Next Article in Journal
Quantitative Characterization of Pore–Fracture Structures in Coal Reservoirs by Using Mercury Injection–Removal Curves and Permeability Variation under Their Constraints
Previous Article in Journal
Special Issue on “Process Monitoring and Fault Diagnosis”
Previous Article in Special Issue
Analysis and Optimization of the Fuel Consumption of an Internal Combustion Vehicle by Minimizing the Parasitic Power in the Cooling System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Economic and Technical Assessing the Hybridization of Solar Combined Cycle System with Fossil Fuel and Rock Bed Thermal Energy Storage in Neom City

by
Abdulrazzak Akroot
1 and
Abdullah Sultan Al Shammre
2,*
1
Mechanical Engineering Department, Faculty of Engineering, Karabük University, 78050 Karabük, Turkey
2
Economics Department, College of Business Administration, King Faisal University, Al-Ahsa 31982, Saudi Arabia
*
Author to whom correspondence should be addressed.
Processes 2024, 12(7), 1433; https://doi.org/10.3390/pr12071433
Submission received: 11 June 2024 / Revised: 3 July 2024 / Accepted: 5 July 2024 / Published: 9 July 2024

Abstract

:
Rising energy demands, the depletion of fossil fuels, and their environmental impact necessitate a shift towards sustainable power generation. Concentrating solar power (CSP) offers a promising solution. This study examines a hybridization of a combined cycle power plant (CCPP) based on solar energy with fossil fuel and energy storage in rock layers to increase Saudi Arabia’s electricity production from renewable energy. The fuel is used to keep the temperature at the inlet of the gas turbine at 1000 °C, ensuring the power produced by the Rankine cycle remains constant. During the summer, the sun is the main source of power generation, whereas in the winter, reliance on fuel increases significantly. The Brayton cycle operates for 10 h during peak solar radiation periods, storing exhaust heat in rock beds. For the remaining 14 h of the day, this stored heat is discharged to operate the Rankine steam cycle. Simulations and optimizations are performed, and the system is evaluated using a comprehensive 4E analysis (energy, exergy, exergoconomic, and environmental) alongside a sustainability assessment. A parametric evaluation examines the effect of key factors on system performance. The rock bed storage system compensates for solar intermittency, enabling power generation even without sunlight. The study reveals that the system generated 12.334 MW in June, achieving an energy efficiency of 37% and an exergy efficiency of 40.35%. The average electricity cost during this period was 0.0303 USD/kWh, and the carbon footprint was 0.108 kg CO2/kWh. In contrast, during January, the system produced 13.276 MW with an energy efficiency of 37.91% and an exergy efficiency of 44.16%. The average electricity cost in January was 0.045 USD/kWh, and the carbon footprint was 0.1 kg CO2/kWh. Interestingly, solar energy played a significant role: it contributed 81.42% of the heat in June, while in January, it accounted for 46.77%. The reduced electricity costs during June are primarily attributed to the abundant sunshine, which significantly powered the system.

1. Introduction

Solar thermal power stations that use thermal energy storage (TES) are an alternative to traditional fossil fuels [1]. Solar thermal power stations use the sun’s limitless energy supplies rather than the limited fossil fuels used by traditional power plants, which cause pollution. These plants play a vital role in combating climate change and reducing greenhouse gas emissions by converting solar heat into electricity. Incorporating TES into solar thermal power plants is a novel approach [2,3,4]. The system efficiently uses solar energy to generate electricity and includes a storage system to provide a consistent power supply, even under low sunshine conditions [5].
Many research projects have focused on thermal energy storage devices and how they might be integrated into power plant operations. Soprani et al. [6] focused on a high-temperature TES prototype using diabase rocks and air for heat transfer, aiming to store heat at 600 °C with 450 kWhth capacity. They analyzed the system’s temperature and flow patterns during charging and discharging, looking into efficiency and heat retention metrics. Desai et al. [7] investigated a new, cost-effective concentrated solar power system employing a micro-structured polymer foil for solar concentration. Their study compared three thermal energy storage options, emphasizing the optimal design for cogeneration systems in terms of cost and efficiency. Marongiu et al. [8] offered insights through a 2-D numerical model of a 450 kWhth rock bed, considering the temperature-dependency of air and rocks to evaluate design factors like the rock size and type on the system efficiency. Zanganeh et al. [9] detailed a 6.5 MWhth air-and-rock thermal storage unit, analyzing a numerical model that aligns closely with experimental data for an accurate heat transfer representation. Nahhas et al. [10] assessed basalt rocks for thermal storage, finding them suitable up to 700 °C due to their thermal and mechanical properties. Kocak and Paksoy [11] tested an affordable and eco-friendly material for thermal storage in low-temperature industrial solar applications. Sharma et al. [12] evaluated packed-bed TES systems with different materials, identifying hybrid particles as the most effective storage medium. Heller and Gauche [13] proposed a rock bed heat transfer model for storing and transmitting energy to steam cycles, using simulations to examine its influence on power plant economics. Abdulla and Reddy [14] modeled the heat transfer in a molten salt-packed-bed system for CSP plants, identifying key parameters for efficient energy discharge. Öztürk et al. [15] introduced a combined gas and steam turbine system with a rock bed TES for diversified heating and power applications, conducting extensive analyses on performance and efficiency under variable conditions. Fernández et al. [16] analyzed operation strategies for packed bed TES systems to enhance efficiency and thermal storage behavior. Pelay et al. [17] assessed the TES systems in concentrated solar power (CSP) plants. They discussed the global status of CSP, various high-temperature TES technologies (200–1000 °C), thermochemical storage, and new integration strategies for TES in CSP. Freeman et al. [18] evaluated the TES solutions for domestic solar combined heat and power systems using ORC engines and solar–thermal collectors. The study discussed how TES enhances system efficiency, aligns with user demands, and reduces costs, especially under variable solar conditions. Çam et al. [19] assessed a solar-assisted heat pump-integrated latent heat TES system for variable-temperature heating. The authors investigated using phase change materials (PCMs) with different characteristics to improve solar energy intermittency and energy efficiency. According to the research, the best PCM settings may save operating expenses considerably, boost heat pump performance by 35% to 80%, and provide a payback time of around 13 years in Izmir. Khamlich et al. [20] conducted a techno-economic analysis of a 100 MWe CSP plant with various TES configurations, assessing their integration into a spot market. Al-Nimr et al. [21] found that a CPV/T solar power system with ORC, a geothermal condenser, and energy storage can achieve up to 22% efficiency in November. Shakouri et al. [22] analyzed a multigeneration system integrating conventional and renewable energy sources, achieving enhanced efficiency and sustainability with significant reductions in fuel use and CO2 emissions. Kuterbekov et al. [23] assessed a hybrid power system utilizing solar and biomass energy to produce electricity, heat, and hydrogen, demonstrating cost efficiencies and exergy improvements through the innovative use of heat exchangers and polymer membrane electrolyzers, with the system’s performance analyzed using the ESS tool (https://fchartsoftware.com/ees/). Zhang et al. [24] analyzed energy and exergy losses during the startup of PTC solar power plants using molten salt. They demonstrated that a molten salt anti-freezing solution can reduce startup energy consumption. Wang et al. [25] proposed a hybrid hydrogen production and gas–solar power system integrating a solar tower, a gas–steam turbine, and an organic Rankine cycle, achieving 103.9 MW power and 41.3% efficiency with stable operations and identifying areas for performance enhancement in exergy losses. Table 1 provides a summary of the reviewed research on thermal energy storage devices and their potential integration into power plant operations.
This work presents a unique study that introduces a 4E analysis on a novel hybridization of a combined cycle based on solar energy with fossil fuels, featuring rock bed energy storage for reliable power supply during low sunlight. The research evaluates the plant’s exergy efficiency, investment, and operational costs, focusing on how varying parameters influence the overall efficiency and cost-effectiveness, marking a distinct approach to analyzing solar-powered combined cycle systems.

2. System Description

The newly developed combined cycle power plant (CCPP) features a primary solar Brayton cycle (SBC), a secondary Rankine cycle (RC), and rock bed thermal energy storage (RBES). This research focused on the Neom region in Saudi Arabia. Figure 1 illustrates the CCPP system, which utilizes both solar and fuel energy. The SBC includes an air compressor, heliostats, solar receivers, a combustion chamber, and a gas turbine. The air compressor increases air pressure to 1469 kPa, which is then heated by a solar central receiver. The combustion chamber adjusts for solar light variations, maintaining the gas turbine’s inlet temperature at 1000 °C using a small amount of CH4 fuel for enhanced efficiency. Exhaust gases from the SBC pass through the RBES, storing energy during the day and enabling the RC to generate power at night without solar radiation. The RC uses hot exhaust gases from the GT stored in the RBES and benefits from the hot air produced by the RBES. The SBC operates for 10 h during peak solar radiation, storing hot exhaust in rocks. The remaining 14 h are used to run the Rankine steam cycle, with the system generating power from the SBC for 10 h and from the RC for 14 h. The temperature profile through the boiler is illustrated in Figure 2.
The simulation in this work utilized Engineering Equation Solver (EES) software (https://fchartsoftware.com/ees/) for modeling and analysis, which calculates thermodynamic properties such as temperature, exergy, pressure, and entropy. The thermodynamic, exergoeconomic, and environmental studies are conducted with the following basic assumptions:
  • The reference state pressure (Po) is 1 bar, and the reference state temperature (To) is 25 °C.
  • There is no change to either the kinetic or potential energy.
  • The system is in a steady-state condition.
  • The fuel is pure methane (CH4).
  • The sun’s surface temperature is 6000 K.
  • The turbines, pump, and compressor operate adiabatically.
The designed system involves individual thermodynamic modeling of each component. Equations specific to mass, energy, exergy, and exergoeconomic analyses are used, with detailed operational and technical parameters provided in Table 2.

3. Thermodynamic Analysis

The system’s mass balance equation is expressed as
m ˙ i n = m ˙ o u t
The following equation represents the system’s overall energy balance [26]:
Q ˙ i n + W ˙ i n + m ˙ i n h in   = Q ˙ o u t + W ˙ o u t + m ˙ o u t h o u t
The system’s exergy balance equation may be written as follows: [27,28]:
( m ˙ e ) i n + E ˙ i n , W + E ˙ i n , Q = ( m ˙ e ) o u t + E ˙ out , W + E ˙ out , Q
E ˙ Q = Q ˙ 1 T 0 T
m ˙ in   S in   + Q ˙ T + S ˙ gen   = m ˙ out   S out  
The heat transfer rate by the solar receiver is determined using the following equations:
Q ˙ Solar   = Q ˙ h Q ˙ rec , loss  
Q ˙ Solar   = m ˙ 3 h 3 m ˙ 2 h 2 = m ˙ 3 c p , 3 T 3 m ˙ 2 c p , 2 T 2
The following equations are used to compute the rate of heat transfer in the heliostat field and the rate of heat loss in the receiver [29]:
Q ˙ rec , loss   = Q ˙ c   + Q ˙ r = A r × h c × T r T 0 + σ × ε × T r 4 T 0 4
Q ˙ h = A h × N × I × η h
The convective heat transfer coefficient of air (hc) can be calculated using the following formula, with units in W/m2K [15]:
h c = 10.45 V + 10 V
The SBC’s outlet temperature and power consumption of the compressor are determined using the following equations [30]:
T out   = T in   1 + 1 η A C P r , A C k a i r 1 k a i r 1
W ˙ A C = m ˙ a i r h out   h i n
In this context, m ˙ a i r and k a i r   denote the mass flow rate and the specific heat ratio of the air, respectively. On the other hand, η A C   represents the isentropic efficiency of the AC, while P r , A C stands for the pressure ratio. The following equations determine the gas turbine’s output temperature and power generation [31]:
T G T ,   out   = T G T ,   in   1 η G T 1 P r , G T 1 k a i r k a i r
W ˙ G T = m ˙ G T , i n h G T ,   in   h G T ,   out  
W ˙ G T , n e t = η G E N W ˙ G T W ˙ A C
The equation uses η G T , P r , G T , η G E N , m ˙ G T , i n to represent the turbine isentropic efficiency, turbine pressure ratio, generator efficiency, and combustion gas flow rate.
Determining the heat loss from the tank to the ambient is a key part of the thermodynamic model used to calculate the RB storage tank. First, we use the below equation to get the total heat loss from the store to the ambient air, and then we compute the total heat loss coefficient (U):
Q ˙ loss   = U × A × T m T 0
where T 0 and T m are the ambient and mean temperatures (in °C), and A is the total surface area of the RB storage tank.
The RBES operates in three modes: charging, discharging, and storage. The following formulas determine the total energy stored when charging [15]:
Q   charging = m 5 h 5 m 6 h 6 × 3600 × t c h a r g i n g
Q charging = m r o c k   c r o c k T 5 T 0
where c r o c k is the rock’s specific heat, t c h a r g i n g is the charging time, and m r o c k is the rock’s mass. To calculate the overall heat loss throughout the storage time, one uses the following equation:
Q charging = Q discharging + Q loss
The heat delivered to the Rankine cycle during the discharge time is determined using the following formula [15]:
Q discharging = m 7 h 7 m 9 h 9 × 3600 × t discharging  
Q discharging = m rock   c rock   T 5 T 0
where t discharging   is the discharging time (in s). Table 3 presents the energy and exergy equations for the CCPP’s components. Table 4 shows the exergy of the fuel and the product for each component.
The efficiency of the CCPP is calculated by analyzing the first and second laws [32]:
η I   = W ˙ net   Q ˙ in  
η II   = W ˙ net   E ˙ i n
The net power of the CCPP is determined by the following equation:
W ˙ net   = W ˙ GT   W ˙ Comp   + W ˙ S T W ˙ Pump   W ˙ Blower  
The CCPP’s heat and exergy input are calculated as follows:
Q ˙ i n = Q ˙ F u e l + Q ˙ Solar  
E ˙ i n = E ˙ Q , F u e l + E ˙ Solar  

