Open access peer-reviewed chapter

Genetic Algorithms for Chemical Engineering Optimization Problems

Written By

Thi Anh-Nga Nguyen and Tuan-Anh Nguyen

Submitted: 20 December 2021 Reviewed: 11 April 2022 Published: 19 May 2022

DOI: 10.5772/intechopen.104884

From the Edited Volume

Genetic Algorithms

Edited by Sebastián Ventura, José María Luna and José María Moyano

Chapter metrics overview

273 Chapter Downloads

View Full Metrics

Abstract

Chemical engineering processes are frequently composed of multiple complex phenomena. These systems can be represented by a set of several equations, which are referred to as mathematical model of the process. Optimization in chemical engineering utilizes specialized techniques to determine the values of the decision variables at which the performance of the process, measured as the objective function(s), is minimum or maximum. The profitability of the process improves remarkably as a result of this selection. This benefit has encouraged the broad application of optimization for important industrial challenges. However, many problems in chemical engineering processes are hard to find the optimum using gradient-based algorithms. For example, the cases when the objective functions of the processes are multimodal, discontinuous, or implicit. Genetic algorithms (GAs) are a kind of metaheuristic searching optimization methods, which are inspired by nature, the mechanics of natural evolution and genetics. Genetic algorithms have received significant attention due to their remarkable advantages over classical algorithms. Compared with traditional optimization approaches, GAs are straightforward, robust, capable of handling the non-differentiable, discontinuous, or multimodal problems. The purpose of this paper is to give several case studies using genetic algorithms in chemical engineering optimization problems.

Keywords

  • optimization
  • genetic algorithm
  • chemical engineering
  • modeling

1. Introduction

Chemical engineering processes are frequently composed of multiple complex phenomena. These systems can be represented by a set of several equations, such as z = f(d; x; p), where z is the vector state of the system. The system of equations is referred to as the mathematical model of the process. The performance of the process is predicted by the model from the assigned data of several “input” variables, d and x, and a group of parameters, p. Among the input variables, some, referred as x, can be changed and are known as design variables, while others, referred as d, are predetermined. The performance of the process can be evaluated through a set of output variables, y = g (z), referred as the function of state of the system. Optimization in chemical engineering utilizes specialized techniques to determine the values of the decision variables, x, at which the performance of the process, measured as the objective function(s), I(x), is minimum or maximum. The profitability of the process improves remarkably as a result of this selection of input/operating/decision variables. This benefit has encouraged the broad application of optimization for important industrial challenges. However, many problems in chemical engineering processes are hard to find the optimum using gradient-based algorithms. For example, the cases when the objective functions of the processes are multimodal, discontinuous, or implicit. Genetic algorithms (GAs) are a kind of metaheuristic searching optimization methods, which are inspired by nature, the mechanics of natural evolution, and genetics [1]. Genetic algorithms have received significant attention due to their remarkable advantages over classical algorithms. Compared with traditional optimization approaches, GAs are straightforward, robust, capable of handling the non-differentiable, discontinuous, or multimodal problems. GAs have been effectively employed in a wide range of various engineering, manufacturing, and management applications [2]. The purpose of this chapter is to give several case studies using genetic algorithms in chemical engineering optimization problems. The case studies include the optimization of an autothermal ammonia synthesis reactor, a separation module using membrane technology, and data-driven modeling optimizations of solid oxide fuel cells.

Advertisement

2. Optimization of an autothermal ammonia synthesis reactor

In the chemical process industries, ammonia is one of the most widely manufactured inorganic compounds [3]. The majority of ammonia produced commercially is consumed in fertilizers, with the rest going into plastics, synthetic fibers and resins, pharmaceuticals, explosives, papers, and refrigeration [4]. As a result, modeling and optimization of ammonia synthesis process have received a significant attention from both the academia and industry. Ammonia is produced predominantly from the combination of elements such as nitrogen and hydrogen in a catalytic process using a promoted iron catalyst firstly established by Haber and Bosch as the reaction [4]:

N2+3H22NH3E1

The reaction is reversible and exothermic, releasing a significant amount of heat. In order to achieve a high conversion, the heat of the reaction should be removed. Therefore, the process is typically carried out in an autothermal synthesis reactor, in which the heat of reaction is utilized to preheat the feed gas and ensure the suitable temperature inside. The production of ammonia depends on several factors such as the reactor length, the operating pressure, temperature of the feed and reacted gas, the flow rate, and composition of the gas mixture. The optimization problem of the process is to maximize the economic return. Many studies discussing the modeling, simulation, and optimization of an autothermal ammonia synthesis reactor can be found in literature. Some of them can be mentioned here as in Babu et al. [5], Babu and Angira [6], Carvalho et al. [7], Edgar et al. [8], Ksasy et al. [9], Murase et al. [10], Upreti and Deb [11], Yusup et al. [12]. However, the model discussed in the studies of Edgar et al. [8], Murase et al. [10] has some minor errors and has been corrected in Upreti and Deb [11]. Moreover, the studies primarily focus on optimizing reactor length for a specific reactor top temperature, usually 694 K [6, 7, 12], or for a limited set of temperatures [9, 11]. However, as reported in some studies [11, 12], the economic return is determined by the top temperature and also the reactor length (the temperature of feed gas entering to the reaction zone). As a result, rather than a single variable problem of reactor length, the optimization problem should be viewed as a multivariable problem.

In the study [13], both the reactor length and the reactor top temperature are considered in the design variables for maximizing the profit return of the process. In order to solve the multivariate optimization problem, the cyclic coordinate search technique was employed. This method alters the value of one decision variable at a time, and for each coordinate direction, the golden section search was utilized to solve the single variable optimum problem. However, this traditional searching approach is prone to get caught in local optima. Therefore, the genetic algorithm has higher chance to obtain the global optimum profit of the process.

2.1 Problem formulation of ammonia synthesis reactor

The system discussed here is an autothermal synthesis reactor, which is described in [10] and contains the correction of the objective function reported in [6, 11]. The feed gas contains 21.75 mole% nitrogen, 65.25 mole% hydrogen, 5.0 mole% ammonia, 4.0 mole% methane, and 4.0 mole% argon. In an autothermal reactor, the feed gas mixture enters from the bottom of the reactor, flows upward, enters the catalyst zone from the top, and moves downward. In the catalyst zone, the reaction takes place at around 500°C and 200 atm of pressure. The heat generated by the reaction is utilized to preheat the feed gas mixture in counter current flow. Figure 1 shows the schematic diagram of an autothermal ammonia synthesis reactor. The considered factors affecting the synthesis process are the temperature of feed gas at the entrance of the reaction zone (top temperature) and the reactor length. The goal of the optimal design is to determine the conditions that will give the highest economic return from the reactor operation.

Figure 1.

Schematic diagram of an autothermal ammonia reactor [10].

2.1.1 Objective function

The return of the process, which is calculated from the value of the product gas (heating value and ammonia value), subtract the cost of feed gas (as a source of heat only) and minus the amortization of reactor capital expenses, is the objective function for maximization (F). Other operating costs are not considered [11].

F=1.3356×1071.708×104NN2+704.09TgT0699.27TfT03.4566×107+1.9837×109L12E2

in which,

Tf is the temperature of the feed gas.

Tg is the temperature of the reacting gas (the gas in the catalyst zone).

T0 is the top temperature or temperature at the inlet of the catalyst zone.

NN2 is the flow rate of nitrogen.

L is the length of the reactor.

2.1.2 Equality constraints

The heat balance for the feed gas and the reacting gas and the mass balance for the nitrogen flow along the catalyst zone, respectively, give the mathematical model for the system:

dTfdx=US1WCpfTgTfE3
dTgdx=US1WCpgTgTf+ΔHS2WCpgdN2dxE4
dNN2dx=fk1pN2pH21.5pNH3k2pNH3pH21.5E5

in which

k1=1.78954×104exp20800RTgE6
k2=2.5714×1016exp47400RTgE7
pH2=3pN2E8
pN2=286NN22.598NN20+2NNE9
pNH3=2862.23NN02NN22.598NN0+2NNE10

The differential equations are valid in the interval [0, L], in which L is the reactor length (or the length of the catalyst zone).

The notations Tf0, Tg0, and NN20 denote the initial value at x = 0 (at the inlet of the catalyst zone) for Tf, Tg, and NN2, respectively, and are given by

Tf0=Tg0=T0,NN20=701.2kmol/hm2E11

Other notations of the system are summarized in Table 1.

Notation
CpfHeat capacity of the feed gas
CpgHeat capacity of the reacting gas
fCatalyst activity
ΔHHeat of reaction
NMass flow of component designed by subscript
kReaction rate constant
pPartial pressure of component designated by subscript
RUniversal gas constant
S1Surface area of catalyst tubes per unit length of reactor
S2Cross-sectional area of catalyst zone
UOverall heat transfer coefficient
WTotal mass transfer flow rate

Table 1.

Notation of the synthesis system.

2.1.3 Box constraints

The variables are subjected to the following physical constraints, as is typical in industries [10]:

0<L10,400Tf800,0NN23220E12

The length of the reactor and the top temperature are chosen as the design variables. The remaining variables (Tf, Tg, and NN2) can be calculated from the model by three differential Eqs. (3), (4), and (5). Then, the objective function is determined by Eq. (2). Due to the constraints of temperatures, the variable T0 is also set to be within 400 and 800.

The optimal design problem is summarized as follows:

maximizeF=FxT0s.t.dTfdx=US1WCpfTgTfdTfdx=US1WCpgTgTf+ΔHS2WCpgdN2dxdN2dx=fk1pN2pH21.5pNH3k2pNH3pH21.5Tf0=Tg0=T0,NN20=701.2400Tf800,0NN232200<L10,400T0800E13

2.2 Optimization strategy

The system of ordinary differential Eqs. (3), (4), and (5) with initial conditions (11) was solved by Runge–Kutta fourth-order method. The system is well defined when the top temperature (T0) and the interval of integration (the reactor length L) are assigned. After that, the economic return is clearly evaluated, and it can be considered as two-variable function of the top temperature and reactor length. The genetic algorithm is employed to find the optimal solution. In order to handle the constraints, a penalty or barrier is defined to the objective function whenever any of variable limits is violated. The proposed strategy is detailed as follows.

2.2.1 Genetic algorithm for optimization problem

The range of the design variables is 0<x10,400T0800. The parameters of GA such as population size, crossover probability, mutation probability values were set to be 50, 1.0, and 0.30, respectively. The selection was chosen as roulette wheel selection with elitism. The number of generations was set to be 500.

2.2.2 Barrier method for constrained optimization

In barrier or penalty methods, the objective function will receive an undesired value when one of the constraints is violated. Therefore, the solution will be kept in the feasible region. The objective function has been modified as

F=Fif400Tf800,0NN232200otherwiseE14

2.2.3 Parameters

The parameters were obtained from the literature [7, 10] and summarized in Table 2.

ParameterValueUnit
Cpf0.707kcal/kg K
Cpg0.719kcal/kg K
f1.0
ΔH−26,000kcal/kmol
R1.987kcal/kmol K
S110M
S20.78m2
U500kcal/h m2 K
W26,400kg/h

Table 2.

Model parameters [7].

2.3 Optimization results

Figure 2 shows the fitness values as a function of generation. As can be observed, the fitness function value achieved the highest after roughly 20 generations and then stayed unchanged. After 100 generations, it was obtained that the reactor length should be 6.772 m, and the top temperature should be 707.09 K. The process produces a profit of 5.018× 106 $ per year. The other parameters of the process are summarized in Table 3 and compared with the findings of a cyclic coordinate search [13]. The profit value is slightly higher than those reported in the literature, which focused solely on reactor length optimization. From the results, the temperature at the entrance of the catalyst zone should be slightly higher, and that the reactor length should also be slightly longer than previously reported.

Figure 2.

Fitness value versus generations.

VariablesIntervalCyclic coordinate [13]Genetic algorithm
x (m)[0,10]6.7246.772
T0 (K)[600,800]700.27707.09
Tf (K)[400,800]400.00401.09
Tg (K)629.94631.12
NN2[0,3220]490.68490.68
F (106$/year)5.0185.018

Table 3.

Maximization results.

Advertisement

3. Economic optimization of membrane module for ultrafiltration of protein solution

The behavior of permeate flux has a significant impact on the performance of cross-flow ultrafiltration. Many factors cause flux declination, such as solution properties, membrane properties, and operation conditions. The majority of current research has centered on increasing membrane performance in terms of permeability and selectivity [14, 15]. Just a few studies have paid attention to the configuration and operation of the membrane module [16].

Various factors determine the decision of membrane module geometry for a given application, including fabrication method, power consumption, and fouling potential [17]. Manufacturers frequently recommend the membrane module design from the fabrication standpoint [17]. There is virtually no evidence that their approach prioritizes the energy efficiency. Currently, with a growing in energy concern and a falling in membrane cost, the membrane module design should place a higher attention on energy efficiency. As a result, it is necessary to propose a module design methodology that takes into account the energy factor.

Furthermore, membrane operating conditions are usually decided by user experience, a handbook, or a manual from the membrane supplier. However, the permeate flux equation governing the performance of the membrane system varies greatly between different situations. In this aspect, for any specific application, a general methodology for the design and operation conditions should be studied.

In cross-flow ultrafiltration of protein solution, Nguyen et al. [18] proposed a simple combined model, which simultaneously considers pore blockage and cake filtration, to describe the flux declination. Then, in the study [19], the correlation between the steady-state permeate flux and operation parameters was reported. From the steady-state operation equation, optimal design and operation conditions for each particular application could be established.

However, just a few reports on the optimization of membrane processes and cost estimation have been published, or the cost estimation is too general. For example, Wiley et al. [17] optimized the membrane module configurations for brackish water desalination. However, the operation mode is single-pass and only the membrane cost and energy cost were taken into account. Sethi and Wiesner [20] developed the cost model for the removal of natural organic matter, but the study has not conducted the optimization. In membrane technology, the feed and bleed operation mode, which combines the batch and the single-pass configurations, is commonly utilized for continuous full-scale filtration [21, 22]. Therefore, the optimization of a membrane module operated in feed and bleed mode for protein ultrafiltration is considered. The membrane geometry dimensions and operating conditions are design variables in the problem. The system is represented by a set of ordinary differential equations. The objective function is the annual cost, which consists of various types of capital investments and an operating expense. The capital investments are classified into several categories, which are individually correlated to plant scale, particularly the membrane area. The operating expense is the power consumption.

3.1 Process configuration and model calculation

3.1.1 Membrane plant configuration

The configuration of filtration system is continuous feed and bleed, which is shown schematically in Figure 3. The notations are summarized in Table 4. There are two main pumps in this operation: the feed pump provides the necessary trans-membrane pressure, while the recirculation pump maintains the cross-flow rate through the modules. The concentrate is continually withdrawn from the system at a flow rate (R).

Figure 3.

Schematic configuration of feed-and-bleed mode membrane system.

NotationName and units
FFeed flow rate [m3/hr]
RRetentate (concentrate) flow rate [m3/hr]
QRecirculation flow rate [m3/hr]
FlowFlow rate in membrane module [m3/hr]
PPermeation flow rate [m3/hr]
PFPressure at outlet of feed pump [kPa]
PiPressure at the inlet of membrane module [kPa]
PoPressure at the inlet of membrane module [kPa]
EPEnergy consumed by the feed pump [kW]
EQEnergy consumed by the recirculation pump [kW]
ϕ0Initial concentration of protein solution [m3/m3]
ϕiInlet concentration of protein solution [m3/m3]
ϕfFinal concentration of protein solution [m3/m3]
ϕPConcentration of protein in permeate flux [m3/m3]
uFluid flow velocity [m/s]
ρFluid density [kg/m3]
μFluid viscosity [kg/(m·s)]
w, h, L, dhWidth, height, length, hydraulic diameter of the membrane module

Table 4.

Summary of system configuration notations.

3.1.2 Modeling of membrane modules

The material balance for total mass and protein give:

F=R+PFϕ0=Rϕf+PϕpFϕ0+Qϕf=F+QϕiE15

The viscosity and density of protein solution correlate to its concentration [23]:

μm=8.94×104exp13.5482ϕmρm=10001ϕm+1360ϕmE16

in which, μm, ϕm are the viscosity and concentration of the mixture (solution), respectively. ρm is the density of protein solution, 1000 kg/m3 is the density of water, and 1360 kg/m3 is the density of dry protein powder.

The permeate flux through the membrane is [19].

Ju=3.66×107Pρmu20.27ρmudhμm0.52E17

in which P is the trans-membrane pressure, u is the cross-flow velocity, dh is the hydraulic diameter of the flow channel.

The equation for permeate flux can be rewritten as:

J=5.124×109Pρmu20.27ρmudhμm0.52udhE18

or in terms of shear rate γ̇=6uh=12udh, where h is the channel height:

J=4.27×1010Pρmu20.27ρmudhμm0.52γ̇E19

The flow rate/velocity drop and channel length change are calculated from the total mass balance and component balance within the control volume dz across the channel. Protein is assumed to be fully rejected by the membrane:

dFlow×ϕm=0E20
dFlow=J×dwz=JwdzE21

The pressure loss is estimated by the Darcy-Weisbach Equation [24, 25].

dP=4fρmdzdhu22E22

in which f is the friction factor.

The set of ordinary equations that describes the membrane module system was established as follows [26].

dFlowdϕm=Flowϕmdzdϕm=FlowϕmwJdPdϕm=4fρm1dhu22Flowϕm1wJμm=8.94×104exp13.5482ϕmρm=10001ϕm+1360ϕmJ=5.124×109Pρmu20.27ρmudhμm0.52udhu=Flowh×wf=24ReifRe<2000f=0.079Re0.25ifRe>2000E23

in the range of concentration [ϕi, ϕf] and the initial condition

Flowϕi=F+Qzϕi=0Pϕi=PiE24

The system of the ordinary equations can be solved numerically by Runge–Kutta fourth-order method [27] to obtain the flow rate, the length, and the pressure. From that, the two important factors determining the total cost, membrane area, and total energy were calculated:

Amembrane=w×LE25
Epumps=EP+EQ=F×Pi+Q×ΔPdropE26

In this equation,

EP, EQ are the power supplied by the feed pump and recirculation pump, respectively.

PF, Pi, Po are the pressure at the outlet of the feed pump, at the inlet and outlet of the membrane module, respectively.

ΔPdrop is the pressure drop in the membrane module.

3.1.3 Cost estimation

3.1.3.1 Operating cost

The operating cost consists of power consumption of the pumps and membrane replacement. The annual energy expense of the pumps is calculated as

Cenergy=Epumpsη×8000×36001000×electricity price$/yearE27

Epumps is from Eq. (26). The system is assumed to work 8000 hr./year (24 hr./day and 333 days/year). η is the efficiency of the pumps, which is set to be 0.7. The electricity price is supposed to be 0.08 $/kWh, which is the price for the industrial sector in the United States [28].

The membrane replacement cost is calculated as

Cmembrane=Amembrane×cmembrane×AP$/yearE28

where Cmembrane [$/m2/year] is the membrane replacement cost calculated per year.

cmembrane [$] is the membrane price per unit area,

AP is the amortization factor, which presents the time value of money [29] and is calculated as a function of interest rate i and the membrane life.

The membrane price is usually about 200 $/m2 [9], and membrane life is 12–18 months. Therefore, the membrane replacement cost per year is roughly estimated as 200 $/m2/year for the interest of i = 8%.

3.1.3.2 Capital cost

It is widely observed that capital costs are correlated to the size in the power-law form [30]:

cost=ksizenE29

In order to achieve higher accuracy, rather than simply predicting the whole capital cost of the membrane plant to capacity, Sethi and Wiesner [20] divided the capital investment into several major categories, which was correlated to the size independently. The major categories include pumps and other manufactured equipment.

  1. Pump capital cost

    The pumps capital cost can be estimated as (Perry et al. [31]):

    Cpump=I×f1×f2×CL×81.27×Q×P0.4E30

    in which.

    I: a cost index ratio for updating the cost to the recent year.

    f1: an adjust factor for pump construction material.

    f2: an adjust factor for suction pressure range.

    CL: a factor used to incorporate labor costs.

    Q: flow capacity of the pump [m3/h].

    P: pressure outlet of the pump [kPa].

    The cost index, I in Eq. (30), can be referred to as the chemical engineering (CE) index to update the cost. It can be obtained from [32], and the value of 2.4 is used. CL = 1.4 with the assumption that 40% of the cost is required to install the equipment [33]. The factors f1 and f2 can be found in [31]. f1 = 1.5 when the material is stainless steel. f2 = 1.0 when the pump pressure is below 10 bar (1 MPa).

    The pump size (Q × P in Eq. (30)) of the two pumps (feed and recirculation pump) is

    pump size=F+Q×PiE31

  2. Capital cost of other equipment

    In membrane application, the membrane area is the key parameter, which determines the plant capacity [34]. Thus, the membrane area is chosen as the basic for the estimation of various components in the capital costs.

    Non-membrane equipment and facilities, excluding the pumps, were grouped into four main categories: (1) pipes and valves; (2) instruments and controls; (3) tanks and frames; and (4) miscellaneous. The capital cost of each is correlated to the membrane area as follows (Sethi and Wiesner [20])

    1. Pipes and valves

      CPV=6000Amembrane0.42$E32

    2. Instruments and controls

      CIC=1500Amembrane0.66$E33

    3. Tanks and frames

      CTF=3100Amembrane0.53$E34

    4. Miscellaneous

      CMI=8000Amembrane0.57$E35

  3. Annual capital cost

    The capital cost can be annualized using the amortization factor as

    Ccapital cost=Ccapital×AP$/yearE36

    For the plant design year of 20 years and the interest rate 8%, the amortization factor will be about 0.1.

3.2 Formulizations of the problem

3.2.1 Fix parameters and design variables

In the problem, some variables, called input variables, are fixed due to the requirement of the design. In membrane design, these are feed flow F, inlet concentration ϕ0, and outlet concentration ϕf. The protein is assumed to be entirely rejected by the membrane (ϕp = 0).

The design variables were: channel geometry (width × length × height), the inlet pressure (Pi), and recirculation flow rate (Q).

3.2.2 Objective function

The objective function is the sum of capital cost and operating cost, which were annualized:

minimumfx=annual total cost=annual capital cost+annual operating cost=Cenergy+Cmembrane+Ccapitalpumps+otherE37

3.2.3 Constraints

The pressure at the outlet point should be positive. This constraint is satisfied by assigning a high value to the objective function if the outlet pressure is negative.

The decision variables are frequently limited on a finite range

xlxxuE38

in which x is the vector of decision variables x = (Pi,Q,w,h), subscripts l and u indicate the lower and upper bound.

The system parameters and variables are summarized in Table 5.

ParametersValue
Feed flow rate (m3/hr)0.02–200
Inlet pressure (kPa)200–1000
Recirculation flow rate (m3/hr)0–50
Initial solid fraction (m3/m3)0.1
Final solid fraction (m3/m3)0.4
Plant design year (year)20
Interest rate (%)8
Energy price ($/kWh)0.08
Efficiency of pumps (%)70
Operating temperature (°C)25
Module height (mm)0–100
Module width (m)0–30

Table 5.

System parameters and variables.

3.2.4 Optimization by genetic algorithm

The parameters of GA such as population size, crossover probability, mutation probability values were set to be, 100, 1.0, and 0.30, respectively. The selection was based on roulette wheel with elitism, which means the most fit individual is guaranteed a place in the next generation. The number of generations was assigned to be 500. Because the problem is to minimize the cost, the fitness function was defined as:

fitness=0iffx>5×1055×105fxotherwiseE39

3.3 Optimum design

For the demonstration of this method, optimum designs of several feed flow rates have been carried out. The lower limit of the membrane width is 0.1 m, the lower limit for the module height is 0.5 mm. The designs are shown in Table 6.

Feed [m3/hr]ϕ0 [−]ϕf [−]Pressure [kPa]Recirculation [m3/hr]width [m]height [mm]total cost [$/yr]
0.020.10.45232.80.15.01.29 × 103
0.20.10.410004.90.18.94.30 × 103
20.10.410000.20.16.91.18 × 104
200.10.49870.21.35.05.60 × 104
2000.10.410000.811.15.03.65 × 105

Table 6.

Optimum designs of membrane module.

Figure 4 presents the optimum total cost per unit of feed flow. The cost per unit of feed flow decreases with an increase in plant capacity. It reflects the economies of scale.

Figure 4.

The behavior of cost per unit flow rate design in optimum condition with plant capacity.

The results also suggest that the membrane module dimensions and operation condition will change greatly depending on the process requirements, such as the required feed capacity. It is challenging to predict the direction. It might be concluded that the permeate flux also greatly affects the geometric design and operation strategy in membrane separation processes. It is difficult to find a general rule for the design, for each specific system, the correlation between the permeate flux and operating conditions and membrane geometry should be investigated.

Advertisement

4. Modeling and optimization of the BSCF-based single-chamber solid oxide fuel cell by artificial neural network and genetic algorithm

Fuel cells that are highly effective and green technology for converting chemical energy stored in fuel to useable power are currently regarded as one of the most promising approaches for future energy requirements [35]. The solid oxide fuel cell (SOFC) has demonstrated an exceptional integration of advantages, such as high efficiency, fuel flexibility, wide contamination acceptance, and low pollution [36, 37]. Modeling and simulation are valuable tools for determining the impact of various design factors and operating conditions on cell performance, as well as for improving fuel cells [38, 39, 40]. Plenty of models have been reported to add to the understanding of fuel cells. Modeling approaches can be categorized into two types: theoretical and empirical one [39, 41, 42]. In the theoretical approach, the spatial dimensions of the models range from simple 0 (0-D) [43, 44] and 1 (1-D) [42, 45, 46, 47], to more complicated 2 (2-D) [48, 49, 50, 51] and 3 (3-D) [52, 53, 54], all with various characteristics and directed at different objectives. The mathematical models, which are based on conservation principles, require a lot of data on parameters and properties of fuel cell, as well as complicated equations and time-consuming calculation.

Empirical or data-driven approach may be more feasible for fuel cell users since the behavior can be quickly and simply deduced without a comprehensive understanding of the internal components, just based on the experimental data [39, 41]. Least squares support vector machine (LS-SVM) [55], Hammerstein model [56, 57] are examples of these approaches. In this approach, artificial neural network (ANN) shows several advantages, including high nonlinearity, rapid computation, a low degree of error in matching experimental data. Using ANNs to model SOFCs appears to be a very promising method.

In this section, an ANN was used to model the performance of the BSCF/GDC-based cathode SOFC. The cell voltage was predicted from cathode sintering temperature, cell operating temperature, and cell current. Several network architectures were examined to find the best structure, and the network was trained using back-propagation methods. The data for training, validation, and testing were taken from our study [58]. The genetic algorithm and the developed ANN were then used to find the best conditions for achieving maximum power.

4.1 Artificial neural network models

Artificial neural networks (ANNs), which were analogous to biological nervous systems, consist of interconnected nodes known as neurons to receive and transfer data [59]. The most basic form, feed-forward architecture, is made up of an input layer, one hidden layers, and an output layer. The input and output layers have the same number of neurons as the number of inputs and outputs in the system to be modeled. Weighted connections connect each neuron to every other neuron in the next layer. In any layer except the input, the weighted sum of data from the previous layer is the input of a neuron. The neuron then activates the data using a function and transfers the response to all neurons in the next layer. The size of the hidden layers is a significant factor that affects the estimation precision because it can make the network become insufficient or overfitting [60]. The number of neurons in hidden layer is generally determined through trials. Figure 5 illustrates a 3–5-1 feed forward artificial neural network with operating temperature, sintering temperature, and current as inputs.

Figure 5.

Artificial neural network (3–5-1) structure.

The activation function employed in this model is the logistic sigmoidfx=1/1+ex.

The input data (xi) are scaled to normalized value (xnorm) to enhance the performance of the network:

xnorm=0.8xxminxmaxxmin+0.1E40

in which xmax and xmin are the bounded interval of the experimental data.

To assess the performance of ANN, the mean squared error (MSE) and coefficient of determination (R2) are usually used [61].

Various factors affect the performance of fuel cells such as cathode and anode structure, electrolyte material and thickness, cell temperature, inlet and outlet gas compositions. Two important factors, cathode sintered temperature and cell operating temperature, were considered in this model. The sintered temperature is from 1000–1050°C, whereas the operating temperature ranges from 625–700°C. The sintered temperature affects the structure of the obtained cathode as reported in [58]. The explanation of the range for the investigated parameters can be found in [62].

An ANN with one input layer, one hidden layer, and one single output layer was proposed. Current density, sintered temperature of the cathode, and cell operating temperature are the inputs. Back-propagation algorithm [63] was used to train the network. The maximum number of iteration and minimum performance gradient were set to 400 and 10−5, respectively, to stop the training. The proper network structure is determined through a series of trial tests. The data were split into three subsets at random: training, validation, and test, each containing 70, 20, 10% of the total samples, respectively. The validation and test sets are necessary for evaluating the validation and power of the networks.

The parameters of the neural network were saved and utilized in the next stage to optimize the power density using genetic algorithms.

4.2 Optimization by genetic algorithm

The objective function is the power density of the fuel cell

P=I×VE41

where P is the power density (mW.cm−2), I is the current density (mA.cm−2) and is the input to the ANN model, and V is the cell voltage (V), which is calculated from the model.

The design variables and their corresponding ranges are summarized as follows:

  • sintered temperature of the cathode, [1000–1050] (°C).

  • operating temperature of the cell, [625–700] (°C).

  • electric current of the cell, [0–1500] (mA.cm−2)

The parameters of GA as population size, mutation probability values were set to be 100 and 0.10, respectively. The survival of the individuals was decided by roulette wheel with elitism. The number of generations was 500.

4.3 Optimization results

Figure 6 depicts the fitness values (maximum and mean) of the population versus generation. As indicated in the figure, after about 20 generations, the value of fitness function attained to a maximum value and then remained unchanged. After 100 generations, the maximum fuel cell power density of 451.64 mW/cm2 could be achieved at the sintered temperature of 1005°C, operating temperature of 668°C, and current density of 777 mA/cm2.

Figure 6.

The fitness values versus generation.

Advertisement

5. Conclusions

The application of genetic algorithm in chemical engineering processes has been illustrated by three case studies. The results suggest that the optimum conditions of complex chemical problems can be easily obtained using genetic algorithm. The successes of genetic algorithm for the challenging problems reported herein, the development of many faster and flexible versions of GA, the improvement of computing ability all suggest the continually increasing impact of metaheuristic methods in chemical engineering systems.

References

  1. 1. Goldberg DE. Genetic Algorithms in Search, Optimization, and Machine Learning. Boston, MA, United States: Addison-Wesley; 1989
  2. 2. Deb K. Introduction to genetic algorithms for engineering optimization. In: Onwubolu GC, Babu BV, editors. New Optimization Techniques in Engineering. Berlin, Heidelberg: Springer Berlin Heidelberg; 2004. pp. 13-51
  3. 3. Nielsen A, Aika K, Christiansen LJ, Dybkjaer I, Hansen JB, Nielsen PEH, et al. Ammonia: Catalysis and Manufacture. Berlin Heidelberg: Springer; 1995
  4. 4. Ammonia AM. Priciples & Industrial Practice. Weinheim, Germany: Wiley; 1999
  5. 5. Babu B, Angira R, Nilekar A. Optimal design of an auto-thermal ammonia synthesis reactor using differential evolution. In: Proceedings of the Eighth World Multi-Conference on Systemics, Cybernetics and Informatics (SCI-2004). Orlando, Florida, USA: International Institute of Informatics and Systemics; 2004
  6. 6. Babu BV, Angira R. Optimal design of an auto-thermal ammonia synthesis reactor. Computers & Chemical Engineering. 2005;29(5):1041-1045. DOI: 10.1016/j.compchemeng.2004.11.010
  7. 7. Carvalho EP, Borges C, Andrade D, Yuan JY, Ravagnani MASS. Modeling and optimization of an ammonia reactor using a penalty-like method. Applied Mathematics and Computation. 2014;237:330-339. DOI: 10.1016/j.amc.2014.03.099
  8. 8. Edgar TF, Himmelblau DM, Lasdon LS. Optimization of Chemical Processes. New York, NY, USA: McGraw-Hill; 2001
  9. 9. Ksasy M, Areed F, Saraya S, Khalik MA. Optimal reactor length of an auto-thermal ammonia synthesis reactor. IJECS: International Journal of Electrical and Computer Sciences. 2010;10(3):6-11
  10. 10. Murase A, Roberts HL, Converse AO. Optimal thermal Design of an Autothermal Ammonia Synthesis Reactor. Industrial & Engineering Chemistry Process Design and Development. 1970;9(4):503-513. DOI: 10.1021/i260036a003
  11. 11. Upreti SR, Deb K. Optimal design of an ammonia synthesis reactor using genetic algorithms. Computers & Chemical Engineering. 1997;21(1):87-92. DOI: 10.1016/0098-1354(95)00251-0
  12. 12. Yusup S, Zabiri H, Yusoff N, Yew YC. Modeling and Optimization of Ammonia Reactor Using Shooting Methods. Bucharest, Romania: Proceedings of the 5th WSEAS international conference on Data networks, communications and computers, World Scientific and Engineering Academy and Society (WSEAS); 2006. pp. 258-268
  13. 13. Nguyen TAN, Nguyen TA, Vu TD, Nguyen KT, Dao TKT, Huynh KPH. Optimization of an auto-thermal ammonia synthesis reactor using cyclic coordinate method. IOP Conference Series: Materials Science and Engineering. 2017;206:012059. DOI: 10.1088/1757-899x/206/1/012059
  14. 14. Mores PL, Arias AM, Scenna NJ, Caballero JA, Mussati SF, Mussati MC. Membrane-based processes: Optimization of hydrogen separation by minimization of power, membrane area, and cost. Processes. 2018;6(11):221. DOI:10.3390/pr6110221
  15. 15. Ramírez-Santos ÁA, Bozorg M, Addis B, Piccialli V, Castel C, Favre E. Optimization of multistage membrane gas separation processes. Example of application to CO2 capture from blast furnace gas. Journal of Membrane Science. 2018;566:346-366. DOI: 10.1016/j.memsci.2018.08.024
  16. 16. Ohs B, Lohaus J, Wessling M. Optimization of membrane based nitrogen removal from natural gas. Journal of Membrane Science. 2016;498:291-301. DOI: 10.1016/j.memsci.2015.10.007
  17. 17. Wiley DE, Fell CJD, Fane AG. Optimisation of membrane module design for brackish water desalination. Desalination. 1985;52(3):249-265. DOI: 10.1016/0011-9164(85)80036-9
  18. 18. Nguyen T-A, Yoshikawa S, Karasu K, Ookawara S. A simple combination model for filtrate flux in cross-flow ultrafiltration of protein suspension. Journal of Membrane Science. 2012;403-404:84-93. DOI: 10.1016/j.memsci.2012.02.026
  19. 19. Nguyen T-A, Yoshikawa S, Ookawara S. Steady state permeate flux estimation in cross-flow ultrafiltration of protein solution. Separation Science and Technology. 2014;49(10):1469-1478. DOI: 10.1080/01496395.2014.893533
  20. 20. Sethi S, Wiesner MR. Performance and cost modeling of ultrafiltration. Journal of Environmental Engineering. 1995;121(12):874-883. DOI: 10.1061/(ASCE)0733-9372(1995)121:12(874)
  21. 21. Cheryan M. Ultrafiltration and Microfiltration Handbook. Boca Raton, FL, USA: Taylor & Francis; 1998
  22. 22. Zeman LJ, Zydney AL. Microfiltration and Ultrafiltration: Principles and Applications. New York, NY, USA: CRC Press; 1996
  23. 23. Karasu K. A Study on Permeation Phenomena in Cross-Flow Ultrafiltration Producing a Compressible Cake Layer. Tokyo, Japan: Tokyo Institute of Technology; 2010
  24. 24. White FM. Fluid Mechanics. 8th ed. New York, NY, USA: McGraw-Hill Education; 2016
  25. 25. Fox RW, McDonald AT, Mitchell JW. Fox and McDonald’s Introduction to Fluid Mechanics. Hoboken, NJ, USA: Wiley; 2020
  26. 26. Nguyen T-A, Yoshikawa S. Modeling and economic optimization of the membrane module for ultrafiltration of protein solution using a genetic algorithm. Processes. 2020;8(1):4. DOI: 10.3390/pr8010004
  27. 27. Dorfman KD, Daoutidis P. Numerical Methods with Chemical Engineering Applications. Cambridge, United Kingdom: Cambridge University Press; 2017
  28. 28. U.S. Energy Information Administration. Average Retail Price of Electricity to Ultimate Customers Total End-Use Sector. U.S., USA: Department of Energy; 2019
  29. 29. Park CS. Fundamentals of Engineering Economics. London, United Kingdom: Pearson Education; 2013
  30. 30. Turton R, Bailie RC, Whiting WB, Shaeiwitz JA. Analysis, Synthesis and Design of Chemical Processes. London, United Kingdom: Pearson Education; 2008
  31. 31. Perry RH, Perry RH, Chilton CH, Perry JH. Chemical Engineers' Handbook. 5th ed. New York, NY, USA: McGraw-Hill; 1973
  32. 32. Green DW, Perry RH. Perry’s Chemical Engineers' Handbook. Eighth ed. New York, NY, USA: McGraw-Hill Education; 2007
  33. 33. Holland FA, Wilkinson JK. Process economics. In: Perry RH, Green DW, editors. Perry’s Chemical Engineers’ Handbook. 7th ed. New York, NY, USA: McGraw-Hill; 1997
  34. 34. Mir L, Michaels SL, Goel V, Kaiser R. Crossflow Microfiltration: Applications, Design, and Cost. In: Ho WSW, Sirkar KK, editors. Membrane handbook: Newyork, NY, USA: Springer; 1992. pp. 571-594
  35. 35. Wachsman ED, Marlowe CA, Lee KT. Role of solid oxide fuel cells in a balanced energy strategy. Energy & Environmental Science. 2012;5(2):5498-5509. DOI: 10.1039/C1EE02445K
  36. 36. Ghezel-Ayagh H, Borglum BP. (invited) review of Progress in solid oxide fuel cell at FuelCell energy. ECS Transactions. 2017;80(9):47-56. DOI: 10.1149/08009.0047ecst
  37. 37. Minh N, Mizusaki J, Singhal SC. Advances in solid oxide fuel cells: Review of Progress through three decades of the international symposia on solid oxide fuel cells. ECS Transactions. 2017;78(1):63-73. DOI: 10.1149/07801.0063ecst
  38. 38. Kakaç S, Pramuanjaroenkij A, Zhou XY. A review of numerical modeling of solid oxide fuel cells. International Journal of Hydrogen Energy. 2007;32(7):761-786. DOI: 10.1016/j.ijhydene.2006.11.028
  39. 39. Hajimolana SA, Hussain MA, Daud WMAW, Soroush M, Shamiri A. Mathematical modeling of solid oxide fuel cells: A review. Renewable and Sustainable Energy Reviews. 2011;15(4):1893-1917. DOI: 10.1016/j.rser.2010.12.011
  40. 40. Ma L, Ingham DB, Pourkashanian M, Carcadea E. Review of the computational fluid dynamics modeling of fuel cells. Journal of Fuel Cell Science and Technology. 2005;2(4):246-257. DOI: 10.1115/1.2039958
  41. 41. Wang K, Hissel D, Péra MC, Steiner N, Marra D, Sorrentino M, et al. A review on solid oxide fuel cell models. International Journal of Hydrogen Energy. 2011;36(12):7212-7228. DOI: 10.1016/j.ijhydene.2011.03.051
  42. 42. Karcz M. From 0D to 1D modeling of tubular solid oxide fuel cell. Energy Conversion and Management. 2009;50(9):2307-2315. DOI: 10.1016/j.enconman.2009.05.007
  43. 43. Costamagna P, Magistri L, Massardo AF. Design and part-load performance of a hybrid system based on a solid oxide fuel cell reactor and a micro gas turbine. Journal of Power Sources. 2001;96(2):352-368. DOI: 10.1016/S0378-7753(00)00668-6
  44. 44. Zabihian F, Fung AS. Macro-level modeling of solid oxide fuel cells, approaches, and assumptions revisited. Journal of Renewable and Sustainable Energy. 2017;9(5):054301. DOI: 10.1063/1.5006909
  45. 45. Ota T, Koyama M, Wen C-j, Yamada K, Takahashi H. Object-based modeling of SOFC system: Dynamic behavior of micro-tube SOFC. Journal of Power Sources. 2003;118(1):430-439. DOI: 10.1016/S0378-7753(03)00109-5
  46. 46. Li P-W, Suzuki K. Numerical modeling and performance study of a tubular SOFC. Journal of The Electrochemical Society. 2004;151(4):A548. DOI: 10.1149/1.1647569
  47. 47. Bove R, Lunghi P, M. Sammes N. SOFC mathematic model for systems simulations—Part 2: Definition of an analytical model. International Journal of Hydrogen Energy. 2005;30(2):189-200. DOI: 10.1016/j.ijhydene.2004.04.018
  48. 48. Ma R, Gao F, Breaz E, Huangfu Y, Briois P. Multidimensional reversible solid oxide fuel cell modeling for embedded applications. IEEE Transactions on Energy Conversion. 2018;33(2):692-701. DOI: 10.1109/TEC.2017.2762962
  49. 49. Ni M. 2D thermal-fluid modeling and parametric analysis of a planar solid oxide fuel cell. Energy Conversion and Management. 2010;51(4):714-721. DOI: 10.1016/j.enconman.2009.10.028
  50. 50. Geisler H, Dierickx S, Weber A, Ivers-Tiffee E. A 2D stationary FEM model for hydrocarbon Fuelled SOFC stack layers. ECS Transactions. 2015;68(1):2151-2158. DOI: 10.1149/06801.2151ecst
  51. 51. Luo XJ, Fong KF. Development of 2D dynamic model for hydrogen-fed and methane-fed solid oxide fuel cells. Journal of Power Sources. 2016;328:91-104. DOI: 10.1016/j.jpowsour.2016.08.005
  52. 52. Nikooyeh K, Jeje AA, Hill JM. 3D modeling of anode-supported planar SOFC with internal reforming of methane. Journal of Power Sources. 2007;171(2):601-609. DOI: 10.1016/j.jpowsour.2007.07.003
  53. 53. Andersson M, Paradis H, Yuan J, Sundén B. Three dimensional modeling of an solid oxide fuel cell coupling charge transfer phenomena with transport processes and heat generation. Electrochimica Acta. 2013;109:881-893. DOI: 10.1016/j.electacta.2013.08.018
  54. 54. Yang C, Yang G, Yue D, Yuan J, Sunden B. Computational fluid dynamics model development on transport phenomena coupling with reactions in intermediate temperature solid oxide fuel cells. Journal of Renewable and Sustainable Energy. 2013;5(2):021420. DOI: 10.1063/1.4798789
  55. 55. Huo H-B, Zhu X-J, Cao G-Y. Nonlinear modeling of a SOFC stack based on a least squares support vector machine. Journal of Power Sources. 2006;162(2):1220-1225. DOI: 10.1016/j.jpowsour.2006.07.031
  56. 56. Huo H-B, Zhong Z-D, Zhu X-J, Tu H-Y. Nonlinear dynamic modeling for a SOFC stack by using a Hammerstein model. Journal of Power Sources. 2008;175(1):441-446. DOI: 10.1016/j.jpowsour.2007.09.059
  57. 57. Jurado F. A method for the identification of solid oxide fuel cells using a Hammerstein model. Journal of Power Sources. 2006;154(1):145-152. DOI: 10.1016/j.jpowsour.2005.04.005
  58. 58. Le M-V, Tsai D-S, Nguyen T-A. BSCF/GDC as a refined cathode to the single-chamber solid oxide fuel cell based on a LAMOX electrolyte. Ceramics International. 2018;44(2):1726-1730. DOI: 10.1016/j.ceramint.2017.10.103
  59. 59. Baughman DR, Liu YA. Neural Networks in Bioprocessing and Chemical Engineering. Amsterdam, Netherlands: Elsevier Science; 1995
  60. 60. Sheela KG, Deepa SN. Review on methods to fix number of hidden neurons in neural networks. Mathematical Problems in Engineering. 2013;2013:425740. DOI: 10.1155/2013/425740
  61. 61. Himmelblau DM. Accounts of experiences in the application of artificial neural networks in chemical engineering. Industrial & Engineering Chemistry Research. 2008;47(16):5782-5796. DOI: 10.1021/ie800076s
  62. 62. Le M-V, Nguyen T-A, Nguyen TA-N. Modeling and optimization of the BSCF-based single-chamber solid oxide fuel cell by artificial neural network and genetic algorithm. Journal of Chemistry. 2019;2019:7828019. DOI: 10.1155/2019/7828019
  63. 63. Wythoff BJ. Backpropagation neural networks: A tutorial. Chemometrics and Intelligent Laboratory Systems. 1993;18(2):115-155. DOI: 10.1016/0169-7439(93)80052-J

Written By

Thi Anh-Nga Nguyen and Tuan-Anh Nguyen

Submitted: 20 December 2021 Reviewed: 11 April 2022 Published: 19 May 2022