Skip to content
Publicly Available Published by De Gruyter January 28, 2016

Electrochemical carbon nanotube filters for water and wastewater treatment

  • Sadia A. Jame and Zhi Zhou EMAIL logo
From the journal Nanotechnology Reviews

Abstract

Electrochemically active carbon nanotube (CNT) filters have been developed as a highly efficient technology for water and wastewater treatment during the last few years. CNT filters have been widely used to adsorb chemical and biological contaminants due to their high stability, great flexibility, and large specific surface area. Electrochemically active CNT filters provide additional electrooxidation of the adsorbed contaminants and have been proven to be a highly effective treatment technology in a few recent lab-scale studies. The working principles, impacting factors, and some of the latest development of electrochemically active CNT filters are reviewed in this paper. The existing challenges and future perspectives are also discussed.

1 Introduction

Carbon nanotube (CNT) filters have been extensively studied for water and wastewater treatment during recent years. Owing to the unique electronic, mechanical, and chemical properties, carbon nanotubes (CNTs) received extensive attention since 1990 [1] and are useful for a variety of applications in energy studies and biomedical devices. The high surface area (30–500 m2 g-1) and electrochemical degradation (104~106 S m-1) of CNT filters result in a high adsorption capability for water and wastewater treatment [2–4]. The application of CNTs in water and wastewater treatment has been extensively reviewed previously [5, 6].

The combination of electrochemistry and CNT filters has further expanded the application of CNT filters for water and wastewater treatment [2, 7–18]. Electrochemistry has been widely applied in separation, corrosion control, electroanalytical sensors, electroplating, and fuel cells, as well as in environmental applications, such as wastewater treatment, environmental sensing, and batteries [19]. The environmental application of electrochemistry had gained more and more popularity [7, 20] and has been extensively reviewed in previous studies [21, 22]. Previous studies have demonstrated that CNT filters were effective to remove aqueous organic pollutants, such as salts [23], proteins [24], viruses [14], azo dyes [16], pharmaceuticals [25], perfluorinated chemicals [26], and phenol [8]. In addition, the application of electrochemistry could reduce filter fouling rate by in situ foulant destruction and biological inactivation [17, 23, 27] and reduce maintenance efforts for optimal membrane permeability [28]. A CNT-polyvinylidene fluoride porous non-Faradaic cathode was shown to significantly reduce organic fouling in ultrafiltration (UF) membranes and has potential to reduce energy requirements by up to twofold compared to the unmodified UF system [18]. The electrochemical applications of CNT filters are related with their good electrocatalytic properties for electrochemical reactions. For example, a CNT-based class electrode showed a very low overpotential and higher peak current values than those observed in bare glass electrodes [13]. As a result, CNT-based electrodes are good candidates for environmental applications, such as wastewater treatment and micropollutant sensors [17].

Although the application of CNT filters for adsorption has been extensively studied and reviewed, a detailed review on the working principles and electrochemical processes of electrochemically active CNT filters is not yet available. The working principles and impacting factors of electrochemically active CNT filters are reviewed in this paper, and a few recently developed CNT-based electrochemical treatment techniques are discussed. The existing challenges and future perspectives are also discussed. A comprehensive review of existing literature on electrochemically active CNT filters will help advance our understanding of the advantages and limitations of electrochemically active CNT filters and help develop an efficient and cost-effective technique for water and wastewater treatment.

2 Working principles of electrochemical CNT filters for water and wastewater treatment

CNT filters are composed of tens of hundreds of individual tubes that are entangled with each other through van der Waals forces of attraction [1], which results in a high surface area (30–500 m2 g-1) that can be used to adsorb chemical and biological contaminants. The configuration of a representative electrochemical CNT filter is shown in Figure 1 [16]. Vecitis et al. modified a commercial polycarbonate filtration casing and inserted stainless steel or titanium cathode and a titanium anode ring so a multiwalled nanotube (MWNT) anode can be used for electrooxidation.

Figure 1: A typical electrochemical CNT filter. (A) Design of the modified commercial polycarbonate filtration casing consisting of (1) a perforated stainless steel cathode, (2) an insulating silicone rubber separator and seal, (3) a titanium anodic ring that is pressed into the carbon nanotube anodic filter, and (4) the MWNT anodic filter supported by a PTFE membrane. (B, C) Images of the modified filtration casing. (D, E) Images of the MWNT network before and after electrochemical filtration, respectively [16]. Copyright © 2011 American Chemical Society.
Figure 1:

A typical electrochemical CNT filter. (A) Design of the modified commercial polycarbonate filtration casing consisting of (1) a perforated stainless steel cathode, (2) an insulating silicone rubber separator and seal, (3) a titanium anodic ring that is pressed into the carbon nanotube anodic filter, and (4) the MWNT anodic filter supported by a PTFE membrane. (B, C) Images of the modified filtration casing. (D, E) Images of the MWNT network before and after electrochemical filtration, respectively [16]. Copyright © 2011 American Chemical Society.

The overall electrochemical filtration includes three processes: (1) hydrodynamically enhanced mass transfer, (2) temperature-dependent physical adsorption/desorption, and (3) voltage-dependent direct electron transfer [16]. Electrochemical oxidation for organic contaminants takes much less time than that of conventional biological wastewater treatment systems [29]. Furthermore, recalcitrant compounds that are often difficult to be treated can be electrochemically oxidized [22]. CNT filters have been widely used for adsorption of various chemical contaminants [30, 31], but previous studies also indicated that breakthrough of CNT filters happened in few hours in fix bed columns [32] or <30 min in a dead end filter [2] after CNT active sites were occupied with adsorbed contaminants. Electrochemical CNT filters can further break down adsorbed contaminants and, therefore, significantly improve their performance in wastewater treatment. Electrochemical CNT filters have been successfully applied to treat various contaminants, and the results of CNT-based electrochemical treatment systems are summarized in Table 1. The results indicated that CNT-based electrochemical treatment systems were highly efficient to treat different chemical and biological contaminants with a typical removal efficiency >90%.

Table 1

Removal efficiencies of CNT-based electrochemical treatment systems and adsorption systems.

ContaminantsAnodeCathodeRemoval efficiencySystem configurationReferences
PhenolSnO2-Sb2O4-CNTStainless steel90%Batch[33]
C.I. basic yellow 2FeCNT-PTFE96%Batch[34]
Methyl orangeSb-doped SnO2 CNTStainless steel95%Filtration[8]
C.I. basic blue 3PtCNT-PTFE94%Batch[35]
4-ChlorophenolLAS-CNT-PbO2Stainless steel99.5%Batch[36]
Methylene blueCNTCNT98%Filtration[16]
E. coliCNTStainless steel75%Filtration[17]
MS2CNTStainless steel99.6%Filtration[17]
MS2CNTStainless steel>99.9%Filtration[14]
m-CresolTi/SnO2-Sb2O5-IrO2CNT82%Batch[37]
C.I. direct red 23PtCNT-PTFE94%Batch[38]
OxalateBTO-CNTTi98%Filtration[39]
EthanolBTO-CNTTi85.7%Filtration[39]
MethanolBTO-CNTTi88.7%Filtration[39]
FormaldehydeBTO-CNTTi33.5%Filtration[39]
FormateBTO-CNTTi25%Filtration[39]
FerrocyanideCNTTi>90%Filtration[15]
p-NitrophenolCNT-Ce-PbO2Stainless steel98.2%Batch[40]
ChromiumCarbon paperCNT96%Batch[41]
TetracyclineGraphene-CNTTi88%Filtration[42]
PhenolGraphene-CNTTi93%Filtration[42]
OxalateGraphene-CNTTi87%Filtration[42]
PhenolCNTCNT87.0%Filtration[43]
TetracyclineCNTCNT90.3%Filtration[43]
Methyl orangeCNTCNT96.0%Filtration[43]
GeosminCNTCNT87.2%Filtration[43]
TetracyclineCNTCNT99%Filtration[2]

3 Impacting factors of electrochemical CNT filters

The treatment efficiency of electrochemical CNT filters may be affected by a number of factors and the main impacting factors are discussed as follows.

3.1 Anode potential

Anode potential is one of the most important impacting factors for electrooxidation with CNT filters. Anode potential needs to be higher than the redox potential for a chemical to be oxidized. However, water splitting may also happen at a high anode potential, and there is a direct competition between organic oxidation and oxygen evolution (Eq. 1).

(1)1/2H2O1/4O2+H++e-, E0=+1.03 V vs. Ag/AgCl (1)

The potentials of oxygen evolution of different anodes, such as Pt, IrO2, PbO2, and SnO2, have been reviewed previously [21]. Anode materials with low oxygen evolution potentials are not efficient for electrochemical treatment of pollutants because competition with oxygen evolution results in low current efficiencies. High anode potential may also result in CNT degradation [44] and limits the applications of CNTs for electrochemical treatment of water and wastewater. To minimize the competition and potential CNT degradation, anode potentials should not be too high, and some high oxygen-overvoltage anode materials, such as a boron-doped diamond (BDD) electrode, could be used to minimize the competition between anoxic oxidation and oxygen evolution [45–47], although these materials are generally more expensive than CNTs.

Surface coating is one technique to improve the overpotential of oxygen evolution for electrochemical applications. In a recent study, Liu et al. fabricated a binder-free, porous, and conductive 3D CNT network uniformly coated with bismuth-doped tin oxide (BTO) nanoparticles via a simple electrosorption-hydrothermal method [39]. The BTO layer is composed of 3.9±1.5-nm-diameter nanoparticles, and further testing of BTO-coated CNT filters showed that anode stability and efficiency for flow-through organic electrooxidation were improved in BTO-coated CNT filters [39]. Around 98% of oxalate was oxidized in BTO-coated CNT filters with relatively high current efficiencies in the range of 32% to >99%. The overpotential of oxygen evolution in the BTO-CNT nanocomposite was determined to be 1.71 V (vs. Ag/AgCl), which is 0.44 V higher than that of an uncoated CNT anode, and therefore, anodic oxidation of organic contaminants on BTO-CNT surface can take place with less competition of oxygen evolution, and anodic stability of CNT filters was also improved [39]. Additional experiments on the electrooxidative filtration of ethanol, methanol, formaldehyde, and formate showed that BTO-CNT anode displayed two to eight times greater TOC removal efficiencies, 1.5–3.5 times greater mineralization current efficiencies, and four to five times less energy consumption than those of the uncoated CNT anode [39].

The effect of applied potentials on anoxic oxidation has been reported in previous studies. Under a total applied potential of 3 volts (V), MS2 virus with an influent concentration of 106 ml-1 was removed below detection limit in the effluent, and viability of Escherichia coli was lost by 99% after electrochemical CNT filtration [17]. Although anode potentials were not reported in these studies, the results suggest that higher anode potentials result in higher anoxic oxidation when total applied potential is below 3 V. Further electrooxidation experiments on ferroxyanide oxidation confirmed this trend, and anoxic oxidation rate increased under higher anode potentials [15, 42]. Current densities were observed to increase with anode potentials [13, 39].

Anode potentials are also related with CNT reactive sites. The sp2-conjugated sidewall CNT sites are dominant when anode potentials are low (≤0.2 V), and CNT tips were also electroactive when anode potentials are high (≥0.3 V). A quantitative 2D electrooxidative CNT filter model has been developed to describe the correlation among target molecule, mass transport, adsorption, and electron transfer and product desorption, and this model has been used to accurately predict effluent concentrations over a much larger range of conditions [48].

3.2 Flow rate

Flow rate is another factor that may affect the electrooxidation rate of CNT filters, and it is related with reaction kinetics. A positive correlation between ferroxyanide electrooxidation and flow rate up to 4 ml min-1 was observed when mass transfer dominated the electrooxidation process at an anode potential of 0.3 V [15]. When the anode potential was below 0.3 V, the electrochemical CNT filtration was under mixed mass transfer and reaction control [15]. Under the typical configuration of CNT filters, electrooxidation rate is mass transfer limited by the convective flow rate through the electrode and subsequent replenishment of the target molecule, but not by the diffusional mass transfer of ferroxyanide from the bulk to the electrode surface [15]. Current efficiency was found to increase slightly with increasing flow rate from J=0.5–4.0 ml min-1 and reached a plateau at the higher flow rates, which was likely due to increased mass transfer to the anode surface with increased flow rates [15].

3.3 Ionic strength

The electrooxidation rate may be affected by ionic strength in the solution. A comparison of three different concentrations (1, 10, and 100 mm) of electrolyte (NaSO4) in the electrochemical CNT filtration system indicated that the Na2SO4 concentration had minimal effect on the ferrocyanide oxidation [15]. Such a result was related with the electrooxidation mechanism of ferrocyanide oxidation, which is a direct electron transfer [8, 10] instead of reactions with other oxidizing species that may be directly affected by ionic strength [15]. For the electrochemical systems where indirect oxidation of the target species in the bulk solution was the primary mechanism, the primary chemical oxidant is a Cl-radical species [15]. Reactor configuration may also affect the effects of ionic strength on electrooxidation in CNT filters. The constant convective renewal of the solution near the CNT surface in the flow-through configuration resulted in faster charge neutralization than the batch system [15], and the hydrodynamic reductions of charge transfer resistances enabled flow-through electrochemical CNT filtration to be a relative robust treatment technology that is less sensitive to the change of ionic strength.

3.4 Cathode material

Many materials have been used in electrochemical systems, such as graphite felt [49], carbon felt [50], and MWNT [17], and the performance of an electrochemical system is directly related with cathode materials because different cathodes resulted in significantly different extents of oxidation under the electron transfer-limited regime. Only 50% of influent ferrocyanide was oxidized at 0.3-V anode potential with Ti cathode, while over 80% of the influent ferrocyanide was oxidized under the same conditions with the CNT cathode [15]. Coating electrodes with other materials may improve the performance of electrochemical CNT filtration system [39] or electrocatalytic performance in other treatment systems [51].

4 Latest development on electrochemical CNT filters

A few recently developed CNT-based electrochemical treatment techniques are also discussed to improve the overall performance of electrochemical CNT filters.

4.1 Surface modification and functionalization of CNT filters

The surface of CNT filters can be modified to improve their functions and performance through covalent attachment of functional groups or the non-covalent adsorption of functional molecules onto the surface of CNTs [52]. Although there are many previous studies on the surface modification and functionalization of CNT filters, most of them focus on electroanalytical applications [53], adsorption [54], sensors [55–57], capacitor devices [58, 59], thermal stability [60], solubility [61], and only limited studies have been done to functionalize CNT filters for electrochemical treatment. Mishra et al. fabricated a new type of flexible carbon fabric-supported MWNT (Fe3O4-MWNTs) nanocomposite-based supercapacitor using a chemical vapor deposition (CVD) technique and applied it for arsenic removal and seawater desalination [62]. The Fe3O4-MWNTs nanocomposite-based electrodes were used to remove both types of arsenic ions (arsenate and arsenite) and salts from seawater and achieved high desalination efficiency (removal of sodium, magnesium, and calcium) of seawater and high repeatability for the simultaneous removal of arsenic (both arsenate and arsenite) and sodium [62]. In addition to the surface coating of CNTs with BTO nanoparticles [39], Liu et al. also coated CNT filters with titanium dioxide (TiO2) via a simple filtration-steam hydrolysis method and evaluated its performance for arsenic removal [63]. The TiO2-coated CNT filters significantly enhanced arsenic sorption kinetics and capacity, which is possibly due to increased mass transport, improved sorption site accessibility, and reduced TiO2 negative surface charge. The TiO2-CNT arsenic sorption kinetics also increased with both increasing flow rate and cell potential. The spent TiO2 filter was successfully regenerated by 5 mm of NaOH [63].

4.2 Anode oxidation coupled with in situ generated H2O2 in the CNT cathode

Within a typical electrochemical CNT filtration system, organic pollutants were only adsorbed and oxidized via a direct/indirect oxidation process on the anodic CNT filters [13], and a Ti ring or CNT filters were used as a counter cathode to provide the required potential [15, 16]. However, the role of a cathode in electrochemical filters beyond a counter electrode has been largely neglected. A typical cathode provides electrons and only supports reduction instead of oxidation, and therefore, it cannot be directly used to oxidize organic pollutants in wastewater, which are mostly reduced hydrocarbons. A recent study utilized the cathode in a CNT filtration system to produce hydrogen peroxide (H2O2) for further oxidation of chemical contaminants [43]. As CNT filers can work as an oxygen reduction reaction catalyst [64–66], a CNT cathode could be used to reduce oxygen to generate H2O2 (Eq. 2) with the counter electrode serving as a functional cathode [67–70].

(2)1/2O2+H++e-1/2H2O2, E0=+0.45 V vs. Ag/AgCl (2)

As a strong oxidant (E0=1.763 V vs. SHE), H2O2 can oxidize various organic pollutants [67, 71, 72] and produce oxygen and water as by-products after oxidation [73–76]. A schematic of an electrochemical CNT filter coupled with in situ generated H2O2 is shown in Figure 2.

Figure 2: Schematic of an electrochemical CNT filter coupled with in situ generated H2O2 [43]. Copyright © 2015 Royal Society of Chemistry.
Figure 2:

Schematic of an electrochemical CNT filter coupled with in situ generated H2O2 [43]. Copyright © 2015 Royal Society of Chemistry.

Under optimized conditions (C-CNT-HCl, applied cathode potential=-0.4 V vs. Ag/AgCl, influent DO flux=1.95 mol l-1 m-2, and [Na2SO4]=10 mmol l-1, flow rate=1.5 ml min-1), H2O2 flux increased linearly with increased flow rates, indicating that the filtration system was under mixed mass transfer and oxygen reduction reaction control. Contrarily, at medium flow rate conditions (1.5–4.0 ml min-1), the system became mass transfer-limited, i.e. the electroreduction rate of O2 was limited by the flow rate of the influent Na2SO4 solution throughout the cathode and subsequent replenishment of O2 to produce H2O2. This was further confirmed by the stable effluent DO of 6.9±0.1 mg l-1 and the stable DO efficiency of 58±2.6%. High flow rates above 4 ml min-1 may be detrimental to the reduction kinetics due to greatly increased pressure within the current filtration casing, which needs delicate engineering design to prevent such a negative effect [15].

H2O2 flux increased with higher initial DO flux, and the maximum H2O2 flux was obtained at an applied cathode potential of -0.4 V (vs. Ag/AgCl) and an O2 flux of 1.95 mol l-1 m-2, which was 1.7-, 2.8-, and 46.8-fold higher than those at the DO flux of 1.33, 0.66, and 0 mol l-1 m-2, respectively. However, the maximum current efficiency for H2O2 generation was only 52.9±1.2%, suggesting the existence of H2O2 decomposition processes or other electron-consuming reactions. For example, H2O2 could undergo chemical decomposition to O2 either on the anode (heterogeneous process) or in the medium (homogeneous process). Reduction from H2O2 to OH- on the CNT cathodes could also consume H2O2 and electrons [77, 78]. Additionally, H2O2 could be consumed indirectly by other reactive species, such as OH˙ to produce HO2˙ at the anode [79] and eventually decomposed into O2 [67, 77]. The electrochemical filter coupled with in situ generated H2O2 was further tested with three additional organic compounds: tetracycline (a typical PPCP), methyl orange (a typical azo-dye), and geosmin (a typical off-flavor compound), and all three organic compounds were efficiently removed with removal efficiencies of 90.3%, 96.0%, and 87.2% for 0.1 mmol l-1 tetracycline, 0.1 mmol l-1 methyl orange, and 0.55 nmol l-1 geosmin, respectively [43].

4.3 Electro-Fenton

A CNT-based electro-Fenton system was developed to further utilize H2O2 to produce a Fenton agent to improve the treatment efficiency of CNT filtration systems [12]. The system is composed of a CNT cathode, a CNT-COOFen+ cathode, a PTFE separator, and a CNT anode. The schematic of the CNT-based electro-Fenton system is shown in Figure 3.

Figure 3: Sandwiched electro-Fenton system based on carbon nanotube membrane stacks. (A) Images of the unfolded sandwich membrane stacks including four layers and (B) schematic of main roles of every layer in membrane stacks, and [P] and [P]m are pollutants and their oxidation intermediates, respectively [12]. Copyright © 2015 American Chemical Society.
Figure 3:

Sandwiched electro-Fenton system based on carbon nanotube membrane stacks. (A) Images of the unfolded sandwich membrane stacks including four layers and (B) schematic of main roles of every layer in membrane stacks, and [P] and [P]m are pollutants and their oxidation intermediates, respectively [12]. Copyright © 2015 American Chemical Society.

In this system, O2 is reduced on the CNT cathode to produce H2O2 first, and H2O2 further reacts with Fe2+ to produce hydroxyl radical (OH˙), which is a strong oxidant and can be used to oxidize contaminants (Eq. 3). The produced Fe3+ will be regenerated in situ in the cathode (Eq. 4), and Fe2+ is reused in the following reactions. The remaining intermediates or metabolites are further oxidized in the CNT anode. As H2O2 is generated on site, the cost and risk associated with H2O2 storage and transportation can be reduced. The system is operated at neutral pH, and therefore, the acid/base required to keep Fe dissolved during the reaction is not needed. Fe is bound to the electrode, and an in situ electroregeneration of Fe2+ reduced the need for the addition of Fe and the disposal of iron sludge.

(3)H2O2+Fe2++H+Fe3++H2O+OH· (3)
(4)Fe3++e-Fe2+ (4)

The results showed that the sequential electro-Fenton oxidation was a highly efficient electrochemical system for contaminant removal. The removal rate for oxalate was 206.8±6.3 mgC m-2 h-1, which is fourfold greater than the sum of the individual electrochemistry (16.4±3.2 mgC m-2 h-1) and Fenton (33.3±1.3 mgC m-2 h-1) reaction fluxes. The removal rates for refractory trifluoroacetic acid (TFA) and trichloroacetic acid (TCA) also reached 11.3±1.2 and 21.8±1.9 mmol m-2 h-1, respectively [12].

5 Energy consumption of electrochemical CNT filters

Energy consumption is one of the most important factors for building and operating water and wastewater treatment systems. The energy consumption of a typical electrochemical CNT filtration system was calculated as 4 kW h kg-1 COD at a total cell potential of 2 V and 15 kW h kg-1 COD at a total cell potential of 3 V for (C-CNT-HNO3 coated with Sb-doped SnO2 particles) [8]. The energy consumption for electrochemical filtration system coupled with in situ generated H2O2 was calculated as 3.75 kW h kg-1 COD at a total cell potential of 1.85 V (corresponding to an optimized cathode potential of -0.4 V vs. Ag/AgCl) and a current of 9.05 mA [43]. Additionally, the pumping energy was considered to pump the liquid solution through the filter. At a common back pressure of 15 kPa [8], a flow rate of 1.5 ml min-1, and a pump efficiency of 75%, the total energy cost for pumping was calculated as 1.35 J [13], which is only 2.2% of the energy used for electrochemical H2O2 production [43]. The energy consumption for the flow-through sequential regenerative electro-Fenton system was calculated as 45.8 kW h kg-1 TOC [12]. These values are lower than those in state-of-the-art electrochemical oxidation processes whose energy consumptions are typically in the range of 5–100 kW h kg-1 COD [20]. Generally speaking, the energy consumption of electrochemical carbon nanotube filters is low. Although there was additional electrical energy input in the filtration system, the benefits of low applied potentials, high oxidation kinetics, short residence time, and long service time can compensate the additional energy input. The low potential could be provided by a solar panel so electrochemical CNT filters may be widely used as a cost-effective point-of-use treatment system in rural areas of developing countries.

6 Existing challenges

Although electrochemical CNT filters have been proven to be an efficient treatment technology, there still remain a few challenges before such a technology can be widely used in large-scale water and wastewater treatment plants. Although the cost of CNTs has decreased significantly during the last few years, the cost of CNTs are still relatively high than those of other commonly used water treatment materials, such as activate carbon. Porous CNT sheets in larger sizes are also expensive to fabricate. Another issue is bubble formation when the applied voltage is higher than 2 V, which is caused by the continuous production of hydrogen in the cathode and oxygen in the anode. A sample picture is shown in Figure 4. Bubble formation may block the active sites of CNT filters, increase back pressure, reduce mass transfer efficiency, and eventually reduce the efficiency of electrochemical CNT filters.

Figure 4: Bubble formation after electrochemical CNT filtration [42]. Copyright © 2014 Royal Society of Chemistry.
Figure 4:

Bubble formation after electrochemical CNT filtration [42]. Copyright © 2014 Royal Society of Chemistry.

The CNT surface could be changed after filtration, and a sample SEM image of the CNT surface is shown in Figure 5. After electrochemical filtration for 3 h, the anodic CNT filter showed some extent of organic, possibly polymer, and inorganic, possibly sodium persulfate, buildup on the CNT surface that may increase resistance to electron transfer and hinder electrochemistry [2].

Figure 5: SEM image of CNT surface after filtration. The white circles and squares showed some extent of inorganic or organic build upon the CNT surface, respectively. Scale bar is 2 μm [2]. Copyright © 2015 American Chemical Society.
Figure 5:

SEM image of CNT surface after filtration. The white circles and squares showed some extent of inorganic or organic build upon the CNT surface, respectively. Scale bar is 2 μm [2]. Copyright © 2015 American Chemical Society.

The existing configuration of dead-end CNT filtration system may limit the overall water flux to be treated, but pilot-scale and full-scale cross-flow CNT filtration systems have not been reported in the literature yet. CNT filters are relatively fragile and can be easily torn apart and fractured. CNT agglomeration is another issue after an extended period of operation and is more obvious under applied potentials higher than 2.5 V. The last but not the least, metabolites can be produced after partial oxidation, and the toxicity of these metabolites are often difficult to characterize, and therefore, their chemical characteristics and toxicity are not well understood.

7 Conclusions and perspectives

Electrochemically active CNT filters have been developed as a highly effective treatment technology for water and wastewater treatment and can be used to treat various chemical and biological contaminants due to its high stability, great flexibility, and large specific surface area. With the continuous reduced costs of CNTs, electrochemical CNT filters could be widely adopted for water purification. Surface modification and functionalization of CNT filters will further improve their applications and treatment efficiency. Fouling control in these electrochemical CNT filters is another area with great potentials, and minimization of membrane fouling will greatly help the development of cost-effective water treatment technologies. Cross-flow configuration provides higher fluxes than that of dead-end filtration, and the development of cross-flow electrochemical CNT systems will build a foundation for scaled-up application of this techniques in pilot-scale and full-scale water and wastewater treatment plants. Because of the simplicity of the reactor configuration, this technology can be easily adopted by developing countries for point-of-use applications as no expensive treatment facilities are needed. The potentials of using these techniques to treat both chemical and biological contaminants can be further explored to provide flexibility to deal with different contamination in water sources. Coupled with solar panels to provide the low applied potentials, electrochemical CNT filters could be one of the most cost-effective treatment technologies to meet ever-increasing needs for adequate clean water.


Corresponding author: Zhi Zhou, Division of Environmental and Ecological Engineering and School of Civil Engineering, Purdue University, 550 Stadium Mall Drive, West Lafayette, IN 47907, USA, e-mail:

Acknowledgments

This work is financially supported by Purdue University.

References

[1] Balasubramanian K, Burghard M. Chemically functionalized carbon nanotubes. Small 2005, 1, 180–192.10.1002/smll.200400118Search in Google Scholar PubMed

[2] Liu YB, Liu H, Zhou Z, Wang TR, Ong CN, Vecitis CD. Degradation of the common aqueous antibiotic tetracycline using a carbon nanotube electrochemical filter. Environ. Sci. Technol. 2015, 49, 7974–7980.Search in Google Scholar

[3] Nardecchia S, Carriazo D, Ferrer ML, Gutiérrez MC, Del Monte F. Three dimensional macroporous architectures and aerogels built of carbon nanotubes and/or graphene: synthesis and applications. Chem. Soc. Rev. 2013, 42, 794–830.Search in Google Scholar

[4] Ryu Y, Yin L, Yu C. Dramatic electrical conductivity improvement of carbon nanotube networks by simultaneous de-bundling and hole-doping with chlorosulfonic acid. J. Mater. Chem. 2012, 22, 6959–6964.Search in Google Scholar

[5] Chaplin BP. Critical review of electrochemical advanced oxidation processes for water treatment applications. Env. Sci. Process. Impact. 2014, 16, 1182–1203.10.1039/C3EM00679DSearch in Google Scholar PubMed

[6] Qu X, Alvarez PJJ, Li Q. Applications of nanotechnology in water and wastewater treatment. Water Res. 2013, 47, 3931–3946.Search in Google Scholar

[7] Brady-Estévez AS, Schnoor MH, Vecitis CD, Saleh NB, Elimelech M. Multiwalled carbon nanotube filter: improving viral removal at low pressure. Langmuir 2010, 26, 14975–14982.10.1021/la102783vSearch in Google Scholar PubMed

[8] Gao GD, Vecitis CD. Electrochemical carbon nanotube filter oxidative performance as a function of surface chemistry. Environ. Sci. Technol. 2011, 45, 9726–9734.Search in Google Scholar

[9] Gao G, Vecitis CD. Doped carbon nanotube networks for electrochemical filtration of aqueous phenol: electrolyte precipitation and phenol polymerization. ACS Appl. Mater. Inter. 2012, 4, 1478–1489.Search in Google Scholar

[10] Gao G, Vecitis CD. Reactive depth and performance of an electrochemical carbon nanotube network as a function of mass transport. ACS Appl. Mater. Inter. 2012, 4, 6096–6103.Search in Google Scholar

[11] Gao G, Vecitis CD. Electrocatalysis aqueous phenol with carbon nanotubes networks as anodes: electrodes passivation and regeneration and prevention. Electrochim. Acta 2013, 98, 131–138.10.1016/j.electacta.2013.02.127Search in Google Scholar

[12] Gao G, Zhang Q, Hao Z, Vecitis CD. Carbon nanotube membrane stack for flow-through sequential regenerative electro-Fenton. Environ. Sci. Technol. 2015, 49, 2375–2383.Search in Google Scholar

[13] Liu H, Vecitis CD. Reactive transport mechanism for organic oxidation during electrochemical filtration: mass-transfer, physical adsorption, and electron-transfer. J. Phys. Chem. C 2012, 116, 374–383.10.1021/jp209390bSearch in Google Scholar

[14] Rahaman MS, Vecitis CD, Elimelech M. Electrochemical carbon-nanotube filter performance toward virus removal and inactivation in the presence of natural organic matter. Environ. Sci. Technol. 2012, 46, 1556–1564.Search in Google Scholar

[15] Schnoor MH, Vecitis CD. Quantitative examination of aqueous ferrocyanide oxidation in a carbon nanotube electrochemical filter: effects of flow rate, ionic strength, and cathode material. J. Phys. Chem. C 2013, 117, 2855–2867.10.1021/jp3112099Search in Google Scholar

[16] Vecitis CD, Gao G, Liu H. Electrochemical carbon nanotube filter for adsorption, desorption, and oxidation of aqueous dyes and anions. J. Phys. Chem. C 2011, 115, 3621–3629.10.1021/jp111844jSearch in Google Scholar

[17] Vecitis CD, Schnoor MH, Rahaman MS, Schiffman JD, Elimelech M. Electrochemical multiwalled carbon nanotube filter for viral and bacterial removal and inactivation. Environ. Sci. Technol. 2011, 45 3672–3679.10.1021/es2000062Search in Google Scholar PubMed

[18] Zhang Q, Vecitis CD. Conductive CNT-PVDF membrane for capacitive organic fouling reduction. J. Membrane Sci. 2014, 459, 143–156.Search in Google Scholar

[19] Rajeshwar K, Ibanez JG, Swain GM. Electrochemistry and the environment. J. Appl. Electrochem. 1994, 24, 1077–1091.Search in Google Scholar

[20] Panizza M, Cerisola G. Direct and mediated anodic oxidation of organic pollutants. Chem. Rev. 2009, 109, 6541–6569.Search in Google Scholar

[21] Chen GH. Electrochemical technologies in wastewater treatment. Sep. Purif. Technol. 2004, 38, 11–41.10.1016/j.seppur.2003.10.006Search in Google Scholar

[22] Martinez-Huitle CA, Ferro S. Electrochemical oxidation of organic pollutants for the wastewater treatment: direct and indirect processes. Chem. Soc. Rev. 2006, 35, 1324–1340.Search in Google Scholar

[23] De Lannoy CF, Jassby D, Gloe K, Gordon AD, Wiesner MR. Aquatic biofouling prevention by electrically charged nanocomposite polymer thin film membranes. Environ. Sci. Technol. 2013, 47 2760–2768.10.1021/es3045168Search in Google Scholar PubMed

[24] Sun X, Wu J, Chen Z, Su X, Hinds BJ. Fouling characteristics and electrochemical recovery of carbon nanotube membranes. Adv. Funct. Mater. 2013, 23, 1500–1506.Search in Google Scholar

[25] Li H, Zhang D, Han X, Xing B. Adsorption of antibiotic ciprofloxacin on carbon nanotubes: pH dependence and thermodynamics. Chemosphere 2014, 95, 150–155.10.1016/j.chemosphere.2013.08.053Search in Google Scholar PubMed

[26] Deng S, Zhang Q, Nie Y, Wei H, Wang B, Huang J, Yu G, Xing B. Sorption mechanisms of perfluorinated compounds on carbon nanotubes. Environ. Pollut. 2012, 168, 138–144.Search in Google Scholar

[27] Ajmani GS, Goodwin D, Marsh K, Fairbrother DH, Schwab KJ, Jacangelo JG, Huang H. Modification of low pressure membranes with carbon nanotube layers for fouling control. Water Res. 2012, 46, 5645–5654.Search in Google Scholar

[28] Boo C, Elimelech M, Hong S. Fouling control in a forward osmosis process integrating seawater desalination and wastewater reclamation. J. Membrane Sci. 2013, 444, 148–156.Search in Google Scholar

[29] Szpyrkowicz L, Kaul S, Neti R. Tannery wastewater treatment by electro-oxidation coupled with a biological process. J. Appl. Electrochem. 2005, 35, 381–390.Search in Google Scholar

[30] Liu X, Wang M, Zhang S, Pan B. Application potential of carbon nanotubes in water treatment: a review. J. Environ. Sci. 2013, 25, 1263–1280.Search in Google Scholar

[31] Pan B, Xing B. Adsorption mechanisms of organic chemicals on carbon nanotubes. Environ. Sci. Technol. 2008, 42, 9005–9013.Search in Google Scholar

[32] Dichiara AB, Harlander SF, Rogers RE. Fixed bed adsorption of diquat dibromide from aqueous solution using carbon nanotubes. RSC Advances 2015, 5, 61508–61512.10.1039/C5RA11167FSearch in Google Scholar

[33] Hu FP, Cui XW, Chen WX. Pulse electro-codeposition of Ti/SnO2-Sb2O4-CNT electrode for phenol oxidation. Electrochem. Solid St. 2010, 13, F20–F23.Search in Google Scholar

[34] Zarei M, Niaei A, Salari D, Khataee AR. Removal of four dyes from aqueous medium by the peroxi-coagulation method using carbon nanotube-PTFE cathode and neural network modeling. J. Electroanal. Chem. 2010, 639, 167–174.Search in Google Scholar

[35] Khataee AR, Zarei M, Khataee AR. Electrochemical treatment of dye solution by oxalate catalyzed photoelectro-Fenton process using a carbon nanotube-PTFE cathode: optimization by central composite design. Clean-Soil Air Water 2011, 39, 482–490.10.1002/clen.201000120Search in Google Scholar

[36] Duan XY, Ma F, Yuan ZX, Chang LM, Jin XT. Lauryl benzene sulfonic acid sodium-carbon nanotube-modified PbO2 electrode for the degradation of 4-chlorophenol. Electrochim. Acta 2012, 76, 333–343.10.1016/j.electacta.2012.05.036Search in Google Scholar

[37] Chu YY, Zhang DM, Liu L, Qian Y, Li LL. Electrochemical degradation of m-cresol using porous carbon-nanotube-containing cathode and Ti/SnO2-Sb2O5-IrO2 anode: kinetics, byproducts and biodegradability. J. Hazard. Mater. 2013, 252, 306–312.Search in Google Scholar

[38] Khataee AR, Vahid B, Behjati B, Safarpour M. Treatment of a dye solution using photoelectro-Fenton process on the cathode containing carbon nanotubes under recirculation mode: investigation of operational parameters and artificial neural network modeling. Environ. Prog. Sustain. 2013, 32, 557–563.Search in Google Scholar

[39] Liu H, Vajpayee A, Vecitis CD. Bismuth-doped tin oxide-coated carbon nanotube network: improved anode stability and efficiency for flow-through organic electrooxidation. ACS Appl. Mater. Inter. 2013, 5, 10054–10066.Search in Google Scholar

[40] Duan XY, Zhao YY, Liu W, Chang LM, Li X. Electrochemical degradation of p-nitrophenol on carbon nanotube and Ce-modified-PbO2 electrode. J. Taiwan Inst. Chem. E 2014, 45, 2975–2985.10.1016/j.jtice.2014.08.031Search in Google Scholar

[41] Wang HT, Na CZ. Binder-free carbon nanotube electrode for electrochemical removal of chromium. ACS Appl. Mater. Inter. 2014, 6, 20309–20316.Search in Google Scholar

[42] Liu Y, Dustin Lee JH, Xia Q, Ma Y, Yu Y, Lanry Yung LY, Xie J, Ong CN, Vecitis CD, Zhou Z. A graphene-based electrochemical filter for water purification. J. Mater. Chem. A 2014, 2, 16554–16562.10.1039/C4TA04006FSearch in Google Scholar

[43] Liu Y, Xie J, Ong CN, Vecitis CD, Zhou Z. Electrochemical wastewater treatment with carbon nanotube filters coupled with in situ generated H2O2. Environ. Sci.: Water Res. Technol. 2015, 1, 769–778.Search in Google Scholar

[44] Ohmori S, Saito T. Electrochemical durability of single-wall carbon nanotube electrode against anodic oxidation in water. Carbon 2012, 50, 4932–4938.10.1016/j.carbon.2012.06.023Search in Google Scholar

[45] Daghrir R, Drogui P, Tshibangu J, Delegan N, El Khakani MA. Electrochemical treatment of domestic wastewater using boron-doped diamond and nanostructured amorphous carbon electrodes. Environ. Sci. Pollut. R 2014, 21, 6578–6589.10.1007/s11356-014-2558-8Search in Google Scholar PubMed

[46] Garcia-Segura S, Keller J, Brillas E, Radjenovic J. Removal of organic contaminants from secondary effluent by anodic oxidation with a boron-doped diamond anode as tertiary treatment. J. Hazard. Mater. 2015, 283, 551–557.Search in Google Scholar

[47] Souza FL, Teodoro TQ, Vasconcelos VM, Migliorini FL, Lima Gomes PC, Ferreira NG, Baldan MR, Haiduke RL, Lanza MR. Electrochemical oxidation of imazapyr with BDD electrode in titanium substrate. Chemosphere 2014, 117, 596–603.10.1016/j.chemosphere.2014.09.051Search in Google Scholar PubMed

[48] Liu H, Liu J, Liu YB, Bertoldi K, Vecitis CD. Quantitative 2D electrooxidative carbon nanotube filter model: insight into reactive sites. Carbon 2014, 80, 651–664.10.1016/j.carbon.2014.09.009Search in Google Scholar

[49] Abda M, Oren Y, Soffer A. The electrodeposition of trace metallic impurities: dependence on the supporting electrolyte concentration – a comparison between bipolar and monopolar porous electrodes. Electrochim. Acta 1987, 32, 1113–1115.10.1016/0013-4686(87)90040-5Search in Google Scholar

[50] Delanghe B, Tellier S, Astruc M. The carbon-felt flow-through electrode in waste water treatment: the case of mercury (II) electrodeposition. Environ. Technol. 1990, 11, 999–1006.Search in Google Scholar

[51] Duan TG, Wen Q, Chen Y, Zhou YD, Duan Y. Enhancing electrocatalytic performance of Sb-doped SnO2 electrode by compositing nitrogen-doped graphene nanosheets. J. Hazard. Mater. 2014, 280, 304–314.Search in Google Scholar

[52] Meng LJ, Fu CL, Lu QH. Advanced technology for functionalization of carbon nanotubes. Prog. Nat. Sci. 2009, 19, 801–810.Search in Google Scholar

[53] Pifferi V, Cappelletti G, Di Bari C, Meroni D, Spadavecchia F, Falciola L. Multi-walled carbon nanotubes (MWCNTs) modified electrodes: effect of purification and functionalization on the electroanalytical performances. Electrochim. Acta 2014, 146, 403–410.10.1016/j.electacta.2014.09.099Search in Google Scholar

[54] Hussain S, Amade R, Jover E, Bertran E. Growth and plasma functionalization of carbon nanotubes. J. Clust. Sci. 2015, 26, 315–336.Search in Google Scholar

[55] Han C, Doepke A, Cho W, Likodimos V, de la Cruz AA, Back T, Heineman WR, Halsall HB, Shanov VN, Schulz MJ, Falaras P, Dionysiou DD. A multiwalled-carbon-nanotube-based biosensor for monitoring microcystin-LR in sources of drinking water supplies. Adv. Funct. Mater. 2013, 23, 1807–1816.Search in Google Scholar

[56] Majzlikova P, Prasek J, Elias M, Jasek O, Pekarek J, Hubalek J, Zajickova L. Comparison of different modifications of screen-printed working electrodes of electrochemical sensors using carbon nanotubes and plasma treatment. Phys. Status Solidi A 2014, 211, 2756–2764.10.1002/pssa.201431438Search in Google Scholar

[57] Wu ZY, Xu YY, Zhang XL, Shen GL, Yu RQ. Microwave plasma treated carbon nanotubes and their electrochemical biosensing application. Talanta 2007, 72, 1336–1341.10.1016/j.talanta.2007.01.052Search in Google Scholar PubMed

[58] Hussain S, Amade R, Jover E, Bertran E. Functionalization of carbon nanotubes by water plasma. Nanotechnology 2012, 23, 385604.10.1088/0957-4484/23/38/385604Search in Google Scholar PubMed

[59] Hussain S, Amade R, Jover E, Bertran E. Water plasma functionalized CNTs/MnO2 composites for supercapacitors. Sci. World J. 2013, 19, 832581.Search in Google Scholar

[60] Scheibe B, Borowiak-Palen E, Kalenczuk RJ. Enhancement of thermal stability of multiwalled carbon nanotubes via different silanization routes. J. Alloy Compd. 2010, 500, 117–124.Search in Google Scholar

[61] Kakade BA, Pillai VK. An efficient route towards the covalent functionalization of single walled carbon nanotubes. Appl. Surf. Sci. 2008, 254, 4936–4943.Search in Google Scholar

[62] Mishra AK, Ramaprabhu S. Magnetite decorated multiwalled carbon nanotube based supercapacitor for arsenic removal and desalination of seawater. J. Phys. Chem. C 2010, 114, 2583–2590.10.1021/jp911631wSearch in Google Scholar

[63] Liu H, Zuo KC, Vecitis CD. Titanium dioxide-coated carbon nanotube network filter for rapid and effective arsenic sorption. Environ. Sci. Technol. 2014, 48, 13871–13879.Search in Google Scholar

[64] Li H, Liu H, Jong Z, Qu W, Geng D, Sun X, Wang H. Nitrogen-doped carbon nanotubes with high activity for oxygen reduction in alkaline media. Int. J. Hydrogen Energy 2011, 36, 2258–2265.10.1016/j.ijhydene.2010.11.025Search in Google Scholar

[65] Zhang M, Yan Y, Gong K, Mao L, Guo Z, Chen Y. Electrostatic layer-by-layer assembled carbon nanotube multilayer film and its electrocatalytic activity for O2 reduction. Langmuir 2004, 20, 8781–8785.10.1021/la048923lSearch in Google Scholar PubMed

[66] Zhang X, Fu J, Zhang Y, Lei L. A nitrogen functionalized carbon nanotube cathode for highly efficient electrocatalytic generation of H2O2 in electro-Fenton system. Sep. Purif. Technol. 2008, 64, 116–123.Search in Google Scholar

[67] Brillas E, Sires I, Oturan MA. Electro-Fenton process and related electrochemical technologies based on Fenton’s reaction chemistry. Chem. Rev. 2009, 109, 6570–6631.Search in Google Scholar

[68] Hsiao YL, Nobe K. Oxidative reactions of phenol and chlorobenzene with in situ electrogenerated Fenton’s reagent. Chem. Eng. Commun. 1993, 126, 97–110.Search in Google Scholar

[69] Pignatello JJ, Oliveros E, MacKay A. Advanced oxidation processes for organic contaminant destruction based on the Fenton reaction and related chemistry. Crit. Rev. Environ. Sci. Technol. 2006, 36, 1–84.Search in Google Scholar

[70] Xie YB, Li XZ. Degradation of bisphenol A in aqueous solution by H2O2-assisted photoelectrocatalytic oxidation. J. Hazard. Mater. 2006, 138, 526–533.Search in Google Scholar

[71] Feng CH, Li FB, Mai HJ, Li XZ. Bio-electro-Fenton process driven by microbial fuel cell for wastewater treatment. Environ. Sci. Technol. 2010, 44, 1875–1880.Search in Google Scholar

[72] Li XZ, Liu HS. Development of an E-H2O2/TiO2 photoelectrocatalytic oxidation system for water and wastewater treatment. Environ. Sci. Technol. 2005, 39, 4614–4620.Search in Google Scholar

[73] Forti JC, Nunes JA, Lanza MRV, Bertazzoli R. Azobenzene-modified oxygen-fed graphite/PTFE electrodes for hydrogen peroxide synthesis. J. Appl. Electrochem. 2007, 37, 527–532.Search in Google Scholar

[74] Guillet N, Roué L, Marcotte S, Villers D, Dodelet JP, Chhim N, Vin ST. Electrogeneration of hydrogen peroxide in acid medium using pyrolyzed cobalt-based catalysts: influence of the cobalt content on the electrode performance. J. Appl. Electrochem. 2006, 36, 863–870.Search in Google Scholar

[75] Isarain-Chávez E, Rosa CDL, Martínez-Huitle CA, Peralta-Hernández JM. On-site hydrogen peroxide production at pilot flow plant. Int. J. Electrochem. Sci. 2013, 8, 3084–3094.Search in Google Scholar

[76] Yuan S, Fan Y, Zhang Y, Tong M, Liao P. Pd-catalytic in situ generation of H2O2 from H2 and O2 produced by water electrolysis for the efficient electro-Fenton degradation of rhodamine B. Environ. Sci. Technol. 2011, 45, 8514–8520.Search in Google Scholar

[77] Brillas E, Bastida RM, Llosa E. Electrochemical destruction of aniline and 4-chloroaniline for waste water treatment using a carbone-PTFE electrode. J. Electrochem. Soc. 1995, 142, 1733–1741.Search in Google Scholar

[78] Zhou M, Yu Q, Lei L. The preparation and characterization of a graphite–PTFE cathode system for the decolorization of C.I. Acid Red 2. Dyes Pigments 2008, 77, 129–136.10.1016/j.dyepig.2007.04.002Search in Google Scholar

[79] Haag WR, David Yao CC. Rate constants for reaction of hydroxyl radicals with several drinking water contaminants. Environ. Sci. Technol. 1992, 26, 1005–1013.Search in Google Scholar

Received: 2015-10-7
Accepted: 2015-12-19
Published Online: 2016-1-28
Published in Print: 2016-2-1

©2016 by De Gruyter

Downloaded on 29.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2015-0056/html
Scroll to top button