of selected substrates, the epitaxial growth method can be divided into the silicon carbide epitaxial growth method and the metal surface epitaxial growth method, respectively. Both methods have their advantages and disadvantages. The silicon carbide epitaxial growth method can produce high–quality single–layer or few–layer graphene, but the energy consumption required is too high, which is not suitable for large–scale production, and the subsequent process of transferring graphene is complicated. On the contrary, the metal surface epitaxial growth method can achieve the purpose of large–scale preparation of graphene by adjusting the parameters. However, the quality of graphene still needs to be improved.

2.2.5. Electrochemical Method

The electrochemical method is a new method for preparing graphene developed in recent years [65]. As demonstrated in Figure 6A, the graphite could act as the working electrode to produce various graphene and graphene–based materials in the presence of electrolytes [66]. Its principle is to drive the charged anisotropic ions through electrolysis to expand the graphite electrode inserted into the mixture, thereby exfoliating graphene. As illustrated in the electrochemical exfoliation of Figure 6B, when the power supply is applied to the graphite electrode, the positively charged ions (e.g., Li+) or negatively charged ions (e.g., SO42−) would be attracted to the graphite electrode in the cathodic or anodic exfoliation, respectively [67]. The ions, electrolyte molecules, or co–intercalating species in electrolytes would be firstly intercalated into the graphene layers of graphite to yield the graphene and graphene–based materials. The main advantages of this method are that the voltage and current can be precisely adjusted, the preparation process is reproducible, the reaction is easy to be controlled, and the cost is low. Nevertheless, the morphology and structures of graphene fabricated by this method are prone to the formation of defects. Therefore, for this method, there are still many technical barriers to be broken through. Tang et al. [68] prepared a few–layers of water–dispersible graphene with ultrafiltration of up to 99%, a thickness of less than six atomic layers, and a lateral size of 1~50 μm by rationally designing the electrolyte (NaOH + p–phthalic acid) and selecting highly oriented pyrolytic graphite (HOPG) as anode material. Meanwhile, the as–prepared graphene exhibited a high dispersibility (2.5 mg mL−1) and excellent stability (over six months).

2.2.6. Organic Synthesis Method

The organic synthesis method is a powerful alternative in fabricating graphene–type structures [69], which uses the traditional organic synthesis method by starting from small organic molecules through a gradual reaction to synthesize graphene. Organic synthesis is a very controllable synthesis method. In this method, graphene with different sizes and thicknesses can be produced by designing synthesis routes. Dössel et al. [70] synthesized graphene nanoribbons with perfect structure by designing a synthetic route in the organic synthesis method. Unfortunately, there are still problems, such as many reaction steps, a long reaction time, and more materials required when graphene is synthesized on a large scale by this method.

3. Graphene–Based Materials for Heavy Metal Removal

3.1. Influencing Factors on Heavy Metal Removal

Some influencing factors on heavy metal removal, including solution pH, adsorbent dose, heavy metal concentration, treatment time, and temperature, are listed below.

3.1.1. Solution pH

Solution pH is an important parameter to affect the adsorption of heavy metals since the heavy metals exhibit pH–dependent chemistry. Taking Cr(VI) as an example, there are several species of Cr(VI) in an aqueous solution with different pH values. In general, as a strong acid, in the aqueous solution, H2CrO4 could be ionized to the form of HCrO4 and CrO42−. In the pH < 2 solutions, H2CrO4 and HCrO4 are present; in the pH range of 2~4, HCrO4 is dominating form; when the solution’s pH is higher than 4, CrO42− appears, and its content rises with the increase of the solution’s pH; when the pH is higher than 9, only CrO42− exists in the solution. Thus, Cr(VI) removal performance by the adsorbents is highly dependent on the existing form of Cr(VI), especially for the adsorbent with surface functional groups. For instance, the amino group functionalized adsorbents favour chelating with HCrO4 rather than H2CrO4 and CrO42−. As a result, the amino group functionalized adsorbents could not provide good Cr(VI) removal performance in a strong acidic and alkaline solution [71].

3.1.2. Adsorbent Dose

In the adsorption process, the adsorbent dose could affect the active sites of adsorbents to the adsorption of heavy metals; thus, the heavy metal removal performance normally increases with coinciding increases in the adsorbent doses [72].

3.1.3. Heavy Metal Concentrations

The heavy metal concentrations have an obvious influence on heavy metal removal. This is because as the heavy metal concentrations increase, the active sites on adsorbents may gradually become saturated by the heavy metals, wherein the adsorbents cannot hold on to the excess heavy metals, leading to the decreased heavy metal removal performance or adsorption equilibrium [73].

3.1.4. Treatment Time and Kinetics

Treatment time is a key factor for heavy metal removal by adsorbents since it is associated with the adsorption kinetics, which is very significant in designing the adsorption reactors in the industry [74]. The kinetics could determine the heavy metal uptake rate for controlling the residence time of adsorbate uptake at the solid–liquid interface. Generally, there are two models to describe the adsorption kinetics: pseudo–first order and pseudo–second order, as listed in Equations (1) and (2), respectively [75].
ln ( Q e Q t ) = ln Q e k 1 × t
( t Q t ) = 1 k 2 Q e 2 + t Q e
where k1 is the adsorption rate constant of the pseudo–first–order model (min−1), k2 is the adsorption rate constant of the pseudo–second–order model (g mg−1 min−1), Qe stands for the adsorption capacity at equilibrium time (mg g−1), and Qt is the amount of metal adsorbed at time t (mg g−1).

3.1.5. Temperature and Thermodynamics

The temperature could affect the adsorption process of heavy metals. As the adsorption process is exothermic, the subsequent increasing temperature is not beneficial for adsorption. Conversely, the rising temperature is favorable to the adsorption in the endothermic adsorption process. The temperature effect on the adsorption is related to the adsorption isotherm [76]. The adsorption isotherm delineates the adsorption equilibrium of adsorbate at the surface of an adsorbent. There are two commonly used models, including Langmuir and Freundlich isotherm models, to explain the solid–liquid adsorption process for thermodynamics. They can be expressed as Equations (3) and (4), respectively [77].
q e = a b C e 1 + b C e
q e = k f C e η
where Ce is the equilibrium concentration (mg L−1) of heavy metal, qe is the adsorbed heavy metal amount at equilibrium (mg g−1), a (mg g−1) and b (L mg−1) are Langmuir isotherm parameters, kf is the Freundlich equilibrium constant, indicating the extent of adsorption, and η is the power term of the Freundlich isotherm and the heterogeneity factor, illustrating the intensity of adsorption. The Langmuir model means the adsorption occurs at the monomolecular layer, whereas the Freundlich model focuses on the adsorption on an energetically heterogeneous surface.

