Next Article in Journal
Integration and Evaluation of a Digital Support Function for Space Claims in Factory Layout Planning
Previous Article in Journal
Application of NACA 6412 Airfoil for Noise and Vibration Reduction in Evaporator Fan Blades
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multi-Objective Optimization and Design for Industrial Vinyl Chloride Reactor by Hybrid Model

School of Chemical Sciences, University of Chinese Academy of Sciences, Beijing 101408, China
*
Authors to whom correspondence should be addressed.
Processes 2024, 12(11), 2378; https://doi.org/10.3390/pr12112378
Submission received: 11 October 2024 / Revised: 24 October 2024 / Accepted: 27 October 2024 / Published: 29 October 2024
(This article belongs to the Section Process Control and Monitoring)

Abstract

:
The acetylene conversion rate and vinyl chloride production capacity are the main economic indexes for vinyl chloride synthesis, and the reaction temperature is an important operating parameter to prevent the Hg active component from being loss. These three factors have not been taken into consideration simultaneously in the traditional optimization process, making it difficult to achieve optimization targets perfectly for industrial application. To overcome this problem, an efficient strategy framework was proposed based on a hybrid model. Compared with conventional paradigms, the proposed framework could not only reduce computational expense but optimize these two economic indexes with a constrained reaction temperature simultaneously. In addition, a machine learning method was used to conduct a feature analysis, which can reveal the potential interaction between different variables so key variables of this reactor could be identified. To demonstrate and verify this framework, multi-objective optimization involving multiple variables with constrained conditions for the industrial reactor was conducted from design and operation perspectives, respectively. The proposed strategy could provide optimal operational direction for the industrial reactor from these design and operation aspects, which contribute to the sustainable and highly efficient process development in this field. For the first section of an industrial vinyl chloride reactor, this strategy could realize significant improvement in the acetylene conversion rate from 81.34% to over 95.00% and in the vinyl chloride production capacity from 2.60 to above 3.40 mol/h in the operation scenarios, which can meet production requirements. So, the second section of the traditional reactor system is not needed at all.

1. Introduction

Polyvinyl chloride (PVC) is widely used in various fields for its excellent physical and chemical properties, including mechanical properties, electrical properties, and easy processability [1,2]. It is mainly produced from vinyl chloride monomer (VCM) by the ethylene method or the acetylene method [3]. In addition, VCM can also be used as a refrigerant to form copolymers via copolymerization with vinyl acetate, vinyl chloride, butadiene, acrylonitrile, acrylic esters, and other monomers. In China, the acetylene method is the most common process due to its abundant coal resource [4], and the most widely used catalysts in plants are mercury catalysts loaded on carbon support [5]. Because of its poor thermal stability, the HgCl2 component will gradually lose at elevated temperatures, reducing catalyst activity and increasing mercury consumption, leading to environmental and health problems [6]. At present, though some studies [7,8,9] have reported mercury-free catalysts, most of them are still in the laboratory stage, and there are few public reports about their industrial application. To optimize a vinyl chloride reactor (VCR), some mathematical models, such as the one-dimensional and two-dimensional pseudo-homogeneous model [10,11], were established to investigate the relations between input parameters and output values, showing that the two models could approximately reflect the industrial reactor. Liang [10] established one-dimensional (1D) and two-dimensional (2D) mathematical models, which showed certain accuracy and reliability compared with the actual industrial data from a factory. The effects of operational parameters on the reactor temperature and acetylene conversion rate (ACR) were also analyzed without optimization analysis and design. Chen [11] established mathematical models based on different types of reactors and used the verified models to conduct a simulation analysis on different factors (space velocity (SV), wall temperature (Tw), and tube diameter (dt)) and system optimization. Yang et al. [12] analyzed the effects of SV, coolant temperature (Tc) on the temperature distribution in the reactor, and ACR. Huang et al. [13] optimized VCR operational parameters via analyzing their effects on the ACR with Aspen Plus. This analysis was a univariate analysis during which the ACR was less than 85% under the optimal condition. These simulations and optimizations are mostly based on a single-variable sensitivity analysis or single-objective function, which could not reveal the underlying influential mechanisms for the multiple-input multiple-output (MIMO) system. In other words, the optimized results may be not the real optimal solutions. To enhance VCR performance, mercury-free catalysts and related reaction mechanisms were also investigated [14,15,16,17,18], including g-C3N4/BiOCl catalysts, single-atom AuI-N3, and Au/CeO2-based catalysts. Though those studies showed good performance, they are still in the laboratory stage and it will take a period of time to achieve industrialization. Currently, the acetylene method based on mercury catalysts is still a common process; thus, the multi-objective optimization of VCR is essential for the industrial production. The models used for the simulation of VCR are summarized in Table 1.
The modeling of the operation unit in the chemical engineering process includes a mechanistic model based on the specific knowledge of the process, a data-driven model (or artificial neural network (ANN), black-box model) based on the historical data from the process, and a gray model integrating mechanistic knowledge with the data-driven model. A mechanistic model, also known as the first principle model (FPM), is derived from the principles of mass, energy, and momentum balance and shows good interpretable and extrapolative performance, but it poses significant challenges to the knowledge of system-specific characteristics and computational capacity involving complex mathematical equations with nonlinearity and multidimensionality characteristics [19]. Data-driven models, such as ANN, can cope with the flaws of the FPM, whose accuracy depends on the quantity and quality of the training dataset. Presently, research on the ANN model for a VCM reactor is still in its infancy. To reduce computational load, this work first proposed a methodology of using a surrogate model instead of FPM to conduct multi-objective optimization for the industrial VCM reactor, where the trained data are obtained from simulation by the FPM as few industrial data are available.
In recent years, the development of machine learning (ML) and artificial intelligence (AI) and their powerful advantages in solving complex system problems have attracted the attention of chemical researchers and engineers [20,21,22,23]. The chemical engineering process, especially chemical reactors, usually involves MIMO coupled with nonlinear problems, which cannot be solved with high precision using traditional methods. With the aid of AI and ML, engineers are exploring new paradigms. Zhao et al. [24] conducted a multi-objective optimization of a radically stirred tank based on CFD and machine learning, which confirmed the accuracy and reliability of the machine learning-based optimization method. Rahimpour et al. [25] used a multilayer perception (MLP) neural network to determine the optimum production of ethylene dichloride, and the error of simulation was found to be less than 5%. Bhakte et al. [26] used a deep neural network (DNN) based on process alarms to assist operators in understanding DNN’s prediction during online process monitoring.
The annual production of vinyl chloride in the world is huge because of its market demand, and the investment and energy cost are very high, so even a small improvement in plant operation can lead to considerable benefits. Thus, there is a requirement of optimal design and operation for the vinyl chloride reactor. Particle swarm optimization (PSO) is a simple and yet powerful metaheuristic search algorithm, widely used to solve various optimization problems [27].
The monitoring, control, and optimization of industrial VC reactors are essential for the plant, but in general, optimization is based on a single-objective optimization, even ignoring the constrained maximum reactor temperature (Tmax) during the optimization process, which may be insufficient for this scenario. In addition, FPM needs more computational time compared with the ANN model, but prediction in advance is essential to the production, based on which timely management could be achieved. To cope with the above obstacles, we propose a new prediction model and optimization strategy for the industrial VC reactor. Firstly, a historical dataset obtained from actual production is used to conduct feature analysis based on machine learning, based on which a surrogate model for the reactor is built, trained, and verified. This surrogate model can be used to conduct the prediction, optimization, control, etc. Here, a surrogate model coupled with an artificial intelligent algorithm is used to conduct multi-objective optimization with constrained conditions. Detailed information on the optimization strategy is shown in Figure 1, which mainly includes three parts: data preprocessing (datasets collection, pretreatment, and feature analysis), predictive model selection, and calibration (the prediction model may be FPM, ANN, or a hybrid model, depending on the specific process and purposes). The development of this optimization strategy for vinyl chloride industrial reactors has important practical applications: (1) a safety production plan based on the calibrated model; (2) operational optimization based on a reliable surrogate and multi-objective optimization algorithm; and (3) the prediction of the reactor performance for designing effective control strategies.

