Next Article in Journal
Analysis of Surrounding Rock Control Technology and Its Application on a Dynamic Pressure Roadway in a Thick Coal Seam
Next Article in Special Issue
Review of Flow Field Designs for Polymer Electrolyte Membrane Fuel Cells
Previous Article in Journal
What China’s Environmental Policy Means for PV Solar, Electric Vehicles, and Carbon Capture and Storage Technologies
Previous Article in Special Issue
Progress of Performance, Emission, and Technical Measures of Hydrogen Fuel Internal-Combustion Engines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogen Production System through Dimethyl Ether Autothermal Reforming, Based on Model Predictive Control

Department of Mechanical Engineering, Chonnam National University, Gwangju 500010, Republic of Korea
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(23), 9038; https://doi.org/10.3390/en15239038
Submission received: 4 November 2022 / Revised: 21 November 2022 / Accepted: 26 November 2022 / Published: 29 November 2022
(This article belongs to the Special Issue Advanced Research on Fuel Cells and Hydrogen Energy Conversion)

Abstract

:
In this study, a thermodynamic analysis of the low temperature autothermal reforming (ATR) of dimethyl ether (DME) for hydrogen production was conducted. The Pd/Zn/γ-Al2O3 catalyst coated on the honeycomb cordierite ceramic was applied to catalyze the reaction, and the optimum activity temperature of this catalyst was demonstrated experimentally and through simulations to be 400 °C. Furthermore, an optimal model predictive control (MPC) strategy was designed to control the hydrogen production rate and the catalyst temperature. Experimental and simulation results indicated that the controller was automated and continuously reliable in the hydrogen production system. By establishing the state-space equations of the autothermal reformer, it can precisely control the feed rates of DME, high-purity air and deionized water. Ultimately, the hydrogen production rate can be precisely controlled when the demand curve changed from 0.09 to 0.23 mol/min, while the catalyst temperature was maintained at 400 °C, with a temporary fluctuation of 4 °C during variations of the hydrogen production rate. Therefore, the tracking performance of the hydrogen production and the anti-disturbance were satisfactory.

1. Introduction

In recent years, the worldwide energy demand has increased significantly as the world’s population has grown, financial conditions have evolved, and urbanization has accelerated [1,2,3,4]. However, the traditional energy supply method relies on the exploitation of fossil energy, which is limited by the uneven geographical distribution and inconsistent exploitation difficulty. At the same time, long-term consumption makes fossil energy increasingly scarce, and reducing carbon emissions has become a global issue [5,6]. In terms of energy supply innovation, the proportion of new clean energy represented by hydrogen, solar energy and biomass (e.g., biogas and biodiesel) in the overall energy supply field is increasing annually, and fuel cell vehicles are cleaner and more efficient than traditional internal combustion engine vehicles [7,8]. Electric vehicles are an eco-friendly transportation method, but due to the relatively long charging time, the risk of explosion, and high cost of replacing batteries, hydrogen vehicles have attracted attention as a next generation transportation method. These situations indicate the urgent need to develop clean and renewable energy to replace the existing energy supply system, leading to greater attention to the research and development of hydrogen energy [9].
Hydrogen power generation is a prospective solution to the energy storage of green energy for a completely clean and efficient hydrogen industry [10], which can provide clean, efficient, reliable, and affordable energy with significant societal benefits [11]. Hydrogen energy thus plays an important role in achieving sustainable development goals (SDGs) on a global scale. However, the hydrogen energy industry has characteristics of multiple components, complex technical routes, diverse application scenarios, and high economic costs. These are the major obstacles that affect the progress of the hydrogen industry. The economics of hydrogen requires the strong cross-chain cooperation to be manifested. With the continuous advancement of technology and the expansion of the scale of the hydrogen energy industry, the marginal cost has been gradually accepted by society.
Types of hydrogen-based systems can be classified, according to the end user [12]. Ahmed et al. [13] systematically studied the use of hydrogen fuel in the Malaysian transportation system for a sustainable and environmentally friendly future. Hou T F et al. [14] presented a facile fabrication process to enhance the photoelectrochemical(PEC) performance of ZnO-based photoelectrodes for the conduct of solar water splitting for hydrogen generation. Bellotti et al. [15] studied the feasibility of an innovative device that generates hydrogen on the grid through a water electrolysis unit and synthesized it with carbon dioxide stored in fossil fuel power plants to produce methanol. In terms of energy supply miniaturization, the proton exchange membrane fuel cells (PEMFCs) can be used for the efficient conversion of chemical and electrical energy. PEMFCs have the characteristics of a high energy density, absence of pollutants, and efficiency at relatively low temperatures and pressures. These conditions make PEMFCs an excellent solution for hydrogen utilization in recent years [7,16,17]. These studies demonstrate the practical solutions for hydrogen applications, but in portable applications, it has many drawbacks, in terms of safety and storage [18]. Therefore, to form a complete hydrogen industry chain, how to produce hydrogen efficiently and economically has become the core issue, which is increasing attention.
Hydrocarbon reforming is a widely used hydrogen production technology. Brown [19] compared the utility of seven common fuels as hydrogen sources for the proton exchange membrane fuel cells for automobile propulsion, which directly confirmed the wide range of feedstock required for the hydrogen production. Among many hydrocarbons, dimethyl ether, as a feedstock for the hydrogen production, has the following advantages:
(1) The sources of dimethyl ether are abundant and diverse. In recent years, hydrogen and carbon monoxide synthesis methods have been developed to replace the original costly methanol dehydration production method, resulting in a significant reduction in the cost of DME preparation and an increase in production efficiency, which is conducive to large-scale applications.
(2) Dimethyl ether is non-toxic and non-corrosive, and is easily degraded in the atmospheric troposphere, which ensures the environmental friendliness of dimethyl ether during use.
(3) Similar to liquefied petroleum gas (LPG), dimethyl ether can be compatible with the existing infrastructure of LPG and natural gas, which reduces the commercialization cost of dimethyl ether reforming to produce hydrogen, and greatly reduces the threshold for this new fuel to enter the energy market.
(4) Dimethyl ether is easily liquefied at about 5–6 atmospheres, making it easy to store and transport, which significantly reduces the transportation cost of the fuel.
(5) It has a higher hydrogen density than other commercial fuels (e.g., diesel) and can be reformed to release more of the hydrogen stored in it.
By integrating these advantages, DME is very attractive as a mobile hydrogen carrier for fuel cell systems [20,21].
Typical hydrocarbon-reforming technologies are steam reforming (SR), autothermal reforming (ATR), and partial oxidation (POX) [22,23,24,25]. Wu Z et al. [26] studied the methane steam reforming in a grille-sphere composite packed bed, and the effect of the diameter distribution of the particles in the packed bed reactor on the efficiency was investigated. Cherif A et al. [27] performed a thermodynamic analysis of the methane autothermal steam reforming for the hydrogen production, and a new configuration with 18 catalyst bed macrostructures alternately installed, was designed. ATR is thermally neutral in the reaction and a muffle furnace is only used to preheat the catalyst to the optimum temperature during the start-up phase, while no external heat source is required to keep the reaction to be carried out during the continuous operation. Furthermore, it has an excellent hydrogen yield among the three methods. Therefore, ATR is an efficient and economical method of hydrogen production [28]. Hydrogen production by autothermal reforming of DME is significantly affected by the catalyst bed temperature; that is, under the same physical operating conditions, a higher reactor temperature theoretically leads to a higher hydrogen yield. However, a sustained high temperature can reduce the catalyst life and hydrogen production efficiency [29]. Many studies showed that excessively high temperatures are destructive to the physical structure of the reactor. Specifically, the prolonged exposure to 20 K over the design temperature halved the lifetime of the reformer tubes [30,31]. Meanwhile, the excessively high temperature in the POX may also lead to catalyst sintering [32]. Therefore, in addition to controlling the hydrogen production rate to meet the specific demands of the load, keeping the reformer operating at the optimum temperature during the autothermal reforming is necessary to supply hydrogen to on-board fuel cells.
In view of the hysteresis of the control action due to the chemical reaction of this system and the situation of the multiple inputs and outputs, many researchers have explored the control strategies. Dolanc et al. [33] developed a PID controller that keeps the reaction at a stable reactor temperature while accurately responding to the hydrogen demand for a specific load in the diesel autothermal reforming process. However, compared to the PID control, the model predictive control (MPC) performs better in the chemical industry, which makes MPC more capable of operating the hydrogen production under strict performance specifications and constraints [34]. From the perspective of the working mechanism of the model predictive control, its advantages are reflected in the following points:
(1) The model is easy to build. The process characterization can be detected by performing simple experiments and does not require a deep exploration of the chemical reaction mechanism.
(2) The discrete convolution and model described by non-minimization are used, and the information redundancy is large, which is beneficial to improve the robustness of the system.
(3) Using the rolling optimization strategy, the online optimization calculation is repeatedly performed, and the rolling execution is conducted to obtain a better dynamic control performance, to compensate for the uncertainty caused by the model mismatch, distortion and disturbance in time.
For instance, the hydrogen production rate and reactor temperature are simultaneously set as system constraints and controlled synchronously in the model predictive control, which has a positive significance for the system performance optimization and long-term stable operation. Similar control strategies have been applied in other devices for the hydrogen production. Kyriakides et al. [35] developed a MPC strategy for low temperature methane SR in membrane reactors, enabling the high methane conversion at relatively low temperature levels. Paliwal N K et al. [36] developed a MPC-based control scheme for a microgrid system, based on photovoltaic power generation, lead-acid battery energy storage and hydrogen production, resulting in a stable and efficient energy supply system. Hu et al. [37] demonstrated a nonlinear multivariable predictive controller with an excellent computational performance for the ATR reactors.
The above studies demonstrate the feasibility and superiority of the MPC in chemical reaction control processes. However, few studies have focused on the application of model predictive control in the autothermal reforming of dimethyl ether for hydrogen production so far, which is a highly integrated and efficient hydrogen production process. Inspired by this research, this study innovatively contracts an efficient DME autothermal reforming system, based on the MPC control and demonstrates its reliability and efficiency through simulation and experiments. Ultimately, this study prospectively applies model predictive control to the hydrogen production using DME, overcoming the hysteresis problem that is characteristic of the chemical reaction control processes. In addition, this system, when integrated with the hydrogen purification module, has the potential to be used in combination with an on-board fuel cell to power new energy vehicles, which is an important step in the promotion of hydrogen energy.

2. Mathematical Modelling and Control Method

2.1. ATR Reformer Model

MATLAB Simulink is applied to the simulation. Simulink plays an important role in the control strategy to linearize the complex chemical reaction system for the MPC control. In chemical kinetic models, Gibbs free energy is a generic function for identifying equilibrium states because of its concise properties. For the multi-matter and multi-phase systems, the equilibrium calculation can be performed by the Gibbs energy minimization method [38]. In this case, the Arrhenius equation to describe the rate of a chemical reaction is appropriate. The system is treated as a closed-loop one so that the equilibrium composition is related to the temperature and concentrations of the individual components in the reactions for a given catalyst performance. To facilitate the model operation, this study adopts the following assumptions:
(1) The overall DME autothermal reforming reaction can be divided into five steps: DME hydrolysis, methanol (MeOH) steam reforming, MeOH decomposition, MeOH partial oxidation, and carbon monoxide(CO) oxidation in which the heat generated by the exothermic reaction can provide energy for the endothermic reaction, making the entire reaction thermoneutral. The water gas shift (WGS) is integrated after the autothermal reformer so that the CO can be converted to carbon dioxide (CO2) and generate more hydrogen. Various catalysts have been extensively studied and commercialized, and the conversion rate of Cu/Zn/γ-Al2O3 is higher than that of other commercial catalysts [39,40]. Therefore, a ternary Cu/Zn/γ-Al2O3 (Cu: 20 wt.%, Zn: 20 wt.%) catalyst was used in the reformer for the WGS.
(2) As the liquid water enters the reactor after passing through a 50 cm-long steel pipe that is heated to 150 °C by a heating belt, the liquid water is completely converted into vapor to participate in the reaction.
The mass balance of the mass transport can be described by the following equation [41]:
ϵ p ρ ω j t + · j j + ρ ( u · ) ω j = R j  
j j = ( ρ ω j k D e , j k d k + D e , j T T T )  
D e , j k = ϵ p τ F D i k ,   D e , j T = ϵ p τ F D j T ,   τ F = ϵ p 1 / 3  
d k = x k + 1 p A [ ( x k w k ) p A ]  
x k = ω k M k M n   ,   M n = ( j ω j M j ) 1  
Meanwhile, the energy balance of the mass flow can be described by the following equations:
A c ( ρ C p ) e f f T t + A c ρ F C p , f u · T + · q = Q  
q = A c k e f f T  
( ρ C p ) e f f = θ s ρ s C p , s + ϵ p ρ f C p , f
k e f f = θ s k s + θ s k f + k d i s p
The boundary conditions at the reactor feed port z = 0 are as follows:
ω j = ω j , o       T = T o       T S = T s , o       P = P o
At the end of the reactor z = L , the boundary conditions can be defined as:
ω j z = 0       T z = 0       T s z = 0
Initial conditions of the reaction are defined as:
ω j = ω j , o       T = T o       T S = T s , o
Annotations of the parameters applied in the mass and energy balance model are shown in Table 1.
Several institutes have developed the chemical kinetics of the autothermal reforming of DME [42,43,44]. Based on these studies, the individual steps of the reaction and kinetic equations used to describe the reaction rates are shown in Table 2, and the parameters can be found in Table 3.
In this multi-input multi-output (MIMO) control system, the catalyst temperature and hydrogen production rate are the objective variables. Therefore, they are fed back to the MPC controller for the feedback correction. The MPC then controls the DME, air, and vapor supply rates, according to the linearized space state equation of the reaction system, to meet specific hydrogen demands and stabilize the reactor temperature. Based on the mathematical description of the hydrogen generation system, the linearized Simulink model is shown in Figure 1.

2.2. State Space Equations Applied in the MPC

Considering that the hydrogen conversion is mainly affected by the catalyst temperature and the flexibility of the hydrogen supply is also crucial in the hydrogen supply of the on-board fuel cells, these two items are set as the control objectives of the control system. Furthermore, MPC is a robust control strategy. Since its inception in the 1970s, MPC has evolved into a new control algorithm characterized by diversity and usefulness. It has a wide range of applications in the chemical and mechanical fields because of its simplicity, richness, and practicality. MPC typically establishes a predictive model first, and then performs a rolling optimization and feedback correction. Once the model of the controlled plant is established, it can control multiple objective variables during its operation.
The implementation of MPC aims to model the optimization problem and then perform a rolling optimization to determine the final output of the controller. Typically, MPC relies on a linear time-invariant, discrete-time, state-space model of the system to predict the future response of the system. MPC requires solving the optimization problem through the repeated prediction and optimization at each time step. When the optimal solution of the optimization problem is obtained, it is used as the control action of the real actuator. Figure 2 illustrates the mechanism of a typical MPC controller.
Following the construction of a predictive model, the MPC can predict the future output characteristics of a control objective over a prediction horizon, and an optimum control sequence is obtained based on certain criteria. To ensure the optimality and reliability, only the first action in the control sequence is implemented [35,45,46]. The current control action is determined by solving a finite-time-domain open-loop optimal control problem at each sampling instant. The current state of the process is taken as the initial state of the optimal control problem and the obtained optimal control sequence only executes the first control action. Therefore, the flow rates of water, DME, and air are continuously optimized over time, based on the space state equation obtained by linearizing the ATR reformer model, which considers the effects of the ideal optimization and practical uncertainty in the future finite time domain. In the simulation, when determining the predictive horizon, it is gradually increased until further increases have a minor impact on the performance. The sampling time is typically determined as 10% of the prediction horizon. Finally, the prediction horizon and sampling time are determined to be 20 s and 2 s, respectively after trying many simulations. We obtained the linearized mathematical model of the plant model using the linear analysis tool in MATLAB. The continuous-time state-space model and the constant matrixes A, B, C, and D are listed as follows.
x ˙ = A x + B u y = C x + D u
where:
A = [ 4.663 × 10 9 9.291 × 10 9 1.398 × 10 8 4.661 × 10 9 4.658 × 10 9 4.915 × 10 13 0.2357 0.4692 0.7064 0.2356 0.2354 3.222 × 10 5 0.0602 0.1205 0.1807 0.0602 0.0603 1.752 × 10 6 0.118 0.2346 0.3535 0.1179 0.1178 1.922 × 10 5 0.2299 0.4653 0.6918 0.2303 0.2308 7.84 × 10 5 16.88 21.91 29.14 13.02 8.054 0.7869 ]
B = [ 1.383 × 10 6 9.384 × 10 7 4.458 × 10 7 69.59 47.64 22.43 18.64 11.71 6.657 36.13 22.83 13.06 68.76 46.78 22.12 3236 2437 1405 ]
C = [ 0 0 0 0 0 1 0.4548 4.222 0 1.431 5.013 0.0007067 ]
D = 0