3.2. Heavy Metal Removal Performance by Graphene–Based Materials

Graphene and GO have attracted much more attention in the field of heavy metal removal because of their large specific surface area. GO especially has various active oxygen–containing functional groups on its surface, which can provide many effective active sites for the adsorption of heavy metals. For instance, Rout et al. [78] used hydrazine monohydrate to reduce GO for producing the reduced GO (RGO) with a mean pore diameter of 3.08 nm and sharp peaks in the pore size distribution between 3.08 and 23.88 nm. They found that the maximum adsorption capacity of phenol by RGO achieved 602.41 mg g−1 with the RGO amount of 0.4 g L−1, solution pH of 8.0, and adsorption time of 75 min at 30 °C. Wang et al. [79] compared the adsorption and co–adsorption of naphthalene, 1–naphthol, and Cd(II) on GO, chemically reduced graphene (CRG), and annealing reduced graphene (ARG) to determine the unique adsorption properties of graphene nanosheets. The results revealed that the order of adsorption capacity to organic pollutants was CRG > ARG > GO, and the adsorption of 1–naphthol by these three adsorbents was stronger than that of naphthalene. GO, with more functional groups on CRG and ARG, exhibited a strong affinity to Cd(II). The adsorption of Cd(II) onto GO and CRG facilitated the co–adsorption of naphthalene and 1–naphthol via a surface bridging mechanism. Notably, even though ARG did not display obvious Cd(II) adsorption, the suppressed co–adsorption of naphthalene on ARG might be from the sieving effect of hydrated Cd(II) binding to the micropore edges on ARG. This study not only reflected the good adsorption properties of graphene materials but also supported the conclusion that graphene materials could simultaneously adsorb heavy metals and organic pollutants in wastewater. Xue et al. [80] explored GO nanosheets for the adsorption of Cd(II) in acidic aqueous solutions. The results illustrated that the removal efficiency of Cd(II) was increased from 6.29 to 96.72% by increasing the adsorbent dosage from 0.02 to 2.00 g L−1. The adsorption capacity of Cd(II) was greatly enhanced when the solution pH was increased from 2.02 to 4.01. Moreover, the reaction reached adsorption equilibrium within 120 min, and the maximum Cd(II) adsorption capacity was about 44.64 mg g−1 at the adsorbent dosage of 0.50 g L−1. The experimental results of this work well demonstrated that graphene materials had good adsorption properties to Cd(II) in suitable conditions.

4. Modified Graphene Materials for Heavy Metal Removal in the Tobacco Industry

4.1. Modification of Graphene Materials

Even though graphene and GO could be widely applied in heavy metal removal, as aforementioned above, the abundant oxygen–containing functional groups on the surface of GO make it well–dispersed in the water. As a result, they are difficult to be separated from wastewater after treatment. Therefore, it is required to modify the graphene and GO materials to solve this problem. Commonly, there are two types of modification methods for graphene and GO materials: covalent bond modification and non–covalent bond modification [81].
The covalent bond modification mainly utilizes the active groups on its surface to chemically react with other molecules to achieve the modification of graphene. According to the different reaction molecules, covalent bond modification can be divided into small organic molecule modification and polymer modification. The small organic molecule modification occurs through the reaction of a small organic molecule with functional groups on the surface of GO. It mainly contains the ring–opening reaction of the hydroxyl and epoxy groups, the diazotization reaction of a conjugated plane, the isocyanate and cycloaddition reactions, etc. Stankovich et al. [82] treated GO with organic isocyanates. The degree of GO modification could be attained by controlling the reactivity of isocyanates or the reaction time. The polymer modification introduces the long polymer chains to change the structural properties of GO.
The non–covalent bond modification is generally accomplished by physical adsorption or polymer wrapping on the surface of GO. This method could maintain the structural properties of GO without destroying its intrinsic structure and improving its solubility. The non–covalent bond modification of GO mostly includes a ππ bond and hydrogen bonding functionalization. Both graphene and GO have highly conjugated systems, which makes it easy to interact with molecules with a π conjugated structure through ππ interaction to achieve modification. Yang et al. [83] manufactured a novel GO–doxorubicin hydrochloride nanohybrid (GO–DXR) by the hydrogen bonding and strong ππ stacking interaction between DXR and GO. Therefore, the loading and release behaviours of DXR on GO disclosed a strong pH–dependent property.
The preparation of magnetic GO nanocomposites is also a good way to solve the separation problem after treatment. By introducing the magnetic nanoparticles onto GO, the adsorbents could be efficiently separated from the solution by a permeant magnet after adsorption of heavy metals. This allows the magnetic GO nanocomposites to be applied in the field of wastewater treatment potentially. The commonly used methods for the fabrication of magnetic GO nanocomposites include the self–assembly method, sol–gel method, chemical co–precipitation method, chemical grafting method, hydrothermal method, high–temperature annealing process, etc.
For example, Adel et al. [84] manufactured the magnetic separable MgFe2O4/crumbled reduced GO nanoparticles by self–assembly method via electrostatic interactions. Gabris et al. [85] synthesized the silica–cyanopropyl functionalized magnetic GO (MGO/SiO2–CN) hybrid nanomaterial derived by the sol–gel method. Cai et al. [86] prepared magnetic carbon microfibers supporting iron/nickel bisalloy (FeNi@CMFs) in a single step by using the electrospinning method combined with the high–temperature annealing method at an optimal condition of 5 wt% mass fractions of metal salts (ferric nitrate and nickel nitrate) annealed at 800 °C for 60 min in 5% (v/v) hydrogen/argon atmosphere.
Wang et al. [41] treated GO with polyacrylic acid (PAA) in order to bring the carboxylic groups to the surface of GO. After that, the GO functionalized with carboxylic groups reacted with the magnetite (Fe3O4) nanoparticles modified by amine groups through chemical grafting to produce the magnetic GO nanocomposites, as shown in Figure 7.
The modification of graphene and GO could solve the dispersion problem of GO in water and assist the GO to be separated from water more easily, which could make GO be better exploited in heavy metal removal and wastewater treatment.

4.2. Modified Graphene Materials for Heavy Metal Removal

In recent years, modified graphene materials have received extensive attention in the field of heavy metal removal. The treatment of heavy metals in wastewater by modified graphene materials mainly contributes to physical adsorption and redox reaction. Physical adsorption directly adsorbs heavy metals by using their physical properties of a large specific surface area. Redox reaction is to reduce the valence state of heavy metals through redox reaction after adsorption, which is mostly applied to treat heavy metals with high chemical valence and high toxicity in wastewater.
Fan et al. [87] used magnetic chitosan (MC) to modify GO for manufacturing a magnetic chitosan/graphene oxide (MCGO) material by a facile and fast process to remove Pb(II) from an aqueous solution. The results illustrated that the MC was assembled on the surface of GO layer in a high–density form and could obviously increase the specific surface area of GO. Furthermore, the adsorption of Pb(II) by MCGO was observed to be strongly solution pH–dependent and mainly ascribed to the abundant functional groups on the surface of MCGO. The maximum adsorption capacity of Pb(II) was estimated to be 76.94 mg g−1, as gained from the Langmuir model. This newly developed MCGO was a good adsorbent material for Pb(II) removal. Kumar et al. [88] produced a novel chitosan functionalized GO material for the efficient adsorption of As(V)/As(III). The results revealed that the maximum adsorption capacity of As(V) and As(III) by chitosan functionalized GO material was 64.27 mg g−1 and 71.9 mg g−1, respectively. Abaszadeh et al. [89] adopted Fe3O4 nanoparticles and 5–amino–1,10–phenanthroline (APhen) to functionalize GO in the presence of N,N′–dicyclohexylcarbodiimide (DCC) to produce magnetic nanocomposites of MGO@APhen, which exhibited a high Pb(II) removal efficiency up to 97.2% and excellent reusability during six cycles after being recycled by a supermagnet.
Chi et al. [90] prepared a novel magnetic surface–ion–imprinted polymer (IIP) by using As(III) as a template and GO/Fe3O4 as support, Figure 8A, which displayed a high selectivity of As(III) in the presence of other competing components, such as As(III)/Mg(II), As(III)/Ca(II), As(III)/HCO3, As(III)/PO43−, As(III)/SO42−, and As(III)/HA, have excellent reusability with an adsorption efficiency of 75% after five cycles of regeneration and reuse (Figure 8B). Xing et al. [91] modified magnetic GO/Fe3O4 (MGO) with ethylenediaminetetraacetic acid (EDTA) groups by the silanization reaction between N–(trimethoxysilylpropyl) ethylenediaminetriacetic acid and –OH on the surface of MGO (MGO–EDTA), as shown in Figure 9, which performed a high Pb(II) adsorption capacity of up to 211.3 mg g−1 coming from the chelating ability of EDTA.
In summary, the modified GO materials express excellent heavy metal adsorption properties relative to GO. Researchers have devoted themselves to the various modified GO materials for heavy metal wastewater treatment. However, the selectivity of heavy metals in the treatment by modified GO materials is rarely discussed. Thus, an extensive investigation on enhancing the selectivity of heavy metals for modified GO materials is still required.

4.3. Modified Graphene Materials for Heavy Metal Removal in the Tobacco Industry

There are more than 8000 kinds of chemical substances that can be detected in the smoke produced by cigarettes during the combustion process [92]. In 1990, Hoffman from the United States proposed 43 kinds of substances, such as As, Cd, Cr, Pb, Ni, Hg, and others, that were most harmful to the human body and were called the Hoffmann list. Among them, these heavy metals could be accumulated in the organs of the human body, causing serious harm to human health [93]. The heavy metals in cigarette smoke mainly come from two aspects. On the one hand, soil and air pollution, improper use of pesticides and fertilizers, irrigation water pollution, and other factors during the planting process of tobacco plants exceed the standard of heavy metals in tobacco leaves. On the other hand, the heavy metals may be from the auxiliary materials used in cigarettes, such as cigarette paper, filter rods, additives, etc. When the cigarette is burned, the heavy metals are released and mixed into the smoke, further entering the human body. Although New Tobacco Products such as Electronic Nicotine Delivery Systems (ENDS), i.e., e–cigarettes, are now emerging, those products still have risks due to the limitations of product technology, form, and appearance. The smoke also contains a large number of heavy metal particles, which will also result in harm to the human body [94]. Lerner et al. [95] found that copper was 6.1 times higher than previously reported for conventional cigarette smoke in e–cigarettes per puff. They believed that the detection and removal of heavy metals from e–cigarettes and their components should raise concerns about the safety of their usage and the disposal of e–cigarette waste.
Graphene and its composite materials as adsorbents or additives have important application prospects in the fields of smoke composition detection, removal of heavy metals in cigarette materials, and New Tobacco Products. For example, Palisoc et al. [96] used graphene to fabricate an effective gold nanoparticle/graphene/Nafion–modified glassy carbon electrode, which had more merits for the detection of Pb(II), Cd(II) and Cu(II) in cigarettes. Guo et al. [97] established an electrochemical sensor consisting of rGO/MoS2/chitosan (CS) nanocomposite–modified glassy carbon electrode, which exhibited wide detection sensitivity for Pb(II) in tobacco leaves in the range of 0.005–0.05–2.0 μM, with excellent performance in reproducibility, stability and anti–interference ability. The limit of detection (LOD) was 0.0016 μM. It provided a new approach to the determination of Pb(II) in tobacco leaves with low content.
Taking the removal of heavy metals from smoke as an example, Yu et al. [98] coated the lyophilized oxidized multiwalled carbon nanotubes/GO composite on the surface of cellulose acetate in the cigarette filter. When the weight ratio of the oxidized multiwalled carbon nanotubes: GO: filter was 1:1:62, the modified filter demonstrated an excellent adsorption performance with the removal efficiency of 83 and 78% to Cd(II) and Cr(VI), respectively, Figure 10.
At present, the research on graphene and its composite materials in the field of the tobacco industry for the detection and removal of heavy metals has been relatively extensive. It is believed that graphene and its composite materials will also have huge application space and development prospects in the tobacco industry in the future.
In order to better understand the removal efficiency of GO and GO–based materials to heavy metals, the synthesis methods and removal efficiency, as well as adsorption conditions, are laid out in Table 1.

5. Conclusions, Challenges, and Perspectives

With the rapid development of society and industry, heavy metal wastewater pollution has become more and more serious, resulting in a tighter supply and demand for water resources. Consequently, it is urgent to solve the heavy metal wastewater pollution problem. The graphene–based materials with unique physicochemical properties have received intensive attention in the field of heavy metal wastewater treatment, especially, in the tobacco industry. Recently, the application of graphene–based materials for heavy metal treatment has made a great progress, but there is still a long way to go to face the heavy metal wastewater treatment problem. In view of the current research status of graphene–based materials in the field of heavy metal wastewater treatment, the main work in future research is to design and synthesize novel functionalized graphene–based composite materials with high heavy metal adsorption capacities. It will be basically developed in three directions. First, the more modified graphene–based materials with different functions should be manufactured to attain relatively stronger adsorption properties. Secondly, the research on the effect of competing for heavy metal ions in wastewater on the adsorption of graphene–based materials should be deepened. At present, researchers have mostly studied the effect of functionalized graphene materials on a certain heavy metal. However, there are often multiple heavy metal ions in real wastewater. In this situation, the functionalized graphene materials for treating a single heavy metal are difficult to be applied in a real wastewater system. It is more practical to develop functionalized graphene–based materials to remove the multi–component heavy metal ions. Thirdly, the cost for the current modification method of graphene–based materials is relatively high, which severely limits the applications of graphene–based materials in practical application. To seek a new method for the preparation of graphene–based materials at a low cost is the future pursuit for heavy metal treatment in the tobacco industry on a large scale.

Author Contributions

Writing—original draft preparation, X.X., J.Z. and Y.W.; conceptualization, writing—review and editing, H.G. and S.W.; supervision, H.G.; project administration, H.G.; funding acquisition, H.G. and Q.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the support and funding from the Fundamental Research Funds for the Central Universities (2022-4-YB-03). This work is supported by Shanghai Science and Technology Commission (19DZ2271500).

Data Availability Statement

Not applicable.

Acknowledgments

We thank the support from the Shanghai Key Lab. of Chemical Assessment and Sustainability.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

CrChromium
Cdcadmium
Asarsenic
Hgmercury
Pblead
Ptplatinum
GOgraphene oxide
CVDchemical vapor deposition
HOPGhighly oriented pyrolytic graphite
RGOreduced GO
CRGchemically reduced graphene
ARGannealing reduced graphene
GO–DXRGO–doxorubicin hydrochloride nanohybrid
MCGOmagnetic chitosan/graphene oxide
APhen5–amino–1,10–phenanthroline
DCCN,N′–dicyclohexylcarbodiimide
MGOmagnetic graphene oxide
EDTAethylenediaminetetraacetic acid
ENDSElectronic Nicotine Delivery Systems
LODlimit of detection

References

  1. Ma, J.; Qin, G.; Zhang, Y.; Sun, J.; Wang, S.; Jiang, L. Heavy metal removal from aqueous solutions by calcium silicate powder from waste coal fly-ash. J. Clean. Prod. 2018, 182, 776–782. [Google Scholar] [CrossRef]
  2. Guo, L.; Zhang, Y.; Zheng, J.; Shang, L.; Shi, Y.; Wu, Q.; Liu, X.; Wang, Y.; Shi, L.; Shao, Q. Synthesis and characterization of ZnNiCr-layered double hydroxides with high adsorption activities for Cr(VI). Adv. Compos. Hybrid Mater. 2021, 4, 819–829. [Google Scholar] [CrossRef]
  3. Wang, B.; Wu, T.; Angaiah, S.; Murugadoss, V.; Ryu, J.E.; Wujcik, E.K.; Lu, N.; Young, D.P.; Gao, Q.; Guo, Z. Development of nanocomposite adsorbents for heavy metal removal from wastewater. ES Mater. Manuf. 2018, 2, 35–44. [Google Scholar] [CrossRef]
  4. Deng, Z.; Sun, S.; Li, H.; Pan, D.; Patil, R.R.; Guo, Z.; Seok, I. Modification of coconut shell-based activated carbon and purification of wastewater. Adv. Compos. Hybrid Mater. 2021, 4, 65–73. [Google Scholar] [CrossRef]
  5. Ouni, L.; Ramazani, A.; Fardood, S.T. An overview of carbon nanotubes role in heavy metals removal from wastewater. Front. Chem. Sci. Eng. 2019, 13, 274–295. [Google Scholar] [CrossRef]
  6. Yin, C.; Wang, C.; Hu, Q. Selective removal of As(V) from wastewater with high efficiency by glycine-modified Fe/Zn-layered double hydroxides. Adv. Compos. Hybrid Mater. 2021, 4, 360–370. [Google Scholar] [CrossRef]
  7. Si, Y.; Li, J.; Cui, B.; Tang, D.; Yang, L.; Murugadoss, V.; Maganti, S.; Huang, M.; Guo, Z. Janus phenol-formaldehyde resin and periodic mesoporous organic silica nanoadsorbent for the removal of heavy metal ions and organic dyes from polluted water. Adv. Compos. Hybrid Mater. 2022, 5, 1180–1195. [Google Scholar] [CrossRef]
  8. Chai, J.; Hu, Q.; Qiu, B. Conductive polyaniline improves Cr(VI) bio-reduction by anaerobic granular sludge. Adv. Compos. Hybrid Mater. 2021, 4, 1137–1145. [Google Scholar] [CrossRef]
  9. Bashir, A.; Malik, L.A.; Ahad, S.; Manzoor, T.; Bhat, M.A.; Dar, G.N.; Pandith, A.H. Removal of heavy metal ions from aqueous system by ion-exchange and biosorption methods. Environ. Chem. Lett. 2019, 17, 729–754. [Google Scholar] [CrossRef]
  10. Khulbe, K.C.; Matsuura, T. Removal of heavy metals and pollutants by membrane adsorption techniques. Appl. Water Sci. 2018, 8, 19. [Google Scholar] [CrossRef]
  11. Bashir, M.S.; Ramzan, N.; Najam, T.; Abbas, G.; Gu, X.; Arif, M.; Qasim, M.; Bashir, H.; Shah, S.S.A.; Sillanpää, M. Metallic nanoparticles for catalytic reduction of toxic hexavalent chromium from aqueous medium: A state-of-the-art review. Sci. Total Environ. 2022, 829, 154475. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, D.S.; Gao, W.Q.; Chang, G.Z.; Luo, S.; Jiao, W.Z.; Liu, Y.Z. Removal of heavy metal lead(II) using nanoscale zero-valent iron with different preservation methods. Adv. Powder Technol. 2019, 30, 581–589. [Google Scholar] [CrossRef]
  13. Nazir, M.A.; Najam, T.; Shahzad, K.; Wattoo, M.A.; Hussain, T.; Tufail MKShah, S.S.A.; Rehman, A.U. Heterointerface engineering of water stable ZIF-8@ZIF-67: Adsorption of rhodamine B from water. Surf. Interfaces 2022, 34, 102324. [Google Scholar] [CrossRef]
  14. Nazir, M.A.; Najam, T.; Jabeen, S.; Wattoo, M.A.; Bashir, M.S.; Shah, S.S.A.; Rehman, A.U. Facile synthesis of Tri-metallic layered double hydroxides (NiZnAl-LDHs): Adsorption of Rhodamine-B and methyl orange from water. Inorg. Chem. Commun. 2022, 145, 110008. [Google Scholar] [CrossRef]
  15. Xu, X.; Zhang, H.; Ma, C.; Gu, H.; Lou, H.; Lyu, S.; Liang, C.; Kong, J.; Gu, J. A superfast hexavalent chromium scavenger: Magnetic nanocarbon bridged nanomagnetite network with excellent recyclability. J. Hazard. Mater. 2018, 353, 166–172. [Google Scholar] [CrossRef] [PubMed]
  16. Anastopoulos, I.; Robalds, A.; Tran, H.N.; Mitrogiannis, D.; Giannakoudakis, D.A.; Hosseini-Bandegharaei, A.; Dotto, G.L. Removal of heavy metals by leaves-derived biosorbents. Environ. Chem. Lett. 2019, 17, 755–766. [Google Scholar] [CrossRef]
  17. Xie, X.; Gao, H.; Luo, X.; Zhang, Y.; Qin, Z.; Ji, H. Polyethyleneimine-modified magnetic starch microspheres for Cd(II) adsorption in aqueous solutions. Adv. Compos. Hybrid Mater. 2022, 5, 2772–2786. [Google Scholar] [CrossRef]
  18. Yu, G.; Lu, Y.; Guo, J.; Patel, M.; Bafana, A.; Wang, X.; Qiu, B.; Jeffryes, C.; Wei, S.; Guo, Z.; et al. Carbon nanotubes, graphene, and their derivatives for heavy metal removal. Adv. Compos. Hybrid Mater. 2018, 1, 56–78. [Google Scholar] [CrossRef]
  19. Yuan, X.; An, N.; Zhu, Z.; Sun, H.; Zheng, J.; Jia, M.; Lu, C.; Zhang, W.; Liu, N. Hierarchically porous nitrogen-doped carbon materials as efficient adsorbents for removal of heavy metal ions. Process Saf. Environ. Prot. 2018, 119, 320–329. [Google Scholar] [CrossRef]
  20. Huang, Y.; Zeng, X.; Guo, L.; Lan, J.; Zhang, L.; Cao, D. Heavy metal ion removal of wastewater by zeolite-imidazolate frameworks. Sep. Purif. Technol. 2018, 194, 462–469. [Google Scholar] [CrossRef]
  21. Dong, Z.; Zhang, F.; Wang, D.; Liu, X.; Jin, J. Polydopamine-mediated surface-functionalization of graphene oxide for heavy metal ions removal. J. Solid State Chem. 2015, 224, 88–93. [Google Scholar] [CrossRef]
  22. Vilela, D.; Parmar, J.; Zeng, Y.; Zhao, Y.; Sánchez, S. Graphene-based microbots for toxic heavy metal removal and recovery from water. Nano Lett. 2016, 16, 2860–2866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Lim, J.Y.; Mubarak, N.M.; Abdullah, E.C.; Nizamuddin, S.; Khalid, M.; Inamuddin. Recent trends in the synthesis of graphene and graphene oxide based nanomaterials for removal of heavy metals—A review. J. Ind. Eng. Chem. 2018, 66, 29–44. [Google Scholar] [CrossRef]
  24. Cai, J.; Tian, J.; Gu, H.; Guo, Z. Amino carbon nanotube modified reduced graphene oxide aerogel for oil/water separation. ES Mater. Manuf. 2019, 6, 68–74. [Google Scholar] [CrossRef]
  25. Deng, H.; Yin, J.; Ma, J.; Zhou, J.; Zhang, L.; Gao, L.; Jiao, T. Exploring the enhanced catalytic performance on nitro dyes via a novel template of flake-network Ni-Ti LDH/GO in-situ deposited with Ag3PO4 NPs. Appl. Surf. Sci. 2021, 543, 148821. [Google Scholar] [CrossRef]
  26. Xie, W.; Yao, F.; Gu, H.; Du, A.; Lei, Q.; Naik, N.; Guo, Z. Magnetoresistive and piezoresistive polyaniline nanoarrays in-situ polymerized surrounding magnetic graphene aerogel. Adv. Compos. Hybrid Mater. 2022, 5, 1003–1016. [Google Scholar] [CrossRef]
  27. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [Green Version]
  28. Menazea, A.A.; Ezzat, H.A.; Omara, W.; Basyouni, O.H.; Ibrahim, S.A.; Mohamed, A.A.; Tawfik, W.; Ibrahim, M.A. Chitosan/graphene oxide composite as an effective removal of Ni, Cu, As, Cd and Pb from wastewater. Comput. Theor. Chem. 2020, 1189, 112980. [Google Scholar] [CrossRef]
  29. Liu, H.; Mao, Y. Graphene oxide-based nanomaterials for uranium adsorptive uptake. ES Mater. Manuf. 2021, 13, 3–22. [Google Scholar] [CrossRef]
  30. Sherlala, A.I.A.; Raman, A.A.A.; Bello, M.M.; Asghar, A. A review of the applications of organo-functionalized magnetic graphene oxide nanocomposites for heavy metal adsorption. Chemosphere 2018, 193, 1004–1017. [Google Scholar] [CrossRef]
  31. Peng, W.J.; Li, H.Q.; Liu, Y.Y.; Song, S.X. A review on heavy metal ions adsorption from water by graphene oxide and its composites. J. Mol. Liq. 2017, 230, 496–504. [Google Scholar] [CrossRef]
  32. Abu-Nada, A.; McKay, G.; Abdala, A. Recent advances in applications of hybrid graphene materials for metals removal from wastewater. Nanomaterials 2020, 10, 595. [Google Scholar] [CrossRef] [Green Version]
  33. De Beni, E.; Giurlani, W.; Fabbri, L.; Emanuele, R.; Santini, S.; Sarti, C.; Martellini, T.; Piciollo, E.; Cincinelli, A.; Innocenti, M. Graphene-based nanomaterials in the electroplating industry: A suitable choice for heavy metal removal from wastewater. Chemospehre 2022, 292, 133448. [Google Scholar] [CrossRef] [PubMed]
  34. Xu, L.; Wang, J. The application of graphene-based materials for the removal of heavy metals and radionuclides from water and wastewater. Crit. Rev. Environ. Sci. Technol. 2017, 47, 1042–1105. [Google Scholar] [CrossRef]
  35. Kong, Q.; Shi, X.; Ma, W.; Zhang, F.; Yu, T.; Zhao, F.; Zhao, D.; Wei, C. Strategies to improve the adsorption properties of graphene-based adsorbent towards heavy metal ions and their compound pollutants: A review. J. Hazard. Mater. 2021, 415, 125690. [Google Scholar] [CrossRef]
  36. Xu, J.; Cao, Z.; Zhang, Y.; Yuan, Z.; Lou, Z.; Xu, X.; Wang, X. A review of functionalized carbon nanotubes and graphene for heavy metal adsorption from water: Preparation, application, and mechanism. Chemosphere 2018, 195, 351–364. [Google Scholar] [CrossRef]
  37. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  38. Liu, R.; Zhang, Y.; Ning, Z.; Xu, Y. A catalytic microwave process for superfast preparation of high-quality reduced graphene oxide. Angew. Chem. Int. Ed. 2017, 56, 15677–15682. [Google Scholar] [CrossRef] [PubMed]
  39. Fallah, S.; Mamaghani, H.R.; Yegani, R.; Hajinajaf, N.; Pourabbas, B. Use of graphene substrates for wastewater treatment of textile industries. Adv. Compos. Hybrid Mater. 2020, 3, 187–193. [Google Scholar] [CrossRef]
  40. He, H.; Klinowski, J.; Forster, M.; Lerf, A. A new structural model for graphite oxide. Chem. Phys. Lett. 1998, 287, 53–56. [Google Scholar] [CrossRef]
  41. Wang, Y.; Xie, W.; Liu, H.; Gu, H. Hyperelastic magnetic reduced graphene oxide three-dimensional framework with superb oil and organic solvent adsorption capability. Adv. Compos. Hybrid Mater. 2020, 3, 473–484. [Google Scholar] [CrossRef]
  42. Osman, A.; Elhakeem, A.; Kaytbay, S.; Ahmed, A. A comprehensive review on the thermal, electrical, and mechanical properties of graphene-based multi-functional epoxy composites. Adv. Compos. Hybrid Mater. 2022, 5, 547–605. [Google Scholar] [CrossRef]
  43. Zhu, Q.; Huang, Y.; Li, Y.; Zhou, M.; Xu, S.; Liu, X.; Liu, C.; Yuan, B.; Guo, Z. Aluminum dihydric tripolyphosphate/polypyrrole-functionalized graphene oxide waterborne epoxy composite coatings for impermeability and corrosion protection performance of metals. Adv. Compos. Hybrid Mater. 2021, 4, 780–792. [Google Scholar] [CrossRef]
  44. Şenel, M.C.; Gürbüz, M. Synergistic effect of graphene/boron nitride binary nanoparticles on aluminum hybrid composite properties. Adv. Compos. Hybrid Mater. 2021, 4, 1248–1260. [Google Scholar] [CrossRef]
  45. Al-Gaashani, R.; Najjar, A.; Zakaria, Y.; Mansour, S.; Atieh, M.A. XPS and structural studies of high quality graphene oxide and reduced graphene oxide prepared by different chemical oxidation methods. Ceram. Int. 2019, 45, 14439–14448. [Google Scholar] [CrossRef]
  46. Backes, C.; Abdelkader, A.M.; Alonso, C.; Andrieux-Ledier, A.; Arenal, R.; Azpeitia, J.; Balakrishnan, N.; Banszerus, L.; Barjon, J.; Bartali, R.; et al. Production and processing of graphene and related materials. 2D Mater. 2020, 7, 022001. [Google Scholar] [CrossRef]
  47. Sumdani, M.G.; Islam, M.R.; Yahaya, A.N.A.; Safie, S.I. Recent advances of the graphite exfoliation processes and structural modification of graphene: A review. J. Nanopart. Res. 2021, 23, 253. [Google Scholar] [CrossRef]
  48. Bohm, S.; Ingle, A.; Bohm, H.L.M.; Fenech-Salerno, B.; Wu, S.; Torrisi, F. Graphene production by cracking. Philos. Trans. A Math. Phys. Eng. Sci. 2021, 379, 20200293. [Google Scholar] [CrossRef]
  49. Sinclair, R.C.; Suter, J.L.; Coveney, P.V. Micromechanical exfoliation of graphene on the atomistic scale. Phys. Chem. Chem. Phys. 2019, 21, 5716–5722. [Google Scholar] [CrossRef] [Green Version]
  50. Deng, S.; Qi, X.-D.; Zhu, Y.-L.; Zhou, H.-J.; Chen, F.; Fu, Q. A facile way to large-scale production of few-layered graphene via planetary ball mill. Chin. J. Polym. Sci. 2016, 34, 1270–1280. [Google Scholar] [CrossRef]
  51. Wang, R.; Liu, Y.; Zhang, Y.; Wang, L.; Yang, G.; Shen, F.; Deng, S.; Zhang, X.; He, Y.; Luo, L. A simplified chemical reduction method for preparation of graphene: Dispersity, reducibility and mechanism. Ceram. Int. 2016, 42, 19042–19046. [Google Scholar] [CrossRef]
  52. Ren, S.; Rong, P.; Yu, Q. Preparations, properties and applications of graphene in functional devices: A concise review. Ceram. Int. 2018, 44, 11940–11955. [Google Scholar] [CrossRef]
  53. Wan, Y.; Chen, L.; Tang, W.C.; Li, J.L. Effect of graphene on tribological properties of Ni based composite coatings prepared by oxidation reduction method. J. Mater. Res. Technol. 2020, 9, 3796–3804. [Google Scholar] [CrossRef]
  54. Li, L.; He, M.; Feng, Y.; Wei, H.; You, X.; Yu, H.; Wang, Q.; Wang, J. Adsorption of xanthate from aqueous solution by multilayer graphene oxide: An experimental and molecular dynamics simulation study. Adv. Compos. Hybrid Mater. 2021, 4, 725–732. [Google Scholar] [CrossRef]
  55. Talyzin, A.V.; Mercier, G.; Klechikov, A.; Hedenström, M.; Johnels, D.; Wei, D.; Cotton, D.; Opitz, A.; Moons, E. Brodie vs. Hummers graphite oxides for preparation of multi-layered materials. Carbon 2017, 115, 430–440. [Google Scholar] [CrossRef] [Green Version]
  56. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved synthesis of graphene oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  57. Ullah, Z.; Riaz, S.; Li, Q.; Atiq, S.; Saleem, M.; Azhar, M.; Naseem, S.; Liu, L.W. A comparative study of graphene growth by APCVD, LPCVD and PECVD. Mater. Res. Express 2018, 5, 035606. [Google Scholar] [CrossRef]
  58. Jones, J.; Pilli, A.; Lee, V.; Beatty, J.; Beauclair, B.; Chugh, N.; Kelber, J. Atomic layer deposition of h-BN(0001) multilayers on Ni(111) and chemical vapor deposition of graphene on h-BN(0001)/Ni(111). J. Vac. Sci. Technol. 2019, 37, 060903. [Google Scholar] [CrossRef]
  59. Nam, J.; Kim, D.-C.; Yun, H.; Shin, D.H.; Nam, S.; Lee, W.K.; Hwang, J.Y.; Lee, S.W.; Weman, H.; Kim, K.S. Chemical vapor deposition of graphene on platinum: Growth and substrate interaction. Carbon 2017, 111, 733–740. [Google Scholar] [CrossRef] [Green Version]
  60. Dato, A.; Radmilovic, V.; Lee, Z.; Phillips, J.; Frenklach, M. Substrate-free gas-phase synthesis of graphene sheets. Nano Lett. 2008, 8, 2012–2016. [Google Scholar] [CrossRef]
  61. Vlassiouk, I.; Fulvio, P.; Meyer, H.; Lavrik, N.; Dai, S.; Datskos, P.; Smirnov, S. Large scale atmospheric pressure chemical vapor deposition of graphene. Carbon 2013, 54, 58–67. [Google Scholar] [CrossRef]
  62. Wu, R.; Ding, Y.; Yu, K.M.; Zhou, K.; Zhu, Z.; Ou, X.; Zhang, Q.; Zhuang, M.; Li, W.-D.; Xu, Z.; et al. Edge-epitaxial growth of graphene on cu with a hydrogen-free approach. Chem. Mater. 2019, 31, 2555–2562. [Google Scholar] [CrossRef]
  63. Imaeda, H.; Koyama, T.; Kishida, H.; Kawahara, K.; Ago, H.; Sakakibara, R.; Norimatsu, W.; Terasawa, T.; Bao, J.F.; Kusunoki, M. Acceleration of photocarrier relaxation in graphene achieved by epitaxial growth: Ultrafast photoluminescence decay of monolayer graphene on SiC. J. Phys. Chem. C 2018, 122, 19273–19279. [Google Scholar] [CrossRef]
  64. Berger, C.; Song, Z.M.; Li, X.B.; Wu, X.S.; Brown, N.; Naud, C.; Mayou, D.; Li, T.B.; Hass, J.; Marchenkov, A.N.; et al. Electronic confinement and coherence in patterned epitaxial graphene. Science 2006, 312, 1191–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Park, S.W.; Jang, B.; Kim, H.; Lee, J.; Park, J.Y.; Kang, S.O.; Choa, Y.H. Highly water-dispersible graphene nanosheets from electrochemical exfoliation of graphite. Front. Chem. 2021, 9, 699231. [Google Scholar] [CrossRef]
  66. Liu, F.; Wang, C.J.; Sui, X.; Riaz, M.A.; Xu, M.Y.; Wei, L.; Chen, Y. Synthesis of graphene materials by electrochemical exfoliation: Recent progress and future potential. Carbon Energy 2019, 1, 173–199. [Google Scholar] [CrossRef] [Green Version]
  67. Yu, P.; Lowe, S.E.; Simon, G.P.; Zhong, Y.L. Electrochemical exfoliation of graphite and production of functional graphene. Curr. Opin. Colloid Interface Sci. 2015, 20, 329–338. [Google Scholar] [CrossRef]
  68. Tang, H.; He, P.; Huang, T.; Cao, Z.; Zhang, P.; Wang, G.; Wang, X.; Ding, G.; Xie, X. Electrochemical method for large size and few-layered water-dispersible graphene. Carbon 2019, 143, 559–563. [Google Scholar] [CrossRef]
  69. Tran-Van, A.F.; Wegner, H.A. Strategies in organic synthesis for condensed arenes, coronene, and graphene. In Polyarenes I; Siegel, J.S., Wu, Y.T., Eds.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 121–157. [Google Scholar] [CrossRef]
  70. Dössel, L.; Gherghel, L.; Feng, X.; Müllen, K. Graphene nanoribbons by chemists: Nanometer-sized, soluble, and defect-free. Angew. Chem. Int. Ed. 2011, 50, 2540–2543. [Google Scholar] [CrossRef]
  71. Gao, F.; Gu, H.; Wang, H.; Wang, X.; Xiang, B.; Guo, Z. Magnetic amine-functionalized polyacrylic acid-nanomagnetite for hexavalent chromium removal from polluted water. RSC Adv. 2015, 5, 60208–60219. [Google Scholar] [CrossRef]
  72. Wu, Q.; Gao, L.; Huang, M.; Mersal, G.A.M.; Ibrahim, M.M.; El-Bahy, Z.M.; Shi, X.; Jiang, Q. Aminated lignin by ultrasonic method with enhanced arsenic (V) adsorption from polluted water. Adv. Compos. Hybrid Mater. 2022, 5, 1044–1053. [Google Scholar] [CrossRef]
  73. Bai, S.; Wang, L.; Ma, F.; Zhu, S.; Xiao, T.; Yu, T.; Wang, Y. Self-assembly biochar colloids mycelial pellet for heavy metal removal from aqueous solution. Chemosphere 2020, 242, 125182. [Google Scholar] [CrossRef] [PubMed]
  74. Fang, L.; Li, L.; Qu, Z.; Xu, H.; Xu, J.; Yan, N. A novel method for the sequential removal and separation of multiple heavy metals from wastewater. J. Hazard. Mater. 2018, 342, 617–624. [Google Scholar] [CrossRef] [PubMed]
  75. Peng, R.; Li, H.; Chen, Y.; Ren, F.; Tian, F.; Gu, Y.; Zhang, H.; Huang, X. Highly efficient and selectivity removal of heavy metal ions using single-layer NaxKyMnO2 nanosheet: A combination of experimental and theoretical study. Chemosphere 2021, 275, 130068. [Google Scholar] [CrossRef]
  76. Zaimee, M.Z.A.; Sarjadi, M.S.; Rahman, M.L. Heavy metals removal from water by efficient adsorbents. Water 2021, 13, 2659. [Google Scholar] [CrossRef]
  77. Chen, X.; Hossain, M.F.; Duan, C.; Lu, J.; Tsang, Y.F.; Islam, M.S.; Zhou, Y. Isotherm models for adsorption of heavy metals from water—A review. Chemosphere 2022, 307, 135545. [Google Scholar] [CrossRef]
  78. Rout, D.R.; Jena, H.M. Removal of phenol from aqueous solution using reduced graphene oxide as adsorbent: Isotherm, kinetic, and thermodynamic studies. Environ. Sci. Pollut. Res. 2022, 29, 32105–32119. [Google Scholar] [CrossRef]
  79. Wang, J.; Chen, B. Adsorption and coadsorption of organic pollutants and a heavy metal by graphene oxide and reduced graphene materials. Chem. Eng. J. 2015, 281, 379–388. [Google Scholar] [CrossRef]
  80. Xue, X.; Xu, J.; Baig, S.A.; Xu, X. Synthesis of graphene oxide nanosheets for the removal of Cd(II) ions from acidic aqueous solutions. J. Taiwan Inst. Chem. Eng. 2016, 59, 365–372. [Google Scholar] [CrossRef]
  81. Kuila, T.; Bose, S.; Mishra, A.K.; Khanra, P.; Kim, N.H.; Lee, J.H. Chemical functionalization of graphene and its applications. Prog. Mater. Sci. 2012, 57, 1061–1105. [Google Scholar] [CrossRef]
  82. Stankovich, S.; Piner, R.D.; Nguyen, S.T.; Ruoff, R.S. Synthesis and exfoliation of isocyanate-treated graphene oxide nanoplatelets. Carbon 2006, 44, 3342–3347. [Google Scholar] [CrossRef]
  83. Yang, X.; Zhang, X.; Liu, Z.; Ma, Y.; Huang, Y.; Chen, Y. High-efficiency loading and controlled release of doxorubicin hydrochloride on graphene oxide. J. Phys. Chem. C 2008, 112, 17554–17558. [Google Scholar] [CrossRef]
  84. Adel, M.; Ahmed, M.A.; Mohamed, A.A. Synthesis and characterization of magnetically separable and recyclable crumbled MgFe2O4/reduced graphene oxide nanoparticles for removal of methylene blue dye from aqueous solutions. J. Phys. Chem. Solids 2021, 149, 109760. [Google Scholar] [CrossRef]
  85. Gabris, M.A.; Jume, B.H.; Rezaali, M.; Shahabuddin, S.; Nodeh, H.R.; Saidur, R. Novel magnetic graphene oxide functionalized cyanopropyl nanocomposite as an adsorbent for the removal of Pb(II) ions from aqueous media: Equilibrium and kinetic studies. Environ. Sci. Pollut. Res. 2018, 25, 27122–27132. [Google Scholar] [CrossRef] [PubMed]
  86. Cai, J.; Wang, W.; Xie, W.; Wei, X.; Liu, H.; Wei, S.; Gu, H.; Guo, Z. Carbon microfibers with tailored surface functionalities supporting iron/nickel bisalloy for highly efficient hexavalent chromium recovery. Carbon 2020, 168, 640–649. [Google Scholar] [CrossRef]
  87. Fan, L.; Luo, C.; Sun, M.; Li, X.; Qiu, H. Highly selective adsorption of lead ions by water-dispersible magnetic chitosan/graphene oxide composites. Colloids Surf. B 2013, 103, 523–529. [Google Scholar] [CrossRef]
  88. Kumar, A.S.K.; Jiang, S.-J. Chitosan-functionalized graphene oxide: A novel adsorbent an efficient adsorption of arsenic from aqueous solution. J. Environ. Chem. Eng. 2016, 4, 1698–1713. [Google Scholar] [CrossRef]
  89. Abaszadeh, M.; Hosseinzadeh, R.; Tajbakhsh, M.; Ghasemi, S. The synthesis of functionalized magnetic graphene oxide with 5-amino-1,10-phenanthroline and investigation of its dual application in C-N coupling reactions and adsorption of heavy metal ions. J. Mol. Struct. 2022, 1261, 132832. [Google Scholar] [CrossRef]
  90. Chi, Z.; Zhu, Y.; Liu, W.; Huang, H.; Li, H. Selective removal of As(III) using magnetic graphene oxide ion-imprinted polymer in porous media: Potential effect of external magnetic field. J. Environ. Chem. Eng. 2021, 9, 105671. [Google Scholar] [CrossRef]
  91. Xing, C.; Xia, A.; Yu, L.; Dong, L.; Hao, Y.; Qi, X. Enhanced removal of Pb(II) from aqueous solution using edta-modified magnetic graphene oxide. Clean-Soil Air Water 2021, 49, 2000272. [Google Scholar] [CrossRef]
  92. Kostelli, G.; Kourea, K.; Ikonomidis, I. Effects of combustible tobacco smoking and novel tobacco products on oxidative stress: Different sides of the same coin? Curr. Opin Toxicol. 2020, 20–21, 41–47. [Google Scholar] [CrossRef]
  93. Nirmala, N.; Shriniti, V.; Aasresha, K.; Arun, J.; Gopinath, K.; Dawn, S.; Sheeladevi, A.; Priyadharsini, P.; Birindhadevi, K.; Chi, N.T.L.; et al. Removal of toxic metals from wastewater environment by graphene-based composites: A review on isotherm and kinetic models, recent trends, challenges and future directions. Sci. Total Environ. 2022, 840, 156564. [Google Scholar] [CrossRef] [PubMed]
  94. Moroz, I.; Scapolio, L.G.B.; Cesarino, I.; Leão, A.L.; Bonanomi, G. Toxicity of cigarette butts and possible recycling solutions-a literature review. Environ. Sci. Pollut. Res. 2021, 28, 10450–10473. [Google Scholar] [CrossRef] [PubMed]
  95. Lerner, C.A.; Sundar, I.K.; Watson, R.M.; Elder, A.; Jones, R.; Done, D.; Kurtzman, R.; Ossip, D.J.; Robinson, R.; McIntosh, S.; et al. Environmental health hazards of e-cigarettes and their components: Oxidants and copper in e-cigarette aerosols. Environ. Pollut. 2015, 198, 100–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Palisoc, S.T.; Valeza, N.C.; Natividad, M.T. Fabrication of an effective gold nanoparticle/graphene/nafion modified glassy carbon electrode for high sensitive detection of trace Cd2+, Pb2+ and Cu2+ in tobacco and tobacco products. Int. J. Electrochem. Sci. 2017, 12, 3859–3872. [Google Scholar] [CrossRef]
  97. Guo, C.; Wang, C.; Sun, H.; Dai, D.; Gao, H.T. A simple electrochemical sensor based on rGO/MoS2/CS modified GCE for highly sensitive detection of Pb(II) in tobacco leaves. RSC Adv. 2021, 11, 29590–29597. [Google Scholar] [CrossRef] [PubMed]
  98. Yu, Y.L.; Zhuang, Y.T.; Song, X.Y.; Wang, J.-H. Lyophilized carbon nanotubes/graphene oxide modified cigarette filter for the effective removal of cadmium and chromium from mainstream smoke. Chem. Eng. J. 2015, 280, 58–65. [Google Scholar] [CrossRef]
Figure 2. Chemical structure of GO. Reprinted with permission from [40]. Copyright@ Elsevier.
Figure 2. Chemical structure of GO. Reprinted with permission from [40]. Copyright@ Elsevier.
Separations 09 00401 g002
Figure 3. Characterizations of graphene (G) and GO: (A) FTIR spectra; (B) Raman spectra; (C) XRD patterns (the black line is GO, the red line is graphene); high–resolution C 1s XPS spectra of (D) graphene and (E) GO; (F) nitrogen adsorption–desorption curve of GO obtained from BET test; (G) and (H) SEM images of GO; and (I) TEM image of GO. Reprinted with permission from [41]. Copyright @Springer–Nature.
Figure 3. Characterizations of graphene (G) and GO: (A) FTIR spectra; (B) Raman spectra; (C) XRD patterns (the black line is GO, the red line is graphene); high–resolution C 1s XPS spectra of (D) graphene and (E) GO; (F) nitrogen adsorption–desorption curve of GO obtained from BET test; (G) and (H) SEM images of GO; and (I) TEM image of GO. Reprinted with permission from [41]. Copyright @Springer–Nature.
Separations 09 00401 g003
Figure 4. Schematic illustration for graphene exfoliation by ball milling. (A) Shear force–induced exfoliation; (B) Compression force–induced fracture. Reprinted with permission from [50]. Copyright@ John & Wiley Inc.
Figure 4. Schematic illustration for graphene exfoliation by ball milling. (A) Shear force–induced exfoliation; (B) Compression force–induced fracture. Reprinted with permission from [50]. Copyright@ John & Wiley Inc.
Separations 09 00401 g004
Figure 5. Preparation procedure of improved Hummer’s method of graphene. Reprinted with permission from [56]. Copyright@ ACS publications.
Figure 5. Preparation procedure of improved Hummer’s method of graphene. Reprinted with permission from [56]. Copyright@ ACS publications.
Separations 09 00401 g005
Figure 6. (A) Illustration for electrochemical exfoliation of graphite to produce various graphene materials; (B) schematic of cathodic and anodic exfoliation mechanisms. Reprinted with permission from [66,67]. Copyright@ Elsevier and John & Wiley Inc., respectively.
Figure 6. (A) Illustration for electrochemical exfoliation of graphite to produce various graphene materials; (B) schematic of cathodic and anodic exfoliation mechanisms. Reprinted with permission from [66,67]. Copyright@ Elsevier and John & Wiley Inc., respectively.
Separations 09 00401 g006
Figure 7. Covalently bonded Fe3O4 nanoparticles onto graphene. Reprinted with permission from [41]. Copyright@ Springer–Nature.
Figure 7. Covalently bonded Fe3O4 nanoparticles onto graphene. Reprinted with permission from [41]. Copyright@ Springer–Nature.
Separations 09 00401 g007
Figure 8. (A) Schematic for the synthesis of MGO–As(III)–IIP; (B) effect of regeneration cycles on the adsorption capacity. Reprinted with permission from [90]. Copyright@ Elsevier.
Figure 8. (A) Schematic for the synthesis of MGO–As(III)–IIP; (B) effect of regeneration cycles on the adsorption capacity. Reprinted with permission from [90]. Copyright@ Elsevier.
Separations 09 00401 g008
Figure 9. Illustration for the preparation of MGO–EDTA. Reprinted with permission from [91]. Copyright@ John Wiley & Sons.
Figure 9. Illustration for the preparation of MGO–EDTA. Reprinted with permission from [91]. Copyright@ John Wiley & Sons.
Separations 09 00401 g009
Figure 10. The effect of the length of o–MWNTs/GO modified cigarette filter on Cd(II) and Cr(VI) removal from mainstream smoke (with a dosage of 4 mg). Reprinted with permission from [98]. Copyright@ Elsevier.
Figure 10. The effect of the length of o–MWNTs/GO modified cigarette filter on Cd(II) and Cr(VI) removal from mainstream smoke (with a dosage of 4 mg). Reprinted with permission from [98]. Copyright@ Elsevier.
Separations 09 00401 g010
Table 1. Heavy metal adsorption capacity/removal efficiency, as reported in the literature.
Table 1. Heavy metal adsorption capacity/removal efficiency, as reported in the literature.
MaterialsSynthesis MethodHeavy MetalTreatment ConditionsMaximum Adsorption Capacity (mg g−1)Ref.
GOModified Hummer’s methodCd(II) Room temperature, pH = 5, 10 mg L−1 Cd(II), 24 h equilibration time35.7[79]
GOHummer’s methodCd(II)T = 303.15 K, pH = 4.00, 0.50 g L−1 adsorbents, 0.02~5 g L−1 Cd(II), 12 h equilibration time44.64[80]
Magnetic chitosan/GOModified Hummer’s methodPb(II)T = 30 ± 0.2 °C, pH = 5, 0.8 g L−1 adsorbents, 0.02~14 mg L−1 Cd(II), 12 h equilibration time76.94[87]
Chitosan–functionalized GOImproved Hummer’s methodAs(V)/As(III)T = 30 °C, pH = 5.5, 8 g L−1 adsorbents, 30~500 ppm As(V)/As(III)71.9/64.2, respectively[88]
MGO@APhen Pb(II) Room temperature, pH = 6.5, 0.5 g L−1 adsorbents, 10 ppm Pb(II)The removal efficiency of 97.2% [89]
MGO ion–imprinted polymerModified Hummer’s methodAs(III)T = 25 °C, pH = 5, 5 g L−1 adsorbents, 0.5~20 mg L−1 As(III), 24 h equilibration time49.42[90]
MGO–EDTAHummer’s and Offeman’s methodsPb(II)T = 20 °C, pH = 1~10, 0.5 g L−1 adsorbents, 10~150 mg L−1 As(III), 3 h equilibration time211.3[91]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, X.; Zeng, J.; Wu, Y.; Wang, Q.; Wu, S.; Gu, H. Preparation and Application of Graphene–Based Materials for Heavy Metal Removal in Tobacco Industry: A Review. Separations 2022, 9, 401. https://doi.org/10.3390/separations9120401

AMA Style

Xu X, Zeng J, Wu Y, Wang Q, Wu S, Gu H. Preparation and Application of Graphene–Based Materials for Heavy Metal Removal in Tobacco Industry: A Review. Separations. 2022; 9(12):401. https://doi.org/10.3390/separations9120401

Chicago/Turabian Style

Xu, Xiaojiang, Junling Zeng, Yue Wu, Qiaoying Wang, Shengchao Wu, and Hongbo Gu. 2022. "Preparation and Application of Graphene–Based Materials for Heavy Metal Removal in Tobacco Industry: A Review" Separations 9, no. 12: 401. https://doi.org/10.3390/separations9120401

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop