Sunteți pe pagina 1din 110

Accepted Manuscript

Carbon nanotube membranes for water purification: Developments, challenges,


and prospects for the future

Ihsan Ullah

PII: S1383-5866(18)31243-7
DOI: https://doi.org/10.1016/j.seppur.2018.07.043
Reference: SEPPUR 14774

To appear in: Separation and Purification Technology

Please cite this article as: I. Ullah, Carbon nanotube membranes for water purification: Developments, challenges,
and prospects for the future, Separation and Purification Technology (2018), doi: https://doi.org/10.1016/j.seppur.
2018.07.043

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Carbon nanotube membranes for water purification: Developments,
challenges, and prospects for the future
Ihsanullah*

Center for Environment and Water, Research Institute, King Fahd University of Petroleum &
Minerals, Dhahran 31261, Saudi Arabia

*Corresponding author. E-mail: engr.ihsan.dir@gmail.com, ihsankhan@kfupm.edu.sa (Ihsanullah)


Tel.: +966-(13)8603868

Abstract
Carbon nanotubes (CNTs) have recently attracted considerable attention for the synthesis of

novel membranes with attractive features for water purification. This paper critically reviews the

recent progress on the synthesis and applications of carbon nanotube (CNT) based membranes in

water treatment. Various synthesis techniques for the preparation of CNT based membranes are

discussed. Furthermore, the effect of incorporating CNTs in the matrix on the membrane

properties has deliberated in detail. The key issues associated with the synthesis of CNT based

membranes for actual applications are highlighted. Finally, research directions are given to

ensure the fabrication and application of CNT membranes in a more effective manner. This

paper may provide an insight for the development of CNT based membranes for water

purification in future. With their tremendous separation performance, low biofouling potential

and ultra-high water flux, CNT membranes have the potential to be a leading technology in water

treatment, especially desalination.

Keywords: Carbon nanotube; membranes; water treatment; mixed matrix membranes; bucky-
paper; desalination; nanocomposite membranes

1
Contents
1. Introduction………………………………………………………………………..3
2. Basics of carbon nanotubes………………………………………………………..5
3. Types of CNT membranes………………………………………………………...8
4. Vertically aligned carbon nanotube (VA-CNT) membranes…………………..10
4.1. Applications of VA-CNT membranes in water treatment………………….....14
5. Bucky paper CNT membranes…………………………………………………..20
5.1. Applications of Bucky paper CNT membranes in water treatment…………...21
6. Mixed (nanocomposite) CNT Membranes………………………………………27
6.1. Synthesis techniques of mixed (nanocomposite) CNT Membranes…………...27
6.1.1. Phase inversion…………………………………………………………………27
6.1.2. Interfacial polymerization…………………………………………………….28
6.1.3. Solution mixing…………………………………………………………...…….30
6.1.4. Spray-assisted layer-by-layer…………………………………………………31
6.1.5. Polymer grafting………………………………………………………….…….32
6.1.6. In-situ polymerization…………………………………………………………..33
6.1.7. In-situ colloidal precipitation…………………………………….………..….34
6.2. Effects of CNTs addition on the nanocomposite membrane properties…………35
6.2.1. Effect on the antifouling and antimicrobial properties…………………..…35
6.2.2. Effect of CNTs on salt rejection…………………………………………..……41
6.2.3. Effect on the surface hydrophilicity and flux………………………..………46
6.2.4. Effect on the mechanical and thermal properties…………………..…….....49
6.3. Electrically conducive CNT membranes………………………………………………53
6.4. Anti-fouling mechanisms of mixed (nanocomposite) CNT membranes……………55
7. Potential of CNT membranes in desalination………………………………………82
8. Current challenges and need for further research ………..………………….……83
9. Conclusion ……………………………………………………………………………85
References…………………………………………………………………………..86

2
1. Introduction
Fresh water is not only important to human health but also serves as a crucial feedstock for

several industries. As per a report from the United Nations (UN), by 2025, two-thirds of the

world population could be under water stress conditions and 1800 million people will be facing

absolute water scarcity [1–3]. According to the alarming projections by United Nation, the

number of people that will suffer from water scarcity will be 7 billion in 60 countries by 2050

[4].

Intensive efforts are underway throughout the world to avert this looming crisis. The reuse,

recycling and recovery of water has proven to be fruitful in creating a new and reliable water

supply while not compromising public health. Membrane filtration is considered to be among the

most promising and widely used processes for the purification of wastewater, seawater and

brackish water [2,5]. Membranes are classified into different categories based on their

configurations and their molecular weight cut-off (MWCO). Membrane processes such as

microfiltration (MF), ultrafiltration (UF), nanofiltration (NF), reverse osmosis (RO),

pervaporation (PV) and membrane distillation (MD) are widely employed in water purification.

Membranes are mostly manufactured from polymeric, ceramic or hybrid materials [3]. Polymeric

membranes are usually used in water purification and desalination, owing to their high selectivity

and excellent mechanical strength. Ceramic membranes, on the other hand, are typically

employed for challenging water purification processes due to their good thermal and chemical

and stabilities. However, both these types of membranes still have limitations and there exists

sufficient space of improvement [3]. For example, the polymeric membranes are less chemically

stable and low fouling-resistant than ceramic membranes, in many water treatment

3
applications [6]. Ceramic membranes, on the other hand, are typically recommended for small-

scale processes only due to their high cost.

Recent years have observed remarkable revolutions toward application of various nanomaterials

such as zeolites, metal/metal-oxide nanoparticles, dendrimers and carbon nanomaterials for water

purification [1,6–8]. However, carbon nanotubes (CNTs) have gained considerable attention

owing to their unique mechanical, thermal, electrical and chemical properties. CNTs have been

extensively used for the elimination of numerous impurities from aqueous media [9–14].

Recently, CNTs have gained significant consideration for the synthesis of innovative membranes

for water decontamination [15–34]. The significant features that make CNTs an emerging

nanomaterial in water purification and desalination devices are their large surface area, ease of

functionalization, high aspect ratio and fast water transport [35]. CNTs can be used either as

direct filters or as a filler to improve the membrane performance. CNTs have proven to be

tremendous fillers in various membranes, due to improved antifouling behavior, strength,

disinfection, rejection and permeability. The flux through CNTs has been estimated to be 3–4

orders of magnitude faster than predicted by the Hagen–Poiseuille equation [16,27,36,37].

This paper reviews the various approaches for devising CNT membranes and evaluates their

potential applications in water treatment. Various synthesis techniques for the preparation of

CNT based membranes are discussed. Furthermore, the effect of incorporating CNTs in the

matrix on the membrane properties has deliberated in detail. The prospect of using CNT

membranes for water purification is also discussed.

4
2. Basics of carbon nanotubes
Carbon exists in many molecular forms, known as allotropes of carbon. CNTs are allotropes of

carbon that are composed of cylindrical graphite sheets rolled up in a tube-like structure [38].

CNTs that are composed of a single graphene sheet are termed as single-walled carbon

nanotubes (SWCNTs or SWNTs). Conversely, multilayers of graphene sheets are known as

multi-walled carbon nanotubes (MWCNTs or MWNTs). The structure of MWCNTs and

SWCNTs are depicted in Fig. 1 [39], while Fig. 2 represents several SWCNT structures based on

the way graphene sheets are rolled [40]. Comparison of SWCNTs and MWCNTs is presented in

Table 1.

Table 1 A Comparison between SWCNTs and MWCNTs [41–43]


Property SWCNTs MWCNTs

Graphene layer Single Multiple


Bulk synthesis Difficult Easy
Purity Poor High
Specific gravity (g/cm3) 0.8 1.8
Thermal conductivity (W/(m K)) 6000 2000
Electrical conductivity (S/cm) 102–106 103–105
Thermal stability in air (°C) >600 >600
Electron mobility (cm2/(Vs)) ~105 104–105

In 1991, Sumio Ijima at the NEC Laboratory in Tsukuba, Japan, discovered CNTs using the arc-

discharge technique. These CNTs were then characterized using a high-resolution transmission

electron microscope (HRTEM) [44]. Chemical bonding in nanotubes is composed entirely of sp2

5
bonds, which are stronger than the sp3 bonds found in alkanes and provide nanotubes with

unique strength [45]. The most common techniques for the synthesis of CNTs are laser ablation,

arc-discharge and chemical vapor deposition (CVD) [46]. CNTs have been extensively employed

for the remediation of various pollutants from aqueous solutions [9–14,47]. Recently, CNTs have

gained significant consideration for the synthesis of innovative membranes for water

decontamination [15,16,27–30].

Fig. 1. Structure representations of (a) MWCNT and (b) SWCNT. Reproduced with permission
from [39]. Copyright (2009) American Chemical Society

6
Fig. 2. Three different structures of CNTs. Reproduced with permission from [40]. Copyright
(2004) John Wiley and Sons

CNTs have emerged at the forefront of nanotechnology research and advancement. However,

there has been a great concern on the potential environmental impacts of CNTs [48]. Release of

CNTs into the environment can lead to serious environmental problems. CNTs can pose an

occupational inhalation exposure hazard as suspended particulate matter in the working

environment. Furthermore, there is a possibility of leaching of CNTs in water treatment plants,

which can have several adverse effects. This not only harms the treatment plant but can also

affect the aquatic life if CNTs enter into the environment. CNTs have been observed to induce

pulmonary inflammation and lung cellular proliferation in rats [9,49,50].

7
However, the toxicity of CNTs depends upon the type of CNTs, physical state (i.e., dispersed or

agglomerated), synthesis technique, presence of impurities, and dimensions. In general, the well-

dispersed CNTs are less toxic as compared to the agglomerated CNTs. Likewise, the soluble

CNTs did not have a substantial toxic effect on cell viability [51–53].

3. Types of CNT membranes

Carbon nanotube membranes can be classified into different categories according to the

fabrication methods, however, the two broad classes are; (1) freestanding CNT membranes, and

(2) mixed (nanocomposite) CNT membranes. Two main types of freestanding CNT membranes,

typically used in desalination and water treatment applications are vertically aligned CNT (VA-

CNT) membranes and bucky-paper membranes [54,55]. In the VA-CNT membranes, CNTs are

aligned as cylindrical pores that forces the fluid to pass only through the hollow CNT interior or

between the CNTs bundles. Bucky-paper CNT membranes, on the other hand, is random

arrangement of CNTs in an extremely large porous 3D network with high specific surface area

[1,2].

Mixed-CNT membranes have a structure analogous to that of the thin-film composite RO

membranes, where the top layer is mixed with CNT and a polymer. Representative images of

VA-CNT and mixed (nanocomposite) CNT membranes are displayed in Fig. 3 [56]. The features

of VA-CNT membranes and mixed (nanocomposite) CNT membranes are summarized in Table

2. Both types of membranes have their own pros and cons. For example, the VA-CNT

membranes have the advantage of high water flux due to the compact nanotube forest and short

nano-channel length. However, the complex fabrication techniques are a big hurdle in fabrication

of these membranes for large-scale applications. On the other hand, the mixed (nanocomposite)

8
CNT membranes has the advantages of simple synthesis procedures, however, the flux through

these membranes is not comparable with one predicted for the VA-CNT membranes.

Fig. 3. Representative images of (a) Vertically aligned (VA-CNT) membranes, and (b) mixed
(nanocomposite) CNT membranes. Reproduced with permission from [56]. Copyright
(2012) The Korean Society of Industrial and Engineering Chemistry & Elsevier B.V.

Table 2 Mixed (nanocomposite) CNT membranes versus vertically aligned (VA-CNT)


membranes [56]
Mixed (nanocomposite) CNT membranes VA-CNT membranes

CNTs are mixed with polymeric materials CNTs are aligned vertically
Simple fabrication process Complicated fabrication process
Moderately fast water flux Drastically fast water flux
Require simple operating system Special operating system may be required

9
The CNT based membranes have the potential to replace or improve the performance of RO, NF,

UF, MF and forward osmosis (FO) membranes in water treatment. The hydrophobic hollow

tubes of CNTs facilitates the fast transport of polar water molecules. This enables the CNT

membranes to replace both RO and UF membranes with little or no consumption of energy. The

three main categories of CNT membranes i.e., VACNT membrane, Bucky-paper CNT

membranes and mixed (nanocomposite) CNT membranes are discussed separately below.

4. Vertically aligned carbon nanotube membranes

Hinds’s research group proposed the aligned multiwalled carbon nanotube membranes for the

first time in 2004 [29]. CNTs were generated using CVD process and CNTs arrays are produced

with polystyrene (PS) embedded between the CNTs. However, the concept of single carbon

nanotube membranes was introduced back in 2000, when Sun et al. [57] reported the mass

transport through carbon nanotubes. The single carbon nanotubes were mounted on a

macroscopic metal wire and then embedded in epoxy matrix. The transport rates of PS probe

particles through the single-pore membrane were reported. Hummer et al. [58] reported the

conduction of water molecules through a single carbon nanotubes by using the molecular

dynamics simulations. Jopesh et al (2003) predicted the transport of certain molecules through

functionalized CNTs using molecular dynamics simulation [59]. Holt et al. (2004) fabricated

VA-CNT membranes with inorganic filler (silicon nitride, Si3N4) among the CNTs arrays and

reported that water flux was >3-fold greater than calculated from the Hagen–Poiseuille equation

[60]. It was also reported that the fluid through the CNT pores and selectivity of CNTs can be

improved by functionalization of the tip of CNTs [55,61].

10
Some other simulation studies showed that water molecules can be conducted fast thorough the

inner core of CNTs [58,62]. The water transport through CNTs depends mainly on the diameter

and uniformity of CNTs [63,64]. Water transport and salt rejection is sensitive to change in

diameter of the nanotubes [65]. The potential of size based separation though CNT membranes

was predicted by various researchers [56,66]. Specific diameter of CNTs can be used to allow

water molecules to pass though while rejects ions like sodium and chloride [56,66]. Verweij et

al. [67] suggested some new models to predict the fast flow of gas and water through CNT

membranes by using molecular dynamics simulation.

Although molecular dynamic simulation is frequently used to present the flow of various

molecules through CNTs, only few experimental studies are available on the applications of

these CNT membranes in water treatment. Majumder et al. (2005) confirmed experimentally that

the water flow rates through the aligned CNTs arrays with a PS embedded between the CNTs is

4- to 5-fold higher than the value estimated from the Hagen–Poiseuille equation [68]. Baek et al.

[69] synthesized vertically aligned carbon nanotubes using epoxy resin as filler between the

CNTs bundles as shown in Fig 4. The produced membranes have a pore density of 6.8 × 1010

pores/cm2 and 4.8 nm of pore diameter. The membrane performance in term of biofouling,

rejection and flux was compared with the typical ultrafiltration (UF) membrane. The flux

through the membrane was determined for different solvents. It was concluded that the flux

through a CNT membrane was 3 times higher than that of typical UF membranes. Polyethylene

oxide (PEO, molecular weight: 100 kDa) was used to predict the rejection rate of the CNT

membranes. The rejection of CNT membrane is comparable to the commercial UF membranes.

The membrane biofouling tendency was analyzed by using an initial concertation of 1×107

CFU/mL of a model bacterial strain (Pseudomonas aeruginosa PAO1 GFP). The CNT membrane

11
showed bacteria attachment reduction of almost 2 log scale and flux drop of approximately 15%

less as compared to UF membranes.

Fig. 4. Schematic of the manufacturing of the VA-CNT membrane; (a) transfer of VA CNTs to
the tape, (b) infiltration of epoxy into the vacant areas of VA-CNTs utilizing the cast, and (c)
fabrication of the VA-CNT membrane utilizing microtome. Reproduced with permission from
[69]. Copyright (2014) Elsevier B.V.

Yu et al. [70] presented a method for the preparation of high-density VA-CNT membranes. The

CVD method was used to prepare the CNT arrays with an average space of ∼3 nm between the

CNTs. The gas molecules were permeated through both interstitial pores and CNTs. The gas

permeation values were 450 times of those predicted for Knudsen diffusion and 4-7 orders of

magnitude higher than reported in the literature. These membranes separated a larger molecule

12
(triisopropyl orthoformate (TIPO)) from a smaller molecule (nhexane) possibly due to

preferential adsorption.

Lee et al. [71] developed a novel filter with both superoleophilic and superhydrophobic

properties synthesizing needle-like, aligned multi-walled carbon nanotubes on a stainless steel

mesh with micro-scale pores as shown in Fig. 5. The nanotube filter was able to separate diesel

and water layers. Majumder et al. [72] reported the transport properties through an aligned

multiwalled CNT membrane structure impermeable PS matrix, consisting of substantially dense

(∼109-10 tubes/cm2), with ∼7 nm pore diameter. The liquid flow through the noninteracting

hydrophobic CNT cores was found to be 1000-10,000 times faster than liquid transport from

conventional no-slip hydrodynamic predictions.

Fig. 5. (a) SEM image of the stainless steel mesh with a size of 400. (b) Optical images of the
mesh before and after the synthesis of nano-tubes. (c) Top view of the synthesized nano-tubes on
the mesh (SEM). (d) Top view of the nano-tubes with an average length of about 100 lm. (e)
Magnified, tilted view of the tip of nano-tubes. (f) Cross-sectional image of the VA-MWNT
filter. (g, h) Magnified SEM images. Reproduced with permission from [71]. Copyright
(2010) Elsevier Ltd.

13
4.1. Applications of VA-CNT membranes in water treatment
Beside the higher flow rate, the VA-CNT membranes exhibited strong antibacterial

characteristics and an excellent removal performance for various salts. Brady-Estevez et al. [73]

proposed a SWNT filter for the removal of microbial pathogens from water. It was found that

the SWNT filter exhibited high antibacterial activity for Escherichia coli K12 (E. coli). Lee et al.

[74] prepared millimeter-thick ultrafiltration membrane with extremely high water permeability.

The antibiofouling potential of the membranes was determined using Pseudomonas Aeruginosa

PA01 as a model bacterium. The CNT membranes were found to resist the biofilm formation and

impede bacterial adhesion.

Du et al. [75] reported a simple method for the synthesis of superlong vertically aligned carbon

nanotubes (SLVA-CNT) and epoxy composite membranes. The prepared membrane exhibited a

significantly high flow rate for numerous liquids including dodecane, hexane and water. A

magnetic coil was used to generate magnetic field to attract and remove iron nanoparticles to and

from the membrane and to block and open the CNT membrane by applying currents. Based on

the SEM and AFM images, the hole packing density of the CNT membrane was estimated to be

around 2.4 × 1010/cm2. Fig. 6a shows a typical optical microscopic image for the SLVA-

CNT/epoxy composite membrane [75]. Fig. 6b and c, respectively, show pores homogeneously

distributed over the whole membrane surface with an average pore size of about 10 nm. The

results of this study were also confirmed by molecular dynamics simulations [75].

14
Fig. 6. SLVA-CNT and epoxy membranes: (a) digital photograph of the 2 cm x 2 cm membrane;
(b, c) AFM and SEM images of the SLVA-CNT and epoxy membrane surface under different
magnifications. Reproduced with permission from [75]. Copyright (2011) American Chemical
Society.

Chan et al. [76] predicted through simulation studies that zwitterion functional groups at the ends

of CNTs exhibited rejection of essentially all ions and allow a high flux of water. Simulations

reveal that the ion rejection ratio is nearly 100%, when two zwitterions are attached to each end

of CNTs having diameters of about 15 Å. The increase in ion rejection for the zwitterion

functionalized CNTs is due to a combination of steric hindrance from the functional groups

partially blocking the tube ends and electrostatic repulsion between functional groups and ions,

with steric effects dominating. A typical simulation cell for single-walled CNT membranes is

presented in Fig. 7.

15
Fig. 7. View of the section for simulation cell containing a membrane composed of four CNTs
embedded between two graphene sheets with saltwater on either side of the membrane. Each end
of each tube is functionalized with two zwitterionic groups. The carbons of the CNTs and
graphene sheets are shown as cyan lines. Water molecules are shown as red and white sticks, Cl –
and Na+ ions are shown as green and blue spheres, respectively, and the atoms of the zwitterions
are shown as space filling models, cyan for C, red for O, white for H, and magenta for N.
Reproduced with permission from [76]. Copyright (2013) American Chemical Society.

Li et al. [16] reported a novel concept for the synthesis of ultra-filtration membrane with high-

flux comprising polyethersulfone (PES) and a pre-aligned MWCNT array. The membrane was

prepared through a simple drop-casting and phase inversion process. As shown in Fig. 8, the

vertically aligned CNTs were uniformly distributed inside a PES matrix [16]. Results

demonstrated that the water speed through the prepared membranes was more than 10 times

higher than that of a pure PES membrane and about 3 times higher than that of a simply mixed

CNT/PES membrane, under the same pressure load.

16
Fig 8. Illustration of possible pathways for water transportation in a CNT/polymer blend membrane
due to (1) hydrophobic effect enhanced transport, (2) nano-confinement enhanced flux, (3) ultrafast
transport through the CNT pores, and (4) direct transport through the membrane matrix. Reproduced
with permission from [16]. Copyright (2014) Royal Society of Chemistry.

Kim et al. [77] prepared VA-CNT/polymer composite films via a novel in-situ bulk

polymerization method. A VA-CNT array was infiltrated with styrene monomer with a certain

amount of polystyrene–polybutadiene (PS–PB) copolymer that acts as a plasticizer. The addition

of PS-b–PB copolymer into the matrix have improved the elongation at break of the CNT/PS

composite film, as confirmed by the micro-indentation measurements. It was found that the

water and gas permeability through the prepared CNTs/polymer composite was higher than for

other VA-CNT composite membranes.

Park et al. [78] reported the synthesis of a VA-CNT membrane by the CVD method. The

prepared VA-CNT membranes had a high water permeability but showed rapid irreversible

fouling due to hydrophobic interactions between the membrane surface and foulants. The

17
rejection capability of the VA-CNT membranes was similar to commercial polymeric UF

membranes. The rejection of nanoparticles larger than 10 nm was very high. However, the

surface modification by the graft polymerization of methacrylic acid (MA) was effective to

change the characteristics of VA-CNT and UF membranes. The contact angle of the membranes

was reduced after the introduction of carboxylic groups. The modified CNT membrane showed

an increased BSA (bovine serum albumin) rejection and a lower fouling propensity. The carbon

nanotube walls of the membrane are observed to resist biofilm formation and impede bacterial

adhesion.

Li et al. [79] developed a simple method to deposit MWCNT forest-like films from

polyacrylonitrile (PAN) and MWCNT solutions by coaxial electrospray. A MWCNT solution

was used as the shell flow to generate vertically aligned tree-like structures at mesoscopic scale

(~10-5 m) on silicon, aluminum, and fiber membrane substrates. These forest-like structures

assembled from highly conductive core-shell polymer/MWCNT particles with diameter of about

1 mm. The prepared films have potential applications in various areas such as conductive filters

and flexible electronics, owing to their high electrical conductivity and porous structure. Hinds

[80] reported that the major mechanisms of mass transport through CNTs are ionic diffusion and

gatekeeper activity. The term ‘gatekeeper’ is referred to a chemical layer only at the pore

entrance that selectively allows chemicals to pass into and through the pores of the membrane.

Srivastava et al. [81] reported the fabrication of freestanding monolithic uniform macroscopic

hollow cylindrical membranes filters via spray pyrolysis of ferrocene/benzene, as shown in Fig.

9 [81]. These filters were efficient in the separation of heavy hydrocarbon and the removal of

bacteria. Fig. 10 shows the experimental setup for the removal of bacteria using nanotube filter

[81]. The water permeability and salt removal efficiency of vertically aligned double-walled

18
carbon nanotube array membrane were lower than the values reported in the literature [82]. The

authors suggested that Hagen-Poiseuille might not be suitable to calculate the flux through

VACNT membranes.

(a) (b)

Fig. 9. (a) SEM image of the aligned tubes with radial symmetry resulting in hollow cylindrical
structure (scale 1 mm), (b) SEM image of the cylindrical macrostructure assembly showing the
wall of the bulk tube consisting of aligned MWNTs. Reproduced with permission from [81].
Copyright (2004) Springer Nature.

Fig. 10. (a) The unfiltered water containing E. coli bacteria, (b) The colonies of E. coli
bacteria (marked by arrows) grown by the culture of the polluted water, (c) The assembly for

19
the filtration experiment, (d) The water filtered through nanotube filter. Reproduced with
permission from [81]. Copyright (2004) Springer Nature.

It is worth mentioning that despite the tremendous fluid transport properties of VA-CNTs, the

alignment control of CNTs in the membrane matrix and agglomeration control of CNTs is still a

huge challenge.

5. Bucky paper CNT membranes

Bucky paper membranes are composed of randomly arranged CNTs in the form of a thin mat

held together by van der Waals interactions. Van der Waals interactions are responsible for the

strong aggregation of CNTs forming the cohesive bucky-paper structure. This type of CNT

membranes has the advantage of a high specific surface area and large porous 3D network.

Different methods such as vacuum assisted filtration, electrospinning and layer-by-layer (LBL)

deposition have been reported for the synthesis of bucky paper CNT membranes.

One of the most important steps in the preparation of bucky-paper CNT membranes is the

purification of CNTs. In general, oxidative treatment is used to remove carbonaceous

impurities. However, these purification treatments can also damage and shorten the CNTs.

However this process leads to functionalize the CNTs with carboxyl and hydroxyl groups

rendering them hydrophilic, which can be advantageous for CNT dispersion into polar solvents

such as water [61].

Dumée et al. [83] synthesized bucky paper CNT membranes by using chemically modified

CNTs by UV/ozone treatment and reacting with alkoxysilane based groups. These membranes

were reported to have a high hydrophobicity and an enhanced durability. Fig. 11 represents a

20
typical process for the manufacturing of bucky paper CNT membranes [61].

Fig. 11. (a) Process for manufacturing Bucky-papers, (b) SEM image showing the Bucky-
paper surface and (c) Bucky-paper origami aeroplane demonstrating their flexibility
mechanical robustness. Reproduced with permission from [61]. Copyright (2010) MDPI
(Basel, Switzerland)

5.1. Applications of Bucky paper CNT membranes in water treatment

Dumée et al. [84] reported the preparation of bucky paper CNT membranes and assessed their

potential and performance in direct contact membrane distillation (DCMD). The bucky paper

CNT membranes were found to have a high thermal conductivity, and they are highly porous and

highly hydrophobic. It was demonstrated that Bucky-Paper CNT membranes have an excellent

21
potential to be used for desalination in DCMD. About 99% salt was rejected, with a water flux of

∼12 kg/m2.h and water vapour partial pressure difference of 22.7 kPa.

A bucky-paper CNT membrane produced by using a polymeric support was also effectively

employed for desalination of seawater via DCMD [85]. The composite CNT membranes gave an

average salt rejection of 95% with permeabilities as high as 3.3 × 10-12 kg/(m.s.Pa) and a lifespan

of up to 39 h of continuous testing [85].

Bucky-paper CNT membranes also demonstrated strong anti-bacterial properties [73,87]. Brady-

Esétvez et al. reported that a SWNT Bucky-paper exhibited high removal of the model virus

MS2 bacteriophage (27 nm diameter) due to depth filtration and was also effective in completely

retaining E. coli cells (2 μm size) due to size exclusion. Moreover, the Bucky-paper was

effective in damaging the cell membrane and inactivate the E. coli cells [73,87].

Ihsanullah et al. [5] prepared CNT membranes using a novel approach. The binder free

membranes comprising of CNT impregnated with silver have been synthesized using a powder

metallurgy route. The silver nanoparticles were impregnated on the surface of the CNTs by a wet

impregnation method. CNTs impregnated with silver particles were uniaxially pressed at

200 MPa in a tool steel die with 27-mm diameter. Compaction yielded disks of φ 27 mm × 3 mm

(thickness) containing 1, 10 or 20 wt.% silver by mass. The disks were then sintered in a

horizontal tube furnace with a programmable temperature controller. Sintering was performed at

a temperature of 800 °C for 5 h with a heating rate of 10 °C/min. A schematic overview of the

membrane synthesis process is presented in Fig. 12. The affinity of membranes to biofouling was

studied using Escherichia coli bacteria. The membranes exhibited excellent anti-bacterial
22
characteristics. The same authors also reported the synthesis of iron oxide impregnated CNT

membranes with excellent antifouling potential for sodium alginate (SA) [3,88]. The aluminum

oxide CNT membrane prepared via a similar route were efficient in removing cadmium ions

from aqueous solution in a continuous flow system [89].

Fig. 12. Schematic of silver doped CNTs membrane synthesis process. Reproduced with
permission from [5]. Copyright (2015) Elsevier B.V.

Table 3 highlights the synthesis techniques, applications and performance of various VA-CNT
and Bucky paper CNT membranes in water treatment.

23
Table 3 Synthesis techniques, applications and performance of various VA-CNT and Bucky paper CNT membrane

Membrane Membrane Material Synthesis Application Membrane performance Reference


Type technique

VA-CNT CNT/PES CVD UF  The water transportation speed for the membrane contains [16]
membrane vertically aligned CNTs (∼ 100 L/m2.h at 60 Psi) was about 3
times higher than the simply mixed CNT/PES membrane with
random orientation and more than 10 times higher than the pure
PES membrane under the same pressure load.
VA-CNT CNT/PS CVD -  Transport of two different sized molecules (ruthenium bipyridine [68]
membrane [Ru-(bipy)32+] and methyl viologen [MV2+]).
 The flux of Ru-(bipy)32+ and MV2+ molecules through CNT-dye
membrane was 9.57 (±0.91) and 21.05 (±2.32) nmol/h,
respectively.
VA-CNT CNT/PS/Epoxy resin CVD UF  The VA CNT membrane showed water flux that was about three [69]
membrane times faster (1100±130 L/m2.h.bar) than the commercial UF
membrane (477±60 L/m2.h.bar).
 The VA CNT membrane showed better biofouling resistance
with approximately 15% less permeate flux reduction and 2 log
less bacterial attachment than the UF membrane.
VA-CNT CNT/Stainless steel CVD -  The nanotube filter was able to separate diesel and water layers. [71]
membrane mesh  The flux for diesel was 4692 kg/m2.h.Pa at 400 Pa and
8415 kg/m2. h. Pa at 800 Pa while the flux for water was zero.
 The flux for water was 85.6 kg/m2. h. Pa which could be obtained
at a significantly higher pressure (1820 Pa).
VA-CNT CNT/PS CVD -  Water flow through the noninteracting hydrophobic CNT cores [72]
membrane was ∼1000-10,000 times faster than predictions of liquid
transport from conventional no-slip hydrodynamic predictions.
SWCNT filter CNT/Poly-Vinyl Di- Vacuum MF  The water permeability was 13,800 L/m2.h.bar and 6500 [73]
Fluoride (PVDF) filtration L/m2.h.bar at SWNT loading of 0.3 mg/cm2 and 0.8 mg/cm2,
respectively.
 Results indicated that only 6% of E. coli cells remained

24
metabolically active after retention by the SWNT filter, in
comparison to over 70% metabolically active cells on the control
filter.
VA-CNT CNT/Polytetrafluoroe CVD UF  The water permeability through carbon nanotubes millimetre- [74]
membrane thylene/Si substrate thick ultrafiltration membrane was 30,000 L/m2.h.bar, compared
with the best water permeability of 2,400 L/m2.h.bar reported for
carbon nanotube membranes.
 The carbon nanotube walls of the membrane were observed to
resist the biofilm formation and impede bacterial adhesion.
VA-CNT Superlong vertically CVD -  Water flow rates for the noncompressed (2 cm × 2 cm) and half- [75]
membrane aligned carbon compressed (2 cm × 1 cm) SLVA-CNT/epoxy composite
nanotubes (SLVA- membranes were found to be 6.75 × 10-2 and 1.21 × 10-1
CNT)/Epoxy mL/cm2.min, respectively.
composite  Significantly high flow rate for numerous other liquids including
dodecane and hexane.
VA-CNT CNT/Syrene In-situ bulk -  CNTs/polymer composite membranes showed high gas (∼3-16 [77]
membrane monomer/ polymerization mm3.µm/cm2.s.atm) and water permeability (∼8
Polystyrene– mm3.µm/cm2.s.atm) comparable to the other VACNT composite
polybutadiene (PS– membranes.
PB) copolymer  The calculated elastic modulus of the CNT/PS/PS-b–PB
membrane ranges between 470 and 980 MPa.
VA-CNT CNT/Fe/Al2O3/Si CVD UF  The pure water permeabilities for the VA-CNT membrane and [78]
membrane substrate UF membrane were 1000 ± 100 L/m2.h.bar and
2
400 ± 10 L/m .h.bar, respectively.
 The BSA rejection increased from 71% to 90% (at pH 7.4)
because of the modification by graft-polymerization of
methacrylic acid (MA).

VA-CNT CNT/PAN Coaxial Conductive  The fabricated MWCNT films possess a porous structure and a [79]
membrane electrospray filters high electrical conductivity.
Radially CNT Spray -  The values for the Young’s modulus of the CNT macrotube was [81]
aligned CNT pyrolysis ∼50 MPa and the tensile strength was ∼2.2 MPa.
membrane  The CNT macrotube exhibited efficient separation of heavy
hydrocarbon and removal of bacteria.
VA double- CNT/Silicon wafer CVD -  The water flux through the membrane was 1.31 × 10-3 - 6.57 × [82]
walled CNT substrate 10-2 L/cm2.day.MPa.
array  The rejection of salt (NaCl) was 41.4 %.
membrane

25
Bucky paper CNT CVD followed DCMD  Water vapour flux across the membranes in DCMD after ozone [83]
CNT by UV/ozone treatment (1hr) was 4.5 ± 0.1 × 1012 kg/m.s.Pa.
membrane treatment of  The salt rejection was >95%.
CNTs and
reacting with
alkoxysilane
based groups
Bucky paper CNT Vacuum DCMD  The membranes exhibited contact angle of 113°, porosity (90%), [84]
CNT filtration and thermal conductivity of 2.7 kW/m2.h.
membrane  The membranes showed 99% salt rejection and a flux rate of
∼12 kg/m2.h at a water vapour partial pressure difference of
22.7 kPa in DCMD setup.
Bucky paper CNT/Poly-propylene Filtration DCMD  The best composite CNT membranes gave permeabilities as high [85]
CNT (PP)/Poly-ether- as 3.3 × 10-12 kg/(m.s.Pa) with an average salt rejection of 95%
membrane sulfone (PES)/Poly- and lifespan of up to 39 h of continuous testing in DCMD setup.
Styrene (PS)/PVDF

26
6. Mixed (nanocomposite) CNT Membranes
The mixed (nanocomposite) CNT membrane has the main aim to improve the performance of

existing membranes (mainly polymeric) with the addition of CNTs. CNTs offer several attractive

features including surface hydrophilicity, thermal and mechanical stability, antimicrobial and

antifouling properties and improved salt rejection capability.

CNTs tends to agglomerate due to van der Waals interactions, which limits their applications in

some cases [87,90]. However, acid treated CNTs showed improved dispersion in the polymer

matrix as compared to pristine CNTs [91]. CNTs are incorporated in a polymer matrix in several

ways. The most common techniques are phase inversion [30,91–100], interfacial polymerization

[8,101–112], layer-by-layer [113,114] and solution-casting [115,116].

6.1.Synthesis techniques of Mixed (nanocomposite) CNT Membranes


Several synthesis techniques were employed to synthesize the mixed-CNT membranes. The most

common methods are phase inversion [91,94,95], interfacial polymerization [8,101], solution

mixing [117], spray-assisted layer-by-layer [114], polymer grafting [118,119], in-

situ polymerization [106,107] and in-situ colloidal precipitation [120].

6.1.1. Phase inversion

Choi et al. [91] prepared MWCNTs/PSf blend membranes via the phase inversion process, using

water as a coagulant. A homogeneous MWCNTs solution was prepared in N- methyl-2-

pyrrolidone (NMP) and blended with PSf solution. Wu et al. [93] prepared novel

MWCNTs/brominated polyphenylene oxide (BPPO) membrane via phase inversion using N-

methyl-2-pyrrolidinone (NMP) as the solvent and water as the coagulant. The BPPO was

dissolved in NMP and a given amount of the tri- ethanolamine (TEOA) and in N- methyl-2-

27
pyrrolidone (NMP) solution was added to the polymer solution. MWCNTs were then added to

the solution followed by stirring and sonication. The resultant casting solution was cast onto a

clean glass plate. Subsequently, the glass plate was horizontally immersed into deionized (DI)

water at a temperature of 30◦ C for at least 24h to remove the solvent and solidify the membrane

structure. Finally, the membranes were washed with DI water repeatedly and stored in wet

environment. Fig. 13 presents the preparation of MWNTs/BPPOqua membranes [93].

Fig. 13. Schematic of the MWCNTs/BPPOqua membrane synthesis. Reproduced with permission
from [93]. Copyright (2010) Elsevier B.V.

6.1.2. Interfacial polymerization

Kim et al. [8] synthesized polyamide (PA) RO membranes with carbon nanotubes (CNTs) via

interfacial polymerization using aqueous solutions of m-phenylenediamine (MPD) containing

functionalized CNTs and trimesoyl chloride (TMC) solutions in n-hexane. The polysulfone

support membrane was pretreated with isopropyl alcohol (IPA) to enlarge pores and was then

placed in the water bath for 3 h to stabilize the pores. A specific amount of TMC and n-hexane

was added into a round-bottom flask equipped with a magnetic stirring bar in a glovebox filled

with argon gas and the solution was stirred at room temperature. The PSf membrane was taken

28
out of the bath and air bubbles and droplets of aqueous solution formed on the membrane surface

were removed by rolling a rubber roller. The membrane was fixed on the acryl flat board with

rubber mold. The TMC solution was poured on the membrane saturated with the aqueous

solution. After that the membrane was placed in an oven at 100 °C for 5 min to induce cross-

linking as well as further polymerization and remove the excess of organic solution. The

resulting membrane was then several times washed with water.

Shen et al. [101] prepared film nanocomposite (TFN) functionalized multiwalled carbon

nanotubes (MWNTs)/poly(methyl methacrylate) (PMMA) membranes using interfacial

polymerization. A schematic overview of the synthesis technique is presented in Fig. 14 [101].

PMMA–MWNTs composites were produced by regular microemulsion polymerization. The PA

TFN membranes were produced by immersing the PSF substrates in a piperazine (PIP) solution.

A solution of PMMA–MWNTs and trimesoyl chloride (TMC) in toluene was then gently poured

onto the PIP-soaked membrane substrate. As a result, an ultra-thin PA rejection layer was formed

on the PSF substrates due to the reaction of TMC and PIP at the interface. The resultant

membranes were dried in air after pouring off the toluene solution. The fresh PA TFN membrane

was placed in an oven at 80oC in air at ambient pressure. Finally, the TFN membranes were

thoroughly rinsed and stored at 20 oC in DI water before being used.

29
Figure 14. Schematic view of the TFN membrane synthesis. Reproduced with permission from
[101]. Copyright (2013) Elsevier B.V.

6.1.3. Solution mixing

Ahmed et al. [117] prepared the antimicrobial polyvinyl-N-carbazole (PVK)/SWCNTs

nanocomposite membrane. Suspensions of PVK and SWNT were prepared by ultrasonication

30
with N-cyclohexyl-2-pyrrolidone (CHP). Later, the CHP-SWNT solution was added to CHP-

PVK solution. The mixture was centrifuged and the pellet of PVK-SWCNT was washed with DI

water and resuspended by ultrasonication to yield a 97:3 weight ratio of well- dispersed PVK-

SWCNT mixture. Nitrocellulose membrane filters were dip-coated with PVK, SWCNT and

PVK-SWCNT suspensions in DI water.

Mehwish et al. [121] used solution blending method to synthesize PVDF/Poly(styrene–

butadiene–styrene) (SBS)/Thiocyanate and silver-modified MWCNTs nanocomposite

membranes. A blend of PVDF and SBS was prepared by dissolving two polymers (1:1) in

tetrahydrofuran (THF) with stirring of 6 h at room temperature. Modified MWCNTs (0.01–0.1

wt%) were added to the as-prepared blend while stirring for 24 h. This blend solution was casted

on a glass plate, and evaporated at room temperature to prepare the membranes.

6.1.4. Spray-assisted layer-by-layer

Liu et al. [114] prepared PES/functionalized MWCNTs (F-MWCNTs) membrane by spray-

assisted layer-by-layer technique. The F-MWCNTs were added to ethanol aqueous solution and

ultrasonicated; then, a poly (sodium 4-styrenesulfonate) (PSS) aqueous solution was mixed with

MWCNTs solution to form a homogeneous PSS solution with MWCNTs content with the aid of

another ultrasonication. A poly (diallyl-dimethylammonium chloride) (PDDA) aqueous solution

was prepared by spiking the PDDA polymer into DI water.

The PES substrates were soaked in 25 oC DI water for 24 h to fully remove the wetting agent of

the membrane. The pretreated PES membrane was mounted on a holder so that only one side of

the PES membrane contacted with the spraying solution. A schematic view of the fabrication of

polyelectrolyte multilayer membrane via spray-assisted technique is shown in Fig. 15 [114].

31
Compressed air (20 psi) was used from a spray pistol. The process was repeated for n cycles, that

is, n bilayers of PSS/MWCNTs-PDDA thin film were formed on the PES membrane. Spraying

was initiated by PSS/MWCNTs on the PES substrate through hydrogen bonding, hydrogen-

hydrogen, and hydrophobic interactions, the positively charged PDDA interact with

PSS/MWCNTs layer via electrostatic and van der Waals forces.

Fig. 15. Schematic view of the fabrication of polyelectrolyte multilayer membrane via spray-
assisted technique. Reproduced with permission from [114]. Copyright (2013) Informa UK
Limited.

6.1.5. Polymer grafting

Shawky et al. [119] prepared MWCNT/aromatic PA nanocomposite membranes by a polymer

grafting process using N,N-dimethylacetamide (DMAc) as a solvent. Benzoyl peroxide (BPO)

was used as the initiator leading to the formation of free-radicals on both PA and MWCNTs.

32
For the preparation of MWCNT-PA composite membranes, MWCNTs were dispersed in DMAc

via ultrasonication. Then, benzoyl peroxide (BPO) was added with the aim to enhance the

dispersion and form a more homogeneous MWCNTs-PA mixture. The mixture was then heated

with stirring for 3 h at 80 °C followed by continuous stirring for 24 h. The solution was casted

onto a dried clean glass Petri dish to form a uniform thin film with thickness of 200 μm. The film

was immediately placed in an oven at 90 °C for 30 min to evaporate the solvent. Finally, the

Petri dish with the membrane was cooled and immersed in a DI water bath for at least 15 h at

room temperature.

6.1.6. In-situ polymerization

Lee et. al [122] reported the application of MWCNTs/polyaniline (PANI)/PES membranes by

incorporation of In-situ polymerization. In the first step, a MWCNTs/PANI complex was formed

by in-situ polymerization, while in the second step; MWCNTs/PANI/PES membrane was

synthesized by the phase inversion method. Schematic of both steps is show in in Fig. 16 [122].

In the first step, MWCNTs were dispersed in N-methyl-2- pyrrolidone (NMP) solution by

sonication. Three substances (i.e., MWCNTs, APS and aniline) were mixed in a glass vessel and

stirred at 4 oC for 48 h. MWCNTs/PANI complex was dispersed in NMP. A casting solution was

prepared by mixing the PES with NMP and then blending with the MWCNTs/PANI composite.

The resultant solution was cast on a glass plate and immersed in the pure water bath, which

allows the polymer to be transformed from a liquid to solid state by phase inversion. The

membranes were stored in a DI water bath for one day to remove residual solvents.

33
Fig. 16. Synthesis of (a) MWCNTs/PANI nanocomposite (b) fabrication procedure of
MWCNTs/PANI/PES composite membrane. Reproduced with permission from [122]. Copyright
(2016) Elsevier B.V.

6.1.7. In-situ colloidal precipitation

Ho et. al [120] reported the synthesis of an oxidized MWCNTs (OMWCNTs)/graphene oxide

(GO)/PVDF membrane via in-situ colloidal precipitation. A membrane polymer solution was

prepared by dissolving PVDF powder in N-N-dimethylacetamide (DMAc) and stirred for 4 h.

The membrane polymer solution was kept idle overnight at room temperature in order to remove

the air bubbles. The prepared membrane polymer solution was spread on a flat nonwoven

polyester membrane support mounted on a glass plate using film applicator. The glass plate with

membrane polymer solution film was then immersed into a coagulation bath and left for a day.

This will ensure the complete removal of residual solvent. Lastly, the membranes were stored in

ultrapure water. For the incorporation of the GO and OMWCNTs into the membrane matrix, the

membranes solution layers were immersed into the coagulation bath with GO and OMWCNTs

colloidal suspension [123].

34
6.2. Effects of CNTs addition on the membrane properties
6.2.1. Effect on the antifouling and antimicrobial properties

CNTs addition has also shown a remarkable enhancement in the antifouling characteristics of the

traditional polymer membranes [95,124–126]. The addition of CNTs has improved the resistance

against BSA, protein and bacteria fouling. Fig. 17a indicate results for fouling test for various

membranes PES, PES with pure CNT, PES with polycitricacid (PCA) modified MWCNTs and

PES with acrylamide (AAm) modified MWCNTs [125]. This test comprises 60 min pure water

passing though the membranes, then whey solution filtration for 90 min and subsequently after-

washing water flux test again for 60 min. No significant change in water flux was observed for

PCA-CNT after getting fouled by whey proteins [125].

In other words, PCA-CNT membrane exhibited the lowest irreversible fouling tendency and

superior ability of membrane surface to be cleaned by simple rinsing with water after getting

fouled by whey protein. A similar trend was reported by Vatanpour et. al [127] by measuring

water flux of the pristine PES and modified MWCNT/PES nanofiltration membranes after

fouling by BSA solution as presented in Fig. 17b. It is obvious that the flux recovery percentage

of the MWCNT embedded membranes was higher than that of the pristine membrane. The

pristine PES membrane exhibited poor anti-biofouling properties and showed a flux recovery

value of only 29.7%. In contrast , membrane with 0.04 wt% MWCNT content exhibited flux

recovery percentage of 87.7% [127]. The best antifouling properties of 0.04 wt% membrane can

be attributed to lower roughness of the membrane surface. The membrane with 0.2 wt%

MWCNT content showed more decline in flux recovery owing to their higher surface roughness.

It can be concluded that a membrane with higher surface roughness has low antifouling abilities,

and membrane fouling can be reduced by decreasing the surface roughness.

35
(a)

(b)

Fig. 17. (a) Initial pure water flux, whey solution flux and after-washing flux for all prepared
membranes. Reproduced with permission from [125]. Copyright (2013) Elsevier B.V., (b) Flux
versus time for functionalized MWCNT blended PES membranes at 4 bar during three steps:
water flux for 120 min, BSA solution (150 ppm, pH = 7 ± 0.1) flux for 120 min, and water flux
for 120 min after 20 min washing with distilled water. Reproduced with permission from [127].
Copyright (2011) Elsevier B.V.

Yin et. al. [98] reported the water flux profiles of MWCNTs/polysulfone (PSU) (15 and 18 wt%

of PSU) membranes during BSA fouling tests as shown in Fig. 18 [98]. The ratios of final flux to

36
their initial values are also presented as percentage. After feeding with BSA solution, water

fluxes decreased gradually with time for all membranes. All other membrane samples containing

OMWCNTs showed a higher initial water flux than their respective pristine membranes

[98]. The lower flux decline for membranes containing OMWCNTs indicated that the mixed

matrix membranes had an improved antifouling property for BSA. This could be due to the

increased hydrophilicity of the membrane surface [98].

Fig. 18. Fouling behavior of MWCNTs/PSU (a)15 wt% of PSU) (b) 18 wt% of PSU with
filtration of BSA. Reproduced with permission from [98]. Copyright (2013) Elsevier B.V.

The antifouling performance of the PSf/MWCNTs membranes, reversible and irreversible

fouling resistances, total fouling resistance and flux recovery were quantified during filtration

experiments with BSA solution by Khalid et. al. [99] as described in Fig. 19. It was observed

that a membrane with 0.5 wt.% dodecylamine functionalized MWNTs (DDA-MWNTs) loading

and pristine PSf membrane showed a total fouling membrane resistance (Rt) of 29 % and 51%,

respectively, during filtration of BSA solutions.

37
The value of Rt during filtration of BSA solutions decreased from 51% for the pristine PSf

membrane to 29% for the nanocomposite membrane, with 0.5 wt.% DDA-MWNTs loading.

However, at lower DDA-MWNTs loading of 0.1–0.25 wt.%, values of total flux losses were

approximately similar. The flux recovery ratio of all PSf/DDA-MWNTs membranes was higher

than for the pristine PSf membrane. The membrane with 0.5 wt.% DDA-MWNTs loading

displayed the lowest total flux loss (29%) with reduced irreversible resistance (17%) and the

highest flux recovery (83%). The fouling resistances of the membrane with 1.0 wt.% DDA-

MWNTs content were higher compared with the membrane of 0.5 wt.% DDA-MWNTs content.

Fig. 19. Fouling resistances and flux recovery of nanocomposite membranes (%). Reproduced
with permission from [99]. Copyright (2015) Elsevier B.V.

Vatanpour et al. [127] also reported that the flux recovery percentage of the MWCNT/PES

membrane is higher than that of the pristine membrane [127]. Fig. 20 shows the irreversible

fouling ratio (Rir), reversible fouling ratio (Rr), and total fouling ratio (Rt), which were calculated

from the water flux before and after BSA fouling [127]. It was observed that the pristine PES

38
membrane had a high irreversible fouling ratio (70%, more than 77% of total fouling) due to its

lower hydrophilicity and surface charge. A membrane with 0.2 wt% MWCNT content exhibited

the highest irreversible fouling, which can be attributed to its higher roughness. An

MWCNT/PES membrane with 0.04 wt% had the lowest irreversible fouling ratio of 12% and the

highest flux recovery value of 87.7%.

Fig. 20. Fouling resistance ratio of MWCNT blended PES membranes. Reproduced with
permission from [127]. Copyright (2011) Elsevier B.V.

Zhao et. al. [128] reported the protein adsorption capability on the surface of

PVDF/MWNTHPAE nanocomposite membranes. As observed from Fig. 21a, PVDF exhibits the

highest protein adsorption as a result of its high hydrophobicity, while the adsorbed amount of

BSA decreased dramatically with an increase of MWNT HPAE content in the casting solution

[128]. These results indicated that the hydrophilicity of membrane surface was improved after

the incorporation of MWNTHPAE. When the ―OH groups of the membrane surface stretched into

the surrounding aqueous environment, hydrogen bonds formed between hydrophilic HPAE

39
groups and water molecules, inducing the formation of hydration layer and the effect of steric

exclusion, which could consequently inhibit protein adsorption.

(a)

(b)

Fig. 21. (a) The amount of BSA adsorbed on PVDF/MWNTHPAE nanocomposite membranes.
Reproduced with permission from [128]. Copyright (2012) Elsevier B.V., (b) BSA protein
rejection of MWCNT/PES membranes. Reproduced with permission from [129]. Copyright
(2011) Elsevier B.V.

A similar trend was reported by Vatanpour et. al [130] for the BSA adsorption on PES/amine

functionalized MWCNTs mixed matrix membranes. The bare PES membrane exhibited the

40
highest protein adsorption, while the adsorbed amount of BSA declined significantly with an

increase of NH2-MWCNTs content. This behavior was due to the high hydrophobicity of the

bare PES membrane. It was confirmed that the incorporation of NH2-MWCNTs improved the

hydrophilicity of the membrane surface [130]. However, Rahimpour et. al. [129] reported that

the BSA rejection of MWCNT/PES increased slightly by the addition of amine functionalized F-

MWCNTs in the casting solution as indicated in Fig. 21b. The electric potential also influences

the membrane performance and fouling resistance [131]. Carbon nanotube–polyvinyl alcohol

composite membranes exhibit a good fouling resistance during emulsified oily wastewater

treatment under electrical assistance [132].

6.2.2. Effect of CNTs on salt rejection

The addition of CNTs has an influence not only on the water flux but it also affects the salt

rejection of the membranes. Zarrabi et. al. [107] studied the effect of CNTs on the flux and salt

rejection of NH2-MWCNT embedded thin film nanocomposite nanofiltration membranes, as

shown in Fig. 22. The water flux increased with increasing concentration of NH 2-MWCNT. The

rejection of Na2SO4 initially decreased with the addition of low amount of the NH2-MWCNT

and afterwards, increased at higher concentration of the NH2-MWCNT. This reduction in

rejection can be attributed to the “trade-off” between the salt rejection and permeability. At low

concentration of NH2-MWCNT, the flux was improved. However, at higher concentration, the

surface charge increased and the repulsion between the solute and the membrane surface is

increased (Donnan effect) and therefore, the salt rejection was improved. The maximum NaCl

and Na2SO4 rejection values were about 36.7 and 95.7%, respectively.

41
Fig. 22. Salt rejection and saline aqueous flux of the prepared NF membranes with different
amounts of NH2-MWCNT. Reproduced with permission from [107]. Copyright (2016) Elsevier
B.V.

Park et al. [113] measured the permeate flux and salt rejection of functionalized MWCNTs-

Poly(allylamine hydrochloride) (PAH)- Poly(acrylic acid) (PAA)/PSf (MWCNTPAA/PAH)n

multilayers using a lab-made RO test system. The effect of the number of MWCNT-PAA/PAH

layers on the flux and salt rejection as a function of operating time is presented in Fig. 23. All the

membranes contained 1 wt % of MWCNTs with the 10, 15, and 20 bilayers. It was observed that

as the number of layers was increased from 10 to 20, the salt rejection increased from 90.4% to

91.2% [113].

The chemical resistance of the nanocomposite (MWCNT-PAA/PAH)n multilayers to chlorine

was also studied. The membranes were immersed in 3,000 ppm sodium hypochloride (NaOCl)
42
solution for 4 h, and then the salt rejections were compared. The salt rejection of PA RO

membranes before and after treating with NaOCl solution is presented in Fig. 24 [113]. The salt

rejection of PA membrane was observed to decrease from 98.3% to 76.5% upon chlorination.

The (MWCNT-PAA/PAH)n multilayers also led to a reduction in salt rejection and the values

were affected by the number of layers. As the number of layers was increased, the percentage

reduction was decreased. A maximum reduction of by 16.2% (from 90.4% to 74.2%) was

observed by (MWCNT-PAA/PAH)n with 10 layers. It was concluded that (MWCNT-

PAA/PAH)n multilayers have a better chlorine resistance than the PA membranes. Furthermore,

an increase in the number of bilayers leads to a lower decrease in the salt rejection after

chlorination, i.e., the better chlorine resistance [113]. The same authors also reported in a

separate study [103] that the addition of MWCNTs to the conventional PA RO membrane has

increased its stability against chlorine resistance. Some other studies also suggested that adding

MWCNTs improves the chlorine resistance of PA membranes [133].

43
Fig. 23. Plot of salt rejection and permeate flux as a function of operation time for (a) (MWCNT-
PAA/PAH)10, (b) (MWCNTPAA/PAH)15, and (c) (MWCNT-PAA/PAH)20 multilayer.
Reproduced with permission from [113]. Copyright (2010) Informa UK Limited.

44
Fig. 24. Plot of salt rejection for PA membrane, (MWCNT- PAA/PAH)10, (MWCNT-
PAA/PAH)15, (MWCNT-PAA/PAH)20 multilayers before and after treatment with 3,000 ppm
NaOCl solutions. Reproduced with permission from [113]. Copyright (2010) Informa UK
Limited.

Shawky et. al. [119] reported that the addition of MWCNTs in an aromatic PA membrane

increases the NaCl rejection from 24 to 76%. Likewise, an increase in MWCNTs loading from 0

to 10 mg/g has increased the humic acid (HA) removal from 54 to 90%, as described in Fig. 25

[119].

45
Fig. 25. Removal efficiency of HA (initial concentration = 20 ppm, UV254 = 0.499) by different
MWCNT-PA nanocomposite membranes at 40 bar. Reproduced with permission from [119].
Copyright (2010) Elsevier B.V.

In another study, Park et. al [103] reported that increasing the MWCNTs loading from 0.1, 0.5,

1% to in a MWCNTs/m-phenylene diamine (MPDA) membrane decreased the NaCl rejection

from 93.4% to 92.5%. This is presumably caused by an increasing number of pores in the RO

membranes with MWCNT loadings that can allow the solution to permeate easily. When the

MWCNTs loading was increased to 5%, the salt rejection was reduced to 46%. This means that

the presence of a large amount of MWCNTs leads to the formation of large pores in the

membrane and leads to a poor salt rejection.

6.2.3. Effect on the surface hydrophilicity and flux

The surface hydrophilicity of the polymeric membranes has been reported to be influenced by

the addition of MWCNTs. The contact angle of PAN ultrafiltration membranes was observed to

46
decrease with an increase in MWCNTs loading (Fig. 26) [91]. This decrease in contact angle is

an indication of the improved surface hydrophilicity and enhanced flux through the membrane

[30,91,93,95,124].

As demonstrated in Fig. 27a, the water flux values for the PAN with various loading of

MWCNTs. PAN nanocomposite membranes PAN0, PAN0.5, PAN1 and PAN2 represent loading

of MWCNTs [91]. The water flux of PAN nanocomposite membranes was observed to increase

with increase in MWCNTs loading until 0.5 wt%. However, the flux decreased for membranes

with loadings of MWCNTs above 0.5 wt%. The water flux of membrane with of 0.5 wt%

MWCNTs was 63% higher compared to neat membranes. The flux of PAN2 membranes,

although smaller than membrane PAN0.5, was still 28% higher compared to the pure PAN

membrane. The increased flux is attributed to increased pore sizes.

The nature of CNTs in the polymeric matrix also affects the flux through the membrane.

MWCNT and functionalized (carboxylated) MWCNT (FMWCNT)/poly(phenylene sulfone)

(PPSU) blend membranes were synthesized via the phase-inversion method. In Fig 27b, the pure

water flux of three different membranes is presented with respect to pressure [30]. The addition

of 0.5 wt% MWCNT to PPSU enhances the flux from 7.9 to 46.6 L/m2. h. However, a higher

flux was observed for PPSU/FMWCNT blend membranes in comparison with PPSU/MWCNT

membranes. The higher flux of PPSU membrane with FMWCNT was due to the addition of

hydrophilic –COOH moieties in the blend membrane.

47
Fig. 26. Surface contact angles of PAN nanocomposite membrane as a function of MWCNTs loading.
Reproduced with permission from [91]. Copyright (2012) Elsevier B.V.

(a)

48
(b)

Fig. 27. (a) Water flux of PAN nanocomposite membranes as function of transmembrane
pressure ΔP. Reproduced with permission from [91]. Copyright (2012) Elsevier B.V. (b) Effect
of pressure on PPSU, PPSU/MWCNT and PPSU/FMWCNT membranes. Reproduced with
permission from [30]. Copyright (2012) Springer Nature.

6.2.4. Effect on the mechanical and thermal properties

The addition of CNTs also affects the mechanical and thermal properties of the polymeric

membranes. The tensile strength, Young's modulus and elongation at break of the various mixed

CNT nanocomposite membranes was found to be influenced after incorporating CNTs in the

matrix [18,95,118,121,134–136]. Majeed et. al. [95] reported that the tensile strength at break of

the PAN increases with MWCNTs addition, as shown in Fig. 28 (a). The tensile strength

increased over 97% as the loading of MWCNTs the increased from 0 to 2 wt.% [95]. This was

due to the good interactions of MWCNTs in the matrix and the decreased porosity of the

nanocomposite membranes with increased MWCNTs loading.

49
Likewise, an increase in N-CNT loadings from 0 to 0.04 wt.% in the PES membrane increased

the modulus from 114 to 135MPa, elongation at break (%) from 22 to 18.2 and the ultimate

tensile strength from 3.8 to 4.5 MPa [135]. Increasing the PEG-g-MWCNTs amount from 0.5

wt.% to 2.0 wt.% increased the tensile strength of MWCNTs/PES membranes from 3.51 to 5.08

[136]. Similarly, the tensile strength of the PS and PA nanocomposite membranes was increased

significantly with the MWCNTs content [118,119]. These improvements are attributed to the

excellent mechanical properties of CNTs and strong interaction (hydrogen bonding) between the

polymeric matrix and the CNTs factions.

The tensile strength of PVDF membrane was increased from 0.84 ± 0.18 MPa to 0.87 ± 0.09

with the addition of 0.2 wt.% oxidized MWCNTs and afterwards declined from 0.87 MPa to

0.66 MPa as the oxidized MWCNT concentration was further increased. However, the

elongation-at-break of hybrid membranes decreased markedly with increased of the oxidized

MWCNT content [134]. The decrease in performance might be due to the aggregation of

MWCNT at higher loadings and weak interface compatibility between hydrophilic oxidized

MWCNTs and the hydrophobic PVDF matrix.

Likewise, increasing the concentration of polyethylene glycol functionalized CNTs (PEG-CNTs)

in PSU nanocomposite membranes decreased the tensile strength and the modulus of the

membranes until a specific loading (0–0.25 wt.% PEG-CNTs); beyond this loading the

mechanical properties increased dramatically, as shown in Fig. 28 (b) [137]. The prepared

nanocomposite membranes have a much higher tensile strength in comparison with other

reported UF membranes. The tensile strength of iron oxide impregnated CNT membrane was

increased with iron oxide loading up to 20% and decreased with a further increase of iron oxide

50
beyond 20%. This decrease was due to the agglomeration of iron oxide particles in the matrix,

which leads to an increase in porosity and a reduction of the mechanical strength.

The addition of CNTs has also a significant effect on the thermal stability of polymeric

membranes. The thermo-gravimetric analysis (TGA) of the PSf membrane is presented in Fig.

29 [97]. The degradation of blend membranes occurs at higher temperatures compared to pure

PSf membranes, as the concentration of CNTs increases from 0–1 wt%. The degradation of a

CNT/PSf blended membrane with 1.0 wt% amide and oxidized MWCNTs content starts at

369 °C and 374 °C, while in PSf membrane without MWCNTs the degradation starts at 312 °C.

This behavior was due to the high heat and mechanical resistance of CNTs, which offers an

increased heat tolerance. A similar behavior of the TGA of pure PMMA and PMMA–CNTs was

reported by Shen et al. [101]. The addition of CNTs shifts the degradation temperature of

PMMA to a higher value than that of pure PMMA [101]. In conclusion, CNTs enhance the

thermal stability of PSf membranes.

51
(a)

(b)

Fig. 28. (a)Tensile strength at break as a function of MWCNTs loading (wt%). Reproduced with
permission from [91]. Copyright (2012) Elsevier B.V. (b) Mechanical properties off the PEG-
CNTs nanocomposite PSU membranes with different loading of PEG-CNTs (M1(0%), M2
(0.1%), M3(0.25%), M4(0.5%), M5(1%). Reproduced with permission from [137].
Copyright (2017) Elsevier B.V.

52
Fig. 29. TGA data composite PSf membranes with different percentages of amide functionalized
MWCNTs. Reproduced with permission from [97]. Copyright (2013) Elsevier B.V.

Table 4 highlights the types, synthesis techniques, application and performance of various mixed
matrix membranes in water treatment.

6.3. Electrically conducive CNT membranes

The electric potential also influences the membrane performance and fouling resistance

[15,131,132,138–140]. Boo et al. [138] reported that applying a high-potential alternating

current to a CNT–polymer composite film provides a self-heating membrane with enhanced

desalination performance of the membrane, as shown in Fig. 30. A thin conductive composite

layer was formed via sequential spray coating of CNTs and poly (vinyl alcohol) (PVA) on a

hydrophobic porous substrate (polytetrafluoroethylene). A high-potential alternating current

was applied to enable Joule heating. For the self-heating membrane, the surface temperature

was high because of Joule heating, leading to an increase in the thermal driving force for self-

heating membranes, which results in an enhanced desalination performance.

Ho et. al [131] fabricated electrically conductive membranes using MWCNTs and GO for

enhanced fouling mitigation using POME as a feed. It was observed that GO/MWCNTs

nanocomposite membranes exhibited a lower fouling tendency due to the stronger electrostatic

repulsion between the negatively charged colloidal particles and negatively charged membrane

surface. Tankus et al. [20] reported that hydrophobic CNT mats can be converted to an efficient

hydrophilic membrane via electrooxidation for facile oil−water separations. Carbon nanotube–

polyvinyl alcohol composite membranes have a good fouling resistance during emulsified oily

wastewater treatment under electrical assistance [132].

53
Fig. 30. Self-heating membranes for thermal desalination via CNT Joule heating [138] .
Reproduced with permission from [138]. Copyright (2018) Springer Nature.

6.4. Antifouling mechanisms of mixed (nanocomposite) CNT membranes

The incorporation of CNTs into mixed nanocomposite membranes also influences the antifouling

properties of mixed (nanocomposite) CNT membrane by altering the physical/chemical

properties of the membrane. Various reports suggest that CNT based nanocomposite membranes

showed improved resistance against BSA, protein and bacteria fouling [95,124,125].

Zhang et. al. [141] studied the fouling mechanisms of BSA on pristine membranes and

GO/OMWCNTs/PVDF membranes; this is presented in Fig. 31 [141]. It seems that, the higher

the adhesion force of the membrane−foulant, the more compact the corresponding structure's

cake layer and the lower pure water flux and antifouling performance. On the contrary, a lower

54
adhesion force of membrane−foulant will result in a looser cake layer and a higher pure water

flux and antifouling performance. Results revealed that the membrane−foulant adhesion force

should be a good indicator of pure water flux and cake layer structure of the membranes [141].

Fig. 31. Schematic illustration of antifouling mechanisms of (a) pristine membranes and (b)
modified membranes. Reproduced with permission from [141]. Copyright (2013) Elsevier B.V.

Ahmed et al. [117] prepared antimicrobial PVK/SWCNTs and reported a high bacterial

inactivation (∼80–90%) for Gram-positive and Gram-negative bacteria and virus (MS2

bacteriophage) removal efficiency (∼2.5 logs) on the filter surface. The high bacterial removal

was due to the combined effect of cell retention and inactivation by SWCNT while passing

through the membranes. Bacterial cell membrane damage was considered a possible mechanism

of cellular inactivation. The mechanism of virus removal was reported to be by depth filtration.

Lee et. al [142] reported the removal mechanism of natural organic matter (NOM) by the

MWCNTs/PANI/PES membrane. The two dominant factors describing the HA (model of NOM)

removal was adsorption capacity and size exclusion. The narrow pore size distribution of the

55
membrane would be one of the key factors for the high removal of HA. Furthermore,

electrostatic and π–π interactions between the negatively charged groups of HA and positively

charged membrane are responsible for the high removal of NOM.

The fouling mechanism HA vary with the concentrations. Rabbani et al. [143] reported that the

mechanisms for membrane fouling at 2 ppm HA and 700 ppm HA were different for

MWCNT/Nano-TiO2/PSf membrane. At 2 ppm, HA was small enough to penetrate and be

dynamically trapped in the membrane pores, while at 700 ppm the HA can form micelles which

were too large to penetrate the pores and deposited on the membrane surface. The fouling with

700 ppm HA was reversible and membrane surface can be cleaned easily by washing. The

reversible fouling of 700 ppm HA might also be due to electrostatic repulsion between the

negatively charged surface of the membrane and negative charge of the HA molecules.

Ho et. al [131] fabricated electrically conductive membranes using MWCNTs and GO for

enhanced fouling mitigation using palm oil mill effluent (POME) as a feed. It was observed that

GO/MWCNTs nanocomposite membranes exhibited a lower fouling tendency due to the

stronger electrostatic repulsion between the negatively-charged colloidal particles and the

likewise negatively-charged membrane surface. The surface roughness also affects the fouling

tendency on the membrane surface. A smooth surface is effective in repelling the foulants

deposited on the surface with the aid of an electric field. However, in case of a rough membrane

surface, the foulants are deposited and trapped in valleys and cannot be repealed easily. This

phenomenon is demonstrated in Fig. 32 [131].

56
Fig. 32. Schematic illustrations of foulant repulsion from (a) smooth and (b) rough membrane
surface. Reproduced with permission from [131]. Copyright (2018) Elsevier B.V.

57
Table 4 Properties of some of the common CNT based mixed matrix membranes

Membrane type CNT amount Synthesis CNT Application Membrane performance Reference
(wt%) technique Properties

CNT/PA 0.2 g Interfacial Diameter: 10– RO  The pure water flux of PA [8]
polymerization 20 nm membrane was increased from
~ 37 to 44 L/m2.h at 15.5 bar of
Length= 10– feed pressure with addition of
20 µm 2 wt.% MWCNTs.
 The water flux of PA-CNT
membraned decreases by only
18.40%, while the flux of PA
membrane decreased by 32.80%
after 48 h.
 Salt rejection of >95% for 2000
ppm NaCl solution.

MWCNTs/PPSU 0.5 Phase inversion - UF  The pure water flux of PPSU [30]
membrane was increased from
7.91 to 56.91 L/m2.h at 345 kPa
with addition of 0.5 wt.% F-
MWCNTs.
 The rejection percentage of
PPSU membrane for pepsin and
trypsin reduced from 97 and 90
% to ~ 90 and 84 % with
addition of 0.5 wt.% F-
MWCNTs.

Zwitterion functionalized 0-20 Interfacial Diameter = -  The flux of water was found to [76]
CNTs/PA nanocomposite polymerization 15.6 Å increase by more than a factor
membranes of 4, from 6.8 to 28.7 GFD
(gallons per square foot per
day), as the fraction of CNTs
was increased from 0 to 20 wt.
%.

MWNCTs/PSf 0–4 Phase inversion Outer UF  Membrane with 1.5 wt.% [91]
diameter: 10– MWCNTs showed the highest
20 nm pure water flux (~21 m3/m2.day
at 4 bar).
Length= 10–  Membrane with 4 wt.%
50 µm MWCNTs showed the lowest
pure water flux (~16 m3/m2.day
Purity =>95%
at 4 bar).
 The contact angle of PSf
membrane was reduced from ~72
to 55 with addition of 4 wt.%
MWCNTs.
 Membrane with 4 wt.%
MWCNTs showed 100%
rejection for 1000 ppm PEO at 1
bar.

MWCNT/PSf (C/P) 0-4 Phase inversion - UF  The amount of foulant on bare [92]
PES membranes was 63% higher
than the C/P blend membrane for
2% MWCNTs content.
 The highest waterer flux (~ 90
L/m2.h at 60 Psi) was obtained
for C/P blend membrane for 2%
MWCNTs content.

MWCNTs/ 0 -7 Phase inversion Outer UF  The pure water flux of BPPOqua [93]
BPPO/Triethanolamine diameter < membrane was increased from
(TEOA) 8nm 197 to 487 L/m2.h at 0.2 MPa
with addition of 5 wt.%
Length= 10– MWCNTs.
30 µm  The MWCNTs (5
Purity =>95% wt.%)/BPPOqua membrane
maintained 94% rejection rate to
egg albumin.

MWCNTs/PSf 4 Phase inversion Outer UF  The water flux increased from [94]
diameter: 10– 24.6 ± 12.6 to 28 ± 10.7 L/m2. h.
40 nm  The surface roughness increased
by 80%.
Length= 50µm
 The elongation to failure
Number of deceased by 73%.
walls= 7–16  No improvement in antibacterial
properties of the membrane.
Purity = 98%

MWCNTs/PAN 0.5-2 Phase inversion Length= 15– UF  The contact angle of PAN [95]
20 µm nanocomposite membrane was
reduced from ~50 to 42 with
addition of 2 wt.% MWCNTs.
 The pure water flux was
increased from ~ 41 to 53
L/m2.h at 2 bar.
 The tensile strength at break
increases over 97% (from 7.6 to
15 MPa) with addition of
2 wt.% MWCNTs.
SWCNT-/PA/PSf 10 mg Phase inversion Diameter = UF/RO  Membrane with SWCNT [96]
0.8 ±0.1 nm exhibited enhanced bacterial
cytotoxicity (60%/h) for E. coli
Purity =>90% cells.

MWCNT/PSf 0-1 Phase inversion Diameter = UF  Amide functionalized CNT/PSf [97]


110-170 nm composite membranes gave
94.2% removal for Cr(VI) and
Length = 5-9 78.2% removal for Cd(II) which
µm was just 10.2% and 9.9%,
respectively, with unblended
plain PSf membranes.
Purity =>90%

MWCNTs/PSf hallow fiber 0-1 Phase inversion Diameter = UF  The water flux of HFM15 [98]
membrane (HFM) 10-20 nm increased from 36.1±4.0 to
70.7±1.8 L/m2.h and then
Length = 5-15 decreased to 38.6±1.0 L/m2.h
µm when MWCNTs concentration
increased from 0% to 0.1% and
Purity =>95%
to 1.0%.
 Improved fouling resistance to
protein (BSA). The ratio of final
flux to initial flux for HFM15
with 0.1 % MWCNTs was higher
(63.6%) than pristine membrane
(57.1%).
 Improved surface hydrophilicity.

DDA-MWNTs/PSf 0.1–1 Phase inversion Outer UF  The pure water flux of PSf [99]
diameter = 10- membrane with 0.25% DDA-
20 nm MWNTs was ~ 12 L/m2.h at 1
bar.
Length = 10-  Total fouling membrane
30 µm resistance (Rt) during filtration of
BSA solutions decreased from
Purity > 95 % 51% for the pristine PSf
membrane to 29% for the
nanocomposite membrane with
0.5 wt.% DDA-MWNTs loading.

MWCNT/PSf 1 - 20 Phase inversion Outer MF  Pure water permeability of [100]


diameter = 10- MWCNT/PSu membranes was
40 nm 1200 L.h− 1 m− 2 bar− 1.

Purity = 98 %

MWNTs/PMMA 0- 4.00 g/L Interfacial Outer NF  The pure water flux through [101]
PMMA– polymerization diameter = 20- PMMA–MWNTs (0.67 wt%)/PA
MWNTs in 30 nm (PIP/TMC) membrane was
toluene ∼1.94×10−3 cm3.cm−2.s−1.
Length <50
 The rejection of Na2SO4 was
µm
high (99%).
Purity =>95%  Greater antifouling activity
against whey proteins.

MWCNTs/Polyester TFN 0-2.0 mg/mL Interfacial Outer NF  The pure water flux of the [102]
membrane polymerization diameter < 8 membranes (without MWNTs)
nm increases from 10.8 L/m2.h to
21.2 L/m2.h when MWNTs
Length = 10- concentration in aqueous phase is
30 µm 0.5 mg/mL.
Purity > 95%

MWCNTs/MPDA 0-1 Interfacial Diameter = 9- RO  For the organic/inorganic [103]


polymerization 12 nm nanocomposite RO membranes,
as MWCNT loadings increased,
Length = 10- the permeate flux increased
15 µm [from 11.1 L/m2.h at 0.1% (w/v)
of MWCNT to 13.6 L/m2.h at
1% (w/v) of MWCNT] but salt
rejection decreased [from 93.4%
at 0.1% (w/v) of MWCNT to
92.5% at 1% (w/v) of MWCNT].

MWCNTs/Polyester TFN 0.05 Interfacial Outer NF  The water flux through TFN [104]
membrane polymerization diameter < 8 membrane with 0.05% (w/v)
nm MWNTs) was 4.7 L/m2. h.
 Presence of MWNT in the
Length = 10- composite membrane enhance
30 µm the salt separation property.
Purity > 95%

Amine functionalized 0.01–0.1 Interfacial Outer FO  The water flux through [105]
MWCNT/PA/PSf polymerization diameter = 5 TFN membrane with 0.1 wt% F-
nm MWCNTs was 95.7 L/m2.h
nearly 160% higher than thin-
Length = 50 film composite (TFC)
µm
membrane.

Carboxy-functionalized 0–0.1 Interfacial Outer RO  Compared to PA membrane [106]


MWNTs/PA polymerization diameter = 20- without MWNTs, increasing the
40 nm amount of MWNTs in a
Length = 1-5 membrane to 0.1 wt% would
µm increase the water flux from
Purity = 95 % 14.86 to 28.05 L/m2. h.
 Enhanced antifouling and
antioxidative properties.

NH2 functionalized MWCNT/ 0-0.01 Interfacial Diameter =5- NF  The NF membrane modified [107]
PA polymerization 20 nm with 0.005 wt% NH2-MWCNT
rejected 36.71 and 95.72% of
Length = 1-10 Na2SO4 and NaCl salts,
µm respectively.
 The average permeation flux of
bare NF membrane was 48.6
and 48.1 L/m2.h for NaCl and
Na2SO4 solutions, respectively.
With addition of 0.001 wt%
NH2-MWCNT, the permeation
flux reached to about 61.7 and
60.8 L/m2. h.

Functionalized MWCNT/ 0.01-0.06 Interfacial Outer NF  For the solute/dye, brilliant blue [108]
Polypropylene (PP) or PES polymerization diameter ∼30 R (826 Da), solute rejections
nm upwards of 91% were achieved
in PES support-based
Length membrane in methanol solution;
(regular) = 10- for aqueous solution the
30 µm rejection was >96%.

Length (short)
= 0.5-2 µm

MWNT/PA 0.005 – 0.2 Interfacial Purity=95% RO  For 2000 ppm NaCl, the water [109]
polymerization permeability increases from 26
L/m2.h without MWNTs to 71
L/m2.h at the acidified MWNTs
loading of 0.1% (w/v).
 For 200 ppm purified
terephthalic acid (PTA)
solution, the pure water flux
increases from 19 L/m2.h up to
49 L/m2 .h, while PTA rejection
is all higher than 98%.
Sulfonated 0–0.02 Interfacial Outer NF/UF  TFN-0.01% membrane showed [110]
MWCNTs/Poly(piperazine polymerization diameter < 8 high water flux of
2
amide) nm 13.2 L/m .h.bar, 1.6 times more
than the pristine TFC
Length = 0.5-2 membrane, maintaining
µm reasonably high rejection of
96.8% to Na2SO4.
Purity > 95%
 TFN-0.01% membrane
exhibited better antifouling
ability to BSA with water flux
recovery ratio as high as 91.2%,
even after two “fouling-
washing” cycles, compared to
82.0% for TFC membrane.

MWCNTs/Aromatic PA 15.5 Interfacial Diameter = RO  The permeate flow (3.5 wt.% [111]
polymerization 80-100 nm NaCl solution) of MWCNT/PA
at 5MPa (1.7m3/m2. day) was
Length = 10- more than double when
20 µm compared to the pure PA
membrane (0.65m3/m2.day).

PMMA modified MWNTs/PA 0.67-2.0 g/L Interfacial Diameter = NF  Membrane with 0.67g/L [112]
polymerization 20-30 nm PMMA-MWNTs exhibited
Na2SO4 rejection above 98%
Purity > 95% and water flux up to 150%
higher than the TFC membrane.

Functionalized MWCNTs- 1 Layer-by-layer Diameter = 9- RO  The salt rejection of (MWCNT- [113]


PAH- PAA/ PSf assembly 12 nm PAA/PAH)n membranes
decreases by 16.2%, 15.2% and
Length = 10- 9.9% for 10, 15 and 20 bilayers,
15 µm respectively, while the
conventional PA membrane
exhibited 21.8% decreases in
the salt rejection.

PES/F-MWCNTs membrane 1 Spray-assisted - UF  The pure water flux of bare PES [114]
layer-by-layer membrane was reduced with
more bilayer deposition of
polyelectrolyte/MWCNTs.
 The irreversible fouling ratio of
BSA for bare PES membrane
was reduced from 49.3± 0.5 to
12.3± 2.9 after 6.5 bilayer
deposition of
polyelectrolyte/MWCNTs.

MWCNTs/PVA 1–5 Solution-casting Diameter = PV  The glass-transition temperature [115]


10-20 nm of the PVA membrane was
increased from 69.21 to 78.53oC
Length = 10- via blending with MWNTs.
50 µm  The crystallinity of the PVA
matrix decreased with
increasing MWNTs up to 5 wt
% from 41 to 36%.

Polycaprolactone modified 0-3 Solution-casting - -  The pure water flux enhanced [116]
MWCNTs (PCL- from 28 L/m2.h (the unmodified
MWCNTs)/PES membrane) to 61 L/m2.h (the
modified membrane including
3 w/v% PCL-MWCNTs).
 The Cd ions rejection (%)
increases from 8.7% to 27%
with addition of 3 w/v% PCL-
MWCNTs.
PVK/ SWCNTs 3 Solution mixing Outer -  Cytotoxic for Gram-positive and [117]
diameter = 1-2 Gram-negative bacteria (~80–
nm 90%).
 Removed MS2
Length =5-30 bacteriophage virus (~2.5 logs).
µm

Purity =>90%

PS/MWCNTs-grafted-GO 0.1-3 Polymer grafting Outer NF  PS-NH2/MWCNT-g-GO hybrid [118]


hybrids diameter = 6- showed increasing trend in
13 nm tensile strength from 45.5-50.2
MPa relative to PS-
Length = 2.5- NH2/MWCNT-g-GO 0.1-3
20 µm (38.1-43.5 MPa).
 Pure water flux (12.3
Purity > 98%
mL/cm2.min and membrane
recovery 89.6 %) for of PS-
NH2/MWCNT-g-GO 3 were
higher than other membranes
prepared.

MWCNTs/Aromatic PA 2.5-15 mg Polymer grafting Diameter = 15 RO  The addition of 15 mg/g [119]


MWCNTs/g PA nm MWCNT in membrane
polymer increased the rejection NaCl
Length = 0.5-1 and HA by factors of 3.17 and
µm
1.67 respectively, while
Purity > 95 % membrane permeability
decreased by 6.5%.
 The increase in MWCNTs
loading from 0 to 15 mg
MWCNT/g PA in the
membrane increased membrane
strength from 11.3 to 34.3 MPa
and the Young's modulus from
97.8 to 491.2 MPa.

Oxidized MWCNTs 0.001-0.1 g/L In-situ colloidal Diameter = UF  The water permeability through [120]
(OMWCNTs)/GO/PVDF GO/OMWCNTs precipitation 12-15 nm membranes with carbon
membrane with different nanomaterials concertation of
Length = 3-15 0.1, 0.001 and 0.01 g/L were
ratio (0-10) of
µm 43.99 L/m2·h·bar,
GO/OMWCNTs
52.62 L/m2·h·bar, and
Purity > 97%
43.38 L/m2·h·bar, respectively
 Membrane with carbon
nanomaterials concertation of
0.1 g/L exhibited improved
rejection of TDS, phosphorus,
hardness, COD, chlorine,
turbidity, color, and TSS with
maximum rejection percentage
of 1.51%, 6.55%, 21.79%,
75.5%, 76%, 81.94%, 86.3%,
and 100%, respectively.

SBS/Thiocyanate and silver- 0.01-0.1 Solution blending - NF  Tensile strength of PVDF/SBS- [121]
modified MWCNTs MWCNTs-SCN 0.01–1
nanocomposite series increased
from 10.2 to 13.9 MPa, while
PVDF/SBS-MWCNTs-SCN-
Ag 0.01–1 had values in the
range of 12.6–20.1 MPa.
 PVDF/SBS-MWCNTs-SCN
revealed maximum
decomposition temperature
around 550–580 °C, while
PVDF/SBS-MWCNTs-SCN-
Ag had T max of 567–599 °C.

TiO2 coated MWCNTs/PES 0.1 -1 Phase inversion Diameter: 5– NF  The pure water flux of PES [124]
induced by 20 nm membrane was increased from
immersion ~3.71 to5.66 kg/m2.h at 5 bar of
precipitation Length= 1–10 feed pressure with addition of
µm 1 wt.% TiO2 coated MWCNTs.
 The total fouling resistance of
Number of
walls= 3–15 the PES membrane was
decreased from 46.9% 21.6 %
with addition of 1 wt.% TiO2
coated MWCNTs.

Polymers (Citric acid (CA), 0.1 Phase inversion Length= 1–10 -  The water flux through PCA- [125]
acrylic acid (AA) and precipitation µm CNT membrane was highest
acrylamide (AAm)) modified (~30 kg/m2.h after 1 h)
MWCNTs/PES Number of compared to bare PES
walls= 3–15 membrane (~10 kg/m2.h after 1
h).
 The flux recovery ratio (FR)
after passing whey solution was
higher (95%) for PCA-CNT
membrane as compared to PES
(44%).
 Smooth and hydrophilic
membrane surface.

Oxidized MWCNTs / 0.04-0.4 Phase inversion Diameter = NF  The pure water flux of PES [127]
PES induced by 10-30 nm membrane was increased from ~
immersion 5.5 to 9 kg/m2.h at 4 bar with
precipitation Purity = 95 % addition of 0.2 wt.% MWCNTs.
 The rejection capability of the
membrane for the rejection of
Na2SO4 increased with the
addition of 0.04 wt.% MWCNTs.

Hyperbranched-polymer multi- 0 –2 Phase inversion Diameter = UF  The pure water flux was at [128]
walled carbon nanotubes 10-20 nm maximum when the
(MWNTHPAE)/PVDF MWNTHPAE/PVDF ratio was
Length = 30 1.5%; beyond 1.5%, the flux
µm gradually decreased.
 The flux recovery increased
Purity > 95 %
from around 82% for PVDF to
95.7% for
PVDF/MWNTHPAE-2.
 The PVDF exhibits the highest
protein adsorption (~70
2
mg/cm ), while the adsorbed
amount of BSA decreased
dramatically (~20 mg/cm2) with
the addition of 2%
MWNTHPAE content in casting
solution.

Amine functionalized 0 -2 Phase inversion Outer UF  A maximum value of pure water [129]
MWCNTs (F-MWCNTs)/PES induced by diameter = 5 flux (184 L/m2.h at 3 bar) was
immersion nm observed for membrane
precipitation. prepared with 1 wt.% F-
Length = 50 MWCNTs.
µm
 Rejection of BSA increased
from 81 to 88% with addition of
2% F-MWCNTs in PES
membrane.

Amine-functionalized 0-0.06 Phase inversion Diameter = 5- NF  Increasing the nanotube dosages [130]
MWCNTs (NH2- 20 nm from 0 wt% to 0.045 wt% led to
MWCNTs)/PES increase of pure water flux from
Length = 1-10 13.6 to 23.7 L/m2. h.
µm  The salt retention sequence for
0.045 wt% NH2-MWCNT
Number of embedded membrane was
walls= 3–15 Na2SO4 (65%)>MgSO4 (45%)>
NaCl (20%) after 180 min
filtration.

PVDF/GO/ MWCNTs 1 Phase inversion Diameter = UF  The pure water flux of modified [141]
10-50 nm membrane
(GO/OMWCNTs=5:5) was
Length = 1-30 251.73% higher compared with
µm that of pristine PVDF
membranes (116.5 L/ m2.h at
0.1 MPa).

MWCNTs/PANI/PES 0-2 In-situ Diameter = 5- UF  The MWCNTs/PANI PES [142]


membranes polymerization/ 15 nm membrane showed high
phase inversion permeability (1400 L/m2.h
Length = 1-5 (LMH)/bar), which is 30 times
µm greater than the PES membrane.
 The NOM rejection rate (80%)
Purity = 98%
for composite membrane was 4
times higher than PES
membrane.
 The membrane demonstrated
100% water flux recovery and
65% total fouling ratio after
cleaning with 0.1 M HCl/0.1 M
NaOH solution for 1 h.

MWCNTs/Nano-TiO2/ PSf 0-1 Phase inversion Outer UF  The pure water flux of PES [143]
diameter = 13 membrane was increased from
nm 10 (L/m2.h) to 210 (L/m2.h) by
addition of 1% MWCNT.
Length = 0.2-1  The rejection of HA increased
µm from 6% for pure PSF to 56%
for the PSF 0.5/0.5 (0.5% TiO2–
0.5% MWCNT) membrane.

MWCNTs/PVA 0-20 Pressure filtering Radius < 8 nm UF  Membranes formed with 20 min [139]
deposition cross-linking curing times and 20
Length = 10- w/% CNT concentration showed
30 µm electrical conductivities as high
as 3.6×103 S/m) and pure water
flux of 1440 L/m2. h at 550 kPa.
 Membranes with 5w/% CNT
concentration showed over 90%
rejection of 100 kDa PEO.

Carbon nanotube/PSf 3.2 Immersion SWCNTs UF  The water permeability was [144]
precipitation found to be same (~ 23-30
Diameter = 1- L/m2.h) in all the nanocomposite
2 nm membranes compared to pure
PSf membrane.
MWCNTs
 Membranes with impregnation of
Diameter single walled CNTs possess
(Type-1) = 7- better antibiofouling behaviour
15 nm as compared to pure PSf as well
as PSf membrane embedded with
Diameter multi walled CNTs.
(Type-2) =
110-170 nm

MWCNT/PSf 0-1 Phase inversion Diameter = UF  Enhanced thermal stability. The [145]
110-170 nm thermal degradation of
membrane starts at 374 °C and
Length = 5-9 312 °C for oxidized MWNTs and
unblended PSf membranes,
µm respectively.
 Amide functionalized CNT/PSf
composite membranes gave
94.2% removal for Cr(VI) and
78.2% removal for Cd(II) at pH
2.6.

MWCNTs/PVDF/ 0.05 Deposition/coating Outer MF  The pure water flux for [146]
Polydimethylsiloxane (PDMS) diameter = 10- membrane coated with 2 wt.%
30 nm PDMS was ~38 kg/m2.h at 4 bar.
 The flux recovery ratio for
Purity = 95 % membrane coated with 5 wt.%
PDMS was 82%.
 Rejection of Na2SO4 was ~ 80%
by membrane coated with 5 wt.%
PDMS at pH 7.

MWCNT/PSf 0–10.55 - Diameter = 8- UF  The water flux of the membrane [147]


15 nm was increased from 2.5 to ~ 5.5
L/m2. h. KPa with addition of
Length = 100 CNT with 6.94% carboxylation
µm degree (wt%).
Purity > 95 %

Functionalized carbon 0-0.5 Phase inversion Diameter = UF  The pure water flux of the blend [148]
nanotube/PSf 20-40 nm membranes increased from ~ 46
L/m2.h to 175 L/m2.h at 100 kPa
Purity = 95 % with the content of
functionalized MWNTs, up to
0.19%, and then gradually
decreased.
 CNTs content suppressed the
adsorption of protein .

Carbon nanotube/PES (C/P) 0-4 Phase inversion Length = 1-3.5 UF  The pure water flux for [149]
µm CNT/PES (C/P) composite
membrane with 0.5% CNT was
the highest (93 L/m2.h).
 The amount of BSA adsorbed at
pH 3 was reduced from ~ 210 to
~75 g/cm2 with addition of 4%
CNT.

MWCNT/PVDF 0 -2 Phase inversion Diameter = UF  The water flux of membrane [134]


10-50 nm increased to 225 L/m2.h (11
times higher than pristine
Length = 01- membrane) by adding 1 wt.%
30 µm oxidized MWCNTs.
 Contact angle of membranes
decreased from 75.8° to 54.7°
(1 wt.% oxidized MWCNTs).
 The BSA rejection increased
significantly (86.0%) when the
MWCNT content was increased
to 0.5 wt.%.
 The tensile strength of PVDF
membrane was increased from
0.84 ± 0.18 MPa to 0.87 ± 0.09
with the addition of 0.2 %
oxidized MWCNTs and
afterwards declined from
0.87 MPa to 0.66 MPa as the
oxidized MWCNT
concentration was further
increased.
PVA/MWNTs/PAN 0 -15 Electrospinning Outer UF  The PVA-MWNT /PAN TFNC [150]
technique diameter = 10- (10 wt% MWNT) membrane
combined with 30 nm showed very high-water flux
solution treatment (270.1 L/m2.h) with high
method Length = 0.5-2 rejection rate (99.5%) in
µm oil/water emulsion separation
even at very low feeding
Purity > 95 %
pressure (0.1 MPa).

PVDF/MWCNTs 1 Phase inversion Diameter = UF  The water flux of [151]


10-50 nm PVDF/MWCNTs membranes
reached a peak value of 620
Length = 1-30 L/m2.h and increased
µm approximately 114% compared
with the pure PVDF membrane.
 The bovine serum albumin
rejection (BSA) of
PVDF/MWCNTs membranes
was enhanced about 31.8%
compared with those of the pure
PVDF membranes.

F-MWCNTs/Polyetherimide 0.3 Electrospinning Diameter ~ 11 FO  The F-CNTs in the nanofibers [152]


(PEI)/PA nm increased the substrate porosity
by 18% and reduced
Length = 10 membrane's structural
µm parameter (S value) by 30%,
and significantly improved the
substrate tensile modulus by
53%.

MWNT/PVDF/PP 10 mg The CNT - -  Presence of CNTs increased [153]


dispersion in permeability of the non-polar
PVDF was solutes and decreased
injected through a permeability of the polar
porous membrane compounds.
 Membrane with CNTs helped
methanol retention.

Amino functionalized 0-2 Casting- Diameter = 5- -  Adsorption capacity (for [154]


MWCNTs/Chitosan/poly(vinyl) evaporation 20 nm Cu(II)) of the membrane
alcohol containing 2 wt.% MWCNTs
Length = 1-10
(20.1 mg/g at 40 °C) was almost
µm
twice as large as that of the
Number of plain membrane (11.1 mg/g);
walls= 3–15 however, capacity enlargement
was not significant (<3 mg/g),
when MWCNTs content raised
from 1 to 2 wt.%.
 No significant adsorption
capacity loss (∼3%) was
observed for the membrane
containing MWCNT-NH2 in
comparison to the capacity loss
of the plain membrane (∼10%)
after four successive
adsorption/regeneration cycles.

Functionalized 0–0.02 Solution blending Outer PV  The swelling degree of the [155]
MWNTs/Chitosan diameter = 20- functionalized MWNTs
40 nm incorporated chitosan
membranes in ethanol/water
Purity = 95% mixtures was 6 times that of the
pristine chitosan membrane.
 The addition of MWNTs
content in the membrane matrix
has decreased the Arrhenius
activation parameters for the
total permeation from 28.15 to
12.91 kJ/mol.

Acid modified MWNTs/ 0–5 Phase inversion Diameter = 5- UF  MWNTs at 5.0 wt.% in the [156]
Nanosilver (nAg)/PSf and interfacial 10 nm support layer and nAg particles
polymerization at 10 wt.% in the thin-film layer
Length = 10- enhanced the pure water
30 µm permeability of the n-TFN
membrane by 23% and 20%,
respectively, compared to
0 wt.% of these components in
their respective layers.

CNT/PVDF 0.5 g Phase inversion - UF  Conductive CNT-PVDF [157]


membrane has potential to
reduce energy requirements by
up to 2-fold as compared to the
unmodified UF system when
challenged with 10 ppm NOM
solutions at low ionic strength.

F-MWCNTs/PES 0-2 Phase inversion Outer UF  The UF membranes modified [158]


induced by diameter = 15- by F-MWCNT exhibited
immersion 30 nm permeate flux; flux decline,
precipitation COD removal and total phenols
rejection of about
21.2 (kg/m2 .h), 12.6%, 72.6%
and 89.5%, respectively based
on Taguchi method.

MWCNTs/PA and 5 and 10 Phase inversion Diameter = 5- MF/UF  PSU membrane with MWCNTs [159]
MWCNTs/PSf and Interfacial 10 nm showed higher permeate flux
polymerization (18.4 L/m2.h at 2.1 bar)
Length = 10- compared to membrane without
30 µm MWCNTs (16.4 L/m2.h at 2.3
bar).

PAA modified MWCNTs 0-0.1 Phase inversion Length = 1-10 NF  The PAA grafted [160]
(PAA-g-MWCNT)/PES µm nanocomposite membrane
exhibited high water flux
Number of (40 kg/m2.h at 0.4 MPa),
walls = 3-15 superior antifouling properties
and more efficient salt rejection.

F-MWCNTs/PES 0-0.5 Phase inversion - UF  The pure water flux increased [161]
from 24.28 L/m2.h to
2
53.91 L/m .h on addition of
0.5 wt.% of F-MWCNTs to
PES.
 Hydrophilic property of PES/f-
MWCNTs, identified by the
contact angle measurement, was
improved by 18.7% more than
that of neat PES membrane.
 PES/F-MWCNTs membrane
exhibited 27–30% rejection,
much higher than that of neat
PES membrane.

Nitrogen doped 0-0.5 Modified phase Outer -  The addition of 0.02 and 0.04 [135]
CNT(NCNT)/PES inversion diameter = 30- wt% N-CNTs to PES membrane
45 nm enhanced the flux from 260 to
375 and 450 L/m2.h,
respectively.
 The increase in N-CNT
loadings from 0 to 0.04 wt% in
the PES membrane increased
the modulus from 114 to 135
MPa and the UTS from 3.8 to
4.5 MPa.

F-MWCNTs/ 0-0.5 Phase inversion Diameter = 10 UF  Compared to the pristine PES [162]
Polyvinylpyrrolidone (PVP)/ nm membrane, the antifouling ability
PES of the PES membrane
Length = 12 incorporated with F-
µm MWCNT/PVP membrane is
greater, recording 81.7% flux
Purity > 98%
recovery and 80.2% total
resistance (>76% were reversible
one).
 The modified membrane was able
to reject 93.4%, 74.7%, 59.4%
and 28.5% for bovine serum
albumin (66 kDa), pepsin (34.6
kDa), trypsin (20 kDa) and (14.6
kDa), respectively.

Oxidized MWCNTs / 0.2 Phase inversion Diameter = 8- UF  Permeation flux (for oil/water [163]
15 nm emulsion) of MWCNT containing
TFNC membranes was about 10-
Length = 10- times higher than those of
50 µm commercial PAN10 (Sepro) UF
membrane, while still maintaining
Purity = 95%
a rejection ratio of ∼99.5%.

PES/Poly (citric acid)-grafted- 0.01 Phase inversion Outer UF/Hemodialysis  The incorporation of 0.1 wt.% [164]
MWCNTs (PCA-g-MWCNTs) diameter =6- of PCA-g-MWCNTs increased
membrane 13 nm the permeability of neat PES
membrane from 22.57 to
Length = 2.5– 149.67 L/m2/h/bar.
20 µm
Purity > 98%

PES mixed matrix NF 0-1 Non-solvent Diameter = 5- NF  The pure water flux was [165]
membrane/ZnO coated induced phase 20 nm maximum (16.7 kg/m2.h) for
MWCNTs inversion method. membrane with 0.5 wt% of
Length = 1-10 ZnO/MWCNTs.
µm  The Direct Red 16 rejection
performance of all the prepared
Number of
membranes was > 90%.
walls= 3–15

Mixed isotactic polypropylene 4 Melt mixing and Diameter = -  The water permeability of [166]
(i-PP) membrane/MWCNTs/ melt pressing 15-35 nm composite membrane with
functionalized PP-chain grafted 4 wt% MWCNT-g-PP
MWCNTs (MWCNT-g-PP) Length ~ 3-5 concentration increases by a
µm factor ∼ 35 compared to the
pure i-PP membranes.

