You are on page 1of 8

Available online at www.sciencedirect.

com
JOURNAL OF
ENVIRONMENTAL
SCIENCES
ISSN 1001-0742
CN 11-2629/X
Journal of Environmental Sciences 2012, 24(9) 1579–1586
www.jesc.ac.cn

Ciprofloxacin adsorption from aqueous solution onto chemically prepared


carbon from date palm leaflets
El-Said Ibrahim El-Shafey ∗, Haider Al-Lawati, Asmaa Soliman Al-Sumri

Department of Chemistry, College of Science, Sultan Qaboos University, P. O. Box 36-Al-Khodh 123, Muscat, Oman

Received 13 October 2011; revised 14 December 2011; accepted 16 December 2011

Abstract
A chemically prepared carbon was synthesized from date palm leaflets via sulphuric acid carbonization at 160°C. Adsorption of
ciprofloxacin (CIP) from aqueous solution was investigated in terms of time, pH, concentration, temperature and adsorbent status
(wet and dry). The equilibrium time was found to be 48 hr. The adsorption rate was enhanced by raising the temperature for both
adsorbents, with adsorption data fitting a pseudo second-order model well. The activation energy, E a , was found to be 17 kJ/mol,
indicating a diffusion-controlled, physical adsorption process. The maximum adsorption was found at initial pH 6. The wet adsorbent
showed faster removal with higher uptake than the dry adsorbent, with increased performance as temperature increased (25–45°C). The
equilibrium data were found to fit the Langmuir model better than the Freundlich model. The thermodynamic parameters showed that
the adsorption process is spontaneous and endothermic. The adsorption mechanism is mainly related to cation exchange and hydrogen
bonding.

Key words: ciprofloxacin; carbon; adsorption; palm leaflets


DOI: 10.1016/S1001-0742(11)60949-2

Introduction (Bhandari et al., 2008; Golet et al., 2002). In addition,


because it is toxic to some microorganisms, CIP can
Ciprofloxacin (CIP) is an antibiotic that is globally used for cause disturbances to the biological balance in the aquatic
the treatment of several bacterial infections in humans and ecosystem. In the literature, several studies have been
animals. Significant sources of wastewater drug contami- conducted for the removal of CIP from water including
nation include drug manufacturers and hospitals (Bhandari photodegradation (Belden et al., 2007), photo-Fenton oxi-
et al., 2008). In addition, the incomplete metabolization dation (Sun et al., 2009), ozonation (De Witte et al., 2010),
of the drug in humans and improper disposal of unused and adsorption on activated charcoal and talc (Ibezim et
or expired drugs through the sink contribute to wastew- al., 1999), dissolved organic carbon (Carmosini and Lee,
ater CIP contamination. In wastewater treatment plants, 2009), montmorillonite (Wang et al., 2010, 2011), surface-
CIP and other antibiotics are only partially eliminated modified carbon materials (Carabineiro et al., 2011a) and
and residual amounts can reach the aquatic ecosystem. goethite (Zhang and Huang, 2007).
CIP has been found in wastewater and surface water at Date palm (Phoenix dactylifera L.) is an important com-
concentrations typically < 1 μg/L (Bhandari et al., 2008; mercial crop in the Middle East, particularly in Oman and
Carmosini and Lee, 2009). Higher concentrations were the Gulf states (Al-Yahyai and Al-Khanjari, 2008) where
reported in effluents from hospitals (up to 150 μg/L) and 180,000 and 3 million tons of palm leaflets are produced
drug production facilities (31 mg/L) (Hartmann et al., annually, respectively. Burning in the field is a common
1999; Martins et al., 2008). CIP can be adsorbed by sewage practice that poses environmental problems. The objective
sludge with concentrations up to 2.42 mg/kg (Golet et al., of this research was to prepare a carbonaceous adsorbent
2002). It can accumulate in soils when CIP-contaminated from date palm leaflets using sulphuric acid carbonization,
sludge is utilized as fertilizer (Golet et al., 2003). CIP and to investigate its capability for the adsorption of CIP
was also found to accumulate in plants such as lettuce, from aqueous solution in terms of kinetics and equilibrium.
cucumber and common barley when cultivated in CIP-
spiked soil (Lillenberg et al., 2010). 1 Materials and methods
The presence of CIP in wastewater and surface water,
even at low concentrations, is considered an environmental 1.1 Materials
hazard as it can develop antibiotic resistance in bacteria All the chemicals used were of analytical grade. Pure
* Corresponding author. E-mail: dr el shafey2004@yahoo.co.uk
ciprofloxacin HCl·H2 O, CIP, was supplied by the National
1580 Journal of Environmental Sciences 2012, 24(9) 1579–1586 / El-Said Ibrahim El-Shafey et al. Vol. 24

Pharmaceutical Industries Company, Muscat, Oman. Palm 1.4 Physicochemical characterization


leaflets were collected from a local farm in Muscat and
Physico-chemical tests were carried out using the dry
were washed with distilled water to remove dirt, dust and
carbon. The surface area was measured using an Autosorb-
surface impurities. The cleaned leaflets were put in plastic
1(Quantachrome Instruments, USA) via nitrogen adsorp-
trays and left to dry at room temperature. A stock CIP
tion at 77 K. Scanning electron microscope (SEM) analysis
solution, 500 mg/L, was prepared by dissolving 0.5 g of
was carried out using a Geol JSM 840-A scanning electron
CIP in 1000 mL deionized water. Other concentrations
microscope (Jeol, Japan). X-ray powder diffraction was
were prepared by suitable dilution with deionized water.
carried out to reveal information about the crystallographic
structure of the carbon. A Philips PW 1830 generator with
1.2 Preparation of the carbon adsorbent
a Philips PW 1050 powder goniometer (Philips, USA) was
used and the diffraction patterns were determined using
Clean dry palm leaflets were cut into small pieces and
copper Kα radiation with a Ni filter. FT-IR spectroscopy
were chemically carbonized via sulphuric acid treatment
was also employed using a FT-IR spectrometer (Spectrun
as follows: 40 g of clean dry palm leaflets was mixed
BX, Berkin Elmer, Germany) in the wavelength range
with 12 mol/L sulphuric acid (100 mL). The mixture was
of 4000–800 cm−1 with background subtraction. Surface
heated to (160 ± 2)°C and was kept in that temperature
functional groups were also determined using Boehm titra-
range for 25 min for complete carbonization. The resulting
tions (Boehm, 1996; El-Shafey, 2007). The point of zero
black mixture was allowed to cool to room temperature and
charge (PZC) of the carbon was determined by following
the spent sulphuric acid was filtered off using a Buchner
procedure (Lu et al., 2009): 0.1 g of the dry carbon was
funnel under vacuum. The produced wet carbon was
placed in a 25 mL Erlenmeyer flask to which 20 mL of
stored under 1 mol/L sulphuric acid solution to avoid any
CO2 -free water was added. The flask was sealed with a
possible bacterial growth. Before use, the wet adsorbent
cap and left with occasional agitation for 48 hr at room
was washed in a Gooch crucible (G2) with distilled water
temperature. The pH of the filtrate was then measured and
to remove residual acid until the wash water did not show
this value is the point of zero charge. This method has
a change in methyl orange color. The sample was then
been used satisfactorily by Moreno-Castilla et al. (2000)
left under suction for 30 min to remove water among
and Leyva-Ramos et al. (2005).
the particles. The wet carbonaceous product was then
used for adsorption experiments and 1 g sample was used
for moisture content determination. Moisture content was 2 Results and discussion
determined in each experiment using the wet adsorbent and
was found to be 76%. For experiments with dry adsorbent, 2.1 Characterization of chemically carbonized adsor-
the acid-free wet carbon was dried at 120°C to constant bent
weight. The dry carbon was transferred to a desiccator Cellulose, hemicelluloses and lignin are the main con-
to cool and then used in the adsorption experiments. The stituents of the agricultural waste of date palm leaflets.
yield of the carbon produced was 96%. Concentrated sulphuric acid acts as an extremely strong
dehydrating agent, carbonizing the cellulose and hemicel-
1.3 Adsorption experiments luloses via the removal of water. In addition, it acts as
a strong oxidizing agent because of its affinity to lose
For the kinetic experiments, samples of 0.4 g of the dry an atom of oxygen to form sulphurous acid, that readily
carbon (or its equivalent mass of wet carbon) were mixed decomposes to sulfur dioxide and water (Parker, 1997).
with CIP (200 mL, 100 mg/L) at initial pH 6. At different Cellulose is susceptible to oxidizing agents due to the
time intervals, an aliquot of supernatant was withdrawn for presence of three hydroxyl groups in each anhydroglucose
analysis. Batch experiments were carried out by mixing 0.1 unit (Nikitin, 1996). Lignin is more readily oxidized than
g of the adsorbent with 50 mL of CIP solution of the de- cellulose and other wood polysaccharides (Nikitin, 1996).
sired concentration, pH and temperature in a shaking water Under the applied preparation conditions, carbonization
bath (100 r/min) until equilibrium was reached. The effect (via dehydration) with partial oxidation took place to the
of pH on CIP adsorption was studied for a concentration cellulose and hemicelluloses, however, for lignin, partial
of 200 mg/L at different pH values (2.0–11.5). Before the oxidation and fragmentation took place (Cox et al., 1999).
addition of the pre-weighed adsorbent to the CIP solution, The scanning electron microscope image in Fig. 1 shows
the solution pH was adjusted by adding 0.1 mol/L NaOH that the carbonaceous adsorbent still maintains the plant’s
or 0.1 mol/L H2 SO4 . Equilibrium isotherm studies were fibrous texture. The X-ray diffraction pattern showed no
carried out at initial pH 6 for different CIP concentra- observable peaks, indicating the amorphous structure of
tions (50–300 mg/L) at different temperatures (25–45°C). the adsorbent (data not shown).
After the equilibrium time was reached, an aliquot of The FT-IR spectrum of the produced carbon (Fig. 2)
the supernatant was analyzed spectrophotometrically using shows the presence of hydroxyl, carboxyl and other
a UV-Visible spectrophotometer (Varian/Cary/50 Conc., carbon-oxygen species as assigned in Table 1. Simi-
Varian Incorporation, The Netherlands) at a wavelength lar functional groups were identified on the surface of
(λ) of 273 nm (Wang et al., 2010). All experiments were chemically-carbonized carbon from flax shive via sul-
carried out at least twice. phuric acid treatment (Cox et al., 1999). Surface acidic
No. 9 Ciprofloxacin adsorption from aqueous solution onto chemically prepared carbon from date palm leaflets 1581

for a carbonaceous adsorbent prepared from flax shive


via sulphuric acid carbonization. The ash content was
determined to be 11.5%.
2.2 Effect of pH
Adsorbed CIP was calculated from the decrease in CIP
concentration in solution by considering the adsorption
volume and amount of the adsorbent (Eq. (1)).

(C0 − Ce )V
qe = (1)
m
where, qe (mg/g) is the amount of CIP (mg) adsorbed onto
a unit mass of the carbon (g) on dry basis at equilibrium;
C 0 (mg/L) and C e (mg/L) are the concentrations of CIP in
Fig. 1 SEM image of the chemically carbonized adsorbent. the initial solution and at equilibrium; m (g) is the amount
of adsorbent used on dry basis and V (L) is the volume of
functional groups were determined by Boehm titra- CIP solution.
tions, where NaHCO3 neutralizes carboxylic, Na2 CO3 The CIP acid dissociation constants pK a1 and pK a2 are
neutralizes carboxylic and lactonic, and NaOH neutral- 6.1 and 8.7 respectively (Gu and Karthikeyan, 2005) as
izes carboxylic, lactonic and phenolic functional groups presented in Fig. 3. At pH < 6.1, CIP molecules mainly
(Boehm, 1996). The number of milliequivalents (meq) exist as cations (CIP+ ) because of the protonation of the
calculated based on the titrations for the surface carboxyl,
lactone and phenol groups were 2.6, 1.5 and 0.82 meq/g, O O
pKa1 6.1
respectively. The carbon pHpzc (Lu et al., 2009) was found
F
to be 3.4. HO
The surface area measured via nitrogen adsorption at 77
K using the BET model was 24.4 m2 /g. The monolayer
N N
capacity and the BET constant were found to be 5.61
cm3 /g and 149 respectively. Such a low surface area NH
can be possibly related to the presence of carbon-oxygen
functional groups that occupy a large fraction of the pKa2 8.7
adsorbent surface (Youssef et al., 1995). A low surface Fig. 3 Molecular structure of ciprofloxacin.
area of 19 m2 /g was also found by Cox et al. (1999)
60
46 pKa1 = 6.1 pKa2 = 8.7
44 55
42
CIP sorbed (mg/g)

40 50
Transmittance (%)

38
36 45
34
32 40
30
28 35 Wet
26 Dry
24 30
0 2 8 4 10 6 12 14
22
Initial pH
4000 3600 3200 2800 2400 20001800 1600 1400 1200 1000 800
Wavenumber (cm-1) Fig. 4 Adsorption of CIP at different initial pH values and 25°C on wet
and dry carbons. Initial concentration 200 mg/L, volume of CIP solution
Fig. 2 FT-IR spectrum of the carbonaceous adsorbent. 50 mL, shaking speed 100 r/min.

Table 1 FT-IR band positions and suggested assignment for the functional groups on the surface of the carbonaceous adsorbent

Band position (cm−1 ) Band assignment Reference

3410 O–H stretching vibrations (H-bonded) Zawadzki, 1980


2920, 2855 C–H stretching vibrations in CH2 George and McIntyre, 1987
1700 –C=O stretching vibrations (lactone and carboxylic acids) George and McIntyre, 1987
1613 –COO− asymmetric and symmetric stretching (carboxylate) Gómez-Serrano et al., 1994; van der Mass, 1969
or skeletal C=C aromatic vibrations
1300–900 stretching C–O vibrations in hydroxyl groups and ether type structures Gmez-Serrano et al., 1994
1582 Journal of Environmental Sciences 2012, 24(9) 1579–1586 / El-Said Ibrahim El-Shafey et al. Vol. 24

amine group in the piperazine moiety (Wang et al., 2010). 2.3 Kinetics of ciprofloxacin adsorption
At pH > 8.7, CIP molecules exist as anions (CIP− ) due to
Based on the results of the pH experiments, initial pH
the loss of a proton from the carboxylic group. In the pH
6 was chosen for the studies of adsorption kinetics and
range 6.1–8.7, the pH values are higher than pK a1 of the
equilibrium. Equilibrium was reached within 48 hr. As
carboxylic group, thus deprotonating the carboxylic group
temperature increased, CIP adsorption appeared to be
to the negatively charged carboxylate. However, the amine
faster, with higher performance for the wet carbon than for
group stays protonated and positively charged as this pH
the dry (Fig. 5).
range is still lower than pK a2 of the amine group. Ac-
CIP uptake in the initial stages of adsorption, qt , varied
cordingly, in the above mentioned pH range, zwitterionic
almost linearly with the square root of time, t0.5 . This
species of CIP molecules dominate in the aqueous solution.
indicates that the initial rate of adsorption is controlled by
Zwitterion formation of CIP molecules in the range of
intraparticle diffusion, based on Eq. (2), where kd is the
pH between pK a1 and pK a2 has been emphasized in other
rate constant for pore diffusion (Weber and Morris, 1963)
studies (Drakopoulos and Ioannou, 1997; Carmosini and
and qt (mg/g) is the amount of adsorbed CIP at time t.
Lee, 2009; Wu et al., 2010, Carabineiro et al., 2011b; Wang
et al., 2010, 2011).
As shown in Fig. 4, CIP adsorption was low at pH below qt = kd × t0.5 (2)
3 and increased with the rise in initial pH, showing a max-
imum adsorption around pH 6, with better performance The calculated values of kd (Table 2) appear higher
for the wet adsorbent than for the dry. Adsorption started for the wet carbon than for the dry. The drying pro-
to decrease at pH higher than 6; however, a significant cess narrowed the adsorbent pores due to shrinkage and
decrease was shown beyond pH 8.7. Thus, pH 6 was compaction, causing less diffusion of CIP into the dry
selected as the optimum pH for the kinetic and equilibrium adsorbent. An increase in kd values was observed with
studies. At pH < 3.4 (pHpzc of the carbon), the carbon rising temperature, as adsorbent swelling is expected with
surface stays protonated (such as –COOH), however, at pH higher temperatures, giving wider pores and, consequently,
> 3.4, the carbon surface becomes negatively charged. At more diffusion of CIP molecules.
pH  3, adsorption was low between the protonated surface Pseudo first- and second-order kinetic models (Eq. (3)
and the CIP cations. However, in the pH range 3.4–6.0, and (4), respectively), were tested for the adsorption kinet-
electrostatic interaction is maximized between the nega- ic data (Zhang et al., 2011).
tively charged surface and the CIP cations. In the pH range
6.1–8.7, electrostatic interaction is believed to take place
between the positively charged amine group in the CIP k1 t
log(qe − qt ) = log qe − (3)
zwitterion and negatively charged carbon-oxygen groups 2.303
on the adsorbent surface. However, a degree of repulsion
60
between negatively charged carboxylate in the CIP zwit-
terion and the negatively charged surface probably takes 50
Ciprofloxacin sorbed (mg/g)

place, leading to a slight decrease in CIP sorption. Beyond


pH 8.7, both the carbon and CIP stay negatively charged 40
in solution and this leads to a further decreased uptake.
Similar results were found by Wang et al. (2010) for the ad- 30
sorption of CIP onto montmorillonite. As shown in Fig. 4,
adsorption still appears significant below pH 3 and above 20
pH 8.7, and this is due to the presence of hydrogen bonding
Wet, 25°C Dry, 25°C
between the CIP molecules and carbon-oxygen functional 10 Wet, 35°C Dry, 35°C
groups on the carbon surface. Electrostatic interactions and Wet, 45°C Dry, 45°C
hydrogen bonding were reported as adsorption forces for 00 10
30 40 20 50 60
CIP adsorption on functionalized silicates (Punyapalakul Time (hr)
and Sitthisorn, 2010) and montmorillonite (Wang et al., Fig. 5 Adsorption of CIP on the wet and dry carbons as a function of
2010). time at different temperatures. Initial pH 6, volume of solution 200 mL,
shaking speed 100 r/min, initial concentration 100 mg/L.

Table 2 Pseudo first-order and second-order rate constants of the kinetics of CIP sorption at different temperatures

Adsorbent Temp. (°C) Diffusion Pseudo first-order model Pseudo second-order model
model (kd ) k1 (hr) qe (mg/g) R2 k2 (g/(mg·hr)) h (mg/(g·hr)) qe (mg/g) R2
Wet 25 16.02 0.0875 24.6 0.9750 0.0121 24.1 44.6 0.9975
35 17.96 0.0940 21.84 0.9674 0.0153 35.1 47.8 0.9992
45 21.71 0.1094 21.632 0.9606 0.0187 52.9 53.2 0.9997
Dry 25 12.64 0.0654 14.6 0.9115 0.0110 7.05 25.3 0.9976
35 13.73 0.0702 16.3 0.9591 0.0134 11.6 29.4 0.9986
45 18.52 0.1073 18.3 0.9938 0.0167 18.0 32.8 0.9994
k1 : rate const of pseudo first-order model; k2 : rate const of pseudo second-order model, h: initial adsorption rate.
No. 9 Ciprofloxacin adsorption from aqueous solution onto chemically prepared carbon from date palm leaflets 1583

1/T (hr-1)
0.0031 0.00315 0.0032 0.00325 0.0033 0.00335 0.0034
t 1 t -3.9
= + (4)
qt k2 q2e qe -4.0
y = -2067.7x + 2.527
where, k1 and k2 are the rate constants for the pseudo -4.1 R2 = 0.999
first-and second-order models, respectively. The initial
adsorption rate, h = k2 qe 2 . -4.2

lnk2
The linear plots of log(qe -qt ) versus t for the pseu- -4.3
do first-order model present low correlation values (R2 ), y = -1985.7x + 2.1487
showing a poor fit for the model; however, the linear -4.4 R2 = 0.998
plots of t/qt versus t for the pseudo second-order model -4.5 Wet
show better fitting with high R2 values (Table 2). This Dry
indicates that the adsorption of CIP conforms well with a -4.6
pseudo second-order kinetic reaction, which agrees with Fig. 6 Plots of lnk2 against 1/T for the adsorption of CIP at different
chemisorption as the rate-limiting mechanism through temperatures on wet and dry carbons.
sharing or exchange of electrons between adsorbent and
adsorbate (Ho, 2006). Both of the rate constants k2 (Fig. 7). The experimental isotherm data conform well
(g/(mg·hr)), and the initial adsorption rate, h (mg/(g·hr)), to the Langmuir equation (Zhang et al., 2011), Eq. (6),
were higher for the wet adsorbent than for the dry (Table 2) showing high correlation values of R2 . However, a poor fit
and this, again, could be related to the shrinkage and with low correlation values was found for the application
compaction of the adsorbent as a result of drying, giving of the Freundlich equation (Zhang et al., 2011), Eq. (7), to
narrower pores for the diffusion of CIP molecules (El- the equilibrium adsorption data, Table 3.
Shafey, 2007). The pseudo first-order kinetic model gave
a poor fit and a trend of results cannot be established. Ce 1 Ce
The Arrhenius equation (Eq. (5)) was applied to calcu- = + (6)
qe b×q q
late the activation energy (E a ) for the adsorption processes.
As a result of the good fitting of the kinetic data to the where, q (mg/g) and b (l/mg) are the Langmuir constants
model, the rate constant of the pseudo second-order model, related to maximum adsorption capacity and the relative
k2 , was selected for E a calculation. energy of adsorption, respectively.

k2 = Ae−Ea /RT (5) 1


log qe = log Ce + log K (7)
n
where, E a (kJ/mol) is the activation energy of CIP adsorp-
tion. A is the pre-exponential factor (frequency factor), where, 1/n and K (L1/n mg1−1/n g−1 ) are constants which
R is the gas constant (8.314 J/(mol·K)) and T (K) is the are considered to be the relative indicators of adsorption
solution temperature. From the linear relationship between intensity and adsorption capacity, respectively.
the logk2 and 1/T (Fig. 6), E a was calculated and found The 1.15 and 1.2 fold increases of CIP adsorption on
to be 17.2 and 16.5 kJ/mol for the wet and dry adsor- wet and dry carbons, respectively, were obtained with
bents, respectively. Low E a values (< 42 kJ/mol) indicate temperature rise from 25 to 45°C (Table 3). This could be
diffusion-controlled processes, while higher E a values (> a result of the carbon swelling, allowing more adsorption
42 kJ/mol) indicate chemically-controlled processes (Liu active sites to become free for CIP. In a previous study,
and Huang, 2003). Accordingly, the rate-limiting process Zhang et al. (2011) found that CIP adsorption on modified
for CIP adsorption is evidently physical. coal ash increased with rising temperature. The adsorption
capacities of CIP from the present study were compared
2.4 Adsorption capacity and temperature effect
with other adsorbents in previous studies (Table 4). It is
The adsorption of CIP showed an ‘L-type’ adsorption clear that the cheap carbonaceous adsorbent produced from
isotherm for both wet and dry adsorbents. As temperature date palm leaflets shows a high adsorption capacity for CIP.
rose, adsorption increased with an advantage for the wet The isotherm shape can be used to predict if an ad-
adsorbent over the one that has been previously dried sorption system is ‘favorable’ or ‘unfavorable’. Hall et
Table 3 Isotherm parameters for CIP sorption onto wet and dry adsorbents at different temperatures

Adsorbent Temp. (°C) Langmuir model Freundlich model


q (mg/g) b (mg−1 ) Rs R2 1/n K R2
Wet 25 116.3 0.0664 0.048–0.23 0.9973 0.3129 25.28 0.9797
35 125.0 0.0758 0.042–0.21 0.9975 0.3281 26.67 0.9870
45 133.3 0.0997 0.032–0.17 0.9953 0.3213 31.62 0.9892
Dry 25 104.2 0.0522 0.060–0.28 0.9996 0.3362 19.10 0.9515
35 113.6 0.0601 0.053–0.25 0.9998 0.3545 20.42 0.9484
45 125.0 0.0719 0.044–0.22 0.9986 0.3748 21.97 0.9593
Rs : separation factor.
1584 Journal of Environmental Sciences 2012, 24(9) 1579–1586 / El-Said Ibrahim El-Shafey et al. Vol. 24

140 temperatures (Eq. (9)).


120
qe
100 Kc = (9)
CIP sorbed (mg/g)

Ce
80
K c (L/g) values were obtained using the method of Khan
60 and Singh (1987) by plotting ln (qe /C e ) versus qe and
40 Wet 25°C Dry 25°C
extrapolating to zero qe . The intercept of the straight
Wet 35°C Dry 45°C line with the vertical axis gives the values of Kc . As
20 Wet 45°C Dry 45°C presented in Table 5, the value of K c increases with rising
0 temperature, indicating an endothermic nature for CIP
0 40 20 60 80 100 120
adsorption. The Gibbs free energy change of the adsorption
Equilibrium Ce concentration (mg/L)
process is related to K c as shown in Eq. (10) (Uslu and Inci,
Fig. 7 Adsorption of CIP on wet and dry carbons at initial pH 6 at
different temperatures. 2009).

ΔG0 = −RT lnKc (10)


al. (1966) pointed out that the essential characteristics of
the Langmuir isotherm can be expressed in terms of a The changes in enthalpy (ΔH 0 ) and entropy (ΔS 0 )
dimensionless constant, Rs , as presented in Eq. (8). for CIP adsorption were calculated from the slope and
1 intercept of the plot of lnK c against 1/T according to the
Rs = (8) van’t Hoff equation (Uslu and Inci, 2009), Eq. (11).
1 + bC0
where, Rs is the separation factor or equilibrium parameter, ΔS 0 ΔH 0
lnKc = − (11)
which is a direct function of the Langmuir constant b. R RT
The values of Rs indicate the shape of the isotherm: Rs > Plotting lnK c versus 1/T for the wet and dry adsorbents
1 (unfavorable), Rs = 1 (linear), 1 > Rs > 0 (favorable) shows straight lines (Fig. 8). From their slope and inter-
and Rs = 0 (irreversible) (Hall et al., 1966). As shown in cept, ΔH 0 and ΔS0 were determined. The negative values
Table 3, Rs values are in the range of 0.032–0.23 for the of ΔG0 (Table 5) indicate a favorite and spontaneous
wet adsorbent, and 0.044–0.28 for the dry, thus, indicating process (Li et al., 2011).
‘favorable adsorption’ and strong binding between the CIP The positive values of ΔH 0 show that CIP adsorption is
and the adsorbent. endothermic on both adsorbents. As ΔH 0 values are below
R2 values appear more satisfactory for the Langmuir 40 kJ/mol (Table 5), it can be concluded that the nature of
isotherm than for the Freundlich isotherm (Table 3). As the adsorption process is physical (Chern and Wu, 2001),
the Langmuir adsorption isotherm is based on monolayer either cation exchange or hydrogen bonding (Li et al.,
coverage of the adsorbate onto the active sites of the
adsorbent surface (Nadeem et al., 2006), CIP sorption onto Table 5 Thermodynamic parameters of CIP adsorption on wet and dry
the investigated carbon (wet and dry) seems to generate adsorbents
monolayer formation. Adsorbent Temp. Kc ΔG0 ΔH 0 ΔS 0
(K) (L/g) (kJ/mol) (kJ/mol) (kJ/mol)
2.5 Thermodynamic calculations
Wet 298 1.71 –1.33 11.68 43.63
Thermodynamic parameters were calculated from the 308 1.95 –1.71
variation of the equilibrium constant, K c , at different 318 2.31 –2.21
Dry 298 1.33 –0.71 13.10 46.4
Table 4 Maximum adsorption capacities of CIP on different adsorbents 308 1.57 –1.15
318 1.86 –1.64
Adsorbent Maximum Reference
adsorption
(mg/g) 1.2

Kaolinite 6.3 Li et al., 2011 1.0


Kaolinite 7.95 Mackay and Seremet, 2008 y = -1576.5x + 5.5748
Rectorite 135 Wang et al., 2011 0.8
R2 = 0.9983
Illite 33 Wang et al., 2011
lnKc

Montmorillonite 395 Wang et al., 2011 0.6


Montmorillonite 330 Wang et al., 2010
Geothite (Syn-FeOOH) 49.8 Zhang and Huang, 2007 0.4
Geothite (Ald-FeOOH) 33.1 Zhang and Huang, 2007 y = -1405.4x + 5.2474
Modified coal fly ash 1.55 Zhang et al., 2011 Wet R2 = 0.9914
0.2
Aluminous oxide 13.6 Gu and Karthikeyan, 2005 Dry
Oxidized xerogel 60 Carabineiro et al., 2011a 0
Chemically prepared 133.3 This study 0.0031 0.00315 0.0032 0.00325 0.0033 0.00335 0.0034
carbon (wet) 1/T (K-1)
Chemically prepared 125.0 This study
Fig. 8 Plots of lnK c against 1/T for the adsorption of CIP at different
carbon (dry)
temperatures.
No. 9 Ciprofloxacin adsorption from aqueous solution onto chemically prepared carbon from date palm leaflets 1585

2011). The positive value of ΔS 0 indicates an increase Carabineiro S A C, Thavorn-amornsria T, Pereiraa M F R, Serp
in randomness at the solid-solution interface during the P, Figueiredo J L, 2011b. Comparison between activated
adsorption of CIP onto the carbon adsorbents. carbon, carbon xerogel and carbon nanotubes for the ad-
Similar thermodynamic results were found for CIP sorption of the antibiotic ciprofloxacin. Catalysis Today,
adsorption onto modified coal fly ash (Zhang et al., 2011), 186(1): 29–34.
Carmosini N, Lee L S, 2009. Ciprofloxacin sorption by dissolved
showing negative ΔG0 in the range of –5.4 to –2.8 kJ/mol,
organic carbon from reference and bio-waste materials.
positive ΔH 0 (14.9–28.6 kJ/mol) and positive ΔS 0 (55.3– Chemosphere, 77(6): 813–820.
107 J/mol). Chern J M, Wu C Y, 2001. Desorption of dye from activated
2.6 Desorption studies carbon beds: effects of temperature, pH and alcohol. Water
Research, 35(17): 4159–4165.
Ciprofloxacin desorption was investigated using 0.5 mol/L Cox M, El-Shafey E I, Pichugin A A, Appleton Q, 1999.
HCl at room temperature. For a sorbent sample (0.5 g) Preparation and characterisation of a carbon adsorbent from
loaded with CIP at 100 mg/g, desorption reached 83% at flax shive by dehydration with sulfuric acid. Journal of
such acidic conditions. The protons replace CIP cations on Chemical Technology and Biotechnology, 74(11): 1019–
the ion exchange sites. The undesorbed fraction of CIP 1029.
under such acidic conditions is probably related to other De Witte B, Van Langenhove H, Demeestere K, Saerens K, De
Wispelaere P, Dewulf J, 2010. Ciprofloxacin ozonation in
sorption forces between CIP molecules and the carbon
hospital wastewater treatment plant effluent: effect of pH
surface, such as hydrogen and hydrophobic bonding. and H2 O2 . Chemosphere, 78(9): 1142–1147.
Drakopoulos A I, Ioannou P C, 1997. Spectrofluorimetric study
3 Conclusions of the acid-base equilibria and complexation behavior
of the fluoroquinolone antibiotics ofloxacin, norfloxacin,
Chemically-prepared carbon is cheap and efficient for ciprofloxacin and pefloxacin in aqueous solution. Analytica
CIP adsorption. The wet carbon shows better diffusion Chimica Acta, 354(10): 197–204.
and more adsorption capacity than the dry carbon. CIP El-Shafey E I, 2007. Sorption of Cd(II) and Se(IV) from aqueous
adsorption shows a maximum at initial pH 6 because of solution using modified rice husk. Journal of Hazardous
Materials B, 147(1-2): 546–555.
the electrostatic interaction between positively charge CIP
George W O, McIntyre P S, 1987. Infrared Spectroscopy. John
and negatively charged carbon. The activation energy E a Wiley and Sons, Chichester.
was 17 kJ/mol, indicating a diffusion-controlled physical Golet E M, Strehler A, Alder A C, Giger W, 2002. Determination
adsorption process. The calculated thermodynamic param- of fluoroquinolone antibacterial agents in sewage sludge
eters showed that the adsorption process of CIP is physical, and sludge-treated soil using accelerated solvent extraction
spontaneous and endothermic, and the adsorption mecha- followed by solid-phase extraction. Analytical Chemistry,
nism is mainly related to cation exchange and hydrogen 74(21): 5455–5462.
bonding. Golet E M, Xifra I, Siegrist H, Alder A C, Giger W, 2003.
Environmental exposure assessment of fluoroquinolone
Acknowledgment antibacterial agents from sewage to soil. Environmental
Science & Technology, 37(15): 3243–3249.
The authors would like to thank the National Pharma- Gómez-Serrano V, Acedo-Ramos M, López-Peinado A J,
ceutical Industries Company, Muscat, for supplying CIP Valenzula-Calahorro C, 1994. Oxidation of activated carbon
samples. by hydrogen peroxide: study of surface functional groups
by FTIR. Fuel, 73(3): 387–395.
Gu C, Karthikeyan K G, 2005. Sorption of the antimicrobial
References ciprofloxacin to aluminum and iron hydrous oxides. Envi-
ronmental Science & Technology, 39(23): 9166–9173.
Al-Yahyai R, Al-Khanjari S, 2008. Biodiversity of date palm Hall K R, Eagleton L C, Acrivos A, Vermeulen T, 1966. Pore-
in the sultanate of Oman. African Journal of Agricultural and solid-diffusion kinetics in fixed-bed adsorption un-
Research, 3(6): 389–395. der constant-pattern conditions. Industrial & Engineering
Belden J B, Maul J D, Lydy M J, 2007. Partitioning and Chemistry Fundamentals, 5(2): 212–223.
photodegradation of ciprofloxacin in aqueous systems in Hartmann A, Golet E M, Gartiser S, Alder A C, Koller T,
the presence of organic matter. Chemosphere, 66(8): 1390– Widmer R M, 1999. Primary DNA damage but not mu-
1395. tagenicity correlates with ciprofloxacin concentrations in
Bhandari A, Close L I, Kim W, Hunter R P, Koch D E, Surampalli German hospital wastewaters. Archives of Environmental
R Y, 2008. Occurrence of ciprofloxacin, sulfamethoxazole, Contamination and Toxicology, 36(2): 115–119.
and azithromycin in municipal wastewater treatment plants. Ho Y S, 2006. Review of second-order models for adsorption
Practice Periodical of Hazardous, Toxic, and Radioactive systems. Journal of Hazardous Materials B, 136(3): 681–
Waste Management, 12(4): 275–281. 689.
Boehm H P, 1996. Chemical identification of surface groups. Ibezim E C, Ofoefule S I, Ejeahalaka C N C, Orisakwe O E, 1999.
Advances in Catalysis, 16: 179–274. In vitro adsorption of ciprofloxacin on activated charcoal
Carabineiro S A C, Thavorn-Amornsri T, Pereira M F R, and talc. American Journal of Therapeutics, 6(4): 199–201.
Figueiredo J L, 2011a. Adsorption of ciprofloxacin on Khan A A, Singh R P, 1987. Adsorption thermodynamics of
surface-modified carbon materials. Water Research, 45(15): carbofuran on Sn(IV) arsenosilicate in H+ , Na+ and Ca2+
4583–4591. forms. Colloids and Surfaces, 24(1): 33–42.
1586 Journal of Environmental Sciences 2012, 24(9) 1579–1586 / El-Said Ibrahim El-Shafey et al. Vol. 24

Leyva-Ramos R, Bernal-Jacome L A, Acosta-Rodriguez I, 2005. and carbamazepine by adsorption on functionalized meso-


Adsorption of cadmium(II) from aqueous solution on nat- porous silicates. World Academy of Science, Engineering
ural and oxidized corncob. Separation and Purification and Technology, 69: 546–550.
Technology, 45(1): 41–49. Sun S P, Guo H Q, Ke Q, Sun J H, Shi S H, Zhang M L et al.,
Li Z H, Hong H L, Liao L B, Ackley C J, Schulz L A, 2009. Degradation of antibiotic ciprofloxacin hydrochloride
MacDonald R A et al., 2011. A mechanistic study of by photo-fenton oxidation process. Environmental Engi-
ciprofloxacin removal by kaolinite. Colloids and Surfaces neering Science, 26(4): 753–759.
B: Biointerfaces, 88(1): 339–344. Uslu H, Inci I, 2009. Adsorption equilibria of L-(+)-tartaric acid
Lillenberg M, Litvin S V, Nei L, Roasto M, Sepp K, 2010. onto alumina. Journal of Chemical & Engineering Data,
Enrofloxacin and ciprofloxacin uptake by plants from soil. 54(7): 1997–2001.
Agronomy Research, 8(1): 807–814. van der Mass J H, 1969. Basic Infrared Spectroscopy. Heyden &
Liu C, Huang P M, 2003. Kinetics of lead adsorption by iron Son Ltd., London.
oxides formed under the influence of citrate. Geochimica et Wang C J, Li Z H, Jiang W T, 2011. Adsorption of ciprofloxacin
Cosmochimica Acta, 67(5): 1045–1054. on 2:1 dioctahedral clay minerals. Applied Clay Science,
Lu D D, Cao Q L, Cao X J, Luo F, 2009. Removal of Pb(II) using 53(4): 723–728.
the modified lawny grass: Mechanism, kinetics, equilibrium Wang C J, Li Z H, Jiang W T, Jean J S, Chuan C C, 2010. Cation
and thermodynamic studies. Journal of Hazardous Materi- exchange interaction between antibiotic ciprofloxacin and
als B, 166(1): 239–247. montmorillonite. Journal of Hazardous Materials B, 183(1-
Mackay A A, Seremet D E, 2008. Probe compounds to quan- 3): 309–314.
tify cation exchange and complexation interactions of Weber W J Jr, Morris J C, 1963. Kinetics of adsorption on
ciprofloxacin with soils. Environmental Science & Technol- carbon from solutions. Journal of the Sanitary Engineering
ogy, 42(22): 8270–8276. Division ASCE, 89(2): 31–60.
Martins A F, Vasconcelos T G, Hariques D M, Frank C S, König Wu Q F, Li Z H, Hong H L, Yin K, Tie L Y, 2010. Adsorption and
A, Kümmerer K, 2008. Concentration of ciprofloxacin in intercalation of ciprofloxacin on montmorillonite. Applied
Brazilian hospital effluent and preliminary risk assessment: Clay Science, 50(2): 204–211.
a case study. Clean-Soil, Air, Water, 36(3): 264–269. Youssef A M, El-Khouly A A, Ahmed A I, El-Shafey E I, 1995.
Moreno-Castilla C, López-Ramón M V, Carrasco-Marı́n F, 2000. Changes in the adsorption properties of activated carbon
Changes in surface chemistry of activated carbons by wet due to partial oxidation of the surface. Adsorption Science
oxidation. Carbon, 38(14): 1995–2001. & Technology, 12: 211–219.
Nadeem M, Mahmood A, Shahid S A, Shah S S, Khalid A M, Zawadzki J, 1980. IR spectroscopic investigations of the mech-
McKay G, 2006. Sorption of lead from aqueous solution anism of oxidation of carbonaceous films with HNO3
by chemically modified carbon adsorbents. Journal of Haz- solution. Carbon, 18(4): 281–285.
ardous Materials B, 138(3): 604–613. Zhang C L, Qiao G L, Zhao F, Wang Y, 2011. Thermodynamic
Nikitin N I, 1966. The Chemistry of Cellulose and Wood, and kinetic parameters of ciprofloxacin adsorption onto
Translated from Russian by Schmorak J. Israel Program for modified coal fly ash from aqueous solution. Journal of
Scientific Translation, Jerusalem. Molecular Liquids, 163(1): 53–56.
Parker S P, 1997. McGraw-Hill Encyclopedia of Science & Zhang H C, Huang C H, 2007. Adsorption and oxidation of
Technology (8th ed.). McGraw-Hill, London. vol. 7, 161. fluoroquinolone antibacterial agents and structurally related
Punyapalakul P, Sitthisorn T, 2010. Removal of ciprofloxazin amines with goethite. Chemosphere, 66(8): 1502–1512.

You might also like