3. Experiment Apparatus

3.1. Catalyst Preparation

For the autothermal reforming, several institutions are dedicated to the development of catalysts. The Pd/Zn/Al2O3 catalyst exhibited a high activity and generated syngas with a CO concentration below 5 vol.% when the catalyst was heated from 350 to 450 °C, where the hydrogen concentration was close to 50 vol.% [47,48,49]. Takeishi et al. [50] proposed a sol-gel method to prepare the superior catalyst for the hydrocarbon reforming. Due to the close existence of the active sites for each reaction of the catalyst, its catalytic performance was better than that of the physically mixed dimethyl ether hydrolysis catalyst and methanol SR catalyst. In our experiments, 3.5 g of the Pd/Zn/Al2O3 catalyst prepared by the sol-gel method is coated on the surface of honeycomb cordierite ceramics with specific loadings of Pd, Zn and γ-Al2O3 of 5, 20 and 75 wt.%, respectively. Entire catalysts are made into a porous cylinder with a diameter of 36 mm and a length of 160 mm.

3.2. MPC Controller Design

Considering the diversity of the hydrogen generation systems, the main subject of this paper is the use of an MPC control scheme for the MIMO ATR reactors. Control of the hydrogen production rate and reactor temperature of the ATR reactor is achieved by specifically feeding air, dimethyl ether, and steam.
As shown in Figure 3, the implementation of the hydrogen production control system consists of three sections: feedstock supply module, autothermal reaction module, and signal acquisition and control module. Dimethyl ether and high-purity air are stored in two cylinders, and when the autothermal reaction proceeds, they are precisely fed to the reactor through two mass flow controllers (Ethernet, MKS Instruments, Andover, MA, USA) attached to the cylinders. At the same time, under the action of the syringe pump (Chemyx.Inc, Fusion100, Stafford, UK), the deionized water stored in the syringe is converted into steam and fed to the reformer through a steel pipe wrapped with an electric heating belt at 150°C to participate in the autothermal reaction. The hydrogen generated system prototype is shown in Figure 4.
The autothermal reaction module is a fixed catalyst bed reactor formed by packing the catalyst mentioned in Section 3.1 in a reactor, which is made of a steel tube with an outer diameter of 40 mm and thickness of 2 mm. The reactor is enclosed in a tubular muffle furnace that provides heat to raise the catalyst temperature for the optimal catalytic activity and isolates the heat losses during the reaction. These measures can efficiently initialize the hydrogen production system and are beneficial to maintain the optimal temperature of the catalyst. A cold trap is attached to the reactor outlet to lower the temperature of the hydrogen-rich gas and condense the excess vapor in it because the subsequent gas flow meter and gas chromatograph (GC) measurements need to be performed in a normal temperature and dehydrated measurement environment. The GC can test the purity of a particular compound, separate the components in a mixture, and detect the relative amounts of each component. In this study, GC and a gas flow meter are used to test the molarity of hydrogen in the syngas and detect the hydrogen yield.
The signal acquisition and control module are developed, based on the model predictive control, where two mass flow controllers (MFCs) and one syringe pump are used to control the feed rates of the feedstocks. The NI device and LabVIEW 2019 are used to collect and process the temperature of the catalyst bed and the hydrogen yield. This type of interaction has a wide range of applications for its efficient and accurate properties. In the closed-loop control system, based on MPC, the reactor temperature T s e t and hydrogen demand H 2 s e t are two control objectives that need to be detected as real-time outputs of the system. Therefore, a K-type thermocouple (0–1300 ℃) is attached to the reactor surface to detect the reaction temperature T m e a s u r e d , and a flow meter (Sensirion AG Inc., SFM3020, Stäfa, Swizerland) is installed at the reactor outlet to detect the flow rate of the syngas F P .
Due to the lack of a mechanism to detect the hydrogen flow rate online, this study obtained the general volume fraction φ H 2 of hydrogen through GC measurements in multiple experiments. Then, the hydrogen production rate can be calculated with Equations (14) and (15):
F H 2 = ( F P F N 2 ) × φ H 2
F N 2 = F A i r × φ N 2
where F H 2   denotes the hydrogen production rate; F P is the flow rate of the syngas, which is detected by the gas flow meter; and F A i r is the flow rate of the high-purity air, which can be measured by the MFC. φ N 2 denotes the nitrogen volume fraction in the high-purity air, which is 79%.
φ H 2 is the hydrogen volume fraction in the syngas. According to previous research results [34], the volume fraction of hydrogen is mainly affected by the reaction temperature and the concentration of each component, which does not change drastically in the operation. The effect of the reaction temperature on the hydrogen production is also fully explored through the experiment and simulation. As shown in Figure 5, the hydrogen content reaches the maximum value of 43% while the reactor temperature is approximately 400 °C. Therefore, according to the measurements of the GC, and considering the necessary hydrogen redundancy, the general hydrogen volume fraction φ H 2 in the syngas is determined as 40%, while the optimal temperature of the autothermal reforming is 400 °C.

4. Result Discussion

The developed MPC was tested to verify the tracking performance of the desired hydrogen yield and the anti-disturbance of the catalyst temperature in the reactor-specific hydrogen demand profile. The dynamic behavior of the hydrogen yield and catalyst temperature with a step change in the set signal was also investigated.
Figure 6 shows the hydrogen demand profile in a control system with two MFCs and a syringe pump as actuators. The set hydrogen demand flow rate presents a step change from 0.09 mol/min to 0.23 mol/min. The corresponding flow rates of DME, high-purity air, and liquid water in both the simulation and experiment are shown in Figure 7. To achieve the control objectives, the MPC control strategy implemented in LabVIEW and the results obtained in the simulation show similar corresponding control actions, according to the changes in the hydrogen demand curve.
Based on the control actions of the MPC and the working characteristics of the actuators, the dynamic response of DME and a high-purity air show certain overshoot fluctuations when the set signal changes, but after approximately 20 s of the MPC modulation, the system reaches a stable state, while the liquid water flow provided by the syringe pump rises more smoothly. The experimental results show a slightly larger overshoot and slightly longer steady-state arrival time during the dynamic changes, compared to the simulation results, but within an acceptable range. The steam-to-carbon ratio (SCR) and oxygen-to-carbon ratio (OCR) in the reaction changed due to the different fluctuations among the components. Thus, the thermal neutrality of the entire autothermal reforming reaction is guaranteed in the dynamic operation and is beneficial for maintaining a stable reactor temperature.
The dashed line in Figure 8 shows the hydrogen flow rate obtained from the simulation, and the solid line represents the hydrogen generation curve in the experiment. This curve is calculated, based on the detection of the mass flow meter and Equations (13) and (14). The simulation result shows that the overshoot of the hydrogen flow rate during the dynamic response process does not exceed 0.004 mol/min. Furthermore, the measured hydrogen generation rate fluctuates, due to the operating error of the flowmeter and the influence of the turbulent flow in the pipeline. However, these factors are not significant with respect to the overall change of hydrogen flow. Based on these results, the traceability of the hydrogen yield is validated, and the dynamic behavior of the controller indicates that the setpoint tracking is satisfactory.
The reactor temperature has a considerable effect on the catalyst activity. In the dynamic control, it should exhibit an excellent anti-disturbance with the continuous change of the hydrogen production rate to ensure that the autothermal reforming reaction is conducted at an optimal temperature. Figure 9a shows the variation profile of the catalyst temperature obtained from the simulation and experimental measurements in the MPC control. Figure 9b presents the experimental temperature result obtained in the PID control, as reported in previous research [41]. Based on the similar changes in the hydrogen demand curve, both control mechanisms can keep the catalyst temperature at the optimum value of 400 °C, but the temperature curve is smoother in the MPC control strategy. Only a slight perturbation occurs when the set hydrogen demand curve changes in steps. This dynamic performance shows that the anti-disturbance of the catalyst temperature is satisfactory.

5. Conclusions

In this study, a rigorous mathematical model was proposed to the simulate and control the DME autothermal reformer. An advanced MPC scheme was implemented, which calculated the optimal sequence of the controlled variables within a specified prediction horizon. The experiment and simulation results indicated that:
1. The expected dynamic response of the hydrogen production rate was achieved by controlling the feed rate of dimethyl ether in the range of 0.09–0.23 mol/min, and a high purity air and steam as feedstocks. The traceability of the hydrogen production rate was satisfactory.
2. The disturbance compensation was obtained for maintaining the catalyst temperature stability. The catalyst temperature shows a good immunity to the feedstock flow rate variations. Although a deviation of the catalyst temperature of about 4 °C occurred during the step change of the feedstock flow rate, it recovered to the control target value within a short time domain (20 s).
Based on these results, the next step of the study is to further improve the purity and production efficiency of hydrogen and to integrate this hydrogen generation system with the PEMFCs to develop a complete energy management system.

Author Contributions

Formal analysis, T.-Q.Z.; Investigation, S.J.; Project administration, Y.-B.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Republic of Korea (22R1I1A3063283).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, J.; Tan, J.; Yang, W.; Li, Y.; Wang, C. Better electrochemical performance of PEMFC under a novel pneumatic clamping mechanism. Energy 2021, 229, 120796. [Google Scholar] [CrossRef]
  2. Chiu, W.C.; Hou, S.S.; Chen, C.Y.; Lai, W.H.; Horng, R.F. Hydrogen-rich gas with low-level CO produced with autothermal methanol reforming providing a real-time supply used to drive a kW-scale PEMFC system. Energy 2022, 239, 122267. [Google Scholar] [CrossRef]
  3. Yang, B.; Li, D.; Zeng, C.; Chen, Y.; Guo, Z.; Wang, J.; Shu, H.; Yu, T.; Zhu, J. Parameter extraction of PEMFC via Bayesian regularization neural network based meta-heuristic algorithms. Energy 2021, 228, 120592. [Google Scholar] [CrossRef]
  4. Karayel, G.K.; Javani, N.; Dincer, I. Effective use of geothermal energy for hydrogen production, A comprehensive application. Energy 2022, 249, 123597. [Google Scholar] [CrossRef]
  5. Kafetzis, A.; Ziogou, C.; Papadopoulou, S.; Voutetakis, S.; Seferlis, P. Nonlinear Model Predictive Control of an Autonomous Power System Based on Hydrocarbon Reforming and High Temperature Fuel Cell. Energies 2021, 14, 1371. [Google Scholar] [CrossRef]
  6. Liu, J.; Yang, Q.; Ou, S.; Liu, J. Factor decomposition and the decoupling effect of carbon emissions in China’s manufacturing high-emission subsectors. Energy 2022, 248, 123568. [Google Scholar] [CrossRef]
  7. Takach, M.; Sarajlić, M.; Peters, D.; Kroener, M.; Schuldt, F.; Maydell, K. Review of Hydrogen Production Techniques from Water Using Renewable Energy Sources and Its Storage in Salt Caverns. Energies 2022, 15, 1415. [Google Scholar] [CrossRef]
  8. Sharma, P.; Sahoo, B.B. Precise prediction of performance and emission of a waste derived Biogas–Biodiesel powered Dual–Fuel engine using modern ensemble Boosted regression Tree: A critique to Artificial neural network. Fuel 2022, 321, 124131. [Google Scholar] [CrossRef]
  9. Shafiei, E.; Davidsdottir, B.; Leaver, J.; Stefansson, H.; Asgeirsson, E.I. Energy, economic, and mitigation cost implications of transition toward a carbon-neutral transport sector, A simulation-based comparison between hydrogen and electricity. J. Clean. Prod. 2017, 141, 237–247. [Google Scholar] [CrossRef]
  10. Dawood, F.; Anda, M.; Shafiullah, G.M. Hydrogen production for energy, An overview. Energy 2020, 45, 3847–3869. [Google Scholar] [CrossRef]
  11. Kalinci, Y.; Hepbasli, A.; Dincer, I. Techno-economic analysis of a stand-alone hybrid renewable energy system with hydrogen production and storage options. Int. J. Hydrog. Energy 2015, 40, 7652–7664. [Google Scholar] [CrossRef]
  12. Acar, C.; Dincer, I. Review and evaluation of hydrogen production options for better environment. J. Clean. Prod. 2019, 218, 835–849. [Google Scholar] [CrossRef]
  13. Ahmed, A.; Al-Amin, A.Q.; Ambrose, A.F.; Saidur, R. Hydrogen fuel and transport system, A sustainable and environmental future. Int. J. Hydrog. Energy 2016, 41, 1369–1380. [Google Scholar] [CrossRef]
  14. Hou, T.F.; Shanmugasundaram, A.; Hassan, M.A.; Johar, M.A.; Ryu, S.W.; Lee, D.W. ZnO/Cu2O-decorated rGO, heterojunction photoelectrode with improved solar water splitting performance. Int. J. Hydrog. Energy 2019, 44, 19177–19192. [Google Scholar] [CrossRef]
  15. Bellotti, D.; Rivarolo, M.; Magistri, L.; Massardo, A.F. Feasibility study of methanol production plant from hydrogen and captured carbon dioxide. J. CO2 Util. 2017, 21, 132–138. [Google Scholar] [CrossRef]
  16. Zou, W.J.; Shen, K.Y.; Jung, S.; Kim, Y.B. Application of thermoelectric devices in performance optimization of a domestic PEMFC-based CHP system. Energy 2021, 229, 120698. [Google Scholar] [CrossRef]
  17. Ou, K.; Wang, Y.X.; Li, Z.Z.; Shen, Y.D.; Xuan, D.J. Feedforward fuzzy-PID control for air flow regulation of PEM fuel cell system. Int. J. Hydrog. Energy 2015, 40, 11686–11695. [Google Scholar] [CrossRef]
  18. Bianchini, M.; Alayo, N.; Soler, L.; Salleras, M.; Fonseca, L.; Llorca, J.; Tarancón, A. Standalone micro-reformer for on-demand hydrogen production from dimethyl ether. J. Power Sources 2021, 506, 230241. [Google Scholar] [CrossRef]
  19. Brown, L.F. A comparative study of fuels for on-board hydrogen production for fuel-cell-powered automobiles. Int. J. Hydrog. Energy 2001, 26, 381–397. [Google Scholar] [CrossRef]
  20. Faungnawakij, K.; Tanaka, Y.; Shimoda, N.; Fukunaga, T.; Kikuchi, R.; Eguchi, K. Hydrogen production from dimethyl ether steam reforming over composite catalysts of copper ferrite spinel and alumina. Appl. Catal. B-Environ. 2007, 74, 144–151. [Google Scholar] [CrossRef]
  21. Choi, S.; Bae, J. Autothermal reforming of dimethyl ether with CGO-based precious metal catalysts. J. Power Sources 2016, 307, 351–357. [Google Scholar] [CrossRef]
  22. Tartakovsky, L.; Mosyak, A.; Zvirin, Y. Energy analysis of ethanol steam reforming for hybrid electric vehicle. Int. J. Energy Res. 2013, 37, 259–267. [Google Scholar] [CrossRef]
  23. Malik, F.R.; Kim, Y.B. Autothermal reforming of n-hexadecane over Rh catalyst to produce syngas in microchannel reactor using finite element method. Int. J. Energy Res. 2019, 43, 779–790. [Google Scholar] [CrossRef]
  24. Nwosu, C.; Ayodele, O.; Ibrahim, H. Optimization of hydrogen production via catalytic autothermal reforming of crude glycerol using response surface methodology and artificial neural network. Int. J. Energy Res. 2021, 45, 18999–19013. [Google Scholar] [CrossRef]
  25. Teng, C.Y.; Chiu, S.J. Hydrogen production by partial oxidation of ethanol over Pt/CNT catalysts. Int. J. Energy Res. 2013, 37, 1689–1698. [Google Scholar] [CrossRef]
  26. Wu, Z.; Guo, Z.; Yang, J.; Wang, Q. Effect of diameter distribution of particles on methane steam reforming in multi-channel grille-sphere composite packed bed. Energy Conv. Manag. 2022, 265, 115764. [Google Scholar] [CrossRef]
  27. Cherif, A.; Nebbali, R.; Sen, F.; Sheffield, J.W.; Doner, N.; Nasseri, L. Modeling and simulation of steam methane reforming and methane combustion over continuous and segmented catalyst beds in autothermal reactor. Int. J. Hydrog. Energy 2022, 47, 9127–9138. [Google Scholar] [CrossRef]
  28. Faungnawakij, K.; Viriya-Empikul, N.; Tanthapanichakoon, W. Evaluation of the thermodynamic equilibrium of the autothermal reforming of dimethyl ether. Int. J. Hydrog. Energy 2011, 36, 5865–5874. [Google Scholar] [CrossRef]
  29. Wu, Z.; Aguirre, A.; Tran, A.; Durand, H.; Ni, D.; Christofides, P.D. Model predictive control of a steam methane reforming reactor described by a computational fluid dynamics model. Ind. Eng. Chem. Res. 2017, 56, 6002–6011. [Google Scholar] [CrossRef]
  30. Wang, H.S.; Chang, C.P.; Huang, Y.J.; Su, Y.C.; Tseng, F.G. A high-yield and ultra-low-temperature methanol reformer integratable with phosphoric acid fuel cell (PAFC). Energy 2017, 133, 1142–1152. [Google Scholar] [CrossRef]
  31. Pantoleontos, G.; Kikkinides, E.S.; Georgiadis, M.C. A heterogeneous dynamic model for the simulation and optimisation of the steam methane reforming reactor. Int. J. Hydrog. Energy 2012, 37, 16346–16358. [Google Scholar] [CrossRef]
  32. Ubago-Pérez, R.; Carrasco-Marín, F.; Moreno-Castilla, C. Methanol partial oxidation on carbon-supported Pt and Pd catalysts. Catal. Today 2007, 123, 158–163. [Google Scholar] [CrossRef]
  33. Dolanc, G.; Pregelj, B.; Petrovčič, J.; Pasel, J.; Kolb, G. Control of autothermal reforming reactor of diesel fuel. J. Power Sources 2016, 313, 223–232. [Google Scholar] [CrossRef]
  34. Khalid, M.; Savkin, A.V. A model predictive control approach to the problem of wind power smoothing with controlled battery storage. Renew. Energy 2010, 35, 1520–1526. [Google Scholar] [CrossRef]
  35. Kyriakides, S.; Seferlis, P.; Voutetakis, S.; Papadopoulou, S. Model predictive control for hydrogen production in a membrane methane steam reforming reactor. Chem. Eng. Trans. 2016, 52, 991–996. [Google Scholar]
  36. Paliwal, N.K.; Singh, A.K.; Singh, N.K. Short-term optimal energy management in stand-alone microgrid with battery energy storage. Arch. Electr. Eng. 2018, 67, 499–513. [Google Scholar]
  37. Hu, Y.; Chmielewski, D.J. Nonlinear multivariable predictive control of an autothermal reforming reactor for fuel cell applications. In Proceedings of the 2009 American Control Conference, St. Louis, MO, USA, 10–12 June 2009; pp. 659–664. [Google Scholar]
  38. Wang, J.; Chen, H.; Tian, Y.; Yao, M.; Li, Y. Thermodynamic analysis of hydrogen production for fuel cells from oxidative steam reforming of methanol. Fuel 2012, 97, 805–811. [Google Scholar] [CrossRef]
  39. Li, Y.; Fu, Q.; Flytzani-Stephanopoulos, M. Low-temperature water-gas shift reaction over Cu-and Ni-loaded cerium oxide catalysts. Appl. Catal. B-Environ. 2000, 27, 179–191. [Google Scholar] [CrossRef] [Green Version]
  40. LeValley, T.L.; Richard, A.R.; Fan, M. The progress in water gas shift and steam reforming hydrogen production technologies–A review. Int. J. Hydrog. Energy 2014, 39, 16983–17000. [Google Scholar] [CrossRef]
  41. Zhang, T.Q.; Malik, F.R.; Jung, S.; Kim, Y.B. Hydrogen production and temperature control for DME autothermal reforming process. Energy 2022, 239, 121980. [Google Scholar] [CrossRef]
  42. Ouzounidou, M.; Ipsakis, D.; Voutetakis, S.; Papadopoulou, S.; Seferlis, P. A combined methanol autothermal steam reforming and PEM fuel cell pilot plant unit: Experimental and simulation studies. Energy 2009, 34, 1733–1743. [Google Scholar] [CrossRef]
  43. Ipsakis, D.; Voutetakis, S.; Papadopoulou, S.; Seferlis, P. Optimal operability by design in a methanol reforming-PEM fuel cell autonomous power system. Int. J. Hydrog. Energy 2012, 37, 16697–16710. [Google Scholar] [CrossRef]
  44. Nilsson, M.; Jozsa, P.; Pettersson, L.J. Evaluation of Pd-based catalysts and the influence of operating conditions for autothermal reforming of dimethyl ether. Appl. Catal. B-Environ. 2007, 76, 42–50. [Google Scholar] [CrossRef]
  45. Wang, J.; Wei, S.; Wang, Q.; Sundén, B. Transient numerical modeling and model predictive control of an industrial-scale steam methane reforming reactor. Int. J. Hydrog. Energy 2021, 46, 15241–15256. [Google Scholar] [CrossRef]
  46. Zecevic, N.; Bolf, N. Advanced operation of the steam methane reformer by using gain-scheduled model predictive control. Ind. Eng. Chem. Res. 2020, 59, 3458–3474. [Google Scholar] [CrossRef]
  47. Ismagilov, I.Z.; Matus, E.V.; Kuznetsov, V.V.; Mota, N.; Navarro, R.M.; Kerzhentsev, M.A.; Ismagilov, Z.R.; Fierro, J.L.G. Nanoscale control during synthesis of Me/La2O3, Me/CexGd1−xOy and Me/CexZr1− xOy (Me = Ni, Pt, Pd, Rh) catalysts for autothermal reforming of methane. Catal. Today 2013, 210, 10–18. [Google Scholar] [CrossRef]
  48. Ledesma, C.; Llorca, J. CuZn/ZrO2 catalytic honeycombs for dimethyl ether steam reforming and autothermal reforming. Fuel 2013, 104, 711–716. [Google Scholar] [CrossRef]
  49. Nilsson, M.; Jansson, K.; Jozsa, P.; Pettersson, L.J. Catalytic properties of Pd supported on ZnO/ZnAl2O4/Al2O3 mixtures in dimethyl ether autothermal reforming. Appl. Catal. B-Environ. 2009, 86, 18–26. [Google Scholar] [CrossRef]
  50. Takeishi, K.; Akaike, Y. Hydrogen production by dimethyl ether steam reforming over copper alumina catalysts prepared using the sol–gel method. Appl. Catal. A-Gen. 2016, 510, 20–26. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the MPC in Simulink.
Figure 1. Schematic diagram of the MPC in Simulink.
Energies 15 09038 g001
Figure 2. MPC controller mechanism.
Figure 2. MPC controller mechanism.
Energies 15 09038 g002
Figure 3. Schematic diagram of the controlled pilot plant unit for the hydrogen production system.
Figure 3. Schematic diagram of the controlled pilot plant unit for the hydrogen production system.
Energies 15 09038 g003
Figure 4. Hydrogen generation system prototype.
Figure 4. Hydrogen generation system prototype.
Energies 15 09038 g004
Figure 5. Schematic diagram of the hydrogen volume fraction under temperature variation.
Figure 5. Schematic diagram of the hydrogen volume fraction under temperature variation.
Energies 15 09038 g005
Figure 6. Schematic diagram of the hydrogen demand profile.
Figure 6. Schematic diagram of the hydrogen demand profile.
Energies 15 09038 g006
Figure 7. Schematic diagram of the feedstocks supply rate profile ((a) DME; (b) High-purity air; (c) Water).
Figure 7. Schematic diagram of the feedstocks supply rate profile ((a) DME; (b) High-purity air; (c) Water).
Energies 15 09038 g007aEnergies 15 09038 g007b
Figure 8. Schematic diagram of the simulation and experiment hydrogen flow rate.
Figure 8. Schematic diagram of the simulation and experiment hydrogen flow rate.
Energies 15 09038 g008
Figure 9. Schematic diagram of the reactor temperature profile ((a) MPC control; (b) PID control).
Figure 9. Schematic diagram of the reactor temperature profile ((a) MPC control; (b) PID control).
Energies 15 09038 g009aEnergies 15 09038 g009b
Table 1. Parameters applied in the Equations (1)–(12).
Table 1. Parameters applied in the Equations (1)–(12).
ParametersAnnotationUnit
ωiMass fraction of component j-
AcCross-sectional area of the channelm2
RjOverall reaction rate for component jmol/s/m3
ϵpCatalyst porosity-
τFEffective transport factor-
uFluid velocitym/s
ρFluid densitykg/m3
Cp,sHeat capacity of solid J/K/kg
Cp,fHeat capacity of gasJ/K/kg
qFlux of heatW/m2
keffEffective thermal conductivityW/m/K
(ρCp)effHeat capacity per unit volume J/m3/K
QHeat production rate per unit volumeW/m3
Table 2. Reactions that take place in the DME autothermal reformer.
Table 2. Reactions that take place in the DME autothermal reformer.
ReactionsEnthalpyKinetics of the Reactions
DME Hydrolysis C H 3 O C H 3 + H 2 O 2 C H 3 O H ∆H = 24 kJ/mol r H = k H   exp ( E H R T )   C D M E
MeOH SR C H 3 O C H 3 + H 2 O C O 2 + 3 H 2 ∆H = 49 kJ/mol r S R = k S R   exp ( E S R R T ) C C H 3 O H C H 2 O
MeOH Decomposition C H 3 O H C O + H 2 ∆H = 90.1 kJ/mol r D = k D exp ( E D R T )
Partial Oxidation C H 3 O H + 1 / 2 O 2 2 C O 2 + 2 H 2 ∆H = −193 kJ/mol r P O = k P O   exp ( E P O R T ) C C H 3 O H C O 2
CO Oxidation C O + 1 / 2 O 2 C O 2 ∆H = −283 kJ/mol r C O X = k C O X   exp ( E C O X R T ) C C O
Water Gas Shift C O + H 2 O C O 2 + H 2 ∆H = −41.1 kJ/mol r W G S = k W G S   exp ( E W G S R T ) ( P C O   P H 2 O   P C O 2   P H 2 k e q , W G S )
k e q , W G S = exp ( E e q . W G S R T C e q , W G S )
Table 3. Reaction kinetic parameters.
Table 3. Reaction kinetic parameters.
ParametersAnnotation ValueUnit
kHPre-coefficient of Reaction i1163.232m3·s/kgcat
kSR154.7m6/(mol·s·kgcat)
kD99.7 × 103mol/(s·kgcat)
kPO400.5m6/(mol·s·kgcat)
kCOX2.1 × 105m3/(s·kgcat)
kWGS5000mol/(kgcat·s·Pa2)
Ceq,WGS3.33-
EHActivation Energy of Reaction i22.237kJ/mol
ESR57.9
ED110.8
EPO54.9
ECOX63.3
EWGS100
Eeq,WGS 40
RUniversal Gas Constant8.314J/K/mol
C i Concentration of Species i-mol/m3
TTemperature of Reactor-K
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, T.-Q.; Jung, S.; Kim, Y.-B. Hydrogen Production System through Dimethyl Ether Autothermal Reforming, Based on Model Predictive Control. Energies 2022, 15, 9038. https://doi.org/10.3390/en15239038

AMA Style

Zhang T-Q, Jung S, Kim Y-B. Hydrogen Production System through Dimethyl Ether Autothermal Reforming, Based on Model Predictive Control. Energies. 2022; 15(23):9038. https://doi.org/10.3390/en15239038

Chicago/Turabian Style

Zhang, Tie-Qing, Seunghun Jung, and Young-Bae Kim. 2022. "Hydrogen Production System through Dimethyl Ether Autothermal Reforming, Based on Model Predictive Control" Energies 15, no. 23: 9038. https://doi.org/10.3390/en15239038

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