Carbón Activado Adsorción de Paracetamol Varios Valores en MG G
Carbón Activado Adsorción de Paracetamol Varios Valores en MG G
PII: S2213-3437(20)30757-0
DOI: https://doi.org/10.1016/j.jece.2020.104408
Reference: JECE 104408
Please cite this article as: Nguyen DT, Tran HN, Juang R-Shin, Dat ND, Tomul F, Ivanets A,
Woo SH, Hosseini-Bandegharaei A, Nguyen VP, Chao H-Ping, Adsorption process and
mechanism of acetaminophen onto commercial activated carbon, Journal of Environmental
Chemical Engineering (2020), doi: https://doi.org/10.1016/j.jece.2020.104408
This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.
activated carbon
Dong Thanh Nguyen1, Hai Nguyen Tran2,3*, Ruey-Shin Juang4,5, Nguyen Duy Dat6, Fatma Tomul7,
Andrei Ivanets8, Seung Han Woo9, Ahmad Hosseini-Bandegharaei10, Van Phuong Nguyen3,11*,
Huan-Ping Chao12*
1
Institute of Environmental Technology, Vietnam Academy of Science and Technology, Hanoi, Vietnam
2
Institute of Fundamental and Applied Sciences, Duy Tan University, Ho Chi Minh City, 700000, Vietnam
of
3
Faculty of Environmental and Chemical Engineering, Duy Tan University, Da Nang, 550000, Vietnam
4
Department of Chemical and Materials Engineering, Chang Gung University, Guishan, Taoyuan, 33302, Taiwan
ro
5
Division of Nephrology, Department of Internal Medicine, Chang Gung Memorial Hospital, Linkou, Taiwan
6
Faculty of Chemical & Food Technology, Ho Chi Minh City University of Technology and Education, Thu Duc,
Ho Chi Minh, 700000, Vietnam
-p
7
Burdur Mehmet Akif Ersoy University, Faculty of Arts and Science, Chemistry Department, 15100 Burdur,
Turkey
8
Institute of General and Inorganic Chemistry of National Academy of Sciences of Belarus, st. Surganova 9/1,
re
220072, Minsk, Belarus
9
Department of Chemical and Biological Engineering, Hanbat National University, 125 Dongseodaero, Yuseong-
Gu, Daejeon 305-719, Republic of Korea
lP
10
Department of Engineering, Kashmar Branch, Islamic Azad University, PO Box 161, Kashmar, Iran
11
Institute of Research and Development, Duy Tan University, Da Nang 550000, Vietnam
12
Department of Environmental Engineering & Center for Environmental Risk Management, Chung Yuan
na
*Corresponding authors:
ur
Hai Nguyen Tran; Institute of Fundamental and Applied Sciences, Duy Tan University, Ho Chi Minh
City, Vietnam; Faculty of Environmental and Chemical Engineering, Duy Tan University, Da
Nang City, Vietnam; [email protected].
Jo
Van Phuong Nguyen; Institute of Research and Development, Duy Tan University, Da Nang, Vietnam;
Faculty of Environmental and Chemical Engineering, Duy Tan University, Da Nang, Vietnam;
[email protected].
Huan-Ping Chao; Department of Environmental Engineering & Center for Environmental Risk
Management, Chung Yuan Christian University, Taoyuan, Taiwan; [email protected].
1
Abstract
The presence of acetaminophen (also known as paracetamol; PRC) micropollutant in water caused
some potential health risks for human. In this study, commercial activated carbon (CAC), which had
been oxidized with HNO3 by supplier to increase the oxygen-functional groups, was applied to
remove PRC from water. Results demonstrated that CAC is a dominantly mesoporous (accounting
for 76.3%) with high surface area (SBET = 1284 m2/g) and total pore volume (VTotal = 0.680 cm3/g).
CAC possessed abundantly oxygen-containing functionalities and low pHPZC (4.95). Raman
of
spectrum of CAC indicated that CAC possessed a more disordered structure with a high intensity
ratio of D band and G band (ID/IG = 2.011). Adsorption study showed that the adsorption capacity of
CAC towards PRC was less affected by solution pH value (2.0–10), and ionic strength (0–1.0 M
ro
NaCl), and different water matrixes (distilled water, tap water, seawater, water treatment plant,
groundwater, and beauty salon). The adsorption process occurred rapidly, with around 52% of PRC
in solution (~517 mg/L) being removed within 5 min of contact. The Langmuir maximum
-p
adsorption capacity of CAC was 221 mg/g under 1.0 g/L of CAC, pH 7.0, 25 °C, and initial
concentration of paracetamol (~100–1200 mg/L). The pore-filling was the most important
re
mechanism. The SBET and VTotal of CAC after adsorption decreased (by approximately 96% for both)
to 45.6 m2/g and 0.039 cm3/g, respectively. The second important mechanism involved in n-π
interaction was established by a remarkably decrease in the band intensity (the FTIR spectrum after
lP
adsorption) at 1630 cm-1 (the C=O group). Weak π-π interaction was confirmed a significant
decrease in the ID/IG ratio from 2.011 to 1.947 after adsorption. Hydrogen bonding formations were
recommended by decreasing band intensity in FTIR spectrum at 3448 cm -1 (–OH) and 1045 cm-1
(C–O). Weak van der Waals force was identified through the study of effect of solution temperature
na
and desorption. Consequently, oxidized CAC can serve as a promising and potential material for
efficiently eliminating PRC from water environments.
Keywords: Acetaminophen; Activated carbon; adsorption mechanism; emerging contaminant;
ur
paracetamol; oxidation.
Jo
2
1. Introduction
pharmaceuticals, personal care products, endocrine disruptors, etc.) present in water environment
has caused an enormous concern among general public. Among them, pharmaceuticals have been
ubiquitously detected in different water matrixes, especially in surface and wastewaters [1]. This is
because they were extensively used not only for human but also veterinary medicine [2].
of
one of the most common antipyretic and analgesic drug frequently used in many countries around
the world for the effective pain relief of several diseases such as headaches, fever, and minor aches
ro
[3, 4]. Paracetamol drug is frequently detected in ground water, surface water, drinking water,
-p
municipal wastewater, and especially in wastewater treatment plant [5]. Therefore, it is essential to
(alone or combined methods). They comprise adsorption [6], biodegradation [7], membrane
filtration [8], and advanced oxidation processes (AOPs). Several common AOPs, which was applied
na
to effectively remove PRC from water, includes Fenton and photo-Fenton reactions [9], UV/chlorine
and UV/H2O2 AOPs [10], photocatalytic oxidation [11], electrochemical oxidation [12], catalytic
ur
ozonation [13], and ozonation [14]. However, during the PRC degradation under AOPs, some by-
products (lower low-molecular weight compounds) is usually generated [15]. For example, Dao et
Jo
al. [10] found that, approximately twenty compound structures (transformation products) were
detected during the PRC degradation by the UV/chlorine AOP. The presence of such by-products
can cause the secondary pollutant. In contrast, adsorption has been acknowledged as the green and
promising method for effectively removing various emerging contaminants from water media [4,
3
16]. Therefore, the low-cost technology of adsorption using activated carbon as adsorbent is
In this study, commercial activated carbon (CAC) was used as a potential adsorbent to
remove PRC emerging pollutant from water media. CAC was characterized by various techniques
Fourier-transform infrared spectroscopy (FTIR), Raman technique, and their electrical state (pHPZC).
The process of PRC adsorption in solution was investigated in a common batch technique to explore
of
the effects of different operational conditions (temperature, solution pH, ionic strength, initial PRC
ro
concentration, contact time, and desorbing agent) on the adsorption capacity of CAC towards PRC.
Three important components of the adsorption study (kinetics, isotherms, and thermodynamics)
-p
were also provided. The adsorption mechanism was discussed and proposed herein.
All chemicals were of reagent grade. Paracetamol (or acetaminophen), methanol, and ethanol
were purchased from Sigma-Aldrich. Some other chemicals (NaCl, NaOH, HCl, H2O2, acrylic acid,
na
and HNO3) were obtained from Scharlau. They were directly used without any further purification.
Commercial activated carbon (CAC) was bought from a local company. According to the
ur
company’s information, CAC has been oxidized to enhance the density of oxygen-functional groups
Jo
on the surface of CAC. Briefly, CAC was well mixed with HNO3 at 40 °C for 24 h. During the
mixture process, a very small amount of iron was transferred into this solution. In this study, CAC
was washed with pure water three times to remove some unexpected impurities and then dried at
105 °C for 24 h. The dried CAC was ground, sieved into the size range of 0.106–0.250 mm, dried at
4
105 °C for 24 to obtain a constant moisture percentage (~5%), and then stored in a desiccator until
further use.
The dried particle sizes of CAC used for the study of its characterization ranged from 0.106 to
0.250 mm. The textural properties of CAC were calculated from the adsorption/desorption isotherm
of nitrogen at 77 K (PMI’s BET-Sorptometer; BET 201-A). The t-plot method was used to calculate
the surface area and volume of the micro-pores. Surface morphology of CAC was explored by
of
SEM-EDS images (JSM-6510 LV). FTIR (IR Nicolet iS10) can measure the range from 400 to
ro
4,000 wavenumber. The CAC pelleted mixing with KBr was used to identify the main functional
groups on the surface of CAC. Raman spectrometer (Xplora Plus; HORIBA) was applied to provide
-p
the information on the crystal structure and molecular vibration of CAC.
re
The pH value of point zero charge (pHPZC) of CAC was obtained from the common drift
method [17, 18]. Briefly, approximately 0.1 g of CAC was transferred onto an Erlenmeyer flask
lP
containing 50 mL of NaCl solution (0.1 M). The pH of NaCl solution (denoted as pHBefore) have
before adjusted from 2.0±0.2 to 10±0.2 using 1.0 M HCl or NaOH. After shaking at 30 °C for 6 h,
na
the mixture of CAC and NaCl solution was separated by a filter paper. The pH value of the filtrate
(pHAfter) was recorded. The pHPZC value of CAC was estimated through the plot of (1) pHBefore
ur
versus pHAfter, (2) pHBefore versus ∆pH (pHBefore – pHAfter), or (3) pHBefore versus ∆pH (pHAfter –
pHBefore).
Jo
The process of PRC adsorption in solution onto CAC was conducted at different operation
conditions. The adsorption experiment was conducted in triplicate, and the result was reported as
average ± standard deviation (SD). In general, the solid/liquid (m/V) ratio of the adsorption study
5
was constant at 2.0 g/L. Approximately 0.5 g of CAC (m) was added into 0.25 L (V) of solution
containing PRC at a target concentration. The solution pH value was adjusted before and during the
adsorption using 1M HCl and 1M NaOH. The mixture of the solid/liquid was shaken at 150 rpm in a
water batch isothermal shaker at a controlled temperature. After reached equilibrium, the
solid/liquid mixture was separated using 0.45 µm filter paper. The laden solid (PRC-adsorbed CAC)
had been collected and dried at 105 °C for 48 h before it was used to characterize. Meanwhile, the
residual liquid was analyzed its concentration using HPLC (Dionex 3000 Thermo, Sunnyvale,
of
California, USA) coupled with a photo-diode array. The analyzing conditions have been published
in our recently previous study [10]. Notably, the blank samples (the PRC solution) were
ro
simultaneously carried out without the presence of CAC. Such samples that were denoted as the
-p
initial PRC concentration (Co; mg/L) played an important role in correctly calculating the amount of
PRC adsorbed by CAC at equilibrium (qe; Equation 1) or any time t (qt; Equation 2). The detail on
re
the adsorption conditions (i.e., pH, ionic strength, initial PRC concentration, time, temperature, etc.)
qe
C o Ce V (1)
m
na
qt
Co Ct V (2)
m
ur
where Co, Ct, and Ce are the concentrations of PRC in solution (mg/L) at the beginning, any time t,
Jo
and equilibrium, respectively; V (L) is the volume of the PRC solution; and m (g) is the mass of used
CAC.
6
2.4. Feasible study of desorption and reuse of laden CAC
Desorption study was conducted after the concentration of equilibrium adsorption reached
673 mg/L, pH 7.0, 25 °C, and the solid/liquid ratio of 2.0 g/L. The laden CAC was desorbed by
different desorbing agents. They included deionized distilled water (pH 2 and pH 12), HCl (0.2 M),
NaOH (0.2 M), NaCl (0.2 M), ethanol, and methanol. The solid/liquid ratio of the desorption study
was also maintained at 2.0 g/L. After shaking with 150rpm at 25 °C for 6 h, the residual
concentration of PRC in solution was determined. The percentage of PRC desorbed (ηDes) from
of
laden CAC was calculated from Equation 3.
ro
qe qd (3)
Des (%) (qe ) 100
qe
-p
where qe (mg/g) is defined from Equation 1, and qd (mg/g) is the amount of PRC desorbed from the
re
laden CAC.
lP
The common carbonization method was applied to explore the potential reuse of laden CAC.
The laden CAC was carbonized at 600 °C for 2 h. It is expected that the PRC adsorbed by CAC was
completely thermally degraded under 600 °C because the endset temperature of pure PRC
na
(determined by thermogravimetric analysis) was around 380 °C [19]. The adsorption was
consecutively conducted at the solid/liquid ratio of 2 g/L, initial PRC concentration of around 500
ur
mg/L, 25 °C, 72 h, and pH = 7.0. The amount of PRC adsorbed by CAC after each cycle was
Jo
The difference between the qe value obtained from experiment (qe,exp) and that predicted from
each applied model (qe,model) was reflected through error functions. To minimize error functions
during modelling, the nonlinear optimization technique was applied using the Origin software. The
7
statistics (Section S1; Supporting Information) includes adjusted coefficient of determination (adj-
R2), chi-square test (χ2), standard deviation of residues (SD), Bayesian information criterion (BIC),
and Marquardt's percent standard deviation (MPSD). The best fitting model was reflected and
selected through the highest value of adj-R2, but the lowest value of the others (SD, BIC, and
MPSD).
of
3.1.1. Textural and morphological properties
ro
The textural property of an adsorbent plays an essential role in estimating its maximum
-p
adsorption capacity and relevant mechanism (i.e., pore-filling). An adsorbent with larger specific
surface and higher total pore volume often exhibits an excellent adsorption capacity to aromatic
re
pollutants (i.e., dye adsorbate) if the pore-filling is primary adsorption mechanism. The textural
parameters (i.e., the characteristics of surface area and pore volume) are calculated from the nitrogen
lP
The gas physisorption isotherm of CAC (Figure 1a) belongs to reversible Type Ib, suggesting
na
that the pore size distribution of CAC comprised both wider micropore and narrowed mesopore
[20]. In addition, a visible wide knee of the isotherm curve (also known as a H4-type hysteresis
ur
loop) appeared at a relative pressure (p/po) higher than 0.4, confirming that CAC was dominant
Jo
mesoporous material with dominant pores diameter centered at 2.44 nm and 4.13 nm in mesopore
range (Figure 1c). Similarly, a porous structure of CAC was visible in a SEM image (Figure S1).
Textural parameters of CAC are provided in Table 1. As expected, CAC exhibited large SBET
(1248 m2/g) and high VTotal (0.680 cm3/g) values along with a dominant mesoporous nature (average
pore diameter ≈ 2.18 nm). Therefore, it is expected that CAC with a larger specific surface area and
8
well-developed internal pore structure can effectively remove paracetamol molecules from water
The presence of main functional groups on the surface of pristine CAC was qualitatively
detected by FTIR. Figure 2 shows some important bands within a wavenumber range of 4000–650
cm-1. Such bands were identified at proximately 3448 cm-1 (–OH groups in the carboxylic, phenols,
or absorbed water), 1630 cm-1 (C=O groups in the carboxylic and lactonic groups); 1386 cm-1 (C=C
of
groups in the aromatic rings); and 1045 cm-1 (C–O groups in ethers or lactones) [21-23]. In essence,
ro
the oxygen-containing functionalities often play a proper role in reacting with aromatic
contaminants (i.e., phenol and dye) in solution through n-π interaction and hydrogen bonding
-p
formation [24-26]. CAC exited a strong band intensity at 1630 cm-1 (the C=O group) because it has
re
been oxidized. This C=O group is often expected as strongly active groups to adsorbing aromatic
contaminants in solution.
lP
The presence of aromatic structures and defective natures in carbonaceous materials was
na
assessed through the Raman spectroscopy. The UV-laser Raman spectrum of CAC (Figure 3)
revealed two pronounced bands of non-graphitic carbon materials at around 1290 cm-1 (D-band) and
ur
1590 cm-1 (G-band). The D-band corresponds to a disordered graphitic lattice or graphene layer
Jo
edges; meanwhile, the G-band is relative to an ideal graphitic lattice [27, 28]. The defective density
present on the graphene structure or the degree of disorder in CAC is defined by an intensity ratio of
D band and G band (ID/IG). As showed in Figure 3, the indicator (ID/IG ratio) of CAC (~2.011) was
higher than unity and some other carbonaceous material in the literature, for example, porous
spheres activated carbon (ID/IG = 1.26) and spherical biochar (ID/IG = 1.14) derived from commercial
9
lignin [29], biochar from bamboo (ID/IG = 0.81) [30], biochar (ID/IG = 1.16) and AC (ID/IG = 1.10)
from tobacco waste [31], and biochar from peanut shell (ID/IG = 0.947) [32]. CAC with a high ID/IG
ratio suggested that it exhibited a more disordered structure. Similarly, Maal-Bared [30] investigated
the effect of pyrolysis temperature from 700 °C to 1000 °C on the ID/IG ratio of AC generated from
bamboo. They found that an increase in the pyrolysis temperature lead to increasing the ID/IG ratio
from 1.05 (700 °C), 1.08 (800 °C), to 1.14 (900 °C), and then to 1.16 (1000 °C). They concluded
that an increase in the ID/IG ratio demonstrated that the extent of disorder in atomic arrangement.
of
Therefore, it is expected that the aromatic C=C bonds in the aromatic rings of CAC can well react
ro
3.2. Performance of paracetamol adsorption process
-p
re
In essence, the adsorption process of adsorbate into an adsorbent is strongly dependent on
two proper operation conditions (pH value of solution and ionic strength). This is because they
lP
greatly affect (1) the property of the charge surface of adsorbent and (2) the species of adsorbate in
solution.
na
First, an adsorbent (i.e., activated carbon) often exhibits an amphoteric nature; it means that
the charge density on its surface often changes within the change of solution pH (pH solution). In this
ur
study, the pHPZC of CAC is used to identify the situation of surface charge of CAC in solution.
Figure 4a indicates three common presentations how to determine pHPZC value from the “drift
Jo
method”. According to Figure 4a, the pHPZC of CAC was established as 4.95. This means that the
surface charge of CAC is predominantly negative when pHsolution > its pHPZC, and vice versa. In this
case, the adsorption of cationic adsorbate onto the negative surface of CAC through electrostatic
10
Second, according to the literature [33], the dissociation constant (pKa) of paracetamol—
similar (i.e., 11.08 and 10.96, respectively). Therefore, PRC exists as an uncharged species (or
molecular form) under pHsolution from 2.0 to 10 (Figure 4b). As a result, the electrostatic attraction
was ruled out in this study. An analogous conclusion was reported in the literature [26, 34].
Because the adsorption process did not involve the electrostatic attraction, the adsorption
efficiency of PRC onto CAC was less affected by the change of pHsolution (Figure 5a) and ionic
of
strength (Figure 5b). Previous studies demonstrated that the process of paracetamol removal from
ro
water by various porous carbonaceous materials was insignificantly dependent on the parameters of
pHsolution and ionic strength, such as activated carbons derived from spent tea leaves [35],
-p
re
As expected, CAC possessed an excellent ability of PRC adsorption under different values of
pHsolution and concentrations of NaCl. Therefore, its outstanding adsorption capacity is expectedly
lP
maintained under real application in removing PRC pollutant from water and wastewater streams.
The contact time-dependent study was conducted from 2 min to 4360 min at ~517 mg/L.
Figure 6 shows that the adsorption process occurred rapidly with approximately 52% PRC in
ur
solution removed within 5 min of contact. The result suggested that CAC had a high affinity to PRC
Jo
in solution. In this study, we applied three adsorption kinetic models to describe the intrinsic
In the literatures, some authors modelled the experimental data of time-dependent adsorption
with the selective models by two different ways: with and without the origin point (0;0). The results
11
of modelling with the contact time from 0 to 4360 min (with the origin point) and from 2 to 4360
min (without the origin point) are shown in Figure 6a and 6b, respectively. The relevant parameters
of the secretive models are listed in Table 2. Clearly, the values of the parameters obtained from
two modelling ways were the same. However, the values of adjust coefficient of determination (adj-
R2) and chi-square statistics (χ2) are remarkably different. The modelling way with the origin point
indicated a higher adj-R2 and lower χ2 values than that without the origin, suggesting that the
important role of the original point in modelling the calcinating the parameters of the adsorption
of
kinetic models.
ro
Furthermore, the adsorption process of organic adsorbate onto porous CAC material often
takes a longer time to reach a true adsorption equilibrium. In contrast, some authors assumed that
-p
the adsorption reached an equilibrium at a shorter time. Therefore, the modelling trials of
experimental data from 0 to 360 min (with the origin point) and from 2 to 360 min (without the
re
origin point) were explored. The results are shown in Figure 6c and 6d and Table 2. A similar
lP
phenomenon was repeated that (1) there was not difference on the values of the parameters of the
kinetic models from two modelling ways and (2) the origin point played an important role in
obtaining a high adj-R2 and low χ2 values. However, a shorter contact time is not enough to correctly
na
modelling whole the adsorption process. Therefore, the parameters of the kinetic models obtained
According to the values of adj-R2, χ2, SD, BIC, and MPSD, it can be concluded that the
Jo
experimental data of adsorption kinetics were better described by the Avrami model (0.997, 13.2,
3.51, 46.2, and 26.4) and Elovich model (0.991, 36.3, 6.17, 65.4, and 45.5) than the pseudo-second
order model (0.864, 521, 22.8, 110, and 239) and pseudo-first-order model (0.758, 929, 30.5, 120,
and 291, respectively). The initial rate constant α of the Elovich model was up to 4039 mg/(g×min).
This is well consistent with the experimental data in Figure 6a, suggesting that CAC exhibited a
12
strong affinity to PRC molecules in the solution. The half-life (t1/2; calculated based on from the
Avrami model) of the adsorption process was 17.79 min. Further experiments were conducted at 48
The isotherm of PRC adsorption onto CAC is provided in Figure 7. As expected, the isotherm
curve is concave downward that can be classified as an H-type shape [37]. The result suggested that
CAC exhibited high affinity of PRC adsorption even at low initial PRC concentrations. In this study,
of
some adsorption isotherm models were applied to model the data of equilibrium adsorption. More
ro
information on such selected models are provided in Section S3 (Supporting Information).
Table 3 provides the modeling result and the calculated parameters of such models. The
-p
results indicated that the experimental data of adsorption equilibrium were reasonably described by
re
the selective models (R2 >0.94), with the fitting model following the decreasing order: Redlich–
Peterson > Langmuir-Freundlich > Freundlich > Langmuir models. The Redlich–Peterson model
lP
was the best fitting model because of the highest Adj-R2 (0.993) value and the lowest others (i.e., χ2
= 43.3, SD = 7.33, BIC = 54.5, and MPSD = 59.8). However, this model does not provide the
na
information on the maximum adsorption capacity of an adsorbent. Although the QLF parameter in
not commonly used and reported ion the literature compared to the Qomax. Therefore, the Qomax
parameter in the Langmuir model was used for comparing the adsorption capacity of CAC and other
Jo
Table S1 lists the Qomax value of CAC and other kinds of CAC in the literature. The
comparative result indicated that CAC (Qomax = 221 mg/g) exhibited a higher adsorption capacity to
PRC compared to some other kinds of CAC (45.4–204 mg/g), but lower than some others (255–322
13
mg/g). Moreover, there was not any relationship between SBET and Qomax (R² = 0.381; n = 17; Table
S1) as well as VTotal and Qomax (R² = 0.320; n = 15; Table S1). This means that an adsorbent with
higher SBET and VTotal value might not means that it will exhibit a higher adsorption capacity of PRC.
In this case, the pore size distribution of adsorbent (i.e., Figure 1c) often plays a determining role in
the amount of organic contaminant (i.e., PRC) adsorbed by a porous adsorbent (i.e., CAC) through
pore-filling mechanism.
of
The effect of different solution temperatures (10 °C, 25 °C, and 50 °C) on the process of PRC
ro
adsorption onto CAC is presented Figure S2. The parameters of adsorption thermodynamics (∆G°,
∆H°, and ∆S°) are calculated based on the laws of thermodynamics using the van’t Hoff equation.
G G RT ln K Equilibriu
o
m
-p (4)
re
When an adsorption process reaches equilibrium, the free energy change (∆G) is nearly zero.
lP
Equation 4 becomes Equation 5 that has been commonly applied to calculate the standard Gibbs
G RT ln K Equilibriu
o
m
(5)
The relationship between ∆G° and the others (∆H° and ∆S°) of an adsorption process is express as
ur
follows:
Jo
It is assumed that the change of ΔS° and ΔH° with the temperature is negligible. After
substitution Equation 5 into 6, the well-known van’t Hoff equation [32, 38, 39] is yield as follows:
14
H 1 S (7)
LnK Equilibrium
R T R
where KL (L/mol) is the Langmuir constant, C° is the selected standard state of adsorbate (C° = 1
mol/L); R is the universal gas constant (0.00831 kJ/mol×K); T is the absolute temperature in degrees
Kelvin (K); and γ (dimensionless) is the activity coefficient of PRC in solution. As provided in
of
Figure 4b, PRC neutral adsorbate under solution pH of 7.0, so its activity coefficient in this study is
ro
expressed as 1.0 [38, 39].
In this study, the Langmuir constant KL (L/mmol; Figure S2) was applied to calculate the
-p
thermodynamics parameters. To consist with physical meanings, the units of adsorption isotherm
re
(the plot of qe vs. Ce) needs to be presented as mol/kg for qe and mol/L for Ce. Equation 8 describes
o
the relationship between KL (with unit) and K Equilibrium (without unit) [32, 38, 40]. Table S2
lP
summaries the calculating performance of the parameters of thermodynamics for whole the
adsorption process (10 °C, 25 °C, and 50° C). The negative values of standard Gibbs free energy (–
na
∆G°) suggest that the process of PRC adsorption onto CAC occurred spontaneously without any
further requirements (i.e., energy or heating). An identical conclusion was reported by some other
ur
scholars for the adsorption of PRC onto porous carbonaceous materials [35, 36, 41]. Because the R2
Jo
value (R2 = 0.271) of the van’t Hoff equation is too low, the ∆H° and ΔS° values cannot be
adequately calculated from such equation. In addition, the adsorption capacity of PRC by CAC and
o
the equilibrium constant K Equilibrium did not followed the order of temperatures, such as 10 °C < 25
°C < 50 °C. Therefore, in this study, the thermodynamics parameters were calculated at two
processes: one at 10 °C and 25 °C (283–298 K) and another at 25 °C and 50 °C (298–323 K). The
15
∆H° value was calculated as the following equation, and the ∆S° value was obtained from Equation
9 [39]. The ∆G° and ∆S° values were calculated from Equations 5 and 6, respectively.
T T KT (9)
H R( 2 1 ) ln( 2 )
T2 T1 KT1
where T2 and T1 are at temperature at two operation temperatures (T2 > T1); and KT1 and KT2 are the
o
equilibrium constant ( K Equilibrium ) at T2 and T1, respectively.
of
The parameters of ∆H° and ∆S° calculated at 283–298 K and 298–323 K are provided in
ro
Table 4. Clearly, an increase in solution temperature from 10 °C to 25 °C, the maximum adsorption
capacity of CAC significantly increased from 1.18 mol/kg to 1.59 mol/kg (Figure S2). This is
-p
because the phenomenon of enlargement pore network of CAC within the increasing temperature
re
simulate to increase the amount of PRC adsorbed onto CAC through common pore-filling
mechanism [42, 43]. Therefore, it can be primarily performed that the important contribution of
lP
pore-filing and van der Waals in the adsorption mechanism. The result is well consistent with the
ΔH° value. A positive value of the standard adsorption enthalpy (ΔH° = 24.1 kJ/mol) confirms that
na
the adsorption was endothermic reaction. However, a further increasing in temperature to 50 °C, the
Qomax value decreased to 1.48 mol/kg (Figure S2). A decrease in adsorption capacity at a higher
ur
temperature might be caused by two feasible reasons that (1) a continuing increase in enlarging pore
network of CAC was not favorable to adsorbed PRC molecules [17, 42] and (2) desorption
Jo
phenomenon resulted from the weak adsorption interaction during the adsorption process (i.e., van
der Waals) [17, 32]. Because the contribution of weak van der Waals in the adsorption process, the
standard adsorption enthalpy change was negative and low magnitude (ΔH° = –6.39 kJ/mol) within
16
In essences, the entropy change of adsorption is generally negative because the randomness of
the adsorbates after adsorption would be decreased compared to those in the bulk liquid phase.
However, the positive value of ΔS° (Table 4) might be because of the fact that the adsorption occurs
in two consecutive steps: desorption of the adsorbed water and adsorption of the PRC adsorbate.
Similarly, Lima et al. [38] found that the positive value of ΔS° [39 and 40 J/(mol×K)] was obtained
for the PRC adsorption process onto two AC samples prepared from Brazil nutshells.
of
The result of desorption efficiency (ɳDes) by the chemical method is showed in Table 5. The
ro
best desorption efficiency was found by using ethanol (53.8%) and methanol (50.1%) a desorbing
agent. In addition, NaOH (0.2 M) can desorbed around 32.5% of PRC from laden CAC. Similarly,
-p
Lima et al. [38] found that the ɳDes value was 66.8% (using 0.1 M NaOH) and 69.9% (using 0.1 M
re
NaOH + 5% ethanol). The results the important contribution of surface interaction (i.e., hydrogen
bonding and n-π interaction) during the adsorption process. Furthermore, the PRC adsorbed by CAC
lP
can be desorbed by deionized distilled water (pH 12) and (pH 2.0), with ɳDes being 18.1% and
11.8%, respectively. The result might result from the minor contribution of weak van der Waals
na
However, the chemical method was less efficient to desorb PRC from laden CAC (Table 5).
ur
According to other studies [44], PRC in the laden CAC sample can be thermally degrade when the
carbonization temperature is higher than 400 °C (determined by thermal analysis). Therefore, the
Jo
thermal method was applied. The carbonization temperature of 600 °C was used to completely
thermally degrade of PRC adsorbed on the surface and in the internal pore network of CAC. Figure
8 shows that after each adsorption/desorption cycle, the adsorption capacity of CAC significantly
increased. A similar finding was reported by Tomul [32] for the adsorption of naproxen by biochar.
The increasing adsorption capacity after each cycle might be because the proper contribution of π-π
17
interaction in the adsorption mechanism of PRC onto 600 °C-carbonized CAC. Similarly, some
authors also found that the 900 ℃-carbonized GAC (Qomax = 197 and 175 mg/g) exhibited an
excellent adsorption capacity towards the aromatic pollutants (i.e., phenol and methylene blue) than
pristine GAC did (156 and 131 mg/g), respectively [45]. A similar conclusion was reported by [46]
for the adsorption of chloramphenicol onto CAC (Qomax = 146 mg/g) and 400 ℃-treated CAC (226
mg/g), and 800 ℃-treated CAC (343 mg/g). However, the reuse of laden CAC by thermal method
has a big problem. This is because the remarkable mass loss of CAC after each cycle (data not
of
showed).
ro
3.2.6. Discussion on adsorption mechanism
In essence, the adsorption process of polar aromatic pollutants (i.e., PRC and dye) in solution
-p
onto porous carbonaceous materials (i.e., biochar and activated carbon) often combines many
re
interactions (or mechanisms) such as electrostatic attraction, van der Waals force, ion exchange,
hydrogen bonding formation, n-π interaction, π-π interaction, and pore-filling [18, 47-50]. However,
lP
the primary adsorption mechanism is often strongly dependent on the conditions of adsorption study
(i.e., solution pH, temperature, and initial adsorbate concentration), adsorbate characteristics (i.e.,
na
molecular size, solubility, pKa, and electron distribution), and adsorbent properties (i.e., surface
area, pore size distribution, and surface functional groups). Notably, Tran et al. [51] investigated the
ur
adsorption of methylene green 5 (MG5) onto porous carbonaceous biochar. They found that
although the removal efficiency of around 99% was reached at an initial low MG5 concentration (Co
Jo
≈ 50 mg/L), the mechanism of MG5 adsorption onto biochar through pore-filling was negligible
compared to the others. This is because the textural parameters (SBET and VTotal) of biochar before
(565 m2/g and 0.236 cm3/g) and after (529 m2/g and 0.264 cm3/g) adsorption were slightly different,
respectively. However, the pore-filling mechanism dominated others when an initial MG5
18
concentration was higher than ~300 mg/L because of a remarkable decease in the SBET and VTotal
In this study, the mechanism of PRC adsorption by CAC was discussed at solution pH 7.0, 25
°C, and 1200 mg/L. As aforementioned in Section 3.2.1, electrostatic attraction was ruled out
because PRC (pKa ≈ 11; Figure 4b) was not ionized under solution pH 7.0; a similar assumption for
ion exchange. In addition, the presence of van der Waals force in the adsorption process was certain,
but this weak force was acknowledged as a less important contributor in the adsorption mechanism
of
(see Section 3.2.4).
ro
The n-π interaction (also known as “donor-acceptor complex mechanism”) was firstly
proposed by Mattson and coworkers [48] for the adsorption of phenol and nitrophenols onto lignin-
-p
based AC. In such mechanism, the carbonyl oxygen in the surface of AC acted as an electron donor,
while the aromatic rings of phenol acted as an electron acceptor. The n-π electron interaction has
re
been reported as an integral adsorption mechanism of various aromatic pollutants onto carbonaceous
lP
materials [18, 49, 52, 53]. In this interaction, the carbonyl oxygen ions on the biochar’s surface
serve as electron donors, while the aromatic ring of PRC acts as an electron acceptor. Figure 2
na
indicated that after adsorption of CAC, the intensity of relevant band at around 1630 cm-1 (the C=O
groups) sustainably decreased. The result suggested that there was a great contribution of n-π
ur
Hydrogen bonding interaction can be categorized into two types [18]. First, dipole−dipole H-
Jo
bonding interaction often occurs between the surface hydrogens of the hydroxyl groups (H-donor)
on the surface of carbonaceous material and the appropriate atoms (i.e., oxygen and nitrogen and
oxygen; H-acceptor) of aromatic adsorbate [54]. Second, Yoshida H-bonding interaction often
occurs between the hydroxyl groups on the surface of carbonaceous material and the aromatic ring
in the adsorbate molecules [55]. Figure 2 indicates that the band intensity of the C–O and –OH
19
groups slightly decreased after adsorption, suggesting that the existence of dipole−dipole H-bonding
and Yoshida H-bonding interactions, respectively. Although the H-bonding interaction played an
integral role in the adsorption process, it did not dominant. A similar conclusion was reported
elsewhere [26].
Regarding π-π interaction (or π-π electron donor-acceptor interaction), such a mechanism that
was initially reported by Coughlin and Ezra [56] occurs between the π-electrons of carbonaceous
material (donor) and the π-electrons of the aromatic rings of adsorbate (acceptor). The π-π
of
interaction has been identified as a primary contributor for the adsorption mechanism for various
ro
aromatic pollutants onto various kinds of carbonaceous materials [18, 49, 50, 53, 57, 58]. Figure 3
shows that after adsorption of PRC, the ID/IG ratio of AC decreased from 2.011 to 1.947, suggesting
-p
the presence of π-π interaction during the adsorption process. Similarly, some other authors applied
the Raman technique to identify the molecular charge-transfer of organic adsorbates onto carbon-
re
based material [32, 59, 60]. They also found that the ID/IG ratio of CAC deceased after adsorption.
lP
Finally, Table 1 shows the change of textural properties of CAC after adsorption. Clearly,
the SBET and VTotal values of CAC dramatically decreased by approximately 96% after adsorption.
na
For example, the SBET value decreased from 1248 m2/g to 45.6 m2/g after adsorption, suggesting the
most important role of pore-filling compared to the other mechanisms in the adsorption process [18,
ur
32]. As expected, CAC exhibited a dominant mesoporous; therefore, it can effectively adsorb PRC
onto its internal porosity network. The result was supported by the un-change of porosity properties
Jo
of CAC before adsorption (Figure 1c) and after adsorption (Figure 1d).
To sum up, the adsorption mechanisms of PRC in aqueous solution onto CAC was
adsorption mechanism might be expressed as following: pore-filling > n-π interaction > π-π
Several kinds of water matrix samples were collected to investigate the removal process of
PRC from water environments. They include distilled water (denoted as a comparative sample), tap
water, groundwater, wastewater (from beauty salon), coastal water, and wastewater (from water
treatment plant). As expected, the applied CAC possessed an excellent capacity of PRC adsorption
in different real water samples (Figure 9). The selected water samples possessed different solution
pH, ionic strength salts, and composition (Table S3). The sweater possesses high salt content and
of
ionic strength. The groundwater could possess the lower pH and higher salt content. The little
ro
organic substance exists in the beauty salon beauty salon and wastewater. However, the qe values
ranging from 192 mg/g to 216 mg/g (average: 203 mg/g, standard deviation: 9.09, standard error:
-p
0.64). The result indicated that adsorption capacity of CAC was less-effected by the characteristic’s
parameters of water. Therefore, CAC is expected to have a great potential application for
re
eliminating PRC micropollutant from water and wastewater streams.
lP
4. Conclusions
The oxidized carbonaceous porous material CAC with well-developed porosity (SBET = 1284
na
m2/g and VTotal = 0.680 cm3/g) and high oxygen-containing groups exhibited high adsorption
capacity towards PRC in solution. The Langmuir maximum adsorption capacity of CAC was
ur
221 milligram of PRC per gram of CAC. CAC was a dominant mesoporous material
(mesopore volume accounting for 76.3%). The pore size distribution of CAC was highly
Jo
feasible to adequately adsorb PRC in solution through pore-filling. This conclusion was
confirmed by the unchanged of pore network after adsorption. The pore-filling played
extremely important role in the adsorption mechanism of PRC, which was confirmed by a
remarkable decrease in SBET and VTotal after adsorption to 45.6 m2/g and 0.039 cm3/g,
respectively. The second important contribution in the adsorption process of PRC onto
21
oxidized CAC (abundant oxygen-containing groups) was n-π interaction. Such interaction
was identified through the FTIR technique. A remarkably decrease in the band intensity at
1630 cm-1 (the C=O group) after adsorption supported this conclusion. The presence of π-π
interaction was established (1) a significant decrease in the ID/IG ratio from 2.011 to 1.947
after adsorption and (2) a significant increase in the adsorption capacity of CAC after each
cycle of adsorption and desorption using the thermal method. Hydrogen bonding formations
and weak van der Waals force played a less-important role in the adsorption mechanism.
of
The Langmuir maximum adsorption capacity of CAC to PRC was 221 mg/g. The excellent
adsorption capacity of CAC was less dependent on some operation parameters (solution pH,
ro
ionic strength, and water matrix).
-p
Oxidized CAC can serve as a promising adsorbent for real water treatment.
Tomul, Andrei Ivanets, Seung Han Woo, Ahmad Hosseini-Bandegharaei, Van Phuong Nguyen,
lP
Declaration of interests
The authors declare that they have no known competing financial interests or personal
ur
relationships that could have appeared to influence the work reported in this paper.
Jo
Acknowledgement
This research is funded by Vietnam National Foundation for Science and Technology Development
of
[4] O.M. Rodriguez-Narvaez, J.M. Peralta-Hernandez, A. Goonetilleke, E.R. Bandala, Treatment
technologies for emerging contaminants in water: A review, Chemical Engineering Journal, 323
ro
(2017) 361-380.
[5] A. Gogoi, P. Mazumder, V.K. Tyagi, G.G. Tushara Chaminda, A.K. An, M. Kumar, Occurrence
-p
and fate of emerging contaminants in water environment: A review, Groundwater for Sustainable
Development, 6 (2018) 169-180.
re
[6] A. Solanki, T.H. Boyer, Physical-chemical interactions between pharmaceuticals and biochar in
synthetic and real urine, Chemosphere, 218 (2019) 818-826.
lP
[7] D. Dionisi, C.C. Etteh, Effect of process conditions on the aerobic biodegradation of phenol and
paracetamol by open mixed microbial cultures, Journal of Environmental Chemical Engineering, 7
(2019) 103282.
na
[8] S. Bahmanzadeh, M. Noroozifar, Fabrication of modified carbon paste electrodes with Ni-doped
Lewatit FO36 nano ion exchange resin for simultaneous determination of epinephrine, paracetamol
and tryptophan, Journal of Electroanalytical Chemistry, 809 (2018) 153-162.
ur
[9] F. Audino, L.O. Conte, A.V. Schenone, M. Pérez-Moya, M. Graells, O.M. Alfano, A kinetic
study for the Fenton and photo-Fenton paracetamol degradation in an annular photoreactor,
Jo
23
[12] Y. He, X. Wang, W. Huang, R. Chen, W. Zhang, H. Li, H. Lin, Hydrophobic networked PbO2
electrode for electrochemical oxidation of paracetamol drug and degradation mechanism kinetics,
Chemosphere, 193 (2018) 89-99.
[13] A. Ziylan-Yavaş, N.H. Ince, Catalytic ozonation of paracetamol using commercial and Pt-
supported nanocomposites of Al2O3: The impact of ultrasound, Ultrasonics Sonochemistry, 40
(2018) 175-182.
[14] N. Villota, J.I. Lombraña, A. Cruz-Alcalde, M. Marcé, S. Esplugas, Kinetic study of colored
species formation during paracetamol removal from water in a semicontinuous ozonation contactor,
Science of The Total Environment, 649 (2019) 1434-1442.
of
[15] W.-C. Yun, K.-Y.A. Lin, W.-C. Tong, Y.-F. Lin, Y. Du, Enhanced degradation of paracetamol
in water using sulfate radical-based advanced oxidation processes catalyzed by 3-dimensional
ro
Co3O4 nanoflower, Chemical Engineering Journal, 373 (2019) 1329-1337.
[16] C. Sophia A, E.C. Lima, Removal of emerging contaminants from the environment by
-p
adsorption, Ecotoxicology and Environmental Safety, 150 (2018) 1-17.
[17] H.N. Tran, F. Tomul, H.T.H. Nguyen, D.T. Nguyen, E.C. Lima, G.T. Le, C.-T. Chang, V.
Masindi, S.H. Woo, Innovative spherical biochar for pharmaceutical removal from water: Insight
re
into adsorption mechanism, Journal of Hazardous Materials, 394 (2020) 122255.
[18] H.N. Tran, S.-J. You, H.-P. Chao, Fast and efficient adsorption of methylene green 5 on
lP
activated carbon prepared from new chemical activation method, Journal of Environmental
Management, 188 (2017) 322-336.
[19] I. Tiffour, A. Dehbi, A.-H.I. Mourad, A. Belfedal, Synthesis and characterization of a new
na
organic semiconductor material, Materials Chemistry and Physics, 178 (2016) 49-56.
[20] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso, J. Rouquerol,
ur
K.S. Sing, Physisorption of gases, with special reference to the evaluation of surface area and pore
size distribution (IUPAC Technical Report), Pure and Applied Chemistry, 87 (2015) 1051-1069.
[21] F.-C. Huang, C.-K. Lee, Y.-L. Han, W.-C. Chao, H.-P. Chao, Preparation of activated carbon
Jo
using micro-nano carbon spheres through chemical activation, Journal of the Taiwan Institute of
Chemical Engineers, 45 (2014) 2805-2812.
[22] K.C. Bedin, A.L. Cazetta, I.P.A.F. Souza, O. Pezoti, L.S. Souza, P.S.C. Souza, J.T.C.
Yokoyama, V.C. Almeida, Porosity enhancement of spherical activated carbon: Influence and
optimization of hydrothermal synthesis conditions using response surface methodology, Journal of
Environmental Chemical Engineering, 6 (2018) 991-999.
24
[23] L. Yu, C. Falco, J. Weber, R.J. White, J.Y. Howe, M.-M. Titirici, Carbohydrate-derived
hydrothermal carbons: A thorough characterization study, Langmuir, 28 (2012) 12373-12383.
[24] Q.A. Binh, P. Kajitvichyanukul, Adsorption mechanism of dichlorvos onto coconut fibre
biochar: The significant dependence of H-bonding and the pore-filling mechanism, Water Science
and Technology, 79 (2018) 866–876.
[25] D.D. Sewu, H. Jung, S.S. Kim, D.S. Lee, S.H. Woo, Decolorization of cationic and anionic
dye-laden wastewater by steam-activated biochar produced at an industrial-scale from spent
mushroom substrate, Bioresource Technology, 277 (2019) 77-86.
[26] I. Villaescusa, N. Fiol, J. Poch, A. Bianchi, C. Bazzicalupi, Mechanism of paracetamol removal
of
by vegetable wastes: The contribution of π–π interactions, hydrogen bonding and hydrophobic
effect, Desalination, 270 (2011) 135-142.
ro
[27] J. Hodkiewicz, T. Scientific, Characterizing carbon materials with Raman spectroscopy,
Application note, 51946 (2010).
-p
[28] A. Fuertes, M.C. Arbestain, M. Sevilla, J.A. Maciá-Agulló, S. Fiol, R. López, R. Smernik, W.
Aitkenhead, F. Arce, F. Macìas, Chemical and structural properties of carbonaceous products
obtained by pyrolysis and hydrothermal carbonisation of corn stover, Soil Research, 48 (2010) 618-
re
626.
[29] Y. Chen, G. Zhang, J. Zhang, H. Guo, X. Feng, Y. Chen, Synthesis of porous carbon spheres
lP
derived from lignin through a facile method for high performance supercapacitors, Journal of
Materials Science & Technology, 34 (2018) 2189-2196.
[30] R. Maal-Bared, What we know about coronavirus and water treatment,
na
https://www.wef.org/wef-waterblog/wef-waterblog/what-we-know-about-coronavirus-and-water-
treatment/, (2020).
ur
[31] H. Chen, Y.-c. Guo, F. Wang, G. Wang, P.-r. Qi, X.-h. Guo, B. Dai, F. Yu, An activated carbon
derived from tobacco waste for use as a supercapacitor electrode material, New Carbon Materials,
32 (2017) 592-599.
Jo
[32] F. Tomul, Y. Arslan, B. Kabak, D. Trak, E. Kendüzler, E.C. Lima, H.N. Tran, Peanut shells-
derived biochars prepared from different carbonization processes: Comparison of characterization
and mechanism of naproxen adsorption in water, Science of The Total Environment, 726 (2020)
137828.
25
[33] M.A. Elbagerma, G. Azimi, H.G.M. Edwards, A.I. Alajtal, I.J. Scowen, In situ monitoring of
pH titration by Raman spectroscopy, Spectrochimica Acta Part A: Molecular and Biomolecular
Spectroscopy, 75 (2010) 1403-1410.
[34] H. Nourmoradi, K.F. Moghadam, A. Jafari, B. Kamarehie, Removal of acetaminophen and
ibuprofen from aqueous solutions by activated carbon derived from Quercus Brantii (Oak) acorn as
a low-cost biosorbent, Journal of Environmental Chemical Engineering, 6 (2018) 6807-6815.
[35] S. Wong, Y. Lim, N. Ngadi, R. Mat, O. Hassan, I.M. Inuwa, N.B. Mohamed, J.H. Low,
Removal of acetaminophen by activated carbon synthesized from spent tea leaves: equilibrium,
kinetics and thermodynamics studies, Powder Technology, 338 (2018) 878-886.
of
[36] A. Mashayekh-Salehi, G. Moussavi, Removal of acetaminophen from the contaminated water
using adsorption onto carbon activated with NH4Cl, Desalination and Water Treatment, 57 (2016)
ro
12861-12873.
[37] C.H. Giles, D. Smith, A. Huitson, A general treatment and classification of the solute
-p
adsorption isotherm. I. Theoretical, Journal of colloid and interface science, 47 (1974) 755-765.
[38] D.R. Lima, A. Hosseini-Bandegharaei, P.S. Thue, E.C. Lima, Y.R.T. de Albuquerque, G.S. dos
Reis, C.S. Umpierres, S.L.P. Dias, H.N. Tran, Efficient acetaminophen removal from water and
re
hospital effluents treatment by activated carbons derived from Brazil nutshells, Colloids and
Surfaces A: Physicochemical and Engineering Aspects, 583 (2019) 123966.
lP
[39] G.G. Hammes, Thermodynamics and kinetics for the biological sciences, New York: Wiley-
Interscience, 2000.
[40] J.-C. Bollinger, Letter to the Editor: Comments on “Adsorption of methylene blue and Cd(II)
na
onto maleylated modified hydrochar from water”, Environmental Pollution, 261 (2019) 113824.
[41] L.A. Al-Khateeb, S. Almotiry, M.A. Salam, Adsorption of pharmaceutical pollutants onto
ur
26
[44] I. Cabrita, B. Ruiz, A.S. Mestre, I.M. Fonseca, A.P. Carvalho, C.O. Ania, Removal of an
analgesic using activated carbons prepared from urban and industrial residues, Chemical
Engineering Journal, 163 (2010) 249-255.
[45] C.A. Sáenz-Alanís, R.B. García-Reyes, E. Soto-Regalado, A. García-González, Phenol and
methylene blue adsorption on heat-treated activated carbon: Characterization, kinetics, and
equilibrium studies, Adsorption Science & Technology, 35 (2017) 789-805.
[46] J. Lach, Adsorption of chloramphenicol on commercial and modified activated carbons, Water,
11 (2019) 1141.
[47] P. Boakye, H.N. Tran, D.S. Lee, S.H. Woo, Effect of water washing pretreatment on property
of
and adsorption capacity of macroalgae-derived biochar, Journal of Environmental Management, 233
(2019) 165-174.
ro
[48] J.A. Mattson, H.B. Mark, M.D. Malbin, W.J. Weber, J.C. Crittenden, Surface chemistry of
active carbon: Specific adsorption of phenols, Journal of Colloid and Interface Science, 31 (1969)
-p
116-130.
[49] Q.-S. Liu, T. Zheng, P. Wang, J.-P. Jiang, N. Li, Adsorption isotherm, kinetic and mechanism
studies of some substituted phenols on activated carbon fibers, Chemical Engineering Journal, 157
re
(2010) 348-356.
[50] H. Liu, W. Ning, P. Cheng, J. Zhang, Y. Wang, C. Zhang, Evaluation of animal hairs-based
lP
activated carbon for sorption of norfloxacin and acetaminophen by comparing with cattail fiber-
based activated carbon, Journal of Analytical and Applied Pyrolysis, 101 (2013) 156-165.
[51] H.N. Tran, Y. Wang, S. You, H. Chao, Sustainable biochar derived from agricultural wastes for
na
and Agricultural Solid Wastes Adsorbents (1st Edition). Boca Raton: CRC Press., (2017) 255-291.
[52] L. Yanyan, T.A. Kurniawan, M. Zhu, T. Ouyang, R. Avtar, M.H. Dzarfan Othman, B.T.
Mohammad, A.B. Albadarin, Removal of acetaminophen from synthetic wastewater in a fixed-bed
Jo
column adsorption using low-cost coconut shell waste pretreated with NaOH, HNO3, ozone, and/or
chitosan, Journal of Environmental Management, 226 (2018) 365-376.
[53] M. Turk Sekulic, N. Boskovic, A. Slavkovic, J. Garunovic, S. Kolakovic, S. Pap, Surface
functionalised adsorbent for emerging pharmaceutical removal: Adsorption performance and
mechanisms, Process Safety and Environmental Protection, 125 (2019) 50-63.
27
[54] R.S. Blackburn, Natural pplysaccharides and their interactions with dye molecules:
Applications in effluent treatment, Environmental Science & Technology, 38 (2004) 4905-4909.
[55] H.L. Parker, A.J. Hunt, V.L. Budarin, P.S. Shuttleworth, K.L. Miller, J.H. Clark, The
importance of being porous: polysaccharide-derived mesoporous materials for use in dye adsorption,
RSC Advances, 2 (2012) 8992-8997.
[56] R.W. Coughlin, F.S. Ezra, Role of surface acidity in the adsorption of organic pollutants on the
surface of carbon, Environmental Science & Technology, 2 (1968) 291-297.
[57] B. Peng, L. Chen, C. Que, K. Yang, F. Deng, X. Deng, G. Shi, G. Xu, M. Wu, Adsorption of
antibiotics on graphene and biochar in aqueous solutions induced by π-π interactions, Scientific
of
Reports, 6 (2016) 31920.
[58] K. Yang, Y. Jiang, J. Yang, D. Lin, Correlations and adsorption mechanisms of aromatic
ro
compounds on biochars produced from various biomass at 700 °C, Environmental Pollution, 233
(2018) 64-70.
-p
[59] R. Voggu, C.S. Rout, A.D. Franklin, T.S. Fisher, C.N.R. Rao, Extraordinary Sensitivity of the
Electronic Structure and Properties of Single-Walled Carbon Nanotubes to Molecular Charge-
Transfer, The Journal of Physical Chemistry C, 112 (2008) 13053-13056.
re
[60] G.K. Ramesha, A. Vijaya Kumara, H.B. Muralidhara, S. Sampath, Graphene and graphene
oxide as effective adsorbents toward anionic and cationic dyes, Journal of Colloid and Interface
lP
28
of
ro
-p
re
Figure 1. Nitrogen adsorption/desorption isotherm of CAC before and after paracetamol adsorption
lP
na
ur
Jo
29
of
ro
-p
re
lP
30
of
ro
-p Figure
re
3. Raman spectrum of CAC before and after adsorption of paracetamol
lP
na
ur
Jo
31
of
ro
-p
re
lP
na
ur
Jo
Figure 4. (a) pH point of zero charge (pHPZC) of CAC; and (b) paracetamol distribution in
different solutions pH
32
of
ro
-p
re
lP
na
ur
Figure 5. Effect of solution pH (a) and ionic strength (b) on the amount of PRC adsorbed onto CAC
Jo
33
of
ro
-p
re
Figure 6. Effect of contact time on the adsorption process and some modelling ways of the
experimental data
lP
na
ur
Jo
34
of
ro
-p
re
Figure
7. Adsorption isotherm of PRC onto CAC at 25 °C and the modelling of different adsorption
lP
isotherms
na
ur
Jo
35
Figure 8. Feasible reuse of PRC-laden CAC through the cycle of adsorption-desorption
of
ro
-p
re
lP
na
ur
Jo
36
of
ro
-p
Figure 9. Comparison of adsorption efficiency of CAC towards different PRC solutions. The
re
abbreviation of water samples: M1 and M2 (coastal water), M3 and M4 (wastewater) from
water treatment plant), M5 and M6 (groundwater) collected nearly industrial zone, M7
(wastewater) from beauty salon, M8 (tap water from laboratory), and M9 (distilled water)
lP
Before After
Unit
adsorption adsorption
1. Surface area
ur
2. Pore volume
37
Mesopore volume VMeso cm3/g 0.518 0.0341
3. Pore width
of
ro
-p
re
lP
na
ur
Jo
38
Table 2. Corresponding parameters of the selective kinetic models of adsorption PRC onto CAC
of
SD — 6.17 — 6.39 — 4.46 — 21.6 —
BIC — 65.4 — 62.7 — 41.8 — 82.8 —
MPSD — 45.5 — 47.1 — 33.3 — 258 —
ro
2. PSO model
qe mg/g 214 7.19 214 7.44 196 6.19 196 6.49
3
k2 (×10 ) g/(mg×min) 1.18 4.1E-01 1.18 4.2E-01 2.48 7.9E-04 2.48 8.3E-04
-p
Adj-R 2
— 0.864 — 0.691 — 0.92 — 0.766 —
χ2 — 521 — 559 — 292 — 321 —
SD — 22.8 — 23.6 — 21.6 — 22.6 —
re
BIC — 110 — 105 — 82.8 — 77.7 —
MPSD — 239 — 248 — 258 — 271 —
3. PFO model
lP
39
Table 3. Corresponding parameters of the selective isotherm models of adsorption PRC onto CAC
Unit Value SE
1. Langmuir model
Qomax mg/g 221 6.31
KL L/mg 0.244 0.057
adj-R2 — 0.946 —
χ2 — 317 —
SD — 20.7 —
BIC — 81.8 —
MPSD — 238 —
of
2. Freundlich model
KF (mg/g)(L/mg)n 90.1 7.15
n — 0.150 0.014
ro
adj-R2 — 0.963 —
χ2 — 219 —
— —
-p
SD 15.5
BIC — 74.1 —
MPSD — 148 —
re
2. Langmuir-Freundlich model
QLF mg/g 260 12.6
KLF L/mg 0.125 0.037
lP
12.6
BIC — 56.1 —
MPSD — 78.2 —
4. Redlich–Peterson model
ur
adj-R2 — 0.993 —
χ2 — 43.3 —
SD — 7.33 —
BIC — 54.5 —
MPSD — 59.8 —
40
Table 4. Thermodynamic parameters (ΔH° and ΔS°) of the PRC adsorption by CAC at two
operation temperatures
∆H° ∆S°
Temperature
[kJ/mol] [J/(mol×K)]
Note: the values of ∆G° (–19.0, –21.3, and –22.5 kJ/mol) at different temperatures (283, 298, and
323 K, respectively) are presented in Table S2.
of
ro
-p
re
lP
na
ur
Jo
41
Table 5. Desorption efficiency of paracetamol from the laden-CAC using different desorbing agents
of
Methanol 50.1
Note: The adsorption experimental conditions: 673 mg/L of PRC, pH 7.0, 25 °C, 2.0 g/L, and 72 h.
ro
-p
re
lP
na
ur
Jo
42