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

Only $11.99/month after trial. Cancel anytime.

Nanotechnology Applications for Clean Water: Solutions for Improving Water Quality
Nanotechnology Applications for Clean Water: Solutions for Improving Water Quality
Nanotechnology Applications for Clean Water: Solutions for Improving Water Quality
Ebook1,374 pages8 hours

Nanotechnology Applications for Clean Water: Solutions for Improving Water Quality

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Nanotechnology is already having a dramatic impact on improving water quality and the second edition of Nanotechnology Applications for Clean Water highlights both the challenges and the opportunities for nanotechnology to positively influence this area of environmental protection. This book presents detailed information on cutting-edge technologies, current research, and trends that may impact the success and uptake of the applications.

Recent advances show that many of the current problems with water quality can be addressed using nanosorbents, nanocatalysts, bioactive nanoparticles, nanostructured catalytic membranes, and nanoparticle enhanced filtration. The book describes these technologies in detail and demonstrates how they can provide clean drinking water in both large scale water treatment plants and in point-of-use systems. In addition, the book addresses the societal factors that may affect widespread acceptance of the applications.

Sections are also featured on carbon nanotube arrays and graphene-based sensors for contaminant sensing, nanostructured membranes for water purification, and multifunctional materials in carbon microspheres for the remediation of chlorinated hydrocarbons.

  • Addresses both the technological aspects of delivering clean water supplies and the societal implications that affect take-up
  • Details how the technologies are applied in large-scale water treatment plants and in point-of-use systems
  • Highlights challenges and the opportunities for nanotechnology to positively influence this area of environmental protection
LanguageEnglish
Release dateMay 15, 2014
ISBN9781455731855
Nanotechnology Applications for Clean Water: Solutions for Improving Water Quality

Related to Nanotechnology Applications for Clean Water

Titles in the series (97)

View More

Related ebooks

Environmental Engineering For You

View More

Related articles

Reviews for Nanotechnology Applications for Clean Water

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Nanotechnology Applications for Clean Water - Anita Street

    1.

    Part 1

    Contaminant Sensing Technologies

    Outline

    Chapter 1 Sensors Based on Carbon Nanotube Arrays and Graphene for Water Monitoring

    Chapter 2 Advanced Nanosensors for Environmental Monitoring

    Chapter 3 Electrochemical Biosensors Based on Nanomaterials for Detection of Pesticides and Explosives

    Chapter 4 Dye Nanoparticle-Coated Test Strips for Detection of ppb-Level Ions in Water

    Chapter 5 Functional Nucleic Acid-Directed Assembly of Nanomaterials and Their Applications as Colorimetric and Fluorescent Sensors for Trace Contaminants in Water

    Chapter 1

    Sensors Based on Carbon Nanotube Arrays and Graphene for Water Monitoring

    Dan Du¹,³, Weiying Zhang¹,³, Abdullah Mohamed Asiri² and Yuehe Lin³,⁴,    ¹Key Laboratory of Pesticide and Chemical Biology of Ministry of Education, College of Chemistry, Central China Normal University, Wuhan, PR China,    ²Chemistry Department, King Abdulaziz University, Jeddah, Saudi Arabia,    ³School of Mechanical and Materials Engineering, Washington State University, Pullman, WA, USA,    ⁴Pacific Northwest National Laboratory, Richland, WA, USA

    Nanomaterials, particularly carbon nanomaterials, have shown great promise in environmental applications. In this chapter, we focus only on the use of two carbon nanomaterials, carbon nanotubes (CNTs) and graphene-based sensors for water monitoring and detection of hazardous chemicals. The excellent electrochemical behaviors of CNTs and graphene indicate that they are promising electrode materials in electroanalysis. Sensors based on CNTs and graphene have shown excellent performance in electrochemical detection of metal ions, pesticides, and other pollutants.

    Keywords

    Sensors; carbon nanotubes; arrays; graphene; metals; pesticides

    Outline

    1.1 Introduction 3

    1.2 CNT-based electrochemical sensors 5

    1.2.1 Various methods for preparation of CNT-based sensors 5

    1.2.2 Fabrication of aligned CNT NEA 6

    1.2.3 Applications of CNT-based sensors for metal ion monitoring 7

    1.3 Graphene-based sensors 8

    1.3.1 Graphene-based electrochemical sensors 8

    1.3.2 Graphene sensors for pesticides 10

    1.3.3 Graphene sensors for other pollutants 14

    1.4 Conclusions and future work 15

    Acknowledgments 16

    References 16

    1.1 Introduction

    Environmental monitoring of toxic pollutants in surface and subsurface water sources and wastewaters presently relies on the collection of discrete liquid samples for subsequent laboratory analysis. Sensors that are field-deployable and able to measure part-per-billion (ppb) or nanomolar levels of toxic pollutants will reduce time and costs associated with environmental monitoring of hazardous chemicals. Electrochemical sensors appear to be a very promising technique that offers desired characteristics such as field-deployability, selectivity, sensitivity, robustness, and inexpensiveness [1–3].

    Nanomaterials, particularly carbon nanomaterials, have a significant role to play in new developments in each of the biosensor size domains [4–6]. They have shown great promise in many applications, such as bioscience and biotechnology, energy storage and conversion, environmental and biomedical applications. In this chapter, we focus only on the use of two carbon nanomaterials, carbon nanotubes (CNTs) and graphene-based sensors for water monitoring and environmental applications.

    CNTs are well-ordered, hollow graphitic nanomaterials made of cylinders of sp²-hybridized carbon atoms. They have high aspect ratios, high mechanical strength, high surface areas, excellent chemical and thermal stabilities, and rich electronic and optical properties [7]. The latter properties make CNTs important transducer materials in biosensors: high conductivity along their length means they are excellent nanoscale electrode materials [8–10]. These materials are classed as single-walled carbon nanotubes (SWCNTs) (Figure 1.1A), which are single sheets of graphene rolled into tubes, or multiwalled carbon nanotubes (MWCNTs), each of which contains several concentric tubes that share a common longitudinal axis [11]. As one-dimensional (1D) carbon allotropes, CNTs have lengths that can vary from several hundred nanometers to several millimeters, but their diameters depend on their class: SWCNTs are 0.4–2 nm in diameter and MWCNTs are 2–100 nm in diameter. An important part of the success of CNTs for these applications is their ability to promote electron transfer in electrochemical reactions.

    Figure 1.1 (A) The ideal structures of an SWCNT and (B) a graphene sheet. From Ref. [4], reproduced by permission of Wiley-VCH.

    Graphene is a 2D sheet of carbon atoms in a hexagonal configuration with atoms bonded by sp² bonds. These bonds and this electron configuration are the reasons for the extraordinary properties of graphene, which include a very large surface area (at 2630 m²/g, it is double that of SWCNTs), a tunable band gap, room-temperature Hall effect, high mechanical strength (200 times greater than steel), and high elasticity and thermal conductivity [12]. Graphene is the most recent member of the multidimensional carbon-nanomaterial family, starting with fullerenes as a 0D material, SWCNTs as 1D nanomaterials, and ending with graphite as a 3D material. Graphene fills the gap for 2D carbon nanomaterials (Figure 1.1B). Isolation of individual graphene sheets was long sought, but only in 2004, it was achieved by a surprisingly simple technique [13]. Since then, fundamental research and research on applications have increased rapidly. Graphene is an ideal material for electrochemistry [14–17] because of its very large 2D electrical conductivity, large surface area, and low cost. The use of graphene in electrochemical sensors and biosensors is particularly interesting, with the first articles emerging in 2008. Since then, their number has grown explosively.

    1.2 CNT-based electrochemical sensors

    1.2.1 Various methods for preparation of CNT-based sensors

    Although CNTs are relatively new in analytical fields, their unique electronic (electron transfer rate similar to edge-plane graphite), chemical (biocompatibility and ability to be covalently functionalized), and mechanical properties (3 times stronger than steel) make them extremely attractive for chemical and biochemical sensors [18,19]. Various methods have been used to prepare CNT-based sensors: (a) casting of CNT thin films, from the suspensions of CNTs in solvents [20–28], such as sulfuric acid [21], Nafion [25], dihexadecyl hydrogen phosphate (DHP) [26], DMF [27], and acetone [28], prior to being coated on electrode surfaces, (b) using CNTs as paste electrodes or electrode composites [29–31], and (c) using aligned CNT as electrode substrates [32–40]. With specific to metal ion sensors, DHP [26] and Nafion [25] have been used to disperse MWCNTs under ultrasonication prior to being drop-coated on glassy carbon electrodes. The CNT film enables the development of mercury-free electrodes that can detect from 10−9 to 10−6 M of Cd and Pb.

    Most CNT-based sensors take advantage of the bulk properties of CNTs, including increased electrode surface area [41], fast electron transfer rate [42], and good electrocatalytivity in promoting electron transfer reactions of many important species [43,44]. Using CNTs as nanoelectrodes have been increasing explored since conventional macroelectrodes (having diameter in millimeters), such as glassy carbon and carbon paste, are known to have slow mass transport. Using nanoelectrodes (with diameter in nanometers) can enhance mass transport [37,38,45]; as electrodes decrease in size, radial (3D) diffusion becomes dominant and results in fast mass transport and fast electron transfer, and promotes rapid electrochemical detection. Nanoelectrodes also have higher responsiveness (or higher mass sensitivity) than macroelectrodes, attributed to their lower background (charging) currents [46]. Additionally, they are less influenced by solution resistance due to lower ohmic drop [38]. The high diffusion rate at nanoelectrodes enables the study of fast electrochemical and chemical reactions [47]. Despite its advantages, a single nanoelectrode offers extremely low capacitive current (in pico-amperes), thereby requiring expensive signal amplifiers. To solve this issue, nanoelectrode arrays (NEA) consisting of millions of nanoelectrodes have been developed at Pacific Northwest National Laboratory (PNNL) in collaboration with the Boston College, in order to provide magnified signals without the need for a signal amplifier.

