You are on page 1of 23

Case Studies in Thermal Engineering 53 (2024) 103902

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


journal homepage: www.elsevier.com/locate/csite

Optimizing power, cooling, and hydrogen generation: A


thermodynamic and exergoeconomic study of an advanced sCO2
trigeneration system
Mohamed S. Yousef a, b, *, Domingo Santana a
a
Department of Thermal and Fluids Engineering, Carlos III University of Madrid, Avda. de la Universidad 30, 28911, Leganés, Madrid, Spain
b
Department of Mechanical Engineering, Benha Faculty of Eng., Benha University, Benha, Egypt

A R T I C L E I N F O A B S T R A C T

Handling Editor: Huihe Qiu This paper presents an innovative trigeneration system designed for efficient power, cooling, and
Keywords:
hydrogen production. It combines a supercritical carbon dioxide (sCO2) power cycle with a Kalina
sCO2 cycle cycle (KC) and an ammonia-water-based absorption refrigeration cycle (ARC), all integrated with
Kalina cycle a PEM electrolyzer (PEME) unit. The system optimally utilizes waste heat from the sCO2 power
ARC cycle to enhance power generation through the KC and provide cooling via the ARC. Additionally,
Hydrogen it leverages the PEME system and KC-generated power for eco-friendly hydrogen production.
Exergoeconomic Mathematical models, thermodynamic, and exergoeconomic analyses were performed, including
Optimization parametric studies, optimization, and comparative analyses. The results indicate that the reactor
experiences the highest exergy destruction rate, while components in the bottoming cycles exhibit
lower exergy destruction. From an exergoeconomic perspective, the reactor and sCO2 turbine are
ranked as the first and second most significant components. Under the optimal conditions, the
system achieved a 9.76 % increase in exergy efficiency and a 6.63 % reduction in total product
unit cost. The system also provides substantial net power output, cooling capacity, and hydrogen
production rates of 261.74 MW, 123.95 MW, and 176.328 kg/h, respectively. These findings
highlight the system’s significant thermodynamic and economic advantages, making it a prom­
ising choice for diverse user needs.

Nomenclature

A Heat transfer area (m2)


Ċ Cost rate ($/h)
c Cost per exergy unit ($/GJ)
cP,total Total product unit cost ($/GJ)
Ė Exergy rate (kW)
e Standard chemical exergy (kJ/kg)
F Faraday constant (c/mol)
fk Exergoeconomic factor

* Corresponding author. Department of Thermal and Fluids Engineering, Carlos III University of Madrid, Avda. de la Universidad 30, 28911, Leganés, Madrid,
Spain.
E-mail address: myousef@ing.uc3m.es (M.S. Yousef).

https://doi.org/10.1016/j.csite.2023.103902
Received 7 September 2023; Received in revised form 30 November 2023; Accepted 8 December 2023
Available online 14 December 2023
2214-157X/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

h Specific enthalpy (kJ/kg)


i Interest rate
J Current density (A/m2)
LMTD Log mean temperature difference (oC)
ṁ mass flow rate (kg/s)
M Molar mass (kg/kmol)
n Number of operation year
P Pressure (bar) & (kPa)
PRc Compressor pressure ratio
Q̇ Heat transfer rate (kW)
R Overall ohmic resistance, (Ω)
rk Relative cost difference
s specific entropy (kJ/kg K)
T Temperature (oC) & Turbine
U Overall heat transfer coefficient (W/m2K)
V Voltage
Ẇ Power (kW)
X Ammonia mass fraction (%)
Z Capital cost of a component ($)
Ż Capital cost rate ($/h)

Greek Symbols
η Efficiency (%)
ϵ Effectiveness
Δ Difference
τ Annual plant operation Hours (h)
γ Maintenance factor
μ Entrainment ratio
λ Content of water
σ Total ionic conductivity

Subscripts and Abbreviations


0 Ambient state
APC Absorption Power Cycle
act activation
CRF Capital recovery factor
CCP Combined cooling and power system
COND Condenser
ch Chemical
c Cathode
COD Cost optimal design
d Destruction & Diffuser section
EES Engineering Equation Solver
ERC Ejector refrigeration cycle
EUF Energy utilization factor
E.V1 Expansion valve 1
E.V2 Expansion valve 2
E.V3 Expansion valve 3
EVAP Evaporator
EEOD Energy efficiency optimal design
ex Exergy
exi exit
F Fuel
Gen1 Generator1
Gen2 Generator2
HTR High temperature recuperator
H2 Hydrogen
Gen2,hot Hot end temperature difference for generator2
in Inlet
KC Kalina cycle
k Component number

2
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

LTR Low temperature recuperator


LiBr Lithium bromide
L loss
MC Main compressor
mf Mixed flow
m Mixing chamber
n Nozzle
ORC Organic Rankine cycle
OM Operation and maintenance
ohm ohmic
out Outlet
PC Pre-cooler
P Product
PEME Proton exchange membrane electrolysis
RC Recompression compressor
R Reactor
ref reference
SCRBC supercritical CO2 recompression Brayton
SR Split ratio
sf Secondary flow
SHX Solution heat exchanger
TUR1 sCO2 Turbine
TUR2 KC Turbine
tot Total
TTD Terminal temperature difference
tCO2 Transcritical CO2 cycle
V.G1 Vapor generator1
V.G2 Vapor generator2
XEOD Exergy efficiency optimal design

1. Introduction
The use of clean energy not only mitigates pollutant emissions but also significantly enhances energy conversion efficiency, offering
solutions to address the growing demand for fossil fuels and environmental degradation. Traditional power cycles, like Rankine and
Brayton cycles, face challenges such as efficiency constraints, size limitations, extended start-up times, material restrictions, and
environmental impacts. Notably, their substantial water demand poses challenges in regions with limited water resources, and steam-
based cycles are constrained by temperature and pressure thresholds, limiting overall efficiency. In contrast, the supercritical carbon
dioxide (sCO2) Brayton cycle overcomes these challenges by operating at elevated temperatures and pressures, resulting in improved
efficiency, a smaller footprint, faster start-up times, and enhanced transient response. Advanced materials and innovative heat
exchanger designs further enhance the sCO2 Brayton cycle’s potential, positioning it as a superior, more efficient, and environmentally
friendly power generation alternative [1]. As Gen IV nuclear reactor technology advances, nuclear energy use is rapidly growing as a
reliable, cost-effective, and environmentally friendly power generation option. However, in the sCO2 Brayton cycle’s cooling process,
around 50 % of energy inputs are dissipated into the environment, resulting in significant energy loss. This heat dissipation is
attributed to the design of the cooling heat exchanger, operating at a higher temperature compared to the condenser in the traditional
Rankine Steam Cycle. As highlighted by the United States Department of Energy (DOE), the sCO2 Brayton cycle’s single-phase
operation eliminates the need for extra energy to convert between liquid and gas phases, enhancing overall energy conversion effi­
ciency. Elevated sCO2 operating temperatures lead to high enthalpies and greater densities than steam, reducing the working fluid
volume and system size, lowering capital costs. Despite these benefits, increased heat rejection in the sCO2 Brayton cycle is a
consequence of the higher cooling heat exchanger operating temperature, necessitating a greater heat rejection rate to maintain a
constant temperature. The heightened temperature difference between the working fluid and the cooling medium contributes to this
elevated heat rejection rate [2].
Researchers are actively exploring a range of methods to effectively harness this wasted thermal energy. One promising option is to
harness the wasted heat from the gas cooler of sCO2 power cycles to generate electricity. This paves the way for energy cascade
utilization, potentially eliminating energy wastage. An organic Rankine cycle (ORC), a transcritical CO2 cycle (tCO2), an absorption
power cycle (APC), and a Kalina cycle (KC) are some of the cycles that show great promise for utilizing low-grade waste heat [3]. The
research paper by Akbari and Mahmoudi [4] presents an intriguing proposal for energy production, namely a combined sCO2/ORC
system. The researchers conducted a comprehensive analysis, focusing on the thermodynamic and exergoeconomic aspects of the
proposed system. Their investigation aimed to identify the optimal operating conditions for achieving maximum cost-effectiveness.
Through their analysis, the researchers determined that utilizing the RC318 refrigerant offers the most economical operation

3
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

condition for the combined sCO2/ORC system. Zhu et al. [5] introduced a simplified boiling temperature type of boundary condition
(BTBC) to assess a printed circuit heat exchanger’s (PCHE) performance with semicircular channels, acting as a precooler in a com­
bined sCO2 Brayton/ORC. Comparing BTBC with the heat flux type of boundary condition (HFBC) and assessing two common heat
transfer correlations (Gnielinski and modified Jackson), the results highlight the significant influence of the PCHE’s different tem­
perature sections on heat transfer coefficients. Moreover, BTBC demonstrates more pronounced effects on heat transfer coefficients at
various precooler junctions than HFBC.In their study, Wu et al. [6] analyzed the SCRB/OFC, a system that combines supercritical
carbon dioxide recompression Brayton and organic flash cycle. This system utilizes the waste heat from the SCRBC to generate power.
The researchers found that the exergy efficiency of the SCRB/OFC was 6.57 % higher, and the total product unit cost was 3.57 % lower
compared to the SCRBC. According to Wang et al. [7,8], a combination of sCO2 and tCO2 systems can enhance the effectiveness of a
standalone recompression sCO2 power system. Their research indicated that the optimized sCO2/tCO2 system was similarly efficient in
terms of exergy compared to the sCO2/ORC system, which was also studied. The potential of the APC system using low-grade heat
sources was explored by Li et al. [9], who proposed a combined sCO2 power cycle and an APC system. Through their comparative
study, they found that this combination significantly improved energy efficiency by up to 5.98 %, while reducing total product unit
costs as much as 4.24 %.
The KC is a modified version of the Rankine cycle that uses a binary ammonia-water mixture as the working fluid. This mixture
offers several advantages over using a pure fluid, such as being able to vary the evaporation and condensation temperatures while
maintaining a constant pressure. This temperature matching in the evaporator and condenser leads to improved system performance.
This has made the KC particularly attractive for power generation from low-grade heat sources compared to traditional Rankine cycles.
Shokati et al. [10] conducted a comparative study of four different cycles utilizing geothermal energy: simple ORC, dual fluid ORC,
dual pressure ORC, and KC. When the systems were optimized for maximum net power with 175 ◦ C heat source as their reference
temperature, the results revealed that compared to its counterparts, the KC had an incredibly low unit cost of produced power at 26.2
%, 66.7 %, and 52.1 % lower than those same values in respective optimal states from all other tested cycles. Because of its superior
efficiency compared to the traditional steam power cycle, there has been an increasing research interest in evaluating the KC as a
bottoming cycle for sCO2 power cycles. Li et al. [11] developed a way to harness and reuse the low-grade heat from sCO2 systems,
resulting in not only improved exergy efficiency of 8.02 %, but also reduced total product unit costs by 5.50 %. By implementing this
KC technique, they were able to reduce both their exergy destruction rate as well as its associated cost rate by 9.75 % and 8.57 %,
respectively. Feng et al. [12] revealed the potential of a combined SCRBC-KC to significantly reduce fuel consumption and increase
energy efficiency in marine low-speed diesel engines; their thermodynamic analysis indicated an annual decrease of up to 16.62 % for
the former and 15.01 % for the latter design index. In another study, Fan and Dai [13] have integrated two sCO2 Brayton cycle
configurations with KC to create a more efficient energy conversion system. The findings consistently demonstrated that the recom­
pression cycle outperformed the simple cycle, exhibiting significant enhancements in both energy conversion efficiency (up to 7.53 %)
and cost effectiveness (up to 5.84 %).
In areas such as distributed energy systems and islands, power demand is not the only important factor to consider [14]. Apart from
this basic need for resources, cooling capacity is also vital [15]. To capitalize upon this potential requirement, many researchers have
invested again in combined cooling and power (CCP) systems powered by nuclear energy based sCO2 Brayton cycle [16]. Akbari et al.
[17] introduced a novel cooling and power supply system comprising SCRBC and tCO2 cycle components, which included an inte­
grated expander. Economic analysis and optimization were then conducted on this system. The optimized model exhibited remarkable
outcomes, with a 3.5 % reduction in the total product unit cost compared to models that solely emphasized exergy efficiency opti­
mization. Manjunath et al. [18] proposed a novel approach for recovering waste heat onboard ships. Their system combines a tran­
scritical compression refrigeration cycle with a regenerative sCO2 to enhance power output by 18 % and achieve a coefficient of
performance (COP) of 2.75. Building upon this work, Yu et al. [19] further improved the system by introducing a recompression
configuration in the power sub-cycle and incorporating a low-temperature recuperator in the refrigeration side, resulting in even better
performance. The absorption refrigeration cycle (ARC) is a thermodynamic cycle that facilitates cooling through the processes of
absorption and evaporation. Unlike conventional refrigeration systems that rely on compressors, the ARC operates using a combination
of a refrigerant and an absorbent fluid. The ammonia-water (NH3–H2O) and lithium bromide-water (LiBr–H2O) are two of the most
common working fluids in ARC. Compared to water, ammonia has a lower evaporation temperature range and can support heat
sources at temperatures between 50 and 200 ◦ C. These properties make it preferable for use with ARC compared to other options like
LiBr–H2O, leading its selection as the primary refrigerant fluid [20]. Wu et al. [21] conducted research on energy conservation and
efficiency by investigating a combination of sCO2 and NH3–H2O based ARC. Through their optimization method, they achieved a
remarkable improvement of over 25 % in system efficiency. This outcome inspired Li et al. [22] to delve deeper into similar systems,
this time combining sCO2 with a LiBr–H2O based ARC. The results demonstrated an even more efficient and low-pressure system, with
an increase in refrigeration coefficient by 0.3112 and a negligible decrease of only 0.0004 in the power coefficient.
Utilizing low-grade waste heat through power-cooling cogeneration cycles like the KC and Ejector Refrigeration Cycle (ERC) has
garnered substantial attention in literature. These cycles enable the simultaneous generation of power and cooling, effectively utilizing
otherwise wasted residual heat. Particularly valuable for industries producing significant low-grade heat, KC and ERC systems offer a
dual advantage: reducing energy wastage while improving operational efficiency. Du and Dai [23] undertook detailed evaluation of a
novel cogeneration system that amalgamates the KC and an ERC. Fitted with a separator, five heat exchangers, an ejector, two pumps,
turbine, an ejector as well as expansion valves all using R134a fluid; simulations revealed impressive power (619.74 kW) and cooling
(71.28 KW) production at base conditions which further improved when manipulating mass flow ratio or inlet water temperature.
Maximum exergy efficiency was achieved through evaporators operating at 6 ◦ C. Barkhordarian et al. [24] presented an innovative
power-cooling cogeneration cycle that combines KC and ERC to achieve adjustable ratios of produced energy to refrigeration. In their

4
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

research, they found that the base case exergy efficiency was 42.75 % with 19 % thermal efficiency simultaneously.
The ARC holds distinct advantages over the ERC in the realm of cooling production. ARC offers notable energy efficiency due to its
utilization of low-grade heat sources, making it particularly suitable for waste heat recovery [25]. Unlike ERC, ARC doesn’t require
moving parts such as compressors, resulting in reduced maintenance and lower noise levels [26]. Additionally, ARC systems operate
with environmentally friendly refrigerants and have the capacity to handle a wide range of cooling capacities. These advantages
position the absorption refrigeration cycle as a compelling choice for efficient and environmentally conscious cooling production [21].
Kalina-cycle based ERC have achieved lower coefficient of performance compared with water-ammonia absorption alternatives in one
dimensional assumptions, thereby reducing their potential for cooling load generation. Consequently, developing an effective and
efficient system for generating a reliable cooling source is paramount in our modern age [27]. Extensive research has been conducted
on absorption power-cooling combined systems featuring KC and ARC, both in practice and in theory. Among these studies, Wang et al.
[28] have contributed by introducing a new hybrid system that combines the principles of ARC and KC. Their approach involves using
the outlet flow of the turbine which then passes through a separator where the ammonia vapor is extracted, leaving behind purer
ammonia. This vapor is then directed towards the evaporator to create a cooling output with greater efficiency. Cao et al. [29] pre­
sented a multigeneration system combining flashed-binary cycle, KC, ARC, and electrolyzer to reduce carbon emissions while
increasing thermal efficiency from 68.73 % to 70.08 %. Azariyan et al. [27] conducted an exergoeconomic analysis of a tri-generation
system producing power, cooling and hydrogen from geothermal energy. This involved utilizing the heat exiting a separator in the KC
to act as a thermal source for an ARC alongside some of its output electricity providing input into PEME electrolysis for hydrogen
production. As demonstrated by these findings, this unique combination yielded notable gains in system efficiency compared with
other equivalents at 3 %.
Hydrogen is a promising green energy source, especially in multi-generation systems, aligning with global efforts for reduced
carbon footprints and sustainable energy. Despite its benefits, challenges like limited natural availability and emissions from con­
ventional methane-based production exist. Current hydrogen production processes raise environmental concerns, urging exploration
of cleaner alternatives [30]. Green energy sources, particularly nuclear-based sCO2 power cycles, offer a more eco-friendly approach.
While techniques like water electrolysis, photo-catalysis, and thermochemical processes are employed for hydrogen generation,
electrolysis emerges as the most efficient method for large-scale industrial applications. Notably, the use of environmentally sus­
tainable energy sources for hydrogen production, like nuclear-based sCO2 power cycles, is underexplored in existing literature. In their
research, Abid et al. [31] conducted an investigation aimed at enhancing the efficiency of solar collectors and amplifying hydrogen
generation. By integrating a parabolic dish solar collector into a SCRBC, they achieved impressive outcomes. Notably, the system
incorporating reheat exhibited a notable net power output of 3177 kW, a significant improvement over the 1800 kW output of the
system lacking reheat. Furthermore, the system’s overall efficiencies were notably elevated with the reheat configuration, positioning
it as a promising choice for efficient hydrogen production. Toker et al. [32] designed a distinctive solar-assisted energy system for
hydrogen generation, electricity production, refrigeration, and hot water supply. Their system includes a solar dish, a sCO2 power
cycle, PEM electrolyzer, ammonia-based ERC system, and hot water preparation. Their analyses evaluated performance, efficiency,
and environmental impact, while parametric assessments gauged outputs and energy loss. Notably, the system yields 0.062 g/s of
hydrogen (0.223 kg/h) and generates 74.86 kW of power using the supercritical closed Brayton cycle. However, the plant’s overall
irreversibility reaches 535.7 kW, primarily due to the solar system contributing 365.5 kW to energy loss. Hadelu et al. [33] introduced
an innovative solar-powered system that seamlessly combines a supercritical carbon dioxide ejector refrigeration cycle, a solar
desalination unit, and a solid oxide steam electrolyzer, all integrated with parabolic dish collectors. Their study extensively assessed
diverse high-temperature phase change materials, revealing Cu–Si as the frontrunner in terms of thermodynamic performance, while
NaF–CaF2 MgF2 displayed superior economic and sustainability attributes, characterized by lower carbon and water footprints.
While the existing literature has emphasized the application of nuclear-based supercritical carbon dioxide (sCO2) power cycles in
combined cooling and power (CCP) systems, our research introduces a distinctive and innovative approach by combining the su­
percritical CO2 power cycle with specific combined cycles, notably the Kalina cycle (KC) and ammonia-water refrigeration cycle (ARC)
system (KC-ARC). Our research introduces a unique system integration by combining the sCO2 power cycle with the KC-ARC system,
presenting a novel configuration that hasn’t been extensively explored in the literature. In particular, the integration of nuclear-based
sCO2 cycles for producing power, cooling, and hydrogen has received limited attention despite the growing significance of sustainable
energy solutions. Green hydrogen, a clean energy carrier produced through water electrolysis using sustainable sources, is a focal point
of our study. The efficient conversion of high-temperature heat from nuclear reactors into electricity by nuclear-based sCO2 cycles
facilitates stable and environmentally friendly hydrogen production. Our research addresses a critical research gap by scrutinizing the
feasibility, challenges, and advantages of this tri-generation concept. This integration not only contributes to establishing a low-carbon
economy and enhancing energy security but also tackles challenges related to the intermittency of renewable energy sources. Through
a comprehensive analysis encompassing thermodynamic, exergoeconomic, and comparative aspects, our study aims to evaluate the
proposed nuclear-based tri-generation system’s performance, emphasizing its unique integration and potential applications for sus­
tainable energy production.

2. Description of the system and assumptions


In this section, an overview of the structure and operational processes of our proposed tri-generation system is presented. Addi­
tionally, the essential assumptions that are vital for the system to function effectively is outlined.

5
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

2.1. Overview of the proposed trigeneration system


Fig. 1 presents the schematic diagram of the SCRBC/KC-ARC-PEME system introduced in this study. This integrated system consists
of two primary subsystems: the SCRBC subsystem, serving as the topping cycle, and the KC-ARC-PEME subsystem, serving as the
bottoming cycle. The bottoming cycle is composed of three distinct subsystems: the Kalina cycle (KC), the water-ammonia absorption
refrigeration cycle (ARC), and the proton exchange membrane electrolyzer (PEME). The topping SCRBC subsystem comprises essential
components, including the main compressor (MC), recompression compressor (RC), reactor, sCO2 turbine, low-temperature recu­
perator (LTR), high-temperature recuperator (HTR), and pre-cooler (PC). Conversely, the KC-ARC-PEME subsystem is formed by
integrating the KC subsystem, which includes the vapor generator 1 (VG1), separator (SEP), Kalina turbine (KT), mixer (MIX),
condenser1 (COND1), expansion valve (E.V1), pump1, and heat exchanger1 (HEX1). Additionally, the ARC subsystem features vapor
generator 2 (VG2), solution heat exchanger (SHX), pump 2, expansion valve 2 (E.V2), expansion valve 3 (E.V3), condenser2 (COND2),
evaporator (EVAP), and absorber (ABS). Lastly, the PEME subsystem consists of heat exchanger2 (HEX2) and the PEM unit. This
integrated system efficiently utilizes energy across these interconnected subsystems to optimize performance.
The stream of CO2, which is initially heated within the reactor (state 2), serves as the driving force for the sCO2 turbine, generating
power. During the expansion phase, state 3, the gas liberates heat orderly in both the HTR and LTR. Upon reaching state 5, the CO2
stream divides into two branches: Stream 5a and Stream 5 b. Stream 5a proceeds through VG1 (state 6), transferring heat to the KC,
subsequently releasing heat in HEX2 (state 7) and PC (state 8). The resulting fluid is then pressurized by the MC at state 9. Meanwhile,
the high-pressure fluid from the RC at state 10 b combines with the high-pressure fluid, proceeding to the LTR (state 10a) for ab­
sorption. The resulting composition of CO2, designated as state 10, undergoes reheating in the HTR before returning to the reactor for
additional heat absorption.
In the KC subsystem, VG1 receives a stream (designated as 5a) consisting of CO2 at a specific temperature and pressure. This stream
contributes heat to the KC. Another stream (designated as 13), which is a vapor-saturated mixture of water and ammonia, flows
through VG1 and reaches the separator. At this point, it separates into two streams: rich ammonia-water saturated vapor (stream 14)
and poor ammonia-water saturated liquid (stream 20). The Kalina turbine expands the saturated vapor, resulting in stream 15 at a
lower pressure. This expansion generates power, which is then fed into the MIX. On the other hand, VG2 in the water-ammonia ARC
receives the poor ammonia-water saturated liquid from the separator, providing the necessary heat input for the ARC cycle to produce
the required refrigeration. The output flow from VG2 is directed towards HEX1 (stream 21) to release some thermal energy before
passing through E.V1, which is at state 22. The two-phase flow observed at state 23 combines with the Kalina turbine exit stream and
enters COND1 at state 16. Subsequently, the saturated flow is pumped at state 18 and reheated using HEX1. Finally, this reheated
mixture returns to VG1 at state 19.
The ARC, which employs a water-ammonia absorption process with ammonia as the solute, follows a specific sequence of steps.
Initially, the rich ammonia solution is vaporized and heated within VG2 using the saturated liquid from the SEP. The resulting
condensed ammonia (stream 24), now in a high-pressure liquid state (stream 25), proceeds through E.V2, where it undergoes evap­
oration (stream 26). The refrigeration cycle continues as the low-pressure ammonia vapor is absorbed in the ABS, transforming into a
rich water-saturated ammonia solution. This rich solution is then circulated by solution pump 2 as streams (28–30), while the less
concentrated ammonia solution outside the system returns to the ABS through streams (31–33). The system’s operation is further
enhanced by an additional heat source, which serves multiple purposes. It provides the necessary power through the KC’s turbine,
produces hydrogen (stream 44) and oxygen (stream 45), and supplies heat to the inlet water of the electrolyzer. This heat transfer is
facilitated by HEX2 and the output CO2 stream from VG2 (stream 6).

Fig. 1. A schematic diagram of the SCRBC/KC-ARC-PEME system.

6
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

2.2. The system assumptions


To assess the performance of the proposed trigeneration system, we’ve created an appropriate computer program utilizing the EES
software [34] and established the following assumptions:
• The SCRBC/KC-ARC-PEME system functions under steady-state conditions.
• The exiting fluids from the generator, absorber, evaporator and condenser are to be taken as being in a saturated state.
• Devices such as pumps, compressors, and turbines are regulated by well-defined isentropic rates to maintain their operating
efficiency.
• Owing to their inconsequential magnitudes, changes in kinetic and potential energy within the system are disregarded.

3. Mathematical frameworks and metrics for assessing performance


This section offers detailed insights into the mathematical models used to assess the energy, exergy, and economic performance of
the SCRBC/KC-ARC-PEME system. Additionally, specific performance parameters are employed to evaluate the system’s effectiveness.

3.1. Thermodynamic analysis


The thermodynamic models for the SCRBC/KC-ARC-PEME system have been formulated based on equations that uphold the
principles of mass and energy conservation. These equations are outlined below [21]:
∑ ∑
ṁin = ṁout (1)
∑ ∑ ∑ ∑
Q̇− Ẇ= ṁout hout − ṁin hin (2)

Table 1 presents the thermodynamic balance equations corresponding to each component within the SCRBC/KC-ARC-PEME sys­
tem. Solving these equations provides valuable thermodynamic data for each state point:
Table 2 presents the exergy balance equations corresponding to each component in the proposed SCRBC/KC-ARC-PEME system.
The overall exergy (E) can be decomposed into its physical exergy (Eph) and chemical exergy (Ech), disregarding any kinetic or po­
tential energy variations in the working fluid [35]. Chemical exergy may be neglected when the concentration of the working medium
remains constant, as is the case with CO2 in the topping subsystem. However, in scenarios where there is a change in concentration
during the process, such as with the NH3–H2O solution in the KC subsystem, the chemical exergy must be taken into account. The
equations for calculating physical exergy, chemical exergy, and total exergy are provided below [21]:

Table 1
Thermodynamic balance equations for SCRBC/KC-ARC-PEME system components.

Component Equation

Reactor Q̇R = ṁCO2 (h2 − h1 )


MC ẆMC = ṁCO2 (1 − SR)(h9 − h8 ), h9 = h8 + (h9s − h8 )/ηMC
RC ẆRC = ṁCO2 SR(h10 − h5 ), h10 = h5 + (h10s − h5 )/ηRC
HTR Q̇HTR = ṁCO2 (h3 − h4 ) = ṁCO2 (h1 − h10 )
LTR Q̇LTR = ṁCO2 (h4 − h5 ) = ṁCO2 (1 − SR)(h10 − h9 )
sCO2 turbine ẆTUR1 = ṁCO2 (h2 − h3 ), h3 = h2 − (h2 − h3s )ηTUR1
Kalina turbine ẆTUR2 = ṁCO2 (h14 − h15 ), h14 = h14 − (h14 − h15s )ηTUR2
Pre-cooler Q̇PC = ṁCO2 (1 − SR)(h7 − h8 ) = ṁPC (h12 − h11 )
Vapor generator 1 Q̇GEN1 = ṁCO2 (1 − SR)(h5a − h6 ) = ṁ13 h13 − ṁ19 h19
TTDGen1 = T5a − T13
Vapor generator 2 ṁ30 = ṁ31 + ṁ24 , ṁ30 X30 = ṁ31 X31 + m24 ,
Q̇GEN1 = ṁ20 (h20 − h21 ) = ṁ24 h24 + ṁ31 h31 − ṁ30 h30
ΔThot,Gen2 = T20 − T31 , ΔTcold,Gen2 = T21 − T30
Pump1 ẆPUMP1 = ṁ17 (h18 − h17 ), ηPUMP1 = (h18s − h17 )/(h18 − h17 )
Pump2 ẆPUMP2 = ṁ28 (h29 − h28 ), ηPUMP2 = (h29s − h28 )/(h29 − h28 )
SHX Q̇SHX = ṁ29 (h30 − h29 ) = ṁ31 (h31 − h32 ),
εSHX = (T31 − T32 )/(T31 − T29 )
Absorber Q̇ABS = ṁ27 h27 + ṁ33 h33 − ṁ28 h28 = ṁ36 (h37 − h36 )
E.V1 h22 = h23
E.V2 h32 = h33
E.V3 h25 = h26
Condenser1 Q̇COND1 = ṁ16 (h16 − h17 ) = ṁ34 (h35 − h34 )
Condenser2 Q̇COND2 = ṁ24 (h24 − h25 ) = ṁ40 (h41 − h40 )
Separator ṁ13 = ṁ14 + ṁ20 , ṁ13 h13 = ṁ14 h14 + ṁ20 h20
Evaporator Q̇EVAP = ṁ26 (h27 − h26 ) = ṁ38 (h39 − h38 )
Mixer ṁ15 + ṁ23 = ṁ16 , ṁ15 h15 + ṁ23 h23 = ṁ16 h16
Heat Exchanger 1 Q̇HEX1 = ṁ18 (h19 − h18 ) = ṁ21 (h21 − h20 )
Heat Exchanger 2 Q̇HEX2 = ṁ6 (h6 − h7 ) = ṁ42 (h43 − h42 )
PEM electrolyzer See electrolyzer modeling (see section 3.2)

7
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Table 2
Exergy balance equations and the definition of fuel exergy and production exergy for each component within the SCRBC/KC-ARC-PEME system.

Component Exergy balance equations Fuel exergy Product exergy


) )
Reactor ( To ( To Ė2
Ė1 + QR · 1 − = Ė2 + Ėd,R Ė1 + QR · 1 −
TR TR
MC Ė8 + ẆMC = Ė9 + Ėd,MC ẆMC Ė9 − Ė8
RC Ė5b + ẆRC = Ė10b + Ėd,RC ẆRC Ė10b − Ė5b
HTR Ė3 + Ė10 = Ė1 + Ė4 + Ėd,HTR Ė3 − Ė4 Ė1 − Ė10
LTR Ė4 + Ė9 = Ė5 + Ė10a + Ėd,LTR Ė4 − Ė5 Ė10a − Ė9
sCO2 turbine Ė2 = Ė3 + ẆTUR1 + Ėd,TUR1 Ė2 − Ė3 ẆTUR1
Kalina turbine Ė14 = Ė15 + ẆTUR2 + Ėd,TUR2 Ė14 − Ė15 ẆTUR2
Pre-cooler Ė7 + Ė11 = Ė8 + Ė12 + Ėd,PC Ė7 − Ė8 Ė12 − Ė11
Vapor generator 1 Ė5a + Ė19 = Ė6 + Ė13 + Ėd,GEN1 Ė5a − Ė6 Ė13 − Ė19
Vapor generator 2 Ė20 + Ė30 = Ė21 + Ė24 + Ė31 + Ėd,GEN2 Ė20 − Ė21 Ė24 + Ė31 − Ė30
Pump1 Ė17 + ẆPUMP1 = Ė18 + Ėd,PUMP1 ẆPUMP1 Ė18 − Ė17
Pump2 Ė28 + ẆPUMP2 = Ė29 + Ėd,PUMP2 ẆPUMP2 Ė29 − Ė28
SHX Ė29 + Ė31 = Ė30 + Ė32 + Ėd,SHX Ė31 − Ė32 Ė30 − Ė29
Absorber Ė27 + Ė33 + Ė36 = Ė28 + Ė37 + Ėd,ABS Ė27 + Ė33 − Ė28 Ė37 − Ė36
E.V1 Ė22 = Ė23 + Ėd,EV1 Ė22 Ė23
E.V2 Ė32 = Ė33 + Ėd,EV2 Ė32 Ė33
E.V3 Ė25 = Ė26 + Ėd,EV2 Ė25 Ė26
Condenser1 Ė16 + Ė34 = Ė17 + Ė35 + Ėd,COND1 Ė16 − Ė17 Ė35 − Ė34
Condenser2 Ė24 + Ė40 = Ė25 + Ė41 + Ėd,COND2 Ė24 − Ė25 Ė41 − Ė40
Separator Ė13 = Ė14 + Ė20 + Ėd,SEP Ė13 Ė14 + Ė20
Evaporator Ė26 + Ė38 = Ė27 + Ė39 + Ėd,EVAP Ė26 − Ė27 Ė39 − Ė38
Mixer Ė15 + Ė23 = Ė16 + Ėd,MIX Ė15 + Ė23 Ė16
Heat Exchanger 1 Ė21 + Ė18 = Ė19 + Ė22 + Ėd,HEX1 Ė21 − Ė22 Ė19 − Ė18
Heat Exchanger 2 Ė42 + Ė6 = Ė7 + Ė43 + Ėd,HEX2 Ė6 − Ė7 Ė43 − Ė42
PEM electrolyzer Ė43 + ẆPEM = Ė44 + Ė45 + Ėd,PEM Ė43 + ẆPEM Ė44 + Ė45

Ė = Ėph + Ėch (3)


[
Ėph = ṁ (h − ho ) − To (s − so ) (4)
[( ) ( ) ]
X 1− X 0
Ėch = ṁ e0ch,NH3 + ech,H2 O (5)
MNH3 MH2 O

where e0ch,NH3 and e0ch,H2 O represent the standard chemical exergies of ammonia and water, respectively, with their values sourced from
Ref. [36].

3.2. Electrolyzer model


To develop a mathematical model of a PEM electrolyzer, we need to consider various factors such as hydrogen generation, power
requirements, and activation potentials. The potential difference between the two electrodes in an electrolyzer cell drives the water
splitting reaction, which results in the production of hydrogen and oxygen. The reactions associated with the anode, cathode, and the
overall process can be represented as follows [37]:

Anode : H2 O→2H + +0.5O2 +2e− (6)

Cathode : 2H + +2e− →H2 (7)

Overall : H2 O + energy→H2 +0.5O2 (8)


Hydrogen production rate (mH2) can be determined using Faraday’s law of electrolysis, which establishes a direct proportionality
between the quantity of substance produced at an electrode and the amount of electricity that passes through it. The equation for the
rate of hydrogen production is [37]:
J
ṁH2 = (9)
2F
Here, F is the Faraday constant and J is the current flowing through the electrolyzer.
The electrical power consumed by the electrolyzer can be calculated as the product of the voltage applied (V) and the current (J)
[37]:

8
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Ẇ PEME = Ẇ KT /2 =J × V (10)
It is noteworthy to mention that in the current study, half of the power output of the KT is utilized to power the electrolyzer and the
electrolyzer voltage V is expressed as follows [37]:
V = Vohm + Vact,a + Vact,c (11)

Here, V0 represents the reversible potential, Vact,a signifies the activation potential on the anode side, Vact,c refers to the activation
potential on the cathode side, and Vohm denotes the ohmic potential.
The V0 represents the thermodynamic potential of the hydrogen generation reaction and is expressed as [37]:

V0 = 1.229 − 8.5 × 10− 4 (TPEME − 298) (12)


The Vact,a and Vact,c can be expressed as follows:
( )
R × TPEME J
Vact,i = sinh− 1 , i = a, c (13)
F 2 × J0,i

Here, J0,i represents the exchange current density of the electrolyzer and is determined as follows [37]:
( )
Eact,i
J0,i = Jiref × exp − , i = a, c (14)
R × TPEME
Also, the ohmic potential is expressed as follows [37]:
Vohm = J × RPEME (15)
Here, RPEME accounts for the resistive losses due to the ionic conductivity of the electrolyte and the electrical resistance of the
membrane and can be expressed as [37]:
∫L
dx
RPEME = (16)
σPEME [λ(x)]
0

The local unique coefficient of conductivity, denoted as σPEME[λ(x)], signifies the ionic conductivity of the PEM electrolyte at a
given position x. The water content at that specific position is represented as λ(x). The formulas for calculating both parameters can be
expressed as [37]:
[ ( )]
1 1
σPEME [λ(x)] = [0.5139λ(x)− 0.326] 1268 − (17)
303 TPEME

λa − λc
λ(x) = x + λc (18)
D

3.3. Exergoeconomic analysis


Exergoeconomic analysis, widely used in engineering and energy systems, assesses both economic and thermodynamic perfor­
mance simultaneously. It combines exergy analysis, which evaluates energy quality and losses, with economic analysis, examining
resource and component costs. This approach examines exergy destruction and costs, revealing inefficiencies and economic viability in
system components. It aids in identifying areas for improvement, optimizing design, and cost-effectively allocating resources [38].
While investing more capital can boost thermodynamic performance in energy conversion systems, it’s crucial to balance capital costs
with efficiency gains [39]. In this section, we introduce exergoeconomic models to evaluate the SCRBC/KC-ARC-PEME system’s
economic efficiency, calculating its cost per unit of exergy to gauge both economic and thermodynamic performance.
In the previous investigation, the determination of energy and exergy parameters is carried out for each state point. Moreover, for
the conduct of an exergoeconomic analysis, it is imperative to specify the exergy of fuel and production for each component within the
suggested system. This information is outlined in Table 2. Furthermore, the establishment of individual cost equations is crucial to
maintain the cost balance among diverse system components. It guarantees that the rate of cost for outgoing exergy streams matches
the total cost of incoming exergy streams. This accounts for capital investment, maintenance, operating expenses, and is depicted
below [39]:
∑ ∑
Ċout,k + Ċw,k = Ċin,k + Ċq,k + Ż k (19)

Where,

Ċ = cĖ (20)

Where, c = the cost of per unit exergy stream [39].


CI
The cost rate for the kth component, denoted by (Żk ) in Eq. (4), comprises the yearly levelized capital investment rate (Żk ), and the

9
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

OM
yearly levelized maintenance and operation cost rate (Żk ), as given below [39].

(21)
CI OM
Ż k = Ż k + Ż k

The capital investment rate of the kth component is given by Ref. [39]:
( )
CRF
(22)
CI
Ż k = Zk
τ
The Capital Recovery Factor (CRF) is a parameter that allows us to assess the impact of the interest rate, where τ represents the total
annual hours of operation for plants within a year. In this study, τ is set to 8000 h. CRF is defined as follows [8]:

ir (1 + ir )n
CRF = (23)
(1 + ir )n − 1
The given equation factors in a 12 % interest rate applied over a 20-year period, aiming to evaluate its influence on capital
investments.
The maintenance and operation cost rate for the kth component is given by Ref. [8]:

(23a)
OM
Ż k = γZk/τ

Given the fixed operational and maintenance costs, a weighting coefficient γ of 0.1 is applied [8].
This study entails the formulation of a set of cost balances and auxiliary equations for the proposed SCRBC/KC-ARC-PEME system
components. By employing EES software, it becomes feasible to ascertain the cost rates of 45 exergy streams. These equations are
elaborated upon and listed in Table 3.
Table 4 presents the empirical formulas used to estimate investment costs for turbomachinery, heat exchangers, and reactor [40].
However, it is crucial to account for the present-day costs as these formulas were developed in the past. To achieve this, the cost rate
calculated using the equations in Table 4 should be multiplied by the ratio of the current year’s CEPCI (Chemical Engineering Plant
Cost Index) to that of the year when the formula was initially derived. This adjustment ensures more accurate cost estimations for
contemporary projects [21]. For reference, the CEPCI for 2022 averaged at 816.0 [41].
Cost in present year = Cost in original year ∗ [(Cost index for present year) / (Cost index for original year)]
The calculation of the heat exchanger area is a vital component in the thermo-economic analysis. In this particular investigation, we
determined the corresponding heat transfer area for each heat exchanger by employing the following equation [40]:

Table 3
Cost balance and auxiliary equations for SCRBC/KC-ARC-PEME system components.

Component Exergy cost rate balance equation Auxiliary equation

Reactor Ċ2 = Ċfuel + Ċ1 + ŻR nil


MC Ċ9 = Ċ8 + ĊW,MC + ŻMC ĊW,MC /ẆMC = ĊW,TUR1 /ẆTUR1
RC Ċ10b = Ċ5b + ĊW,RC + ŻRC Ċ5b = SR × Ċ5 , ĊW,RC /ẆRC = ĊW,TUR1 /ẆTUR1
HTR Ċ1 + Ċ4 = Ċ3 + Ċ10 + ŻHTR Ċ3 /Ė3 = Ċ4 /Ė4
LTR Ċ5 + Ċ10a = Ċ4 + Ċ9 + ŻLTR Ċ10 = Ċ10a + Ċ10b , Ċ4 /Ė4 = Ċ5 /Ė5
sCO2 turbine Ċ3 + ĊW,TUR1 = Ċ2 + ŻTUR1 Ċ2 /Ė2 = Ċ3 /Ė3
Kalina turbine Ċ15 + ĊW,TUR2 = Ċ14 + ŻTUR2 Ċ14 /Ė14 = Ċ15 /Ė15
Pre-cooler Ċ8 + Ċ12 = Ċ7 + Ċ11 + ŻPC Ċ7 /Ė7 = Ċ8 /Ė8 , Ċ11 = 0
Vapor generator 1 Ċ6 + Ċ13 = Ċ5a + Ċ19 + ŻGEN1 Ċ5a /Ė5a = Ċ6 /Ė6 , Ċ5a = (1 − SR) × Ċ5
Vapor generator 2 Ċ21 + Ċ31 + Ċ24 = Ċ20 + Ċ30 + ŻGEN2 (Ċ24 − Ċ30 )/(Ė24 − Ė30 ) = (Ċ31 − Ċ30 )/(Ė31 − Ė30 )
Ċ20 /Ė20 = Ċ21 /Ė21
Pump1 Ċ18 = Ċ17 + ĊW,PUMP1 + ŻPUMP1 ĊW,PUMP1 /ẆPUMP1 = ĊW,TUR2 /ẆTUR2
Pump2 Ċ29 = Ċ28 + ĊW,PUMP2 + ŻPUMP2 ĊW,PUMP2 /ẆPUMP2 = ĊW,TUR2 /ẆTUR2
SHX Ċ30 + Ċ32 = Ċ29 + Ċ31 + ŻSHX Ċ31 /Ė31 = Ċ32 /Ė32
Absorber Ċ28 + Ċ37 = Ċ27 + Ċ33 + Ċ36 + ŻABS Ċ28 /Ė28 = (Ċ27 + Ċ33 )/(Ė27 + Ė33 ), Ċ36 = 0
E.V1 Ċ23 = Ċ22 + ŻEV1 ŻEV1 = 0
E.V2 Ċ33 = Ċ32 + ŻEV2 ŻEV2 = 0
E.V3 Ċ26 = Ċ25 + ŻEV3 ŻEV3 = 0
Condenser1 Ċ17 + Ċ35 = Ċ16 + Ċ34 + ŻCOND1 Ċ16 /Ė16 = Ċ17 /Ė17 , Ċ34 = 0
Condenser2 Ċ25 + Ċ41 = Ċ24 + Ċ40 + ŻCOND2 Ċ24 /Ė24 = Ċ25 /Ė25 , Ċ40 = 0
Separator Ċ14 + Ċ20 = Ċ13 + ŻSEP Ċ14 /Ė14 = Ċ20 /Ė20 , ŻSEP = 0
Evaporator Ċ27 + Ċ39 = Ċ38 + Ċ26 + ŻEVAP Ċ26 /Ė26 = Ċ27 /Ė27 , Ċ38 /Ė38 = Ċ39 /Ė39
Mixer Ċ16 = Ċ15 + Ċ23 + ŻMIX ŻMIX = 0
Heat Exchanger 1 Ċ19 + Ċ22 = Ċ18 + Ċ21 + ŻHEX1 Ċ21 /Ė21 = Ċ22 /Ė22
Heat Exchanger 2 Ċ7 + Ċ43 = Ċ6 + Ċ42 + ŻHEX2 Ċ6 /Ė6 = Ċ7 /Ė7 , Ċ42 = 0
PEM electrolyzer Ċ44 + Ċ45 = Ċ43 + ĊW,PEM + ŻPEM ĊW,PEM /ẆPEM = ĊW,TUR2 /ẆTUR2 , (Ċ44 − Ċ43 )/(Ė44 − Ė43 ) = (Ċ45 − Ċ43 )/(Ė45 − Ė43 )

10
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Table 4
Functions for valuing investment costs of the components [40].

Component Cost function CEPCI

Reactor ZR = C1 × Q̇R , C1 = 283$/kWth 402


Compressors (71.1m ) 368.1
CO2
ZMC,RC = × PRc × ln(PRc )
0.93 − ηc
sCO2 turbine (479.3m ) 368.1
CO2 54.4)
ZT1 = × ln(PRc ) × (1 + e(0.036∗Tin − )
0.93 − ηT
Kalina turbine ZT2 = 4405 ×
0.7
ẆKT 368.1
HTR, LTR ZHTR,LTR = 2681 × A0.59
k
318.4
Pre-cooler ZPr = 2143 × A0.514
k
318.4
)
Absorber, SHX, generator, condenser, and evaporator ( AK 0.6 394.1
Zk = Zref,k ×
100
Pumps ZPUMP = 1120 × W0.8PUMP
468.2

Qk
Ak = (24)
Uk · LTMDk
The logarithmic mean temperature difference, denoted as LTMDk, can be determined using the overall heat transfer coefficient, Uk,
obtained from relevant literature sources [40]:

3.4. Metrics for assessing performance


In this study, an extensive evaluation is carried out to analyze the thermodynamic and economic performance of the SCRBC/KC-
ARC-PEME system using various performance metrics. The quantification of the system’s effectiveness is crucial, and therefore, energy
efficiency (ηen), exergy efficiency (ηex), and total product unit cost (cP,total) are considered. Below, we present the expressions for these
criteria, which will be employed in our analysis [42]:

Ẇ net + Q̇EVAP + ṁH2 LHVH2


ηen = (25)
Q̇R

where, QEVAP is the cooling load, mH2 is the hydrogen production rate, and Wnet is the net power output which can be expressed as:

Ẇ net = Ẇ TUR1 + Ẇ TUR2 − Ẇ MC − Ẇ RC − Ẇ Pump1 − Ẇ Pump2 − Ẇ PEME (26)

The second law of thermodynamics serves as the foundation for assessing a system’s performance, particularly through exergetic
efficiency. The exergy input into the SCRBC/KC-ARC-PEME system can be mathematically represented as follows [43]:
( )
To
ĖR = Q̇R 1 − (27)
TR
The exergetic efficiency of the SCRBC/KC-ARC-PEME system can be expressed as:
( )
Ẇ net + Q̇EVAP 1 − TT26o + Ė44
ηex = (28)
ĖR
The total product unit cost is defined as:

Ċw,net + Ċ39 + Ċ44


cP,tot = (29)
Ẇ net + Ė39 + Ė44

3.5. Model validation


The proposed trigeneration system encompasses the SCRBC, KC, water-ammonia ARC, and PEM electrolyzer subsystems. To ensure
the reliability and relevance of this research, it is imperative to validate the mathematical models employed for analyzing the per­
formance of the proposed SCRBC/KC-ARC-PEME system. Hence, validation is conducted for each of the aforementioned subsystems to
establish the accuracy of the modeling. The validation study commences with assessing the topping SCRBC cycle, where the obtained
results are compared against data from published literature, as presented in Table 5 [8]. Similarly, for the validation of the KC, the

Table 5
Comparison between current results and those of Ref. [8] for the SCRBC cycle.

Parameter Previous Work [8] Present work Relative error (%)

Net produced power (kW) 248.84 248.89 0.02


Thermal efficiency (%) 41.47 41.48 0.026
Exergy efficiency (%) 57.48 57.53 0.086

11
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

results of this study are compared with the data from the work of Hettiarachchi et al. [44] under varying generator pressures, as shown
in Table 6. Moreover, a validation is performed for the ammonia-water ARC, where the attained results are juxtaposed with data from
published literature, as displayed in Table 7 [37]. The high precision of the validation process is affirmed by the comparison of the
calculated thermodynamic parameters listed in Tables 5–7 with the findings of other researchers.
Furthermore, Fig. 2 demonstrates the results of the PEM electrolyzer validation, revealing a highly accurate voltage output in
comparison with the work of Ioroi et al. [46].

4. Results and discussion


This section discusses the findings of our study, which primarily evaluates the SCRBC/KC-ARC-PEME system’s energy, exergy, and
exergoeconomic performance. Our approach involves a parametric study to systematically analyze the impact of various variables on
the system’s thermodynamics and exergoeconomics. The insights gained from this analysis guide our parameter optimization efforts,
achieved through optimization techniques. Finally, we conduct a comprehensive comparative evaluation, comparing the proposed
trigeneration system’s performance with the conventional SCRBC system, highlighting its unique advantages.

4.1. Thermodynamic and exergoeconomic performance


To holistically assess the efficiency of the newly introduced SCRBC/KC-ARC-PEME system, a rigorous analysis was conducted using
EES software. The input parameters employed for this study are briefly summarized in Table 8. Furthermore, a detailed representation
of the properties associated with each stream within the proposed system is meticulously presented in Table 9. This comprehensive
overview encompasses various thermodynamic and thermoeconomic aspects, including pressure, temperature, mass flow rate,
enthalpy, entropy, ammonia mass fraction, total exergy rate, cost rate, and cost per exergy unit. These detailed data points are crucial
for a thorough assessment of the overall system efficiency.
Table 10 provides a comprehensive comparison between the conventional SCRBC system and the novel SCRBC/KC-ARC-PEME
system, evaluated through energy, exergy, and cost approaches. The standalone SCRBC cycle, focused on power generation, yields
a net output of 237.89 MW, accompanied by 362.1 MW of waste heat in the pre-cooler. Notably, the energy efficiency and exergy
efficiency of this system stand at 39.6 % and 54.38 %, respectively. However, with the integration of the KC-ARC-PEME subsystem into
the SCRBC setup, noteworthy improvements are realized. The combined system produces an additional 19.475 MW of cooling load and
an extra 16.134 MW of power output, while concurrently reducing pre-cooler heat transfer to 204.802 MW. It’s pertinent to highlight
that half of the net power output of the KC is allocated for the PEME to facilitate hydrogen production, achieving a mass flow rate of
115.452 kg/h. These enhancements translate into a 14.7 % boost in the energy efficiency of the SCRBC/KC-ARC-PEME system,
elevating it to 45.44 %. Furthermore, there’s an 8.03 % escalation in exergy efficiency, reaching 58.75 %. Importantly, the incor­
poration of the bottoming KC-ARC-PEME subsystem also brings about a significant reduction in the total product unit cost of the
SCRBC system, from $12.272/GJ to $11.675/GJ, a reduction of 5.11 %. However, these outcomes are established under base design
conditions, signifying the potential for further improvements in energy, exergy, and cost performance through optimization tech­
niques, which will be expounded upon in subsequent discussions.
Table 11 provides a comprehensive overview of our analysis concerning the exergy and exergoeconomic aspects of the SCRBC/KC-
ARC-PEME system. To assess the relative significance of each system component in terms of exergoeconomic impact, we employ (Zk +
CD + CL). A higher value of (Zk + CD + CL) indicates a greater influence of that specific component. Furthermore, our study utilizes the
exergoeconomic factor (fk) to pinpoint the primary sources of costs within the system. This factor is calculated as the percentage of non-
exergy-correlated cost rate to the total cost rate [35]. Lower fk values for major components signify potential for broader cost savings
through efficiency enhancements, even if these components require increased capital investment. Conversely, higher fk values for
major components suggest the potential for cost reduction by compromising on component efficiency. The findings reveal that the
reactor and sCO2 turbine exhibit the highest costs of exergy destruction, correlating with their relatively high fuel costs or elevated
exergy destruction rates. As expected, the reactor carries the highest investment cost among all components, with the sCO2 turbine
following closely behind. In contrast, the mixer, E. Vs, and separator have negligible investment costs compared to the other com­
ponents. From an exergoeconomic point of view, the data presented in Table 11 underscores the reactor and sCO2 turbine as the most
pivotal components within the system. This observation is substantiated by their notably high Zk + CD + CL values, surpassing those of
other components. Remarkably, the sCO2 turbine and the reactor exhibit particularly high fk values of 75.7 % and 73.2 %, respectively.
This implies that the capital investment costs linked with these components are the primary contributors to exergy destruction costs.
To alleviate capital investment expenses, there may be a need to balance exergy efficiency compromises for these components.
Conversely, the components within the bottoming KC-ARC-PEME subsystem, excluding pump1 and pump2, demonstrate compara­
tively lower fk values. This suggests that investing more in capital could potentially enhance their exergy efficiencies. Notably, the V.
G1, condenser1, and absorber play pivotal roles within the bottoming KC-ARC-PEME, evident from their high Zk + CD + CL values. A
deeper analysis reveals that the reactor possesses the largest ED,k, thereby indicating that enhancing its efficiency could yield im­
provements in overall system performance.

Table 6
Comparison between current results and those of Ref. [44] for the KC cycle.

Parameter Previous Work [44] Present work Error (%)

Energy efficiency 6.94 % 7.04 % 1.44 %

12
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Table 7
Comparison between current results and those of Ref. [45] for the ARC cycle.

Parameter Previous Work [45] Present work Error (%)

COP 0.616 0.602 2.3

Fig. 2. Validation of Electrolyzer Model Using Ioroi et al.’s Work [46].

Table 8
Basic required input parameters for thermodynamic simulation.

Parameter Value Parameter Value


o o
To ( C) 25 T41 ( C) 32
Po (bar) 1 Tevap (oC) 5
P8 (bar) 74 T17 (oC) 30
PPump2 (bar) 35 T25 (oC) 30
PRc 3 T28 (oC) 30
T8 (oC) 35 TTD,VG1 (oC) 20
T2 (oC) 550 ΔTPPT,VG1 (oC) 10
TR (oC) 800 ΔTVG2,hot (oC) 15
T38 (oC) 12 ΔTVG2,cold (oC) 3
T39 (oC) 7 ηMC&ηRC 0.85
T34 (oC) 25 ηT1&ηT2 0.9
T35 (oC) 32 ϵ HTR & ϵ LTR 0.86
T36 (oC) 25 ϵ SHX 0.82
T37 (oC) 32 X13 (%) 70
T40 (oC) 25 QR (MW) 600

4.2. Parametric study


Parametric studies conducted on tri-generation systems, such as the proposed SCRBC/KC-ARC-PEME system, offer insights into
how crucial parameters influence both the system’s thermal performance and its economic aspects. For assessing the system’s ther­
modynamic efficiency, we’ve chosen energy efficiency (ηen) and exergetic efficiency (ηex). Meanwhile, the economic evaluation
centers around the unit cost of product output (cp,total). In this study, specific key parameters are selected as decision variables to gauge
their impact. These parameters encompass the sCO2 compressor pressure ratio (PRc), the evaporating pressure in the KC (P VG1), the
terminal temperature difference of vapor generator 1 (TTDVG1), the condenser temperature of the KC (T17), the ammonia-water
concentration (X13), the temperature difference at the hot end of vapor generator 2 (ΔTVG2,hot), the condenser temperature of the
ARC subsystem (T25), and the evaporator temperature (T26). The sCO2 compressor inlet pressure is held constant at 7400 kPa to
maintain CO2 in a supercritical state. PRc, linked to the maximum pressure of the system, relates to the ratio of sCO2 compressor outlet
pressure to inlet pressure. PVG1 sets the upper operational pressure of the bottoming KC, playing a pivotal role in heat recovery and
thereby influencing KC performance. TTDVG1 characterizes the difference in temperature between the CO2 inlet and the outlet of the
ammonia-water solution in VG1. This parameter affects the inlet temperature of the Kalina turbine and, consequently, the turbine’s
thermodynamic efficiency. T17 has the potential to influence the backpressure of the KC turbine, thus impacting the Kalina turbine’s
power output. Moreover, X13 significantly affects heat transfer capacity in VG1. Additionally, ΔTVG2,hot, T25, and T26 wield substantial
influence over the performance of the ARC subsystem in terms of refrigeration capacity. Therefore, these variables are designated as
decision parameters. It’s noteworthy that each parameter is evaluated individually, with its counterparts held constant, to ascertain its

13
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Table 9
Thermodynamic properties and costs of exergy streams for the SCRBC/KC-ARC-PEME system.

State Fluid P (bar) T (oC) ṁ (kg/s) h (kJ/kg) s (kJ/kg.K) X (%) Ė (MW) Costs

Ċ ($/h) c ($/GJ)

1 CO2 222 385.4 2943 322.9 − 0.2966 – 1211 47,058 10.57


2 CO2 222 550 2943 526.8 − 0.02022 – 1569 56,483 9.8
3 CO2 74 420.1 2943 384.8 0.002764 – 1131 40,713 9.8
4 CO2 74 282.6 2943 227.7 − 0.2497 – 889.995 32,042 9.8
5 CO2 74 138.3 2943 64.46 − 0.5902 – 708.281 25,500 9.8
5a CO2 74 138.3 2144 64.46 − 0.5902 – 516.112 18,582 9.8
5b CO2 74 138.3 799 64.46 − 0.5902 – 192.169 6919 9.8
6 CO2 74 80.34 2144 − 8.835 − 0.7826 – 481.983 17,353 9.8
7 CO2 74 80.32 2144 − 8.864 − 0.7827 – 481.973 17,352 9.8
8 CO2 74 35 2144 − 104.4 − 1.076 – 464.425 16,721 9.8
9 CO2 222 115.2 2144 − 58.23 − 1.058 – 551.930 21,279 10.495
10 CO2 222 260.4 2943 165.8 − 0.5613 981.278 38,367 10.642
10a CO2 222 260.4 2144 165.8 − 0.5613 – 715.040 27,855 10.6036
10 b CO2 222 260.4 799 165.8 − 0.5613 – 266.238 10,513 10.75
11 Water 1.013 25 2130 104.9 0.3672 – 0 0 0
12 Water 1.013 48 2130 201.1 0.6778 – 7.522 651.5 23.578
13 NH3–H2O 35 118.3 205.9 773.4 2.625 – 2927 131,853 12.259
14 NH3–H2O 35 118.3 81.3 1478 4.369 – 1609 72,458 12.259
15 NH3–H2O 7.718 44.37 81.3 1280 4.439 – 1591 71,655 12.259
16 NH3–H2O 7.718 48.47 205.9 472 2.027 – 2902 130,735 12.259
17 NH3–H2O 7.718 30 205.9 − 55.52 0.3377 – 2897 130,514 12.259
18 NH3–H2O 35 30.51 205.9 − 51.17 0.3398 – 2898 130,571 12.269
19 NH3–H2O 35 43.72 205.9 10.21 0.5377 – 2898 130,618 12.269
20 NH3–H2O 35 118.3 124.6 313.7 1.487 – 1319 59,396 12.259
21 NH3–H2O 35 62.86 124.6 46.64 0.7524 – 1313 59,126 12.259
22 NH3–H2O 35 40.51 124.6 − 54.78 0.4401 – 1312 59,080 12.259
23 NH3–H2O 7.718 40.98 124.6 − 54.78 0.4509 – 1311 59,080 12.269
24 NH3 11.67 103.3 17.3 1686 5.857 – 349.505 23,802 18.541
25 NH3 11.67 30 17.3 341.6 1.488 – 348.783 23,752 18.541
26 NH3 5.16 5 17.3 341.6 1.509 – 348.670 23,752 18.541
27 NH3 5.16 5 17.3 1467 5.556 – 347.268 23,657 18.541
28 NH3–H2O 5.16 30 47.13 − 97.36 0.2943 0.5763 546.256 37,190 18.531
29 NH3–H2O 11.67 30.06 47.13 − 96.54 0.2943 0.5763 546.294 37,192 18.531
30 NH3–H2O 11.67 59.86 47.13 71.71 0.8186 0.5763 546.856 37,268 18.551
31 NH3–H2O 11.67 103.3 29.83 250.9 1.304 0.3306 202.067 13,741 18.512
32 NH3–H2O 11.67 43.24 29.83 − 14.98 0.5357 0.3306 200.974 13,667 18.512
33 NH3–H2O 5.16 43.37 29.83 − 14.98 0.538 0.3306 200.953 13,667 18.512
34 Water 1.013 25 3713 104.9 0.3672 – 0 0 0
35 Water 1.013 32 3713 134.2 0.4642 – 1.313 228.3 47.324
36 Water 1.013 25 1009 104.9 0.3672 – 0 0 0
37 Water 1.013 32 1009 134.2 0.4642 – 0.3569 140.1 106.82
38 Water 1.013 12 928.2 50.51 0.1806 – 1.107 106.1 26.087
39 Water 1.013 7 928.2 29.53 0.1064 – 2.165 207.4 26.087
40 Water 1.013 25 795 104.9 0.3672 – 0 0 0
41 Water 1.013 32 795 134.2 0.4642 – 0.2812 50.96 49.323
42 Water 1.013 25 0.2647 104.9 0.3672 – 0 0 0
43 Water 1.013 80 0.2647 335.1 1.076 – 0.005 0.3966 21.403
44 H2 1.013 80 0.03207 4723 55.86 – 0.00216 187.4 23,567
45 O2 1.013 80 0.2543 50.7 6.56 – 0.00137 238.6 47,248

specific impact on the system’s outputs.

4.2.1. Impact of compressor pressure ratio (PRC)


This section examines how changes in the PRC affect the SCRBC/KC-ARC-PEME system. Notably, the PRC dictates the maximum
system pressure and significantly influences the topping SCRBC subsystem. Thus, variations in the overall system’s performance with
increased PRC are mainly driven by the topping SCRBC subsystem, with contributions from the bottoming KC-ARC-PEME subsystem
influencing these variations. The results have revealed that an increase in the PRC invariably leads to a rise in its turbine work output.
Coupled with this, however, is a decrease of mass flow rate owing to set conditions such as fixed heat input rates and turbine inlet
temperature. As such, power output growth becomes more sluggish over time even though it initially rises at higher PRC. Ultimately,
the net power output reaches peak potential before diminishing due to elevated levels of power consumption of sCO2 compressors
relative to power generation levels of sCO2 turbine (see Fig. 3a).
As the PRc rises, the CO2 stream’s mass flow rate declines. This is due to the interplay of factors where the temperature at the
turbine outlet drops with higher PRC, resulting in lower temperatures at the outlet of the cold side of the HTR. As a result, the increase
in temperature of the CO2 stream as it passes through the reactor increases due to a fixed heat input. Despite reduced CO2 flow through

14
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Table 10
Comparison between a conventional SCRBC system and the SCRBC/APC-PEME system.

Performance parameters SCRBC system SCRBC/KC-ARC-PEME system

Main compressor power (ẆMC ) 98.946 MW 98.946 MW


Recompression compressor power (ẆRC ) 80.927 MW 80.927 MW
sCO2 turbine power (ẆTUR1 ) 417.9 MW 417.9 MW
KC turbine net power (ẆTUR2 ) – 8.067 MW
Pump 1 power (ẆPUMP1 ) – 0.8961 MW
Pump 2 power (ẆPUMP2 ) – 0.0387 MW
Net power output of the cycle (Ẇnet ) 237.89 MW 245.586 MW
Pre-cooler output (Q̇PC ) 362.1 MW 204.802 MW
Cooling output (Q̇EVAP ) – 19.475 MW
Hydrogen production (kg/hr) – 115.452
Energy efficiency (ηen ) 39.6 (%) 45.44 (%)
Exergy efficiency(CP;tot ) 54.38 (%) 58.75 (%)
Total product cost of the system (cP;tot ) 12.272 ($/GJ) 11.675 ($/GJ)

Table 11
Exergy and exergoeconomic findings of the SCRBC/KC-ARC-PEME system for the basic design case.

