Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Nanoscale Materials in Water Purification
Nanoscale Materials in Water Purification
Nanoscale Materials in Water Purification
Ebook1,915 pages13 hours

Nanoscale Materials in Water Purification

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Novel nanoscale materials are now an essential part of meeting the current and future needs for clean water, and are at the heart of the development of novel technologies to desalinate water. The unique properties of nanomaterials and their convergence with current treatment technologies present great opportunities to revolutionize water and wastewater treatment. Nanoscale Materials for Water Purification brings together sustainable solutions using novel nanomaterials to alleviate the physical effects of water scarcity.

This book covers a wide range of nanomaterials, including noble metal nanoparticles, magnetic nanoparticles, dendrimers, bioactive nanoparticles, polysaccharidebased nanoparticles, nanocatalysts, and redox nanoparticles for water purification. Significant properties and characterization methods of nanomaterials such as surface morphology, mechanical properties, and adsorption capacities are also investigated

  • Explains how the unique properties of a range of nanomaterials makes them important water purification agents
  • Shows how the use of nanotechnology can help create cheaper, more reliable, less energy-intensive, more environmentally friendly water purification techniques
  • Includes case studies to show how nanotechnology has successfully been integrated into water purification system design
LanguageEnglish
Release dateNov 14, 2018
ISBN9780128139271
Nanoscale Materials in Water Purification

Related to Nanoscale Materials in Water Purification

Titles in the series (97)

View More

Related ebooks

Science & Mathematics For You

View More

Related articles

Reviews for Nanoscale Materials in Water Purification

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Nanoscale Materials in Water Purification - Sabu Thomas

    China

    Chapter 1

    Nanomaterials—State of Art, New Challenges, and Opportunities

    Deepu A. Gopakumar⁎; Avinash R. Pai†; Daniel Pasquini‡; Shao-Yuan (Ben)Leu§; Abdul Khalil H.P.S.⁎; Sabu Thomas†    ⁎ School of Industrial Technology, Universiti Sains Malaysia (USM), Gelugor, Penang, Malaysia

    † International and Inter University Centre for Nanoscience and Nanotechnology, Mahatma Gandhi University, Kottayam, India

    ‡ Chemistry Institute, Federal University of Uberlandia-UFU, Uberlandia, Brazil

    § Department of Civil and Environmental Engineering ZS923, The Hong Kong Polytechnic University Hung Hom, Kowloon, Hong Kong

    Abstract

    Nanomaterials have a number of key physicochemical properties that make them particularly. Attractive as separation media for water purification. On a mass basis, they have much larger surface areas than bulk particles. Nanomaterials can also be functionalized with various chemical groups to increase their affinity toward a given, particular pollutant in water. Due to their effective surface area and diverse functionalities, nanomaterials are promising candidates for effective water purification. In this context, this chapter gives an overview of the use of nanomaterials in water purification. We discuss the current developments in novel nanoscale materials and techniques for the purification of wastewater contaminated by toxic metal ions, organic and inorganic solutes, bacteria, and viruses.

    Keywords

    Nanotechnology; Nanomaterials; Membranes; Water purification

    1 Introduction

    Nanotechnology is considered an interdisciplinary branch of science integrating engineering aspects with biology, chemistry, and physics [1]. It descends from the present trend of miniaturization in devices and technology (as described by Moore's Law) and combining all other branches of science. Nanotechnology, however, has its drawbacks, and there is a strong need for developing new approaches in manufacturing (bottom-up approaches), which have to be developed to accomplish anticipated milestones. Nanotechnology is generally considered a tool of science that steps across the limit of miniaturization, where the materials exhibit different behavior as compared to macroscopic scale. More specifically, when the dimensions of a solid material approach nanoscale, its physical and chemical properties can become very different from those of the same material at the macrolevel. This is the main highlight of nanotechnology, which can be described as a research area in which this limit of new properties is reached, and strategies are developed to exploit the regime of size-controlled properties.

    In the past few decades, the term nanotechnology has been extensively discussed and has almost become synonymous for things that are innovative and highly promising. On the other hand, the area of nanoscience is a subject of considerable debate regarding the open question of toxic and health hazards of nanoparticles and other nano-objects [2, 3]. In this argument, a definition of nanotechnology and its underlying sectors, applications, and markets is important for the purpose of risk assessment and risk communication. Many researchers refer to the length scale (nano) of this new science, but not all focus on the novel functionalities and properties of materials at the nanodimension. In general, the most accepted definition refers to the size and varying properties of materials in the size range between 10 and 100 nm, but this gives rise to many uncertainties and discrepancies which need to be resolved.

    2 Nanomaterials for Water Filtration

    Conventional water treatment techniques like sedimentation, flocculation, coagulation, and activated carbon are not able to eliminate organic pollutants from water effectively. In this context, membrane techniques play a vital role for water purification [4, 5]. Among the membrane techniques, microfiltration (MF), ultrafiltration (UF), and nanofiltration (NF) have gained considerable interest due to their ease of operation [6, 7]. Because membrane processes are considered important components of advanced water purification and desalination technologies, there is a need for a continuous search for new advanced materials for water purification technologies. In this context, nanomaterials (e.g., carbon nanotubes, nanoparticles, and dendrimers) are contributing to the development of more efficient and cost-effective water filtration processes. Fig. 1 highlights four classes of nanoscale materials that are being evaluated as functional materials for water purification: (1) metal-containing nanoparticles, (2) carbonaceous nanomaterials, (3) zeolites, and (4), dendrimers. These have a broad range of physicochemical properties that make them particularly attractive as separation and reactive media for water purification.

    Fig. 1 Selected nanomaterials currently being evaluated as functional materials for water purification. Reproduced with permission from N. Savage, M.S. Diallo, Nanomaterials and water purification: opportunities and challenges, J. Nanopart. Res. 7 (2005) 331–342, Springer Nature.

    2.1 Classification of Nanomaterials for Water Purification

    Nanomaterials can be classified into 0D (fullerenes), 1D (graphene nanoribbons and carbon nanotubes), 2D (graphene sheets), and 3D (nanoparticles) hybrid architectures based on their special confinement in a three-dimensional space. They can exist in single, fused, and hybridized forms with spherical, tubular, and irregular shapes as shown in Fig. 2A–G. Common types of nanomaterials include nanotubes, nanowires, nanofibers, quantum dots, and fullerenes. These hybrid 3D networks of nanomaterials possess several advantages such as highly accessible surface area, minimal aggregation or re-stacking, enhanced thermal and electrical transport, and robust mechanical properties as compared to their nanoscale 1D or 2D building blocks.

    Fig. 2 (A) 0D nanomaterials, (B, C) 1D nanomaterials, (D) 2D nanomaterials, and (E–G) 3D hybrid assemblies of CNT and graphene. Reproduced with permission from A. Dasgupta, L.P. Rajukumar, C. Rotella, Y. Lei, M. Terrones, Covalent three-dimensional networks of graphene and carbon nanotubes: synthesis and environmental applications, Nano Today (2017) 116–135, Elsevier Ltd.

    0D nanomaterials—fullerenes

    The existence of carbon clusters was proposed and came into existence in 1959. It actually took decades to expound the structure and elucidate the properties of fullerenes [8]. The nature and chemical reactivity of fullerenes was first reported by Kroto et al., which attracted researchers’ attention to this new, interesting class of carbon allotrope [9]. In 1990, Kratschmer et al. made another breakthrough for synthetic chemists by formulating a process for synthesizing macroscopic amounts of C60 [10]. Much attention has been received by C60 structures in recent times, as many other carbon allotropes may have developing potential applications [11, 12]. The name buckminsterfullerene was coined as an honorific for Buckminster Fuller, an American architect and constructor of the geodesic dome. In the family of fullerenes, C60 (buckminsterfullerene) and C70 (falmarene) have been extensively studied by researchers because of their outstanding properties and cage structure. Following the availability of macroscopic amounts of fullerenes, a variety of chemical modifications, such as, for example, arylation, halogenation, hydroxylation, alkoxylation, and osmylation, have been performed on C60 and its analogues [13]. Progressively, fullerenes have gained much interest from polymer chemists as building blocks for the construction of novel materials with fascinating properties.

    Initial attempts to synthesize fullerene were made by irradiating graphite with a laser beam, thereby producing long-chain carbon molecules formed in the interstellar space and producing a remarkably stable cluster consisting of 60 carbon atoms. This experiment led to an outburst of interest in the chemistry of C60. After 5 years, a method for producing large quantities of C60 was developed by vaporizing graphite rods in a helium atmosphere by a resistive heating technique [10]. This method made a significant breakthrough in fullerene science by conducting further modification in the molecule for potential applications in diverse areas of technology. Many different methods have been proposed with the aim of synthesizing C60. Unfortunately, only a few of them have been successful. For instance, one particularly attractive strategy is the assembly of two identical hemispherical hydrocarbons [14]. However, methods for fusing such molecules do not currently exist. The major challenge in the synthesis of fullerenes is the introduction of curvatures or pyramidalizations in the carbon network. Fig. 3 shows an array of applications which have been developed with the advancement of fullerene science.

    Fig. 3 A schematic representation of potential applications of fullerenes.

    Fullerenes are molecules with 60 atoms and have received much attention because of their wide application [15]. Due to their insolubility in water, solvent exchange and ultrasonification have been applied to disperse fullerenes as nanoparticles [16]. C60 nanoparticles have a broad spectrum antimicrobial activity, although the mode of action is debatable [17]. The toxicity of C60 is also under investigation. The use of C60 as antifouling agents for ceramic membranes was determined by evaluating the effect on the attachment and respiratory activity of E. coli K12. A decrease in bacterial numbers and increase in inactive E. coli cells confirmed the bactericidal effect of C60. The increase in membrane hydrophobicity might also play a role in its antimicrobial activity [18].

    1D nanomaterial—carbon nanotubes

    Carbon nanotubes (CNTs), also known as bucky tubes, are made of cylindrical carbon moieties with superior physical and chemical properties that make them possibly useful in a wide variety of applications, from sensors to water purification membranes. These include applications in nano-electronics,opto-electronic devices, and other engineered applications. CNTs exhibit extraordinary mechanical strength as well as distinctive electrical and thermal properties. CNTs are considered a sub class of the fullerene family, which was discovered by Kroto et al. in 1985 [19]. Bucky balls are commonly known as spherical fullerenes, whereas CNTs are cylindrical, with at least one end typically capped with a hemisphere that has the bucky ball structure. The name CNT derives from the size, as the diameter of a nanotube is on the order of a few nanometers. The synthesis of multiwalled carbon nanotubes (MWNTs) was first reported by Iijima and his colleagues in 1991 using a simple arc-evaporation method [20]. However, in the early 1950s, predictions were made about the existence of CNTs by Radushkevich and Lukyanovich, who reported the discovery of worm-like carbon formations [21]. The formation of long filaments of needle-like carbon crystals with a diameter of about 50 nm was achieved during their study of the soot formed by the decomposition of carbon monoxide (CO) on iron particles at 600°C. The discovery of carbonaceous nanotubes was left unnoticed due to lack of attention to the nanotechnology sector at that time, like a number of other areas before and after [22]. Fig. 4 shows 1D and 3D nanostructured materials derived from 2D graphene sheets.

    Fig. 4 A graphical representation of various forms of 0D (fullerenes), 1D (carbon nanotubes), and 3D (graphene sheets stacked to graphite) architectures derived from 2D graphene. Reproduced with permission from A.K. Geim, K.S. Novoselov, The rise of graphene, Nat. Mater. 6(3) (2007) 183–191, Springer Nature.

    Carbon nanotubes can be synthesized by various physicochemical methods such as the arc discharge method, chemical vapor deposition, vapor phase growth, and the flame synthesis method. In 1991, Iijima and colleagues reported the existence of a novel tubular-shaped carbonaceous species, which are now known as CNTs [20]. A simple arc-discharge evaporation technique similar to the method used for the production of fullerene was used to prepare CNTs. Long, continuous, needle-like carbon (4–30 nm diameter) and lengths up to 1 mm were grown on the negative end (cathode) of a carbon electrode by means of the direct current (DC) arc-discharge evaporation of carbon in an argon-filled vessel at a pressure of 100 Torr. A method for large-scale synthesis of MWNTs was reported by Ebbesen, and Ajayan reported using the same technique [23]. The arc-discharge assembly consists of two vertical thin electrodes placed at the center of a chamber. The lower electrode (cathode) has a narrow dip to hold a small piece of iron during the evaporation phase. The arc discharge can be generated by running a DC current of 200 A at 20 V between the two electrodes. The ratio of three compositions, that is, argon, iron, and methane, is very crucial for the synthesis of single-walled CNTs. Some of the potential applications of CNTs in diverse technological sectors are given in Fig. 5.

    Fig. 5 Few potential applications of carbon nanotubes in diverse technological sectors.

    CNTs are composed of cylindrical graphite sheets (allotropic form of carbon) rolled up in a tube-like structure with the appearance of latticework fence [24]. Single-walled carbon nanotubes (SWCNTs) have cylindrical shape consisting of a single shell of graphene. On the other hand, multiwalled carbon nanotubes (MWCNTs) are composed of multiple layers of graphene sheets. Both SWCNTs and MWCNTs have been used for direct water desalination [25–27] or indirectly to remove troublemaking compounds that complicate the desalination processes [28]. CNTs are fascinating in advanced membrane technologies for water desalination because they provide a low-energy solution for water treatment. CNT membranes provide near-frictionless water flow through themselves with the retention of a broad spectrum of water pollutants. The inner hollow cavity of CNTs provides a great possibility for desalinating water. The high aspect ratios, smooth hydrophobic walls, and inner pore diameter of CNTs allow ultra-efficient transport of water molecules. Some prototypes of CNT-based membranes are shown in Fig. 6.

    Fig. 6 Structures of some CNT membranes. Shown are (A) a cross-sectional scanning electron microscope (SEM) image of a pristine CNT membrane. (B) A CNT-based water filter with cylindrical geometry. (C) Movement of water molecules through a CNT channel. (D) A SEM image of scattered NaCl nanocrystals on CNT membrane surface. (E) Movement of pure water molecules through a CNT membrane in osmotically imbalanced compartments, and (F), Engineered CNT membranes in an industrial setup. Reproduced with permission from R. Das, M.E. Ali, S.B.A. Hamid, S. Ramakrishna, Z.Z. Chowdhury, Carbon nanotube membranes for water purification: a bright future in water desalination, Desalination (2014) 97–109, Elsevier Ltd.

    2D nanomaterial—graphene

    Graphene is basically a flat monolayer of carbon atoms closely packed into a two-dimensional (2D) honeycomb lattice, and is a basic building block for graphitic materials of all other dimensionalities. One can make any kind of nanostructure from this fascinating material by wrapping it up into 0D fullerenes, rolling it into 1D nanotubes, or stacking it into 3D graphite. Theoretically, graphene and its various configurations have been studied for 60 years, and graphene is widely used for elucidating the properties of different carbonaceous materials [29]. Forty years later, it was realized that graphene also provides an excellent condensed-matter analogue of (2 + 1)-dimensional quantum electrodynamics, which thrust graphene into a thriving theoretical toy model [30]. On the other hand, although known as an integral part of 3D materials, graphene was presumed not to exist in the free state, being described as an academic material, and was believed to be unstable with respect to the formation of curved structures such as soot, fullerenes, and nanotubes. Suddenly, the vintage model turned into reality, when freestanding graphene was unexpectedly found 3 years ago, and especially when follow-up experiments confirmed that its charge carriers were indeed massless Dirac fermions [31].

    Graphene sheets can be synthesized on a small scale by chemical vapor deposition (CVD) [32], epitaxial growth [33], and mechanical exfoliation (the Scotch tape method) [34]. Large-scale industrial production of graphene is primarily accomplished by oxidation of graphite and the subsequent reduction of graphene oxide (GO). The synthesis of GO was first reported by Brodie in 1859. Currently, three representative synthesis methods and their variations have been developed, attributable to Brodie [35], Staudenmaier [36], and Hummers [37]. In these three methods, the layers in graphite are thoroughly oxidized by oxidative treatment in fuming acids, as follows: KClO3 in HNO3 (Brodie), KClO3 in HNO3/H2SO4 (Staudenmaier), and KMnO4 in H2SO4 (Hummers). Consequently, the graphene sheets become functionalized with hydroxyl and epoxide groups on their basal plane, while the edges are decorated with carbonyl and carboxyl groups [38].

