Вы находитесь на странице: 1из 9

Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468

Contents lists available at ScienceDirect

Journal of the Taiwan Institute of Chemical Engineers


journal homepage: www.elsevier.com/locate/jtice

Influence of acidic and basic treatments of activated carbon derived from


waste rubber tires on adsorptive desulfurization of thiophenes
Tawfik A. Saleh a,∗, Gaddafi I. Danmaliki b
a
Chemistry Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
b
Environmental Science Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

a r t i c l e i n f o a b s t r a c t

Article history: This work reports on the influence of treatment conditions on the waste tire-derived activated carbon for ad-
Received 6 June 2015 sorptive desulfurization. The rubber tires were carbonized and activated. The obtained activated carbon (AC)
Revised 8 November 2015
was treated with HNO3 or NaOH at a temperature range of 30–90 °C. The morphology and surface proper-
Accepted 13 November 2015
ties of AC were characterized by surface pH, Boehm’s titration, N2 adsorption–desorption isotherms, Fourier-
Available online 7 December 2015
transform infrared spectroscopy, X-ray diffraction, and scanning electron microscope. The AC sample, treated
Keywords: with HNO3 at 90 °C, possess the highest surface oxygen containing functional groups (2.39 mmol/g), surface
Adsorptive desulfurization area (473.35 m2 /g) and pore volume (0.70 cm3 /g) and the more adsorption capacity to the refractory sulfur
Fixed bed compounds. The Boehm’s titration experiments indicated that the amount of surface oxygen containing func-
Waste tires tional groups on the surface of the acid-treated AC increases with treatment temperatures. Acid-treated AC
Carbon at 90 °C proves to be optimum for adsorptive desulfurization with the order of dibenzothiophene > benzoth-
iophene > thiophene.
© 2015 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction larger surface areas and pore size distributions [12]. Mesoporous and
microporous volumes of AC also play a crucial role in the adsorption
Combustion of fossil fuels is associated with the release of toxic of larger molecules. In general, the selective adsorption of inorganic
gases that pose serious environmental and health related impacts. molecules on the surface of AC directly depends on the amount of
Sulfur and its refractory compounds found in fossil fuels are ex- oxygen containing complexes. The oxygen containing complexes are
tremely reactive. They cause catalyst poisoning and corrode refining mostly created using dry or wet oxidation methods [13–15]. Adsorp-
equipment. In order to reduce air pollution and technological prob- tive capability of AC modified with HNO3 at 120 °C was increased sig-
lems, EU and US have enacted laws to regulate sulfur emissions to nificantly and modification using HNO3 even at high temperatures
10 and 15 ppm respectively [1,2]. Various technologies are currently showed a promising result in the removal of thiophenes from oil
in use to lower the organic sulfur compounds in fuels, these meth- [16,17].
ods include: hydrodesulphurization (conventional method), oxida- The aim of this work was to produce AC from end of life tires as a
tive desulfurization, ionic liquids desulfurization, adsorptive desul- free-of-cost source, and to treat the AC using HNO3 and NaOH at var-
furization and biodesulfurization [3–6]. Of all these technologies, ad- ious temperatures. The influence of acid/base treatment of AC for the
sorptive desulfurization, one of the most promising technologies is removal of thiophenes was examined. The adsorption performance of
receiving much attention because of its facile and mild operating con- the AC was evaluated.
ditions. Much attention is focused on the development of porous ad-
sorbents such as activated carbon, zeolites, alumina, and zirconia that
2. Experimental
are cheap, easily regenerated and possesses high selectivity for sulfur
compounds [7–11].
2.1. Development of the adsorbent
The surface areas and porosities of AC are greatly influenced by the
parent material and the method employed in the production. Better
Waste rubber tires were cleaned by removing the iron wires from
adsorption capacities and adsorption rates of AC are directly linked to
the tire. Then, the waste rubber was cut into small pieces and thor-
oughly washed with deionized water, and then dried in an oven at

Corresponding author. Tel.: +966 13 860 1734; fax: +966 13 860 1734. 110 °C. The elemental analysis of the waste rubber indicated their ini-
E-mail address: tawfik@kfupm.edu.sa, tawfikas@hotmail.com (T.A. Saleh). tial elemental composition as 82.3 carbon, 9.2 oxygen, 0.7 nitrogen,
URL: http://faculty.kfupm.edu.sa/CHEM/tawfik/ (T.A. Saleh) 2.8 sulfur, 3.8 silicon and 1.2 silicon and silicon with different wt. %.

http://dx.doi.org/10.1016/j.jtice.2015.11.008
1876-1070/© 2015 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468 461

The granules were heated up to 300 °C to separate the black tire the aliquot taken from the original sample. [NaOH]VNaOH is the con-
crude oil, distilled diesel oil and produced oil. Pyrolysis was carried centration and volume of the titrant in the back titration. Va is the
out in a muffle furnace at 500 °C for 5 h to isolate the ash and carbon volume of the aliquot taken from VB .
black from pyro-gas and other oils, followed by treatment with hy- Surface pH of the carbon samples was tested by HI 3512 bench-top
drogen peroxide solution to oxidize all adhering organic impurities. pH meter equipped with a graphic LCD display to gain insight about
The material was washed with deionized water and dried in vacuum the acidity and basicity of the samples. A suspension of 0.2 g of each
oven at 110 °C overnight. The char was then placed in the stainless sample was added to 10 ml of water and the suspension was stirred
tube, inserted into the furnace, heated under nitrogen flow through overnight to reach equilibrium. The samples were then filtrated and
the sample bed at 200 ml/min at 20 °C/min to 900 °C and a holding the pH of the each solution was measured.
time of 5 h. Then the reactor was cooled down to room temperature
before the sample was removed, and kept in an oven overnight, then 2.3. Characterization of the adsorbent
grounded to small particles using pestle and mortar.
The produced black granules were sieved and the powdered car- The FT-IR spectra of the samples were recorded using Nicolet 6700
bon with a particle size of < 0.1 mm was used to allow good surface spectrometer (Thermo electron, USA) with a resolution of 2.0 cm–1
contact with acid or base reagent solution. Chemical activation was well equipped with deuterated triglycine sulfate detector and an OM-
conducted on the prepared material using 4 M HNO3 and 4 M NaOH NIC program. The experiments were conducted on the powdered
at three different temperatures respectively (30, 60, and 90 °C) and a samples ground in an agate mortar to produce KBr pellets and spec-
duration of 3, 6 and 12 h. Then, the mixture was filtered and the ob- tra were obtained by adding 64 scans and corrected for the back-
tained activated carbon (AC) was washed with deionized water and ground noise. The spectra of the samples were recorded in transmis-
dried overnight. sion mode and the wavenumber range of 4000–400 cm–1 .
It should be mentioned that increasing the refluxing time from Scanning electron microscopy (SEM) analysis of the treated sam-
3 h to 12 h has an overall net negative effect on the removal of the ples was conducted by using Hitachi model S- 3500 N instrument.
refractory sulfur compounds (see supporting information file, section Energy dispersive X-ray spectroscopy (EDX) analysis was conducted
S3.1.). Therefore, 3 h refluxing time was selected. using Oxford Instrument (England) and X-Max detector to determine
the elemental composition of the treated samples.
2.2. Boehm’s method for surface functional groups X ray diffraction patterns of the adsorbents were taken using
(Rigaku Miniflex II desktop X-ray diffractometer) using Cu-Kά radi-
The amount of oxygen surface functional groups (both acidic and ation and an X-ray gun operated at 40 kV (voltage) and 200 mA cur-
basic) was determined according to modified Boehm’s method [18]. rent. Data was collected from 2θ = 10°– 80° at a scan rate of 4°/min.
0.5 g of raw and treated carbon samples were added to beakers each The porous structure of the AC samples was characterized by ad-
containing 25 ml of the following 0.05 M solutions: NaOH, Na2 CO3 , sorption/desorption of nitrogen at (−196 °C) on a Micromeritics ASAP
NaHCO3 and HCl. The beakers were sealed and shaken for 24 h and 2020 surface area and porosimetry analyzer (Micromeritics, USA) to
then 10 ml of each filtrate was pipetted in an excess of 20 ml 0.05 M determine the surface area (BET) and pore volume of the treated
HCl for the determination of acidic functional groups or NaOH for the sorbents.
basic functional groups. The filtrate was titrated with 0.05 M NaOH
or HCl, respectively, using phenolphthalein indicator and the volume 2.4. Adsorption experiment
required to reach the endpoint was noted. For Na2 CO3 reaction bases
an excess of 30 ml 0.05 M HCl was added rather than 20 ml due to 2.4.1. Batch mode
diprotic property of the base to ensure complete reaction with acid. In a typical run in the batch mode adsorption studies, various
The number of acidic sites was calculated under the assumption that amounts, in the range between 0.1 and 0.5 g of adsorbents were intro-
NaOH neutralizes carboxylic, phenolic, and lactonic groups; Na2 CO3 duced into 15 ml of the fuel solution containing thiophene (T), ben-
neutralizes carboxylic and lactonic groups; and NaHCO3 neutralizes zothiophene (BT) and dibenzothiophene (DBT) in hexane (85%) and
only carboxylic groups. The number of surface basic sites was calcu- toluene (15%) with initial concentration of 50 ppm each. The sulfur
lated from the amount of hydrochloric acid required. The following removal efficiency of the AC was calculated using the formula below:
formula was used to calculate the amount of surface acidic and basic
groups neutralized with NaOH and HCl respectively [19]:
x = (C0 − Ce )/C0 ∗ 100% (1)
Moles of Carbon functionality
where x x = sulfur removal percentage (%) C0 = initial concentration
([R a or b] Vr − ([X] Vx − [T ] Vt )) of sulfur in model is (ppmw), Ce = final sulfur concentration.
=
m The amount of sulfur compounds adsorbed per unit mass of the
where [R a or b] and Vr denote the concentration and volume of the adsorbent at equilibrium (qe mg/g) and at any time t (qt mg/g) termed
reaction base or acid in mol/L and liters respectively. [X] and Vx de- as the adsorption capacity was calculated from the formulas:
note the concentration and volume of the excess acid or base added
V (C0 − Ce )
to the aliquot. [T ] and Vt denote the concentration and volume of the qe = (2)
m
titrant required to reach the endpoint. m stands for the mass of the
carbon used in grams. V (C0 − Ct )
The amount of surface acidic groups neutralized by Na2 CO3 and qt = (3)
m
NaHCO3 was calculated using the following formula [20,21]:
where V (L) is the volume of the liquid phase, C0 (mg/L) is the initial
nHCl VB
ncsf = [B]VB − ([HCl]VHCl − [NaOH]VNaOH ) concentration of the sulfur compounds before they come in contact
nB Va with the adsorbent, Ce and Ct (mg/L) are the concentration at equilib-
where ncsf denotes the moles of carbon surface functionalities on rium and at any time t, and m (g) is the amount of the adsorbent.
the surface of the carbon that reacted with base during the mix-
n
ing step. nHCl = molar ratio of acid to base [B] and VB are the con- 2.4.2. Break-through experiments
B
centration and volume of the reaction base mixed with the carbon. Fixed-bed flow system was used to test the simultaneous adsorp-
[HCl] and VHCl are the concentration and volume of the acid added to tive desulfurization of the treated AC samples using a column system.
462 T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468

Table 1
Surface pH, total basic and acidic groups on the raw and treated AC.

Sample Surface pH Total basic groups (mmol/g) Surface acidic groups (mmol/g)

Phenol Lactone Carboxyl Total

Raw AC 4.92 1.62 0.81 0.22 0.68 1.71


AC-NaOH 30 °C 10.66 2.48 1.07 0.16 0.00 1.23
AC-NaOH 60 °C 10.18 1.98 1.21 0.33 0.18 1.72
AC-NaOH 90 °C 10.69 2.46 1.13 0.25 0.00 1.38
AC-HNO3 30 °C 4.21 1.37 0.80 0.64 0.67 2.11
AC-HNO3 60 °C 4.05 1.36 0.82 0.71 0.69 2.22
AC-HNO3 90 °C 3.09 1.33 0.69 0.81 0.89 2.39

The adsorbent was packed inside the column. The model fuel sam- 2.11 to 2.22 and finally 2.39 mmol/g for AC-HNO3 30, 60 and 90 °C
ple was then passed through the column by a peristaltic pump with respectively, and the total basic groups slightly reduced from 1.37 to
a controlled flow rate (150 rpm) at a temperature of 25 °C. Once the 1.36 and finally to 1.33 mmol/g for the same samples respectively.
adsorption process started, treated fuels were periodically sampled at
different time intervals for analysis. The experiments were repeated
3.1.2. FTIR spectroscopy
three times and the relative standard deviation was calculated and FTIR spectra of the raw, HNO3 , and NaOH treated AC at various
found to be < 1.5%. temperatures (30, 60 and 90 °C) were taken to understand the func-
tional groups present in the AC and they are given in Fig. 1. The Raw
2.5. Analysis AC, NaOH and HNO3 treated AC have displayed varying degrees and
intensities of their bands though they all have similar patterns but
The concentrations of the refractory sulfur compounds were an- the bands increase with in treatment temperature. This implies that
alyzed by gas chromatography sulfur chemiluminescence detector temperature has an effect on the development of oxygen containing
(GC-SCD) (Model 7890A) system (Agilent) equipped with auto sam- functional groups and acidic and basic sites on the surface of the AC
pler (7693) and splitless injector. The column used was Agilent [24,25]. It can be seen that the samples showed a broad peak cen-
19091S-001: 2638-45555 50 m × 0.2 mm dimensions and a film tering around 3430 cm−1 that can be ascribed to O–H stretching vi-
thickness of 0.5 μm methyl siloxane stationary phase. The injection brations of hydroxyl groups or to chemisorbed water. There is a no-
volume was 1 μl with a flow rate of 0.839 ml/min. Inlet temperature ticeable development peak at 2854 cm−1 and 2924 cm−1 as the tem-
was a 60 °C and oven program of 60 °C for 1 min, then 10 °C/min to perature increases and these peaks can be due to the presence of
230 °C for 2 min. Other parameters and conditions of the method are aliphatic C–H stretch of CH, CH2 and CH3 . The peak at 1720 cm−1
listed in the Supporting information file, Table S1. is attributed to the carboxyl group. The peak at 1640 cm-1 is usu-
ally attributed to the carbonyl group. The peak around 1400 cm−1
3. Results and discussion appeared in HNO3 treated AC, may be attributed to the presence of
C=O stretching vibrations of the C=O, and C–O [19,26]. This confirms
3.1. Characterization the influence of temperature on the development of carbonyl groups
[27]. The NaOH treated AC at (30, 60 and 90 °C) have shown similar
3.1.1. Boehm’s titration patterns but the intensities are not well developed as compared to
The acidic functional groups are due to the presence of phenols, the HNO3 treated AC. The results also confirm the influence of tem-
lactols, lactones and carboxylic acids. The acidic groups are said to dif- perature and HNO3 on the development of surface functional groups
fer in their acidities that can be determined by reaction with 0.05 M such as nitrate, carboxyl, and carbonyl groups. In addition to the ab-
NaOH, Na2 CO3 and NaHCO3 respectively [17,22]. Phenols are neutral- sorption peaks already discussed, the peaks around 2200–2400 cm−1
ized with NaOH, while lactones and carboxylic acids are neutralized are almost certainly due to residual CO2 , either in the chamber or ad-
by Na2 CO3 , and NaHCO3 neutralizes only carboxyl groups. The raw sorbed to the surface. The peaks at 500–700 cm–1 are noisy and dif-
AC is acidic with relatively high amount of phenolic groups and total ficult to attribute.
acidic groups of 1.71 mmol/g. However, upon treatment with NaOH
at 30 °C the total acidic groups decreased to 1.23 mmol/g but the
3.1.3. SEM/EDX result on the treated AC
phenolic groups were greater than that of the raw samples (Table 1).
The increase in phenolic groups can be attributed to the deforma- The SEM/EDX result of the treated AC is depicted in Fig. 2. The
tion of the lactone groups resulting in the formation of more phenolic treatment with HNO3 at 30, 60, and 90 °C showed the presence of car-
bon and oxygen as the only elemental composition of the adsorbents
groups [23]. This can be clearly observed by comparing the amount
of the lactone groups on the base-treated ACs with that of the acid- and they showed a significant improvement trend in the weight% of
treated ACs. Moreover, for the AC treated with NaOH at 90 °C, the oxygen from 11.01 to 11.86 and finally 14.39 meaning that oxygen con-
taining functional groups kept on increasing with temperature upon
carboxylic groups were reduced due to its significant deformation;
acid treatment. This result is in accordance with the result obtained
but only the phenolic and lactonic groups were determined which
contributed less acidity. The total basic functional groups on the sur- from the FT-IR, Boehm’s titration experiment and XRD results. EDX
face of the AC-NaOH 30 °C were higher in all the samples modified. results on the base treated samples showed the presence of carbon,
oxygen and sodium in different proportions. Increase in treatment
Although the influence of the temperature was not that significant,
it could be noticed that the treatment at 60 °C seems to be the best temperature shows significant reduction in the wt. % of oxygen from
temperature for NaOH modification giving the highest surface acidic 18.33 to 17.72 and finally 13.63 for AC-NaOH 30 °C, 60 °C and 90 °C
functional groups (1.72 mmol/g) and the lowest surface basic groups respectively.
(1.98 mmol/g). On the other side, the increase in the total acidity
of the acid-treated ACs can be ascribed to the introduction of the 3.1.4. XRD patterns of the treated AC
new acidic functional groups on the surface of the ACs [17,22]. Total X-ray diffraction was conducted for the raw obtained AC, base-
acidic functional groups increased with treatment temperature from treated AC and acid-treated AC samples (Figs. 3 and 4). The X-ray
T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468 463

Fig. 1. FTIR spectra of raw AC and AC samples treated with sodium hydroxide or nitric acid at various temperatures.

diffraction profiles show two broad diffraction peaks of graphitic car- can be seen that the removal is very rapid, nearly 95% of all the ad-
bon with maxima at around 25° and 42° two theta. The two peaks sorption process occur in the first 5 min of the process. This is because
are indexed to (002) and (100) or (101) planes of crystalline hexag- of the available free sites for adsorption of the sulfur compounds at
onal graphite lattice respectively (JCPDS card no. 41-1487) [28,29]. the beginning of the adsorption process. Equilibrium was achieved
The relative intensities increased upon treatment signifying the im- at around 30 min of adsorption process and afterwards the adsorp-
provement in the degree of graphitization structure and crystallinity tion of the refractory sulfur compounds was noticed to decline due to
of the treated AC in addition to the removal of amorphous carbons the desorption process of the analytes from the adsorbent. The size
[7,30]. The peak, at approximately 25° two theta, was broadened on of the adsorbent, concentration of the sulfur compounds, degree of
increase of temperature, and this change in the width of the peak mixing, the affinity of the adsorbent to the sulfur compounds and the
can be attributed to the widening or expansion of the graphite lay- diffusion coefficient of the adsorbent in bulk and solid phases play
ers in the individual crystalline (disintegration to relatively smaller a cardinal role in the percentage of sulfur compounds adsorbed on
crystallites). It can be seen in Fig. 3 that the intensity of the peak, at the surface of the adsorbent [34]. The adsorption capacity followed
approximately 25° two theta, of AC treated with the base decreased the order: DBT > BT > T. The vast majority of the area of the sam-
with increasing temperature from 60 to 90 °C indicating the induced ples is mesoporous, which makes it easier for adsorption at the initial
irregularity into the layered graphitic structure of the AC after the ac- time, however, when the contact time increases the refractory sulfur
tivation process [31]. compounds are faced with much larger resistance to pass through the
micropores.

3.1.5. Brunauer–Emmett–Teller (BET) surface area and porosity studies


3.2.2. Influence of adsorbent dosage on the performance of
Nitrogen adsorption desorption isotherm is a useful tool in the
AC-HNO3 -90 °C and AC-NaOH-60 °C
characterization of carbon materials. AC samples express Type I
Adsorbent dosage is an important parameter in adsorptive studies
isotherm according to Brunauer et al., 1940 [32,33], See Fig. S1 in
because it determines the maximum adsorptive capacity of a given
the Supporting information file. A hysteresis loop can be observed in
adsorbent using a starting concentration of the analyte of interest.
all the samples at the relative high pressure indicating the presence
Best NaOH and HNO3 treated samples were selected for testing the ef-
of mesopores. As shown in Table 2, the surface area and pore vol-
fect of adsorbent dosage. The effect of AC-HNO3 -90 °C and AC-NaOH-
umes were significantly enhanced upon NaOH and HNO3 treatments.
60 °C dosages on the uptake of refractory sulfur compounds (Thio-
The N2 adsorption capacity followed the order: Raw < AC-NaOH
phene 52 ppm, BT 50 ppm and DBT 51 ppm at room temperature)
30 °C < AC-NaOH 90 °C < AC-NaOH 60 °C < AC-HNO3 -30 °C < AC-
were studied and the results are shown in Fig. 6a–b. The mass of the
HNO3 -60 °C < AC-HNO3 -90 °C. Temperature also enhanced the de-
adsorbent was varied from 0.1 g to 0.5 g and the amount of refractory
velopment of the surface area in the acid treated AC. The change in
sulfur compounds adsorbed was found to be increasing with adsor-
surface area caused by different treatments is summarized in Table 2.
bent dosage. The increase in adsorption capacity of sulfur compounds
It can be clearly seen that AC-NaOH-60 °C sample has the highest
with dosage is expected because more active surface areas and sites
surface area (369 m2 /g), pore volume (0.69 cm3 /g) and pore size
for sulfur adsorption are introduced in the medium [35,36]. The re-
(75.03 Å) in the base treated AC. However, AC-HNO3 -90 °C has the
moval efficiency after 0.25 g was less, becoming almost constant after
highest surface area (473.35 m2 /g) and pore volume (0.70 cm3 /g) of
0.5 g; due to this, we fixed the rest of the experiments at an adsorbent
all the samples used which confirms the significance of acid treat-
dosage of 0.5 g. The trend of adsorption of the refractory sulfur com-
ment of AC in the development of surface area and porosity. The re-
pounds followed the order of DBT > BT > T. DBT was removed higher
sults are in agreement with the results obtained from the XRD, FT-IR
than all other refractory sulfur compounds which can be attributed to
and Boehm’s experiment.
the dispersive interactions formed between the AC and DBT and due
to the presence of double benzene rings in DBT [37,38].
3.2. Adsorption evaluation
3.2.3. Effect of treatment reagent
3.2.1. Contact time The raw and treated AC both with NaOH and HNO3 were com-
The influence of contact time on the efficiency of AC-HNO3 -90 °C pared for their desulfurization efficiency. 0.5 g of each sample was
on the removal of refractory sulfur compounds is depicted in Fig. 5. It taken and added to 50 ppm mixture of thiophene, BT and DBT in
464 T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468

Fig. 2. SEM image and EDX spectrum with an elemental analysis table of the prepared carbon treated with nitric acid at 30 °C (a), 60 °C (b), 90 °C (c), sodium hydroxide at 30 °C
(d), 60 °C (e), 90 °C (f).
T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468 465

Fig. 3. XRD pattern of Raw AC and NaOH treated AC at 30, 60 and 90 °C.

Fig. 4. XRD pattern of AC treated with HNO3 at 30, 60 and 90 °C.

Table 2
Surface area and pore volume analysis.

Adsorbents BET surface area (m2 /g) Pore volume (cm3 /g) Pore size (Å)

Raw AC 183.21 0.30 77.54


AC-NaOH 30 °C 220.48 0.44 79.83
AC-NaOH 60 °C 369.27 0.69 75.03
AC-NaOH 90 °C 268.11 0.47 70.47
AC-HNO3 30 °C 429.26 0.66 62.10
AC-HNO3 60 °C 454.80 0.67 59.60
AC-HNO3 90 °C 473.35 0.70 60.22

20 ml (85% hexane and 15% toluene). The set ups were stirred and in % transmittance in FT-IR. It also possesses the highest surface area
samples were taken after 30 mins and analyzed with GC-SCD. The and pore volume. It should be noted that the reagent type plays a
result is shown in Fig. 7 and the percentage removals of thiophene, role while the temperature has a subtle effect. The adsorbents per-
BT and DBT from the raw and treated samples at various temper- formance and adsorptive capacities followed the order: AC-HNO3 -
atures were reported. It can be clearly seen that raw sample per- 90 °C > AC-HNO3 -60 °C > AC-HNO3 -30 °C > AC-NaOH-60 °C > AC-
formed the least. AC-NaOH-60 °C showed the best performance in NaOH-90 °C > AC-NaOH-30 °C > Raw AC. This order is consistent
the base treated samples and this can be attributed to the presence with the increase in the surface area and pore volume of the ad-
of oxygen containing functional groups, high surface area, pore vol- sorbents and their total acidic groups [40,41]. This implies that both
ume, and pore size exhibited by the sample compared with other surface area and surface chemistry play an important role in adsorp-
base treated samples. AC-HNO3 -90 °C showed the best performance tion capacity. The adsorption capacities were calculated and listed in
regardless of treatment conditions because it showed the highest to- Table S2.
tal acidic groups, the largest intensity in XRD analysis, and the highest
466 T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468

Fig. 5. Influence of contact time on removal of thiophene, BT and DBT by AC-HNO3 -


90 °C (adsorbent dosage 0.5 g, 20 ml volume of model fuel containing 52 ppm thio-
phene, 50 ppm BT and 51 ppm DBT).
Fig. 7. Percentage removal of thiophene, BT, and DBT (initial concentration of 50 ppm
each) on acid and base treated AC samples after 30 min (dosage of each sample was
0.5 g in 20 ml initial solution volume) at 25 °C.

The percentage removal of the refractory sulfur compounds in all


the ACs followed the order: DBT > BT > T. The treated ACs showed at
least 80% percentage removal or greater of DBT and this is due to the
size of the molecule and acidic functional groups on the surfaces of
the larger pores which contributes to polar interactions [42,43]. Dis-
persive interactions between the delocalized π -electrons within the
benzene rings of DBT and the electron rich region of the nanoporous
carbon aromatic ring also play a major role in the adsorption. The re-
sults also confirm the influence of treatment conditions on the overall
performance in adsorptive studies.

3.3. Breakthrough curves

The performance of AC-NaOH-60 °C and AC-HNO3 -90 °C were fur-


ther studied in a fixed bed mode. Breakthrough curves were gener-
ated by plotting transient total sulfur concentration normalized by
the feed total sulfur concentration (C/C0 ) versus cumulative time. The
breakthroughs curves for thiophene, BT and DBT in the base and acid
treated samples are depicted in Fig. 8a–b. The starting concentrations
of thiophene, BT and DBT were 50 ppm, 51 ppm and 52 ppm respec-
tively. The acid treated sample showed higher breakthrough times
compared to the base treated sample. DBT was lowered to 0 ppm after
5 min of adsorption in all the samples tested; however, breakthrough
was not achieved even after 400 min of adsorption. This indicates the
suitability of acid treatment in the adsorption of this compound. A
breakthrough was achieved after 180 min for DBT using AC-NaOH-
60 °C. The breakthroughs of thiophene and BT in both AC-NaOH-60 °C
and AC-HNO3 -90 °C were achieved after 5 min and 10 min respec-
tively. For AC-NaOH-60 °C, thiophene returned to its initial concen-
Fig. 6. Influence of adsorbent dosage AC-HNO3 -90 °C (a) and AC-NaOH-60 °C (b) on
tration after the first 120 min while BT returned to its initial concen-
the adsorptive removal of thiophene, BT and DBT (initial concentrations 51, 50 and
52 ppm respectively, 20 ml volume of solution and 30 min contact time). tration after 50 min signifying no further adsorption that has taken
place after this time. This is in contrast with the AC-HNO3 -90 °C per-
formance, both thiophene and BT did not reach their initial concen-
The decrease in the efficiency of the AC-NaOH-90 °C compared tration even after 400 min of the column adsorption process. The ad-
with AC-NaOH-60 °C can be attributed to the increase in irregularity sorption capacities in column mode for all the adsorbents followed
of graphitic layer structures of AC treated with the base at 90 °C as the order DBT > BT > T. This means that AC treated with HNO3 is
it has been discussed under the section of XRD results (see Fig. 3). good in both batch and fixed bed mode.
On the other side, the adsorption efficiency of the AC treated with
acid increases with treatment temperature. This trend is in consis- 3.4. Regeneration of the used adsorbent
tence with the XRD patterns of these samples, which indicated the
improvement in the crystallinity and graphitic structure of the AC The synthesized activated carbon (AC-HNO3 -90C) was evalu-
(see Fig. 4). From this point of view, it is conceivable that increas- ated for its regeneration abilities. The regeneration was done by
ing behaviors of the adsorption by acid-treated AC could reflect the heating the adsorbent in air for 3 h at 350 °C. The temperature
contribution of π –π interaction between the basal plane of carbon is said to vaporize all the sulfur compounds from the adsorbent.
and the aromatic ring of the adsorbate [39,40]. As results showed in Fig. 9, the percentage removal of the sulfur
T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468 467

obtained for AC-HNO3 at 90 °C. The adsorption of dibenzothiophene,


benzothiophene and thiophene by the AC-HNO3 90 °C were higher
than that of the raw AC by about 2.2, 3.4 and 3.8 times respectively.
This can be explained according to Lewis acid–base theory, most thio-
phene sulfur compounds are Lewis base, which are easy to be ad-
sorbed at acid sites. The study has the advantages of solving dou-
bly environmental problems by converting such negative-value waste
tires to value-added activated carbon for adsorptive desulfurization
toward clean fuel.

Acknowledgments

The authors gratefully acknowledge the support provided by King


Fahd University of Petroleum & Minerals (KFUPM). This project was
funded by the National Plan for Science, Technology and Innova-
tion (MAARIFAH)—King AbdulAziz City for Science and Technology—
through the Science & Technology Unit at King Fahd University of
Petroleum & Minerals (KFUPM)—the Kingdom of Saudi Arabia, award
number (13-PET393-04).

Supplementary materials

Supplementary material associated with this article can be found,


in the online version, at doi:10.1016/j.jtice.2015.11.008.

References

[1] Khan NA, Hasan Z, Jhung SH. Adsorptive removal of hazardous materials using
metal–organic frameworks (MOFs): a review. J Hazard Mater 2013;244–245:444–
56.
[2] Srivastava VC. An evaluation of desulfurization technologies for sulfur removal
Fig. 8. Breakthrough curve for AC-NaOH-60 °C (a) and AC-HNO3 -90 °C (b). from liquid fuels. RSC Adv 2012;2:759–83.
[3] Pawelec B, Navarro RM, Castano P, Alvarez-Galvan MC, Fierro JLG. Energy Fuels
2009;23:1364–72.
[4] Hasan Z, Jeon J, Jhung SH. Oxidative desulfurization of benzothiophene and thio-
phene with WOx/ZrO2 catalysts: effect of calcination temperature of catalysts.
J Hazard Mater 2012;205–6:216–21.
[5] Rodríguez-Cabo B, Rodríguez H, Rodil E, Arce A, Soto A. Extractive and oxidative-
extractive desulfurization of fuels with ionic liquids. Fuel 2014;117:882–9.
[6] Martinez I, Santos VE, Alcon A, Garcia-Ochoa F. Enhancement of the biodesulfu-
rization capacity of Pseudomonas putida CECT5279 by co-substrate addition. Pro-
cess Biochem 2015;50:119–24.
[7] Ma X, Yang H, Yu L, Chen Y, Li Y. Preparation, surface and pore structure of high
surface area activated carbon fibers from bamboo by steam activation. Materials
2014;7:4431–41.
[8] Zhang HX, Huang HL, Li CX, Meng H, Lu YZ, Zhong CL. Adsorption behavior of
metal–organic frameworks for thiophenic sulfur from diesel oil. Ind Eng Chem
Res 2012;51:12449–55.
[9] Srivastava VC. An evaluation of desulfurization technologies for sulfur removal
from liquid fuels. RSC Adv 2012;2:759–83.
[10] Martinez I, Santos VE, Alcon A, Garcia-Ochoa F. Enhancement of the biodesulfu-
rization capacity of Pseudomonas putida CECT5279 by co-substrate addition. Pro-
cess Biochem 2015;50:119–24.
[11] Xiao J, Li Z, Liu B, Xia Q, Yu M. Adsorption of benzothiophene and dibenzothio-
phene on ion-impregnated activated carbons and ion-exchanged Y zeolites. En-
ergy Fuels 2008;22:3858–63.
[12] Chiang HL, Huang CP, Chiang PC, You JH. Effect of metal additives on the physico-
Fig. 9. Regeneration efficiency of AC-HNO3 -90 °C on the adsorption of sulfur
chemical characteristics of activated carbon exemplified by benzene and acetic
compounds. acid adsorption. Carbon 1999;37:1919–28.
[13] Jiang Z, Liu Y, Sun X, Tian F, Sun F, Liang C, et al. Activated carbons chem-
ically modified by concentrated h2 so4 for the adsorption of the pollutants
compounds showed some decrease in the adsorptive capacity of the from wastewater and the dibenzothiophene from fuel oils. Langmuir
adsorbent. However, it still showed great promise in the removal 2003;19:731–6.
[14] Jung BK, Jhung SH. Adsorptive removal of benzothiophene from model fuel, using
DBT even after three cycles of regeneration. The order of desulfur- modified activated carbons, in presence of diethylether. Fuel 2015;145:249–55.
ization after three regeneration cycles followed the sequence: DBT [15] Boehm HP. Surface oxides on carbon and their analysis: a critical assessment.
(59%) > T (6%) > BT (5%). Carbon 2002;40:145–9.
[16] Moreno-Castilla C, Lo´pez-Ramo´n MV, Carrasco-Marin F. Changes in
surface chemistry of activated carbons by wet oxidation. Carbon 2000;38:1995–
4. Conclusion 2001.
[17] Yu V, Qiu JS, Sun YF, Li XH, Chen G, Zhao ZB. Adsorption removal of thiophene and
dibenzothiophene from oils with activated carbon as adsorbent: effect of surface
Waste tires can be effectively used as a raw material for the
chemistry. J. Porous Mater. 2008;15:151–7.
preparation of high-surface area AC after treatment. Upon treatment [18] Boehm HP. Chemical identification of surface groups. Adv Catal. 1966;16:179.
with NaOH or HNO3 , the surface area and functional groups, such [19] Chen PJ, Shunnian W. Acid/base-treated activated carbons: characterization of
as carboxyl, lactone and phenol groups were enhanced. The effect functional groups and metal adsorptive properties. Langmuir 2004;20:2233–42.
[20] Oickle AM, Goertzen SL, Hopper KR, Abdalla YO, Andreas HA. Standardization of
of the treatment temperature on the adsorption performance was the Boehm titration: Part II. Method of agitation, effect of filtering and dilute
limited. The highest total acidic functional groups of 2.39 mmol/g was titrant. Carbon 2010;48:3313–22.
468 T.A. Saleh, G.I. Danmaliki / Journal of the Taiwan Institute of Chemical Engineers 60 (2016) 460–468

[21] Goertzen SL, Theriault KD, Oickle AM, Tarasuk AC, Andreas HA. Standardization [33] Deng H, Yang L, Tao G, Dai J. Preparation and characterization of activated
of the Boehm titration. Part I. CO2 expulsion and endpoint determination. Carbon carbon from cotton stalk by microwave assisted chemical activation: application
2010;48:1252–61. in methylene blue adsorption from aqueous solution. J Hazard Mater 2009;166(2–
[22] Boehm HP. Surface properties of carbons. Studies Surface Sci Catal 1989;48:145– 3):1514–21.
57. [34] Srivastav A, Srivastava VC. Adsorptive desulfurization by activated alumina. J Haz-
[23] Park S, Seo M-K. Interface Science and Composites. Academic Press; 2011. p. 197– ard Mater 2009;170:1133–40.
8. [35] Xu X, Zhang S, Li P, Shen Y. Adsorptive desulfurization of liquid Jet-A fuel at ambi-
[24] Saleh TA. The influence of treatment temperature on the acidity of MWCNT oxi- ent conditions with an improved adsorbent for on-board fuel treatment for SOFC
dized by HNO3 or a mixture of HNO3 /H2 SO4 . Appl Surface Sci 2011;257(17):7746– applications. Fuel Process Technol 2014;124:140–6.
51. [36] Hernández-Maldonado AJ, Yang RT. Desulfurization of transportation fuels by ad-
[25] Guo JX, Liu XL, Luo DM, Yin HQ, Li JJ, Chu YH. Influence of calcination tem- sorption. Catal Rev 2004;46(2):111–50.
peratures on the desulfurization performance of fe supported activated carbons [37] Bu J, Loh G, Gwie CG, Dewiyanti S, Tasrif M, Borgna A. Desulfurization of diesel fu-
treated by HNO3 . Ind Eng Chem Res 2015;54:1261–70. els by selective adsorption on activated carbons: competitive adsorption of poly-
[26] Wu XH, Hong XT, Luo ZP, Hui KS, Chen HY, Wu JW. The effects of surface mod- cyclic aromatic sulfur heterocycles and polycyclic aromatic hydrocarbons. Chem
ification on the super-capacitive behaviors of novel mesoporous carbon derived Eng J 2011;166:207–17.
from rod-like hydroxyapatite template. Electrochim Acta 2013;89:400–6. [38] Takahashi A, Yang FH, Yang RT. New sorbents for desulfurization by
[27] Tazibet S, Boucheffa Y, Lodewyckx P. Heat treatment effect on the textural, hy- π -complexation: thiophene/benzene adsorption. Ind Eng Chem Res.
drophobic and adsorptive properties of activated carbons obtained from olive 2002;41(10):2487–96.
waste. Microporous Mesoporous Mater 2013;170:293–8. [39] Wang Y, Yang RT. Desulfurization of liquid fuels by adsorption on carbon-
[28] Babu VS, Seehr MS. Modeling of disorder and X-ray diffraction in coal-based based sorbents and ultrasound-assisted sorbent regeneration. Langmuir
graphitic carbons. Carbon 1996;34(10):1259–65. 2007;23(7):3825–31.
[29] Ji Y, Li T, Zhu L, Wang X, Lin Q. Preparation of activated carbons by microwave [40] Saleh TA. Nanocomposite of carbon nanotubes/silica nanoparticles and their use
heating KOH activation. Appl Surface Sci 2007;254:506–12. for adsorption of Pb (II): from surface properties to sorption mechanism. Desalin
[30] Viswanathan B, Neel PI, Varadarajan TK. Methods of activation and specific ap- Water Treatment 2015:1–15. doi:10.1080/19443994.2015.1036784.
plications of carbon materials. Nat Centre Catal Res., Indian Inst Technol Madras [41] Saleh TA. Isotherm, kinetic, and thermodynamic studies on Hg(II) adsorption from
Chennai 2009;600:036. aqueous solution by silica- multiwall carbon nanotubes. Environ Sci Pollut Res
[31] Kim BJ, Lee YS, Park SJ. A study on pore-opening behaviors of graphite nanofibers 2015;22(21):16721–31.
by a chemical activation process. J Colloid Interface Sci 2007;306:454. [42] Triantafyllidis KS, Deliyanni EA. Desulfurization of diesel fuels: adsorption of 4,
[32] Brunauer S, Deming LS, Demomg WE, Teller DE. On a theory of the van der Waals 6-DMDBT on different origin and surface chemistry nanoporous activated car-
adsorption of gases. J Am Chem Soc. 1940;62:1723–32. bons. Chem Eng J 2014;236:406–14.
[43] Saleh TA, Alhooshani KR, Abdelbassit MSA. Evaluation of AC/ZnO composite for
sorption of dichloromethane, trichloromethane and carbon tetrachloride: kinet-
ics and isotherms. J. Taiwan Inst Chem Eng 2015;55:159–69.

Вам также может понравиться