Brought to you by:
Paper

Adsorption uptake of synthetic organic chemicals by carbon nanotubes and activated carbons

, and

Published 28 June 2012 © 2012 IOP Publishing Ltd
, , Citation A J Brooks et al 2012 Nanotechnology 23 294008 DOI 10.1088/0957-4484/23/29/294008

0957-4484/23/29/294008

Abstract

Carbon nanotubes (CNTs) have shown great promise as high performance materials for adsorbing priority pollutants from water and wastewater. This study compared uptake of two contaminants of interest in drinking water treatment (atrazine and trichloroethylene) by nine different types of carbonaceous adsorbents: three different types of single walled carbon nanotubes (SWNTs), three different sized multi-walled nanotubes (MWNTs), two granular activated carbons (GACs) and a powdered activated carbon (PAC). On a mass basis, the activated carbons exhibited the highest uptake, followed by SWNTs and MWNTs. However, metallic impurities in SWNTs and multiple walls in MWNTs contribute to adsorbent mass but do not contribute commensurate adsorption sites. Therefore, when uptake was normalized by purity (carbon content) and surface area (instead of mass), the isotherms collapsed and much of the CNT data was comparable to the activated carbons, indicating that these two characteristics drive much of the observed differences between activated carbons and CNT materials. For the limited data set here, the Raman D:G ratio as a measure of disordered non-nanotube graphitic components was not a good predictor of adsorption from solution. Uptake of atrazine by MWNTs having a range of lengths and diameters was comparable and their Freundlich isotherms were statistically similar, and we found no impact of solution pH on the adsorption of either atrazine or trichloroethylene in the range of naturally occurring surface water (pH = 5.7–8.3). Experiments were performed using a suite of model aromatic compounds having a range of π-electron energy to investigate the role of π–π electron donor–acceptor interactions on organic compound uptake by SWNTs. For the compounds studied, hydrophobic interactions were the dominant mechanism in the uptake by both SWNTs and activated carbon. However, comparing the uptake of naphthalene and phenanthrene by activated carbon and SWNTs, size exclusion effects appear to be more pronounced with activated carbon materials, perhaps due to smaller pore sizes or larger adsorption surface areas in small pores.

Export citation and abstract BibTeX RIS

1. Introduction

As a result of their unique physical, chemical and electrical properties, carbon nanotubes (CNTs) have received much attention as promising new materials since their discovery in 1991 (Iijima 1991, Baughman et al 2002, Upadhyayula et al 2009, Mauter and Elimelech 2008). One application of interest in environmental engineering is the use of CNTs as high performance sorbents for water purification. The adsorptive capacity of CNTs comes from their high surface area (although generally lower than that of activated carbon), and the presence of graphene sheets that offer a hydrophobic surface with a unique electronic structure. In single walled nanotubes (SWNTs), a single graphene sheet is rolled into a tubule, whereas multi-walled nanotubes (MWNTs) are multiple concentric graphene sheets that form a tubule. Several studies have demonstrated the promise of CNTs for water purification or have elucidated factors governing contaminant removal. Peng et al (2003) showed that CNTs are good adsorbents for the removal of 1,2-dichlorobenzene from wastewater (2003). Lu et al (2006) showed that CNTs had a greater adsorption capacity for trihalomethanes (THMs) than a commercially available powdered activated carbon (PAC). Pan et al (2008) demonstrated high uptake of two endocrine disrupting chemicals, 17α-ethinyl estradiol and bisphenol A. Yang et al (2008) showed that nitro-substituted functional groups of aniline or phenols could enhance the adsorption by CNTs more than chloride and methyl groups. Moreover, a higher degree of substitution increased the adsorption affinity. Among aromatic compounds, aromatic ring planarity, ring conformation, and size of the ring system could influence the adsorption affinity due to the curvature of the nanotube surfaces or access to preferential adsorption sites (Lin and Xing 2008). Zhang et al (2010) suggested that the interstitial channels and longitudinally parallel exterior surfaces of SWNTs may provide more homogeneous adsorption sites for rigid planar molecules, whereas more flexible nonplanar molecules can pack into the inner cavities of MWNTs. Although these studies provide important insights into the use of CNTs for removal of contaminants from water, there is a need for additional data related to specific compounds regulated for drinking water consumption.

An attractive feature of CNTs for water purification is the presence of graphene sheets that offer a hydrophobic surface and potential to interact with solutes having π-electrons in their structure. As summarized by Zhu and Pignatello (2005), the π–π complex is formed mainly by electrostatic forces between sigma-π quadrupoles of opposing ring systems; enthalpies may approach those of hydrogen bonds when both the donor and acceptor are strong. Donor ability correlates with polarizability, which increases with the number of rings in a fused-ring system, and the number and type of electron donating substituents, such as phenolic and alkyl groups. Aromatic compounds such as phenanthrene and naphthalene are electron-rich π-donors (Zhu et al 2004), whereas acceptor ability is enhanced by electron withdrawing substituents.

Long and Yang (2001) found that dioxin exhibited a higher desorption activation energy from MWNTs, as compared to an activated carbon, during temperature programmed desorption, and cited potentially strong interactions between the two benzene rings of dioxin and the MWNT surface. Gotovac et al (2006) hypothesized π–π coupling between the π electrons of the aromatic rings of phenanthrene and the surface of the CNTs (2006). The higher uptake of tetracene was attributed to the potential for π–π interactions along the nanotube axis (Gotovac et al 2007). Chen et al (2007) showed that, in addition to hydrophobic effects, sorption mechanisms may also involve π-electron polarizability. The high adsorption affinity of nitroaromatics was attributed to strong interactions between the π-electron rich aromatic ring and the π-region of the CNT surface (Chen et al 2008). Pan et al (2008) postulated that π–π electron donor–acceptor interactions were operative during the uptake of 17α-ethinyl estradiol and bisphenol A. Lin and Xing (2008) found that sorption affinity increased with the number of aromatic rings. They proposed that the CNTs act as amphoteric adsorbents, which allows them to attract either π-electron acceptors or donors to their surface, and that the strength of the π–π interactions increases with increasing aromaticity. Ji et al (2009a) attributed strong sorption of two sulfonamide antibiotics by MWNTs to π–π electron coupling with the graphene surface. Zhang et al (2010) compared the adsorption affinities of phenanthrene, biphenyl and 2-phenylphenol, and suggested that the major mechanisms for adsorption include a combination of hydrophobic interactions, electrostatic repulsion, and water cluster formation on the CNT surface. These studies provide valuable insights into the mechanisms of uptake by CNTs. However, few studies attempt to control the hydrophobic effect when elucidating such mechanisms, and the hydrophobic effect is likely to be quite important for organic compounds. For example, Yang et al (2006b) showed that adsorption affinity of polycyclic aromatic hydrocarbons (PAHs) was consistent with their hydrophobicity (Kow) in the order of naphthalene < phenanthrene < pyrene.

Several studies have investigated the effects of solution chemistry on the uptake of organics by CNTs. Peng et al (2003) showed that CNTs have near constant adsorption capacity for 1,2-dichlorobenzene over a wide pH range (3–10). Ji et al (2009a, 2009b) showed that uptake of two sulfonamide antibiotics was insensitive to a pH between 3 and 8, and to an ionic strength between 0.01 and 0.1 M (using NaCl and CaCl2) (Ji et al 2009b). Zhang et al (2010) showed that changes in solution pH and ionic strength within the range typical of natural waters did not affect the adsorption of phenanthrene, biphenyl or 2-phenylphenol by CNTs. However, there are few data for the effects of solution chemistry on uptake of compounds of interest in the drinking water industry.

In the context of water treatment, it is of interest to compare the efficacy of different carbon adsorbents, including single and multi-wall nanotubes, and activated carbons, the carbon adsorbent most widely used in practice. This comparison is of interest, in part, because it is likely that the nature of adsorption sites on CNTs and activated carbons is significantly different. Zhang et al (2009) concluded, based on nitrogen adsorption isotherms, that only a small fraction of adsorption sites was attributable to the inner cavities of the CNTs; rather, the majority of the sites were found in the interstitial channels and grooves that formed as a result of aggregation. In contrast, activated carbons are known to have wide pore size distributions, with a large fraction of the total surface area in 'micropores' approaching adsorbate dimensions. Pores smaller than adsorbate dimensions may also be present, leading to size exclusion. Finally, at higher loadings, available micropore volume may become saturated, limiting uptake. Ji et al (2009a) found that whereas tetracycline uptake by CNTs was greater than by activated carbon, the finding that GAC had the highest adsorption affinity for naphthalene was attributed to a pore-filling mechanism. Zhang et al (2009) found that uptake from aqueous solution correlated with surface area rather than pore volume, and aggregation reduced adsorption uptake. Thus, although adsorption uptake by CNTs is likely to increase when dispersed as individual nanotubes, such dispersal is difficult to accomplish without surface treatment (Chen et al 2004, Piao et al 2008, Zhang et al 2009). Such treatment often adds surface oxygen groups that can interact strongly with water, thus lowering the uptake (Kilduff and King 1997, Cho et al 2008).

The overarching goal of this research was to investigate the efficacy of CNTs to adsorb atrazine and trichloroethylene, two contaminants of interest in drinking water treatment, and to compare uptake by CNTs and activated carbon, a widely used adsorbent for water purification. In addition, several different carbon and nanotube materials were screened for atrazine uptake to provide data on the potential range of performance over a wide cross section of commercially available CNT materials. Several commercially available SWNTs and several MWNT materials, having a range of diameter and length, were selected for this purpose. Both as-produced and purified materials were investigated, in an attempt to determine how purification can enhance adsorption uptake. Rigorous purification can produce materials with purity levels above 90%; however, this can lead to a significant increase in the price of the purified material (Itkis et al 2003), an important consideration for water treatment applications.

CNT materials were compared to two different water treatment activated carbons having different starting materials and methods of activation. The effect of pH, over a range of 5.7–8.3, was also examined to encompass the range expected for natural waters. The role of specific (including π–π) interactions was assessed by comparing uptake of several model compounds after controlling for the hydrophobic effect, which was done by expressing uptake data measured in the aqueous phase in terms of concentration in an inert solvent (i.e. n-hexadecane). The inert solvent reference state eliminates or minimizes differences in solute–solvent interactions for different solutes, enabling a comparison between sorbates in terms of their interactions with the sorbent.

2. Materials and methods

2.1. Adsorbate selection

A number of different compounds were selected to compare uptake between traditional activated carbon and carbon nanotubes, and to determine the role that π–π interactions play in their uptake by CNTs. The physical and chemical properties of these compounds are tabulated in table 1, and compound structures are tabulated in supplemental information (available at stacks.iop.org/Nano/23/294008/mmedia). Atrazine (2-chloro-4-ethylamino-6-isopropylamino-s-triazine) is one of the most frequently applied and widely used herbicides in the world (Graymore et al 2001, Lazorko-Connon and Achari 2009, Chen et al 2009). Once applied in the environment, atrazine leaches into groundwater and surface water, and has thus been widely detected in drinking water supplies. It is recognized as an endocrine disruptor, toxic to humans and aquatic organisms (Graymore et al 2001, Lazorko-Connon and Achari 2009). The USEPA has suggested that exposure to the pesticide was associated with several birth-related health risks (Erickson 2010). Trichloroethylene (TCE) is also a common groundwater and surface water pollutant, and is a suspected carcinogen (Lavin et al 2000). These contaminants are representative of a wide variety of pollutants that are frequently identified in both surface and ground water drinking water sources, and for this reason, they have been chosen for this study.

Table 1.  Physicochemical properties of adsorbates. (Note: vapor pressure, solubility, Henry's constant, and log (octanol water partition coefficient) (log Kow) values were obtained from EPI Suite software (US Environmental Protection Agency); reported experimental database values are shown. Pi energy values calculated from the Hückel molecular orbital theory using ChemAxon software. Refractivity values calculated using the atomic contribution method of Viswanadhan et al (1989) as implemented in the ChemAxon software (www.chemaxon.com).)

Compound Formula MW (g mol−1) Vapor pressure (Pa), 25 °C Solubility (mg l−1), 25 °C Henry constant (Pa m3 mol−1) Log Kow (—) Pi energy (β) Refractivity 106 (m3 mol−1) Polarizability (Å3)
Cyclohexene C6H10 82.15 1.19 × 104 213 4.61 × 103 2.86 2.00 28.72 10.34
Benzene C6H6 78.11 1.26 × 104 1790 5.62 × 102 2.13 8.00 26.06 8.89
Naphthalene C10H8 128.18 1.13 × 101 31 4.46 × 101 3.30 13.68 42.51 14.56
Phenanthrene C14H10 178.24 1.16 × 10−2 1.15 4.29 × 100 4.46 19.45 58.96 20.43
Chlorobenzene C6H5Cl 112.56 1.60 × 103 498 3.15 × 102 2.84 11.10 30.86 11.06
1,4-dichlorobenzene C6H4Cl2 147 2.32 × 102 81.3 2.44 × 102 3.44 14.20 35.67 13.32
Methylbenzene C7H8 92.14 3.79 × 103 526 6.73 × 102 2.73 8.00 31.10 10.97
1,4-dimethylbenzene C8H10 106.17 1.18 × 103 162 6.99 × 102 3.15 8.00 36.14 13.12
Trichloroethylene C2HCl3 131.39 9.20 × 103 1280 9.98 × 102 2.42 11.31 35.18 9.59
Atrazine C8H14N5Cl 215.69 3.85 × 10−5 34.7 2.39 × 10−4 2.61 19.23 62.22 22.58

To improve our understanding of π–π interactions between adsorbate aromatic rings and the CNT surface, which are expected to depend on the electron density in the carbon and in the aromatic structure of the organic compound, four compounds possessing significantly different π electron densities were selected: cyclohexene, benzene, naphthalene, and phenanthrene. Cyclohexene and benzene have one and three double bonds, respectively, which are expected to promote π–π interactions. Both naphthalene and phenanthrene are polycyclic aromatic hydrocarbons containing two and three rings, respectively. The adsorption trends of these molecules will provide insight into the significance of π–π interactions and the significance of hydrophobic versus π–π interactions during adsorption.

Four additional compounds were selected in order to examine the effects of methyl group and chlorine substitution in aromatic compounds, and to assess the effects of electron withdrawing and donating groups on π–π interactions: chlorobenzene, 1,4-dichlorobenzene, methylbenzene (toluene), and 1,4-dimethylbenzene (p-xylene). The presence of electron withdrawing groups on the benzene ring (e.g. –Cl) would be expected to increase the interactions of the adsorbate with carbon surfaces having high electron density, whereas the presence of an electron donating group (e.g.–CH3) would exhibit the opposite trend. The inductive and resonance effects of a substituent can be quantified by their Hammett constants, which are tabulated in table 2.

Table 2.  Hammett constants for adsorbate substituents. (Note: Hammett constants taken from Hansch et al (1991). Sigma values (σ) refer to either the meta (m) or para (p) position. The σp value can be expressed in terms of a field/inductive effect and a resonance effect, σp = R + F, where F is found from 1.297σm − 0.385 σp + 0.033, and R is determined by difference. Constants for NO2 and NH2 are shown for comparison purposes.)

Group σm σp F R
H 0 0 0.03 0
Cl 0.37 0.23 0.42 −0.19
OH 0.12 −0.37 0.33 −0.7
CH3 −0.07 −0.17 0.01 −0.18
NO2 0.71 0.78 0.65 0.13
NH2 −0.16 −0.66 0.08 −0.74

2.2. Adsorbents

Nine different carbon sorbents were used to define the range of performance available for these materials, and to compare uptake of activated carbons and carbon nanotubes for compounds of interest in drinking water treatment. These adsorbents were used, as received, to simulate how they would be employed in practice. Elemental analyses of the adsorbents were performed by Huffman Laboratories, Inc. (Golden, CO) and are given in table 3. Additional physical and chemical properties of the CNT materials are tabulated in table 4. The activated carbons include coal based granular materials (TOG, Calgon) from two different batches and having two different particle sizes, 80 × 325 (TOG80 × 325) and 20 × 50 (TOG 20 × 50). Zimmerman et al (2005) report an elemental carbon content of 87% and a specific surface area of 938 m2 g−1 for this material. A wood-based powdered carbon (Aqua Nuchar PAC, MeadWestvaco, Corington, VA USA) was also investigated. The manufacturer reports that this carbon is phosphoric acid activated followed by a hot water wash, which yields a mesoporous porosity (a predominance of pores in the 2–50 nm range) and a 0.4 ml g−1 micropore volume. The acid activation results in a slightly acidic surface, having a pH of 4–6. The manufacturer reports a specific surface area of 1200 m2 g−1 and an ash content of less than 10%. Three different SWNTs were used with different degrees of purity.

Table 3.  Adsorbent elemental analysis.

Adsorbent Type C (%) H (%) N (%) O (%) S (%) Ash (%)
Calgon TOG 20 × 50 GAC 90.07 0.55 0.59 2.61 0.73 5.78
Calgon TOG 80 × 325 GAC 89.44 0.51 0.54 2.86 0.66 6.38
MeadWestvaco Aqua Nuchar PAC 82.76 1.86 0.19 9.95 0.04 8.14
Cheap Tubes SWNT 94.8 0.31 0.1 2.26 0.06 2.99
Carbon Solutions SWNT 68.31 0.33 0.22 2.48 1.76 38.72
Carbolex AP SWNT 62.71 0.45 0.31 4.47 0.02 40.67
NanoLab PD15L1-5 MWNT 96.03 0.21 0.05 0.61 0.01 3.99
NanoLab PD30L1-5 MWNT 95.91 0.17 0.05 0.75 0.02 4.13
NanoLab PD30L5-20 MWNT 94.94 0.18 0.06 0.99 0.02 5.42

Table 4.  Carbon nanotube characterization data. (Note: 1. Process: CVD = chemical vapor deposition; EA = electric arc discharge; HiPCO = high pressure CO conversion process. 2. Purity: as reported by the manufacturer. 3. Purity method (analysis technique) R = Raman spectroscopy; XRD = energy dispersive x-ray spectroscopy; NIR = near-infrared spectroscopy, see Itkis et al (2005); TGA = thermogravimetric analysis. 4. SSA = specific surface area. The value for Cheap Tubes was reported by the manufacturer. The value for carbon solutions SWNTs was calculated using the approach defined by Peigney et al (2001) based on a nanotube bundle to nanotube diameter ratio of 7. The value for carbolex was calculated using the approach defined by Peigney et al (2001) based on a nanotube bundle of N = 100 tubes, as estimated by the manufacturer. The value for carbon nanotechnologies is the BET area measured using nitrogen at 77 K. The value for the Nanolab materials is at the low end of the range reported by the manufacturer, and is consistent with the calculated values from Peigney et al (2001) for a MWNT having 10 walls (as reported by the manufacturer). 5. Tube and bundle dimensions: as reported by the manufacturer except carbon nanotechnologies, reported in Parra-Vasquez et al (2007). 6. Raman D:G ratio: Cheap Tubes, as reported by Jubete et al (2011); carbon solutions, as reported in Itkis et al (2005); carbolex, Raman spectra provided by manufacturer; carbon nanotechnologies, Raman spectra measured in our laboratory; Nanolab, Raman spectra provided by the manufacturer. 7. Metal content reported by the manufacturer except carbon nanotechnologies, reported in Zhou et al (2001). NR: not reported.)

Adsorbent Type Process Purity (%) Purity Method Catalyst/impurity SSA (m2 g−1) Tube D(nm) Bundle D(nm) Length (µm) Raman D:G Metal (%)
Cheap Tubes SWNT CVD 90 R, XRD Al, Co, S 407 1–2 2–10 5–30 0.06 <2
Carbon Solutions SWNT EA 40–60 NIR Ni/Y 355 1.55 2–10 1–5 0.12 30
Carbolex AP SWNT EA 66 TGA Ni/Y 240 1.40 10–20 3–5 0.05 33
Carbon Nanotechnologies SWNT HiPCO 85 TGA Fe 633 1–2 2–8 0.4–0.7 0.11 <6
NanoLab PD15L1-5 MWNT CVD 95 TGA Fe, S 200 15 NR 1–5 0.61  ≈ 1
NanoLab PD30L1-5 MWNT CVD 95 TGA Fe, S 200 30 NR 1–5 0.61  ≈ 1
NanoLab PD30L5-20 MWNT CVD 95 TGA Fe, S 200 30 NR 5–20 0.61  ≈ 1

Cheap Tubes nanotubes (Cheap Tubes, Brattleboro VT, USA) were made using the catalyzed chemical vapor deposition process (CVD). The as-produced material is purified by the manufacturer to a reported purity of 90%, a MWNT content of 5% and an amorphous carbon content of 3%. Jubete et al (2011) reported characterization data for the Cheap Tubes SWNTs. Raman spectra showed a strong tangential G band at ∼1580 cm−1. Broad features at about 1345 cm−1 are due to the disordered sp2 carbon D-band of non-nanotube graphitic components, and indicates the presence of impurities. For this reason, the ratio of the amplitudes of the G-band and the D-band has been used as a measure of the purity (Itkis et al 2005). A low D : G peak intensity ratio of 0.06 was reported for the Cheap Tubes material. A bimodal G'-band was observed at ∼2680 cm−1, suggesting a mixture of several types of SWCNTs in the starting material.

Carbon Solutions (Carbon Solutions, Inc., Riverside, CA, USA) nanotubes were electric arc synthesized using a Ni/Y catalyst. Carbonaceous purity of the as-produced materials is reportedly between 40 and 60%, with a metal content of 30%. Individual tube lengths range from 0.5 to 3 µm and have an average diameter of 1.4 nm. These SWNTs tend to occur as bundles with lengths of 1–5 µm and average bundle diameters of 2–10 nm. The density of the AP-SWNT is 1.2–1.5 g cm−3, and the ratio of semiconducting to metallic SWNTs produced using the electric arc discharge method is 2–1 (manufacturer's data).

Carbolex SWNTs obtained from Sigma-Aldrich were also produced by the electric arc method, with a diameter range of 1.2–1.5 nm (manufacturer's data). The carbonaceous purity of these nanotubes is reportedly between 50 and 70%; a metal content of <33% was reported by the manufacturer.

The NanoLab multi-wall nanotubes (NanoLab, Inc., Waltham, MA, USA) were produced by CVD to various length and diameter specifications, and were purified to >95%, as measured by TGA. Residuals may contain iron and sulfur, and the reported specific surface area is in the range of 200–400 m2 g−1. Three samples were investigated to probe effects of the MWNT length (L) and diameter (D), having D = 15 ± 5 nm, L = 1–5 µm (PD15L1-5); D = 30 ± 15 nm, L = 1–5 µm (PD30L1-5); or D = 30 ± 15 nm, L = 5–10 µm (PD30L5-20).

To investigate the role of π–π interactions, a SWNT (Carbon Nanotechnologies, Inc., Houston, TX, USA) was used. These were synthesized by high pressure CO conversion synthesis (HiPCO) and were purified before use. First, the SWNTs were heat treated in air at 300 °C for 30 min to reduce amorphous carbon, and facilitate the separation of the nanotubes from amorphous carbon coatings via surface cracking. This was followed by soaking in 18 wt% HCl for 20 min to solubilize the metals used as catalysts.

2.3. Isotherm experiments

Isotherm experiments were conducted (1) to compare different CNTs and to compare CNTs with activated carbon, and (2) to investigate π–π interactions between CNTs and adsorbates having various electron donor/acceptor capabilities. Constant carbon dose aqueous adsorption isotherm experiments were conducted in completely mixed batch reactors (CMBR, 250 ml amber colored glass bottles with Teflon-lined screw caps). Each point of the adsorption isotherm represents an individual experiment in one CMBR. Stock solutions of each adsorbate were prepared in methanol. For single solute experiments, approximately 10 mg of carbon adsorbent (nanotube or activated carbon) was carefully added to each reactor, which was then filled approximately halfway with 1.0 mM phosphate buffer solution made with MQ water containing 100 ppm NaN3. The pH of each CMBR was then adjusted to 5.7, 7.0, or 8.3 using HCL and NaOH, and the ionic strength was adjusted to 0.01 M using NaCl. Adsorbate was then added to each reactor via a microliter aliquot of stock solution. The bottle was then carefully filled without headspace with 1.0 mM phosphate buffer solution. The CMBRs were then placed on a jar mill roller/tumbler (Norton/St Gobain) and rotated end over and for a period of two weeks. Preliminary experiments indicated that this time was sufficient to reach an equilibrium conditions as measured by a lack of change in solution concentration with additional contact time (see supplemental information available at stacks.iop.org/Nano/23/294008/mmedia). At the end of the second week, the bottles were removed and allowed to sit for 24 h to allow the suspended carbon to settle out of the solution. Supernatant solutions were analyzed for TCE using gas chromatography (GC) with a μECD detector (Model 6890N, Agilent Technologies, Inc., Santa Clara, CA, USA). Atrazine was determined by high performance liquid chromatography (HPLC) (Model 1100 Agilent Technologies, Inc., Santa Clara, CA, USA) equipped with an UV-diode array detector set at a wavelength of 220 nm; the mobile phase was 60% methanol and 40% MQ water.

3. Results and discussion

3.1. Data analysis/isotherm modeling

The organic compound sorption data was fitted using the Freundlich model:

Equation (1)

where qe is the solid-phase concentration (µg —adsorbate/g—sorbent), Ce is the aqueous-phase concentration (mg—adsorbate/m3—water), KF is a capacity factor ((µg g−1) (m3 mg−1)n), and n is an exponent related to the intensity of adsorption; lower values indicate a site energy distribution encompassing high energy sites (dimensionless). The Freundlich model has been widely used for describing the sorption of organic chemicals to carbonaceous materials such as activated carbon and CNT materials (Kilduff et al 1996, Yang et al 2006a) and other heterogeneous sorbents, including natural soils and sediments.

3.2. Atrazine uptake

Single solute experiments were conducted to compare the uptake of atrazine by several different activated carbons and carbon nanotube materials; the results are plotted on an adsorbent mass basis in figure 1, and Freundlich isotherm parameters are tabulated in table 5. For the TOG 80 × 325 and Cheap Tubes SWNT isotherms, the data represent pooled data taken at three different pH values (pH 5.7, 7.0, and 8.3), and will be discussed in more detail below. For the other adsorbents, uptake was measured at pH 5.7. Over the concentration range examined, the three activated carbons exhibited higher uptake for atrazine when compared to the nanotube materials, although at higher concentrations the isotherms appear likely to converge. The smaller particle size TOG 80 × 325 carbon exhibited the highest uptake, followed by the TOG 20 × 50 and the Aqua Nuchar powdered carbon. The uptake was favorable, with Freundlich n-values between 0.30 (TOG 20 × 50) and 0.38 (Aqua Nuchar). The Cheap Tubes SWNT material performed the next best, with uptake lower than the activated carbons, but higher than the other nanotube materials at concentrations above about 500 mg m−3; however, the uptake by the Cheap Tubes SWNTs is similar to or lower than the other SWNTs at concentrations less than 100 µg l−1. Additionally, the Cheap Tubes SWNTs had the highest Freundlich n-value of all adsorbents investigated, 0.54, although still quite favorable. The Carbolex and Carbon Solutions SWNTs exhibited similar uptake to each other, whereas the MWNT samples exhibited the lowest uptake, especially at concentrations less than 1000 mg m−3.

Figure 1.

Figure 1. Atrazine uptake by a range of carbon materials including coal based granular activated carbon (GAC), wood-based powdered activated carbon (PAC), single walled nanotubes (SWNT) and multi-walled nanotubes (MWNT).

Standard image

Table 5.  Freundlich parameters for atrazine uptake. (Note: parameters are for the Freundlich model, ${q}_{\mathrm{e}}={K}_{\mathrm{F}}{C}_{\mathrm{e}}^{n}$, where qe is the solid-phase concentration (µ g—adsorbate/g—sorbent), Ce is the aqueous-phase concentration (mg—adsorbate/m3—water), KF is a capacity factor ((µg g−1) (m3 mg−1)n), and n is a dimensionless parameter.)

Adsorbent KF n r2
Calgon TOG 20 × 50 25.37 0.299 0.950
Calgon TOG 80 × 325 42.44 0.293 0.885
Westvaco Aqua Nuchar 12.99 0.379 0.945
Cheap Tubes 2.38 0.541 0.825
Carbon solutions 7.01 0.314 0.936
Carbolex AP 11.70 0.214 0.985
Nanolab PD15L1-5 2.20 0.457 0.997
Nanolab PD30L1-5 2.29 0.465 0.974
Nanolab PD30L5-20 2.78 0.439 0.992

The Cheap Tubes and Carbolex SWNTs have similar reported Raman D:G ratios (0.05 and 0.06, respectively); therefore, this parameter does not appear to correlate with adsorption from solution. The lower uptake (on a mass basis) of the Carbolex and Carbon Solutions SWNTs, as compared to the Cheap Tubes SWNTs, is most likely caused by their metal impurities leading to a lower carbon content (of the order of 60–70%, versus nearly 95% for the Cheap Tubes). The impurities contribute a significant mass but do not provide strong adsorption sites. Although the Nanolab MWNTs have high purity, it is likely that their concentric sheets contribute to adsorbent mass but are not accessible for adsorption. The MWNTs also have a significantly higher D:G ratio; this may indicate a greater degree of carbonaceous impurities, which are lower energy adsorption sites. Shown in greater detail in figure 2, uptake of atrazine by the MWNTs was comparable, and their Freundlich isotherms were statistically similar, providing evidence that there is little effect of the MWNT length or diameter on the uptake of atrazine over the range studied here.

Figure 2.

Figure 2. Atrazine uptake by multi-walled nanotubes (MWNT) of varying diameter (D = 15 or 30 nm) and length (L = 1–5 or 5–20). Freundlich isotherms (lines) are statistically similar: KF = 2.44 (2.09, 2.85); n = 0.452 (0.425, 0.479).

Standard image

The potential effects of carbon content, surface area, and the combined effect of these were investigated by plotting the atrazine uptake normalized by these properties. The normalized isotherms are shown in figure 3, panels (A)–(C). The isotherms normalized by carbon content (panel (A)) appear quite similar to those in figure 1, with the exception that the normalized uptake increased significantly for the lower purity Carbolex and Carbon Solutions SWNTs. Still, the rank order of the adsorbents did not change appreciably, and the overall range of uptake among the different materials is about an order of magnitude. Normalizing by surface area (panel (B)) collapses the isotherms significantly, uptake is now within about a factor of two. Notably, most of the CNT data is comparable to or above the wood-based PAC. This could be explained by either a high surface area that is inaccessible to the adsorbate, a surface chemistry effect, or both. The acid activation of the wood-based material leads to oxygen containing functional groups on the surface which have been shown to reduce uptake (Kilduff and King 1997, Karanfil and Kilduff 1999, Karanfil et al 1999). Normalizing uptake by both carbon content and surface area collapses the data even further (panel (C)), indicating that these two characteristics drive many of the observed differences between activated carbons and CNT materials.

Figure 3.

Figure 3. Atrazine uptake by a range of carbon materials including coal based granular activated carbon (GAC), wood-based powdered activated carbon (PAC), single walled nanotubes (SWNT) and multi-walled nanotubes (MWNT). Panel (A): uptake is normalized for carbon content (from table 3). Lines are Freundlich isotherm fits to TOG 80 × 325 and MWNT data to illustrate the range in uptake. Panel (B): uptake is normalized for specific surface area (from table 4 and as described in the text). Lines are Freundlich isotherm fits to TOG 80 × 325 and Carbon Solutions SWNT data to illustrate the range in uptake. Panel (C): uptake is normalized for both carbon content and specific surface area. Lines are Freundlich isotherm fits to TOG 80 × 325 and Carbon Solutions SWNT data to illustrate the range in uptake.

Standard image

The data plotted in figure 4 represents single solute experiments conducted at three different pH values (5.7, 7.0, and 8.3) for SWNT (Cheap Tubes) and for the TOG activated carbon; Freundlich model parameters are tabulated in table 6. As discussed above, on a mass basis, the activated carbon exhibited higher uptake for atrazine, as compared to the SWNT, over the range of concentrations examined. Also, the data plotted in figure 4 establishes that over the pH range from 5.7 to 8.3, there is no significant effect of pH on the uptake of atrazine by either the TOG 80 × 325 carbon or the Cheap Tubes SWNT material. This was verified by statistical analysis, quantified by overlapping 95% confidence intervals. Therefore, the Freundlich isotherm parameters reported in table 6 represent pooled data.

Figure 4.

Figure 4. Effect of pH on single solute atrazine uptake by (a) TOG activated carbon; Freundlich model (solid line): ${q}_{\mathrm{e}}=42.44{C}_{\mathrm{e}}^{0.293}$, R2 = 0.885 and (b) SWNTs (Cheap Tubes); Freundlich model (solid line): ${q}_{\mathrm{e}}=2.38{C}_{\mathrm{e}}^{0.541}$, R2 = 0.825.

Standard image

Table 6.  Freundlich parameters for atrazine uptake by TOG 80 × 325 GAC and Cheap Tubes SWNT. (Note: parameters are for the Freundlich model, ${q}_{\mathrm{e}}={K}_{\mathrm{F}}{C}_{\mathrm{e}}^{n}$, where qe is the solid-phase concentration (µg—adsorbate/g—sorbent), Ce is the aqueous-phase concentration (mg—adsorbate/m3—water), KF is a capacity factor ((µg g−1) (m3 mg−1)n), and n is a dimensionless parameter.

Carbon KF n r2
Calgon TOG 80 × 325 42.44 (30.40, 59.23) 0.293 (0.232, 0.353) 0.885
Cheap Tubes SWNT 2.38 (1.05, 5.40) 0.541 (0.425, 0.658) 0.825

Atrazine does not ionize over the pH range studied; therefore, changes in pH will not have significant effects on solute activity (fugacity) or solubility, and the lowest pH value is sufficiently high that sorption competition from H+ is not expected. Consequently, the only potential effect of pH is an effect on the adsorbent surface chemistry, such as ionization of surface functional groups. The results presented in figure 4 confirm that if such changes occur, they are negligibly small and subsequently have little effect on atrazine adsorption. This result was expected for TOG, because it is a 'basic' type activated carbon, meaning it is not exposed to air or other oxidizing gas during cooling after activation. Such carbons, in general, do not contain a high concentration of surface oxygen groups, in contrast to acid-activated wood-based carbons. The fact that pH had little effect on atrazine uptake by the SWNT suggests that, like the activated carbon, its surface does not contain a high concentration of ionizable groups. In this regard, it can be concluded that the surface chemistry of the SWNT and the (basic) activated carbon are similar.

Uptake by the TOG 80 × 325 activated carbon was well described by the Freundlich model; individual isotherms had r2 values greater than 0.87, and pooled data yield an r2 value of 0.91. The Freundlich model also fits the SWNT uptake data, but lower r2 values were found due to greater variability inherent in the SWNT data; pooled data yield an r2 value of 0.681. Because the same adsorbent masses were used, and the same protocols were followed, we conclude that the SWNT material is inherently more heterogeneous than the activated carbon material. One explanation for the lower uptake of atrazine by SWNTs on an adsorbent mass basis suggests that the material has a lower available surface area than the activated carbon. This may be due to a lower degree of microporosity, due to the formation of bundles or aggregates that reduce the available surface area, or both (Ji et al 2009a, 2009b, Wang et al 2010, Zhang et al 2009).

3.3. TCE uptake under single solute conditions

Single solute experiments were conducted at three different pH values (5.7, 7.0, and 8.3) to measure the uptake of TCE by SWNT (Cheap Tubes) and TOG 80 × 325 activated carbon. The data are plotted in figure 5; Freundlich isotherm model parameters are tabulated in table 7. As was observed for atrazine, there is no effect of pH on TCE uptake over the range studied, by either the TOG 80 × 325 or the Cheap Tubes SWNTs. The major potential effect of pH would be on the adsorbent surface chemistry, which could include ionization of surface functional groups. The results presented in figure 5 confirm that such pH effects, if present, are small and have a negligible effect on TCE uptake by either adsorbent. This is consistent with what was observed for atrazine, and further suggests that the surface chemistry of the Cheap Tubes SWNTs and the TOG active carbon are similar in their response to pH.

Figure 5.

Figure 5. Effect of pH on single solute TCE uptake by (a) TOG activated carbon; Freundlich model (solid line): ${q}_{\mathrm{e}}=1.033{C}_{\mathrm{e}}^{0.592}$, R2 = 0.889 and (b) SWNTs (Cheap Tubes); Freundlich model (solid line): ${q}_{\mathrm{e}}=0.241{C}_{\mathrm{e}}^{0.808}$, R2 = 0.879.

Standard image

Table 7.  Freundlich model parameters for TCE uptake by TOG 80 × 325 GAC and Cheap Tubes SWNT.

  KF n r2
Calgon TOG 80 × 325 1.032 (0.620, 1.719) 0.592 (0.492, 0.693) 0.889
Cheap Tubes SWNT 0.157(0.087, 0.0283) 0.881 (0.770, 0.992) 0.879

The TCE uptake measured here was similar to that reported earlier for a similar water treatment carbon (Karanfil and Kilduff 1999, Kilduff and Karanfil 2002). The TOG activated carbon exhibited greater uptake than SWNT at low concentrations and comparable uptake at higher equilibrium concentrations. TCE uptake by the TOG activated carbon is well described by the Freundlich model, with an r2 value of 0.889. The SWNT uptake is also well described by the Freundlich model with an r2 value of 0.828; the lower correlation coefficient may be attributable to the greater variability that has been found in the SWNT data.

The potential effects of surface area were investigated by plotting TCE uptake normalized by this property, as was done for atrazine in figure 3. Because the carbon content of TOG GAC and Cheap Tubes SWNTs is similar, normalization by this parameter did not yield useful results. However, as shown in figure 6, after normalization by surface area, the data for TCE uptake by TOG GAC and Cheap Tubes SWNT correspond rather closely, consistent with our findings for atrazine.

Figure 6.

Figure 6. TCE uptake, normalized by specific surface area, by TOG 80 × 325 activated carbon and Cheap Tubes SWNTs.

Standard image

3.4. Evidence for specific interactions

To assess specific interactions, uptake data, which was measured in the aqueous phase, was expressed in terms of concentration in an inert solvent (i.e. n-hexadecane). The inert solvent reference state minimizes differences in solute–solvent interactions for different solutes, thus enabling a comparison between sorbates in terms of their interactions with the sorbent (Borisover and Graber 2003):

Equation (2)

where Cn-hexa is the solution concentration in an inert solvent (i.e. n-hexadecane) that would exist at the same fugacity as the measured aqueous-phase concentration; Caq is the aqueous solution concentration; Sn-hexa and Saq are the compound solubility in n-hexadecane and water, respectively. This approach is used to normalize differences in solubility and hydrophobic effects, and therefore provides a better comparison of interactions between the adsorbates and the sorbent surfaces. The ratio of solubilities could also be replaced by the ratio of air/water Henry's constant, Hw, and the air/hexadecane Henry's constant, Hn-hexa (Borisover and Graber 2003):

Equation (3)

Equivalently, the normalization factor could be expressed as the n-hexadecane–water partition coefficient (KHW,i (m3 m−3)) as done by Zhu and Pignatello (2005). This reveals more clearly the fact that the normalization factor is related to the change in chemical potential of a compound upon transfer from the infinitely dilute standard state in water to the infinitely dilute standard state in n-hexadecane (Borisover and Graber 2003). Freundlich model parameters for probe compounds are tabulated in table 8.

Table 8.  Freundlich parameters for ring and aromatic compounds. (Note: qe is in mmol g−1, and Ce is in units of mmol l−1, normalized for hydrophobic effects as described in the text. Adsorbent: HipCo SWNTs (Carbon Nanotechnologies).)

Compound KF n r2
Cyclohexene 0.118 (0.074, 0.186) 0.550 (0.462, 0.638) 0.928
Benzene 0.067 (0.031, 0.146) 1.083 (0.854, 1.312) 0.830
Naphthalene 0.034 (0.027, 0.043) 0.849 (0.703, 0.995) 0.937
Phenanthrene 0.030 (0.024, 0.038) 0.936 (0.816, 1.06) 0.964
Chlorobenzene 0.187 (0.091, 0.385) 0.586 (0.423, 0.749) 0.796
1,4-dichlorobenzene 0.657 (0.401, 1.08) 0.168 (0.068, 0.268) 0.355
Methylbenzene 0.198 (0.132, 0.296) 0.690 (0.576, 0.804) 0.929
1,4-dimethylbenzene 0.045 (0.035, 0.058) 0.655 (0.601, 0.709) 0.975

The uptake of model compounds, with the exception of chlorinated benzenes, is shown in figure 7. The data appear to cluster into two groups: over the range of concentration studied, benzene and toluene comprise one group, and exhibit higher uptake than the group consisting of cyclohexene, p-xylene, naphthalene, and phenanthrene. When normalized for hydrophobic effects, there does not appear to be any significant role of π-electron energy or of polarizability within a single group. Therefore, the hydrophobic effect appears to be the dominant mechanism controlling uptake. The higher uptake of benzene and toluene cannot be explained by specific interactions. It is possible that these compounds can access micropores that the other compounds cannot; or perhaps they have a higher affinity for amorphous carbon present (e.g. an ability to partition into such phases). The chlorinated benzenes exhibit flatter isotherms, which do not fit into either of the groups apparent in figure 7; their data tends to overlap both groups.

Figure 7.

Figure 7. Uptake of model compounds by HipCo SWNTs (carbon nanotechnologies). Solution phase concentrations normalized for hydrophobic effects as described in the text. Solid lines are empirical correlating functions.

Standard image

For comparison, the data for uptake of similar compounds on activated carbon are plotted in figure 8 (Speth and Miltner 1990, Walters and Luthy 1984). With the exception of phenanthrene, the adsorbates are rather tightly grouped after normalizing for hydrophobic effects, although benzene exhibits somewhat lower uptake in the low concentration range. The lower uptake of benzene is in contrast to the findings for the SWNTs. The observation that compounds with either electron withdrawing or electron donating groups appear to exhibit somewhat higher uptake than benzene may indicate enhanced sorption caused by specific interactions, but the effect is not dramatic. The significantly lower uptake of phenanthrene by activated carbon is likely a result of size exclusion effects. Taken as a whole, however, it appears that hydrophobic interactions govern the uptake of the chemicals studied here by both activated carbons and CNTs.

Figure 8.

Figure 8. Uptake of model compounds by F400 activated carbon. Data for naphthalene and phenanthrene are from Walters and Luthy (1984). Data for other compounds is from Speth and Miltner (1990).

Standard image

It is possible that uptake by SWNTs also involves size exclusion effects, which mask possible specific interactions with naphthalene and phenanthrene; i.e. it is possible that, in general, the observed coincidence of the isotherms in both figures 7 and 8 is a result of enhanced sorption due to the effect of functional groups on the electron density of the ring, partially canceled by size exclusion effects. This is difficult to assess in any further detail, and is not considered likely. However, we can conclude that at a minimum, size exclusion effects appear to be less significant for the SWNTs than for the GAC.

4. Conclusions

As-received CNT materials were shown to be good adsorbents for organic compounds of interest in drinking water treatment. Uptake of atrazine by both CNTs and activated carbons was higher than that of TCE, as expected, based on solubility and logKow values. However, the difference was greater for activated carbon, possibly due to the effects of greater microporosity. For single solute uptake of both TCE and atrazine by both SWNTs and activated carbon, there was no effect of pH over the range of pH values 5.7–8.3. As-received CNTs did not offer advantages over commercially available water treatment activated carbons, when compared on a mass basis. This may be related to how these materials aggregate in aqueous suspension; work is currently underway to develop ways to better utilize CNT properties. However, normalizing uptake by both carbon content (to account for purity) and by specific surface area, collapses the data significantly, and indicates that these two characteristics drive much of the observed differences between activated carbons and CNT materials.

Despite aggregation, adsorption sites on CNTs do appear to be more accessible than activated carbons, especially for larger molecules. Based on a comparison to literature data, especially comparing the uptake of naphthalene and phenanthrene by activated carbon and SWNTs, size exclusion effects appear more pronounced with activated carbon materials, perhaps due to smaller pore sizes or larger adsorption surface area in small pores. The uptake of model compounds, with the exception of chlorinated benzenes, appears to cluster into two groups: over the range of concentration studied, benzene and toluene comprise one group, and exhibit higher uptake than the group consisting of cyclohexene, p-xylene, naphthalene, and phenanthrene. When normalized for hydrophobic effects, there does not appear to be any significant role of π-electron density on the aromatic ring, or of polarizability within a single group. Hydrophobic interactions appear to be the dominant mechanism controlling uptake; in this regard, the mechanism dominating uptake by SWNTs is similar to that of activated carbons. Higher uptake of benzene and toluene was observed, and cannot be explained by specific interactions. It is possible that these compounds can access micropores that the other compounds cannot; or perhaps they have a higher affinity for the amorphous carbon present (e.g. an ability to partition into such phases).

Acknowledgments

The Rensselaer Polytechnic Institute thanks the Water Research Foundation (WaterRF) for its financial, technical, and administrative assistance in funding the project through which this information was discovered. The Rensselaer Polytechnic Institute acknowledges that the WaterRF is the joint owner of certain technical information upon which this manuscript is based. This document was reviewed by a panel of independent experts selected by WaterRF prior to submission to the journal. Mention of trade names or commercial products does not constitute WaterRF endorsement or recommendations for use. Similarly, omission of products or trade names indicates nothing concerning WaterRF's position regarding product effectiveness or applicability. The comments and views detailed herein may not necessarily reflect the views of the WaterRF, its officers, directors, affiliates or agents.

Please wait… references are loading.
10.1088/0957-4484/23/29/294008