Next Article in Journal
Brands’ Green Activism: An Empirical Comparison between Posts of Digital Influencers and Brands
Previous Article in Journal
Assessing Urban Ventilation in Common Street Morphologies for Climate-Responsive Design toward Effective Outdoor Space Regeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Study on Hydrogen Production via Waste Heat Recovery of a Natural Gas-Fueled Internal Combustion Engine in Cogeneration Power-Hydrogen Layouts: 4E Study and Optimization

by
Mohammad Zoghi
1,*,
Nasser Hosseinzadeh
2,
Saleh Gharaie
1 and
Ali Zare
1,*
1
School of Engineering, Deakin University, Geelong, VIC 3216, Australia
2
Department of Renewables and Distributed Energy, Network Connections, Energy Queensland, 26 Reddacliff St., Newstead, QLD 4006, Australia
*
Authors to whom correspondence should be addressed.
Sustainability 2024, 16(16), 6860; https://doi.org/10.3390/su16166860
Submission received: 24 May 2024 / Revised: 15 July 2024 / Accepted: 22 July 2024 / Published: 9 August 2024

Abstract

:
Internal combustion engines (ICEs) are one of the significant sources of wasted energy, with approximately 65% of their input energy being wasted and dissipated into the environment. Given their wide usage globally, it is necessary to find ways to recover their waste energies, addressing this inefficiency and reducing environmental pollution. While previous studies have explored various aspects of waste energy recovery, a comparative analysis of different bottoming configurations has been lacking. In this research, an extensive review of the existing literature was conducted by an exploration of four key bottoming cycles: the steam Rankine cycle (SRC), CO2 supercritical Brayton cycle, inverse Brayton cycle (IBC), and air bottoming cycle. In addition, these four main bottoming systems are utilized for the waste energy recovery of natural gas-fired ICE with a capacity of 584 kW and an exhausted gas temperature of 493 °C. For the efficient waste heat recovery of residual exhausted gas and heat rejection stage of the main bottoming system, two thermoelectric generators are utilized. Then, the produced power in bottoming systems is sent to a proton exchange membrane electrolyzer for hydrogen production. A comprehensive 4E (energy, exergy, exergy-economic, and environmental) optimization is conducted to find the best main bottoming system for hydrogen production. Results showed that the SRC-based system has the highest exergy efficiency (21.93%), while the IBC-based system results in the lowest efficiency (13.72%), total cost rate (25.58 $/h), and unit cost of hydrogen production (59.91 $/GJ). This combined literature review and research article underscore the importance of finding an economically efficient bottoming cycle in the context of waste energy recovery and hydrogen production.

1. Introduction

In the present day, the issues of energy scarcity and environmental pollution have become increasingly important [1,2]. Consequently, the value of energy conservation cannot be overstated. In a country with a thriving industrial sector, a significant portion of the total fossil fuel consumption—roughly 60–70%—is attributed to internal combustion engines (ICEs), with automotive ICEs accounting for 40–50% of this figure [3,4]. However, a considerable amount of the total heat generated by the combustion of fuel in ICEs—approximately 65%—is squandered through exhaust and engine coolant, resulting in energy waste and emission-related predicaments [5,6,7]. If this waste heat can be recuperated, there could be a substantial improvement in engine fuel efficiency and a marked reduction in environmental pollution [8]. Thus, waste heat recovery has gained paramount importance. In this regard, various bottoming systems have been proposed for the purpose of utilizing the waste heat generated by ICEs [9,10]. Rankine cycle (RC), CO2 supercritical Brayton cycle (SBC), and inverse Brayton cycle (IBC) are common proposed configurations for the waste energy recovery of ICEs. Although the above-mentioned systems were investigated in previous studies as bottoming systems of ICEs, there is a need in the literature for a comprehensive review and comparison of different systems for waste heat utilization of ICE. Hence, in the present study, these three cycles—steam RC (SRC), CO2 SBC, and IBC—in addition to the air bottoming cycle (ABC) are investigated and compared as bottoming systems for a natural gas ICE from 4E (energy, exergy, exergy-economic, and environmental) perspectives. It should be mentioned that ABC has not been used for the waste heat recovery of an ICE, and it is used here as bottoming system for ICE for the first time.
One of the primary challenges is to address the increasing global energy demand and ascertain sustainable and environmentally sound energy solutions to meet this demand. Currently, the predominant portion of the energy demand is met by fossil fuel supplies, but these materials are limited by geography. Fossil fuels possess the disadvantage of faster depletion and greenhouse gas emissions. Due to the restricted supplies, fossil fuels are not expected to meet the cumulative demand, and further research is being conducted on renewable energy systems [11,12]. Hydrogen is widely recognized as a potential fuel due to its ability to function as both an energy carrier and storage medium, in addition to offering carbon-free solutions [13,14]. In comparison to electricity, hydrogen possesses several advantages, including high energy conversion effectiveness, ample sources, the capability to be produced with zero emissions from water, long-distance transportability, availability of storage options, conversion into fuels, the potential to be generated through renewable energies, and the elimination of environmental damage [15,16]. Hydrogen also offers superior heating values in comparison to traditional fossil fuels. Conversely, electricity carries the disadvantages of transmission and heat losses caused by high voltages and electrical resistance. Hence, in the present research, the produced power in the waste recovery bottoming systems is sent to an efficient type of hydrogen production unit named proton exchange membrane electrolyzer (PEME). Therefore, the complete system is converted to an efficient power-hydrogen production unit.
This paper is structured in a way that it first conducts a thorough review of the papers related to combined ICE with Rankine cycle (RC), SBC, IBC, and ABC (Section 1.1, Section 1.2, Section 1.3 and Section 1.4) and related to hydrogen production using PEME (Section 1.5). Section 2 introduces the considered systems. Section 3 presents the basic assumptions and mathematical modeling of the systems. In the Result section, Section 4, after validation (Section 4.1) and base case inputs (Section 4.2), a thorough parametric study is conducted to investigate the effect of influential parameters on the systems (Section 4.3), and the results of the multi-objective optimization for choosing the best configuration are presented in Section 4.4. Finally, the conclusion and recommendations are presented in Section 5.

1.1. The Waste Heat Recovery of Engines by Rankine Cycles

Rankine cycles are favorable choices for the unused energy utilization of ICEs that has been used in many studies and different configurations. Some review papers are available about the waste energy recovery of engines by organic Rankine cycles (ORCs). In this regard, Tian et al. [17] examined the use of 20 various organic fluids in ORC for efficient waste heat utilization of a diesel engine (235 kW power with an exhaust temperature of 519). Using R141b, R123, and R245fa led to the highest efficiency between 13.3% and16.6%. The electricity production cost was estimated as 0.3–0.35 $/kWh. Yu et al. [18] used R245fa as a passing flow stream for ORC in an ICE unused energy recovery system. A thermal oil circuit was used between ICE exhausted gas and ORC to prevent the decomposition of ORC fluid in a high temperature of exhausted gas. Energy efficiency ( η e n ) and exergy efficiency ( η e x ) were 9.2% and 21.7%, respectively. Shu et al. [19] embedded a dual-loop Rankine cycle for unused energy utilization of a diesel engine with a capacity from 58.8 kW to 235.8 kW. Water was used in the high-temperature (HT) loop and six different organic fluids were used in the low-temperature (LT) loop and in subcritical or trans-critical mode. The HT loop heat source was exhausted gas, while the residual exhaust gas, engine coolant, and the heat rejection of the HT cycle were sources of energy for the LT loop. R134a was the best fluid in the LT loop with net electricity of 39.91 kW and η e x of 48.42%. Zhang et al. [20] introduced a novel dual-loop ORC with R245fa and R134a as working fluids in HT and LT loops. The exhausted gas of a 105 kW engine was the heat source of the HT loop, while intake air, heat rejection of the HT loop, and cooling water of engine were the sources of energy for the LT loop. The output electricity improvement was from 14% to 16% due to the waste heat recovery. Yang [21] investigated a dual-loop ORC with R245fa as a passing stream in both loops. The waste energy of exhausted gas, intake air path, and jacket cooling water of a 247 kW engine were used as energy sources of the dual-loop layout. The net output electricity of the bottoming system was 27.85 kW, and efficiency improvement was reported at 13%. Shu et al. [22] introduced a dual-loop Rankine system for the unused energy utilization of an engine with a maximum power of 235.8 KW and an exhausted temperature of 519. The HT loop fluid was water, and six organic fluids were used in the LT loop. R1234yf led to the highest output with a generated electricity of 36.77 kW and a η e x of 55.05%. Song and Gu [23] used a bottoming ORC for unused energy elimination of exhausted gas and jacket cooling water of a 996 kW engine. They used pure and zeotropic mixtures as working fluids of ORC. Using cyclohexane/R141b resulted in 13.3% more output power in comparison with pure cyclohexane.
Yue et al. [24] compared using ORC and Kalina cycle for the unused heat elimination of a 200–2000 kW diesel engine. The ORC passing fluids worked in trans-critical mode. N-nonane led to the highest efficiency at 64.1%. Song et al. [25] used exhausted gas and jacket cooling water of a diesel engine (996 kW power) as energy sources for HT ORC and LT ORC. The exhausted gas and cooling water were at 573.15 K and 363.15 K. Five and seven different working fluids were used for HT and LT loops. The highest generated power of the bottoming system was 101.1 kW. This led to the whole system’s efficiency improvement by 10.2%. Song and Gu [26] surveyed the energy–exergy outputs of a cascade partial evaporation Rankine cycle (PERC) (water) and ORC for the unused heat elimination of a 996 kW diesel engine. They used exhausted gas and cooling water from the engine as energy sources. Bottoming system power production was 115.1 kW.
Daghigh and Shafieian [27] used exhaust gas and jacket cooling water of a diesel engine as the heat source of an ORC, a hot water unit (HWU), and an absorption chiller (ACH) in a complex configuration. The bottoming system could produce 53 kW of power and 176.8 kW of cooling. Salek [28] et al. recovered the exhaust heat of a 250 kW engine in bottoming ORC and ACH. The exhaust gas of the engine was sent to the ORC evaporator and ACH generator in a series configuration. The output power and cooling were obtained as 6.73 kW and 14.2 kW, respectively. Habibi et al. [29] suggested using a cascade PERC/ORC/LNG subsystem for unused energy elimination of a 200 kW diesel engine. The exhausted gas acted as the energy source for the PERC evaporator and a HWU. Using isopentane led to the highest output. In this case, η e x and total cost rate ( C ˙ t o t ) were 38.74% and 19.3 $/h, respectively. Mohammadkhani et al. [30] used the exhausted energy of a 98.9 kW diesel engine as an energy source for an HT Kalian cycle. Jacket cooling water exhausted gas was used for preheating and evaporation of the NH3–H2O mixture of the Kalina cycle. The bottoming layout could generate 21.74 kW of power. This contributed to η e x and unit cost ( c p ) of 55.52% and 15.51 cent/kWh. Jafarzad et al. [31] proposed a new configuration comprising an ORC, a steam generator, a hot water unit, and a reverse osmosis desalination unit (RODU) for efficient unused energy elimination of a 603 kW capacity diesel engine. η e n and η e x were obtained as 82.82% and 54.1%.
A comprehensive review of papers based on the combination of ICE and RC is provided in Table 1. As was described in the literature review, dual-loop RCs were mostly used for waste heat elimination in ICEs. A few studies have been done about tri-generation systems using HWU, ACH, and RODU as bottoming systems for ICE. There is no study about the full waste heat recovery of ICE for hydrogen production, and the present study tries to fill this gap by introducing a new configuration.

1.2. The Waste Heat Recovery of Engine by CO2 SBC

As a promising alternative power block, the employment of SBCs has recently been proposed to replace the conventional water/steam Rankine cycle, as well as the ideal gas (air or helium) Brayton cycles, in waste heat recovery systems [57]. One of the key benefits of SBCs is their lower level of corrosiveness when compared to water, as well as the decreased compression work required in comparison to ideal gas Brayton cycles, ultimately leading to improved efficiency. Additionally, operating SBCs beyond the critical point (7.38 MPa, 31 °C) results in a smaller size of both turbo-machinery and heat exchangers when compared to that in conventional power cycles [58,59]. Hence, some studies have been done about the waste energy recovery of ICEs by CO2 SBCs.
For example, Liang et al. [60] proposed using a cascade regenerative CO2 SBC/ORC for the unused energy elimination of a diesel–natural gas engine (power between 126.8 kW and 251.1 kW). They used seven working fluids in the ORC. The output power of the bottoming system was 40.88 kW, leading to a 6.78% improvement in the whole system’s efficiency. Wu et al. [61] used the exhausted gas of a 2928 kW diesel engine as an energy source of a regenerative CO2 SBC. Then, the waste heat of SBC was recovered by an ACH. η e n and η e x were 39.85% and 53.54%. Furthermore, c p was calculated as 3.41 $/GJ. Feng et al. [62] used the exhausted gas of a 41,840 kW diesel engine as an energy source for a recompression CO2 SBC and Kalia cycle in a series configuration. The bottoming system’s output power was 1976.1 kW. Pan et al. [63] used regenerative CO2 SBC for unused energy elimination of a diesel engine. The generated electricity of SBC was fed to the compressor of an ejector refrigeration cycle for cooling production. Output cooling and coefficient of performance were 225.5 kW and 2.05. Almohana et al. [64] used a regenerative CO2 SBC and a modified Kalina cycle for the unused energy utilization of an ICE. Furthermore, an ORC and thermoelectric generator 1 (TEG 1) were used as bottoming systems for SBC, and TEG 2 was used to recuperate the unused energy of the Kalina cycle. The net output power was 2657 kW. This resulted in η e n and η e x of 26.33% and 51.69%. Wang et al. [65] used the waste gas of a 235.8 kW diesel engine (with a temperature of 519) in a cascade partial heating CO2 SBC and trans-critical CO2 Rankine cycle. η e n and electricity production cost were calculated as 29.25% and 7.43 cent/kWh. Wang et al. [66] recovered the exhausted waste heat of a 1323.1 kW engine by CO2 regenerative SBC and an ORC with a mixture of Benzene/R365mfc as a working fluid. Then, the unused heat of SBC was recuperated by an ejector refrigeration cycle. η e x and c p were estimated as 61.93% and 5.34 cent/kWh. Dadpour et al. [67] sent the exhausted gas of a ship engine to a new regenerative CO2 SBC for power and cooling and trans-critical CO2 ORC systems. Moreover, the ORC heat sink medium was a LNG subsystem. The proposed system could generate 9.06 kW power and 19.52 kW cooling, bringing about a η e x and total price of 42.3% and 2.5 M$.
A comprehensive review of papers based on the combination of ICE and SBC is provided in Table 2. In this case, the waste heat elimination of ICE was mainly conducted by SBC and ORC in series and cascade configurations. In some studies, other systems such as the ACH and Kalina cycles were used for utilization of the heat rejection stage in the SBC, and in the present research, the whole waste heat of bottoming systems is recovered by two TEG and the output power is directed to a PEME for hydrogen production.

1.3. The Waste Heat Recovery of Engines by IBC

IBC is another bottoming cycle that although it has a good potential for recovery of high-temperature exhausted gas, less attention is devoted to it in comparison with other bottoming cycles. The basic IBC has three main components: a gas turbine that expands the hot gas to sub-atmospheric pressure, a cooling heat exchanger to cool down gas before entering the compressor, and a compressor for increasing the pressure of gas to atmosphere pressure [78,79].
In this regard, Zhu et al. [80] recovered the waste energy of an asymmetric twin-scroll turbocharged diesel engine by an IBC. Output power was increased by about 5% due to using the IBC bottoming system. Battista et al. [81] used IBC for the unused energy recovery of a turbocharged diesel engine. The effect of important parameters such as turbine and compressor efficiencies and pressure drops were investigated on the combined system. The waste heat recovery leads to the production of an extra 1.5–2% of engine output power. Chagnon-Lessard et al. [82] proposed using different configurations of IBC as bottoming systems for a methane-fueled engine. Inverted Brayton cycle with liquid water drainage, steam turbine, and refrigeration cycle had the highest amount of specific work among different configurations. The highest thermal indicator of 25% was obtained for the considered layout. Abrosimov et al. [83] used the IBC cycle as the bottoming system for a methane engine with a capacity between 500 and 1400 kW. An ORC and a regenerative ORC were coupled with the IBC heat rejection heat exchanger through a thermal oil loop. Using pentane in regenerative ORC led to the output power of 152.9 kW for the IBC/ORC system. η e n and LCOE were obtained as 11.09% and 159.5 $/MWh. Salek et al. [84] used cascade IBC/ORC for unused energy elimination of a turbo-charged industrial diesel engine. They used R245fa as a passing stream in the ORC. The bottoming IBC/ORC enhanced the generated electricity of the layout by 18%.
A comprehensive review of papers based on the combination of ICE and IBC is provided in Table 3. As clarified in the literature review, limited studies have been conducted on the waste energy recovery of ICE by IBC. In some studies, an ORC was used for waste heat elimination in the heat rejection stage of ICE, while the exhaust gas was not recovered for extra power generation. In the present research, the energy of both heat rejection and exhausted gas in IBC was recovered by two TEGs for the first time. Then, the output power of bottoming systems was sent to a PEME for hydrogen output.

1.4. The Waste Heat Recovery of Engines by ABC

ABC is composed of three main components: air compressor, heating air heat exchanger, and air turbine. Some studies have been done on the combination of the gas turbine cycle (GTC) and ABC [87,88]. These studies showed that ABC can be an economically competitive option to a conventional system (GTC-SRC), especially in small-scale power generation [89]. On the other hand, despite the high potential of ABC for exhausted gas recovery, there is no study in the literature on using ABC for engine waste heat recovery. Therefore, this is the first study that compares using ABC with other common unused heat elimination bottoming systems for an NG-fueled engine.