    Han et al. reported graphene nanofiltration membranes (uGNMs) (≈ 22–53 nm thick) on microporous substrates presented for efficient water purification using chemically converted graphene (CCG) as shown in Fig. 7. The prepared uGNMs exhibited well-packed layer structure formed by CCG sheets, which was confirmed by microscopic techniques. The performance of the uGNMs for water treatment was evaluated on a dead-end filtration device, and the pure water flux of uGNMs was high (21.8 L m− 2 h− 1 bar− 1). The uGNMs showed high retention (> 99%) for organic dyes and moderate retention (≈ 20%–60%) for ion salts. The rejection mechanism of this kind of negatively charged membranes was intensively studied, and the results revealed that the physical sieving and electrostatic interaction dominated the rejection process. Finally they concluded that the integration of high performance, low cost, and simple solution-based fabrication process promises uGNMs’ great potential application in practical water purification [39].

    Fig. 7 (A) Digital photo of an uGNM coated on an AAO disk (left) and a twisted uGNM coated on a PVDF membrane (right). (B) Schematic representation of a brGO: graphene sheet with a certain quantity of holes and most of the oxidized groups located on the edges, and the periphery of the holes on it. Note that the real graphene sheets extend further than depicted. (C) Schematic view for possible permeation route: water molecules go through the nanochannels of the uGNMs and the holes on the graphene sheets and at last reach the pores of supporting membranes. The blank squares present the holes on the graphene sheets (black line) . The edges of the brGO and the periphery of the holes are negatively charged. Reproduced with permission from Y. Han, Z. Xu, C. Gao, Ultrathin graphene nanofiltration membrane for water purification ultrathin graphene nanofi ltration membrane for water purifi cation, Adv. Funct. Mater. (2013), John Wiley and Sons.

    Nanocellulose

    With the increase in water quality regulations and decrease in available fresh water supplies, scientists across the world focus on renewable raw materials and environmentally friendly, cost-effective materials for water purification. Among these, nanocellulose has gained considerable interest because of its excellent properties. Many laboratory operations are dependent upon the filtration processes that are performed with cellulose-based filter papers. Nanocellulose's inherent fibrous nature and remarkable mechanical properties, including its low cost and biocompatibility, make nanocellulose a huge potential component in water filtration membranes [40]. Nanocelluloses are very promising adsorbents for removing heavy metals, viruses, dyes, etc., due to their high surface-area-to-volume ratio, low cost, high natural abundance, and inherent environmental inertness [41]. Moreover, the surface of nanocellulose has easily functionalizable OH groups, and this facilitates the incorporation of chemical moieties that may increase the binding efficiency of pollutants to the nanocellulosic materials [40]. Cellulose nanofibers (CNF) are one type of cellulose nanomaterials. CNFs are a promising substitute adsorbent due to their high surface-areato-volume ratio, low cost, high natural abundance, and inherent environmental inertness. Furthermore, CNFs readily functionalized surface facilitates the incorporation of chemical moieties that may enhance the binding efficiency of pollutants to the CNF. Mainly, nanocelluloses are classified into (1) cellulose nanofibers (CNFs) and (2) cellulose nanocrystals (CNCs). The CNCs have nanodimension in both length and diameter, whereas CNFs have length in microdimension and diameter in the nanodimension is shown in Fig. 8. Isolation of nanocellulose from renewable source is becoming an important area for the researchers all over the world. A lot of procedures have been reported for the extraction of CNFs from the plant's cell wall. There are several mechanical treatments which have been used to extract CNFs from various sources. These processes mainly include high-pressure homogenization, grinding, cryocrushing. and high-intensity ultrasonic treatments which lead to the transverse cleavage of cellulose fibers.

    Fig. 8 Hierarchical structure of wood cellulose, forming cellulose nanofibers and nanocrystals. Reproduced with permission from A. Isogai, Wood nanocelluloses: fundamentals and applications as new bio-based nanomaterials, J. Wood Sci. (2013) 449–459, Springer.

    Nanochitin

    Chitin is the second most abundant natural polymer after cellulose. Chitin occurs in the structural components of arthropod exoskeletons or in the cell walls of fungi and yeast [42]. Chitin and cellulose are almost similar polysaccharide compounds; cellulose contains a hydroxyl group, whereas chitin contains an acetamide group. Chitin is a high molecular weight polymer, specifically β (1-4)(N-acetyl-d-glucoamine), as shown in Fig. 9. The high crystallinity and mechanical strength of nanochitin make it a suitable candidate for reinforcement in polymer nanocomposites. Chitin is natural, nontoxic, nonallergic, antimicrobial, and biodegradable, and it is insoluble in water. There are several reports in the literature about the extraction of nanochitin from different sources. Acid hydrolysis is a well-known procedure for extracting nanochitin from seafood wastes. Chitin, the second most abundant polysaccharide, is synthesized by an enormous number of living organisms, including fungi and insects. These biopolymers have found many applications in different areas such as packaging material, membranes for removal of metal ions, and dyes and pigments in wastewater engineering. Chitin has been reported to be useful in heavy metal chelating of industrial wastewater. Chitin is inexpensive, nontoxic, and can bind to the common water pollutants (Fig. 10). This binding consists of mercury, copper, iron, nickel, chromium, lead, zinc, cadmium, silver, and cobalt.

    Fig. 9 Schematic representations of hierarchical structure of cuticles and existence of chitin in them. Reproduced with permission from M. Barikani, E. Oliaei, H. Seddiqi, H. Honarkar, Preparation and application of chitin and its derivatives: a review, Iran. Polym. J. Eng. Ed. (2014) 307–326, Springer Nature.

    Fig. 10 Biosorption of heavy metals by chitin. Reproduced with permission from G. Camci-Unal, N.L.B. Pohl, Quantitative determination of heavy metal contaminant complexation by the carbohydrate polymer chitin, J. Chem. Eng. Data 55(3) (2010) 1117–1121, American Chemical Society.

    The results showed that the strongest binding takes place with mercury and the weakest with cobalt [43].

    Noble metal nanoparticles

    Noble metal nanoparticle-based chemistry has immense advantages in the area of drinking water purification. The noble metal nanoparticles have multipronged use in drinking water purification (removal of organic compounds, heavy metals, and microorganisms). While the chemistry of noble metal nanoparticles is unique in the removal of many contaminants, their high energy surface results in the system attempting to minimize the surface energy through protection or chemical transformation or agglomeration. Therefore, nanoparticles are likely to adsorb a number of other species onto their surface. The same properties of nanomaterials (increased surface area, chemical reactivity, etc.) which confer the unique ability to degrade toxic species present in the environment may also render them toxic to humans. Among all the recently developed nano-adsorbents for drinking water purification, the chemistry of noble metal nanoparticles is truly unique. The chemistry of silver nanoparticles has been utilized to remove a number of toxic contaminants found in drinking water including pesticides, heavy metals, and microorganisms. Gold nanoparticles have been effectively used to detect ultra-low concentrations of pesticides in water via changes in the signature properties of a functional group attached to the gold nanoparticle surfaces in the presence of organic molecules like pesticides. Gold nanoparticle surfaces can be modified with the OPH enzyme, and fluorophore molecules are positioned at the OPH active sites. The addition of organophosphorus compounds, leading to the release of a fluorophore molecule, is measured through changes in the fluorescence intensity [44]. The AChE (bound on gold nanoparticle surfaces)-mediated hydrolysis reaction of acetylthiocholine has also been utilized as a pesticide sensor through measurement of electrochemical current, which depends on the extent of formation of thiocholine in the presence of pesticides [45].

    Metal-basednano-adsorbents

    In recent years, the synthesis and utilization of iron oxide metal nanoparticles with novel properties and functions have been widely studied due to their size in nano-range, high surface-area-to-volume ratios, and super paramagnetism [46–48]. Additionally, iron oxide metal nanoparticles with low toxicity, chemical inertness, and biocompatibility show a tremendous potential in combination with biotechnology [49–51]. The stability of iron oxide colloid suspensions could be greatly augmented by surface modification with suitable functional groups, such as phosphonic acids, carboxylic acid, and amines, as shown in Fig. 11 [52, 53].

    Fig. 11 Common chemical moieties for the anchoring of polymers and functional groups at the surface of iron oxide magnetic nanoparticles. Reproduced with permission from P. Xu, G.M. Zeng, D.L. Huang, C.L. Feng, S. Hu, M.H. Zhao, C. Lai, Z. Wei, C. Huang, G.X. Xie, et al., Use of iron oxide nanomaterials in wastewater treatment: a review, Sci. Total Environ. (2012) 1–10, Elsevier Ltd.

    Mak and Chen [54] investigated the adsorption of methylene blue (MB) from an aqueous solution by polyacrylic acid-bound, iron oxide, magnetic nanoparticle. They demonstrated that the novel magnetic nano-adsorbent was quite efficient for the adsorption/desorption of MB. They showed that the adsorption capacity increased with the increase in solution pH (2–10), and the adsorption process was endothermic in nature. By using the methanol solution of acetic acid, they demonstrated that the adsorbed MB could be desorbed. Moreover, they concluded that the both adsorption and desorption of MB were quite fast and could be completed within 2 min due to the absence of internal diffusion resistance [54].

    Gupta et al. [55] investigated the chromium removal by combining magnetic properties of iron oxide with adsorption properties of carbon nanotubes. The adsorption features of multiwall carbon nanotubes (MWCNTs) with the magnetic properties of iron oxides have been combined in a composite to produce a magnetic adsorbent as shown in Fig. 12. They performed adsorption capability of the composites in batch and fixed-bed modes. They observed that the composites have demonstrated an adsorption capability superior to that of activated carbon. The results also showed that the adsorptions of Cr (III) on the composites are strongly dependent on the pH as shown in Fig. 13. They concluded that the combination of magnetic iron oxide nanoparticles with high surface area carbon nanotubes can be used as adsorbent for contaminants in water and can be subsequently controlled and removed from the medium by a simple magnetic process [55].

    Fig. 12 e SEM images of (A) oxidized MWCNTs (B) MWCNTs/nano-iron oxide. Reproduced with permission from V.K. Gupta, S. Agarwal, T.A. Saleh, Chromium removal by combining the magnetic properties of Iron oxide with adsorption properties of carbon nanotubes, Water Res. 45(6) (2011) 2207–2212, Elsevier Ltd.

    Fig. 13 The effect of pH on the amount of Cr (III) adsorbed on the MWCNT/nano-iron oxide. Reproduced with permission from V.K. Gupta, S. Agarwal, T.A. Saleh, Chromium removal by combining the magnetic properties of Iron oxide with adsorption properties of carbon nanotubes, Water Res. 45(6) (2011) 2207–2212, Elsevier Ltd.

    Titanium dioxide is an attractive photocatalyst for water treatment as it is resistant to corrosion and nontoxic when ingested [56]. The basic mechanism of TiO2 photoactivation and reactive oxygen species (ROS) generation is well known [57].

    Liga et al. [58] prepared the photocatalytic silver-doped titanium dioxide nanoparticles (nAg/TiO2) and investigated their capability to inactivate bacteriophage MS2 in aqueous media. They deposited nanosized Ag onto commercial TiO2 nanopowders using a photochemical reduction method. The MS2 inactivation kinetics of nAg/TiO2 were compared to the base TiO2 material and silver ions leached from the catalyst. They found that the inactivation rate of MS2 was enhanced by more than fivefold, depending on the base TiO2 material, and the inactivation efficiency increased with increasing silver content. The increased production of hydroxyl free radicals was found to be responsible for the enhanced viral inactivation [58].

    Aluminum oxide (Al2O3) is one of the most stable inorganic materials, and thus it is used as ultra-filtration membranes [59]. In addition, it is inexpensive, nontoxic, highly abrasive, and resistant. Remarkably, the use of alumina nanoparticles in membranes has been extensively studied and reported.

    Saleh and Gupta [60] fabricated polyamide (PA) nanocomposite membrane containing alumina nanoparticles synthesized via in situ interfacial polymerization as shown in Fig. 14. In the first part of the work, aluminum oxide (Al2O3) nanoparticles with an average size of 14 nm were synthesized by an aqueous sol-gel method using precursors of aluminum nitrate and citric acid mixed solution. In the second part, the Al2O3 were used to prepare a nanoparticle-entrapped PA membrane via interfacial polymerization. SEM analysis demonstrated that nanoparticles were dispersed in the membrane and embedded in polyamide chains. They found that the existence of nanoparticles in the membrane improved the permeate flux and maintained the salt rejection. It also resulted in enhanced hydrophilicity of the membranes, as proven by the decreased water contact angle [60].

    Fig. 14 Schematic representation of polysulfone the polyamide membrane embedded with alumina nanoparticles. Reproduced with permission from T.A. Saleh, V.K. Gupta, Synthesis and characterization of alumina nano-particles polyamide membrane with enhanced flux rejection performance, Sep. Purif. Technol. 89 (2012) 245–251, Elsevier Ltd.

    Nanodendrimers

    Dendrimers are highly branched macromolecules with controlled composition and architecture and monodispersity. They consist of a central core, repeating units, and terminal functional groups [61, 62]. The nature of the internal repeating units determines the microenvironment of the interior and consequently its solubilization properties, while the external groups determine the chemical behavior of dendrimers in the external medium [62, 63].

    Reverse osmosis-assisted membranes have pore sizes in the range of 0.1–1 nm, and thus, they are very effective at retaining dissolved inorganic and organic solutes with molar mass below 1000 Da [64]. On other hand, nanofiltration membranes are very promising candidates for removing hardness and organic solutes with molar mass between 1000 and 3000 Da [64]. However, these reverse-osmosis membranes require high pressures to operate. On the other hand, UF membranes require lower pressure (200–700 kPa). Unfortunately, they are not very effective at removing dissolved organic and inorganic solute with molar mass below 3000 Da. Advances and recent developments in macromolecular chemistry, such as the invention of dendritic polymers, are providing unprecedented opportunities to develop effective UF processes for purification of water contaminated by toxic metal ions, radionuclides, organic and inorganic solutes, bacteria, and viruses. Dendritic polymers, which include random hyperbranched polymers, dendrigraft polymers, dendrons, and dendrimers, are relatively monodispersed and highly branched macromolecules with controlled composition and architecture consisting of three components: a core, interior branch cells, and terminal branch cell [63].

    Arkas et al. [65] synthesized organosilicon dendrimeric and hyperbranched networks and impregnated these networks into porous ceramic filters, which proved to be promising devices for water purification. They prepared triethoxysilyl-functionalized poly (propylene imine) (DAB32) dendrimers and poly (ethylene imine) (PEI5) hyperbranched polymers in chloroform, which were subsequently allowed to impregnate porous ceramic filters. They showed that the demonstrated filters could encapsulate toxic polycyclic aromatic compounds dissolved in water. They showed that the concentration of polycyclic aromatic compounds in water was lowered to the level of few ppb via continuous filtration of contaminated water at different flow rates. Moreover, the filters were effectively regenerated by treatment with acetonitrile. Finally, they concluded that this demonstrated dendrimer-based ceramic filter had huge potential in water purification, especially for the removal of water-soluble toxic pollutants [65].

    3 Nanostructured Membranes

    Advanced technologies for water purification are an essential part of meeting the current and future needs for water. Innovations in the development of novel technologies to desalinate water are among the most exciting and promising. Additionally, nanotechnology-derived products that reduce the concentrations of toxic compounds to subppb levels can assist in the attainment of water quality standards and health advisories. Nanotechnology for water and wastewater treatment is gaining momentum globally. The unique properties of nanomaterials and their convergence with current treatment technologies present great opportunities to revolutionize water and wastewater treatment. Advances in nanoscale science and engineering are providing unprecedented opportunities to develop more cost-effective and environmentally acceptable water purification processes. However, the application of nanotechnology to water purification is currently faced with the issue of how to design nanomaterials that are capable of collecting and preconcentrating a large number of contaminants per unit volume. Specifically, it is not clear how to interface nanoparticles with contaminants because direct addition of nanoparticles into drinking water may require extra separation steps to recover the expensive nanomaterials. Traditional membrane filters cannot be used for eliminating smaller particles like viruses, etc., from water as the size of these particles is in the range below 50 nm. In order to rectify these issues, nanostructured membranes have been extensively used and investigated in order to remove submicron particles below 100 nm.

    Krieg et al. [66] have demonstrated the fabrication and use of a supramolecular ultrafiltration membrane. The demonstrated membrane can be used in both filtration and size-selective chromatography regimes, which is shown in the Fig. 15, and allowed the filtration of nanoparticles with a cutoff of 5 nm. A thicker layer of the supramolecular material was capable of chromatographic size separation in the sub-5-nm range, which was particularly useful for quantum dots. Moreover, they showed that the supramolecular structure retains its adaptivity, allowing for facile recycling of the membrane material, as well as the retrieval of retained nanoparticles. They proved that the simple fabrication of the membrane, as well as its performance, versatility, and recyclability, represent a significant advantage over conventional membranes with similar rejection properties and performances [66].

    Fig. 15 Supramolecular filtration membrane. (A) Photograph showing the preparation of the membrane by filtration of a supramolecular solution of PP2b (5 × 1024 M) in water over a CA filter (pore size, 0.45 mm). (B) Photograph of the supramolecular membrane deposited on top of the CA support in a commercial syringe filter. (C) Cryo-SEM image of the cross-section of a 1 × 1 mm piece of the supramolecular membrane (0.65 mg PP2b cm22) on the CA support. (D) Magnified image showing the sharp border between the coarse CA and the smooth PP2b layer (dashed line). (E) High-magnification image of the supramolecular PP2b layer. Reproduced with permission from E. Krieg, H. Weissman, E. Shirman, E. Shimoni, B. Rybtchinski, A recyclable supramolecular membrane for size-selective separation of nanoparticles, Nat. Nanotechnol. 6(3) (2011) 141–146, Springer Nature.

    Asper et al. [67] demonstrated nanocellulose-based filter paper to remove xenotropic leukemia virus by size exclusion. The filter paper was composed of 100% naturally derived cellulose. The filter paper was fabricated using cellulose nanofibers derived from Cladophora algae. They obtained good retention value (LRV > 5.25) against the virus. They also showed that 100 nm latex beads and 50 nm gold nanoparticles can be filtered through the demonstrated nanocellulose based filters as shown in Fig. 16. They concluded that the demonstrated nanocellulose-based filter paper could be useful for the removal of endogenous rodent retroviruses and retrovirus-like particles during the production of recombinant proteins [67].

    Fig. 16 SEM images of 100 nm latex beads (A) and 50 nm gold nanoparticles (B) retained on the nanocellulose filter paper. Reproduced with permission from M. Asper, T. Hanrieder, A. Quellmalz, A. Mihranyan, Removal of xenotropic murine leukemia virus by nanocellulose based filter paper, Biologicals 43(6) (2015) 452–456, Elsevier Ltd.

    Quellmalz and Mihranyan [68] fabricated citric acid, crosslinked, nanocellulose-based papers for size-exclusion nanofiltration. They crosslinked the cellulose nanopapers with the citric acid in order to increase the wet strength of the paper. Their particle retention studies revealed that the fabricated paper was capable of removing tracer particles as small as 20 nm, as shown in Fig. 17. They concluded that citric acid crosslinking of nanocellulose was beneficial for developing paper-based sterile (virus) removal industrial filters [68].

    Fig. 17 A photograph of crosslinked nanocellulose paper filter (diameter 26 mm) covered with a precipitate of Au nanoparticles following the filtration. Reproduced with permission from A. Quellmalz, A. Mihranyan, Citric acid cross-linkednanocellulose-based paper for size-exclusion nanofiltration, ACS Biomat. Sci. Eng. 1(4) (2015) 271–276, American Chemical Society.

    4 Nanofiber Membranes

    Currently, adsorption is considered to be an effective method for water purification, and many polymeric and inorganic materials have been developed as adsorbents [69]. In particular, membrane adsorption processes are probably the most attractive method in wastewater decontamination [70]. The adsorption process can be limited by diffusion into the pores of common adsorbents. However, in the membrane adsorption process, convection mass transport through membrane pores may overcome this problem and allow very rapid adsorption and separation processes.

    Liang et al. [71] demonstrated a simple filtration process to decontaminate water by employing a freestanding fibrous membrane fabricated from highly uniform carbonaceous nanofibers (CNFs). They demonstrated a process which combines the excellent adsorption behavior of CNFs and the advantages of membrane filtration over conventional adsorption techniques. Their membrane filtration experiments proved that the CNF membranes could remove methylene blue (MB) efficiently at a very high flux of 1580 L m− 2 h− 1, which is 10–100 times higher than that of commercial nano- or ultrafiltration membranes with similar rejection properties. In addition, they showed that the CNF membranes are easily regenerated and remain unaltered in adsorption performance over six successive cycles of dye adsorption, desorption, and washing. The concluded that the demonstrated membrane has potential applications in water purification [71].

    Ma et al. [72] fabricated a new class of thin-film nanofibrous composite (TFNC) membranes for water purification by using ultrafine polysaccharide nanofibers (i.e., cellulose and chitin) with 5–10 nm diameters. In this case, ultrafine polysaccharide nanofibers were employed as barrier layers in the polyacrylonitrile (PAN) electrospun membrane as shown in Fig. 18. They compared the demonstrated membrane with two commercial UF membranes (PAN10 and PAN400), and they observed that the demonstrated showed tenfold higher permeation flux with above 99.5% rejection ratio when compared to the commercial membranes. The demonstrated membrane also showed high virus adsorption capacity, as verified by MS2 bacteriophage testing. This high virus adsorption capacity was due to the negatively charged surface of cellulose nanofibers which were coated onto the PAN electrospun membrane. Finally, they concluded that the impressive high-flux performance of the demonstrated membrane indicates that such ultrafine polysaccharide nanofiber-based TFNC membranes can surpass conventional membrane systems in many different water applications [72].

    Fig. 18 SEM images of (A) cross-sectioned view of a TFNC membrane with cellulose nanofiber barrier layer. (B) Top view of a TFNC membrane and (C) magnified cellulose nanofiber barrier layer of TFNC membrane. Reproduced with permission from H. Ma, C. Burger, B.S. Hsiao, B. Chu, Ultrafine polysaccharide nanofibrous membranes for water purification, Biomacromolecules 12(4) (2011) 970–976, American Chemical Society.

    5 Challenges and Limitations of Nanotechnology in Water Treatment

    Nanotechnology is actively tracked to both enhance the performance of existing technologies and develop new methodologies. Nanomaterial properties desirable for water and wastewater applications include high surface area for adsorption, high activity for (photo) catalysis, antimicrobial properties for disinfection and biofouling control, superparamagnetism for particle separation, and other unique optical and electronic properties that find use in novel treatment processes and sensors for water quality monitoring. However, there are still some limitations for nanotechnology in water purification. Among them, the dispersion and retention of nanomaterials and the sustainability of antimicrobial activity are major drawbacks. Although nanoparticles provide very high specific surface area, a primary reason for their high reactivity, aggregation in water, negates this benefit. Nanoparticles such as TiO2 aggregate severely when added to water. If nanomaterials are applied in the form of slurry, an efficient separation process downstream such as membrane filtration is needed to retain and recycle the nanomaterials. Immobilization of nanomaterials on the reactor surfaces or membrane filters eliminates the need for separation. Nanoparticles may also escape from the treatment system and enter the product water. In addition, because all nanoparticles must be retained in the treatment system, no disinfectant residual is provided. Therefore, they must be used in conjunction with a secondary disinfectant that provides residual through the distribution system. Retention of nanomaterials is critical not only because of the cost associated with loss of nanomaterials, but also, and more importantly, because of the potential impacts of nanomaterials on human health and ecosystems. In the past few years, there has been increase in the number of nanotoxicity studies, which have enabled better understanding of the effect of the nanomaterials on human health and environmental implication.

    Another major challenge is the retain ability of nanomaterials. Effective and reliable techniques are needed to anchor the nanoparticles to reactor surfaces or the selective layers of filtration membranes, or to separate and retain suspended nanoparticles in order to reduce costs associated with premature material loss and to prevent potential human health and environmental impacts. This is comprised of developing better surface coating techniques, perhaps through nanoparticle surface functionalization, minimizing membrane fouling by the nanomaterial suspension, and impregnating nanoparticles into filter packing materials, etc. Another strategy is to make use of nanomaterials with different properties allowing them to be mounted on the nanomagnetite core to create a multifunctional nanocomposite material that could be retained and recycled via magnetic separation. Future research needs to address the scalability of nanomaterial production as well as the nanomaterial-based water treatment systems. Low-cost nanomaterials should be explored for potential applications in water treatment. For example, iron-coated sands and microscale iron powder have been reported to adsorb viruses via electrostatic attraction and cause viruses to disintegrate or become noninfective [73].

    6 Summary and Conclusion

    Hygienic water is vital to human health and is a critical feedstock in various industries including electronics, pharmaceuticals, and food. The entire world is facing tough global challenges in providing clean water due to extended droughts, enhancement in population, more stringent health-basedregulations, and competing demands from variety of users. Nanomaterials play a significant role in developing new materials with high efficiency at low cost for water pollution. Nanomaterials have a number of key physicochemical properties that make them particularly attractive as separation media for water purification. On a mass basis, they have much larger surface areas than bulk particles. Nanomaterials can also be functionalized with various chemical groups to increase their affinity toward a given compound. They can also serve as high capacity/selectivity and recyclable ligands for toxic metal ions, radionuclides, and organic and inorganic solutes/anions in aqueous solutions. Nanomaterials also provide unprecedented opportunities to develop more efficient water-purification catalysts and redox active media due their large surface areas and their size and shape-dependent optical, electronic, and catalytic properties.

    Nanotechnology for water and wastewater treatment is gaining momentum globally. Most of the nanotechnologies that have been reported in the literature are still in the laboratory research stage; some have succeeded in setting pilot plants. The versatile properties of nanomaterials and their conjunction with advanced treatment techniques present great openings for modernizing water and wastewater treatment. Among the nanomaterials, three categories show the most promising candidates in large-scale application in the coming future due to their commercial availability, cost, and compatibility with the existing infrastructure. They are nanoadsorbents, nanotechnology enabled membranes, and nanophotocatalysts. All three categories have commercial products, although they have not been applied in large-scale water or wastewater treatment. The challenges faced by water/wastewater treatment nanotechnologies are important, but many of these challenges are perhaps only temporary, including technical hurdles, high cost, and potential environmental and human risk. In order to fix these issues, effective collaboration between academia, industry, government is very crucial. We have confidence that progressing nanotechnology through sensible direction, while avoiding unintentional consequences, can continuously deliver efficient solutions to the challenges that we are facing globally in wastewater treatment.

    References

    [1] Lehn J.M. Toward self-organization and complex matter. Science. 2002;295(5564):2400–2403.

    [2] Donaldson K., Stone V., Tran C.L., Kreyling W., Borm P.J.A. Nanotoxicology. Occup. Environ. Med. 2004;727–728.

    [3] Oberdörster G., Oberdörster E., Oberdörster J. Nanotoxicology: an emerging discipline evolving from studies of ultrafine particles. Environ. Health Perspect. 2005;113:823–839.

    [4] Theron J., Walker J.A., Cloete T.E. Nanotechnology and water treatment: applications and emerging opportunities. Crit. Rev. Microbiol. 2008;34(1):43–69.

    [5] Qu X., Alvarez P.J.J., Li Q. Applications of nanotechnology in water and wastewater treatment. Water Res. 2013;47(12):3931–3946.

    [6] Ngwenya N., Ncube E.J., Parsons J. Recent advances in drinking water disinfection: successes and challenges. Rev. Environ. Contam. Toxicol. 2013;222:111–170.

    [7] Van Der Bruggen B., Vandecasteele C. Removal of pollutants from surface water and groundwater by nanofiltration: overview of possible applications in the drinking water industry. Environ. Pollut. 2003;122(3):435–445.

    [8] Drnenburg E., Hintenberger H. Das Auftreten Vielatomiger Kohlenstoffmoleküle Im Hochfrequenzfunken Zwischen Graphitelektroden. Zeitschrift fur Naturforschung—Section A. J. Phys. Sci. 1959;14(8):765–767.

    [9] Kroto H.W., Heath J.R., O’Brien S.C., Curl R.F., Smalley R.E. C 60: buckminsterfullerene. Nature. 1985;318:162.

    [10] Krätschmer W., Lamb L.D., Fostiropoulos K., Huffman D.R. Solid C60: a new form of carbon. Nature. 1990;347(6291):354–358.

    [11] Dresselhaus M.S., Dresselhaus G., Eklund P.C. Fullerenes. J. Mater. Res. 1993;8:2054–2097.

    [12] Darwish A.D. Fullerenes. Annu. Rep. Prog. Chem., Sect. A: Inorg. Chem. 2008;104:360–378.

    [13] Giacalone F., Martin N. Fullerene polymers: synthesis and properties. Chem. Rev. 2006;106(12):5136–5190.

    [14] Scott L.T. Methods for the chemical synthesis of fullerenes. Angew. Chem. Int. Ed. 2004;43(38):4994–5007.

    [15] Ju-Nam Y., Lead J.R. Manufactured nanoparticles: an overview of their chemistry, interactions and potential environmental implications. Sci. Total Environ. 2008;400(1–3):396–414.

    [16] Deguchi S., Alargova R.G., Tsujii K. Stable dispersions of fullerenes, C60 and C70, in water. Prep. Charact. Langmuir. 2001;17(19):6013–6017.

    [17] Lyon D.Y., Brown D.A., Alvarez P.J.J. Implications and potential applications of bactericidal fullerene water suspensions: effect of Nc60 concentration, exposure conditions and shelf life. Water Sci. Technol. 2008;57(10):1533–1538.

    [18] Chae S.R., Wang S., Hendren Z.D., Wiesner M.R., Watanabe Y., Gunsch C.K. Effects of fullerene nanoparticles on Escherichia coli K12 respiratory activity in aqueous suspension and potential use for membrane biofouling control. J. Membr. Sci. 2009;329(1–2):68–74.

    [19] Kroto H.W., Heath J.R., O’Brien S.C., Curl R.F., Smalley R.E. C60: buckminsterfullerene. Nature. 1985;318(6042):162–163.

    [20] Iijima S. Helical microtubules of graphitic carbon. Nature. 1991;354(6348):56–58.

    [21] Radushkevich L. O Strukture Ugleroda, Obrazujucegosja Pri Termiceskom Razlozenii Okisi Ugleroda Na Zeleznom Kontakte. Zhurnal Fizicheskoi Khimii. 1952;26(1):88–95.

    [22] Lau A.K.-T., Hui D. The revolutionary creation of new advanced materials—carbon nanotube composites. Compos. Part B. 2002;33(4):263–277.

    [23] Ebbesen T.W., Ajayan P.M. Large-scale synthesis of carbon nanotubes. Nature. 1992;358(6383):220–222.

    [24] Thostenson E.T., Ren Z., Chou T.-W. Advances in the science and technology of carbon nanotubes and their composites: a review. Compos. Sci. Technol. 2001;61(13):1899–1912.

    [25] Dai K., Shi L., Fang J., Zhang D., Yu B. NaCl adsorption in multi-walled carbon nanotubes. Mater. Lett. 2005;59(16):1989–1992.

    [26] Li H., Zou L. Ion-exchange membrane capacitive deionization: a new strategy for brackish water desalination. Desalination. 2011;275(1–3):62–66.

    [27] Nasrabadi A.T., Foroutan M. Ion-separation and water-purification using single-walled carbon nanotube electrodes. Desalination. 2011;277(1–3):236–243.

    [28] Joseph L., Heo J., Park Y.G., Flora J.R.V., Yoon Y. Adsorption of bisphenol a and 17α-Ethinyl estradiol on single walled carbon nanotubes from seawater and brackish water. Desalination. 2011;281(1):68–74.

    [29] Wallace P.R. The band theory of graphite. Phys. Rev. 1947;71(9):622–634.

    [30] Semenoff G.W. Condensed-matter simulation of a three-dimensional anomaly. Phys. Rev. Lett. 1984;53(26):2449–2452.

    [31] Zhang Y.B., Tan Y.W., Stormer H.L., Kim P. Experimental observation of the quantum hall effect and Berry's phase in graphene. Nature. 2005;438(7065):201–204.

    [32] Li X., Cai W., An J., Kim S., Nah J., Yang D., Piner R., Velamakanni A., Jung I., Tutuc E., et al. Large-area synthesis of high-quality and uniform graphene films on copper foils. Science. 2009;324(5932):1312–1314.

    [33] Sutter P.W., Flege J.I., Sutter E.A. Epitaxial graphene on ruthenium. Nat. Mater. 2008;7(5):406–411.

    [34] Novoselov K.S., Geim A.K., Morozov S.V., Jiang D., Zhang Y., Dubonos S.V., Grigorieva I.V., Firsov A.A. Electric field effect in atomically thin carbon films. Science. 2004;306(5696):666–669.

    [35] Brodie B.C. On the atomic weight of graphite. Philos. Trans. R. Soc. Lond. 1859;149:249–259.

    [36] Staudenmaier L. Verfahren Zur Darstellung Der Graphits??Ure. Ber. Dtsch. Chem. Ges. 1898;31(2):1481–1487.

    [37] Hummers W.S., Offeman R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958;80(6):1339.

    [38] Gao W. The chemistry of graphene oxide. In: Graphene Oxide: Reduction Recipes, Spectroscopy, and Applications. Springer; 2015:61–95.

    [39] Han Y., Xu Z., Gao C. Ultrathin graphene nanofiltration membrane for water purification ultrathin graphene nanofi ltration membrane for water purifi cation. Adv. Funct. Mater. 2013;23(29):3693–3700.

    [40] Carpenter A.W., De Lannoy C.F., Wiesner M.R. Cellulose nanomaterials in water treatment technologies. Environ. Sci. Technol. 2015;49(9):5277–5287.

    [41] Gopakumar D.A., Thomas S., Grohens Y. Nanocelluloses as innovative polymers for membrane applications. In: Multifunctional Polymeric Nanocomposites Based on Cellulosic Reinforcements. Elsevier; 2016:253–275.

    [42] Barikani M., Oliaei E., Seddiqi H., Honarkar H. Preparation and application of chitin and its derivatives: a review. Iran. Polym. J. Eng. Ed. 2014;23(4):307–326.

    [43] Camci-Unal G., Pohl N.L.B. Quantitative determination of heavy metal contaminant complexation by the carbohydrate polymer chitin. J. Chem. Eng. Data. 2010;55(3):1117–1121.

    [44] Simonian A.L., Disioudi B.D., Wild J.R. An enzyme based biosensor for the direct determination of diisopropyl fluorophosphate. Anal. Chim. Acta. 1999;389(1–3):189–196.

    [45] Pradeep T. Anshup. Noble metal nanoparticles for water purification: a critical review. Thin Solid Films. 2009;517(24):6441–6478.

    [46] Liu J.P., Fullerton E., Gutfleisch O., Sellmyer D.J., eds. Nanoscale Magnetic Materials and Applications. LLC: Springer Science+ Business Media; 2009.

    [47] Afkhami A., Saber-Tehrani M., Bagheri H. Modified Maghemite nanoparticles as an efficient adsorbent for removing some cationic dyes from aqueous solution. Desalination. 2010;263(1–3):240–248.

    [48] Pan B., Qiu H., Pan B., Nie G., Xiao L., Lv L., Zhang W., Zhang Q., Zheng S. Highly efficient removal of heavy metals by polymer-supported nanosized hydrated Fe(III) oxides: behavior and XPS study. Water Res. 2010;44(3):815–824.

    [49] Huang S.H., Liao M.H., Chen D.H. Direct binding and characterization of lipase onto magnetic nanoparticles. Biotechnol. Prog. 2003;19(3):1095–1100.

    [50] Roco M., Nanotechnology C. Convergence with modern biology and medicine. Curr. Opin. Biotechnol. 2003;14(3):337–346.

    [51] Gupta A.K., Gupta M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials. 2005;26(18):3995–4021.

    [52] Boyer C., Whittaker M.R., Bulmus V., Liu J., Davis T.P. The design and utility of polymer-stabilizediron-oxide nanoparticles for nanomedicine applications. NPG Asia Materials. 2010;2(1):23–30.

    [53] Assa F., Jafarizadeh-Malmiri H., Ajamein H., Anarjan N., Vaghari H., Sayyar Z., Berenjian A. A biotechnological perspective on the application of iron oxide nanoparticles. Nano Res. 2016;9(8):2203–2225.

    [54] Mak S.-Y., Chen D.-H. Fast adsorption of methylene blue on polyacrylic acid-bound Iron oxide magnetic nanoparticles. Dyes Pigments. 2004;61(1):93–98.

    [55] Gupta V.K., Agarwal S., Saleh T.A. Chromium removal by combining the magnetic properties of Iron oxide with adsorption properties of carbon nanotubes. Water Res. 2011;45(6):2207–2212.

    [56] Hashimoto K., Irie H., Fujishima A. A historical overview and future prospects. AAPPS Bull. 2007;17(6):12–28.

    [57] Hoffmann M.R., Martin S.T., Choi W., Bahnemann D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995;95(1):69–96.

    [58] Liga M.V., Bryant E.L., Colvin V.L., Li Q. Virus inactivation by silver doped titanium dioxide nanoparticles for drinking water treatment. Water Res. 2011;45(2):535–544.

    [59] Wang B., Lee M., Li K. YSZ-reinforced alumina multi-channel capillary membranes for micro-filtration. Membranes. 2015;6(1):5.

    [60] Saleh T.A., Gupta V.K. Synthesis and characterization of alumina nano-particles polyamide membrane with enhanced flux rejection performance. Sep. Purif. Technol. 2012;89:245–251.

    [61] Bosman A.W., Janssen H.M., Meijer E.W. About dendrimers: structure, physical properties, and applications. Chem. Rev. 1999;99(7):1665–1688.

    [62] Malkoch M., Malmström E., Nyström A.M. Dendrimers: properties and applications. In: Polymer Science: A Comprehensive Reference, 10 Volume Set. 113–176. 2012;vol. 6.

    [63] Tomalia D.A., Fréchet J.M.J. Discovery of dendrimers and dendritic polymers: a brief historical perspective. J. Polym. Sci. A Polym. Chem. 2002;40(16):2719–2728.

    [64] Cheryan M. Ultrafiltration and Microfiltration Handbook. CRC Press; 1998.

    [65] Arkas M., Tsiourvas D., Paleos C.M. Organosilicon dendritic networks in porous ceramics for water purification. Chem. Mater. 2005;17(13):3439–3444.

    [66] Krieg E., Weissman H., Shirman E., Shimoni E., Rybtchinski B. A recyclable supramolecular membrane for size-selective separation of nanoparticles. Nat. Nanotechnol. 2011;6(3):141–146.

    [67] Asper M., Hanrieder T., Quellmalz A., Mihranyan A. Removal of xenotropic murine leukemia virus by nanocellulose based filter paper. Biologicals. 2015;43(6):452–456.

    [68] Quellmalz A., Mihranyan A. Citric acid cross-linkednanocellulose-based paper for size-exclusion nanofiltration. ACS Biomat. Sci. Eng. 2015;1(4):271–276.

    [69] Crittenden J.C., Rhodes Trussell R., Hand D.W., Howe K.J., Tchobanoglous G. MWH's Water Treatment: Principles and Design. John Wiley & Sons; 2012.

    [70] Vandezande P., Gevers L.E.M., Vankelecom I.F.J. Solvent resistant nanofiltration: separating on a molecular level. Chem. Soc. Rev. 2008;37(2):365–405.

    [71] Liang H., Cao X., Zhang W., Lin H., Zhou F., Chen L. Robust and highly efficient free-standing carbonaceous nanofiber membranes for water Purifi cation. Adv. Funct. Mater. 2011;21:3851–3858.

    [72] Ma H., Burger C., Hsiao B.S., Chu B. Ultrafine polysaccharide nanofibrous membranes for water purification. Biomacromolecules. 2011;12(4):970–976.

    [73] Ryan J.N., Harvey R.W., Metge D., Elimelech M., Navigato T., Pieper A.P. Field and laboratory investigations of inactivation of viruses (PRD1 and MS2) attached to Iron oxide-coated quartz sand. Environ. Sci. Technol. 2002;36(11):2403–2413.

    Further Reading

    [74] Savage N., Diallo M.S. Nanomaterials and water purification: opportunities and challenges. In J. Nanopart. Res.. 2005;7:331–342.

    [75] Dasgupta A., Rajukumar L.P., Rotella C., Lei Y., Terrones M. Covalent three-dimensional networks of graphene and carbon nanotubes: synthesis and environmental applications. Nano Today. 2017;12:116–135.

    [76] Geim A.K., Novoselov K.S. The rise of graphene. Nat. Mater.. 2007;6(3):183–191.

    [77] Das R., Ali M.E., Hamid S.B.A., Ramakrishna S., Chowdhury Z.Z. Carbon nanotube membranes for water purification: a bright future in water desalination. Desalination. 2014;336:97–109.

    [78] Isogai A. Wood nanocelluloses: fundamentals and applications as new bio-based nanomaterials. J. Wood Sci.. 2013;59(6):449–459.

    [79] Xu P., Zeng G.M., Huang D.L., Feng C.L., Hu S., Zhao M.H., Lai C., Wei Z., Huang C., Xie G.X., et al. Use of iron oxide nanomaterials in wastewater treatment: a review. Sci. Total Environ.. 2012;424:1–10.

    Chapter 2

    Introduction to Nanostructured and Nano-enhanced Polymeric Membranes: Preparation, Function, and Application for Water Purification

    Jaydevsinh M. Gohil⁎,†; Rikarani R. Choudhury†    ⁎ Advanced Polymer Design and Development Research Laboratory (SARP-APDDRL), Central Institute of Plastics Engineering & Technology (CIPET), Bengaluru, India

    † Laboratory for Advanced Research in Polymeric Materials (SARP-LARPM), Central Institute of Plastics Engineering & Technology (CIPET), Bhubaneswar, India

    Abstract

    The science and engineering of nanomaterial-based membranes is a present interest for various separation industries associated with purification of water. Removal of micropollutants, and separation and recovery of value-added materials from water are highly important from the viewpoint of water purification and reuse. In this chapter, the fundamentals of membrane technologies like microfiltration, ultrafiltration, nanofiltration, and reverse osmosis are elaborated. Key governing equations for membrane transport, and membrane preparation techniques (phase inversion, interfacial polymerization, electrospinning) are explained. In addition, preparation of nanostructured polymeric membranes from new polymers and copolymers with controlled morphology in flat-sheet and hollow-fiber configuration through self-assembly of block copolymers and phase separation are described.

    Keywords

    Nanomaterials; Nanostructured membrane; Nanoporous membrane; Nano-enhanced membrane; Water purification; Disinfection; Clarification; Nanofiltration; Reverse osmosis; Ultrafiltration; Microfiltration

    Acknowledgment

    One of the authors (R. Choudhury) acknowledges support from the Science and Engineering Research Board (SERB) under Project No. ECR/2017/000028.

    1 Introduction

    The terms water treatment and water purification are extensively used for any unit operations and processes that involve methods and processing steps for the removal of detrimental contaminants such as particulates, bacteria, minerals, organic pollutants, chemicals, pharmaceuticals, dissolved salts, etc., found in water. These contaminants may be physical, chemical, biological, or radiological in nature. There are three established water purification technologies: physical separation (settling, thermal distillation, filtration, and membrane separation), chemical processes (disinfection, coagulation, ion exchange), and biological processes (slow sand filtration). Among all these techniques, membrane processes alone contribute about 53% toward production of clean water, wastewater treatment/purification, and water recycling. Membrane technology is the science and engineering of separation, recognition, conversion, and recovery processes. The current boom in technological revolution in developing and developed nations, as well as increasing awareness of human beings regarding sustainable development, inspires industrialists, researchers, and entrepreneurs to adopt more economical, efficient, and environmentally benign methods for water purification. In this context, membranes and membrane processes such as microfiltration (MF), ultrafiltration (UF), nanofiltration (NF), and reverse osmosis (RO) have been used in many types of separation methods related to water treatment processes and for the recovery of some important chemical ingredients, solvents, and byproducts from polluted water streams, thereby helping to minimize environmental pollution by maximizing product output. In addition, thin film composite membranes (TFCMs) are dominantly being employed at the commercial scale for the removal of salt ions from brackish water and seawater, the creation of potable water, and wastewater treatment [1]. Specifically, membrane technologies are also employed in the biotechnology, bioprocessing, and petrochemicals industries for process intensification through improving chemical and any other manufacturing and processing unit operation by decreasing equipment size, energy consumption, and/or waste generation. Membrane systems excel because they are precisely tailored to the specific requirements of these separation processes.

    Membranes with nanosized pores, usually known as nanoporous membranes (NPMs) or nanostructured membranes (NSMs), and nano-enhanced membranes (NEMs) have been used to remove suspended solids, bacteria, macromolecules, viruses, colloids, organic compounds, and salt and mineral ions from water streams [2]. Thus, membrane technologies cover the entire spectrum of water clarification, disinfection, and desalination. Water clarification is pretreatment to remove turbidity-causing suspended solids and particulate matter from industrial effluents, municipal wastewater, and municipal potable water. Water disinfection refers to the process of removal, deactivation, and/or killing of pathogenic microorganisms in potable water. Additionally, the membrane process is capable of eliminating organic contaminants from water, contaminants that otherwise serve as a food source for microbial agents. Desalination is the practice of separating dissolved salts and minerals from water by means of thermal distillation, chemical treatment, or membrane processing.

    This chapter introduces the state-of-the-art in pressure-driven NSMs and NEMs such as MF, UF, NF, and RO used in water clarification, disinfection and desalination. The introduction briefly describes membranes and membrane processes, and membrane materials and types. Subsequent sections elaborate conventional polymeric membrane preparation and advanced membrane preparation using nanoscale materials, and the applications and functional attributes of NEMs.

    2 Pressure-Driven Membranes and Membrane Processes

    2.1 Definition of Membrane

    Conventionally, a membrane is described as a selective barrier-layer that permits some molecules, ions, or particles to pass through but constrains others. By nature, membranes can be biological or synthetic. Therefore, a precise and complete definition of the membrane covering all its functional features is quite challenging. In the most general sense, a biological membrane is a complex, multifunctional, and dynamically structured multicomponent system acting as an enclosing or separating barrier within or around the cell. A synthetic membrane is an interphase, which separates two phases and impedes the transport of diverse components in a definite fashion. Also, according to the International Union of Pure and Applied Chemistry (IUPAC), a membrane is a structure possessing lateral dimensions much greater than its thickness, through which mass transfer may occur under a variety of driving forces [3].

    2.2 Membrane Processes

    Water treatment through a membrane involves the removal of unwanted contaminants from water under the influence of pressure, osmotic pressure (concentration), or potential (electrical) driving forces. Fundamentally, as shown in Fig. 1, pressure-driven water treatment membranes can be classified as MF, UF, NF, and RO membranes. The classification spectrum for these water and wastewater treatment membranes relies on their pore size rating and applied pressure. The nanostructured polymer-based membrane is a porous membrane having nanoscale pores (1–100 nm) or a membrane prepared using nanomaterials/nanoparticles (NMs/NPs), or both [4]. These NSMs and membrane processes are well-defined by IUPAC as [3, 5]:

    Fig. 1 Pressure-driven membranes processes for water treatment.

    MF membrane: Pressure-driven membrane-based separation procedure; particles and dissolved macromolecules larger than 100 nm are excluded. Permeate transport across MF membranes is best described as occurring by convective flow through pores by a pore-flow/sieving mechanism. Separation of particles and dissolved components takes place by a sieving mechanism, based on a size exclusion principle. MF membranes possess symmetric or asymmetric porous structure.

    UF membrane: A separation process in which a solution containing a solute of molecular size notably greater than that of the solvent molecule is removed from the solvent by the application of a hydraulic pressure, forcing only the solvent to flow through a suitable membrane, usually having a pore size in the range of 2–50 nm. In UF, the pressure gradient governs mass transport through the convective flux within membrane pores. Therefore, permeation through these types of membranes is inversely propositional to membrane thickness and directly propositional to pore size and applied pressure.

    NF membrane: Pressure-driven membrane-based separation procedure wherein particles and dissolved macromolecules smaller than 2 nm are rejected. The extended Nernst-Planck equation has been extensively employed to explain permeates transport across NF-type membranes. Operating pressure of the NF membrane is in between that of the UF and RO processes, and sometimes, it is also referred to as a loose RO membrane.

    RO membrane: Liquid-phase

    Enjoying the preview?
    Page 1 of 1