Accepted Manuscript Selective adsorption of phenanthrene dissolved in Tween 80 solution using activated carbon derived from walnut shells Xin Zheng, Heng Lin, Yufang Tao, Hui Zhang PII:
S0045-6535(18)31099-3
DOI:
10.1016/j.chemosphere.2018.06.025
Reference:
CHEM 21557
To appear in:
ECSN
Received Date: 9 January 2018 Revised Date:
23 April 2018
Accepted Date: 3 June 2018
Please cite this article as: Zheng, X., Lin, H., Tao, Y., Zhang, H., Selective adsorption of phenanthrene dissolved in Tween 80 solution using activated carbon derived from walnut shells, Chemosphere (2018), doi: 10.1016/j.chemosphere.2018.06.025. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
Graphical Abstract
ACCEPTED MANUSCRIPT 1
Selective adsorption of phenanthrene dissolved in Tween 80 solution using activated
2
carbon derived from walnut shells
3
Xin Zhenga,b, Heng Lina,b, Yufang Taoa,b, Hui Zhanga,b,*
4
a
5
Remediation Material Engineering Technology Research Center, Wuhan University,
6
Wuhan 430079, China
7
b
SC
RI PT
Department of Environmental Science and Engineering, Hubei Environmental
Shenzhen Research Institute of Wuhan University, Shenzhen 518057, China
9
M AN U
8
* Corresponding author: Tel: + 86-27-68775837; Fax: +86 27 68778893. E-mail:
[email protected]
AC C
EP
TE D
10
1
ACCEPTED MANUSCRIPT 11
Abstract
12
In order to remove phenanthrene (PHE) from surfactant solution, activated carbon
13
(AC)
14
Brunauer-Emmett-Teller (BET), field-emission scanning electron microscopy
15
(FESEM), Fourier transform infrared (FTIR) spectroscopy and X-ray photoelectron
16
spectroscopy (XPS).
17
effectively removed and the latter could be economically recovered after adsorption
18
by the prepared AC.
19
of AC play important roles in the PHE adsorption process.
20
process could best be described using the pseudo-second-order model and adsorption
21
isotherm results indicated that the Langmuir model best fitted the data.
22
thermodynamic parameters, including enthalpy change, Gibbs free energy change and
23
entropy change were calculated.
24
80 recovery reached 95% and 90%, respectively.
25
provided an efficient alternative for selective adsorption of PHE and recovery of
26
Tween 80 after the soil washing processes.
27
regenerated with ethanol and even if AC were regenerated twice PHE removal
28
reached 80%.
from
waste
walnut
shells
and
characterized
by
RI PT
prepared
SC
For solutions containing PHE and Tween 80, the former was
The π-π interactions and oxygen containing functional groups
TE D
M AN U
was
The adsorption kinetics
Adsorption
The results suggest that AC
After adsorption AC could be
AC C
EP
Under optimal conditions, PHE removal and Tween
29 30
Keywords: Activated carbon; Selective adsorption; Polycyclic aromatic hydrocarbons;
31
Tween 80; Recovery
32 2
ACCEPTED MANUSCRIPT 1. Introduction
34
Polycyclic aromatic hydrocarbons (PAHs) which are organic pollutants composed of
35
two or more fused aromatic rings (Gharibzadeh et al., 2016) are always found in
36
groundwater and soils at sites involved in coal processing, coal storage, coke oven
37
plants, and as a result of coal tar spillage (Paria and Yuet, 2006).
38
their hydrophobicity, have a very low solubility in water and are easily adsorbed by
39
clay minerals and organic matter in contaminated soils (Gómez et al., 2010a).
40
Moreover, PAHs are known to be mutagenic and carcinogenic (Chen and Liao, 2006;
41
Liu et al., 2016a) and can be harmful to human health (Gan et al., 2009; Gharibzadeh
42
et al., 2016) which makes remediation of PAH-contaminated soils highly desirable.
RI PT
33
M AN U
SC
PAHs, because of
43
A variety of methods have been used for the remediation of PAH-contaminated soils,
45
including incineration (Acharya and Ives, 1994), thermal desorption (Kuppusamy et
46
al., 2017), chemical oxidation (Usman et al., 2012; Lemaire et al., 2013),
47
bioremediation (Xiong et al., 2017; Yu et al., 2017) and soil washing (Ahn et al.,
48
2008a; Gong et al., 2010; Vizcaíno et al., 2012; Kuppusamy et al., 2017).
49
the processes used thus far, surfactant-enhanced soil washing has proved to be an
50
effective technology for the remediation of soils contaminated by PAHs (Lau et al.,
51
2014; Trellu et al., 2016; Cheng et al., 2017).
52
of PAHs in aqueous solution to a great degree by trapping the targeted hydrophobic
53
molecules into the hydrophobic cores of surfactant micelles (Ahn et al., 2008a;
AC C
EP
TE D
44
3
Among
Surfactants can increase the solubility
ACCEPTED MANUSCRIPT 54
Gharibzadeh et al., 2016).
However, soil washing generates large amounts of
55
effluent containing surfactant and PAHs, which can cause secondary contamination if
56
the effluent is not treated appropriately and the high cost of the surfactant restricts the
57
widespread application of soil washing (Zhou et al., 2013; Li et al., 2014).
58
be more advantageous, environmentally friendly and economical if the surfactant
59
could be recovered and reused after the soil washing process.
60
treatment (Gharibzadeh et al., 2016), solvent extraction (Lee et al., 2002),
61
electrochemical treatment (Gómez et al., 2010b) and adsorption (Ahn et al., 2007;
62
Rosas et al., 2013; Zhou et al., 2013; Li et al., 2014) have all been employed to
63
recover the surfactant from the soil washing effluent.
64
methods, adsorption may have the advantages of lower cost and decreased pollution,
65
higher efficiency and ease of operation (Ahn et al., 2010; Zhou et al., 2013; Liu et al.,
66
2014a).
67
organic wastewater (Yang et al., 2009), which has been used in the recovery of
68
surfactants from a number of PAH-contaminated soil washing effluents.
69
(2007, 2008a, 2008b, 2010) used commercial AC to adsorb PAHs from soil washing
70
effluents and obtained about 91% of phenanthrene (PHE) removal.
71
(2015a) employed a fixed-bed with commercial AC to recover surfactant solutions
72
from soil washing effluents and 80% surfactant retention was attained.
73
(2016b) separated PAHs from rhamnolipid solution using commercial AC and the
74
maximum adsorption capacity for PHE was around 43.9 mg g−1.
RI PT
It would
M AN U
SC
Therefore, biological
TE D
Compared with other recovery
AC C
EP
Additionally, activated carbon (AC) is a common type of adsorbent to treat
4
Ahn et al.
Zhou et al.
Liu et al.
As mentioned
ACCEPTED MANUSCRIPT 75
above, all the AC used in the recovery of surfactant was commercial and the
76
adsorption capacity was limited (up to 58.84 mg g−1).
77
spent AC was not investigated extensively.
Moreover, the regeneration of
RI PT
78 79
Walnut shells are a major agricultural waste with a production of 100,000 ton/year in
80
China alone (Yang and Qiu, 2010).
81
preparation of AC and has been utilized, for example, in the adsorption of dyes (Yang
82
and Qiu, 2010; Alimohammadi et al., 2016; Ashrafi et al., 2017), heavy metals (Yi et
83
al., 2015) and antibiotics (Nazari et al., 2016).
84
the use of walnut shells AC for the adsorption of PAHs from soil washing effluents,
85
and consequently, in this study PHE was selected as a model PAH and Tween 80 as
86
the nonionic surfactant (Zhao et al., 2016).
87
prepare AC, with a high adsorption capacity, from waste walnut shells; (2) to explore
88
the feasibility of selective adsorption of PHE from simulated soil washing effluents;
89
(3) to investigate the factors which influence PHE adsorption; (4) to measure the
90
adsorption kinetics, determine the adsorption isotherms and calculate thermodynamic
91
parameters; (5) to regenerate the spent AC; and (6) to explore the PHE adsorption
92
mechanism.
93
process to remove PHE from the surfactant solution.
94
recovered and could potentially be reused in soil washing, and at the same time
95
greatly reduce the cost of soil washing.
However, there has been no report on
The objectives of this work were: (1) to
AC C
EP
TE D
M AN U
SC
It is an excellent and cheap precursor for the
In summary, the significance of this study was to provide a feasible
5
The surfactant solution is
ACCEPTED MANUSCRIPT 96 2. Materials and methods
98
2.1 Materials
99
Phenanthrene was supplied by Sun Chemical Technology (Shanghai) Co. Ltd and
100
Tween 80, phosphoric acid (H3PO4), hydrochloric acid (HCl), potassium bromide
101
(KBr), sodium hydroxide (NaOH), sodium carbonate (Na2CO3), sodium bicarbonate
102
(NaHCO3), ammonium thiocyanate (NH4SCN), chloroform (CHCl3), cobalt nitrate
103
hexahydrate (Co(NO3)2·6H2O), ethanol (C2H5OH) and acetonitrile (CH3CN) were
104
supplied by Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China).
105
chemicals were used without further purification.
M AN U
SC
RI PT
97
106
All
2.2 Preparation of AC
108
Walnut shells were washed with deionized water, dried at 60 ºC, crushed using a
109
grinder and sieved to thirty mesh.
110
was then impregnated with 30 wt% phosphoric acid at the ratio of 2 g H3PO4
111
solution/g crushed shells for 24 h at room temperature.
112
was then thermally activated under a nitrogen atmosphere in a tube furnace at 700 ºC
113
for 2 h.
114
hydrochloric acid solution for 30 min and then washed with deionized water to neutral
115
pH.
116
sealed in a hermetic vessel for further use.
TE D
107
AC C
EP
This material was dried at 80 ºC overnight and
The impregnated material
After being cooled to room temperature, the material was soaked in 0.1 M
Finally, the AC was dried at 105 ºC for 24 h and sieved (200 mesh) and then
6
ACCEPTED MANUSCRIPT 117 118
2.3 Characterization of AC
119
The morphology of the AC was observed with a Zeiss SIGMA field-emission
120
scanning electron microscopy (FESEM).
121
by nitrogen adsorption at 77 K via the Brunauer-Emmett-Teller (BET) equation on an
122
ASAP 2020 analyzer.
The pHPZC was measured by a batch equilibration technique
123
(Zhou et al., 2015b).
The Fourier transform infrared (FTIR) spectroscopy was
124
performed on a Nicolet 5700 FTIR Spectrometer using KBr pellets.
125
was employed to quantify the acidic groups present on the AC surface (Ge et al.,
126
2015).
127
were determined by X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi,
128
Thermo Fisher).
RI PT
SC
M AN U
Boehm titration
Detailed information regarding surface functionalities present on the AC
TE D
129
The specific surface area was calculated
2.4 Batch sorption experiments
131
A stock of solution was prepared by dissolving PHE in Tween 80 solution.
132
influence factor experiments, 100 mL solution containing PHE (20 mg L−1) and
133
Tween 80 (1-20 g L−1) was applied into the 250 mL conical flask.
134
required amount of AC (0.05-0.6 g L−1), the conical flasks were placed on orbital
135
shakers with a speed of 200 rpm at 25 ± 1 ºC for 24 h, which was the equilibrium time
136
as determined from the sorption kinetics experiments.
137
withdrawn with a syringe and filtered through a 0.45 µm nylon filter for the analysis
AC C
EP
130
7
In the
After adding the
Then the samples were
ACCEPTED MANUSCRIPT 138
of PHE and Tween 80.
The PHE removal and Tween 80 loss were calculated
139
according to the difference between the initial and equilibrium concentrations in
140
aqueous solutions as expressed in Eqs. (1) and (2). PHE removal % =
142
Tween 80 loss % =
[PHE]0
× 100%
[Tween 80]0 ‒ [Tween 80]e [Tween 80]0
(1)
RI PT
[PHE]0 ‒ [PHE]e
141
× 100%
(2)
where [PHE]0 (mg L−1) and [Tween 80]0 (g L−1) are the initial concentration of PHE
144
and Tween 80, respectively; and [PHE]e (mg L−1) and [Tween 80]e (g L−1) are the
145
equilibrium concentration of PHE and Tween 80, respectively.
M AN U
SC
143
146
In the sorption kinetics experiments, 100 mL solution containing PHE (10, 20 and 30
148
mg L−1) and Tween 80 (5 g L−1) was dosed into the 250 mL conical flask.
149
adding 0.3 g L−1 AC, the conical flasks were placed on orbital shakers with a speed of
150
200 rpm at 25 ± 1 ºC.
151
withdrawn and filtered through a 0.45 µm nylon filter before analysis.
152
of PHE adsorbed qt (mg g−1) at a certain time t (h) was calculated from Eq. (3) (Li et
153
al., 2017):
After
At selected time intervals of 0-24 h, the samples were The amount
EP
AC C
154
TE D
147
qt =
[PHE]0 ‒ [PHE]t V
(3)
m
155
where [PHE]t (mg L−1) is the concentration of PHE at time t (h) in solution; V (L) is
156
the total volume of the solution; and m (g) is the mass of AC added.
157 158
The procedure of adsorption isotherms experiments was similar to that sorption 8
ACCEPTED MANUSCRIPT 159
kinetics experiments, and the PHE concentrations were ranged from 2 to 100 mg L−1
160
while the Tween 80 concentration and AC dosage were fixed at 5 and 0.3 g L−1,
161
respectively.
162
calculated from Eq. (4): qe =
163
RI PT
The amount of PHE adsorbed at the equilibrium qe (mg g−1) was
[PHE]0 ‒ [PHE]e V
(4)
m
SC
164 2.5 Analytical methods
166
The concentrations of PHE in aqueous solution were determined with a Shimadzu
167
LC-20AB high-performance liquid chromatography (HPLC) fitted with a diode array
168
detector (SPD-M20A) and a Shimadzu Shim-pack VP-ODS column (4.6 mm × 150
169
mm, 5 µm) using acetonitrile/water (v/v, 70/30) as the mobile phase at a flow rate of 1
170
mL min−1 at a wavelength of 254 nm.
171
determined by the cobalt ammonium thiocyanate color-developing method (Smullin
172
et al., 1971).
The concentrations of Tween 80 were
EP
TE D
M AN U
165
AC C
173 174
3
Results and discussion
175
3.1 Characterization of the prepared AC
176
The characterization results are given in Table S1, Figure 1, Figure S1 and Figure S2,
177
respectively.
178
Figure 1(a) show a Type IV isotherm which indicates the presence of a mesoporous
179
surface on the AC.
N2 adsorption-desorption isotherms and pore diameter distribution in
Moreover, an obvious hysteresis loop (H1 Type), due to 9
ACCEPTED MANUSCRIPT 180
capillary condensation, further proved the existence of a mesoporous structure and the
181
auxiliary chart in Figure 1(a) shows that the pore size is uniform with an average pore
182
diameter of 5.92 nm.
183
can be seen that it is irregular with a clear pore structure.
184
surface analysis results, it can be concluded that AC consists of a mesoporous rather
185
than microporous structure.
Figure 1(b) exhibits the surface morphology of the AC and it
SC
RI PT
Thus, along with the BET
186
Figure S1 shows that the value of pHPZC was obtained when pHinitial was equal to
188
pHfinal and corresponds to about pH 3.
189
surface of the AC will be negatively charged when the pH of the solution is greater
190
than 3.
M AN U
187
TE D
191
This low value of pHPZC indicates that the
Figure S2 shows the XPS survey spectrum of fresh AC and the elements C, O and P
193
could be detected; the presence of small amounts of P resulted from the H3PO4
194
impregnation process.
195
Figure 1(d) includes four signals which are attributed to graphitic type of C atoms (C1,
196
284.0 eV) (Li et al., 2017); carbon species of alcohol and ether groups (C2, 285.5 eV)
197
(Puziy et al., 2008); carboxylic/ester/lactone groups (C3, O=C-O, 288.5 eV) (Fang et
198
al., 2014); and π-π* transitions in aromatic rings (C4, 290.4 eV) (Li et al., 2017).
EP
192
AC C
Also, the C1s photoelectron spectrum of the fresh AC in
199 200
It can be seen from Figure 1(e) that in the fresh AC FTIR spectrum there are seven 10
ACCEPTED MANUSCRIPT 201
main peaks at 3426 cm−1, 2919 cm−1, 2845 cm−1, 2357 cm−1, 1640 cm−1, 1385 cm−1
202
and 1092 cm−1.
203
O-H stretching vibration of adsorbed water on the surface of AC ) (Salehi et al.,
204
2017); 2919 cm−1 and 2845 cm−1 (C-H symmetric and asymmetric stretching
205
vibration) (Salehi et al., 2017); 2357 cm−1 (the C-O vibrations of adsorbed carbon
206
dioxide); 1640 cm−1 (C=C stretching of aromatic rings) (Bernard et al., 2018); 1385
207
cm−1 (carboxyl O=C−O stretching vibration) (Wang et al., 2014); and 1092 cm−1
208
(alkoxy C−O stretching vibration) (Wang et al., 2014).
M AN U
SC
RI PT
The absorption bands are assigned as: 3426 cm−1 (hydroxyl groups
209 210
In the Boehm titration process, NaOH solution was used to titrate carboxyl, lactone
211
and phenolic groups; Na2CO3 titrated carboxyl and lactone groups; and NaHCO3 only
212
titrated carboxyl groups.
213
the different consumptions of the three basic solutions and the results are shown in
214
Table 1.
215
groups account for the main surface modification, lactone groups are much less
216
prevalent and there are very few phenolic groups present.
TE D
EP
The result is consistent with the work of Ge et al. (2015) that carboxyl
AC C
217
The amount of acidic groups was calculated according to
218
3.2 The factors influencing selective adsorption
219
In order to evaluate the efficiency and practicability of removing PHE and recovering
220
Tween 80 by selective adsorption, parameter selectivity (S) was introduced which was
221
expressed as the ratio of PHE removal percentage to Tween 80 loss percentage: 11
ACCEPTED MANUSCRIPT 222
S=
PHE removal percentage
(5)
Tween 80 loss percentage
223
An S value larger than 1 indicates that more PHE relative to Tween 80 was removed
224
which suggested a selective adsorption process and the higher the S value the better.
RI PT
225 Solution pH is usually an important factor in any adsorption process because pH
227
could influence the electrostatic and dispersive interactions between the adsorbate and
228
the adsorbent (Shi et al., 2013).
229
values used there was over 95% PHE removal and Tween 80 loss percentage was low
230
(less than 10%) under most situations.
231
little influence on the PHE removal and Tween 80 recovery so, for cost and ease of
232
operations, all further experiments were carried out at the natural solution pH of about
233
6.4.
SC
226
M AN U
Figure 2(a) shows clearly in that at each of the pH
TE D
This result indicates that solution pH had
234
Figure 2(b) shows that as the AC dose was increased from 0.05 to 0.3 g L−1, the
236
percentage of PHE removed increased from 49.2% to 95.1% and reached 98.8% when
237
the AC dose was 0.6 g L−1.
238
less than 10% in all instances indicating that larger AC doses promoted the PHE
239
removal.
240
that there are more effective adsorption sites.
241
removal and the full use of AC, the AC dose of 0.3 g L−1 was selected in all further
242
experiments.
AC C
EP
235
It was noteworthy that Tween 80 loss percentages were
This result is easy to understand since the higher adsorbent dose means
12
In order to make sure of high PHE
ACCEPTED MANUSCRIPT 243 Figure 2(c) shows the effect of changing the concentration of Tween 80 and as the
245
concentration of Tween 80 is increased the percentage of PHE removed decreased
246
gradually from 99.6% to 80.7% though the Tween 80 loss percentages were less than
247
10% in all situations.
248
could obstruct the adsorption process of PHE.
249
higher surfactant concentration, more surfactant micelles would remain in the aqueous
250
state, which can solubilize more PAH and decrease PAH sorption.
251
the entrance of the micropores of AC would be blocked by more surfactant molecules
252
(Rosas et al., 2013; Liu et al., 2016b), and consequently, fewer pores are available for
253
PHE adsorption.
Thus it can be concluded that higher Tween 80 concentrations
SC
According to Liu et al. (2016b), at
M AN U
In the meanwhile,
TE D
254
RI PT
244
Overall, it was found that all the S values were much greater than 1 which meant that
256
the selective adsorption was a good method for removal of PHE and recovery of
257
Tween 80 from soil washing effluent.
AC C
258
EP
255
259
3.3 Adsorption mechanism
260
To investigate adsorption mechanism, the AC surface before and after adsorption was
261
first characterized by FESEM.
262
covered by a thin film from the surface-adsorbed Tween 80 after adsorption.
263
would block the micropores of AC, and consequently decrease the surface area of AC
As can be seen from Figure 1(c), the AC surface was
13
This
ACCEPTED MANUSCRIPT 264
(Ahn et al., 2007; Yang et al., 2009).
Table S1 indicates that the specific surface area
265
and pore volume of AC reduced from 410.84 to 324.52 m2 g−1 and 0.61 to 0.50 cm3
266
g−1, respectively, while the average pore diameter of AC rose from 5.92 to 6.21 nm.
RI PT
267 268
The effects of surface chemistry of AC on PHE adsorptive removal can be studied by
269
XPS, FTIR and Boehm titration.
270
confirm the oxygen containing functional groups on the surface of AC, which is
271
consistent with the results obtained by Boehm titration.
272
in aromatic rings (peak C4 in Figure 1d) indicates the existence of π electron on the
273
surface of AC which may be related to PHE adsorption mechanism.
274
spectrum of spent AC (Figure 1e) shows the characteristic peaks of PHE at 811 cm−1
275
and 726 cm−1 (C-H out of plane bend) (Gupta and Gupta, 2016) indicating the
276
adsorption of PHE on the AC.
277
intensity of C=C stretching of aromatic rings at 1640 cm−1 was shifted to 1667 cm−1
278
and was strengthened. This proves the important role of π-π interactions in the
279
adsorption process between PHE and AC (Pei et al., 2013; Wang et al., 2014) and it
280
was obvious that π-π interactions strengthened the double bands and resulted in
281
blue-shift of C=C stretching vibration bands.
282
1385 cm−1 was slightly enhanced; and the alkoxy C−O band shifted from 1092 cm−1
283
to 1073 cm−1 after adsorption.
284
confirm the interactions between the oxygen containing functional groups of the AC
SC
The XPS peaks of C2 and C3 in Figure 1(c)
The FTIR
TE D
M AN U
In addition, π-π* transitions
AC C
EP
Additionally, the spent AC spectrum shows that the
The O=C−O stretching vibration at
These clear changes in the spent AC FTIR spectrum
14
ACCEPTED MANUSCRIPT 285
and PHE (Wang et al., 2014).
286 287
Based on the result and the literature, the selective adsorption of PHE from the Tween
288
80 solution can be considered to proceed via the following steps.
289
would be formed when the Tween 80 concentration is higher than critical micelle
290
concentration (CMC) (Haigh, 1996; Mousset et al., 2014).
291
solubility can be increased by being partitioned into the hydrophobic cores of Tween
292
80 (Chun et al., 2002; Zhou and Zhu, 2005).
293
80 solution containing PHE, the micelles with PHE partitioned into their hydrophobic
294
core (PHE-containing micelles) would be sorbed onto the solid/liquid interface (Liu et
295
al., 2016b), and form hemi-micelles on the surface of AC (Lanzon and Brown, 2013).
296
The formation of hemi-micelles also provides an additional PHE partitioning site
297
(Guha and Jaffé, 1996; Liu et al., 2016b).
298
will diffuse into the pores of AC, and the “empty micelles” will then be exchanged
299
with new PHE-containing micelles (Guha and Jaffé, 1996).
300
can be recovered through activated carbon adsorption.
RI PT
SC
As a result, the PHE
The PHE molecules in the hemi-micelle
EP
TE D
M AN U
When the AC was dispersed in Tween
As a result, Tween 80
AC C
301
Briefly, micelles
302
3.4 Adsorption kinetics of PHE
303
Firstly, pseudo-first-order and pseudo-second-order models (Zheng et al., 2016) were
304
applied to the adsorption kinetic process of the AC.
305
adsorbed quickly during the initial 5 h and the adsorption is virtually completed after 15
As seen in Figure 3(a), PHE is
ACCEPTED MANUSCRIPT 306
24 h under three different initial concentrations of PHE.
307 The non-linear and linear forms of pseudo-first-order model are represented by:
309
qt = qe 1 ‒ e‒k1 t
310
lnqe ‒ qt = lnqe ‒ k1 t
(7)
where k1 (h−1) is the adsorption rate constant of pseudo-first-order model.
SC
311
(6)
RI PT
308
312
314
The non-linear and linear forms of pseudo-second-order model are represented by:
M AN U
313
k2 q2 t
qt = 1 + k eq t 2 e
316 317
t qt
= q + k t
1
e
2 2 qe
(9)
where k2 (g mg−1 h−1) is the adsorption rate constant of pseudo-second-order model.
TE D
315
(8)
Figures 3(b) and (c) show the data plotted using the two kinetic models and it is
319
evident that the pseudo-second-order model (Figure 3c) best describes the adsorption
320
process with the kinetic parameters shown in Table S2.
321
pseudo-second-order adsorption rate constant k2 ranged from 0.26 to 0.06 g mg‒1 h‒1
322
with 10–30 mg L‒1 PHE.
323
when 120 mg L‒1 PHE was removed from TX100 solution using commercial AC (Liu
324
et al., 2014a), and the k2 value changed to 0.12 g mg‒1 h‒1 when 60 mg L‒1 PHE was
325
removed from rhamnolipid solution (Liu et al., 2016b).
AC C
EP
318
It can be seen that the
The corresponding k2 value was only 5.04×10‒3 g mg‒1 h‒1
326 16
ACCEPTED MANUSCRIPT 327
To further investigate the diffusion mechanism between solutes and particles, the
328
intra-particle diffusion model proposed by Weber and Morris is introduced and the
329
equation is shown below (Liu et al., 2015):
330
qt = kdi t0.5 + Cdi
RI PT
(10)
where kdi (mg g−1 h−0.5) is the adsorption rate constants of Weber and Morris
332
intra-particle diffusion model; Cdi (mg g−1) represents the effect of the boundary layer
333
on diffusion.
SC
331
M AN U
334
Generally speaking, an adsorption process on a porous adsorbent can be divided into
336
three stages: (1) external diffusion: the adsorbates go through the liquid film to the
337
external surface of the adsorbent and this stage is also known as boundary layer
338
diffusion or film diffusion; (2) intra-particle diffusion: the adsorbates move from the
339
exterior surface of the adsorbent to the internal structure of the adsorbent; (3) the last
340
stage: the adsorbates are quickly adsorbed to the active sites on the adsorbent and this
341
step is too fast to be a rate-limiting step (Yu et al., 2012 and 2016).
342
above description, it can be seen that the adsorption rate may be controlled by the first
343
stage and/or the second stage.
344
parameters of intra-particle diffusion model are shown in Table S2.
345
the intra-particle diffusion is the rate-limiting step in the first phase since the linear
346
fitting line passes through the origin (Yu et al., 2016).
347
fitting line deviates from the origin, indicating the adsorption process is controlled by
AC C
EP
TE D
335
According to the
The fitting plot is demonstrated in Figure 3(d) and the
17
As can be seen,
In the second phase, the
ACCEPTED MANUSCRIPT 348
both external and intra-particle diffusion.
The adsorption equilibrium reached in the
349
third phase (Li et al., 2016), and less time is required to reach equilibrium at lower
350
initial PHE concentration.
RI PT
351 3.5 The Adsorption isotherm
353
The classical Freundlich and Langmuir isotherm models were used to describe the
354
PHE adsorption isotherm (Foo and Hameed, 2010).
355
Freundlich fitting curve and the Langmuir fitting curve are given in Figure 4.
356
The non-linear and linear forms of Langmuir isotherm model are represented by:
SC
352
L
1
358
qe
e
= q + K 1
1
L qm []e
m
M AN U
q K []
e m L qe = 1 + K []
357
The original data, the
(11) (12)
where qm (mg g−1) is the maximum adsorption capacity of PHE; KL (L mg−1) is
360
Langmuir constant related to free energy of adsorption.
TE D
359
363 364
The non-linear and linear forms of Freundlich isotherm model are represented by:
AC C
362
EP
361
1
qe =KF [PHE]ne
(13)
1
lnqe = n ln[PHE]e + lnKF 1
365
where KF mg g1 L mg1 n and
366
intensity.
(14) 1 n
are Freundlich constants related to adsorption
367 368
It is evident that Langmuir model fits the data better than Freundlich model and the R2 18
ACCEPTED MANUSCRIPT value of the Langmuir model was 0.994 (shown in Table S3), indicating a very good
370
fit with the data, and this adsorption model represents monolayer adsorption and
371
assumes that adsorption sites have a definite number and location and every adsorbent
372
molecule possesses constant enthalpy and activation energy (Foo and Hameed, 2010).
373
Based on the Langmuir model, the maximum adsorption capacity qm was calculated to
374
be 247.54 mg g‒1, which is much higher than the literature report. The qm value was
375
only 43.9 mg g−1 when PHE was removed from rhamnolipid solution with
376
commercial AC (Liu et al., 2016b) and this value changed to 58.84 mg g−1 when
377
TX100 was used as the surfactant instead (Liu et al., 2014a).
378
M AN U
SC
RI PT
369
3.6 Adsorption thermodynamics
380
Thermodynamic parameters including Gibbs free energy change (∆G0 (kJ mol−1)),
381
enthalpy change (∆H0 (kJ mol−1)) and entropy change (∆S0 (J K−1 mol−1)) were
382
calculated using the following equations (Liu et al., 2014b):
383
EP
TE D
379
lnKd =
(15)
384
∆G0 = ∆H0 ‒ T∆S0
(16)
∆S0
∆H0 RT
AC C
R
−
385
where R is the gas constant, and T is the temperature (K).
386
coefficient (L g−1) that was calculated from the following equation:
387
Kd =
qe
Kd is the distribution
(17)
[PHE]e
388
Experiments were conducted at temperatures of 293.15, 303.15 and 313.15 K and the
389
linear regression of lnKd ∼
1 T
and the fitting equation based on Eq. (15) are shown in 19
ACCEPTED MANUSCRIPT The values of ∆H0 and ∆S0 were calculated from the equation and the
390
Figure S3.
391
data is given in Table 2.
392
adsorption process is exothermic and higher temperatures decrease adsorption.
393
positive value of ∆S0 reflected the affinity of the AC for PHE and the system
394
randomness increased after the adsorption process and the negative values of ∆G0
395
indicated that the PHE adsorption process was spontaneous (Murugesan et al., 2014).
The negative value of ∆H0 suggests that the PHE
SC
RI PT
The
396 3.7 Regeneration of the AC
398
It would be wasteful and cause secondary pollution if AC was no re-used, so
399
regeneration of the AC was investigated using the cheap and nontoxic solvent, ethanol.
400
After use, the spent AC was filtered and regenerated by shaking with ethanol in a
401
conical flask on an orbital stirrer, filtered, dried, cooled and retained for future use.
402
As seen in Figure 1(e), the FTIR spectrum of AC after regeneration indicated that the
403
two characteristic peaks at 811 cm−1 and 726 cm−1 of PHE had disappeared showing
404
that complete desorption of PHE had occurred.
405
and Figure 5 illustrates that PHE removal decreased slightly after regeneration but it
406
still reached 80% which was an acceptable result.
The regeneration was repeated twice
AC C
EP
TE D
M AN U
397
407 408
4
Conclusions
409
An effective adsorption process to remove PHE and recover Tween 80 surfactant
410
solution from soil washing effluent was investigated using activated carbon 20
ACCEPTED MANUSCRIPT 411
synthesized from waste walnut shells.
412
concentrations were optimized.
413
influence on PHE removal and the optimal conditions for selective adsorption
414
occurred when the solution was at natural pH, AC dose was 0.3 g L−1 and Tween 80
415
concentration was 5 g L−1.
416
and PHE removal and Tween 80 recovery percentage reached 95% and 90%,
417
respectively.
418
functional groups of AC and PHE are both important in the adsorption process.
419
Adsorption of PHE fitted fit well to the pseudo-second-order kinetic model and the
420
Weber and Morris intra-particle diffusion model.
421
best isotherm model which represents monolayer absorption and the PHE adsorption
422
process was exothermic and spontaneous as ∆H0 and ∆G0 were both negative.
423
could be efficiently regenerated with ethanol and subsequent reuse (twice) resulted in
424
80% of PHE removal.
425
effluent and the Tween 80 surfactant solution could be reused.
RI PT
The results showed that solution pH had no
SC
Good selectivity was displayed under optimal conditions
M AN U
The π-π interactions and the interactions between oxygen containing
TE D
Langmuir isotherm model is the
AC
EP
Overall, PHE was effectively removed from soil washing
AC C
426
The solution pH, AC dose and Tween 80
427
Acknowledgements
428
This work was supported by Wuhan Applied Basic Research Project (Grant No.
429
2016060101010074) and Shenzhen Basic Research Plan Project (Grant No.
430
JCYJ20150508152951667).
431
polishing this manuscript is also greatly appreciated.
The generous help of Professor David H. Bremner in
21
ACCEPTED MANUSCRIPT 432 433
References
434
Acharya, P., Ives, P., 1994. Incineration at Bayou Bounfouca remediation project. Waste Manage. 44, 13-26.
RI PT
435 436
Ahn, C.K., Kim, Y.M., Woo, S.H., Park, J.M., 2007. Selective adsorption of
437
phenanthrene dissolved in surfactant solution using activated carbon.
438
Chemosphere 69, 1681-1688.
Ahn, C.K., Kim, Y.M., Woo, S.H., Park, J.M., 2008a. Soil washing using various
440
nonionic surfactants and their recovery by selective adsorption with activated
441
carbon. J. Hazard. Mater. 154, 153-160.
M AN U
SC
439
442
Ahn, C.K., Lee, M.W., Lee, D.S., Woo, S.H., Park, J.M., 2008b. Mathematical
443
evaluation of activated carbon adsorption for surfactant recovery in a soil
444
washing process. J. Hazard. Mater. 160, 13-19.
Ahn, C.K., Woo, S.H., Park, J.M., 2010. Selective adsorption of phenanthrene in
446
nonionic-anionic surfactant mixtures using activated carbon. Chem. Eng. J. 158,
447
115-119.
EP
TE D
445
Alimohammadi, Z., Younesi, H., Bahramifar, N., 2016. Batch and column adsorption
449
of reactive red 198 from textile industry effluent by microporous activated
450
AC C
448
carbon developed from walnut shells. Waste Biomass Valori. 7, 1255-1270.
451
Ashrafi, M., Chamjangali, M.A., Bagherian, G., Goudarzi, N., 2017. Application of
452
linear and non-linear methods for modeling removal efficiency of textile dyes
453
from aqueous solutions using magnetic Fe3O4 impregnated onto walnut shell.
454
Spectrochim. Acta A. 171, 268-279.
455
Bernard, F.L., Duczinski, R.B., Rojas, M.F., Fialho, M.C., Carreño, L.A., Chaban,
456
V.V., Vecchia, F.D., Einloft, S., 2018. Cellulose based poly (ionic liquids): tuning 22
ACCEPTED MANUSCRIPT 457
cation-anion interaction to improve carbon dioxide sorption. Fuel 211, 76-86.
458
Chen, S.C., Liao, C.M., 2006. Health risk assessment on human exposed to
459
environmental polycyclic aromatic hydrocarbons pollution sources. Sci. Total
460
Environ. 366, 112-123. Cheng, M., Zeng, G.M., Huang, D.L., Yang, C.P., Lai, C., Zhang, C., Liu, Y., 2017.
462
Advantages and challenges of Tween 80 surfactant-enhanced technologies for
463
the remediation of soils contaminated with hydrophobic organic compounds.
464
Chem. Eng. J. 314, 98-113.
SC
466
Chun, C.L., Lee, J.J., Park, J.W., 2002. Solubilization of PAH mixtures by three different anionic surfactants. Environ. Pollut. 118, 307-313.
M AN U
465
RI PT
461
Fang, G.D., Liu, C., Gao, J., Zhou, D.M., 2014. New insights into the mechanism of
468
the catalytic decomposition of hydrogen peroxide by activated carbon:
469
implications for degradation of diethyl phthalate. Ind. Eng. Chem. Res. 53,
470
19925-19933.
471 472
TE D
467
Foo, K.Y., Hameed, B.H., 2010. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 156, 2-10.
Gan, S., Lau, E.V., Ng, H.K., 2009. Remediation of soils contaminated with
474
polycyclic aromatic hydrocarbons (PAHs). J. Hazard. Mater. 172, 532-549.
475
Ge, X.Y., Tian, F., Wu, Z.L., Yan, Y.J., Cravotto, G., Wu, Z.S., 2015. Adsorption of
476
naphthalene from aqueous solution on coal-based activated carbon modified by
AC C
477
EP
473
microwave induction: microwave power effects. Chem. Eng. Process. 91, 67-77.
478
Gharibzadeh, F., Kalantary, R.R., Nasseri, S., Esrafili, A., Azari, A., 2016. Reuse of
479
polycyclic aromatic hydrocarbons (PAHs) contaminated soil washing effluent by
480
bioaugmentation/biostimulation process. Sep. Purif. Technol. 168, 248-256.
481
Gómez, J., Alcántara, M.T., Pazos, M., Sanromán, M.A., 2010a. Remediation of 23
ACCEPTED MANUSCRIPT 482
polluted soil by a two-stage treatment system: desorption of phenanthrene in soil
483
and electrochemical treatment to recover the extraction agent. J. Hazard. Mater.
484
173, 794-798. Gómez, J., Alcántara, M.T., Pazos, M., Sanromán, M.A., 2010b. Soil washing using
486
cyclodextrins and their recovery by application of electrochemical technology.
487
Chem. Eng. J. 159, 53-57.
RI PT
485
Gong, Z.Q., Wang, X.G., Tu, Y., Wu, J.Y., Sun, Y.F., Li, P., 2010. Polycyclic
489
aromatic hydrocarbon removal from contaminated soils using fatty acid methyl
490
esters. Chemosphere 79, 138-143.
SC
488
Guha, S., Jaffé, P.R., 1996. Bioavailability of hydrophobic compounds partitioned
492
into the micellar phase of nonionic surfactants. Environ. Sci. Technol. 30, 1382–
493
1391.
495 496 497
Gupta, H., Gupta, B., 2016. Adsorption of polycyclic aromatic hydrocarbons on banana peel activated carbon. Desalin. Water Treat. 57, 9498-9509.
TE D
494
M AN U
491
Haigh, S.D., 1996. A review of the interaction of surfactants with organic contaminants in soil. Sci. Total Environ. 185, 161-170. Kuppusamy, S., Thavamani, P., Venkateswarlu, K., Lee, Y.B., Naidu, R., Megharaj,
499
M., 2017. Remediation approaches for polycyclic aromatic hydrocarbons (PAHs)
501 502
AC C
500
EP
498
contaminated soils: technological constraints, emerging trends and future
directions. Chemosphere 168, 944-968.
Lanzon, J.B., Brown, D.G., 2013. Partitioning of phenanthrene into surfactant
503
hemi-micelles
on
the
bacterial
cell
surface
and
implications
504
surfactant-enhanced biodegradation. Water Res. 47, 4612-4620.
for
505
Lau, E.V., Gan, S., Ng, H.K., Poh, P.E., 2014. Extraction agents for the removal of
506
polycyclic aromatic hydrocarbons (PAHs) from soil in soil washing technologies. 24
ACCEPTED MANUSCRIPT 507 508 509
Environ. Pollut. 184, 640-649. Lee, D.H., Cody, R.D., Kim, D.J., 2002. Surfactant recycling by solvent extraction in surfactant-aided remediation. Sep. Purif. Technol. 27, 77-82. Lemaire, J., Buès, M., Kabeche, T., Hanna, K., Simonnot, M.O., 2013. Oxidant
511
selection to treat an aged PAH contaminated soil by in situ chemical oxidation. J.
512
Environ. Chem. Eng. 1, 1261-1268.
RI PT
510
Li, H.L., Qu, R.H., Li, C., Guo, W.L., Han, X.M., He, F., Ma, Y.B., Xing, B.S., 2014.
514
Selective removal of polycyclic aromatic hydrocarbons (PAHs) from soil
515
washing effluents using biochars produced at different pyrolytic temperatures.
516
Bioresource Technol. 163, 193-198.
M AN U
SC
513
517
Li, H.Y., Li, G.X., Wu, J., Zhong, X., Zhang, H., 2016. Removal of tetracycline from
518
aqueous solution by hydrothermal method derived titanate nanotubes. Desalin.
519
Water Treat. 57, 19965-19974.
Li, J., Du, Y., Deng, B., Zhu, K.M., Zhang, H., 2017. Activated carbon adsorptive
521
removal of azo dye and peroxydisulfate regeneration: from a batch study to
522
continuous column operation. Environ. Sci. Pollut. R. 24, 4932-4941.
TE D
520
Liu, J.F., Chen, J.J., Jiang, L., Yin, X., 2014a. Adsorption of mixed polycyclic
524
aromatic hydrocarbons in surfactant solutions by activated carbon. J. Ind. Eng.
AC C
525
EP
523
Chem. 20, 616-623.
526
Liu, J.F., Chen, J.J., Jiang, L., Wang, X.W., 2014b. Adsorption of fluoranthene in
527
surfactant solution on activated carbon: equilibrium, thermodynamic, kinetic
528
studies. Environ. Sci. Pollut. R. 21, 1809-1818.
529
Liu, J.F., Zhang, Y.S., Sun, X.J., Hu, W.L., 2016b. Separation of polycyclic aromatic
530
hydrocarbons from rhamnolipid solution by activated carbon adsorption. Environ.
531
Earth Sci. 75, 1453. 25
ACCEPTED MANUSCRIPT 532
Liu, N., Charrua, A.B., Weng, C.H., Yuan, X.L., Ding, F., 2015. Characterization of
533
biochars derived from agriculture wastes and their adsorptive removal of atrazine
534
from aqueous solution: a comparative study. Bioresource Technol. 198, 55-62. Liu, W., Cai, Z.Q., Zhao, X., Wang, T., Li, F., Zhao, D.Y., 2016a. High-capacity and
536
photoregenerable composite material for efficient adsorption and degradation of
537
phenanthrene in water. Environ. Sci. Technol. 50, 11174-11183.
RI PT
535
Mousset, E., Oturan, N., Hullebusch, E.D., Guibaud, G., Esposito, G., Oturan, M.A.,
539
2014. Influence of solubilizing agents (cyclodextrin or surfactant) on
540
phenanthrene degradation by electro-Fenton process - Study of soil washing
541
recycling possibilities and environmental impact. Water Res. 48, 306-316.
M AN U
SC
538
Murugesan, A., Vidhyadevi, T., Kalaivani, S.S., Thiruvengadaravi, K.V., Ravikumar,
543
L., Anuradha, C.D., Sivanesan. S., 2014. Modelling of lead (II) ion adsorption
544
onto poly (thiourea imine) functionalized chelating resin using response surface
545
methodology (RSM). J. Water Process. Eng. 3,132-143.
TE D
542
Nazari, G., Abolghasemi, H., Esmaieli, M., 2016. Batch adsorption of cephalexin
547
antibiotic from aqueous solution by walnut shell based-activated carbon. J.
548
Taiwan Inst. Chem. E. 58, 357-365.
550
Paria, S., Yuet, P.K., 2006. Solubilization of naphthalene by pure and mixed
AC C
549
EP
546
surfactants. Ind. Eng. Chem. Res. 45, 3552-3558.
551
Pei, Z.G., Li, L.Y., Sun, L.X., Zhang, S.Z., Shan, X.Q., Yang, S., Wen, B., 2013.
552
Adsorption characteristics of 1,2,4-trichlorobenzene, 2,4,6-trichlorophenol,
553
2-naphthol and naphthalene on graphene and graphene oxide. Carbon 51,
554
156-163.
555 556
Puziy, A.M., Poddubnaya, O.I., Socha, R.P., Gurgul, J., Wisniewski, M., 2008. XPS and NMR studies of phosphoric acid activated carbons. Carbon 46, 2113-2123. 26
ACCEPTED MANUSCRIPT 557
Rosas, J.M., Santos, A., Romero, A., 2013. Soil-washing effluent treatment by
558
selective adsorption of toxic organic contaminants on activated carbon. Water Air
559
Soil Poll. 224, 1506. Salehi, A.M., Moussavi, G., Yaghmaeian, K., 2017. Preparation, characterization and
561
catalytic activity of a novel mesoporous nanocrystalline MgO nanoparticle for
562
ozonation of acetaminophen as an emerging water contaminant. Chem. Eng. J.
563
310, 157-169.
RI PT
560
Shi, Q.Q., Li, A.M., Zhu, Z.L., Liu, B., 2013. Adsorption of naphthalene onto a
565
high-surface-area carbon from waste ion exchange resin. J. Environ. Sci. 25, 1,
566
188-194.
568
M AN U
567
SC
564
Smullin, C.F., Wetterau, F.P., Olsanski, V.L., 1971. The determination of poly sorbate 60 in foods. J. Am. Oil. Chem. Soc. 48, 1, 18-20.
Trellu, C., Mousset, E., Pechaud, Y., Huguenot, D., van Hullebusch, E.D., Esposito,
570
G., Oturan, M.A., 2016. Removal of hydrophobic organic pollutants from soil
571
washing/flushing solutions: a critical review. J. Hazard. Mater. 306, 149-174.
572
Usman, M., Faure, P., Ruby, C., Hanna, K., 2012. Application of magnetite-activated
573
persulfate oxidation for the degradation of PAHs in contaminated soils.
574
Chemosphere 87, 234-240.
AC C
EP
TE D
569
575
Vizcaíno, R.L., Sáez, C., Cañizares, P., Rodrigo, M.A., 2012. The use of a combined
576
process of surfactant-aided soil washing and coagulation for PAH-contaminated
577
soils treatment. Sep. Purif. Technol. 88, 46-51.
578
Wang, J., Chen, Z.M., Chen, B.L., 2014. Adsorption of polycyclic aromatic
579
hydrocarbons by graphene and graphene oxide nanosheets. Environ. Sci. Technol.
580
48, 4817-4825.
581
Xiong, B.J., Zhang, Y.C., Hou, Y.W., Arp, H.P., Reid, B.J., Cai, C., 2017. Enhanced 27
ACCEPTED MANUSCRIPT 582
biodegradation of PAHs in historically contaminated soil by M. Gilvum
583
inoculated biochar. Chemosphere 182, 316-324. Yang, J.S., Baek, K., Kwon, T.S., Yang, J.W., 2009. Adsorption of chlorinated
585
solvents in nonionic surfactant solutions with activated carbon in a fixed bed. J.
586
Ind. Eng. Chem. 15, 777-779.
RI PT
584
Yang, J., Qiu, K.Q., 2010. Preparation of activated carbons from walnut shells via
588
vacuum chemical activation and their application for methylene blue removal.
589
Chem. Eng. J. 165, 209-217.
SC
587
Yi, Z.Y., Yao, J., Kuang, Y.F., Chen, H.L., Wang, F., Yuan, Z.M., 2015. Removal of
591
Pb (II) by adsorption onto Chinese walnut shell activated carbon. Water Sci.
592
Technol. 72, 6, 983-989.
M AN U
590
Yu, B., Tian, J., Feng, L., 2017. Remediation of PAH polluted soils using a soil
594
microbial fuel cell: influence of electrode interval and role of microbial
595
community. J. Hazard. Mater. 336, 110-118.
TE D
593
Yu, F., Wu, Y.Q., Li, X.M., Ma, J., 2012. Kinetic and thermodynamic studies of
597
toluene, ethylbenzene, and m-xylene adsorption from aqueous solutions onto
598
KOH-activated multi-walled carbon nanotubes. J. Agric. Food Chem. 60,
599
12245-12253.
AC C
EP
596
600
Yu, F., Sun, S.N., Han, S., Zheng, J., Ma, J., 2016. Adsorption removal of
601
ciprofloxacin by multi-walled carbon nanotubes with different oxygen contents
602
from aqueous solutions. Chem. Eng. J. 285, 588-595.
603
Zhao, B.W., Che, H.L., Wang, H.F., Xu, J., 2016. Column flushing of phenanthrene
604
and copper (II) Co-contaminants from sandy soil using Tween 80 and citric acid.
605
Soil Sediment Contam. 25, 1, 50-63.
606
Zheng, T.L., Wang, Q.H., Shi, Z.N., Zhang, Z.H., Ma, Y.H., 2016. Microwave 28
ACCEPTED MANUSCRIPT 607
regeneration of spent activated carbon for the treatment of ester-containing
608
wastewater. RSC Adv. 6, 60815-60825. Zhou, J., Yang, B., Li, Z.J., Lei, L.C., Zhang, X.W., 2015b. Selective adsorption of
610
naphthalene in aqueous solution on mesoporous carbon functionalized by
611
task-specific ionic liquid. Ind. Eng. Chem. Res. 54, 2329-2338.
RI PT
609
Zhou, W.J., Zhu, L.Z., 2005. Distribution of polycyclic aromatic hydrocarbons in
613
soil-water system containing a nonionic surfactant. Chemosphere 60, 1237-1245.
614
Zhou, W.J., Wang, X.H., Chen, C.P., Zhu, L.Z., 2013. Removal of polycyclic
615
aromatic hydrocarbons from surfactant solutions by selective sorption with
616
organo-bentonite. Chem. Eng. J. 233, 251-257.
M AN U
SC
612
Zhou, W.J., Yang, Q., Chen, C.P., Wu, Q.Q., Zhu, L.Z., 2015a. Fixed-bed study and
618
modeling of selective phenanthrene removal from surfactant solutions. Colloid
619
Surf. A 470, 100-107.
AC C
EP
TE D
617
29
ACCEPTED MANUSCRIPT Table 1 AC surface functional groups from Boehm titration Amount (mmol g−1)
Carboxyl groups
0.9275
Lactones
0.1000
RI PT
Functional groups
0.0025
AC C
EP
TE D
M AN U
SC
Phenolic groups
ACCEPTED MANUSCRIPT Table 2 Thermodynamic parameters of PHE adsorption
∆H0 (kJ mol−1)
∆G0 (kJ mol−1)
∆S0 (J K−1 mol−1)
20.473
303.15 K
313.15 K
‒7.402
‒7.607
‒7.812
AC C
EP
TE D
M AN U
SC
RI PT
‒1.400
293.15 K
AC C
EP
TE D
M AN U
(a)
SC
RI PT
ACCEPTED MANUSCRIPT
(b)
AC C
EP
TE D
M AN U
(c)
SC
RI PT
ACCEPTED MANUSCRIPT
(d)
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
(e)
Figure 1. Characterization of AC: (a) N2 adsorption-desorption isotherm and pore diameter distribution of fresh and spent AC; (b) FESEM image of fresh AC; (c)
TE D
FESEM image of spent AC; (d) C1s spectrum of fresh AC; (e) FTIR spectra ([PHE] =
AC C
EP
100 mg L−1, [AC] = 0.3 g L−1, [Tween 80] = 5 g L−1)
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
(a)
(b)
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
(c)
Figure 2. Effects of: (a) pH ([PHE] = 20 mg L−1, [AC] = 0.3 g L−1, [Tween 80] = 5 g L−1); (b) AC dose ([PHE] = 20 mg L−1, [Tween 80] = 5 g L−1, pH 6.4); and (c)
AC C
EP
TE D
Tween 80 concentration ([PHE] = 20 mg L−1, [AC] = 0.3 g L−1, pH 6.4)
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
(a)
(b)
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
(c)
(d)
Figure 3. Adsorption kinetics of PHE: (a) adsorption equilibrium time; (b) pseudo-first-order model; (c) pseudo-second-order model; (d) Weber and Morris intra-particle diffusion model
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
Figure 4. Adsorption isotherm of PHE: Langmuir model and Freundlich model.
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 5. Reuse of AC ([PHE] = 20 mg L−1, [AC] = 0.3 g L−1, [Tween 80] = 5 g
AC C
EP
TE D
L−1, pH 6.4)
ACCEPTED MANUSCRIPT Highlights
•
AC was prepared from walnut shell to adsorb phenanthrene from Tween 80
•
RI PT
solution 247.54 mg g‒1 adsorption capacity, 95% phenanthrene removal and 90% Tween 80 recovery
π-π interactions and oxygen containing groups play important roles in adsorption
•
The activated carbon could be regenerated and reused at least two times
AC C
EP
TE D
M AN U
SC
•