1.5. Hydrogen Production by PEME

PEMEs possess a multitude of advantages in comparison to alternative electrolysis technologies [90,91]. Said benefits encompass superior energy efficacy, heightened hydrogen production velocity, a more condensed configuration, augmented hydrogen purity, and diminished power utilization [92,93].
Emadi et al. [94] utilized geothermal energy as the primary source of heat, while the LNG subsystem functioned as the heat sink in a cascade Kalina cycle/Stirling engine. The system’s output power was subsequently consumed in a PEME. The hydrogen produced by the system amounted to a rate of 204.77 kg/h, and η e x was 43.46%. Nami et al. [95] successfully recuperated the residual energy from a gas turbine modular helium reactor setup by means of two ORCs. The resultant residual energy from the ORC was then conveyed to a PEME for hydrogen production. The optimal conditions resulted in an η e x of 49.21% and an output of hydrogen totaling 56.2 kg/h. Akrami et al. [96] employed a cascade ORC/absorption cooling and heating system coupled with a heat pump-driven geothermal heat source. The resultant power generated by the ORC was subsequently directed to a PEME to facilitate the production of hydrogen. η e n and η e x were determined to be 34.98% and 49.18%, respectively. The minimum c p was computed to be 22.73 $/GJ. Habibollahzade et al. [97] employed the residual energy generated by a solid oxide fuel cell to power a bottoming Stirling engine. The resulting power output of the Stirling engine was directed toward a PEME unit. The most favorable η e x and c p were computed to be 38.03% and 24.93 $/GJ, respectively. Ishaq and Dincer [98] conducted a comparative analysis of three renewable energy-based systems, namely biomass, solar, and geothermal, for the purpose of hydrogen production using PEME. The exergy of the aforementioned systems, specifically the biomass gasification system, geothermal power system, and solar photovoltaic system, were determined to be 49.8%, 10.2%, and 16.95%, respectively.
Akrami et al. [99] employed the thermal and electrical properties of photovoltaic/thermal cells to operate an air conditioning system and PEME. The suggested system underwent scrutiny from both an energy and exergy-economic standpoint. The system’s all-encompassing η e x came to 11.28%. Additionally, photovoltaic/thermal cells constituted ~80% of the total exergy depletion. Habibollahzade et al. [100] employed PTSC as an energy source for an SRC. The recovery of waste energy from the SRC was accomplished through a TEG. Thereafter, the resultant electricity output from the SRC was transmitted to a PEME unit. The utilization of multi-objective optimization yielded a η e x and c p of 12.76% and 61.69 $/GJ, respectively, for the system. Ahmadi et al. [101] used an ORC in a hybrid combination of solar and ocean thermal energy conversion systems. The resultant power output of the ORC was then transmitted to a PEME unit for the purpose of hydrogen production. The integrated system’s η e n and η e x were calculated to be 3.6% and 22.7%, respectively. Moharramian et al. [102] employed a combined ORC/PEME system as a bottoming cycle in a GTC fueled by natural gas and hydrogen, with a post-combustion chamber fed by biomass. The hydrogen produced was subsequently introduced into the combustion chamber of the GTC. The integration of hydrogen injection resulted in a noteworthy decrease of 37% and 27% in the exergy and CO2 emissions, respectively.

2. Systems Description

Figure 1, Figure 2, Figure 3 and Figure 4 show the schematic diagrams of the suggested setups. An NG-fired ICE with a 584 kW capacity and a 493 °C exhaust gas temperature serve as a topping system in all versions. SRC, regenerated CO2 SBC, IBC, and ABC play the primary bottoming system roles in Configurations 1–4. Two TEGs are used in bottoming systems to recover all waste energy. The remaining heat of exhausted gas is converted by a TEG to 70 °C, which is the lowest temperature of exhausted gas. For the recovery of heat rejection stages in main bottoming systems, another TEG is employed. A generator uses the output power of the topping GTCs to generate electricity, while a PEME makes hydrogen using the output power of the main bottoming systems plus TEG 1 and TEG 2. The Appendix has more information on the calculations and formulas for each point, and a point-to-point explanation of systems is provided below.
In Figure 1, exhaust gas (Point 1) exits the ICE at high temperature, entering the HRSG. The exhaust gas transfers heat to the working fluid (water) in the HRSG, generating high-pressure steam (Point 6). The high-pressure steam flows to the steam turbine. The steam expands through the turbine, generating mechanical work and converting it to electrical power. The low-pressure steam exits the turbine (Point 7) and moves to TEG 2. In TEG 2, the steam releases heat and condenses back into water (Point 8). The condensate water is pumped to increase its pressure before returning to the HRSG (Points 8 to 9). Then, any remaining heat from the exhaust gas (Point 2) is transferred to TEG 1, converting temperature differences directly into electrical energy, enhancing overall energy recovery efficiency.
In this study, the focus was on the SRC due to its superior performance for high-temperature waste heat recovery (above 400 °C). The SRC is particularly effective in utilizing the high-temperature exhaust gases from the ICE, allowing for more efficient energy recovery compared to ORCs, which are generally more suitable for lower temperature ranges. It is believed that the SRC offers a better match for the temperature conditions in this application, providing an optimized solution for maximizing energy recovery.
Throughout the electrolysis process, the PEME is supplied with the heat and electricity required to fuel the electrochemical operations. Liquid H2O is supplied to the system at the reference environment parameters of 25 °C and 1 atm (Point 10). Before H2O enters, the TEG (Points 10 to 11) and a counter-flow heat exchanger (Points 11 to 12) boost its temperature to that of the PEME. As the H2 exits the cathode (Point 14), heat is released into the surrounding air, where it cools to the temperature of the reference environment. The O2 gas produced at the anode is cooled to the temperature of the reference environment (Point 16) following its isolation from the H2O/O2 mixture (Point 15). The residual hot H2O (Point 17) is recycled into the H2O supply stream (Point 12) for the subsequent cycle of H2 production. Throughout the entire PEME reaction, H2O is transformed into H2 and O2, along with energy and heat. PEME has a surface area of 1 m2.
In Figure 2, the process begins with the hot exhaust gases exiting the ICE at Point 1, carrying high temperature and pressure from the combustion process. As these gases flow through the HRHX at Point 2, they transfer heat to the supercritical CO2 working fluid, causing the exhaust gases to cool down. The supercritical CO2 enters the cycle at Point 6, absorbing heat from the exhaust gases and expanding through the turbine (Points 6 to 7), where it undergoes adiabatic expansion, converting thermal energy into mechanical work that drives a turbine generator. The CO2 exits the turbine at Point 7, now at a lower pressure and temperature, but still in the supercritical state. In the recuperator, the high-temperature CO2 (Points 7 to 8) transfers heat to low-temperature CO2 (Points 10 to 11), improving overall cycle efficiency by preheating the incoming CO2. Exiting the recuperator at Point 8, CO2 is at a lower temperature and ready for compression in the compressor, where it is recompressed to high pressure, requiring mechanical work input. Before that, the temperature of CO2 is reduced in TEG 2 (Points 8 to 9), converting the waste energy to useful power. At point 10 and after compressor, CO2 enters the high-pressure side of the cycle again, now at a higher temperature and pressure, prepared to absorb heat from the heat exchangers in the recuperator and HRHX. Finally, at Point 2, TEG 1, located at the exhaust system exit, converts part of the waste heat directly into electricity through the Seebeck effect, utilizing the temperature difference between the exhaust gases and the ambient water to generate additional electrical power, thereby maximizing energy recovery efficiency.
In the waste heat recovery system using an IBC (Figure 3), the process begins at point 1, where hot exhaust gases exit the ICE at high temperature and pressure. The exhaust gases then enter the turbine at Point 1, still retaining substantial thermal energy. As the gases expand to sub-atmospheric pressure through the turbine, they produce mechanical work by driving the turbine. The gases exit the turbine at Point 2, now at a lower pressure and temperature due to the energy extracted. In TEG 1, the exhaust gases are further cooled, producing extra electricity. Finally, at point 3, the cooled, low-pressure gases are drawn into the compressor, where they are pressurized to ambient pressure. Furthermore, at Point 4, these gases transfer their heat to TEG 2, which completely converts this waste heat into electricity via the Seebeck effect. This system efficiently recovers waste heat from the engine, converting it into useful electrical energy.
In the waste heat recovery system utilizing an ABC (Figure 4), the process begins at point 1, where hot exhaust gases exit the ICE. At this stage, the hot gases enter the AHX, transferring heat to the working fluid (air) of the bottoming ABC. In the ABC, ambient air is compressed by the ABC compressor, moving from points 6 to 7. The hot, high-pressure air at point 7 then absorbs energy from the exhaust gases, becoming sufficiently hot to enter the ABC turbine at point 8 and produce bottoming cycle output power. The remaining exhaust gas at point 2 and exhaust air at point 9 enter thermoelectric generators (TEG 1 and TEG 2, respectively) to fully utilize the system’s residual energy, thereby generating additional power.

3. Modeling with Mathematics

The simulations of energy, exergy, exergy-economic, and environment related to the previously specified configurations have been carried out by means of suitable correlations included in the Engineering Equation Solver (EES). Following that, the EES software’s (Version 11.644) outputs have been coupled to the MATLAB NSGA-II toolbox through the use of artificial neural network (ANN) modeling. The following list of simplifying assumptions has been taken into consideration when the simulation process was started [29,77,103,104,105]:
1.
The equations for conversion have been inscribed while adhering to the conditions of a constant state.
2.
In heat exchangers, the gas streams have a 3% pressure drop.
3.
The energy and exergy equations’ kinetic and potential components have been ignored.
4.
It is assumed that the isentropic efficiencies do not change for any turbo-machines.
5.
Correlations pertaining to the mixture of ideal gases are utilized to compute the characteristics of gas and air mixtures.
6.
All components are fully insulated and have adiabatic performance.
7.
The mole fraction of ICE exhausted gas is N2: 72.5%, O2: 9.5%, H2O: 9.8%, CO2: 8.2%.
8.
The minimum temperature of the exhaust gas that is exiting the TEGs is regarded as being equivalent to 70 °C.
9.
In the bottoming SRC, the pump inlet fluid is saturated liquid.
10.
The values of ambient pressure and temperature are deemed as the reference standards for any analysis involving exergy.
Table 4 shows the essential equations for the TEG and PEME simulation.

3.1. Thermodynamic and Exergy-Economic Analyses

The formulas pertaining to mass, energy, exergy, and exergy-economic simulations of entire configurations are delineated below [110,111,112]:
m ˙ i = m ˙ e
Q ˙ + m ˙ i h i = W ˙ + m ˙ e h e
E x ˙ Q + m ˙ i e x i = E x ˙ w + m ˙ e e x e + E x ˙ dest
c Q E x ˙ Q + c i E x ˙ i + Z ˙ = c e E x ˙ e + c w W ˙
According to the Refs. [111,113], Appendix A contains the fuel and product exergies as well as the energy balance equations for each of the various devices in the arrangements. Furthermore, Appendix B presents the exergy-economic and associated complimentary relationships in a tabular style for each of the configurations’ devices. Moreover, Appendix C provides an exact demonstration of the principal expenses required for the procurement of the component parts. The variables related to the cost of procurement for each component are then listed in Appendix D.

3.2. Output Parameters

The exergy efficiency is computed by means of the following equation:
η e x = E x ˙ h y d r o g e n E x ˙ 1
Then, for the purposes of exergy-economic evaluation, the total cost rate is calculated in the following manner:
C ˙ t o t = Z ˙ k + C ˙ d e s t , k +   C ˙ env
Finally, the environmental cost rate is determined as below [114]:
C ˙ env = 0.024   m ˙ C O 2

3.3. Optimization

It should be remembered that energy conversion systems frequently face competing goals that need to be resolved at the same time. Building an efficient system usually costs more than building a less efficient one. It is possible to obtain a group of optimal solutions while working with several goals. The NSGA-II sub-program in MATLAB software (R2023b) has been used in the current study to identify a set of optimal spots. The parametric study in Section 4.3 reveals that three output parameters, η e x , C ˙ tot , and c hydrogen , have opposite trends. Therefore, they are convenient for a three-objective optimization problem. It is worth mentioning that η e x should be maximized and C ˙ tot and c hydrogen should be minimized:
O b j f u n 1 = max ( η e x )
O b j f u n 2 = min ( C ˙ t o t )
O b j f u n 3 = min ( c h y d r o g e n )
The proper input parameters and their considered range in the optimization problem are tabulated in Table 5.
The process of transferring data from the EES stage to the optimization stage involves several steps. First, numerical data are extracted from EES, which includes the necessary parameters and performance metrics of the systems under study. These data are then imported into MATLAB, where it is used to train an ANN. The ANN is trained to establish a relationship between the design parameters and the objective functions. Once the ANN is trained, it produces m-files that encapsulate this relationship. These m-files are used in the optimization stage, where the NSGA-II algorithm in MATLAB optimizes the design parameters based on the ANN’s predictions. The optimization process is performed entirely in MATLAB, utilizing the ANN’s output to efficiently navigate the design space and identify optimal solutions. This approach ensures that the complex, nonlinear relationships captured by the ANN are accurately reflected in the optimization process.

4. Results and Discussion

Section 4.1 of this part contains the systems’ validation reports. Section 4.2 shows the simulation’s input values as well as the entire work flowchart. Section 4.3 reports on a thorough parametric investigation of four configurations, and Section 4.4 presents the optimization results.

4.1. Validation

The entire investigation is simulated using the EES. Table 6 first presents a comparison of the outcomes of regenerative SBC with those of earlier research. Furthermore, Figure 5 compares the cell potential–current density of PEME to a related study. The fact that the output parameters only differ slightly suggests that the simulation is correct.
Table 6. Regenerative SBC results validation using the same input data *.
Table 6. Regenerative SBC results validation using the same input data *.
State PointTemperature (°C)Temperature (°C)
[115]
Pressure   ( kPa ) Pressure   ( kPa )
[115]
14323273607360
1562.6965.920,00020,000
16323.7323.920,00020,000
1755055020,00020,000
18435.4434.720,00020,000
1972.1975.473607360
* Isentropic efficiency of compressor = 0.86, Isentropic efficiency of gas turbine = 0.93, Inlet temperature to compressor = 32 °C, Inlet temperature to turbine = 500 °C, Recuperator effectiveness = 0.86.
Figure 5. PEME’s cell potential–current density in comparison with [116].
Figure 5. PEME’s cell potential–current density in comparison with [116].
Sustainability 16 06860 g005

4.2. Baseline Condition Inputs

Table 7 displays the fundamental requirements for conducting the simulation, while Figure 6 outlines the complete process of simulation and optimization using EES and MATLAB software.

4.3. Parametric Study

In this part, the impact of changes in influential parameters in four bottoming systems is investigated on η e x , Ex ˙ dest , tot , C ˙ tot , and c hydrogen . It is of great importance to note that solely the parameter under consideration experiences alterations within the desired range, while the remaining parameters remain unaltered, as demonstrated in Table 7.

4.3.1. SRC Parametric Analysis

The influence of fluctuation in the evaporation temperature of the HRSG, spanning from 170 °C to 300 °C, on the overall performance of the system (Configuration 1) is demonstrated in Figure 7. An increase in evaporation temperature, with the consideration of the energy balance equation in HRSG, results in a decrease in m ˙ S R C . Conversely, an increase in enthalpy difference in the SRC turbine leads to a rise in SRC output power. In TEG 2, a decrease in passing flow stream reduces power output, whereas in TEG 1, an increase in T2 contributes to an enhancement in power. An increase in the power of SRC and TEG 1 has a positive impact on W ˙ P E M E , m ˙ h y d r o g e n , and η e x . This leads to a notable rise in η e x from 18.89 to 20.82% (Figure 7a). Additionally, a reduction in E x ˙ d e s t of HRSG leads to a decrease in Ex ˙ dest , tot from 144.1 to 139.6 kW. Despite an increase in investment cost rate, Z ˙ , of SRC turbine and PEME, a decrease in C ˙ dest of HRSG results in a drop in C ˙ tot from 32.97 $/h to 31.43 $/h. Subsequently, c hydrogen is reduced from 117 to 101.7 $/GJ (Figure 7b).
The illustration of the system’s performance (Configuration 1) is shown in Figure 8, with regard to the temperature at the pinch point, which spans from 5 to 50. A rise in the pinch point temperature leads to a decrease in m ˙ S R C , resulting in a decline in W ˙ S R C and W ˙ T E G   2 . Conversely, an increase in T2 results in a surge in W ˙ T E G   1 . The former holds more power and causes a decrease in W ˙ P E M E and m ˙ h y d r o g e n . Therefore, η e x decreases from 22.67% to 19.82%. The primary reason for the increase in Ex ˙ dest , tot from 135.3 kW to 141.9 kW (Figure 8a) is the rise in E x ˙ d e s t of TEG 1. Furthermore, a reduction in C ˙ dest of the PEME and a decrease in Z ˙ of HRSG are the critical factors contributing to a decline in C ˙ tot , from 32.67 $/h to 31.62 $/h. It is worth noting that c hydrogen has experienced an increase from 104.5 $/GJ to 105.9 $/GJ, as illustrated in Figure 8b.
Figure 9 illustrates the impact of varying the superheating temperature of the HRSG within the range of 50 °C to 150 °C on the system’s performance, as noted in Configuration 1. An increase in the superheating temperature results in a decline in m ˙ S R C , which subsequently causes a reduction in W ˙ S R C and W ˙ T E G   2 . Conversely,   W ˙ T E G   1 increases due to the expansion of T2. This trend, however, results in a minimum for m ˙ h y d r o g e n and η e x at a superheating temperature of 88 °C. Additionally, a maximum for Ex ˙ dest , tot occurs at this point, as depicted in Figure 9a. Furthermore, according to Figure 9b, C ˙ tot decreases from 32 $/h to 31.81 $/h, and c hydrogen drops from 105.4 to 104.1 $/GJ.
Figure 10 illustrates the impact of alterations in the condenser temperature difference, ranging from 30 to 50, on the system’s performance in accordance with Configuration 1. In this instance, m ˙ S R C remains constant, while a decrease in the enthalpy difference of the turbine leads to a decline in the SRC output power. Conversely, the output power of TEG 1 and TEG2 increases, with the former exhibiting greater power, resulting in a reduction in W ˙ P E M E and m ˙ h y d r o g e n . Consequently, η e x experiences a drop from 21.82% to 20.34% (as indicated in Figure 10a). The primary factor contributing to the increase in Ex ˙ dest , tot from 137.3 kW to 140.7 kW is the rise in TEG 2 exergy destruction. Then, a reduction in Z ˙ and C ˙ dest of SRC turbine and PEME is the main reason for a fall in C ˙ tot from 32.05 $/h to 21.77 $/h. Ultimately, c hydrogen goes down from 106.2 $/GJ to 103.5 $/GJ (Figure 10b).

4.3.2. SBC Parametric Analysis

Figure 11 illustrates the impact of varying inlet temperature of the compressor on the system’s performance (Configuration 2). An elevation in temperature results in an increase in m ˙ S B C . Consequently, the power of the SBC turbine, TEG 1, TEG 2, and SBC compressor also increases. The SBC compressor’s effect is dominant, leading to a decline in W ˙ P E M E and m ˙ h y d r o g e n . As a result, η e x experiences a reduction from 16.53% to 13.96% (Figure 11a). An increase in E x ˙ d e s t of TEG 1 and TEG 2 contributes to a rise in Ex ˙ dest , tot from 149.6 kW to 155.6 kW. The primary cause for the decrease in C ˙ tot from 32.93 $/h to 30.07 $/h can be attributed to the decline in C ˙ dest of the SBC recuperator and PEME, as well as Z ˙ of PEME. It is worth noting that c hydrogen has also experienced a decline, decreasing from 102.5 $/GJ to 87.74 $/GJ (Figure 11b).
The effects of fluctuation in the inlet temperature of the compressor, ranging from 7400 kPa to 11,000 kPa, on the system’s performance (Configuration 2) are illustrated in Figure 12. In this particular scenario, a decline in m ˙ S B C results in a decrease in the power of SBC turbine, TEG 2, and SBC compressor. Additionally, a reduction in T2 leads to a decline in TEG 1 output power. The aforementioned trends, however, lead to a maximum η e x of 15.98% in Pin,com = 9400 kPa (Figure 12a). Furthermore, a decrease in TEG 1 and an increase in recuperator exergy destruction results in a minimum value for Ex ˙ dest , tot at this juncture. Furthermore, an increase in C ˙ dest of both PEME and SBC recuperator, along with a surge in Z ˙ of PEME, contributes to the escalation of C ˙ tot from 30.3 $/h to 31.39 $/h. Additionally, it is noteworthy that c hydrogen attains its maximum value of 95.14 $/GJ at Pin,com = 10,200 kPa, as illustrated in Figure 12b.
Figure 13 reveals the impact of variations in compressor pressure ratios between 1.2 and 3 on the performance of the system (Configuration 2). An increase in pressure ratio leads to a rise in the SBC turbine, TEG 2, and SBC compressor power. Conversely, a decrease in T2 results in a reduction in TEG 1 power. In conclusion, an increase in turbine output power has a significant effect, resulting in an improvement in W ˙ P E M E and m ˙ h y d r o g e n . As a result, η e x increases from 6.46% to 16.11%. Ex ˙ dest , tot decreases from 173 kW to 150.6 kW due to a decrease in E x ˙ d e s t of the recuperator and TEG 1 (Figure 13a). Moreover, the primary cause for the decrease in C ˙ tot from 47.86 $/h to 29.67 $/h can be attributed to a dip in Z ˙ and C ˙ dest of these two components. Additionally, within the range of consideration, c hydrogen experiences a decline from 195.4 $/GJ to 86.41 $/GJ (Figure 13b).
Figure 14 depicts the ramifications of altering the outputs consequent to an increase in the hot side temperature differential of the heater, ranging from 10 °C to 50 °C. This escalation in temperature disparity leads to a surge in m ˙ S B C , which in turn leads to a rise in the power of turbine, compressor, and TEG 2. Subsequently, there is a decrease in the TEG 1 power due to a decline in T2. The increase in the turbine of SBC and TEG 2 leads to an improvement in PEME power, thereby resulting in an advancement in η e x from 14.77% to 15.22% (as shown in Figure 14a). The pivotal factor for the decrement in Ex ˙ dest , tot from 153.7 kW to 152.6 kW is the reduction in E x ˙ d e s t of TEG 1. Furthermore, it should be noted that an increase in SBC cost rates and a decrease in TEG 1 cost rate result in a minimum point of 30.36 $/h for C ˙ tot at a temperature differential of 21.1 °C. It is interesting to observe that at this point, c hydrogen reaches its minimum value of 90.16 $/GJ (as depicted in Figure 14b).

4.3.3. IBC Parametric Analysis

The impact of alterations in inverse Brayton turbine back pressure (IBTBP) ranging from 30 kPa to 80 kPa on the output indicators is shown in Figure 15. An escalation in IBTBP results in a decline in IBC turbine power, TEG 2 power, and IBC compressor power. Consequently, the output power of TEG 1 surges owing to an increase in T2. These aforementioned parameters collectively lead to a decrease in PEME input power, thereby causing a reduction in η e x from 13.02% to 11.39% (Figure 15a). Furthermore, E x ˙ d e s t , t o t experiences an upsurge from 158.2 kW to 162 kW due to the amplification of TEG 1 exergy destruction. A decrease in C ˙ dest of TEG 2 as well as a reduction in Z ˙ of IBC turbine, TEG 2, and PEME are the principal factors contributing to the diminution of C ˙ tot from 26.81 $/h to 25.74 $/h. Subsequently, c hydrogen declines from 60.15 $/GJ to 58.33 $/GJ within the examined range (Figure 15b).
Figure 16 presents an analysis of the impact of varying the inlet temperature of the IBC compressor within the range of 30 °C to 100 °C on the system’s performance. Notably, this increase leads to a rise in the power of both the IBC compressor and TEG 2. However, the TEG 1 output power experiences a decrease. Consequently, the input electricity to PEME decreases, which results in a reduction in η e x from 14.89% to 12.11% (as shown in Figure 16a). Moreover, Ex ˙ dest , tot increases from 153.8 kW to 160.3 kW due to the rise in TEG 2 exergy destruction. The upsurge in Z ˙ and C ˙ dest of TEG 2 is the main contributing factor to the enhancement of C ˙ tot from 25.55 $/h to 26.2 $/h (as illustrated in Figure 16b). Ultimately, c hydrogen decreases from 60.73 $/GJ to 59.14 $/GJ.

4.3.4. ABC Parametric Analysis

Figure 17 illustrates the impact of varying the pressure ratio of ABC from 3 to 7 on the system’s outputs. An increase in the pressure ratio results in a corresponding increase in the ABC compressor, ABC turbine, and TEG 1. Conversely, the output power of TEG 2 decreases. These aforementioned parameters collectively contribute to a reduction in W ˙ P E M E and m ˙ h y d r o g e n . Consequently, η e x reduces from 13.68% to 8.46% (as shown in Figure 17a). The primary cause of the increase in Ex ˙ dest , tot from 156.2 kW to 168.4 kW is due to the increase in E x ˙ d e s t of the ABC turbine and TEG 1. Moreover, an increase in C ˙ dest of the ABC compressor, ABC turbine, and TEG 1 results in a rise in C ˙ tot from 46.3 $/h to 55.64 $/h. Lastly, c hydrogen surges by 53.07% within the examined range (Figure 17b).
Figure 18 illustrates the impact of the variation in hot side temperature difference of the AHX, ranging from 10 °C to 50 °C, on the output indicators. An increase in temperature difference leads to an increase in the ABC passing stream, resulting in a rise in the ABC compressor and turbine work. Consequently, the output power of TEG 2 decreases due to a decline in T9. Overall, the input power to the electrolyzer and hydrogen production decreases, leading to a dip in η e x from 12.35% to 10.52% (Figure 18a). Ex ˙ dest , tot increases from 159.3 kW to 163.6 kW because of an increase in E x ˙ d e s t of AHX. A decrease in C ˙ dest of TEG 2 and PEME, coupled with a reduction in Z ˙ of AHX, leads to a notable decrement of C ˙ tot from 54.27 $/h to 47.44 $/h. As a result, c hydrogen experiences a decline from 145.5 $/GJ to 132.2 $/GJ, as illustrated in Figure 18b.

4.4. Optimization Results

The present section showcases the outcomes linked with the NSGA-II algorithm and the three-objective optimization. In view of the fact that the objectives in question, namely, η ex , C ˙ t o t , and c h y d r o g e n , are in conflict with each other, the Pareto frontier and its associated numerical achievements are depicted in Figure 19, Figure 20, Figure 21 and Figure 22 and Table 8 and Table 9, respectively. Table 8 is related to design variables, and Table 9 is regarding objective functions. Then, all of the red circles within this frontier display the optimal points that are not dominated by any other points. Therefore, the results for the tradeoff point are reported in Table 9. Accordingly, the highest η e x is related to the SRC-based system (21.93%), while the lowest C ˙ tot (25.58 $/h) and c hydrogen (59.91 $/GJ) are for the IBC-bases system. Furthermore, the specific cost of the system (=total cost rate/exergy of product), which is a tradeoff between exergy and economic parameters, is minimum for the SRC-based system (0.6159 $/kWh). Hence, the SRC-based and IBC-based systems have been chosen as the best configurations for the waste energy recovery of high-temperature exhaust heat of ICE.
Then, the main output parameters of each layout at the optimum point are tabulated in Table 10. The lowest amount of hydrogen production and exergy efficiency is related to the ABC-based system. Then, the IBC-based system has the second stage. Despite the IBC-based system producing low hydrogen and having low η e x , the lowest C ˙ tot and c hydrogen are related to this configuration. As is clear from Table 10, the addition of TEGs to the systems considerably improves the output power, especially in cases of TEG 1 in the IBC-based system (25.84 kW output power) and TEG2 in the ABC-based system (16.16 kW output power). Therefore, the waste energy recovery of systems by TEGs is justifiable. Furthermore, in an IBC-based system, for example, the exergy of exhausted gas is 235.5 kW, of which 32.31 kW is converted to useful output productions, bringing about a η e x of 13.72%. Then, 66.49% (156.6 kW) of fuel exergy is turned out to exergy destruction and 19.81% (46.67 kW) is changed to exergy loss. In addition, Z ˙ tot , C ˙ dest , tot , and C ˙ env rate make up 7.74%, 53.32%, and 38.93% of C ˙ tot of IBC-based system (25.58 $/h). As is obvious, the IBC-based system has the lowest amounts of Z ˙ tot (1.98 $/h) and C ˙ dest , tot (13.64 $/h) among different configurations. This occurs because of the lowest number of components of IBC in comparison with other bottoming systems. Finally, the IBC-based system leads to the lowest c el in the bottoming cycles (25.61 $/GJ) and, consequently, the lowest c hydrogen (59.91 $/h).
The primary energy-exergy-economic parameters in the various system components at the optimal point are shown in Table 11. As is clear from Table 11, important components in terms of exergy destruction are PEME (Configuration 1 and Configuration 2), TEG 1 (Configuration 3), and TEG 2 (Configuration 4). Hence, exergy efficiencies of all components are higher than 50% except for TEG 1, TEG 2, and PEME (in Configuration 1 and Configuration 2). Moreover, the biggest investment cost rate is related to SRC in Configuration 1 and PEME in Configurations 2 and 3, while in Configuration 4, the AHX has the highest investment cost rate. This is the main reason why Configuration 4 has the highest C ˙ tot among the four considered configurations. Regarding the exergy destruction cost rate, in Configurations 1 and 2, the PEME has the highest values. Then, in Configurations 3 and 4, TEG 1 and TEG 2 are important components. This occurs because of the high values of TEG1 power in Configuration 3 and TEG 2 power in Configuration 4.

5. Practical Implementation, Limitations, and Future Research

The practical implementation of waste heat recovery systems, such as the SRC, SBC, IBC, and ABC, in ICEs presents a promising opportunity to enhance energy efficiency, reduce operational costs, and contribute to environmental sustainability. These systems can be scaled and adapted to various engine sizes, integrated into existing infrastructures with minimal disruption, and be economically viable with proper financial incentives. To ensure reliability and widespread adoption, it is essential to focus on robust designs, regular maintenance, and training for maintenance personnel. Additionally, supportive government policies and regulations will play a crucial role in facilitating the implementation of these technologies. Standardized interfaces and collaboration with engine manufacturers can further ease the integration process, ensuring compatibility and efficiency. By converting waste heat into useful energy forms, such as electricity and hydrogen, these technologies align with global sustainability goals, reducing greenhouse gas emissions and fossil fuel consumption while promoting the use of cleaner energy sources. Our study demonstrates the potential of these systems to make significant contributions to energy efficiency and environmental sustainability, providing a strong foundation for future research and practical applications.
It is acknowledged that simulations may not capture all real-world variables and challenges. However, the insights gained from theoretical models are crucial in the early stages of research, guiding future experimental studies to identify key parameters and potential issues before large-scale implementation. It is recognized that the systems presented, particularly the SBC, involve high implementation complexity and may entail significant costs. This feasibility analysis highlights these aspects and aims to provide an understanding of the potential benefits and challenges associated with these advanced systems. The high complexity is intrinsic to the innovative nature of the technologies explored. However, it is through detailed theoretical studies that can pave the way for more practical and cost-effective solutions in the future. While this study focuses on theoretical validation, it sets the stage for future experimental research. This study emphasizes the importance of experimental studies to further validate the findings and address real-world challenges. The current work provides a necessary first step, offering a detailed theoretical framework that can inform and guide experimental designs. It is anticipated that future research will build on these findings, incorporating experimental data to refine and enhance the models.
The advancement of waste energy recovery systems in ICEs holds significant potential for future research and practical applications. Future studies can focus on the integration of novel materials and advanced manufacturing techniques to enhance the efficiency and cost-effectiveness of bottoming cycles. Additionally, exploring hybrid systems that combine multiple waste heat recovery technologies could provide even greater efficiency gains. Experimental validations and real-world pilot projects will be crucial to demonstrating the viability and benefits of these systems. Furthermore, developing standardized protocols and fostering collaboration between industry stakeholders and policymakers will be essential to accelerate the adoption of waste energy recovery technologies. Ultimately, continued innovation and optimization in this field will contribute to achieving global sustainability goals, reducing greenhouse gas emissions, and promoting the transition to cleaner energy sources.

6. Conclusions

This study extensively reviewed the existing literature on four key bottoming cycles: SRC, CO2 SBC, IBC, and ABC. In addition, these four main bottoming systems were utilized for the waste energy recovery of natural gas-fired ICE with a capacity of 584 kW and an exhausted gas temperature of 493 °C. This was undertaken with the objective of determining the optimal configuration for waste recovery from high-temperature exhausted gas produced by an ICE. Four distinct cycles—namely, SRC, SBC with CO2 working fluid, IBC, and ABC—were selected as the primary bottoming systems. Furthermore, to ensure complete utilization of the waste heat generated by the bottoming systems, two TEGs were incorporated into the design. One TEG was utilized to harness residual heat from the exhausted gas, while the other was employed to recover unused energy from the main bottoming system. The power generated by the bottoming systems was subsequently directed toward a PEME for hydrogen production, thereby prompting the main inquiry: which main bottoming system yields the most cost-effective hydrogen production, as determined by a 4E optimization? The key findings of this study are presented below:
  • At the optimum mode, the SRC-based system resulted in the highest exergy efficiency (21.93%), while it brought about the second worst total cost rate (31.82 $/h) and unit cost of hydrogen (101.3 $/GJ). On the other hand, the IBC-based system conduced to the lowest exergy efficiency (13.72%), total cost rate (25.58 $/h), and unit cost of hydrogen (59.91 $/GJ).
  • Using TEGs for complete waste heat recovery of the topping and main bottoming system was a good choice, especially for TEG 1 in the IBC-based system (25.84 kW output power) and TEG 2 in the ABC-based system (16.16 kW output power). Although TEG 2 had a good performance in Configuration 4, it did not have a good performance overall in comparison with other configurations because of the low power generation of the main system and the high cost of AHX.
  • PEME in Configurations 1 and 2, TEG 1 in Configuration 3, and TEG 2 in Configuration 4 had the highest amount of exergy destruction among the different components of the systems. Those led to the highest exergy destruction cost rate of the above-mentioned components. Regarding investment cost rate, SRC turbine in Configuration 1, PEME in Configurations 2 and 3, and AHX in Configuration 4 were important components. The high cost of AHX in Configuration 4 is the chief reason for the worst economic performance of this configuration.

Author Contributions

Methodology, M.Z.; Software, M.Z.; Validation, M.Z.; Formal analysis, M.Z.; Investigation, M.Z.; Writing—original draft, M.Z.; Writing—review & editing, M.Z., N.H., S.G. and A.Z.; Supervision, N.H., S.G. and A.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

Optimization

Nomenclature
Aarea ( m 2 )
C ˙ cost rate ( $   h 1 )
ccost per exergy unit ( $   GJ 1 )
E a c t , j activation energy ( J   mol 1 )
E x ˙ exergy flow rate ( kW )
exspecific exergy ( kJ   kg 1 )
FFaraday constant
hspecific enthalpy ( kJ   kg 1 )
iinterest rate
Jcurrent density ( A   m 2 )
J 0 , j exchange current density ( A   m 2 )
J j r e f pre-exponential factor ( A   m 2 )
m ˙ mass flow rate ( kg   s 1 )
Nannual operating hours
n ˙ molar flow rate ( mol   s 1 )
Ppressure (kPa)
Q ˙ heat transfer rate ( kW )
sspecific entropy ( kJ   kg   1 K 1 )
Ttemperature (°C)
Vvoltage ( V )
V 0 reversible potential ( V )
W ˙ power ( kW )
Ymolar fraction of gas
Z T M figure of merit multiplied by a mean temperature
Z ˙ investment cost rate ( $   h 1 )
Zinvestment cost ( $ )
Subscripts
ananode
actactivation overvoltage
cacathode
comcompressor
concconcentration overvoltage
destdestruction
elelectricity
envenvironmental
eexit
exexergy
ffuel
iinlet
ohmohmic overvoltage
pproduct
pumpump
turturbine
Greek letters
δ fuel-air molar ratio
τlifetime of the proposed system
ηefficiency
ρdensity ( kg   m 3 )
φmaintenance factor
ωPEMElocal ionic conductivity of the electrolyzer
Abbreviation
ABC air bottoming cycle
ACHabsorption chiller
AHXair heat exchanger
ANNartificial natural network
CNGcompressed natural gas
CRFcapital recovery factor
EESengineering equation solver
GTCgas turbine cycle
HRHXheat recovery heat exchanger
HRSGheat recovery steam generator
HThigh temperature
HWUhot water unit
IBCinverse Brayton cycle
IBTBPinverse Brayton turbine back pressure
ICEinternal combustion engine
LTlow temperature
NGnatural gas
ORCorganic Rankine cycle
PERCpartial evaporation Rankine cycle
PEMEproton exchange membrane electrolyzer
PPpayback period
PRpressure ratio
RCRankine cycle
RHXrecovery heat exchanger
RODUreverse osmosis desalination unit
SBCsupercritical Brayton cycle
SRCsteam Rankine cycle
TEGthermoelectric generator

Appendix A

Table A1 offers the fuel-product definition and energy equation for each component.
Table A1. Fuel-product energy fluxes and energy equations in bottoming systems.
Table A1. Fuel-product energy fluxes and energy equations in bottoming systems.
DeviceEnergy RelationFuelProduct
Steam Rankine cycle
SRC HRSG Q ˙ H R S G = m ˙ g a s ( h 1 h 2 ) = m ˙ S R C ( h 6 h 9 ) E x ˙ 1 E x ˙ 2 E x ˙ 6 E x ˙ 9
SRC turbine W ˙ S R C , t u r = m ˙ S R C ( h 6 h 7 ) E x ˙ 6 E x ˙ 7 W ˙ S R C , t u r
SRC pump W ˙ S R C , p u m = m ˙ S R C ( h 9 h 8 ) W ˙ S R C , p u m E x ˙ 9 E x ˙ 8
TEG 1Table 1 E x ˙ 2 E x ˙ 3 E x ˙ 5 E x ˙ 4 + W ˙ T E G ,   1
TEG 2Table 1 E x ˙ 7 E x ˙ 8 E x ˙ 11 E x ˙ 10 + W ˙ T E G ,   2
PEMETable 1 W ˙ e l e c t r o l y z e r E x ˙ 15
Supercritical Brayton cycle
Heat recovery heat exchanger (HRHX) Q ˙ H R H X = m ˙ g a s ( h 1 h 2 ) = m ˙ S B C ( h 6 h 11 ) E x ˙ 1 E x ˙ 2 E x ˙ 6 E x ˙ 11
SBC turbine W ˙ S B C , t u r = m ˙ S B C ( h 6 h 7 ) W ˙ S B C , t u r E x ˙ 6 E x ˙ 7
Recuperator (Rec) Q ˙ R e c = ( h 7 h 8 ) = ( h 11 h 10 ) E x ˙ 7 E x ˙ 8 E x ˙ 11 E x ˙ 10
SBC compressor W ˙ S B C , c o m = m ˙ S B C ( h 10 h 9 ) W ˙ S B C , c o m E x ˙ 10 E x ˙ 9
TEG 1Table 1 E x ˙ 2 E x ˙ 3 E x ˙ 5 E x ˙ 4 + W ˙ T E G ,   1
TEG 2Table 1 E x ˙ 8 E x ˙ 9 E x ˙ 13 E x ˙ 12 + W ˙ T E G ,   2
PEMETable 1 W ˙ e l e c t r o l y z e r E x ˙ 17
Inverse Brayton cycle
IBC turbine W ˙ I B C ,   c o m = m ˙ g a s ( h 1 h 2 ) E x ˙ 1 E x ˙ 2 W ˙ I B C ,   t u r
IBC compressor W ˙ I B C , t u r = m ˙ g a s ( h 3 h 4 ) W ˙ I B C , c o m E x ˙ 3 E x ˙ 4
TEG 1Table 1 E x ˙ 2 E x ˙ 3 E x ˙ 7 E x ˙ 6 + W ˙ T E G ,   1
TEG 2Table 1 E x ˙ 4 E x ˙ 5 E x ˙ 9 E x ˙ 8 + W ˙ T E G ,   2
PEMETable 1 W ˙ e l e c t r o l y z e r E x ˙ 13
Air bottoming cycle
ABC compressor W ˙ A B C ,   c o m = m ˙ A B C ( h 7 h 6 ) W ˙ A B C ,   c o m E x ˙ 7 E x ˙ 6
Air heat exchanger (AHX) Q ˙ A H X = m ˙ g a s ( h 1 h 2 ) = m ˙ A B C ( h 8 h 7 ) E x ˙ 1 E x ˙ 2 E x ˙ 8 E x ˙ 7
ABC turbine W ˙ A B C , t u r = m ˙ A B C ( h 8 h 9 ) E x ˙ 8 E x ˙ 9 W ˙ A B C , t u r
TEG 1Table 1 E x ˙ 2 E x ˙ 3 E x ˙ 5 E x ˙ 4 + W ˙ T E G ,   1
TEG 2Table 1 E x ˙ 9 E x ˙ 10 E x ˙ 12 E x ˙ 11 + W ˙ T E G ,   2
PEMETable 1 W ˙ e l e c t r o l y z e r E x ˙ 16

Appendix B

The cost balance formulas can be found in Table A2.
Table A2. Cost class of bottoming systems formulas.
Table A2. Cost class of bottoming systems formulas.
DeviceMain EquationSupplementary Equation
Steam Rankine cycle
SRC HRSG C ˙ 1 + C ˙ 9 + Z ˙ H R S G = C ˙ 2 + C ˙ 6 c 1 = 23.59 $ GJ   c 1 = c 2 [29]
SRC turbine C ˙ 6 + Z ˙ S R C , t u r = c e l , S R C W ˙ S R C , t u r + C ˙ 7 c 6 = c 7
SRC pump C ˙ 8 + c e l , S R C W ˙ S R C , p u m + Z ˙ S R C , p u m = C ˙ 9
TEG 1 C ˙ 2 + C ˙ 4 + Z ˙ T E G , 1 = C ˙ 3 + C ˙ 5 + c e l , S R C W ˙ T E G , 1 c 2 = c 3 , c 4 = 0
TEG 2 C ˙ 7 + C ˙ 10 + Z ˙ T E G , 2 = C ˙ 8 + C ˙ 11 + c e l , S R C W ˙ T E G , 2 c 7 = c 8 , c 10 = 0
PEME c e l , S R C W ˙ P E M E + Z ˙ P E M E = C ˙ 15
Supercritical Brayton cycle
HRHX C ˙ 1 + C ˙ 11 + Z ˙ H R H X = C ˙ 2 + C ˙ 6 c 1 = 23.59   $ / GJ ,
c 1 = c 2 [29]
SBC turbine C ˙ 6 + Z ˙ SBC , tur = C ˙ 7 + c el , SBC   W ˙ SBC , tur c 6 = c 7
Recuperator C ˙ 7 + C ˙ 10 + Z ˙ Rec = C ˙ 8 + C ˙ 11 c 7 = c 8
SBC compressor C ˙ 9 + Z ˙ SBC , com + c el , SBC   W ˙ SBC , com = C ˙ 10
TEG 1 C ˙ 2 + C ˙ 4 + Z ˙ T E G , 1 = C ˙ 3 + C ˙ 5 + c e l , S B C W ˙ T E G , 1 c 2 = c 3 , c 4 = 0
TEG 2 C ˙ 8 + C ˙ 12 + Z ˙ T E G , 2 = C ˙ 9 + C ˙ 13 + c e l , S B C W ˙ T E G , 2 c 8 = c 9 , c 12 = 0
PEME c e l , S B C W ˙ P E M E + Z ˙ P E M E = C ˙ 17
Inverse Brayton cycle
IBC turbine C ˙ 1 + Z ˙ IBC , tur = C ˙ 2 + c el , IBC   W ˙ IBC , tur c 1 = 23.59   $ / GJ ,
c 1 = c 2 [29]
IBC compressor C ˙ 3 + Z ˙ IBC , com + c el , IBC   W ˙ IBC , com = C ˙ 4
TEG 1 C ˙ 2 + C ˙ 6 + Z ˙ T E G , 1 = C ˙ 3 + C ˙ 7 + c e l , I B C W ˙ T E G , 1 c 2 = c 3 , c 6 = 0
TEG 2 C ˙ 4 + C ˙ 8 + Z ˙ T E G , 2 = C ˙ 5 + C ˙ 9 + c e l , I B C W ˙ T E G , 2 c 4 = c 5 , c 8 = 0
PEME c e l , I B C W ˙ P E M E + Z ˙ P E M E = C ˙ 13
Air bottoming cycle
ABC compressor C ˙ 6 + Z ˙ ABC , com + c el , ABC   W ˙ ABC , com = C ˙ 7
Air heat exchanger (AHX) C ˙ 1 + C ˙ 7 + Z ˙ A H X = C ˙ 2 + C ˙ 8 c 1 = 23.59   $ / GJ ,
c 1 = c 2 [29]
ABC turbine C ˙ 8 + Z ˙ ABC , tur = C ˙ 9 + c el , ABC   W ˙ ABC , tur c 8 = c 9
TEG 1 C ˙ 2 + C ˙ 4 + Z ˙ T E G , 1 = C ˙ 3 + C ˙ 5 + c e l , A B C W ˙ T E G , 1 c 2 = c 3 , c 4 = 0
TEG 2 C ˙ 9 + C ˙ 11 + Z ˙ T E G , 2 = C ˙ 10 + C ˙ 12 + c e l , A B C W ˙ T E G , 2 c 9 = c 10 , c 11 = 0
PEME c e l , A B C W ˙ P E M E + Z ˙ P E M E = C ˙ 16

Appendix C

Table A3 demonstrates the relationship between investment costs.
Table A3. Equations for bottoming system purchases.
Table A3. Equations for bottoming system purchases.
DeviceEquationReference
Steam Rankine cycle
SRC HRSG Z H R S G = 6570   ( Q ˙ j Δ T l m t d , j ) 0.8 + 21276   m ˙ S R C + 1184.4 ( m ˙ g a s ) 1.2 [122]
SRC turbine Z S R C , t u r = 3880.5     W ˙ S R C , t u r 0.7 [ 1 + ( 0.05 1 η S R C , t u r ) 3 ]   [ 1 + 5   e x p ( T 11 866 10.42 ) ] [122]
SRC pump Z S R C , p u m = 705.48   W ˙ S R C , p u m 0.71   . [ 1 + ( 0.2 1 η S R C , p u m ) ] [122]
TEG 1 Z T E G , 1 = 1500   W ˙ T E G , 1 [128]
TEG 2 Z T E G .2 = 1500   W ˙ T E G , 2 [128]
PEME Z P E M E = 1000   W ˙ P E M E [128]
Supercritical Brayton cycle
HRHX Z H R H X = 2681   A H R H X 0.59 [106]
SBC turbine Z S B C , t u r = 479.34   m ˙ S B C ( 1 0.92 η S B C , t u r ) ln ( P 6 P 7 ) ( 1 + exp ( 0.036 T 6 54.4 ) ) [129]
Recuperator Z R e c = 2681   A R e c 0.59 [106]
SBC compressor Z S B C , c o m = 71.1   m ˙ S B C ( 1 0.9 η S B C , c o m ) ( P 10 P 9 ) ln ( P 10 P 9 ) [129]
TEG 1 Z T E G , 1 = 1500   W ˙ T E G , 1 [128]
TEG 2 Z T E G .2 = 1500   W ˙ T E G , 2 [128]
PEME Z P E M E = 1000   W ˙ P E M E [128]
Inverse Brayton cycle
IBC turbine Z I B C , t u r = 479.34   m ˙ I B C ( 1 0.92 η I B C , t u r ) ln ( P 1 P 2 ) ( 1 + exp ( 0.036 T 1 54.4 ) ) [129]
IBC compressor Z I B C , c o m = 71.1   m ˙ I B C ( 1 0.9 η I B C , c o m ) ( P 4 P 3 ) ln ( P 4 P 3 ) [129]
TEG 1 Z T E G , 1 = 1500   W ˙ T E G , 1 [128]
TEG 2 Z T E G .2 = 1500   W ˙ T E G , 2 [128]
PEME Z P E M E = 1000   W ˙ P E M E [128]
Air bottoming cycle
ABC compressor Z A B C , c o m = 71.1   m ˙ A B C ( 1 0.9 η A B C , c o m ) ( P 7 P 6 ) ln ( P 7 P 6 ) [129]
Air heat exchanger (AHX) Z A H X = 4122 A A H X 0.6 [129]
ABC turbine Z A B C , t u r = 479.34   m ˙ A B C ( 1 0.92 η A B C , t u r ) ln ( P 8 P 9 ) ( 1 + exp ( 0.036 T 8 54.4 ) ) [129]
TEG 1 Z T E G , 1 = 1500   W ˙ T E G , 1 [128]
TEG 2 Z T E G .2 = 1500   W ˙ T E G , 2 [128]
PEME Z P E M E = 1000   W ˙ P E M E [128]

Appendix D

The investment cost rate, denoted by Z ˙ k , is finally obtained as follows [130]:
Z ˙ k = Z k   C R F   φ N
Table A4 presents the other economic parameters associated with the abovementioned equation.
Table A4. The parameters of Equation (A1) [131].
Table A4. The parameters of Equation (A1) [131].
ParameterValue/Equation
N (System hours of operation in a year)7446
φ (Maintenance factor) 1.06
CRF (Capital recovery factor) i ( 1 + i ) τ ( 1 + i ) τ 1
i (Interest rate)10%
τ (Lifetime of the proposed system)20

References

  1. Gupta, A.; Kumar, R.; Maurya, A.; Ahmadi, M.H.; Ongar, B.; Yegzekova, A.; Sharma, J.P.; Kanchan, S.; Shelare, S. A comparative study of the impact on combustion and emission characteristics of nanoparticle-based fuel additives in the internal combustion engine. Energy Sci. Eng. 2024, 12, 284–303. [Google Scholar] [CrossRef]
  2. Zhao, X.; Shi, J.; Liu, J.; Han, Z.; Tan, Z.; Xiang, L. A response coordination design technology for the main loaded structures of the internal combustion engines based on NSEHO. Energy Sci. Eng. 2023, 12, 637–656. [Google Scholar] [CrossRef]
  3. Salek, F.; Zamen, M.; Hosseini, S.V.; Babaie, M. Novel hybrid system of pulsed HHO generator/TEG waste heat recovery for CO reduction of a gasoline engine. Int. J. Hydrogen Energy 2020, 45, 23576–23586. [Google Scholar] [CrossRef]
  4. Parsa, S.; Jafarmadar, S.; Neshat, E. Application of waste heat in a novel trigeneration system integrated with an HCCI engine for power, heat and hydrogen production. Int. J. Hydrogen Energy 2022, 47, 26303–26315. [Google Scholar] [CrossRef]
  5. Thaddaeus, J.; Unachukwu, G.O.; Mgbemene, C.A.; Pesyridis, A.; Usman, M.; Alshammari, F.A. Design, size estimation, and thermodynamic analysis of a realizable organic Rankine cycle system for waste heat recovery in commercial truck engines. Therm. Sci. Eng. Prog. 2021, 22, 100849. [Google Scholar] [CrossRef]
  6. Naeimi, A.; Bidi, M.; Ahmadi, M.H.; Kumar, R.; Sadeghzadeh, M.; Nazari, M.A. Design and exergy analysis of waste heat recovery system and gas engine for power generation in Tehran cement factory. Therm. Sci. Eng. Prog. 2019, 9, 299–307. [Google Scholar] [CrossRef]
  7. Chiriac, R.; Racovitza, A.; Podevin, P.; Descombes, G. On the possibility to reduce CO2 emissions of heat engines fuelled partially with hydrogen produced by waste heat recovery. Int. J. Hydrogen Energy 2015, 40, 15856–15863. [Google Scholar] [CrossRef]
  8. Ghazikhani, M.; Hatami, M.; Safari, B. The effect of alcoholic fuel additives on exergy parameters and emissions in a two stroke gasoline engine. Arab. J. Sci. Eng. 2014, 39, 2117–2125. [Google Scholar] [CrossRef]
  9. Yağlı, H.; Koç, Y.; Köse, Ö.; Koç, A.; Yumrutaş, R. Optimisation of simple and regenerative organic Rankine cycles using jacket water of an internal combustion engine fuelled with biogas produced from agricultural waste. Process Saf. Environ. Prot. 2021, 155, 17–31. [Google Scholar] [CrossRef]
  10. Demir, M.E.; Çıtakoğlu, F. Design and modeling of a multigeneration system driven by waste heat of a marine diesel engine. Int. J. Hydrogen Energy 2022, 47, 40513–40530. [Google Scholar] [CrossRef]
  11. Chen, F.; Zhang, W.; Liu, Y.; Cai, J.; Zhang, J.; Wang, X.; Su, Q. Simulation and 4E analysis of a novel trigeneration process using a gas turbine cycle combined with a geothermal-driven multi-waste heat recovery method. Process Saf. Environ. Prot. 2023, 176, 1026–1047. [Google Scholar] [CrossRef]
  12. Prato-Garcia, D.; Robayo-Avendaño, A.; Vasquez-Medrano, R. Hydrogen from natural gas and biogas: Building bridges for a sustainable transition to a green economy. Gas Sci. Eng. 2023, 111, 204918. [Google Scholar] [CrossRef]
  13. Incer-Valverde, J.; Lugo-Mayor, C.; Tsatsaronis, G.; Morosuk, T. Evaluation of the large-scale hydrogen supply chain and perspectives on LH2 regasification cogeneration systems. Gas Sci. Eng. 2023, 115, 205005. [Google Scholar] [CrossRef]
  14. Shamsi, M.; Moghaddas, S.; Naeiji, E.; Farokhi, S. Techno-economic, energy, exergy, and environmental comparison of hydrogen production from natural gas, biogas, and their combination as feedstock. Arab. J. Sci. Eng. 2023, 48, 8971–8987. [Google Scholar] [CrossRef]
  15. Yin, K.; Zhang, R.; Yan, M.; Sun, L.; Ma, Y.; Cui, P.; Zhu, Z.; Wang, Y. Thermodynamic and Economic Analysis of A Hydrogen Production Process From Medical Waste by Plasma Gasification. Process Saf. Environ. Prot. 2023, 178, 8–17. [Google Scholar] [CrossRef]
  16. Shehata, M.F.; El-Shamy, A.M. Hydrogen-based failure in oil and gas pipelines a review. Gas Sci. Eng. 2023, 115, 204994. [Google Scholar] [CrossRef]
  17. Tian, H.; Shu, G.; Wei, H.; Liang, X.; Liu, L. Fluids and parameters optimization for the organic Rankine cycles (ORCs) used in exhaust heat recovery of Internal Combustion Engine (ICE). Energy 2012, 47, 125–136. [Google Scholar] [CrossRef]
  18. Yu, G.; Shu, G.; Tian, H.; Wei, H.; Liu, L. Simulation and thermodynamic analysis of a bottoming Organic Rankine Cycle (ORC) of diesel engine (DE). Energy 2013, 51, 281–290. [Google Scholar] [CrossRef]
  19. Shu, G.; Liu, L.; Tian, H.; Wei, H.; Xu, X. Performance comparison and working fluid analysis of subcritical and transcritical dual-loop organic Rankine cycle (DORC) used in engine waste heat recovery. Energy Convers. Manag. 2013, 74, 35–43. [Google Scholar] [CrossRef]
  20. Zhang, H.G.; Wang, E.H.; Fan, B.Y. A performance analysis of a novel system of a dual loop bottoming organic Rankine cycle (ORC) with a light-duty diesel engine. Appl. Energy 2013, 102, 1504–1513. [Google Scholar] [CrossRef]
  21. Yang, F.; Dong, X.; Zhang, H.; Wang, Z.; Yang, K.; Zhang, J.; Wang, E.; Liu, H.; Zhao, G. Performance analysis of waste heat recovery with a dual loop organic Rankine cycle (ORC) system for diesel engine under various operating conditions. Energy Convers. Manag. 2014, 80, 243–255. [Google Scholar] [CrossRef]
  22. Shu, G.; Liu, L.; Tian, H.; Wei, H.; Yu, G. Parametric and working fluid analysis of a dual-loop organic Rankine cycle (DORC) used in engine waste heat recovery. Appl. Energy 2014, 113, 1188–1198. [Google Scholar] [CrossRef]
  23. Song, J.; Gu, C.-W. Analysis of ORC (Organic Rankine Cycle) systems with pure hydrocarbons and mixtures of hydrocarbon and retardant for engine waste heat recovery. Appl. Therm. Eng. 2015, 89, 693–702. [Google Scholar] [CrossRef]
  24. Yue, C.; Han, D.; Pu, W.; He, W. Comparative analysis of a bottoming transcritical ORC and a Kalina cycle for engine exhaust heat recovery. Energy Convers. Manag. 2015, 89, 764–774. [Google Scholar] [CrossRef]
  25. Song, J.; Song, Y.; Gu, C.-W. Thermodynamic analysis and performance optimization of an Organic Rankine Cycle (ORC) waste heat recovery system for marine diesel engines. Energy 2015, 82, 976–985. [Google Scholar] [CrossRef]
  26. Song, J.; Gu, C.-W. Performance analysis of a dual-loop organic Rankine cycle (ORC) system with wet steam expansion for engine waste heat recovery. Appl. Energy 2015, 156, 280–289. [Google Scholar] [CrossRef]
  27. Daghigh, R.; Shafieian, A. An investigation of heat recovery of submarine diesel engines for combined cooling, heating and power systems. Energy Convers. Manag. 2016, 108, 50–59. [Google Scholar] [CrossRef]
  28. Salek, F.; Moghaddam, A.N.; Naserian, M.M. Thermodynamic analysis of diesel engine coupled with ORC and absorption refrigeration cycle. Energy Convers. Manag. 2017, 140, 240–246. [Google Scholar] [CrossRef]
  29. Habibi, H.; Zoghi, M.; Chitsaz, A.; Javaherdeh, K.; Ayazpour, M. Thermo-economic analysis and optimization of combined PERC-ORC-LNG power system for diesel engine waste heat recovery. Energy Convers. Manag. 2018, 173, 613–625. [Google Scholar] [CrossRef]
  30. Mohammadkhani, F.; Yari, M.; Ranjbar, F. A zero-dimensional model for simulation of a Diesel engine and exergoeconomic analysis of waste heat recovery from its exhaust and coolant employing a high-temperature Kalina cycle. Energy Convers. Manag. 2019, 198, 111782. [Google Scholar] [CrossRef]
  31. Jafarzad, A.; Asgari, N.; Ranjbar, F.; Mohammadkhani, F. Thermodynamic assessment and optimization of the influences of the steam-assisted turbocharging and organic Rankine cycle on the overall performance of a diesel engine-based cogeneration integrated with a reverse osmosis desalination unit. Sustain. Energy Technol. Assess. 2021, 46, 101175. [Google Scholar] [CrossRef]
  32. Shu, G.; Liu, L.; Tian, H.; Wei, H.; Liang, Y. Analysis of regenerative dual-loop organic Rankine cycles (DORCs) used in engine waste heat recovery. Energy Convers. Manag. 2013, 76, 234–243. [Google Scholar] [CrossRef]
  33. Galindo, J.; Ruiz, S.; Dolz, V.; Royo-Pascual, L.; Haller, R.; Nicolas, B.; Glavatskaya, Y. Experimental and thermodynamic analysis of a bottoming Organic Rankine Cycle (ORC) of gasoline engine using swash-plate expander. Energy Convers. Manag. 2015, 103, 519–532. [Google Scholar] [CrossRef]
  34. Yang, F.; Zhang, H.; Bei, C.; Song, S.; Wang, E. Parametric optimization and performance analysis of ORC (organic Rankine cycle) for diesel engine waste heat recovery with a fin-and-tube evaporator. Energy 2015, 91, 128–141. [Google Scholar] [CrossRef]
  35. Mastrullo, R.; Mauro, A.W.; Revellin, R.; Viscito, L. Modeling and optimization of a shell and louvered fin mini-tubes heat exchanger in an ORC powered by an internal combustion engine. Energy Convers. Manag. 2015, 101, 697–712. [Google Scholar] [CrossRef]
  36. Yu, G.; Shu, G.; Tian, H.; Huo, Y.; Zhu, W. Experimental investigations on a cascaded steam-/organic-Rankine-cycle (RC/ORC) system for waste heat recovery (WHR) from diesel engine. Energy Convers. Manag. 2016, 129, 43–51. [Google Scholar] [CrossRef]
  37. Shu, G.; Yu, G.; Tian, H.; Wei, H.; Liang, X.; Huang, Z. Multi-approach evaluations of a cascade-Organic Rankine Cycle (C-ORC) system driven by diesel engine waste heat: Part A–Thermodynamic evaluations. Energy Convers. Manag. 2016, 108, 579–595. [Google Scholar] [CrossRef]
  38. Zhao, M.; Wei, M.; Song, P.; Liu, Z.; Tian, G. Performance evaluation of a diesel engine integrated with ORC system. Appl. Therm. Eng. 2017, 115, 221–228. [Google Scholar] [CrossRef]
  39. Chen, T.; Zhuge, W.; Zhang, Y.; Zhang, L. A novel cascade organic Rankine cycle (ORC) system for waste heat recovery of truck diesel engines. Energy Convers. Manag. 2017, 138, 210–223. [Google Scholar] [CrossRef]
  40. Yang, F.; Zhang, H.; Yu, Z.; Wang, E.; Meng, F.; Liu, H.; Wang, J. Parametric optimization and heat transfer analysis of a dual loop ORC (organic Rankine cycle) system for CNG engine waste heat recovery. Energy 2017, 118, 753–775. [Google Scholar] [CrossRef]
  41. Seyedkavoosi, S.; Javan, S.; Kota, K. Exergy-based optimization of an organic Rankine cycle (ORC) for waste heat recovery from an internal combustion engine (ICE). Appl. Therm. Eng. 2017, 126, 447–457. [Google Scholar] [CrossRef]
  42. Uusitalo, A.; Honkatukia, J.; Turunen-Saaresti, T. Evaluation of a small-scale waste heat recovery organic Rankine cycle. Appl. Energy 2017, 192, 146–158. [Google Scholar] [CrossRef]
  43. Wang, F.; Wang, L.; Zhang, H.; Xia, L.; Miao, H.; Yuan, J. Design and optimization of hydrogen production by solid oxide electrolyzer with marine engine waste heat recovery and ORC cycle. Energy Convers. Manag. 2021, 229, 113775. [Google Scholar] [CrossRef]
  44. Zhao, R.; Zhang, H.; Song, S.; Yang, F.; Hou, X.; Yang, Y. Global optimization of the diesel engine–organic Rankine cycle (ORC) combined system based on particle swarm optimizer (PSO). Energy Convers. Manag. 2018, 174, 248–259. [Google Scholar] [CrossRef]
  45. Scaccabarozzi, R.; Tavano, M.; Invernizzi, C.M.; Martelli, E. Comparison of working fluids and cycle optimization for heat recovery ORCs from large internal combustion engines. Energy 2018, 158, 396–416. [Google Scholar] [CrossRef]
  46. da Silva, J.A.M.; Seifert, V.; de Morais, V.O.B.; Tsolakis, A.; Herreros, J.; Torres, E. Exergy evaluation and ORC use as an alternative for efficiency improvement in a CI-engine power plant. Sustain. Energy Technol. Assess. 2018, 30, 216–223. [Google Scholar] [CrossRef]
  47. Zhu, Y.; Li, W.; Sun, G.; Li, H. Thermo-economic analysis based on objective functions of an organic Rankine cycle for waste heat recovery from marine diesel engine. Energy 2018, 158, 343–356. [Google Scholar] [CrossRef]
  48. Liu, P.; Shu, G.; Tian, H. How to approach optimal practical Organic Rankine cycle (OP-ORC) by configuration modification for diesel engine waste heat recovery. Energy 2019, 174, 543–552. [Google Scholar] [CrossRef]
  49. Lin, S.; Zhao, L.; Deng, S.; Ni, J.; Zhang, Y.; Ma, M. Dynamic performance investigation for two types of ORC system driven by waste heat of automotive internal combustion engine. Energy 2019, 169, 958–971. [Google Scholar] [CrossRef]
  50. Mohammadkhani, F.; Yari, M. A 0D model for diesel engine simulation and employing a transcritical dual loop Organic Rankine Cycle (ORC) for waste heat recovery from its exhaust and coolant: Thermodynamic and economic analysis. Appl. Therm. Eng. 2019, 150, 329–347. [Google Scholar] [CrossRef]
  51. Fang, Y.; Yang, F.; Zhang, H. Comparative analysis and multi-objective optimization of organic Rankine cycle (ORC) using pure working fluids and their zeotropic mixtures for diesel engine waste heat recovery. Appl. Therm. Eng. 2019, 157, 113704. [Google Scholar] [CrossRef]
  52. Wang, Z.; Hu, Y.; Xia, X.; Zuo, Q.; Zhao, B.; Li, Z. Thermo-economic selection criteria of working fluid used in dual-loop ORC for engine waste heat recovery by multi-objective optimization. Energy 2020, 197, 117053. [Google Scholar] [CrossRef]
  53. Khodadadi, F.; Deymi-Dashtebayaz, M.; Lakzian, E. Parametric analysis of combined power and freshwater producing system for natural gas engine heat recovery. Energy Convers. Manag. 2020, 225, 113464. [Google Scholar] [CrossRef]
  54. Boodaghi, H.; Etghani, M.M.; Sedighi, K. Performance analysis of a dual-loop bottoming organic Rankine cycle (ORC) for waste heat recovery of a heavy-duty diesel engine, Part I: Thermodynamic analysis. Energy Convers. Manag. 2021, 241, 113830. [Google Scholar] [CrossRef]
  55. Li, T.; Gao, R.; Gao, X. Energy, exergy, economic, and environment (4E) assessment of trans-critical organic Rankine cycle for combined heating and power in wastewater treatment plant. Energy Convers. Manag. 2022, 267, 115932. [Google Scholar] [CrossRef]
  56. Xing, C.; Ping, X.; Guo, R.; Zhang, H.; Yang, F.; Yu, M.; Yang, A.; Wang, Y. Machine learning-based multi-objective optimization and thermodynamic evaluation of organic Rankine cycle (ORC) system for vehicle engine under road condition. Appl. Therm. Eng. 2023, 231, 120904. [Google Scholar] [CrossRef]
  57. Hadelu, L.M.; Noorpoor, A.; Boyaghchi, F.A.; Mirjalili, S. Exergoeconomic, carbon, and water footprint analyses and optimization of a new solar-driven multigeneration system based on supercritical CO2 cycle and solid oxide steam electrolyzer using various phase change materials. Process Saf. Environ. Prot. 2022, 159, 393–421. [Google Scholar] [CrossRef]
  58. Guo, J.-Q.; Li, M.-J.; He, Y.-L.; Jiang, T.; Ma, T.; Xu, J.-L.; Cao, F. A systematic review of supercritical carbon dioxide (S-CO2) power cycle for energy industries: Technologies, key issues, and potential prospects. Energy Convers. Manag. 2022, 258, 115437. [Google Scholar] [CrossRef]
  59. Sun, L.; Wang, D.; Xie, Y. Energy, exergy and exergoeconomic analysis of two supercritical CO2 cycles for waste heat recovery of gas turbine. Appl. Therm. Eng. 2021, 196, 117337. [Google Scholar] [CrossRef]
  60. Liang, Y.; Bian, X.; Qian, W.; Pan, M.; Ban, Z.; Yu, Z. Theoretical analysis of a regenerative supercritical carbon dioxide Brayton cycle/organic Rankine cycle dual loop for waste heat recovery of a diesel/natural gas dual-fuel engine. Energy Convers. Manag. 2019, 197, 111845. [Google Scholar] [CrossRef]
  61. Wu, C.; Xu, X.; Li, Q.; Li, J.; Wang, S.; Liu, C. Proposal and assessment of a combined cooling and power system based on the regenerative supercritical carbon dioxide Brayton cycle integrated with an absorption refrigeration cycle for engine waste heat recovery. Energy Convers. Manag. 2020, 207, 112527. [Google Scholar] [CrossRef]
  62. Feng, Y.; Du, Z.; Shreka, M.; Zhu, Y.; Zhou, S.; Zhang, W. Thermodynamic analysis and performance optimization of the supercritical carbon dioxide Brayton cycle combined with the Kalina cycle for waste heat recovery from a marine low-speed diesel engine. Energy Convers. Manag. 2020, 206, 112483. [Google Scholar] [CrossRef]
  63. Pan, M.; Bian, X.; Zhu, Y.; Liang, Y.; Lu, F.; Xiao, G. Thermodynamic analysis of a combined supercritical CO2 and ejector expansion refrigeration cycle for engine waste heat recovery. Energy Convers. Manag. 2020, 224, 113373. [Google Scholar] [CrossRef]
  64. Almohana, A.I.; Almojil, S.F.; Alali, A.F.; Anqi, A.E.; Rajhi, A.A.; Alamri, S.; Dhahad, H.A.; Ahmed, A.N.; Mohamed, A. Mathematical simulation and optimization of a waste energy recovery for an internal combustion engine integrated with ScCO2 cycle and modified Kalina cycle. Sustain. Energy Technol. Assess. 2022, 53, 102650. [Google Scholar]
  65. Wang, Z.; Jiang, Y.; Ma, Y.; Han, F.; Ji, Y.; Cai, W. A partial heating supercritical CO2 nested transcritical CO2 cascade power cycle for marine engine waste heat recovery: Thermodynamic, economic, and footprint analysis. Energy 2022, 261, 125269. [Google Scholar] [CrossRef]
  66. Wang, Z.; Jiang, Y.; Han, F.; Yu, S.; Li, W.; Ji, Y.; Cai, W. A thermodynamic configuration method of combined supercritical CO2 power system for marine engine waste heat recovery based on recuperative effects. Appl. Therm. Eng. 2022, 200, 117645. [Google Scholar] [CrossRef]
  67. Dadpour, D.; Gholizadeh, M.; Estiri, M.; Poncet, S. Multi objective optimization and 3E analyses of a novel supercritical/transcritical CO2 waste heat recovery from a ship exhaust. Energy 2023, 278, 127843. [Google Scholar] [CrossRef]
  68. Song, J.; Li, X.-s.; Ren, X.-d.; Gu, C.-w. Performance improvement of a preheating supercritical CO2 (S-CO2) cycle based system for engine waste heat recovery. Energy Convers. Manag. 2018, 161, 225–233. [Google Scholar] [CrossRef]
  69. Liang, Y.; Sun, Z.; Dong, M.; Lu, J.; Yu, Z. Investigation of a refrigeration system based on combined supercritical CO2 power and transcritical CO2 refrigeration cycles by waste heat recovery of engine. Int. J. Refrig. 2020, 118, 470–482. [Google Scholar] [CrossRef]
  70. Pan, M.; Zhu, Y.; Bian, X.; Liang, Y.; Lu, F.; Ban, Z. Theoretical analysis and comparison on supercritical CO2 based combined cycles for waste heat recovery of engine. Energy Convers. Manag. 2020, 219, 113049. [Google Scholar] [CrossRef]
  71. Song, J.; Li, X.; Wang, K.; Markides, C.N. Parametric optimisation of a combined supercritical CO2 (S-CO2) cycle and organic Rankine cycle (ORC) system for internal combustion engine (ICE) waste-heat recovery. Energy Convers. Manag. 2020, 218, 112999. [Google Scholar] [CrossRef]
  72. Zhang, R.; Su, W.; Lin, X.; Zhou, N.; Zhao, L. Thermodynamic analysis and parametric optimization of a novel S–CO2 power cycle for the waste heat recovery of internal combustion engines. Energy 2020, 209, 118484. [Google Scholar] [CrossRef]
  73. Ouyang, T.; Zhang, M.; Mo, X.; Qin, P. Transient characteristic evaluation and optimization of supercritical CO2 Brayton cycle driven by waste heat of automotive gasoline engine. J. Clean. Prod. 2021, 329, 129796. [Google Scholar] [CrossRef]
  74. Maestre-Cambronel, D.; Barros, J.G.; Gonzalez-Quiroga, A.; Bula, A.; Duarte-Forero, J. Thermoeconomic analysis of improved exhaust waste heat recovery system for natural gas engine based on Vortex Tube heat booster and supercritical CO2 Brayton cycle. Sustain. Energy Technol. Assess. 2021, 47, 101355. [Google Scholar] [CrossRef]
  75. Sakalis, G. Investigation of supercritical CO2 cycles potential for marine Diesel engine waste heat recovery applications. Appl. Therm. Eng. 2021, 195, 117201. [Google Scholar] [CrossRef]
  76. Liu, H.; Wang, M.; Li, S. Investigation of the polygeneration system integrated with gas engine-driven heat pump system and CO2 Brayton cycle for waste heat recovery. Appl. Therm. Eng. 2023, 221, 119872. [Google Scholar] [CrossRef]
  77. Fang, Z.; Dong, X.; Tang, X.; Lv, Z.; Qiao, X.; Wang, L.; Sun, C.; Yu, X. Study on supercritical CO2 power cycles for natural gas engine energy cascade utilization. Appl. Therm. Eng. 2023, 225, 120255. [Google Scholar] [CrossRef]
  78. Goodarzi, M. Usefulness analysis on regenerator and heat exchanger in Brayton & inverse Brayton cycles at moderate pressure ratio operation. Energy Convers. Manag. 2016, 126, 982–990. [Google Scholar]
  79. Besarati, S.M.; Atashkari, K.; Jamali, A.; Hajiloo, A.; Nariman-Zadeh, N. Multi-objective thermodynamic optimization of combined Brayton and inverse Brayton cycles using genetic algorithms. Energy Convers. Manag. 2010, 51, 212–217. [Google Scholar] [CrossRef]
  80. Zhu, D.; Zheng, X. Potential for energy and emissions of asymmetric twin-scroll turbocharged diesel engines combining inverse Brayton cycle system. Energy 2019, 179, 581–592. [Google Scholar] [CrossRef]
  81. Di Battista, D.; Fatigati, F.; Carapellucci, R.; Cipollone, R. Inverted Brayton Cycle for waste heat recovery in reciprocating internal combustion engines. Appl. Energy 2019, 253, 113565. [Google Scholar] [CrossRef]
  82. Chagnon-Lessard, N.; Copeland, C.; Mathieu-Potvin, F.; Gosselin, L. Maximizing specific work output extracted from engine exhaust with novel inverted Brayton cycles over a large range of operating conditions. Energy 2020, 191, 116350. [Google Scholar] [CrossRef]
  83. Abrosimov, K.A.; Baccioli, A.; Bischi, A. Techno-economic analysis of combined inverted Brayton–Organic Rankine cycle for high-temperature waste heat recovery. Energy Convers. Manag. 2020, 207, 112336. [Google Scholar] [CrossRef]
  84. Salek, F.; Babaie, M.; Naserian, M.M.; Ahmadi, M.H. Power enhancement of a turbo-charged industrial diesel engine by using of a waste heat recovery system based on inverted Brayton and organic Rankine cycles. Fuel 2022, 322, 124036. [Google Scholar] [CrossRef]
  85. Di Battista, D.; Carapellucci, R.; Cipollone, R. Integrated evaluation of Inverted Brayton cycle recovery unit bottomed to a turbocharged diesel engine. Appl. Therm. Eng. 2020, 175, 115353. [Google Scholar] [CrossRef]
  86. Abrosimov, K.; Baccioli, A.; Bischi, A. Extensive techno-economic assessment of combined inverted Brayton–Organic Rankine cycle for high-temperature waste heat recovery. Energy 2020, 211, 118406. [Google Scholar] [CrossRef]
  87. Zoghi, M.; Habibi, H.; Chitsaz, A.; Shamsaiee, M. Exergoeconomic and environmental analyses of a novel trigeneration system based on combined gas turbine-air bottoming cycle with hybridization of solar power tower and natural gas combustion. Appl. Therm. Eng. 2021, 188, 116610. [Google Scholar] [CrossRef]
  88. Ahmadi, A.; Ehyaei, M.A.; Jamali, D.H.; Despotovic, M.; Esmaeilion, F.; Abdalisousan, A.; Hani, E.B. Energy, exergy, and economic analyses of integration of heliostat solar receiver to gas and air bottom cycles. J. Clean. Prod. 2021, 280, 124322. [Google Scholar] [CrossRef]
  89. Mohammadi, K.; Saghafifar, M.; McGowan, J.G. Thermo-economic evaluation of modifications to a gas power plant with an air bottoming combined cycle. Energy Convers. Manag. 2018, 172, 619–644. [Google Scholar] [CrossRef]
  90. Inac, S.; Unverdi, S.O.; Midilli, A. A study on sustainability analysis of solid oxide fuel cell and proton exchange membrane electrolyzer integrated hybrid multi-generation system. Int. J. Hydrogen Energy 2022, 47, 26600–26631. [Google Scholar] [CrossRef]
  91. Abdel-Motagali, A.; Al Bacha, S.; El Rouby, W.M.A.; Bigarré, J.; Millet, P. Evaluating the performance of hybrid proton exchange membrane for PEM water electrolysis. Int. J. Hydrogen Energy 2023, 49, 87–102. [Google Scholar] [CrossRef]
  92. Haghghi, M.A.; Hasanzadeh, A.; Nadimi, E.; Rosato, A.; Athari, H. An intelligent thermodynamic/economic approach based on artificial neural network combined with MOGWO algorithm to study a novel polygeneration scheme using a modified dual-flash geothermal cycle. Process Saf. Environ. Prot. 2023, 173, 859–880. [Google Scholar] [CrossRef]
  93. Krishnan, S.; Koning, V.; de Groot, M.T.; de Groot, A.; Mendoza, P.G.; Junginger, M.; Kramer, G.J. Present and future cost of alkaline and PEM electrolyser stacks. Int. J. Hydrogen Energy 2023, 48, 32313–32330. [Google Scholar] [CrossRef]
  94. Emadi, M.A.; Pourrahmani, H.; Moghimi, M. Performance evaluation of an integrated hydrogen production system with LNG cold energy utilization. Int. J. Hydrogen Energy 2018, 43, 22075–22087. [Google Scholar] [CrossRef]
  95. Nami, H.; Mohammadkhani, F.; Ranjbar, F. Utilization of waste heat from GTMHR for hydrogen generation via combination of organic Rankine cycles and PEM electrolysis. Energy Convers. Manag. 2016, 127, 589–598. [Google Scholar] [CrossRef]
  96. Akrami, E.; Chitsaz, A.; Nami, H.; Mahmoudi, S.M.S. Energetic and exergoeconomic assessment of a multi-generation energy system based on indirect use of geothermal energy. Energy 2017, 124, 625–639. [Google Scholar] [CrossRef]
  97. Habibollahzade, A.; Gholamian, E.; Houshfar, E.; Behzadi, A. Multi-objective optimization of biomass-based solid oxide fuel cell integrated with Stirling engine and electrolyzer. Energy Convers. Manag. 2018, 171, 1116–1133. [Google Scholar] [CrossRef]
  98. Ishaq, H.; Dincer, I. Comparative assessment of renewable energy-based hydrogen production methods. Renew. Sustain. Energy Rev. 2021, 135, 110192. [Google Scholar] [CrossRef]
  99. Akrami, E.; Nemati, A.; Nami, H.; Ranjbar, F. Exergy and exergoeconomic assessment of hydrogen and cooling production from concentrated PVT equipped with PEM electrolyzer and LiBr-H2O absorption chiller. Int. J. Hydrogen Energy 2018, 43, 622–633. [Google Scholar] [CrossRef]
  100. Habibollahzade, A.; Gholamian, E.; Ahmadi, P.; Behzadi, A. Multi-criteria optimization of an integrated energy system with thermoelectric generator, parabolic trough solar collector and electrolysis for hydrogen production. Int. J. Hydrogen Energy 2018, 43, 14140–14157. [Google Scholar] [CrossRef]
  101. Ahmadi, P.; Dincer, I.; Rosen, M.A. Energy and exergy analyses of hydrogen production via solar-boosted ocean thermal energy conversion and PEM electrolysis. Int. J. Hydrogen Energy 2013, 38, 1795–1805. [Google Scholar] [CrossRef]
  102. Moharamian, A.; Soltani, S.; Rosen, M.A.; Mahmoudi, S.M.S.; Morosuk, T. Exergoeconomic analysis of natural gas fired and biomass post-fired combined cycle with hydrogen injection into the combustion chamber. J. Clean. Prod. 2018, 180, 450–465. [Google Scholar] [CrossRef]
  103. Zoghi, M.; Habibi, H.; Chitsaz, A.; Javaherdeh, K.; Ayazpour, M. Exergoeconomic analysis of a novel trigeneration system based on organic quadrilateral cycle integrated with cascade absorption-compression system for waste heat recovery. Energy Convers. Manag. 2019, 198, 111818. [Google Scholar] [CrossRef]
  104. Zoghi, M.; Habibi, H.; Chitsaz, A.; Ayazpour, M.; Mojaver, P. Thermo-economic assessment of a novel trigeneration system based on coupling of organic Rankine cycle and absorption-compression cooling and power system for waste heat recovery. Energy Convers. Manag. 2019, 196, 567–580. [Google Scholar] [CrossRef]
  105. Azizi, N.; Esmaeilion, F.; Moosavian, S.F.; Yaghoubirad, M.; Ahmadi, A.; Aliehyaei, M.; Soltani, M. Critical review of multigeneration system powered by geothermal energy resource from the energy, exergy, and economic point of views. Energy Sci. Eng. 2022, 10, 4859–4889. [Google Scholar] [CrossRef]
  106. Khan, M.N.; Zoghi, M.; Habibi, H.; Zanj, A.; Anqi, A.E. Waste heat recovery of two solar-driven supercritical CO2 Brayton cycles: Exergoeconomic analysis, comparative study, and monthly performance. Appl. Therm. Eng. 2022, 214, 118837. [Google Scholar] [CrossRef]
  107. Zhang, M.; Chen, H.; Zoghi, M.; Habibi, H. Comparison between biogas and pure methane as the fuel of a polygeneration system including a regenerative gas turbine cycle and partial cooling supercritical CO2 Brayton cycle: 4E analysis and tri-objective optimization. Energy 2022, 257, 124695. [Google Scholar] [CrossRef]
  108. Zhou, Z.; Cao, Y.; Anqi, A.E.; Zoghi, M.; Habibi, H.; Rajhi, A.A.; Alamri, S. Converting a geothermal-driven steam flash cycle into a high-performance polygeneration system by waste heat recovery: 3E analysis and Genetic-Fgoalattain optimization. Renew. Energy 2022, 186, 609–627. [Google Scholar] [CrossRef]
  109. Li, J.; Zoghi, M.; Zhao, L. Thermo-economic assessment and optimization of a geothermal-driven tri-generation system for power, cooling, and hydrogen production. Energy 2022, 244, 123151. [Google Scholar] [CrossRef]
  110. Habibi, H.; Zoghi, M.; Chitsaz, A.; Javaherdeh, K.; Ayazpour, M.; Bellos, E. Working fluid selection for regenerative supercritical Brayton cycle combined with bottoming ORC driven by molten salt solar power tower using energy–exergy analysis. Sustain. Energy Technol. Assess. 2020, 39, 100699. [Google Scholar] [CrossRef]
  111. Bejan, A.; Tsatsaronis, G.; Moran, M.J. Thermal Design and Optimization; John Wiley & Sons: Hoboken, NJ, USA, 1995. [Google Scholar]
  112. Maurya, A.; Kumar, R.; Gupta, A.; Ahmadi, M.H.; Al-Bahrani, M. Energy, exergy, economic, exergoeconomic, exergoenvironment, and enviroeconomic (6E) analysis of solar stills—A critical review. Energy Sci. Eng. 2024, 12, 267–283. [Google Scholar] [CrossRef]
  113. Zaresharif, M.; Vatani, A.; Ghasemian, M. Evaluation of different flare gas recovery alternatives with exergy and exergoeconomic analyses. Arab. J. Sci. Eng. 2022, 47, 5501–5520. [Google Scholar] [CrossRef]
  114. Kaviri, A.G.; Jaafar, M.N.M.; Lazim, T.M. Modeling and multi-objective exergy based optimization of a combined cycle power plant using a genetic algorithm. Energy Convers. Manag. 2012, 58, 94–103. [Google Scholar] [CrossRef]
  115. Manente, G.; Lazzaretto, A. Innovative biomass to power conversion systems based on cascaded supercritical CO2 Brayton cycles. Biomass Bioenergy 2014, 69, 155–168. [Google Scholar] [CrossRef]
  116. Ni, M.; Leung, M.K.H.; Leung, D.Y.C. Energy and exergy analysis of hydrogen production by a proton exchange membrane (PEM) electrolyzer plant. Energy Convers. Manag. 2008, 49, 2748–2756. [Google Scholar] [CrossRef]
  117. Chen, Y.; Feng, L.; Li, X.; Zoghi, M.; Javaherdeh, K. Exergy-economic analysis and multi-objective optimization of a multi-generation system based on efficient waste heat recovery of combined wind turbine and compressed CO2 energy storage system. Sustain. Cities Soc. 2023, 96, 104714. [Google Scholar] [CrossRef]
  118. Hai, T.; Zoghi, M.; Javaherdeh, K. 4E analysis and optimization of a biomass-fired waste-to-energy plant integrated with a compressed air energy storage system for the multi-generation purpose. Fuel 2023, 348, 128457. [Google Scholar] [CrossRef]
  119. Zhou, J.; Zoghi, M.; Abed, H. Efficient waste heat recovery of a hybrid solar-biogas-fueled gas turbine cycle for poly-generation purpose: 4E analysis, parametric study, and multi-objective optimization. Fuel 2023, 333, 126502. [Google Scholar] [CrossRef]
  120. Zoghi, M.; Habibi, H.; Chitsaz, A.; Holagh, S.G. Multi-criteria analysis of a novel biomass-driven multi-generation system including combined cycle power plant integrated with a modified Kalina-LNG subsystem employing thermoelectric generator and PEM electrolyzer. Therm. Sci. Eng. Prog. 2021, 26, 101092. [Google Scholar] [CrossRef]
  121. Habibi, H.; Zoghi, M.; Chitsaz, A.; Javaherdeh, K.; Ayazpour, M. Thermo-economic performance comparison of two configurations of combined steam and organic Rankine cycle with steam Rankine cycle driven by Al2O3-therminol VP-1 based PTSC. Sol. Energy 2019, 180, 116–132. [Google Scholar] [CrossRef]
  122. Ahmadi, P. Modeling, Analysis and Optimization of Integrated Energy Systems for Multigeneration Purposes. 2013. Available online: https://library-archives.canada.ca/eng/services/services-libraries/theses/Pages/item.aspx?idNumber=1032938246 (accessed on 25 January 2024).
  123. Al-Sulaiman, F.A.; Atif, M. Performance comparison of different supercritical carbon dioxide Brayton cycles integrated with a solar power tower. Energy 2015, 82, 61–71. [Google Scholar] [CrossRef]
  124. Akbari, A.D.; Mahmoudi, S.M.S. Thermoeconomic analysis & optimization of the combined supercritical CO2 (carbon dioxide) recompression Brayton/organic Rankine cycle. Energy 2014, 78, 501–512. [Google Scholar]
  125. Prajapati, P.; Patel, V.; Raja, B.D.; Jouhara, H. Thermal efficiency and specific work optimization of combined Brayton and inverse Brayton cycle: A multi-objective approach. Therm. Sci. Eng. Prog. 2023, 37, 101624. [Google Scholar] [CrossRef]
  126. Goodarzi, M. Energy and exergy analyses of a new atmospheric regenerative Brayton and Inverse Brayton cycle. Energy Rep. 2021, 7, 4530–4539. [Google Scholar] [CrossRef]
  127. Ahmadi, A.; Jamali, D.H.; Ehyaei, M.A.; Assad, M.E.H. Energy, exergy, economic and exergoenvironmental analyses of gas and air bottoming cycles for production of electricity and hydrogen with gas reformer. J. Clean. Prod. 2020, 259, 120915. [Google Scholar] [CrossRef]
  128. Zoghi, M.; Habibi, H.; Choubari, A.Y.; Ehyaei, M.A. Exergoeconomic and environmental analyses of a novel multi-generation system including five subsystems for efficient waste heat recovery of a regenerative gas turbine cycle with hybridization of solar power tower and biomass gasifier. Energy Convers. Manag. 2021, 228, 113702. [Google Scholar] [CrossRef]
  129. Zoghi, M.; Habibi, H.; Chitsaz, A.; Ayazpour, M. Multi-criteria performance comparison between a novel and two conventional configurations of natural gas–driven combined cycle power plant based on a hybrid multi-objective optimization. Therm. Sci. Eng. Prog. 2020, 19, 100597. [Google Scholar] [CrossRef]
  130. Habibi, H.; Zoghi, M.; Chitsaz, A.; Shamsaiee, M. Thermo-economic performance evaluation and multi-objective optimization of a screw expander-based cascade Rankine cycle integrated with parabolic trough solar collector. Appl. Therm. Eng. 2020, 180, 115827. [Google Scholar] [CrossRef]
  131. Cao, Y.; Zoghi, M.; Habibi, H.; Raise, A. Waste heat recovery of a combined solid oxide fuel cell-gas turbine system for multi-generation purposes. Appl. Therm. Eng. 2021, 198, 117463. [Google Scholar] [CrossRef]
Figure 1. Configuration 1: SRC-based combined hydrogen production system.
Figure 1. Configuration 1: SRC-based combined hydrogen production system.
Sustainability 16 06860 g001
Figure 2. Configuration 2: SBC-based combined hydrogen production system.
Figure 2. Configuration 2: SBC-based combined hydrogen production system.
Sustainability 16 06860 g002
Figure 3. Configuration 3: IBC-based combined hydrogen production system.
Figure 3. Configuration 3: IBC-based combined hydrogen production system.
Sustainability 16 06860 g003
Figure 4. Configuration 4: ABC-based combined hydrogen production system.
Figure 4. Configuration 4: ABC-based combined hydrogen production system.
Sustainability 16 06860 g004
Figure 6. The flow chart of simulation and optimization.
Figure 6. The flow chart of simulation and optimization.
Sustainability 16 06860 g006
Figure 7. The effect of THRSG,SRC on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 7. The effect of THRSG,SRC on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g007
Figure 8. The effect of pinch point temperature difference on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 8. The effect of pinch point temperature difference on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g008
Figure 9. The effect of HRSG superheating temperature on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 9. The effect of HRSG superheating temperature on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g009
Figure 10. The effect of Tcon on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 10. The effect of Tcon on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g010
Figure 11. The effect of Tin,com on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 11. The effect of Tin,com on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g011
Figure 12. The effect of Pin,com on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 12. The effect of Pin,com on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g012
Figure 13. The effect of PRcom on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 13. The effect of PRcom on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g013
Figure 14. The effect of hot side temperature difference of HRHX on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 14. The effect of hot side temperature difference of HRHX on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g014
Figure 15. The effect of IBTBP on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 15. The effect of IBTBP on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g015
Figure 16. The effect of Tin,com,IBC on the output parameters.(a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 16. The effect of Tin,com,IBC on the output parameters.(a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g016
Figure 17. The effect of PRcom.ABC on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 17. The effect of PRcom.ABC on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g017
Figure 18. The effect of hot side temperature difference of AHX on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Figure 18. The effect of hot side temperature difference of AHX on the output parameters. (a) On exergy efficiency and total exergy destruction; (b) on total cost rate and unit cost of product.
Sustainability 16 06860 g018
Figure 19. The SRC-based system’s Pareto front diagram as a result (Configuration 1).
Figure 19. The SRC-based system’s Pareto front diagram as a result (Configuration 1).
Sustainability 16 06860 g019
Figure 20. The SBC-based system’s Pareto front diagram as a result (Configuration 1).
Figure 20. The SBC-based system’s Pareto front diagram as a result (Configuration 1).
Sustainability 16 06860 g020
Figure 21. The IBC-based system’s Pareto front diagram as a result (Configuration 1).
Figure 21. The IBC-based system’s Pareto front diagram as a result (Configuration 1).
Sustainability 16 06860 g021
Figure 22. The ABC-based system’s Pareto front diagram as a result (Configuration 1).
Figure 22. The ABC-based system’s Pareto front diagram as a result (Configuration 1).
Sustainability 16 06860 g022
Table 1. ICE as the topping system and Rankine cycle as the main bottoming system.
Table 1. ICE as the topping system and Rankine cycle as the main bottoming system.
Ref. and YearSystems DescriptionAnalyses
Topping
System Description
Bottoming
System(s) Description
OptimizationEnergy
Outputs
Exergy
Outputs
Economic and Environmental
Outputs
[17] in 2012235 kW diesel engine with an exhaust temperature of 519 °C ORC with 20 different working fluids ×R141b, R123, R245fa
13.3 < η en < 16.6 %
0.3 < Electricity production cost < 0.35 $/kWh
[18] in 2013117.7 < Diesel engine power < 258.3 KW,
693.15 < Exhaust gas temperature < 808.15 K,
363.75 < Jacket cooling water temperature < 366.15 K
Thermal oil circuit/ORC with R245fa working fluid× η en = 9.2 % η ex = 21.7 %
[19] in 201358.8 < Diesel engine power < 235.8 KW,
326 < Exhaust gas temperature < 519 °C,
80.7 < Jacket cooling water temperature < 83.3 °C
SRC + Subcritical
and trans-critical ORC with 6 working fluids
×R134a is the best fluid
W ˙ n e t = 39.91   kW
η ex = 48.42 %
[20] in 2013light-duty diesel engine power = 105 kW with exhausted gas, intake air, and coolant water waste heat recoveryHT ORC (R245fa) + LT ORC (R134a) ×The output power improved by 14–16%
[32] in 2013Diesel engine power = 235.8 kW,
Exhaust gas temperature = 792.2 K,
Jacket cooling water temperature = 356.5 K
SRC or siloxane ORC at topping system + R134a trans-critical ORC bottoming system× W ˙ n e t = 39.67   kW η ex = 42.98 %
[21] in 2014Diesel engine power of 247 kW, With recovery of exhaust gas energy, waste heat from the coolant system, and released heat from turbocharged air in the intercoolerHT ORC (R245fa) + LT ORC (R245fa)× W ˙ n e t = 27.85   kW , Thermal efficiency improvement
=13%.
[22] in 201458.8 < Diesel engine power < 235.8 KW,
326 < Exhaust gas temperature < 519 °C,
80.7 < Jacket cooling water temperature < 83.3 °C
SRC + Subcritical
ORC with six working fluids
×R1234yf was the best fluid,
W ˙ n e t = 36.77   kW
η ex = 55.05 %
[23] in 2015Diesel engine power = 996 kW, Exhaust gas temperature = 573.15 K, Jacket cooling water temperature = 363.15 KORC with zeotropic mixtures
cyclohexane/R141b (0.5/0.5)
×Increasing
the net power output of the system by 13.3% compared to pure cyclohexane
[24] in 2015200 < Diesel engine power < 2000 KW,
419 < Exhaust gas temperature < 712 K
ORC (6 working fluids) or Kalina cycle×N-nonane was the best,
η en = 64.1 %
[25] in 2015Diesel engine power = 996 kW, Exhaust gas temperature = 573.15 K, Jacket cooling water temperature = 363.15 KHT ORC (5 working fluids) + LT ORC (7 working fluids)× W ˙ n e t = 101.1   kW ,
Efficiency improvement of ICE = 10.2%
[26] in 2015Diesel engine power = 996 kW, Exhaust gas temperature = 573.15 K, Jacket cooling water temperature = 363.15 KCascade partial evaporation Rankine cycle (PERC) (water)/ORC (3 working fluids)× W ˙ n e t = 115.1   kW ,
Efficiency improvement of ICE = 11.6%
[33] in 2015
(Experimental)
12.9 < Gasoline engine power < 44.8 kW,
429 < Exhaust gas temperature < 673 °C
ORC (ethanol)× η en = 6 %
[34] in 2015Diesel engine power = 247 kW,
Exhaust gas temperature = 667 K
ORC (R245fa) W ˙ n e t = 13.84   kW , Net power output per unit heat
transfer area = 0.74 kW / m 2
[35] in 2015Diesel and gasoline engine ORC (R245fa)×An increase of the overall system efficiency up to 9%
[36] in 2016 (Experimental)Diesel engine power = 243 kW,
Exhaust gas temperature = 480 °C
Cascade SRC/ORC (R123)× W ˙ n e t = 12.7   kW ,
Power enhancement = 5.6%
[37] in 2016Diesel engine power = 243 kW,
200 < Exhaust gas temperature < 450 °C
74 < Jacket cooling water temperature < 84 °C
Cascade HT ORC (4 working fluids)/LT ORC (4 working fluids)×Toluene and R134a were best fluids for HT and LT cycles, W ˙ n e t = 33.9   kW ,
η en = 9.9 %
η ex = 39.1 %
[27] in 2016Diesel engineORC (R245fa) + HWU + ACH× W ˙ n e t = 53   kW ,
Q ˙ c h i l l e r = 176.8   kW
[38] in 2017Diesel engine power = 258 kW redundant breakORC (R245fa)×Increase in power = 4.13 kW, Increase in efficiency = 0.66%
[39] in 2017Truck diesel engine power = 258.9 kW, Exhaust gas temperature = 405 °C, Jacket cooling water temperature = 88 °C Simple ORC, dual-loop ORC or cascade expansion ORC (Cyclopentane as working fluid)×dual-loop ORC power = 26.8 kW,
cascade expansion ORC power = 29 kW
[40] in 2017CNG (compressed natural gas) engine power = 210 kWCascade HT ORC (R245fa)/LT ORC (R245fa) W ˙ n e t = 23.62   k W ,
8.79 % < η en < 10.17 %
[28] in 2017180 < Diesel engine power < 250 kW,
698 < Exhaust gas temperature < 758 K
ORC (R245fa) + ACH× W ˙ n e t = 6.73   kW ,
Q ˙ c h i l l e r = 14.2   kW
[41] in 2017 Natural gas fired engine power of 2928 kW
Exhaust gas temperature = 470 °C, Jacket cooling water temperature = 79.7 °C
RC (Water, R123, R134a)× η ex , R 123 = 21.01 ,
η ex , water = 13.16 ,
η ex , R 134 a = 8.7
[42] in 2017
(Experimental)
Diesel engine power = 200 kW,
Exhaust gas temperature = 400 °C
ORC (4 working fluids) ×MDM is the best fluid,
W ˙ n e t = 8.1 9.8   kW
[43] in 2017Diesel engine power = 80,080 kW,
Exhaust gas temperature = 308 °C
ORC (R123) +
Solid oxide electrolyzer
× η en = 53.56 %
[44] in 2018Diesel engine power = 80 kWORC3.24% improvement in the output power
[45] in 2018Two different diesel engines,
245 < Exhaust gas temperature < 354 °C
ORC (23 working fluids)× 22.7 % < η en < 23.76 %
[29] in 2018Diesel engine power = 200 kW,
Exhaust gas temperature = 400 °C
Cascade PERC/ORC (6 working fluids)/LNG subsystem + HWUIso-pentane
as ORC working fluid,
W ˙ n e t = 178.6   kW
η ex = 38.74 C ˙ t o t = 19.3 $ h
[46] in 2018300 MW turbo-aspirated
compression ignition
RC (7 working fluids)×R123 in super-critical mode was the best fluid,
W ˙ n e t = 715.2   kW
[47] in 2018Diesel engineORC (7 working fluids)×R141b was the best fluid,
W ˙ n e t = 97   kW
[48] in 2019Diesel engine power = 243 kWOptimal practical ORC, Basic ORC, trans-critical ORC, regenerative ORC, split regenerative ORC, cascade ORC, Dual pressure ORC×Cyclo-pentane working fluid, Dual pressure ORC efficiency was the highest at 14.23
[49] in 2019117.7 < Automotive internal combustion engine power < 258.3 kW,
693.15 < Exhaust gas temperature < 808.15 K,
363.75 < Jacket cooling water temperature < 366.15 K
Basic ORC, Basic ORC with oil storage for higher stability×Dynamic behavior was analyzed for two systems,
W ˙ n e t = 28.61   kW ,
W ˙ n e t ,   w i t h   O S = 23.34   kW
[50] in 2019Diesel engine power = 98.9 kW
Exhaust gas temperature = 524.9 °C, Jacket cooling water temperature = 86.8 °C
Cascade HT trans-critical ORC (Toluene or cyclohexane)/LT trans-critical ORC (CO2 or R134a)×Using toluene and R134a was the best, W ˙ n e t = 24.93   kW PP = 9.24 years, Specific cost = 4361 $/kW
[51] in 2019Diesel engine power = 247 kW
Exhaust gas temperature = 635.15 K
ORC with zeotropic mixtures working fluids×0.9 toluene/0.1 decane
was the best, W ˙ n e t = 26.9   kW
Electricity production cost = 0.5975 $/kWh
[30] in 2019Diesel engine power = 98.9 kW
Exhaust gas temperature = 524.9 °C, Jacket cooling water temperature = 86.8 °C
High-temperature Kalina cycle× W ˙ n e t = 21.74   kW ,
η en = 25.55 %
η ex = 55.52 % Unit cost of electricity = 15.52 cent/kWh
[52] in 2020Diesel engine power = 235.8 kW
Exhaust gas temperature = 519 °C, Jacket cooling water temperature = 83.5 °C
Cascade HT ORC/LT ORC with 24 candidate working fluid pairstoluene/R124 was the best pair,
η en = 39 %
PP = 1.26 years
[53] in 2020Natural gas engine power = 85 Kw, Exhaust gas temperature = 850 °C, Jacket cooling water temperature = 95 °C ORC + RODUFreshwater mass flow rate =2.926 kg/s η ex = 30.7 %
[54] in 2021
(Experimental)
Diesel engine power = 1000 kW
Exhaust gas temperature = 530 °C, Jacket cooling water temperature = 84 °C
Cascade HT ORC (R245fa)/LT ORC (R134a)× W ˙ n e t = 310   kW ,
η en = 9.5 %
η ex = 43 %
[31] in 2021Diesel engine power = 603 kW
Exhaust gas temperature = 659.3 °C, Jacket cooling water temperature = 78.6 °C
ORC + HWU + HRSG + RODU η en = 82.82 % η ex = 54.1 %
[55] in 2022Biogas fueled internal combustion engine,
Exhaust gas temperature = 587.08 °C
Trans-critical ORC (9 working fluids)n-decane was the best fluid, W ˙ n e t = 156.4   kW PP=5.537 years
[56] in 2023Vehicle engine under road conditionORC (R245fa) η en = 82.82 % ,
power output per unit heat transfer area = 0.55 kW / m 2
The present study NG fueled ICE
( W ˙ I C E = 0.584   MW )
SRC (Single pressure level HRSG)/TEG 2 + TEG 1 + PEME η ex = 21.93 % C ˙ t o t = 31.82 $ h ,
Unit cost of outputs = 101.3 $/GJ
Table 2. ICE as the topping system and SBC as the main bottoming system.
Table 2. ICE as the topping system and SBC as the main bottoming system.
Ref. and YearSystems DescriptionAnalyses
Topping
System Description
Bottoming
System(s) Description
OptimizationEnergy
Outputs
Exergy
Outputs
Economic and Environmental
Outputs
[68] in 2018Diesel engine power = 996 kW
Exhaust gas temperature = 300 °C, Jacket cooling water temperature = 90 °C
Regenerative CO2 SBC or
Improved
regenerative CO2 SBC
× W ˙ n e t = 68.4   kW
[60] in 2019Diesel/natural gas dual-fuel engine,
126.8 < engine power < 251.1 kW
423.2 < Exhaust gas temperature < 488.3 °C
Cascade Regenerative CO2 SBC/ORC (7 working fluids)× W ˙ n e t = 40.88   kW ,
improved the dual-fuel engine power output by 6.78%
[61] in 2020Diesel engine power = 2928 kW
Exhaust gas temperature = 470 °C
Cascade Regenerative CO2 SBC/ACH 248.1 < W ˙ n e t < 253.9 ,
70.5 < Cooling capacity < 168.8,
η en = 39.85 %
η ex = 53.54 % Unit cost of outputs = 3.41 $/GJ
[62] in 2020Diesel engine power = 41,840 kWRecompression CO2 SBC + Kalina cycle W ˙ n e t = 1976.1   kW ,
reduced the annual fuel consumption by 16.62%
[69] in 2020Diesel engineRegenerative CO2 SBC coupled with trans-critical CO2 refrigeration cycle× W ˙ n e t = 20.88   k W , Cooling capacity = 59.47 kW η ex = 12.52 %
[70] in 2020184.8 < Diesel/natural gas dual-fuel engine power < 197 kW,
Exhaust gas temperature = 557.6 K
Four configurations of CO2 SBC/ORC × W ˙ n e t = 33.77   kW ,
η en = 32.92 %
η ex = 65.81 %
[63] in 2020Diesel/natural gas dual-fuel engine power = 197 kW, Exhaust gas temperature = 557.5 °C Regenerative CO2 SBC couple with ejector expansion refrigeration cycle (zeotropic mixtures)×Cooling capacity = 225.5 kW,
COP = 2.05
[71] in 2020Diesel engine power = 1170 kW
Exhaust gas temperature = 457 °C, Jacket cooling water temperature = 89 °C
Cascade Regenerative CO2 SBC/ORC (R245fa)× W ˙ n e t = 215   kW
[72] in 2020Diesel engine power = 235.8 kW
Exhaust gas temperature = 519 °C
Novel CO2 SBC W ˙ n e t = 39.49   kW ,
η en = 35.86 % ,
η ex = 67.9 %
[73] in 2021 Automobile engineRegenerative CO2 SBC×Transient characteristic evaluation of the system,
W ˙ n e t = 2.97   kW
Electricity production cost = 0.38 $/kWh
[74] in 20211071 < Natural gas engine power < 2108 kW,
590 < Exhaust gas temperature < 690 °C
Vortex Tube (VT) heat booster coupled with Regenerative CO2 SBC and Recompression CO2 SBC×Implementing VT boosts energy and exergy efficiencies by up to 1.85% Implementing VT
decreases exergy destruction by around 8–12%
Electricity price = 0.3 $/kWh,
8 < PP < 12 years,
Adding VT contributed to reducing LCOE by 10–15% and the payback period by around 3–5 years
[75] in 2021Marine Diesel engineDifferent Confs. of CO2 SBC×The best CO2 cycle improved the efficiency by 6.6–7.25%
[64] in 2022 Internal combustion engine, Exhaust gas temperature = 519 °C Cascade Regenerative CO2 SBC/TEG 1/ORC + Modified Kalina cycle/TEG 2 W ˙ n e t = 2657   kW , η en = 26.33 % η ex = 51.69 %
[65] in 2022Diesel engine power = 235.8 kW
Exhaust gas temperature = 519 °C,
Cascade partial heating CO2 SBC/Trans-critical CO2 ORC W ˙ n e t = 841.84   kW , η en = 29.25 % Electricity production cost = 7.43 cent/kWh
[66] in 2022Engine power = 1323.1 kW
Exhaust gas temperature = 633.1 °C
Cascade Regenerative CO2 SBC/Ejector refrigeration cycle + ORC (Benzene/R365mfc) η en = 33.17 % , η ex = 61.93 % Electricity production cost = 5.34 cent/kWh
[76] in 2023Gas engine power = 45 kW
Exhaust gas temperature = 350 °C, Jacket cooling water temperature = 90 °C
Gas engine coupled with heat pump and HWU + Regenerative CO2 SBC × W ˙ n e t , S B C = 2.14   kW η ex , SBC = 36.61 %
[67] in 2023Ship engine,
Exhaust gas temperature = 572 °C
New Regenerative CO2 SBC for power and cooling + Trans-critical CO2 ORC + LNG subsystem η en = 69.6 % ,
W ˙ n e t = 9.06   kW ,
Cooling capacity = 19.52 kW
η ex = 42.3 % Total cost= 2.5 M$
[77] in 2023584 < Natural gas (NG) engine power < 1167 kW,
445 < Exhaust gas temperature < 493 °C
Partial heating CO2 SBC η en = 48.94 % , η ex = 52.01 % Unit cost of electricity = 13.19 euro/kWh
Present studyNG fueled ICE
( W ˙ I C E = 0.584   M W )
Regenerative SBC (CO2)/TEG 2 + TEG 1 + PEME η ex = 18.05 % C ˙ t o t = 30.75 $ h ,
Unit cost of outputs = 93.38 $/GJ
Table 3. ICE as the topping system and IBC as the main bottoming system.
Table 3. ICE as the topping system and IBC as the main bottoming system.
Ref. and YearSystems DescriptionAnalyses
Topping
System Description
Bottoming
System(s) Description
OptimizationEnergy
Outputs
Exergy
Outputs
Economic and Environmental
Outputs
[80] in 2019Diesel engine power = 351 kWIBC×Power improvement due to IBC = 5%
[81] in 201963.6 < Diesel engine power < 91.1 kW,
200 < Exhaust gas temperature < 500 °C
IBC×Waste heat recovery of 1.5–2% of engine power by IBC
[85] in 2020Diesel engine power = 130 kWIBC×Efficiency improvement due to IBC = 3.4%
[82] in 2020Engine with CH2 fuel, Exhaust gas temperature = 1140 K,Five different configurations of IBC (with liquid water drainage, steam turbine, and refrigeration cycle)× η en = 25 %
[86] in 2020Methane engine power = 1.4 MW, 470 < Exhaust gas temperature < 570 °C IBC/Thermal oil loop/Regenerative ORC η en = 13.3 % LCOE = 146.1 $/MWh
[83] in 2020500 < Methane engine power < 1400 kW, 400 < Exhaust gas temperature < 600 °C IBC/Thermal oil loop/Basic ORC or regenerative ORC (3 different working fluids)Pentane was the best fluid, W ˙ n e t = 152.9   k W , η en = 11.09 % LCOE = 159.5 $/MWh
[84] in 2022Diesel engine maximum power = 257 kW
1056.9 < Exhaust gas temperature < 1071.5 K
IBC/ORC (R245fa)×IBC/ORC system leads to enhancement of the system power by about 18%
Present studyNG fueled ICE
( W ˙ I C E = 0.584   M W )
IBC/TEG 1 + TEG 2 + PEME η ex = 13.72 % C ˙ t o t = 25.58 $ h ,
Unit cost of outputs = 59.91 $/GJ
Table 4. Fundamental correlations for a few configuration components.
Table 4. Fundamental correlations for a few configuration components.
CorrelationNoteNumber
Thermoelectric generator [106,107]
W ˙ T E G = η T E G · Q ˙ c o l d s i d e Output power(1)
Q ˙ c o l d s i d e = m ˙ c o l d ( h o u t h i n ) Cold side energy rate(2)
η T E G = η C a r n o t 1 + Z T M 1 1 + Z T M + T L T H Efficiency of TEG(3)
η C a r n o t = 1 T L T H Carnot efficiency(4)
Proton exchange membrane electrolyzer [108,109]
n ˙ H 2 , o u t = J 2 F Molar rate of hydrogen(5)
E n i = J . V = W ˙ b o t t o m i n g , m a i n + W ˙ T E G , 1 + W ˙ T E G , 2 The energy provided to the PEME(6)
V = V 0 + V a c t , a n + V a c t , c a + V o h m Electric potential in PEME(7)
V 0 = 1.229 0.00085 ( T P E M 298 ) Nernst equation(8)
V a c t , i = R T P E M F s i n h 1 ( J 2 . J 0 , i ) Activation over potential(9)
J 0 , i = J i r e f e x p ( E a c t , i R T P E M ) Exchange current density(10)
V o h m = J . R P E M Ohmic over potential(11)
R P E M = 0 D d x ω P E M . θ x Overall resistance of the PEME(12)
θ ( x ) = θ a θ c D . x + θ c Membrane surface water(13)
ω P E M = ( 0.5139   θ ( x ) 0.326 ) · e x p ( 1268 ( 1 303 1 T P E M ) ) Local ionic conductivity(14)
Table 5. The parameters in the optimization process, together with their range.
Table 5. The parameters in the optimization process, together with their range.
ParameterValue/Description
SRC (Configuration 1)
T H R S G   ( ° C ) 170–300
Δ T p i n c h , H R S G   ( ° C ) 5–50
Δ T s u p e h e a t , H R S G   ( ° C ) 50–150
T c o n ( ° C ) 30–50
SBC (Configuration 2)
T i n , c o m ( ° C ) 30–50
P i n , c o m   ( k P a ) 7400–11,000
P R c o m 1.2–3
Δ T H S , H R H X   ( ° C ) 10–50
IBC (Configuration 3)
I B T B P   ( kPa ) 30–80
T i n , c o m   ( ° C ) 30–100
ABC (Configuration 4)
P R c o m 3–7
Δ T H S , A H X   ( ° C ) 10–50
Table 7. Input data for the baseline condition.
Table 7. Input data for the baseline condition.
ParameterValue
Reference temperature ( T 0 )20 °C
Reference pressure ( P 0 )101.3 kPa
NG ICE (All Configurations) [77]
Mass flow rate of exhausted gas ( m ˙ g a s )3303 kg   h 1
Temperature of the exhaust gas ( T 1 )493 °C
Pressure of the exhaust gas ( P 1 ) 105 kPa
TEG (All Configurations) [117,118]
Temperature of TEGs water inlet 25 °C
Temperature of TEGs water outlet35 °C
Z T M 0.9
PEME (All Configurations) [119,120]
T P E M E 80 °C
E a c t , a n 76 kJ / mol
E a c t , c a 18 kJ / mol
θ a n 14
θ c a 10
D 100 μ m
F 96,486 C / mol
J a n r e f 1.7 × 10 5   A / m 2
J c a r e f 4.6 × 10 3   A / m 2
SRC (Configuration 1) [121,122]
HRSG evaporation temperature ( T H R S G )250 °C
HRSG pinch temperature difference ( Δ T p p , H R S G ) 30 °C
HRSG degree of superheating ( Δ T s u p , H R S G ) 100 °C
SRC condensation temperature ( T c o n )40 °C
SRC pump’s isentropic efficiency0.8
SRC turbine’s isentropic efficiency0.85
SBC (Configuration 2) [123,124]
Compressor inlet temperature ( T 9 )40 °C
Compressor inlet pressure ( P 9 )8000 kPa
Pressure ratio related to compressor ( P R S B C )2.5
Hot side temperature difference of HRHX ( Δ T H S , H R H X ) 25 °C
The isentropic efficiency of SBC compressor0.86
The isentropic efficiency of SBC turbine0.86
IBC (Configuration 3) [125,126]
Inverse Brayton turbine back pressure (IBTBP) ( P 2 )50 kPa
Compressor inlet temperature ( T 3 )65 °C
The isentropic efficiency of IBC compressor0.86
The isentropic efficiency of IBC turbine0.86
ABC (Configuration 4) [87,127]
Pressure ratio related to compressor ( P R A B C )5
Hot side temperature difference of AHX ( Δ T H S , A H X ) 25 °C
Minimum temperature of air ( T 10 )30 °C
The isentropic efficiency of ABC compressor0.86
The isentropic efficiency of ABC turbine0.86
Table 8. The four layouts’ optimal operating points.
Table 8. The four layouts’ optimal operating points.
Decision VariablesValues
SRC (Configuration 1)
T H R S G , S R C 284 °C
Δ T p i n c h , e v a , S R C 42.46 °C
Δ T s u p , e v a , S R C 14.6 °C
T c o n , S R C 131.2 °C
SBC (Configuration 2)
T i n , c o m ,   S B C 34.47 °C
P i n , c o m ,   S B C 40.84 °C
P R c o m , S B C 8360 kPa
Δ T H S , H R H X , S B C 2.992
IBC (Configuration 3)
P i n , t u r ,   I B C 59.71 kPa
T i n , c o m ,   I B C 41.27 °C
ABC (Configuration 4)
P R c o m , A B C 3.018
Δ T H S , A H X , A B C 40.48 °C
Table 9. Values of the objective functions for Configurations 1–4 at the optimal condition.
Table 9. Values of the objective functions for Configurations 1–4 at the optimal condition.
ParametersSRC-Based System
(Configuration 1)
SBC-Based System
(Configuration 2)
IBC-Based System
(Configuration 3)
ABC-Based System
(Configuration 4)
η e x ( % ) 21.9318.0513.7213.26
C ˙ t o t ( $   h 1 ) 31.8230.7525.5843.94
c h y d r o g e n ( $   G J 1 ) 101.393.3859.91108.6
S p e c i f i c   c o s t   o f   s y s t e m = C ˙ t o t E x ˙ p r o d u c t ( $ / k W h ) 0.61590.72350.79191.407
Table 10. System performances at their optimal condition.
Table 10. System performances at their optimal condition.
ParameterSRC-Based System
(Configuration 1)
SBC-Based System
(Configuration 2)
IBC-Based System
(Configuration 3)
ABC-Based System
(Configuration 4)
W ˙ I C E ( K W ) 584584584584
W ˙ m a i n , b o t t o m i n g ( K W ) 102.575.1436.9141.73
W ˙ T E G , 1 ( K W ) 4.957.3825.843.22
W ˙ T E G , 2 ( K W ) 1.644.3140.7116.16
W ˙ P E M E ( K W ) 109.186.8563.7461.16
m ˙ h y d r o g e n ( k g h ) 1.5871.3060.9920.959
E x ˙ f u e l , t o t ( k W ) 235.5235.5235.5235.5
E x ˙ p r o d u c t , t o t ( k W ) 51.6642.5132.3131.24
E x ˙ d e s t , t o t ( k W ) 137146156.6157.2
E x ˙ l o s s , t o t ( k W ) 46.8646.9846.6747.11
η e x ( % ) 21.9318.0513.7213.26
Z ˙ t o t ( $ / h ) 5.522.7321.989.348
C ˙ d e s t , t o t ( $ / h ) 16.3418.0613.6424.63
C ˙ e n v ( $ / h ) 9.969.969.969.96
C ˙ t o t ( $ / h ) 31.8230.7525.5843.94
c e l , b o t t o m i n g   ( $ / G J ) 43.140.8325.6150.63
c h y d r o g e n ( $ / G J ) 101.393.3959.91108.6
Table 11. Component performances at their optimal state.
Table 11. Component performances at their optimal state.
System/Components W ˙   or   Q ˙ ( kW ) E x ˙ f ( kW ) E x ˙ p ( kW ) E x ˙ D ( kW ) η e x ( % ) Z ˙ ( $ / h ) C ˙ d e s t ( $ / h )
Bottoming systems Configuration 1
HRSG317.2162.6131.830.881.060.8262.61
SRC turbine 103.4120.6103.417.2285.722.561.92
TEG 21.6411.983.518.4629.320.0440.946
SRC pump0.9020.9020.7310.17181.030.0320.026
TEG 14.9529.315.9323.3820.240.1341.98
PEME109.1109.152.1156.9647.781.978.83
Bottoming systems Configuration 2
HRHX281.3149.1139.79.4493.670.3320.801
SBC turbine101.4109.4101.48.0592.640.1961.08
Recuperator257.6101.967.2534.6266.020.2164.65
TEG 24.31418.666.0812.5732.630.1171.68
SBC compressor26.2426.2423.043.2087.80.1420.47
TEG 17.3842.848.6534.1820.210.22.90
PEME86.8585.8542.8843.9749.371.526.46
Bottoming systems Configuration 3
IBC turbine92.1498.7992.146.6493.270.1080.564
TEG 125.8413828.85109.220.90.7019.27
IBC compressor55.2355.2348.966.2788.640.0380.578
TEG 20.7194.520.9533.5721.060.0190.376
PEME63.4763.4732.5930.8851.351.112.84
Bottoming systems Configuration 4
ABC compressor141.4141.4127.314.1789.980.1672.58
Air heat exchanger347.2172.6160.811.893.167.331.00
ABC turbine183.2199.2183.215.9791.980.2432.65
TEG 13.2219.393.9415.4520.340.0871.31
TEG 216.1688.8418.6570.1920.990.43811.68
PEME61.1661.1131.5129.5951.571.0745.39
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

Zoghi, M.; Hosseinzadeh, N.; Gharaie, S.; Zare, A. A Comprehensive Study on Hydrogen Production via Waste Heat Recovery of a Natural Gas-Fueled Internal Combustion Engine in Cogeneration Power-Hydrogen Layouts: 4E Study and Optimization. Sustainability 2024, 16, 6860. https://doi.org/10.3390/su16166860

AMA Style

Zoghi M, Hosseinzadeh N, Gharaie S, Zare A. A Comprehensive Study on Hydrogen Production via Waste Heat Recovery of a Natural Gas-Fueled Internal Combustion Engine in Cogeneration Power-Hydrogen Layouts: 4E Study and Optimization. Sustainability. 2024; 16(16):6860. https://doi.org/10.3390/su16166860

Chicago/Turabian Style

Zoghi, Mohammad, Nasser Hosseinzadeh, Saleh Gharaie, and Ali Zare. 2024. "A Comprehensive Study on Hydrogen Production via Waste Heat Recovery of a Natural Gas-Fueled Internal Combustion Engine in Cogeneration Power-Hydrogen Layouts: 4E Study and Optimization" Sustainability 16, no. 16: 6860. https://doi.org/10.3390/su16166860

APA Style

Zoghi, M., Hosseinzadeh, N., Gharaie, S., & Zare, A. (2024). A Comprehensive Study on Hydrogen Production via Waste Heat Recovery of a Natural Gas-Fueled Internal Combustion Engine in Cogeneration Power-Hydrogen Layouts: 4E Study and Optimization. Sustainability, 16(16), 6860. https://doi.org/10.3390/su16166860

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