PVDF/Fe2O3/ MWCNTs 0.2 In Diameter = 5- -  The inclusion of MWCNTs [167]


situ polymerization 10 nm resulted in higher permeate
fluxes as compared to bare
Length = 10- membranes.
30 µm  Batch studies revealed that 48%
of cyclohexanoic acid (CHA)
was degraded after 24 h of
membrane exposure with H2O2.
For HAs, removal with
H2O2 addition was significantly
higher than without at
53.1±4.4% and 28.1±4.1%,
respectively.

PA/MWCNTs RO membrane 0-0.01 Interfacial Diameter =20- RO  The saline solution fluxes were [168]
polymerization 30 nm reached from 20.3 to 25.9 in the
raw and 28.9 L/m2.h in the
Length = 10- oxidized MWCNTs embedded
30 µm membranes.
 Salt (NaCl) rejection did not
Purity = 95% changed remarkably by
embedding MWCNTs and was
>96% for all concentrations.
 The membranes with all
concentrations of raw or
oxidized MWCNTs had better
antifouling performance for
BSA/salt solution rather than
the unfilled membrane.

MWCNTs/PES membranes 0-3 Phase inversion Diameter < 8 UF  With PEG-g-MWCNTs from 0 [136]
method nm to 1.5 wt%, contact angle of
hybrid membranes declined
Length = 10- from 82.3o to 53.9o, water flux
30 µm enhanced from 5.18 L/m2.h to
71.26 L/m2.h, flux recovery
ratio improved from 41.6% to
94.1% and the adsorption
amount of BSA decreased from
58.96 µg/cm2 to 41.63 µg/cm2.
 The tensile strength increased
from 3.51 to 5.08 MPa with the
increase of PEG-g-MWCNTs
content from 0.5 wt% to 2.0 wt%.

PVA/Carboxylated multi- 0.01 Sequential Diameter =13- UF  The PVA–CNT–COOH modified [169]
walled carbon nanotubes deposition and 18 nm membranes demonstrated lower
(PVA–CNT–COOH) cross-linking fouling rates compared to
method Length = 3-30 standard PS-35 supports in high
µm concentrations of alginic acid
(AA) solutions as well as AA
Purity > 99% solution in synthetic wastewater.

Piperazine–PA/ MWCNTs 0.001-0.01 Non-solvent Outer NF  Water flux was maximum for [170]
nanocomposite NF membranes induced phase diameter =20- membrane with 0.005 wt. %
inversion 30 nm oxidized MWCNTs.
technique  The salt rejection performance
Length = 10– exhibited that by embedding of
30 µm the raw and oxidized
Purity = 95% MWCNTs, improvement in
rejecting of Na2SO4 salt can be
observed.

MWCNTs -titania 0.01–0.05 Solution mixing Outer RO  The highest water permeability [171]
nanotube/PSf TFN membrane diameter =16- was achieved at 1.13 L/m2.h.bar
20 nm when nanoparticles loading of
0.03% was incorporated into the
PA layer and PS support treated
at 1.5 M H2SO4 for 48 h.
 The NaCl rejection of all the
synthesized membranes was
found to be above 96%.
7. Potential of CNT membranes in desalination

Although the performance of VA-CNT membranes in desalination has not been studied

practically yet, the potential performance of these membranes in desalination can be roughly

evaluated. The major factor that affects the separation of ions is the size of the nanotubes.

Molecular simulations studies revealed that increasing the diameter of the tube has an adverse

effect on the removal of salts [56,65,172]. In general, for the VA-CNT membranes to have

desalination capacity comparable to that of an RO membrane, the MD simulation suggested that

the inner diameter of nanotubes should be ~ 0.6 nm [56]. However, with the currently available

technologies, it is not practical yet to produce nanotubes with this small diameter. It is worth

mentioning that size is not the only factor that affect the desalination performance of the

membrane. Rather, other factors such as the pH of the solution and electrostatic interactions also

play a significant rule. The salt rejection efficiency of VA-CNT membrane can also be improved

by functionalization of CNT surface with various groups. The CNT membrane has the advantage

of a low fouling potential compared to the traditional RO and NF membranes and this require

less maintenance. The inactivation of bacterial cells by CNTs make them ideal candidates for

future desalination technology. Another unique feature of CNT membrane is the low energy

consumption during the flow of molecules through the tubes. The flow of molecules through

smooth and hydrophobic inner walls of CNTs does not require a high-pressure pump.

Mixed (nanocomposite) CNT membranes and Bucky paper CNT membranes were found to have

an excellent desalination performance in DCMD in various lab scale studies

[85,99,107,168,170]. A more detailed exploration is, however, necessary to fill the gaps between

vision and reality, with the fast-technological transfer from academic to industries.

82
8. Current challenges and need for further research

The CNT membranes have an excellent potential to be the future membrane technology for water

purification. However, the synthesis of these membranes and applications in water treatment is

still at premature stage and numerous critical issues are yet to be effectively addressed.

Commercial readiness, improving manufacturing scalability, cutting down the CNTs cost, and

assessment of potential toxic effects of CNTs are the potential challenges that are on the way

forward to be overcome [173].

The synthesis of CNTs on large scale with suitable pore size and distribution is still a key

obstacle in the applications of CNTs on commercial scale. Researchers must investigate and

explore new methods for the economical production of CNTs on large scale. Another hurdle that

limits the applications of CNTs in large-scale operation is the cost of CNTs, especially

SWCNTs. Although researches are in quest of developing an economical process for the mass

production of CNTs, the current cost does not suggest the applications of CNTs on large scale.

However, it is expected that the due to increase in commercial production of CNTs, their price

will be greatly reduced in future.

Furthermore, the potential hazardous effects of CNTs on the on human health and environment

put serious questions to be answered. It is generally believed that raw CNTs are more toxic as

compared to chemical functionalized CNTs. This may be due to presence of metal catalyst in raw

CNTs. The potential toxic effects of CNTs on human health and environment need to be

investigated in detail and necessary precautions must be suggested.

Another obstacle is the difficulties in CNT growth with proper alignment in VA-CNT

membranes. The irregular alignment affects the membrane properties such as salt rejection and

83
flux. Further research is essential to effectively address this issue to enhance the membrane

performance. Effective methods to functionalize the tip of CNTs without compromising the CNT

properties must also be explored.

Besides, there is a need to standardize the terminology for assessing the performance of various

membranes reported in the literature. Currently, an accurate comparison of the membrane

performance reported in various studies is not possible due to inconsistency in membrane

dimension and experimental operating conditions [173]. Furthermore, since most of the CNT

based membranes are currently assembled on a ceramic or polymeric membrane that may

affect/weaken the CNT properties. Hence, there is a need of more focus on fabrication of

freestanding membrane to fully utilize the extraordinary features of CNTs. As most of the

current applications of CNT based membrane is limited to enhance the performance of pressure

driven membrane, extensive research is needed to explore the other possible applications of CNT

based membranes such as membrane distillation and capacitive deionization. Most of the current

studies of VA-CNT based membranes are focused on the simulation and modeling. The

experimental proof of this approach is essential to assess the CNT membrane performance at

viable scale. The mechanism of separations of various pollutants by CNT bases membranes

needs to be investigated in detail. More efforts on developing fabrication technique and scaling

up the fabrication process to pilot plant from laboratory-scale experiment are required to advance

current CNT-based membrane process.

9. Conclusion

The carbon nanotube membranes are ideal candidates for water purification in the future

membrane. Although there are still various challenges, the current research trends suggest that

84
many improvements in the synthesis techniques and applications of CNT membranes is expected

in near future. For the CNT membranes to be pioneer in membrane technology for water

purification, further research is essential to overcome the current hurdles. There is no suspicion

that CNTs-based membranes have a bright future in water purification and desalination

technology.

Acknowledgments

The author gratefully acknowledges the support provided by King Fahd University of Petroleum

and Minerals (KFUPM), Saudi Arabia. The author would also like to acknowledge the support of

Center for Environment and Water (CEW), Research Institute, at KFUPM.

85
References

[1] S. Kar, R.C. Bindal, P.K. Tewari, Carbon nanotube membranes for desalination and water
purification: Challenges and opportunities, Nano Today. 7 (2012) 385–389.
doi:10.1016/j.nantod.2012.09.002.

[2] X. Qu, P.J.J. Alvarez, Q. Li, Applications of nanotechnology in water and wastewater
treatment, Water Res. 47 (2013) 3931–3946. doi:10.1016/j.watres.2012.09.058.

[3] Ihsanullah, A.M. Al Amer, T. Laoui, A. Abbas, N. Al-Aqeeli, F. Patel, M. Khraisheh, M.


Ali, N. Hilal, Fabrication and antifouling behaviour of a carbon nanotube membrane,
Mater. Des. 89 (2016) 549–558. doi:10.1016/j.matdes.2015.10.018.

[4] P.S. Goh, A.F. Ismail, B.C. Ng, Carbon nanotubes for desalination: Performance
evaluation and current hurdles, Desalination. 308 (2013) 2–14.
doi:10.1016/j.desal.2012.07.040.

[5] Ihsanullah, T. Laoui, A.M. Al-Amer, A.B. Khalil, A. Abbas, M. Khraisheh, M.A. Atieh,
Novel anti-microbial membrane for desalination pretreatment : A silver nanoparticle-
doped carbon nanotube membrane, Desalination. 376 (2015) 82–93.

[6] M.M. Pendergast, E.M. V Hoek, A review of water treatment membrane


nanotechnologies, Environ. Sci. 4 (2011) 1946–1971. doi:10.1039/c0ee00541j.

[7] J. Kim, B. Van Der Bruggen, The use of nanoparticles in polymeric and ceramic
membrane structures: Review of manufacturing procedures and performance improvement
for water treatment, Environ. Pollut. 158 (2010) 2335–2349.
doi:10.1016/j.envpol.2010.03.024.

[8] H.J. Kim, K. Choi, Y. Baek, D. Kim, J. Shim, J. Yoon, J. Lee, High-Performance reverse
osmosis CNT/polyamide nanocomposite membrane by controlled interfacial interactions,
ACS Appl. Mater. Interfaces. 6 (2014) 2819−2829.

[9] Ihsanullah, A. Abbas, A.M. Al-Amer, T. Laoui, M.J. Al-Marri, M.S. Nasser, M.
86
Khraisheh, M.A. Atieh, Heavy metal removal from aqueous solution by advanced carbon
nanotubes: Critical review of adsorption applications, Sep. Purif. Technol. 157 (2016)
141–161. doi:10.1016/j.seppur.2015.11.039.

[10] Ihsanullah, F.A. Al-Khaldi, B. Abusharkh, M. Khaled, M.A. Atieh, M.S. Nasser, T. Laoui,
T.A. Saleh, S. Agarwal, I. Tyagi, V.K. Gupta, Adsorptive removal of cadmium(II) ions
from liquid phase using acid modified carbon-based adsorbents, J. Mol. Liq. 204 (2015)
255–263. doi:10.1016/j.molliq.2015.01.033.

[11] Ihsanullah, F.A. Al-khaldi, B. Abu-sharkh, A. Mahmoud, M.I. Qureshi, T. Laoui, M.A.
Atieh, Effect of acid modification on adsorption of hexavalent chromium (Cr(VI)) from
aqueous solution by activated carbon and carbon nanotubes, Desalin. Water Treat. 57
(2016) 7232–7244. doi:10.1080/19443994.2015.1021847.

[12] Ihsanullah, H.A. Asmaly, T.A. Saleh, T. Laoui, V.K. Gupta, M.A. Atieh, Enhanced
adsorption of phenols from liquids by aluminum oxide/carbon nanotubes: Comprehensive
study from synthesis to surface properties, J. Mol. Liq. 206 (2015) 176–182.
doi:10.1016/j.molliq.2015.02.028.

[13] H.A. Asmaly, B. Abussaud, Ihsanullah, T.A. Saleh, V.K. Gupta, M.A. Atieh, Ferric oxide
nanoparticles decorated carbon nanotubes and carbon nanofibers: From synthesis to
enhanced removal of phenol, J. Saudi Chem. Soc. 19 (2015) 511–520.
doi:10.1016/j.jscs.2015.06.002.

[14] H.A. Asmaly, B. Abussaud, Ihsanullah, T.A. Saleh, A. Alaadin, T. Laoui, A.M. Shemsi,
V.K. Gupta, M.A. Atieh, H.A. Asmaly, B. Abussaud, T.A. Saleh, A. Alaadin, Evaluation
of micro- and nano-carbon-based adsorbents for the removal of phenol from aqueous
solutions, Toxicol. Environ. Chem. 97 (2015) 1164–1179.
doi:10.1080/02772248.2015.1092543.

[15] B.S. Lalia, F.E. Ahmed, T. Shah, N. Hilal, R. Hashaikeh, Electrically conductive
membranes based on carbon nanostructures for self-cleaning of biofouling, Desalination.
360 (2015) 8–12. doi:10.1016/j.desal.2015.01.006.

[16] S. Li, G. Liao, Z. Liu, Y. Pan, Q. Wu, Y. Weng, X. Zhang, Z. Yang, O.K.C. Tsui,

87
Enhanced water flux in vertically aligned carbon nanotube arrays and polyethersulfone
composite membranes, J. Mater. Chem. A. 2 (2014) 12171–12176.
doi:10.1039/C4TA02119C.

[17] C. Rizzuto, G. Pugliese, M.A. Bahattab, S.A. Aljlil, E. Drioli, E. Tocci, Multiwalled
carbon nanotube membranes for water purification, Sep. Purif. Technol. 193 (2018) 378–
385. doi:https://doi.org/10.1016/j.seppur.2017.10.025.

[18] M.H.D.A. Farahani, V. Vatanpour, A comprehensive study on the performance and


antifouling enhancement of the PVDF mixed matrix membranes by embedding different
nanoparticulates: Clay, functionalized carbon nanotube, SiO 2 and TiO2, Sep. Purif.
Technol. 197 (2018) 372–381. doi:https://doi.org/10.1016/j.seppur.2018.01.031.

[19] C. Thamaraiselvan, S. Lerman, K. Weinfeld-Cohen, C.G. Dosoretz, Characterization of a


support-free carbon nanotube-microporous membrane for water and wastewater filtration,
Sep. Purif. Technol. 202 (2018) 1–8. doi:https://doi.org/10.1016/j.seppur.2018.03.038.

[20] K.A. Tankus, L. Issman, M. Stolov, V. Freger, Electrotreated carbon nanotube membranes
for facile oil–water separations, ACS Appl. Nano Mater. 1 (2018) 2057–2061.
doi:10.1021/acsanm.8b00442.

[21] W. Chen, S. Chen, T. Liang, Q. Zhang, Z. Fan, H. Yin, K.-W. Huang, X. Zhang, Z. Lai, P.
Sheng, High-flux water desalination with interfacial salt sieving effect in nanoporous
carbon composite membranes, Nat. Nanotechnol. 13 (2018) 345–350.
doi:10.1038/s41565-018-0067-5.

[22] J. Saththasivam, W. Yiming, K. Wang, J. Jin, Z. Liu, A novel srchitecture for carbon
nanotube membranes towards fast and efficient oil/water separation, Sci. Rep. 8 (2018)
7418. doi:10.1038/s41598-018-25788-9.

[23] T.H. Kim, I. Lee, K.-M. Yeon, J. Kim, Biocatalytic membrane with acylase stabilized on
intact carbon nanotubes for effective antifouling via quorum quenching, J. Memb. Sci. 554
(2018) 357–365. doi:https://doi.org/10.1016/j.memsci.2018.03.020.

[24] Y. Wang, Y. Liu, Y. Yu, H. Huang, Influence of CNT-rGO composite structures on their

88
permeability and selectivity for membrane water treatment, J. Memb. Sci. 551 (2018)
326–332. doi:https://doi.org/10.1016/j.memsci.2018.01.031.

[25] K.-J. Lee, H.-D. Park, The most densified vertically-aligned carbon nanotube membranes
and their normalized water permeability and high pressure durability, J. Memb. Sci. 501
(2016) 144–151. doi:https://doi.org/10.1016/j.memsci.2015.12.009.

[26] V. Vatanpour, M. Safarpour, A. Khataee, H. Zarrabi, M.E. Yekavalangi, M. Kavian, A


thin film nanocomposite reverse osmosis membrane containing amine-functionalized
carbon nanotubes, Sep. Purif. Technol. 184 (2017) 135–143.
doi:https://doi.org/10.1016/j.seppur.2017.04.038.

[27] M. Majumder, N. Chopra, R. Andrews, B.J. Hinds, Nanoscale hydrodynamics: Enhanced


flow in carbon nanotubes, Nature. 438 (2005) 44. doi:10.1038/43844a.

[28] C. Wang, M. Li, S. Pan, H. Li, Well-aligned carbon nanotube array membrane synthesized
in porous alumina template by chemical vapor deposition, Chinese Sci. Bull. 45 (2000)
1373–1376. doi:10.1007/BF02886240.

[29] B.J. Hinds, N. Chopra, T. Rantell, R. Andrews, V. Gavalas, L.G. Bachas, Aligned
multiwalled carbon nanotube membranes, Science. 303 (2004) 62–65.
doi:10.1126/science.1092048.

[30] D. Lawrence Arockiasamy, J. Alam, M. Alhoshan, Carbon nanotubes-blended


poly(phenylene sulfone) membranes for ultrafiltration applications, Appl. Water Sci. 3
(2012) 93–103. doi:10.1007/s13201-012-0063-0.

[31] R.S. Hebbar, A.M. Isloor, Inamuddin, A.M. Asiri, Carbon nanotube- and graphene-based
advanced membrane materials for desalination, Environ. Chem. Lett. 15 (2017) 643–671.
doi:10.1007/s10311-017-0653-z.

[32] L. Ma, X. Dong, M. Chen, L. Zhu, C. Wang, F. Yang, Y. Dong, Fabrication and water
treatment application of carbon nanotubes (CNTs)-based composite membranes: A
review, Membranes (Basel). 7 (2017) 16. doi:10.3390/membranes7010016.

[33] M.H. Rashid, S.F. Ralph, Carbon nanotube membranes: Synthesis, properties, and future

89
filtration applications, Nanomaterials. 7 (2017). doi:10.3390/nano7050099.

[34] S. Begum, A. Kausar, H. Ullah, M. Siddiq, Potential of Polyvinylidene Fluoride/carbon


nanotube composite in energy, electronics, and membrane technology: An overview,
Polym. Plast. Technol. Eng. 55 (2016) 1949–1970. doi:10.1080/03602559.2016.1185630.

[35] J. Lee, S. Jeong, Z. Liu, Progress and challenges of carbon nanotube membrane in water
treatment, Crit. Rev. Environ. Sci. Technol. 46 (2016) 999–1046.
doi:10.1080/10643389.2016.1191894.

[36] J.K. Holt, H.G. Park, Y. Wang, M. Stadermann, A.B. Artyukhin, C.P. Grigoropoulos, A.
Noy, O. Bakajin, Fast mass transport through sub-2-nanometer carbon nanotubes, Science.
312 (2006) 1034–1037. doi:10.1126/science.1126298.

[37] R.L. McGinnis, K. Reimund, J. Ren, L. Xia, M.R. Chowdhury, X. Sun, M. Abril, J.D.
Moon, M.M. Merrick, J. Park, K.A. Stevens, J.R. McCutcheon, B.D. Freeman, Large-
scale polymeric carbon nanotube membranes with sub–1.27-nm pores, Sci. Adv. 4 (2018)
e1700938. http://advances.sciencemag.org/content/4/3/e1700938.abstract.

[38] H. Dai, Carbon nanotubes: Opportunities and challenges, Surf. Sci. 500 (2002) 218–241.
doi:10.1016/S0039-6028(01)01558-8.

[39] Y.L. Zhao, J.F. Stoddart, Noncovalent functionalization of single-walled carbon


nanotubes, Acc. Chem. Res. 42 (2009) 1161–1171. doi:10.1021/ar900056z.

[40] K. Balasubramanian, M. Burghard, Chemically functionalized carbon nanotubes, Small. 1


(2005) 180–192. doi:10.1002/smll.200400118.

[41] N. Saifuddin, A.Z. Raziah, A.R. Junizah, Carbon nanotubes: A review on structure and
their interaction with proteins, J. Chem. 2013 (2013) Article ID 676815, 18 pages.
doi:10.1155/2013/676815.

[42] P. Ma, N.A. Siddiqui, G. Marom, J. Kim, Dispersion and functionalization of carbon
nanotubes for polymer-based nanocomposites: A review, Compos. Part A. 41 (2010)
1345–1367. doi:10.1016/j.compositesa.2010.07.003.

90
[43] G. Mittal, V. Dhand, K. Yop, S. Park, W. Ro, A review on carbon nanotubes and graphene
as fillers in reinforced polymer nanocomposites, J. Ind. Eng. Chem. 21 (2015) 11–25.
doi:10.1016/j.jiec.2014.03.022.

[44] S. Iijima, Helical microtubules of graphitic carbon, Nature. 354 (1991) 56–58.
doi:10.1038/354056a0.

[45] X. Wang, Q. Li, J. Xie, Z. Jin, J. Wang, Y. Li, K. Jiang, S. Fan, Fabrication of ultralong
and electrically uniform single-walled carbon nanotubes on clean substrates, Nano Lett. 9
(2009) 3137–3141. doi:10.1021/nl901260b.

[46] Y.M. Manawi, Ihsanullah, A. Samara, T. Al-Ansari, M.A. Atieh, A review of carbon
nanomaterials’ synthesis via the chemical vapor deposition (CVD) method, Materials
(Basel). 11 (2018). doi:10.3390/ma11050822.

[47] A. Abbas, B.A. Abussaud, Ihsanullah, N.A.H. Al-baghli, M. Khraisheh, M.A. Atieh,
Benzene removal by iron oxide nanoparticles decorated carbon nanotubes, J. Nanomater.
2016 (2016) Article ID 5654129, 10 pages http://dx.doi.org/10.

[48] R. Das, B.F. Leo, F. Murphy, The toxic truth about carbon nanotubes in water
purification: A perspective view, Nanoscale Res. Lett. 13 (2018) 183.
doi:10.1186/s11671-018-2589-z.

[49] V.K.K. Upadhyayula, S. Deng, M.C. Mitchell, G.B. Smith, Application of carbon
nanotube technology for removal of contaminants in drinking water: A review, Sci. Total
Environ. 408 (2009) 1–13. doi:10.1016/j.scitotenv.2009.09.027.

[50] C.R. of C.N.T. and A. of P.O. and E.H.R. Lam, J.T. James, R. McCluskey, S. Arepalli,
R.L. Hunter, A review of carbon nanotube toxicity and assessment of potential
occupational and environmental health risks, Crit. Rev. Toxicol. 36 (2006) 189–217.
doi:10.1080/10408440600570233.

[51] P. Wick, P. Manser, L.K. Limbach, U. Dettlaff-Weglikowska, F. Krumeich, S. Roth, W.J.


Stark, A. Bruinink, The degree and kind of agglomeration affect carbon nanotube
cytotoxicity, Toxicol. Lett. 168 (2007) 121–131.

91
doi:https://doi.org/10.1016/j.toxlet.2006.08.019.

[52] H. Dumortier, S. Lacotte, G. Pastorin, R. Marega, W. Wu, D. Bonifazi, J.-P. Briand, M.


Prato, S. Muller, A. Bianco, Functionalized carbon nanotubes are non-cytotoxic and
preserve the functionality of primary immune cells, Nano Lett. 6 (2006) 1522–1528.
doi:10.1021/nl061160x.

[53] A. Magrez, S. Kasas, V. Salicio, N. Pasquier, J.W. Seo, M. Celio, S. Catsicas, B.


Schwaller, L. Forró, Cellular toxicity of carbon-based nanomaterials, Nano Lett. 6 (2006)
1121–1125. doi:10.1021/nl060162e.

[54] R. Das, E. Ali, S. Bee, A. Hamid, S. Ramakrishna, Z. Zaman, Carbon nanotube


membranes for water purification: A bright future in water desalination, Desalination. 336
(2014) 97–109.

[55] Y. Manawi, V. Kochkodan, M.A. Hussein, M.A. Khaleel, M. Khraisheh, N. Hilal, Can
carbon-based nanomaterials revolutionize membrane fabrication for water treatment and
desalination ?, Desalination. 391 (2016) 69–88. doi:10.1016/j.desal.2016.02.015.

[56] C. Hoon, Y. Baek, C. Lee, S. Ouk, S. Kim, S. Lee, S. Kim, S. Seek, J. Park, J. Yoon,
Carbon nanotube-based membranes: Fabrication and application to desalination, J. Ind.
Eng. Chem. 18 (2012) 1551–1559. doi:10.1016/j.jiec.2012.04.005.

[57] L. Sun, R.M. Crooks, Single carbon nanotube membranes: A well-defined model for
studying mass transport through nanoporous materials, J. Am. Chem. Soc. 122 (2000)
12340–12345.

[58] B. A, G. Hummer, J.C. Rasaiah, J.P. Noworyta, Water conduction through the
hydrophobic channel of a carbon nanotube, Nature. 414 (2001) 188–190.

[59] S. Joseph, R.J. Mashl, E. Jakobsson, N.R. Aluru, Electrolytic transport in modified carbon
nanotubes, Nano Lett. 3 (2003) 1399–1403. doi:10.1021/nl0346326.

[60] J.K. Holt, A. Noy, T. Huser, D. Eaglesham, O. Bakajin, Fabrication of a carbon nanotube-
embedded silicon nitride membrane for studies of nanometer-scale mass transport, Nano
Lett. 4 (2004) 2245–2250. doi:10.1021/nl048876h.

92
[61] K. Sears, L. Dumée, J. Schütz, M. She, C. Huynh, Recent developments in carbon
nanotube membranes for water purification and gas separation, Materials (Basel). (2010)
127–149. doi:10.3390/ma3010127.

[62] M.E. Suk, N.R. Aluru, Modeling water flow through carbon nanotube membranes with
entrance/exit effects, Nanoscale Microscale Thermophys. Eng. 21 (2017) 247–262.
doi:10.1080/15567265.2017.1355949.

[63] J.A. Thomas, A.J.H. Mcgaughey, Water flow in carbon nanotubes: Transition to
subcontinuum transport, Phys. Rev. Lett. 102 (2009) 184502.
doi:10.1103/PhysRevLett.102.184502.

[64] R.H. Tunuguntla, R.Y. Henley, Y.-C. Yao, T.A. Pham, M. Wanunu, A. Noy, Enhanced
water permeability and tunable ion selectivity in subnanometer carbon nanotube porins,
Science (80-. ). 357 (2017) 792–796. doi:10.1126/science.aan2438.

[65] B. Corry, Designing carbon nanotube membranes for efficient water desalination, J. Phys.
Chem. B. 112 (2008) 1427–1434. doi:10.1021/jp709845u.

[66] Y. Chan, J.M. Hill, Ion selectivity using membranes comprising functionalized carbon
nanotubes, J Math Chem. 51 (2013) 1258–1273. doi:10.1007/s10910-013-0142-y.

[67] H. Verweij, M.C. Schillo, J. Li, Fast mass transport through carbon nanotube membranes,
Small. 12 (2007) 1996–2004. doi:10.1002/smll.200700368.

[68] M. Majumder, N. Chopra, B.J. Hinds, Effect of tip functionalization on transport through
vertically oriented carbon nanotube membranes, J. Am. Chem. Soc. 127 (2005) 9062–
9070.

[69] Y. Baek, C. Kim, D. Kyun, T. Kim, J. Seok, Y. Hyup, K. Hyun, S. Seek, S. Cheol, J. Lim,
K. Lee, J. Yoon, High performance and antifouling vertically aligned carbon nanotube
membrane for water purification, J. Memb. Sci. 460 (2014) 171–177.
doi:10.1016/j.memsci.2014.02.042.

[70] M. Yu, H.H. Funke, J.L. Falconer, R.D. Noble, High density, vertically-aligned carbon
nanotube membranes, Nano Lett. 9 (2009) 225–229.

93
[71] C. Lee, S. Baik, Vertically-aligned carbon nano-tube membrane filters with
superhydrophobicity and superoleophilicity, Carbon N. Y. 48 (2010) 2192–2197.
doi:10.1016/j.carbon.2010.02.020.

[72] M. Majumder, N. Chopra, B.J. Hinds, Mass transport through carbon nanotube
membranes in three diff erent regimes: Ionic diff usion and gas and liquid flow, ACS
Nano. 5 (2011) 3867–3877.

[73] A. S. Brady-Estevez, S. Kang, M. Elimelech, A single-walled-carbon-nanotube filter for


removal of viral and bacterial pathogens, Small. 4 (2008) 481–484.
doi:10.1002/smll.200700863.

[74] B. Lee, Y. Baek, M. Lee, D.H. Jeong, H.H. Lee, J. Yoon, Y.H. Kim, A carbon nanotube
wall membrane for water treatment, Nat. Commun. 6 (2015) 7109.
doi:10.1038/ncomms8109.

[75] F. Du, L. Qu, Z. Xia, L. Feng, L. Dai, Membranes of vertically aligned superlong carbon
nanotubes, Langmuir. 27 (2011) 8437–8443.

[76] W. Chan, H. Chen, A. Surapathi, M.G. Taylor, X. Shao, E. Marand, C.E.T. Al, Zwitterion
functionalized carbon nanotube/polyamide nanocomposite, ACS Nano. 7 (2013) 5308–
5319.

[77] S. Kim, F. Fornasiero, H. Gyu, J. Bin, E. Meshot, G. Giraldo, M. Stadermann, M.


Fireman, J. Shan, C.P. Grigoropoulos, O. Bakajin, Fabrication of flexible, aligned carbon
nanotube/polymer composite membranes by in-situ polymerization, J. Memb. Sci. 460
(2014) 91–98. doi:10.1016/j.memsci.2014.02.016.

[78] S. Park, J. Jung, S. Lee, Y. Baek, J. Yoon, D. Kyun, Y. Hyup, Fouling and rejection
behavior of carbon nanotube membranes, Desalination. 343 (2014) 180–186.
doi:10.1016/j.desal.2013.10.005.

[79] T.I. Qing Li, Christina W. Kartikowati, Takashi Ogi, K. Okuyama, Facile fabrication of
carbon nanotube forest-like films via coaxial electrospray, Carbon N. Y. 115 (2017) 116–
119. doi:10.1016/j.carbon.2016.12.095.

94
[80] B. Hinds, Dramatic transport properties of carbon nanotube membranes for a robust
protein channel mimetic platform, Curr. Opin. Solid State Mater. Sci. 16 (2012) 1–9.
doi:10.1016/j.cossms.2011.05.003.

[81] A. Srivastava, O.N. Srivastava, S. Talapatra, R. Vajtai, P.M. Ajayan, Carbon nanotube
filters, Nat. Mater. 3 (2004) 610–614. doi:10.1038/nmat1192.

[82] H. Matsumoto, S. Tsuruoka, Y. Hayashi, K. Abe, K. Hata, S. Zhang, Y. Saito, M. Aiba, T.


Tokunaga, T. Iijima, T. Hayashi, H. Inoue, G.A.J. Amaratunga, Water transport
phenomena through membranes consisting of vertically-aligned double-walled carbon
nanotube array, Carbon N. Y. 120 (2017) 358–365.
doi:https://doi.org/10.1016/j.carbon.2017.05.034.

[83] L. Dumée, V. Germain, K. Sears, J. Schütz, N. Finn, M. Duke, S. Cerneaux, D. Cornu, S.


Gray, Enhanced durability and hydrophobicity of carbon nanotube bucky paper
membranes in membrane distillation, J. Memb. Sci. 376 (2011) 241–246.
doi:10.1016/j.memsci.2011.04.024.

[84] L.F. Dumée, K. Sears, J. Schütz, N. Finn, C. Huynh, S. Hawkins, M. Duke, S. Gray,
Characterization and evaluation of carbon nanotube bucky-paper membranes for direct
contact membrane distillation, J. Memb. Sci. 351 (2010) 36–43.
doi:10.1016/j.memsci.2010.01.025.

[85] L. Dumée, K. Sears, N. Finn, M. Duke, L. Dumée, K. Sears, N. Finn, M. Duke, Carbon
nanotube based composite membranes for water desalination by membrane distillation,
Desalin. Water Treat. 17 (2010) 72–79. doi:10.5004/dwt.2010.1701.

[86] B.G. Viswanathan, D.B. Kane, P.J. Lipowicz, High efficiency fine particulate filtration
using carbon nanotube coatings, Adv. Mater. 16 (2004) 2045–2049.
doi:10.1002/adma.200400463.

[87] S. Kang, M. Pinault, L.D. Pfefferle, M. Elimelech, Single-walled carbon nanotubes exhibit
strong antimicrobial activity, Langmuir. 23 (2007) 8670–8673.

[88] M.A. Atieh, Ihsanullah, and T. Laoui., Fabrication of carbon nanotube membranes,

95
(2017) U.S. Patent 9,776,140.

[89] Ihsanullah, F. Patel, M. Khraisheh, M.A. Atieh, T. Laoui, Novel aluminum oxide-
impregnated carbon nanotube membrane for the removal of cadmium from aqueous
solution, Materials (Basel). 10 (2017). doi:10.3390/ma10101144.

[90] P.M. Ajayan, Nanotubes from carbon, Chem. Rev. 99 (1999) 1787−1799.

[91] J. Choi, J. Jegal, W. Kim, Fabrication and characterization of multi-walled carbon


nanotubes/polymer blend membranes, J. Membr. Sci. 284 (2006) 406–415.
doi:10.1016/j.memsci.2006.08.013.

[92] E. Celik, H. Park, H. Choi, H. Choi, Carbon nanotube blended polyethersulfone


membranes for fouling control in water treatment, Water Res. 45 (2011) 274–282.
doi:10.1016/j.watres.2010.07.060.

[93] H. Wu, B. Tang, P. Wu, Novel ultrafiltration membranes prepared from a multi-walled
carbon nanotubes/polymer composite, J. Memb. Sci. 362 (2010) 374–383.
doi:10.1016/j.memsci.2010.06.064.

[94] L. Brunet, D.. Lyon, K. Zodrow, J.-C. Rouch, B. Caussat, P. Serp, J.-C. Remigy, M..
Wiesner, P.J.. Alvarez, Properties of membranes containing semi- dispersed carbon
nanotubes, Environ. Eng. Sci. 25 (2008) 565–575. doi:10.1089/ees.2007.0076.

[95] S. Majeed, D. Fierro, K. Buhr, J. Wind, B. Du, A. Boschetti-de-Fierro, V. Abetz, Multi-


walled carbon nanotubes (MWCNTs) mixed polyacrylonitrile (PAN) ultrafiltration
membranes, J. Memb. Sci. 403–404 (2012) 101–109. doi:10.1016/j.memsci.2012.02.029.

[96] A. Tiraferri, C.D. Vecitis, M. Elimelech, Covalent binding of single-walled carbon


nanotubes to polyamide membranes for antimicrobial surface properties, ACS Appl.
Mater. Interfaces. 3 (2011) 2869–2877.

[97] P. Shah, C.N. Murthy, Studies on the porosity control of MWCNT/polysulfone composite
membrane and its effect on metal removal, J. Memb. Sci. 437 (2013) 90–98.
doi:10.1016/j.memsci.2013.02.042.

96
[98] J. Yin, G. Zhu, B. Deng, Multi-walled carbon nanotubes(MWNTs)/polysulfone(PSU)
mixed matrix hollow fi ber membranes for enhanced water treatment, J. Membr. Sci. 437
(2013) 237–248.

[99] A. Khalid, A.A. Al-Juhani, O.C. Al-Hamouz, T. Laoui, Z. Khan, M.A. Atieh, Preparation
and properties of nanocomposite polysulfone/multi-walled carbon nanotubes membranes
for desalination, Desalination. 367 (2015) 134–144. doi:10.1016/j.desal.2015.04.001.

[100] Y. Medina-gonzalez, J. Remigy, Sonication-assisted preparation of pristine MWCNT –


polysulfone conductive microporous membranes, Mater. Lett. 65 (2011) 229–232.
doi:10.1016/j.matlet.2010.10.016.

[101] J. nan Shen, C. chao Yu, R. Hui min, Gao, Cong jie, B. Van Der Bruggen, Preparation and
characterization of thin-film nanocomposite membranes embedded with poly(methyl
methacrylate) hydrophobic modified multiwalled carbon nanotubes by interfacial
polymerization, J. Memb. Sci. 442 (2013) 18–26. doi:10.1016/j.memsci.2013.04.018.

[102] H. Wu, B. Tang, P. Wu, Optimization, characterization and nanofiltration properties test
of MWNTs/polyester thin film nanocomposite membrane, J. Memb. Sci. 428 (2013) 425–
433. doi:10.1016/j.memsci.2012.10.042.

[103] J. Park, W. Choi, S.H. Kim, B.H. Chun, J. Bang, K.B. Lee, J. Park, W. Choi, S.H. Kim,
B.H. Chun, J. Bang, K.B. Lee, Enhancement of chlorine resistance in carbon nanotube
based nanocomposite reverse osmosis membranes, Desalin. Water Treat. 15 (2010) 198–
204. doi:10.5004/dwt.2010.1686.

[104] H. Wu, B. Tang, P. Wu, MWNTs/Polyester thin film nanocomposite membrane: An


approach to overcome the trade-off effect between permeability and selectivity, J. Phys.
Chem. C 2010,. 114 (2010) 16395–16400.

[105] M. Amini, M. Jahanshahi, A. Rahimpour, Synthesis of novel thin film nanocomposite


(TFN) forward osmosis membranes using functionalized multi-walled carbon nanotubes,
J. Memb. Sci. 435 (2013) 233–241. doi:10.1016/j.memsci.2013.01.041.

[106] H. Zhao, S. Qiu, L. Wu, L. Zhang, H. Chen, C. Gao, Improving the performance of

97
polyamide reverse osmosis membrane by incorporation of modified multi-walled carbon
nanotubes, J. Memb. Sci. 450 (2014) 249–256. doi:10.1016/j.memsci.2013.09.014.

[107] H. Zarrabi, M. Ehsan, V. Vatanpour, A. Shockravi, M. Safarpour, Improvement in


desalination performance of thin film nanocomposite nanofiltration membrane using
amine-functionalized multiwalled carbon nanotube, Desalination. 394 (2016) 83–90.
doi:10.1016/j.desal.2016.05.002.

[108] S. Roy, S. Addo, S. Mitra, K.K. Sirkar, Facile fabrication of superior nanofiltration
membranes from interfacially polymerized CNT-polymer composites, J. Memb. Sci. 375
(2011) 81–87. doi:10.1016/j.memsci.2011.03.012.

[109] L. Zhang, H. Chen, Preparation of high-flux thin film nanocomposite reverse osmosis
membranes by incorporating functionalized multi-walled carbon nanotubes, Desalin.
Water Treat. 34 (2011) 19–24.

[110] J. Zheng, M. Li, K. Yu, J. Hu, X. Zhang, L. Wang, Sulfonated multiwall carbon nanotubes
assisted thin-film nanocomposite membrane with enhanced water flux and anti-fouling
property, J. Memb. Sci. 524 (2017) 344–353. doi:10.1016/j.memsci.2016.11.032.

[111] S. Inukai, R. Cruz-silva, J. Ortiz-medina, A. Morelos-gomez, High-performance


multifunctional reverse osmosis membranes obtained by carbon nanotube·polyamide
nanocomposite, Sci. Rep. 5 (2015) 13562. doi:10.1038/srep13562.

[112] Y. Chang-chao, Y. Hong-wei, C. Yue-xia, R. Hui-min, S. Jiang-nan, Preparation thin film


nanocomposite membrane incorporating PMMA modified MWNT for nanofiltration, Key
Eng. Mater. 562–565 (2013) 882–886. doi:10.4028/www.scientific.net/KEM.562-
565.882.

[113] J. Park, W. Choi, J. Cho, B.H. Chun, S.H. Kim, K.B. Lee, J. Bang, Carbon nanotube-
based nanocomposite desalination membranes from layer-by-layer assembly, Desalin.
Water Treat. 15 (2010) 76–83. doi:10.5004/dwt.2010.1670.

[114] L. Liu, M. Son, S. Chakraborty, C. Bhattacharjee, Fabrication of ultra-thin


polyelectrolyte/carbon nanotube membrane by spray-assisted layer-by- layer technique:

98
Characterization and its anti- protein fouling properties for water treatment, Desalin.
Water Treat. 51 (2013) 6194–6200. doi:10.1080/19443994.2013.780767.

[115] J. Choi, J. Jegal, W. Kim, H. Choi, Incorporation of multiwalled carbon nanotubes into
Poly (vinyl alcohol) membranes for use in the pervaporation of water/ethanol mixtures, J.
Appl. Polym. Sci. 111 (2009) 2186–2193. doi:10.1002/app.

[116] Y. Mansourpanah, S.S. Madaeni, A. Rahimpour, M. Adeli, M.Y. Hashemi, M.R.


Moradian, Fabrication new PES-based mixed matrix nanocomposite membranes using
polycaprolactone modified carbon nanotubes as the additive: Property changes and
morphological studies, Desalination. 277 (2011) 171–177.
doi:10.1016/j.desal.2011.04.022.

[117] F. Ahmed, C.M. Santos, J. Mangadlao, R. Advincula, D.F. Rodrigues, Antimicrobial


PVK:SWNT nanocomposite coated membrane for water purification: Performance and
toxicity testing, Water Res. 47 (2013) 3966–3975. doi:10.1016/j.watres.2012.10.055.

[118] A. Kausar, Novel water purification membranes of polystyrene/multi-walled carbon


nanotube- grafted -graphene oxide hybrids, Am. J. Polym. Sci. 4 (2014) 63–72.
doi:10.5923/j.ajps.20140403.01.

[119] H.A. Shawky, S. Chae, S. Lin, M.R. Wiesner, Synthesis and characterization of a carbon
nanotube/polymer nanocomposite membrane for water treatment, Desalination. 272
(2011) 46–50. doi:10.1016/j.desal.2010.12.051.

[120] K.C. Ho, Y.H. Teow, W.L. Ang, A.W. Mohammad, Novel GO/OMWCNTs mixed-matrix
membrane with enhanced antifouling property for palm oil mill effluent treatment, Sep.
Purif. Technol. 177 (2017) 337–349.

[121] N. Mehwish, A. Kausar, M. Siddiq, High‑performance polyvinylidene fluoride/poly


(styrene – butadiene – styrene)/ functionalized MWCNTs‑SCN‑Ag nanocomposite
membranes, Iran. Polym. J. 24 (2015) 549–559. doi:10.1007/s13726-015-0346-z.

[122] J. Lee, Y. Ye, A.J. Ward, C. Zhou, V. Chen, A.I. Minett, S. Lee, Z. Liu, S. Chae, J. Shi,
High flux and high selectivity carbon nanotube composite membranes for natural organic

99
matter removal, Sep. Purif. Technol. 163 (2016) 109–119.
doi:10.1016/j.seppur.2016.02.032.

[123] Y.H. Teow, A.L. Ahmad, J.K. Lim, B.S. Ooi, Preparation and characterization of
PVDF/TiO2 mixed matrix membrane via in situ colloidal precipitation method,
Desalination. 295 (2012) 61–69. doi:https://doi.org/10.1016/j.desal.2012.03.019.

[124] V. Vatanpour, S. Siavash, R. Moradian, S. Zinadini, B. Astinchap, Novel antibifouling


nanofiltration polyethersulfone membrane fabricated from embedding TiO 2 coated
multiwalled carbon nanotubes, Sep. Purif. Technol. 90 (2012) 69–82.
doi:10.1016/j.seppur.2012.02.014.

[125] P. Daraei, S. Siavash, N. Ghaemi, M. Ali, B. Astinchap, R. Moradian, Enhancing


antifouling capability of PES membrane via mixing with various types of polymer
modified multi-walled carbon nanotube, J. Memb. Sci. 444 (2013) 184–191.
doi:10.1016/j.memsci.2013.05.020.

[126] V.K.K. Upadhyayula, V. Gadhamshetty, Appreciating the role of carbon nanotube


composites in preventing biofouling and promoting biofilms on material surfaces in
environmental engineering: A review, Biotechnol. Adv. 28 (2010) 802–816.
doi:https://doi.org/10.1016/j.biotechadv.2010.06.006.

[127] V. Vatanpour, S. Siavash, R. Moradian, S. Zinadini, B. Astinchap, Fabrication and


characterization of novel antifouling nanofiltration membrane prepared from oxidized
multiwalled carbon nanotube/polyethersulfone nanocomposite, J. Membr. Sci. 375 (2011)
284–294. doi:10.1016/j.memsci.2011.03.055.

[128] X. Zhao, J. Ma, Z. Wang, G. Wen, J. Jiang, F. Shi, L. Sheng, Hyperbranched-polymer


functionalized multi-walled carbon nanotubes for poly (vinylidene fluoride ) membranes:
From dispersion to blended fouling-control membrane, Desalination. 303 (2012) 29–38.
doi:10.1016/j.desal.2012.07.009.

[129] A. Rahimpour, M. Jahanshahi, S. Khalili, A. Mollahosseini, A. Zirepour, B. Rajaeian,


Novel functionalized carbon nanotubes for improving the surface properties and
performance of polyethersulfone (PES) membrane, Desalination. 286 (2012) 99–107.

100
doi:10.1016/j.desal.2011.10.039.

[130] V. Vatanpour, M. Esmaeili, M. Hossein, D. Abadi, Fouling reduction and retention


increment of polyethersulfone nano filtration membranes embedded by amine-
functionalized multi-walled carbon nanotubes, J. Memb. Sci. 466 (2014) 70–81.
doi:10.1016/j.memsci.2014.04.031.

[131] K.C. Ho, Y.H. Teow, A.W. Mohammad, W.L. Ang, P.H. Lee, Development of graphene
oxide (GO)/multi-walled carbon nanotubes (MWCNTs) nanocomposite conductive
membranes for electrically enhanced fouling mitigation, J. Memb. Sci. 552 (2018) 189–
201. doi:https://doi.org/10.1016/j.memsci.2018.02.001.

[132] G. Yi, S. Chen, X. Quan, G. Wei, X. Fan, H. Yu, Enhanced separation performance of
carbon nanotube–polyvinyl alcohol composite membranes for emulsified oily wastewater
treatment under electrical assistance, Sep. Purif. Technol. 197 (2018) 107–115.
doi:https://doi.org/10.1016/j.seppur.2017.12.058.

[133] J. Ortiz-Medina, S. Inukai, T. Araki, A. Morelos-Gomez, R. Cruz-Silva, K. Takeuchi, T.


Noguchi, T. Kawaguchi, M. Terrones, M. Endo, Robust water desalination membranes
against degradation using high loads of carbon nanotubes, Sci. Rep. 8 (2018) 2748.
doi:10.1038/s41598-018-21192-5.

[134] J. Ma, Y. Zhao, Z. Xu, C. Min, B. Zhou, Y. Li, B. Li, J. Niu, Role of oxygen-containing
groups on MWCNTs in enhanced separation and permeability performance for PVDF
hybrid ultra filtration membranes, Desalination. 320 (2013) 1–9.
doi:10.1016/j.desal.2013.04.012.

[135] N. Phao, E.N. Nxumalo, B.B. Mamba, S.D. Mhlanga, A nitrogen-doped carbon nanotube
enhanced polyethersulfone membrane system for water treatment, Phys. Chem. Earth. 66
(2013) 148–156. doi:10.1016/j.pce.2013.09.009.

[136] W. Wang, Preparation and characterization of PEG-g-MWCNTs/PSf nano-hybrid


membranes with hydrophilicity and antifouling properties, RSC Adv. 5 (2015) 84746–
84753. doi:10.1039/C5RA16077D.

101
[137] A. Khalid, A. Abdel-Karim, M. Ali Atieh, S. Javed, G. McKay, PEG-CNTs
nanocomposite PSU membranes for wastewater treatment by membrane bioreactor, Sep.
Purif. Technol. 190 (2018) 165–176. doi:https://doi.org/10.1016/j.seppur.2017.08.055.

[138] C. Boo, M. Elimelech, Carbon nanotubes keep up the heat, Nat. Nanotechnol. 12 (2017)
501. http://dx.doi.org/10.1038/nnano.2017.114.

[139] C.F. De Lannoy, D. Jassby, D.D. Davis, M.R. Wiesner, A highly electrically conductive
polymer–multiwalled carbon nanotube nanocomposite membrane, J. Memb. Sci. 416
(2012) 718–724.

[140] F. Ahmed, B.S. Lalia, V. Kochkodan, N. Hilal, R. Hashaikeh, Electrically conductive


polymeric membranes for fouling prevention and detection: A review, Desalination. 391
(2016) 1–15. doi:https://doi.org/10.1016/j.desal.2016.01.030.

[141] J. Zhang, Z. Xu, M. Shan, B. Zhou, Y. Li, B. Li, Synergetic effects of oxidized carbon
nanotubes and graphene oxide on fouling control and anti-fouling mechanism of
polyvinylidene fluoride ultra filtration membranes, J. Memb. Sci. 448 (2013) 81–92.
doi:10.1016/j.memsci.2013.07.064.

[142] J. Lee, Y. Ye, A.J. Ward, C. Zhou, V. Chen, I. Andrew, S. Lee, Z. Liu, S. Chae, J. Shi,
High flux and high selectivity carbon nanotube composite membrane for natural organic
matter removal, Sep. Purif. Technol. 163 (2016) 109–119.
doi:10.1016/j.seppur.2016.02.032.

[143] M. Rabbani, J.L. Tyler, H.A. Stretz, M.J.M. Wells, Effects of a dual nano filler, nano-
TiO2 and MWCNT, for polysulfone-based nanocomposite membranes for water
purification, Desalination. 372 (2015) 47–56. doi:10.1016/j.desal.2015.06.014.

[144] S. Kar, M. Subramanian, A. Pal, A.K. Ghosh, R.C. Bindal, S. Prabhakar, J. Nuwad,
C.G.S. Pillai, S. Chattopadhyay, P.K. Tewari, Preparation, characterisation and
performance evaluation of anti-biofouling property of carbon nanotube-polysulfone
nanocomposite membranes, AIP Conf. Proc. 1538 (2013) 181–185.
doi:10.1063/1.4810053.

102
[145] P. Shah, C.N. Murthy, Studies on the porosity control of MWCNT/polysulfone composite
membrane and its effect on metal removal, J. Memb. Sci. 437 (2013) 90–98.
doi:10.1016/j.memsci.2013.02.042.

[146] S.S. Madaeni, S. Zinadini, V. Vatanpour, Preparation of superhydrophobic nanofiltration


membrane by embedding multiwalled carbon nanotube and polydimethylsiloxane in pores
of microfiltration membrane, Sep. Purif. Technol. 111 (2013) 98–107.
doi:10.1016/j.seppur.2013.03.033.

[147] C. De Lannoy, E. Soyer, M.R. Wiesner, Optimizing carbon nanotube-reinforced


polysulfone ultra filtration membranes through carboxylic acid functionalization, J.
Memb. Sci. 447 (2013) 395–402.

[148] S. Qiu, L. Wu, X. Pan, L. Zhang, H. Chen, C. Gao, Preparation and properties of
functionalized carbon nanotube/PSF blend ultrafiltration membranes, J. Memb. Sci. 342
(2009) 165–172. doi:10.1016/j.memsci.2009.06.041.

[149] E. Celik, L. Liu, H. Choi, Protein fouling behavior of carbon nanotube/polyethersulfone


composite membranes during water filtration, Water Res. 45 (2011) 5287–5294.
doi:10.1016/j.watres.2011.07.036.

[150] H. You, X. Li, Y. Yang, B. Wang, Z. Li, X. Wang, M. Zhu, B.S. Hsiao, High flux low
pressure thin film nanocomposite ultrafiltration membranes based on nanofibrous
substrates, Sep. Purif. Technol. 108 (2013) 143–151. doi:10.1016/j.seppur.2013.02.014.

[151] Y. Zhao, Z. Xu, M. Shan, C. Min, B. Zhou, Y. Li, B. Li, L. Liu, X. Qian, Effect of
graphite oxide and multi-walled carbon nanotubes on the microstructure and performance
of PVDF membranes, Sep. Purif. Technol. 103 (2013) 78–83.
doi:10.1016/j.seppur.2012.10.012.

[152] M. Tian, Y. Wang, R. Wang, Synthesis and characterization of novel high-performance


thin film nanocomposite (TFN) FO membranes with nano fibrous substrate reinforced by
functionalized carbon nanotubes, Desalination. 370 (2015) 79–86.
doi:10.1016/j.desal.2015.05.016.

103
[153] O. Sae-khow, S. Mitra, Fabrication and characterization of carbon nanotubes immobilized
in porous polymeric membranes, J. Mater. Chem. 19 (2009) 3713–3718.
doi:10.1039/b822879e.

[154] E. Salehi, S.S. Madaeni, L. Rajabi, V. Vatanpour, A.A. Derakhshan, S. Zinadini, S.


Ghorabi, Novel chitosan/poly (vinyl ) alcohol thin adsorptive membranes modified with
amino functionalized multi-walled carbon nanotubes for Cu(II) removal from water:
Preparation, characterization, adsorption kinetics and thermodynamics, Sep. Purif.
Technol. 89 (2012) 309–319. doi:10.1016/j.seppur.2012.02.002.

[155] S. Qiu, L. Wu, G. Shi, L. Zhang, H. Chen, C. Gao, Preparation and pervaporation property
of chitosan membrane with functionalized multiwalled carbon nanotubes, Ind. Eng. Chem.
Res. (2010) 11667–11675.

[156] E. Kim, G. Hwang, M.G. El-din, Y. Liu, Development of nanosilver and multi-walled
carbon nanotubes thin-film nanocomposite membrane for enhanced water treatment, J.
Memb. Sci. 394–395 (2012) 37–48. doi:10.1016/j.memsci.2011.11.041.

[157] Q. Zhang, C.D. Vecitis, Conductive CNT-PVDF membrane for capacitive organic fouling
reduction, J. Memb. Sci. 459 (2014) 143–156. doi:10.1016/j.memsci.2014.02.017.

[158] A. Zirehpour, A. Rahimpour, M. Jahanshahi, M. Peyravi, Mixed matrix membrane


application for olive oil wastewater treatment: Process optimization based on Taguchi
design method, J. Environ. Manage. 132 (2014) 113–120.
doi:10.1016/j.jenvman.2013.10.028.

[159] E. Kim, Y. Liu, M.G. El-din, An in-situ integrated system of carbon nanotubes
nanocomposite membrane for oil sands process-affected water treatment, J. Memb. Sci.
429 (2013) 418–427. doi:10.1016/j.memsci.2012.11.077.

[160] P. Daraei, S. Siavash, N. Ghaemi, H. Ahmadi, M. Ali, Fabrication of PES nanofiltration


membrane by simultaneous use of multi-walled carbon nanotube and surface graft
polymerization method: Comparison of MWCNT and PAA modified MWCNT, Sep.
Purif. Technol. 104 (2013) 32–44. doi:10.1016/j.seppur.2012.11.004.

104
[161] R. Saranya, G. Arthanareeswaran, D.D. Dionysiou, Treatment of paper mill effluent using
polyethersulfone/functionalised multiwalled carbon nanotubes based nanocomposite
membranes, Chem. Eng. J. 236 (2014) 369–377. doi:10.1016/j.cej.2013.09.096.

[162] M. Irfan, H. Basri, W. Lau, An acid functionalized MWCNT/PVP nanocomposite as a


new additive for fabrication of an ultrafiltration membrane with improved anti-fouling
resistance, RSC Adv. 5 (2015) 95421–95432. doi:10.1039/C5RA11344J.

[163] H. Ma, K. Yoon, L. Rong, M. Shokralla, A. Kopot, X. Wang, D. Fang, B.S. Hsiao, B.
Chu, Thin-film nanofibrous composite ultrafiltration membranes based on polyvinyl
alcohol barrier layer containing directional water channels, Ind. Eng. Chem. Res. (2010)
11978–11984.

[164] M. Nidzhom, Z. Abidin, P. Sean, A. Fauzi, M. Ha, D. Othman, H. Hasbullah, N. Said, S.


Hamimah, S. Abdul, F. Kamal, Development of biocompatible and safe polyethersulfone
hemodialysis membrane incorporated with functionalized multi-walled carbon nanotubes,
Mater. Sci. Eng. C. 77 (2017) 572–582. doi:10.1016/j.msec.2017.03.273.

[165] S. Zinadini, S. Rostami, V. Vatanpour, E. Jalilian, Preparation of antibiofouling


polyethersulfone mixed matrix NF membrane using photocatalytic activity of
ZnO/MWCNTs nanocomposite, J. Memb. Sci. 529 (2017) 133–141.
doi:10.1016/j.memsci.2017.01.047.

[166] G. Bounos, K.S. Andrikopoulos, H. Moschopoulou, G.C. Lainioti, D. Roilo, Enhancing


water vapor permeability in mixed matrix polypropylene membranes through carbon
nanotubes dispersion, J. Memb. Sci. 524 (2017) 576–584.
doi:10.1016/j.memsci.2016.11.076.

[167] A. Alpatova, M. Meshref, K.N. Mcphedran, M.G. El-din, Composite polyvinylidene


fluoride (PVDF) membrane impregnated with Fe2O3 nanoparticles and multiwalled
carbon nanotubes for catalytic degradation of organic contaminants, J. Memb. Sci. 490
(2015) 227–235. doi:10.1016/j.memsci.2015.05.001.

[168] J. Farahbakhsh, M. Delnavaz, V. Vatanpour, Investigation of raw and oxidized


multiwalled carbon nanotubes in fabrication of reverse osmosis polyamide membranes for

105
improvement in desalination and antifouling properties, Desalination. 410 (2017) 1–9.
doi:10.1016/j.desal.2017.01.031.

[169] A. V Dudchenko, J. Rolf, K. Russell, W. Duan, D. Jassby, Organic fouling inhibition on


electrically conducting carbon nanotube–polyvinyl alcohol composite ultra filtration
membranes, J. Memb. Sci. 468 (2014) 1–10. doi:10.1016/j.memsci.2014.05.041.

[170] M.R. Mahdavi, M. Delnavaz, V. Vatanpour, Fabrication and water desalination


performance of piperazine–polyamide nanocomposite nanofiltration membranes
embedded with raw and oxidized MWCNTs, J. Taiwan Inst. Chem. Eng. 75 (2017) 189–
198. doi:https://doi.org/10.1016/j.jtice.2017.03.039.

[171] I. Wan Azelee, P.S. Goh, W.J. Lau, A.F. Ismail, Facile acid treatment of multiwalled
carbon nanotube-titania nanotube thin film nanocomposite membrane for reverse osmosis
desalination, J. Clean. Prod. 181 (2018) 517–526.
doi:https://doi.org/10.1016/j.jclepro.2018.01.212.

[172] E. Environ, B. Corry, Water and ion transport through functionalised carbon nanotubes:
implications for desalination technology, Energy Environ. Sci. (2011) 751–759.
doi:10.1039/c0ee00481b.

[173] Y. Ying, W. Ying, Q. Li, D. Meng, G. Ren, R. Yan, X. Peng, Recent advances of
nanomaterial-based membrane for water purification, Appl. Mater. Today. 7 (2017) 144–
158. doi:https://doi.org/10.1016/j.apmt.2017.02.010.

106
Water
Salt

VA-CNT membranes

CNTs

Mixed CNT membranes

SWCNTs MWCNTs

Bucky-paper CNT membrane


Highlights

 A broad overview of synthesis of CNT membranes is presented.

 The potential application of CNT membranes in water purification is critically reviewed.

 Effect of CNT addition on various membrane properties is discussed.

 Current challenges and direction for future research is highlighted.

107

S-ar putea să vă placă și