4. Exergy Economic Analysis

The following equation describes the overall system-wide exergy-cost balance for all the system components [33]:
C ˙ i n , k + C ˙ Q , k + Z ˙ k = C ˙ o u t , k + C ˙ w , k
The heat and power exergy cost flows are denoted as C ˙ Q k and C ˙ w , k correspondingly, whereas Z ˙ k represents the capital cost. The calculation of Z ˙ k is based on the formula shown below [34,35]:
Z ˙ k = Z k C R F ϕ t
The referenced equations used to determine the investment cost of the components, denoted by Z k , are derived from references. ϕ denotes the maintenance factor, and t stands for the yearly running time of the CCPP [13,36,37]. This study assumes a total annual working time of 8640 h. CRF represents the return on capital factor, and it is calculated using the following equation [38]:
C R F = i ( 1 + i ) N ( 1 i ) N 1
The interest rate, denoted as i, is fixed at 12%, while N denotes the CCPP’s overall lifespan, which is 20 years [39]. The exergy cost flow for each location in the CCPP is determined using the following formula:
C ˙ = c E ˙
where c represents the specific exergy cost, and E ˙ indicates the exergy current. The CCPP’s performance is assessed using the exergy-economic factor, designated as f k , which is determined using the following equation [40]:
f k = Z ˙ k Z ˙ k + C ˙ D , k
The exergy destruction cost, denoted as C ˙ D , k , is determined using the following equation [40]:
C ˙ D , k = c f E ˙ D , k
where E ˙ D , k is the exergy destruction. Table 5 contains all the auxiliary equations that are linked to the cost equilibrium equations. The total cost of the CCPP system ( C ˙ s y s t e m ) can be calculated utilizing the following equation [41,42]:
C ˙ s y s t e m = Z ˙ k + C ˙ D , k
The system’s unit cost of electricity produced ( C ˙ e l e c t r i c i t y ) is another important aspect in the cost analysis. The equation provided calculates the overall cost of the power generated [43,44]:
C ˙ e l e c t r i c i t y = C ˙ s y s t e m W ˙ n e t

5. Environmental Assessment

The C O 2 emission rate ( ϵ C O 2 ) measures the amount of C O 2 that is emitted into the atmosphere as a result of human activities [45,46]. The value of ϵ C O 2 for electricity production is calculated by dividing the total amount of C O 2 emissions produced during a certain period of time by the total amount of energy generated during the same period. The rate is calculated using the following equation [47,48]:
ϵ C O 2 = m ˙ C O 2 W ˙ n e t
The estimation of the mass flow rate of C O 2 is determined using the following equation [49]:
m ˙ C O 2 = y C O 2 m ˙ g , 5 M ¯ C O 2 M ¯ g
where m ˙ C O 2 represents the molecular weight of C O 2 , whereas M ¯ C O 2 represents the mole fraction of C O 2 . Furthermore, m ˙ g represents the molecular weight of the exhaust gases, whereas M ¯ g denotes the mole fraction of the exhaust gases after the combustion chamber.

6. Results and Discussion

The results of analyzing the energy, exergy, and cost balance equations for each of the developed cycle’s constituent components are shown in this section. Table 6 provides details of the calculated and assumed properties for all the state points. The calculations related to exergy destruction and capital investment costs are performed.
The monthly fluctuations in the CCPP’s power output and energy cost rate will be provided. The thermodynamic model employed in this research for the developed cycle has undergone cross-verification using the findings from a study conducted by Oztürk et al. [15], as demonstrated in Figure 3. The outcomes of this study closely align with the existing literature.
Exergy destruction inside the CCPP components under ideal operating circumstances is explained in Table 7 and Figure 4. High-temperature variances correlate with more significant exergy destruction. Solar receivers experience the most significant temperature difference, and they exhibit the highest destruction rate, amounting to 7.346 MW. GT and CC also contribute significantly to the overall destruction, accounting for 13.98% and 10.41%, respectively. Figure 3 highlights the pronounced exergy destruction in the GT. This phenomenon is attributed to the impact of hot air on its turbine blades. Rock bed energy storage (RBES) contributes 10.13% due to charging and discharging processes. AC, condensers, and ST played roles in the overall destruction. The pump exhibits minimal contributions. The boiler achieves an impressive exergy efficiency of 97.6%. These findings shed light on the intricate dynamics of exergy destruction and utilization within the system.
The designed cycle’s exergoeconomic characteristics are shown in Table 8. Among all the components, the solar receiver incurs the greatest capital cost rate, amounting to 105.8 USD/h. The steam turbine and combustion chamber follow the solar receiver, with capital costs of 56.09 USD/h and 55.23 USD/h, respectively. The boiler, air compressor, and gas turbine also have relatively high capital costs. The CCPP’s exergy destruction cost rate is crucial for exergoeconomic analysis. The CC has the highest exergy destruction cost rate at 55.07 USD/h, followed by the GT. Interestingly, the capital cost rate of the entire system surpasses the total cost rate of exergy destruction. The components with the highest values are considered the most significant. The solar receiver (SR) stands out with the maximum value of Z ˙ k + C ˙ D , while the steam turbine (ST) comes in second. The CCPP demonstrates an exceptionally high exergoeconomic factor of 63.07%. These insights provide valuable information for assessing the economic and energetic aspects of the system.
Figure 5a,b illustrate the influence of ambient temperature ( T 1 ) on various performance metrics. As ( T 1 ) increases, a decrease is observed in the network output ( W ˙ n e t ) , the system power cost rate ( C ˙ e l e c t r i c i t y ) , the overall system efficiency ( η I ), and the second-law efficiency ( η I I ). This trend is mainly attributed to the rising enthalpy ( h 2 ) and the power consumption of the air compressor with increasing T 1 . A portion of W ˙ n e t powers the compressor, and the rise in W ˙ A C negatively impacts both the W ˙ n e t and overall efficiencies. Figure 5a quantifies this effect, showing a decrease in W ˙ n e t from 13,382 kW to 11,641 kW when T 1 increases from 290 K to 315 K. When the ambient temperature decreases, both the work consumed by the compressor and the exit temperature of the compressor decrease. Consequently, the air temperature entering the combustion chamber also decreases. This decrease in inlet temperature increases the cost of exergy destruction in the combustion chamber and the entire cycle, which raises the specific energy cost. Figure 5a highlights this trend, with C ˙ e l e c t r i c i t y decreasing from 31.21 USD/MWh to 30.26 USD/MWh as T 1 increases from 290 K to 315 K. Furthermore, Figure 5b demonstrates the corresponding decline in efficiencies. η I , representing the system’s overall efficiency, decreases from 38.01% to 36.2%, and η I I , reflecting the second-law efficiency, decreases from 41.9% to 39.31%.
Figure 6a,b illustrate how the pressure ratio (PR) impacts various aspects of the combined cycle power plant. As the PR increases, the system’s power output decreases. Specifically, when the PR rises from 8 to 20, the W ˙ n e t decreases significantly from 14.04 MW to 10.445 MW (a reduction of around 3.594 MW), as seen in Figure 6a. This reduction is primarily due to the increased work of the air compressor. Elevating the PR to 15 results in a reduction in the system’s electricity cost rate. The cost rate decreases from 34.38 USD/MWh to 30.49 USD/MWh. This reduction is attributed to the decrease in the Ẇnet and the fuel input rate. Interestingly, when the PR surpasses 15, the cycle’s energy cost rate begins to climb. This upward trend is primarily due to the significant drop in the W ˙ n e t . Figure 6b demonstrates that as the pressure ratio (PR) increases, the system efficiencies improve until they peak. Beyond this point, the efficiency declines. At higher PR values, a decrease in the W ˙ n e t negatively affects the overall system efficiencies. At 13 bars, the maximum thermal efficiency reached 36.9%, while the maximum exergy efficiency of 40.66% was achieved at 11.4 bar.
The effect of the GTIT ( T 4 ) on the power output, the cycle’s energy cost rate, and the overall efficiencies of the system according to the first and second laws are depicted in Figure 7a,b. Increasing T 4 improves the cycle’s power output and overall efficiencies. The CCPP achieved its greatest energy and exergy overall efficiencies of 38.39% and 42.294%, respectively, at 1050 K, as presented in Figure 7b. As T 4 rises, the energy and exergy inputs to the gas turbine increase. Notably, an elevated T 4 significantly enhances energy storage in the rock bed energy storage (RBES), improving the overall efficiency and power output. When T 4 rises from 800 K to 1050 K, the power output experiences a substantial increase. Specifically, the power output increased from 5590 kW to 13,922 kW (an increase of approximately 8332 kW). The study reveals that T 4 significantly affects the CCPP’s electricity cost rate. Figure 7a demonstrates that as T 4 increases, the CCPP’s power cost rate decreases. The findings suggest that the system’s electricity cost was reduced from 32.15 to 30.47 USD/MWh. These insights provide valuable information for optimizing the system’s performance and cost-effectiveness.
Figure 8 depicts how the sun and backup fuel contribute to the CCPP’s yearly heat supply. The data reveal that the summer months benefit from the sun’s increased intensity, with solar energy providing a whopping 81.42% of the heat in June. Conversely, the winter presents a different version. Due to lower solar radiation, the system relies heavily on fuel for heat, with solar only accounting for 44.41% of the needs in December. In essence, the sun plays the leading role in heat generation during the summer, while fuel becomes more crucial in the winter.
Figure 9 illustrates the CCPP’s power output variation across the year. Interestingly, January sees the highest power generation, reaching 13,276 kW. This can be attributed to two factors. First, fuel delivers higher quality energy compared to solar energy. Second, colder environments favor the gas turbine cycle’s efficiency. Due to these reasons, the system produces more power in January than in August, which sees a lower output of 12,206 kW.
Figure 10 shows how this CCPP’s electricity cost changes throughout the year. Interestingly, it is cheaper in the summer compared to the winter. This is because summer sunshine provides most of the power, and sunshine is free! The data present that the energy cost rate is as low as 30.36 USD/MWh in June but increases to 46.76 USD/MWh in December.
Figure 11 illustrates the annual variation in   C O 2 emissions for this system. Overall, the system maintains a very low carbon footprint due to its reliance on clean energy sources and minimal fuel usage. Interestingly, the carbon footprint is lower during the winter months compared to the summer. This is attributed to the system generating more power in January than in August. Specifically, the data reveal C O 2 emission rates as low as 0.1 kg/kWh in January, rising to 0.11 kg/kWh in August.