    High-density aligned CNT electrode arrays (i.e., dense CNT forest) have been reported to show fast electron transfer and electrocatalytic characteristics [48], but they do not maintain the properties of individual nanoelectrodes due to the overlapping of their diffusion layers [49–53]. To make each CNT on the array work as an individual nanoelectrode, the spacing must be sufficiently larger than the diameter of the nanotubes. A nanoscale in size of the electrode and the million numbers of electrodes will potentially result in an improved signal-to-noise ratio (and hence improved detection limits) by three to four orders of magnitude compared with that of single nanoelectrode.

    1.2.2 Fabrication of aligned CNT NEA

    The fabrication procedure for low-site density aligned CNT NEA has been refined [54,55]. Figure 1.2 shows the scheme of the fabrication procedure. After the electrochemical deposition of Ni nanoparticles (shown in Figure 1.2A), an aligned CNT array was grown (shown in Figure 1.2B), and then a thin layer of SiO2 was coated on the surface by magnetron sputtering to insulate the Cr layer. After that, 5 m thick M-Bond 610 (two-component, solvent-thinned, epoxy-phenolic adhesive from Vishay Intertechnology, Inc.) was coated and cured at 170°C for 2 hours, which further insulates the Cr and also provides mechanical support to the CNTs. After these steps, the CNTs were half embedded in the polymer resin as shown in Figure 1.2C. In the next step, a fiber-free cloth was used to polish the surface that mechanically breaks the top part of the CNTs and exposes the tip of the CNTs as shown in Figure 1.2D. Finally, the sample surface was rinsed in deionized water, and an insulated copper wire (0.5 mm in radius) was attached to the corner of the substrate by applying a drop of conductive silver epoxy followed by insulating epoxy. The copper wire NEAs assembly was left to cure in air at room temperature for several hours. Figure 1.3 shows the structure of the final CNT NEAs.

    Figure 1.2 Fabrication scheme of the NEAs. (A) Ni nanoparticles electrodeposition; (B) aligned CNT growth; (C) coating of SiO2 and M-Bond; and (D) polishing to expose CNTs. From Ref. [54], reproduced by permission of American Chemical Society.

    Figure 1.3 Structure of CNT NEAs.

    1.2.3 Applications of CNT-based sensors for metal ion monitoring

    The CNT NEAs have been applied for detection of toxic metal ions. With Hg-coated CNT NEA, linear relationship between the Pb signals and the Pb concentration in the solutions ranging from 1 to 100 ppb (g/L) has been obtained [55]. Because Hg is highly toxic and not suitable for field-deployable use, the relatively benign bismuth, Bi(III), has been evaluated as a Hg substitute [56,57]. Bi-based electrodes performed as well as Hg-based electrode for Cr(VI) quantitation after the accumulation of Cr(VI)-diethylenetriammine pentaacetic acid (DTPA) chelate [56]. Highly responsive voltammetric analysis of Cd and Pb at the Bi-based CNT NEA was obtained [34]. Figure 1.4A shows well-defined peaks of 5 g/L of Cd and Pb obtained after only 2 min of accumulation, in which Bi was accumulated in situ with the target metals at −1.2 V. Figure 1.4B shows a linear calibration curve of Cd achieved at a very useful concentration range of 0.5–8 g/L of Cd. The detection limit of 0.04 g/L was obtained under optimum experimental conditions.

    Figure 1.4 (A) Typical square wave voltammogram of 5 µg/L cadmium and lead on the CNT NEAs in 0.1 M acetate buffer (pH 4.5) in the presence of 500 µg/L bismuth. (B) SWV response for increasing the concentration of cadmium (0.5, 1, 2, 4, 6, 8SWV µg/L). Also shown (inset), the resulting calibration plot. Other conditions include in situ plated bismuth-coated CNT-NEA with 1 min pretreatment at 0.3 V; 2 min accumulation at −1.2 V; 10 s rest period (without stirring); square-wave voltammetric scan with a step potential of 5 mV; amplitude, 20 mV; frequency, 25 Hz. From Ref. [34], reproduced by permission of The Royal Society of Chemistry.

