Next Article in Journal
Effects of Catalyst Ink Storage on Polymer Electrolyte Fuel Cells
Previous Article in Journal
Anomaly Detection for Wind Turbines Using Long Short-Term Memory-Based Variational Autoencoder Wasserstein Generation Adversarial Network under Semi-Supervised Training
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Maximizing Liquid Fuel Production from Reformed Biogas by Kinetic Studies and Optimization of Fischer–Tropsch Reactions

by
Firas K. Al-Zuhairi
1,
Zaidoon M. Shakor
1 and
Ihsan Hamawand
2,*
1
Chemical Engineering Department, University of Technology-Iraq, Baghdad 10066, Iraq
2
Wide Bay Water Process Operations, Fraser Coast Regional Council, Urangan, QLD 4655, Australia
*
Author to whom correspondence should be addressed.
Energies 2023, 16(19), 7009; https://doi.org/10.3390/en16197009
Submission received: 14 August 2023 / Revised: 26 September 2023 / Accepted: 28 September 2023 / Published: 9 October 2023
(This article belongs to the Section A4: Bio-Energy)

Abstract

:
In the current work, the operating conditions for the Fischer–Tropsch process were optimized using experimental testing, kinetic modelling, simulation, and optimization. The experiments were carried out using a Ce-Co/SiO2 catalyst to examine how operating parameters affected the conversion of CO and product selectivity. A power-law kinetic model was used to represent the reaction rates in a mathematical model that was created to replicate the Fischer–Tropsch synthesis (FTS). It was decided to estimate the kinetic parameters using a genetic optimization technique. The developed model was validated for a range of operating conditions, including a temperature range of 200–240 °C, a pressure range of 5–25 bar, a H2/CO ratio of 0.5–4, and a space velocity range of 1000–5000 mL/gcat·h. The mean absolute relative error (MARE) between the experimental and predicted results was found to be 11.7%, indicating good agreement between the experimental data and the predicted results obtained by the mathematical model. Optimization was applied to maximize the production of liquid biofuels (C5+). The maximum C5+ selectivity was 91.66, achieved at an operating temperature of 200 °C, reactor total pressure of 6.29 bar, space velocity of 1529.58 mL/gcat·h, and a H2/CO feed ratio of 3.96. The practical implications of the present study are maximizing liquid biofuel production from biomass and municipal solid waste (MSW) as a renewable energy source to meet energy requirements, reducing greenhouse gas emissions, and waste management.

1. Introduction

In recent years, increasing energy consumption has resulted from evolution in different transportation and industrial sectors, expansion in the human community, and a changing role for fossil fuels in the climate crisis. Therefore, the prospect of strict regulations on fossil fuel exploration has opened up opportunities for increasing the production of clean energy [1,2]. The Fischer–Tropsch reaction (FTR) system is considered an appealing method of producing liquid fuels and feedstock materials for petrochemicals from unconventional sources of petroleum, including coal, natural gas, and biomass. This process involves converting a syngas mixture of H2 and CO over many types of catalysts [3]. The distribution of products in the Fischer–Tropsch reaction depends mainly on the reaction conditions, including operating temperature, total pressure, gas hourly space velocity, and H2/CO ratio, as well as catalyst types and their physical characteristics [4].
Anaerobic digestion (AD) of municipal solid waste and biomass is an important technology that allows for the generation of extra value in the form of biogas. This process is part of circular economic creativity for power production [5,6,7]. Biogas is a stimulating carbon source due to its CO2 content, which can be used to convert it into different ratios of syngas (a mixture of H2 and CO) through dry reforming after upgrading and transformation [8].
The liquid fuels produced via FTR are clean fuels with low contents of nitrogen, aromatics, and sulfur compounds. To maximize the production of liquid fuel, it is necessary to choose a suitable catalyst, optimize reaction conditions, and develop a mathematical model to describe the reaction kinetics to achieve higher reactant conversion while minimizing the production of light hydrocarbons (C1–C4) and carbon dioxide gas [9,10,11]. Cobalt catalysts have been considered the most active catalysts for the production of high molecular weight hydrocarbons (C5+) and have high resistance to deactivation [12]. Moreover, the characteristics of catalysts can be improved by adding promoters to modify the porosity and texture of the support, which leads to beneficial mechanical properties and attrition resistance. Furthermore, this improves the dispersion and reducibility of metal species [13].
The main prerequisite factors in the design, optimization, and simulation of industrial processes are kinetic modeling and kinetic description [14]. The power law kinetic model for Fischer–Tropsch reactions has been proposed in many studies, including CO, H2, H2O, CO2, CH4, C2H4, C2H6, C3H8, n-C4H10, i-C4H10, and C5+ over different catalysts. Additionally, each reaction’s activation energy is shown in (Table 1).
For Fischer–Tropsch systems, many researchers have studied reaction kinetics and optimization of reaction conditions extensively. Using the power law rate equation or an expression based on certain mechanistic hypotheses, these investigations seek to define the reaction rate. Mosayebi and Haghtalab [18] developed a detailed kinetic model for Fischer–Tropsch synthesis over a Co Ru/γ-Al2O3 catalyst using a fixed-bed reactor at a range of reaction conditions. The Langmuir–Hinshelwood–Hougen–Watts (LHHW) method was used to derive elementary reactions for FTS. Kinetic expressions for the formation of n-paraffin and olefin were developed by combining alkyl and alkenyl mechanisms without considering the readsorption and secondary reactions of olefins. The Levenberg–Marquardt method and a genetic algorithm approach were used to optimize the kinetic model parameters. They showed that the computed values of activation energy have good agreement with values obtained by other studies.
Atashi et al. [19] proposed a kinetic model for the hydrogenation of CO over Fe/Ce catalyst using a fixed-bed microreactor at a range of reaction conditions. The proposed Langmuir–Hinshelwood kinetic mechanism tried about 16 kinetic expressions for the consumption of CO and the interchange among adsorbed molecules of H2 and CO. They found that the CO conversion increased with an increase in temperature, pressure, and higher molecular hydrocarbon due to the temperature increase.
Moazami et al. [14] developed a sequence of collective Fischer–Tropsch (FT) and water–gas shift reaction mechanisms to investigate the kinetics of FT synthesis over a 37% Co/SiO2 catalyst at twelve different operating conditions. They used different dual methodologies to improve the reaction network model for the FT synthesis. They found that the originally developed kinetic model, which was built on the combination alkyl/alkenyl mechanism, aimed at FT reactions (for the production of paraffin and α-olefins).
Hatami et al. [20] prepared and characterized a cobalt catalyst held on carbon nanotubes. The catalyst was tested at a wide range of operating conditions for liquid hydrocarbon products by the Fischer–Tropsch reaction in a fixed-bed reactor. Eleven kinetic expressions for CO consumption were developed based on several mechanisms. The nonlinear regression method using the Levenberg–Marquardt method was used to estimate kinetic parameters and to obtain the best optimization. They showed that the best models for suggested elementary reactions included the formation of surface species as rate-determining steps instead of syngas adsorption.
Haghtalab et al. [21] utilized a Langmuir–Hinshelwood–Hougen–Watson (LHHW) alkyl mechanism for kinetic modeling of FTS. A fixed-bed reactor (FBR) was used for the FTS reactions over a Co–Ru/CNTs catalyst with a nanostructure. The parameters of the kinetic model were estimated under different operating conditions such as pressure, temperature, gas hourly space velocity, and the H2/CO ratio. They noticed that the outcomes of the experimental results are consistent with the developed kinetic model.
Pandey et al. [1] developed a kinetic model for the Fischer–Tropsch reaction to describe reactant consumption and product distribution using a 20Co-0.5Re/γ-Al2O3 catalyst. The model was optimized against an experimental database over a variety of operating conditions, with and without water present in the feed. They displayed an average absolute relative variation of almost 23.3%. In addition, the model successfully predicted the effects of water in the feed on the kinetics rate.
Although there have been many studies dealing with the kinetic modeling of the Fischer–Tropsch reaction, little information is presented in the literature focused on maximizing liquid biofuel products for the FT process. Accordingly, the main topic of the present study is to modify the kinetic model to describe FT accurately and then optimize the operating conditions of the process (i.e., pressure, temperature, H2/CO feed ratio, and space velocity) to increase the selectivity of desired liquid biofuel products (C5+).

2. Experimental Section

2.1. Catalyst Preparation

The Ce-Co/SiO2 catalyst was prepared using the incipient wetness impregnation (IWI) method, which involves similar steps to those described by [22]. First, a suitable quantity of support (Silica Engineering, Maidstone, UK) was calcined under airflow at 500 °C for 4 h to achieve morphological stability during subsequent applications. Then, an aqueous solution was derived from a mixture of Co (NO3)2·6H2O (97% Merck purity) and Ce (NO3)3·6H2O (98.5% Merck purity), and the target loading of 1 wt. % Ce and 10 wt. % Co was achieved by adjusting the solution concentration. Impregnation was carried out at 80 °C for 4 h with continuous stirring until 10 wt. % Co and 1 wt. % Ce were loaded onto the SiO2 support. The catalyst was then dried at 110 °C for 12 h and calcined at 600 °C for 6 h.

2.2. The Characterization Techniques

The BET surface area, pore radius, and pore volume of the prepared catalysts and support (SiO2) were obtained by N2 physisorption using a Micromeritics ASAP 2020 plus (USA), which is a high-performance adsorption analyzer. Before testing, the temperature was gradually raised to approximately 200 °C, at which point the samples were evacuated to 0.07 bar and held for ten hours. Afterward, the properties of the samples were recorded.
An analysis was performed on a sample of calcined catalyst using a TCD-equipped TP-5000 analytical machine (Tianjin Xianquan Co., Ltd., Tianjin City, China) to conduct temperature-programmed reduction (TPR). To eliminate any water, 0.1 g of each sample was first purged at 400 °C in an argon flow, and was then cooled to room temperature. The samples were then subjected to TPR testing, where the temperature was raised by 10 °C/min in steps from 25 °C to 900 °C while the argon flow rate was 30 mL/min. The catalysts and their precursors’ morphology were examined using a Scanning Electron Microscope (SEM) equipped with a Silicon Drift Detector (SDD)—X-Max from Oxford Instruments Group (Oxford Instruments plc, Abingdon, UK).

2.3. FTS Reaction

A fixed-bed (F-B) reactor (ID: 10 mm, L: 900 mm) filled with two grams of catalyst (Figure 1a,b) was used to evaluate the catalyst for long-chain hydrocarbon production over a range of different parameters, including 200–240 °C, 5–25 bar, 1000–5000 mL/gcat·h, and a 0.5–4 H2/CO feed ratio. In the first step, the catalyst was heated with a continuous stream of argon and then subjected to in situ reduction between 200 and 600 °C at a rate of 5 °C/min under a flow rate of 30 m/min for twelve hours using 5% H2/Ar.
The system was then gradually cooled to ambient temperature. After reduction, argon gas was used to heat and pressurize the system to the desired reaction conditions. A Brooks 5850-mass flow controller was used to introduce the reformed biogas into the reactor at the desired space velocity and H2/CO ratio. Every time the reaction conditions changed, the catalyst bed was thermally treated with argon at a flow rate of 41.6 mL/min for three hours at 400 °C to eliminate any carbon atoms that had accumulated there and may have reduced the catalyst activity. Two types of gas chromatography were used to analyze the FTR composition products: an online gas chromatograph (Shimadzu-2014, Shimadzu Corp. Ltd., Tokyo, Japan) equipped with FID and TCD for the analysis of the gas phase and an offline gas chromatograph (Varian CP 3800, SpectraLab Scientific Inc., Markham, ON, Canada) equipped with FID for the analysis of the liquid phase.
The reactant conversions (i.e., CO) and product selectivity were calculated using Equations (1)–(4).
CO   Conversion   % =   CO mole   input   mole   CO mole   output   CO mole   input × 100
Low Chain   Hydrocarbon   Selectivity =   S i   C 1 C 4 % = mole i   CO mole   input   CO mole   output × 100
Carbon   Dioxide   Selectivity   =   S   CO 2 % = Mole CO 2   CO mole   input   CO mole   output × 100
Total   Liquid   Selectivity   % = S C 5 + % = 100 S i   C 1 C 4   S   CO 2  
where:
i = C H 4 ,   C 2   H 4 , C 2   H 6 , C 3   H 6 , n C 4   H 10 , i C 4   H 10

3. Mathematical Modelling

3.1. Model Assumptions

The mathematical model was simplified using a number of simplification assumptions. The following are these presumptions:
One-dimensional steady state model [14,23];
Reactor’s isothermal plug flow [24];
Reactions in the homogeneous gas phase [25,26,27];
Power law kinetic model [28,29].

3.2. Kinetic Model

The difference in the amounts of component i with respect to catalyst weight in the reactor bed is described by the following ordinary differential equation [30,31]:
d F i d w = j m s i , j r j
The power law reaction kinetic model shown in Table 1 was taken into consideration in this study, and the reaction rate was estimated using Equations (6) and (7):
r j = k j P C O n j P H 2 m j   For   reactions   1   to   7
r j = k j P C O n j P H 2 O m j   For   reaction   8
Eleven ordinary differential equations make up the reactor model, and Equation (5) was generalized for every component in the reaction mixture. The set of differential Equations (8)–(18) expresses the change in the molar flow rates with respect to catalyst weight within the fixed-bed reactor for CO, H2, CH4, C2H4, C2H6, C3H8, n-C4H10, i-C4H10, C5+, CO2, and H2O, components, respectively [32].
d F C O d W = r 1 2 r 2 2 r 3 3 r 4 4 r 5 4 r 6 6.05 r 7 r 8
d F H 2 d W = 3 r 1 4 r 2 5 r 3 7 r 4 9 r 5 9 r 6 12.235 r 7 + r 8
d F C H 4 d W = + r 1
d F C 2 H 4 d W = + r 2
d F C 2 H 6 d W = + r 3
d F C 3 H 8 d W = + r 4
d F i C 4 H 10 d W = + r 5
d F n C 4 H 10 d W = + r 6
d F C 5 + d W = + r 7
d F C O 2 d W = r 8
d F H 2 O d W = + r 1 + 2 r 2 + 2 r 3 + 3 r 4 + 4 r 5 + 4 r 6 + 6.05 r 7 r 8
The partial pressure of the ith component can be calculated according to Dalton’s law:
P i = y i × P t = F i i N F i × P t
The fluctuation of the reaction rate constants with temperature was represented using the Arrhenius equation (Equation (20)).
k j = k j o e x p E A j R T

3.3. Assessment of Kinetic Parameters

To reduce the objective function (f) in Equation (21), which measures the mean relative deviation between the experimental and projected gas compositions of the reactor effluent, the kinetic parameters were calculated using an integration-optimization technique [33,34,35].
f = 1 n e x p i = 1 n e x p 1 n C o m j = 1 n C o m y i , j e x p y i , j p r e d y i , j e x p × 100
The eleven ordinary differential equations were numerically integrated using the 4th. The Runge–Kutta integration method for each set of kinetic parameters, and a global optimal set of kinetic parameters was estimated using a genetic algorithm. The “ga” and “ode45” Matlab subroutines were used for optimization and integration, respectively. The major steps of the genetic algorithm are depicted in a flowchart in (Figure 2).

4. Results and Discussion

This part is presented in two sections. In the first Section 4.1, the results of the experimental studies are presented, while in Section 4.2, the focus is on the results of the parametric optimization and kinetic studies for converting reformed biogas to liquid fuel.

4.1. Results of Experimental Studies

4.1.1. Catalyst Characterization

The textural characteristics of the prepared catalysts and support, which include surface area, pore radius, and pore volume, are presented in Table 2. It can be observed that the loading of metals (Ce and Co) on the support (SiO2) reduces the surface area by around 9.2% and 11.5% for 10Co/SiO2 and 1Ce-10Co/SiO2, respectively. The reduction in surface area of the support, due to metals partially blocking its pores, implies that the metals were effectively incorporated into its porous structure [36,37].
In order to predict the reducibility behavior of the prepared catalysts, H2-TPR analysis was performed for the 10Co/SiO2 and 1Ce-10Co/SiO2 catalysts under the same conditions as shown in (Figure 3). The results showed a pair of reduction peaks for the 10Co/SiO2 catalyst at temperatures of 335 °C and 535 °C, corresponding to the reduction of Co3O4 to CoO and CoO to Co metal, respectively. Moreover, the addition of Ce metal as a promoter reduced the reduction peaks by about 25 °C and 18 °C, respectively. These results demonstrate that Ce metal can reduce Co3+ to Co2+ ions at a lower temperature than the cobalt oxides in the 10Co/SiO2 catalyst [38]. Similar patterns of behavior were observed by [39,40].
Accordingly, the H2-TPR results show that the chosen conditions (5 vol. % H2, flow rate of 30 m/min, and a duration of 12 h) are suitable for reducing Co3O4 to Co metal prior to the beginning of FTR.
A comprehensive scanning electron microscope (SEM) study was conducted to examine the morphological changes of the catalysts and their precursors at different stages: before and after the test. The SEM images captured during the analysis are presented in Figure 4. Notably, distinct differences were observed in the morphology of the catalyst precursor (Figure 4a) and the calcined catalyst (Figure 4b). The electron micrograph of the catalyst precursor displayed numerous agglomerates of particles, indicating a less dense structure. However, after calcination at 600 °C, the morphology significantly transformed, with reduced agglomeration evident in the calcined catalyst (Figure 4b). This transformation was attributed to the coverage of the calcined catalyst surface with small crystallites of cobalt and cerium oxide [41]. Moreover, following the Fischer–Tropsch synthesis (FTS) chemical reaction, the catalyst underwent further changes in texture and morphology (Figure 4c). Notably, the grain size of the tested catalyst increased, primarily due to agglomeration resulting from sintering after the reactions. These SEM observations provided valuable insights into the structural modifications and behavior of the catalysts during the various stages of the study.

4.1.2. Effects of Variables on FTS

The effects of different reaction conditions, such as operating temperature (200–240 °C), total pressure (5–25 bar), space velocity (1000–5000 mL/gcat·h), and H2/CO (0.5–4) on the experimental results for CO conversion and product selectivity are presented in (Figure 5, Figure 6, Figure 7 and Figure 8) and summarized in (Table 3).
As shown in (Figure 5), the CO conversion and low-chain hydrocarbons selectivity (CH4, C2–C4) and CO2 increased with increases in reaction temperature from 200 to 240 °C. In contrast, the selectivity of total liquid biofuel products (C5+) decreased by about 21.4%. The decrease in total liquid products is attributed to the fact that the Fischer–Tropsch reaction is an exothermic polymerization reaction. Furthermore, at higher reaction temperatures, the rate of hydrogenation of “CH2” units accelerated, shifting the product selectivity towards lower-chain hydrocarbons [42,43]. Given these important findings, a lower reaction temperature of 200 °C is recommended for obtaining the best selectivity of liquid biofuel products (C5+).
Based on thermodynamic principles, increasing the total pressure of the reaction unit shifts the equilibrium reactions towards the products, leading to an increase in reactant conversion. Therefore, the effect of total pressure on the activity of the catalyst and the product selectivity was investigated and presented in Figure 6. The results indicate that the conversion of CO and the formation of desired products (C5+) increased by about 58.9% and 2.1%, respectively, when the total pressure was increased from 5 to 25 bar. On the other hand, the selectivity of lower chain hydrocarbons and CO2 decreased, which is consistent with an increase in chain growth probability with increasing pressure [44]. Using the outcomes of these trials, it can be said that high pressure (25 bar) is the optimal choice for producing liquid biofuel products (C5+) in the range of reaction pressures tested (5–25 bar), with a reaction temperature of 200 °C, a space velocity of 1000 mL/gcat·h, and an inlet feed ratio of 0.5 (H2/CO).
To understand the effect of residence time on reactant conversion and product selectivity, a series of experiments were conducted at different space velocities, as shown in (Figure 7). It can be observed that increasing the space velocity from 1000 to 5000 mL/gcat·h, resulted in decreases in the formation of desired hydrocarbons (C5+) and the conversion of CO by around 6% and 43.8%, respectively. Conversely, the formation of lower chain hydrocarbons (CH4 and C2–C4) increased. The low production of CO2 is in agreement with the lower WGS activity typically observed with cobalt catalysts [45]. Increasing the space velocity reduces the residence time of the reaction, which in turn reduces chain growth and reactant conversion, leading to a hasty increase in the production of low-chain hydrocarbons and a decrease in the production of total liquid products (C5+) [46]. According to the experimental results, a lower space velocity of 1000 mL/gcat·h is the optimal choice for producing favorable liquid biofuel (C5+) in the range of space velocities tested (1000–5000 mL/gcat·h), with a reaction temperature of 200 °C, a total pressure of 25 bar, and an inlet feed ratio of 0.5 (H2/CO).
Figure 8 presents the effect of different inlet feed ratios on catalyst activity and product selectivity at 200 °C, 25 bar, and 1000 mL/gcat·h The results show that increasing the proportion of inlet feed from 0.5 to 2 reduces the formation of low-chain hydrocarbons and increases selectivity for the desired product (C5+) and CO conversion by about 3% and 31.3%, respectively. However, increasing the proportion of inlet feed from two to four results in a decrease in reactant conversion and desired product selectivity (C5+) of about 18.2% and 2.5%, respectively. The change in reactant concentration within the proportion of inlet feed has a considerable effect on its partial pressure. At a proportion of feed less than two, the H2 partial pressure will decrease, causing some CO molecules to escape without being hydrogenated. At a proportion greater than two, the H2 partial pressure will increase, leading to a decrease in catalyst activity [47,48]. Therefore, the best option for producing the desired products is an inlet feed ratio of two.

4.2. Results of Parametric Optimization and Kinetic Studies

4.2.1. Model Validation

The procedure for estimating kinetic parameters depends on minimizing the mean absolute relative error (MARE) obtained between the experimental and predicted results. The genetic optimization method initially generates a number of chromosomes, with each chromosome representing a different set of kinetic parameters. Each chromosome is evaluated by applying its kinetic parameters in a mathematical model to simulate the response corresponding to sixteen experiments and predict the MARE for each one of these chromosomes. The optimization process continues until the MARE reaches its minimal value, and this chromosome represents the optimum kinetic parameters. The parity plot for the comparison of the experimental data from 16 experiments with the expected findings for CO, CH4, C2H4, C2H6, C3H8, n-C4H10, i-C4H10, C5+, and CO2 is shown in Figure 9. The simulation results at the optimum kinetic parameters are also shown in Table 4. Bench assesses the accuracy of the fitted model by utilizing a quantitative analysis approach and the MARE (11.7%) to compare the acquired results by the fitted model with those estimated by experiments (Table 3), with respect to CO conversion and product selectivity. The good accuracy of the fitted model is in comparison with the values reported by [1] (23.1%); [49] (20%); and [50] (48.4%). The minimum error is recognized for C5+ (MARE = 2.5%), while the maximum error obtained is for remaining CO (MARE = 27.7%). These confirm that the modified model accurately estimates the production rate of liquid biofuels (C5+). The MARE for CO is relatively higher than that for other components because the conversion of CO is high, and the remaining quantities of CO in the reactor effluent are too low. Therefore, higher levels of relative errors occur when comparing the experimental CO concentrations with the predicted CO concentrations in the reactor effluent. The modeling results agree very well with the experimental data, especially in predicting C5+ selectivity.
The optimized values of the kinetic model parameters are given in (Table 5). A higher activation energy (80.214 kJ/mol.) was found for methane formation, while a lower activation energy was found for C5+ formation. By comparing the activation energies obtained from this study with those obtained from previous studies, it can be observed that a similar range of activation energies (12.86–80.21 kJ/mol.) was obtained in the present study compared with that obtained in previous kinetic works [14,15,16], which was in the range of 10–101.15 kJ/mol. The estimated activation energies for C5+ formation (12.86) are also within those reported in the literature [14,15,17], which are in the range of 12.44–23.56. The estimated values of the kinetic parameters are shown in Table 5.

4.2.2. Process Optimization

The goal of the optimization study was to increase the selectivity of liquid biofuel, in particular the selectivity of higher-molecular-weight hydrocarbons like C5+, as well as to increase synthesis gas conversions (particularly CO conversion) and decrease the forming of unwanted products like carbon dioxide and methane. For each of these variables, Table 6 includes a set of randomly chosen values together with the best result from the optimization process. In particular, C5+ hydrocarbon productivity was increased by the optimization of the working conditions. In light of this, the objective function equation is as follows:
Obj = maximize S C 5 + / Undesired =   F C 5 + F Co 2 + F CH 4 + F C 2 H 4 + F C 2 H 6 + F C 3 H 6 + F n C 4 H 8 + F i C 4 H 8
Table 6 lists the outcomes of genetic optimization, and (Figure 10) shows how the objective function changes as the number of iterations increases (generation number). After the highest possible generation count (1000) is achieved, the optimizer ends. At these operating conditions: 200 °C temperature, 6.29 bar reactor total pressure, 1529.58 mL/gcat·h space velocity, and a H2/CO feed ratio of 3.96, the maximum C5+ selectivity was achieved at 91.66.
We should monitor here that liquid biofuel generated by integration of the Fisher–Tropsch process with anaerobic digestion reduces greenhouse gas emissions in comparison with conventional fossil fuels and contributes to meeting renewable transportation fuel standards [51]. Additionally, capital, operating costs, and the lower trade price of the system are directly proportional to the feed capacity of the plant; if the feed capacity is increased, the selling price is decreased. The lowermost trade price of FT liquid fuels derived from biogas for a plant capacity of 2000 m3/h was around $5.29/GGE, but at a larger plant size, the trade price was $2.37/GGE and $1.92/GGE for plant capacities of 10,000 and 20,000 m3/h, respectively. In other words, the economy of this technology is based mainly on operating costs, biogas costs, and the conversion rate of FTS [52]. Through investigating high- and low-temperature operating modes, [8], revealed that tri-reforming for the invention of FT liquid fuels from biogas, taking into account the biogas as rubbish with zero cost and from a plant supplied with waste, enhanced syngas production. The cost of the generation for a plant that receives 12 Mm3 of biogas/year is $3/gallon for HTFT and $1.7/gallon for LTFT. Naqi et al. [53], reported the liquid fuel derived from integration of the AD and FTR requires a capital expenditure of around $210/ton of biomass and operational expenses of around $110/ton of biomass for a plant capacity of 90,000 tons/year. In another, a number of studies were carried out by [54,55,56] to evaluate the techno-economic assessment for converting various kinds of biomass to liquid fuel, they established that the product cost alternated from $2.86 to $4.61 per GGE.
In order to evaluate the environmental effect of liquid fuel generated from biogas using FTS, [51] accomplished a life cycle analysis to create liquid fuel (diesel) from biogas. Crop cultivation as feedstocks for the anaerobic digestion system, harvest, and transport to the bio-refinery, biomass to Fischer–Tropsch liquid (BTL), and combustion of liquid fuel for a passenger car were all considered in the analysis. By comparison with classic fossil-derived diesel, they confirmed how biogas-derived liquid fuel might decrease greenhouse gas emissions by about 73%. Generally, in comparison to liquid fuel conversion practices from biomass to their traditional fossil fuel equivalent, they display lesser climate variation capacity and GHG emissions. The normal reduction in GHG emissions from biofuels prepared from first-generation feedstock is between 50 and 75% [57,58]. However, compared with first-generation biofuels, second-generation biofuels have a higher possibility of cutting GHG emissions by 50–100% [59]. GHG emissions may be decreased by approximately 76% using third-generation feedstock [60].
The work of the dynamic metals employed in FT catalysts, and the comparison of anticipated lifetimes, is another critical highlight of FTS. Practically, 25–30% of the CO consumed in FTR was used by press catalysts, which are motivated by the generation of CO2 through the water–gas shift reaction. Also, the water vapor delivered amid the union prevents CO from being changed over to hydrocarbons, which brings down the reaction rate within the press catalyst’s lower layers [61]. These problems are not shown with cobalt catalysts. Water has a small impact on the reaction, and the selectivity of CO2 generation on it often does not surpass a number of percent [62,63]. Moreover, the presence of water vapor makes steps in the selectivity of cobalt catalysts for the desired higher hydrocarbons simpler; advancement of Cobalt catalysts with respectable metals, such as Zr and Ce, essentially expanded the activity and lifetime of the catalyst [64]. According to the above, the high selectivity for C5+ was about 91.66% when the prepared cobalt catalysts promoted with Ce were used in the present study.

5. Conclusions

In the current investigation, catalytic tests for Fischer–Tropsch synthesis were carried out in a fixed-bed reactor over a Ce-Co/SiO2 catalyst. The impregnation process was used to create the catalyst. At various operating temperatures, pressures, gas hourly space velocities (GHSV), and syngas ratios (H2/CO), sixteen experiments were conducted. The proposed kinetic model was logically investigated using statistical and correlation tests. As a result, the results show that the estimated activation energies’ values were consistent with those provided by earlier investigations. The Matlab 2020a programming language was used to code all of the computations in this investigation. An isothermal plug flow reactor model was used to simulate the lab reactor, and a MATLAB ordinary differential equation solver was used to resolve a set of species mass balance equations. A genetic algorithm was utilized to estimate the ideal set of kinetic parameters. The statistical test found the estimated parameters and thorough kinetic model to be statistically significant. The established model predicts a fraction of the reactants and products over the length of the reactor bed.

Author Contributions

Validation, I.H. Writing—original draft, F.K.A.-Z.; Writing—review & editing, F.K.A.-Z., Z.M.S. and I.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

All data used in the present study appear in the submitted article.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

CiConcentration of component imol.cm−3
EaActivation energykJ/mol
FiMolar flow rate of component imole/s
kjRate constant of reaction jmol/gcat·h·bar
kj0Pre-exponential factormol/gcat·h·bar
MmAverage molecular weightkg/mole
nj, mjOrders of reaction rate j-
PtTotal pressurebar
PiPartial pressure of component ibar
RUniversal gas constant8.3145 J/mole. K
rjRate of reaction jmole/gm cat·h
SiSelectivity for component i mole%
si,jStoichiometric coefficient for the ith component participating in the jth reaction
TTemperatureK
wAmount of catalyst in the reactorGm
ximole fraction of component i in the liquid phase-
yimole fraction of component i in the gas phase-
Abbreviations
FTSFischer–Tropsch Synthesis
GAGenetic Algorithm
LHHWLangmuir–Hinshelwood–Hougen–Watts
MAREMean absolute relative error
catCatalyst
comComponent
expExperimental value
HHydrogen

References

  1. Pandey, U.; Runningen, A.; Gavrilović, L.; Jørgensen, E.A.; Putta, K.R.; Rout, K.R.; Rytter, E.; Blekkan, E.A.; Hillestad, M. Modeling Fischer–Tropsch kinetics and product distribution over a cobalt catalyst. AIChE J. 2021, 67, e17234. [Google Scholar] [CrossRef]
  2. Al-Zuhairi, F.K.; Azeez, R.A.; Mahdi, S.A. Synthesis of long-chain hydrocarbons from Syngas over promoted Co/SiO2 catalysts using Fischer–Tropsch reaction. AIP Conf. Proc. 2022, 2443, 030028. [Google Scholar]
  3. Steynberg, A. Introduction to fischer-tropsch technology. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 2004; Volume 152, pp. 1–63. [Google Scholar]
  4. Van Santen, R.; Markvoort, A.; Filot, I.; Ghouri, M.; Hensen, E. Mechanism and microkinetics of the Fischer–Tropsch reaction. Phys. Chem. Chem. Phys. 2013, 15, 17038–17063. [Google Scholar] [CrossRef] [PubMed]
  5. Al-Zuhairi, F.; Micoli, L.; Florio, C.; Ausiello, A.; Turco, M.; Pirozzi, D.; Toscano, G. Anaerobic co-digestion of municipal solid wastes with giant reed under mesophilic conditions. J. Mater. Cycles Waste Manag. 2019, 21, 1332–1340. [Google Scholar] [CrossRef]
  6. Sukkar, K.A.; Al-Zuhairi, F.K.; Dawood, E.A. Evaluating the Influence of Temperature and Flow Rate on Biogas Production from Wood Waste via a Packed-Bed Bioreactor. Arab. J. Sci. Eng. 2021, 46, 6167–6175. [Google Scholar] [CrossRef]
  7. León, E.; Martín, M. Optimal production of power in a combined cycle from manure based biogas. Energy Convers. Manag. 2016, 114, 89–99. [Google Scholar] [CrossRef]
  8. Hernandez, B.; Martin, M. Optimization for biogas to chemicals via tri-reforming. Analysis of Fischer-Tropsch fuels from biogas. Energy Convers. Manag. 2018, 174, 998–1013. [Google Scholar] [CrossRef]
  9. Xie, J.; Paalanen, P.P.; van Deelen, T.W.; Weckhuysen, B.M.; Louwerse, M.J.; de Jong, K.P. Promoted cobalt metal catalysts suitable for the production of lower olefins from natural gas. Nat. Commun. 2019, 10, 167. [Google Scholar] [CrossRef]
  10. Zhang, Q.; Cheng, K.; Kang, J.; Deng, W.; Wang, Y. Fischer–Tropsch catalysts for the production of hydrocarbon fuels with high selectivity. ChemSusChem 2014, 7, 1251–1264. [Google Scholar] [CrossRef]
  11. Moazami, N.; Mahmoudi, H.; Panahifar, P.; Rahbar, K.; Tsolakis, A.; Wyszynski, M.L. Mathematical modeling and performance study of Fischer-Tropsch synthesis of liquid fuel over cobalt-silica. Energy Procedia 2015, 75, 62–71. [Google Scholar] [CrossRef]
  12. Zhu, C.; Bollas, G.M. Gasoline selective Fischer-Tropsch synthesis in structured bifunctional catalysts. Appl. Catal. B Environ. 2018, 235, 92–102. [Google Scholar] [CrossRef]
  13. Pardo-Tarifa, F.; Cabrera, S.; Sanchez-Dominguez, M.; Boutonnet, M. Ce-promoted Co/Al2O3 catalysts for Fischer–Tropsch synthesis. Int. J. Hydrog. Energy 2017, 42, 9754–9765. [Google Scholar] [CrossRef]
  14. Moazami, N.; Wyszynski, M.L.; Rahbar, K.; Tsolakis, A.; Mahmoudi, H. A comprehensive study of kinetics mechanism of Fischer-Tropsch synthesis over cobalt-based catalyst. Chem. Eng. Sci. 2017, 171, 32–60. [Google Scholar] [CrossRef]
  15. Egiebor, N.; Ungar, K.; Wojciechowski, B. A classification of fischer–tropsch synthesis product distributions. Can. J. Chem. Eng. 1984, 62, 425–430. [Google Scholar] [CrossRef]
  16. Irani, M. Modeling of Fischer-Tropsch Synthesis Packed Bed Reactor for Producing Liquid Fuels from Natural Gas. Pet. Coal 2013, 55, 82–92. [Google Scholar]
  17. Moazami, N.; Mahmoudi, H.; Rahbar, K.; Panahifar, P.; Tsolakis, A.; Wyszynski, M.L. Catalytic performance of cobalt–silica catalyst for Fischer–Tropsch synthesis: Effects of reaction rates on efficiency of liquid synthesis. Chem. Eng. Sci. 2015, 134, 374–384. [Google Scholar] [CrossRef]
  18. Mosayebi, A.; Haghtalab, A. The comprehensive kinetic modeling of the Fischer–Tropsch synthesis over Co@ Ru/γ-Al2O3 core–shell structure catalyst. Chem. Eng. J. 2015, 259, 191–204. [Google Scholar] [CrossRef]
  19. Atashi, H.; Gholizadeh, J.; Tabrizi, F.F.; Tayebi, J. Modeling selectivity of ethylene and propylene in the Fischer-Tropsch synthesis with artificial neural network and response surface methodology. ChemistrySelect 2016, 1, 3271–3275. [Google Scholar] [CrossRef]
  20. Hatami, B.; Tavasoli, A.; Asghari, A.; Zamani, Y.; Zamaniyan, A. Kinetics Modeling of Fischer–Tropsch Synthesis on the Cobalt Catalyst Supported on Functionalized Carbon Nanotubes. Kinet. Catal. 2018, 59, 701–709. [Google Scholar] [CrossRef]
  21. Haghtalab, A.; Shariati, J.; Mosayebi, A. Experimental and kinetic modeling of Fischer–Tropsch synthesis over nano structure catalyst of Co–Ru/carbon nanotube. React. Kinet. Mech. Catal. 2019, 126, 1003–1026. [Google Scholar] [CrossRef]
  22. Zohdi-Fasaei, H.; Atashi, H.; Farshchi Tabrizi, F.; Mirzaei, A.A. Modeling and optimization of Fischer-Tropsch synthesis over Co-Mn-Ce/SiO2 catalyst using hybrid RSM/LHHW approaches. Energy 2017, 128, 496–508. [Google Scholar] [CrossRef]
  23. Barkalov, K.; Gubaydullin, I.; Kozinov, E.; Lebedev, I.; Faskhutdinova, R.; Faskhutdinov, A.; Enikeeva, L. On Solving the Problem of Finding Kinetic Parameters of Catalytic Isomerization of the Pentane-Hexane Fraction Using a Parallel Global Search Algorithm. Mathematics 2022, 10, 3665. [Google Scholar] [CrossRef]
  24. Nestler, F.; Müller, V.; Ouda, M.; Hadrich, M.; Schaadt, A.; Bajohr, S.; Kolb, T. A novel approach for kinetic measurements in exothermic fixed bed reactors: Advancements in non-isothermal bed conditions demonstrated for methanol synthesis. React. Chem. Eng. 2021, 6, 1092–1107. [Google Scholar] [CrossRef]
  25. Parsafard, N.; Garmroodi, A.; Mirzaei, S. Gas-phase catalytic isomerization of n-heptane using Pt/(CrOx/ZrO2)-HMS catalysts: A kinetic modeling. Int. J. Chem. Kinet. 2021, 53, 971–981. [Google Scholar] [CrossRef]
  26. Shakor, Z.M.; AbdulRazak, A.A.; Sukkar, K.A. A detailed reaction kinetic model of heavy naphtha reforming. Arab. J. Sci. Eng. 2020, 45, 7361–7370. [Google Scholar] [CrossRef]
  27. Shakor, Z.M.; Ramos, M.J.; AbdulRazak, A.A. A detailed reaction kinetic model of light naphtha isomerization on Pt/zeolite catalyst. J. King Saud. Univ. Eng. Sci. 2022, 34, 303–308. [Google Scholar] [CrossRef]
  28. Regali, F.; Boutonnet, M.; Järås, S. Hydrocracking of n-hexadecane on noble metal/silica–alumina catalysts. Catal. Today 2013, 214, 12–18. [Google Scholar] [CrossRef]
  29. Ancheyta-Juárez, J.; Villafuerte-Macías, E. Kinetic modeling of naphtha catalytic reforming reactions. Energy Fuels 2000, 14, 1032–1037. [Google Scholar] [CrossRef]
  30. Al-Shathr, A.; Shakor, Z.M.; Majdi, H.S.; AbdulRazak, A.A.; Albayati, T.M. Comparison between artificial neural network and rigorous mathematical model in simulation of industrial heavy naphtha reforming process. Catalysts 2021, 11, 1034. [Google Scholar] [CrossRef]
  31. Aammer, H.; Shakor, Z.M.; Al-Sheikh, F.; Al-Naimi, S.A.; Anderson, W.A. Simulation and Optimization of the Ethane Cracking Furnace Using ASPEN PLUS and MATLAB: A Case Study from Petrochemical Complexes. Combust. Sci. Technol. 2022, 195, 2634–2654. [Google Scholar] [CrossRef]
  32. Hájek, J.; Murzin, D.Y. Liquid-Phase Hydrogenation of Cinnamaldehyde over a Ru−Sn Sol−Gel Catalyst. 1. Evaluation of Mass Transfer via a Combined Experimental/Theoretical Approach. Ind. Eng. Chem. Res. 2004, 43, 2030–2038. [Google Scholar] [CrossRef]
  33. Khalaf, Y.H.; Al-Zaidi, B.Y.S.; Shakour, Z.M. Experimental and Kinetic Study of the Effect of using Zr-and Pt-loaded Metals on Y-zeolite-based Catalyst to Improve the Products of n-heptane Hydroisomerization Reactions. Orbital Electron. J. Chem. 2022, 14, 153–167. [Google Scholar] [CrossRef]
  34. Al-Shathr, A.; Shakor, Z.M.; Al-Zaidi, B.Y.; Majdi, H.S.; AbdulRazak, A.A.; Aal-Kaeb, S.; Shohib, A.A.; McGregor, J. Reaction Kinetics of Cinnamaldehyde Hydrogenation over Pt/SiO2: Comparison between Bulk and Intraparticle Diffusion Models. Int. J. Chem. Eng. 2022, 2022, 8303874. [Google Scholar] [CrossRef]
  35. Al-Karim, A.A.; Shakor, Z.M.; Al-Sheikh, F.; Anderson, W.A. Experimental and kinetic study of vacuum residue cracking over zirconium based catalysts. React. Kinet. Mech. Catal. 2022, 135, 847–865. [Google Scholar] [CrossRef]
  36. Khalaf, A.L.; Al-Zuhairi, F.K.; Kadhim, W.A.; Ab Rahim, M.H. The effects of cerium promoter on the performance of cobalt-based catalysts in fischer tropsch synthesis for liquid fuel production. IIUM Eng. J. 2020, 21, 1–11. [Google Scholar] [CrossRef]
  37. Gnanamani, M.K.; Jacobs, G.; Graham, U.M.; Pendyala, V.R.R.; Martinelli, M.; MacLennan, A.; Hu, Y.; Davis, B.H. Effect of sequence of P and Co addition over silica for Fischer-Tropsch synthesis. Appl. Catal. A Gen. 2017, 538, 190–198. [Google Scholar] [CrossRef]
  38. Xue, L.; Zhang, C.; He, H.; Teraoka, Y. Catalytic decomposition of N2O over CeO2 promoted Co3O4 spinel catalyst. Appl. Catal. B Environ. 2007, 75, 167–174. [Google Scholar] [CrossRef]
  39. Xiaoping, D.; Changchun, Y.; Ranjia, L.; Haibo, S.; Shikong, S. Role of CeO2 promoter in Co/SiO2 catalyst for Fischer-Tropsch synthesis. Chin. J. Catal. 2006, 27, 904–910. [Google Scholar]
  40. Eliseev, O.; Tsapkina, M.; Davydov, P.; Dzhukasheva, L.; Lapidus, A. Synthesis of hydrocarbons from carbon monoxide and hydrogen in the presence of cobalt catalysts promoted with cerium oxide. Solid Fuel Chem. 2017, 51, 166–169. [Google Scholar] [CrossRef]
  41. Mansouri, M.; Atashi, H.; Mirzaei, A.A.; Jangi, R. Kinetics of the Fischer-Tropsch synthesis on silica-supported cobalt-cerium catalyst. Int. J. Ind. Chem. 2013, 4, 1. [Google Scholar] [CrossRef]
  42. Mansouri, M.; Atashi, H.; Tabrizi, F.F.; Mansouri, G.; Setareshenas, N. Fischer–Tropsch synthesis on cobalt–manganese nanocatalyst: Studies on rate equations and operation conditions. Int. J. Ind. Chem. 2014, 5, 14. [Google Scholar] [CrossRef]
  43. Dry, M.E. Practical and theoretical aspects of the catalytic Fischer-Tropsch process. Appl. Catal. A Gen. 1996, 138, 319–344. [Google Scholar] [CrossRef]
  44. Zolfaghari, Z.; Tavasoli, A.; Tabyar, S.; Pour, A.N. Enhancement of bimetallic Fe-Mn/CNTs nano catalyst activity and product selectivity using microemulsion technique. J. Energy Chem. 2014, 23, 57–65. [Google Scholar] [CrossRef]
  45. Pendyala, V.R.R.; Jacobs, G.; Bertaux, C.; Khalid, S.; Davis, B.H. Fischer–Tropsch synthesis: Effect of ammonia on supported cobalt catalysts. J. Catal. 2016, 337, 80–90. [Google Scholar] [CrossRef]
  46. Zhang, L.; Chu, H.; Qu, H.; Zhang, Q.; Xu, H.; Cao, J.; Tang, Z.; Xuan, J. An investigation of efficient microstructured reactor with monolith Co/anodic γ-Al2O3/Al catalyst in Fischer-Tropsch synthesis. Int. J. Hydrog. Energy 2018, 43, 3077–3086. [Google Scholar] [CrossRef]
  47. Akbarzadeh, O.; Mohd Zabidi, N.A.; Wang, G.; Kordijazi, A.; Sadabadi, H.; Moosavi, S.; Amani Babadi, A.; Hamizi, N.A.; Abdul Wahab, Y.; Ab Rahman, M. Effect of Pressure, H2/CO Ratio and Reduction Conditions on Co–Mn/CNT Bimetallic Catalyst Performance in Fischer–Tropsch Reaction. Symmetry 2020, 12, 698. [Google Scholar] [CrossRef]
  48. Tristantini, D.; Lögdberg, S.; Gevert, B.; Borg, Ø.; Holmen, A. The effect of synthesis gas composition on the Fischer–Tropsch synthesis over Co/γ-Al2O3 and Co–Re/γ-Al2O3 catalysts. Fuel Process. Technol. 2007, 88, 643–649. [Google Scholar] [CrossRef]
  49. Pant, K.K.; Upadhyayula, S. Detailed kinetics of Fischer Tropsch synthesis over Fe-Co bimetallic catalyst considering chain length dependent olefin desorption. Fuel 2019, 236, 1263–1272. [Google Scholar]
  50. Ghouri, M.M.; Afzal, S.; Hussain, R.; Blank, J.; Bukur, D.B.; Elbashir, N.O. Multi-scale modeling of fixed-bed Fischer Tropsch reactor. Comput. Chem. Eng. 2016, 91, 38–48. [Google Scholar] [CrossRef]
  51. Okeke, I.J.; Sahoo, K.; Kaliyan, N.; Mani, S. Life cycle assessment of renewable diesel production via anaerobic digestion and Fischer-Tropsch synthesis from miscanthus grown in strip-mined soils. J. Clean. Prod. 2020, 249, 119358. [Google Scholar] [CrossRef]
  52. Okeke, I.J.; Mani, S. Techno-economic assessment of biogas to liquid fuels conversion technology via Fischer-Tropsch synthesis. Biofuels Bioprod. Biorefining 2017, 11, 472–487. [Google Scholar] [CrossRef]
  53. Naqi, A.; Kuhn, J.N.; Joseph, B. Techno-economic analysis of producing liquid fuels from biomass via anaerobic digestion and thermochemical conversion. Biomass Bioenergy 2019, 130, 105395. [Google Scholar] [CrossRef]
  54. Tijmensen, M.J.; Faaij, A.P.; Hamelinck, C.N.; van Hardeveld, M.R. Exploration of the possibilities for production of Fischer Tropsch liquids and power via biomass gasification. Biomass Bioenergy 2002, 23, 129–152. [Google Scholar] [CrossRef]
  55. Swanson, R.M.; Platon, A.; Satrio, J.A.; Brown, R.C. Techno-economic analysis of biomass-to-liquids production based on gasification. Fuel 2010, 89, S11–S19. [Google Scholar] [CrossRef]
  56. Hamelinck, C.N.; Faaij, A.P.; den Uil, H.; Boerrigter, H. Production of FT transportation fuels from biomass; technical options, process analysis and optimisation, and development potential. Energy 2004, 29, 1743–1771. [Google Scholar] [CrossRef]
  57. Kauffman, N.; Hayes, D.; Brown, R. A life cycle assessment of advanced biofuel production from a hectare of corn. Fuel 2011, 90, 3306–3314. [Google Scholar] [CrossRef]
  58. De Jong, S.; Antonissen, K.; Hoefnagels, R.; Lonza, L.; Wang, M.; Faaij, A.; Junginger, M. Life-cycle analysis of greenhouse gas emissions from renewable jet fuel production. Biotechnol. Biofuels 2017, 10, 64. [Google Scholar] [CrossRef] [PubMed]
  59. Dang, Q.; Yu, C.; Luo, Z. Environmental life cycle assessment of bio-fuel production via fast pyrolysis of corn stover and hydroprocessing. Fuel 2014, 131, 36–42. [Google Scholar] [CrossRef]
  60. Fortier, M.-O.P.; Roberts, G.W.; Stagg-Williams, S.M.; Sturm, B.S. Life cycle assessment of bio-jet fuel from hydrothermal liquefaction of microalgae. Appl. Energy 2014, 122, 73–82. [Google Scholar] [CrossRef]
  61. Dry, M.E. Fischer-Tropsch synthesis over iron catalysts. Catal. Lett. 1990, 7, 241–251. [Google Scholar] [CrossRef]
  62. Rytter, E.; Tsakoumis, N.E.; Holmen, A. On the selectivity to higher hydrocarbons in Co-based Fischer–Tropsch synthesis. Catal. Today 2016, 261, 3–16. [Google Scholar] [CrossRef]
  63. Khodakov, A.Y. Fischer-Tropsch synthesis: Relations between structure of cobalt catalysts and their catalytic performance. Catal. Today 2009, 144, 251–257. [Google Scholar] [CrossRef]
  64. Borg, Ø.; Eri, S.; Blekkan, E.A.; Storsæter, S.; Wigum, H.; Rytter, E.; Holmen, A. Fischer–Tropsch synthesis over γ-alumina-supported cobalt catalysts: Effect of support variables. J. Catal. 2007, 248, 89–100. [Google Scholar] [CrossRef]
Figure 1. (a) View of the FTR experimental apparatus. (b) Schematic diagram of experimental setup.
Figure 1. (a) View of the FTR experimental apparatus. (b) Schematic diagram of experimental setup.
Energies 16 07009 g001aEnergies 16 07009 g001b
Figure 2. Flowchart of computational procedure of genetic algorithm.
Figure 2. Flowchart of computational procedure of genetic algorithm.
Energies 16 07009 g002
Figure 3. H2-TPR patterns for cobalt-based catalysts.
Figure 3. H2-TPR patterns for cobalt-based catalysts.
Energies 16 07009 g003
Figure 4. SEM images of the catalyst: (a) precursor, (b) before the test, and (c) after the test.
Figure 4. SEM images of the catalyst: (a) precursor, (b) before the test, and (c) after the test.
Energies 16 07009 g004
Figure 5. Effects of temperatures on the CO conversion and products selectivity at 5 bar, 1000 mL/gcat·h, and H2/CO = 0.5.
Figure 5. Effects of temperatures on the CO conversion and products selectivity at 5 bar, 1000 mL/gcat·h, and H2/CO = 0.5.
Energies 16 07009 g005
Figure 6. Effects of total pressure on the conversion of CO and products selectivity at 200 °C, 1000 mL/gcat·h, and H2/CO = 0.5.
Figure 6. Effects of total pressure on the conversion of CO and products selectivity at 200 °C, 1000 mL/gcat·h, and H2/CO = 0.5.
Energies 16 07009 g006
Figure 7. Effect of space velocity on the conversion of CO and products selectivity at 200 °C, 25 bar, and H2/CO = 0.5.
Figure 7. Effect of space velocity on the conversion of CO and products selectivity at 200 °C, 25 bar, and H2/CO = 0.5.
Energies 16 07009 g007
Figure 8. Effect of proportion of inlet feed (H2/CO) on the conversion of CO and products selectivity at 200 °C, 25 bar, and 1000 mL/gcat. h.
Figure 8. Effect of proportion of inlet feed (H2/CO) on the conversion of CO and products selectivity at 200 °C, 25 bar, and 1000 mL/gcat. h.
Energies 16 07009 g008
Figure 9. Parity plot comparing experimental data and results predicted by a power-law rate model for CO2, CO, CH4, C2H4, C2H6, C3H8, n-C4H10, and i-C4H10.
Figure 9. Parity plot comparing experimental data and results predicted by a power-law rate model for CO2, CO, CH4, C2H4, C2H6, C3H8, n-C4H10, and i-C4H10.
Energies 16 07009 g009aEnergies 16 07009 g009b
Figure 10. Objective function changes as the number of iterations.
Figure 10. Objective function changes as the number of iterations.
Energies 16 07009 g010
Table 1. Fischer–Tropsch reactions and activation energy.
Table 1. Fischer–Tropsch reactions and activation energy.
ReactionsEj
(kJ/mol)
[15]
Ej
(kJ/mol)
[16]
Ej
(kJ/mol)
[11]
Ej
(kJ/mol)
[17]
Ej
(kJ/mol)
[14]
CO + 3 H 2   CH 4 + H 2 O 83.423915.69383.599.79101.15
2 CO + 4 H 2 C 2 H 4 + 2 H 2 O 65.01820.38465.072.5178.79
2 CO + 5 H 2 C 2 H 6 + 2 H 2 O 49.78215.60749.848.5159.95
3 CO + 7 H 2 C 3 H 8 + 3 H 2 O 34.885516.40634.931.0333.73
4 CO + 9 H 2 n C 4 H 10 + 4 H 2 O 27.728986.93427.727.5923.04
4 CO + 9 H 2 i C 4 H 10 + 4 H 2 O 25.730181.75325.710.1417.83
6.5 CO + 12.23 H 2 C 6.05 H 12.36 C 5 + + 6.05 H 2 O 23.564373.87823.612.4421.25
CO + H 2 O CO 2 + H 2 58.826355.32958.810.0050.24
Table 2. Textural characteristics of support and catalysts.
Table 2. Textural characteristics of support and catalysts.
SampleBET SA (m2/g)Pore Volume
(cm3/g)
Pore Radius
(nm)
SiO22260.519
10Co/SiO22050.438.6
1 Ce10Co/SiO22000.428.4
Table 3. Experimental results at different operation conditions using the 1Ce-10Co/SiO2 catalyst.
Table 3. Experimental results at different operation conditions using the 1Ce-10Co/SiO2 catalyst.
T (°C)P (bar)Sv (mL/gcat·h)H2/CO%XCOExperimental Products Selectivity (%)
CH₄C2H₄C2H6C₃H6n-C₄H10i-C₄H10CO2C5+
200510000.514.94.3681.1762.1501.4281.1761.6801.60086.422
210510000.517.36.9541.4642.3421.5561.2811.8301.90082.673
220510000.521.69.4051.6723.0931.6721.4631.8812.30078.514
230510000.524.813.1082.1923.3901.8081.5822.0342.50073.386
240510000.527.917.8502.6273.2901.7851.6582.0662.80067.926
200510000.514.94.3681.1762.1501.4281.1761.6801.60086.422
2001010000.523.45.1701.3681.7281.2240.9941.5701.46886.479
2001510000.527.55.4801.5071.3841.1650.9591.3701.41086.726
2002010000.529.95.0001.6001.2631.0630.8751.2501.40087.550
2002510000.536.34.7201.5101.1921.0030.8261.1801.30088.269
2002510000.536.34.7201.5101.1921.0030.8261.1801.30088.269
2002520000.533.55.0001.6001.2631.0630.8751.1251.70087.375
2002530000.530.56.3201.7381.1061.3430.9481.5801.73085.235
2002540000.526.66.9601.7401.1310.8700.9571.7401.73584.867
2002550000.520.48.3201.7681.1440.8321.0402.1011.73483.061
2002510000.536.34.7201.5101.1921.0030.8261.1801.30088.269
200251000147.94.3341.0751.0670.7140.5880.8401.20090.182
200251000252.83.9730.9861.0010.6550.5390.7781.20090.869
200251000347.44.5921.1391.1570.7300.8010.7651.22089.595
200251000443.25.1081.2671.2870.8420.6930.9901.23088.583
Table 4. Predicted results at different operation conditions using the 1Ce-10Co/SiO2 catalyst.
Table 4. Predicted results at different operation conditions using the 1Ce-10Co/SiO2 catalyst.
T (°C)P (bar)Sv (mL/gcat·h)H2/CO%XCOPredicted Products Selectivity (%)
CH₄C2H₄C2H6C₃H6n-C₄H10i-C₄H10CO2C5+
200510000.515.34.2711.1762.1531.4301.1801.7661.67486.351
210510000.518.16.4341.4622.3181.5471.2811.8111.90083.247
220510000.521.09.4061.7842.4741.6621.3771.8542.15379.289
230510000.523.813.3092.1302.6201.7771.4681.9012.45474.340
240510000.525.418.1732.4752.7331.8811.5411.9502.88868.359
200510000.515.34.2711.1762.1531.4301.1801.7661.67486.351
2001010000.525.04.5711.3981.7441.2961.0581.5831.34487.005
2001510000.525.14.7501.5071.4611.1460.9411.3741.42187.399
2002010000.525.54.8781.5961.3051.0670.8761.2701.40087.608
2002510000.525.74.9801.6671.1921.0030.8271.1801.36787.785
2002520000.524.75.0311.6601.1470.9460.7971.0770.60288.740
2002530000.524.25.0491.6551.1300.9260.7851.0440.34389.069
2002540000.520.55.0631.6201.0690.8610.7440.9490.25689.437
2002550000.517.25.0691.6021.0430.8330.7250.9100.21289.606
2002510000.525.74.9801.6671.1921.0030.8271.1801.36787.785
200251000149.74.8841.4581.1430.8700.7560.9691.03988.882
200251000254.94.7091.1981.1090.7410.6840.7820.86689.911
200251000347.64.5921.0921.1570.7300.6840.7650.92890.051
Table 5. The optimal values of the kinetic parameters for the current kinetic model.
Table 5. The optimal values of the kinetic parameters for the current kinetic model.
Reaction No.Ao (mol/gmcat.h)E (kJ/mol)njmj
1148.88968880.2140.820.60
20.00541360.4240.250.08
30.00288645.8140.690.55
40.00014732.4950.160.13
50.00002928.9320.660.54
60.00001227.8710.090.42
70.00007112.8610.550.63
80.00012537.8500.130.22
Table 6. Genetic optimization table.
Table 6. Genetic optimization table.
Iteration No.Temperature (°C)Pressure (bar)GHSV (mL/gcat·h)H2/COC5+ Selectivity
1203.629.962431.982.5548.144
2203.6222.414679.281.0848.144
3239.8522.414679.282.8348.144
4203.6222.414679.281.0848.144
5239.1717.002720.281.8256.916
9203.6216.834648.611.8257.596
13201.9222.314648.611.5260.792
16227.5617.613854.652.8961.268
18201.9213.981626.902.5577.112
20204.1622.411914.133.8980.784
28211.326.591914.133.4181.396
41206.516.591798.413.4181.464
57202.5018.794808.263.7082.28
59202.5022.821171.163.7082.348
94201.0013.841023.513.8488.264
100202.9115.113885.253.8588.672
120216.155.201023.513.8589.284
124237.446.101067.013.6689.556
240200.656.211638.913.8189.692
286200.1115.961012.162.7891.12
446200.116.213206.023.0691.188
851200.096.211513.963.9091.188
852230.416.211513.963.7091.46
1000200.096.291529.583.9691.664
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Al-Zuhairi, F.K.; Shakor, Z.M.; Hamawand, I. Maximizing Liquid Fuel Production from Reformed Biogas by Kinetic Studies and Optimization of Fischer–Tropsch Reactions. Energies 2023, 16, 7009. https://doi.org/10.3390/en16197009

AMA Style

Al-Zuhairi FK, Shakor ZM, Hamawand I. Maximizing Liquid Fuel Production from Reformed Biogas by Kinetic Studies and Optimization of Fischer–Tropsch Reactions. Energies. 2023; 16(19):7009. https://doi.org/10.3390/en16197009

Chicago/Turabian Style

Al-Zuhairi, Firas K., Zaidoon M. Shakor, and Ihsan Hamawand. 2023. "Maximizing Liquid Fuel Production from Reformed Biogas by Kinetic Studies and Optimization of Fischer–Tropsch Reactions" Energies 16, no. 19: 7009. https://doi.org/10.3390/en16197009

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop