Next Article in Journal
The Influence of Water Sorption of Dental Light-Cured Composites on Shrinkage Stress
Previous Article in Journal
On the Material Characterisation of Wind Turbine Blade Coatings: The Effect of Interphase Coating–Laminate Adhesion on Rain Erosion Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Aluminum Oxide-Impregnated Carbon Nanotube Membrane for the Removal of Cadmium from Aqueous Solution

1
Center for Environment and Water (CEW), Research Institute, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
2
Department of Mechanical Engineering, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
3
Department of Chemical Engineering, Qatar University, Doha 2713, Qatar
4
College of Science and Engineering, Hamad Bin Khalifa University, Qatar Foundation, Doha 5825, Qatar
*
Authors to whom correspondence should be addressed.
Materials 2017, 10(10), 1144; https://doi.org/10.3390/ma10101144
Submission received: 13 August 2017 / Revised: 11 September 2017 / Accepted: 12 September 2017 / Published: 28 September 2017

Abstract

:
An aluminum oxide-impregnated carbon nanotube (CNT-Al2O3) membrane was developed via a novel approach and used in the removal of toxic metal cadmium ions, Cd(II). The membrane did not require any binder to hold the carbon nanotubes (CNTs) together. Instead, the Al2O3 particles impregnated on the surface of the CNTs were sintered together during heating at 1400 °C. Impregnated CNTs were characterized using XRD, while the CNT-Al2O3 membrane was characterized using scanning electron microscopy (SEM). Water flux, contact angle, and porosity measurements were performed on the membrane prior to the Cd(II) ion removal experiment, which was conducted in a specially devised continuous filtration system. The results demonstrated the extreme hydrophilic behavior of the developed membrane, which yielded a high water flux through the membrane. The filtration system removed 84% of the Cd(II) ions at pH 7 using CNT membrane with 10% Al2O3 loading. A maximum adsorption capacity of 54 mg/g was predicted by the Langmuir isotherm model for the CNT membrane with 10% Al2O3 loading. This high adsorption capacity indicated that adsorption was the main mechanism involved in the removal of Cd(II) ions.

Graphical Abstract

1. Introduction

Cadmium is a well-known highly toxic metal found in drinking water, and is associated with major negative health impacts. The World Health Organization guidelines suggest an allowable limit of cadmium ions, Cd(II), in water of 0.003 mg/L [1]. Cadmium primarily accumulates in the kidneys, and has a relatively long biological half-life of 10 to 35 years in humans [2]. A potential source of cadmium contamination in drinking water is industrial wastewater, such as that produced by the manufacturing processes for smelting, pesticides, fertilizers, dyes, pigments, refining, and textile operations. Cadmium contamination of drinking water might also be caused by the presence of Cd(II) ions as an impurity in the zinc of galvanized pipes and certain metal fittings [3,4].
Various toxic metal decontamination techniques (such as adsorption, precipitation, reduction, ion exchange, precipitation, solvent extraction, electrolytic recovery, and chemical oxidation) have been applied for the removal of toxic metals from water [3,5]. However, the majority of these methods have limited applications due to economical or technical constraints. Water treatment by adsorption offers the most practical and economical treatment alternative. Numerous adsorbents have been used in the removal of metal ions from water, including activated carbon, fly ash, biomaterials, zeolite, recycled alum sludge, algal biomass, peanut hulls, resins, kaolinite, manganese oxides, and carbon nanotubes [6]. However, the majority of the adsorption techniques are batch-level processes and are unable to process large amounts of contaminated water. Carbon nanotube-based novel membranes are a promising candidate for toxic metal removal.
CNTs have attracted considerable attention in recent years as a novel adsorbent for the adsorption of numerous pollutants from water, including toxic metal ions [3,6] and organic chemicals [7,8,9,10]. CNTs have also emerged as an ideal candidate for the synthesis of unique membranes with excellent properties for applications in water treatment [11,12,13]. The friction-less and smooth graphitic walls of CNTs are considered ideal channels for enhanced molecule transport [13,14,15]. CNTs can be used either as fillers to improve the mechanical, electrical, and thermal properties of various polymeric membranes or as a direct filter [16,17,18,19].
Different types of CNT-based membranes have been reported in the published literature. The most common categories include mixed-matrix membranes [16,17,18,19], vertically aligned CNT membranes [20,21], bucky paper membranes [22,23,24], and template-assisted open-ended CNT membranes [25,26].
In this work, a novel approach for the synthesis of CNT membranes is presented using a powder metallurgy technique. To avoid heating at extremely high temperatures (approximately 3000 °C) to bond the CNTs together, aluminum oxide (Al2O3) particles are impregnated onto the surface of CNTs and used to bond the 3D CNT network together during a sintering process performed at only 1400 °C to produce the CNT membrane. Moreover, surface impregnation of CNTs has been reported to yield an enhanced surface area and better adsorption capacity for the removal of different contaminants from water [27]. CNTs were impregnated via a wet chemistry technique with different loadings of aluminum oxide. Impregnated CNTs were characterized using XRD, and the developed membranes were analyzed to measure the contact angle, porosity, and water flux. The potential of the membrane for cadmium removal was investigated using a continuous flow system. The effect of pH, aluminum oxide loading onto the CNTs, initial cadmium concentration of the solution, and time, on the removal efficiency of Cd(II) ions, was investigated.

2. Materials and Methods

2.1. Materials

CNTs with a purity >95% were acquired from Chengdu Organic Chemicals Co. Ltd., China (Chengdu, China) with an outside diameter of 10–20 nm and a length of 1–10 μm. Aluminum isopropoxide [C9H21O3Al] (purity ≥ 98%, Sigma Aldrich, Saint Louis, MO, USA) was used as a precursor for aluminum oxide.

2.2. Impregnated Aluminum Oxide-Carbon Nanotubes

The surfaces of the CNTs were impregnated with Al2O3 nanoparticles via the wet impregnation method [28,29]. The loading content of Al2O3 ranged from 1 to 20% (by weight, wt %). In this method, to prepare 10% Al2O3 loading, 7.5 g of aluminum isopropoxide [C9H21O3Al] was dissolved in 500 mL of ethanol (98% purity). Nine grams of CNTs were dispersed in a separate 500 mL of ethanol. These mixtures were separately ultrasonicated for 1 h and subsequently mixed together. The mixture was further ultrasonicated for 1 h at ambient temperature, to ensure a homogeneous dispersion of the CNTs. To remove the ethanol solvent, the mixture was stored in an oven overnight at the required temperature. The residue was calcinated in a tube furnace under argon at 350 °C for 4 h. This process results in the attachment of Al2O3 particles onto the surface of CNTs. A similar procedure was used to prepare CNTs with various Al2O3 loading percentages (between 1% and 20%).

2.3. Membrane Preparation

The Al2O3-impregnated CNTs, prepared as described in Section 2.1, were uniaxially pressed at 200 MPa compaction pressure using a circular metallic die in a hydraulic bench-type press (4387 N.E.L, Carver, Inc., Wabash, IN, USA). This process resulted in a compact disc (27 mm in diameter and 3 mm thick) containing 1%, 10%, and 20% Al2O3. The compacted discs were sintered at 1400 °C for 5 h under argon (300–400 mL/min) in a tube furnace (GSL-1700X, MTI Corporation, Richmond, CA, USA). A schematic representation of the synthesis route is shown in Figure 1.

2.4. Characterization Analysis of Raw and Impregnated CNTs and Membranes

2.4.1. SEM Analysis

SEM studies of the membranes were performed using field emission scanning electron microscopy (TESCAN MIRA 3 FEG-SEM, TESCAN, Brno-Kohoutovic, Czech Republic).

2.4.2. X-ray Diffraction (XRD)

The XRD patterns for raw and impregnated CNTs were measured at a rate of 1.0°/min in the range of 10°–80° (2α) using an X-ray diffractometer equipped with a Cu Kα radiation source (40 kV, 20 mA).

2.4.3. Porosity Measurement

The dry-wet method [16] was used to determine the porosity of the membranes using Equation (1):
Porosity = W 2 W 1 ρ · V × 100 %
where W1 (g) and W2 (g) are the weights of the dry and wet membranes, respectively; V (cm3) is the volume of the membrane; and ρ (g/cm3) is the density of distilled water at ambient temperature. The membrane was immersed in distilled water for 24 h, and its wet weight was measured. The membrane was subsequently dried in an oven at 90 °C for 24 h, and the dry weight of the membrane was measured. The experiment was performed in triplicate, and the data are presented as the mean value of all experiments.

2.4.4. Contact Angle Measurement

A contact angle analyzer (model DM-301, KYOWA, Niiza, Japan) was used to measure the contact angle of the membrane surface and hence the hydrophobicity/hydrophilicity of the membrane.

2.4.5. Zeta Potential Measurement

The zeta potential for a suspension of 0.5 g/L CNTs-Al2O3 in distilled water was determined using a Malvern ZEN2600 Zetasizer Nano Z (Malvern, Worcestershire, UK) at pH 2.0–10 (adjusted with 0.1 M NaOH or HNO3).

2.5. Continuous Filtration System

The main components of the continuous filtration system used in this study (presented in Figure 2) were a membrane cell with an effective surface area of 5.7 cm2, a 10 L feed tank, and the required pressure pump and flow meter. The membrane was placed in a circular housing with a mesh underneath it as a support structure to maintain the stability of the membrane during the flow experiments. The pure water flux analysis was performed before the Cd(II) remediation studies were performed. The pure water flux was measured using Equation (2) [16,28,29]:
J = V / ( A · t )
where J (L·m−2·h−1) is the water flux, t (h) is the time required for permeate water to pass through the membrane, and V (L) is the volume of permeate water.
For cadmium removal, the experimental runs began with the circulation of a 1 ppm solution of Cd(II) from the feed tank through the system. An initial volume of approximately 10 L was added to the feed tank, and the pH was adjusted using 1 M NaOH or 1 M HNO3, as required. The pressure and flow rate were adjusted to the desired values. Permeate (purified water) passing through the membrane was collected from the sample collection point (shown in Figure 2) at different time intervals using sample bottles with volumes of approximately 20 mL. The effects of the initial concentration, time, pH, and membrane Al2O3 loading on cadmium removal were studied.
Table 1 shows the experimental conditions for Cd(II) removal using different CNT-Al2O3 membranes. First, a CNT membrane with 10% Al2O3 loading was used to determine the optimum pH for the maximum removal of Cd(II). The transmembrane pressure difference and concentration of Cd(II) were held constant during these experiments. The optimum pH (pH = 7) was held constant during the remainder of the experiments, and the effects of Al2O3 loading and the initial Cd(II) concentration in the solution (water) on the removal efficiency of Cd(II) ions were determined.

2.6. Analytical Methods

Inductively coupled plasma mass spectrometry (X-Series 2 Q-ICP-MS, Thermo Fisher Scientific, Waltham, MA, USA) was used to measure the concentration of Cd(II) before and after the experiments.

3. Results and Discussion

3.1. SEM and EDS Analysis

Figure 3 shows the SEM images of the prepared membranes with various Al2O3 contents. The particles are well dispersed at low Al2O3 loadings, whereas the particles tend to agglomerate at higher Al2O3 loadings (i.e., 20% or above).

3.2. X-ray Diffraction (XRD)

Figure 4 displays the XRD patterns of the raw and impregnated CNTs powders. In the XRD pattern of the raw CNTs, the characteristic peaks at 2θ = 27° and 44° correspond to the CNTs. However, the XRD pattern of the impregnated CNTs presents new peaks at 2θ = 17°, 33°, and 40° in addition to the two apparent peaks associated with the CNTs. These peaks correspond to the Al2O3 nanoparticles and indicate the successful impregnation of Al2O3 particles onto the surface of the CNTs. The XRD spectrum shows a minor shift of CNTs peaks for the CNT-Al2O3, possibly due to the residual stresses imposed by the Al2O3 particles embedded onto the CNTs [30,31].

3.3. Measurement of the Zeta Potential and Point of Zero Electric Charge (pHPZC)

The zeta potentials of CNT-10% Al2O3 in distilled water were determined in a pH range of 2.0–10. As displayed in Figure 5, the surface charge of the membrane surface is positive at pH < 6.5 and negative at pH > 6.5. The zero electric charge (pHPZC) value of the membrane was noted at 6.5. The membrane is expected to have a relatively higher removal of Cd(II) ions at pH 7, as discussed in Section 3.6, because of the electrostatic interactions between the negatively charged membrane surface and cationic Cd(II) ions.

3.4. Membrane Characterization

3.4.1. Porosity Measurement

The dry-wet method was used to determine the porosity of the membranes. Figure 6 displays the porosity versus Al2O3 loading for loadings of 1 to 20%. Considering the standard deviation reported for the measured porosity values, the variation in porosity with Al2O3 content is rather minor, implying that the Al2O3 loading does not have a significant effect on the membrane porosity.

3.4.2. Contact Angle Measurement

Contact angle measurement is an index of the hydrophobicity/hydrophilicity of the membrane surface. As shown in Figure 7, the contact angle of the membrane decreased with increasing Al2O3 content. In other words, the hydrophilicity of the membrane increases with increasing Al2O3 loading. The decrease in contact angle with increase in Al2O3 loading might be due to change in membrane pore size due to agglomeration at higher Al2O3 loading. This agglomeration leads to enhanced water flux (as shown in Figure 8), and hence, the contact angle value is decreased. This hydrophilic nature of the membrane is primarily responsible for the enhanced flux through the membrane [28,29].

3.5. Water Flux Measurements: Effect of Transmembrane Pressure Difference and Aluminum Oxide Loading

The effects of transmembrane pressure difference, aluminum oxide loading, and time, on the water flux through the membranes, were studied, as shown in Figure 8. The transmembrane pressure difference was varied from 1 to 40 psi. A nearly linear relationship exists between the pressure and flux for all membranes with different Al2O3 loadings. The water permeate flux was measured by holding the transmembrane pressure difference constant at 20 psi for 30 min. The permeate flux values for different pressures were obtained using the same procedure. For each reading, the pressure was maintained for 10 min before noting the readings.
Figure 8 illustrates that the permeate water flux increased as the Al2O3 content increased from 1 to 20%. The increased permeate flux for membranes with high Al2O3 loading can be explained based on two mechanisms. First, the hydrophilic surface of the membrane at high Al2O3 loading facilitates the transport of water through the membrane (Figure 7). Second, the agglomeration of Al2O3 particles (Figure 3c) at high loading (20%) results in the formation of relatively large pores in the membrane, thus contributing to the higher permeate flux.

3.6. Cadmium Removal

The Cd(II) ion removal studies were performed in the flow loop system, as shown in Figure 2. The cadmium solution was passed through the CNT-Al2O3 membrane. Cd(II) ions are retained in the feed side while purified water permeates the membrane. The cadmium removal (R) can be determined using the following equation:
R = 1 C p / C f
where Cp and Cf are the concentrations of the solute in the permeate and feed, respectively.

3.6.1. Effect of Feed pH

Solution pH is an important parameter that determines the removal of toxic metal by carbon-based materials. The Cd(II)removal experiments were performed for the membrane with 10% Al2O3 loading in the pH range of 3–10 (experimental set 1). The results of the analysis are shown in Figure 9.
Cadmium species are present in deionized (DI) water in the form of Cd2+, Cd(OH)+, and Cd(OH)2(s) [32,33]. At pH < 8, the dominant cadmium species is Cd2+ in the form of complex [Cd(H2O)6]2+ [34]. The pHPZC value for Al2O3-doped CNTs is pH 6.5, as shown in Figure 4. This value demonstrates that Al2O3 is basic in DI water [27]. At pH < pHPZC, the membrane surface is positively charged, and repulsion exists between the Cd(II) ions and surface, causing a low removal rate for Cd(II) ions. In addition, competition between H+ and Cd2+ ions for the active sites decreases the Cd(II) ions adsorption rate. At pH > pHPZC, the surface of the membrane becomes more negatively charged, and thus, additional Cd(II) ions are attracted to the surface, due to electrostatic interactions.
A maximum removal of 84% was observed at pH 8. However, cadmium might have precipitated as Cd(OH)2 at pH > 8, as reported elsewhere [35,36], and the removal might be due to both adsorption and precipitation. Therefore, pH 7 was used in all experiments as an optimum value to avoid the precipitation of cadmium ions. Moreover, at pH > 8, the concentration of Cd(II) ions is low, and the predominant ions are HCO 3 . The repulsion between the negatively charged surface of the membrane at pH > pHPZC, and the HCO 3 ions, causes a decrease in the removal of cadmium ions. In addition to electrostatic interactions, the van der Waals interactions occurring between the cadmium ions and carbon atoms could also induce the adsorption of cadmium ions [27,37].

3.6.2. Effect of Time

To study the effect of time on the removal of Cd(II) ions, the experiments were performed at constant pH 7, and an initial concentration of 1 ppm. The samples were collected every 30 min and analyzed. The results of the analysis are presented in Figure 10. The percentage removal of Cd(II) increases with time until 2 h of operation, after which no significant increase in removal was observed. Equilibrium was reached within nearly 2 h for all membranes. The maximum removal rate of 84% was achieved with the CNT-10% Al2O3 membrane. Membranes with 1% and 20% Al2O3 loadings achieved Cd(II) removal rates of 80% and 74%, respectively, under similar conditions.
The relatively higher removal of Cd(II) ions by the CNT-10% Al2O3 membrane might be due to the greater number of adsorption sites than the membrane with 1% Al2O3 loading. The relatively lower removal of Cd(II) ions by the membrane with 20% Al2O3 loading could be attributed to the agglomeration of Al2O3 particles at higher loading i.e., 20% Al2O3 (as discussed in Section 3.1 (Figure 3c)). This agglomeration might create large pores in the membrane that leads to poor separation of Cd(II) ions.

3.6.3. Effect of the Initial Concentration

Figure 11 presents the effect of the initial concentration of the solution on the percentage removal of Cd(II) ions. The initial concentration was varied from 0.5 to 10 ppm, and the other experimental parameters were pH 6, a contact time of 2 h, and a transmembrane pressure difference of 15 psi.
The removal increased slightly from 78 to 84%, as the concentration increased from 0.5 to 1 ppm, and remained nearly constant until approximately 5 ppm. At an initial concentration of 0.5 ppm, the membrane still had available adsorption sites and did not reach saturation; thus, more Cd(II) ions could be adsorbed. The percentage removal remained constant as the concentration was increased from 1 to 5 ppm. However, as the concentration increased further beyond 5 ppm, the removal decreased slightly, likely because all available adsorption sites were covered by the Cd(II) ions and the membrane had reached its adsorption equilibrium limit. The membrane was effective in removing a low concentration of Cd(II) ions.

3.6.4. Adsorption Isotherms

The nonlinear forms of the Langmuir and Freundlich adsorption isotherms for the adsorption of cadmium ions on the CNT-Al2O3 membrane surface are presented in Figure 12. Representative equations and the results of the analysis are summarized in Table 2.
In the Langmuir model, qe (mg/g) represents the concentration of adsorbate on the surface of adsorbent, Ce (mg/L) indicates the concentration of adsorbate in solution when equilibrium was reached, qm (mg/g) is the maximum adsorption capacity, and KL is the Langmuir adsorption equilibrium constant (L/mg). In the Freundlich isotherm model, KF (mg/g)·(L/mg)1/n and n (dimensionless) are Freundlich constants. Mathematica version 10 (Wolfram 2015, Long Hanborough Oxfordshire, UK) was used to plot the isotherm data and determine the values of various parameters.
Both models can describe the experimental data satisfactorily, but the correlation coefficient (R2) value for the Langmuir isotherm model is slightly higher than that for the Freundlich isotherm model.

3.7. Mechanism of Cadmium Ion Removal by the CNT-10% Al2O3 Membrane

The possible mechanism underlying Cd(II) ion interactions with the CNT-Al2O3 membrane is presented in Figure 13. As discussed in Section 3.6.1, the dominant cadmium species in deionized (DI) water is Cd(II), or Cd2+, in the form of complex [Cd(H2O)6]2+ at pH 7, used as an optimum value in all experiments. When pH < pHPZC, the membrane surface is positively charged, and the low removal of Cd(II) ions can be attributed to electrostatic repulsions between the Cd(II) ions and the surface of the CNT-Al2O3 membrane. Similarly, the higher removal of Cd(II) ions at pH > pHPZC might be due to the strong electrostatic interactions between the negatively charged CNT-Al2O3 membrane surface and the cationic Cd(II) ions [27].
This observation suggests that electrostatic interaction is the main mechanism involved in the sorption of Cd(II) ions onto the CNT-Al2O3 membrane surface. In addition to electrostatic interaction, Cd(II) ions might also adsorb on the surface of CNT-Al2O3 membrane due to van der Waals interactions (physical adsorption) occurring between Cd(II) ions and carbon atoms in the CNT-Al2O3 composite. At the adsorption saturation of the membrane surface and internal structure, certain pores among the CNTs might be covered by Cd(II) ions, thus blocking further ions from passing through and acting as sieves (size exclusion).

3.8. Comparative Analysis

The adsorption capacity and removal efficiency of Cd(II) ions from water for the membrane developed in this study are compared to those of similar studies reported in the literature in Table 3. The estimated maximum Cd(II) adsorption capacity of the CNT-10% Al2O3 membrane is 54.42 mg/g, which is higher than those of related adsorbents (used in batch experiments), such as raw CNTs (1.661 mg/g) [38], acid-modified CNTs (4.35 mg/g) [3], ethylenediamine-functionalized multi-walled carbon nanotubes (MWCNTs) (25.70 mg/g) [33] and nano-alumina on single-walled carbon nanotube (SWCNTs) (2.18 mg/g) [39]. This result suggests that the CNT-Al2O3 membrane is effective in removing low concentrations of Cd(II) ions from water.

4. Conclusions

A novel approach was developed to synthesize an aluminum oxide-impregnated CNT membrane. No binder was used in the membrane synthesis; instead, aluminum oxide particles served as a binder to hold the 3D CNT network together. The membrane surface demonstrated extreme hydrophilic behavior and yielded a high water flux. The membrane was able to remove low concentrations of Cd(II) ions from aqueous solution. The removal was affected by the aluminum oxide loading, initial cadmium solution concentration, pH, and time. The maximum Cd(II) removal of 84% was obtained at pH 7 and an initial concentration of 1 ppm using the CNT membrane with 10% Al2O3 loading and 2 h of operation. Membranes with 1% and 20% Al2O3 loadings were able to remove 80% and 74% of Cd(II) ions, respectively, under similar experimental conditions. These results suggest that the CNT-Al2O3 membrane can be effectively used in a continuous filtration system for the removal of Cd(II) ions.

Acknowledgments

The authors acknowledge support from King Abdulaziz City for Science and Technology (KACST) through the Science & Technology Unit at King Fahd University of Petroleum & Minerals (KFUPM) and funding of this work through Project No: 13-ADV161-04 as part of the National Science Technology and Innovation Plan (NSTIP).

Author Contributions

Ihsanullah, Muataz Ali Atieh and Tahar Laoui conceived and designed the experiments; Ihsanullah performed the experiments; Ihsanullah and Tahar Laoui analysed the data; Faheemuddin Patel contributed reagents/materials/analysis tools; Ihsanullah, Tahar Laoui and Majeda Khraisheh wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization. International Standards for Drinking Water; WHO: Geneva, Switzerland, 2008. [Google Scholar]
  2. Li, Y.; Wang, S.; Luan, Z.; Ding, J.; Xu, C. Adsorption of cadmium(II) from aqueous solution by surface oxidized carbon nanotubes. Carbon 2003, 41, 1057–1062. [Google Scholar] [CrossRef]
  3. Ihsanullah; Al-Khaldi, F.A.; Abusharkh, B.; Khaled, M.; Atieh, M.A.; Nasser, M.S.; Laoui, T.; Saleh, T.A.; Agarwal, S.; Tyagi, I.; et al. Adsorptive removal of cadmium(II) ions from liquid phase using acid modified carbon-based adsorbents. J. Mol. Liq. 2015, 204, 255–263. [Google Scholar] [CrossRef]
  4. Wang, F.Y.; Wang, H.; Ma, J.W. Adsorption of cadmium(II) ions from aqueous solution by a new low-cost adsorbent—Bamboo charcoal. J. Hazard. Mater. 2010, 177, 300–306. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Q.; Wu, S.; Liu, G.; Liao, X.; Deng, X.; Sun, D.; Hu, Y.; Huang, Y. Simultaneous biosorption of cadmium(II) and lead(II) ions by pretreated biomass of phanerochaete chrysosporium. Sep. Purif. Technol. 2004, 34, 135–142. [Google Scholar] [CrossRef]
  6. Ihsanullah; Abbas, A.; Al-Amer, A.M.; Laoui, T.; Al-Marri, M.J.; Nasser, M.S.; Khraisheh, M.; Atieh, M.A. Heavy metal removal from aqueous solution by advanced carbon nanotubes: Critical review of adsorption applications. Sep. Purif. Technol. 2016, 157, 141–161. [Google Scholar] [CrossRef]
  7. Asmaly, H.A.; Abussaud, B.; Ihsanullah; Saleh, T.A.; Gupta, V.K.; Atieh, M.A. Ferric oxide nanoparticles decorated carbon nanotubes and carbon nanofibers: From synthesis to enhanced removal of phenol. J. Saudi Chem. Soc. 2015, 19, 511–520. [Google Scholar] [CrossRef]
  8. Asmaly, H.A.; Abussaud, B.; Ihsanullah; Saleh, T.A.; Bukhari, A.A.; Laoui, T.; Shemsi, A.M.; Gupta, V.K.; Atieh, M.A. Evaluation of micro- and nano-carbon-based adsorbents for the removal of phenol from aqueous solutions. Toxicol. Environ. Chem. 2015, 97, 1164–1179. [Google Scholar] [CrossRef]
  9. Abbas, A.; Abussaud, B.A.; Al-baghli, N.A.H.; Khraisheh, M.; Atieh, M.A. Benzene removal by iron oxide nanoparticles decorated carbon nanotubes. J. Nanomater. 2016, 2016, 5654129. [Google Scholar] [CrossRef]
  10. Abbas, A.; Abussaud, B.A.; Ihsanullah; Al-baghli, N.A.H.; Redhwi, H.H. Adsorption of Toluene and Paraxylene from aqueous solution using pure and iron oxide impregnated carbon nanotubes: Kinetics and isotherms study. Bioinorg. Chem. Appl. 2017, 2017, 2853925. [Google Scholar] [CrossRef] [PubMed]
  11. Ho, K.C.; Teow, Y.H.; Ang, W.L.; Mohammad, A.W. Novel GO/OMWCNTs mixed-matrix membrane with enhanced antifouling property for palm oil mill effluent treatment. Sep. Purif. Technol. 2017, 177, 337–349. [Google Scholar] [CrossRef]
  12. Li, S.; Liao, G.; Liu, Z.; Pan, Y.; Wu, Q.; Weng, Y.; Zhang, X.; Yang, Z.; Tsui, O.K.C. Enhanced water flux in vertically aligned carbon nanotube arrays and polyethersulfone composite membranes. J. Mater. Chem. A 2014, 2, 12171–12176. [Google Scholar] [CrossRef]
  13. Majumder, M.; Chopra, N.; Andrews, R.; Hinds, B.J. Nanoscale hydrodynamics: Enhanced flow in carbon nanotubes. Nature 2005, 438, 44. [Google Scholar] [CrossRef] [PubMed]
  14. Holt, J.K.; Park, H.G.; Wang, Y.; Stadermann, M.; Artyukhin, A.B.; Grigoropoulos, C.P.; Noy, A.; Bakajin, O. Fast mass transport through sub-2-nanometer carbon nanotubes. Science 2006, 312, 1034–1037. [Google Scholar] [CrossRef] [PubMed]
  15. Srivastava, A.; Srivastava, O.N.; Talapatra, S.; Vajtai, R.; Ajayan, P.M. Carbon nanotube filters. Nat. Mater. 2004, 3, 610–614. [Google Scholar] [CrossRef] [PubMed]
  16. Majeed, S.; Fierro, D.; Buhr, K.; Wind, J.; Du, B.; Boschetti-de-Fierro, A.; Abetz, V. Multi-walled carbon nanotubes (MWCNTs) mixed polyacrylonitrile (PAN) ultrafiltration membranes. J. Membr. Sci. 2012, 403–404, 101–109. [Google Scholar] [CrossRef]
  17. Deb, A.K.S.; Dwivedi, V.; Dasgupta, K.; Ali, S.M.; Shenoy, K.T. Novel amidoamine functionalized multi-walled carbon nanotubes for removal of mercury(II) ions from wastewater: Combined Experimental and density functional theoretical approach. Chem. Eng. J. 2017, 313, 899–911. [Google Scholar]
  18. Lee, J.; Ye, Y.; Ward, A.J.; Zhou, C.; Chen, V.; Andrew, I.; Lee, S.; Liu, Z.; Chae, S.; Shi, J. High flux and high selectivity carbon nanotube composite membrane for natural organic matter removal. Sep. Purif. Technol. 2016, 163, 109–119. [Google Scholar] [CrossRef]
  19. Zarrabi, H.; Ehsan, M.; Vatanpour, V.; Shockravi, A.; Safarpour, M. Improvement in desalination performance of thin film nanocomposite nano filtration membrane using amine-functionalized multiwalled carbon nanotube. Desalination 2016, 394, 83–90. [Google Scholar] [CrossRef]
  20. Hinds, B.J.; Chopra, N.; Rantell, T.; Andrews, R.; Gavalas, V.; Bachas, L.G. Aligned multiwalled carbon nanotube membranes. Science 2004, 303, 62–65. [Google Scholar] [CrossRef] [PubMed]
  21. Andrews, R.; Jacques, D.; Rao, A.M.; Derbyshire, F.; Qian, D.; Fan, X.; Dickey, E.C.; Chen, J. Continuous production of aligned carbon nanotubes: A step closer to commercial realization. Chem. Phys. Lett. 1999, 303, 467–474. [Google Scholar] [CrossRef]
  22. Pham, G.T.; Park, Y.-B.; Wang, S.; Liang, Z.; Wang, B.; Zhang, C.; Funchess, P.; Kramer, L. Mechanical and electrical properties of polycarbonate nanotube buckypaper composite sheets. Nanotechnology 2008, 19, 325705. [Google Scholar] [CrossRef] [PubMed]
  23. Dumée, L.F.; Sears, K.; Schütz, J.; Finn, N.; Huynh, C.; Hawkins, S.; Duke, M.; Gray, S. Characterization and evaluation of carbon nanotube Bucky-Paper membranes for direct contact membrane distillation. J. Membr. Sci. 2010, 351, 36–43. [Google Scholar] [CrossRef] [Green Version]
  24. Cooper, S.M.; Chuang, H.F.; Cinke, M.; Cruden, B.A.; Meyyappan, M. Gas permeability of a buckypaper membrane. Nano Lett. 2003, 3, 189–192. [Google Scholar] [CrossRef]
  25. Wang, C.; Li, M.; Pan, S.; Li, H. Well-aligned carbon nanotube array membrane synthesized in porous alumina template by chemical vapor deposition. Chin. Sci. Bull. 2000, 45, 1373–1376. [Google Scholar] [CrossRef]
  26. Tofighy, M.A.; Shirazi, Y.; Mohammadi, T.; Pak, A. Salty water desalination using carbon nanotubes membrane. Chem. Eng. J. 2011, 168, 1064–1072. [Google Scholar] [CrossRef]
  27. Liang, J.; Liu, J.; Yuan, X.; Dong, H.; Zeng, G.; Wu, H.; Wang, H.; Liu, J.; Hua, S.; Zhang, S.; et al. Facile synthesis of alumina-decorated multi-walled carbon nanotubes for simultaneous adsorption of cadmium ion and trichloroethylene. Chem. Eng. J. 2015, 273, 101–110. [Google Scholar] [CrossRef]
  28. Ihsanullah; Al Amer, A.M.; Laoui, T.; Abbas, A.; Al-aqeeli, N.; Patel, F.; Khraisheh, M.; Ali, M.; Hilal, N. Fabrication and antifouling behaviour of a carbon nanotube membrane. Mater. Des. 2016, 89, 549–558. [Google Scholar] [CrossRef]
  29. Ihsanullah; Laoui, T.; Al-Amer, A.M.; Khalil, A.B.; Abbas, A.; Khraisheh, M.; Atieh, M.A. Novel anti-microbial membrane for desalination pretreatment: A silver nanoparticle-doped carbon nanotube membrane. Desalination 2015, 376, 82–91. [Google Scholar] [CrossRef]
  30. Wei, T.; Fan, Z.; Luo, G.; Wei, G. A new structure for multi-walled carbon nanotubes reinforced alumina nanocomposite with high strength and toughness. Mater. Lett. 2008, 62, 641–644. [Google Scholar] [CrossRef]
  31. Wang, K.; Wang, Y.; Fan, Z.; Yan, J.; Wei, T. Preparation of graphene nanosheet/alumina composites by spark plasma sintering. Mater. Res. Bull. 2011, 46, 315–318. [Google Scholar] [CrossRef]
  32. Leyva-Ramos, R.; Rangel-Mendez, J.R.; Mendoza-Barron, J.; Fuentes-Rubio, L.; Guerrero-Coronado, R.M. Adsorption of cadmium(II) from aqueous solution onto activated carbon. Water Sci. Technol. 1997, 35, 205–211. [Google Scholar] [CrossRef]
  33. Vukovic, G.D.; Marinkovic, A.D.; Coli, M.; Risti, M.Ð.; Aleksić, R.; Perić-Grujić, A.A.; Uskoković, P.S. Removal of cadmium from aqueous solutions by oxidized and ethylenediamine-functionalized multi-walled carbon nanotubes. Chem. Eng. J. 2010, 157, 238–248. [Google Scholar] [CrossRef]
  34. Ozutsumi, K.; Takamuku, T.; Shin-ichi, I.; Ohtaki, H. An X-ray diffraction study on the structure of solvated cadmium(II) ion and tetrathiocyanatocadmate(II) complex in N,N-dimethylformamide. Bull. Chem. Soc. Jpn. 1989, 62, 1875–1879. [Google Scholar] [CrossRef]
  35. Ahn, C.K.; Kim, Y.M.; Woo, S.H.; Park, J.M. Removal of cadmium using acid-treated activated carbon in the presence of nonionic and/or anionic surfactants. Hydrometallurgy 2009, 99, 209–213. [Google Scholar] [CrossRef]
  36. Ahn, C.K.; Park, D.; Woo, S.H.; Park, J.M. Removal of cationic heavy metal from aqueous solution by activated carbon impregnated with anionic surfactants. J. Hazard. Mater. 2009, 164, 1130–1136. [Google Scholar] [CrossRef] [PubMed]
  37. Gao, Z.; Bandosz, T.J.; Zhao, Z.; Han, M.; Qiu, J. Investigation of factors affecting adsorption of transition metals on oxidized carbon nanotubes. J. Hazard. Mater. 2009, 167, 357–365. [Google Scholar] [CrossRef] [PubMed]
  38. Al-Khaldi, F.A.; Abu-Sharkh, B.; Abulkibash, A.M.; Atieh, M.A. Cadmium removal by activated carbon, carbon nanotubes, carbon nanofibers, and carbon fly ash: A comparative study. Desalin. Water Treat. 2015, 53, 1417–1429. [Google Scholar] [CrossRef]
  39. Kalfa, O.M.; Yalçinkaya, Ö.; Türker, A.R. Synthesis and characterization of nano-scale alumina on single walled carbon nanotube. Inorg. Mater. 2009, 45, 988–992. [Google Scholar] [CrossRef]
  40. Moradi, O. The removal of ions by functionalized carbon nanotube: Equilibrium, isotherms and thermodynamic studies. Chem. Biochem. Eng. Q. 2011, 25, 229–240. [Google Scholar]
Figure 1. Flowchart for the synthesis of aluminum oxide-impregnated carbon nanotube (CNTs–Al2O3) membrane.
Figure 1. Flowchart for the synthesis of aluminum oxide-impregnated carbon nanotube (CNTs–Al2O3) membrane.
Materials 10 01144 g001
Figure 2. Schematic diagram of the flow loop system.
Figure 2. Schematic diagram of the flow loop system.
Materials 10 01144 g002
Figure 3. SEM micrographs of CNT membranes impregnated with (a) 1%; (b) 10%; and (c) 20% Al2O3 particles.
Figure 3. SEM micrographs of CNT membranes impregnated with (a) 1%; (b) 10%; and (c) 20% Al2O3 particles.
Materials 10 01144 g003
Figure 4. XRD plots of raw and Al2O3-impregnated CNTs.
Figure 4. XRD plots of raw and Al2O3-impregnated CNTs.
Materials 10 01144 g004
Figure 5. Variation of the zeta potential values with pH for the CNT-Al2O3 membrane.
Figure 5. Variation of the zeta potential values with pH for the CNT-Al2O3 membrane.
Materials 10 01144 g005
Figure 6. Porosity of the Al2O3-impregnated CNT membranes.
Figure 6. Porosity of the Al2O3-impregnated CNT membranes.
Materials 10 01144 g006
Figure 7. Contact angle of the membrane versus Al2O3 content.
Figure 7. Contact angle of the membrane versus Al2O3 content.
Materials 10 01144 g007
Figure 8. Effect of transmembrane pressure difference on the water flux of the CNT-Al2O3 membranes.
Figure 8. Effect of transmembrane pressure difference on the water flux of the CNT-Al2O3 membranes.
Materials 10 01144 g008
Figure 9. Effect of pH on the percentage removal of Cd(II) ions by CNT-10% Al2O3 membrane (initial concentration = 1 ppm, time = 3 h).
Figure 9. Effect of pH on the percentage removal of Cd(II) ions by CNT-10% Al2O3 membrane (initial concentration = 1 ppm, time = 3 h).
Materials 10 01144 g009
Figure 10. Effect of time on the percentage removal of cadmium ions by the CNT-Al2O3 membranes (initial concentration = 1 ppm, pH = 7).
Figure 10. Effect of time on the percentage removal of cadmium ions by the CNT-Al2O3 membranes (initial concentration = 1 ppm, pH = 7).
Materials 10 01144 g010
Figure 11. Effect of the initial concentration on the percentage removal of cadmium ions by the CNT-10% Al2O3 membrane (contact time = 2 h, pH = 7).
Figure 11. Effect of the initial concentration on the percentage removal of cadmium ions by the CNT-10% Al2O3 membrane (contact time = 2 h, pH = 7).
Materials 10 01144 g011
Figure 12. Langmuir and Freundlich adsorption isotherm model fits for the removal of cadmium ions by the CNT-10% Al2O3 membrane.
Figure 12. Langmuir and Freundlich adsorption isotherm model fits for the removal of cadmium ions by the CNT-10% Al2O3 membrane.
Materials 10 01144 g012
Figure 13. Possible mechanisms of Cd(II) ion interactions with the CNT-Al2O3 membrane.
Figure 13. Possible mechanisms of Cd(II) ion interactions with the CNT-Al2O3 membrane.
Materials 10 01144 g013
Table 1. Experimental matrix of parameters used in the removal of cadmium with the developed membranes.
Table 1. Experimental matrix of parameters used in the removal of cadmium with the developed membranes.
Experimental SetAl2O3 Loading (wt %) in the MembraneTransmembrane Pressure Difference (psi)pH of the SolutionInitial Concentration of Cd(II) (ppm)
1CNT-10% Al2O153, 5, 7, 8, 101
2CNT-1% Al2O31570.5, 1, 5, 10
3CNT-10% Al2O31570.5, 1, 5, 10
4CNT-20% Al2O31570.5, 1, 5, 10
Table 2. Langmuir and Freundlich isotherm model parameters for the adsorption of Cd(II) ions on the CNT-10% Al2O3 membrane.
Table 2. Langmuir and Freundlich isotherm model parameters for the adsorption of Cd(II) ions on the CNT-10% Al2O3 membrane.
ModelParametersCNT-10% Al2O3 Membrane
Langmuir ( q e = q m   K L C e 1 + K L C e )KL (L/mg)0.09
qm (mg/g)54.42
R20.997
Freundlich ( q e = K F · C e 1 / n )KF (mg/g)·(L/mg)1/n4.89
N1.35
R20.99
Table 3. Comparative analysis of the adsorption capacity and removal efficiency of Cd(II) ion removal.
Table 3. Comparative analysis of the adsorption capacity and removal efficiency of Cd(II) ion removal.
AdsorbentExperimental ConditionsPercentage RemovalAdsorption Capacity (mg/g)Reference
As-grown CNTspH 5.5, initial concentration = 4 mg/L-1.1[2]
H2O2 oxidized CNTspH 5.5, initial concentration = 4 mg/L-2.6[2]
HNO3 oxidized CNTspH 5.5, initial concentration = 4 mg/L-5.1[2]
Acid-modified CNTspH 7, initial concentration = 1 ppm934.35[3]
Raw CNTspH 7, initial concentration = 1 ppm271.661[38]
Ethylenediamine-functionalized MWCNTspH 8, initial concentration = 5 mg/L-25.70[33]
SWCNTspH 8-1.97[39]
Nano-alumina/SWCNTspH 8-2.18[39]
SWCNTs-COOHpH 5, initial concentration = 20 mg/L69.9755.89[40]
CNT-10% Al2O3 membranepH 7, initial concentration = 1 mg/L, contact time = 2 h8454.42This study

Share and Cite

MDPI and ACS Style

Ihsanullah; Patel, F.; Khraisheh, M.; Atieh, M.A.; Laoui, T. Novel Aluminum Oxide-Impregnated Carbon Nanotube Membrane for the Removal of Cadmium from Aqueous Solution. Materials 2017, 10, 1144. https://doi.org/10.3390/ma10101144

AMA Style

Ihsanullah, Patel F, Khraisheh M, Atieh MA, Laoui T. Novel Aluminum Oxide-Impregnated Carbon Nanotube Membrane for the Removal of Cadmium from Aqueous Solution. Materials. 2017; 10(10):1144. https://doi.org/10.3390/ma10101144

Chicago/Turabian Style

Ihsanullah, Faheemuddin Patel, Majeda Khraisheh, Muataz Ali Atieh, and Tahar Laoui. 2017. "Novel Aluminum Oxide-Impregnated Carbon Nanotube Membrane for the Removal of Cadmium from Aqueous Solution" Materials 10, no. 10: 1144. https://doi.org/10.3390/ma10101144

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop