Next Article in Journal
On the Complexity of the Bipartite Polarization Problem: From Neutral to Highly Polarized Discussions
Previous Article in Journal
Augmented Dataset for Vision-Based Analysis of Railroad Ballast via Multi-Dimensional Data Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification of Crude Distillation Unit: A Comparison between Neural Network and Koopman Operator

by
Abdulrazaq Nafiu Abubakar
1,
Mustapha Kamel Khaldi
1,*,
Mujahed Aldhaifallah
1,
Rohit Patwardhan
2 and
Hussain Salloum
2
1
Control and Instrumentation Engineering Department, King Fahd University of Petroleum and Minerals, P.O. Box 5067, Dhahran 31261, Saudi Arabia
2
Process & Control Systems Department, Saudi Aramco, Dhahran 31311, Saudi Arabia
*
Author to whom correspondence should be addressed.
Algorithms 2024, 17(8), 368; https://doi.org/10.3390/a17080368 (registering DOI)
Submission received: 25 June 2024 / Revised: 28 July 2024 / Accepted: 2 August 2024 / Published: 21 August 2024

Abstract

:
In this paper, we aimed to identify the dynamics of a crude distillation unit (CDU) using closed-loop data with NARX−NN and the Koopman operator in both linear (KL) and bilinear (KB) forms. A comparative analysis was conducted to assess the performance of each method under different experimental conditions, such as the gain, a delay and time constant mismatch, tight constraints, nonlinearities, and poor tuning. Although NARX−NN showed good training performance with the lowest Mean Squared Error (MSE), the KB demonstrated better generalization and robustness, outperforming the other methods. The KL observed a significant decline in performance in the presence of nonlinearities in inputs, yet it remained competitive with the KB under other circumstances. The use of the bilinear form proved to be crucial, as it offered a more accurate representation of CDU dynamics, resulting in enhanced performance.

1. Introduction

Crude distillation units (CDUs) play a significant role in the refining sector, responsible for fractionating crude oil into intermediate products that are subsequently processed in downstream units to meet the market specifications. The quality of these intermediate products, primarily influenced by the operating conditions of the CDUs, is crucial for ensuring the quality of the final refinery products [1]. Currently, the oil refining industry is confronting several challenges, including a significant increase in crude oil prices, fluctuations in product demand driven by market dynamics, and specific regulatory constraints imposed on industrial activities [2]. Improving these units may lead to significant improvements in the efficiency and reliability of oil refining processes, ultimately reducing operational costs and increasing overall savings. Developing suitable process control for the CDU is necessary to achieve these objectives. Nevertheless, the existing body of literature primarily focuses on the design of control algorithms, rather than the identification of a CDU to a sufficient accuracy.
In recent years, significant advancements in artificial intelligence have facilitated the development of dynamic models through various data-driven techniques, such as polynomial regressions [3,4,5,6], support vector regression (SVR) [7], and Artificial Neural Networks (ANNs). For example, Liau et al. [8] aimed at optimizing product outputs by using an ANN to predict the yield of kerosene, diesel, and atmospheric gas oil. Motlaghi et al. [9] employed an ANN to predict products flow rate that were optimized based on market values. Gueddar et al. [10] developed an ANN model to optimize energy efficiency by considering crude oil properties, such as boiling point and crude flow rate. Building on this approach, Durrani et al. [11] developed a multi-output ANN model to address variations in crude composition and predict optimum cut-point temperatures, using a hybrid Taguchi and genetic algorithm for more energy-efficient operations. Ochoa-Estopier et al. [12] developed an ANN model for a CDU and employed a Simulated Annealing (SA) optimizer to enhance revenue while reducing energy usage. This work was further extended by the same authors in [13,14], who incorporated a heat exchanger network model to enhance operational optimization, aiming to boost net profit while adhering to practical constraints. Shi et al. [15] modeled a CDU process using a wavelet neural network which was combined with the line-up competition algorithm (LCA) for the economic optimization of the CDU operation. More recently, a Long Short-Term Memory network (LSTM) was developed by [16] to predict and analyze energy efficiency in the CDU under different operating conditions. A hybrid ANN-SVM model was developed by [17] to simulate the performance of the CDU accurately and efficiently within an optimization framework. Li et al. [18] developed a hybrid Fuzzy Logic–ANN model to construct a knowledge-based strategy to adapt to different feedstock properties. Bootstrap ANN models were used by Osuolale and Zhang in [19] and Muhsin et al. in [20] to develop a model of a CDU process, with the former authors focusing on energy efficiency and the latter on maximizing the production rate. A comparison between different data-driven models for predicting CDU product properties was investigated by [21], including PCA-ResNet, SOM-ResNet, Feedforward Neural Networks (FNNs), Partial Least Squares (PLS), and LASSO. The study concluded that incorporating prior knowledge and employing appropriate dimensionality reduction techniques, such as PCA-ResNet, greatly improved model accuracy. Using machine learning can offer more accurate and sometimes computationally efficient solutions compared to the complex and resource-intensive nature of ANNs [22]. Fadzil et al. [23] explored five machine learning models for optimizing product yields based on varying feed properties and operating conditions. These models included decision tree regression, support vector regression, ANN, random forest regression, and extreme gradient boosting (XGBoost), with XGBoost demonstrating superior performance.
While ANNs have proven to be suitable for modeling a CDU, they rely on static mapping of outputs from inputs using data; this limits our understanding of the physical or chemical mechanisms governing the CDU. Linear predictors can be utilized for nonlinear systems to effectively capture and model nonlinear behavior while also identifying the dynamic response of a linear system. From a control engineering perspective, this approach provides valuable insights into the process by analyzing its time-domain characteristics [24]. In this context, Bernard Koopman [25] introduced his operator, which describes the evolution of measurements in Hamiltonian systems over time. Mezic and his collaborators [26,27] expanded these concepts to include nonconservative systems, extending its utility to a wide range of applications, solar panels [28], power systems [29], robotics [30], autonomous driving [31], biology [32], and traffic flow [33], to list a few.
The quality of a model is usually evaluated by its ability to generalize to new, unseen data. In regression tasks, if the model is properly selected in terms of structure and hyperparameters, and overfitting is avoided, the problem can be considered conceptually solved. However, when dealing with real data from advanced process control units, inputs to the model may fall outside the training domain due to factors like model-plant mismatches, poor tuning, or stringent constraints in the closed-loop system. In such situations, it is essential for the regression model to make acceptable predictions or, at the very least, not fail. This subject is the topic of this paper. In this work, a comparative analysis of modeling a CDU under different experimental conditions is conducted. This includes the Koopman operator in both linear (KL) and bilinear (KB) forms, as well as a NARX−NN model. The performance of these models is tested in real-world settings, such as gain and delay mismatches, nonlinearities, and disturbances. Bayesian optimization is used for hyperparameter tuning to ensure a fair comparison. The remainder of the paper is structured as follows: Section 2 provides preliminaries of the methodologies used. Section 3 covers the process description and data generation. The results and discussion are presented in Section 4. Finally, the conclusion and suggestions for future work are outlined in Section 5.

2. Modeling and Methodology

2.1. Nonlinear Autoregressive Exogenous Model

The Nonlinear AutoRegressive Exogenous model (NARX) was proposed as a novel framework for representing a diverse range of nonlinear systems. The fundamental principle behind this model is the incorporation of past inputs and outputs, which helps to mitigate the effects of both nonlinearities and delays in the output [34]. This model is defined as:
y ( k ) = F [ y ( k 1 ) , y ( k 2 ) , , y ( k n y ) , u ( k d ) , u ( k d 1 ) , , u ( k d n u ) ]
where y ( k ) and u ( k ) represent the output and the input of the system, respectively, with n y and n u the maximum lags for the output and input, whereas d is the time delay. F [ . ] denotes a nonlinear mapping function. In practical applications, there are several methods that can be used to approximate the unknown function F [ . ] ; these include polynomial series, wavelet expansions, Fuzzy Logic, Neural Networks, and many others.
While the formulation presented in Equation (1) is designed for a Single-Input Single-Output (SISO) system, it is straightforward to extend these models to accommodate Multi-Input Multi-Output (MIMO) systems. For example, consider a generic MIMO scenario with r inputs { u 1 ( k ) , u 2 ( k ) . . , u r ( k ) } and s outputs { y 1 ( k ) , y 2 ( k ) . . , y s ( k ) } , as given by Equation (2).
u i [ k 1 ] = [ u i ( k 1 ) , u i ( k 2 ) , , u i ( k n u ) ] y j [ k 1 ] = [ y j ( k 1 ) , y j ( k 2 ) , , y j ( k n y ) ] i = 1 , 2 , 3 , , r j = 1 , 2 , , s
The formulation of the NARX model for the MIMO system (2) can be expressed as
y 1 ( k ) = F 1 [ y 1 ( k 1 ) , , y s ( k 1 ) , u 1 ( k 1 ) , , u r ( k 1 ) ] y 2 ( k ) = F 2 [ y 1 ( k 1 ) , , y s ( k 1 ) , u 1 ( k 1 ) , , u r ( k 1 ) ] y s ( k ) = F s [ y 1 ( k 1 ) , , y s ( k 1 ) , u 1 ( k 1 ) , , u r ( k 1 ) ]
where F 1 [ . ] , F 2 [ . ] ,..., F s [ . ] are nonlinear functions. ANNs are widely used in signal processing, pattern recognition, data fitting, analysis, and control due to their ability to learn and represent mathematical descriptions of systems [35]. This can be either Single-Layer Networks (SLNs) or Multi-Layer Networks (MLNs). SLNs, in particular, have been extensively applied to reconstruct the mapping function F [ . ] , resulting in more accurate NARX models in different applications [36,37,38,39]. Therefore, this approach was adopted in our work. Figure 1 depicts the conventional configuration of a NARX−NN model. Combining the past inputs and outputs into a single vector x ( k ) :
x ( k ) = [ u 1 ( k 1 ) , , u 1 ( k n u ) , u r ( k 1 ) , , u r ( k n u ) , y 1 ( k 1 ) , , y 1 ( k n y ) , y s ( k 1 ) , , y s ( k n y ) ] T
The NARX−NN model can be mathematically expressed in a matrix form as follows:
y ( k + 1 ) = ϕ ( W x ( k ) + b )
where y ( k + 1 ) is the output vector { y 1 ( k + 1 ) , , y s ( k + 1 ) } , W is the weight matrix, b is the bias vector, and ϕ is typically a nonlinear activation function applied element-wise (e.g., sigmoid, tanh, ReLU). A pseudo-code for the identification of the NARX−NN model is presented in Algorithm 1.
Algorithm 1 NARX−NN model identification algorithm
 1:
Input Data D = Y , U
 2:
Initialize W , b
 3:
for all epochs do
 4:
    for all batches of D do
 5:
    Construct input vector from past inputs/outputs Equation (4)
 6:
     Forward pass to obtain predicted output y ^ ( k + 1 ) Equation (5)
 7:
        Compute loss L = 1 N k = 1 N ( y ( k + 1 ) y ^ ( k + 1 ) ) 2
 8:
        Update W and b using backpropagation
 9:
    end for
10:
end for

2.2. Koopman Operator Theory

In this section, a brief introduction to the Koopman operator is provided; for more details, readers should refer to [40].
First, consider a discrete nonlinear autonomous system in the following form:
x k + 1 = f ( x k )
where x k X R n is the vector of state variables, X is the state space, and f : X X is a map function that represents the evolution of the state. The infinite-dimensional Koopman operator K acting on a set of observation functions g is defined as
K g ( x k ) = g ( f ( x k ) ) = g ( x k + 1 )
where K : G G and g : X G . G denotes the lifted space. The Koopman operator advances the value of the observable function forward in time. One might obtain a finite approximation of the Koopman operator by restricting it to an invariant subspace G N G , which is spanned by a set of Koopman eigenfunctions that satisfy the following property
λ ϕ x = φ ( f x )
where the eigenvalue λ corresponds to the eigenfunction φ ( x ) . Using the Koopman Mode Decomposition (KMD) [27], the observation function g x k can be represented as
g x k = K k g x 0 = λ j k φ j x 0 v j
where v j is the jth Koopman mode associated with the eigenfunction φ j and the eigenvalue  λ j .
For a discrete-time control system, where u U R m ,
x k + 1 = f ( x k , u k )
We define an extended state x ˜ : = [ x T , u T ] T , and let g ( x , u ) : X × U G be a set of scalar observable function of the extended state space. The Koopman operator advances measurement functions according to
K g ( x k , u k ) = g ( f ( x k , u k ) , u k ) = g ( x k + 1 , u k + 1 )
The Koopman eigenpairs ( φ , λ ) associated with Equation (11) satisfy
K φ ( x k , u k ) = φ ( f ( x k , u k ) , u k ) = φ ( x k + 1 , u k + 1 ) = λ φ ( x k , u k )
Assuming that the observation function can be separated into two parts:
g x , u = g x x , u ; g u x , u
Then, the expression for the Koopman operator would have the following form [41,42]:
g x x k + 1 , u k + 1 u k + 1 = A B C D g x x , u g u x , u
where A , B , C , D are submatrices of the Koopman operator matrix K . We assume that g x x , u depends only on the system’s state, g x x , u = g x x . An embedding ANN, ϑ x k : R n R d , where d is a tunable hyperparameter, shall be used to learn the observation function g x ( x ) . Furthermore, the control inputs in the lifted space are assumed to remain unchanged compared to those in the nonlinear space; hence, g u x , u = u . Denoting z k = g x x k , u k as the state within the lifted dynamics, a linear relationship of the state variable z can be obtained as follows
z k + 1 = A z k + B u k
By employing the Koopman Canonical Transformation (KCT) [43], the Koopman bilinear form (KBF) [44] is given by,
z k + 1 = A z k + i = 1 m B i u i
A schematic representation of the Koopman dynamics based on an ANN is presented in Figure 2.
Furthermore, the original states are concatenated with the output of ϑ , allowing for the direct retrieval of the original states:
z k = x k ϑ x k x k = C x z k , C x = I n 0
To learn the Koopman dynamics jointly with the embedding network, we formulate the following loss function
L = L p r e d + α L l i n + β L L 1
where
L p r e d = x k + 1 C x z ^ k + 1 2
L l i n = z k + 1 z ^ k + 1 2
L L 1 = A 1 + B 1
where . 2 represents the Mean Squared Error (MSE), averaged over the data points, then over the number of outputs. The term z ^ denotes the estimated lifted state obtained from Equation (15) if a linear model is used, or from Equation (16) if a bilinear model is employed. The loss term L p r e d represents the one-step prediction error in the nonlinear state-space, which ensures the consistency of the Koopman model with the underlying nonlinear system as it progresses over time. L l i n represents the one-step prediction error in the lifted space. Additionally, L L 1 encourages sparsity in the Koopman dynamical system, facilitating better generalization and reducing overfitting. The coefficients α and β determine the weighting of L l i n and L L 1 with respect to L p r e d . The algorithm for training Koopman dynamics is summarized as Algorithm 2, where ϑ θ are the weights of the embedding network ϑ .
Algorithm 2 Training procedure of the Koopman dynamics
  • Input Dataset D = X , U
  • Initialize A , B , ϑ θ
  • for all epochs do
  •     for all batches of D do
  •         lift the state z k [ x k , ϑ ( x k ) ]
  •         lift the state z k + 1 [ x k + 1 , ϑ ( x k + 1 ) ]
  •        Forward z k one step in time to obtain the estimated lifted state z ^
  •         Compute prediction loss Equation (19)
  •         Compute linear loss Equation (20)
  •         Compute the total loss Equation (18)
  •         Update A , B , ϑ θ by back propagation
  •     end for
  • end for