Component Ẇ or Q̇ (MW) ĖF,k (MW) ĖP,k (MW) ĖD,k (MW) cF,k ($/GJ) cP,k ($/GJ) ĊD,k ($/hr) Żk ($/hr) rk fk

Reactor 600 1645 1569 76 8.32 9.8 2276.35 6219 0.177 0.732
MC 98.946 98.946 87.5 11.44 11.71 14.18 482.26 300.1 0.21 0.383
RC 80.927 80.927 74.07 6.859 11.71 13.21 289.14 111.7 0.128 0.278
HTR 462.296 240.8 230 10.8 9.8 10.29 381.02 19.96 0.05 0.049
LTR 480.416 181.7 163.1 18.6 9.8 10.97 656.2 33.56 0.119 0.048
sCO2 turbine 417.9 438 417.9 20.1 9.8 11.71 709.12 2212 0.194 0.757
Kalina turbine 16.134 17.82 16.134 1.686 12.259 15.24 74.407 100.8 0.243 0.575
Pre-cooler 204.802 17.55 7.522 10.027 9.8 23.57 353.75 19.7 1.405 0.0527
Vapor generator 1 157.151 34.13 28.98 5.146 9.8 11.603 181.55 6.33 0.183 0.033
Vapor generator 2 33.275 5.979 4.716 1.263 12.259 15.86 55.74 5.455 0.293 0.089
Pump1 0.8961 0.8961 0.7641 0.132 15.24 20.1 7.242 6.245 0.318 0.463
Pump2 0.0387 0.0387 0.0386 0.0001 15.24 18.8 0.0054 0.5056 0.233 0.689
SHX 7.929 1.092 0.5619 0.5305 18.51 36.52 35.35 1.1 0.97 0.03
Absorber 29.526 1.966 0.3569 1.609 18.53 106.8 107.33 6.258 4.76 0.055
Condenser1 108.637 4.896 1.313 3.583 12.259 47.3 158.12 7.721 2.86 0.0465
Condenser2 23.263 0.7228 0.2812 0.4416 18.54 49.32 29.47 1.729 1.66 0.055
Separator – 2927 2927 0 12.259 12.259 0 0 0 0
Evaporator 19.475 1.402 1.058 0.3447 18.54 26.08 23.006 5.817 0.406 0.201
Mixer – 2902 2901.9 0.107 12.259 12.259 4.72 0 0 0
Heat Exchanger 1 12.639 1.033 0.4905 0.453 12.259 26.45 19.99 1.115 1.157 0.052
Heat Exchanger 2 0.0609 0.00954 0.005 0.0045 0 21.4 0 0.0532 – –
PEM electrolyzer 8.067 0.005 0.0035 0.0015 15.24 32,767 0.082 0.00015 2149.06 0.0018

Fig. 3. Impacts of PRc on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

the generator due to this, the hotter compressor outlet temperature raises the entering CO2 stream’s temperature into the generator,
elevating the generator’s heat flux. Notably, the impact of rising elevated fluid temperature on the generator’s heat flux outweighs the
influence of decreased hot fluid mass flow rate. This leads to increased waste heat recovery in the KC generator. Consequently, the mass
flow rate of the vapor exiting the separator in a saturated state and the enthalpy drop across the KC turbine both increase, amplifying
KC turbine’s power output. This drives higher net power output for the trigeneration system (Fig. 3a). Additionally, it is worth noting
that the production of hydrogen mass flow (mh2) demonstrates a direct correlation with the power generation of the KC. As depicted in
Fig. 3a, the mH2 exhibits a linear growth as the PRC increases, which aligns with the observed trend in the power generation of the KC as

15
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

a function of PRC. Conversely, heightened PRC leads to a significant reduction in the mass flow rate of saturated liquid exiting the
separator towards the ARC, reducing fluid mass flow through the evaporator and causing decreased refrigeration capacity.
It can be inferred that the net power output of the combined SCRBC/KC-ARC-PEME system rises with an increase in the PRC until
reaching an optimal point where the maximum net power output is achieved. After this point, the net power output starts to decline.
Moreover, the combined SCRBC/KC-ARC-PEME system exhibits superior net power output magnitude and optimum pressure ratio
compared to the standalone SCRBC system. This improvement can be attributed to the escalating additional net power output
generated by the KC subsystem in response to increasing PRC. Hence, with a constant system heat input scenario, the trends of system
energy and exergy efficiencies align with the net power output of the SCRBC/KC-ARC-PEME system, as illustrated in Fig. 3b.
Furthermore, the higher exergy efficiency in our proposed system, compared to its energy efficiency counterpart, arises from the
significant disparity between input energy and input exergy. Unlike energy efficiency, which only considers energy quantity, exergy
efficiency incorporates both quantity and quality of energy, along with system thermodynamic imperfections. This distinction is
apparent in the denominators of the energy efficiency equation (24) and exergy efficiency equation (27). In equation (27), the heat
input is scaled by the exergy factor (1- To/TR), revealing that the exergy input is around 37 % lower than the energy input, resulting in
the observed higher exergy efficiency. Additionally, components operating under higher pressures necessitate greater total investment.
As a result, the total product unit cost decreases as the PRC rises until the impact of the increasing total investment outweighs that of the
escalating net power output, at which point it begins to decline for the SCRBC/KC-ARC-PEME system (see Fig. 3a).

4.2.2. Impact of KC vapor generator pressure (PVG1)


The positive correlation between PVG1 and the level of mass flow rate for the saturated vapor exiting the separator (ṁ14) is evident.
It is established that higher-pressure vapor, when propelled through turbines at a fixed back pressure, exhibits a greater power capacity
relative to the unit working flow rate. The power output from the KC turbine, however, diminishes due to the reduced mass flow rate of
the vapor stream. As a result, the net power output of the KC turbine increases with increasing PVG1 until the impact of expansion ratio
on power output becomes less significant than that of vapor mass flow rate reduction, resulting in a subsequent decline. Optimal power
generation is achieved at the PVG1 pressure of 2400 kPa (see Fig. 4a). Similarly, as PVG1 increases, the mass flow rate of the liquid
exiting the separator in a saturated state (ṁ20) exhibits an ascending trend. This augmentation in mass flow rate contributes to
increased heat input to the refrigeration cycle ARC, consequently leading to a higher cooling load. The relationship between the
production of hydrogen (ṁ H2) and power generation of the KC is evidently correlated. As demonstrated in Fig. 4a, an initial rise in ṁ H2
is observed with increasing PVG1, followed by a subsequent decline. This trend mirrors the pattern seen in power generation of the KC
relative to PVG1.
The constancy of decision variables in the topping SCRBC power cycle contributes to the dependence of net power output and total
product unit cost trends on those of the KC-ARC-PEME subsystems. Linear growth in energy efficiency occurs even at high PVG1, as the
increase in cooling capacity surpasses the decrement in net output power. This continuous enhancement in energy efficiency is
illustrated in Fig. 4b. Moreover, the decrease in power output is greater in magnitude than the increase in refrigeration exergy,
resulting in a similar trend between exergy efficiency of the combined systems and net power output from the KC subsystem as
indicated in Fig. 4b. On the contrary, the total product unit cost demonstrates an inverse variation pattern compared to exergy
efficiency.

4.2.3. Impact of KC generator terminal temperature difference (TTDVG1)


Fig. 5 (a,b) intricately portrays the profound impact of TTDVG1 on the diverse output parameters within the SCRBC coupled with the
KC-ARC-PEME subsystem. TTDVG1, defined as the difference between the LTR exit temperature (T5) and the separator temperature
(Tsep), holds a pivotal role in this context. An escalation in TTDVG1 precipitates a concurrent decrease in Tsep, owing to the consistent
maintenance of T5. This decrement in Tsep engenders a subsequent augmentation in the mass flow rate of saturated vapor discharged
from the separator, accompanied by a decline in the enthalpy difference across the KC turbine. As this augmented vapor mass flow rate
supersedes the enthalpy losses incurred by the KC turbine, it emerges as the dominant factor propelling the enhancement of KC power
output concomitant with heightened TTDVG1. This finding, substantiated by Fig. 5a, is aligned with the fact that the net power output
of the KC-ARC subsystem diminishes as TTDVG1 increases. This observation stems from the marked augmentation in the NH3–H2O mass
flow rate within the cycle, a consequence of elevated TTDVG1, which, in turn, escalates the power consumption of the KC pump. The
prominence of this increase in pump power surpasses the rise in turbine power output, thereby culminating in a discernible reduction

Fig. 4. Impacts of PVG1 on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

16
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Fig. 5. Impact of TTDVG1 on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

in the net power output of the KC-ARC subsystem. Furthermore, an inherent direct correlation exists between mH2 and the net power
output of the KC; hence, an elevation in TTDVG1 corresponds to a diminution in mH2. On the other hand, as the temperature difference
(Tsep) decreases, the quality of the NH3–H2O mixture entering the separator also decreases. This results in a notable increase in the
mass flow rate of the liquid leaving the separator compared to the vapor. Consequently, more thermal energy is added to the ARC
refrigeration cycle, enhancing the cooling capacity, as depicted in Fig. 5a.
Illustrated in Fig. 5b, it becomes evident that with the augmentation of the TTDVG1, there is a concomitant elevation in the energy
and exergy efficiencies observed within the SCRBC/KC-ARC-PEME system. This enhancement is attributed to the substantial increase
in the cooling load, surpassing the relatively minor reduction in output power. Paradoxically, this favorable impact on efficiencies
contrasts with the economic aspect of the SCRBC/KC-ARC-PEME system. The escalating TTDVG1, as visually represented in Fig. 5b,
leads to a corresponding rise in the total product unit cost. This observation underscores the intricate interplay between efficiency
enhancements and the resultant economic implications arising from variations in TTDVG1 within the system.

4.2.4. Impact of condenser temperatures


The impacts of different condenser temperatures on the output parameters of the SCRBC/KC-ARC-PEME system are visually
depicted in Figs. 6 and 7. Fig. 6a illustrates that an increase in the KC’s condenser temperature (T17) leads to a proportional rise in the
specific enthalpy at the turbine outlet. Consequently, this elevation results in a decline in the net output power generated by the KC
turbine. This reduction in generated power leads to a corresponding decrease in the input power supplied to the PEME, subsequently
resulting in a reduction of hydrogen production within the system. Interestingly, the alteration in T17 doesn’t influence the cooling
output of the ARC, thereby making the energy efficiency and exergy efficiency variations reliant solely on the KC’s net power output.
This correlation aligns with the observed declining trends in energy and exergy efficiencies, as portrayed in Fig. 6b, with respect to T17
fluctuations. Consequently, it becomes evident that the system’s cost is primarily influenced by the reduction in the net power cost
rate. As a result, an increase in T17 corresponds to a decrease in the overall cost of the system.
On the other hand, when the condenser temperature is elevated within the ARC subsystem (T25), this instigates an increase in the
output enthalpy of VG2. As a direct consequence, the mass flow rate traversing the evaporator diminishes, consequently resulting in a
reduction of the cooling load, as illustrated in Fig. 7a. However, it’s important to highlight that the alteration in T25 has a marginal
impact on the net power output of the KC. As a result, the oscillations observed in energy efficiency and exergy efficiency are pre­
dominantly influenced by the cooling output of the ARC, thus aligning with the observed downward trends in energy and exergy
efficiencies, showcased in Fig. 7b, in response to variations in T25. Significantly, the product cost is significantly influenced by the
decrease in the cooling cost rate. This intricate relationship signifies that as T25 increases, there is a corresponding decrease in the
overall cost of the system, thereby underscoring the economic implications tied to T25 adjustments.

4.2.5. Impact of vapor generator 2 hot terminal temperature difference (Δ TVG2, hot)
Fig. 8 (a, b) provides a comprehensive insight into the influence of ΔTVG2,hot on the crucial performance parameters characterizing
the SCRB/ARC cycle. Increasing ΔTVG2,hot maintains the separator outlet temperature while enhancing the heat exchange within the
generator, leading to heightened heat absorption by the ARC generator. Concurrently, Tgen (Tgen = T20 - ΔTVG2,hot) decreases alongside

Fig. 6. Impact of T17 on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

17
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Fig. 7. Impact of T25 on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

Fig. 8. Impact of ΔTVG2,hot on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

ΔTVG2,hot, resulting in an augmented refrigerant mass flow rate and subsequent enhancement of refrigeration capacity, as demon­
strated in Fig. 8a. Furthermore, the same Fig. 8a reveals a marginal decline in the net power output (Wnet) of the combined cycle as
ΔTVG2,hot increases, mainly due to a slight elevation in pump work within the ARC. Consequently, this effect contributes to a subtle
reduction in hydrogen production.
As the increase in cooling load adequately compensates for the reduction in net power output with the escalation of ΔTVG2,hot, there
is a noticeable enhancement in energy efficiency, as illustrated in Fig. 8b. Similarly, the trend in exergy efficiency exhibits a marginal
improvement due to the modest surpassing of the increase in cooling exergy over the decrease in net power output. This marginal net
gain in exergy efficiency results in only a minor enhancement, thus accounting for its almost negligible improvement, as reflected in
the figure. Furthermore, as ΔTVG2,hot undergoes variations, the total investment cost exhibits minimal change, and the fuel cost re­
mains constant. Consequently, alterations in the total product cost primarily stem from the combined impact of changes in cooling
exergy and the net power output of the cycle. Specifically, within the range of 10–20 ◦ C for ΔTVG2,hot, the increase in cooling exergy
effectively counterbalances the decrease in net power output as ΔTVG2,hot increases. This harmonic results in a decrease in the total
product cost, as evidenced in Fig. 8b.

4.2.6. Impact of ammonia concentration (X13)


Fig. 9 (a, b) provide a comprehensive depiction of the effects of ammonia concentration (X13) on the evaluation parameters within
the proposed tri-generation system. Fig. 9 (a) distinctly portrays that an elevated X13 translates into enhanced power extraction across
all employed turbines, consequently leading to an upward trajectory in power generation and hydrogen production. This phenomenon
is attributed to the concurrent increase in the mass flow rate of saturated vapor departing from the separator. Although there is a
reduction in enthalpy losses across the KC turbine, this reduction is outweighed by the amplified mass flow rate, resulting in a net
increase in KC’s power output. Conversely, the cooling load exhibits a declining trend due to the drop in mass flow rate of saturated

Fig. 9. Impact of X13 on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

18
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

liquid exiting the separator. Furthermore, the escalation of X13 leads to a considerable decrease in cooling capacity, which outweighs
the amplified power output. Consequently, this interplay fosters an augmentation in energy efficiency, as underscored in Fig. 9b. The
outcomes of the exergy analysis are consistent, revealing that the increase in exergy efficiency correlates with the rise in X13. This
observation stems from the dominance of the increase in net power output over the concurrent decrease in cooling exergy as X13
increases, as shown in Fig. 9b. Moreover, the findings of the cost analysis highlight a reduction in the overall cost of the proposed tri-
generation system, attributable to the ascending trend in power generation.

4.2.7. Impact of evaporator temperature (T26)


Fig. 10(a–b) serves as a visualization of the influence of evaporator temperature on the diverse assessing parameters within the
proposed tri-generation system. The evaporator’s role within the ARC subsystem becomes evident, where elevating the T26 translates
to an augmented cooling load, leaving the power generation of the KC unaffected. The consequence of elevating the T26 becomes
twofold. Primarily, this temperature rise enhances the heat influx to the ARC, thereby elevating the mass flow rate from VG2. This
intricate relationship is illustrated in Fig. 10a, where an amplified cooling output becomes manifest. Simultaneously, the escalation in
T26 leads to a decrease in the mass flow rate of the saturated liquid exiting the water intake. Consequently, pump 2’s power con­
sumption and mass flow rate exhibit reductions, ultimately resulting in marginal increments in power generation and hydrogen
production rates.
As the T26 rises, a parallel increase in thermal efficiency is observed. This outcome stems from the compounding effect, wherein the
augmented evaporator temperature precipitates higher power generation, an increased cooling load, and an elevated hydrogen
production rate. However, the scrutiny of exergy efficiency presents an intriguing revelation. Despite the enhanced cooling load and
power generation, a decline in exergy efficiency is indicated. This phenomenon is rooted in the interplay between the terms Qeva and
cooling exergy, which are integral to exergy efficiency. While Qeva shows an incremental pattern with increasing T26, cooling exergy,
on the contrary, exhibits a negative relationship. In instances where the positive impact of Qeva on cooling exergy supersedes the
negative influence of T26, cooling exergy rises in tandem with T26. However, in this context, the overall effect of T26 is negative, leading
to a reduction in exergy efficiency, as indicated in Fig. 10b. Considering the implications of T26 variations on product cost, it is
observed that as T26 increases, the total investment cost registers a slight increment while maintaining relative constancy. The fuel
cost, however, remains steady. Consequently, the variance in total cost becomes primarily linked to the combined outcome of Wnet and
cooling exergy, both of which exhibit increments, as depicted in Fig. 10b.

4.3. Optimization
The enhancement of both the SCRBC and SCRBC/KC-ARC-PEME systems’ performance hinges on the imperative of optimization.
Our study is dedicated to this optimization endeavor with the twin objectives of maximizing energy efficiency (ηen) and exergy ef­
ficiency (ηex), or alternatively, minimizing the total product cost (cP;tot). This optimization is executed through the application of the
genetic method, a robust technique embedded within the EES software platform. In the pursuit of optimizing the SCRBC system, the
process revolves around incorporating the pressure ratio across the compressor (PRc) as the designated decision parameter. In parallel,
the optimization of the SCRBC/KC-ARC-PEME system encompasses a broader spectrum of parameters. This includes not only PRc, but
also P13, TTDVG1, condenser temperature of the ARC (T25), condenser temperature of the KC (T17), evaporator temperature (T26),
ΔTVG2,hot, and (X13). These parameters are subjected to constraints meticulously detailed in Table 12, capturing the detailed behaviors
of each part. Notably, all other parameters maintain consistency with the comprehensive framework outlined in Table 8.
Table 13 presents a comprehensive overview of the decision variables and objective function values corresponding to three distinct
optimization cases: energy efficiency optimal design (EEOD), exergy efficiency optimal design (XEOD), and cost optimal design (COD)
scenarios for both the SCRBC and SCRBC/KC-ARC-PEME systems. In the case of the SCRBC, it is noteworthy that the optimal pressure
ratio (PRc) for the EEOD and XEOD instances exceeds that of the COD configuration. In comparing the COD situation to the EEOD or
XEOD instances for the SCRBC, a trade-off becomes evident: while the COD configuration demonstrates a modest 1.63 % reduction in
cP;tot, it comes at the cost of a more substantial 1.84 % decrease in both energy and exergy efficiencies. Turning our attention to the
SCRBC/KC-ARC-PEME system, a similar trade-off exists in the COD case against the EEOD or XEOD scenarios. Specifically, the COD
configuration exhibits a greater reduction of about 3.62 % and 4.28 % in cP;tot, leading to concurrent drops of 44.19 % and 6.38 % in
energy efficiency, and 0.77 % and 5.87 % in exergy efficiency, respectively. Comparing the optimization outcomes between the SCRBC
and SCRBC/KC-ARC-PEME systems unveils a consistent trend: the incorporation of the KC-ARC-PEME subsystem within the SCRBC
framework consistently enhances all three objective functions. Highlighting this improvement, Table 13 reveals substantial
enhancement percentages for energy efficiency: approximately 53.56 % for EEOD, 13.3 % for XEOD, and 8.45 % for COD scenarios.
Furthermore, the exergy efficiency demonstrates increments of around 4.46 % for EEOD, 9.76 % for XEOD, and 5.58 % for COD
scenarios. Conversely, a reduction in cP;tot is observed across all three cases, with decrease percentages of approximately 4.59 % for
EEOD, 3.92 % for XEOD, and 6.63 % for COD instances. Remarkably, the SCRBC/KC-ARC-PEME system excels in multiple aspects. It
not only achieves a substantial net power output increase of up to 22.56 MW compared to the SCRBC alone, but it also delivers
remarkable cooling capacity, particularly evident in the EEOD case with a capacity of 123.949 MW. Furthermore, the optimization
process highlights the potential of the SCRBC/KC-ARC-PEME system to yield substantial hydrogen production from the PEME,
reaching up to 176.328 kg/h.

4.4. Comparison with previous studies


To establish the superiority of our study, we compared it with other similar combined systems, specifically those based on sCO2
power cycles. By reviewing literature on cogeneration or trigeneration systems, we extracted improvement results from various

19
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

Fig. 10. Impact of T26 on: (a) Qevap, Wnet,sCO2,Wnet,KC, Wnet, Cycle, and mH2, and (b) ηen, ηex & cp,tot.

Table 12
Practical ranges of decision variables.

Decision parameters PRc PVG1 (KPa) TTDVG1 (oC) T25 (oC) T17 (oC) T26 (oC) ΔTVG2,hot (oC) X13

Range 2.4–4.6 2200–3800 10–30 26–42 26–42 0–10 10–20 0.64–0.82

Table 13
Optimization outcomes for the SCRBC and SCRBC/APC- PEME systems.

Parameters SCRBC SCRBC/KC-ARC- PEME

Base case EEOD XEOD COD Base case EEOD XEOD COD

PRc 3 3.457 3.457 2.68 3 2.768 3.836 2.897


PVG1 (KPa) – – – – 3500 3800 3488 2200
TTDVG1 (oC) – – – – 20 25.87 30 10
T25 (oC) – – – – 30 26 26 42
T17 (oC) – – – – 30 28.47 26 42
T26 (oC) – – – – 5 10 0 10
ΔTVG2,hot (oC) – – – – 15 20 20 10
X13 – – – – 0.7 0.64 0.82 0.82
ηen (%) 39.6 39.86 39.86 39.14 45.44 61.21 45.16 42.45
ηex (%) 54.38 55.13 55.13 54.13 58.75 57.59 60.51 57.15
mH2 (kg/hr) – – – – 115.452 55.728 176.328 79.128
Q̇EVAP (MW) – – – – 19.475 123.949 9.216 7.112
Ẇnet (MW) 237.886 239.174 239.174 234.837 245.586 243.249 261.737 247.566
cP,tot ($/GJ) 12.27 12.42 12.42 12.22 11.675 11.875 11.951 11.46

combined systems and compared them with our findings. The developed system outperforms previous ones in several aspects. Our
comprehensive simulations and analysis reveal a significant performance enhancement, as indicated in Table 14, with the integration
of KC-ARC-PEME subsystems resulting in an 8.03 % improvement in exergy efficiency and a 5.11 % reduction in total product unit cost,
surpassing similar studies in the literature. Furthermore, our research highlights contributions to sustainability by achieving a very low
total product unit cost (11.67 $/GJ) through exergoeconomic analysis. Prioritizing low product cost aligns with sustainability,
fostering efficient resource use, minimized waste, and a reduced environmental footprint, establishing our system as eco-friendly and
sustainable.

5. Conclusions
This paper presents an energy-efficient trigeneration system that simultaneously provides cooling, power, and hydrogen. It’s a
hybrid system comprising a supercritical carbon dioxide recompression Brayton cycle (SCRBC) as the primary cycle and three
interconnected subsystems: a Kalina cycle (KC) combined with an ammonia-water-based absorption refrigeration cycle (ARC) and a
PEM electrolyzer (PEME) unit. We conducted a comprehensive assessment of the system’s performance, combining thermodynamic

Table 14
Comparison between the current study with previous studies.

Ref. System type Improvement in ηex (%) Improvement in cP,tot (%)

Present work SCRBC/KC-ARC 8.03 5.11


[47] SCRBC/ARC-MED 4.02 –
[21] SCRBC/ARC 2.73 2.03
[20] RSCBC/ARC 2.54 –
[13] SCRBC/KC 6.37 3.51

20
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

and exergoeconomic analysis and optimizing its operation. Additionally, we performed a parametric analysis to understand the impact
of key parameters on both the thermodynamics and economics. Furthermore, we compared this trigeneration system to a conventional
SCRBC system. Our research findings and discussions have led to the following conclusions:
• For the base design, the integration of the KC-ARC-PEME subsystem with the SCRBC cycle resulted in a notable 14.7 % increase in
energy efficiency, reaching a value of 45.44 %. Furthermore, there was a significant 8.03 % improvement in exergy efficiency,
elevating it to 58.75 %. These enhancements correspondingly led to a reduction in the total product unit cost of the SCRBC system,
decreasing it from $12.272/GJ to $11.67/GJ, which represents a 5.11 % cost reduction.
• According to an exergoeconomic analysis, the reactor and sCO2 turbine are the critical components of the SCRBC/KC-ARC-PEME
system. This is indicated by their significantly higher values of Zk + CD + CL, surpassing those of other components.
• The SCRBC/KC-ARC-PEME system experiences the highest rate of exergy destruction in the reactor which is caused by a significant
temperature difference. However, the KC-ARC-PEME cycle’s components demonstrate lower levels of exergy destruction.
Enhancing the performance of the combined cycle is possible by boosting the outlet temperature of the reactor, thus improving both
thermodynamic and exergoeconomic outcomes.
• The optimization outcomes indicated that integrating the KC-ARC-PEME subsystem with the SCRBC yields improvements across all
three objective functions. The enhancement percentages for energy efficiency: approximately 53.56 % for EEOD, 13.3 % for XEOD,
and 8.45 % for COD scenarios. Furthermore, the exergy efficiency demonstrates increments of around 4.46 % for EEOD, 9.76 % for
XEOD, and 5.58 % for COD scenarios. Conversely, a reduction in cP;tot is observed across all three cases, with decrease percentages
of approximately 4.59 % for EEOD, 3.92 % for XEOD, and 6.63 % for COD instance.
• Across all optimal scenarios, the SCRBC/KC-ARC-PEME system not only achieves a substantial net power output increase of up to
22.56 MW compared to the standalone SCRBC, but also demonstrates remarkable cooling capacity, particularly evident in the
EEOD case with a capacity of 123.949 MW. Furthermore, the optimization results underscore the potential of the SCRBC/KC-ARC-
PEME system to yield hydrogen production from PEME, reaching levels as high as 176.328 kg/h.
• Analysis of parameters indicates the presence of an optimal compressor pressure ratio for maximizing energy and exergy efficiency
or minimizing the total product unit cost.
• In a future study, multi-objective optimization needs to be carried out to comprehensively analyze and optimize the proposed
system while considering trade-offs between energy efficiency, exergy efficiency, and total product cost.

CRediT authorship contribution statement


Mohamed S. Yousef: Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration,
Software, Writing – original draft, Writing – review & editing. Domingo Santana: Funding acquisition, Supervision, Writing – review
& editing.

Declaration of competing interest


I disclose any actual or potential conflict of interest including any financial, personal or other relationships with other people or
organizations within three years of beginning the submitted work that could inappropriately influence, or be perceived to influence,
this work.
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgments
The researcher (Mohamed S. Yousef) acknowledges support from the CONEX-Plus program funded by Universidad Carlos III de
Madrid and the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement
No. 801538.

References
[1] R. Sun, M. Liu, X. Chen, K. Yang, J. Yan, Thermodynamic optimization on supercritical carbon dioxide Brayton cycles to achieve combined heat and power
generation, Energy Convers. Manag. 251 (2022), 114929, https://doi.org/10.1016/j.enconman.2021.114929.
[2] Y. Yang, Y. Huang, P. Jiang, Y. Zhu, Multi-objective optimization of combined cooling, heating, and power systems with supercritical CO2 recompression
Brayton cycle, Appl. Energy 271 (2020), 115189, https://doi.org/10.1016/j.apenergy.2020.115189.
[3] M. Ebadollahi, H. Rostamzadeh, P. Seyedmatin, H. Ghaebi, M. Amidpour, Thermal and exergetic performance enhancement of basic dual-loop combined cooling
and power cycle driven by solar energy, Therm. Sci. Eng. Prog. 18 (2020), 100556, https://doi.org/10.1016/j.tsep.2020.100556.
[4] A.D. Akbari, S.M.S. Mahmoudi, Thermoeconomic analysis & optimization of the combined supercritical CO2 (carbon dioxide) recompression Brayton/organic
Rankine cycle, Energy 78 (2014) 501–512, https://doi.org/10.1016/j.energy.2014.10.037.
[5] H. Zhu, G. Xie, A.S. Berrouk, S. Ni, Performance of Printed Circuit Heat Exchangers in SCO 2 Brayton/Organic Rankine Combined Cycle : Assessment of
Simplified Boiling Temperature and Heat Flux Type of Boundary Conditions, vol. 236, 2024, https://doi.org/10.1016/j.applthermaleng.2023.121543.
[6] C. Wu, S sen Wang, J. Li, Exergoeconomic analysis and optimization of a combined supercritical carbon dioxide recompression Brayton/organic flash cycle for
nuclear power plants, Energy Convers. Manag. 171 (2018) 936–952, https://doi.org/10.1016/j.enconman.2018.06.041.

21
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

[7] X. Wang, Y. Dai, Exergoeconomic analysis of utilizing the transcritical CO2 cycle and the ORC for a recompression supercritical CO2 cycle waste heat recovery: a
comparative study, Appl. Energy 170 (2016) 193–207, https://doi.org/10.1016/j.apenergy.2016.02.112.
[8] X. Wang, Y. Yang, Y. Zheng, Y. Dai, Exergy and exergoeconomic analyses of a supercritical CO2 cycle for a cogeneration application, Energy 119 (2017)
971–982, https://doi.org/10.1016/j.energy.2016.11.044.
[9] H. Li, M. Xu, X. Yan, J. Li, W. Su, J. Wang, et al., Preliminary conceptual exploration about performance improvement on supercritical CO2 power system via
integrating with different absorption power generation systems, Energy Convers. Manag. 173 (2018) 219–232, https://doi.org/10.1016/j.
enconman.2018.07.075.
[10] N. Shokati, F. Ranjbar, M. Yari, Exergoeconomic analysis and optimization of basic, dual-pressure and dual-fluid ORCs and Kalina geothermal power plants: a
comparative study, Renew. Energy 83 (2015) 527–542, https://doi.org/10.1016/j.renene.2015.04.069.
[11] H. Li, M. Wang, J. Wang, Y. Dai, Exergoeconomic analysis and optimization of a supercritical CO2 cycle coupled with a Kalina cycle, J. Energy Eng. 143 (2017)
1–13, https://doi.org/10.1061/(asce)ey.1943-7897.0000411.
[12] Y. Feng, Z. Du, M. Shreka, Y. Zhu, S. Zhou, W. Zhang, 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. 206 (2020), 112483, https://doi.org/
10.1016/j.enconman.2020.112483.
[13] G. Fan, Y. Dai, Thermo-economic optimization and part-load analysis of the combined supercritical CO2 and Kalina cycle, Energy Convers. Manag. 245 (2021),
114572, https://doi.org/10.1016/j.enconman.2021.114572.
[14] M. Mohsenipour, M. Ebadollahi, H. Rostamzadeh, M. Amidpour, Design and evaluation of a solar-based trigeneration system for a nearly zero energy
greenhouse in arid region, J. Clean. Prod. 254 (2020), 119990, https://doi.org/10.1016/j.jclepro.2020.119990.
[15] H. Rostamzadeh, M. Ebadollahi, H. Ghaebi, M. Amidpour, R. Kheiri, Energy and exergy analysis of novel combined cooling and power (CCP) cycles, Appl.
Therm. Eng. 124 (2017) 152–169, https://doi.org/10.1016/j.applthermaleng.2017.06.011.
[16] J. Yuan, C. Wu, X. Xu, C. Liu, Multi-mode analysis and comparison of four different carbon dioxide-based combined cooling and power cycles for the distributed
energy system, Energy Convers. Manag. 244 (2021), 114476, https://doi.org/10.1016/j.enconman.2021.114476.
[17] A.D. Akbari, S.M.S. Mahmoudi, Thermoeconomic performance and optimization of a novel cogeneration system using carbon dioxide as working fluid, Energy
Convers. Manag. 145 (2017) 265–277, https://doi.org/10.1016/j.enconman.2017.04.103.
[18] K. Manjunath, O.P. Sharma, S.K. Tyagi, S.C. Kaushik, Thermodynamic analysis of a supercritical/transcritical CO2 based waste heat recovery cycle for shipboard
power and cooling applications, Energy Convers. Manag. 155 (2018) 262–275, https://doi.org/10.1016/j.enconman.2017.10.097.
[19] A. Yu, W. Su, X. Lin, N. Zhou, L. Zhao, Thermodynamic analysis on the combination of supercritical carbon dioxide power cycle and transcritical carbon dioxide
refrigeration cycle for the waste heat recovery of shipboard, Energy Convers. Manag. 221 (2020), 113214, https://doi.org/10.1016/j.enconman.2020.113214.
[20] C. Wu, X. Xu, Q. Li, J. Li, S. Wang, C. Liu, 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. 207 (2020), 112527, https://
doi.org/10.1016/j.enconman.2020.112527.
[21] C. Wu, S sen Wang, Feng X. jia, J. Li, Energy, exergy and exergoeconomic analyses of a combined supercritical CO2 recompression Brayton/absorption
refrigeration cycle, Energy Convers. Manag. 148 (2017) 360–377, https://doi.org/10.1016/j.enconman.2017.05.042.
[22] H. Li, W. Su, L. Cao, F. Chang, W. Xia, Y. Dai, Preliminary conceptual design and thermodynamic comparative study on vapor absorption refrigeration cycles
integrated with a supercritical CO2 power cycle, Energy Convers. Manag. 161 (2018) 162–171, https://doi.org/10.1016/j.enconman.2018.01.065.
[23] Y. Du, Y. Dai, Off-design performance analysis of a power-cooling cogeneration system combining a Kalina cycle with an ejector refrigeration cycle, Energy 161
(2018) 233–250, https://doi.org/10.1016/j.energy.2018.07.106.
[24] O. Barkhordarian, A. Behbahaninia, R. Bahrampoury, A novel ammonia-water combined power and refrigeration cycle with two different cooling temperature
levels, Energy 120 (2017) 816–826, https://doi.org/10.1016/j.energy.2016.11.127.
[25] M. Ebadollahi, M. Amidpour, O. Pourali, H. Ghaebi, Flexibility concept in design of advanced multi-energy carrier systems driven by biogas fuel for sustainable
development, Sustain. Cities Soc. 86 (2022), 104121, https://doi.org/10.1016/j.scs.2022.104121.
[26] M.S. Yousef, D. Santana, Energy and exergy analyses of a recompression supercritical CO2 cycle combined with a double-effect parallel absorption refrigeration
cycle, Energy Rep. 9 (2023) 195–201, https://doi.org/10.1016/j.egyr.2023.09.161.
[27] H. Azariyan, M. Vajdi, H. Rostamnejad Takleh, Assessment of a high-performance geothermal-based multigeneration system for production of power, cooling,
and hydrogen: thermodynamic and exergoeconomic evaluation, Energy Convers. Manag. 236 (2021), 113970, https://doi.org/10.1016/j.
enconman.2021.113970.
[28] J. Wang, J. Wang, P. Zhao, Y. Dai, Thermodynamic analysis of a new combined cooling and power system using ammonia-water mixture, Energy Convers.
Manag. 117 (2016) 335–342, https://doi.org/10.1016/j.enconman.2016.03.019.
[29] L. Cao, J. Lou, J. Wang, Y. Dai, Exergy analysis and optimization of a combined cooling and power system driven by geothermal energy for ice-making and
hydrogen production, Energy Convers. Manag. 174 (2018) 886–896, https://doi.org/10.1016/j.enconman.2018.08.067.
[30] S.M. Alirahmi, E. Assareh, N.N. Pourghassab, M. Delpisheh, L. Barelli, A. Baldinelli, Green hydrogen & electricity production via geothermal-driven multi-
generation system: thermodynamic modeling and optimization, Fuel 308 (2022), 122049, https://doi.org/10.1016/j.fuel.2021.122049.
[31] M. Abid, M.S. Khan, T.A.H. Ratlamwala, Comparative energy, exergy and exergo-economic analysis of solar driven supercritical carbon dioxide power and
hydrogen generation cycle, Int. J. Hydrogen Energy 45 (2020) 5653–5667, https://doi.org/10.1016/j.ijhydene.2019.06.103.
[32] S.C. Toker, G. Soyturk, O. Kizilkan, Development of a sustainable multi-generation system with re-compression sCO2 Brayton cycle for hydrogen generation, Int.
J. Hydrogen Energy 47 (2022) 19397–19410, https://doi.org/10.1016/j.ijhydene.2021.11.138.
[33] L.M. Hadelu, A. Noorpoor, F.A. Boyaghchi, S. Mirjalili, 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. Protect.
159 (2022) 393–421, https://doi.org/10.1016/j.psep.2022.01.013.
[34] S. Klein, F. Alvarado, Engineering Equation Solver, vols. 1–2, F-Chart Software, Box, 2002.
[35] Y. Cao, F. Rostamian, M. Ebadollahi, M. Bezaatpour, H. Ghaebi, Advanced exergy assessment of a solar absorption power cycle, Renew. Energy 183 (2022)
561–574, https://doi.org/10.1016/j.renene.2021.11.039.
[36] R. Palacios-Bereche, R. Gonzales, S.A. Nebra, Exergy calculation of lithium bromide-water solution and its application in the exergetic evaluation of absorption
refrigeration systems LiBr-H2O, Int. J. Energy Res. 36 (2012) 166–181, https://doi.org/10.1002/er.1790.
[37] M.A. Emadi, J. Mahmoudimehr, Modeling and thermo-economic optimization of a new multi-generation system with geothermal heat source and LNG heat sink,
Energy Convers. Manag. 189 (2019) 153–166, https://doi.org/10.1016/j.enconman.2019.03.086.
[38] M. Sharaf, M.S. Yousef, A.S. Huzayyin, Year-round energy and exergy performance investigation of a photovoltaic panel coupled with metal foam/phase change
material composite, Renew. Energy (2022), https://doi.org/10.1016/j.renene.2022.03.071.
[39] M. Noaman, G. Saade, T. Morosuk, G. Tsatsaronis, Exergoeconomic analysis applied to supercritical CO2 power systems, Energy 183 (2019) 756–765, https://
doi.org/10.1016/j.energy.2019.06.161.
[40] J. Tang, Q. Zhang, Z. Zhang, Q. Li, C. Wu, X. Wang, Development and performance assessment of a novel combined power system integrating a supercritical
carbon dioxide Brayton cycle with an absorption heat transformer, Energy Convers. Manag. 251 (2022), 114992, https://doi.org/10.1016/j.
enconman.2021.114992.
[41] D. Mignard, Correlating the chemical engineering plant cost index with macro-economic indicators, Chem. Eng. Res. Des. 92 (2014) 285–294, https://doi.org/
10.1016/j.cherd.2013.07.022.
[42] M.S. Yousef, D. Santana, Thermodynamic and exergoeconomic optimization of a new combined cooling and power system based on supercritical CO2
recompression Brayton cycle, Energy Convers. Manag. 295 (2023), 117592, https://doi.org/10.1016/j.enconman.2023.117592.
[43] M.S. Yousef, D. Santana, Thermodynamic and exergoeconomic analysis of utilizing a modified Kalina cycle for a recompression supercritical CO 2 cycle waste
heat recovery, 2023 6th Int Conf Electr Eng Green Energy (2023) 253–259, https://doi.org/10.1109/ceege58447.2023.10246643.

22
M.S. Yousef and D. Santana Case Studies in Thermal Engineering 53 (2024) 103902

[44] H.D.M. Hettiarachchi, M. Golubovic, W.M. Worek, Y. Ikegami, The performance of the Kalina cycle system 11(kcs-11) with low-temperature heat sources,
J Energy Resour Technol Trans ASME 129 (2007) 243–247, https://doi.org/10.1115/1.2748815.
[45] D.-W. Sun, COMPARISON of the performances OF NH3-H20, NH3-LiNO3 and NH3-NaSCN absorption refrigeration systems absorption ammonia-water
refrigeration air conditioning ammonia-lithium nitrate ammonia-sodium thiocyanate aqua-ammonia computer simulation mathematical mo, Energy Convers
Mgmt 39 (1998) 357–368.
[46] T. Ioroi, K. Yasuda, Z. Siroma, N. Fujiwara, Y. Miyazaki, Thin film electrocatalyst layer for unitized regenerative polymer electrolyte fuel cells, J. Power Sources
112 (2002) 583–587, https://doi.org/10.1016/S0378-7753(02)00466-4.
[47] H.W. Li, Y. Sun, Y.Y. Pan, C.H. Du, D. Wang, Preliminary design, thermodynamic analysis and optimization of a novel carbon dioxide based combined power,
cooling and distillate water system, Energy Convers. Manag. 255 (2022), 115367, https://doi.org/10.1016/j.enconman.2022.115367.

23

You might also like