2. Materials and Methods

Due to the fact that large and high-quality training data are needed and yet it is difficult to obtain sufficient historical data from the actual industrial production in machine learning, we used FPM to generate a large dataset replacing industrial historical data, which could reflect the effectiveness of this proposed idea. The methodology in this research consists of two main sections. In the first section, first principle models (FPMs) were verified using industrial data, and then a dataset generated with the calibrated FPMs was used to train and test the relevant surrogate models. In the second section, the calibrated models coupled with PSO were used to optimize the VCM reactor from design and operation aspects, respectively.

2.1. Mechanistic Model and Data Generation

At present, the acetylene method is most commonly used in China for the production of vinyl chloride monomer [28]. The main reaction of the acetylene method is as below [29].
C H C H + H C l C a t C H 2 = C H C l       Δ H = 124.8   k J / m o l
This study is based on a real industrial reactor, in which the catalyst is mercury chloride loaded on carbon support. Reactors for acetylene hydrochlorination are usually fixed beds with multiple tubes. Because the tube diameter (dt) is small and gas velocity is high, the axial dispersion inside the tubes can be neglected. By merging the effect of industrial catalysts and pore diffusion into the reaction kinetics, a homogeneous model is used instead of a heterogeneous model. In this study, one-dimensional and two-dimensional homogeneous mathematical models are developed, where reaction kinetics [30] is adopted.
One-dimensional model:
d x A d z = ( r A ) ( 1 ε ) F A 0 π d t 2 4
d T d Z = Δ H 4 r A 1 ε + U T w T π d t F t C p
Two-dimensional model:
x A z = a 1 2 x A r 2 + 1 r x A r + b 1 r A
T z = a 2 ( 2 T r 2 + 1 r T r ) + b 2 ( r A )
where:
a 1 = E r u , b 1 = ρ B u C A 0
a 2 = λ e r G C P ¯ , b 2 = ρ B Δ H R G C P ¯
r = k k H P A 0 2 a 3.6 ( 1 x A ) ( R M I x A ) ( P P A 0 x A ) 2 + k H P A 0 ( P P A 0 x A ) ( R M I x A )     mol/(s·kg-cat)
(RMI represents the mole ratio of HCl to C2H2)
where:
η = 57.0 ( T 300 ) 0.12 × e 0.1 x A
k = 25.6 × η × e 4065 R T
k H = 3.31 × 10 6 × e 10630 R T
The industrial reactor data [11] were used to verify the models’ accuracy.
Firstly, based on the above 1D and 2D mathematical models, two types of VCR simulation framework were built with the Python program. To compare with industrial data, the input parameters of the above two models should agree with the industrial VCR (dt = 0.044 m; RMI = 1.1; SV = 34.12 h−1; a = 0.209; Tin = 362.8 K), based on which VCR simulation calculations were conducted, where differential equations were solved using Runge–Kutta and difference methods. Table 2 shows that the results of the two FPMs agree well with the real industrial data, demonstrating that these two models are both credible; the values of Tmax, Tout, and xout are average values for the radial direction position at the axial position in the tube reactor. To simplify the simulation and reduce the calculation cost, the following work was based on the one-dimensional model. About 39,000 sets of data were generated via simulation, and feature analysis was firstly conducted based on Pearson correlation [31].
R = i = 1 n   ( x i X ¯ ) ( y i Y ¯ ) i = 1 n   ( x i X ¯ ) 2 i = 1 n   ( x i Y ¯ ) 2
Figure 2 shows the R between different variables, demonstrating that input variables have different effects on all three target variables. The absolute value of R represents the strength of the relationship between two variables, and the sign represents the effect tendency (positive or negative). For example, as to ACR(x), the effect order of input variables is a, l, Tw (temperature of tube wall), RMI, SV, dt, and Tin (inlet temperature), while as to vinyl chloride production capacity (VCPC, Pro), the order of influence is l, dt, SV, a, Tw, RMI, and Tin. Meanwhile, there exists a contrary correlation between the same input variables and different target variables, for example, there is a negative correlation (R = −0.14) between SV and ACR(x), and a positive correlation (R = 0.24) between SV and VCPC(Pro). Because SV decides the retention time of reactants in the reactor, an elevated SV means a shorter reaction period, and thus SV has a negative effect on the ACR, whereas SV contributes to Pro based on Equation (10), and thus the effect of SV on Pro from Figure 2 is a comprehensive result based on multiple input variables. Therefore, multi-objective optimization involving MIMO is necessary for the industrial reactor.

2.2. Surrogate Prediction Model

In this study, an ANN was employed to accurately predict target functions; its architecture contains an input layer, multiple hidden layers, and an output layer (Figure 3).
Before the ANN training phase, data pre-processing is needed. Here, min-max normalization [32] was used to deal with inputs, scaling the range of each input to [0,1], since the order of magnitude difference among these features will affect the prediction performance.
x = ( x m i n ) / ( m a x m i n )
where min and max represent the minimum and maximum value corresponding to input variables, respectively.
During the ANN training process, the acetylene conversion rate (xA), the maximum temperature in the reactor (Tmax), and vinyl chloride production (Pro) were selected as target functions. In the design process, feature variables were RMI, SV (/h), a, Tw (K), Tin (K), dt (m), and l (m), while in the operation process, dt and l were removed from the above feature variables. The optimization algorithm and loss function were Adam (Adaptive Moment Estimation) [33] and MSE (mean square error).
MSE = 1 n i = 1 n ( y i y p r e _ i ) 2
In the operation scenarios, except for the feature variables, the others were similar to those in the design scenarios. The inputs consisted of RMI, SV, a, Tw, and Tin. In this study, vinyl chloride production is expressed as shown below:
P r o = C A 0   *   x A   *   d t 2   *   π 4   *   S V   *   l   (mol/h)
where C A 0 represents acetylene mole concentration (mol/m3), Pro represents vinyl chloride production capacity per tube, and xA represents the conversion rate of acetylene.

2.3. Multi-Objective Optimization

To optimize the vinyl chloride industrial reactor, multi-objective optimization algorithm coupled with surrogate prediction model was constructed. Because the HgCl2 component will lose at elevated temperature, Tmax should be constrained in a range; in this study, the value was set to 423.15 K. PSO mimics the foraging behavior of a group of birds, where the birds are treated as particles and the foraging behavior is treated as a process of merit-seeking in a particle swarm. Each particle has its own direction and velocity, and the particle position is the decision variable for the optimization problem. The velocity and position of the particles are updated as follows [34]:
V i = w × V i + c 1 × r a n d 0,1 × P b e s t X i + c 1 × r a n d 0,1 × g b e s t X i
X i = X i + V i
where P b e s t and g b e s t represent the particle’s historical optimal position and the optimal position sought by all particles, respectively, both of which control the direction of motion and position of the particles. PSO can also be used for multi-objective optimization problems, for which not a single optimal solution but a series of optimal solutions exist. A set of nondominant solutions, not dominated by others in the solution, is called the Pareto fronts solution. In the optimization process for the vinyl chloride reactor, the optimized goals involve two variables, which are multi-objective optimization problems; the detailed expression is as shown below:
max f(X) = [xA,Pro]s.t Tmax ≤ 423.15
The numerical solution and the optimization algorithm were both implemented through Python programming language.

3. Results and Discussion

3.1. Surrogate Model Training and Validation

A dataset containing 390,625 sets of data was obtained from the simulation based on the one-dimensional model. Given that the dataset came from simulation, data cleansing was ignored, which is necessary for industrial datasets. First, data preprocessing was conducted, where the values of all feature variables were scaled to the range of 0 to 1. Then, the dataset was divided into training and testing sets; here, the testing set size was 20% of the dataset. This is a MIMO problem, and thus a parallel ANN structure is needed. Here, a parallel ANN consisting of three sub-ANNs was constructed, where all sub-ANNs share inputs. The number of ANN layers, activation functions, and nodes are essential parameters, directly influencing prediction accuracy.
Based on Section 2, the surrogate prediction models of the VCM reactor in respect to the design and operation scenarios were trained using the training dataset. In the design scenarios, for xA, the ANN structure consisted of ReLU of activation function, one hidden layer, 100 neural nodes, and a learning_rate of 0.005, while those for Tmax and Pro were ReLu, one, 600, and 0.0001; and ReLU, one, 200, and 0.0001, respectively. Similarly, in the operation scenarios, those for xA, Tmax, and Pro were ReLu, one, 80, and 0.0005; ReLu, one, 600, and 0.0008; and ReLu, one, 300, and 0.00008. To further test the prediction accuracy of these two models, the outputs between FPM and the surrogate model were compared (Table 3 and Table 4).
Five sets of inputs were chosen randomly to further verify the accuracy of the surrogate models in design and operation scenarios (Table 3 and Table 4). Table 3 shows that the prediction data from the surrogate models agreed well with those from FPM, and all relative deviations for these three outputs were below 1.5%. In addition, the accuracy of the surrogate models in operation scenarios was also verified, where all relative deviations for these three outputs were lower than 1%. Therefore, the trained models above can be used to optimize the VCM reactor with high fidelity.

3.2. Sensitivity Analysis of Input Variables

In this section, the single variable method was adopted to determine the range of input variables. For example, to determine the range of RMI, all other variables were kept constant, and SV, a, dt, Tw, l, and Tin were set to 34.13 h−1, 0.209, 0.044 m, 372.2 K, 3.0 m, and 362.2 K, respectively. The results of RMI, SV, a, dt, Tw, Tin, and l are shown in Figure 4, Figure 5, Figure 6, Figure 7, Figure 8, Figure 9 and Figure 10.
Figure 4 shows that the ultimate mole conversion rate of acetylene (Pro_xa) and Pro first increase and then decrease with increasing RMI values (see the left graph). The reason for this is that the term of (m-xA) dominates the value of rA when RMI is less than 1.2, while the term of PA0 dominates rA when RMI is greater than 1.2 (see Equation (6)). In addition, Tmax decreases with RMI increment (see the right graph), because more hydrogen chloride will share the effect of reaction heat on the reaction temperature. Thus, in view of operation cost and production efficiency, the range of RMI was chosen to be from 1.01 to 1.20.
Figure 5 shows that Pro_xa decreases with SV increasing, while Pro trends downward (see the left graph). An increasing SV means a rising linear velocity of C2H2 and an increasing mass flux of C2H2, leading to increased Pro. The reason for the different variation trend in Pro is the shortened reaction time at an elevated SV. In addition, an elevated linear velocity will improve the total thermal conductive coefficient, and thus Tmax will decrease with increasing SV (see the right graph). Thus, considering Pro_xa and production efficiency, the range of SV was chosen to be 33−40 h−1.
Catalysis activity directly affects the catalytic reaction rate; its sensitivity analysis is shown in Figure 6. It was found that Pro_xa, Pro, and Tmax all increase with catalysis activity increasing, because a directly contributes to the reaction rate, which will accelerate the exothermic reaction, thus leading to a reactor temperature increase. Due to the dependance of catalyst properties on the reaction temperature, the range of a was chosen as 0.2 to 0.5.
Figure 7 shows that Pro_xa, Pro, and Tmax all increase with increasing dt. Because a rising dt reduces the total thermal conductivity and weakens the heat removal via coolants, Tmax increases (right graph). Now that the reaction rate correlates with Tmax, Pro_xa increases with increasing dt whereas a high Tmax will lower the catalyst performance, so the range of dt was set as 0.04 to 0.05 m.
As Tw represents the heat removal capacity via the tube wall, a higher Tw means less heat which can be removed and an increased reaction temperature. As shown in Figure 8 regarding the effect of Tw on Pro_xa, Pro, and Tmax, the reliance of catalyst properties on temperature corresponds with a range of Tw from 371 to 393 K.
The reactor tube length affects the residence time of the reactants, and thus increasing the tube length (l) contributes to Pro_xa and Pro. Though the heat generated by the reaction increases with l increasing, the heat removal area also expands. The overall result is that Tmax changes little when l is varied in the range of 2.6 to 5.0 m. Taking the operation cost and equipment occupation space into consideration, the range of l was chosen to be from 3 to 5 m. In addition, the result from the sensitivity analysis indicates that Tin varying from 360 to 380 K has a small effect on Pro_xa, Pro, and Tmax. Taking into account the industrial condition, the range of Tin was chosen to be from 362.2 to 373.2 K.

3.3. Optimization in the Design Scenarios

In the design scenarios, equipment parameters and process parameters are adjustable and they all can be selected as feature variables except for the operation pressure, according to Section 2.1. The ranges of RMI, a, SV, Tw, dt, l, and Tin are listed in Table 5.
The Pareto fronts obtained using the PSO algorithm for the vinyl chloride industrial reactor in the design scenarios are marked in blue in Figure 11. There is a contrary changing trend between xA and Pro with constrained Tmax in the tube reactor. In the analysis ranges for all variables, Pro can approach 6.81 mol h−1 when xA is about 0.986, while a Pro value of about 5.99 mol h−1 gives rise to an xA value of 0.990. In practice, a trade-off between Pro and xA is essential according to the commodity market, and appropriate input conditions from the Pareto fronts should be chosen. For example, if 0.98 satisfies the upstream and downstream requirements, the input condition can choose the leftmost set, where Pro is highest in the range. Some Pareto fronts and their corresponding inputs are shown in Table 6.

3.4. Optimization in the Operation Scenarios

In the operation scenarios, equipment parameters are fixed (dt = 0.044 m; l = 3.0 m), and thus they should be removed from the feature variables. The ranges of RMI, a, SV, Tw, and Tin are shown in Table 7.
The Pareto fronts obtained using the PSO algorithm for the vinyl chloride industrial reactor in the operation scenarios are marked in blue in Figure 12. There is a contrary changing trend between xA and Pro with constrained Tmax in the tube reactor. In the analysis ranges for all variables, Pro can approach 3.47 mol h−1 when xA is about 0.952, while xA can approach 0.953 when Pro is about 3.42 mol h−1. In practice, industries should make a trade-off between Pro and xA according to market requirement, choosing the appropriate input conditions from the Pareto fronts. For example, if 0.95 satisfies the upstream and downstream needs, the leftmost set can be chosen for the input condition, where Pro is highest in the range. Some Pareto fronts and their corresponding inputs are shown in Table 8.

3.5. Comparison Between Single-Objective and Multi-Objective Optimization

As mentioned in the introduction section, the optimization of VCR mainly focuses on single-objective optimization without taking all related variables into consideration simultaneously. To further verify the advantages of the proposed multi-objective optimization frame over conventional paradigms, single-objective optimization involving multiple inputs with a constrained condition of Tmax was conducted, in which the ranges of different feature variables were the same as in multi-objective optimization. The results of the design scenario are shown in Figure 13. It can be seen that xA could approach 99.4% with the following parameters: dt 0.045 m, l 5.0 m, RMI 1.20, SV 33.0 h−1, a 0.417, Tw 372.3 K, and Tin 362.8 K. In order to verify the optimum performance, the just-mentioned parameters were substituted into the model, showing a Pro of 5.36 mol/h. In contrast, Pro could approach 5.76 mol/h or above with an xA of 98.5% using the multi-objective optimization proposed in this work. This comparison demonstrates the obvious advantage of multi-objective optimization over single-objective optimization for industrial VCR, and the former adapts better to industrial demand.

4. Conclusions

The optimization of vinyl chloride industrial reactors faces multi-objective optimization issues, and to maintain its optimal operational performance, an efficient optimization control strategy based on machine learning and hybrid model was proposed. The trained and verified surrogate model could be used to monitor and predict the reactor operation performance timely and accurately. A multi-objective optimization algorithm coupled with a prediction model could provide optimal operational direction for the industrial reactor, which contributes to sustainable development in this field and can even be extended to relative fields.
Taking the commonly used first section of a VC reactor as an example, the adoption of this optimization strategy could realize significant improvement for ACR and VCPC, rising from 81.34% to over 95% for ACR and from 2.60 to over 3.40 mol h−1 for VCPC in the operation scenarios based on the reference value obtained from industrial data. For example, with the parameters of the first section being dt 0.044 m, l 3.0 m, RMI 1.20, SV 38.73 h−1, a 0.336, Tw 385 K, and Tin 369.5 k, ACR and VCP could approach 95.26% and 3.44 mol/h, respectively, which are exactly the same as those of the traditional two-section reactor. So, in the future, the first section could be used alone for industrial production, with large savings in device investment.
The surrogate model can be a data-driven model or hybrid model. This work chooses a data-driven model owing to insufficient available industrial data, thus only demonstrating the practical application for the proposed strategy. In addition, an ML model can also be used to substitute FPM to compensate for the drawbacks of the latter in monitoring and control, fault diagnosis, etc., in complex chemical engineering industry. Finally, it is generally true that a pure ML model only has sufficient fidelity in the range of the training dataset, which is reliable for industrial processes. If prediction accuracy and interpretability are to be further enhanced, a hybrid model is necessary. The combination of ANN with FPMs becomes powerful as ANN can compensate for the gap between FPMs and real industry while FPMs can enhance the interpretability and extrapolation of surrogate ANN models.

Author Contributions

Conceptualization, E.S.; methodology, E.S.; software, X.Z.; validation, E.S.; formal analysis, M.W.; investigation, E.S.; resources, E.W.; data curation, M.W.; writing—original draft preparation, E.S.; writing—review and editing, E.W. and C.-Y.G.; visualization, X.Z.; supervision, E.W. and C.-Y.G.; project administration, E.W.; funding acquisition, E.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Beijing Natural Science Foundation (2222037); the Special Educating Project of the Talent for Carbon Peak and Carbon Neutrality of University of Chinese Academy of Sciences (Innovation of talent cultivation model for “dual carbon” in chemical engineering industry, E3E56501A2); and the “Low-carbon transformation technologies and demonstrations in chemical engineering and metallurgical processing”, Strategic Priority Research Program of the Chinese Academy of Sciences (XDA0390000).

Data Availability Statement

The data and code that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

xAconversion of acetylene (-)
rreaction rate, mol/(s·kg-cat)
rAreaction rate of acetylene, mol/(s·m3)
εbed voidage (-)
dtinner diameter of tube, m
FA0inlet molar flow of acetylene, mol/s
ΔHheat of reaction, J/mol
Ttemperature of reaction mixture reactor, K
Uoverall heat transfer coefficient, J/(m2·s·K)
Twtemperature of tube wall, K
Fttotal molar flow of reaction mixture, mol/s
Cpspecific heat capacity of reaction mixture, J/(mol·K)
Erradial diffusion coefficient, m2/s
ulinear velocity of acetylene, m/s
ρBbulk density of catalyst bed, kg/m3
λerradial effective thermal conductivity of the bed, w/(m·k)
Gtotal mass rate, kg/(m2·s)
C P ¯ specific heat capacity of reaction mixture, J/(kg·K)
Pinlet pressure of the fixed bed, atm
PA0partial pressure of acetylene, atm
acatalyst activity (-)
ηeffective coefficient of the catalyst (-)
kreaction rate constant (-)
kHadsorption equilibrium constant of HCL(-)

References

  1. Li, D.; Li, P. Trends and prospects for thermal stabilizers in polyvinyl chloride. J. Vinyl Addit. Technol. 2022, 28, 669–688. [Google Scholar] [CrossRef]
  2. Meng, X.; Li, W.Y.; Tian, Y.X.; Sun, C.Y.; Huang, J.; Liu, X.F. Synthesis of trioctyl polyvinyl chloride ammonium, membrane extraction properties, and electrodriven mass transfer behavior of Chromium (VI). Sep. Purif. Technol. 2023, 320, 124191. [Google Scholar] [CrossRef]
  3. Li, F.J. Design and analysis of production process of polyvinyl chloride by acetylene method. Shanxi Chem. Ind. 2022, 204, 66–67+80. (In Chinese) [Google Scholar] [CrossRef]
  4. Xue, X.C.; Zhou, C.X.; Ding, J. Progress in preparation of vinyl chloride. Chem. Enterp. Manag. 2017, 26, 52. [Google Scholar]
  5. Li, J.; Fan, J.T.; Ali, S.; Lan, G.J.; Tang, H.D.; Han, W.F.; Liu, H.Z.; Li, B.; Li, Y. The origin of the extraordinary stability of mercury catalysts on the carbon support: The synergy effects between oxygen groups and defects revealed from a combined experimental and DFT study. Chin. J. Catal. 2019, 40, 141–146. [Google Scholar] [CrossRef]
  6. Li, C.H.; Zhou, J.; Yu, X.Y.; Li, X.H. Summary on the running of converter with circulating fluidized bed for the synthesis of 3000 t/a vinyl chloride. Polyvinyl Chloride 2008, 1, 35–37. [Google Scholar]
  7. Jin, Y.H.; Li, G.B.; Zhang, J.L.; Pu, Y.F.; Li, W. Effects of potassium additive on the activity of Ru catalyst for acetylene hydrochlorination. RSC Adv. 2015, 5, 37774–37779. [Google Scholar] [CrossRef]
  8. Gu, J.J.; Gao, Y.M.; Zhang, J.L.; Li, W.; Dong, Y.Z.; Han, Y. Hydrochlorination of acetylene catalyzed by an activated carbon-supported ammonium hexachlororuthenate complex. Catalysts 2017, 12, 17. [Google Scholar] [CrossRef]
  9. Dong, X.B.; Chao, S.L.; Wan, F.F.; Guan, Q.X.; Wang, G.H.; Li, W. Sulfur and nitrogen co-doped mesoporous carbon with enhanced performance for acetylene hydrochlorination. J. Catal. 2018, 359, 161–170. [Google Scholar] [CrossRef]
  10. Liang, Z.Q. Modeling and analysis of fixed bed reactor for vinyl chloride synthesized from acetylene method. Master’s Thesis, Qingdao University of Science and Technology, Qingdao, China, 2014. [Google Scholar]
  11. Chen, J. Study on Simulation and Optimization of Vinyl Chloride Reactors from Hydrochlorination of Acetylene Process; Zhejiang University: Hangzhou, China, 2012. [Google Scholar]
  12. Yang, X.; Liang, Z.; Lu, W.; Zheng, S. Modeling and analysis of fixed bed reactor for vinyl chloride synthesized from acetylene method. Comput. Appl. Chem. 2015, 32, 115–118. [Google Scholar]
  13. Huang, L.; Xu, Z.; Zhu, M.; Wang, X.; Dai, B. Mercury-free catalytic kinetics of acetylene hydrochlorination and its reactor optimization. J. Shihezi Univ. (Nat. Sci.) 2021, 39, 529–536. [Google Scholar]
  14. Zhao, Z.; Wang, H.; Li, J.; Qiao, X.; Liu, Z.; Ren, Z.; Yuan, M.; Zhang, J. Photocatalytic acetylene hydrochlorination by pairing proton reduction and chlorine oxidation over g-C3N4/BiOCl catalysts. Am. Chem. Soc. 2024. Online ahead of print. [Google Scholar]
  15. Zhong, J.; Xu, Y.; Liu, Z. Heterogeneous non-mercury catalysts for acetylene hydrochlorination: Progress, challenges, and opportunities. Green Chem. 2018, 20, 2412–2427. [Google Scholar] [CrossRef]
  16. Chen, Z.; Chen, Y.; Chao, S.; Dong, X.; Chen, W.; Luo, J.; Liu, C.; Wang, D.; Chen, C.; Li, W.; et al. Single-atom AuI-N3 site for acetylene hydrochlorination reaction. ACS Catal. 2020, 10, 1865–1870. [Google Scholar] [CrossRef]
  17. Ye, L.; Duan, X.; Wu, S.; Wu, T.S.; Zhao, Y.; Robertson, A.W.; Chou, H.L.; Zheng, J.; Ayvali, T.; Day, S.; et al. Self-regeneration of Au/CeO2 based catalysts with enhanced activity and ultra-stability for acetylene hydrochlorination. Nat. Commun. 2019, 10, 914. [Google Scholar] [CrossRef]
  18. Kaiser, S.K.; Fako, E.; Surin, I.; Krumeich, F.; Kondratenko, V.A.; Kondratenko, E.V.; Clark, A.H.; Lopez, N.; Perez-Ramirez, J. Performance descriptors of nanostructured metal catalysts for acetylene hydrochlorination. Nat. Nanotechnol. 2022, 17, 606–612. [Google Scholar] [CrossRef]
  19. Peterson, L.; Bremer, J.; Sundmacher, K. Challenges in data-based reactor modeling: A critical analysis of purely data-driven and hybrid models for a CSTR case study. Comput. Chem. Eng. 2024, 184, 108643. [Google Scholar] [CrossRef]
  20. Gao, H.Y.; Struble, T.J.; Coley, C.W.; Wang, Y.R.; Green, W.H.; Jensen, K.F. Using machine learning to predict suitable conditions for organic reactions. ACS Cent. Sci. 2018, 4, 1465–1476. [Google Scholar] [CrossRef] [PubMed]
  21. Schwaller, P.; Gaudin, T.; Lányi, D.; Bekasa, C.; Lainoa, T. “Foundin Translation”: Predicting outcomes of complex organic chemistry reactions using neural sequence-to-sequence models. Chem. Sci. 2018, 28, 6091–6098. [Google Scholar] [CrossRef]
  22. Schweidtmann, A.M.; Esche, E.; Fischer, A.; Kloft, M.; Repke, J.U.; Sager, S.; Mitsos, A. Machine learning in chemical engineering: A perspective. Chem. Ing. Tech. 2021, 93, 2029–2039. [Google Scholar] [CrossRef]
  23. Chiang, L.H.; Braun, B.; Wang, Z.Y.; Castillo, I. Towards artificial intelligence at scale in the chemical industry. AIChE J. 2022, 68, e17644. [Google Scholar] [CrossRef]
  24. Zhao, X.Z.; Fan, H.A.; Lin, G.B.; Fang, Z.C.; Yang, W.L.; Li, M.; Wang, J.H.; Lu, X.Y.; Li, B.L.; Wu, K.J.; et al. Multi-objective optimization of radially stirred tank based on CFD and machine learning. AIChE J. 2024, 70, e18324. [Google Scholar] [CrossRef]
  25. Rahimpour, M.R.; Shayanmehr, M.; Nazari, M. Modeling and simulation of an industrial ethylene oxide (EO) reactor using artificial neural networks (ANN). Ind. Eng. Chem. Res. 2011, 50, 6044–6052. [Google Scholar] [CrossRef]
  26. Bhakte, A.; Chakane, M.; Srinivasan, R. Alarm based explanations of process monitoring results from deep neural networks. Comput. Chem. Eng. 2023, 179, 108442. [Google Scholar] [CrossRef]
  27. Karim, A.A.; Isa, N.A.M.; Lim, W.H. Hovering swarm particle swarm optimization. IEEE Access 2021, 9, 115719–115749. [Google Scholar] [CrossRef]
  28. Xu, H.; Si, J.K.; Luo, G.H. The kinetics model and fixed bed reactor simulation of Cu catalyst for acetylene hydrochlorination. Int. J. Chem. React. Eng. 2017, 15, 20160165. [Google Scholar] [CrossRef]
  29. Liu, G.Q.; Ma, L.X.; Xiang, S.G. Database Handbook of Physical Properties in Chemistry and Chemical Engineering; Chemical Industry Press: Beijing, China, 2012. [Google Scholar]
  30. Joint Institute of Chemical Reaction Engineering. Study on the kinetics and deactivation mechanics of acetylene hydrogen chloride reaction on HgCl2-(industrial) activated carbon catalyst. Chem. React. Eng. Technol. 1985; 1–13. (In Chinese) [Google Scholar]
  31. Schober, P.; Boer, C.; Schwarte, L.A. Correlation coefficients: Appropriate use and interpretation. Anesth Analg. 2018, 126, 1763–1768. [Google Scholar] [CrossRef]
  32. Gao, H.Y.; Zhu, L.T.; Luo, Z.H.; Fraga, M.A.; Hsing, I.M. Machine learning and data science in chemical engineering. Ind. Eng. Chem. Res. 2022, 24, 8357–8358. [Google Scholar] [CrossRef]
  33. Kingma, D.P.; Ba, J. Adam: A method for stochastic optimization. In Proceedings of the 3rd International Conference on Learning Representations, San Diego, CA, USA, 7–9 May 2015. [Google Scholar]
  34. Coello, C.A.C.; Lechuga, M.S. MOPSO: A proposal for multiple objective particle swarm optimization. In Proceedings of the 2002 Congress on Evolutionary Computation. CEC’02 (Cat. No.02TH8600), Honolulu, HI, USA, 12–17 May 2002; IEEE: Piscataway, NJ, USA, 2002; Volume 2, pp. 1051–1056. [Google Scholar] [CrossRef]
Figure 1. Schematic view of the monitor and optimization control for the VCM reactor.
Figure 1. Schematic view of the monitor and optimization control for the VCM reactor.
Processes 12 02378 g001
Figure 2. Pearson correlation matrix between any two of input variables and target variables (x, Tmax, and Pro) based on the total database.
Figure 2. Pearson correlation matrix between any two of input variables and target variables (x, Tmax, and Pro) based on the total database.
Processes 12 02378 g002
Figure 3. Architecture of ANN.
Figure 3. Architecture of ANN.
Processes 12 02378 g003
Figure 4. Sensitivity analysis of mole ratio of HCl to C2H2 (RMI).
Figure 4. Sensitivity analysis of mole ratio of HCl to C2H2 (RMI).
Processes 12 02378 g004
Figure 5. Sensitivity analysis of space velocity (SV).
Figure 5. Sensitivity analysis of space velocity (SV).
Processes 12 02378 g005
Figure 6. Sensitivity analysis of catalysis activity (a).
Figure 6. Sensitivity analysis of catalysis activity (a).
Processes 12 02378 g006
Figure 7. Sensitivity analysis of inner diameter of reactor tube (dt).
Figure 7. Sensitivity analysis of inner diameter of reactor tube (dt).
Processes 12 02378 g007
Figure 8. Sensitivity analysis of the temperature of the tube wall (Tw).
Figure 8. Sensitivity analysis of the temperature of the tube wall (Tw).
Processes 12 02378 g008
Figure 9. Sensitivity analysis of the input temperature (Tin).
Figure 9. Sensitivity analysis of the input temperature (Tin).
Processes 12 02378 g009
Figure 10. Sensitivity analysis of the tube length (l).
Figure 10. Sensitivity analysis of the tube length (l).
Processes 12 02378 g010
Figure 11. Pareto fronts for the design process.
Figure 11. Pareto fronts for the design process.
Processes 12 02378 g011
Figure 12. Pareto fronts for the operation process.
Figure 12. Pareto fronts for the operation process.
Processes 12 02378 g012
Figure 13. Single-objective optimization in the design scenario.
Figure 13. Single-objective optimization in the design scenario.
Processes 12 02378 g013
Table 1. Models used for the simulation of VCR.
Table 1. Models used for the simulation of VCR.
Activate ComponentTypes of ModelsOperational ParameterRef.
Mercury1D and 2DTw~95 °C, RMI~1.05[10]
Mercury2DSV~30 h−1, Tw~98 °C, dt[11]
Mercury1DSV, Tw[12]
Gold-based catalyst1DRMI, SV, Tin, Tw[13]
Table 2. Comparison of simulation results and industrial data.
Table 2. Comparison of simulation results and industrial data.
ItemsSV, h−1Tmax, KTout, Kxout, %
Industrial data34.12393.20372.2081.48
One-dimensional model34.15392.08374.3081.58
Two-dimensional model34.12391.56374.2981.35
Table 3. Comparison between FPM and surrogate model in design scenarios.
Table 3. Comparison between FPM and surrogate model in design scenarios.
NO.RMISVadtTwlTinFPMSurrogate Model
xATmaxProxATmaxPro
11.0335.20.230.041375.13.52366.20.876397.6742.9040.876397.8312.939
21.0638.930.310.05382.23.88368.20.950422.9405.7740.951423.5105.787
31.1134.930.410.045385.14.7368.20.995438.1665.3240.994437.3945.324
41.1539.930.480.042389.24.85371.20.997443.0065.3730.998443.3335.493
51.1937.930.210.042377.63.13362.40.828397.1562.7600.831397.5572.804
Table 4. Comparison between FPM and surrogate model in operation scenarios.
Table 4. Comparison between FPM and surrogate model in operation scenarios.
NO.RMISVaTwTinFPMSurrogate Model
xATmaxProxATmaxPro
11.1937.930.21377.6362.40.820398.9112.9060.821398.7092.916
21.0335.20.23375.1366.20.845400.1002.7920.848399.9062.793
31.0638.930.31382.2368.20.924418.1033.3450.920417.9233.351
41.1134.930.41385.1368.20.970436.5383.1690.975436.4083.189
51.1535.930.43388.1370.20.981441.3863.2970.983441.3453.306
Table 5. The ranges of different feature variables in the design scenarios.
Table 5. The ranges of different feature variables in the design scenarios.
ItemsMinimum ValueMaximum Value
RMI1.011.20
a0.200.50
SV, h−133.0040.00
Tw, K371.00393.00
dt, m0.040.05
l, m3.005.00
Tin, K362.20373.20
Table 6. Some Pareto fronts and their input conditions in design scenarios.
Table 6. Some Pareto fronts and their input conditions in design scenarios.
NO.RMISVadtTwlTinxProTmax
11.18635.9720.3270.045384.944.96366.240.9905.763422.95
21.18636.1270.3310.046383.534.99365.680.9896.038422.93
31.19036.5910.3130.048383.845.00365.330.9886.569422.37
Table 7. The ranges of different feature variables in the operation scenarios.
Table 7. The ranges of different feature variables in the operation scenarios.
ItemsMin ValueMax Value
RMI1.011.20
a0.200.50
SV, h−133.0040.00
Tw, K371.00393.00
Tin, K362.20373.20
Table 8. Some Pareto fronts and their input conditions in operation scenarios.
Table 8. Some Pareto fronts and their input conditions in operation scenarios.
NO.RMISVaTwTinxProTmax
11.2038.7280.336385.05369.540.95263.447423.14
21.2038.2010.337384.70369.550.95333.406423.12
31.2038.8570.334385.29369.490.95213.456423.10
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

Song, E.; Wang, M.; Zhou, X.; Wang, E.; Guo, C.-Y. Multi-Objective Optimization and Design for Industrial Vinyl Chloride Reactor by Hybrid Model. Processes 2024, 12, 2378. https://doi.org/10.3390/pr12112378

AMA Style

Song E, Wang M, Zhou X, Wang E, Guo C-Y. Multi-Objective Optimization and Design for Industrial Vinyl Chloride Reactor by Hybrid Model. Processes. 2024; 12(11):2378. https://doi.org/10.3390/pr12112378

Chicago/Turabian Style

Song, Erwei, Mingmei Wang, Xiaoyu Zhou, Erqiang Wang, and Cun-Yue Guo. 2024. "Multi-Objective Optimization and Design for Industrial Vinyl Chloride Reactor by Hybrid Model" Processes 12, no. 11: 2378. https://doi.org/10.3390/pr12112378

APA Style

Song, E., Wang, M., Zhou, X., Wang, E., & Guo, C. -Y. (2024). Multi-Objective Optimization and Design for Industrial Vinyl Chloride Reactor by Hybrid Model. Processes, 12(11), 2378. https://doi.org/10.3390/pr12112378

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