7. Conclusions

This study explored an innovative power generation system integrating a solar combined cycle system with fossil fuel and rock bed thermal energy storage in Neom City, Saudi Arabia. The investigation centered on thermodynamics, specifically energy, exergy, exergoeconomic, and environmental analysis. This distinctive configuration merges RBES with two power cycles: a solar Brayton cycle and a steam Rankine cycle. Over ten hours, the heated exhaust gas from the solar Brayton cycle powers the RBES, which uses this stored energy to run the RC for fourteen hours. The research scrutinized how vital parameters such as ambient temperature, pressure ratios within the cycles, and the GTIT influence electricity production’s comprehensive performance and cost. The study led to several noteworthy conclusions, as follows:
  • In the CCPP powered by SE, the system’s power output reaches a peak of 13,276 kW in January and decreases to its lowest output of 12,206 kW in August.
  • The CCPP’s power cost is greater during the winter than in the summer. Specifically, it was 30.35 USD/MWh in June, while it escalated to 46.76 USD/MWh in December.
  • The exergoeconomic factor (fk) performs a crucial role in the analysis of exergy economics. The designed CCPP system boasts an impressive exergoeconomic factor, peaking at 63.07%. This significant figure will likely motivate the Saudi Arabian government to construct this CCPP.
  • The designed system has a lower carbon impact due to clean energy and efficient fuel use. CO2 emission rates vary from 0.1 kg CO2/kWh in January to 0.11 kg CO2/kWh in August.
  • The system’s performance and cost are notably influenced by the ambient temperature, a factor that warrants careful consideration during the system’s design phase.
  • An increase in the GTIT notably enhances the system’s performance and efficiencies while concurrently reducing the system’s electricity cost rate.

Author Contributions

Conceptualization, A.A. and A.S.A.S.; Methodology, A.A.; Formal analysis, A.A.; Investigation, A.A. and A.S.A.S.; Resources, A.A. and A.S.A.S.; Data curation, A.A.; Writing—original draft, A.A.; Writing—review and editing, A.A. and A.S.A.S.; Funding acquisition, A.S.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia (Grant No. KFU241351).

Data Availability Statement

The data are contained within the article.

Acknowledgments

All the authors are grateful to King Faisal University and Karabuk University for providing their software (https://fchartsoftware.com/ees/) and hardware infrastructure to realize the current study.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

E ˙ exergy rate (kJ/s)
h c convective heat transfer coefficient of air (W/m2K)
W ˙ power (kW)
Z ˙ capital cost (USD/h)
A a p area of the solar field (m2)
C ˙ cost rate (USD/h)
Q ˙ heat transfer rate (kW)
S ˙ entropy rate (W/K)
T s u n sun temperature
m ˙ mass flow rate (kg/s)
cspecific heat (kJ kg−1 K−1)
DNIdirect normal irradiance of the sun
especific exergy (kJ/kg)
hspecific enthalpy (kJ. kg−1)
kspecific heat ratio
L H V fuel’s lower heating value (kJ/kg)
mmass (kg)
N number of operating hours
Ppressure (kPa)
Qheat transfer (kJ)
s specific entropy (kJ kg−1 K−1)
ttime (s)
Ttemperature
Uoverall heat transfer coefficient (W/m2K)
Vwind velocity (m/s)
Greek Symbols
i interest rate
η energy efficiency
Ψ exergy efficiency
f exergy-economic factor
ε emissivity
σ Stefan–Boltzmann constant
φ maintenance factor
Subscripts
cconvection
Ddestruction
ffuel
hheliostat
ininlet
kcomponent
Oreference state
outexit
pproduct
qrelated to heat
rradiation
tottotal
wrelated to work
Abbreviations
ACair compressor
CC combustion chamber
CRFcapital recovery factor
CSPconcentration solar power
GTgas turbine
GTC gas turbine cycle
GTITgas turbine inlet temperature
HEheat exchanger
Prpressure ration
RBESrock bed energy storage
SRsolar receiver
ST steam turbine
TESthermal energy storage

References

  1. Ma, Z.; Glatzmaier, G.C.; Kutscher, C.F. Thermal energy storage and its potential applications in solar thermal power plants and electricity storage. In Proceedings of the ASME 2011 5th International Conference on Energy Sustainability, ES, Washington, DC, USA, 7–10 August 2011; Volume 2011, pp. 447–456. [Google Scholar] [CrossRef]
  2. Talal, W.; Akroot, A. An Exergoeconomic Evaluation of an Innovative Polygeneration System Using a Solar-Driven Rankine Cycle Integrated with the Al-Qayyara Gas Turbine Power Plant and the Absorption Refrigeration Cycle. Machines 2024, 12, 133. [Google Scholar] [CrossRef]
  3. Strielkowski, W.; Tarkhanova, E.; Tvaronavi, M.; Petrenko, Y. Renewable Energy in the Sustainable Development of Electrical. Energies 2021, 14, 8240. [Google Scholar] [CrossRef]
  4. Delibaş, H.M.; Kayabaşı, E. Energy, Environment And Economy Assessment Of Waste Heat Recovery Technologies In Marine Industry. Int. J. Mater. Eng. Technol. 2019, 4, 39–45. [Google Scholar]
  5. Kuravi, S.; Trahan, J.; Goswami, D.Y.; Rahman, M.M.; Stefanakos, E.K. Thermal energy storage technologies and systems for concentrating solar power plants. Prog Energy Combust Sci 2013, 39, 285–319. [Google Scholar] [CrossRef]
  6. Soprani, S.; Marongiu, F.; Christensen, L.; Alm, O.; Dinesen, K.; Ulrich, T.; Engelbrecht, K. Design and testing of a horizontal rock bed for high temperature thermal energy storage. Appl. Energy 2019, 251, 113345. [Google Scholar] [CrossRef]
  7. Desai, N.B.; Mondejar, M.E.; Haglind, F. Techno-economic analysis of two-tank and packed-bed rock thermal energy storages for foil-based concentrating solar collector driven cogeneration plants. Renew. Energy 2022, 186, 814–830. [Google Scholar] [CrossRef]
  8. Marongiu, F.; Soprani, S.; Engelbrecht, K. Modeling of high temperature thermal energy storage in rock beds—Experimental comparison and parametric study. Appl. Therm. Eng. 2019, 163, 114355. [Google Scholar] [CrossRef]
  9. Zanganeh, G.; Pedretti, A.; Zavattoni, S.; Barbato, M.; Steinfeld, A. Packed-bed thermal storage for concentrated solar power—Pilot-scale demonstration and industrial-scale design. Sol. Energy 2012, 86, 3084–3098. [Google Scholar] [CrossRef]
  10. Nahhas, T.; Py, X.; Sadiki, N. Experimental investigation of basalt rocks as storage material for high- temperature concentrated solar power plants. Renew. Sustain. Energy Rev. 2019, 110, 226–235. [Google Scholar] [CrossRef]
  11. Kocak, B.; Paksoy, H. Performance of laboratory scale packed-bed thermal energy storage using new demolition waste based sensible heat materials for industrial solar applications. Sol. Energy 2020, 211, 1335–1346. [Google Scholar] [CrossRef]
  12. Sharma, A.; Pandey, P.K.; Didwania, M. Techno-economic optimization of packed-bed thermal energy storage system combined with CSP plant using DOE: Design of experiment technique and Taguchi method. Int. J. Energy Environ. Eng. 2022, 14, 287–308. [Google Scholar] [CrossRef]
  13. Allen, K.; Heller, L.; von Backström, T. Cost Optimum Parameters for Rock Bed Thermal Storage at 550–600 C: A Parametric Study. J. Sol. Energy Eng. 2016, 138, 061004. [Google Scholar] [CrossRef]
  14. Abdulla, A.; Reddy, K.S. Effect of operating parameters on thermal performance of molten salt packed-bed thermocline thermal energy storage system for concentrating solar power plants. Int. J. Therm. Sci. 2017, 121, 30–44. [Google Scholar] [CrossRef]
  15. Ozturk, M.; Dincer, I. Thermodynamic analysis of a solar-based multi-generation system with hydrogen production. Appl. Therm. Eng. 2013, 51, 1235–1244. [Google Scholar] [CrossRef]
  16. Cabrera, F.J.; Fernández-García, A.; Silva, R.M.P.; Pérez-García, M. Use of parabolic trough solar collectors for solar refrigeration and air-conditioning applications. Renew. Sustain. Energy Rev. 2013, 20, 103–118. [Google Scholar] [CrossRef]
  17. Pelay, U.; Luo, L.; Fan, Y.; Stitou, D.; Rood, M. Thermal energy storage systems for concentrated solar power plants. Renew. Sustain. Energy Rev. 2017, 79, 82–100. [Google Scholar] [CrossRef]
  18. Freeman, J.; Guarracino, I.; Kalogirou, S.A.; Markides, C.N. A small-scale solar organic Rankine cycle combined heat and power system with integrated thermal energy storage. Appl. Therm. Eng. 2017, 127, 1543–1554. [Google Scholar] [CrossRef]
  19. Çam, N.Y.; Alptekin, E.; Bilir, L.; Ezan, M.A. Thermal behavior of a solar-assisted latent heat thermal energy storage unit on the heating season under variable weather conditions. J. Energy Storage 2022, 52, 104934. [Google Scholar] [CrossRef]
  20. Khamlich, I.; Zeng, K.; Flamant, G.; Baeyens, J.; Zou, C.; Li, J.; Yang, X.; He, X.; Liu, Q.; Yang, H.; et al. Technical and economic assessment of thermal energy storage in concentrated solar power plants within a spot electricity market. Renew. Sustain. Energy Rev. 2021, 139, 110583. [Google Scholar] [CrossRef]
  21. Al-Nimr, M.A.; Bukhari, M.; Mansour, M. A combined CPV/T and ORC solar power generation system integrated with geothermal cooling and electrolyser/fuel cell storage unit. Energy 2017, 133, 513–524. [Google Scholar] [CrossRef]
  22. Shakouri, A.; Gorjian, S.; Ghobadian, B. Energy, exergy, and exergoeconomic (3E) evaluation of a hybrid multigeneration system based on a solar tower. Appl. Therm. Eng. 2024, 252, 123660. [Google Scholar] [CrossRef]
  23. Kuterbekov, K.; Kabyshev, A.; Bekmyrza, K.; Kubenova, M. Energy, exergy and thermo-economics analyses of hybrid solar, steam turbine and biomass gasification system for hydrogen production by polymer membrane electrolyzer. Int. J. Thermofluids 2024, 21, 100556. [Google Scholar] [CrossRef]
  24. Zhang, S.; Liu, M.; Zhao, Y.; Liu, J.; Yan, J. Energy and exergy analyses of a parabolic trough concentrated solar power plant using molten salt during the start-up process. Energy 2022, 254, 124480. [Google Scholar] [CrossRef]