    The attractive behavior of the new CNT NEA sensing platform holds great promise for on-site environmental monitoring and biomonitoring of toxic metals.

    1.3 Graphene-based sensors

    1.3.1 Graphene-based electrochemical sensors

    The excellent electrochemical behaviors of graphene indicate that graphene is a promising electrode material in electroanalysis. Graphene, together with its various derivatives, such as graphene oxide (GO), graphene nanoribbon, chemically reduced graphene oxide (rGO), or nitrogen doped graphene [58–62], has shown fascinating advantages in electrochemistry such as electrochemical devices, capacitors, or transistors due to its remarkable electrochemical properties [63–66]. The graphene-based materials have been used to construct various biosensors based on different sensing mechanisms. Because their 2D structure provides a large area for the immobilization of enzyme, and enriched oxygen-containing groups are apt to react with amino-acid residues, graphene derivatives are ideal substrates for immobilization of certain biomolecules. Enzymes can be immobilized on graphene surfaces through the following three methods: (1) electrostatic interaction between positively charged graphene and a negatively charged enzyme, (2) covalent linkage of the enzyme molecule to the graphene surface, and (3) entrapment of the enzyme using the polysaccharide chitosan or another polymer. Electrochemical approach is considered as one of the best methods for biomolecule detection due to its high sensitivity, low cost, quick response, and easy operation. The excellent electrochemical behaviors of graphene make it a promising electrode material to improve the detection of biomolecules [67]. Graphene-based materials enhance the direct electron transfer that is characteristic of some redox enzymes and retain bioactivity because of their favorable electronic properties and biocompatibility [68–70].

    1.3.2 Graphene sensors for pesticides

    Because of the high toxicity of organophosphorus pesticides (OPs), they still remain a serious health threat to humans across the world. Rapid detection of these toxic agents has become increasingly important for health protection [71,72]. Although gas or liquid chromatography and mass spectrometry are accurate and routinely performed for analyzing OPs [73,74], these methods have a number of disadvantages, limiting their applications for rapid analyses under field conditions. Development of a simple, rapid, sensitive, and inexpensive method with real-time output to detect OPs is of considerable interest. Graphene has been developed as an enhanced sensing platform for on-site detection of OPs, based on immobilization of acetylcholinesterase (AChE) or OP hydrolase (OPH). Our group has reported a self-assembly of AChE on a gold nanoparticles (Au NPs)–graphene nanosheet hybrid using a long-chain polyelectrolyte, poly-(diallyldimethylammonium chloride) (PDDA) as a linker for detection of OP [75]. As shown in Figure 1.5, GO was used as the precursor to produce chemically reduced graphene oxide nanosheets (cr-Gs) and PDDA strands could terminate at the cr-Gs surface due to the electrostatic interaction between positive charged PDDA and negative charged GO. By adding NaBH4, chemical reduction and Au NPs decoration were realized in one step. By using PDDA as a linker, AChE could be enriched and stabilized on the nanohybrid. Self-assembling was easy and fast to be realized between negatively charged AChE and positively charged PDDA. As a result, nanoassembly AChE/Au NPs/cr-Gs was formed in the presence of PDDA and was utilized as the catalyst platform of the electrochemical biosensor. Under the optimized condition, ultrasensitive detection of paraoxon (a model compound of organophosphate pesticides) was realized with lab-built flow injection analysis system. Figure 1.6 is a typical it curve acquired at AChE/Au NPs/cr-Gs incubated with different concentrations of paraoxon. Signal (a) is from AChE/Au NPs/cr-Gs for 2 mM acetylthiocholine (ATCh), signals (b to h) are generated from 2 mM ATCh after the AChE/Au NPs/cr-Gs incubated with different concentrations of paraoxon: 0.1 pM, 0.25 pM, 2.5 pM, 25 pM, 0.25 nM, 2.5 nM, and 5 nM paraoxon for 15 min, respectively. Catalysis activity of AChE to ATCh is inhibited apparently after exposure to paraoxon, and even a very low concentration of paraoxon (0.1 pM) could cause the apparent inhibition. Oxidation current of enzymatic product decreased sharply as paraoxon concentration increased. The results demonstrate that the developed approach provides a promising strategy to improve the sensitivity and enzyme activity of electrochemical biosensors.

    Figure 1.5 Schematic illustration of Au NP/cr-Gs hybrid synthesis and AChE/Au NP/cr-Gs nanoassembly generation by using PDDA. Graphite was used for producing GO with Hummer’s method and Au NPs/cr-Gs was obtained by reducing HAuCl4 on GO. AChE was stabilized on the surface of Au NP/cr-Gs hybrid by self-assembling. From Ref. [75], reproduced by permission of The Royal Society of

    Enjoying the preview?
    Page 1 of 1