Study and optimization of adsorption of sulfur compounds present in fuel

Biswajit Saha, Anshu Singh and Sonali Sengupta*
Chemical Engineering Department IIT, Kharagpur, India. E-mail: sonalis.iitkgp@gmail.com

Received 24th June 2016 , Accepted 2nd August 2016

First published on 2nd August 2016


Abstract

The present work investigates the performance of fly ash, coal dust, bentonite, laterite and sodium zeolite as adsorbents for desulfurization of synthetic fuel by batch adsorption experiments at 50 °C and under atmospheric pressure. A series of synthetic fuel samples containing known concentrations of thiophenol, benzothiophene and di-benzothiophene, as singular and multi component mixtures in isooctane, were prepared. The performance of coal dust was found to be the best among all the adsorbents mentioned, in removal of both single and multi sulfur compounds from synthetic fuel. The adsorbents were characterized by BET surface area analysis, XRF, XRD, FTIR, SEM and NH3-TPD. The effects of parameters such as stirrer speed, adsorbent particle size, temperature, adsorbent amount, initial sulfur concentration and contact time between adsorbate and adsorbent were investigated for desulfurization of synthetic fuel using coal dust as adsorbent. The sulfur adsorption equilibrium with this adsorbent was fitted to Langmuir, Freundlich, BET, Halsey and Temkin isotherms. Finally Response Surface Methodology (RSM) was applied for optimizing the adsorption process parameters. The four factor Box–Behnken design (BBD) was performed, aimed at developing second order polynomial models at the optimum conditions. The objective was to find out how the output, sulfur adsorption per gram of adsorbent (qt), is related to the inputs, initial sulfur concentration, adsorbent amount, time and temperature, in order to get a clear picture for further research.


Introduction

In the last twenty five years, the energy usage of the world has increased to an average of 404 quadrillion BTUs per annum, where more than 40% of this energy comes from petroleum.1,2 Gasoline, jet fuel and diesel are the three major types of transportation fuels which contain significant amounts of sulfur and combustion of these fuels results in the production of sulfur-oxide gases which cause severe harm to human health and the environment.3,4 In India, the existing two tier system for liquid fuel quality called Bharat Stage specification introduced the BS(IV) grade diesel in April 2010, which required a reduction of organic sulfur content from 350 ppm to 50 ppm.5 The conventional refinery-practiced desulfurization technology is hydrodesulfurization (HDS), which is an intensive process, and hence requires a high investment cost. Moreover, this process significantly reduces octane number by saturating alkenes and arenes by hydrogen used at high temperature and pressure.6 Therefore, there is an urgent need to find out new and alternative pathways for desulfurization which would be cost effective, efficient and can meet the environmental regulations. Adsorptive desulfurization is one of the alternatives to HDS used for production of ultra clean fuel. Adsorption is a mass transfer process where the molecules in free phase become bound to a surface by intermolecular forces.7 This process would be accomplished relatively at lower temperature and pressure. Several research works have been done on adsorptive desulfurization using various adsorbents. Shi et al.8 studied the adsorptive desulfurization of dibenzothiophene with 95% removal by using mesoporous carbon. Adekanmi et al.7 reported the adsorption efficiency of manganese-di-oxide and zinc-di-oxide on desulfurization of crude oil, where manganese dioxide showed good result compared to zinc dioxide with 0.816 g of sulfur adsorption per gram of adsorbent. Das Gupta et al.5 showed the adsorption efficiency of nickel based adsorbents (NiMCM-41) on desulfurization of diesel oil by bringing down the sulfur concentration from 450 ppm to 50 ppm. Adsorptive desulfurization of kerosene and diesel oil by using zinc impregnated montmorillonite clay as adsorbent has been reported which removed 76 and 77% sulfur from kerosene and diesel oil respectively.9 The adsorption efficiency of nano copper oxide on diesel fuel was reported to be 70% by Khodadadi et al.10 Kumar et al.11 studied the adsorptive desulfurization of di-benzothiophene in isooctane phase by using zirconia based adsorbents with 53 mg g−1 to 60 mg g−1 sulfur adsorption.

In the present study, the performance of fly ash, coal dust, bentonite, laterite and sodium zeolite as adsorbents for desulfurization of synthetic fuel by batch process has been studied. Table 1 showed a comparative study of efficiency of our adsorbents and adsorbents which were found in literature. The effect of process variables such as time, temperature, initial sulfur concentration, stirrer speed, adsorbent amount and adsorbent particle size on the efficiency of adsorptive desulfurization has been studied with the best adsorbent. Finally Langmuir, Freundlich, BET, Halsey and Temkin adsorption isotherm models were fitted with adsorption equilibrium data and four factor BBD (Box–Behnken design) for optimizing the adsorption process parameters was implemented.

Table 1 Comparison of efficiency of some adsorbents from literature and present ones
Ref no. Model fuel Adsorbent Adsorption efficiency
1 DBT in n-heptane Magnetite nanoparticle loaded bentonite 95 mg g−1, sulfur adsorbed
4 JP-8 fuel Ag-mesoporous silica nanoparticle 32.6 mg g−1 sulfur adsorbed
11 DBT in isooctane Zirconia based adsorbent 55 mg g−1 sulfur adsorbed
12 (a) Thiophene in n-hexane Zeolites from fly ash (a) 67.8% sulfur removal
(b) BT in n-hexane (b) 94.2% sulfur removal
13 DBT in n-hexane Activated alumina 16 mg g−1 sulfur adsorbed
Our research adsorbate (a) Thiopheneol, BT & DBT in isooctane Untreated coal dust (a) 84.14 mg sulfur per g or 64% sulfur removal
(b) Gasoline range fuel (b) 82% sulfur removal


Materials and method

Materials

Bentonite powder and thiophenol were purchased from Loba Chemie Pvt Ltd, Mumbai, India. Benzothiophene was procured from Himedia Lab. Pvt Ltd, Mumbai, India. Dibenzothiophene was purchased from Merck Schuchardt OHG, Germany. Fly ash was obtained from Kolaghat thermal power plant, India. Laterite soil was taken from Midnapur District, West Bengal, India. Sodium zeolite was purchased from SRL Pvt Ltd, Kolkata, India. Coal was obtained from Jharkhand Coal Suppliers, India, which was made dust of the required size in our laboratory.

Model fuel preparation

Five types of model fuels were prepared by adding sulfur compounds such as, thiophenol, benzothiophene and dibenzothiophene in isooctane at different concentrations. The compositions of model fuels are as follows; model fuel 1 [MF-1]: thiophenol (300 ppm) in iso-octane, model fuel 2 [MF-2]: benzothiophene (300 ppm) in iso-octane, model fuel 3 [MF-3]: dibenzothiophene (300 ppm) in iso-octane, model fuel 4 [MF-4]: mixture of thiophenol (300 ppm), benzothiophene (300 ppm) and di-benzothiophene (300 ppm) in iso-octane with total sulfur concentration of 900 ppm, model fuel 5 [MF-5]: mixture of thiophenol (300 ppm), benzothiophene (300 ppm) and di-benzothiophene (300 ppm) with total sulfur content of 900 ppm added with two aromatic compounds (benzene and toluene) with total aromatic content of 600 ppm.

Batch adsorption experiment

A typical adsorption experiment was conducted using 40 mL model fuel and 0.27 g of adsorbent, taken in a three necked glass flask housed in a water bath whose temperature was maintained at 50 °C with a digital temperature controller cum indicator with an accuracy of ±1 °C. The experiment was conducted for 2 h under stirring at 1000 rpm and samples were withdrawn at every 10 min intervals. The samples were filtered to remove any trace of adsorbent particle before they were analyzed in HPLC [Perkin Elmer, Series 200] with reversed phase Agilent SB C-18 column and a Perkin Elmer Series 200 UV/VIS detector set at 254 nm. The mobile phase used was 90% methanol in water.

Adsorbent characterization

Adsorbents were characterised by BET surface area analyser, (Model: AS1 MP/Chemi-LP, USA) instrument, X-ray fluorescence spectroscope (XRF), (Model Panalytical, Axios, Netherlands), X-ray diffraction (XRD), (Model Panalytical 3050/60) with beryllium filtered Cu Kα (1.5418 A), operating at 40 kV and 30 mA, scanning electron micrograph (SEM), (Model Zeol and Zeiss with Oxford EDS detector). NH3-TPD (Model Chembet-3000) and Fourier Transform Infrared Spectroscopy (FTIR), (Model Perkin Elmer spectrum 100) in the spectral range of 400 to 4000 cm−1.

The sulfur adsorption per unit mass of the adsorbent (mg g−1) was calculated using eqn (1) and (2), as,

 
image file: c6ra16367j-t1.tif(1)
 
image file: c6ra16367j-t2.tif(2)

Results and discussion

XRD analysis of adsorbents

Powder X-ray diffraction patterns of coal dust, bentonite, fly ash, laterite soil and sodium zeolite are shown in Fig. 1. In coal dust the characteristic peaks of silicon-di-oxide (2θ = 12.469°) can be assigned as (220) with the standard pattern (JCPDS no. 89-1421), aluminium oxide (2θ = 31.052°, 50.237°) can be assigned as (400), (331) with the standard pattern (JCPDS nos 79-1559 & 79-1558), sulphur-tri-oxide (2θ = 24.969°) can be assigned as (121) with the standard pattern (JCPDS no. 72-1664) and calcium oxide (2θ = 26.740°) can be assigned with the standard pattern (JCPDS no. 17-0912). In bentonite the characteristic peaks of sodium oxide (2θ = 19.855°) which can be assigned as (001) with the standard pattern (JCPDS no. 74-0111), silicon-di-oxide (2θ = 26.674°, 50.231°, 62.361°) is as (321), (063), (1310) with the standard pattern (JCPDS nos 89-1813, 89-0735 & 89-1668). In fly ash the characteristic peaks of silicon-di-oxide (2θ = 24.07°, 29.31°, 31.27°, 44.454°) which can be assigned as (314), (132), (015), (201) with the standard pattern (JCPDS nos 88-1461, 89-0735 & 89-3433), aluminium oxide (2θ = 39.444°) can be assigned as (206) with the standard pattern (JCPDS no. 88-1609), ferrous oxide (2θ = 21.059°) can be assigned as (021) with the standard pattern (JCPDS no. 89-6466), ferric oxide (2θ = 48.465°) can be assigned as (334) with the standard pattern (JCPDS no. 89-5894) and magnesium oxide (2θ = 45.89°) can be assigned as (211) with the standard pattern (JCPDS no. 76-1363). In laterite soil the characteristic peaks of ferric or iron oxide (2θ = 26.674°, 50.170°) which can be assigned as (022), (3110) with the standard pattern (JCPDS nos 89-6466 & 89-5894), silicon-di-oxide (2θ = 21.359°) can be assigned as (006) with the standard pattern (JCPDS no. 89-1349). In zeolite the characteristic peaks of silicon-di-oxide (2θ = 21.554°, 23.893°, 27.002°, 29.809°, 34.054°) can be assigned as (111), (222), (402), (314), (512) with the standard pattern (JCPDS nos 89-3435, 88-1155, 89-1350, 89-5416 & 89-1349), aluminium oxide (2θ = 52.503°) can be assigned as (024) with the standard pattern (JCPDS no. 75-1862), sodium oxide (2θ = 69.081°) can be assigned with the standard pattern (JCPDS no. 15-0068).
image file: c6ra16367j-f1.tif
Fig. 1 XRD peak of all adsorbent. image file: c6ra16367j-u1.tif ferrous oxide (Fe2O3); image file: c6ra16367j-u2.tif sulfur-tri-oxide (SO3); image file: c6ra16367j-u3.tif aluminium oxide (Al2O3); image file: c6ra16367j-u4.tif sodium oxide (Na2O); image file: c6ra16367j-u5.tif calcium oxide (CaO); image file: c6ra16367j-u6.tif ferric oxide (Fe3O4); image file: c6ra16367j-u7.tif magnesium oxide (MgO); * silicon-di-oxide (SiO2).

Scanning electron micrograph analysis

Fig. 2 shows the SEM image of coal dust sample before and after adsorption.
image file: c6ra16367j-f2.tif
Fig. 2 SEM images of coal before and after adsorption.

FTIR analysis of adsorbents

Fig. 3 shows the FTIR analysis of coal dust sample before and after adsorption.
image file: c6ra16367j-f3.tif
Fig. 3 FTIR of fresh and used coal.

BET surface area analysis

BET surface area of all the adsorbents were determined by BET analysis and tabulated in Table 2. It has been observed that coal dust possesses largest surface area (149.4 m2 g−1) compared to the other adsorbents.
Table 2 BET surface area of adsorbents
Adsorbent BET surface area (m2 g−1)
Coal dust 149.40
Laterite 60.00
Bentonite 33.06
Sodium zeolite 25.24
Fly ash 19.25


X-ray fluorescence analysis

XRF of the samples tabulated in Table 3, shows different compositional analysis of all the adsorbents.
Table 3 Elemental analysis (wt%) of all adsorbents using XRF
Adsorbents SiO2 Al2O3 SO3 CaO MgO Na2O Fe2O3
Coal dust 42.43 22.169 9.571 5.099 0 0 13.491
Bentonite 53.28 18.27 0 0 3.027 7.311 12.975
Fly ash 25.85 10.68 0 47.34 7.604 0 4.074
Laterite soil 30.551 17.28 0 0 0 0 49.37
Sodium zeolite 41.58 32.96 0 0 0 23.67 0


NH3-temperature programmed desorption analysis

NH3-TPD of all the adsorbent samples was done to determine their acidic nature. It is observed that coal dust possesses a significant acidic nature with acidity 8.22 mmol g−1. Although fly ash shows better acidity (10.65 mmol g−1), yet because of its very low surface area, its acidic nature cannot prominently take part in attracting large amount of adsorbate compared to coal dust. Table 4 and Fig. 4 show the acidity of coal dust before adsorption and after adsorption with MF-4 and it is observed that there is a drastic decrease of acidity after adsorption (8.22 to 6.22 mmol g−1).
Table 4 NH3-TPD of coal before and after adsorption
Adsorbent Acidity (mmol g−1)
Coal before adsorption 8.22
Coal after adsorption 6.22



image file: c6ra16367j-f4.tif
Fig. 4 NH3-TPD curve of coal before and after adsorption.

Adsorptive desulfurization of MF-1, MF-2, MF-3, MF-4 and MF-5 using different adsorbents

Fig. 5A–E showed the adsorptive desulfurization of MF-1, 2, 3, 4 and 5 by using coal dust, bentonite, fly ash, laterite and zeolite. The highest sulfur adsorption per gram adsorbent (mg g−1) were 44.4, 29.3, 18.66, 85 and 76.3 for MF-1, 2, 3, 4 and 5 respectively using coal dust, 43.92, 25.9, 17.9, 84.1 and 74 for MF-1, 2, 3, 4 and 5 respectively using bentonite, 43.86, 21.8, 2.66, 64.45 and 59.2 for MF-1, 2, 3, 4 and 5 respectively using fly ash, 43.43, 18.7, 1.93, 58.6 and 47.7 for MF-1, 2, 3, 4 and 5 respectively using laterite and 43.29, 18.8, 1.91, 59.2 and 48.8 for MF 1, 2, 3, 4 and 5 respectively using sodium zeolite.
image file: c6ra16367j-f5.tif
Fig. 5 (A) Adsorptive desulfurization of MF-1 (thiophenol: 300 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel, stirrer speed: 1000 rpm). (B) Adsorptive desulfurization of MF-2 (benzothiophene: 300 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel, stirrer speed: 1000 rpm). (C) Adsorptive desulfurization of MF-3 (dibenzothiophene: 300 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel, stirrer speed: 1000 rpm). (D) Adsorptive desulfurization of MF-4 (sulfur: 900 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel, stirrer speed: 1000 rpm). (E) Adsorptive desulfurization of MF-5 (sulfur: 900 ppm; aromatic: 600 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel, stirrer speed: 1000 rpm).

Table 5 showed the comparative studies of all adsorbents on MF-4 and MF-5 to find out the influence of aromatic compounds present in the model fuel in MF-5. It has been observed that, presence of aromatic compounds decreases the overall removal of sulfur compounds in MF-5 from 64 to 57% for coal dust, 63 to 55% for bentonite, 48 to 44.4% for fly ash, 44 to 35.7% for laterite and 44.4 to 36.6% for zeolite compared to MF-4. This may be due to the competition between sulfur compounds and aromatic compounds to avail the active sites of adsorbents.

Table 5 Efficiency of all adsorbents on model fuel 5 (MF-5)a
Adsorbent MF-4 MF-5
qt (mg g−1) R (%) qt (mg g−1) R (%)
a qt (mg g−1) sulfur adsorbed per gram of adsorbent. R (%) is percentage of adsorption of sulfur compounds.
Coal 85 64 76.3 57
Bentonite 84.1 63 74 55
Fly ash 64.45 48 59.2 44.4
Laterite soil 58.6 44 47.7 35.7
Zeolite 59.2 44.4 48.8 36.6


The adsorption result shows that coal dust is the best adsorbent and the effect of different process parameters of adsorptive desulfurization has been studied using coal dust. In this work, MF-4 was taken as a fuel sample of choice. The reason for best performance of coal dust as adsorbent, compared to other four adsorbents is its highest surface area which causes larger amount of mass transfer between adsorbate and adsorbent.

Parametric study on adsorptive desulfurization of model fuel 4

MF-4 was abruptly chosen as a sample of choice for parametric study with coal dust as adsorbent.

Effect of stirrer speed variation

Fig. 6 shows the effect of stirrer speed on adsorption efficiency (mg g−1) of coal dust. It is observed that increase in stirrer speed from 400 rpm to 1000 rpm increases the adsorption efficiency from 53.87 to 84.29. This may be due to increase in the mass transfer between adsorbate and adsorbent. But when the stirrer speed was increased from 1000 rpm to 1200 rpm the change in adsorption efficiency was not significant, which may be due to the reason that resistance to mass transfer has no prominent influence on adsorption at that range of speed.14 Hence, all the subsequent parametric studies were done at 1000 rpm.
image file: c6ra16367j-f6.tif
Fig. 6 Stirrer speed variation (total sulfur in model fuel: 900 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel).

Effect of adsorbent particle size variation

Fig. 7 shows the adsorbent particle size effect, the size ranging from 0.125 mm to 0.052 mm on its adsorption efficiency (mg g−1). It has been observed from the figure that, as particle size decreases from 0.125 to 0.052 mm, the efficiency of adsorption increases. Due to the decrease in particle size, the surface area is expected to increase, which in turn may increase the efficiency from 58.4 to 86.34.14 Moreover, when the size of adsorbent is reduced to less than 0.052 mm, the adsorption capacity is not significantly increased. This may be due to the fact that, the increase in surface area after that size was not significant; hence the increase in adsorption capacity was not prominent. The rest of the experiments were performed with 0.052 mm size as internal mass transfer resistance may be negligible at this stage.
image file: c6ra16367j-f7.tif
Fig. 7 Adsorbent particle size variation (total sulfur in model fuel: 900 ppm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel; stirrer speed: 1000 rpm).

Influence of initial sulfur concentration

The adsorption experiments were performed with different initial sulfur concentration of the model fuel, ranging from 300 to 1800 ppm and Fig. 8 shows the effect of initial sulfur concentration on the adsorption capacity of coal dust. As the initial concentration of sulfur increases, intake capacity of adsorbent or adsorption rate also increases at the same time. Increase in initial concentration leads to increase in driving force (concentration gradient) for mass transfer. From the figure it is observed that, sulfur removal on the adsorption increases with time, reaches its maximum and then forms a plateau with a little decreasing trend for all the concentration. The maximum removal of sulfur reaches faster for higher initial concentration of sulfur containing model fuel compared to the lower ones, and after that, no improvement in adsorption with time was observed. The maximum removal of sulfur for initial concentration of 1500 and 1800 ppm obtained are 105 mg g−1 at 40 min and 120 mg g−1 at 30 min respectively. This may be due to saturation of active sites of coal dust or pore blocking/active site blocking at higher sulfur concentration.1,14
image file: c6ra16367j-f8.tif
Fig. 8 Initial sulfur concentration variations (stirrer speed: 1000 rpm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel).

Effect of operating temperature

Temperature for adsorption was varied from 30 to 70 °C and the effect of temperature on adsorption (mg g−1) is shown in Fig. 9. It has been observed from the figure that, the adsorption of sulfur increases with increase in temperature from 30 to 50 °C in a regular manner, after that it decreases as the temperature increases from 50 to 70 °C. This nature may be explained by the logic that, with increase in temperature, pore size of adsorbent as well as diffusion rate may be increased up to 50 °C but desorption of sulfur compounds from adsorbent surface may take place at temperature higher than 50 °C then the overall adsorption decreases.1,9,10
image file: c6ra16367j-f9.tif
Fig. 9 Adsorption temperature variation (stirrer speed: 1000 rpm; adsorbent particle size: 0.052 mm; sulfur concentration in model fuel: 900 ppm; adsorbent amount: 6.75 g L−1 of model fuel).

Effect of variation of adsorbent amount

Fig. 10 shows the effect of adsorbent (coal dust) quantity on adsorbent capacity. It is observed from the figure that the adsorption capacity increases from 62 to 92 as the amount of adsorbent increases from 2 to 10 g L−1. This may be due to the availability of more adsorbent which provides more active sites for adsorption.1,9,10
image file: c6ra16367j-f10.tif
Fig. 10 Adsorbent amount variation (stirrer speed: 1000 rpm; adsorbent particle size: 0.052 mm; sulfur concentration in model fuel: 900 ppm; temperature: 50 °C).

Reusability of adsorbent

Fig. 11 shows the study of reusability of all adsorbents on sulfur adsorption. It is observed from the figure that the reusability efficiency of coal dust is quite high compared to the other four adsorbents. After 3rd cycle, the sulfur adsorption decreased only 5% for coal dust, whereas for bentonite, fly ash, laterite and zeolite the sulfur adsorption decreased by 15%, 30%, 32% and 37% respectively.
image file: c6ra16367j-f11.tif
Fig. 11 Reusability of adsorbents (sulfur: 900 ppm; adsorbent particle size: 0.052 mm; temperature: 50 °C; adsorbent amount: 6.75 g L−1 of model fuel, stirrer speed: 1000 rpm).

Adsorption equilibrium isotherm

From different adsorption equilibrium isotherms, Langmuir, Freundlich, BET, Halsey and Temkin have been chosen to fit the experimental data and the best fitted one was determined. The experimental condition was constant for all experiments (stirrer speed: 1000 rpm; temperature: 50 °C; time: 2 h; adsorbent particle size: 0.052 mm; adsorbent amount: 6.75 g L−1 of adsorbate; sulfur conc: 900 ppm).

Langmuir isotherm

The Langmuir15 is one of the most extensively used isotherm models in all adsorption processes. It assumes the formation of monolayer adsorbate on the outer surface of the adsorbent and all sorption sites have equal affinity towards the adsorbate.16 The theoretical equation of this isotherm is written in linear form as,
 
image file: c6ra16367j-t3.tif(3)

Fig. 12 shows the straight line plot of Ce/qe vs. Ce. From slope and intercept of the plot Qm and K are calculated.


image file: c6ra16367j-f12.tif
Fig. 12 Langmuir isotherm.

Freundlich isotherm

Freundlich isotherm is based on the equilibrium sorption on heterogeneous surfaces involving large adsorbate concentration range.17 The main assumption of this isotherm is that the adsorption sites are distributed exponentially with respect to heat of adsorption.16 This isotherm is expressed by the following equation.
qe = KFCe1/n

This can be linearised as,

 
image file: c6ra16367j-t4.tif(4)
Which gives a linear plot with a slope of 1/n and an intercept of log(KF). This plot with the present experimental data is shown in Fig. 13.


image file: c6ra16367j-f13.tif
Fig. 13 Freundlich isotherm.

BET isotherm

The BET model is an useful model for multi layer adsorption.18 Here, the assumption is that the solid surface influences the adsorption for very first layer of adsorbed material.19 The mathematical form of this isotherm is as follows,
 
image file: c6ra16367j-t5.tif(5)

The plot of Ce/qe (CsCe) vs. Ce/Cs is shown in Fig. 14 and from the slope and intercept the values of CBET and qs are calculated.


image file: c6ra16367j-f14.tif
Fig. 14 BET isotherm.

Halsey isotherm

The Halsey model20 is applicable for multilayer adsorption. This model involves the heteroporous nature of the adsorbent which makes its difference from Freundlich isotherm.21 The linear form of this isotherm is as follows
 
image file: c6ra16367j-t6.tif(6)

A plot of ln[thin space (1/6-em)]qs vs. ln[thin space (1/6-em)]Ce with experimental data is shown in Fig. 15 and the values of KHa and nHa from intercept and slope were determined.


image file: c6ra16367j-f15.tif
Fig. 15 Halsey isotherm.

Temkin isotherm

This isotherm contains a factor that comes from adsorbent–adsorbate interaction and the heat of sorption of all molecules is assumed to decrease linearly.22 Its derivation is characterized by a uniform distribution of binding energies.18 The general equation of this isotherm is as follows,
 
image file: c6ra16367j-t7.tif(7)

Fig. 16 shows a plot of qe vs. ln[thin space (1/6-em)]Ce and the values of B and AT were obtained from slope and intercept of the straight line.


image file: c6ra16367j-f16.tif
Fig. 16 Temkin isotherm.

The values of all adsorption equilibrium isotherm constants for different models are showed in Table 6. It has been observed that the Langmuir adsorption isotherm is the best fitted isotherm model with highest R2 value. This reveals that the adsorption is physically in nature.

Table 6 Values of adsorption equilibrium isotherm constants
Isotherm model Parameters Value at 50 °C R2
(1) Langmuir K 0.00183 0.991
Qm 200
(2) Freundlich KF 1.945 0.985
1/n 1.67
(3) BET qs 16.95 0.932
CBET 9.83
(4) Halsey KHa 0.321 0.957
nHa −1.66
(5) Temkin AT 0.0198 0.959
B 38.94


Thermodynamic analysis

The temperature effect on desulfurization of model fuel by adsorption using coal dust is explained further by thermodynamic parameters. Thermodynamic parameters i.e. change in free energy (ΔG), change in enthalpy (ΔH) and change in entropy (ΔS) were investigated by the following equations,23,24
 
ΔG = −RT[thin space (1/6-em)]ln[thin space (1/6-em)]KD (8)
 
ΔG = ΔHTΔS (9)
 
image file: c6ra16367j-t8.tif(10)
and,
 
image file: c6ra16367j-t9.tif(11)

Fig. 17 shows the plot of ln[thin space (1/6-em)]KD against 1/T and the values of ΔH and ΔS were found from slope and intercept of this linear plot. This is van't Hoff plot.


image file: c6ra16367j-f17.tif
Fig. 17 van't Hoff plot.

From the figure, ΔH and ΔS values for adsorption obtained are 4.67 kJ mol−1 and 0.0163 kJ mol−1 respectively. The low value of ΔH confirms that this adsorption is physical adsorption.25

(−ΔG) values at 10, 20, 30, 40 and 50 °C are 0.05, 0.1, 0.26, 0.43 and 0.59 kJ mol−1 respectively and the negative values prove the spontaneity of the system.26 The ΔS value is positive which indicates that the sulfur was randomly adsorbed on the surface of adsorbent, while positive value of ΔH confirms that the adsorption process was endothermic.1,27

Mechanism of adsorption

The coal dust used in the present study contains some degree of surface chemical heterogeneity which has been originated from the presence of heteroatoms in the sample. The acidity obtained from NH3-TPD analysis of all adsorbents shows that the acidity of coal dust is reasonably high (8.22 mmol g−1), which is the second highest acidity among all the adsorbents. Although fly ash is showing higher acidity than coal dust, but owing to very low surface area (19 m2 g−1), fly ash cannot compete with coal dust having high surface area (149.4 m2 g−1). The high acidic nature of coal dust may be originated from surface acidic oxygen functional groups which attracts basic thiophenic ring28 and causes a reduction in acidity as revealed from Table 3 and Fig. 4. The thiophenic ring contains two lone pairs of electrons and one of those pairs is utilized to retain the aromaticity of the thiophenic ring. The other one is free and increases the interaction between adsorbent acid site and the sulphur compounds.

Comparing the FTIR spectra of coal dust sample before and after adsorption (Fig. 3), it is observed that there is no change in functional group of the coal dust sample after adsorption and also no peak shift is located. Hence, no evidence of formation of any new bond with organic sulfur compound with coal dust sample is obtained and so it can be said that the sulfur compound was attached on coal surface with weak electrostatic or van der Waals force and it was a physical adsorption. No change in surface morphology in SEM pictures of coal dust sample after adsorption compared to the sample before adsorption proves the same.29,30

XRD analysis of coal dust samples before and after adsorption also proves that the nature of adsorption is physical. All these proofs confirm that the governing isotherm model Langmuir isotherm.

Optimization of the process parameters of sulfur adsorption by using Box–Behnken model (BBD)

Response surface methodology (RSM) is a powerful statistical-based technique that can be used to predict the response of a system to any new condition and determine the optimum conditions.31 Box–Behnken Design (BBD) in RSM is an important design tool used for optimization of process parameters and consists of an equal number of factors of all possible combinations.32 It provides comprehensive conclusions and detailed information for smaller number of experiments and interactive effects of operating parameters on all responses.33 The choice of BBD amongst other RSM design was due to its feasibility, simplicity, efficiency and its application where extreme treatment combinations are to be avoided. In addition, BBD also offers the fewer experimental runs for three factors which can be used to explore a quadratic response surfaces as well as, creating a second order polynomial model.34 The experimental design for the optimization study was performed employing a four-factor BBD with input variables; initial sulfur concentration, adsorbent amount, time and temperature.

A second-order polynomial equation was fitted in the experimental data to make a relationship between the input variables and the responses. The BBD analysis was developed by using design expert software 7.0.0.

 
image file: c6ra16367j-t10.tif(12)

This equation represents the second order polynomial equation where y is the response variable of sulfur removal. x1, x2, x3xi are the coded parameters and β0, βi, βii and βij are the interception, regression co-efficients of linear effect, squared effect and interaction effect, ε represents an error. The levels of the independent variables and their limits are presented in Tables 7 and 8.

Table 7 Experimental values and levels of variables
Internal parameter Coded value Actual value Range
Initial sulfur concentration (mg L−1) C1 R1 300–1800
Adsorbent weight or mass (g L−1) C2 R2 3–10
Time (min) C3 R3 10–120
Temperature (°C) C4 R4 40–70


Table 8 Output response
SL no. Output
1 Sulfur adsorbed per gram of adsorbent (mg g−1)


The observed and the predicted values for sulfur adsorption process are presented in Table 9 and plotted in Fig. 18. The following quadratic model was obtained to establish a relationship between four independent process parameters and the dependent response [sulfur adsorption per gram of adsorbent].

 
Sulfur adsorption per gram of adsorbent (mg g−1) = 79.98 + 37.72 × A + 13.32 × B + 15.90 × C − 5.33 × D − 0.42 × A × B + 5.00 × A × C − 6.75 × A × D − 0.31 × B × C + 0.75 × B × D − 0.50 × C × D − 7.30 × A2 − 6.15 × B2 − 15.57 × C2 − 16.47 × D2 (13)

Table 9 Experimental design, observed and predicted response values for BBD
Run Coded value Real value Sulfur adsorbed (mg g−1)
C1 C2 C3 C4 R1 R2 R3 R4 Observed Predicted
1 −1 −1 0 0 300 3 65 55 11.34 15.06
2 1 −1 0 0 1800 3 65 55 90 91.34
3 −1 1 0 0 300 10 65 55 39 42.54
4 1 1 0 0 1800 10 65 55 116 117.15
5 0 0 −1 −1 1050 6.5 10 40 35 36.87
6 0 0 1 −1 1050 6.5 120 40 70 69.67
7 0 0 −1 1 1050 6.5 10 70 22 27.21
8 0 0 1 1 1050 6.5 120 70 55 58
9 −1 0 0 −1 300 6.5 65 40 18 17.07
10 1 0 0 −1 1800 6.5 65 40 100 106.01
11 −1 0 0 1 300 6.5 65 70 25 19.90
12 1 0 0 1 1800 6.5 65 70 80 81.84
13 0 −1 −1 0 1050 3 10 55 30 28.73
14 0 1 −1 0 1050 10 10 55 54.22 55.99
15 0 −1 1 0 1050 3 120 55 62 61.14
16 0 1 1 0 1050 10 120 55 85 87.18
17 −1 0 −1 0 300 6.5 10 55 10 8.49
18 1 0 −1 0 1800 6.5 10 55 80 73.93
19 −1 0 1 0 300 6.5 120 55 30 30.28
20 1 0 1 0 1800 6.5 120 55 120 115.73
21 0 −1 0 −1 1050 3 65 40 52 50.12
22 0 1 0 −1 1050 10 65 40 80 75.26
23 0 −1 0 1 1050 3 65 70 39 37.95
24 0 1 0 1 1050 10 65 70 70 66.10
25 0 0 0 0 1050 6.5 65 55 80 79.98
26 0 0 0 0 1050 6.5 65 55 79.98 79.98
27 0 0 0 0 1050 6.5 65 55 79 79.98
28 0 0 0 0 1050 6.5 65 55 81 79.98
29 0 0 0 0 1050 6.5 65 55 79.90 79.98



image file: c6ra16367j-f18.tif
Fig. 18 Predicted vs. experimental values of sulfur adsorption.

ANOVA analysis

The second order polynomial equation was fitted satisfactorily for sulfur adsorption. The significance of any process parameter is high if F-value is high and p-value (probability value) is lesser than 0.05. The significance of any process parameter can also be analyzed using the sum of square value and its higher value implies more importance of corresponding variables.35,36 Here the quadratic model has a high F value [103.78] and p-value is less than 0.0001, which signify the reliability of the model. Similarly the lower p-value of the independent variables indicates the higher significance.37–39 Tables 10 and 11 show that the A, B, C, D, AC, AD, A2, B2, C2 and D2 are significant parameters and as their p-values are less than 0.05. Finally a high R2 value of 0.990 with adjusted R2 0.980 indicate that the proposed polynomial model is reasonably well fitted with the data.
Table 10 Analysis of variances (ANOVA) for sulfur adsorption (mg g−1)a
Source SS df MS F value p value Prob > F
a A: Initial sulfur Concentration; B: Adsorbent mass; C: Time; D: Temperature; SS: Sum of squares; MS: Mean square; df: degree of freedom.
Model 25[thin space (1/6-em)]706.15 14 1836.15 103.78 <0.0001 Significant
A 17[thin space (1/6-em)]075.09 1 17[thin space (1/6-em)]075.09 965.06 <0.0001  
B 2130.13 1 2130.13 120.39 <0.0001  
C 3033.08 1 3033.08 171.43 <0.0001  
D 341.33 1 341.33 19.29 0.0006  
AB 0.69 1 0.69 0.039 0.8464  
AC 100 1 100 5.65 0.0322  
AD 182.25 1 182.25 10.30 0.0063  
BC 0.37 1 0.37 0.021 0.8868  
BD 2.25 1 2.25 0.13 0.7267  
CD 1 1 1 0.057 0.8155  
A2 345.87 1 345.87 19.55 0.0006  
B2 245.31 1 245.31 13.86 0.0023  
C2 1571.91 1 1571.91 88.84 <0.0001  
D2 1759.46 1 1759.46 99.44 <0.0001  
Residual 247.71 14 17.69      
Lack of fit 245.70 10 24.57 48.96 0.0010 Significant
Pure error 2.01 4 0.50      
Corrected total 25[thin space (1/6-em)]953.85 28        


Table 11 Estimated regression coefficient and corresponding error
Factor Coefficient estimate Standard error
Intercept 79.98 1.88
A 37.72 1.21
B 13.32 1.21
C 15.90 1.21
D −5.33 1.21
AB −0.42 2.10
AC 5.00 2.10
AD −6.75 2.10
BC −0.31 2.10
BD 0.75 2.10
CD −0.50 2.10
A2 −7.30 1.65
B2 −6.15 1.65
C2 −15.57 1.65
D2 −16.47 1.65


Table 12 represents the optimization of process parameters. After minimization of all process parameters, the highest predicted adsorption is found to be of 70.76% which is close to the observed value of 68%.

Table 12 Optimal operating conditions maximizing adsorption of sulfur
Sulfur concentration (mg L−1) Adsorbent amount (g L−1) Time (min) Temperature (°C) Sulfur adsorption (mg g−1)
Observed Predicted
1800 3 34.33 40 68 70.76


Table 12 represents the optimal conditions when all independent variables are minimized and conversion is maximized with desirability 0.810.

3-D plot of process parameters

Central composite design can be visualized as 3-D plots that represent the variation of the response with two parameters, keeping the other parameter constant. The resulting 3-D surface response plots shown in Fig. 19 describe sulfur adsorption as a function of surface plots of (a) adsorbent mass and initial sulfur concentration, (b) time and initial sulfur concentration, (c) temperature and initial sulfur concentration, (d) time and adsorbent mass, (e) temperature and adsorbent mass and (f) temperature and time. From 3D plots it is concluded that all process parameters have great influence on the sulfur adsorption.
image file: c6ra16367j-f19.tif
Fig. 19 Surface plots of sulfur adsorption as a function of (a) adsorbent mass and initial sulfur concentration, (b) time and initial sulfur concentration, (c) temperature and initial sulfur concentration, (d) time and adsorbent mass, (e) temperature and adsorbent mass, (f) temperature and time.

Conclusion

Adsorptive desulfurization of model fuels was performed by using coal dust, bentonite, fly ash, laterite and zeolite. The best result was performed by coal dust [84.22 mg sulfur adsorbed per g of adsorbent] with a high surface area of 149.40 m2 g−1. There was a significant effect of process parameters such as time, temperature, initial sulfur concentration, stirrer speed, adsorbent amount and adsorbent particle size on the adsorptive desulfurization process by using best adsorbent. The Langmuir isotherm was best fitted with equilibrium isotherm data. The adsorption is physical adsorption in nature which is confirmed by van't Hoff plot. Finally, the Box–Behnken design model was proposed for optimization of the process parameters and it was observed that initial sulfur concentration, adsorbent amount, time and temperature have significant influence on sulfur adsorption process.

XRD of coal dust sample before and after adsorption has been provided in ESI.

SEM micrograph analysis of coal dust, bentonite, fly ash, laterite and sodium zeolite has been done and the images are provided as ESI. From those images it is clear that coal dust sample is least agglomerated compared to the other four adsorbents which in turn are responsible for best results of adsorption.

FTIR spectral analysis of coal dust, bentonite, fly ash, laterite and zeolite in the wave range of 4000–450 cm−1 has been done. The spectral image and discussion part are provided as ESI.

NH3-TPD of all the adsorbent samples was reported in ESI.

The proximate and ultimate analysis of coal dust sample has been provided as ESI.

The HPLC analysis of MF-4 samples, before and after adsorption is provided as ESI. It was showed that the peak areas of all sulfur compounds significantly decreased after adsorption compared the peak areas of sulfur compounds of feed stock. In the case of coal dust, peak area of all sulfur compounds decreased more compared the other four adsorbents.

Finally, we also checked the efficiency of adsorbents on real fuels; gasoline and light gas oil obtained from Haldia Refinery, Indian Oil Corporation, India. The adsorption efficiency of all adsorbents for desulfurization of gasoline and light gas oil are tabulated in ESI along with the properties of both the fuels before and after adsorption. It is observed that there is no significant change in fuel properties due to adsorption rather a significant decrease in sulfur level.

Nomenclature

qtAdsorption capacity at time t (mg g−1)
qeAdsorption capacity at equilibrium (mg g−1)
qsAmount of sulfur adsorbed per gram of adsorbent at saturation condition or theoretical isotherm saturation capacity (mg g−1)
CeResidual concentration of sulfur compound in model fuel at equilibrium (mg L−1)
CiInitial concentration of sulfur compound in model fuel (mg L−1)
CtResidual concentration of sulfur compound in model fuel at time t (mg L−1)
VVolume of model fuel (L)
WMass of adsorbent (g)
QmMaximum monolayer coverage capacities (mg g−1)
KLangmuir isotherm constant (L mg−1)
KFFreundlich isotherm constant related to adsorption capacity (mg g−1)
NAdsorption intensity
CsAdsorbate monolayer saturation concentration (mg L−1)
CBETBET adsorption isotherm relating to the energy of surface interaction (L mg−1)
KHa, nHaHalsey isotherm constants (mg L−1)
ATTemkin isotherm equilibrium binding constant (L g−1)
RUniversal gas constant (J mol−1 K−1)
TAbsolute temperature (K)
bTTemkin isotherm constant
KDDistribution coefficient
ΔHEnthalpy difference (kJ mol−1)
ΔSEntropy difference (kJ mol−1)

References

  1. M. Ishaq, S. Sultan, I. Ahmad, H. Ullah, M. Yaseen and A. Amir, J. Saudi Chem. Soc., 2015 DOI:10.1016/j.jscs.2015.02.003.
  2. M. Muzic, K. S. Bionda and Z. Gomzi, Chem. Eng. Technol., 2008, 31(3), 355–364 CrossRef CAS.
  3. S. Velu, X. Ma and C. Song, Prepr.–Am. Chem. Soc., Div. Fuel Chem., 2002, 47(2), 447–448 CAS.
  4. J. M. Palomino, D. T. Tran, J. L. Hauser, H. Dong and S. R. J. Oliver, J. Mater. Chem. A, 2014, 2, 14890–14895 CAS.
  5. S. Das Gupta, P. Gupta, Aarti, A. Nanoti, A. N. Goswami, M. O. Garg, E. Tangstad, O. B. Vistad, A. Karlsson and M. Stocker, Fuel, 2013, 108, 184–189 CrossRef CAS.
  6. X. Xu, S. Zhang, P. Li and Y. Shen, Fuel, 2013, 111, 172–179 CrossRef CAS.
  7. A. A. Adekanmi and A. Folorunsho, Res. J. Chem. Sci., 2012, 2(8), 14–20 Search PubMed.
  8. Y. Shi, X. Zhang and G. Liu, Fuel, 2015, 158, 565–571 CrossRef CAS.
  9. W. Ahmad, I. Ahmad, M. Ishaq and K. Ihsan, Arabian J. Chem., 2014 DOI:10.1016/j.arabjc.2013.12.025.
  10. A. Khodadadi, M. T. Angaji, A. T. Rafsanjani, and A. Yonesi, 4th International Conference on Nanostructures, 2012, pp. 1197–1199 Search PubMed.
  11. S. Kumar, V. C. Srivastava and R. P. Badoni, Fuel, 2011, 90, 3209–3216 CrossRef CAS.
  12. C. Ngamcharussrivichai, C. Chatratananon, S. Nuntang and P. Prasassarakich, Fuel, 2008, 87, 2347–2351 CrossRef CAS.
  13. A. Srivastav and V. C. Srivastava, J. Hazard. Mater., 2009, 170, 1133–1140 CrossRef CAS PubMed.
  14. M. S. Patil, Y. C. Bhattacharyulu and S. R. Kulkarni, Journal of Engineering Research and Studies, 2011, 2(1), 81–98 Search PubMed.
  15. Y. S. Ho, W. T. Chiu and C. C. Wang, Biotechnology, 2005, 96, 1285–1291 CAS.
  16. P. Sampranpiboon, P. Charnkeitkong and X. Feng, WSEAS Trans. Environ. Dev., 2014, 10, 35–47 Search PubMed.
  17. O. Moradi, Arabian J. Chem., 2013 DOI:10.1016/j.arabjc.2011.12.014.
  18. K. Y. Foo and B. H. Hameed, Chem. Eng. J., 2010, 156, 2–10 CrossRef CAS.
  19. G. Buckton, Interfacial Phenomena in Drug Delivery and Targeting, CRC Press, 2000, ISBN No: 3718656337, 9783718656332 Search PubMed.
  20. O. Moradi, A. Fakhri, S. Adami and S. Adami, J. Colloid Interface Sci., 2013, 395, 224–229 CrossRef CAS PubMed.
  21. E. Lichtfouse, J. Schwarzbauer and D. Robert, Green Materials for Energy, Products and Depollution, Springer Science & Business Media, 2013, ISBN no.: 9400768362, 9789400768369 Search PubMed.
  22. A. O. Dada, A. P. Olalekan, A. M. Olatunya and O. Dada, J. Appl. Chem., 2012, 3(1), 38–45 Search PubMed.
  23. T. M. Elmorsi, J. Environ. Prot., 2011, 2(6), 817–827 CrossRef CAS.
  24. R. Kobiraj, N. Gupta, A. K. Kushwaha and M. C. Chattopadhyaya, Indian J. Chem. Technol., 2012, 19, 26–31 CAS.
  25. E. Worch, Adsorption Technology in Water Treatment: Fundamentals, Processes, and Modeling, Walter de Gruyter, 2012, ISBN no.: 3110240238, 9783110240238 Search PubMed.
  26. A. E. Ofomaja and Y. S. Ho, Biotechnology, 2008, 99, 5411–5417 CAS.
  27. I. Ibrahim, Y. M. A. Obaidi and S. M. Hussin, Am. Chem. Sci. J., 2015, 9(3), 1–7 CrossRef CAS.
  28. M. Yu, N. Zhang, L. Fan, C. Zhang, X. He, M. Zheng and Z. Li, Rev. Chem. Eng., 2015, 31(1), 27–43 CAS.
  29. K. Muthukumaran and S. S. Beulah, Asian J. Chem., 2010, 22(10), 7857–7864 CAS.
  30. S. Mandal, M. K. Sahu and R. K. Patel, Water Resources and Industry, 2013, 4, 51–67 CrossRef.
  31. B. Y. Tak, B. S. Tak, Y. J. Kim, Y. J. Park, Y. H. Yoon and G. H. Min, J. Ind. Eng. Chem., 2015, 28, 307–315 CrossRef CAS.
  32. M. A. Tekindal, H. Bayrak, B. Ozkaya and Y. Genc, Turkish Journal of Field Crops, 2012, 17(2), 115–123 Search PubMed.
  33. K. K. Garg and B. Prasad, J. Environ. Chem. Eng., 2016, 4, 178–190 CrossRef CAS.
  34. B. V. Ayodele and C. K. Cheng, J. Ind. Eng. Chem., 2015, 32, 246–258 CrossRef CAS.
  35. S. H. Dhawane, T. Kumar and G. Halder, Energy Convers. Manage., 2015, 100, 277–287 CrossRef CAS.
  36. A. S. Reshad, P. Tiwari and V. Goud, Fuel, 2015, 150, 636–644 CrossRef CAS.
  37. Z. Shayegan, M. Razzaghi, A. Niaei, D. Salari, M. T. S. Tabar and A. N. Akbari, Korean J. Chem. Eng., 2013, 30(9), 1751–1759 CrossRef CAS.
  38. A. A. Meybodi, A. Ebadi, S. Shafiei, A. R. Khataee and M. Rostampour, J. Taiwan Inst. Chem. Eng., 2015, 48, 40–48 CrossRef.
  39. A. Sengupta, P. D. Kamble, J. K. Basu and S. Sengupta, Ind. Eng. Chem. Res., 2012, 51, 147–157 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra16367j

This journal is © The Royal Society of Chemistry 2016