Next Article in Journal
Temperature-Related N2O Emission and Emission Potential of Freshwater Sediment
Next Article in Special Issue
Integrated Optimization for the Coupling Network of Refinery and Synthetic Plant of Chemicals
Previous Article in Journal
Dynamics of Pressure Variation in Closed Vessel Explosions of Diluted Fuel/Oxidant Mixtures
Previous Article in Special Issue
Life Cycle Energy Consumption and GHG Emissions of the Copper Production in China and the Influence of Main Factors on the above Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Agent-Based and Stochastic Optimization Incorporated with Machine Learning for Simulation of Postcombustion CO2 Capture Process

1
Department of Energy and Process Engineering, NTNU—Norwegian University of Science and Technology, Kolbjørn Hejes vei 1 B, 7491 Trondheim, Norway
2
Department of Electrical Engineering, Information Technology and Cybernetics, USN—University of South-Eastern Norway, Campus Porsgrunn, 3199 Borre, Norway
*
Author to whom correspondence should be addressed.
Processes 2022, 10(12), 2727; https://doi.org/10.3390/pr10122727
Submission received: 24 October 2022 / Revised: 5 December 2022 / Accepted: 12 December 2022 / Published: 16 December 2022
(This article belongs to the Special Issue Research on Process System Engineering)

Abstract

:
In this paper, a novel method is proposed for the incorporation of data-driven machine learning techniques into process optimization. Such integration improves the computational time required for calculations during optimization and benefits the online application of advanced control algorithms. The proposed method is illustrated via the chemical absorption-based postcombustion CO2 capture process, which plays an important role in the reduction of CO2 emissions to address climate challenges. These processes simulated in a software environment are typically based on first-principle models and calculate physical properties from basic physical quantities such as mass and temperature. Employing first-principle models usually requires a long computation time, making process optimization and control challenging. To overcome this challenge, in this study, machine learning algorithms are used to simulate the postcombustion CO2 capture process. The extreme gradient boosting (XGBoost) and support vector regression (SVR) algorithms are employed to build models for prediction of carbon capture rate (CR) and specific reboiler duty (SRD). The R2 (a statistical measure that represents the fitness) of these models is, on average, greater than 90% for all the cases. XGBoost and SVR take 0.022 and 0.317 s, respectively, to predict CR and SRD of 1318 cases, whereas the first-principal process simulation model needs 3.15 s to calculate one case. The models built by XGBoost are employed in the optimization methods, such as an agent-based approach represented by the particle swarm optimization and stochastic technique indicated by the simulated annealing, to find specific optimal operating conditions. The most economical case, in which the CR is 72.2% and SRD is 4.3 MJ/kg, is obtained during optimization. The results show that computations with the data-driven models incorporated in the optimization technique are faster than first-principle modeling approaches. Thus, the application of machine learning techniques in the optimization of carbon capture technologies is demonstrated successfully.

1. Introduction

According to the World Meteorological Organization (WMO), the world experienced the warmest weather on record because of greenhouse gas emissions (GHG) in the last decade (2011–2020). The temperature difference between the preindustrial baseline (1850–1900) and global mean surface temperature in 2020 has reached 1.2 °C [1]. The action is on call for addressing climate changes as reported in this report. To avoid the worst climate impact, carbon capture and storage (CCS) technologies have been developed. CCS is the process of capturing CO2 before it enters the atmosphere, transporting and storing it for centuries or millennia [2]. Currently, the most mature technology is CO2 capture, which can be realized through precombustion capture, oxy-fuel combustion capture, and postcombustion capture (PCC) [3]. From an economic and practical perspective, PCC is preferred, as it can be retrofitted to an existing plant. Many research efforts now focus on topics within the PCC process, such as process simulation, solvent degradation, and equipment design. This study also contributes to the simulation of PCC process in order to better control and optimize the process efficiency. This technology is a process that consists of the chemical absorption and thermal-stripping route. The CO2 in the flue gas stream is captured by an absorbent solvent, such as monoethanolamine (MEA) and diethanolamine (DEA) in the chemical absorption process. In the stripping process, the CO2-loaded solvent is sent to a stripper, the pure CO2 is released, and the absorbent solvent is regenerated [4].
The PCC process studied in this paper is with 30 wt% MEA chemical solvent because it is regarded as the benchmark solvent in academic work and industrial research. In order to study the process behavior of the PCC process, the first-principal model simulation is widely used. The process simulation software, such as Aspen HYSYS®, Aspen Plus®, gPROMS® and CO2SIM® are used to build models to calculate the mass flow rate of stream, temperature, pressure, and other physical quantities in PCC for analytical purposes [5,6,7,8]. Øi investigated the difference between equilibrium and rate-based models built by Aspen HYSYS® and Aspen Plus®. It is reported that the rate-based models in software are close to the equilibrium model when calculating CO2-removal efficiency as a function of circulation rate, number of column stages, and inlet temperature [9]. In general, PCC process simulations are developed by using first-principle models and compared for the accuracy of these models with real plants. In addition, some studies improve the accuracy of simulation models. To improve a simulation model’s predictions for experimental and pilot results, Spek et al. upgraded the software package by using the regression of activated 2-amino-2-methyl-1-propanol(AMP)–piperazine (PZ) binary interaction parameters [10]. Luo et al. built a steady-state model, which is rate-based for the MEA-based PCC process. This model is validated against thermodynamic and physical properties over a wide range of pressures, temperatures, and CO2 loadings [11]. Dutta et al. built an equilibrium-based absorber model by using the reduced stage efficiencies in order to reduce the computational time for simulation. This model is used to predict operating conditions within an accepted range [12]. Nonetheless, the implementation of further optimization and control based on first-principle simulation model is challenged as the time required for solving differential and algebraic equations in this simulation model is approximately 4 s for one case. Therefore, the alternative modeling method must be investigated and machine learning turned out to be a promising one. Fernandez et al. has proven that the machine learning techniques can be used to recognize high-performing metal-organic framework materials in the CO2 capture process [13]. Venkatraman et al. used the machine learning technique to model structure–property relationships between molecular structures of cations and anions and their CO2 solubilities in the ionic liquid (IL)-based CO2 capture process [14]. Ahmadi et al. investigated the possibility of using an artificial neural network (ANN) and least squares support vector machine (LSSVM) approaches to predict the viscosity and thermal conductivity of CO2. Both methods are proven to be effective [15]. In order to change the CO2 capture rate quickly and smoothly in a wide operating range, Wu et al. performed a nonlinearity analysis of the multimodel predictive control strategy for solvent-based postcombustion CO2 capture plants [16]. Rahimi et al. discussed how machine learning implementations have improved the carbon capture process in absorption- and adsorption-based approaches, ranging from the molecular to the process level. In particular, the application of reinforcement learning to obtain the optimized operating parameters is highlighted [17]. Shalaby et al. applied Matérn Gaussian process regression (GPR), rational quadratic GPR, squared exponential GPR models, and a feed-forward ANN model in predicting the output parameters of the PCC process simulation in gPROMS®. The system’s energy requirement, capture rate, and the purity of the condenser outlet stream are accurately predicted [18]. Shahsavand et al. applied the backpropagation multilayer perceptron (BPMLP) and the radial basis function (RBF) neural networks to understand packed absorption processes, especially the absorption of CO2 from the air by various alkanolamine solutions [19]. The result showed that the RBF networks perform more adequately than the MLP network for filtering noise and capturing the real trend. Chan et al. used the neural network rule extraction algorithm to reveal that the steam flow rate through reboiler, reboiler pressure, and the CO2 concentration in the flue gas are the three most significant parameters of the CO2 production rate [20]. Li et al. predicted the CO2 capture rate by applying the bootstrap aggregated extreme learning machine (ELM) to build models based on the data collected from the gPROMS® simulation model. The ELM model performs well as it has a small mean square error (MSE) and reduces computational time [21]. The possibility of the application of machine learning technique has been demonstrated through research in the literature. However, the extreme gradient boosting (XGBoost) and support vector regression (SVR) have not been investigated yet. This study aims to fill this research gap of application of new algorithms. A mature machine learning-based modeling method is introduced in this paper. Additionally, these advanced machine learning techniques are applied in the context of process optimization. In this study, the process modeling approach is data driven with the application of new algorithms (XGBoost and SVR) that improve the prediction accuracy of PCC process parameters. In addition, the faster computational time of these data-driven techniques to model the PCC process is advantageous during optimization and control studies.
In the literature, the application of data-driven models for the optimization and control of the energy systems has been investigated [22,23]. Mirlekar et al. employed the autoregressive model with the exogenous inputs (ARX) method in the development of biologically inspired optimal control strategy (BIOCS) for implementation on a subsystem of a CO2 capture process [23]. Saboori et al. studied the application of particle swarm optimization (PSO) in a multistage generation expansion planning (GEP) which includes nuclear units, renewable energy units, and different fossil fuel-fired units equipped with CCS [24]. Amar et al. used the genetic algorithm (GA), PSO, and artificial bee colony (ABC) to optimize the radial basis function neural network (RBFNN) models employed for modeling of CO2 solubility in brine [25]. Simulated annealing (SA) has also been investigated in the study of the PCC process. SA is an optimization method based on the analogy between the simulation of the annealing of solids and solving large combinatorial optimization problems [26]. Mores et al. developed an optimized radial basis function based on grid search coupled with SA and tenfold cross-validation algorithms. This intelligent correlation model was used to predict the mass transfer coefficient for CO2 capture with NaOH solution in different types of rotating packed beds [27]. Dashti et al. presented unique computational models to estimate CO2 solubility in commonly used amines. A series of models, including the genetic algorithm-adaptive neurofuzzy inference system (GA-ANFIS), PSO-ANFIS, coupled simulated annealing-least squares support vector machine (CSA-LSSVM) and RBF neural networks were developed to estimate CO2 equilibrium absorption capacity in 12 aqueous amine solutions [28]. In this paper, a systematic method is proposed to study the application of the machine learning technique for reducing computational time in the simulation of the PCC process. The PCC process simulation model is developed in Aspen HYSYS® in order to generate original data for the study. The characteristics of the datasets are summarized and analyzed before the application of machine learning. XGBoost and SVR algorithms are employed for building data-driven models with the objective of predicting the CO2 CR and specific reboiler duty (SRD) in the PCC process. These models are further incorporated into optimization techniques represented by PSO and SA. Thus, the optimal process operating conditions are obtained. In general, the SRD increases with the CR, whereas the cost-effective situation deals with higher CR and lower SRD. Therefore, the optimized operating condition would be the one with relatively high CR and low SRD. Such an integration would help in improving online advanced control implementation and dynamic optimization of the process. The paper is organized as follows. The method is described in Section 2; the implementation results are discussed in Section 3; and in Section 4, the summary and conclusions are presented.

2. Method

In this section, the PCC process simulation, the machine learning algorithms incorporated into the optimization methods and the optimization techniques are described. The workflow diagram considered in this study is shown in Figure 1. The large amount of data required for machine learning from a pilot plant is unavailable. To overcome this challenge, a steady-state process simulation model of the PCC is set up in Aspen HYSYS® [12]. This model is used for generating the raw data for data-driven modeling purposes. These data are summarized and analyzed, and then the XGBoost and SVR algorithms are applied to develop data-driven models. The assessment indicators, such as coefficient of determination (R2) and MSE, are used to measure model performance. In the model validation step, a new dataset is generated specifically for validation purposes. The data-driven models are further used in optimization for searching the optimal operating conditions, such as the largest CR or lowest SRD. The PSO and SA are adopted to perform optimization.

2.1. PCC Process Simulation Model

The steady-state simulation model of the PCC process used in this study is shown in Figure 2. There are two main process operations in the PCC. One is the absorption chemical process with the solvent, and the other one is the desorption of CO2 from the loaded solvent and regeneration of the lean solvent. The flue gas stream is absorbed by the lean amine stream in the absorber column. The depleted flue gas, which does not contain CO2, is then released from the top of the column into the atmosphere and the CO2-loaded MEA stream (shown as the rich MEA stream in Figure 2) leaves the absorber from the bottom, followed by a heat exchanger and a stripper column. In the stripper, the stripping vapor goes up and condenses at the cooler and condenser on the top. The condensate containing the regenerated solvent is recycled back to the stripper. The CO2 separated (shown as CO2 for compression in Figure 2) is sent to the compressor and further processed for transportation or storage. Lean amine is recycled back to the absorber. This lean amine stream heats the CO2-loaded MEA stream from the absorber in the heat exchanger.
Flue gas mass flow rate, lean amine mass flow rate, CO2 molar fraction in flue gas, lean amine loading, and rich amine loading are initially selected as independent variables. The control variable method is used for data generation. Each variable is selected as the control variable, and its value is changed in a stepwise manner while the other variables are held constant. The capture rate and specific reboiler duty of simulated cases are recorded and collected for further analysis and application. A base case, the main parameters of which are calibrated with industrial data, are listed in Table 1 and Table 2 [29]. Here, lean amine loading is referred to the mole ratio of acid gas to the solvent from the bottom of the stripping column which is lean in acid gas. Rich amine loading is referred to the mole ratio of acid gas to the solvent from the bottom of the absorber column which is rich in acid gas. The reboiler duty is the amount of heat required to regenerate solvent.
The main components in the lean amine stream are H2O, CO2 and MEA. In the flue gas stream, the main compositions are H2O, CO2 and N2. Mass fractions of lean amine stream and molar fraction of compositions in the flue gas are listed in Table 3. A script in Python 3.8 connected to the process simulation model in the Aspen HYSYS® file is specifically written for the implementation of changing values of controlled variables. The step size, upper and lower limits, and number of collected data samples for each variable are listed in Table 4. In total, 4393 cases are simulated to generate the raw data by controlling and changing values of different variables. In the next subsection, the machine learning techniques employed to develop data-driven PCC process models are discussed.

2.2. Machine Learning Techniques

2.2.1. Extreme Gradient Boosting

XGBoost, also known as extreme gradient boosting, was launched in 2016. It is a decision tree-based ensemble machine learning algorithm that uses a gradient boosting framework [30]. The evolution route of the XGBoost algorithm is shown in Figure 3, and the instruction starts from the decision tree.
Decision tree is a nonparametric supervised learning method used for classification and regression. A decision tree usually consists of three main parts: the root node, decision nodes, and leaf nodes. The root node is a point at which to start the model. The decision node is a judgment node, a conditional judgement; for example, “lean amine loading is above 0.3 or not”, and “flue gas flow rate is larger than 0.38 kg/s or smaller than 0.25 kg/s” exist in a decision node. The different answer to the judgment node leads to different nodes in the next level. The calculation is perfomed to cross-decision nodes until it reaches the leaf node, the last level of the decision tree. The leaf node is the result of the calculation.
Bagging, short for “bootstrap aggregating”, is an ensemble learning method. In this method, a certain number of samples are randomly picked from the sample pool with replacement. After multiple selections, (for regression problems) the average of the multiple results, (for classification problems) or the result most voted, is the final result. A sample can be chosen more than one time in the bagging method. Bagging can effectively reduce variance in noisy datasets.
Random forest is a bagging-based algorithm. As known from previous decision tree introduction, characteristics are chosen to be the judgement base in the decision node. In random forest, bagging is used to choose characteristics to form decision trees. All these trees come together to be the “forest”, and the final result comes from the average or the most voted of all tree results.
Boosting is a family of algorithms that convert weak learners into strong learners. At first, training data are used to build a basic learner, and then the wrong prediction samples are focused. The errors are corrected and used to build a stronger learner. This “correct” step is repeated until a model emerges that has an error within acceptable range.
Gradient boosting, which is also called gradient boosting or gradient tree boosting, was developed by Friedman [31]. Three basic elements, a loss function to be optimized, a weak learner to make predictions, and an additive model to add weak learners to minimize the loss function, are included in the gradient-boosting algorithm. Squared error is used as loss function for regression problem and logarithmic loss is used for classification problems. In gradient boosting, the decision tree is the weak learner. A decision tree could be constrained in different ways, including maximum number of layers, nodes, and leaf nodes. Each time, a new decision tree is added for minimizing the loss function and the existing trees remain unchanged.
XGBoost is a decision-tree-based ensemble machine learning algorithm. It takes the bootstrap sample 1 to build model 1, then takes the bootstrap sample 2 to build model 2, which leads a smaller regularized objective function than model 1. Then it iterates until the final model and result are produced. The regularized objective function is
L ( ϕ ) = i l ( y ^ i , y i ) + k Ω ( f k ) ,
where Ω ( f ) = γ T + 1 2 λ w 2 . L ( ϕ ) is the objective in minimization problem. It is the summation of convex loss function represented as i l ( y ^ i , y i ) and regularization item denoted as k Ω ( f k ) . Here, y i and y ^ i stand for targeted value and predicted value, respectively. λ and γ are hyperparameter constants. T is the number of leaf nodes. w is the predicted value of the leaf node. The data from the PCC process simulation is used in this machine learning method. The tasks of predicting capture rate and SRD are regression problems solvable by the XGBoost algorithm.

2.2.2. Support Vector Regression

Support vector machine (SVM) is commonly used for solving classification problems. In a few regression applications, SVR is used. In an SVR model, a line or a hyperplane (when the data has higher dimensions) is searched to fit the data. The users have the freedom to decide the acceptable error. In a two-dimensional space, assume the line to fit the data is
| Y i α X i | ε .
For a classical linear regression, the objective is to minimize the squared error of original values and predicted values. In the SVR model, the objective (f) is to minimize the coefficients, the l 2 -norm of the coefficient vector, which is shown as follows [32]:
f = M I N ( 1 2 α 2 + i = 1 n C | ξ i | ) .
Here, Yi stands for the targeted value, Xi represents the feature value, and α denotes coefficients. ξ is the differences between y value of outliers to the closest hyperplane. In SVR models, kernel functions are used to transform input data to the required form of processing data. There are different kinds of kernels: the Gaussian kernel, the radial basis function, the sigmoid kernel, and the polynomial kernel. The advantages of SVR is that its computational complexity does not depend on the dimensionality of the input space. Moreover, it has excellent generalization capability with high prediction accuracy [33]. The SVR algorithm is an alternative way to build models for predicting CR and SRD in the PCC process.

2.2.3. Model Performance Assessment

Three indicators, R2, MSE, and cross-validation score, are used to evaluate the performance of data-driven models:
A.
Coefficient of determination (R2)
R-squared (R2) is a statistical measure that represents how much the variance for a dependent variable can be explained by independent variables in a regression model. In general, a higher R2 indicates a better regression effect. The formula of R2 is
R 2 = 1 S S R S S T = 1 ( y i y ^ i ) 2 ( y i y ¯ ) 2 .
SSR represents the sum squared regression, SST represents the total sum of squares. In Equations (4) and (5), y i and y ^ i have the same meaning as in Equation (1). The y i is real value, y ^ i is the predicted value, y ¯ is the average of the actual y, and q is the number of observations.
B.
Mean Square Error
MSE is the average squared distance between the actual and predicted values. The formula is
M S E = ( y t y ^ t ) 2 q .
C.
Cross-validation score
Cross-validation is a technique for detecting overfitting or assessing the generalized ability of the model. Overfitting means a model fits well and has high accuracy with the training data, while it has low accuracy against the test or unseen data. Generalization refers to a model’s ability to adapt properly to new or unseen data that are drawn from the same distribution as the one used to create the model. A model that is overfitting has a low generalized ability. When cross-validation is used, datasets are separated into several folds evenly. At each turn, one fold is selected as a test fold whereas the others are training folds. This method is useful as it gives the algorithm different subsets of data to train on. In this study, the cross-validation score is the average value of R2 of all testing folds. In the next subsection, optimization techniques associated with machine learning incorporation are explained. Specifically, the agent-based and stochastic optimization methods used in this study are illustrated.

2.3. Optimization Algorithms

2.3.1. Particle Swarm Optimization

PSO is an agent-based search optimization technique inspired by the migration behavior of birds proposed by Kennedy et al. [34]. Assume there are N particles subject to random initialization in a swarm. These particles have random positions and velocities in a D-dimensional search space, and they move at a certain speed toward the objective location inside the whole space. For each particle, the new velocity is updated based on its own historical experience and the group general experience. Assume the D-dimensional position vector of the i -th particle is
S i = ( S i 1 , S i 2 , S i 3 , , S i N ) , i = 1 , 2 , , N .
The velocity vector of i-th particle is
V i = ( V i 1 , V i 2 , V i 3 , , V i N ) , i = 1 , 2 , , N .
The equations used to update the position and velocity of each particle are
v i d k + 1 = m v i d k + c 1 r 1 ( p i d k s i d k ) + c 2 r 2 ( p g d k s i d k )
s i d k + 1 = s i d k + r v i d k + 1 .
The equation to update the speed of each particle consists of three parts. The v i d and s i d are the velocity and position of the i-th particle in the d-th dimension, k and k + 1 represent the current and next iterations. The first part, m v i d k , is the exploration, and m is an inertia factor of fixed value. The second part c 1 r 1 ( p i d k s i d k ) represents self-learning, and the third part, c 2 r 2 ( p g d k s g d k ) , represents group learning. c 1 and c 2 are learning factors, r 1 and r 2 are random numbers within the range [0, 1]. p i d and p g d are the most satisfied positions searched by the i-th particle and the whole group. The p i d and p g d are recorded as
P b e s t = ( P i 1 , P i 2 , P i 3 , , P i N ) , i = 1 , 2 , , N
G b e s t = ( G i 1 , G i 2 , G i 3 , , G i N ) , i = 1 , 2 , , N .
The PSO is a heuristic algorithm and available for this case to search for the optimized operating conditions in PCC within the XGBoost models.

2.3.2. Simulated Annealing

SA is a stochastic optimization algorithm based on Monte Carlo iterative solution strategy introduced by S. Kirkpatrick et al. in 1983 [35]. The main idea of this method is to introduce the annealing in solids into traditional optimization problems. A temperature, τ , is set for the control of the whole optimal solution searching process. Assume there is a initial solution ψ for the goal function E ( ψ ) , and random perturbations are added to this solution to generate the new solution. For these two solutions, if the new solution is better than old one, it is accepted. If it is not, it is not to be given up. Instead, Metropolis criteria is used to decide whether the new solution will be accepted or not. In Metropolis criteria, the probability of accepting the new solution is
P = 1 E ( ψ + 1 ) < E ( ψ ) e x p ( ( E ( ψ + 1 ) E ( ψ ) ) τ ) E ( ψ + 1 ) E ( ψ ) .
After deciding to accept the new solution or not, the number of iterations are confirmed. If the end condition is not met, the temperature is adjusted slowly, and a new solution is regenerated. The SA optimization method is versatile as it does not rely on restrictive properties of the models. It is also used for optimization in this study to search for the optimal operating conditions. Note that all the data-driven model development and optimization studies in this paper are carried out by using Python script and subroutines on an Intel Core i7 (Sandy bridge) 2.30 GHz processor. In the next section, the data analysis result, the prediction performance of built models, and the optimization results are shown and discussed.

3. Results

3.1. Data Analysis and Model Development

Data analysis aims at having an overview of the datasets by investigating the relations between selected variables. The Pearson correlation coefficient which measures the linear correlation between variables is used. Given the paired dataset (Z11,Z21),(Z12,Z22),…,(Z1n,Z2n), the formula of Pearson correlation coefficient R Z 1 , Z 2 is
R Z 1 , Z 2 = i = 1 n ( Z 1 i Z ¯ 1 ) ( Z 2 i Z ¯ 2 ) i = 1 n ( Z 1 i Z ¯ 1 ) 2 i = 1 n ( Z 2 i Z ¯ 2 ) 2 ,
where Z ¯ 1 and Z ¯ 2 are the average values of variables Z 1 and Z 2 . Figure 4 shows Pearson correlation coefficients between distinct variables in this study.
In Figure 4, the correlation coefficients of different variables considered in this study are shown in the grid format. The grid color is proportional to the correlation coefficient. The color tends to be greener with smaller values, whereas the blue color corresponds to a larger value. For example, the grids intersected by lean amine flow rate and the flue gas flow rate show the correlation coefficient is −0.17, and the color is light green. This means that these two variables are slightly inverse related. The correlation coefficient between rich amine loading and lean amine loading is 0.62, which shows they have positive linear relationships and that the rich amine loading is influenced by the lean amine loading. Therefore, rich amine loading is not included as an input variable. A general understanding of the datasets is obtained by this Pearson correlation coefficient method. Note that nonlinear relations are not taken into account here.
A detailed analysis showing the relationship between dependent and independent variables is shown in Figure 5. In this section, Depleted flue gas(To atmosphere), CO2 for compression, and Rich MEA refer to the gas stream from the absorber, the captured CO2 stream out from the stripper, and the loaded amine stream out from absorber, as shown in Figure 2, which depicts the simulation model. In Figure 5, the solid line denotes smoothed data and solid dots representing original data obtained from process simulation models. Figure 5a includes the capture rate and the flue gas flow rate of 1543 process simulation cases with a downward trend. This shows that the capture rate and the flue gas flow rate are inversely proportional. The highest point represents a case in which the flue gas flow rate is 0.108 kg/s and the capture rate is 0.874; the lowest one indicates a case in which the flue gas flow rate is 0.262 kg/s and the capture rate is 0.556. The CO2 entering the absorber in a time unit increases with the flue gas mass flow rate. Due to the solvent quantity limitation and increased inflow of CO2, the capture rate is lower, and more CO2 is released into the atmosphere through Depleted flue gas (To atmosphere) stream. Therefore, the capture rate is smaller. Toward the left side of this plot, where the flue gas flow rate is close to 0.262, the capture rate fluctuates. This could be because of the model instability. In Figure 5a, the relation between the SRD and the flue gas flow rate is inversely proportional. Because SRD is the reboiler duty divided by the mass of CO2 in CO2 for compression stream, the SRD decreases as the CO2 in CO2 for compression stream goes up and the reboiler duty goes down. In Figure 5b, the capture rate grows with the lean amine flow rate. This is because the CO2 entering the rich stream increases with the lean amine flow rate. In Figure 5b, the trend of CO2 in the rich MEA stream is upward due to increase in the lean MEA flow rate and CR. Consequently, the reboiler duty is larger as the stream enters the stripper that contains more MEA. The SRD increases in a similar manner. The rise in CO2 molar fraction in flue gas results in the increase in CO2 in the stream that is released into the atmosphere and the Rich MEA stream, but the increment in Depleted flue gas (To atmosphere) is larger while the increment in Rich MEA stream is smaller. This provides an explanation for the lower CR values in this case. The reboiler duty of the stripper is lower, resulting in a lower SRD. Lean amine loading is changed with the mass flow rate of CO2 in the lean amine. Because the lean amine loading is larger, CO2 in the Depleted flue gas (To atmosphere) stream increases, and the capture rate decreases. Reboiler duty and SRD are lower accordingly. From Figure 5, it is observed that the SRD fluctuates due to the instability in the reboiler duty. In general, data smoothing is a way to remove the fluctuations or disturbances. The data smoothing is implemented via methods such as simple exponential, moving average, exponential moving average, and Holt–Winters smoothing methods. However, the expected improvement is not observed while using the smoothed data to develop the model. For the consideration of keeping original information, the following models are developed by using the raw data without smoothing.

3.2. Validation Results

The 4393 data samples are randomly divided into training datasets and test datasets with a division ratio of 7:3. The data-driven models are built based on training datasets. The predicted values based on independent variables in test datasets are compared with the original values to measure the model performance. The four indicators, cross-validation score, MSE, and R2, are used for assessing model performances. The model which is built by XGBoost for predicting capture rate had a cross-validation average score of 0.9995, R2 of 0.9996, and MSE of 0.0000. Figure 6a depicts the plot of the model results obtained by employing the XGBoost machine learning algorithm. In particular, the difference between predicted CR and tested CR in terms of the error represented by black solid dots is shown. Most points are distributed around the horizontal line y = 0. This means that the predicted values are close to the original values, which indicates the model has an accurate prediction. This aligns well with the high R2 of 0.9996. One possibility to lead to the high R2 is the overfitting. To validate this accuracy, cross-validation is performed to check the existence of the overfitting. In addition, the SVR algorithm is implemented to develop the other model for comparison purposes. The SVR models show accurate/similar prediction performance when compared with XGBoost model. If the SVR models show different prediction performances, then it can be concluded that the XGBoost models show better prediction abilities. Consequently, four models are built to predict capture rate and SRD individually by employing XGBoost and SVR algorithms, respectively. The closer the dots are to the y = 0, the better performance the model has. The qualities of both XGBoost and SVR models are observed in Figure 6. In particular, the XGBoost model shows better performance in predicting CR and SRD when compared with the SVR model as the error plot is less scattered. When using the XGBoost to predict the capture rate, the model has the highest prediction accuracy among the four models. The model using the SVR algorithm to predict the capture rate has the most scattered pattern, which illustrates the lowest accuracy. In Table 5, the performance of distinct models are summarized.

3.3. Verification Results

The verification of the developed models is performed by simulating 78 cases with the consideration of variables within the same range, as shown in Table 4 during process simulation. These cases are different from the cases in training datasets and test datasets. This verification dataset can include more data samples; however, the number of data samples is restricted due to the operational range where the process model is developed. The verification results of models are shown in Table 5.
It is assumed that the models are available when the independent variables are within certain ranges shown in Table 4. The availability of developed models in all cases or outside the range is a subject investigation. From Table 5, it is seen that the XGBoost models show R2 of 91.70% in predicting capture rate and 80.77% in predicting SRD, whereas the R2 of SVR models are 87.93% in predicting capture rate and 87.16% in predicting SRD. The model results are discussed as follows.
(1) In general, data-driven modeling is implemented via data collection, data analysis, variables selection, modeling and verification. It is noted that data collection and data analysis are steps that precede the variable selection. In this order, the related variables are kept, and unrelated variables are discarded. However, in this study, the order is slightly different due to data availability. The data are not preprovided and are generated by using process simulation. Hence, it is unrealistic to produce datasets that include all variables in process simulation models. Thus, the variable selection is conducted first for ensuring the working efficiency. The variables are chosen mainly based on previous research and industrial experience.
(2) Figure 5b–d have fluctuations, and Figure 5c has breakpoints. This is due to the instability of the process simulation models. The variables in these first-principal models include multiple differential and algebraic equations. Therefore, the same inputs lead to outputs with slight differences due to steady-state multiplicity.
(3) The amount of the original data samples is more than 4000, only 70% of the data samples are used to train the model and the step size between different samples are small, which may not differentiate much. Therefore, the model still has room for improvement if more data samples are included in training datasets.
(4) In many applications, the XGBoost algorithm has shown high accuracy of predictions. In this study, training and test datasets are synthetic data generated by the steady-state simulation. Therefore, these data have certain mathematic expressions because they come from the same simulation process model. Consequently, it is possible that the R2 of the model can reach 99%. The other reasons for this high accuracy are the complexity of the prediction task, the number of training datasets, and the possibility of overfitting, which has mentioned before.

3.4. Optimization Results

In this section, the optimization results of the PCC process simulation are explained. In general, the SRD increases with the capture rate. However, the lower SRD and higher capture rate are favorable. Therefore, three cases are investigated corresponding to the maximum capture rate, the lowest SRD, and the one with a tradeoff between capture rate and SRD respectively. The objective functions are shown in Equations (14)–(16). We have
f ( x ) = C R ( x )
f ( x ) = S R D ( x )
f ( x ) = S R D ( x ) S R D u p p e r l i m i t C R ( x ) C R u p p e r l i m i t ,
where the SRD(x) and CR(x) are the SRD and CR predicted by the XGBoost model based on the vector x, which represents the operating conditions. In the tradeoff case, the capture rate and the SRD are scaled to have a similar unit as they are not within the same ranges. The implementation of the agent-based and stochastic optimization involve many iterations. At each iteration, the values of different variables obtained by running process simulation are desired. It takes the process simulation model 3.15 s to predict the CR and SVR for 1 case. On the other hand, the data-driven models developed by using the XGBoost and SVR algorithms require 0.022 s and 0.217 s for predicting the CR and SVD for 1318 cases. The reduction of the prediction time has a positive effect on the optimization. During the application of the PSO algorithm, a swarm of 10 particles and 1500 number of iterations are considered. The operating conditions for the three specific objectives mentioned above are searched. The maximum capture rate reaches 95.98%, and the lowest SRD is 3.720 MJ/kg. The tradeoff case shows the minimal value of fitness in Equation (16) as −0.530. The detailed results are shown in Table 6. For the tradeoff case, the CR is lower than the CR of 76.3% in the base case and the SRD is 1.7 MJ/kg less than the 5.9 MJ/kg in the base case with the decrease of 28.2%.
The results obtained from the SA algorithm implementation are different from the PSO algorithm. For the SA algortihm, the initial temperature τ of 1000 °C, 300 iterations are considered. The searched operating points are depicted in Table 7. The largest capture rate is 94.87%, smaller than the 95.98% in the PSO algorithm. The SA method is better in searching the smallest SRD operating point. The operating point found by SA is 3.624 MJ/kg, is smaller than the 3.720 MJ/kg in the PSO algorithm. In the tradeoff case, a CR of 70.08% and SRD of 4.079 MJ/kg are obtained by employing SA. Compared with the base case, which has a CR of 76.3% and an SRD of 5.9 MJ/kg, although the CR was 6.22% lower, the SRD decreased by 1.83 MJ/kg, which accounts for 31.02% of 5.9 MJ/kg. The fitness f(x) in Equation (16) with this operating point is −0.518. As this value is larger than −0.530 in the PSO algorithm, the tradeoff operating points in the PSO performs better.
The operating points with the largest CR found by PSO and SA are quite different, whereas those with the smallest SRD and tradeoff case are similar. For the operating conditions corresponding to the lowest SRD, despite the other conditions being similar, lean amine loading reduces the value from 0.45 to 0.409, and the SRD reduces from 3.720 MJ/kg to 3.624 MJ/kg. This may not align with the trend shown in Figure 5c. These two operating points happen to be in the fluctuation part, where the decrease of lean amine loading leads to a lower SRD. This is possible in the case that the change of lean amine loading is small. The reason for this trend reversal is due to the three other parameters (flue gas flow rate, lean amine flow rate, CO2 molar fraction in flue gas) in these operating cases being similar rather than exactly the same. The change of the other three parameters may also influence the SRD. The tradeoff cases searched by PSO and SA are similar, the differences between values are less than 0.05, for example, and the flue gas flow rate is 0.18 kg/s and 0.162 kg/s, less than 0.02 kg/s. By combining these two algorithms, it is seen that the largest capture rate is 95.98%, and the lowest SRD is 3.624 MJ/kg. The tradeoff point, which has a balance between the larger capture rate and the lower SRD, is with flue gas flow rate 0.18 kg/s, lean amine flow rate 0.46 kg/s, CO2 molar fraction 0.201, and lean amine loading 0.309. The corresponding capture rate is 72.2% and specific reboiler duty is 4.3 MJ/kg. In the next section, the conclusions of this paper are presented.

4. Conclusions

In this paper, the application of machine learning techniques such as XGBoost and SVR for the PCC process model simulation was demonstrated successfully. First, the energy efficiency indicators and parameters associated with the PCC process model were identified. Machine learning algorithms were then applied to build models to predict CR and SRD. The models were used in optimization routines, and adequate operating conditions are characterized. The data-driven models showed high accuracy in predicting the capture rate and energy requirement in the reboiler of the PCC model. The XGBoost model had the accuracy of 99.9% and 99.3% for predicting CR and SRD based on the validation datasets. The SVR model showed 84.1% and 96.2% in CR and SRD prediction, respectively. The computation time for XGBoost and SVR models was approximately 90% lower when compared with the first-principle-based process model. Thus, the data-driven models showed improved performance in the time-efficient aspect. The goal of developing time-efficient models by machine learning techniques was achieved. The data-driven model were incorporated in optimization methods such as PSO and SA. The largest capture rate reached was found to be 96.0%, and the corresponding specific reboiler duty was 19.0 MJ/kg. The lowest specific reboiler duty was 3.6 MJ/kg; however, it had a low capture rate of 55.0%. The ideal operating condition was flue gas flow rate as 0.18 kg/s, lean amine flow rate as 0.46 kg/s, CO2 molar fraction in flue gas as 0.20, and lean amine loading as 0.31. The corresponding CR and SRD are 72.2% and 4.3 MJ/kg, respectively. In this case, the CR decreases 4.1% lower and SRD dwindles 1.7 MJ/kg compared with the base case. Thus, the application of machine learning techniques was demonstrated as useful in process optimization considering PCC process simulation. The advantage of fast model predictions by using data-driven techniques can be extended in the application of advanced online control and optimization methods. Processes with fast dynamics, in which the control makes calculations within a second, are required to benefit from the proposed method.

Author Contributions

H.Z. performed all simulations and wrote the first draft of the paper. G.M. helped with the writing, reviewing, editing, analysis, conceptualization, supervision and verification of the technical aspects of the research work. L.O.N. helped with the paper writing, funding acquisition and oversaw all the aspects of the research work. All authors have read and agreed to the published version of the manuscript.

Funding

This publication has been funded by the CleanExport project—Planning Clean Energy Export from Norway to Europe, Project no. 308811. The authors gratefully acknowledge the financial support from the Research Council of Norway and the user partners Agder Energi, Air Liquide, Equinor Energy, Gassco, and Total E&P Norge.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All datasets used in this work are generated by simulation software and available from the corresponding author on request.

Acknowledgments

The authors acknowledge the Department of Energy and Process Engineering at NTNU for facilitating with computer resources and software package. The authors would also like to thank NTNU Ph.D. candidate Valentin Formont for help during the work.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature and Abbreviations

The following nopmenclature and abbreviations are used in this manuscript:
XGBoosteXtreme Gradient Boosting
SVRSupport Vector Regression
CRCapture Rate
SRDSpecific Reboiler Duty
R2Coefficient of determination
WMOWorld Meteorological Organization
GHGGreenhouse Gas Emissions
CCSCarbon capture and storage
PCCPost-combustion capture
MEAMonoethanolamine
DEADi-EthanolAmine
AMP2-amino-2-methyl-1-propanol
PZPiperazine
ILIonic Liquid
ANNArtificial Neural Network
LSSVMLeast Squares Support Vector Machine
GPRGaussian Process Regression
BPMLPBack Propagation Multi-Layer Perceptron
RBFRadial Basis Function
ELMExtreme Learning Machine
MSEMean squared error
ARXAuto-Regressive model with eXogenous inputs
BIOCSBiologically Inspired Optimal Control Strategy
PSOParticle Swarm Optimization
GEPGeneration Expansion Planning
GAGenetic Algorithm
ABCArtificial Bee Colony
RBFNNRadial Basis Function Neural Network
SASimulated Annealing
GA-ANFISGenetic Algorithm-Adaptive Neuro Fuzzy Inference System
PSO-ANFISParticle Swarm Optimization ANFIS
CSA-LSSVMCoupled Simulated Annealing-Least Squares Support Vector Machine
SVMSupport Vector Machine
SSTThe total sum of squares
SSRThe sum squared regress
MSEMean squared error
SASimulated Annealing

References

  1. World Meteorological Organization. 2020 Was One of Three Warmest Years on Record. 2021. Available online: https://public.wmo.int/en/media/press-release/2020-was-one-of-three-warmest-years-record (accessed on 13 May 2022).
  2. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Dowell, N.M.; Fernández, J.R.; Ferrari, M.-C.; Gross, R.; Hallett, J.P.; et al. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [Google Scholar] [CrossRef]
  3. Rubin, E.S.; Mantripragada, H.; Marks, A.; Versteeg, P.; Kitchin, J. The outlook for improved carbon capture technology. Prog. Energy Combust. Sci. 2012, 38, 630–671. [Google Scholar] [CrossRef]
  4. Svendsen, H.F.; Hessen, E.T.; Mejdell, T. Carbon dioxide capture by absorption, challenges and possibilities. Chem. Eng. J. 2011, 171, 718–724. [Google Scholar] [CrossRef]
  5. Aspen HYSYS® (Version 10) [Software]. 2022. Available online: https://www.aspentech.com/en (accessed on 22 August 2022).
  6. Aspen Plus® (Version 12) [Software]. 2022. Available online: https://www.aspentech.com/en (accessed on 22 August 2022).
  7. gPROMS® (Version 2) [Software]. 2022. Available online: https://www.psenterprise.com/products/gproms (accessed on 22 August 2022).
  8. Co2sim® [Software]. 2022. Available online: https://www.sintef.no/en/software/co2sim-flowsheet-simulator-for-co2-absorption-proc/ (accessed on 22 August 2022).
  9. Øi, L.E. Comparison of aspen hysys and aspen plus simulation of CO2 absorption into mea from atmospheric gas. Energy Procedia 2012, 23, 360–369. [Google Scholar] [CrossRef] [Green Version]
  10. van der Spek, M.; Arendsen, R.; Ramirez, A.; Faaij, A. Model development and process simulation of postcombustion carbon capture technology with aqueous amp/pz solvent. Int. J. Greenh. Gas Control 2016, 47, 176–199. [Google Scholar] [CrossRef]
  11. Luo, X.; Wang, M. Improving prediction accuracy of a rate-based model of an MEA-based carbon capture process for large-scale commercial deployment. Engineering 2017, 3, 232–243. [Google Scholar] [CrossRef]
  12. Dutta, R.; Nord, L.O.; Bolland, O. Applicability and validation of use of equilibrium-based absorber models with reduced stage efficiency for dynamic simulation of post-combustion CO2 capture processes. Energy Procedia 2017, 114, 1424–1433. [Google Scholar] [CrossRef]
  13. Fernandez, M.; Boyd, P.G.; Daff, T.D.; Aghaji, M.Z.; Woo, T.K. Rapid and accurate machine learning recognition of high performing metal organic frameworks for CO2 capture. J. Phys. Chem. Lett. 2014, 5, 3056–3060. [Google Scholar] [CrossRef]
  14. Venkatraman, V.; Alsberg, B.K. Predicting CO2 capture of ionic liquids using machine learning. J. Co2 Util. 2017, 21, 162–168. [Google Scholar] [CrossRef]
  15. Ahmadi, M.; Kashiwao, T.; Rozyn, J.; Bahadori, A. Accurate prediction of properties of carbon dioxide for carbon capture and sequestration operations. Pet. Sci. Technol. 2016, 34, 97–103. [Google Scholar] [CrossRef]
  16. Wu, X.; Shen, J.; Li, Y.; Wang, M.; Lawal, A. Flexible operation of post-combustion solvent-based carbon capture for coal-fired power plants using multi-model predictive control: A simulation study. Fuel 2018, 220, 931–941. [Google Scholar] [CrossRef] [Green Version]
  17. Rahimi, M.; Moosavi, S.M.; Smit, B.; Hatton, T.A. Toward smart carbon capture with machine learning. Cell Rep. Phys. Sci. 2021, 2, 100396. [Google Scholar] [CrossRef]
  18. Shalaby, A.; Elkamel, A.; Douglas, P.L.; Zhu, Q.; Zheng, Q.P. A machine learning approach for modeling and optimization of a CO2 post-combustion capture unit. Energy 2021, 215, 119113. [Google Scholar] [CrossRef]
  19. Shahsavand, A.; Fard, F.D.; Sotoudeh, F. Application of artificial neural networks for simulation of experimental CO2 absorption data in a packed column. J. Nat. Gas Sci. Eng. 2011, 3, 518–529. [Google Scholar] [CrossRef]
  20. Chan, V.; Chan, C. Learning from a carbon dioxide capture system dataset: Application of the piecewise neural network algorithm. Petroleum 2017, 3, 56–67. [Google Scholar] [CrossRef]
  21. Li, F.; Zhang, J.; Oko, E.; Wang, M. Modelling of a post-combustion CO2 capture process using extreme learning machine. Int. J. Coal Sci. Technol. 2017, 4, 33–40. [Google Scholar] [CrossRef] [Green Version]
  22. Mirlekar, G.; Gebreslassieb, B.; Diwekar, U.; Lima, F.V. Biomimetic model-based advanced control strategy integrated with multi-agent optimization for nonlinear chemical processes. Chem. Eng. Res. Des. 2018, 140, 229–240. [Google Scholar] [CrossRef]
  23. Mirlekar, G.; Al-Sinbol, G.; Perhinschi, M.; Lima, F.V. A biologically-inspired approach for adaptive control of advanced energy systems. Comput. Chem. Eng. 2018, 117, 378–390. [Google Scholar] [CrossRef]
  24. Saboori, H.; Hemmati, R. Considering carbon capture and storage in electricity generation expansion planning. IEEE Trans. Sustain. Energy 2016, 7, 1371–1378. [Google Scholar] [CrossRef]
  25. Menad, N.A.; Hemmati-Sarapardeh, A.; Varamesh, A.; Shamshirband, S. Predicting solubility of CO2 in brine by advanced machine learning systems: Application to carbon capture and sequestration. J. CO2 Util. 2019, 33, 83–95. [Google Scholar] [CrossRef]
  26. Van Laarhoven, P.J.; Aarts, E.H. Simulated annealing. In Simulated Annealing: Theory and Applications; Springer: Dordrecht, The Netherlands, 1987; pp. 7–15. [Google Scholar] [CrossRef]
  27. Zhao, B.; Su, Y.; Tao, W. Mass transfer performance of CO2 capture in rotating packed bed: Dimensionless modeling and intelligent prediction. Appl. Energy 2014, 136, 132–142. [Google Scholar] [CrossRef]
  28. Dashti, A.; Raji, M.; Alivand, M.S.; Mohammadi, A.H. Estimation of CO2 equilibrium absorption in aqueous solutions of commonly used amines using different computational schemes. Fuel 2020, 264, 116616. [Google Scholar] [CrossRef]
  29. Dutta, R.; Nord, L.O.; Bolland, O. Prospects of using equilibrium-based column models in dynamic process simulation of post-combustion CO2 capture for coal-fired power plant. Fuel 2017, 202, 85–97. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, T.; Guestrin, C. Xgboost: A scalable tree boosting system. In Proceedings of the 22nd ACM Sigkdd International Conference on Knowledge Discovery and Data Mining, San Francisco, CA, USA, 13–17 August 2016; pp. 785–794. [Google Scholar] [CrossRef] [Green Version]
  31. Friedman, J.H. Greedy function approximation: A gradient boosting machine. Ann. Stat. 2001, 29, 1189–1232. [Google Scholar] [CrossRef]
  32. Smola, A.J.; Schölkopf, B. A tutorial on support vector regression. Stat. Comput. 2004, 14, 199–222. [Google Scholar] [CrossRef] [Green Version]
  33. Awad, M.; Khanna, R. Support vector regression. In Efficient Learning Machines; Apress: Berkeley, CA, USA, 2015; pp. 67–80. [Google Scholar] [CrossRef] [Green Version]
  34. Kennedy, J.; Eberhart, R. Particle swarm optimization. In Proceedings of the ICNN’95-International Conference on Neural Networks, Perth, WA, Australia, 27 November–1 December 1995; Volume 4, pp. 1942–1948. [Google Scholar] [CrossRef]
  35. Kirkpatrick, S.; Gelatt, C.D., Jr.; Vecchi, M.P. Optimization by simulated annealing. Science 1983, 220, 671–680. [Google Scholar] [CrossRef]
Figure 1. Workflow diagram of the proposed method for data-driven process modeling and optimization.
Figure 1. Workflow diagram of the proposed method for data-driven process modeling and optimization.
Processes 10 02727 g001
Figure 2. Schematic of chemical absorption process representing postcombustion CO2 capture (PCC).
Figure 2. Schematic of chemical absorption process representing postcombustion CO2 capture (PCC).
Processes 10 02727 g002
Figure 3. Schematic of evolution from decision tree to XGBoost.
Figure 3. Schematic of evolution from decision tree to XGBoost.
Processes 10 02727 g003
Figure 4. Plot of data correlation coefficients.
Figure 4. Plot of data correlation coefficients.
Processes 10 02727 g004
Figure 5. Plots of original data collected and smoothed curve for each independent variable and dependent variable. (a) Flue gas flow rate vs. capture rate/specific reboiler duty. (b) Lean amine flow rate vs. capture rate/specific reboiler duty. (c) CO2 molar fraction in flue gas vs. capture rate/specific reboiler duty. (d) Lean amine loading vs. capture rate/specific reboiler duty.
Figure 5. Plots of original data collected and smoothed curve for each independent variable and dependent variable. (a) Flue gas flow rate vs. capture rate/specific reboiler duty. (b) Lean amine flow rate vs. capture rate/specific reboiler duty. (c) CO2 molar fraction in flue gas vs. capture rate/specific reboiler duty. (d) Lean amine loading vs. capture rate/specific reboiler duty.
Processes 10 02727 g005
Figure 6. Plots of errors for model prediction. (a) Error of capture rate prediction (XGBoost). (b) Error of SRD prediction (XGBoost). (c) Error of capture rate prediction (SVR). (d) Error of SRD prediction (SVR). All errors for model prediction are based on original data.
Figure 6. Plots of errors for model prediction. (a) Error of capture rate prediction (XGBoost). (b) Error of SRD prediction (XGBoost). (c) Error of capture rate prediction (SVR). (d) Error of SRD prediction (SVR). All errors for model prediction are based on original data.
Processes 10 02727 g006
Table 1. Parameters in base case.
Table 1. Parameters in base case.
StreamMass Flow Rate (kg/s)Temperature (K)Pressure (bar)
Lean amine0.642313.71.703
Flue gas0.158332.41.033
Table 2. Base case values.
Table 2. Base case values.
StreamValue
Lean amine loading (-)0.281
Rich amine loading (-)0.488
Capture rate (-)0.763
Reboiler duty (MJ/kg)5.958
Table 3. Stream composition in base case.
Table 3. Stream composition in base case.
CompositionsLean Amine Mass FractionFlue Gas Molar Fraction
H2O0.6330.033
CO20.0620.167
MEA0.305-
N2-0.801
Table 4. Data collection information.
Table 4. Data collection information.
UnitLower LimitUpper LimitStep SizeTotal Samples
Flue gas mass flow ratekg/s0.1080.2620.00011543
Lean amine mass flow ratekg/s0.4522.0040.011553
CO2 molar fraction in flue gas-0.1670.2260.0004150
Lean loading-0.0280.5240.00051147
Table 5. Model assessment in terms of model validation and verification. The cross-validation score is the average value of R2 of all testing folds. Validation R2 and Validation MSE are the values during the model validation. After validation, 78 cases different from training and test datasets are used for verification. The verification R2 and verification MSE are the results of model performance assessment during verification.
Table 5. Model assessment in terms of model validation and verification. The cross-validation score is the average value of R2 of all testing folds. Validation R2 and Validation MSE are the values during the model validation. After validation, 78 cases different from training and test datasets are used for verification. The verification R2 and verification MSE are the results of model performance assessment during verification.
AssessmentXGBoost Predict CRXGBoost Predict SRDSVR Predict CRSVR Predict SRD
cross-validation score0.99950.99340.85730.9620
Validation R20.99960.99300.84180.9622
Validation MSE0.00000.05950.00410.3231
Verification R20.91700.80770.87930.8716
Verification MSE0.00290.40430.00420.2700
Table 6. Optimized operating conditions for PSO algorithm implementation. Case 1: Find the operating conditions that resulted in the largest CR without leveraging the value of SRD. Case 2: Find the operating conditions corresponding to the smallest SRD without leveraging the value of CR. Case 3: A trade-off case, it aims to find a relatively high CR and low SRD.
Table 6. Optimized operating conditions for PSO algorithm implementation. Case 1: Find the operating conditions that resulted in the largest CR without leveraging the value of SRD. Case 2: Find the operating conditions corresponding to the smallest SRD without leveraging the value of CR. Case 3: A trade-off case, it aims to find a relatively high CR and low SRD.
Operating ConditionsCase 1: Largest CRCase 2: Smallest SRDCase 3: Trade-Off Case
flue gas flow rate0.1580.260.18
lean amine flow rate0.6060.460.46
CO2 molar fraction0.1670.210.200
lean amine loading0.0280.450.309
CR0.9600.5500.722
SRD19.0293.7204.3
Table 7. Optimized operating conditions for SA algorithm implementation. Case 1: Find the operating conditions that resulted in the largest CR without leveraging the value of SRD. Case 2: Find the operating conditions corresponding to the smallest SRD without leveraging the value of CR. Case 3: A trade-off case, aims to find a relatively high CR and low SRD.
Table 7. Optimized operating conditions for SA algorithm implementation. Case 1: Find the operating conditions that resulted in the largest CR without leveraging the value of SRD. Case 2: Find the operating conditions corresponding to the smallest SRD without leveraging the value of CR. Case 3: A trade-off case, aims to find a relatively high CR and low SRD.
Operating ConditionsCase 1: Largest CRCase 2: Smallest SRDCase 3: Trade-Off Case
flue gas flow rate0.1430.2610.163
lean amine flow rate1.9140.4600.474
CO2 molar fraction0.1660.2180.201
lean amine loading0.0790.4090.339
CR0.9490.5500.701
SRD13.5403.6244.079
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, H.; Mirlekar, G.; Nord, L.O. Agent-Based and Stochastic Optimization Incorporated with Machine Learning for Simulation of Postcombustion CO2 Capture Process. Processes 2022, 10, 2727. https://doi.org/10.3390/pr10122727

AMA Style

Zheng H, Mirlekar G, Nord LO. Agent-Based and Stochastic Optimization Incorporated with Machine Learning for Simulation of Postcombustion CO2 Capture Process. Processes. 2022; 10(12):2727. https://doi.org/10.3390/pr10122727

Chicago/Turabian Style

Zheng, Huilan, Gaurav Mirlekar, and Lars O. Nord. 2022. "Agent-Based and Stochastic Optimization Incorporated with Machine Learning for Simulation of Postcombustion CO2 Capture Process" Processes 10, no. 12: 2727. https://doi.org/10.3390/pr10122727

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