Equilibrium, Kinetic, and Thermodynamic Analysis - Energy & Fuels

Feb 19, 2018 - ... experimental results indicated that the adsorption efficiency reaches to more than 96%, when the initial concentration of the organ...
1 downloads 14 Views 921KB Size
Subscriber access provided by the Henry Madden Library | California State University, Fresno

Article

Experimental and Modeling Study of Organic Chloride Compounds Removal from Naphtha Fraction of Contaminated Crude Oil Using Sintered # -Al2O3 Nanoparticles: Equilibrium, Kinetic and Thermodynamic Analysis Samad Arjang, Kazem Motahari, and Majid Saidi Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b03845 • Publication Date (Web): 19 Feb 2018 Downloaded from http://pubs.acs.org on February 25, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Experimental and Modeling Study of Organic Chloride Compounds Removal from Naphtha Fraction of Contaminated Crude Oil Using Sintered γ−Al2O3 Nanoparticles: Equilibrium, Kinetic and Thermodynamic Analysis Samad Arjang a, Kazem Motahari a*, Majid Saidi b† a

b

Department of Chemical Engineering, Arak University, Arak, 38156−8−8349, Iran

School of Chemistry, College of Science, University of Tehran, PO Box 14155−6455, Tehran, Iran

Abstract Because of destructive effects of organic chlorides, elimination of these compounds from contaminated crude oil and its distillates is very crucial for oil refiners. Despite the importance of the subject, limited researches in this field are performed and most of the proposed removal methods are operationally difficult and costly. In the present study, adsorption process as an efficient technology is proposed and employed to eliminate the organic chlorides compounds from naphtha fraction of contaminated crude oil. The γ−Al2O3 nanoparticles as an adsorbent is prepared and characterized by XRD, SEM−EDS, and BET analyses. Adsorption experiments are carried out at different operating temperatures and also various initial organic chloride concentrations. The experimental results indicated that the adsorption efficiency reaches to more than 96%, when the initial concentration of the organic chloride in the sample is 8.5 mg/L. The adsorption equilibrium analysis revealed that the Freundlich isotherm model provided the best fit and prediction of experimental data. It was * †

Corresponding Author: (Kazem Motahari) [email protected] Corresponding Author: (Majid Saidi) Email: [email protected] , [email protected]

1 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 68

also found that for all of verified samples, the adsorption kinetic followed pseudo−second order equation well. According to experimental investigations reported in the present work, the sintered γ−Al2O3 nanoparticles would be an effective adsorbent for elimination of organic chloride as a hazardous material from the naphtha distillate of crude oil. Keywords: Organic chloride; Adsorption; Crude oil; Naphtha distillate; γ−Al2O3 nanoparticles.

1. Introduction Due to harmful effects of crude oil impurities on the refinery processes and facilities, elimination of these contaminants and control of their level in the crude oil samples is very important. Recently, according to environmental and technical regulations, the concentration of organic chlorides as one of the most important impurities should be monitored and the crude oil containing high amount of these compounds do not be allowed to enter as feed to refineries. Therefore, this type of impurity could make a considerable additional cost for oil producers and consumers. There are two types of chloride compounds in the crude oil including inorganic and organic chlorides

1, 2

. Both of these compounds could be harmful and cause many problems for

refining processes and facilities 1 such as corrosion of equipments including crude distillation unit (CDU) overhead 3, 4, heat exchangers and other downstream facilities 5, 6. Fouling in the processing vessels 7, catalyst poisoning and deactivation

8, 9

and equipment blockage

1

are

some other important problems caused by these compounds. The most important refinery units reported to be damaged by organic chlorides are distillation towers

1, 10

, naphtha

hydro−treating (NHT) units 11, catalytic reforming units 12 and catalytic cracking units 13. 2 ACS Paragon Plus Environment

Page 3 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Inorganic chlorides are mainly separated using an electric desalting unit

2, 14

. But separation

of organic chlorides is very difficult because most of them have more solubility in the crude oil with respect to water and also they could not be easily extracted 2. This type of chloride compounds reacts and forms hydrogen chloride (HCl) under high temperature and pressure operating conditions during refining processes 1, 15-17. Dissolution of HCl in small amount of water in the distillation towers overhead leads to serious acidic corrosion in the downstream facilities

1, 2, 18

. In addition, in some other units like diesel and fuel hydrogenation, these

dangerous compounds hydrolyzed and form HCl

18

. In another situation, the chemical

reaction of this acid and nitrogen compounds leads to formation of NH4Cl via hydrogenation process 19. The presence of ammonium chloride causes the blockage of heat exchangers and serious under−deposit corrosion in some units like Fluid Catalytic Cracking Unit (FCCU) 13. Also the presence of organic chlorides contamination in the final refinery products like fuels will be harmful for the last user equipment (like engines) and also leads to environment pollution 18. It is well documented that even trace amounts of these compounds could disturb processes and facilities 1, 8, 10, 18. Therefore, as the first strategy these compounds must not be allowed to enter or form in the crude oil. It is obvious that elimination, controlling or mitigation of the main sources of these compounds at production and primary processing plants are much easier than separation or elimination of them from contaminated crude oil. It is necessary to mention that under some operating conditions and also during production, storage and transportation processes, crude oil could be contaminated with these harmful compounds. In this case, these compounds must be removed from the crude oil in order to reach the allowable concentration for refining processes which is below 1.0 mg/L in the naphtha fraction of the crude (Initial Boiling Point, IBP, to 477.59 K)

20, 21

. Some previous works were performed to eliminate the organic

3 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 68

chlorides from the contaminated crude oil samples containing high amount of these compounds

18, 22-24

. But all of these proposed methods are very difficult and costly. Many of

them focused on breaking C−Cl bond and convert these compounds to harmless molecules. It is known that this bond is very strong and its average bond dissociation energy is about 83.0 kcal mole-1

25, 26

. Therefore, all of these methods need sever operating conditions

27

and also

catalytic processes are required for cleavage of this bond and convert them to other harmless compounds

28, 29

. A few other methods are based on the separation and removal of organic

chloride compounds from crude oil and its distillates, especially naphtha important methods of these proposed techniques are hydrogenation

29-32

. The most

29, 33

, nucleophilic

substitution 31 and adsorption 18, 32, 34, 35. In hydrogenation process, different forms of organic chlorides are converted to HCl over a specific hydrogenation catalyst 9. It is important to note that this removal method needs high temperature and pressure operating conditions and also this method has high operating cost because of its catalytic reactions 29. The most important problem of this method is the formation of hydrochloric acid which could make considerable corrosion in the reactor vessel and other facilities of this process. In the substitution method, the nucleophilic reagent such as sodium cyanide, sodium thiocyanate or sodium hydroxide reacts with organic chlorides to substitute chloride atom in the presence of phase transfer catalyst, like polyethylene glycol or dimethyl sulfoxide

31, 36-38

. This method also has some

drawbacks such as long reaction time and high temperature condition and operating cost. Considering the previous proposed methods represents that the mentioned converting methods could not be industrially applicable because of high operating and investment costs. Meanwhile, adsorption method 18, 35, 39, 40 does not need sever operating conditions and also it can be executed easily 41. Ma et al.

18

employed Na−LSX zeolite as an adsorbent of organic

chlorides from a model jet fuel. Although their results indicated that Na−LSX zeolite could

4 ACS Paragon Plus Environment

Page 5 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

act as an appropriate adsorbent for 5−chloro−2−methylaniline, but the main limitation of their work is high operating time of adsorption process (about 12 hours) 18. Also they performed their experiments only for a simple model sample not a real industrial one. In a few researches, it is reported that alumina could absorb some organohalide compounds 42

. Also it is well documented that when the adsorbent grain size decreases to nano order, the

adsorption capacity of the adsorbent increases significantly

43

. Alumina has perfectly

controlled mesoporosity associated with hardness, hydrolytic stability, amphoteric character, and thermal stability of the γ−transition−oxide phases

44, 45

. On the other hand, γ−phase of

alumina shows better characteristics and higher specific surface area for adsorption process 46

. Thus, in the present research, sintered γ−Al2O3 nanoparticles are employed for the first

time as the adsorbent of organic chloride compounds from naphtha distillate as a hydrocarbon media. Sintering temperature affects the final characteristics of sintered alumina. In this case the sintering temperature must be lower than 1023 K in order to prevent any phase transition of γ−Al2O3

47

and also must not be too high for maintaining the basic characteristics of

alumina nanoparticles

48

. Also, in the present study, the adsorption of different types of

organic chloride compounds from the industrial naphtha distillates of contaminated crude oil samples by sintered γ−Al2O3 is investigated. For this purpose, five crude oil samples containing different initial concentration of organic chlorides in the range of 8.5 to 105 mg/L are selected. The adsorption performance of prepared adsorbents for removal of these compounds from naphtha distillates at seven adsorption times in the range of 1 to 10 h and seven different constant temperatures including 298, 308, 318, 328, 338, 348 and 358 K are considered. 2. Experimental Methods 2.1. Materials 5 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

All of the nanoparticles were used as received. Alumina nanoparticles (γ−Al2O3) basically produced by SSNano (Sky Spring Nano) Inc. were bought with the specified characterizations as: Average Particle Size (APS), 20 nm; Specific Surface Area (SSA), 230−400 m2/g and Purity, 99.9%. The basic characteristics of crude oil used in this research are summarized in Table 1. These industrial samples were contaminated with organic chloride compounds during production or primary processing. The initial concentration of the organic chlorine compounds in the crude oil samples varied from 8.5 to 105 mg/L. Table 1

2.2. Alumina Nanoparticles Sintering and Adsorbent Preparation To prepare the adsorbent for the main experiments, 10 g of γ−Al2O3 nanoparticles was placed in the porcelain and let it to be sintered at temperature of 773 K in an enclosed furnace for about 4 hours. Then, it is allowed to be cooled slowly to room temperature. Thirty five same adsorbents were prepared based on this procedure and each of them is used for a specified sample at a certain temperature. Also, in order to investigate the effect of adsorbent quantity on the adsorption process, different amount of adsorbents including 10, 15, 20, 25, 30 and 35 g of alumina nanoparticles were employed.

2.3. Apparatus In this study, a batch reactor was designed to verify the effect of time and temperature on the adsorption performance of sintered alumina nanoparticles. A schematic diagram of the set−up is shown in Figure S−1 (Supporting Information). This reactor contains a batch 1200 ml 6 ACS Paragon Plus Environment

Page 6 of 68

Page 7 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

volumetric flask vessel, electrical heating jacket, temperature sensor (one platinum resistance thermometer (Pt 100) with a precision of ±0.1 K), pressure sensor, flanged sealing door and temperature controller. Due to safety consideration during the experiments, the entire set−up was placed on an electrically insulated platform.

2.4. Measurement Methods 2.4.1. Characterization of Adsorbent X−Ray−Diffraction (XRD) pattern for sintered γ−Al2O3 nanoparticles was obtained using Philips diffractometer model PW 1800. Scanning−Electron−Microscopy (SEM) Philips model XL30 was used for taking SEM micrograph and Energy Dispersive X−ray Spectroscopy (EDS) was employed to analyze the nano−adsorbent. The Brunauer–Emmett– Teller (BET) specific surface area, pore size, pore volume and micropore volume for the sintered alumina nanoparticles sample were determined by nitrogen adsorption using ASAP2010 instrument.

2.4.2. Measuring Method for Organic Chlorides Concentration 2.4.2.1. Measuring analyzers Organic chloride concentration of all samples measured using two different methods to ensure about the accuracy of final experimental results. The first method of measurement was based on coulometery while the second one was done with X−Ray fluorescence (XRF). Coulometery method was performed with Analytik Jena model multi EA 5000 analyzer while XRF method was done with CLORA 2XP model XOS instrument with a measurement range 7 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 68

of 0−3000 mg/L. Both of these methods lead to the same results and confirmed each other. But the coulometery technique is more precise than XRF method for low concentration especially for concentrations below 1 mg/L. In both measurement methods, the final result of each sample was recorded based on arithmetic mean calculated from five different measurements.

2.4.2.2. Preparation of naphtha distillate samples In order to measure the concentration of organic chloride in the crude oil samples, naphtha distillate of these samples must be prepared according to the standard procedure. At the first step, the crude oil sample was distilled and its naphtha cut at temperature of 477 K was obtained based on ASTM D86 standard method

49

. Then, according to ASTM D4929

standard procedure, this distillate was washed three times with equal volume of caustic solution (1 M KOH) followed by equal volume of washing water for each time. The washing process with caustic solution removes hydrogen sulfide, while the washing with water removes traces of inorganic chlorides either originally present in the crude oil or from impurities in the caustic solution 50. Finally, the organic chloride concentration of the washed distillate was measured using two different analyzers based on the methods mentioned in the above section.

2.5. Experimental Procedure In this study, five crude oil samples with different initial organic chloride concentration were selected. These samples were taken from the outlet stream of same primary process plant at different times through two months (eight weeks). In fact, samples were taken at week 1, 3, 5, 8 ACS Paragon Plus Environment

Page 9 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

7 and 8, respectively. For each time, about 1000 liters processed crude oil was taken from the outlet stream of industrial plant to check the repeatability of experiments. The organic chloride concentrations of these five samples were measured using the mentioned methods and reported in Table 2. Based on reported data in this table, it seems that an unwanted source of organic chlorides entered the crude oil and then decreased gradually. Table 2 As presented in Table 2, for simplicity of the samples verification, they were named and ordered from the maximum organic chloride level to the minimum one. Therefore, samples with initial organic chloride concentration of 105, 82, 55, 27 and 8.5 mg/L were named sample no. 1, 2, 3, 4 and 5, respectively. Adsorption experiments were done at seven different constant temperatures of 298, 308, 318, 328, 338, 348 and 358 K and seven adsorption time intervals including 1, 2, 4, 6, 8, 9 and 10 h. For each sample and adsorption temperature, the adsorbent bed was made by sintering 10 g of alumina nanoparticles. In each experiment, 2200 ml of the sample was distilled based on ASTM D86 procedure and its naphtha distillate was obtained at 477 K. Then, the prepared adsorbent was placed in the reactor vessel and in the following, 280 ml of the obtained naphtha fraction was directly poured in the vessel. The vessel was completely sealed and temperature was adjusted on the determined adsorption temperature. After elapsing specified adsorption time interval, the electrical heating jacket was turned off and the vessel and its content let to be cooled to the room temperature. Organic chlorides concentration of treated sample was measured and reported based on the mentioned procedures. In order to recycle the adsorbent for repeating the experiments, it was heated to 478 K for 25 minutes to vaporize the remaining of naphtha distillate and adsorbed species. The recovery temperature was selected based on the final boiling point of naphtha distillate. To ensure the complete evaporation of the distillate compounds adsorbed on the 9 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 68

sintered alumina nanoparticles, this temperature was adjusted one degree higher than boiling point of the obtained naphtha fraction which was 477 K.

3. Modeling The amount of adsorbed organic chlorides, qt (mg/g) at time t, was calculated by the following equation 51, 52:

 =

 −   m

(1)

where C0 is the initial concentration of the organic chloride (mg/L), Ct is the concentration of the organic chloride at specified time t (mg/L), V is the volume of naphtha fraction (L), and m is the mass of adsorbent (g). Also, the adsorption capacity of sintered γ−Al2O3 nanoparticles at equilibrium condition, qe (mg/g), was calculated by the following mass balance equation 5356

:

 =

 −   m

(2)

where Ce is the equilibrium concentration of the organic chloride compounds in the naphtha fraction (mg/L). The removal percentage (%) of organic chloride compounds from naphtha distillate by adsorption process can be calculated using the following equation 54, 57:

 % =

 −  × 100% 

3.1. Adsorption Equilibrium Modeling

10 ACS Paragon Plus Environment

(3)

Page 11 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

In order to investigate the adsorption equilibrium, the most famous isotherm models including Langmuir, Temkin, Freundlich and Dubinin–Radushkevich isotherms were selected to describe the relationship between the adsorbed amount of organic chloride compounds and its equilibrium concentration in the naphtha distillate.

3.1.1. Langmuir Isotherm Model Langmuir model was proposed by Hall et al. 58. The main characteristic of this model is the monolayer adsorption of the solute on a homogenous surface. In fact, in Langmuir model it is assumed that the adsorption process occurred at uniform adsorption energies and there is no transmigration on the adsorbent surface. This model was described by the following equation 59-61

:

 =

 !  1 + ! 

(4)

This equation could be rewritten as the following linear form 55, 62, 63:

1 = 

1 1 1 . +  !  

(5)

where qe (mg/g) is the adsorptive capacity of sintered γ−Al2O3 nanoparticles at equilibrium condition, and k (L/mg) is the Langmuir adsorption equilibrium constant.

3.1.2. Temkin Isotherm Model In this model, the effect of indirect interaction of adsorbent−adsorbate on the adsorption isotherm process was considered

64-66

. This interaction involves the linear decrease of 11

ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 68

adsorption heat with increasing the adsorption coverage. The following equation describes the linear form of Temkin isotherm model 67, 68:

%& %&  = $ ( )  + $ ( )  ' '

(6)

where b and A are Temkin constants. This isotherm model was based on two important assumptions. Firstly, the heat of adsorption of all the molecules in the layer decreases linearly with coverage due to adsorbent–adsorbate interaction, and secondly, the adsorption is characterized by a uniform distribution of binding energies, up to some maximum binding energy 64, 69 .

3.1.3. Freundlich Isotherm Model Freundlich model is more realistic than Langmuir isotherm model. In this model, it is assumed that the adsorbent surface is heterogeneous and adsorption sites are different with each other. But all of these sites obey the Langmuir isotherm

51, 67

. In fact this model

describes the multilayer adsorption of adsorbate on the adsorbent surface 67. The describing equation for this model is as follows 70:

 =

* 

+

(7)

where kF and n are the capacity and the intensity of adsorption, respectively. This formula can be reformed to the following linear equation 71-73:

) = )

*

+ )

(8)

3.1.4. Dubinin−Radushkevich Isotherm Model 12 ACS Paragon Plus Environment

Page 13 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Dubinin−Radushkevich isotherm model is generally applied to express the adsorption mechanism with a Gaussian energy distribution instead of heterogeneous surface

74, 75

. Also,

the present model described the adsorption process performance as a pore filling mechanism instead of an energetically non−uniform surface of the adsorbent

76

. The basic equation of

this isotherm model is as follows 74: 3

(9)

0%& )1 + 1⁄ 2  =  exp / 6 −25 3

where R is the gas constant (8.314 J/mol K), T (K) is the absolute temperature and E is the adsorption energy (kJ/mol). The linear form of Dubinin−Radushkevich equation is as follow 77, 78

:

)  = )  − 78 3

(10)

where,

8 = %& ) $1 + 7=

1 25 3

1 ( 

(11)

(12)

3.2. Kinetic Modeling To evaluate the efficiency and mechanism of the adsorption process, kinetic investigation is necessary 54, 79. In this study, the adsorption kinetic model was developed based on the most popular

models.

Pseudo−first

order

66,

80

,

pseudo−second

13 ACS Paragon Plus Environment

order

81,

82

and

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

intra−particle−diffusion

83

Page 14 of 68

as the most widely used models were employed to investigate the

kinetic mechanism of organic chloride adsorption on sintered γ−Al2O3 nanoparticles.

3.2.1. Pseudo−First Order Model Lagergren presented the first order rate equation for the adsorption of oxalic acid and malonic acid into charcoal to explain the kinetics of adsorption process on solid surfaces. The pseudo−first order kinetic model is expressed as follows 80, 84, 85:

 = 

9 

(13)

−  

where k1 is the equilibrium rate constant of pseudo−first order kinetic (1/min). After integration and by applying the boundary conditions (qt=0 at t=0), Eq. (13) can be rewritten as 63, 86-88:

) −   = )   −

9

(14)

3.2.2. Pseudo−Second Order Model The equation of pseudo−second order kinetic model is written as follow 70, 81, 89:

 = 

3 

(15)

−  3

By integrating the above equation, the linear form of this equation is obtained as below 55, 62, 90, 91

:

14 ACS Paragon Plus Environment

Page 15 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

 = 

1 1 +  3  3 

(16)

where k2 (g/ (mg .min)) is the pseudo−second order rate constant for the adsorption process.

3.2.3. Intra−Particle Diffusion Model 83

This model proposed by Weber and Morris

can explain the possibility of intra particle

diffusion in an adsorption process. The adsorbate transport from the solution phase to the surface of the adsorbent particles occurs in several steps. The overall adsorption process may be controlled either by one or more steps, e.g. film or external diffusion, pore diffusion, surface diffusion and adsorption on the pore surface, or a combination of more than one step 84

. The equation of this model is expressed as below 83, 84:

 = :;+ 

9< 3

(17)

+=

where kint is the intra−particle diffusion rate constant (mg g−1min−0.5). According to this equation, after plotting qt versus t1/2, if the fitting line passes through the origin (i.e., Intercept I=0), it means that the controlling step of adsorption process would be intra−particle diffusion mechanism, unless otherwise, the intra−particle diffusion would not be the only rate−controlling step mechanism, indeed other factors may affect the rate of adsorption process and it is called hybrid control mechanism 92, 93.

3.3. Thermodynamic of Adsorption Process To investigate the possibility and the type of adsorption process, thermodynamic analysis should be performed. For this purpose, Gibbs free energy change, ∆G0, which is the most 15 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 68

important thermodynamic parameter, was employed. The describing equation for Gibbs free energy is as follows 88, 94, 95:

∆?  = −%& ): 

(18)

where Ke (L/g) is obtained by the below equation 60, 94:

: =

 

(19)

The relationship between other thermodynamic parameters including enthalpy change (∆H0) and entropy change (∆S0) with Gibbs free energy change (∆G0) can be described as the following equation 72, 96, 97:

∆?  = ∆@ − &∆A 

(20)

and this equation can be rewritten as 60, 95, 98, 99:

): = −

∆@ ∆A  + %& %

(21)

Based on Eq. 21, ∆S0 and ∆H0 can be obtained from the intercept and slope of linear fitting line of Ln Ke plot versus 1/T, respectively.

4. Results and Discussion 4.1. Characterization of Sintered Alumina Nanoparticles XRD pattern for sintered γ−Al2O3 nanoparticles is presented in Figure S−2 (Supporting Information). This pattern represents the high pure alumina existence in the final calcined

adsorbent without any phase change. The diffractograms for the sample display three distinct 16 ACS Paragon Plus Environment

Page 17 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

reflections at 38.3o (311 reflection), 45.7o (400) and 66.6o (440) which is in agreement with the database standard (JCPDS reference no.00−010−0425) and also with the reported results in the literatures for gamma alumina 100-103. SEM micrograph of this nano−adsorbent is shown in Figure S−3 (Supporting Information). Analysis of this image with Image J 1.51j8 software revealed that the adsorbent surface covered by fine nanoparticles with an average particle size less than 90 nm. According to the EDS analysis which is reported in Table S−1 (Supporting Information), the mass content of Al and O in this nano−adsorbent is 51.81% and 48.19% respectively. The BET analysis of surface area for this adsorbent indicated that the average specific surface area of this sample is 287.32 m2/g. Also its average pore size, pore volume and micropore volume were 1.23 nm, 0.42 cm3/g and 0.36 cm3/g, respectively.

4.2. Adsorbent Surface Characterization after Adsorption Process and after Regeneration It is well documented that there are different kinds of organic chlorides in the crude oil

10, 19

and some of these compounds enter this fluid during production, primary processing, storage and transportation 2. Wu et al. 1, 10 reported the existence of 17 kinds of organic chlorides in a specific crude oil in china including carbon tetrachloride, trichlorethylene, tetrachlorethylene, 1,2,3−trichloropropene, hexachloroethane, 1,1−dichloroethane,

1,1,2,2−tetrachloroethane,

methylene

chloride,

1,2−dichloropropane,

1,1,1,3−

tetrachloropropane,

2−chloropropane,

1,2−dichloroethane,

2,3−dichlorobutane,

1,2−dichlorobutane,

1,2,3−trichloropropane, 1,2−dichlorobenzene, and 1,2,4−trichlorobenzene. In this study, in order to ensure about the adsorption of organic chlorides on the prepared adsorbent, the EDS 17 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

analysis of adsorbent was performed after the adsorption process. The analysis of sample no.1 (with average initial organic chloride concentration of 105 mg/L) was performed at 308 K and adsorption time of 10 h. The results reported in Table 3 confirmed the adsorption of such compounds on sintered alumina nanoparticles. Table 3 After the regeneration process, the EDS analyses of absorbents were performed. The EDS analyses of adsorbents surfaces after regeneration process indicated the complete elimination of organic chlorides via the evaporation process. In fact, no elements other than aluminum and oxygen were detected by this qualitative analysis after heating the adsorbents. The results of adsorbents analysis after regeneration process were the same with the analysis taken before adsorption process. These results also revealed that the adsorption process of organic chloride compounds on sintered γ−Al2O3 nanoparticles is a physisorption type, because the adsorbate−adsorbent bond cleaved at low temperature 104, 105.

4.3. Effect of Initial Adsorbate Concentration and Adsorption Time The organic chloride concentrations of final processed distillates after contacting with the adsorbent plotted as a function of adsorption time at constant adsorption temperatures in Figure 1. According to Figure 1, it is deduced that the sintered γ−Al2O3 nanoparticles could act as an appropriate adsorbent for the organic chloride compounds adsorption in the crude oil naphtha distillate. For each sample, the organic chlorides content of final processed naphtha fractions (after adsorption process) via the adsorption time showed the same behavior. According to this figure, the organic chloride content of samples continuously

18 ACS Paragon Plus Environment

Page 18 of 68

Page 19 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

decreased with increasing the adsorption time to reach the equilibrium point where no considerable change observed in the organic chlorides content of naphtha distillates. Figure 1 It seemed that for the first and second samples, adsorption equilibrium obtained after 8 h, because according to the experimental results plotted in Figure 1 (a, b and c) no considerable change in the organic chlorides concentration of final processed samples observed after 480 min. The equilibrium time decreased to 360 min for sample no. 4 and it reached to 240 min for sample no. 5. In all of these experiments, the amount of adsorbent and subsequently, the number of vacant sites of adsorption kept constant. It can be concluded that for the samples with lower organic chlorides concentration, the required time to reach the equilibrium condition is shorter than the samples with higher concentration.

4.3. Effect of Adsorption Temperature Temperature is an important parameter for the adsorption process 106-109. The organic chloride concentration of each sample at a certain adsorption time was plotted as a function of process temperature as shown in Figure 2 (sample no.1 and 5) and Figure S−4 (samples no.2 to 4). Figure 2 According to EDS analyses presented in the previous section, we inferred that due to physisorption adsorption, the chemical bond between adsorbent surface and adsorbate is a Wan der Waals type 110, 111. Therefore the higher temperature leads to higher liquid mobility. It means that the activation of liquid molecules increases and subsequently, the contact of these molecules with adsorbent surface enhances. This phenomenon could improve the 19 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

physical adsorption process. But, according to Figure 2, increase the distillate temperature more than a certain value has a negative effect on the adsorption process performance and some adsorbed organic chloride compounds desorbed from adsorbent. For naphtha distillates with lower initial organic chlorides content, this optimum condition occurred at higher temperatures, because for these samples, the chlorides molecules must be more energetic to reach the adsorbent surface and this is occurred at higher temperatures. The optimum temperature for the samples with initial concentration of 105, 82 and 55 mg/L was 318 K while this temperature rose to 328 K for the sample with initial concentration of 27 mg/L, and 338 K for the sample with initial organic chlorides concentration of 8.5 mg/L.

4.4. Adsorption Performance The amount of organic chlorides adsorption qt (mg/g) from naphtha distillate at different temperatures is presented in Figure 3 (sample no.1 and 5) and Figure S−5 (samples no.2 to 4). According to the obtained results, it is concluded that when the initial concentration of organic chlorides increased from 8.5 to 105 mg/L, the organic chloride adsorption per unit mass of sintered γ−Al2O3 nanoparticles enhanced. For example, qt increased from 0.22 to 1.51 mg/g as the initial organic chloride concentration increased from 8.5 to 105 mg/L at temperature of 298 K and adsorption time of 10 h. This manner is related to enhancement of mass transfer driving force due to higher concentration of organic chloride compounds in the naphtha distillate. In fact, with increasing the initial concentration of organic chloride in the naphtha fraction, the mass transfer driving force increased and therefore more organic chloride molecules transferred from the solution to the surface of the adsorbent. As shown in Figure 3 and Figure S−5, it can be seen that the highest qt value was obtained at the optimum temperature. Also, the maximum adsorption percentages of organic chloride compounds are 20 ACS Paragon Plus Environment

Page 20 of 68

Page 21 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

shown in Figure 4. It can be observed that at the constant amount of adsorbent, the final removal percentage increased with decreasing the initial organic chloride concentration from 105 to 8.5 mg/L. Figure 3 Figure 4 Figure S−6 (Supporting Information) illustrates the percentage removal of organic chloride compounds at different adsorption times and temperatures. According to this figure, the rate of organic chloride compounds removal from naphtha distillate was high at the beginning of adsorption process which is related to the large number of available adsorption sites. By passing the time, the repulsive force between organic chloride molecules on the surface of adsorbent and in the liquid phase enhances, and this phenomenon leads to difficult adsorption of organic chloride molecules on the remaining vacant sites of adsorbent. Therefore, in the following of process, the adsorption rate slowed down gradually.

4.5. Effect of Adsorbent Quantity The amount of sintered alumina nanoparticles is a determinative parameter on the adsorbent capacity and adsorption equilibrium. In this section, the sample with initial concentration of 105 mg/L was selected to investigate the effect of adsorbent quantity on the adsorption process. Six different quantities of γ−Al2O3 nanoparticles including 10, 15, 20, 25, 30 and 35 g were used. As shown in Figure 5 at different temperatures, the removal percentage of organic chloride compounds improved with increasing the amount of adsorbent. For example, the removal percentage of organic chloride compounds increased from 51.3% to a considerable value of 91.5% with increasing the amount of adsorbent from 10 to 35 g at 21 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

temperature of 298 K. Also, this increasing trend was observed for other adsorption temperatures. Especially, the maximum removal percentage was seen at optimum temperature of 318 K and the adsorbent amount of 35 g. Figure 5 Figure 6 shows the equilibrium adsorption capacity of the adsorbent, qe versus the amount of adsorbent at different temperatures. According to the plotted data in Figure 6, the equilibrium adsorption capacity of organic chloride per unit mass of sintered γ−Al2O3 nanoparticles decreased with increasing the amount of adsorbent. Indeed, the adsorption efficiency enhances with increasing the amount of adsorbent, due to increasing the effective absorbent surface area. Figure 6

4.6. Adsorption Isotherms Adsorption isotherms of organic chloride compounds on sintered γ−Al2O3 nanoparticles were determined for adsorption experimental data of five samples with initial organic chloride concentration of 8.5, 27, 55, 82 and 105 mg/L at different adsorption temperatures. The Langmuir, Temkin, Freundlich and Dubinin–Radushkevich (D–R) isotherm parameters for adsorption of organic chloride compounds on sintered γ−Al2O3 nanoparticles were obtained by linear fitting the experimental data. Figure 7 illustrates the fitting line for each of these models at different temperatures. Figure 7

22 ACS Paragon Plus Environment

Page 22 of 68

Page 23 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

As presented in Figure 7, at adsorption temperature of 298 K, the correlation coefficient R2 was 0.9409, 0.9793, 0.8547 and 0.7386 for Langmuir model, Freundlich isotherm model, Temkin isotherm model and Dubinin−Radushkevich model, respectively. The model with R2 parameter closer to 1.0 has the better performance to fit the experimental data 112, therefore it can be concluded that Dubinin−Radushkevich isotherm model exhibited the weakest fitting for the experimental data with respect to other used isotherm models. Normalized standard deviation (SD) is another important parameter that could be obtained using the following equation 113, 114: (22)

3

AB =

 ,GHI. −  ,JKL.  M N ,GHI. C∑ E −1



where qe,exp is the adsorption capacity of organic chloride at equilibrium (mg/g) derived from the experimental results, qe,cal is the adsorption capacity calculated by the corresponding models and n is the number of data. For better comparison between the isotherm models, their basic parameters were calculated according to the fitted lines shown in Figure 7 and the results reported in Table 4. Based on the calculated parameters of adsorption isotherms in Table 4, it can be deduced that Freundlich isotherm seemed to be the best one to fit the experimental data due to its largest R2 and the smallest SD values than that of Langmuir and Temkin isotherm models. This result is also obviously confirmed by Figure 8 for temperatures of 298 and 358 K, where the calculated equilibrium adsorption capacities qe,calc, of each isotherm model were compared with respect to experimental values, qexp. It is necessary to mention that the plots for the other five temperatures including 308, 318, 328, 338 and 348 K were shown in Figure S−7 (Supporting Information). As shown in these figures, the nearest qe,calc. values to experimental ones were obtained with Freundlich 23 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 68

isotherm model. Therefore, Freundlich adsorption isotherm model could be used to predict the adsorption of organic chloride compounds onto sintered γ−Al2O3 nanoparticles well. Table 4 Figure 8 4.7. Adsorption Kinetics The adsorption mechanism could be deeply understood using kinetic studies 79, 115. Eqs. (14), (16)

and

(17)

express

the

pseudo−first−order,

pseudo−second−order

and

intra−particle−diffusion kinetic models, respectively. Based on their linearized equations form, the experimental data were fitted and the final results were presented in Figures 9 for sample no. 1 with different initial organic chloride concentration of 105 mg/L. These graphs were plotted for the rest of the samples and shown in the Supporting Information file (Figures S−8 to S−11). Figure 9 When the pseudo−first−order kinetic model was employed to describe the experimental data, the values of qe, k1 and the correlation coefficient R2 were calculated from the linear plot of ln(qe− qt) versus t. Whereas, for fitting kinetic data using the linear form of the pseudo−second−order kinetic model, t/qt must be plotted versus t. According to the equation of this model, qe, k2 and correlation coefficient R2 could be calculated from the slope and intercept of the straight line obtained from the fitting line. In the case of intra−particle−diffusion model, qt plotted versus t1/2 and model parameters including Kint and I and also correlation coefficient R2 could be determined based on the fitted line to the experimental data. The mentioned kinetic models parameters and correlation coefficients R2 were calculated for different samples with different initial organic chloride concentration at 24 ACS Paragon Plus Environment

Page 25 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

various temperatures and the results were reported in Table 5. As reported in this table, the correlation coefficients R2, for intra−particle diffusion model and pseudo−second−order model are greater and closer to number 1.0 than that of pseudo−first−order kinetic model. For instance, for sample with initial organic chloride concentration of 55 mg/L, the correlation coefficients of intra−particle−diffusion model and pseudo−second−order model are 0.9772 and 0.9532 respectively, while this parameter is 0.8851 for pseudo−first−order kinetic model. Therefore, it can be resulted that for all of the samples, the intra−particle diffusion and pseudo−second−order kinetic models could better describe the kinetic of organic chloride adsorption on the sintered alumina nanoparticles. Based on the calculated results, for samples with lower initial concentrations, the pseudo−second−order model could better predict kinetic behavior of organic chloride compounds adsorption. Also, the reported parameters of intra−particle−diffusion kinetic model for the samples with different initial organic chloride concentrations showed that this diffusion model is not the only controlling step of organic chloride adsorption on sintered alumina nanoparticles. This is resulted from the fact that intra−particle−diffusion model intercept (I) is not zero for all of the samples data. Therefore, it can be said eventually that the kinetic of this adsorption process could not be described with only one kinetic model. Table 5

4.8. Adsorption Thermodynamics The thermodynamic parameters were calculated based on Eq. 21 using the slope and intercept of Ln Ke versus 1/T at different temperatures. More details about calculation procedure of thermodynamic parameters are presented in the Supporting Information. The final results were

25 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 68

reported in Table 6. The negative ∆G0 values implied that adsorption process of organic chloride onto the sintered γ−Al2O3 nanoparticles was possible and spontaneous. Moreover, the values of Gibbs free energy for physisorption change between −20 and 0 kJ/mole, while chemical adsorption ∆G0 is between −400 and −80 kJ/mole

116, 117

. Therefore, it can be

concluded that the organic chloride compounds were adsorbed physically to the surface of sintered alumina nanoparticles that is in accordance with experimental results. The enthalpy change for physical adsorption was reported to be in the range of -20 to 40 kJ/mol 118. Hence, the values of -1.05 to -4.98 kJ/mol for ∆H0 showed that physisorption is the dominant mechanism for the adsorption process of organic chloride compounds on the sintered alumina nanoparticles. This important conclusion confirmed the previous results. Furthermore, the negative value of ∆H0 for this process revealed that it is an exothermic process that is in agreement with the decreasing adsorption capacity with increasing temperature. The positive values of ∆S0 implied that the molecular randomness of the solution increased during adsorption of organic chloride compounds that probably caused by the increased mobility of organic chloride on the surface of adsorbent compared with that of naphtha distillate. Table 6

5. Conclusions In this research, the removal of organic chlorides compounds from naphtha fraction of contaminated crude oil via adsorption process onto sintered γ−Al2O3 nanoparticles was investigated. Different crude oil samples with various initial concentrations of organic chlorides were taken from an industrial processing plant. The adsorption process performed at different adsorption time and temperatures. The EDS analyses before and after 26 ACS Paragon Plus Environment

Page 27 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

regeneration of nano−alumina adsorbent showed that the adsorption of organic chlorides is a physisorption type. Thermodynamic calculation of Gibbs free energy and enthalpy change also confirmed the physical adsorption of organic chloride compounds on the adsorbent. In order to investigate the effect of adsorbent dosage, different quantities of sintered alumina nanoparticles were employed. The results indicated that with increasing the amount of adsorbent, the removal percentage enhanced considerably while adsorption capacity, qe, decreased. Application of different adsorption isotherm models revealed that based on the correlation coefficients factor (R2) and standard deviation (SD), the Freundlich isotherm model could best describe the adsorption equilibrium of organic chloride compounds. Also, the kinetic mechanism of organic chloride adsorption process was investigated using the most applicable kinetic models including pseudo−first−order, pseudo−second−order and intra−particle−diffusion models. The correlation coefficients R2 of fitting these models to the experimental data indicated that intra−particle−model exhibited the best fitting for all of the verified samples. Finally, according to the experimental and theoretical interpretation results of this study, it can be concluded that sintered γ−Al2O3 nanoparticles could act as an effective adsorbent for adsorption removal of organic chloride compounds from naphtha fraction of crude oil.

27 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References 1.

Wu, B.; Li, X.; Li, Y.; Zhu, J.; Wang, J., Hydrolysis Reaction Tendency of Low-

Boiling Organic Chlorides To Generate Hydrogen Chloride in Crude Oil Distillation. Energy & Fuels 2016, 30, (2), 1524-1530. 2.

National Association of Corrosion Engineers (NACE). Effect of Nonextractable

Chlorides on Refinery Corrosion and Fouling; NACE International: Houston, TX, 2004; Paper 34105. 3.

Chambers, B.; Srinivasan, S., Yap, K. M., Yunovich, M. Corrosion in Crude

Distillation Unit Overhead Operations: A Comprehensive Review; NACE International: Houston, TX, 2011; Paper 11360. 4.

Elnour, M.; Gasmelseed, G. A.; Abdalla, B. K., The Effects of High HCl and Changes

in pH Levels in CDU Overhead Corrosion. Journal of Applied and Industrial Sciences 2014, 2, (5), 238-243. 5.

Seadat-Talab, M.; Allahkaram, S. R., Failure analysis of overhead flow cooling

systems of a light naphtha separator tower at a petrochemical plant. Engineering Failure Analysis 2013, 27, 130-140. 6.

Santhana Prabha, S.; Joseph Rathish, R.; Dorothy, R.; Brindha, G.; Pandiarajan, M.;

Al-Hashem, A.; Rajendran, S., Corrosion problems in petroleum industry and their solutions. European Chemical Bulletin 2014, 3, (3), 300-307. 7.

Mozdianfard, M. R.; Behranvand, E., Fouling at post desalter and preflash drum heat

exchangers of CDU preheat train. Applied Thermal Engineering 2015, 89, 783-794. 8.

Ancheyta, J. Deactivation of Heavy Oil Hydroprocessing Catalysts: Fundamentals

and Modeling; Wiley: New Jersey, 2016. 9.

Huang, H.; Wang, S.; Wang, S.; Cao, G., Deactivation Mechanism of Cu/Zn Catalyst

Poisoned by Organic Chlorides in Hydrogenation of Fatty Methyl Ester to Fatty Alcohol. Catalysis Letters 2010, 134, (3), 351-357. 10.

Wu, B.; Li, Y.; Li, X.; Zhu, J., Distribution and Identification of Chlorides in

Distillates from YS Crude Oil. Energy & Fuels 2015, 29, (3), 1391-1396. 11.

Alvisi, P. P.; de Freitas Cunha Lins, V., Acid salt corrosion in a hydrotreatment plant

of a petroleum refinery. Engineering Failure Analysis 2008, 15, (8), 1035-1041. 12.

Yao, J. B., Chlorine Corrosion in Catalytic Reforming Units and Prevention.

Corrosion & Protection in Petrochemical Industry 2008, 25, (1), 56−58. 28 ACS Paragon Plus Environment

Page 28 of 68

Page 29 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

13.

Adan Sun, D. F. Prediction, Monitoring, and Control of Ammonium Chloride

Corrosion in Refining Processes; 10359; SanAntonio, TX, USA, 2010; p 17. 14.

Wu, F.; Li, H., Study on the divided-wall electric desalting technology for Suizhong

crude oil. Desalination 2012, 307, 20-25. 15.

Kaur, H.; Eaton, P.; Gray, M. R., The Kinetics and Inhibition of Chloride Hydrolysis

in Canadian Bitumen. Petroleum Science and Technology 2012, 30, (10), 993-1003. 16.

Kaur, H. Kinetics & Inhibition of Chloride Hydrolysis in Canadian Bitumens.

Alberta, Canada, 2009. 17.

Gogotov, A. F.; Silinskaya, Y. N.; Kolotov, V. Y.; Tomin, V. P., Influence of various

additives on base hydrolysis of chlorine-containing organic impurities in naphtha fractions. Russian Journal of Applied Chemistry 2004, 77, (11), 1902-1903. 18.

Ma, R.; Zhu, J.; Wu, B.; Li, X., Adsorptive removal of organic chloride from model

jet fuel by Na-LSX zeolite: Kinetic, equilibrium and thermodynamic studies. Chemical Engineering Research and Design 2016, 114, 321-330. 19.

Ma, R.; Zhu, J.; Wu, B.; Hu, J.; Li, X., Distribution and Qualitative and Quantitative

Analyses of Chlorides in Distillates of Shengli Crude Oil. Energy & Fuels 2017, 31, (1), 374378. 20.

Groysman, A., Physicochemical Basics of Corrosion at Refineriesʼ Units. In

Corrosion Problems and Solutions in Oil Refining and Petrochemical Industry, Springer International Publishing: Cham, 2017; pp 17-36. 21.

Gutzeit, J., Effect of organic chloride contamination of crude oil on refinery

corrosion. In CORROSION 2000, NACE International: Orlando, Florida, USA, 2000. 22.

Zhan, B. Z.; Driver, M.; Timken, H. K. Catalytic dechlorination processes to upgrade

feedstock containing chloride as fuels. US 8795515 B2, 2014. 23.

Zhan, B. Z.; Timken, H. K. C.; DRIVER, M. S. Process for reducing chloride in

hydrocarbon products using an ionic liquid catalyst. US 8969645, 2014. 24.

Aramendı́a, M. A.; Boráu, V.; Garcı́a, I. M.; Jiménez, C.; Lafont, F.; Marinas, A.;

Marinas, J. M.; Urbano, F. J., Liquid-phase hydrodechlorination of chlorobenzene over palladium-supported catalysts: Influence of HCl formation and NaOH addition. Journal of Molecular Catalysis A: Chemical 2002, 184, (1–2), 237-245. 25.

Luo, Y. R., Handbook of Bond Dissociation Energies in Organic Compounds. CRC

Press: New Yourk, 2002.

29 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

26.

Li, X. H.; Tang, Z. X.; Zhang, X. Z., DFT study of the C-Cl bond dissociation

enthalpies and electronic structure of substituted chlorobenzene compounds. Journal of Structural Chemistry 2009, 50, (1), 34. 27.

MARVE, M. G.; PAREKH, A. A.; DAS, A. K.; Rajeshwar, D.; Rana, D. P. S.;

BISHT, H.; HITESH, K. S.; SINGH, J. A. I. K.; KALYAN, N.; Yadav, M. A method for removing chlorides from hydrocarbon stream by steam stripping. WO2014033733 A1, 2014. 28.

McWilliams, J. P.; Nemet-Mavrodin, M. I.; Sigal, C. T.; Wilson, R. C. Guard bed

catalyst for organic chloride removal from hydrocarbon feed. US4721824 A, 1988. 29.

Kim, H.; Lee, J. J.; Moon, S. H., Hydrodesulfurization of dibenzothiophene

compounds using fluorinated NiMo/Al2O3 catalysts. Applied Catalysis B: Environmental 2003, 44, (4), 287-299. 30.

Girgis, M. J.; Gates, B. C., Reactivities, reaction networks, and kinetics in high-

pressure catalytic hydroprocessing. Industrial & Engineering Chemistry Research 1991, 30, (9), 2021-2058. 31.

Johnson, T. H. Reduction of residual organic chlorine in hydrocarbyl amines.

USH1143 H, 1993. 32.

Maglio, A.; McCaffrey, R. T. Organic Chloride Adsorbent. US8551328 B2, 2012.

33.

Lacher, J. R.; Kianpour, A.; Oetting, F.; Park, J. D., Reaction calorimetry. The

hydrogenation of organic fluorides and chlorides. Transactions of the Faraday Society 1956, 52, (0), 1500-1508. 34.

Reusser, R. E. Removal of chemically combined chlorine and other impurities from

hydrocarbons. US3864243 A, 1975. 35.

Lanin, S. N.; Bannykh, A. A.; Kovaleva, N. V., Adsorption of chlorobenzenes on

ultrafine diamond modified with palladium and nickel nanoparticles. Russian Journal of Physical Chemistry A 2013, 87, (9), 1550-1555. 36.

Hill, J. W.; Fry, A., Chlorine Isotope Effects in the Reactions of Benzyl and

Substituted Benzyl Chlorides with Various Nucleophiles. Journal of the American Chemical Society 1962, 84, (14), 2763-2769. 37.

Kauppila, T. J.; Haack, A.; Kroll, K.; Kersten, H.; Benter, T., Nucleophilic Aromatic

Substitution Between Halogenated Benzene Dopants and Nucleophiles in Atmospheric Pressure Photoionization. Journal of The American Society for Mass Spectrometry 2016, 27, (3), 422-431.

30 ACS Paragon Plus Environment

Page 30 of 68

Page 31 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

38.

Landini, D.; Maia, A. M.; Montanari, F.; Pirisi, F. M., Phase-transfer mechanism and

nucleophilicity of halide ions in an aqueous-organic two-phase system. Journal of the Chemical Society, Chemical Communications 1975, (23), 950-951. 39.

Asnin, L. D.; Fedorov, A. A.; Chekryshkin, Y. S., Adsorption of chlorobenzene and

benzene on γ-Al2O3. Russian Chemical Bulletin 2001, 50, (1), 68-72. 40.

Wang, H.; Schwieger, W.; Huang, Y., An investigation of the conformational

behavior of a chlorinated hydrocarbon, 1,1,2-trichloroethane, adsorbed in zeolites by Raman spectroscopy. Canadian Journal of Chemistry 2014, 93, (1), 118-125. 41.

Khaleghi Abbasabadi, M.; Rashidi, A.; Khodabakhshi, S., Benzenesulfonic acid-

grafted graphene as a new and green nanoadsorbent in hydrogen sulfide removal. Journal of Natural Gas Science and Engineering 2016, 28, (Supplement C), 87-94. 42.

Tegge, B. R.; Weary, F. G.; Sakaguchi, Y. Removal of organic halides from

hydrocarbon solvents. US 4713413 A, 1987. 43.

Chinnakoti, P.; Chunduri, A. L. A.; Vankayala, R. K.; Patnaik, S.; Kamisetti, V.,

Enhanced fluoride adsorption by nano crystalline γ-alumina: adsorption kinetics, isotherm modeling and thermodynamic studies. Applied Water Science 2016, 1-11. 44.

Kuemmel, M.; Grosso, D.; Boissière, C.; Smarsly, B.; Brezesinski, T.; Albouy, P. A.;

Amenitsch, H.; Sanchez, C., Thermally Stable Nanocrystalline γ-Alumina Layers with Highly Ordered 3D Mesoporosity. Angewandte Chemie International Edition 2005, 44, (29), 4589-4592. 45.

Zhu, L.; Zhang, X.; Zhu, L.; Li, X.; Meng, L., Synthesis and characterization of

mesoporous alumina, and adsorption performance for n-butane. Research on Chemical Intermediates 2015, 41, (6), 3637-3648. 46.

Men, Y.; Gnaser, H.; Ziegler, C., Adsorption/desorption studies on nanocrystalline

alumina surfaces. Analytical and Bioanalytical Chemistry 2003, 375, (7), 912-916. 47.

Wang, Y.; Suryanarayana, C.; An, L., Phase Transformation in Nanometer-Sized γ-

Alumina by Mechanical Milling. Journal of the American Ceramic Society 2005, 88, (3), 780-783. 48.

Amirsalari, A.; Farjami Shayesteh, S., Effects of pH and calcination temperature on

structural and optical properties of alumina nanoparticles. Superlattices and Microstructures 2015, 82, 507-524.

31 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

49.

ASTM, Standard Test Method for Distillation of Petroleum Products and Liquid Fuels

at Atmospheric Pressure. In ASTM D86-16a, ASTM International: West Conshohocken, PA, 2016. 50.

ASTM, Standard Test Method for Determination of Organic Chloride Content in

Crude Oil. In ASTM D4929-16, ASTM: West Conshohocken, USA, 2016. 51.

Doulia, D.; Leodopoulos, C.; Gimouhopoulos, K.; Rigas, F., Adsorption of humic

acid on acid-activated Greek bentonite. Journal of Colloid and Interface Science 2009, 340, (2), 131-141. 52.

Wu, Z.; Joo, H.; Lee, K., Kinetics and thermodynamics of the organic dye adsorption

on the mesoporous hybrid xerogel. Chemical Engineering Journal 2005, 112, (1–3), 227-236. 53.

Ma, J.; Jia, Y.; Jing, Y.; Yao, Y.; Sun, J., Kinetics and thermodynamics of methylene

blue adsorption by cobalt-hectorite composite. Dyes and Pigments 2012, 93, (1–3), 14411446. 54.

S.E, S.; N, T., Isotherm, kinetic and thermodynamic studies on the adsorption

behaviour of textile dyes onto chitosan. Process Safety and Environmental Protection 2017, 106, 1-10. 55.

Khalighi Sheshdeh, R.; Khosravi Nikou, M. R.; Badii, K.; Mohammadzadeh, S.,

Evaluation of Adsorption Kinetics and Equilibrium for the Removal of Benzene by Modified Diatomite. Chemical Engineering & Technology 2013, 36, (10), 1713-1720. 56.

Aleghafouri, A.; Hasanzadeh, N.; Mahdyarfar, M.; SeifKordi, A.; Mahdavi, S. M.;

Zoghi, A. T., Experimental and theoretical study on BTEX removal from aqueous solution of diethanolamine using activated carbon adsorption. Journal of Natural Gas Science and Engineering 2015, 22, (Supplement C), 618-624. 57.

Kumar, P. S.; Vincent, C.; Kirthika, K.; Kumar, K. S., Kinetics and equilibrium

studies of Pb2+ in removal from aqueous solutions by use of nano-silversol-coated activated carbon. Brazilian Journal of Chemical Engineering 2010, 27, 339-346. 58.

Hall, K. R.; Eagleton, L. C.; Acrivos, A.; Vermeulen, T., Pore- and Solid-Diffusion

Kinetics in Fixed-Bed Adsorption under Constant-Pattern Conditions. Industrial & Engineering Chemistry Fundamentals 1966, 5, (2), 212-223. 59.

Wahab, M. A.; Jellali, S.; Jedidi, N., Ammonium biosorption onto sawdust: FTIR

analysis, kinetics and adsorption isotherms modeling. Bioresource Technology 2010, 101, (14), 5070-5075.

32 ACS Paragon Plus Environment

Page 32 of 68

Page 33 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

60.

Elmorsi, T. M.; Elsayed, M. H.; Bakr, M. F., Enhancing the removal of methylene

blue by modified ZnO nanoparticles: kinetics and equilibrium studies. Canadian Journal of Chemistry 2017, 95, (5), 590-600. 61.

Ebadi, A.; Soltan Mohammadzadeh, J. S.; Khudiev, A., Adsorption of Methyl tert-

butyl Ether on Perfluorooctyl Alumina Adsorbents – High Concentration Range. Chemical Engineering & Technology 2007, 30, (12), 1666-1673. 62.

Mahmoud, M. E.; Amira, M. F.; Seleim, S. M.; Mohamed, A. K., Adsorption

Isotherm Models, Kinetics Study, and Thermodynamic Parameters of Ni(II) and Zn(II) Removal from Water Using the LbL Technique. Journal of Chemical & Engineering Data 2017, 62, (2), 839-850. 63.

Hassan, H. S.; Elmaghraby, E. K., Preparation of graphite by thermal annealing of

polyacrylamide precursor for adsorption of Cs(I) and Co(II) ions from aqueous solutions. Canadian Journal of Chemistry 2012, 90, (10), 843-850. 64.

Temkin, M. J.; Pyzhev, V., Recent Modifications to Langmuir Isotherms. Acta

Physiochim URSS 1940, 12, 6. 65.

Kim, Y.; Kim, C.; Choi, I.; Rengaraj, S.; Yi, J., Arsenic Removal Using Mesoporous

Alumina Prepared via a Templating Method. Environmental Science & Technology 2004, 38, (3), 924-931. 66.

Febrianto, J.; Kosasih, A. N.; Sunarso, J.; Ju, Y.-H.; Indraswati, N.; Ismadji, S.,

Equilibrium and kinetic studies in adsorption of heavy metals using biosorbent: A summary of recent studies. Journal of Hazardous Materials 2009, 162, (2–3), 616-645. 67.

Freundlich, H., Über die adsorption in Lösungen. Zeitschrift für Physikalische Chemie

1906, 57, 6. 68.

Zhou, Q.; Duan, Y.; Chen, M.; Liu, M.; Lu, P., Studies on Mercury Adsorption

Species and Equilibrium on Activated Carbon Surface. Energy & Fuels 2017. 69.

Zhang, H.; Lan, X.; Bai, P.; Guo, X., Adsorptive removal of acetic acid from water

with metal-organic frameworks. Chemical Engineering Research and Design 111, 127-137. 70.

Pal, P.; Banat, F.; AlShoaibi, A., Adsorptive removal of heat stable salt anions from

industrial lean amine solvent using anion exchange resins from gas sweetening unit. Journal of Natural Gas Science and Engineering 2013, 15, (Supplement C), 14-21. 71.

Aziam, R.; Chiban, M.; Eddaoudi, H.; Soudani, A.; Zerbet, M.; Sinan, F., Kinetic

modeling, equilibrium isotherm and thermodynamic studies on a batch adsorption of anionic

33 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 68

dye onto eco-friendly dried Carpobrotus edulis plant. The European Physical Journal Special Topics 2017, 226, (5), 977-992. 72.

Hill, J. M.; Ng, F. T. T., Adsorption of etioporphyrin and Ni-etioporphyrin on a fractal

silica. Canadian Journal of Chemistry 2001, 79, (5-6), 817-822. 73.

Manique, M. C.; Silva, A. P.; Alves, A. K.; Bergmann, C. P., TITANATE

NANOTUBES PRODUCED FROM MICROWAVE-ASSISTED HYDROTHERMAL SYNTHESIS:

CHARACTERIZATION,

ADSORPTION

AND

PHOTOCATALYTIC

ACTIVITY. Brazilian Journal of Chemical Engineering 2017, 34, 331-339. 74.

Günay, A.; Arslankaya, E.; Tosun, Đ., Lead removal from aqueous solution by natural

and pretreated clinoptilolite: Adsorption equilibrium and kinetics. Journal of Hazardous Materials 2007, 146, (1–2), 362-371. 75.

Radushkevich, L. V.; Dubinin, M. M., Equation of the characteristic curve of

activated charcoal. . Chem. Zentr. 1947, 1. 76.

Dubinin, M. M., The Potential Theory of Adsorption of Gases and Vapors for

Adsorbents with Energetically Nonuniform Surfaces. Chemical Reviews 1960, 60, (2), 235241. 77.

Kilic, M.; Apaydin-Varol, E.; Pütün, A. E., Adsorptive removal of phenol from

aqueous solutions on activated carbon prepared from tobacco residues: Equilibrium, kinetics and thermodynamics. Journal of Hazardous Materials 2011, 189, (1–2), 397-403. 78.

Saleh, T. A.; Sarı, A.; Tuzen, M., Effective adsorption of antimony(III) from aqueous

solutions by polyamide-graphene composite as a novel adsorbent. Chemical Engineering Journal 2017, 307, 230-238. 79.

Monte Blanco, S. P. D.; Scheufele, F. B.; Módenes, A. N.; Espinoza-Quiñones, F. R.;

Marin, P.; Kroumov, A. D.; Borba, C. E., Kinetic, equilibrium and thermodynamic phenomenological modeling of reactive dye adsorption onto polymeric adsorbent. Chemical Engineering Journal 2017, 307, 466-475. 80.

Lagergren, S., Zur Theorie der Sogenannten Adsorption Gelöster Stoffe, Kungliga

Svenska Vetenskapsakademiens. Handlingar 1898, 24, (4), 39. 81.

Ho, Y. S.; McKay, G., Pseudo-second order model for sorption processes. Process

Biochemistry 1999, 34, (5), 451-465. 82.

Ho, Y. S.; McKay, G.; Wase, D. A. J.; Forster, C. F., Study of the Sorption of

Divalent Metal Ions on to Peat. Adsorption Science & Technology 2000, 18, (7), 639-650.

34 ACS Paragon Plus Environment

Page 35 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

83.

Weber, W. J.; Morris, J. C., Kinetics of adsorption carbon from solutions. . Journal

Sanitary Engeering Division Proceedings 1963, 89, 31-60. 84.

Asuquo, E.; Martin, A.; Nzerem, P.; Siperstein, F.; Fan, X., Adsorption of Cd(II) and

Pb(II) ions from aqueous solutions using mesoporous activated carbon adsorbent: Equilibrium, kinetics and characterisation studies. Journal of Environmental Chemical Engineering 2017, 5, (1), 679-698. 85.

de Godoi, F. C.; Rabelo, R. B.; Silva, M. A.; Rodríguez-Castellón, E.; Guibal, E.;

Beppu, M. M., Introduction of copper nanoparticles in chitosan matrix as strategy to enhance chromate adsorption. Chemical Engineering and Processing: Process Intensification 2014, 83, 43-48. 86.

Triki, M.; Tanazefti, H.; Kochkar, H., Design of β-cyclodextrin modified TiO2

nanotubes for the adsorption of Cu(II): Isotherms and kinetics study. Journal of Colloid and Interface Science 2017, 493, 77-84. 87.

Alwary, L.; Gafar, M.; Rumie, A., Liquid Phase Adsorption of Phenol and

Chloroform by Activated Charcoal. Chemical Engineering & Technology 2011, 34, (11), 1883-1890. 88.

Pal, P.; Banat, F., Comparison of heavy metal ions removal from industrial lean amine

solvent using ion exchange resins and sand coated with chitosan. Journal of Natural Gas Science and Engineering 2014, 18, (Supplement C), 227-236. 89.

Nassar, N. N., Asphaltene Adsorption onto Alumina Nanoparticles: Kinetics and

Thermodynamic Studies. Energy & Fuels 2010, 24, (8), 4116-4122. 90.

Wibowo, E.; Rokhmat, M.; Sutisna; Khairurrijal; Abdullah, M., Reduction of

seawater salinity by natural zeolite (Clinoptilolite): Adsorption isotherms, thermodynamics and kinetics. Desalination 2017, 409, 146-156. 91.

Ismail, A. I., Thermodynamic and kinetic properties of the adsorption of 4-

nitrophenol on graphene from aqueous solution. Canadian Journal of Chemistry 2015, 93, (10), 1083-1087. 92.

Srivastava, V. C.; Swamy, M. M.; Mall, I. D.; Prasad, B.; Mishra, I. M., Adsorptive

removal of phenol by bagasse fly ash and activated carbon: Equilibrium, kinetics and thermodynamics. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2006, 272, (1–2), 89-104.

35 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

93.

Zhang, B.; Ma, Z.; Yang, F.; Liu, Y.; Guo, M., Adsorption properties of ion

recognition rice straw lignin on PdCl42−: Equilibrium, kinetics and mechanism. Colloids and Surfaces A: Physicochemical and Engineering Aspects 2017, 514, 260-268. 94.

Çelekli, A.; Đlgün, G.; Bozkurt, H., Sorption equilibrium, kinetic, thermodynamic, and

desorption studies of Reactive Red 120 on Chara contraria. Chemical Engineering Journal 2012, 191, 228-235. 95.

Ge, X.; Tian, F.; Wu, Z.; Yan, Y.; Cravotto, G.; Wu, Z., Adsorption of naphthalene

from aqueous solution on coal-based activated carbon modified by microwave induction: Microwave power effects. Chemical Engineering and Processing: Process Intensification 2015, 91, 67-77. 96.

Liu, Y.; Tian, S.; Meng, X.; Dai, X.; Liu, Z.; Meng, M.; Han, J.; Wang, Y.; Chen, R.;

Yan, Y.; Ni, L., Synthesis, characterization, and adsorption properties of a Ce(III)-imprinted polymer supported by mesoporous SBA-15 matrix by a surface molecular imprinting technique. Canadian Journal of Chemistry 2014, 92, (3), 257-266. 97.

Mafra, M. R.; Igarashi-Mafra, L.; Zuim, D. R.; Vasques, É. C.; Ferreira, M. A.,

Adsorption of remazol brilliant blue on an orange peel adsorbent. Brazilian Journal of Chemical Engineering 2013, 30, 657-665. 98.

Rambabu, N.; Guzman, C. A.; Soltan, J.; Himabindu, V., Adsorption Characteristics

of Atrazine on Granulated Activated Carbon and Carbon Nanotubes. Chemical Engineering & Technology 2012, 35, (2), 272-280. 99.

Deng, Y.; Liu, N.-Y.; Tsow, F.; Xian, X.; Forzani, E. S., Adsorption Thermodynamic

Analysis of a Quartz Tuning Fork Based Sensor for Volatile Organic Compounds Detection. ACS Sensors 2017, 2, (11), 1662-1668. 100.

Rozita, Y.; Brydson, R.; Scott, A. J., An investigation of commercial gamma-Al 2 O 3

nanoparticles. Journal of Physics: Conference Series 2010, 241, (1), 012096. 101.

Piriyawong, V.; Thongpool, V.; Asanithi, P.; Limsuwan, P., Preparation and

Characterization of Alumina Nanoparticles in Deionized Water Using Laser Ablation Technique. Journal of Nanomaterials 2012, 2012, 6. 102.

Munhoz, A. H.; Paiva, H. d.; Miranda, L. F. d.; Oliveira, E. C. d.; Andrades, R. C.;

Ribeiro., R. R., Study of Gamma Alumina Synthesis – Analysis of the Specific Surface Area. Advances in Science and Technology 2014, 87, 54-60. 103.

Dubey, S.; Singh, A.; Nim, B.; Singh, I. B., Optimization of molar concentration of

AlCl3 salt in the sol–gel synthesis of nanoparticles of gamma alumina and their application in 36 ACS Paragon Plus Environment

Page 36 of 68

Page 37 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

the removal of fluoride of water. Journal of Sol-Gel Science and Technology 2017, 82, (2), 468-477. 104.

Chowanietz, V.; Pasel, C.; Luckas, M.; Eckardt, T.; Bathen, D., Desorption of

Mercaptans and Water from a Silica–Alumina Gel. Industrial & Engineering Chemistry Research 2017, 56, (2), 614-621. 105.

Kulkarni, S.; Kaware, J., Regeneration and Recovery in Adsorption- a Review.

IJISET 2014, 1, (8), 61-64. 106.

López Valdivieso, A.; Reyes Bahena, J. L.; Song, S.; Herrera Urbina, R., Temperature

effect on the zeta potential and fluoride adsorption at the α-Al2O3/aqueous solution interface. Journal of Colloid and Interface Science 2006, 298, (1), 1-5. 107.

Hongxia, Z.; Xiaoyun, W.; Honghong, L.; Tianshe, T.; Wangsuo, W., Adsorption

behavior of Th(IV) onto illite: Effect of contact time, pH value, ionic strength, humic acid and temperature. Applied Clay Science 2016, 127–128, 35-43. 108.

Santos, R. M. M. d.; Gonçalves, R. G. L.; Constantino, V. R. L.; Santilli, C. V.;

Borges, P. D.; Tronto, J.; Pinto, F. G., Adsorption of Acid Yellow 42 dye on calcined layered double hydroxide: Effect of time, concentration, pH and temperature. Applied Clay Science 2017, 140, 132-139. 109.

Marczewski, A. W.; Seczkowska, M.; Deryło-Marczewska, A.; Blachnio, M.,

Adsorption equilibrium and kinetics of selected phenoxyacid pesticides on activated carbon: effect of temperature. Adsorption 2016, 22, (4), 777-790. 110.

Ruiz, V. G.; Liu, W.; Tkatchenko, A., Density-functional theory with screened van

der Waals interactions applied to atomic and molecular adsorbates on close-packed and nonclose-packed surfaces. Physical Review B 2016, 93, (3), 035118. 111.

Wei, L.; Victor, G. R.; Guo-Xu, Z.; Biswajit, S.; Xinguo, R.; Matthias, S.; Alexandre,

T., Structure and energetics of benzene adsorbed on transition-metal surfaces: densityfunctional theory with van der Waals interactions including collective substrate response. New Journal of Physics 2013, 15, (5), 053046. 112.

Dehghani, M. H.; Mohammadi, M.; Mohammadi, M. A.; Mahvi, A. H.; Yetilmezsoy,

K.; Bhatnagar, A.; Heibati, B.; McKay, G., Equilibrium and Kinetic Studies of Trihalomethanes Adsorption onto Multi-walled Carbon Nanotubes. Water, Air, & Soil Pollution 2016, 227, (9), 332.

37 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

113.

Saadi, R.; Saadi, Z.; Fazaeli, R.; Fard, N. E., Monolayer and multilayer adsorption

isotherm models for sorption from aqueous media. Korean Journal of Chemical Engineering 2015, 32, (5), 787-799. 114.

Fallou, H.; Cimetière, N.; Giraudet, S.; Wolbert, D.; Le Cloirec, P., Adsorption of

pharmaceuticals onto activated carbon fiber cloths – Modeling and extrapolation of adsorption isotherms at very low concentrations. Journal of Environmental Management 2016, 166, 544-555. 115.

Lin, X.; Huang, Q.; Qi, G.; Xiong, L.; Huang, C.; Chen, X.; Li, H.; Chen, X.,

Adsorption behavior of levulinic acid onto microporous hyper-cross-linked polymers in aqueous solution: Equilibrium, thermodynamic, kinetic simulation and fixed-bed column studies. Chemosphere 2017, 171, 231-239. 116.

Almeida, C. A. P.; Debacher, N. A.; Downs, A. J.; Cottet, L.; Mello, C. A. D.,

Removal of methylene blue from colored effluents by adsorption on montmorillonite clay. Journal of Colloid and Interface Science 2009, 332, (1), 46-53. 117.

Wang, F.; Pan, Y.; Cai, P.; Guo, T.; Xiao, H., Single and binary adsorption of heavy

metal ions from aqueous solutions using sugarcane cellulose-based adsorbent. Bioresource Technology 2017, 241, 482-490. 118.

Lian, L.; Guo, L.; Guo, C., Adsorption of Congo red from aqueous solutions onto Ca-

bentonite. Journal of Hazardous Materials 2009, 161, (1), 126-131.

38 ACS Paragon Plus Environment

Page 38 of 68

Page 39 of 68 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

List of Tables: Table 1. Main characteristics of the verified crude oil. Table 2. Organic chloride concentration of five different contaminated crude oil samples. Table 3. EDS analysis of adsorbent surface after adsorption process for sample no.1 after adsorption time of 10 h for sample no. 1. Table 4. Calculated Langmuir, Temkin and Freundlich isotherm models parameters. Table 5. Comparison of the pseudo−first−order, pseudo−second−order and intra particle diffusion kinetic models for the adsorption of organic chloride onto sintered γ−Al2O3 nanoparticles for various initial concentrations and temperatures. Table 6. Thermodynamic parameters for adsorption process of organic chloride compounds onto sintered γ−Al2O3 nanoparticles under different temperatures.

39 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 68

Table 1 Characteristics

Units

Result

Test Method

Specific Gravity @ 15.56°C

−−−

0.8750

ASTM D5002

API Gravity

°API

31

ASTM D4294

Sulfur Content(Total)

wt.%

1.83

IP 170

H2S Content

ppm

1

UOP 163

Mercaptan Content

ppm

185

ASTM D5762

Nitrogen Content (Total)

wt.%

0.23

ASTM D4007

Water & Sediment

vol.%