Next Article in Journal
Evaluation of Host Cell Impurity Effects on the Performance of Sterile Filtration Processes for Therapeutic Viruses
Previous Article in Journal
Wearable Crop Sensor Based on Nano-Graphene Oxide for Noninvasive Real-Time Monitoring of Plant Water
Previous Article in Special Issue
Electrodialysis Can Lower the Environmental Impact of Hemodialysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Literature Review of Modelling and Experimental Studies of Water Treatment by Adsorption Processes on Nanomaterials

Institute of Technology Sligo, Ash Lane, F91 YW50 Sligo, Ireland
*
Author to whom correspondence should be addressed.
Membranes 2022, 12(4), 360; https://doi.org/10.3390/membranes12040360
Submission received: 2 February 2022 / Revised: 11 March 2022 / Accepted: 18 March 2022 / Published: 24 March 2022
(This article belongs to the Special Issue Membrane Filtration for Water Reclamation)

Abstract

:
A significant growth in the future demand for water resources is expected. Hence researchers have focused on finding new technologies to develop water filtration systems by using experimental and simulation methods. These developments were mainly on membrane-based separation technology, and photocatalytic degradation of organic pollutants which play an important role in wastewater treatment by means of adsorption technology. In this work, we provide valuable critical review of the latest experimental and simulation methods on wastewater treatment by adsorption on nanomaterials for the removal of pollutants. First, we review the wastewater treatment processes that were carried out using membranes and nanoparticles. These processes are highlighted and discussed in detail according to the rate of pollutant expulsion, the adsorption capacity, and the effect of adsorption on nanoscale surfaces. Then we review the role of the adsorption process in the photocatalytic degradation of pollutants in wastewater. We summarise the comparison based on decomposition ratios and degradation efficiency of pollutants. Therefore, the present article gives an evidence-based review of the rapid development of experimental and theoretical studies on wastewater treatment by adsorption processes. Lastly, the future direction of adsorption methods on water filtration processes is indicated.

Graphical Abstract

1. Introduction

Water is arguably the main life source on planet earth and is vital for preservation of the modern world. In many parts of the world, water exploitation in agriculture and land development has led to significant economic progress [1]. On the other hand, water pollution is the biggest challenge facing the world and causing the destruction of water resources around the world [2]. This is because of its pollution with oil, chemicals, industrialization, human consumption, leakage of sewage lines, or the absence of these lines [3]. In addition, according to the United Nations (UN) report, some countries in the Middle East region suffer from water scarcity [1]. Moreover, human activities such as overgrazing, cultivation in wet areas, and discharging waste into rivers has changed the water flow patterns in rivers in terms of quantity, quality and flow times [4]. Changing flow quantities can affect water quality significantly in ground and surface water such as rivers [5]. An example of this is in Rajasthan, India due to a very large increase in groundwater pumping [6]. This led to a decrease in water quality due to an increase in the concentration of fluoride above normal levels causing disease in bones and teeth [6]. In other cases, when the consumption of surface water sources is very high, such as rivers and lakes, the decrease in water levels leads to an increase in the concentration of minerals [7]. These problems have led to a lack of safe drinking water for more than a third of the world population [8]. In addition, it creates the need to look for high quality water by converting low quality water to pure water. Some of the latest technologies that revolutionised the field of water filtration in recent years are membrane-based methods and photocatalysis degradation of organic pollutants. In this review, we mainly focused on the role of the adsorption process in both membrane technology and photocatalytic degradation of organic pollutants by experimental and simulation methods.

1.1. Water Filtration

Filtration is the process of removing solid particles or gases from liquids through filter media which allow the liquid to pass through and blocking solid and gases particles [9]. In filtration processes, filtered water goes one way and other particles collected through the filter area go in a different route [10]. Filtration processes have received wide attention since the 18th century to purify polluted water to obtain freshwater sources for drinking and commercial uses [11]. Unfortunately, many systems which have been developed to meet the needs of water purification were unsuccessful due to drawbacks in each method. In the 18th century, filtration was adopted to remove undesirable particles from water such as copper and lead [12]. Sponge filter was the first water filtration device discovered around the mid-1700s, and Joseph Amy obtained a patent for that device [13]. Furthermore, water purification devices were provided for the first time for domestic use in 1750 [14]. In addition, the first known slow sand filtration system was made in 1804 by John Gibb in Scotland and was very effective in the removal of bacteria, viruses, and heavy metals to produce drinking water [15]. There were great developments in water filtration systems in the 19th and 20th centuries such as the development of the reverse osmosis (RO) method [16], membrane technology [17], and photocatalytic degradation of organic pollutants [18]. The RO method forces water to move through a semi-permeable membrane by an external pressure [19,20]. A larger volume of water subsequently passes through the membrane compared to the volume of dissolved salts or organic molecules [19]. It is capable of removing 99% of ions, particles, collides, and bacteria from the feed water [19].
Energy consumption is a key factor which affects the freshwater production cost in the RO process [21]. One of the main factors for the high energy consumption in the RO method is the high-pressure pump which consumes almost 74% of total energy consumption in RO systems [22]. It has become more efficient for water desalination with the advance of nanotechnology [23]. The level of RO energy consumption has declined in the past 40 years due to membrane technological improvements offering higher permeability [24]. Membrane technology is a general term for several different separation processes that depend on a semi-permeable membrane for gas separation and the removal of undesirable ions or molecules from liquids [25]. The semi-permeable membranes with specific pore size are fixed between two media to block particles at the molecular or ionic level by a driving force based on the pressure difference between the two sides [26,27].
Nanocomposite membranes have attracted researchers to develop better materials to achieve remarkable properties such as selectivity, mechanical properties, and enhancement of membrane performance in water desalination [28]. For instance, mechanical properties (including Young’s modulus, bulk modulus, and shear modulus) and selectivity of graphene membrane have improved after the combination with titanium dioxide (TiO2) as indicated in a previous study [29,30,31]. Nanofiltration is a process depending on an external pressure force in which molecules and particles in a range between 1–2 nm are rejected by the membrane which make it one of the most widely used membrane processes in desalination and wastewater treatment [32]. However, researchers have focused on developing a variety of nanoporous membranes which have open pores diameter ranging between 1 to 100 nm for water desalination such as graphene and molybdenum disulphide (MoS2) [33]. These two nanoporous membranes achieved a high water permeability and salt rejection performance [34,35]. Graphene membrane has contributed to other applications in the water filtration field. In 2016, Cohen-Tanugi and colleagues synthesized a multilayer graphene membrane and it proved to be more economical than a single layer graphene membrane [36]. In addition, the multilayer graphene membrane can offer higher salt rejection than a single membrane with high permeability [36]. However, these promising results attracted researchers to synthesize other nanocomposite materials such as multilayer MoS2 membrane, and multilayer graphene oxide (GO) membrane [37]. The multilayer membranes are expected to offer more flexibility in terms of membrane productivity, membrane lifetime, and their performance in filtration processes [38]. In addition, a lot of studies mentioned the contribution of spinel ferrites (MFe2O4) with metal oxides, MFe2O4/carbon-based materials, MFe2O4/polymers, and MFe2O4/metal nanoparticles for the photocatalytic degradation of dyes and other inorganic pollutants as mentioned in previous studies [39,40,41]. In addition, Zeolite nanostructured membranes have been widely used for the removal of hazardous chemicals from a contaminated water solution as shown in previous studies [42,43,44].
On the other hand, nanocomposite materials such as MoS2 showed distinctive properties in the photocatalytic degradation of organic pollutants [45]. This is because of its unique electronic [46] and optical properties [47], and its size-dependent bandgap [48]. In 2020, Yan et al. synthesized MoS2/TiO2 nanotube composite for efficient water disinfection using anodic oxidation method and hydrothermal method [49]. The results showed an excellent photocatalytic disinfection under visible light irradiation for the removal of bacteria up to 98.5% with the possibility of reusing the nanocomposite and recycling it efficiently [49]. Photocatalysis is an eco-friendly technique for the removal of many different pollutants such as nitrogen oxides (NOx), pesticides and other organic pollutants using ultraviolet (UV) light or sunlight [50,51]. TiO2 is considered to be the most promising material for the photocatalytic removal of organic pollutants [52]. This is because of the high photocatalytic activity of TiO2 which make it environmentally friendly [53]. Furthermore, the number of pores located in the TiO2 surface that are generated by photons make it easy to be harvested by free electrons from outside the surface of TiO2 [53]. In addition to wastewater treatment, photocatalysis has many other applications such as storing energy [54], air purification [55], and antifouling by preventing the nonspecific interaction between membrane surface and foulants [56]. However, photocatalytic reactions proved their effectiveness in degradation of organic pollutants in non-toxic filtration medias, without using chemicals [57]. One of the important steps in the photocatalytic process is the adsorption between the reactant substances with the catalyst surface [58]. The adsorption process affects the efficiency of the photocatalysis process because it mainly depends on the catalyst absorption of light source [59,60]. Therefore, it is one of the most important water filtration processes under development that controls filtration [61]. Moreover, there are many advantages of adsorption process such as being inexpensive [62], fast [63], and simple in operation and implementation [64]. It has many applications in wastewater treatment to remove pollutants such as bacteria and heavy metals and has received considerable attention from researchers [65].

1.2. Adsorption Techniques

In water filtration, adsorption is the process of removing organic pollutants from wastewater in which a binding energy is present between the molecules of the substance (ions or atoms) with other surfaces by chemical or physical attraction [66]. It is an important process in catalysis [67], chemical engineering [68], and material science [69]. There are two main types of adsorption process: physical adsorption (physisorption) and chemical adsorption (chemisorption) [70]. Physical adsorption occurs by adsorbing gas molecules onto a solid surface using low intensity forces called van der Waals force [71,72]. London-van der Waals is a dispersion force acting between microscopic non-deformable bodies such as atoms and molecules [73]. There are many applications of physisorption process such as: hydrogen storage [74], acoustic wave sensors [74], gas sorption [75] and water filtration [76]. Furthermore, physical adsorption improves membrane efficiency in terms of water flux, hydrophilicity, and antifouling [77]. For instance, Peng and colleagues modified the performance of PVDF (polyvinylidene fluoride) microfiltration membrane for water filtration by a strong physisorption of amphiphilic copolymers experimentally [78]. The results showed a high improvement in the permeability of PVDF membrane with better antifouling properties without any effect on the membrane structure [78].
On the other hand, in chemical adsorption the bonding occurs between the surface molecules of a metal with high energy and another substance (adsorbate) in contact with it, which may be a liquid or a gas [79]. The bonds formed are comparable in strength to ordinary chemical bonds and are much stronger than the van der Waals forces characteristic of physical adsorption [79]. Chemisorption has been widely used in industrial wastewater treatment for the removal of heavy metals [80,81,82]. For instance, Liu et al. investigated the chemical adsorption behaviour in the removal of arsenic by experimental and simulation techniques using microporous metal-organic framework (MIL-125(Ti)). The results showed a fast and efficient removal of arsenic with low concentrations by chemical adsorption [83]. Most studies have not focused on adsorption kinetics by either ignore it, excluded it, or by assuming the adsorption interaction approaches the equilibrium [84,85,86]. However, Luo et al. have studied the photocatalytic degradation kinetics of graphitic carbon nitrate (g-C3N4) for contaminant removal by using a 36 W LED light (λ = 400 nm) in a dark chamber [87].
There are many factors that affect the performance of chemical and physical adsorption processes in water filtration such as: atmospheric and experimental conditions [88], contact time between the adsorbate and adsorbent [89], and particle size [90]. It was noted that atmospheric conditions significantly affect the effectiveness of physical adsorption process in terms of absorption capacity [88]. The adsorption process generally is an exothermic process where the energy is expelled in the form of heat or light [91]. Therefore, physical adsorption is more efficient at lower temperatures because when the temperature rises, the ability of the material to absorb reduces as indicated in previous studies [92,93,94]. On the other hand, chemical adsorption range increases with increasing temperature to a certain extent and then begins to decrease [95]. In addition, chemical and physical adsorption increases with increasing gas pressure to a certain extent until saturation is reached [96]. Contact time and particle size have a major effect on the efficiency of both types of adsorption process [89]. Zhang et al. noted that the adsorption of methylene blue (MB) from aqueous solution increased with contact time with the adsorbent [89]. Along the same lines, Laabd et al. studied the adsorption capacity of polyaniline (PANi) film for the removal of trimellitic and pyromellitic acids by experimental and density functional theory (DFT) methods [97]. They estimated the optimum experimental conditions for the adsorption of the acids in terms of contact time, pH, initial concentration, and temperature. Meanwhile, they investigated the results by using first principle DFT calculations and studied the physical interactions between the adsorbate and adsorbent surface molecules [97]. In addition, Bergaoui et al. studied the adsorption mechanism of methylene blue (MB) onto organo-bentonite [98]. The results showed a high MB removal with a maximum adsorption capacity up to equal to 321 mg/g. However, adsorption technique enhanced the performance of photocatalysts based membranes in wastewater treatment as reported in previous studies [99,100,101]. For instance, Zhang et al. modified Ag@BiOBr/AC/GO membrane system for efficient removal of rhodamine B (RhB) by membrane separation and high adsorption capacity [99].
So far, molecular dynamic (MD) and DFT calculations have been widely used by researchers to investigate and predict experimental results and save money, as well as acquiring faster results [102]. Adsorption process and its contribution in water and wastewater treatment has been studied experimentally and investigated by simulation in 70 publications which will be reviewed in this paper. Figure 1 shows the rapid increase in recent years in the number of publications related to water treatment by adsorption process using experimental and simulation methods.

1.3. Computational Methods

Computer simulation methods are compatible with experimental work in a laboratory to serve as a bridge between laboratory experiments and theoretical calculations [103]. Furthermore, simulation is an approximate emulation for an integrated complex system to analyse the behaviour and performance of the system over time [104]. There are many applications for simulation such as: manufacturing [105], economics [106], safety engineering [106], and simulation of technology for performance or optimization [107]. Simulation has many advantages such as validating the results obtained by other analytical methods, and sometimes finding unexpected phenomenon while studying the behaviour of the system [108]. MD and DFT simulation methods have been used in many studies in fabrications and modifications of membrane structure for water desalination, gas separation, electrolysis, and many other applications [35,109,110]. In addition, they have been used to investigate the experimental results of the adsorption process as indicated in previous studies [111,112,113].

Calibration and Validation

Due to the significant increase in the use of computer simulation methods in the 21st century, it was necessary to find ways to verify the data issued by simulation programs. Therefore, many models have been modified to measure the error ratio between the data obtained experimentally and data obtained by simulation such as SWAT (Soil and Water Assessment Tool) [114]. SWAT is used to calibrate and validate the data by a procedure using the shuffled complex evolution method [114]. Calibration “is a major element to this evaluation and refers to the estimation and adjustment of model parameters to improve the agreement between model output and a data set” [115]. However, validation “is a model using parameters that were determined during the calibration process and comparing the predictions to observed data not used in the calibration” [115]. Recently, researchers’ dependence on calibration and validation of their results have been gradually increased. Therefore, the number of publications that studied adsorption process in wastewater treatment using combined experimental and simulation methods increased at the beginning of 2017. Figure 2 shows the number of publications in water treatment by adsorption using experimental and simulation methods in the last 10 years.
For instance, de Oliveira and colleagues studied the adsorption process of 17β-estradiol in graphene oxide through methanol co-solvent experimentally and through simulation using SIESTA code (Spanish Initiative for Electronic Simulations with Thousands of Atoms) [113]. The DFT model was calibrated using the experimental findings and was in good agreement after studying the electronic density of state (DOS) and the interactions between graphene oxide with methanol molecules [113]. Along the same lines, Zhu et al. studied the degradation of bisphenol using highly efficient heterogeneous Fenton catalysts (CNTs/Fh) by experiment and simulation [116]. The main aim of this study was to accelerate Fe(III)/Fe(II) cycling by carbon nanotubes (CNTs) and they had an excellent agreement in the experimental and DFT results [116]. More detail about the fabrication and characterization of the nanocomposite materials experimentally and by simulation are given in Section 2.

2. Synthesis and Simulation of Nanomaterials

In the last decade, synthesis and simulation of nanomembranes have received vast attention and have been widely studied by researchers [117]. Figure 3 shows the most common methods used in the synthesis of nanomaterials.

2.1. Nanomembranes

Synthesis and characterization of nanomembranes have received wide attention since the 18th century [118]. During the 18th century, membranes were under fabrication, functionalization, and modification at the laboratories without any commercial use [119]. Since 2004, membrane experimental designs have increased and the number of materials available for these experiments has increased [118]. One of the famous designs was by Jani and colleagues which designed nanoporous anodic aluminium oxide membranes with desired functions, parameters and properties [120]. Similarly, Mei and colleagues fabricated ultrathin AlN/GaN porous crystalline nanomembranes with different layouts including tubes, spirals, and curved sheets [121]. The structural, morphological and chemical properties of nanomembranes will be characterizing using analyses such as X-ray diffraction (XRD), scanning electron microscope (SEM), thermal electron microscope (TEM) and X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, etc. [122]. The optical and electrical properties of the nanomembranes also will be analysed using UV-Vis diffuse reflectance spectroscopy (UV-Vis DRS) [123]. “Nanomembranes are synthetic structures with a thickness less than 100 nm and the aspect of surface-area-volume ratio increases to at least a few orders of magnitude” [124]. Nanomembranes can be classified based on surface chemistry, bulk structure, morphology, and production method [119]. Nanomembranes have been widely used in many applications such as water and wastewater treatment [125], biomedical applications [126], and smart energy storage devices [127]. In this section, we demonstrate the most common methods used in the synthesis and simulation of nanomembranes.

2.1.1. Synthesis of Nanomembranes

Synthesis is a term for producing nanostructured materials including organic, inorganic, and hybrid nanomembranes [128]. It exploits the special physicochemical properties of ionic fluids to control transit and growth [129]. Many methods have been used for the synthesis of nanomembranes such as modified Hummers’ method [130], solvothermal method [131], and solvothermal chemical deposition [132]. Modified Hummers’ method is one of the most common methods used for the synthesis of nanomembranes such as graphene oxide (GO) [133]. It was developed in 1958 with many advantages such as being safer, faster, and a more efficient method for producing graphite oxide [134]. The chemical method can generate graphite oxide through the addition of potassium permanganate to a solution of graphite, sodium nitrate, and sulfuric acid [129]. However, for the synthesis of other nanomembranes such as molybdenum disulphide, the microwave-assisted route has been used [135]. The microwave-assisted route is “a unique and simple technique for fast and efficient processing of materials with higher reproducibility” [136]. It has drawn attention due to its homogeneous heating, fast kinetics, high phase purity, and high yield rate of products in relatively short time [136]. Table 1 shows nanomembranes synthesized by different synthesis methods.
As shown in Table 1, 11 types of nanomembranes have been synthesized by using different materials. Graphene and graphene oxide (GO) were the most synthesized nanomembranes by using Hummers’ method because of their widespread use in water and wastewater treatment. However, other nanocomposites such as nitrogen doped carbon (CNs), are synthesized by using chlorination of Ti(C0.7N0.3) at various temperatures resulting in well-developed micro-pores and small meso-pores with uniform pore structures.

2.1.2. Simulation of Nanomembranes

Density functional theory (DFT) is a computational simulation method used in chemistry, physics, and materials science for the calculation of the mechanical and electronic properties of atoms and molecules [155]. There are many simulation software used for DFT calculations such as Material Studio, Vienna Ab initio Simulation Package (VASP), and GAMESS, etc. The simulation software have been used by researchers and engineers to improve the performance of materials in many applications including pharmaceuticals, catalysts, polymers and composites, metals and alloys, batteries and fuel cells [156]. They have many advantages such as developing new cost-effective materials with better performance and more efficiently than with test and experimentation alone [157]. Material studio is a three-dimensional (3D) modelling and simulation software developed and distributed by BIOVIA to allow researchers in material science and chemistry to understand the behaviour and relationships of a material’s atomic and molecular structure [156]. Similarly, VASP, Gaussian 09, and GAMESS have been used for atomic scale materials modelling using DFT with different functional groups including (B3LYP) and different methods such as the projector augmented wave method (PAW), and Perdew-Burke-Ernzerhof (PBE) method [158]. PAW and PBE methods are both efficient for the electronic structure calculations of large systems [159]. Furthermore, they are used to improve the accuracy of the electrical and electronic calculations for magnetic materials, alkali and alkali earth elements [160].
Figure 4 shows simulation software used to produce nanomembranes with the number of publications using each software.
As shown in Figure 4, Gaussian 09 and VASP contributed to the simulation of 15 nanomembranes out of 22 nanomembranes in this review paper. 8 of the 15 nanomembranes were graphene or graphene oxide (GO). This is due to the high accuracy in the simulation of graphene and GO nanomembranes by these simulation software as indicated in previous studies [161,162]. In addition, molybdenum disulphide (MoS2) nanosheet has been simulated by using VASP with PAW simulation method, while other nanocomposites such as Zn–Fe LDH, and (CF/BiOBr/Ag3PO4) cloth, have been simulated by using Material Studio with DMol3 and GGA-PBE codes, respectively. Table 2 shows the simulation software and methods used for simulation of nanomembranes.
As shown in Table 2, PAW, PBE, and B3LYP are the most common methods used for the DFT calculations of nanomembranes. These calculations are performed based on the solution of Kohn-Sham equations by PAW method. On the other hand, the exchange-correlation functional model, and the Thomas, Yoon–Nelson, and Adams–Bohart model have been solved by B3LYP and PBE method, respectively. Along the same lines, these methods (PAW, PBE, and B3LYP) have been used for the simulation of nanocomposite materials as explained later in Section 2.2.2.

2.2. Nanocomposites

Nanocomposite materials “are composed of several multiple nanomaterials entrapped within a bulk material, which may comprise a combination of a soft and a hard nanomaterial, two soft nanomaterials, or two hard nanomaterials” [169]. They are characterized by their very small size, measured in nanometres [170]. Nanocomposite materials have attractive properties resulting from the combination of inorganic or organic components at the molecular level [171,172]. There are many applications of nanocomposite materials in wastewater treatment [173], energy storage [174], drug delivery [175], and for biomedical purposes [176]. In wastewater treatment, nanocomposite materials have been widely used to treat surface water, sewage, and ground water [177]. By 2009, nano-processing technologies were documented at 44 cleaning sites around the world, most of them in the United States [178]. The synthesis of these nanocomposites received wide attention by the researchers in the last decade [179]. In this section, we demonstrate the most common methods used in the synthesis of nanocomposite materials.

2.2.1. Synthesis of Nanocomposites

For synthesis and characterization of these nanomaterials, many methods have been used including the hydrothermal method [180], chemical vapor deposition (CVD) [181], and one-pot synthesis [182]. The hydrothermal method is one of the most common methods used in the synthesis of nanocomposites [180]. Figure 5 shows the percentages of the number of publications reviewed in this review paper by each experimental method. As shows in Figure 5, the hydrothermal method has been used in more than 56% of the publications reviewed.

The Hydrothermal Method

Hydrothermal synthesis is a method that uses very high temperatures ranging from room temperature to much higher temperatures to synthesize nanomaterials [183]. It was given the name “hydrothermal” because water is used as the solvent [184]. The hydrothermal method was first discovered in the 19th century [185]. It has been widely used by researchers and the first publication on this method appeared in 1813 [185]. The publication was about “Synthesis and Characterization of Zinc Tin Sulphide (ZTS) Thin Films via Chemical Bath Deposition Route” [185]. Hydrothermal synthesis has many advantages over other synthesis methods including “top down” method, “bottom up” method, and sol-gel method such as being an environmentally friendly, low-cost synthesis method, its simplicity, and the production of high-quality one-dimensional (1D) nanostructures [186,187,188,189]. However, there are some disadvantages for this method: taking a long time in the production process, corrosion, and difficulty in recycling and regenerating the catalysts [184,190]. Recently, hydrothermal synthesis has been used in several applications in science such as food and nutrition, organic chemistry, environmental safety, and energy applications [191,192]. For instance, Zhu et al. synthesized a highly efficient heterogeneous Fenton catalyst (CNTs/Fh) for the degradation of (bisphenol A) by using a hydrothermal method [116]. Similarly, Wang et al. synthesized a pyridinic-N doped graphene/BiVO4 nanocomposite (N-rGO/BiVO4) by hydrothermal method with a great potential for the removal of pollutants from wastewater [193]. Table 3 shows the nanocomposite materials synthesized by the hydrothermal method in the last decade.
As shown in Table 3, the hydrothermal method has been used in the synthesis of different nanocomposite materials including titanium dioxide (TiO2) nanoflowers, nanomaterials with carbon nanotubes (CNTs), and metal oxides with carbon. The reason for the wide use of the hydrothermal method is its advantages over others in the ability to create crystalline phases, even those which are not stable at the melting point [218]. For instance, Zhao and colleagues synthesized TiO2 nanoflowers (TNFs) using hydrothermal and calcination treatments [199]. The results showed a strong photocatalytic capability, and satisfactory recycled stability of the TNFs, which enhances their value for practical applications in water purification [199]. Along the same lines, Cheng et al. synthesized a titanate nanotube supported TiO2 (TiO2/TiNTs) using the hydrothermal method [200]. The results showed that TiO2/TiNTs significantly eliminated the toxicity of phenanthrene and can greatly decrease the potential risks of phenanthrene to aquatic organisms [200].

Chemical Vapor Deposition

Chemical vapor deposition (CVD) is a coating process that is defined as a method to produce solids with high purity by using thermally induced chemical reactions at the surface of a heated substrate [219]. CVD has many applications in medicine [220], electronic applications [221], and chemical industries [222]. It has many advantages over other synthesis methods such as the ability to deposit a wide variety of materials with very high purity [223]. The CVD method started in the 19th century with the production of lamp filaments. Then, Van Arkel in the 20th century deposited metals from the gas phase for application in the lamp industry [224,225]. The CVD method has three different types based on the conditions of the process classified by applied pressure [226], physical properties of the vapor [227], and plasma methods [228]. It has been used in the production of several materials including monocrystalline, polycrystalline, amorphous, preparation of carbon nanotubes (CNTs) and carbon nanofibers [221,229]. In addition, CVD is famous for producing semiconductors such as the synthesis of 2D Tungsten disulphide (WS2) monolayer [230]. Table 4 shows nanocomposite materials synthesized by the chemical vapor deposition (CVD) method.
In this review paper, the CVD method contributed in 7% of the total number of publications as shown in Figure 5. Mainly, it has been used for the synthesis of carbon nanotubes (CNTs). As shown in Table 4, the CVD method has been used for the combination of Co3O4, and COOH with CNTs. At a sufficiently high temperature, carbon source (hydrocarbon gas) decomposes with the catalyst in a tubular reactor [233]. By using CVD, Yang et al. synthesized a vertically aligned carbon nanotube hybrid membrane for gas separation [164]. The results showed an excellent separation membrane with high conductivity and resistance stability after 50 cycles of tensile deformation [164]. In addition, Zhang et al. studied the adsorption of lead (Pb2+) on oxidized (O-CNTs) and graphitized multi-walled carbon nanotubes (G-CNTs) synthesized by the CVD method [163]. The results showed a high stability in the adsorption mechanism of Pb2+ [163].

One-Pot Synthesis

One-pot synthesis is a process that is used to improve the efficiency of chemical reactions and focuses on the reduction of number of steps of chemical reactions in one single reaction flask [234]. It is a hydrothermal approach based on a general phase transfer and separation mechanism which occurs at interface of water, solution, and solid phases [190]. Das et al. reported in their book that the one-pot synthesis method has many advantages such as “saving time and resources, improves the efficiency of a chemical reaction, and offers better chemical yield” [190]. An example of one-pot synthesis is the synthesis of highly stable CsPbBr3@SiO2 Core–Shell Nanoparticles [182]. The reported method showed that the formation rates, determined by reaction temperature, precursor species, pH value, etc., of both CsPbBr3 and SiO2 are critical for the successful preparation of core–shell NPs [182]. Table 5 shows nanocomposite materials synthesized by the one-pot synthesis method.
As shown in Table 5, Ren and his colleagues synthesized a 3D porous sulphur and nitrogen co-doped graphene aerogel (SN-rGO-A) for the degradation of Rhodamine B (RhB) [165]. The results showed that sulphur and nitrogen co-doping could synergistically enhance the catalytic performance for activating peroxydisulfate (PDS) compared to the original and N doped graphene aerogels [165]. Along the same lines, Li and colleagues synthesized a series of catalysts by using ZIF-67 for the removal of bisphenol A (BPA) and total organic carbon (TOC) [235]. The results showed an excellent degradation efficiency for TOC and BPA.

Other Synthesis Methods

In addition to the methods presented in this section, there are other known methods for the synthesis of nanocomposite materials such as solvothermal methods. Solvothermal methods “offer a simple, direct, and low-temperature route to obtain nanometric particles with narrow size dispersions, and represent an alternative to calcinations for promoting crystallization under milder temperatures” [237]. It is mainly used for the synthesis of highly crystallized lanthanide (UCNPs) at relatively low temperature [238]. It has many advantages such as being a simple, economical and efficient method [239]. On the other hand, there are some disadvantages of this method such as the long processing time and the contraction that occurs during processing [240]. However, solvothermal method is very similar to the hydrothermal route, the only difference being that the precursor solution is usually non-aqueous [238]. Table 6 shows nanocomposite materials synthesized by other synthesis methods.

2.2.2. Simulation of Nanocomposites

For the simulation of nanocomposites, VASP was the most used simulation software with 27 publications as shown in Figure 6. 10 of these publications were for the simulation of carbon-based material including CNTs and g-C3N4. However, 14 publications were for the simulation of other nanocomposites such as MnFe2O4 nanocubes using the PAW method, Fe3O4-HBPA-ASA using B3LYP functional group, and granular TiO2-La using the PBE method as shown in Table 7.
Figure 6 shows simulation software used to produce nanoparticles with the number of publications by each software.
As shown in Table 7, DFT (PBE), and DFT (PAW) have been mainly used for the simulation of the nanocomposite materials. This is due to the high accuracy of these two simulation methods. The augmented plane wave (PAW), and PBE was used to describe the electron–ion interactions [256]. For instance, Maji and colleagues used DFT (PBE) simulation method through VASP software for the simulation of Fe2O3-PC nanohybrids [203]. The results showed an agreement between the experimental and simulation results [203]. Similarly, Regmi and colleagues simulated an N-doped BiVO4 model using the DFT (PBE) method through VASP simulation software [208]. The results showed good agreement with experimental results in terms of the electronic property calculations such as the band structure and density of state (DOS) [208].

3. Water Filtration by Membrane Technology

Due to its important role in water purification, membrane technology is rapidly developing. The ability of rejection/adsorption may differ from one membrane to another due to the membrane pore size, surface charge, hydrophobicity/hydrophilicity, and surface morphology [257]. As shown in Figure 7, highly porous zeolitic imidazolate frameworks (ZIFs) have the highest adsorption rate for the removal of Uranium as ZIF-8 (540.4 mg/g) > Zn/Co-ZIF (527.5 mg/g) > ZIF-9 (448.6 mg/g) > ZIF-67 (368.2 mg/g). This is due to the high elimination capacity for Uranium because of (ZIF-8) large surface area and active metal ion [247]. However, porous graphitic carbon nitride (g-C3N4) has shown a high adsorption rate of Uranium with 149.7 mg/g. The results showed a strong interaction between uranyl and g-C3N4 (Ead = 156.83 kcal/mol) and the most effective sorption site was inside the holes of g-C3N4 [214]. Similarly, as shown in Figure 7, molybdenum disulphide (MoS2) showed a good elimination capacity of Uranium with 117.9, 45.7 and 37.1 (mg/g) [150,215]. This is due to the binding energy through U-S bond which improved by the molybdenum group [150].
In this section, we discuss the rejection rate and adsorption capacity of highly effective membranes including carbon-based membranes, metal oxides, and other nanocomposite membranes.

3.1. Carbon-Based Membranes

Carbon-based membranes including graphene, graphene oxide (GO), and carbon nanotubes (CNTs) are some of the most important nanocomposite membranes for the adsorption of pollutants from wastewater. Carbonaceous nanofiber (CNF), and graphene oxide have shown the highest adsorption rate for the removal of toxic metals of 341.2 and 288 mg/g, respectively, as shown in Figure 8. However, porous graphitic carbon nitride (g-C3N4), and carbon nanotubes (CNTs) have also shown a good adsorption rate for the removal of toxic metals such as Uranium and Indigo carmine (IC) dye of 149.7 and 136 mg/g, respectively. Figure 8 shows the rate of adsorption for toxic metals by carbon-based membranes. In addition, carbon-based membranes have been used for the removal of other toxic metals such as chloroform molecule (CHCl3), oils and organic solvents, and typical pharmaceuticals as shown in Table 8.
As shown in Table 8, carbon-based materials have been mainly used for the rejection/adsorption of toxic metals with high adsorption capacity (qe). Graphene oxide (GO) has been used for the removal of Ni(II), methylene blue (MB), arsenic (As(III)), 17 β- estradiol, and nitrate with high adsorption capacity ranging between 169–288 mg/g. However, carbon nanotubes (CNTs) have been used for the removal of Pb2+, sulfamethoxazole (SMZ) and ketoprofen (KET), bisphenol A, 17a-ethinyl estradiol (EE2), and typical pharmaceuticals, phosphate, indigo carmine (IC) dye, phenol separation, and binary CO2/N2 mixture. The rate of adsorption by carbon nanotubes (CNTs) was also high as it ranged between 0.64–136 mg/g. It can be observed that carbonaceous nanofiber/Ni-Al layered double hydroxide (CNF/LDH) showed an excellent adsorption capacity for the removal of Cu(II) and Cr(VI) with 219.6 and 341 mg/g, respectively.

3.2. Metal Oxides

Binary and ternary metal oxides have been widely studied for supercapacitors, and wastewater treatment because of its high structure stability, low cost, and high electronic conductivity [258,259]. Metal oxides including ZnO, and MnFe2O4 have shown a good adsorption capacity for the removal of barium ions and uranium with 64.4 and 119.9 mg/g, respectively. The high adsorption capacity of metal oxides is due to the ionic bonding which has important consequences for the adsorption of molecules at their surfaces [260]. In addition, Oct-Cu2O NCs showed an excellent adsorption capacity with 1112.6 mg/g of tetracycline on octahedral Cu2O nanocrystals. Table 9 shows rejection/adsorption capacity of metal oxide.

3.3. Other Nanocomposite Membranes

In addition to carbon-based materials and metal oxides, there are other nanocomposite membranes that can be used in the adsorption of pollutants from water. For instance, MoS2 has shown a high adsorption capacity for the removal of toxic metals such as lead (Pb2+) with 638 mg/g as shown in Figure 9. However, MOF-545, M-ATP, and Fe3O4-HBPA-ASA showed similar values of adsorption for Pb2+ as 73, 53.88, 88.36 mg/g, respectively. The high adsorption capacity of MoS2 is due to the strong ionic capture and electrostatic attractions which improve Pb2+ removal capacity under sunlight irradiation [151]. Table 10 shows rejection/adsorption capacity and the role for each nanocomposite membrane in the adsorption process.
As shown in Table 10, nanocomposite materials have shown a good rejection/adsorption capacity on toxic materials. For instance, Granular TiO2-La shows a good adsorption capacity of Arsenic (As III), and Fluoride (F) equal to 114 and 78.4 mg/g, respectively. Similarly, zirconium-based highly porous metal-organic framework (MOF-545) showed an excellent removal of Pb2+ with adsorption capacity up to 73 mg/g. In addition, the developed SPE-MOF-545 can be reused for up to 42 extraction cycles without a significant loss of extraction efficiency. Furthermore, Amino-modified attapulgite (M-ATP) also showed a high removal of Pb2+, and Cu2+ as follows: 53.58 and 28.86 mg/g, respectively.

4. Photocatalytic Degradation of Organic Pollutants

Photocatalytic degradation of organic pollutants has been widely used by researchers because of its great role in removing undesirable contaminants from water and wastewater [261]. The earliest mention of photocatalysis dates back to 1911, when German chemist Alexander Ebner incorporated the concept into his research on the illumination of zinc oxide (ZnO) on the bleaching of the dark blue dye, Prussian blue [262]. Around this time, Brunner and Kosack published an article discussing the degradation of oxalic acid in the presence of uranyl salts under illumination [263], while in 1913, Landau published an article explaining the phenomenon of photocatalysis [264]. Photo-Fenton oxidation is an advanced process uses hydroxyl radicals which increase the rate of degradation of organic pollutants [265]. While heterogeneous photocatalysis is one of the most studied processes for environmental purposes such as water purification and emission cleaning [266]. In this section, we discuss the role of titanium dioxide (TiO2), carbon nanomaterials, metal oxides, and other nanocomposites in the degradation of organic pollutants.

4.1. Titanium Dioxide (TiO2)

Titanium dioxide (TiO2) contributes significantly in the water purification process through degradation of organic pollutants such as Methylene blue (MB), Methyl orange, and Bisphenol A (BPA). In this process, TiO2 acts as a catalyst to accelerate photoreaction for the removal of organic pollutants. Table 11 shows the decomposition rate/degradation efficiency of titanium dioxide (TiO2).
As shown in Table 11, TiO2 has shown a high efficiency in the degradation of organic pollutants. For instance, Wang and colleagues synthesized C, N, F/TiO2NTs nanocomposite material for the degradation of methyl orange under UV-light and simulated sunlight [197]. The results showed a high photocatalytic activity under UV-light for C/TiO2NTs with degradation efficiency up to 100% [197]. Along the same lines, De Santiago et al. synthesized Cr-TiO2 nanocomposite supported on Fe3O4 for the degradation of malachite green dye (MG), and total organic carbon (TOC) [242]. The results showed a high removal under solar radiation up to 100% of MG, and 60% of TOC [242]. Furthermore, Cheng et al. synthesized titanate nanotubes supported on TiO2 (TiO2/TiNTs) for the removal of Cu(II), and phenanthrene under UV-light [200]. The results showed a high adsorption capacity of Cu(II) up to 115 mg/g, while for phenanthrene the degradation efficiency was more than 95% [200].

4.2. Carbon Nanomaterials

Carbon nanomaterials have showed an excellent contribution in the degradation of toxic materials with high efficiency for reduction of highly toxic contaminants up to 99%. As shown in Table 12, carbon nanomaterials have been used for the degradation of many toxic metals such as bisphenol A, Norfloxacin (NX), Methylene blue, and Tetracycline hydrochloride (TCH). For the removal of bisphenol A, Zhu and colleagues synthesized a heterogeneous Fenton catalyst (CNTs/Fh) [116], while Kim et al. synthesized ternary nanocomposites of Fe3O4 nanoparticles@ graphene–poly-N-phenylglycine nanofibers for the adsorption of Cu2+ with high degradation efficiency up to 95% [241]. Similarly, Shi et al. synthesized a CF/BiOBr/Ag3PO4 cloth for the degradation of tetracycline hydrochloride (TCH) up to 90% [154].

4.3. Metal Oxides

In addition to titanium dioxide (TiO2) and carbon nanomaterials, metal oxides can be used in the adsorption of pollutants from water. For instance, ZnO has shown a high adsorption capacity for the removal of toxic metals such as methylene blue, methyl orange dye (MO), and TOC with high degradation efficiency for TOC up to 80.4%. However, Cerium zirconium oxide (CexZryO2) also showed a high degradation efficiency for the removal of sulphonamides of 91.33% as shown in Table 13. Furthermore, Kitchamsetti and colleagues synthesized NiO nanobelt composite material for the removal of organic pollutants such as RhB, CV, MB, and MO. The results showed a high degradation efficiency up to 89%, 76.7%, 82.7%, and 79.1%, respectively [204].

4.4. Other Nanocomposites

For other nanocomposites, we demonstrate the decomposition rate/degradation efficiency for the most effective compounds for the degradation of organic pollutants such as nickel (Ni(II)), methylene blue (MB), sulfamethoxazole (SDZ), and trimethoprim (TMP). As shown in Table 14, for the degradation of MB, sodium titanate nanotubes (Na-TNT) showed a high degradation efficiency up to 99.5%, while Fe2O3-PC removed 75% of MB. For the degradation of sulfamethoxazole (SDZ), carbon dots/g-C3N4 (C-CN) heterostructures showed a high degradation efficiency up to 92.8% for the 3C-CN heterostructure. However, other nanocomposites such as black phosphorus quantum dots/Tubular g-C3N4 (BPQDs/TCN), CdSe-Ag-WO3-Ag photocatalyst, and 1D/2D W18O49/g-C3N4 nanocomposites, also showed a high degradation efficiency for the degradation of organic pollutants up to 96%.

5. Future Direction of Nanomembrane Adsorption Processes in Wastewater Treatment

In this review, we have discussed the role of the adsorption process in water and wastewater treatment through nanomembranes experimentally and by simulation. Below we will discuss the latest expected developments of the adsorption process in the near future with some recommendations.
  • With the rapid development of simulation software in some important DFT codes such as PBE, B3LYP and PAW, it will be easier to understand the physical and chemical properties of the adsorption process to fill the scientific gaps in realizing the adsorption mechanism, isotherm, kinetics, thermodynamics and other aspects of the adsorption process.
  • Further economic feasibility studies should be conducted on adsorbents including the cost effectiveness of the choices of the materials, which is an important aspect of adsorption investigations.
  • A huge improvement in the synthesis of nanomaterials using simulation will become possible by linking the density functional theory (DFT) codes using software such as Material Studio and Reactive forefield (ReaxFF) with the molecular dynamic (MD) simulation which will give more realism in acquiring accurate results before starting the experimental work. This step can reduce costs of conducting trials and save time.
  • Evolution in the ability and durability of nanomembranes in selectivity of undesirable materials by adsorption which increase the adsorption capacity (qe). This is possible by improving the mechanical properties of the nanomembranes by creating special nanocomposites such as graphene/TiO2, and graphene/MoS2. These two nanocomposites have proven their ability to expel salt and permeate water with high efficiency, so we expect a high adsorption capacity (qe) from them.
  • More comprehensive studies should be conducted on the effect of multi-layer membranes in the adsorption process, which is expected to increase the adsorption capacity (qe) due to the increase in attractions between organic pollutants and membranes. In addition, it is possible to use different layers in the same system which can adsorb different pollutants at the same time. We recommend simulating the system using a molecular dynamic simulation software using two different layers and then testing the possibility of adsorption on different organic pollutants.
  • More studies should be conducted on the possibility of developing more effective forcefields which is highly required in some molecular dynamics simulation software. Creating and developing high effective forcefields will increase the possibility of simulating all kinds of atoms and molecules with high accuracy without errors.

6. Conclusions

Based on the unique properties of the adsorption technique and its excellent contribution in water and wastewater treatment, it has attracted researchers with great interest in the membrane-based separation field. In this review paper, the preparation of nanomembranes and nanocomposite materials were summarized by experimental and simulation methods. Then we focused on water treatment by membrane technology using the adsorption technique. The main characteristics related to adsorption technology are discussed including the adsorbed/rejected material, rate of rejection/adsorption capacity, and the role of the nanocomposite membrane. In synthesis of nanocomposite materials, we have discussed the latest development techniques such as the hydrothermal method, chemical vapor deposition (CVD), and one-pot synthesis. In this review, the hydrothermal method was the most common method used as it contributed in 56% of the total papers reviewed, while chemical vapor deposition and one-pot synthesis methods contributed 7% and 6%, respectively. This is because of the hydrothermal method’s low-cost synthesis, being simple and easy. Graphene oxide (GO) showed promising results in water and wastewater treatment by using the adsorption technique. GO membrane has showed a high adsorption capacity up to 288 mg/g for the removal of As(III), while for the removal of Ni (II), the adsorption capacity equalled 197.8 mg/g. In addition, GO has been used for the removal of 17 β- estradiol with high adsorption capacity up to 169.49 mg/g. In addition, GO contributed to the degradation of organic pollutants such as methylene blue (MB) with a degradation efficiency up to 99.3%. In addition to GO, metal oxides and other nanocomposites has also showed a high adsorption capacity for the removal of toxic metals such as lead (Pb2+), barium ions and uranium. Highly porous zeolitic imidazolate frameworks (ZIFs) has the highest adsorption rate for the removal of Uranium of 540.4 mg/g. For the removal of bisphenol A, PyTTA-Dva-COF membrane had the highest adsorption capacity of 285 mg/g. For the degradation of organic pollutants, TiO2 nanoflowers showed an excellent contribution for the degradation of Bisphenol A (BPA), Diphenyl phenol, P-tert-butyl phenol, and Resorcinol under UV-light up to 95%. Moreover, Cr-TiO2 degraded 100% of Malachite green dye (MG), and 60% of total organic carbon (TOC). In a comparison with other nanocomposite materials, metal oxides including ZnO@C, Cerium zirconium oxide (CexZryO2), ZnO/Al2O3, and NiO nanobelt have also showed good results for the degradation of MB, sulphonamides up to 91.33%, TOC, and other organic pollutants. Finally, special attention has been paid to the future direction of nanomembrane adsorption processes in water and wastewater treatment. The need for an economic feasibility study on adsorbents was mainly highlighted with the need to connect DFT with MD simulation for more realism before starting the experimental work.

Author Contributions

Q.I., writing, methodology, reviewing; L.C., supervision, editing, validation; S.G., concept, supervision, reviewing, editing and validation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the IT Sligo President’s bursary.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to acknowledge the funding of the IT Sligo President’s bursary.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Connor, R. The United Nations World Water Development Report 2015: Water for a Sustainable World; UNESCO Publishing: Paris, France, 2015; Volume 1. [Google Scholar]
  2. Cosgrove, W.J.; Loucks, D.P. Water management: Current and future challenges and research directions. Water Resour. Res. 2015, 51, 4823–4839. [Google Scholar] [CrossRef] [Green Version]
  3. Daud, M.; Nafees, M.; Ali, S.; Rizwan, M.; Bajwa, R.A.; Shakoor, M.B.; Arshad, M.U.; Chatha, S.A.S.; Deeba, F.; Murad, W. Drinking water quality status and contamination in Pakistan. BioMed Res. Int. 2017, 2017, 7908183. [Google Scholar] [CrossRef] [PubMed]
  4. United Nations Environment Programme. Let It Flow: Improving Water Quantity and Quality in Tanzania’s Rufiji River Basin. Available online: https://www.unep.org/news-and-stories/story/let-it-flow-improving-water-quantity-and-quality-tanzanias-rufiji-river (accessed on 9 June 2021).
  5. Costa, D.; Burlando, P.; Priadi, C. The importance of integrated solutions to flooding and water quality problems in the tropical megacity of Jakarta. Sustain. Cities Soc. 2016, 20, 199–209. [Google Scholar] [CrossRef]
  6. Choubisa, S.L. Fluoride distribution in drinking groundwater in Rajasthan, India. Curr. Sci. 2018, 114, 1851–1857. [Google Scholar] [CrossRef]
  7. Edokpayi, J.N.; Odiyo, J.O.; Durowoju, O.S. Impact of wastewater on surface water quality in developing countries: A case study of South Africa. In Water Quality; IntechOpen: London, UK, 2017; pp. 401–416. [Google Scholar]
  8. Martínez-Santos, P. Does 91% of the world’s population really have “sustainable access to safe drinking water”? Int. J. Water Resour. Dev. 2017, 33, 514–533. [Google Scholar] [CrossRef]
  9. Hotza, D.; Di Luccio, M.; Wilhelm, M.; Iwamoto, Y.; Bernard, S.; da Costa, J.C.D. Silicon carbide filters and porous membranes: A review of processing, properties, performance, and application. J. Membr. Sci. 2020, 610, 118193. [Google Scholar] [CrossRef]
  10. Liu, H.; Cao, C.; Huang, J.; Chen, Z.; Chen, G.; Lai, Y. Progress on particulate matter filtration technology: Basic concepts, advanced materials, and performances. Nanoscale 2020, 12, 437–453. [Google Scholar] [CrossRef]
  11. Abejón, R.; Garea, A. A bibliometric analysis of research on arsenic in drinking water during the 1992–2012 period: An outlook to treatment alternatives for arsenic removal. J. Water Process Eng. 2015, 6, 105–119. [Google Scholar] [CrossRef]
  12. Boyd, C.E. Water quality protection. In Water Quality; Springer: Berlin/Heidelberg, Germany, 2020; pp. 379–409. [Google Scholar]
  13. Symons, G.E. Water treatment through the ages. J.-Am. Water Work. Assoc. 2006, 98, 87–98. [Google Scholar] [CrossRef]
  14. Crittenden, J.C.; Trussell, R.R.; Hand, D.W.; Howe, K.; Tchobanoglous, G. MWH’s Water Treatment: Principles and Design; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  15. Phu, N. Evaluating Pilot Scale Slow Sand Filtration Columns to Effectively Remove Emerging Contaminants in Recycled Water. Ph.D. Thesis, University of California, Los Angeles, CA, USA, 2016. [Google Scholar]
  16. Judd, S.J. Membrane technology costs and me. Water Res. 2017, 122, 1–9. [Google Scholar] [CrossRef]
  17. Baker, R.W. Membrane Technology and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2012. [Google Scholar]
  18. Umar, M.; Aziz, H.A. Photocatalytic degradation of organic pollutants in water. Org. Pollut.-Monit. Risk Treat. 2013, 8, 196–197. [Google Scholar]
  19. Abbasi-Garravand, E.; Catherine, N.M.; Laflamme, C.B.; Clairet, G. Investigation of the fouling effect on a commercial semi-permeable membrane in the pressure retarded osmosis (PRO) process. Sep. Purif. Technol. 2018, 193, 81–90. [Google Scholar] [CrossRef] [Green Version]
  20. Puretec Industrial Waters. What is Reverse Osmosis? Available online: https://puretecwater.com/reverse-osmosis/what-is-reverse-osmosis (accessed on 9 June 2021).
  21. Alsarayreh, A.A.; Al-Obaidi, M.; Al-Hroub, A.; Patel, R.; Mujtaba, I. Evaluation and minimisation of energy consumption in a medium-scale reverse osmosis brackish water desalination plant. J. Clean. Prod. 2020, 248, 119220. [Google Scholar] [CrossRef]
  22. Gude, V.G. Energy consumption and recovery in reverse osmosis. Desalination Water Treat. 2011, 36, 239–260. [Google Scholar] [CrossRef]
  23. Lee, K.P.; Arnot, T.C.; Mattia, D. A review of reverse osmosis membrane materials for desalination—Development to date and future potential. J. Membr. Sci. 2011, 370, 1–22. [Google Scholar] [CrossRef] [Green Version]
  24. Homaeigohar, S.; Elbahri, M. Graphene membranes for water desalination. NPG Asia Mater. 2017, 9, e427. [Google Scholar] [CrossRef] [Green Version]
  25. Padaki, M.; Murali, R.S.; Abdullah, M.S.; Misdan, N.; Moslehyani, A.; Kassim, M.; Hilal, N.; Ismail, A. Membrane technology enhancement in oil–water separation. A review. Desalination 2015, 357, 197–207. [Google Scholar] [CrossRef]
  26. Belfort, G. Membrane filtration with liquids: A global approach with prior successes, new developments and unresolved challenges. Angew. Chem. Int. Ed. 2019, 58, 1892–1902. [Google Scholar] [CrossRef]
  27. Yang, Z.; Zhou, Y.; Feng, Z.; Rui, X.; Zhang, T.; Zhang, Z. A review on reverse osmosis and nanofiltration membranes for water purification. Polymers 2019, 11, 1252. [Google Scholar] [CrossRef] [Green Version]
  28. Shon, H.; Phuntsho, S.; Chaudhary, D.; Vigneswaran, S.; Cho, J. Nanofiltration for water and wastewater treatment—A mini review. Drink. Water Eng. Sci. 2013, 6, 47–53. [Google Scholar] [CrossRef] [Green Version]
  29. Ibrahim, Q.; Akbarzadeh, R. A photocatalytic TiO2/graphene bilayer membrane design for water desalination: A molecular dynamic simulation. J. Mol. Modeling 2020, 26, 165. [Google Scholar] [CrossRef] [PubMed]
  30. Zhang, Y.; Tang, Z.-R.; Fu, X.; Xu, Y.-J. Engineering the unique 2D mat of graphene to achieve graphene-TiO2 nanocomposite for photocatalytic selective transformation: What advantage does graphene have over its forebear carbon nanotube? ACS Nano 2011, 5, 7426–7435. [Google Scholar] [CrossRef] [PubMed]
  31. Bhanvase, B.; Shende, T.; Sonawane, S. A review on graphene–TiO2 and doped graphene–TiO2 nanocomposite photocatalyst for water and wastewater treatment. Environ. Technol. Rev. 2017, 6, 1–14. [Google Scholar] [CrossRef]
  32. Madhura, L.; Kanchi, S.; Sabela, M.I.; Singh, S.; Bisetty, K.; Inamuddin. Membrane technology for water purification. Environ. Chem. Lett. 2018, 16, 343–365. [Google Scholar] [CrossRef]
  33. Heiranian, M.; Farimani, A.B.; Aluru, N.R. Water desalination with a single-layer MoS2 nanopore. Nat. Commun. 2015, 6, 8616. [Google Scholar] [CrossRef] [Green Version]
  34. Cohen-Tanugi, D.; Grossman, J.C. Water desalination across nanoporous graphene. Nano Lett. 2012, 12, 3602–3608. [Google Scholar] [CrossRef]
  35. Kou, J.; Yao, J.; Wu, L.; Zhou, X.; Lu, H.; Wu, F.; Fan, J. Nanoporous two-dimensional MoS2 membranes for fast saline solution purification. Phys. Chem. Chem. Phys. 2016, 18, 22210–22216. [Google Scholar] [CrossRef]
  36. Cohen-Tanugi, D.; Lin, L.-C.; Grossman, J.C. Multilayer nanoporous graphene membranes for water desalination. Nano Lett. 2016, 16, 1027–1033. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, Z.; Sim, A.; Urban, J.J.; Mi, B. Removal and recovery of heavy metal ions by two-dimensional MoS2 nanosheets: Performance and mechanisms. Environ. Sci. Technol. 2018, 52, 9741–9748. [Google Scholar] [CrossRef]
  38. Nunes, S.P.; Culfaz-Emecen, P.Z.; Ramon, G.Z.; Visser, T.; Koops, G.H.; Jin, W.; Ulbricht, M. Thinking the future of membranes: Perspectives for advanced and new membrane materials and manufacturing processes. J. Membr. Sci. 2020, 598, 117761. [Google Scholar] [CrossRef]
  39. Suresh, R.; Rajendran, S.; Kumar, P.S.; Vo, D.-V.N.; Cornejo-Ponce, L. Recent advancements of spinel ferrite based binary nanocomposite photocatalysts in wastewater treatment. Chemosphere 2021, 274, 129734. [Google Scholar] [CrossRef] [PubMed]
  40. Marcolongo, D.M.S.; Nocito, F.; Ditaranto, N.; Comparelli, R.; Aresta, M.; Dibenedetto, A. Opto-electronic characterization of photocatalysts based on p, n-junction ternary and quaternary mixed oxides semiconductors (Cu2O-In2O3 and Cu2O-In2O3-TiO2). Catalysts 2022, 12, 153. [Google Scholar] [CrossRef]
  41. Dineshbabu, N.; Jayaprakash, R.N.; Karuppasamy, P.; Arun, T.; Judith Vijaya, J.; Esther Nimshi, R.; Pandian, M.S.; Maria Packiam, S.; Ramasamy, P. Investigation on Tetracycline degradation and bactericidal properties of binary and ternary ZnO/NiO/g-C3N4 composites prepared by a facile co-precipitation method. J. Environ. Chem. Eng. 2022, 10, 107368. [Google Scholar] [CrossRef]
  42. Rassoulinejad-Mousavi, S.M.; Azamat, J.; Khataee, A.; Zhang, Y. Molecular dynamics simulation of water purification using zeolite MFI nanosheets. Sep. Purif. Technol. 2020, 234, 116080. [Google Scholar] [CrossRef]
  43. Guan, W.; Dai, Y.; Dong, C.; Yang, X.; Xi, Y. Zeolite imidazolate framework (ZIF)-based mixed matrix membranes for CO2 separation: A review. J. Appl. Polym. Sci. 2020, 137, 48968. [Google Scholar] [CrossRef]
  44. Aloulou, W.; Aloulou, H.; Khemakhem, M.; Duplay, J.; Daramola, M.; Ben Amar, R. Synthesis and characterization of clay-based ultrafiltration membranes supported on natural zeolite for removal of heavy metals from wastewater. Environ. Technol. Innov. 2020, 18, 100794. [Google Scholar] [CrossRef]
  45. Zalfani, M.; van der Schueren, B.; Mahdouani, M.; Bourguiga, R.; Yu, W.-B.; Wu, M.; Deparis, O.; Li, Y.; Su, B.-L. ZnO quantum dots decorated 3DOM TiO2 nanocomposites: Symbiose of quantum size effects and photonic structure for highly enhanced photocatalytic degradation of organic pollutants. Appl. Catal. B Environ. 2016, 199, 187–198. [Google Scholar] [CrossRef]
  46. Chacko, L.; Jayaraj, M.; Aneesh, P. Excitation-wavelength dependent upconverting surfactant free MoS2 nanoflakes grown by hydrothermal method. J. Lumin. 2017, 192, 6–10. [Google Scholar] [CrossRef]
  47. Abdelaziz Aboelazm, E.A.; Mohammed Ali, G.A.; Algarni, H.; Chong, K.F. Flakes size-dependent optical and electrochemical properties of MoS2. Curr. Nanosci. 2018, 14, 416–420. [Google Scholar] [CrossRef]
  48. Li, B.L.; Zou, H.L.; Lu, L.; Yang, Y.; Lei, J.L.; Luo, H.Q.; Li, N.B. Size-dependent optical absorption of layered MoS2 and dna oligonucleotides induced dispersion behavior for label-free detection of single-nucleotide polymorphism. Adv. Funct. Mater. 2015, 25, 3541–3550. [Google Scholar] [CrossRef]
  49. Yan, H.; Liu, L.; Wang, R.; Zhu, W.; Ren, X.; Luo, L.; Zhang, X.; Luo, S.; Ai, X.; Wang, J. Binary composite MoS2/TiO2 nanotube arrays as a recyclable and efficient photocatalyst for solar water disinfection. Chem. Eng. J. 2020, 401, 126052. [Google Scholar] [CrossRef]
  50. Jiang, Q.; Wang, S.; Li, X.; Han, Z.; Zhao, C.; Di, T.; Liu, S.; Cheng, Z. Controllable growth of MoS2 nanosheets on TiO2 burst nanotubes and their photocatalytic activity. RSC Adv. 2020, 10, 40904–40915. [Google Scholar] [CrossRef]
  51. Khan, S.H.; Pathak, B. Zinc oxide based photocatalytic degradation of persistent pesticides: A comprehensive review. Environ. Nanotechnol. Monit. Manag. 2020, 13, 100290. [Google Scholar] [CrossRef]
  52. Cai, Z.; Dwivedi, A.D.; Lee, W.-N.; Zhao, X.; Liu, W.; Sillanpää, M.; Zhao, D.; Huang, C.-H.; Fu, J. Application of nanotechnologies for removing pharmaceutically active compounds from water: Development and future trends. Environ. Sci. Nano 2018, 5, 27–47. [Google Scholar] [CrossRef]
  53. Zheng, L.; Teng, F.; Ye, X.; Zheng, H.; Fang, X. Photo/electrochemical applications of metal sulfide/TiO2 heterostructures. Adv. Energy Mater. 2020, 10, 1902355. [Google Scholar] [CrossRef]
  54. Devi, L.G.; Kavitha, R. A review on plasmonic metal-TiO2 composite for generation, trapping, storing and dynamic vectorial transfer of photogenerated electrons across the Schottky junction in a photocatalytic system. Appl. Surf. Sci. 2016, 360, 601–622. [Google Scholar] [CrossRef]
  55. Duan, Y.; Liang, L.; Lv, K.; Li, Q.; Li, M. TiO2 faceted nanocrystals on the nanofibers: Homojunction TiO2 based Z-scheme photocatalyst for air purification. Appl. Surf. Sci. 2018, 456, 817–826. [Google Scholar] [CrossRef]
  56. Liu, C.; Faria, A.F.; Jackson, J.; He, Q.; Ma, J. Enhancing the anti-fouling and fouling removal properties of thin-film composite membranes through an intercalated functionalization method. Environ. Sci. Water Res. Technol. 2021, 7, 1336–1347. [Google Scholar] [CrossRef]
  57. Lu, F.; Astruc, D. Nanocatalysts and other nanomaterials for water remediation from organic pollutants. Coord. Chem. Rev. 2020, 408, 213180. [Google Scholar] [CrossRef]
  58. Zou, W.; Gao, B.; Ok, Y.S.; Dong, L. Integrated adsorption and photocatalytic degradation of volatile organic compounds (VOCs) using carbon-based nanocomposites: A critical review. Chemosphere 2019, 218, 845–859. [Google Scholar] [CrossRef]
  59. Reza, K.M.; Kurny, A.; Gulshan, F. Parameters affecting the photocatalytic degradation of dyes using TiO2: A review. Appl. Water Sci. 2017, 7, 1569–1578. [Google Scholar] [CrossRef] [Green Version]
  60. Patnaik, S.; Kumar Das, K.; Mohanty, A.; Parida, K. Enhanced photo catalytic reduction of Cr(VI) over polymer-sensitized g-C3N4/ZnFe2O4 and its synergism with phenol oxidation under visible light irradiation. Catal. Today 2018, 315, 52–66. [Google Scholar] [CrossRef]
  61. Cai, X.; Li, J.; Liu, Y.; Yan, Z.; Tan, X.; Liu, S.; Zeng, G.; Gu, Y.; Hu, X.; Jiang, L. Titanium dioxide-coated biochar composites as adsorptive and photocatalytic degradation materials for the removal of aqueous organic pollutants. J. Chem. Technol. Biotechnol. 2018, 93, 783–791. [Google Scholar] [CrossRef]
  62. Tereshchenko, A.; Bechelany, M.; Viter, R.; Khranovskyy, V.; Smyntyna, V.; Starodub, N.; Yakimova, R. Optical biosensors based on ZnO nanostructures: Advantages and perspectives. A review. Sens. Actuators B Chem. 2016, 229, 664–677. [Google Scholar] [CrossRef] [Green Version]
  63. Tran, H.N.; You, S.J.; Chao, H.P. Fast and efficient adsorption of methylene green 5 on activated carbon prepared from new chemical activation method. J. Environ. Manag. 2017, 188, 322–336. [Google Scholar] [CrossRef]
  64. Awual, M.R.; Khraisheh, M.; Alharthi, N.H.; Luqman, M.; Islam, A.; Karim, M.R.; Rahman, M.M.; Khaleque, M.A. Efficient detection and adsorption of cadmium (II) ions using innovative nano-composite materials. Chem. Eng. J. 2018, 343, 118–127. [Google Scholar] [CrossRef]
  65. Rashid, R.; Shafiq, I.; Akhter, P.; Iqbal, M.J.; Hussain, M. A state-of-the-art review on wastewater treatment techniques: The effectiveness of adsorption method. Environ. Sci. Pollut. Res. 2021, 28, 9050–9066. [Google Scholar] [CrossRef]
  66. Rashed, M.N. Adsorption technique for the removal of organic pollutants from water and wastewater. Org. Pollut.-Monit. Risk Treat. 2013, 7, 167–194. [Google Scholar]
  67. Clark, A. The Theory of Adsorption and Catalysis; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  68. Wang, C.; Lin, G.; Zhao, J.; Wang, S.; Zhang, L. Enhancing Au (III) adsorption capacity and selectivity via engineering MOF with mercapto-1,3,4-thiadiazole. Chem. Eng. J. 2020, 388, 124221. [Google Scholar] [CrossRef]
  69. Deb, A.; Debnath, A.; Saha, B. Sono-assisted enhanced adsorption of eriochrome Black-T dye onto a novel polymeric nanocomposite: Kinetic, isotherm, and response surface methodology optimization. J. Dispers. Sci. Technol. 2021, 42, 1579–1592. [Google Scholar] [CrossRef]
  70. Liu, J.; Wei, Y.; Li, P.; Zhao, Y.; Zou, R. Selective H2S/CO2 separation by metal–organic frameworks based on chemical-physical adsorption. J. Phys. Chem. C 2017, 121, 13249–13255. [Google Scholar] [CrossRef]
  71. Ruiz, V.G.; Liu, W.; Tkatchenko, A. Density-functional theory with screened van der Waals interactions applied to atomic and molecular adsorbates on close-packed and non-close-packed surfaces. Phys. Rev. B 2016, 93, 035118. [Google Scholar] [CrossRef] [Green Version]
  72. Yang, S.; Evans, J. Metering and dispensing of powder; the quest for new solid freeforming techniques. Powder Technol. 2007, 178, 56–72. [Google Scholar] [CrossRef]
  73. Li, Q.; Rudolph, V.; Peukert, W. London-van der Waals adhesiveness of rough particles. Powder Technol. 2006, 161, 248–255. [Google Scholar] [CrossRef]
  74. Welker, R.W. Basics and sampling of particles for size analysis and identification. In Developments in Surface Contamination and Cleaning; Elsevier: Amsterdam, The Netherlands, 2012; pp. 1–80. [Google Scholar]
  75. Sotomayor, F.J.; Cychosz, K.A.; Thommes, M. Characterization of micro/mesoporous materials by physisorption: Concepts and case studies. Acc. Mater. Surf. Res. 2018, 3, 34–50. [Google Scholar]
  76. Pan, J.H.; Zhang, X.; Du, A.J.; Bai, H.; Ng, J.; Sun, D. A hierarchically assembled mesoporous ZnO hemisphere array and hollow microspheres for photocatalytic membrane water filtration. Phys. Chem. Chem. Phys. 2012, 14, 7481–7489. [Google Scholar] [CrossRef] [PubMed]
  77. Zhou, R.; Ren, P.-F.; Yang, H.-C.; Xu, Z.-K. Fabrication of antifouling membrane surface by poly(sulfobetaine methacrylate)/polydopamine co-deposition. J. Membr. Sci. 2014, 466, 18–25. [Google Scholar] [CrossRef]
  78. Peng, B.; Li, Y.; Zhao, Z.; Chen, Y.; Han, C.C. Facile surface modification of PVDF microfiltration membrane by strong physical adsorption of amphiphilic copolymers. J. Appl. Polym. Sci. 2013, 130, 3112–3121. [Google Scholar] [CrossRef]
  79. Bikerman, J.J. Surface Chemistry: Theory and Applications; Elsevier: Amsterdam, The Netherlands, 2013. [Google Scholar]
  80. Matouq, M.; Jildeh, N.; Qtaishat, M.; Hindiyeh, M.; Al Syouf, M.Q. The adsorption kinetics and modeling for heavy metals removal from wastewater by Moringa pods. J. Environ. Chem. Eng. 2015, 3, 775–784. [Google Scholar] [CrossRef]
  81. Zhang, F.; Li, J.; Tan, J.; Wang, B.; Huang, F. Advance of the treatment of heavy metal wastewater by adsorption. Chem. Ind. Eng. Prog. 2013, 32, 2749–2756. [Google Scholar]
  82. Renu, M.A.; Singh, K.; Upadhyaya, S.; Dohare, R. Removal of heavy metals from wastewater using modified agricultural adsorbents. Mater. Today Proc. 2017, 4, 10534–10538. [Google Scholar] [CrossRef]
  83. Liu, Z.; Wang, C.; Wu, Y.; Geng, L.; Zhang, X.; Zhang, D.; Hu, H.; Zhang, Y.; Li, X.; Liu, W. Synthesis of uniform-sized and microporous MIL-125 (Ti) to boost arsenic removal by chemical adsorption. Polyhedron 2021, 196, 114980. [Google Scholar] [CrossRef]
  84. Martínez, C.M.; Canle L., M.; Fernández, M.I.; Santaballa, J.A.; Faria, J. Kinetics and mechanism of aqueous degradation of carbamazepine by heterogeneous photocatalysis using nanocrystalline TiO2, ZnO and multi-walled carbon nanotubes–anatase composites. Appl. Catal. B Environ. 2011, 102, 563–571. [Google Scholar] [CrossRef]
  85. Son, H.-S.; Lee, S.-J.; Cho, I.-H.; Zoh, K.-D. Kinetics and mechanism of TNT degradation in TiO2 photocatalysis. Chemosphere 2004, 57, 309–317. [Google Scholar] [CrossRef]
  86. Ji, P.; Zhang, J.; Chen, F.; Anpo, M. Study of adsorption and degradation of acid orange 7 on the surface of CeO2 under visible light irradiation. Appl. Catal. B Environ. 2009, 85, 148–154. [Google Scholar] [CrossRef]
  87. Luo, Y.; Wei, X.; Gao, B.; Zou, W.; Zheng, Y.; Yang, Y.; Zhang, Y.; Tong, Q.; Dong, L. Synergistic adsorption-photocatalysis processes of graphitic carbon nitrate (g-C3N4) for contaminant removal: Kinetics, models, and mechanisms. Chem. Eng. J. 2019, 375, 122019. [Google Scholar] [CrossRef]
  88. Lan, L.; Huang, Y.; Dan, Y.; Jiang, L. Conjugated porous polymers for gaseous toluene adsorption in humid atmosphere. React. Funct. Polym. 2021, 159, 104804. [Google Scholar] [CrossRef]
  89. Zhang, J.; Cai, D.; Zhang, G.; Cai, C.; Zhang, C.; Qiu, G.; Zheng, K.; Wu, Z. Adsorption of methylene blue from aqueous solution onto multiporous palygorskite modified by ion beam bombardment: Effect of contact time, temperature, pH and ionic strength. Appl. Clay Sci. 2013, 83, 137–143. [Google Scholar] [CrossRef]
  90. Han, Y.; Cao, X.; Ouyang, X.; Sohi, S.P.; Chen, J. Adsorption kinetics of magnetic biochar derived from peanut hull on removal of Cr (VI) from aqueous solution: Effects of production conditions and particle size. Chemosphere 2016, 145, 336–341. [Google Scholar] [CrossRef]
  91. Thakur, K.; Kandasubramanian, B. Graphene and graphene oxide-based composites for removal of organic pollutants: A review. J. Chem. Eng. Data 2019, 64, 833–867. [Google Scholar] [CrossRef]
  92. Zhang, S.; Zhang, Q.; Wang, H.; Ni, Y.; Zhu, Z. Absorption behaviors study on doped Li4SiO4 under a humidified atmosphere with low CO2 concentration. Int. J. Hydrogen Energy 2014, 39, 17913–17920. [Google Scholar] [CrossRef]
  93. Ben-Mansour, R.; Habib, M.A.; Bamidele, O.E.; Basha, M.; Qasem, N.A.A.; Peedikakkal, A.; Laoui, T.; Ali, M. Carbon capture by physical adsorption: Materials, experimental investigations and numerical modeling and simulations–a review. Appl. Energy 2016, 161, 225–255. [Google Scholar] [CrossRef]
  94. Chen, M.; He, Y.; Zhu, J.; Shuai, Y.; Jiang, B.; Huang, Y. An experimental investigation on sunlight absorption characteristics of silver nanofluids. Sol. Energy 2015, 115, 85–94. [Google Scholar] [CrossRef]
  95. Reimhult, E.; Höök, F.; Kasemo, B. Intact vesicle adsorption and supported biomembrane formation from vesicles in solution: Influence of surface chemistry, vesicle size, temperature, and osmotic pressure. Langmuir 2003, 19, 1681–1691. [Google Scholar] [CrossRef]
  96. Woellner, M.; Hausdorf, S.; Klein, N.; Mueller, P.; Smith, M.W.; Kaskel, S. Adsorption and detection of hazardous trace gases by metal–organic frameworks. Adv. Mater. 2018, 30, 1704679. [Google Scholar] [CrossRef]
  97. Laabd, M.; El Jaouhari, A.; Bazzaoui, M.; Albourine, A.; El Jazouli, H. Adsorption of benzene-polycarboxylic acids on the electrosynthesized polyaniline films: Experimental and DFT calculation. J. Polym. Environ. 2017, 25, 359–369. [Google Scholar] [CrossRef]
  98. Bergaoui, M.; Nakhli, A.; Benguerba, Y.; Khalfaoui, M.; Erto, A.; Soetaredjo, F.E.; Ismadji, S.; Ernst, B. Novel insights into the adsorption mechanism of methylene blue onto organo-bentonite: Adsorption isotherms modeling and molecular simulation. J. Mol. Liq. 2018, 272, 697–707. [Google Scholar] [CrossRef]
  99. Zhang, J.; Tong, H.; Pei, W.; Liu, W.; Shi, F.; Li, Y.; Huo, Y. Integrated photocatalysis-adsorption-membrane separation in rotating reactor for synergistic removal of RhB. Chemosphere 2021, 270, 129424. [Google Scholar] [CrossRef]
  100. Zhou, Q.; Huang, W.; Xu, C.; Liu, X.; Yang, K.; Li, D.; Hou, Y.; Dionysiou, D.D. Novel hierarchical carbon quantum dots-decorated BiOCl nanosheet/carbonized eggshell membrane composites for improved removal of organic contaminants from water via synergistic adsorption and photocatalysis. Chem. Eng. J. 2021, 420, 129582. [Google Scholar] [CrossRef]
  101. Huang, J.; Huang, D.; Zeng, F.; Ma, L.; Wang, Z. Photocatalytic MOF fibrous membranes for cyclic adsorption and degradation of dyes. J. Mater. Sci. 2021, 56, 3127–3139. [Google Scholar] [CrossRef]
  102. Schmidt, J.; Shi, J.; Borlido, P.; Chen, L.; Botti, S.; Marques, M.A. Predicting the thermodynamic stability of solids combining density functional theory and machine learning. Chem. Mater. 2017, 29, 5090–5103. [Google Scholar] [CrossRef]
  103. Allen, M.P.; Tildesley, D.J. Computer Simulation of Liquids; Oxford University Press: Oxford, UK, 2017. [Google Scholar]
  104. Tu, C.-H.; Hsu, H.-H.; Chen, J.-H.; Chen, C.-H.; Hung, S.-H. Performance and power profiling for emulated android systems. ACM Trans. Des. Autom. Electron. Syst. 2014, 19, 1–25. [Google Scholar] [CrossRef]
  105. Mourtzis, D.; Doukas, M.; Bernidaki, D. Simulation in manufacturing: Review and challenges. Procedia Cirp 2014, 25, 213–229. [Google Scholar] [CrossRef] [Green Version]
  106. Brandimarte, P. Handbook in Monte Carlo Simulation: Applications in Financial Engineering, Risk Management, and Economics; John Wiley & Sons: Hoboken, NJ, USA, 2014. [Google Scholar]
  107. Hensen, J.L.; Lamberts, R. Building Performance Simulation for Design and Operation; Routledge: London, UK, 2012. [Google Scholar]
  108. Vo, H.T. On Efficient Algorithms for Stochastic Simulation of Biochemical Reaction Systems. Ph.D. Thesis, University of Trento, Trento, Italy, 2013. [Google Scholar]
  109. Fatemi, S.M.; Abbasi, Z.; Rajabzadeh, H.; Hashemizadeh, S.A.; Deldar, A.N. A review of recent advances in molecular simulation of graphene-derived membranes for gas separation. Eur. Phys. J. D 2017, 71, 194. [Google Scholar] [CrossRef]
  110. Yang, P.-Y.; Ju, S.-P.; Hsieh, H.-S.; Lin, J.-S.; Hsieh, J.-Y. Electrolytic molecule in-pore structure and capacitance of supercapacitors with nanoporous carbon electrodes: A coarse-grained molecular dynamics study. Comput. Mater. Sci. 2019, 166, 293–302. [Google Scholar] [CrossRef]
  111. Hu, Q.; Xue, M.; Shen, C.; Zhang, Z.; Guo, W. Graphynes for water desalination and gas separation. Adv. Mater. 2019, 42, 1803772. [Google Scholar]
  112. Rebelo, P.; Pacheco, J.G.; Voroshylova, I.V.; Melo, A.; Cordeiro, M.N.D.; Delerue-Matos, C. Rational development of molecular imprinted carbon paste electrode for Furazolidone detection: Theoretical and experimental approach. Sens. Actuators B Chem. 2021, 329, 129112. [Google Scholar] [CrossRef]
  113. de Oliveira, P.V.; Zanella, I.; Bulhoes, L.O.S.; Fagan, S.B. Adsorption of 17 β-estradiol in graphene oxide through the competing methanol co-solvent: Experimental and computational analysis. J. Mol. Liq. 2021, 321, 114738. [Google Scholar] [CrossRef]
  114. Arnold, J.G.; Moriasi, D.N.; Gassman, P.W.; Abbaspour, K.C.; White, M.J.; Srinivasan, R.; Santhi, C.; Harmel, R.D.; van Griensven, A.; van Liew, M.W.; et al. SWAT: Model use, calibration, and validation. Trans. ASABE 2012, 55, 1491–1508. [Google Scholar] [CrossRef]
  115. Malleson, N. Calibration of simulation models. Encycl. Criminol. Crim. Justice 2014, 40, 115–118. [Google Scholar]
  116. Zhu, R.; Zhu, Y.; Xian, H.; Yan, L.; Fu, H.; Zhu, G.; Xi, Y.; Zhu, J.; He, H. CNTs/ferrihydrite as a highly efficient heterogeneous Fenton catalyst for the degradation of bisphenol A: The important role of CNTs in accelerating Fe(III)/Fe(II) cycling. Appl. Catal. B Environ. 2020, 270, 118891. [Google Scholar] [CrossRef]
  117. Deng, Y.; Wu, Y.; Chen, G.; Zheng, X.; Dai, M.; Peng, C. Metal-organic framework membranes: Recent development in the synthesis strategies and their application in oil-water separation. Chem. Eng. J. 2021, 405, 127004. [Google Scholar] [CrossRef]
  118. Jakšić, Z.; Jakšić, O. Biomimetic nanomembranes: An overview. Biomimetics 2020, 5, 24. [Google Scholar] [CrossRef] [PubMed]
  119. Nosheen, S. Nanomembrane applications in environmental engineering. In Nanotechnology Applications in Environmental Engineering; IGI Global: Hershey, PA, USA, 2019; pp. 103–120. [Google Scholar]
  120. Jani, A.M.; Anglin, E.J.; McInnes, S.J.P.; Losic, D.; Shapter, J.G.; Voelcker, N.H. Nanoporous anodic aluminium oxide membranes with layered surface chemistry. Chem. Commun. 2009, 21, 3062–3064. [Google Scholar] [CrossRef]
  121. Mei, Y.; Thurmer, D.J.; Deneke, C.; Kiravittaya, S.; Chen, Y.-F.; Dadgar, A.; Bertram, F.; Bastek, B.; Krost, A.; Christen, J.; et al. Fabrication, self-assembly, and properties of ultrathin AlN/GaN porous crystalline nanomembranes: Tubes, spirals, and curved sheets. Acs Nano 2009, 3, 1663–1668. [Google Scholar] [CrossRef]
  122. De Wolf, I.; Senez, V.; Balboni, R.; Armigliato, A.; Frabboni, S.; Cedola, A.; Lagomarsino, S. Techniques for mechanical strain analysis in sub-micrometer structures: TEM/CBED, micro-Raman spectroscopy, X-ray micro-diffraction and modeling. Microelectron. Eng. 2003, 4, 425–435. [Google Scholar] [CrossRef]
  123. Dhanabal, R.; Naveena, D.; Velmathi, S.; Bose, A.C. Reduced graphene oxide supported molybdenum oxide hybrid nanocomposites: High performance electrode material for supercapacitor and photocatalytic applications. J. Nanosci. Nanotechnol. 2020, 20, 4035–4046. [Google Scholar] [CrossRef]
  124. Agboola, O.; Sadiku, E.R.; Mokrani, T. Nanomembrane materials based on polymer blends. In Design and Applications of Nanostructured Polymer Blends and Nanocomposite Systems; William Andrew Publishing: Norwich, NY, USA, 2016; pp. 101–123. [Google Scholar]
  125. Won, Y.-J.; Lee, J.; Choi, D.-C.; Chae, H.R.; Kim, I.; Lee, C.-H.; Kim, I.-C. Preparation and application of patterned membranes for wastewater treatment. Environ. Sci. Technol. 2012, 46, 11021–11027. [Google Scholar] [CrossRef]
  126. Pérez-Madrigal, M.M.; Armelin, E.; Puiggalí, J.; Alemán, C. Insulating and semiconducting polymeric free-standing nanomembranes with biomedical applications. J. Mater. Chem. B 2015, 3, 5904–5932. [Google Scholar] [CrossRef] [Green Version]
  127. Lu, P.; Chen, W.; Zhu, M.; Murray, S. Embedding lauric acid into polystyrene nanofibers to make high-capacity membranes for efficient thermal energy storage. ACS Sustain. Chem. Eng. 2017, 5, 7249–7259. [Google Scholar] [CrossRef]
  128. Saleh, T.A. Protocols for synthesis of nanomaterials, polymers, and green materials as adsorbents for water treatment technologies. Environ. Technol. Innov. 2021, 24, 101821. [Google Scholar] [CrossRef]
  129. Gracheva, I.E.; Moshnikov, V.; Maraeva, E.; Karpova, S.S.; Alexsandrova, O.A.; Alekseyev, N.I.; Kuznetsov, V.V.; Olchowik, G.; Semenov, K.; Startseva, A.V.; et al. Nanostructured materials obtained under conditions of hierarchical self-assembly and modified by derivative forms of fullerenes. J. Non-Cryst. Solids 2012, 358, 433–439. [Google Scholar] [CrossRef]
  130. Chen, J.; Yao, B.; Li, C.; Shi, G. An improved Hummers method for eco-friendly synthesis of graphene oxide. Carbon 2013, 64, 225–229. [Google Scholar] [CrossRef]
  131. Yahaya, N.Z.S.; Paiman, S.H.; Abdullah, N.; Mahpoz, N.M.; Raffi, A.A.; Rahman, M.A.; Abas, K.H.; Aziz, A.A.; Othman, M.H.D.; Jaafar, J. Synthesis and characterizations of MIL-140B-Al2O3/YSZ ceramic membrane using solvothermal method for seawater desalination. J. Aust. Ceram. Soc. 2020, 56, 291–300. [Google Scholar] [CrossRef]
  132. Maina, J.W.; Gonzalo, C.P.; Merenda, A.; Kong, L.; Schütz, J.A.; Dumée, L.F. The growth of high density network of MOF nano-crystals across macroporous metal substrates–Solvothermal synthesis versus rapid thermal deposition. Appl. Surf. Sci. 2018, 427, 401–408. [Google Scholar] [CrossRef]
  133. Ahn, E.; Gaiji, H.; Kim, T.; Abderrabba, M.; Lee, H.-W.; Kim, B.-S. Graphene oxide nanosheet as a two-dimensional polyelectrolyte: pH-responsive behavior of a multilayered nanomembrane. J. Membr. Sci. 2019, 585, 191–198. [Google Scholar] [CrossRef]
  134. Kohn, J. Small-scale membrane filter electrophoresis and immuno-electrophoresis. Clin. Chim. Acta 1958, 3, 450–454. [Google Scholar] [CrossRef]
  135. Lagashetty, A.; Havanoor, V.; Basavaraja, S.; Balaji, S.D.; Venkataraman, A. Microwave-assisted route for synthesis of nanosized metal oxides. Sci. Technol. Adv. Mater. 2007, 8, 484. [Google Scholar] [CrossRef]
  136. Dahiya, M.S.; Tomer, V.K.; Duhan, S. Metal–ferrite nanocomposites for targeted drug delivery. In Applications of Nanocomposite Materials in Drug Delivery; Woodhead Publishing: Sawston, UK, 2018; pp. 737–760. [Google Scholar]
  137. Al-Hamadani, Y.A.; Jung, C.; Im, J.-K.; Boateng, L.K.; Flora, J.R.; Jang, M.; Heo, J.; Park, C.M.; Yoon, Y. Sonocatalytic degradation coupled with single-walled carbon nanotubes for removal of ibuprofen and sulfamethoxazole. Chem. Eng. Sci. 2017, 162, 300–308. [Google Scholar] [CrossRef]
  138. Abdulkhair, B.; Salih, M.; Modwi, A.; Adam, F.; Elamin, N.; Seydou, M.; Rahali, S. Adsorption behavior of barium ions onto ZnO surfaces: Experiments associated with DFT calculations. J. Mol. Struct. 2021, 1223, 128991. [Google Scholar] [CrossRef]
  139. Hu, Y.; Zhao, C.; Yin, L.; Wen, T.; Yang, Y.; Ai, Y.; Wang, X. Combining batch technique with theoretical calculation studies to analyze the highly efficient enrichment of U(VI) and Eu(III) on magnetic MnFe2O4 nanocubes. Chem. Eng. J. 2018, 349, 347–357. [Google Scholar] [CrossRef]
  140. Tian, Y.L.; Hua, H.L.; Yue, W.W.; Chen, M.N.; Hu, G.C.; Ren, J.F.; Yuan, X.B. Adsorption properties of chloroform molecule on graphene: Experimental and first-principles calculations. Mod. Phys. Lett. B 2017, 31, 1750335. [Google Scholar] [CrossRef]
  141. Kang, W.; Cui, Y.; Yang, Y.; Zhao, Z.; Wang, X.; Liu, X. An acid induction strategy to construct an ultralight and durable amino-functionalized graphene oxide aerogel for enhanced quinoline pollutants extraction from coking wastewater. Chem. Eng. J. 2021, 412, 128686. [Google Scholar] [CrossRef]
  142. Hao, J.; Zhang, Q.; Liu, Y.; Chen, P.; Zheng, X.; Zhuang, X.; Fu, D.; Liu, H.; Liu, G.; Lv, W. A novel nitrogen-containing covalent organic framework adsorbent for the efficient removal of bisphenol A from aqueous solution. J. Taiwan Inst. Chem. Eng. 2020, 113, 204–213. [Google Scholar] [CrossRef]
  143. Khajouei, M.; Najafi, M.; Jafari, S.A. Development of ultrafiltration membrane via in-situ grafting of nano-GO/PSF with anti-biofouling properties. Chem. Eng. Res. Des. 2019, 142, 34–43. [Google Scholar] [CrossRef]
  144. Han, J.; Lee, S.; Choi, K.; Kim, J.; Ha, D.; Lee, C.-G.; An, B.; Lee, S.-H.; Mizuseki, H.; Choi, J.-W.; et al. Effect of nitrogen doping on titanium carbonitride-derived adsorbents used for arsenic removal. J. Hazard. Mater. 2016, 302, 375–385. [Google Scholar] [CrossRef]
  145. Reynosa-Martínez, A.C.G.; Tovar, N.; Gallegos, W.R.; Rodríguez-Meléndez, H.; Torres-Cadena, R.; Mondragón-Solórzano, G.; Barroso-Flores, J.; Alvarez-Lemus, M.A.; Montalvo, V.G.; López-Honorato, E. Effect of the degree of oxidation of graphene oxide on As(III) adsorption. J. Hazard. Mater. 2020, 384, 121440. [Google Scholar] [CrossRef] [Green Version]
  146. Zhang, L.; Hu, X.; Zhou, Q. Sunlight-assisted tailoring of surface nanostructures on single-layer graphene nanosheets for highly efficient cation capture and high-flux desalination. Carbon 2020, 161, 674–684. [Google Scholar] [CrossRef]
  147. Lawal, I.A.; Lawal, M.M.; Akpotu, S.O.; Azeez, M.A.; Ndungu, P.; Moodley, B. Theoretical and experimental adsorption studies of sulfamethoxazole and ketoprofen on synthesized ionic liquids modified CNTs. Ecotoxicol. Environ. Saf. 2018, 161, 542–552. [Google Scholar] [CrossRef]
  148. Chen, Y.; Zhang, W.; Yang, S.; Hobiny, A.; Alsaedi, A.; Wang, X. Understanding the adsorption mechanism of Ni(II) on graphene oxides by batch experiments and density functional theory studies. Sci. China Chem. 2016, 59, 412–419. [Google Scholar] [CrossRef]
  149. Zhang, J.; Lu, X.; Shi, C.; Yan, B.; Gong, L.; Chen, J.; Xiang, L.; Xu, H.; Liu, Q.; Zeng, H. Unraveling the molecular interaction mechanism between graphene oxide and aromatic organic compounds with implications on wastewater treatment. Chem. Eng. J. 2019, 358, 842–849. [Google Scholar] [CrossRef]
  150. Liu, Y.; Fang, C.; Zhong, W.; Wei, Q.; Wang, Y.; Dai, Y.; Wang, Y.; Zhang, Z.; Liu, Y. Effective adsorption of uranyl ions with different MoS2-exposed surfaces in aqueous solution. Surf. Interfaces 2020, 18, 100409. [Google Scholar] [CrossRef]
  151. Zhang, L.; He, X.; Zhou, Q.; Hu, X. Fabrication of 1T-MoS2 nanosheets and the high-efficiency removal of toxic metals in aquatic systems: Performance and mechanisms. Chem. Eng. J. 2020, 386, 123996. [Google Scholar] [CrossRef]
  152. Younes, H.A.; Khaled, R.; Mahmoud, H.M.; Nassar, H.F.; Abdelrahman, M.M.; El-Ela, F.I.A.; Taha, M. Computational and experimental studies on the efficient removal of diclofenac from water using ZnFe-layered double hydroxide as an environmentally benign absorbent. J. Taiwan Inst. Chem. Eng. 2019, 102, 297–311. [Google Scholar] [CrossRef]
  153. Huang, L.; Yang, Z.; Lei, D.; Liu, F.; He, Y.; Wang, H.; Luo, J. Experimental and modeling studies for adsorbing different species of fluoride using lanthanum-aluminum perovskite. Chemosphere 2021, 263, 128089. [Google Scholar] [CrossRef]
  154. Shi, Z.; Zhang, Y.; Liu, T.; Cao, W.; Zhang, L.; Li, M.; Chen, Z. Synthesis of BiOBr/Ag3PO4 heterojunctions on carbon-fiber cloth as filter-membrane-shaped photocatalyst for treating the flowing antibiotic wastewater. J. Colloid Interface Sci. 2020, 575, 183–193. [Google Scholar] [CrossRef]
  155. Neugebauer, J.; Hickel, T. Density functional theory in materials science. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2013, 3, 438–448. [Google Scholar] [CrossRef] [Green Version]
  156. Sharma, S.; Kumar, P.; Chandra, R. Applications of BIOVIA materials studio, LAMMPS, and GROMACS in various fields of science and engineering. In Molecular Dynamics Simulation of Nanocomposites Using BIOVIA Materials Studio, Lammps and Gromacs; Elsevier: Amsterdam, The Netherlands, 2019; pp. 329–341. [Google Scholar]
  157. Robinson, S. Simulation: The Practice of Model Development and Use; Wiley: Chichester, UK, 2004; Volume 50. [Google Scholar]
  158. Krogel, J.T. Nexus: A modular workflow management system for quantum simulation codes. Comput. Phys. Commun. 2016, 198, 154–168. [Google Scholar] [CrossRef] [Green Version]
  159. Fan, Z.; Zhang, W.-B.; Tang, B.-Y. Electronic structures and elastic properties of monolayer and bilayer transition metal dichalcogenides MX2 (M = Mo, W; X = O, S, Se, Te): A comparative first-principles study. Chin. Phys. B 2015, 24, 097103. [Google Scholar]
  160. Lehner, A.J.; Fabini, D.H.; Evans, H.A.; Hébert, C.-A.; Smock, S.R.; Hu, J.; Wang, H.; Zwanziger, J.W.; Chabinyc, M.L.; Seshadri, R. Crystal and electronic structures of complex bismuth iodides A3Bi2I9 (A = K, Rb, Cs) related to perovskite: Aiding the rational design of photovoltaics. Chem. Mater. 2015, 27, 7137–7148. [Google Scholar] [CrossRef] [Green Version]
  161. Lindsay, L.; Li, W.; Carrete, J.; Mingo, N.; Broido, D.A.; Reinecke, T.L. Phonon thermal transport in strained and unstrained graphene from first principles. Phys. Rev. B 2014, 89, 155426. [Google Scholar] [CrossRef]
  162. Repetsky, S.P.; Vyshyvana, I.G.; Kruchinin, S.P.; Melnyk, R.M.; Polishchuk, A.P. The energy spectrum and the electrical conductivity of graphene with substitution impurity. arXiv 2020, arXiv:2003.02084v1. [Google Scholar] [CrossRef]
  163. Zhang, J.; Li, T.; Li, X.; Liu, Y.; Li, N.; Wang, Y.; Li, X. A key role of inner-cation-π interaction in adsorption of Pb(II) on carbon nanotubes: Experimental and DFT studies. J. Hazard. Mater. 2021, 412, 125187. [Google Scholar] [CrossRef] [PubMed]
  164. Yang, D.; Tian, D.; Cheng, C.; Liu, Y.; Zhao, Z.; Liu, Y.; Bao, Y.; Xue, C. Carbon nanotube arrays hybrid membrane with excellent separation performance and conductivity. J. Membr. Sci. 2021, 620, 118874. [Google Scholar] [CrossRef]
  165. Ren, X.; Feng, J.; Si, P.; Zhang, L.; Lou, J.; Ci, L. Enhanced heterogeneous activation of peroxydisulfate by S, N co-doped graphene via controlling S, N functionalization for the catalytic decolorization of dyes in water. Chemosphere 2018, 210, 120–128. [Google Scholar] [CrossRef] [PubMed]
  166. Wang, Y.; Zhao, W.; Qi, Z.; Zhang, L.; Peng, Y. Phosphate removal by ZIF-8@ MWCNT hybrids in presence of effluent organic matter: Adsorbent structure, wastewater quality, and DFT analysis. Sci. Total Environ. 2020, 745, 141054. [Google Scholar] [CrossRef] [PubMed]
  167. Yu, S.; Liu, Y.; Ai, Y.; Wang, X.; Zhang, R.; Chen, Z.; Chen, Z.; Zhao, G.; Wang, X. Rational design of carbonaceous nanofiber/Ni-Al layered double hydroxide nanocomposites for high-efficiency removal of heavy metals from aqueous solutions. Environ. Pollut. 2018, 242, 1–11. [Google Scholar] [CrossRef]
  168. Joseph, L.; Boateng, L.K.; Flora, J.R.; Park, Y.-G.; Son, A.; Badawy, M.; Yoon, Y. Removal of bisphenol A and 17α-ethinyl estradiol by combined coagulation and adsorption using carbon nanomaterials and powdered activated carbon. Sep. Purif. Technol. 2013, 107, 37–47. [Google Scholar] [CrossRef]
  169. do Céu Teixeira, M.; Santini, A.; Souto, E.B. Delivery of antimicrobials by chitosan-composed therapeutic nanostructures. In Nanostructures for Antimicrobial Therapy; Elsevier: Amsterdam, The Netherlands, 2017; pp. 203–222. [Google Scholar]
  170. Zhang, S.; Sun, G.; He, Y.; Fu, R.; Gu, Y.; Chen, S. Preparation, characterization, and electrochromic properties of nanocellulose-based polyaniline nanocomposite films. ACS Appl. Mater. Interfaces 2017, 9, 16426–16434. [Google Scholar] [CrossRef]
  171. Malhotra, B.D.; Ali, A. Nanomaterials in biosensors: Fundamentals and applications. In Nanomaterials for Biosensors; Elsevier: Amsterdam, The Netherlands, 2018; pp. 1–74. [Google Scholar]
  172. Chen, Y.; Lai, Z.; Zhang, X.; Fan, Z.; He, Q.; Tan, C.; Zhang, H. Phase engineering of nanomaterials. Nat. Rev. Chem. 2020, 4, 243–256. [Google Scholar] [CrossRef]
  173. Chkirida, S.; Zari, N.; El Kacem Qaiss, A.; Bouhfid, R. Nanocomposite materials based on TiO2/clay for wastewater treatment. In Advanced Research in Nanosciences for Water Technology; Springer: Cham, Switzerland, 2019; pp. 363–380. [Google Scholar]
  174. Wang, J.; Li, Y.; Zheng, D.; Mikulčić, H.; Vujanović, M.; Sundén, B. Preparation and thermophysical property analysis of nanocomposite phase change materials for energy storage. Renew. Sustain. Energy Rev. 2021, 151, 111541. [Google Scholar] [CrossRef]
  175. Jacob, J.; Haponiuk, J.T.; Thomas, S.; Gopi, S. Biopolymer based nanomaterials in drug delivery systems: A review. Mater. Today Chem. 2018, 9, 43–55. [Google Scholar] [CrossRef]
  176. Hasnain, M.S.; Nayak, A.K. Nanocomposites for improved orthopedic and bone tissue engineering applications. In Applications of Nanocomposite Materials in Orthopedics; Woodhead Publishing: Sawston, UK, 2019; pp. 145–177. [Google Scholar]
  177. Singh, S.; Kumar, V.; Romero, R.; Sharma, K.; Singh, J. Applications of nanoparticles in wastewater treatment. In Nanobiotechnology in Bioformulations; Springer: Cham, Switzerland, 2019; pp. 395–418. [Google Scholar]
  178. Karn, B.; Kuiken, T.; Otto, M. Nanotechnology and in situ remediation: A review of the benefits and potential risks. Environ. Health Perspect. 2009, 117, 1813–1831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Zhang, S.; Qian, L.; Zhao, Q.; Wang, Z.; Lin, D.; Liu, W.; Chen, Y.; Zhang, J. Carbon nanotube: Controlled synthesis determines its future. Sci. China Mater. 2020, 63, 16–34. [Google Scholar] [CrossRef] [Green Version]
  180. Rane, A.V.; Kanny, K.; Abitha, V.K.; Thomas, S. Methods for synthesis of nanoparticles and fabrication of nanocomposites. In Synthesis of Inorganic Nanomaterials; Woodhead Publishing: Sawston, UK, 2018; pp. 121–139. [Google Scholar]
  181. Manawi, Y.M.; Samara, A.; Al-Ansari, T.; Atieh, M.A. A review of carbon nanomaterials’ synthesis via the chemical vapor deposition (CVD) method. Materials 2018, 11, 822. [Google Scholar] [CrossRef] [Green Version]
  182. Zhong, Q.; Cao, M.; Hu, H.; Yang, D.; Chen, M.; Li, P.; Wu, L.; Zhang, Q. One-pot synthesis of highly stable CsPbBr3@ SiO2 core–shell nanoparticles. Acs Nano 2018, 12, 8579–8587. [Google Scholar] [CrossRef]
  183. Gan, Y.X.; Jayatissa, A.H.; Yu, Z.; Chen, X.; Li, M. Hydrothermal synthesis of nanomaterials. J. Nanomater. 2020, 2020, 8917013. [Google Scholar] [CrossRef]
  184. Kaflé, B.P. Chemical Analysis and Material Characterization by Spectrophotometry; Elsevier: Amsterdam, The Netherlands, 2019. [Google Scholar]
  185. Adeniji, Q.A.; Odunaike, R.K.; Bamijoko, B.A.; Adeleke, A.T.; Dahunsi, K.T. Synthesis and characterization of Zinc Tin sulphide (ZTS) thin films via chemical bath deposition route. J. Appl. Sci. Inf. Comput. 2020, 1, 22–31. [Google Scholar]
  186. Yang, G.; Park, S.-J. Conventional and microwave hydrothermal synthesis and application of functional materials: A review. Materials 2019, 12, 1177. [Google Scholar] [CrossRef] [Green Version]
  187. Li, M.; Magdassi, S.; Gao, Y.; Long, Y. Hydrothermal synthesis of VO2 polymorphs: Advantages, challenges and prospects for the application of energy efficient smart windows. Small 2017, 13, 1701147. [Google Scholar] [CrossRef]
  188. Wu, Y.Z.; Chen, M.; Yan, X.H.; Ren, J.; Dai, Y.; Wang, J.J.; Pan, J.M.; Wang, Y.P.; Cheng, X.N. Hydrothermal synthesis of Fe3O4 nanorods/graphitic C3N4 composite with enhanced supercapacitive performance. Mater. Lett. 2017, 198, 114–117. [Google Scholar] [CrossRef]
  189. Modan, E.M.; Plăiașu, A.G. Advantages and disadvantages of chemical methods in the elaboration of nanomaterials. Ann. “Dunarea de Jos” Univ. Galati. Fascicle IX Met. Mater. Sci. 2020, 43, 53–60. [Google Scholar] [CrossRef]
  190. Das, S.; Dhara, S. Chemical solution synthesis for materials design and thin film device applications. In Chemical Solution Synthesis for Materials Design and Thin Film Device Applications; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  191. Yang, G.; Xie, J.; Deng, Y.; Bian, Y.; Hong, F. Hydrothermal synthesis of bacterial cellulose/AgNPs composite: A “green” route for antibacterial application. Carbohydr. Polym. 2012, 87, 2482–2487. [Google Scholar] [CrossRef]
  192. Byrappa, K.; Yoshimura, M. Handbook of Hydrothermal Technology; William Andrew: Norwich, NY, USA, 2012. [Google Scholar]
  193. Wang, X.L.; Li, J.; Liu, W.M. Synthesizing pyridinic-N dominate-doped graphene/BiVO4 nanocomposite as a superior photocatalyst for degradation under visible-irradiation. Opt. Mater. 2021, 114, 110922. [Google Scholar] [CrossRef]
  194. Yang, C.; Wang, X.; Ji, Y.; Ma, T.; Zhang, F.; Wang, Y.; Ci, M.; Chen, D.; Jiang, A.; Wang, W. Photocatalytic degradation of methylene blue with ZnO@ C nanocomposites: Kinetics, mechanism, and the inhibition effect on monoamine oxidase A and B. NanoImpact 2019, 15, 100174. [Google Scholar] [CrossRef]
  195. Li, P.; Guo, M.; Wang, Q.; Li, Z.; Wang, C.; Chen, N.; Wang, C.-C.; Wan, C.; Chen, S. Controllable synthesis of cerium zirconium oxide nanocomposites and their application for photocatalytic degradation of sulfonamides. Appl. Catal. B Environ. 2019, 259, 118107. [Google Scholar] [CrossRef]
  196. Nie, Q.; Xie, Y.; Ma, J.; Wang, J.; Zhang, G. High piezo–catalytic activity of ZnO/Al2O3 nanosheets utilizing ultrasonic energy for wastewater treatment. J. Clean. Prod. 2020, 242, 118532. [Google Scholar] [CrossRef]
  197. Wang, X.; Wang, L.L.; Guo, D.; Ma, L.L.; Zhu, B.L.; Wang, P.; Wang, G.C.; Zhang, S.M.; Huang, W.P. Fabrication and photocatalytic performance of C,N,F-tridoped TiO2 nanotubes. Catal. Today 2019, 327, 182–189. [Google Scholar] [CrossRef]
  198. Ke, T.; Shen, S.; Rajavel, K.; Yang, K.; Lin, D. In situ growth of TiO2 nanoparticles on nitrogen-doped Ti3C2 with isopropyl amine toward enhanced photocatalytic activity. J. Hazard. Mater. 2021, 402, 124066. [Google Scholar] [CrossRef]
  199. Zhao, Y.; Chi, Y.; Tian, C.; Liu, Y.; Li, H.; Wang, A. Recycling of titanium-coagulated algae-rich sludge for enhanced photocatalytic oxidation of phenolic contaminants through oxygen vacancy. Water Res. 2020, 177, 115789. [Google Scholar] [CrossRef]
  200. Cheng, K.; Cai, Z.; Fu, J.; Sun, X.; Sun, W.; Chen, L.; Zhang, D.; Liu, W. Synergistic adsorption of Cu(II) and photocatalytic degradation of phenanthrene by a jaboticaba-like TiO2/titanate nanotube composite: An experimental and theoretical study. Chem. Eng. J. 2019, 358, 1155–1165. [Google Scholar] [CrossRef]
  201. Wang, W.; Niu, Q.; Zeng, G.; Zhang, C.; Huang, D.; Shao, B.; Zhou, C.; Yang, Y.; Liu, Y.; Guo, H.; et al. 1D porous tubular g-C3N4 capture black phosphorus quantum dots as 1D/0D metal-free photocatalysts for oxytetracycline hydrochloride degradation and hexavalent chromium reduction. Appl. Catal. B Environ. 2020, 273, 119051. [Google Scholar] [CrossRef]
  202. Saleh, R.; Zaki, A.H.; El-Ela, F.I.A.; Farghali, A.A.; Taha, M.; Mahmoud, R. Consecutive removal of heavy metals and dyes by a fascinating method using titanate nanotubes. J. Environ. Chem. Eng. 2021, 9, 104726. [Google Scholar] [CrossRef]
  203. Maji, T.K.; Hasan, N.; Ghosh, S.; Wulferding, D.; Bhattacharya, C.; Lemmens, P.; Karmakar, D.; Pal, S.K. Development of a magnetic nanohybrid for multifunctional application: From immobile photocatalysis to efficient photoelectrochemical water splitting: A combined experimental and computational study. J. Photochem. Photobiol. A Chem. 2020, 397, 112575. [Google Scholar] [CrossRef]
  204. Kitchamsetti, N.; Ramteke, M.S.; Rondiya, S.R.; Mulani, S.R.; Patil, M.S.; Cross, R.W.; Dzade, N.Y.; Devan, R.S. DFT and experimental investigations on the photocatalytic activities of NiO nanobelts for removal of organic pollutants. J. Alloys Compd. 2021, 855, 157337. [Google Scholar] [CrossRef]
  205. Duan, Y.; Deng, L.; Shi, Z.; Liu, X.; Zeng, H.; Zhang, H.; Crittenden, J. Efficient sulfadiazine degradation via in-situ epitaxial grow of Graphitic Carbon Nitride (g-C3N4) on carbon dots heterostructures under visible light irradiation: Synthesis, mechanisms and toxicity evaluation. J. Colloid Interface Sci. 2020, 561, 696–707. [Google Scholar] [CrossRef]
  206. Cai, Z.; Song, Y.; Jin, X.; Wang, C.-C.; Ji, H.; Liu, W.; Sun, X. Highly efficient AgBr/h-MoO3 with charge separation tuning for photocatalytic degradation of trimethoprim: Mechanism insight and toxicity assessment. Sci. Total Environ. 2021, 781, 146754. [Google Scholar] [CrossRef]
  207. Zhang, X.; Zhao, R.; Zhang, N.; Su, Y.; Liu, Z.; Gao, R.; Du, C. Insight to unprecedented catalytic activity of double-nitrogen defective metal-free catalyst: Key role of coal gangue. Appl. Catal. B Environ. 2020, 263, 118316. [Google Scholar] [CrossRef]
  208. Regmi, C.; Kshetri, Y.K.; Kim, T.-H.; Dhakal, D.; Lee, S.W. Mechanistic understanding of enhanced photocatalytic activity of N-doped BiVO4 towards degradation of ibuprofen: An experimental and theoretical approach. Mol. Catal. 2019, 470, 8–18. [Google Scholar] [CrossRef]
  209. Wei, Z.; Hu, J.; Zhu, K.; Wei, W.; Ma, X.; Zhu, Y. Self-assembled polymer phenylethnylcopper nanowires for photoelectrochemical and photocatalytic performance under visible light. Appl. Catal. B Environ. 2018, 226, 616–623. [Google Scholar] [CrossRef]
  210. Huang, L.-Z.; Wei, X.; Gao, E.; Zhang, C.; Hu, X.-M.; Chen, Y.; Liu, Z.; Finck, N.; Lützenkirchen, J.; Dionysiou, D. Single Fe atoms confined in two-dimensional MoS2 for sulfite activation: A biomimetic approach towards efficient radical generation. Appl. Catal. B Environ. 2020, 268, 118459. [Google Scholar] [CrossRef]
  211. Li, Z.; Chen, Q.; Lin, Q.; Chen, Y.; Liao, X.; Yu, H.; Yu, C. Three-dimensional P-doped porous g-C3N4 nanosheets as an efficient metal-free photocatalyst for visible-light photocatalytic degradation of Rhodamine B model pollutant. J. Taiwan Inst. Chem. Eng. 2020, 114, 249–262. [Google Scholar] [CrossRef]
  212. Deng, Y.; Feng, C.; Tang, L.; Zhou, Y.; Chen, Z.; Feng, H.; Wang, J.; Yu, J.; Liu, Y. Ultrathin low dimensional heterostructure composites with superior photocatalytic activity: Insight into the multichannel charge transfer mechanism. Chem. Eng. J. 2020, 393, 124718. [Google Scholar] [CrossRef]
  213. Wang, P.; Tian, Y.; Wang, H.; Zhang, J.; Kong, L.; Zuo, W.; Li, D.; Yin, L. Strong adsorption of tetracycline on octahedral Cu2O nanocrystals exposed with {111} facets: Adsorption behavior and mechanism insight. Appl. Surf. Sci. 2021, 542, 148545. [Google Scholar] [CrossRef]
  214. Zhang, C.; Liu, Y.; Li, X.; Chen, H.; Wen, T.; Jiang, Z.; Ai, Y.; Sun, Y.; Hayat, T.; Wang, X. Highly uranium elimination by crab shells-derived porous graphitic carbon nitride: Batch, EXAFS and theoretical calculations. Chem. Eng. J. 2018, 346, 406–415. [Google Scholar] [CrossRef]
  215. Gu, P.; Zhao, C.; Wen, T.; Ai, Y.; Zhang, S.; Chen, W.; Wang, J.; Hu, B.; Wang, X. Highly U(VI) immobilization on polyvinyl pyrrolidine intercalated molybdenum disulfide: Experimental and computational studies. Chem. Eng. J. 2019, 359, 1563–1572. [Google Scholar] [CrossRef]
  216. Wang, J.; Zhang, R.; Lu, Z.; Ai, Y. Experimental and theoretical studies of spherical β-cyclodextrin modified titanium dioxide composites for uranium removal. Ecol. Eng. 2020, 149, 105835. [Google Scholar] [CrossRef]
  217. Tokalıoğlu, Ş.; Yavuz, E.; Demir, S.; Patat, Ş. Zirconium-based highly porous metal-organic framework (MOF-545) as an efficient adsorbent for vortex assisted-solid phase extraction of lead from cereal, beverage and water samples. Food Chem. 2017, 237, 707–715. [Google Scholar] [CrossRef]
  218. Senapati, S.; Maiti, P. Emerging bio-applications of two-dimensional nanoheterostructure materials. In 2D Nanoscale Heterostructured Materials; Elsevier: Amsterdam, The Netherlands, 2020; pp. 243–255. [Google Scholar]
  219. Carlsson, J.-O.; Martin, P.M. Chemical vapor deposition. In Handbook of Deposition Technologies for Films and Coatings; William Andrew Publishing: Norwich, NY, USA, 2010; pp. 314–363. [Google Scholar]
  220. Badv, M.; Jaffer, I.H.; Weitz, J.I.; Didar, T.F. An omniphobic lubricant-infused coating produced by chemical vapor deposition of hydrophobic organosilanes attenuates clotting on catheter surfaces. Sci. Rep. 2017, 7, 11639. [Google Scholar] [CrossRef]
  221. Zhang, Y.; Zhang, L.; Zhou, C. Review of chemical vapor deposition of graphene and related applications. Acc. Chem. Res. 2013, 46, 2329–2339. [Google Scholar] [CrossRef]
  222. Pourmand, S.; Abdouss, M.; Rashidi, A. Fabrication of nanoporous graphene by chemical vapor deposition (CVD) and its application in oil spill removal as a recyclable nanosorbent. J. Ind. Eng. Chem. 2015, 22, 8–18. [Google Scholar] [CrossRef]
  223. Creighton, J.R.; Ho, P. Introduction to chemical vapor deposition (CVD). Chem. Vap. Depos. 2001, 2, 1–22. [Google Scholar]
  224. Haubner, R. The history of hard CVD coatings for tool applications at the University of Technology Vienna. Int. J. Refract. Met. Hard Mater. 2013, 41, 22–34. [Google Scholar] [CrossRef]
  225. Tietjen, J.J. Chemical vapor deposition of electronic materials. Annu. Rev. Mater. Sci. 1973, 3, 317–326. [Google Scholar] [CrossRef]
  226. Gao, L.; Ren, W.; Zhao, J.; Ma, L.-P.; Chen, Z.; Cheng, H.-M. Efficient growth of high-quality graphene films on Cu foils by ambient pressure chemical vapor deposition. Appl. Phys. Lett. 2010, 97, 183109. [Google Scholar]
  227. Park, H.J.; Meyer, J.; Roth, S.; Skákalová, V. Growth and properties of few-layer graphene prepared by chemical vapor deposition. Carbon 2010, 48, 1088–1094. [Google Scholar] [CrossRef] [Green Version]
  228. Vasudev, M.C.; Anderson, K.D.; Bunning, T.J.; Tsukruk, V.V.; Naik, R.R. Exploration of plasma-enhanced chemical vapor deposition as a method for thin-film fabrication with biological applications. ACS Appl. Mater. Interfaces 2013, 5, 3983–3994. [Google Scholar] [CrossRef]
  229. Cai, Z.; Liu, B.; Zou, X.; Cheng, H.-M. Chemical vapor deposition growth and applications of two-dimensional materials and their heterostructures. Chem. Rev. 2018, 118, 6091–6133. [Google Scholar] [CrossRef]
  230. Cong, C.; Shang, J.; Wu, X.; Cao, B.; Peimyoo, N.; Qiu, C.; Sun, L.; Yu, T. Synthesis and optical properties of large-area single-crystalline 2D semiconductor WS2 monolayer from chemical vapor deposition. Adv. Opt. Mater. 2014, 2, 131–136. [Google Scholar] [CrossRef]
  231. Liu, B.; Song, W.; Wu, H.; Liu, Z.; Teng, Y.; Sun, Y.; Xu, Y.; Zheng, H. Degradation of norfloxacin with peroxymonosulfate activated by nanoconfinement Co3O4@CNT nanocomposite. Chem. Eng. J. 2020, 398, 125498. [Google Scholar] [CrossRef]
  232. Dastgerdi, Z.H.; Meshkat, S.S.; Esrafili, M.D. Enhanced adsorptive removal of Indigo carmine dye performance by functionalized carbon nanotubes based adsorbents from aqueous solution: Equilibrium, kinetic, and DFT study. J. Nanostruct. Chem. 2019, 9, 323–334. [Google Scholar] [CrossRef] [Green Version]
  233. Shah, K.A.; Tali, B.A. Tali. Synthesis of carbon nanotubes by catalytic chemical vapour deposition: A review on carbon sources, catalysts and substrates. Mater. Sci. Semicond. Process. 2016, 41, 67–82. [Google Scholar] [CrossRef]
  234. Zhang, Z.; Li, Q.; Wang, F.; Li, R.; Yu, X.; Kang, L.; Zhao, J.; Li, A. One-pot biosynthesis of 1, 6-hexanediol from cyclohexane by de novo designed cascade biocatalysis. Green Chem. 2020, 22, 7476–7483. [Google Scholar] [CrossRef]
  235. Li, H.; Tian, J.; Zhu, Z.; Cui, F.; Zhu, Y.-A.; Duan, X.; Wang, S. Magnetic nitrogen-doped nanocarbons for enhanced metal-free catalytic oxidation: Integrated experimental and theoretical investigations for mechanism and application. Chem. Eng. J. 2018, 354, 507–516. [Google Scholar] [CrossRef]
  236. Du, X.; Fu, W.; Su, P.; Cai, J.; Zhou, M. Internal-micro-electrolysis-enhanced heterogeneous electro-Fenton process catalyzed by Fe/Fe3C@ PC core–shell hybrid for sulfamethazine degradation. Chem. Eng. J. 2020, 398, 125681. [Google Scholar] [CrossRef]
  237. Mingos, D.; Crabtree, M.P.; Robert, H. Comprehensive Organometallic Chemistry, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  238. Andrews, D.; Nann, T.; Lipson, R.H. Comprehensive Nanoscience and Nanotechnology; Academic Press: Cambridge, MA, USA, 2019. [Google Scholar]
  239. Sun, H.; Cao, L.; Lu, L. Magnetite/reduced graphene oxide nanocomposites: One step solvothermal synthesis and use as a novel platform for removal of dye pollutants. Nano Res. 2011, 4, 550–562. [Google Scholar] [CrossRef]
  240. Jamkhande, P.G.; Ghule, N.W.; Bamer, A.H.; Kalaskar, M.G. Metal nanoparticles synthesis: An overview on methods of preparation, advantages and disadvantages, and applications. J. Drug Deliv. Sci. Technol. 2019, 53, 101174. [Google Scholar] [CrossRef]
  241. Kim, H.J.; Choi, H.; Sharma, A.K.; Hong, W.G.; Shin, K.; Song, H.; Kim, H.Y.; Hong, Y.J. Recyclable aqueous metal adsorbent: Synthesis and Cu(II) sorption characteristics of ternary nanocomposites of Fe3O4 nanoparticles@ graphene–poly-N-phenylglycine nanofibers. J. Hazard. Mater. 2021, 401, 123283. [Google Scholar] [CrossRef]
  242. de Santiago Colín, D.M.; Martínez-Chávez, L.A.; Cuán, Á.; Elizalde-Peña, E.A.; Rivera, J.A.; Guzmán, C.; Escobar-Alarcón, L.; Esquivel, K. Sonochemical coupled synthesis of Cr-TiO2 supported on Fe3O4 structures and chemical simulation of the degradation mechanism of Malachite Green dye. J. Photochem. Photobiol. A Chem. 2018, 364, 250–261. [Google Scholar] [CrossRef]
  243. Li, S.; Wang, J.; Xia, Y.; Li, P.; Wu, Y.; Yang, K.; Song, Y.; Jiang, S.; Zhang, T.; Li, B. Boosted electron-transfer by coupling Ag and Z-scheme heterostructures in CdSe-Ag-WO3-Ag for excellent photocatalytic H2 evolution with simultaneous degradation. Chem. Eng. J. 2021, 417, 129298. [Google Scholar] [CrossRef]
  244. Gong, J.; Lee, C.-S.; Chang, Y.-Y. Novel self-assembled bimetallic structure of Bi/Fe0: The oxidative and reductive degradation of hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX). J. Hazard. Mater. 2015, 286, 107–117. [Google Scholar] [CrossRef] [PubMed]
  245. Shan, D.; Deng, S.; Zhao, T.; Yu, G.; Winglee, J.; Wiesner, M.R. Preparation of regenerable granular carbon nanotubes by a simple heating-filtration method for efficient removal of typical pharmaceuticals. Chem. Eng. J. 2016, 294, 353–361. [Google Scholar] [CrossRef] [Green Version]
  246. Wang, H.; Wang, Z.; Yue, R.; Gao, F.; Ren, R.; Wei, J.; Wang, X.; Kong, Z. Functional group-rich hyperbranched magnetic material for simultaneous efficient removal of heavy metal ions from aqueous solution. J. Hazard. Mater. 2020, 384, 121288. [Google Scholar] [CrossRef] [PubMed]
  247. Liu, Y.; Huo, Y.; Wang, X.; Yu, S.; Ai, Y.; Chen, Z.; Zhang, P.; Chen, L.; Song, G.; Alharbi, N.S.; et al. Impact of metal ions and organic ligands on uranium removal properties by zeolitic imidazolate framework materials. J. Clean. Prod. 2021, 278, 123216. [Google Scholar] [CrossRef]
  248. Yan, L.; Tu, H.; Chan, T.; Jing, C. Mechanistic study of simultaneous arsenic and fluoride removal using granular TiO2-La adsorbent. Chem. Eng. J. 2017, 313, 983–992. [Google Scholar] [CrossRef]
  249. Song, Q.; Liang, J.; Fang, Y.; Guo, Z.; Du, Z.; Zhang, L.; Liu, Z.; Huang, Y.; Lin, J.; Tang, C. Nickel(II) modified porous boron nitride: An effective adsorbent for tetracycline removal from aqueous solution. Chem. Eng. J. 2020, 394, 124985. [Google Scholar] [CrossRef]
  250. Liu, H.; Chen, P.; Yuan, X.; Zhang, Y.; Huang, H.; Wang, L.; Dong, F. Pivotal roles of artificial oxygen vacancies in enhancing photocatalytic activity and selectivity on Bi2O2CO3 nanosheets. Chin. J. Catal. 2019, 40, 620–630. [Google Scholar] [CrossRef]
  251. Xu, L.; Liu, Y.; Wang, J.; Tang, Y.; Zhang, Z. Selective adsorption of Pb2+ and Cu2+ on amino-modified attapulgite: Kinetic, thermal dynamic and DFT studies. J. Hazard. Mater. 2021, 404, 124140. [Google Scholar] [CrossRef]
  252. Shen, C.; Chen, C.; Wen, T.; Zhao, Z.; Wang, X.; Xu, A. Superior adsorption capacity of g-C3N4 for heavy metal ions from aqueous solutions. J. Colloid Interface Sci. 2015, 456, 7–14. [Google Scholar] [CrossRef]
  253. Liang, H.; Liu, R.; Hu, C.; An, X.; Zhang, X.; Liu, H.; Qu, J. Synergistic effect of dual sites on bimetal-organic frameworks for highly efficient peroxide activation. J. Hazard. Mater. 2021, 406, 124692. [Google Scholar] [CrossRef]
  254. Valadi, F.M.; Gholami, M.R. Synthesis of CuCo2O4/BiVO4 composites as promise and efficient catalysts for 4-nitrophenol reduction in water: Experimental and theoretical study. J. Environ. Chem. Eng. 2021, 9, 105408. [Google Scholar] [CrossRef]
  255. Zaher, A.; Taha, M.; Farghali, A.A.; Mahmoud, R.K. Zn/Fe LDH as a clay-like adsorbent for the removal of oxytetracycline from water: Combining experimental results and molecular simulations to understand the removal mechanism. Environ. Sci. Pollut. Res. 2020, 27, 12256–12269. [Google Scholar] [CrossRef] [PubMed]
  256. Yu, J.; Li, N.; Zhu, L.; Xu, X. Application of ZnTiO3 in quantum-dot-sensitized solar cells and numerical simulations using first-principles theory. J. Alloys Compd. 2016, 681, 88–95. [Google Scholar] [CrossRef]
  257. Zahid, M.; Abd-Elsalam, K.A. Applications of nanomaterials in water remediation: A note from the Editors. In Aquananotechnology; Elsevier: Amsterdam, The Netherlands, 2021; pp. 15–114. [Google Scholar]
  258. Zhang, Y.; Li, L.; Su, H.; Huang, W.; Dong, X. Binary metal oxide: Advanced energy storage materials in supercapacitors. J. Mater. Chem. A 2015, 3, 43–59. [Google Scholar] [CrossRef]
  259. Abdelwahab, A.; Carrasco-Marín, F.; Pérez-Cadenas, A. Binary and ternary 3D nanobundles metal oxides functionalized carbon xerogels as electrocatalysts toward oxygen reduction reaction. Materials 2020, 13, 3531. [Google Scholar] [CrossRef]
  260. Parkinson, G.S.; Diebold, U. Adsorption on metal oxide surfaces. Surf. Interface Sci. Solid-Gas Interfaces II 2016, 6, 793–817. [Google Scholar]
  261. Ani, I.J.; Akpan, U.G.; Olutoye, M.A.; Hameed, B.H. Photocatalytic degradation of pollutants in petroleum refinery wastewater by TiO2-and ZnO-based photocatalysts: Recent development. J. Clean. Prod. 2018, 205, 930–954. [Google Scholar] [CrossRef]
  262. Zhu, S.; Wang, D. Photocatalysis: Basic principles, diverse forms of implementations and emerging scientific opportunities. Adv. Energy Mater. 2017, 7, 1700841. [Google Scholar] [CrossRef] [Green Version]
  263. Coronado, J.M. A historical introduction to photocatalysis. In Design of Advanced Photocatalytic Materials for Energy and Environmental Applications; Springer: London, UK, 2013; pp. 1–4. [Google Scholar]
  264. Landau, M. Le phénomène de la photocatalyse. Compt. Rend. 1913, 156, 1894–1896. [Google Scholar]
  265. Kim, S.-M.; Vogelpohl, A. Degradation of organic pollutants by the photo-Fenton-process. Chem. Eng. Technol. Ind. Chem.-Plant Equip.-Process Eng.-Biotechnol. 1998, 21, 187–191. [Google Scholar] [CrossRef]
  266. Bonelli, B.; Freyria, F.S.; Rossetti, I.; Sethi, R. (Eds.) Nanomaterials for the Detection and Removal of Wastewater Pollutants; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
Figure 1. Number of publications in water and wastewater treatment by adsorption (experimentally validated by simulation) between 2011 and 2021 (topic keywords “adsorption”, “nanomaterials” “DFT” “simulation” and “wastewater treatment” searched from ScienceDirect and Springer), data updated 16 June 2021.
Figure 1. Number of publications in water and wastewater treatment by adsorption (experimentally validated by simulation) between 2011 and 2021 (topic keywords “adsorption”, “nanomaterials” “DFT” “simulation” and “wastewater treatment” searched from ScienceDirect and Springer), data updated 16 June 2021.
Membranes 12 00360 g001
Figure 2. Number of publications in water and wastewater treatment by adsorption (experimentally, simulation (DFT), and experimentally validated by simulation) between 2011 and 2021. Data updated 23 June 2021 using ScienceDirect database.
Figure 2. Number of publications in water and wastewater treatment by adsorption (experimentally, simulation (DFT), and experimentally validated by simulation) between 2011 and 2021. Data updated 23 June 2021 using ScienceDirect database.
Membranes 12 00360 g002
Figure 3. The most common methods used in the synthesis of nanomaterials.
Figure 3. The most common methods used in the synthesis of nanomaterials.
Membranes 12 00360 g003
Figure 4. Simulation software used to produce nanomembranes with the number of publications by each software.
Figure 4. Simulation software used to produce nanomembranes with the number of publications by each software.
Membranes 12 00360 g004
Figure 5. The percentages of the number of publications by each experimental method used in the synthesis of nanocomposite materials.
Figure 5. The percentages of the number of publications by each experimental method used in the synthesis of nanocomposite materials.
Membranes 12 00360 g005
Figure 6. Simulation software used to produce nanoparticles with the number of publications by each software.
Figure 6. Simulation software used to produce nanoparticles with the number of publications by each software.
Membranes 12 00360 g006
Figure 7. Rate of adsorption for different nanomembranes for the removal of Uranium.
Figure 7. Rate of adsorption for different nanomembranes for the removal of Uranium.
Membranes 12 00360 g007
Figure 8. Rate of adsorption for toxic metals by carbon-based membranes. The adsorption rates are difficult to distinguish in the Figure for g-C3N4 for Cd(II), Cu (II), Ni(II) and Pb(II) which equal 1.00, 2.09, 0.64, and 1.36 mmol/g, respectively.
Figure 8. Rate of adsorption for toxic metals by carbon-based membranes. The adsorption rates are difficult to distinguish in the Figure for g-C3N4 for Cd(II), Cu (II), Ni(II) and Pb(II) which equal 1.00, 2.09, 0.64, and 1.36 mmol/g, respectively.
Membranes 12 00360 g008
Figure 9. Rate of adsorption for Pb2+ by other nanocomposite membranes.
Figure 9. Rate of adsorption for Pb2+ by other nanocomposite membranes.
Membranes 12 00360 g009
Table 1. Nanomembranes synthesized by different synthesis methods.
Table 1. Nanomembranes synthesized by different synthesis methods.
MembraneMaterial TypeSynthesis MethodReference
SWCNTsCarbon nanotube (CNT)Obtained from Cheap Tubes, Inc.[137]
Graphene oxideOxidized graphene oxideObtained commercially from Sigma Aldrich[113]
ZnO surfaceZinc oxide (ZnO)Evaporation methods[138]
MnFe2O4 nanocubesManganese ferrite nanoparticles (MnFe2O4)Co-precipitation phase inversion method[139]
Graphene3D foam grapheneObtained commercially[140]
MGOAGraphene oxide (GO), ammonium (NH4+)Modified Hummers’ method[141]
PyTTA-Dva-COFNitrogen (N), covalent organic frameworkSolvent-thermal method[142]
Ultrafiltration PSF/GO membraneGraphene oxide (GO), polysulfone (PSF)Phase inversion method[143]
Nitrogen doped carbon (CNs)Carbon (C), nitrogen (N), titanium (Ti)Chlorination[144]
Graphene oxideGraphene oxideImproved Hummers’ method[145]
Single-layer graphene nanosheetsGraphiteSolution-phase exfoliation integrating bath sonication and microwave irradiation in organic solvents[146]
Carbon nanotubes (CNTs)Carbon nanotube (CNT)Nuclear magnetic resonance (1H and 13C NMR) and high resolution-mass spectrometry (HR-MS)[147]
Graphene oxideGraphene oxideModified Hummers’ method[148]
Graphene oxideGraphene oxideModified Hummers’ method[149]
MoS2 nanosheetsMolybdenum disulphideMolten salt electrolysis method[150]
MoS2 nanosheetsMolybdenum disulphideMicrowave-assisted route[151]
Zn–Fe LDHZinc (Zn), iron (Fe)Co-precipitation method[152]
Lanthanum-aluminium perovskite (La2Al4O9)Lanthanum (La), aluminium (Al)Obtained commercially from Aladdin company[153]
CF/BiOBr/Ag3PO4 clothCarbon fibre (CF), bismuth oxybromide (BiOBr), silver phosphate (Ag3PO4)Solvothermal-chemical deposition[154]
Table 2. Nanomembranes simulated by different simulation software.
Table 2. Nanomembranes simulated by different simulation software.
MembraneSoftware Simulation MethodMathematical Model Reference
(O-CNTs), (G-CNTs)Gaussian 09WDFT (B3LYP functional group)Integral Equation Formalism Polarized Continuum Model (IEFPCM)[163]
GrapheneVASPDFT (PAW)Kohn-Sham equations[140]
Graphene oxideSIESTA codeDFT (LDA)Kohn-Sham equations[113]
MGOAGaussian 09 DFT (B3LYP functional group)Thomas, Yoon–Nelson, and Adams–Bohart models[141]
PyTTA-Dva-COFGaussian 09 DFT (B3LYP functional group)ONIOM model[142]
Vertically aligned (VA) CNT (open-end) hybrid membraneDMOL3 packageDFT (PW91)Exchange-Correlation functional[164]
Ultrafiltration PSF/GO membraneOPEN-MX softwareDFT (LDA)Hoffmann’s model[143]
Graphene oxide Gaussian 09DFT (Gaussian-Lorentzian function)Exchange-Correlation functional[145]
S, N co-doped graphene aerogel (SN-rGO-A)Gaussian 09 DFT (B3LYP functional group)Thomas, Yoon–Nelson, and Adams–Bohart models[165]
ZIF8@carbon nanotubeVASPDFT (PBE)Exchange-Correlation functional[166]
Carbonaceous nanofiber/Ni-Al layered double hydroxide (CNF/LDH)VASPDFT (PAW)Kohn-Sham equations[167]
SWCNTs, MWCNTs, and PACGAMESS DFT (B3LYP5 functional)Exchange-Correlation functional[168]
Single-layer graphene nanosheetsVASPDFT (PAW)Kohn-Sham equations[146]
Graphene oxide Gaussian 09DFT (PBE1PBE functional model)Exchange-Correlation functional[148]
Graphene oxideGaussian 09DFT (B3LYP/6-31G* level)Exchange-Correlation functional[149]
ZnO surfaceVASPDFT (PBE)Exchange-Correlation functional[138]
MoS2 nanosheetsVASPDFT (PAW)Kohn-Sham equations[150]
Zn–Fe LDHMaterials Studio (BIOVIA, 2017)DFT (DMol3) codeExchange-Correlation functional[152]
Lanthanum-aluminium perovskite (La2Al4O9)Materials StudioDFT (PBE)Exchange-Correlation functional[153]
MoS2 nanosheetsVASPDFT (PAW)Kohn-Sham equations[151]
SWCNTsGaussviewDFT (B3LYP5) functionalExchange-Correlation functional[137]
CF/BiOBr/Ag3PO4 clothMaterials StudioDFT (GGA-PBE)Exchange-Correlation functional[154]
Table 3. Nanocomposite materials synthesized by the hydrothermal method.
Table 3. Nanocomposite materials synthesized by the hydrothermal method.
Nanocomposite MaterialMaterial TypeReference
Heterogeneous Fenton catalysts (CNTs/Fh)Oxidized carbon nanotubes (CNTs), ferrihydrite (Fh)[116]
(N-rGO/BiVO4)Bismuth vanadate (BiVO4), reduced graphene oxide (rGO), nitrogen (N)[193]
ZnO@CZinc Oxide (ZnO), carbon (C)[194]
Cerium zirconium oxide (CexZryO2)Cerium (Ce), zirconium oxide (ZrO2)[195]
ZnO/Al2O3Zinc oxide (ZnO), aluminium oxide (Al2O3).[196]
C, N, F/TiO2NTsCarbon (C), nitrogen (N), fluoride (F), titanium dioxide nanotubes (TiO2NTs)[197]
iN-Ti3C2/TiO2 hybridTitanium carbide (Ti3C2), titanium dioxide (TiO2), isopropyl amine, nitrogen (N)[198]
TiO2 nanoflowers
(TNFs)
Titanium dioxide (TiO2)[199]
Titanate nanotubes supported TiO2 (TiO2/TiNTs)Titanium dioxide (TiO2), titanate nanotubes[200]
Black phosphorus quantum dots/Tubular g-C3N4
(BPQDs/TCN)
Black phosphorus (BP), tubular g-C3N4[201]
Sodium titanate nanotubes (Na-TNT)Sodium (Na), titanate nanotubes (TNT)[202]
Fe2O3-PC nanohybridsIron oxide (Fe2O3)[203]
NiO nanobeltNickel oxide (NiO)[204]
Carbon dots/g-C3N4 (C-CN) heterostructuresGraphitic Carbon Nitride (g-C3N4)[205]
AgBr/h-MoO3Silver bromide (AgBr), hexagonal molybdenum oxide (h-MoO3)[206]
Hybrid catalysts (CN-CGs)Coal gangue (CG),
graphitic carbon nitride g-C3N4 (CN)
[207]
N-doped BiVO4Nitrogen (N), bismuth vanadate (BiVO4)[208]
PPECu thin film electrodeCopper (Cu), phenylacetylene (PPE)[209]
FexMo1-xS2 catalystsIron (Fe), Molybdenum disulfide (MoS2)[210]
P-doped porous g-C3N4Graphitic carbon nitride (g-C3N4), phosphorus (P)[211]
1D/2D W18O49/g-C3N4 nanocompositesGraphitic carbon nitride
(g-C3N4), oxygen-deficient tungsten oxide (W18O49)
[212]
Oct-Cu2O NCsCuprous oxide (Cu2O)[213]
g-C3N4Graphitic carbon nitride (g-C3N4)[214]
ZIF8@carbon nanotubeCarbon nanotube (CNT), zeolitic imidazole framework-8 (ZIF8)[166]
CNF/LDHCarbonaceous nanofiber (CNF), nickel (Ni), aluminium (Al)[167]
PVP/MoS2Molybdenum disulphide, polyvinylpyrrolidone[215]
β-CD/TiO2Titanium dioxide (TiO2), β-cyclodextrin C42H70O35[216]
MOF-545Zirconyl chloride octahydrate, Sigma-Aldrich; porphyrin, H4-Tcpp-H2, TCl [217]
Table 4. Nanocomposite materials synthesized by the chemical vapor deposition (CVD) method.
Table 4. Nanocomposite materials synthesized by the chemical vapor deposition (CVD) method.
Nanocomposite MaterialMaterial TypeReference
Co3O4/CNTsCarbon nanotubes (CNTs), cobalt tetra-oxide (Co3O4)[231]
O-CNTs, G-CNTsOxidized carbon nanotubes (O-CNTs), graphitized carbon nanotubes (G-CNTs).[163]
Vertically aligned (VA) CNT (open-end) hybrid membraneCarbon nanotube (CNT), polydimethylsiloxane
(PDMS) membrane
[164]
COOH/CNTsCarbon nanotubes (CNTs), carboxylic functionalized groups (COOH)[232]
Table 5. Nanocomposite materials synthesized by the one-pot synthesis method.
Table 5. Nanocomposite materials synthesized by the one-pot synthesis method.
Nanocomposite MaterialMaterial TypeReference
S, N co-doped graphene aerogel (SN-rGO-A)Graphene oxide (GO), sulfur (S), nitrogen (N).[165]
ZIF-67 Carbocatalysts, Nitrogen-doped magnetic carbon (Co@N-C)Cobalt (Co), nitrogen (N), carbon (C)[235]
Fe/Fe3C@PCGraphitized porous carbon (PC), Fe-based nanoparticle core (Fe/Fe3C)[236]
Table 6. Nanocomposite materials synthesized by other synthesis methods.
Table 6. Nanocomposite materials synthesized by other synthesis methods.
Nanocomposite MaterialMaterial TypeSynthesis MethodReference
Ternary nanocomposites of Fe3O4 nanoparticles@
graphene–poly-N-phenylglycine nanofibers
Graphene oxide (GO), nitrogen (N), iron oxide (Fe3O4), phenylglycine (C6H5CHCO2H).Wet chemical process[241]
Cr-TiO2 supported on Fe3O4Titanium dioxide (TiO2), chromium (Cr), iron oxide black (Fe3O4).Sonochemical method[242]
CdSe-Ag-WO3-Ag photocatalystCadmium selenide (CdSe), silver (Ag), tungsten trioxide (WO3).Continuous photo-assisted process[243]
Bi/Fe0Bismuth (Bi), iron (Fe)Simple chemical reactions [244]
Granular carbon nanotubes (CNTs)Carbon nanotubes (CNTs)Simple heating-filtration method[245]
SWCNTs, MWCNTs, and PACCarbon nanotubes (CNTs)SWCNTs: Obtained commercially from Cheap Tubes, Inc.
MWCNTs: Obtained commercially from Sigma Aldrich.
[168]
Fe3O4-HBPA-ASAMagnetite (Fe3O4)Solvothermal method[246]
Highly porous zeolitic imidazolate frameworks (ZIFs)Highly porous zeolitic imidazolate frameworks Simple stirring method[247]
Granular TiO2-LaTitanium dioxide (TiO2), lanthanum (La)Hydrolysis[248]
Ni (II) modified porous BNNickel (Ni), boron nitride (BN)Precursor pyrolysis method[249]
Bi2O2CO3 nanosheetsBismuth carbonateSimple stirring method[250]
Amino-modified attapulgite (M-ATP)Attapulgite clay, the 3-aminopropyltriethoxysilane, Pb (NO3)2 and Cu(NO3)2 Simple stirring method[251]
g-C3N4Graphitic carbon nitride (g-C3N4)Salt melt method[252]
MIL-101(Fe) and MIL-101(Fe,Co)MIL-101(Fe) Solvothermal method[253]
CuCo2O4/BiVO4Bismuth vanadate (BiVO4)Solvothermal method[254]
Zn/Fe LDHZinc (Zn), iron (Fe)Co-precipitation method[255]
Table 7. Nanocomposite materials simulated by different simulation software.
Table 7. Nanocomposite materials simulated by different simulation software.
Nanocomposite Material Software Simulation MethodMathematical ModelReference
Nitrogen doped carbon (CNs)VASPDFT (PAW)Kohn-Sham equations[144]
COOH/CNTsDMol3 programDFT (PBE)Exchange-Correlation functional[232]
Porous graphitic carbon nitride (g-C3N4)VASPDFT (PAW)Kohn-Sham equations[214]
Granular carbon nanotubes (CNTs)Not suppliedDFTThe Langmuir model [245]
Carbon nanotubes (CNTs)Gaussian 09DFT (Minnesota dispersion functional, M06-2×/6–31G(d) level)Exchange-Correlation functional [147]
MnFe2O4 nanocubesVASPDFT (PAW)Kohn-Sham equations[139]
Oct-Cu2O NCsVASPDFT (PW91)Exchange-Correlation functional[213]
Amino-modified attapulgite (M-ATP)VASPDFT (PBE)Exchange-Correlation functional[251]
β-CD/TiO2VASPDFT (PAW)Kohn-Sham equations[216]
Fe3O4-HBPA-ASAGaussian 16 packageDFT (B3LYP)Exchange-Correlation functional[246]
PVP/MoS2VASPDFT (PAW)Kohn-Sham equations[215]
Highly porous zeolitic imidazolate frameworks (ZIFs)Gaussian 09DFT (B3LYP) Exchange-Correlation functional[247]
Ni (II) modified porous BNVASPDFT (PAW)Kohn-Sham equations[249]
CuCo2O4/BiVO4Materials Studio 6.0 (2011)DFT (PBE) Exchange-Correlation functional[254]
Granular TiO2-LaMaterials Studio 7.0DFT (PBE)Exchange-Correlation functional[248]
g-C3N4Not suppliedDFTLangmuir model, and Freundlich model[252]
MOF-545Not suppliedDFTExchange-Correlation functional[217]
MIL-101(Fe) and MIL-101(Fe, Co)DMol3 codeDFT (PBE)Exchange-Correlation functional[253]
Bi2O2CO3 nanosheetsVASP 5.4DFT (HSE06) Exchange-Correlation functional[250]
Zn/Fe LDHMaterials Studio (BIOVIA, 2017)DFT (GGA-RPBE) Exchange-Correlation functional[255]
Heterogeneous Fenton catalysts (CNTs/Fh)VASPDFT (PAW)Kohn-Sham equations[116]
Co3O4/CNTsMaterial studio 2017DFT (PBE)Exchange-Correlation functional[231]
(ZIF-67 Carbocatalysts), Nitrogen-doped
magnetic carbon (Co@N-C)
VASPDFT (PBE)Exchange-Correlation functional[235]
ternary nanocomposites of Fe3O4 nanoparticles@
graphene–poly-N-phenylglycine nanofibers
VASPDFT (RPBE) Exchange-Correlation functional[241]
(N-rGO/BiVO4)Not suppliedDFTExchange-Correlation functional[193]
Cerium zirconium oxide (CexZryO2)VASPDFT (PBE)Exchange-Correlation functional[195]
NiO nanobeltVASPDFT (PAW)Kohn-Sham equations[204]
ZnO/Al2O3VASP, COMSOL DFT (PBE)Exchange-Correlation functional[196]
ZnO@CMolecular
Operating Environment software (MOE, 2008.10)
DFTExchange-Correlation functional[194]
C, N, F/TiO2NTsVASPDFTExchange-Correlation functional[197]
iN-Ti3C2/TiO2 hybridVASPDFT (PBE)Exchange-Correlation functional[198]
TiO2 nanoflowers
(TNFs)
VASPDFT (PBE)Exchange-Correlation functional[199]
Cr-TiO2 supported on Fe3O4Not suppliedDFT (M06 L)a Langmuir-Hinshelwood model[242]
Titanate nanotubes supported TiO2 (TiO2/TiNTs)Gaussian 03DFT (B3LYP)Exchange-Correlation functional[200]
Black phosphorus quantum dots/Tubular g-C3N4
(BPQDs/TCN)
Materials StudioDFT (PBE)Exchange-Correlation functional[201]
CdSe-Ag-WO3-Ag photocatalystVASPDFT (PBE)Exchange-Correlation functional[243]
Sodium titanate nanotubes (Na-TNT)Materials StudioDFT (RPBE)Exchange-Correlation functional[202]
Fe2O3-PC nanohybridsVASPDFT (PBE)Exchange-Correlation functional[203]
Carbon dots/g-C3N4 (C-CN) heterostructuresVASPDFT (PBE)Exchange-Correlation functional[205]
AgBr/h-MoO3Toxicity Estimation Software Tool (T.E.S.T.)DFT (QSAR) Exchange-Correlation functional[206]
Hybrid catalysts (CN-CGs)VASPDFT (GGA-PBE)Exchange-Correlation functional[207]
Fe/Fe3C@PCVASP, Version 5.4.1DFT (PAW)Kohn-Sham equations[236]
N-doped BiVO4VASPDFT (PBE)Exchange-Correlation functional[208]
Bi/Fe0Materials StudioDFT (PBE)Exchange-Correlation functional[244]
PPECu thin film electrodeVASPDFT (PAW)Kohn-Sham equations[209]
FexMo1-xS2 catalystsVASPDFT (PBE)Exchange-Correlation functional[210]
P-doped porous g-C3N4VASPDFT (PBE)Exchange-Correlation functional[211]
1D/2D W18O49/g-C3N4 nanocompositesVASPDFT (PAW)Kohn-Sham equations[212]
Table 8. Rejection/adsorption capacity of carbon-based membranes.
Table 8. Rejection/adsorption capacity of carbon-based membranes.
MembraneRole of Carbon-Based MembraneRejected/Adsorbed MaterialRate of Rejection (%)/Adsorption Capacity (qe) (mg/g)Reference
O-CNTs, G-CNTsAdsorption of Pb2+ on O-CNTs and G-CNTsPb2+<9.03%[163]
Vertically aligned (VA) CNT (open-end) hybrid membraneGas separation
  • Phenol separation
  • binary CO2/N2 mixture separation
Not supplied[164]
COOH/CNTsAdsorptive removal of Indigo carmine (IC) dye onto nanotube carbon (CNTs)Indigo carmine (IC) dyeCNT: (88.5 mg/g)
COOH-CNT: (136 mg/g)
[232]
Granular carbon nanotubes (CNTs)Efficient removal of typical pharmaceuticalsTypical pharmaceuticalsCBZ: 0.3695 mg/g
TC: 0.2842 mg/g
DS: 0.2031 mg/g
[245]
ZIF8@carbon nanotubeAdsorption of Phosphate on ZIF-8@MWCNTPhosphate(92.8–100%)[166]
SWCNTs, MWCNTs, and PACAdsorption of bisphenol A and 17a-ethinyl estradiol (EE2) using carbon nanomaterials and powdered activated carbonbisphenol A,
17a-ethinyl estradiol (EE2)
90% removal of both BPA and EE2[168]
Carbon nanotubes (CNTs)Adsorption of Sulfamethoxazole (SMZ) and ketoprofen (KET) on modified carbon nanotubes (CNTs)Sulfamethoxazole (SMZ) and ketoprofen (KET)Adsorption percentage:
SMZ: >70%
KET >80%
Removal percentage:
SMZ: 30%
KET: >50%
[147]
GrapheneAdsorption of CHCl3 on grapheneChloroform molecule (CHCl3)Not supplied[140]
Single-layer graphene nanosheetsDesalination and ion capture by sunlight single layer graphene nanosheetNa+, Pb2+ and Fe3+Na+: 86.1%
Pb2+: 77.3%
Fe3+: 46.1%
[146]
Graphene oxideAdsorption of 17 β- estradiol on graphene oxide 17 β- estradiol169.49 mg/g[113]
Graphene oxide Adsorption of As(III) on graphene oxideAs(III)288 mg/g[145]
Graphene oxide Removal of Ni(II) from wastewater by adsorption on graphene oxide surface Ni(II)197.8 mg/g[148]
Graphene oxideAdsorption of Methylene blue (MB) on graphene oxide surface Methylene blue (MB)Not supplied[149]
MGOAAdsorption of quinoline in wastewaterQuinoline pollutants 103 mg/g[141]
Ultrafiltration PSF/GO membraneNitrate rejection, antifouling propertyNitrate 22.5% at 0.5 weight percent of GO[143]
SN-rGO-AAdsorb oils and organic solvents by SN-rGO-AOils and organic solventsqe: 65–192 times its weight[165]
Nitrogen doped carbon (CNs)Adsorbent for the removal of anionic heavy metals from wastewater and sewageArsenic31.08 mg/g[144]
g-C3N4Adsorptive removal of uranyl by porous graphitic carbon nitride (g-C3N4)Uranium149.70 mg/g[214]
g-C3N4Removal of heavy metal ions from aqueous solutionsPb(II), Cu(II), Cd(II) and Ni(II)) Pb(II): 1.36 mmol/g
Cu(II): 2.09 mmol/g
Cd(II): 1.00 mmol/g
Ni(II): 0.64 mmol/g
[252]
Carbonaceous nanofiber/Ni-Al layered double hydroxide (CNF/LDH)Removal of heavy metals from aqueous solutionsCu(II), Cr(VI)Cu(II): 219.6 mg/g
Cr(VI): 341.2 mg/g
[167]
Table 9. Rejection/adsorption capacity of metal oxides.
Table 9. Rejection/adsorption capacity of metal oxides.
MembraneRole of Metal Oxide Rejected MaterialAdsorption Capacity (qe) (mg/g)Reference
ZnO surfaceRemoval of barium (Ba2+) ions on ZnO spherical nanoparticlesBarium ions64.6 mg/g[138]
MnFe2O4 nanocubesHigh adsorption capacity of U(VI) and Eu(III) on magnetic MnFe2O4 nanocubesUranium U(VI)
Eu(III)
U(VI): 119.90 mg/g
Eu(III): 473.93 mg/g
[139]
Oct-Cu2O NCsAdsorption of tetracycline on octahedral Cu2O nanocrystalsTetracycline1112.6 mg/g[213]
Table 10. Rejection/adsorption capacity of other nanocomposite membranes.
Table 10. Rejection/adsorption capacity of other nanocomposite membranes.
MembraneRole of Nanocomposite Membrane Rejected MaterialRate of Rejection (%)/Adsorption Capacity (qe) (mg/g)Reference
PyTTA-Dva-COFRemoval of bisphenol A from aqueous solutionbisphenol A285 mg/g[142]
Zn–Fe LDHRemoval of diclofenac from water using Zn–Fe LDHDiclofenac74.50 mg/g[152]
Lanthanum-aluminium perovskite (La2Al4O9)Adsorption mechanisms for removing fluoride using lanthanum-aluminum perovskiteFluoride (F)87.75 mg/g[153]
β-CD/TiO2Adsorption mechanisms for uranium removal by β-CD/TiO2U(VI)129.8 mg/g[216]
Fe3O4-HBPA-ASARemoval of heavy metal ions from aqueous solution by Fe3O4-HBPA-ASAHeavy metal ionsCu(II): 136.66 mg/g
Pb(II): 88.36 mg/g
Cd(II): 165.46 mg/g
[246]
ZIFsHighly efficient removal of U(IV)U(VI)ZIF-8: 540.4 mg/g
Zn/Co-ZIF: 527.5 mg/g
ZIF-9: 448.6 mg/g
ZIF-67: 368.2 mg/g
[247]
Granular TiO2-LaAdsorption of arsenic and fluoride using granular TiO2-La Arsenic (As III), fluoride (F)As(III): 114 mg/g
F: 78.4 mg/g
[248]
Ni (II) modified porous BNRemoval of tetracycline from aqueous solution Tetracycline (Tc)429.582 mg/g
Removal percentage: 99.769%
[249]
Bi2O2CO3 (BOC) nanosheets with oxygen vacancies Removal of (NO) by BOC nanosheets Nitric oxide (NO)Removal percentage: 50.2%[250]
Amino-modified attapulgite (M-ATP)Removal of Pb2+, and Cu2+ by adsorption on Amino-modified attapulgite (M-ATP)Pb2+, Cu2+Pb2+: 53.58 mg/g
Cu2+: 28.86 mg/g
[251]
MIL-101(Fe) and MIL-101(Fe,Co)Removal of Ciprofloxacin (CIP) by MIL-101(Fe) and MIL-101(Fe,Co)Ciprofloxacin (CIP)Removal percentage: 97.8%[253]
CuCo2O4/BiVO4Removal of 4-Nitrophenol 4-NitrophenolNot supplied[254]
MOF-545Removal of lead by adsorption on (MOF-545)Pb(II)Pb(II): 73 mg/g[217]
Zn/Fe LDHRemoval of oxytetracycline hydrochloride (OTC) by adsorption on Zn/Fe LDHOxytetracycline hydrochloride (OTC)Removal percentage: 77.23%[255]
MoS2Removal of uranyl ions U(VI) by adsorption on MoS2U(VI)MoS2 nanosheets: 45.7 mg/g
MoS2 nanoflowers: 37.1 mg/g
[150]
MoS2 nanosheetsRemoval of Pb2+ in aquatic systems by MoS2 nanosheetsToxic metals (Pb2+), (Cd2+), Pb2+: 638 mg/g under 1 sun illuminations, 902 mg/g under 4 sun illuminations
Cd2+: 489 mg/g under 1 sun illuminations, 719 mg/g under 4 sun illuminations
[151]
PVP/MoS2Removal of uranyl ions by adsorption on PVP/MoS2U(VI)U(VI): 117.9 mg/g[215]
Table 11. Decomposition rate/degradation efficiency of titanium dioxide (TiO2).
Table 11. Decomposition rate/degradation efficiency of titanium dioxide (TiO2).
Nanocomposite MaterialRole of TiO2Degraded MaterialDecomposition Rate (min−1)/Degradation Efficiency (%)Reference
C, N, F/TiO2NTsHigh photocatalytic activity under UV-lightMethyl orangeUnder UV-light:
  • TiO2NTs: 60%
  • C/TiO2NTs: 100%
Under simulated sunlight:
  • N,F/TiO2NTs: high activity
  • TiO2NTs: low activity (high bandgap)
  • C/TiO2NTs: high activity
[197]
iN-Ti3C2/TiO2 hybridAchieved a high photocatalytic performance in degrading MB.Methylene blue (MB)Under UV-light:0.02642 min−1[198]
TiO2 nanoflowers(TNFs)high photocatalytic performance for the degradation of diverse phenolic organic contaminantsBisphenol A (BPA), diphenyl phenol, P-tert-butyl phenol, and resorcinolUnder UV-light:>95%[199]
(TiO2/TiNTs)TiO2/TiNTs showed about 10 times higher degradation for phenanthrene compared to the unmodified TiNTsCu(II), phenanthrene
  • Cu(II) adsorption capacity: 115.0 mg/g
Under UV-light:
  • Removal of >95% phenanthrene
[200]
Cr-TiO2 supported on Fe3O4 High photocatalytic activity under solar radiationMalachite green dye (MG), total organic carbon (TOC)Under solar radiation:
  • 100% removal of MG
  • 60% removal of TOC
[242]
Table 12. Decomposition rate/degradation efficiency of carbon nanomaterials.
Table 12. Decomposition rate/degradation efficiency of carbon nanomaterials.
Nanocomposite MaterialRole of Carbon NanomaterialsDegraded MaterialDecomposition Rate (min−1)/Degradation Efficiency (%)Reference
Heterogeneous Fenton catalysts (CNTs/Fh)Degradation of bisphenol Abisphenol A3% CNTs/Fh: 79.1%[116]
Co3O4/CNTsDegradation of norfloxacin (NX)NX97.5%[231]
ZIF-67 Carbocatalysts, Nitrogen-doped
magnetic carbon (Co@N-C)
Degradation of BPABPA60%[235]
Ternary nanocomposites of Fe3O4 nanoparticles@
graphene–poly-N-phenylglycine nanofibers
Adsorption of Cu2+Cu2+ 95%[241]
SWCNTsDegradation of pharmaceutical: PhACs, ibuprofen (IBP) and sulfamethoxazole (SMX)ibuprofen and sulfamethoxazoleAt pH = 3.5:
99% for IBP and SMX
[137]
CF/BiOBr/Ag3PO4 clothDegradation of tetracycline hydrochloride (TCH)TCH90%[154]
(N-rGO/BiVO4)Degradation of methylene blue (MB)MB99.3%[193]
Table 13. Decomposition rate/degradation efficiency of metal oxides.
Table 13. Decomposition rate/degradation efficiency of metal oxides.
Nanocomposite MaterialRole of Metal OxideDegraded MaterialDecomposition Rate (min−1)/Degradation Efficiency (%)Reference
ZnO@CPhotocatalytic degradation of methylene blueMethylene blue (MB)99.8%[194]
Cerium zirconium oxide (CexZryO2)Photocatalytic degradation of sulfonamidesSulfonamides91.33%[195]
ZnO/Al2O3Wastewater treatmentMethyl orange dye (MO), TOCTOC: 80.4%[196]
NiO nanobeltRemoval of organic pollutants such as RhB, MO, MB, and CVRemoval of organic pollutantsRhB: 89%
CV: 76.7%
MB: 82.7%
MO: 79.1%
[204]
Table 14. Decomposition rate/degradation efficiency of other nanocomposites.
Table 14. Decomposition rate/degradation efficiency of other nanocomposites.
Nanocomposite MaterialRole of the NanocompositeDegraded MaterialDecomposition Rate (min−1)/Degradation Efficiency (%)Reference
Black phosphorus quantum dots/Tubular g-C3N4
(BPQDs/TCN)
Facilitates the charge spatial separation in the photocatalytic process which improves the process efficiencyOxytetracycline hydrochloride, hexavalent chromium reductionOxytetracycline hydrochloride: 0.0276 min−1, Hexavalent chromium: 0.0404 min−1[201]
CdSe-Ag-WO3-Ag photocatalystStrong redox capacity, enhanced optical absorption and accelerated transfer and separation of carriersCefazolin (CFZ)CFZ: 96.32% in 30 min[243]
Sodium titanate nanotubes (Na-TNT)Photocatalytic degradation of nickel (Ni(II)), methylene blue (MB)Nickel (Ni(II)), methylene blue (MB)90% of Ni(II) ions within the first 15 min.
Removed 99.5% of MB
[202]
Fe2O3-PC nanohybridsPhotocatalytic degradation of methylene blue (MB)Methylene blue (MB)Fe2O3: Removed 56% of MB
Fe2O3-PC: Removed 75% of MB
[203]
Carbon dots/g-C3N4 (C-CN) heterostructuresPhotocatalytic degradation of sulfamethoxazole (SDZ)Sulfamethoxazole (SDZ)0.5C-CN: 62.7%
1C-CN: 75.2%
3C-CN: 92.8%
5C-CN: 85.7%
[205]
AgBr/h-MoO3 compositePhotocatalytic degradation of trimethoprim (TMP)Trimethoprim (TMP)TMP: 97%[206]
CN-CGsPhotocatalytic degradation of Total organic carbon (TOC), bisphenol A (BPA)Total organic carbon (TOC),
bisphenol A (BPA)
BPA: 90%
TOC: 80%
[207]
Fe/Fe3C@PCPhotocatalytic degradation of sulfamethazine (SMT)Sulfamethazine (SMT)SMT: 99.2%[236]
N-doped BiVO4Photocatalytic degradation of ibuprofen (IBP)Ibuprofen (IBF)IBP: 90%[208]
Bi/Fe0Photocatalytic degradation of hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX)hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX)RDX: The best RDX degradation
was achieved using 4%-Bi/Fe0 (atomic ratio) NPs
[244]
PPECu thin film electrodePhotocatalytic degradation of phenol and 2,4-DCPPhenol, 2,4-DCPPhotocatalytic
degradation of phenol and 2,4-DCP activity was 2.52 and 3.85 times
higher than g-C3N4
[209]
FexMo1-xS2 catalystsPhotocatalytic degradation of propranololPropranolol90% at pH = 4.0[210]
P-doped porous g-C3N4Photocatalytic degradation of rhodamine B (RhB)Rhodamine B (RhB)RhB: 99.5%[211]
1D/2D W18O49/g-C3N4 nanocompositesPhotocatalytic degradation of ibuprofen (IBF)Ibuprofen (IBF)IBF: 96.3%[212]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ibrahim, Q.; Creedon, L.; Gharbia, S. A Literature Review of Modelling and Experimental Studies of Water Treatment by Adsorption Processes on Nanomaterials. Membranes 2022, 12, 360. https://doi.org/10.3390/membranes12040360

AMA Style

Ibrahim Q, Creedon L, Gharbia S. A Literature Review of Modelling and Experimental Studies of Water Treatment by Adsorption Processes on Nanomaterials. Membranes. 2022; 12(4):360. https://doi.org/10.3390/membranes12040360

Chicago/Turabian Style

Ibrahim, Qusai, Leo Creedon, and Salem Gharbia. 2022. "A Literature Review of Modelling and Experimental Studies of Water Treatment by Adsorption Processes on Nanomaterials" Membranes 12, no. 4: 360. https://doi.org/10.3390/membranes12040360

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