  25. Wang, G.; He, D.; Lin, J.; Jiang, T.; Chen, Z. Exergy estimate of a novel hybrid solar-gas power and organic Rankine cycle-based hydrogen-production system. Case Stud. Therm. Eng. 2022, 33, 124480. [Google Scholar] [CrossRef]
  26. Moran, M.J.; Shapiro, H.N.; Boettner, D.D.; Bailey, M.B. Fundamentals of Engineering Thermodynamics; John Wiley & Sons: Hoboken, NJ, USA, 2020. [Google Scholar] [CrossRef]
  27. Aghaziarati, Z.; Aghdam, A.H. Thermoeconomic analysis of a novel combined cooling, heating and power system based on solar organic Rankine cycle and cascade refrigeration cycle. Renew. Energy 2021, 164, 1267–1283. [Google Scholar] [CrossRef]
  28. Tozlu, A.; Kayabasi, E.; Ozcan, H. Thermoeconomic analysis of a low-temperature waste-energy assisted power and hydrogen plant at off-NG grid region. Sustain. Energy Technol. Assess. 2022, 52, 102104. [Google Scholar] [CrossRef]
  29. Yan, M.; Yao, Z.; Nutakki, T.U.K.; Agrawal, M.K.; Muhammad, T.; Albani, A.; Zhao, Z. Design and evaluation of a novel heliostat-based combined cooling, heating, and power (CCHP) system: 3E analysis and multi-criteria optimization by response surface methodology (RSM). Energy 2023, 285, 129389. [Google Scholar] [CrossRef]
  30. Barakat, S.; Ramzy, A.; Hamed, A.M.; El Emam, S.H. Enhancement of gas turbine power output using earth to air heat exchanger (EAHE) cooling system. Energy Convers Manag. 2016, 111, 137–146. [Google Scholar] [CrossRef]
  31. Abudu, K.; Igie, U.; Roumeliotis, I.; Hamilton, R. Impact of gas turbine flexibility improvements on combined cycle gas turbine performance. Appl. Therm. Eng. 2021, 189, 116703. [Google Scholar] [CrossRef]
  32. Akroot, A.; Namli, L. Performance assessment of an electrolyte-supported and anode-supported planar solid oxide fuel cells hybrid system. J. Ther. Eng. 2021, 7, 1921–1935. [Google Scholar]
  33. ABejan GTsatsaronis, M.M. Thermal Design and Optimization; John Wiley Sons Inc.: New York, NY, USA, 1996; Volume 21. [Google Scholar] [CrossRef]
  34. Elmorsy, L.; Morosuk, T.; Tsatsaronis, G. Exergy-based analysis and optimization of an integrated solar combined-cycle power plant. Entropy 2020, 22, 655. [Google Scholar] [CrossRef] [PubMed]
  35. Ahmadi, G.; Jahangiri, A.; Toghraie, D. Design of heat recovery steam generator (HRSG) and selection of gas turbine based on energy, exergy, exergoeconomic, and exergo-environmental prospects. Process Saf. Environ. Prot. 2023, 172, 353–368. [Google Scholar] [CrossRef]
  36. Talal, W.; Akroot, A. Exergoeconomic Analysis of an Integrated Solar Combined Cycle in the Al-Qayara Power Plant in Iraq. Processes 2023, 11, 656. [Google Scholar] [CrossRef]
  37. Nourpour, M.; Khoshgoftar Manesh, M.H. Evaluation of novel integrated combined cycle based on gas turbine-SOFC-geothermal-steam and organic Rankine cycles for gas turbo compressor station. Energy Convers. Manag. 2022, 252, 115050. [Google Scholar] [CrossRef]
  38. Ozcan, H.; Kayabasi, E. Thermodynamic and economic analysis of a synthetic fuel production plant via CO2 hydrogenation using waste heat from an iron-steel facility. Energy Convers. Manag. 2021, 236, 114074. [Google Scholar] [CrossRef]
  39. Mohammadkhani, F.; Shokati, N.; Mahmoudi, S.M.S.; Yari, M.; Rosen, M.A. Exergoeconomic assessment and parametric study of a Gas Turbine-Modular Helium Reactor combined with two Organic Rankine Cycles. Energy 2014, 65, 533–543. [Google Scholar] [CrossRef]
  40. Kareem, A.F.; Akroot, A.; Abdul Wahhab, H.A.; Talal, W.; Ghazal, R.M.; Alfaris, A. Exergo–Economic and Parametric Analysis of Waste Heat Recovery from Taji Gas Turbines Power Plant Using Rankine Cycle and Organic Rankine Cycle. Sustainability 2023, 15, 9376. [Google Scholar] [CrossRef]
  41. Elsafi, A.M. Exergy and exergoeconomic analysis of sustainable direct steam generation solar power plants. Energy Convers. Manag. 2015, 103, 338–347. [Google Scholar] [CrossRef]
  42. Miao, X.; Zhang, H.; Zhao, S.; Zhang, Q.; Xia, Y. An innovative S–CO2 recompression Brayton system and its thermodynamic, exergoeconomic and multi-objective analyses for a nuclear spacecraft. Case Stud. Therm. Eng. 2024, 53, 103805. [Google Scholar] [CrossRef]
  43. Besevli, B.; Kayabasi, E.; Akroot, A.; Talal, W.; Alfaris, A.; Assaf, Y.H.; Nawaf, M.Y.; Bdaiwi, M.; Khudhur, J. Technoeconomic Analysis of Oxygen-Supported Combined Systems for Recovering Waste Heat in an Iron-Steel Facility. Appl. Sci. 2024, 14, 2563. [Google Scholar] [CrossRef]
  44. Alsunousi, M.; Kayabasi, E. Techno-economic assessment of a floating photovoltaic power plant assisted methanol production by hydrogenation of CO2 captured from Zawiya oil refinery. Int. J. Hydrog. Energy 2024, 57, 589–600. [Google Scholar] [CrossRef]
  45. Akroot, A.; Al Shammre, A.S. Techno-Economic and Environmental Impact Analysis of a 50 MW Solar-Powered Rankine Cycle System. Processes 2024, 12, 1059. [Google Scholar] [CrossRef]
  46. Abdelhay, A.O.; Fath, H.E.S.; Nada, S.A. Case Studies in Thermal Engineering Enviro-exergo-economic analysis and optimization of a nanofiltration-multi effect desalination, power generation and cooling in an innovative trigeneration plant. Case Stud. Therm. Eng. 2022, 31, 101857. [Google Scholar] [CrossRef]
  47. Tian, H.; Li, R.; Salah, B.; Thinh, P.H. Bi-objective optimization and environmental assessment of SOFC-based cogeneration system: Performance evaluation with various organic fluids. Process Saf. Environ. Prot. 2023, 178, 311–330. [Google Scholar] [CrossRef]
  48. Akroot, A.; Almaktar, M.; Alasali, F. The Integration of Renewable Energy into a Fossil Fuel Power Generation System in Oil-Producing Countries: A Case Study of an Integrated Solar Combined Cycle at the Sarir Power Plant. Sustainability 2024, 16, 4820. [Google Scholar] [CrossRef]
  49. Nondy, J.; Gogoi, T.K. 4E analyses of an intercooled-recuperative gas turbine-based CCHP system: Parametric analysis and tri-objective optimization. Therm. Sci. Eng. Prog. 2023, 39, 101719. [Google Scholar] [CrossRef]
Figure 1. Combined cycle power plant (CCPP) schematic diagram.
Figure 1. Combined cycle power plant (CCPP) schematic diagram.
Processes 12 01433 g001
Figure 2. Temperature profile in the boiler.
Figure 2. Temperature profile in the boiler.
Processes 12 01433 g002
Figure 3. Comparative analysis between the present thermodynamic model vs. Oztürk et al.’s published work [15].
Figure 3. Comparative analysis between the present thermodynamic model vs. Oztürk et al.’s published work [15].
Processes 12 01433 g003
Figure 4. Exergy degradation rates of system components.
Figure 4. Exergy degradation rates of system components.
Processes 12 01433 g004
Figure 5. The impact of the ambient temperature (T1) change on the system’s work output, power cost rate, and overall efficiency.
Figure 5. The impact of the ambient temperature (T1) change on the system’s work output, power cost rate, and overall efficiency.
Processes 12 01433 g005
Figure 6. The impact of pressure ratio (PR) changes on the system’s work output, power cost rate, and overall efficiency.
Figure 6. The impact of pressure ratio (PR) changes on the system’s work output, power cost rate, and overall efficiency.
Processes 12 01433 g006
Figure 7. The impact of the gas turbine inlet temperature (T4) changes on the system’s work output, energy cost rate, and overall efficiencies.
Figure 7. The impact of the gas turbine inlet temperature (T4) changes on the system’s work output, energy cost rate, and overall efficiencies.
Processes 12 01433 g007
Figure 8. Changes in solar and fuel shares across each month.
Figure 8. Changes in solar and fuel shares across each month.
Processes 12 01433 g008
Figure 9. Changes in the CCPP’s work net across each month.
Figure 9. Changes in the CCPP’s work net across each month.
Processes 12 01433 g009
Figure 10. Changes in the CCPP’s electricity cost rate across each month.
Figure 10. Changes in the CCPP’s electricity cost rate across each month.
Processes 12 01433 g010
Figure 11. Changes in CO2 emissions across each month.
Figure 11. Changes in CO2 emissions across each month.
Processes 12 01433 g011
Table 1. Summary of reviewed research on thermal energy storage devices and their potential integration into power plant operations.
Table 1. Summary of reviewed research on thermal energy storage devices and their potential integration into power plant operations.
Author(s)FocusKey Findings
Soprani et al. [6]High-temperature TES prototype using diabase rocks and air for heat transferAnalyzed temperature and flow patterns during charging and discharging for a system with 450 kWh capacity and operated at 600 °C
Desai et al. [7]Cost-effective concentrated solar power system with micro-structured polymer foilCompared three thermal energy storage options, optimal design for cogeneration systems
Marongiu et al. [8]2-D numerical model of a rock bed considering temperature-dependency of air and rocksEvaluated design factors like rock size and type on system efficiency
Sharma et al. [12]Packed-bed TES systems with different materials, identifying hybrid particles as the most effectiveIdentified hybrid particles as the most effective storage medium
Heller et al. [13]Rock bed heat transfer model for energy storage and transmission to steam cyclesSimulations examined the influence on power plant economics
Abdulla et al. [14]Heat transfer in molten salt-packed-bed system for CSP plantsIdentified key parameters for efficient energy discharge
Öztürk et al. [15]Combined gas and steam turbine system with rock bed TES for diversified heating and power applicationsConducted extensive analyses on performance and efficiency under variable conditions
Fernández et al. [16]Operation strategies for packed bed TES systems to enhance efficiency and thermal storage behaviorEnhanced efficiency and thermal storage behavior
Pelay et al. [17]TES systems in CSP plants, discussing high-temperature TES technologies and integration strategiesDiscussed global status of CSP, various high-temperature TES technologies, and new integration strategies
Freeman et al. [18]TES solutions for domestic solar combined heat and power systems using ORC engines and solar–thermal collectorsTES enhances system efficiency, aligns with user demands, and reduces costs
Çam et al. [19]Solar-assisted heat pump-integrated latent heat TES system for variable-temperature heating using PCMsInvestigated PCMs to improve solar energy intermittency and energy efficiency
Khamlich et al. [20]Techno-economic analysis of a 100 MWe CSP plant with various TES configurationsAssessed TES integration into spot market for economic performance
Shakouri et al. [22]Multigeneration system integrating conventional and renewable energy sourcesEnhanced efficiency and sustainability with significant reductions in fuel use and CO2 emissions
Zhang et al. [24]Energy and exergy losses during the startup of PTC solar power plants using molten saltMolten salt anti-freezing solution reduces startup energy consumption
Wang et al. [25]Hybrid hydrogen production and gas-solar power system integrating solar tower, gas-steam turbine, and ORCAchieved 103.9 MW power and 41.3% efficiency with stable operations
Table 2. Operation conditions employed in the designed system.
Table 2. Operation conditions employed in the designed system.
ComponentParameterValue
ACCompression ratio14.5
Ambient temperature33.65 °C
Ambient pressure101.3 kPa
Air flow rate52 kg/s
Isentropic efficiency84%
Heliostats fieldArea53,935 m2
DNI8.02 (kW-h/m2/day)
LocationNeom, Saudi Arabia
Latitude28.0064°
Longitude35.2025°
GTInlet temperature1000 °C
Isentropic efficiency85%
RBESTime of charge10 h
Time of discharge14 h
STInlet pressure5000 kPa
Isentropic efficiency80%
PumpIsentropic efficiency90%
CondenserCondenser temperature60 °C
Table 3. The equations related to energy and exergy balance in a CCPP.
Table 3. The equations related to energy and exergy balance in a CCPP.
ComponentEnergy Balance EquationExergy Balance Equation
AC m ˙ 1   h 1 + W ˙ A C = m ˙ 2   h 2 E ˙ D , A C = E ˙ 1 E ˙ 2 + W ˙ A C
SR m ˙ 2   h 2 + Q ˙ S R = m ˙ 3   h 3 E ˙ D , S R = E ˙ 2 E ˙ 3 + E ˙ Q , S o l a r
CC m ˙ 3   h 3 + η C C m ˙ f L H V f = m ˙ 4 h 4 E ˙ D , C C = E ˙ 3 + E ˙ Q , F u e l E ˙ 4
GT m ˙ 4 h 4 = m ˙ 5 h 5 + W ˙ G T E ˙ D , G T = E ˙ 4 E ˙ 5 W ˙ G T
RBES
Charging
m 5 h 5 m 6 h 6 × 3600 × t c h a r g i n g = m r o c k   c r o c k T 5 T 0 E ˙ 5 E ˙ 6 × 3600 × t charging   = m rock   c rock   T 5 T 0   T 0 × ln T 5 T 0 + E ˙ D , charging × 3600 × t charging  
RBES
Discharging
m 7 h 7 m 9 h 9 × 3600 × t discharging   = m r o c k   c r o c k T 5 T 0 E ˙ 7 E ˙ 9 × 3600 × t discharging   + E ˙ D , discharging × 3600   ×   t discharging   = m rock   c rock   T 5 T 0 T 0 × ln T 5 / T 0
Blower m ˙ 8   h 8 + W ˙ b l o w e r = m ˙ 9   h 9 E ˙ D , b l o w e r = E ˙ 8 E ˙ 9 + W ˙ b l o w e r
Boiler m ˙ 7 h 7 h 8 = m ˙ 10 h 10 h 13 E ˙ D , B o i l e r = E ˙ 7 E ˙ 8 + E ˙ 13 E ˙ 10
ST m ˙ 10 h 10 = m ˙ 11 h 11 + W ˙ S T E ˙ D , S T = E ˙ 10 E ˙ 11 W ˙ S T
Condenser m ˙ 11 h 11 h 12 = m ˙ 14 h 15 h 14 E ˙ D , c o n d = E ˙ 11 E ˙ 12 + E ˙ 14 E ˙ 15
Pump m ˙ 12   h 12 + W ˙ p u m p = m ˙ 13   h 13 E ˙ D , p u m p = E ˙ 12 E ˙ 13 + W ˙ p u m p
Table 4. The equations related to product and fuel exergy.
Table 4. The equations related to product and fuel exergy.
ComponentFuel Exergy EquationProduct Exergy Equation
AC W ˙ A C E ˙ 2 E ˙ 1
SR E ˙ Q , S o l a r E ˙ 3 E ˙ 2
CC E ˙ Q , F u e l E ˙ 4 E ˙ 3
GT E ˙ 4 E ˙ 5 W ˙ G T
RBES charging E ˙ 5 E ˙ 6 m rock   c rock   T 5 T 0
T 0 × ln T 5 T 0 / ( 3600 × t charging )
RBES discharging m rock   c rock   T 5 T 0
T 0 × ln T 5 T 0 / ( 3600 × t discharging )
E ˙ 7 E ˙ 9
Blower W ˙ b l o w e r E ˙ 9 E ˙ 8
Boiler E ˙ 7 E ˙ 8 E ˙ 10 E ˙ 13
ST E ˙ 10 E ˙ 11 W ˙ S T
Condenser E ˙ 11 E ˙ 12 E ˙ 15 E ˙ 14
Pump W ˙ p u m p E ˙ 13 E ˙ 12
Table 5. Equations for cost balancing system elements.
Table 5. Equations for cost balancing system elements.
ComponentCost Flow EquationsAuxiliary Equations
AC C ˙ 1 + C ˙ A C + Z ˙ A C = C ˙ 2 c 1 = 0
SR C ˙ 2 + Z ˙ h e l = C ˙ 3 c 2 = c 3
CC C ˙ 3 + C ˙ f u e l + Z ˙ C C = C ˙ 4 c 3 = c 4
c f = 12
GT C ˙ 4 + Z ˙ G T = C ˙ 5 + C ˙ G T c 4 = c 5
c A C = c G T
RBES C ˙ 5 + C ˙ 9 + Z ˙ R B E S = C ˙ 6 + C ˙ 7 c 5 = c 6
Blower C ˙ 9 + C ˙ B l o w e r + Z ˙ B l o w e r = C ˙ 10 c b l o w e r = c G T
Boiler C ˙ 7 + C ˙ 13 + Z ˙ B o i l e r = C ˙ 8 + C ˙ 10 c 7 = c 8
ST C ˙ 10 + Z ˙ S T = C ˙ 11 + C ˙ S T c 10 = c 11
Condenser C ˙ 11 + C ˙ 14 + Z ˙ C o n d e n s e r = C ˙ 12 + C ˙ 15 c 12 = c 13
c 14 = 0
Pump C ˙ 12 + C ˙ P u m p + Z ˙ p u m p = C ˙ 13 c P u m p = c S T
Table 6. Thermodynamic characteristics of the solar-powered combined cycle in each of its ideal states.
Table 6. Thermodynamic characteristics of the solar-powered combined cycle in each of its ideal states.
State m ˙
(kg/s)
T
(K)
P
(kPa)
h
(kJ/kg)
S
(KJ/kg. K)
E ˙
(MW)
C ˙
($/h)
c ($/GJ)
152306.7101.3307.15.7240.063300
252711.71469726.25.82320.33117.41.604
3521172146912456.38338.64223.11.604
4521273146913646.4843.31427.52.742
552752.1108.87706.6310.199.682.742
652382.3101.3383.45.9460.58485.7732.742
737.51742.1107759.16.626.986113.24.503
837.51343.2107343.95.8220.33345.4044.503
937.51375.3140376.35.8351.40319.043.769
105.092722.1500033156.8186.557155.66.592
115.09233319.8124597.461.22229.016.592
125.09233319.81250.60.82940.04070.9666.592
135.092333.25000256.20.83110.06672.6110.86
14269306.7101.3140.40.48470.13900
15269316.7101.3182.20.61890.631528.1912.4
Table 7. An examination of the exergy analysis of the CCPP.
Table 7. An examination of the exergy analysis of the CCPP.
Component E ˙ i n p u t
(MW)
E ˙ o u t p u t
(MW)
E ˙ d e s t r u c t i o n
(MW)
E ˙ d e s t r u c t i o n
(%)
Exergy Efficiency (%)
AC21.9720.261.5289.23792.99
Blower1.2151.070.14550.8888.03
Boiler6.6536.490.16250.9897.56
CC6.3924.6681.72410.41173.04
Condenser1.1820.49250.68934.16541.67
GT33.2130.92.31413.9893.03
Pump0.028630.02610.002560.01591.06
RBES15.1513.471.67710.13288.9
RBES charging9.5147.891.6249.81282.93
RBES discharging5.6355.5830.0530.3299.07
SR25.6618.327.34644.443.94
ST5.3344.3590.95725.881.72
Total 16.55
Table 8. Designed cycle’s exergoeconomic characteristics.
Table 8. Designed cycle’s exergoeconomic characteristics.
Component c f
($/GJ)
c p
($/GJ)
C ˙ D
($/h)
Z ˙ K
($/h)
Z ˙ K + C ˙ D
($/h)
f
(%)
AC1.0221.6095.62537.18542.8186.86
Blower3.0613.541.6030.2441.84713.18
Boiler4.5036.5492.63445.33647.9794.94
CC8.87512.1655.070.1655.230.302
Condenser6.59215.916.360.1416.50.8516
GT2.7243.06122.8412.6235.4635.59
Pump10.1717.520.09370.59620.689986.42
RBES1.7221.94210.390.310.692.816
SR-1.604-105.8105.8100
ST6.59210.1723.1532.9456.0958.74
Total system 137.76235.32373.0863.07
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Akroot, A.; Al Shammre, A.S. Economic and Technical Assessing the Hybridization of Solar Combined Cycle System with Fossil Fuel and Rock Bed Thermal Energy Storage in Neom City. Processes 2024, 12, 1433. https://doi.org/10.3390/pr12071433

AMA Style

Akroot A, Al Shammre AS. Economic and Technical Assessing the Hybridization of Solar Combined Cycle System with Fossil Fuel and Rock Bed Thermal Energy Storage in Neom City. Processes. 2024; 12(7):1433. https://doi.org/10.3390/pr12071433

Chicago/Turabian Style

Akroot, Abdulrazzak, and Abdullah Sultan Al Shammre. 2024. "Economic and Technical Assessing the Hybridization of Solar Combined Cycle System with Fossil Fuel and Rock Bed Thermal Energy Storage in Neom City" Processes 12, no. 7: 1433. https://doi.org/10.3390/pr12071433

APA Style

Akroot, A., & Al Shammre, A. S. (2024). Economic and Technical Assessing the Hybridization of Solar Combined Cycle System with Fossil Fuel and Rock Bed Thermal Energy Storage in Neom City. Processes, 12(7), 1433. https://doi.org/10.3390/pr12071433

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

Article Metrics

Back to TopTop