3. Application to CDU

3.1. Process Description

The crude distillation unit (CDU) is the initial processing unit in a petroleum refinery, responsible for separating crude oil into various distillate streams that serve as essential raw materials for downstream refining processes. This process can be divided into several units based on the level of output, including the pre-flash unit (PFU), atmospheric distillation unit (ADU), vacuum distillation unit (VDU), splitting unit (SPU), stabilizer unit (SBU), and heat exchanger network (HEN). Figure 3 illustrates a basic representation of the ADU used in this work. In the CDU, crude oil is first heated, and water is injected to dissolve salts. The mixture then undergoes electrostatic precipitation in a desalter drum to separate the salts. The crude oil is subsequently routed to the distillation tower’s flash zone, where it is heated in a fired heater to vaporize distillate products. Overflash is added to ensure effective reflux streams within the tower. The heated crude oil enters the fractionation tower in the flash zone, where the distillate vapors ascend the tower, countering a colder liquid reflux stream. Distillate products are then removed from selected trays, stream stripped, and sent to storage [45,46].

3.2. Data Collection

The data were generated using a simulated CDU unit under a Model Predictive Controller (MPC) within the Aspen DMC Plus software. The data included three control variables (CVs), three manipulated variables (MVs), and one feedforward (FF) variable, as listed in Table 1. The sampling period was set to 1 min, with a time to steady state of 30 min. A prediction horizon of 30 and a control horizon of 10 were used for the MPC. Various experimental conditions were simulated to depict real-life situations, including gain, delay, and time-constant mismatches, tight constraints, nonlinearities, and poor PID tuning. Each run lasted for 1440 samples, equivalent to 1 day. Table 2 describe the settings for each case. To validate the efficiency of the proposed methods in capturing the dynamics of the CDU, training and validation were performed on the nominal dataset (Case 1), and each model was tested on the remaining datasets.

4. Results and Discussion

To ensure a fair comparison among the models, the hyperparameters of each model were fine-tuned using Bayesian optimization, implemented with the KerasTuner API. The objective was to minimize the validation MSE. Table 3 outlines the specific hyperparameters and their respective search ranges. The results are summarized in Table 4. In what follows, the Koopman linear and bilinear models are referred to as KL and KB, respectively.
The results based on the training dataset are summarized in Figure 4, and those for the test datasets are provided within Figure 5, Figure 6, Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11. The plots in Figure A1, Figure A2, Figure A3, Figure A4, Figure A5, Figure A6, Figure A7, Figure A8, Figure A9, Figure A10, Figure A11, Figure A12, Figure A13, Figure A14, Figure A15, Figure A16, Figure A17, Figure A18, Figure A19, Figure A20, Figure A21, Figure A22, Figure A23 and Figure A24 in Appendix A depict the predicted outputs ( y 1 , y 2 , y 3 ) vs. the actual values for each case. As can be noticed from Figure 4, the NARX−NN model had the lowest average MSE (0.0081) on the training dataset, suggesting that NARX−NN fitted the training data well and indicating good initial performance in the identification of the CDU. In comparison, the KB model performed slightly better than the KL model, where the latter showed the highest average MSE of 0.0127. However, across all test scenarios, KB consistently achieved superior performance with the lowest average MSE, indicating better generalization ability and robustness in capturing the dynamics of the CDU across different experimental conditions. The KL model performed competitively with KB and often outperformed NARX−NN. While benefiting from exogenous inputs and past outputs is powerful for NARX−NN, it also adds more complexity to the model. This may be sensitive in terms of model parameters and can lead to overfitting, causing poorer generalization in capturing the dynamics under different experimental conditions.
In cases where nonlinearities were present in the inputs, both the KL and KB models experienced a noticeable degradation in performance. This degradation was primarily due to the assumption that the control inputs remained unchanged within the lifted dynamics. However, the KL model exhibited the poorest performance; conversely, the inclusion of a bilinear term in the KB model allowed for a more flexible representation of the system dynamics. This flexibility enabled the KB model to capture certain nonlinear behaviors that the linear KL model could not, leading to its superior performance. Based on these observations, one may conclude that under different experimental conditions, the NARX−NN model dramatically failed to provide predictions, while the KB model continued to make good predictions.

5. Conclusions

This article presented a comparison of three different models in identifying the CDU process, namely NARX−NN, KL, and KB, conducted under different experimental conditions, such as gain, delay mismatches, nonlinearities, and disturbances. During the training process, the NARX−NN model had the lowest average Mean Squared Error (MSE), indicating a high level of initial competency. Nevertheless, KB demonstrated higher performance and robustness across different datasets used for testing, surpassing the performance of both NARX−NN and KL. Although KL showed comparable performance to KB and sometimes outperformed NARX−NN, its performance significantly deteriorated when faced with nonlinearities.
The use of a bilinear term in KB was the main differentiating factor, providing a more flexible representation of system dynamics compared to the linear assumption in the KL model. This flexibility allowed KB to capture nonlinear phenomena that other models could not, resulting in superior performance, as evidenced by lower MSE values. The KB stands out as a potential option because of its improved capacity to generalize.

Author Contributions

Conceptualization, A.N.A., M.K.K., M.A., R.P. and H.S.; Data curation, M.K.K. and R.P.; Formal analysis, A.N.A., M.K.K. and M.A.; Funding acquisition, M.A.; Methodology, A.N.A., M.K.K., M.A., R.P. and H.S.; Project administration, M.A., R.P. and H.S.; Software, A.N.A., M.K.K. and R.P.; Supervision, M.A., R.P. and H.S.; Validation, M.K.K., M.A., R.P. and H.S.; Visualization, A.N.A., M.K.K. and R.P.; Writing—original draft, A.N.A. and M.K.K.; Writing—review and editing, A.N.A., M.K.K., M.A., R.P. and H.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge the support of King Fahd University of Petroleum & Minerals and the Interdisciplinary Research Center for Sustainable Energy Systems (IRC-SES).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

Author Rohit Patwardhan and Hussain Salloum are employed by the company Saudi Aramco. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Appendix A

The figures provided herein illustrate the performance of each model, NARX−NN, KL and KB, across various cases. For each case, the predicted outputs ( y 1 , y 2 , y 3 ) are compared against the actual values to evaluate the accuracy and robustness of the models.
Figure A1. Case 1—predicted vs. actual output y 1 for each model.
Figure A1. Case 1—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a1
Figure A2. Case 1—predicted vs. actual output y 2 for each model.
Figure A2. Case 1—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a2
Figure A3. Case 1—predicted vs. actual output y 3 for each model.
Figure A3. Case 1—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a3
Figure A4. Case 2—predicted vs. actual output y 1 for each model.
Figure A4. Case 2—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a4
Figure A5. Case 2—predicted vs. actual output y 2 for each model.
Figure A5. Case 2—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a5
Figure A6. Case 2—predicted vs. actual output y 3 for each model.
Figure A6. Case 2—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a6
Figure A7. Case 3—predicted vs. actual output y 1 for each model.
Figure A7. Case 3—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a7
Figure A8. Case 3—predicted vs. actual output y 2 for each model.
Figure A8. Case 3—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a8
Figure A9. Case 3—predicted vs. actual output y 3 for each model.
Figure A9. Case 3—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a9
Figure A10. Case 4—predicted vs. actual output y 1 for each model.
Figure A10. Case 4—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a10
Figure A11. Case 4—predicted vs. actual output y 2 for each model.
Figure A11. Case 4—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a11
Figure A12. Case 4—predicted vs. actual output y 3 for each model.
Figure A12. Case 4—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a12
Figure A13. Case 5—predicted vs. actual output y 1 for each model.
Figure A13. Case 5—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a13
Figure A14. Case 5—predicted vs. actual output y 2 for each model.
Figure A14. Case 5—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a14
Figure A15. Case 5—predicted vs. actual output y 3 for each model.
Figure A15. Case 5—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a15
Figure A16. Case 6—predicted vs. actual output y 1 for each model.
Figure A16. Case 6—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a16
Figure A17. Case 6—predicted vs. actual output y 2 for each model.
Figure A17. Case 6—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a17
Figure A18. Case 6—predicted vs. actual output y 3 for each model.
Figure A18. Case 6—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a18
Figure A19. Case 7—predicted vs. actual output y 1 for each model.
Figure A19. Case 7—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a19
Figure A20. Case 7—predicted vs. actual output y 2 for each model.
Figure A20. Case 7—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a20
Figure A21. Case 7—predicted vs. actual output y 3 for each model.
Figure A21. Case 7—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a21
Figure A22. Case 8—predicted vs. actual output y 1 for each model.
Figure A22. Case 8—predicted vs. actual output y 1 for each model.
Algorithms 17 00368 g0a22
Figure A23. Case 8—predicted vs. actual output y 2 for each model.
Figure A23. Case 8—predicted vs. actual output y 2 for each model.
Algorithms 17 00368 g0a23
Figure A24. Case 8—predicted vs. actual output y 3 for each model.
Figure A24. Case 8—predicted vs. actual output y 3 for each model.
Algorithms 17 00368 g0a24

References

  1. Achaw, O.-W.; Danso-Boateng, E. Crude Oil Refinery and Refinery Products; Springer International Publishing: Cham, Switzerland, 2021; pp. 235–265. [Google Scholar]
  2. Nanovsky, S. The impact of oil prices on trade. Rev. Int. Econ. 2019, 27, e0001. [Google Scholar]
  3. Mahecha, C.A.; López, D.C.; Hoyos, L.J.; Acevedo, L.; Villamizar, J.F. Optimization model of a system of crude oil distillation units with heat integration and metamodeling. CT&F-Cienc. Tecnol. Futuro 2009, 3, 159–173. [Google Scholar]
  4. López, C.D.C.; Hoyos, L.J.; Mahecha, C.A.; Arellano-Garcia, H.; Wozny, G. Optimization model of crude oil distillation units for optimal crude oil blending and operating conditions. Ind. Eng. Chem. Res. 2013, 52, 12993–13005. [Google Scholar]
  5. Franzoi, R.; Menezes, B.; Kelly, J.; Gut, J.; Grossmann, I. Cutpoint temperature surrogate modeling for distillation yields and properties. Ind. Eng. Chem. Res. 2020, 59, 18616–18628. [Google Scholar]
  6. H’ng, S.X.; Ng, L.Y.; Ng, D.K.S.; Andiappan, V. Optimisation of vacuum distillation units in oil refineries using surrogate models. Process. Integr. Optim. Sustain. 2024, 8, 351–373. [Google Scholar]
  7. Yao, H.; Chu, J. Operational optimization of a simulated atmospheric distillation column using support vector regression models and information analysis. Chem. Eng. Res. Des. 2012, 90, 2247–2261. [Google Scholar]
  8. Liau, L.C.-K.; Yang, T.C.-K.; Tsai, M.-T. Expert system of a crude oil distillation unit for process optimization using neural networks. Expert Syst. Appl. 2004, 26, 247–255. [Google Scholar]
  9. Motlaghi, S.; Jalali, F.; Ahmadabadi, M.N. An expert system design for a crude oil distillation column with the neural networks model and the process optimization using genetic algorithm framework. Expert Syst. Appl. 2008, 35, 540–1545. [Google Scholar]
  10. Gueddar, T.; Dua, V. Novel model reduction techniques for refinery-wide energy optimisation. Spec. Issue Therm. Energy Manag. Process. Ind. 2012, 89, 117–126. [Google Scholar]
  11. Durrani, M.A.; Ahmad, I.; Kano, M.; Hasebe, S. An artificial intelligence method for energy efficient operation of crude distillation units under uncertain feed composition. Energies 2018, 11, 2993. [Google Scholar] [CrossRef]
  12. Ochoa-Estopier, L.M.; Jobson, M.; Smith, R. Operational optimization of crude oil distillation systems using artificial neural networks. In Proceedings of the ESCAPE-22 (European Symposium on Computer Aided Process Engineering-22), London, UK, 17–20 June 2012; Volume 59, pp. 178–185. [Google Scholar]
  13. Ochoa-Estopier, L.M.; Jobson, M. Optimization of heat-integrated crude oil distillation systems. part i: The distillation model. Ind. Eng. Chem. Res. 2015, 54, 4988–5000. [Google Scholar]
  14. Ochoa-Estopier, L.M.; Jobson, M. Optimization of heat-integrated crude oil distillation systems. part iii: Optimisation framework. Ind. Eng. Chem. Res. 2015, 54, 5018–5036. [Google Scholar]
  15. Shi, B.; Yang, X.; Yan, L. Optimization of a crude distillation unit using a combination of wavelet neural network and line-up competition algorithm. Chin. J. Chem. Eng. 2017, 25, 1013–1021. [Google Scholar]
  16. Zhang, Y.; Cui, Z.; Wang, M.; Liu, B.; Fan, X.; Tian, W. An energy-efficiency prediction method in crude distillation process based on long short-term memory network. Processes 2023, 11, 1257. [Google Scholar] [CrossRef]
  17. Ibrahim, D.; Jobson, M.; Li, J.; Guillén-Gosálbez, G. Optimization-based design of crude oil distillation units using surrogate column models and a support vector machine. Chem. Eng. Res. Des. 2018, 134, 212–225. [Google Scholar]
  18. Li, S.; Zheng, Y.; Li, S.; Huang, M. Knowledge-based operation optimization of a distillation unit integrating feedstock property considerations. Eng. Appl. Artif. Intell. 2022, 107, 104496. [Google Scholar]
  19. Osuolale, F.N.; Zhang, J. Energy efficient control and optimisation of distillation column using artificial neural network. Chem. Eng. Trans. 2014, 39, 37–42. [Google Scholar]
  20. Muhsin, W.; Zhang, J. Multi-objective optimization of a crude oil hydrotreating process with a crude distillation unit based on bootstrap aggregated neural network models. Processes 2022, 10, 1438. [Google Scholar] [CrossRef]
  21. Zhu, J.; Fan, C.; Yang, M.; Qian, F.; Mahalec, V. Data-driven models of crude distillation units for production planning and for operations monitoring. Comput. Chem. Eng. 2023, 177, 108322. [Google Scholar]
  22. Mowbray, M.; Vallerio, M.; Perez-Galvan, C.; Zhang, D.; Chanona, A.D.R.; Navarro-Brull, F.J. Industrial data science—A review of machine learning applications for chemical and process industries. React. Chem. Eng. 2022, 7, 1471–1509. [Google Scholar]
  23. fadzil, M.A.M.; Razali, A.A.; Zabiri, H. Machine learning-based modeling and optimization analysis for an integrated industrial base oil production complex. Ind. Eng. Chem. Res. 2023, 62, 20280–20299. [Google Scholar]
  24. Brunton, S.L.; Budišić, M.; Kaiser, E.; Kutz, J.N. Modern koopman theory for dynamical systems. SIAM Rev. 2022, 64, 229–340. [Google Scholar]
  25. Koopman, B.O. Hamiltonian systems and transformation in hilbert space. Proc. Natl. Acad. Sci. USA 1931, 17, 315–318. [Google Scholar] [PubMed]
  26. Mezić, I.; Banaszuk, A. Comparison of systems with complex behavior. Phys. D Nonlinear Phenom. 2004, 197, 101–133. [Google Scholar]
  27. Mezić, I. Spectral properties of dynamical systems, model reduction and decompositions. Nonlinear Dyn. 2005, 41, 309–325. [Google Scholar]
  28. Gholaminejad, T.; Khaki-Sedigh, A. Stable deep koopman model predictive control for solar parabolic-trough collector field. Renew. Energy 2022, 198, 492–504. [Google Scholar]
  29. Susuki, Y.; Mezic, I. Nonlinear koopman modes and coherency identification of coupled swing dynamics. IEEE Trans. Power Syst. 2011, 26, 1894–1904. [Google Scholar]
  30. Abraham, I.; Torre, G.D.L.; Murphey, T.D. Model-based control using koopman operators. arXiv 2017, arXiv:1709.01568. [Google Scholar]
  31. Yu, S.; Shen, C.; Ersal, T. Autonomous driving using linear model predictive control with a koopman operator based bilinear vehicle model. IFAC-PapersOnLine 2022, 55, 254–259. [Google Scholar]
  32. Sootla, A.; Ernst, D. Pulse-based control using koopman operator under parametric uncertainty. arXiv 2017, arXiv:1708.00232. [Google Scholar]
  33. Avila, A.M.; Mezić, I. Data-driven analysis and forecasting of highway traffic dynamics. Nat. Commun. 2020, 11, 2090. [Google Scholar] [PubMed]
  34. Billings, S.A. Models for Linear and Nonlinear Systems; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2013; Chapter 2; pp. 17–59. [Google Scholar]
  35. Hornik, K.; Stinchcombe, M.; White, H. Multilayer feedforward networks are universal approximators. Neural Netw. 1989, 2, 359–366. [Google Scholar]
  36. Manonmani, A.; Thyagarajan, T.; Elango, M.; Sutha, S. Modelling and control of greenhouse system using neural networks. Trans. Inst. Meas. Control. 2018, 40, 918–929. [Google Scholar]
  37. Jaleel, E.A.; Aparna, K. Identification of realistic distillation column using hybrid particle swarm optimization and NARX based artificial neural network. Evol. Syst. 2019, 10, 149–166. [Google Scholar]
  38. Heidari, E.; Daeichian, A.; Sobati, M.A.; Movahedirad, S. Prediction of the droplet spreading dynamics on a solid substrate at irregular sampling intervals: Nonlinear auto-regressive exogenous artificial neural network approach (narx-ann). Chem. Eng. Res. Des. 2020, 156, 263–272. [Google Scholar]
  39. Zilio, A.; Zuanna, F.D.; Biadene, D.; Caldognetto, T.; Mattavelli, P. On the design of narx-anns for the black-box modeling of power electronic converters. In Proceedings of the 2023 IEEE Energy Conversion Congress and Exposition (ECCE), Nashville, TN, USA, 29 October–2 November 2023; pp. 2776–2782. [Google Scholar]
  40. Mauroy, A.; Mezić, I.; Susuki, Y. (Eds.) The Koopman Operator in Systems and Control: Concepts, Methodologies, and Applications; Lecture Notes in Control and Information Sciences; Springer International Publishing: Cham, Switzerland, 2020; Volume 484. [Google Scholar]
  41. Proctor, J.; Brunton, S.; Kutz, J. Generalizing koopman theory to allow for inputs and control. Appl. Dyn. Syst. 2018, 17, 909–930. [Google Scholar]
  42. Liu, Z.; Kundu, S.; Chen, L.; Yeung, E. Decomposition of nonlinear dynamical systems using koopman gramians. In Proceedings of the 2018 Annual American Control Conference (ACC), Milwaukee, WI, USA, 27–29 June 2018; pp. 4811–4818. [Google Scholar]
  43. Surana, A. Koopman operator based observer synthesis for control-affine nonlinear systems. In Proceedings of the 2016 IEEE 55th Conference on Decision and Control (CDC), Las Vegas, NV, USA, 12–14 December 2016; pp. 6492–6499. [Google Scholar]
  44. Goswami, D.; Paley, D.A. Global bilinearization and controllability of control-affine nonlinear systems: A Koopman spectral approach. In Proceedings of the 2017 IEEE 56th Annual Conference on Decision and Control (CDC), Melbourne, VIC, Australia, 12–15 December 2017; pp. 6107–6112. [Google Scholar]
  45. Jones, D.S.J. Atmospheric and Vacuum Crude Distillation Units in Petroleum Refineries; Springer International Publishing: Cham, Switzerland, 2015; pp. 125–198. [Google Scholar]
  46. Fraser, S. Chapter 4—Distillation in refining. In Distillation; Górak, A., Schoenmakers, H., Eds.; Academic Press: Boston, MA, USA, 2014; pp. 155–190. [Google Scholar]
Figure 1. The NARX−NN model representation.
Figure 1. The NARX−NN model representation.
Algorithms 17 00368 g001
Figure 2. Koopman dynamics based on an NN.
Figure 2. Koopman dynamics based on an NN.
Algorithms 17 00368 g002
Figure 3. Crude distillation unit diagram.
Figure 3. Crude distillation unit diagram.
Algorithms 17 00368 g003
Figure 4. Summry of MSE values on the Case 1 dataset.
Figure 4. Summry of MSE values on the Case 1 dataset.
Algorithms 17 00368 g004
Figure 5. Summary of MSE values on the Case 2 dataset.
Figure 5. Summary of MSE values on the Case 2 dataset.
Algorithms 17 00368 g005
Figure 6. Summary of MSE values on the Case 3 dataset.
Figure 6. Summary of MSE values on the Case 3 dataset.
Algorithms 17 00368 g006
Figure 7. Summary of MSE values on the Case 4 dataset.
Figure 7. Summary of MSE values on the Case 4 dataset.
Algorithms 17 00368 g007
Figure 8. Summary of MSE values on the Case 5 dataset.
Figure 8. Summary of MSE values on the Case 5 dataset.
Algorithms 17 00368 g008
Figure 9. Summary of MSE values on the Case 6 dataset.
Figure 9. Summary of MSE values on the Case 6 dataset.
Algorithms 17 00368 g009
Figure 10. Summary of MSE values on the Case 7 dataset.
Figure 10. Summary of MSE values on the Case 7 dataset.
Algorithms 17 00368 g010
Figure 11. Summary of MSE values on the Case 8 dataset.
Figure 11. Summary of MSE values on the Case 8 dataset.
Algorithms 17 00368 g011
Table 1. List of inputs and outputs.
Table 1. List of inputs and outputs.
Inputs
u 1 Column reflux
u 2 Middle draw
u 3 Column bottom tray temperature
dFeed to column
Outputs
y 1 Overhead composition
y 2 Middle-draw composition
y 3 Bottoms’ composition
Table 2. Description of each experimental condition.
Table 2. Description of each experimental condition.
Case #Description
1No mismatch with aggressive tuning
2Gain mismatch, u 1 y 1 , u 1 y 2
3Gain mismatch, u 2 y 2
4Delay mismatch, u 1 y 1 , y 2 , y 3
5Poor PID tuning for all three MVs
6Sqrt-type nonlinearity in u 3 y 1 , y 2 , y 3 channels
7Tight constraints on u 1 , u 2 , u 3
8Unmeasured disturbance affecting y 1 , y 2 , y 3 equally
Table 3. Hyperparameter search space.
Table 3. Hyperparameter search space.
ModelHyperparameterRange
KL / KBNetwork width[16, 256]
Network depth[1, 32]
No. of lifted dynamics[4, 64]
NARXNetwork width[4, 256]
Network depth[1, 32]
Delay[1, 4]
Table 4. Hyperparameters tuning results.
Table 4. Hyperparameters tuning results.
HyperparameterNARXKLKB
Network width121622
Network depth1810
Delay2//
No. of lifted dynamics/4264
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Abubakar, A.N.; Khaldi, M.K.; Aldhaifallah, M.; Patwardhan, R.; Salloum, H. Identification of Crude Distillation Unit: A Comparison between Neural Network and Koopman Operator. Algorithms 2024, 17, 368. https://doi.org/10.3390/a17080368

AMA Style

Abubakar AN, Khaldi MK, Aldhaifallah M, Patwardhan R, Salloum H. Identification of Crude Distillation Unit: A Comparison between Neural Network and Koopman Operator. Algorithms. 2024; 17(8):368. https://doi.org/10.3390/a17080368

Chicago/Turabian Style

Abubakar, Abdulrazaq Nafiu, Mustapha Kamel Khaldi, Mujahed Aldhaifallah, Rohit Patwardhan, and Hussain Salloum. 2024. "Identification of Crude Distillation Unit: A Comparison between Neural Network and Koopman Operator" Algorithms 17, no. 8: 368. https://doi.org/10.3390/a17080368

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop