11
12 Adsorption of lisinopril and chlorpheniramine from aqueous solution on dehydrated and activated carbons El-Said I. El-Shafey , Haider A. J. Al-Lawati and Wafa S. H. Al-Saidi Chemistry Department, College of Science, Sultan Qaboos University, Muscat, P.O. Box 36, Sultanate of Oman Received 5 October 2015 Accepted 15 May 2016 Corresponding Author E-mail: [email protected] Tel: +968-99822317 Open Access pISSN: 1976-4251 eISSN: 2233-4998 Carbon Letters Vol. 19, 12-22 (2016) Original Articles Article Info Copyright © Korean Carbon Society http://carbonlett.org Abstract Date palm leaflets were used as a precursor to prepare dehydrated carbon (DC) via phos- phoric acid treatment at 150°C. DC, acidified with H 3 PO 4 , was converted to activated carbon (AC) at 500°C under a nitrogen atmosphere. DC shows very low surface area (6.1 m 2 /g) while AC possesses very high surface area (829 m 2 /g). The removal of lisinopril (LIS) and chlorpheniramine (CP) from an aqueous solution was tested at different pH, contact time, concentration, and temperature on both carbons. The optimal initial pH for LIS removal was 4.0 and 5.0 for DC and AC, respectively. However, for CP, initial pH 9.0 showed maxi- mum adsorption on both carbons. Adsorption kinetics showed faster removal on AC than DC with adsorption data closely following the pseudo second order kinetic model. Adsorption increases with temperature (25°C–45°C) and activation energy (E a ) is in a range of 19–25 kJ mol/L. Equilibrium studies show higher adsorption on AC than DC. Thermodynamic pa- rameters show that drug removal is endothermic and spontaneous with physical adsorption dominating the adsorption process. Column adsorption data show good fitting to the Thomas model. Despite its very low surface area, DC shows ~70% of AC drug adsorption capacity in addition of being inexpensive and easily prepared. Key words: adsorption, lisinopril, chlorpheniramine, dehydrated, activated carbon 1. Introduction Hospital wastewater is considered a serious source of pollution as it contains various pol- lutants such as pharmaceuticals and metabolites, chlorinated organic compounds, endocrine chemicals, radionuclides, heavy metals, and other chemicals. Hospitals generate significant amounts of wastewater in a range of 400–1200 L/day/bed [1]. Point sources of pharmaceu- ticals in the aquatic environment include hospitals, medical institutions, health care facili- ties, pharmaceutical manufacturers, and animal farms. The disposal of unused or expired pharmaceuticals through sinks or toilets is considered a major diffuse source of municipal wastewater pollution [2]. Adverse effects that result from the presence of pharmaceuticals in the environment include aquatic toxicity, biological imbalance, development of resistance in pathogenic bacteria, genotoxicity, and endocrine disruption [3]. Lisinopril (LIS) is an active angiotensin-converting enzyme inhibitor used for the treatment of hypertension, heart fail- ure, and acute myocardial infarction [4]. Chlorpheniramine (CP) is an anti-histamine drug, commonly used to treat allergies [5]. Adsorption is an efficient method to remove a wide range of pharmaceuticals [2,6]. In this paper, dehydrated and activated carbons were prepared from date palm leaflets, an ag- ricultural byproduct that is available in the Gulf States in large quantities (~3 million tons/ year and ~180,000 tons per year in Oman [2]), using phosphoric acid treatments. Both types of carbon were characterized and investigated for the removal of LIS and CP from aqueous solutions. DOI: http://dx.doi.org/ DOI:10.5714/CL.2016.19.012 This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/ by-nc/3.0/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited. Phenylethynyl-terminated polyimide, exfoliated graphite nanoplatelets, and the composites: an overview Donghwan Cho and Lawrence T. Drzal KCS Korean Carbon Society carbonlett.org pISSN: 1976-4251 eISSN: 2233-4998 REVIEWS VOL. 19 July 31 2016

Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

  • Upload
    hathu

  • View
    219

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

12

Adsorption of lisinopril and chlorpheniramine from aqueous solution on dehydrated and activated carbonsEl-Said I. El-Shafey♠, Haider A. J. Al-Lawati and Wafa S. H. Al-Saidi

Chemistry Department, College of Science, Sultan Qaboos University, Muscat, P.O. Box 36, Sultanate of Oman

Received 5 October 2015Accepted 15 May 2016

♠Corresponding AuthorE-mail: [email protected]: +968-99822317

Open Access

pISSN: 1976-4251 eISSN: 2233-4998

Carbon Letters Vol. 19, 12-22 (2016)Original Articles

Article Info

Copyright © Korean Carbon Society

http://carbonlett.org

AbstractDate palm leaflets were used as a precursor to prepare dehydrated carbon (DC) via phos-phoric acid treatment at 150°C. DC, acidified with H3PO4, was converted to activated carbon (AC) at 500°C under a nitrogen atmosphere. DC shows very low surface area (6.1 m2/g) while AC possesses very high surface area (829 m2/g). The removal of lisinopril (LIS) and chlorpheniramine (CP) from an aqueous solution was tested at different pH, contact time, concentration, and temperature on both carbons. The optimal initial pH for LIS removal was 4.0 and 5.0 for DC and AC, respectively. However, for CP, initial pH 9.0 showed maxi-mum adsorption on both carbons. Adsorption kinetics showed faster removal on AC than DC with adsorption data closely following the pseudo second order kinetic model. Adsorption increases with temperature (25°C–45°C) and activation energy (Ea) is in a range of 19–25 kJ mol/L. Equilibrium studies show higher adsorption on AC than DC. Thermodynamic pa-rameters show that drug removal is endothermic and spontaneous with physical adsorption dominating the adsorption process. Column adsorption data show good fitting to the Thomas model. Despite its very low surface area, DC shows ~70% of AC drug adsorption capacity in addition of being inexpensive and easily prepared.

Key words: adsorption, lisinopril, chlorpheniramine, dehydrated, activated carbon

1. Introduction

Hospital wastewater is considered a serious source of pollution as it contains various pol-lutants such as pharmaceuticals and metabolites, chlorinated organic compounds, endocrine chemicals, radionuclides, heavy metals, and other chemicals. Hospitals generate significant amounts of wastewater in a range of 400–1200 L/day/bed [1]. Point sources of pharmaceu-ticals in the aquatic environment include hospitals, medical institutions, health care facili-ties, pharmaceutical manufacturers, and animal farms. The disposal of unused or expired pharmaceuticals through sinks or toilets is considered a major diffuse source of municipal wastewater pollution [2]. Adverse effects that result from the presence of pharmaceuticals in the environment include aquatic toxicity, biological imbalance, development of resistance in pathogenic bacteria, genotoxicity, and endocrine disruption [3]. Lisinopril (LIS) is an active angiotensin-converting enzyme inhibitor used for the treatment of hypertension, heart fail-ure, and acute myocardial infarction [4]. Chlorpheniramine (CP) is an anti-histamine drug, commonly used to treat allergies [5].

Adsorption is an efficient method to remove a wide range of pharmaceuticals [2,6]. In this paper, dehydrated and activated carbons were prepared from date palm leaflets, an ag-ricultural byproduct that is available in the Gulf States in large quantities (~3 million tons/year and ~180,000 tons per year in Oman [2]), using phosphoric acid treatments. Both types of carbon were characterized and investigated for the removal of LIS and CP from aqueous solutions.

DOI: http://dx.doi.org/DOI:10.5714/CL.2016.19.012

This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License (http://creativecommons.org/licenses/by-nc/3.0/) which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Phenylethynyl-terminated polyimide, exfoliated graphite nanoplatelets, and the composites: an overviewDonghwan Cho and Lawrence T. Drzal

KCS Korean Carbon Society

carbonlett.org

pISSN: 1976-4251 eISSN: 2233-4998

REVIEWS

VOL. 19 July 31 2016

Page 2: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Adsorption of lisinopril and chlorpheniramine on activated carbons

13 http://carbonlett.org

charge (pHzpc) for both carbons was determined following the procedure of Moreno-Castilla et al. [7]. The base neutralization capacity for both carbons was determined using Boehm titra-tions [8]. Using standard methods, cation exchange capacity (CEC), apparent density, and ash content of both carbons were determined [9-11]. All experiments and analyses were carried out at least twice.

2.3. Adsorption experiments

Pharmaceutical stock solutions (500 mg/L each) were kept in a refrigerator and used within a week after preparation. Test and standards solutions were prepared by dilution in deionized wa-ter. Kinetic, equilibrium, and column sorption experiments were carried out at initial pH values, at which maximum sorption of drugs took place. For the kinetic experiments, 0.15 g of DC or AC was added to 50 mL (50 mg/L) of LIS at initial pH 4.0 and 5.0, respectively, and at initial pH 9.0 for CP on both carbons. At different time intervals, aliquots of supernatant were withdrawn for drug analysis. The adsorption process was followed for 70 h at 25°C, 35°C, and 45°C under continuous agitation (100 rpm/min). To investigate the effect of initial pH on drug adsorption, ~0.06 g carbon was mixed with 25 mL (100 mg/L) of drug so-lution in glass vials at different initial pH values (2.0–11.0). The initial pH was adjusted using drops of dilute HCl or NaOH prior to the addition of carbon. Under the same conditions of drug concentration and initial pH, reference drug samples were separated as control samples. Adsorption solutions were shak-en mechanically (100 rpm/min) at 25°C until the equilibrium was reached. Both the initial and the final drug concentrations were analyzed. Isotherm studies at different temperature (25°C–45°C) were carried out by mixing 0.06 g of carbon with 25 mL of drug solution (10–250) mg/L at respective initial pH values of maximum drug adsorption under mechanical shaking (100 rpm/min) until the equilibrium was reached. Residual drug concen-trations were analyzed.

2.4. Desorption studies

Drug solutions (250 mg/L, 100 mL) at initial pH of maxi-mum drug adsorption, were mixed with carbons (0.25 g). After equilibrium, the drug samples were filtered and residual drug was analyzed. Wet drug-loaded carbon was carefully transferred to another clean vial containing 100 mL of aqueous solution at initial pH 2 for drug desorption. After 24 h, samples were with-drawn for drug analysis.

2.5. Column studies

A fixed bed of the carbon adsorbent was prepared. A glass column with a glass wool layer at the bottom was used. The pre-weighed carbon samples were left in deionized water for 3 h for wetting. The column was filled with deionized water before adding the carbon mass and the column was gently tapped to facilitate uniform packing of the bed. For CP (pH 9, 51.5 mg/L), carbon samples (~1.50 g of DC or AC), after being wetted, were loaded in a column of 1 cm diameter, with length of 7.4 cm (bed volume, 5.8 cm3) for DC and length of 6.9 cm (bed volume, 5.43 cm3) for AC. The CP solution was allowed to pass through the

2. Experimental

All chemicals used were of analytical grade. Pure pharma-ceuticals (LIS and CP) were provided as powder samples by the National Pharmaceutical Industries Company (Muscat, Oman). Dry date palm leaflets (Phoenix dactylifera L.) were collected from a local farm in Al-Khodh, Muscat. The leaflets were thor-oughly washed with deionized water to remove dirt, dust, and other impurities and were allowed to dry in open air at room temperature to constant weight. The clean dry leaflets were cut into small pieces (1-cm length) before further use in carbon preparation.

2.1. Carbon preparation

Approximately 20 g of the clean dry leaflets was added to 200 mL distilled water, followed by the addition of 80 g of concen-trated phosphoric acid with stirring. The mixture was left over-night in an oven (Hobersal Mon X B2-125 furnace; Hobersal, Barcelona, Spain) at 150°C to chemically carbonize via dehy-dration. The produced dehydrated carbon (DC) was left to cool at room temperature, and an amount of DC was used to prepare AC as follows. DC with its residual phosphoric acid (without washing) was transferred to a quartz tube (internal diameter 5.1 cm and length 61 cm) to be in the heated zone of the tube furnace (GSL-1100X-110V; MTI Corp., Richmond, CA, USA). Under a nitrogen atmosphere, the temperature was raised from room temperature to 500°C in 90 min at a heating rate of ~5.6°C/min and was kept at 500°C for 1 h to produce AC. The produced AC was left to cool under a nitrogen atmosphere. Both carbons were washed thoroughly with hot deionized water to remove residual phosphoric acid followed by washing with ~1% NaOH solution for 1 h, to release humic substances from the carbons. Both car-bons were then washed with deionized water until neutrality. To retain the ion exchange groups on the surfaces of both carbons in their H-form, the carbons were washed again with ~1% HCl solution followed by washing with deionized water until neu-trality. The carbon samples were allowed to dry at 120°C till constant weight was achieved. After cooling in a desiccator and grinding, carbon with a size range between two sieves of 1.19 mm and 0.25 mm was selected for characterization and adsorp-tion experiments.

2.2. Physico-chemical characterization

The surface area of both carbons was determined using Auto-sorb-1 (Quantachrome Instruments, Boynton Beach, FL, USA) via nitrogen adsorption at 77 K. The carbon samples were tested using a JEOL/EO JSM 5600 scanning electron microscope (To-kyo, Japan) that was subjected to a 20 kV accelerating voltage. Energy dispersive X-ray (EDX) microanalysis of both carbons was carried out using a JEOL/O JSM 5600 editor energy dis-perse analysis system. X-ray powder diffraction was carried out using a Philips PW 1830 generator with a Philips PW 1050 powder goniometer (Philips, USA) and copper Kα was used as the incident radiation. An infrared analysis was carried out for DC and AC using a FT-IR spectrometer (Spectrun BX; Perki-nElmer, Germany) after drying at 120°C for 2 h. Zero point of

Page 3: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Carbon Letters Vol. 19, 12-22 (2016)

DOI: http://dx.doi.org/10.5714/CL.2016.19.012 14

stretching vibration of C=O, COO– or to skeletal C=C aromatic vibrations. Other bands in the range (1400–1000 cm−1) are as-signed to O−H bending and C−O stretching vibrations such as phenols and carboxylic acids [15].

DC possesses more CEC, carboxylic, lactonic, phenolic, and surface acidity (or less pHzpc) than AC (Table 1). EDX analysis shows more carbon and less oxygen content for AC than DC, indicating the presence of more carbon-oxygen groups on the DC surface (Table 1). DC and AC formation can be explained as follows. When the mixture of date palm leaflets and phosphoric acid is heated at 150°C, some hydrolysis to hemicelluloses takes place. As water evaporates, phosphoric acid concentrates, lead-ing to carbonization of the leaflets via dehydration of cellulose and hemicelluloses with partial oxidation and fragmentation of lignin. Concentrated phosphoric acid is capable of dehydrating plant material, but with a smaller oxidation effect than concen-trated sulfuric acid [16]. The produced DC with residual phos-phoric acid was subjected to pyrolysis at 500°C under a nitrogen atmosphere, during which volatile, tarry, and waxy compounds

column at a rate of 1.87 mL/min on DC remaining in contact with DC for 3.10 min and with AC for 2.90 min. For LIS (65 mg/L, pH 4 for DC; pH 5 for AC), carbon samples (1.8 g of DC or AC) were loaded in a column of 1 cm diameter, with length of 10.2 cm (bed volume, 8.01 cm3) for DC and length of 9.8 cm (bed volume, 7.69 cm3) for AC. The LIS solution was allowed to pass through the column at a rate of 1.87 mL/min staying in contact with DC for 4.28 min and AC for 4.11 min. Aliquots of the effluent solution were collected in a fraction collector (Frac-920; Sweden) and drug concentration was analyzed.

2.6. Drug analysis

Drugs were analyzed using a tris(2,2-bipyridyl)-ruthenium (II) peroxydisulphate chemiluminescence system in a two chip device following the procedure of Al Lawati et al. [12]. The system consists of serpentine and teardrop microfluidic chips, fluidic connect 4515, and fused silica capillaries (Micronit, En-schede, the Netherlands). The syringe pumps were purchased from Basi Bee (USA). The detector was a photomultiplier tube (PMT; H7155-2; Hamamatsu Photonics, Hamamatsu, Japan) connected to a PC via a Counting Unit (C8855; Hamamatsu Photonics). The experiments and analysis were carried out at least twice.

3. Results and Discussion

3.1. Carbon physico-chemical properties

The surface properties of the carbons are presented in Table 1. The AC surface area is almost 134 times larger using the Brunauer-Emmett-Teller method and 139 times larger using the α-S method, relative to that of DC. In the α-S method, the amount of nitrogen adsorbed as a function of P/Po is graphically compared with a normalized reference isotherm of a non-porous carbon [13]. This method enables the estimation of the micro-surface area and non-microsurface area. However, this method is influenced by the nature of the selected nonporous reference. The micro surface area, obtained using the α-S method, appears lower for both carbons (Table 1) than the non-micro surface area. In a previous study [6], for AC and DC prepared from date palm leaflets using sulfuric acid, surface area was higher for AC (405 m2/g) than DC (48 m2/g).

The apparent density and ash content are presented in Table 1. Scanning electron microscope photographs show that DC retains the fibrous structure of the leaflets while AC does not due to the pyrolysis at 500°C (Fig. 1a and b). X-ray diffraction patterns for both carbons (Fig. 2a and b) show an amorphous structure with a common peak at 2θ of 22° for amorphous silica [14]. For AC, the two small peaks at 2θ of 26° and 44° repre-sent the degree of graphitization [14]. Fig. 3 shows the infra-red spectra of both carbons. The broad bands at 3440 cm–1 for DC and 3400 cm–1 for AC correspond to hydrogen bonded OH stretching vibrations [15]. Stretching C−H vibrations in CH3 and CH2 groups appear in the DC spectrum at 2946 and 2876 cm–1, respectively. However, such bands are not available in the AC spectrum, likely because they were lost during pyroly-sis. The bands at 1648 cm–1 for both DC and AC correspond to

Table 1. Surface characterization of DC and AC

Property DC AC

pHzpc 3.18 4.17

CEC (meq100 g) 109 81

Surface functionality (meq/g)

Carboxyl 2.40 1.99

Lactone 1.26 0.81

Hydroxyl 1.30 0.49

EDX analysis (%)

Carbon 56.93 73.7

Oxygen 30.58 22.9

Phosphor 0.72 0.90

Surface properties

BET-method

Vm (cm3/g)* 1.41 190

Surface area (SBET) (m2/g) 6.12 829

BET-constant 120 154

α-S Method

Total surface area (m2/g) 6.04 838

Micro surface area (m2/g) 2.52 265

Non-micro surface area (m2/g) 3.52 573

Apparent density (g/cm) 0.27 0.24

Ash content (%) 12.4 14.2

DC, dehydrated carbon; AC, activated carbon; pHzpc, zero point of charge; CEC, cation exchange capacity; EDX, energy dispersive X-ray; BET, Brunauer-Emmett-Teller.*Vm is the monolayer capacity.

Page 4: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Adsorption of lisinopril and chlorpheniramine on activated carbons

15 http://carbonlett.org

are released. The release of tarry compounds from the gaps among crystallites by activation provides a porous structure with a high surface area of AC. Because phosphoric acid is an oxidiz-ing agent, carbon oxygen functional groups such as carboxyls, lactones, and phenols are usually formed on the AC surface. The very low surface area of DC compared to that of AC is likely related to the higher content of carbon-oxygen hydrophilic func-tional groups on its surface. Such hydrophilic functional groups occupy a large fraction of the DC surface, restricting the adsorp-tion of the non-polar nitrogen molecules. In addition, the pres-ence of lignin material within the DC structure can also block the available pores and limit the access of nitrogen gas on the DC surface, thus decreasing the surface area.

3.2. pH effect on drug adsorption

pHzpc is the pH at which the electrical charge density on the carbon surface is zero. For the carbons, at pH values below the pHzpc (3.18 for DC and 4.17 for AC), the surface functional groups become protonated in their non-dissociated form; how-ever, beyond that value, the carbon becomes negatively charged. The pKa values of LIS at 25°C are 2.5 (central –COOH), 4.0 (Prolyl COOH), 6.7 (secondary amine group), and 10.1 (lysyl primary amine group) [17], as presented in Fig. 4a. The central COOH is more acidic than the prolyl COOH due to the proxim-ity of the secondary amine group. The secondary amine group is more acidic than the lysyl primary amine due to the proximity of the electron-withdrawing amide group [17].

The effect of pH on adsorption of LIS is presented in Fig. 4b. At pH lower than the pKa1 value, i.e., <2.5, both amine groups remain protonated and become positively charged; however, the carboxylic groups on the drug and carbons become protonated and neutral (COOH), showing low adsorption and weak interac-tion with the drug cations (Fig. 4b). At pH between 2.5 and 4.0, the central carboxylic group becomes deprotonated, carrying a negative charge; however, the other groups remain protonated; pyrolyl carboxylic (neutral) and the amine groups (positively charged). DC retains its protonated form below pH 3.18; how-ever, beyond pH 3.18, the surface becomes negatively charged. Thus, electrostatic interaction between the negatively charged DC surface and positively charged drug molecules takes place, with maximal adsorption at pH 4.0. However, in this pH range,

Fig. 1. Scanning electron microscope photographs of (a) dehydrated carbon and (b) activated carbon.

Fig. 2. X-ray powder diffraction of (a) dehydrated carbon and (b) acti-vated carbon.

Fig. 3. Fourier transform infrared spectroscopy spectra of dehydrated carbon (DC) and activated carbon (AC).

Page 5: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Carbon Letters Vol. 19, 12-22 (2016)

DOI: http://dx.doi.org/10.5714/CL.2016.19.012 16

Fig. 5b. For DC, at initial pH <3.18, adsorption of CP is low on DC due to the weak interaction between the neutral carbon sur-face and the double positively charged CP ions. In the pH range of 3.2–4.0, a sharp rise in CP adsorption occurs due to the elec-trostatic attraction between the double positively charged CP cations and the negatively charged surface of DC. For AC, at pH below 4.17, the pHzpc of AC, weak interaction occurs between the positively charged drug ions (double positively charged be-low pH 4.0, mono-positively charged beyond pH 4.0) and the neutral carbon surface, showing less CP adsorption. In the pH range of 4.0–9.2 for DC and 4.17–9.2 for AC, there is a gradual slight rise in CP sorption due to the attractive forces between the mono-positively charged CP cations and the negatively charged surface, showing maximum CP adsorption at initial pH 9.0. At pH >9.2, adsorption of CP on both carbons decreases as a result of the weak interaction between the negatively charged surface and neutral CP molecules. However, at pH values >10.1 for LIS and >9.2 for CP, due to the presence of other adsorption forces such as hydrogen bonding and van der Waals forces, the carbons still show significant extents of drug adsorption at such high pH values. AC, in general, shows higher drug adsorption than DC and this is related to its high surface area. However, DC, even with its very small surface area, still shows a good interaction with the drug. This reflects the effectiveness of other adsorption forces such as ion exchange and H-bonding. Furthermore, the presence of lignin material within the pores of DC likely en-hances the uptake of drugs via hydrophobic interactions.

3.3. Kinetics of drug adsorption

From the studies of pH effect, initial pH 4.0 and 5.0 were selected as the optimum pH for LIS adsorption on DC and AC, respectively, while for CP, pH 9.0 was selected for both carbons for adsorption experiments. Equilibrium is reached faster for AC

the AC surface remains protonated and adsorption slightly in-creases as protons decrease. At pH between 4.0 and 6.7, both carboxylic groups become negatively charged; however, the amine groups remain positively charged on the drug molecule. DC remains negatively charged, whereas AC becomes nega-tively charged at pH higher than its pHzpc 4.18. Electrostatic in-teraction between positively charged drug cations and the nega-tively charged surface leads to maximum adsorption of LIS on AC at pH 5.0. Repulsion between the negative charges on the drug ions and the negatively charged carbon surfaces can also take place, and adsorption thus starts to decrease as the initial pH further increases beyond pH 5. At pH between 6.7 and 10.1 for LIS molecules, both carboxylic groups become negatively charged and the secondary amine group becomes deprotonated but the primary amine group remains positively charged. Elec-trostatic attraction exists between the negatively charged carbon surface and the positively charged primary amine group on the drug ions; however, the degree of repulsion appears to increase between the negatively charged carboxylic groups on the drug ions and the negatively charged carbon surface. At pH higher than 10.1, both carboxylic groups become negatively charged; however, the primary amine group becomes deprotonated. Due to the repulsion between the negatively charged drug ions and the negatively charged surface, adsorption further decreases.

CP molecule (Fig. 5a) possesses two basic groups with pKa1 of 4.0 for the pyridyl nitrogen and 9.2 for the tertiary amine group [18]. At pH below 4.0 (pKa1 value for CP), CP is protonated on both nitrogen atoms. In the pH range between that of pKa1 and pKa2 (pH 4.0–9.2), CP remains monoprotonated on its aliphatic nitrogen; however, at pH beyond pKa2, the drug molecule be-comes neutral. The effect of pH on CP adsorption is presented in

Fig. 4. Molecular structure (a) and effect of initial pH (b) on lisinopril (LIS) adsorption (initial concentration, 100 mg/L; volume of LIS solution, 25 mL; shaking speed, 100 rpm). DC, dehydrated carbon, AC activated carbon.

Fig. 5. Molecular structure (a) and effect of initial pH (b) on chlorphe-niramine (CP) adsorption (initial concentration, 100 mg/L; volume of LIS solution, 25 mL; shaking speed, 100 rpm). DC, dehydrated carbon, AC activated carbon.

Page 6: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Adsorption of lisinopril and chlorpheniramine on activated carbons

17 http://carbonlett.org

than DC for both drugs (Fig. 6). The adsorption of drugs varies almost linearly with the half power of time, in the adsorption early stages (eq 1) [19]:

qt = kd t0.5 (1)

where qt is the amount of drug sorbed per gram of carbon (mg/g) and kd is the diffusion rate constant. As the temperature increas-es, kd increases (Table 2) for both drugs on both carbons. The kinetic adsorption data were examined for the pseudo second order kinetic model (eq 2).

t/qt =1/k qe2 + t/qe (2)

where k is a rate constant of the pseudo second order model and qe is the amount of drug adsorbed per unit mass of sorbent (mg/g) at equilibrium. The initial adsorption rate is given by h = k2qe

2. The linear plots of t/qt versus t for the pseudo second order model present a good fitting (high R2 values), as seen in Table 2, indicating that the adsorption of drugs complies well with a pseudo second order kinetic reaction. This indicates that the rate-limiting step in the adsorption process involves both the carbon surface and drug via sharing or exchange of electrons between the carbon surface and the drug [20].

As presented in Table 2, the kinetic constants kd, k, and h show higher values for AC than DC for both drugs. This is per-haps related to the high surface area of AC. The values of kd, k, qe, and h rise with temperature and this may be because of the desolvation of the adsorbing species and the decrease in the thickness of the boundary layer surrounding the carbon as tem-perature increases [21].

For LIS adsorption, raising the temperature from 25°C to 45°C led to an increase in the values of kd, k, and h, by 1.89-,

Fig. 6. Adsorption kinetics of (a) LIS and (b) CP on DC and AC at dif-ferent temperature. LIS, lisinopril; CP, chlorpheniramine; DC, dehydrated carbon, AC activated carbon.

Table 2. Pore diffusion and rate constants for the kinetics of LIS and CP sorption on DC and AC at different temperature

Drug Sorbent Temperature (°C)

Pore diffusion constant, kd

(mg/g/ hr0.5)

Pseudo second order model

Rate constant k(g/mg/hr)

Initial adsorption rate, h(mg/g /hr)

Monolayer capacity, qe (mg/g) R2

Lis DC 25 5.83 0.00513 5.038 31.35 0.9991

35 9.81 0.00717 9.149 35.71 0.9996

45 11.01 0.00975 15.974 40.49 0.9998

AC 25 10.88 0.0158 23.585 38.61 0.9999

35 12.88 0.0213 35.46 40.82 0.9999

45 13.44 0.0284 51.02 42.37 0.9999

CP DC 25 8.54 0.0131 9.569 27.02 0.9996

35 10.02 0.0167 18.28 33.11 0.9993

45 11.77 0.0221 28.57 35.97 0.9998

AC 25 11.21 0.0166 21.60 36.10 0.9996

35 13.04 0.0212 33.33 39.53 0.9999

45 14.43 0.0271 49.51 42.02 0.9999

LIS, Lisinopril; CP, chlorpheniramine; DC, dehydrated carbon; AC, activated carbon.

Page 7: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Carbon Letters Vol. 19, 12-22 (2016)

DOI: http://dx.doi.org/10.5714/CL.2016.19.012 18

1.9-, and 3.17-fold, respectively, for DC and by 1.23-, 1.80-, and 2.16-fold, respectively, for AC. However, raising the tempera-ture from 25°C to 45°C, for CP adsorption, led to a rise in the values of kd, k, and h by 1.38-, 1.69-, and 2.98-fold, respectively, for DC and by 1.29-, 1.64-, and 2.29-fold, respectively, for AC. The relative increase in rate constants at 45°C is higher for DC than for AC and this might be related to an expected swelling of DC in water. El-Shafey et al. [2] reported that the increase of ciprofloxacin adsorption on DC with temperature was related to an obvious swelling of the carbon. However, AC is a rigid sub-stance and does not show swelling in water [22].

The activation energies (Ea, kJ/mol) for LIS and CP adsorp-tion on carbons were calculated using the values of the rate con-stant k using the Arrhenius equation (eq 3).

k2 = Ae-Ea/RT (3)

where R is the gas constant (8.314 J/mol/K), T is the temperature in Kelvin (K), and A is the pre-exponential factor. A plot of ln k2

versus 1/T gives a straight line from which Ea was calculated. Low Ea values (5–40 kJ/mol) correspond to physical adsorp-tion, while higher Ea values (40–800 kJ/mol) are related to che-misorption [23]. The activation energies of drug adsorption on DC and AC, in this study, are 25.3 and 22.5 kJ/mol for LIS, re-spectively, and 20.2 and 19.4 kJ/mol for CP, respectively. These Ea values indicate that physical adsorption forces dominate the removal of both drugs on both carbons.

3.4. Equilibrium studies

The adsorption data of LIS and CP on DC and AC presented in Fig. 7 follow an L-type isotherm, with increased drug uptake as temperature increases (25°C–45°C). The equilibrium data were tested using the Langmuir and Freundlich models (eqs 4

Fig. 7. Adsorption isotherms of (a) LIS and (b) CP on DC and AC at dif-ferent temperature. LIS, lisinopril; CP, chlorpheniramine; DC, dehydrated carbon, AC activated carbon.

Table 3. Langmuir and Freundlich parameters for the sorption of LIS and CP at different temperatures

Drug Sorbent Temperature (°C)

Langmuir constantsCorrelation value, R2

Freundlich constants Correlation value, R2

q (mg/g) b (L/mg) 1/n K

LIS DC 25 64.10 0.0559 0.9988 0.3969 9.25 0.9706

35 72.57 0.0771 0.9994 0.3729 12.59 0.9616

45 84.75 0.0997 0.9993 0.3312 18.67 0.9716

AC 25 90.91 0.2450 0.9996 0.3708 21.32 0.9102

35 99.01 0.3961 0.9996 0.2717 34.50 0.9322

45 108.7 0.6970 0.9989 0.2388 46.23 0.9430

CP DC 25 69.44 0.0862 0.9976 0.384 11.95 0.9589

35 83.33 0.1364 0.9998 0.357 18.32 0.9285

45 98.04 0.1885 0.9951 0.311 28.05 0.9589

AC 25 92.59 0.189 0.9987 0.438 18.95 0.9471

35 107.5 0.225 0.9995 0.399 26.69 0.9485

45 117.6 0.301 0.9995 0.377 35.84 0.9661

LIS, Lisinopril; CP, chlorpheniramine; DC, dehydrated carbon; AC, activated carbon.

Page 8: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Adsorption of lisinopril and chlorpheniramine on activated carbons

19 http://carbonlett.org

and 75%–84%, respectively, of AC adsorption monolayer ca-pacity for both drugs. This reflects the effectiveness of other ad-sorption forces such as ion exchange and H-bonding besides Van der Waals forces, which likely dominate drug adsorption on DC. The adsorption monolayer capacity of CP on AC and DC in this study (Table 3) is comparable to previous studies on rectorite (122 mg/g) [25] and montmorillonite (190 mg/g) [5].

3.5. Thermodynamic parameters

∆Gº, ∆Hº, and ∆Sº were calculated from the equilibrium constant values, Kc, at different temperatures. Kc is presented in eq 6:

Kc = CAe/Ce (6)

where CAe is the amount of LIS adsorbed on carbon (mg) per L of solution and Ce is the equilibrium concentration of the drug solution (mg/L). Kc is estimated from the initial part of the ad-sorption isotherm in which qe versus Ce is linear. The calculated thermodynamic parameters for drug adsorption are shown in Table 4. Kc values increase with temperature, indicating an en-dothermic process of drug adsorption [3]. The Gibbs free energy change of the adsorption process, ΔGº, is related to Kc, as given in eq 7. The enthalpy (ΔHº) and the entropy (ΔSº) are calculated by plotting lnKc versus 1/T according to Van’s Hoff equation (eq 8).

∆Gº = -RT ln Kc (7)

ln Kc = (∆S°/R) +( ∆H°/RT) (8)

Plotting ln Kc versus 1/T gives a linear relationship, from which ΔHº and ΔSº were determined. The negative values of ∆Gº (Table 4) indicate a favorable and spontaneous process for

and 5, respectively).

Ce/qe = 1/b.q + Ce/q (4)

log qe = 1/n (log Ce) + log K (5)

where Ce is the equilibrium drug concentration (mg/L), b is Langmuir constant (L/mg), q is the adsorption monolayer ca-pacity from Langmuir model (mg/g), and 1/n and K (L1/n mg1-1/n /g) are Freundlich constants related to adsorption intensity and adsorption capacity, respectively.

As presented in Table 3, the adsorption equilibrium data show better fitting for the Langmuir model than the Freundlich model for both drugs. The basic assumption of the Langmuir adsorp-tion isotherm is monolayer coverage of the adsorbate onto the adsorbent surface active sites at equilibrium [24]. The increase in drug monolayer adsorption by raising the temperature is ob-vious and may be related to the desolvation of the adsorbing species and the decrease in the thickness of the boundary layer surrounding the carbon with increasing temperature [21]. Rais-ing the temperature from 25°C to 45°C led to an increase in LIS monolayer adsorption by 1.32-fold on DC and 1.20-fold on AC, and in CP monolayer adsorption by 1.41-fold on DC and 1.27-fold on AC.

The slight increase in drug adsorption on DC compared with AC by raising the temperature is likely due to the swelling of DC with greater temperature, leading to the development of wider pores and giving more access to adsorption sites. Similar results were obtained for the adsorption of ciprofloxacin antibiotic on DC prepared via sulfuric acid dehydration [2]. On the other hand, AC is a rigid material and its swelling in water or other solvents is unexpected [22]. CP shows higher uptake on both carbons than LIS and this is related to the smaller molecular size of the former. Even with its very low surface area, DC still shows effective LIS and CP adsorption in a range of 70%–78%

Table 4. Thermodynamic parameters of LIS and CP adsorption on DC and AC

Drug Carbon Temperature (K) Kc ΔGo (kJ/mol) ΔHo (kJ/mol) ΔSo (J/mol)

LIS DC 298 8.37 –5.264 29.04 115.2

308 12.70 –6.509

318 17.48 –7.564

AC 298 53.75 –9.872 28.14 127.4

308 74.57 –11.041

318 109.90 –12.425

CP DC 298 16.30 –6.915 18.51 85.38

308 21.35 –7.839

318 26.06 –8.620

AC 298 50.78 –9.730 17.08 92.09

308 63.32 –10.62

318 79.64 –11.57

LIS, Lisinopril; CP, chlorpheniramine; DC, dehydrated carbon; AC, activated carbon.

Page 9: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Carbon Letters Vol. 19, 12-22 (2016)

DOI: http://dx.doi.org/10.5714/CL.2016.19.012 20

the temperature range evaluated, as is usually the case for many adsorption systems in solution [25,26]. The positive values of ∆Hº show an endothermic nature for CP and LIS adsorption. The values of ∆Hº for the adsorption of both drugs are <40 kJ/mol, indicating that physical adsorption processes are dominant in drug removal [27]. The positive ΔSº values show an increase in randomness at the solid-solution interface during the adsorption process of the drug on both carbons [6].

3.6. Drug desorption

Desorption from LIS loaded carbons is lower than that of CP, reaching ~63.3% from loaded DC and ~40.2% from loaded AC. On the other hand, desorption of CP from loaded DC shows ~81.8% recovery; however, for desorption from loaded AC, drug recovery is only ~61.6%. This indicates a stronger affinity of the carbon surface to LIS molecules than CP molecules. LIS molecule possesses multifunctional groups on its surface where-as CP molecule possesses a tertiary amine group and a pyridyl group (Fig. 5). It is clear that desorption from loaded DC shows better performance than that from AC. This is mainly related to the different dominating forces of adsorption onto both carbons. AC possesses a high extent of physical adsorption forces (van der Waals forces) due its higher surface area but smaller content of surface functional groups. DC possesses much lower surface area, indicating less van der Waals forces. However, DC pos-sesses high content of surface functional groups such as COOH, OH, C=O, and C–O–C that act as active sites for drug adsorption involving other adsorption forces such as cation exchange and H-bonding. Thus, at pH 2, more release of LIS and CP mol-ecules takes place from DC than from AC. In a previous study, desorption of ciprofloxacin from loaded DC was ~83% [2]. A previous study [28] concluded that desorption of organic adsor-bates including pharmaceuticals from loaded ACs is limited due to the high affinity of such compounds to the AC surface.

3.7. Dynamic adsorption

The initial concentration of CP used was ~50.0 mg/L, whereas that of LIS was 65 mg/L. Dynamic adsorption data are presented in Fig. 8a and b. LIS adsorption shows breakthrough curves at ~180 bed volumes with adsorption capacity (qexp) of 68 mg/g for DC and ~270 bed volumes with adsorption capacity of 91 mg/g for AC. CP shows a breakthrough at ~220 bed volumes on DC with adsorption capacity of 72 mg/g; however, for AC, the breakthrough appears at ~290 bed volumes with adsorption capacity of 89 mg/g. The dynamic adsorption capacity of both drugs correlates well with the monolayer capacities calculated from the Langmuir equations for both drugs on both carbon ad-sorbents (Table 3).

The Thomas model is frequently used to estimate the ad-sorption capacity and predict breakthrough curves. The break-through column data were processed using the Thomas Model [29]. The model assumes a negligible axial and radial dispersion in the fixed bed column with the adsorption data following pseu-do second-order reaction kinetics and the Langmuir isotherm at equilibrium [30]. The model is useful in estimating the adsorp-tion process, in which external and internal diffusion resistances are negligible [31]. The Thomas model can be expressed as fol-

Fig. 8. Column data for (a) LIS and (b) CP adsorption on DC and AC. LIS, lisinopril; CP, chlorpheniramine; DC, dehydrated carbon, AC activated carbon.

Fig. 9. The Thomas model application for the column sorption of (a) LIS and (b) CP on DC and AC. LIS, lisinopril; CP, chlorpheniramine; DC, dehy-drated carbon, AC activated carbon.

Page 10: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Adsorption of lisinopril and chlorpheniramine on activated carbons

21 http://carbonlett.org

References

[1] Gupta P, Mathur N, Bhatnagar P, Nagar P, Srivastava S. Genotoxic-ity evaluation of hospital wastewaters. Ecotoxicol Environ Saf, 72, 1925 (2009). http://dx.doi.org/10.1016/j.ecoenv.2009.05.012.

[2] El-Shafey EI, Al-Lawati H, Al-Sumri AS. Ciprofloxacin adsorp-tion from aqueous solution onto chemically prepared carbon from date palm leaflets. J Environ Sci, 24, 1579 (2012). http://dx.doi.org/10.1016/s1001-0742(11)60949-2.

[3] Sim WJ, Lee JW, Lee ES, Shin SK, Hwang SR, Oh JE. Occurrence and distribution of pharmaceuticals in wastewater from house-holds, livestock farms, hospitals and pharmaceutical manufactures. Chemosphere, 82, 179 (2011). http://dx.doi.org/10.1016/j.chemo-sphere.2010.10.026.

[4] Parfitt K, Martindale M. The Complete Drug Reference, 32nd ed., Pharmaceutical Press, 898 (1999).

[5] Li Z, Chang PH, Jean JS, Jiang WT, Hong H. Mechanism of chlorpheniramine adsorption on Ca-montmorillonite. Colloids Surf A: Physicochem Eng Aspects, 385, 213 (2011). http://dx.doi.org/10.1016/j.colsurfa.2011.06.013.

[6] El-Shafey EI, Al-Lawati HAJ, Al-Hussaini AY. Adsorption of fexofenadine and diphenhydramine on dehydrated and activated carbons from date palm leaflets. Chem Ecol, 30, 765 (2014). http://dx.doi.org/10.1080/02757540.2014.894986.

[7] Moreno-Castilla C, López-Ramón MV, Carrasco-Marı́n F. Changes in surface chemistry of activated carbons by wet oxidation. Carbon, 38, 1995 (2000). http://dx.doi.org/10.1016/s0008-6223(00)00048-8.

[8] Boehm HP. Chemical identification of surface groups. Adv Catal, 16, 179 (1966). http://dx.doi.org/10.1016/S0360-0564(08)60354-5.

[9] Thorpe VA. Collaborative study of the cation exchange capacity of peat materials. J Assoc Off Anal Chem, 56, 154 (1973).

[10] American Society for Testing and Materials. Standard Test Method for Apparent Density of Activated Carbon, ASTM, West Con-shohocken, PA, D2854 (1996).

[11] American Society for Testing and Materials. Standard Test Method for Total Ash Content of Activated Carbon, ASTM, West Con-shohocken, PA, D2866 (1996).

[12] Al Lawati HAJ, Al-Azwani M,Varma GB, Suliman FEO, Al Kin-dy SMZ. Towards an ideal method for analysis of lisinopril in pharmaceutical formulations using a tris(2,2′-bipyridyl)-rutheni-um(II)-peroxydisulfate chemiluminescence system in a two chip device. Anal Methods, 4, 773 (2012). http://dx.doi.org/10.1039/c2ay05616j.

[13] Sellés-Pérez MJ, Martín-Martínez JM. Application of α and n plots to N2 adsorption isotherms of activated carbons. J Chem

lows (eq 9):

ln(Co/Ct −1) = kThqom/F − kThCot (9)

where kTh (L/min/mg) is the Thomas rate constant; qo (mg/g) is the equilibrium drug uptake per gram of carbon; Co and Ct (mg/L) are the drug concentrations in the influent and the efflu-ent at time t, respectively; and m (g) is the mass of adsorbent and F is (L/min) the flow rate. A linear plot of ln[(Co/Ct)−1] against time t is presented in Fig. 9, from which qo and kTh are obtained. The parameters from the linear regression analysis are presented in Table 5.

qo values calculated from the Thomas model are very close to the amount determined experimentally for drug adsorption in the column system qexp. In addition, R2 values range from 0.9959 to 0.9843, indicating good fitting to the Thomas model.

4. Conclusions

AC prepared in this study shows high surface area and low content of surface functional groups, in contrast with DC, which possesses much lower surface area with high content of carbon-oxygen surface groups. Despite its very low surface area, DC shows effective drug removal from aqueous solutions, reaching ~70% of the removal capacity of AC. This reflects that the high-ly loaded carbon-oxygen groups on DC serve as active sites for LIS adsorption via ion exchange and hydrogen bonding. Drug adsorption follows the pseudo second order kinetic model and Langmuir at equilibrium. In addition, dynamic adsorption of drugs follows the Thomas model well. DC and AC from date palm leaflets are effective in drug removal from aqueous solu-tions.

Conflict of Interest

No potential conflict of interest relevant to this article was reported.

Acknowledgements

The authors would like to thank the National Pharmaceuti-cal Industries Company (Muscat, Oman) for supplying lisinopril and chlorpheniramine samples that enabled this research work to be carried out.

Table 5. Thomas model parameters at different conditions for the adsorption of LIS and CP on DC and AC using linear regression analysis

Drug Carbon Co (mg/L) ν (L/min) kTh (L/min/mg) 10–4 qo (mg g–1) R2 qexp (mg/g)

LIS DC 48.2 1.87 10–3 0.16 10–3 73.5 0.9908 71.6

AC 51.5 1.87 10–3 0.18 10–3 90.4 0.9843 91.1

CP DC 65 1.87 10–3 0.206 10–3 67.3 0.9877 68.0

AC 65 1.87 10–3 0.218 10–3 89.3 0.9959 91.0

LIS, Lisinopril; CP, chlorpheniramine; DC, dehydrated carbon; AC, activated carbon.

Page 11: Adsorption of lisinopril and chlorpheniramine from aqueous …carbonlett.org/Upload/files/CARBONLETT/[012-022]-02.pdf · 12 Adsorption of lisinopril and chlorpheniramine from aqueous

Carbon Letters Vol. 19, 12-22 (2016)

DOI: http://dx.doi.org/10.5714/CL.2016.19.012 22

(2001). http://dx.doi.org/10.1016/s0008-6223(00)00268-2.[23] Nollet H, Roels M, Lutgen P, Van der Meeren P, Verstraete W. Re-

moval of PCBs from wastewater using fly ash. Chemosphere, 53, 655 (2003). http://dx.doi.org/10.1016/s0045-6535(03)00517-4.

[24] Yasui-Furukori N, Uno T, Sugawara K, Tateishi T. Different ef-fects of three transporting inhibitors, verapamil, cimetidine, and probenecid, on fexofenadine pharmacokinetics. Clin Pharmacol Ther, 77, 17 (2005). http://dx.doi.org/10.1016/j.clpt.2004.08.026.

[25] Wang X, Guocheng L, Wang Q, Zhu Z, Li Z, Mao F, Wu L. Ad-sorption of chlorpheniramine from water by rectorite. Chem Ind Eng Prog, 4, 938 (2012).

[26] Wu Z, Joo H, Lee K. Kinetics and thermodynamics of the organic dye adsorption on the mesoporous hybrid xerogel. Chem Eng J, 112, 227 (2005). http://dx.doi.org/10.1016/j.cej.2005.07.011.

[27] Chern JM, Wu CY. Desorption of dye from activated carbon beds: effects of temperature, pH, and alcohol. Water Res, 35, 4159 (2001). http://dx.doi.org/10.1016/s0043-1354(01)00127-0.

[28] Rege SU, Yang RT, Cain CA. Desorption by ultrasound: phenol on activated carbon and polymeric resin. AIChE J, 44, 1519 (1998). http://dx.doi.org/10.1002/aic.690440706.

[29] Al-Ghouti MA, Khraisheh MAM, Ahmad MN, Allen SJ. Micro-column studies of dye adsorption onto manganese oxides modi-fied diatomite. J Hazard Mater, 146, 316 (2007). http://dx.doi.org/10.1016/j.jhazmat.2006.12.024.

[30] Han R, Ding D, Xu Y, Zou W, Wang Y, Li Y, Zou L. Use of rice husk for the adsorption of congo red from aqueous solution in column mode. Bioresour Technol, 99, 2938 (2008). http://dx.doi.org/10.1016/j.biortech.2007.06.027.

[31] Aksu Z, Gönen F. Biosorption of phenol by immobilized acti-vated sludge in a continuous packed bed: prediction of break-through curves. Process Biochem, 39, 599 (2004). http://dx.doi.org/10.1016/s0032-9592(03)00132-8.

Soc Faraday Trans, 87, 1237 (1991). http://dx.doi.org/10.1039/ft9918701237.

[14] Lua AC, Yang T. Effect of activation temperature on the textural and chemical properties of potassium hydroxide activated carbon prepared from pistachio-nut shell. J Colloid Interface Sci, 274, 594 (2004). http://dx.doi.org/10.1016/j.jcis.2003.10.001.

[15] Gómez-Serrano V, Acedo-Ramos M, López-Peinado AJ, Valenzu-ela-Calahorro C. Oxidation of activated carbon by hydrogen per-oxide: study of surface functional groups by FT-i.r. Fuel, 73, 387 (1994). http://dx.doi.org/10.1016/0016-2361(94)90092-2.

[16] Roberts JD, Caserio MC. Basic Principles of Organic Chemistry, 2nd ed., W.A. Benjamin, Inc., Menlo Park, CA, 599 (1977).

[17] Wang, SL, Chuang CH, Lin SY. pH-dependent coordination of metal-lisinopril complex investigated by attenuated total reflec-tion/Fourier transform infrared spectroscopy. Chem Pharm Bull, 50, 78 (2002). http://dx.doi.org/10.1248/cpb.50.78.

[18] Moreno-Villoslada I, González F, Rivas BL, Shibue T, Nishide H. Tuning the pKa of the antihistaminic drug chlorpheniramine maleate by supramolecular interactions with water-soluble poly-mers. Polymer, 48, 799 (2007). http://dx.doi.org/10.1016/j.poly-mer.2006.12.015.

[19] Weber WJ, Morris JC. Kinetics of adsorption on carbon from solu-tion. J Sanit Eng Div, 89, 31 (1963).

[20] HoYS. Review of second-order models for adsorption systems. J Hazard Mater, 136, 681 (2006). http://dx.doi.org/10.1016/j.jhazmat.2005.12.043.

[20] Meena AK, Kadirvelu K, Mishra GK, Rajagopal C, Nagar PN. Ad-sorption of Pb(II) and Cd(II) metal ions from aqueous solutions by mustard husk. J Hazard Mater, 150, 619 (2008). http://dx.doi.org/10.1016/j.jhazmat.2007.05.011.

[22] Moreno-Castilla C, Carrasco-Marı́n F, López-Ramón MV, Al-varez-Merino MA. Chemical and physical activation of olive-mill waste water to produce activated carbons. Carbon, 39, 1415